Skip to main content
Springer logoLink to Springer
. 2023 Feb 7;9(1):4. doi: 10.1007/s40818-022-00144-3

Naked Singularities in the Einstein-Euler System

Yan Guo 1, Mahir Hadzic 2,, Juhi Jang 3,4
PMCID: PMC9905257  PMID: 36778526

Abstract

In 1990, based on numerical and formal asymptotic analysis, Ori and Piran predicted the existence of selfsimilar spacetimes, called relativistic Larson-Penston solutions, that can be suitably flattened to obtain examples of spacetimes that dynamically form naked singularities from smooth initial data, and solve the radially symmetric Einstein-Euler system. Despite its importance, a rigorous proof of the existence of such spacetimes has remained elusive, in part due to the complications associated with the analysis across the so-called sonic hypersurface. We provide a rigorous mathematical proof. Our strategy is based on a delicate study of nonlinear invariances associated with the underlying non-autonomous dynamical system to which the problem reduces after a selfsimilar reduction. Key technical ingredients are a monotonicity lemma tailored to the problem, an ad hoc shooting method developed to construct a solution connecting the sonic hypersurface to the so-called Friedmann solution, and a nonlinear argument to construct the maximal analytic extension of the solution. Finally, we reformulate the problem in double-null gauge to flatten the selfsimilar profile and thus obtain an asymptotically flat spacetime with an isolated naked singularity.

Keywords: Self-similar solutions, Einstein-Euler system, Implosion, Naked singularities

Introduction

We study the Einstein-Euler system which couples the Einstein field equations to the Euler equations of fluid mechanics. The unknowns are the 4-dimensional Lorentzian spacetime (M,g), the fluid pressure p, the mass-density ρ, and the 4-velocity uα. In an arbitrary coordinate system, the Einstein-Euler equations read

Ricαβ-12Rgαβ=Tαβ,(α,β=0,1,2,3), 1.1
αTαβ=0,(β=0,1,2,3), 1.2
gαβuαuβ=-1, 1.3

where Ricαβ is the Ricci curvature tensor, R the scalar curvature of gαβ, and Tαβ is the energy momentum tensor given by the formula

Tαβ=(ρ+p)uαuβ+pgαβ,(α,β=0,1,2,3). 1.4

To close the system, we assume the linear equation of state

p=ερ, 1.5

where 0<ε<1 corresponds to the square of the speed of sound.

The system (1.1)–(1.5) is a fundamental model of a selfgravitating relativistic gas. We are interested in the existence of selfsimilar solutions to (1.1)–(1.5) under the assumption of radial symmetry. This amounts to the existence of a homothetic Killing vector field ξ with the property

Lξg=2g, 1.6

where the left-hand side is the Lie derivative of the metric g. The presence of such a vector field induces a scaling symmetry, which allows us to look for selfsimilar solutions to (1.1)–(1.5). Study of selfsimilar solutions to Einstein-matter systems has a rich history in the physics literature. They in particular provide a way of constructing spacetimes with so-called naked singularities, a notion intimately tied to the validity of the weak cosmic censorship of Penrose [30], see the discussion in [7, 10, 32]. Naked singularities intuitively correspond to spacetime singularities that are “visible" to far away observers, which informally means that there exists a future outgoing null-geodesic “emanating" from the singularity and reaching the asymptotically flat region of the spacetime. We adopt here a precise mathematical definition from the work of Rodnianski and Shlapentokh-Rothman [32, Definition 1.1], which in turn is related to a formulation of weak cosmic censorship by Christodoulou [7].

In the absence of pressure (ε=0 in (1.5)) and under the assumption of radial symmetry, the problem simplifies considerably. The corresponding family of solutions was studied by Lemaître [21] and Tolman [34] in early 1930s (see also [1]). In their seminal work from 1939, Oppenheimer and Snyder [25] studied the causal structure of a subclass of Lemaître-Tolman solutions with space-homogeneous densities, thus exhibiting the first example of a dynamically forming (what later became known as) black hole. However, in 1984 Christodoulou [5] showed that, within the larger class of Lemaître-Tolman solutions with space-inhomogeneous densities, black holes are exceptional and instead naked singularities form generically.

Of course, in the context of astrophysics, one expects the role of pressure to be very important in the process of gravitational collapse for relativistic gases. In the late stages of collapse, the core region is expected to be very dense and the linear equation of state (1.5) is commonly used in such a setting, as it is compatible with the requirement that the speed of sound is smaller than the speed of light, ε<1. In their pioneering works, Ori and Piran [2628] found numerically selfsimilar solutions to (1.1)–(1.5), which are the relativistic analogues of the Larson-Penston selfsimilar collapsing solutions to the isothermal Euler-Poisson system, see [15, 20, 31]. Through both numerical and asymptotic analysis methods Ori and Piran investigated the causal structure of such relativistic Larson-Penston solutions, ascertaining the existence of spacetimes with naked singularities when ε is smaller than a certain value. Our main goal is to justify the findings of Ori and Piran on rigorous mathematical grounds.

Broadly speaking, this manuscript consists of two parts. In the first part, which constitutes the bulk of our work, we construct a selfsimilar solution of the Einstein-Euler system (Sections 38), assuming that ε – the square of the speed of sound – is sufficiently small.

Theorem 1.1

(Existence of the relativistic Larson-Penston spacetimes) For any sufficiently small 0<ε1 there exists a radially symmetric real-analytic selfsimilar solution to the Einstein-Euler system with a curvature singularity at the scaling origin and an outgoing null-geodesic emanating from it all the way to infinity. The resulting spacetime is called the relativistic Larson-Penston (RLP) spacetime.

It is not hard to see that the selfsimilar solution constructed in Theorem 1.1 is not asymptotically flat. In the second step (Section 9), using PDE techniques, we flatten the selfsimilar RLP-profile in a region away from the singularity and thus obtain an asymptotically flat solution with a naked singularity. Thus, our main theorem states that in the presence of pressure there do exist examples of naked singularities which form from smooth data.

Theorem 1.2

(Existence of naked singularities) For sufficiently small 0<ε1 there exist radially symmetric asymptotically flat solutions to the Einstein-Euler system that form a naked singularity in the sense of [32, Definition 1.1].

Other than the dust-Einstein model mentioned above, we are aware of two other rigorous results on the existence of naked singularities. In 1994 Christodoulou [6] provided a rigorous proof of the existence of radially symmetric solutions to the Einstein-scalar field system, which contain naked singularities (see also [8] for the proof of their instability). Very recently, Rodnianski and Shlapentokh-Rothman [32] proved the existence of solutions to the Einstein-vacuum equations which contain naked singularities and are (necessarily) not radially symmetric.

In the physics literature much attention has been given to selfsimilar solutions and naked singularities for the Einstein-Euler system, see for example [2, 14]. A selfsimilar reduction of the problem was first given in [33]. As explained above, a detailed analysis of the resulting equations, including the discussion of naked singularities, was given in [2628]. Subsequent to [28], a further analysis of the causal structure, including the nonradial null-geodesics was presented in [19], see also [4]. There exist various approaches to the existence of solutions to the selfsimilar problem, most of them rely on numerics [3, 17, 28]. A dynamical systems approach with a discussion of some qualitative properties of the solutions was developed in [3, 13]. Numerical investigation of the stability of the RLP-spacetimes can be found in [17, 18]. selfsimilar relativistic perfect fluids play an important role in the study of the so-called critical phenomena - we refer to reviews [14, 24].

The proof of Theorem 1.1 relies on a careful study of the nonlinear invariances of the finite-dimensional non-autonomous dynamical system obtained through the selfsimilar reduction. The solutions we construct are real-analytic in a suitable choice of coordinates. A special role is played by the so-called sonic line (sonic point), the boundary of the backward sound cone emanating from the scaling origin O. Many difficulties in the proof of Theorem 1.1 originate from possible singularities across this line, which together with the requirement of smoothness, puts severe limitations on the possible space of smooth selfsimilar solutions. We are aware of no general ODE theory for global existence in the presence of singular sonic points. Therefore, our proofs are all based on continuity arguments, where we extract many delicate invariant properties of the nonlinear flow, specific to the ODE system at hand. In particular, the discovery of a crucial monotonicity lemma enables us to apply an ad hoc shooting method, which was developed for the limiting Larson-Penston solution in the non-relativistic context by the authors. From the point of view of fluid mechanics, the singularity at O is an imploding one, as the energy density blows up on approach to O. It is in particular not a shock singularity.

The asymptotic flattening in Theorem 1.2 requires solving a suitable characteristic problem for the Einstein-Euler system formulated in the double-null gauge. We do this in a semi-infinite characteristic rectangular domain wherefrom the resulting solution can be glued smoothly to the exact selfsimilar solution in the region around the singularity O. The proof of Theorem 1.2 is given in Section 9.6.

Due to the complexity of our analysis, in Section 2 we give an extensive overview of our methods and key ideas behind the detailed proofs in Sections 39.

Methodology and Outline

Formulation of the Problem (Section 3)

Following [28] it is convenient to work with the comoving coordinates

g=-e2μ(τ,R)dτ2+e2λ(τ,R)dR2+r2(τ,R)γ, 2.7

where γ=γABdxAdxB is the standard metric on S2, xA, A=2,3 are local coordinates on S2, and r is the areal radius. The vector field τ is chosen in such a way that the four velocity uν is parallel to τ. The normalisation condition (1.3) then implies

u=e-μτ. 2.8

The coordinate R acts as a particle label. The coordinates (τ,R) are then uniquely determined by fixing the remaining gauge freedoms in the problem, the value of μ(τ,R)|R=0 and by setting r(-1,R)=R, which states that on the hypersurface τ=-1 the comoving label R coincides with the areal radius.

Introduce the radial velocity

V:=e-μτr, 2.9

the Hawking (also known as Misner-Sharp) mass

m(τ,R):=4π0r(τ,R)ρs2ds=4π0Rρr2RrdR¯, 2.10

and the mean density

G(τ,R):=m(τ,R)4π3r(τ,R)3=3r(τ,R)30Rρ(τ,R¯)r(τ,R¯)2Rr(τ,R¯)dR¯. 2.11

Recalling the equation of state (1.5), the spherically symmetric Einstein-Euler system in comoving coordinates reads (see [12, 23])

τρ+(1+ε)ρRVRr+2Vreμ=0, 2.12
τλ=eμRVRr, 2.13
e-μτV+ε1+εRre-2λρRρ+4πr13G+ερ=0, 2.14
(Rr)2e-2λ=1+V2-8π3Gr2, 2.15

where we recall (2.11). The well-known Tolman-Oppenheimer-Volkov relation reads (ρ+p)Rμ+Rp=0, which after plugging in (1.5) further gives the relation

Rμ=-ε1+εRρρ. 2.16

Comoving Selfsimilar Formulation

It is straightforward to check that the system (2.12)–(2.15) is invariant under the scaling transformation

ρa-2ρ(s,y),rar(s,y),VV(s,y),λλ(s,y),μμ(s,y), 2.17

where the comoving “time" τ and the particle label R scale according to

s=τa,y=Ra,a>0. 2.18

Motivated by the scaling invariance (2.17)–(2.18), we look for selfsimilar spacetimes of the form

ρ(τ,R)=12πτ2Σ(y), 2.19
r(τ,R)=-ετr~(y), 2.20
V(τ,R)=εV(y), 2.21
λ(τ,R)=λ(y), 2.22
μ(τ,R)=μ(y), 2.23
G(τ,R)=14πτ2G~(y)=32πτ2r~30yΣ(y~)r~2r~dy~, 2.24

where

y=R-ετ. 2.25

Associated with the comoving selfsimilar coordinates are the two fundamental unknowns:

d:=Σ1-ε1+ε, 2.26
w:=(1+ε)eμV+r~r~-ε. 2.27

For future use it is convenient to sometimes consider the quantity

χ(y):=r~(y)y, 2.28

instead of r~. Quantity d corresponds to the selfsimilar number density, while w is referred to as the relative velocity. It is shown in Proposition 3.4 that the radial Einstein-Euler system under the selfsimilar ansatz above reduces to the following system of ODE:

d=-2(1-ε)d(d-w)(1+ε)y(e2μ-2λy-2-1), 2.29
w=(w+ε)(1-3w)(1+ε)y+2w(d-w)y(e2μ-2λy-2-1). 2.30

This formulation of the selfsimilar problem highlights the danger from possible singularities associated with the vanishing of the denominators on the right-hand side of (2.29)–(2.30). Such points play a distinguished role in our analysis, and as we shall show shortly, are unavoidable in the study of physically interesting selfsimilar solutions.

Definition 2.1

(The sonic point) For any smooth solution to (2.29)–(2.30) we refer to a point y(0,) satisfying

y2=e2μ(y)-2λ(y) 2.31

as the sonic point.

Schwarzschild Selfsimilar Formulation

The comoving formulation (2.29)–(2.30) as written does not form a closed system of ODE. To do so, we must express the metric coefficients μ,λ as functions of d,w, which can be done at the expense of working with r~ (or equivalently χ) as a further unknown. To avoid this, it is possible to introduce the so-called Schwarzschild selfsimilar coordinate:

x:=r~(y) 2.32

so that

dxdy=r~=x(w+ε)y(1+ε). 2.33

In this coordinate system the problem takes on a form analogous to the Eulerian formulation of the selfsimilar Euler-Poisson system from [15]. It is shown in Lemma 3.5 that the new unknowns

D(x):=d(y),W(x):=w(y), 2.34

solve the system

D(x)=-2x(1-ε)D(W+ε)(D-W)B, 2.35
W(x)=(1-3W)x+2x(1+ε)W(W+ε)(D-W)B, 2.36

where

B=B[x;D,W]:=D-η-(W+ε)2-ε(W-1)2+4εDWx2, 2.37

and

η:=2ε1-ε. 2.38

The Greek letter η will always be used to mean (2.38) in the rest of the paper. In this formulation, sonic points correspond to zeroes of B=B[x;D,W], i.e. if y is a sonic point in the sense of Definition 2.1, then x:=r~(y) is a zero of the denominator B.

Friedmann, Far-Field, and the Necessity of the Sonic Point

There are two exact solutions to (2.35)–(2.36). The far-field solution

Df(x)=(1-ε)-21+ηx-21+η,Wf(x)=1, 2.39

features a density D that blows up at x=0 and decays to 0 as x. On the other hand, the Friedmann solution

DF(x)=13,WF(x)=13, 2.40

is bounded at x=0, but the density does not decay as x. Our goal is to construct a smooth solution to (2.35)-(2.36) which qualitatively behaves like the far-field solution as x and like the Friedmann solution as x0+. We can therefore think of it as a heteroclinic orbit for the dynamical system (2.35)-(2.36). It is then easy to see that any such solution has the property limxB=- and limx0+B(x)>0. By the intermediate value theorem there must exist a point where B vanishes, i.e. a sonic point.

It is important to understand the formal Newtonian limit, which is obtained by letting ε=0 in (2.35)–(2.36). This yields the system

D~(x)=-2xD~W~(D~-W~)1-x2W~2, 2.41
W~(x)=(1-3W~)x+2xW~2(D~-W~)1-x2W~2, 2.42

which is precisely the selfsimilar formulation of the isothermal Euler-Poisson system. In [15] we showed that there exists a solution to (2.41)–(2.42) satisfying W~(0)=13, limxW~(x)=1, D~(0)>13, and D~xx-2.1 The behaviour of the relativistic solutions when 0<ε1 in the region x[0,) is modelled on this solution, called the Larson-Penston (LP) solution.

The system (2.35)–(2.37) is a non-autonomous 2×2 system of ODE which is at the heart of the proof of Theorem 1.1 and is used to show the existence of the RLP-solution in the region x[0,). As x, we are forced to switch back to a version of the comoving variables in order to extend the solution beyond x= in a unique way, see the discussion in Section 2.5. A version of the comoving formulation (2.29)–(2.30) plays a crucial role in that extension. In our analysis of the radial null-geodesics (Section 8) and the nonradial ones (Appendix B), we often switch between different choices of coordinates to facilitate our calculations.

Even though the smallness of ε is crucial for the validity of our estimates, we emphasise that we do not use perturbation theory to construct the relativistic solution, by for example perturbing away from the Newtonian one. Such an argument is a priori challenging due to the singular nature of the sonic point, as well as complications arising from boundary conditions.

We mention that selfsimilar imploding flows for the compressible Euler system, featuring a sonic point, were constructed recently in the pioneering work of Merle, Raphaël, Rodnianski, and Szeftel [22] - here the associated 2×2 dynamical system is autonomous. In the context of the Euler-Poisson system with polytropic gas law, selfsimilar collapsing solutions featuring a sonic point were recently constructed in [16].

The Sonic Point Analysis (Section 4)

The solution we are trying to construct is on one hand assumed to be smooth, but it also must feature an a priori unknown sonic point, which we name x. This is a singular point for the dynamical system and the assumption of smoothness therefore imposes a hierarchy of constraints on the Taylor coefficients of a solution in a neighbourhood of x. More precisely, we look for solutions (D,W) to (2.35)-(2.36) of the form

D=N=0DN(x-x)N,W=N=0WN(x-x)N. 2.43

It is clear that the first constraint reads D0=W0, as the numerator in (2.35) must vanish at x for the solution to be smooth. Together with the condition B(x)=0, we can show that for any ε>0 sufficiently small, D0=W0 is a function of x, which converges to 1x as ε0 in accordance with the limiting Newtonian problem (2.41)–(2.42). Our goal is to express DN,WN recursively in terms of D0,,DN-1, W0,,WN-1 and thus obtain a hierarchy of algebraic relations that allows to compute the Taylor coefficients up to an arbitrary order.

However, an intriguing dichotomy emerges. At the next order, one obtains a cubic equation for W1, see Lemma 4.7. One of the solutions is a “ghost" solution and therefore unphysical, while the remaining two roots, when real, are both viable candidates for W1, given as a function of D0 (and therefore x). This is related to an analogous dichotomy in the Newtonian case - where one choice of the root leads to so-called Larson-Penston-type (LP type) solutions, while the other choice of the root leads to Hunter-type solutions. Based on this, for any 0<ε1 sufficiently small, we select W1=W1(ε) to correspond to the choice of the branch converging to the LP-type coefficient as ε0. This is a de facto selection principle which allows us to introduce formal Taylor expansions of the relativistic Larson-Penston-type (RLP type), see Definition 4.11. Upon fixing the choice of W1 (and thereby D1), all the higher-order coefficients (DN,WN), N2, are then uniquely determined through a recursive relation, see Section 4.3. We mention that D1 and W1 cease to exist as real numbers before x reaches 2 from above, which led us to define xcrit(ε), see Lemma 4.10. This should be contrasted to the Newtonian problem where D1 and W1 are real-valued as x passes below 22. The existence of a forbidden range for x is related to the band structure in the space of all smooth solutions, see [13, 28].

Guided by the intuition developed in the construction of the (non-relativistic) Larson-Penston solution [15], our next goal is to identify the so-called sonic window - a closed interval [xmin,xmax] within which we will find a sonic point for the global solution of the ODE system on [0,). In fact, by Lemma 5.12 and Lemma 5.9, the set W<13 and the set W>12-2η=12+O(ε) are invariant under the flow to the left of the sonic point. Motivated by this, we choose xmax=xmax(ε)<3 so that the zero order coefficient W0 coincides with the Friedmann solution (2.40): W0|x=xmax=13 (see (4.245)) and fix xmin=2+δ0>xcrit for δ0>0 sufficiently small but independent of ε, so that W(x;xmin)>12-2η for some x<xmin (see (4.285)).

The main result of Section 4 is Theorem 4.18, which states that there exists an 0<ε01 sufficiently small such that for all 0<εε0 and for any choice of x in the sonic window, there in fact exists a local-in-x real analytic solution around x=x. The proof of this theorem relies on a delicate combinatorial argument, where enumeration of indices and N-dependent growth bounds for the coefficients (DN,WN) are moved to Appendix A.

Having fixed the sonic window [xmin,xmax][2,3], our strategy is to determine what values of x[xmin,xmax] allow for RLP-type solutions that exist on the whole real line. We approach this problem by splitting it into two subquestions. We identify those x which give global solutions to the left, i.e. all the way from x=x to x=0, and separately to the right, i.e. on [x,).

The Friedmann Connection (Section 5)

The main goal of Section 5 is to identify a value x¯[xmin,xmax], so that the associated local solution (D(·;x¯),W(·;x¯)) exists on [0,x¯] and is real analytic everywhere.

From the technical point of view the main obstruction to our analysis is the possibility that the flow features more than one sonic point. By the precise analysis around x=x in Section 4 we know that B is strictly positive/negative locally around x to the left/right respectively (and vanishes at x=x). Our strategy is to propagate these signs dynamically. To do so we develop a technical tool, referred to as the monotonicity lemma, even though it is an exact identity, see Lemma 3.7. It is a first order differential equation for a quantity f(x)=J[x;D]-xD with a source term depending on the solution, but with good sign properties in the regime we are interested in, hence – monotonicity lemma. Here J[x;D]-xW is a factor of B in (2.37): B=(1-ε)(J-xW)(J+2η(1+D)x+xW) (see Lemmas 3.63.7). Roughly speaking, the function f allows us to relate the sign of B to the sign of the difference (D-W) in a precise dynamic way, so that we eventually show that away from the sonic point D>W and B>0 to the left, while D<W and B<0 to the right of the sonic point for the relativistic Larson-Penston solution.

To construct the solution to the left of the sonic point we develop a shooting-type method, which we refer to as shooting toward Friedmann. Namely, the requirement of smoothness at x=0 is easily seen to imply W(0)=13, which precisely agrees with the value of WF, see (2.40).

The key idea is to separate the sonic window [xmin,xmax] into the sets of sonic points x that launch solutions W(·;x) which either stay above the Friedmann value WF=13 on its maximum interval of existence (s(x),x] or cross it, see Figure 1. This motivates the following definition.

Fig. 1.

Fig. 1

Schematic depiction of the shooting argument. Here x1X>13, x2X13. The critical point x¯ is obtained by sliding to the left in X13 until we reach the boundary of its first connected component X

Definition 2.2

(X>13, X13, X<13, and X) Let ε0>0 be a small constant introduced in Section 2.2 (see also Theorem 4.18). For any ε(0,ε0] and x[xmin,xmax] we consider the associated RLP-type solution (D(·;x),W(·;x)). We introduce the sets

X>13:=x[xmin,xmax]|infx(s(x),x)W(x;x)>13, 2.44
X13:=x[xmin,xmax]|x(s(x),x)such thatW(x;x)=13, 2.45
X<13:=x[xmin,xmax]|W(x;x)>13for allx(s(x),x)andinfx(s(x),x)W(x;x)13. 2.46

Finally, we introduce the fundamental set XX13 given by

X:=x[xmin,xmax]|x~X13for allx~[x,xmax]. 2.47

The basic observation is that solutions that correspond to the set X13 have the property that once they take on value 13, they never go back up above it. This is a nonlinear invariance of the flow, which guides our shooting argument idea. It is possible to show that both sets X>13 and X13 are non-empty. In Lemma 5.9 we show that xminX>13 and that for some κ>0, (xmax-κ,xmax]X13. Intuitively, we then slide down x starting from xmax, until we reach the first value of x that does not belong to X13, i.e. we let

x¯:=infxXx, 2.48

see Figure 1. This is the candidate for the value of x which gives a real-analytic solution on [0,x¯]. Using the nonlinear invariances of the flow and its continuity properties, in Proposition 5.15, we show that the solution (D(·;x¯),W(·,x¯)) exists on the semi-open interval (0,x¯].

To show that the solution is indeed analytic all the way to x=0, W(0;x¯)=13, and D(0;x¯)>13, we adapt the strategy developed for the classical LP solution in [15]. Using the method of upper and lower solutions (Definition 5.23), we show that there exists a choice of D0>13 such that the solution (D(·;x¯),W(·,x¯)) coincides with a unique real analytic solution to (2.35)–(2.36), with data D(0)=D0, W(0)=13, solving from x=0 to the right. Detailed account of this strategy is contained in Sections 5.35.4. Finally, combining the above results we can prove the central statement of Section 5:

Theorem 2.3

There exists an ε0>0 sufficiently small, such that for any 0<εε0, the solution of RLP-type to (2.35)–(2.36) launched at x¯ (defined by (2.48)) extends (to the left) to the closed interval [0,x¯], is real analytic, and satisfies W(0;x¯)=13, D(0;x¯)>13.

This theorem is formally proved at the very end of Section 5.4.

The Far-Field Connection (Section 6)

By contrast to establishing the existence of the Friedmann connection in Section 5, we show that for any choice of x in our sonic window [xmin,xmax], there exists a global solution to the right, i.e. we prove the following theorem:

Theorem 2.4

Let x[xmin,xmax]. There exists an 0<ε01 such that the unique RLP-type solution (W,D) exists globally to the right for all ε(0,ε0].

The proof relies on a careful study of nonlinear invariances of the flow, and once again the monotonicity properties encoded in Lemma 3.7 play a critical role in our proof. This is highlighted in Lemma 6.3, which allows us to propagate the negativity of the sonic denominator B to the right of the sonic point. Importantly, we may now let x=x¯ defined in (2.48) to obtain a real-analytic RLP-type solution defined globally on [0,). As it turns out, the obtained spacetime is not maximally extended, and to address this issue we need to understand the asymptotic behaviour of our solutions as x.

In Lemma 6.5 we show that solutions from Theorem 2.4 honour the asymptotic behaviour

limxW(x;x)=1,limxD(x;x)x21+η>0; 2.49

hence the name far-field connection, see (2.39). This is however not enough for the purposes of extending the solution beyond x=, as we also need sharp asymptotic behaviour of the relative velocity W. Working with the nonlinear flow (2.35)–(2.36), in Proposition 6.8 we show that the leading order behaviour of W is given by the relation

1-Wxx-11+η. 2.50

Maximal Analytic Extension (Section 7)

Asymptotic relations (2.49)–(2.50) suggest that our unknowns are asymptotically “regular" only if thought of as functions of x-11+η. In fact, it turns out to be more convenient to interpret this in the original comoving selfsimilar variable y, see (2.25). Due to (2.49) and (2.32)–(2.33), it is easy to see that asymptotically

xxy. 2.51

Moreover, by (2.49) and (2.34), we have d(y)yy-21+η. Furthermore by (3.101) we have the relation

e2μ=1(1+ε)2Σ-η1+η=1(1+ε)2d-η. 2.52

As a consequence of (2.49) we then conclude that

e2μyy4ε1+ε, 2.53

which implies that the metric g given by (2.7) becomes singular as y (or equivalently x). In Section 7 we show that this is merely a coordinate singularity, and the spacetime extends smoothly (in fact analytically in a suitable choice of coordinates) across the surface {(τ0,R)|R>0}.

Motivated by the above considerations, we switch to an adapted comoving chart (τ~,R), defined through

τ~(τ,R)=-ε-ε1+εR2ε1+ε(-τ)1-ε1+ε,τ<0,R>0. 2.54

We introduce the selfsimilar variable

Y:=-ετ~R, 2.55

which is then easily checked to be equivalent to the change of variables

Y=y-11+η,y=Y-1-η, 2.56

where we recall η=η(ε)=2ε1-ε. To formulate the extension problem, it is natural to define the new variables

χ(Y):=χ(y),d(Y):=d(y),w(Y):=w(y), 2.57

where we recall the fundamental variables d,w,χ from (2.26)–(2.28). Note that by (2.28) and (2.56) we have χ(Y)=Y1+ηr~(y). It is shown in Lemma 7.1 that the original system (2.29)–(2.30) in the new variables reads

d=2χ2Yd(w+ε)2(d-w)C, 2.58
w=-(w+ε)(1-3w)(1-ε)Y-2(1+ε)χ2(1-ε)Yw(w+ε)2(d-w)C, 2.59
χ=1-w(1-ε)Yχ, 2.60

where

C:=dY-2-ηY2-χ2(w+ε)2-ε(w-1)2+4εwd. 2.61

The first main result of Section 7 is Theorem 7.4, where we prove the local existence of a real analytic solution in an open neighbourhood of Y=0, which provides the local extension of the solution from Y=0+ to Y=0-. Initial data at Y=0 are read off from the asymptotic behaviour (2.49)–(2.50), see Remark 7.2. The most important result of the section is the maximal extension theorem of the solution to the negative Y-s:

Theorem 2.5

(Maximal extension) There exists an 0<ε01 sufficiently small such that for any ε(0,ε0] there exists a Yms<0 such that the unique solution to the initial value problem (2.58)–(2.60) exists on the interval (Yms,0], and

limY(Yms)-w(Y)=limY(Yms)-d(Y)=, 2.62
χ(Y)>0,Y(Yms,0], 2.63
limY(Yms)-χ(Y)=0. 2.64
Fig. 2.

Fig. 2

Schematic depiction of the behaviour of χ(Y) and w(Y) in the maximal extension. As Y approaches Yms from the right, χ approaches 0 and w blows up to

We see from the statement of the theorem that the maximal extension is characterised by the simultaneous blow-up of w,d, and 1χ at the terminal point Yms. By (2.55), in the adapted comoving chart, the point Yms coincides with the hypersurface

MSε:=(τ~,R)|τ~=1ε|Yms|R\{(0,0)},

which we refer to as the massive singularity, following the terminology in [28]. The proof of Theorem 2.5 relies on a careful understanding of nonlinear invariances associated with the dynamical system (2.58)–(2.60) and the key dynamic “sandwich" bound

1dw<1,forY<Y0<0,

where Y0<0 is small. This is shown in Lemma 7.9. The blow-up proof finally follows from a Ricatti-type ordinary differential inequality for the relative velocity w.

In Section 7.4 we compute the sharp asymptotics of w,d, and χ on approach to the massive singularity Yms. This result is stated in Proposition 7.16, which is later crucially used in the study of the causal structure of such a maximally extended solution, where it is in particular shown that the spacetime curvature blows up on approach to the massive singularity.

Remark 2.6

The maximal selfsimilar extension makes sense in the Newtonian limit ε0, which is one of the key observations of Ori and Piran [28]. Our proof of Theorem 2.5 easily extends to the simpler case ε=0, which in particular shows that the LP-solutions constructed in [15] have a natural maximal extension in the Lagrangian (comoving) coordinates.

The RLP-Spacetime and Its Causal Structure (Section 8)

As a consequence of Theorems 2.32.4, and 2.5, we can now formally introduce the exactly selfsimilar solution of the Einstein-Euler system by patching the solutions in the subsonic region x[0,x¯], the supersonic region x[x¯,), and the extended region Y(Yms,0]. Following Ori and Piran [28], we call this spacetime the relativistic Larson-Penston (RLP) solution3.

Definition 2.7

(gRLP,ε-metric) We refer to the 1-parameter family of spherically symmetric selfsimilar spacetimes (MRLP,ε,gRLP,ε) constructed above as the relativistic Larson-Penston spacetimes. In the adapted comoving coordinates (τ~,R) the metric takes the form

gRLP,ε=-e2μ~dτ~2-4ε1+εYe2μ~dτ~dR+e2λ~-4ε(1+ε)2Y2e2μ~dR2+r2γ, 2.65

where the metric coefficients are defined on the connected component of the (τ~,R) coordinate plane given by

D~RLP,ε:=(τ~,R)|R>0,Y(Yms,). 2.66

Here

r(τ~,R)=χ(Y)R,μ~(τ~,R)=μ~(Y),λ~(τ~,R)=λ~(Y), 2.67

where Y=-ετ~R,

e2μ~(Y)=(1+ε)2(1-ε)2Y2ηe2μ(y),y>0, 2.68

and

e2λ~(Y)=e2λ(y),y>0. 2.69

Metric coefficients μ~(Y),λ~(Y) for Y0 are then defined by expressing them as appropriate functions of d,w,χ and extending to Y0, see Proposition 7.16.

The RLP-spacetime in the original comoving coordinates. In the original comoving coordinates (τ,R), the metric takes the form

gRLP,ε=-e2μdτ2+e2λdR2+r2γ. 2.70

It is clear from (2.53) that the metric (2.70) becomes singular across the surface τ=0 (equivalently y=). Nevertheless, away from this surface (i.e. when τ>0 and τ<0) we can keep using comoving coordinates, whereby we define

τ=|Y|ητ~,τ~>0, 2.71

where we recall (2.55). Therefore the metric coefficients are defined on the union of the connected components of the (τ,R) coordinate plane given by

DRLP,ε:=(τ,R)(-,0)×(0,)(τ,R)(0,)×(0,)|y(-,yms), 2.72

where

yms:=-|Yms|-(1+η).

Here

r(τ,R)=-ετr~(y),μ(τ,R)=μ(y),λ(τ,R)=λ(y),forτ<0,

and

r=-ετr~(y),r~(y)=χ(y)y,y=R-ετ=-|Y|-1-η,forτ>0,

where χ(y)=χ(Y), y<0.

Remark 2.8

It is of interest to understand the leading order asymptotic behaviour of the radius r~ as a function of the comoving selfsimilar variable y. It follows from (2.34) and the boundary condition W(0)=13 that limy0+w(y)=13. Therefore, using (2.33) it is easy to see that the leading order behaviour of r~(y) at y=0 is of the form

r~(y)=r~0y1+3ε3(1+ε)+oy0+(y1+3ε3(1+ε)),r~(y)y0+y-23(1+ε), 2.73

for some r~0>0. It follows in particular that yr~(y) is only C0,1+3ε3(1+ε) at y=0. The selfsimilar reduction of the constraint equation (2.15) reads r~2e-2λ=1+εV2-23εG~r~2 (see (3.95)). Using this and (2.73), it then follows

eλ(y)y0+r~(y)y0+, 2.74

which shows that the metric g (2.7) is singular at y=0.

This singularity is not geometric, but instead caused by the Friedmann-like behaviour of the relative velocity W at y=0, see (2.40). It captures an important difference between the comoving and the Schwarzschild coordinates at the centre of symmetry {(τ,R)|τ<0,R=0}. It can be checked that the space-time is regular at the centre of symmetry by switching to the (τ,r) coordinate. The same phenomenon occurs in the Newtonian setting, where it can be shown that the map χ=r~y associated with the LP-solution is exactly C0,13 and therefore χ is not smooth at the labelling origin y=0.

Remark 2.9

Note that the trace of Tμν is easily evaluated

gμνTμν=(ρ+p)gμνuμuν+4p=-ρ+3p=-(1-3ε)ρ.

On the other hand, the trace of the left-hand side of (1.1) is exactly -R, and therefore the Ricci scalar of any classical solution of (1.1)–(1.5) satisfies the relation

R=(1-3ε)ρ. 2.75

This relation also implies that the blow-up of the Ricci scalar is equivalent to the blow up of the mass-energy density when ε13.

Remark 2.10

In Christodolou’s work [6], across the boundary of the backward light cone N emanating from the first singularity, the selfsimilar solution has finite regularity (measured in the Hölder class). This is to be contrasted with the RLP solution which remains real analytic across both characteristic cones - the sonic line and the light cone.

The Outgoing Null-Geodesic

The maximally extended RLP spacetime constructed above has two singular boundary components - the scaling origin O and the massive singularity MSε introduced in Section 2.5, see Figure 3. They are very different in nature, as the density (and the curvature components) blow up at two distinct rates.

Fig. 3.

Fig. 3

Schematic depiction of the outgoing null-geodesics B1, B2, and the massive singularity MSε

The main result of Section 8 states that there exist an outgoing radial null-geodesic (RNG) emanating from the scaling origin O and reaching infinity. Following Ori and Piran we look for a so-called simple RNG:

Definition 2.11

(Simple radial null-geodesics) An RNG of the form

R(τ~)=στ~,σR\{0}, 2.76

is called a simple radial null-geodesic (simple RNG).

Then the key result we prove is the following theorem.

Theorem 2.12

(Existence of global outgoing simple RNG-s) There exists an 0<ε01 sufficiently small so that for any ε(0,ε0] there exist at least two and at most finitely many outgoing simple RNG-s emanating out of the singularity (0, 0). In other words, there exist

Yms<Yn<Y1<0,n2, 2.77

so that the associated simple RNG-s are given by

Bi:={(τ~,R)MRLP,ε|-ετ~R=Yi},i=1,n. 2.78

The proof relies on a beautiful idea of Ori and Piran [28], which we make rigorous. Namely, one can show that the slopes of outgoing simple RNG-s must correspond to roots of a certain real-analytic function, see Lemma 8.2. Using the sharp asymptotic behaviour of the metric coefficients, brought about through our analysis in Sections 67, we can prove that this function converges to negative values at Y=0 and Y=Yms. On the other hand, by the local existence theory for the ODE-system (2.58)–(2.60), we can also ascertain the function in question peaks above 0 for a Y0(Yms,0). Therefore, by the intermediate value theorem, we conclude the proof of Theorem 2.12, see Section 8, immediately after Proposition 8.4.

Informally, the null-hypersurface B1 is the “first" outgoing null-curve emanating from the singular scaling origin and reaching the infinity. It is easy to see that r grows to + along B1. Since the spacetime is not asymptotically flat, we perform a suitable truncation in Section 9 in order to interpret O as a naked singularity.

The maximal extension we construct is unique only if we insist on it being selfsimilar, otherwise there could exist other extensions in the causal future of O. Nevertheless, in our analysis the role of the massive singularity MSε is important, as we use the sharp blow-up asymptotics of our unknowns to run the intermediate value theorem-argument above. Conceptually, MSε has a natural Newtonian limit as ε0, see Remark 2.6, which makes it a useful object for our analysis.

In the remainder of Section 8 we give a detailed account of radial null-geodesics, showing in particular that there is a unique ingoing null-geodesic N emanating from the scaling origin to the past, see Figure 4. This is the boundary of the backward light cone “emanating" from the scaling origin. Following the terminology in [6], it splits the spacetime into the exterior region (in the future of N) and the interior region (in the past of N), see Definition 8.6 and Figure 5. Moreover, the sonic line is contained strictly in the interior region. The complete analysis of nonradial null-geodesics is given in Appendix B.

Fig. 4.

Fig. 4

Schematic depiction of the ingoing null-curve N and the sonic line. The spacetime is smooth across both surfaces

Fig. 5.

Fig. 5

Schematic depiction of the interior (dark grey) and the exterior (light grey) region, see Definition 8.6

Double Null Gauge, Asymptotic Flattening, and Naked Singularities (Section 9)

The final step in the proof of Theorem 1.2 is to truncate the profile away from the scaling origin O and glue it to the already constructed selfsimilar solution. To do that we set up this problem in the double-null gauge:

g=-Ω2dpdq+r2γ, 2.79

where p=const. corresponds to outgoing null-surfaces and q=const. to the ingoing null-surfaces. A similar procedure for the scalar field model was implemented in [6], however due to the complications associated with the Euler evolution, a mere cut-off argument connecting the “inner region" to pure vacuum as we approach null-infinity is hard to do. Instead, we carefully design function spaces that capture the asymptotic decay of the fluid density toward future null-infinity in a way that is both consistent with the asymptotic flatness, and can be propagated dynamically.

Formulation of the Problem in Double-Null Gauge

In addition to the unknowns associated with the fluid, the metric unknowns are the conformal factor Ω=Ω(p,q) and the areal radius r=r(p,q). Clearly,

gpp=gqq=0,gpq=-12Ω2,gpq=-2Ω-2. 2.80

Let u=(up,uq,0,0) be the components of the velocity 4-vector u in the frame {p,q,E2,E3}, where {EA}, A=2,3, describe the generators of some local coordinates on S2. We have g(p,EA)=g(q,EA)=0, A=2,3. The normalisation condition (1.3) in the double-null gauge reads

-1=gμνuμuν=gpqupuq=-4Ω-2upuq, 2.81

which therefore equivalently reads

upuq=14Ω2,upuq=Ω-2. 2.82
Lemma 2.13

(Einstein-Euler system in double-null gauge) In the double-null gauge (2.79), the Einstein field equations take the form

pqr=-Ω24r-1rprqr+πrΩ4Tpq, 2.83
pqlogΩ=-(1+η)πΩ4Tpq+Ω24r2+1r2prqr, 2.84
qΩ-2qr=-πrΩ2Tpp, 2.85
pΩ-2pr=-πrΩ2Tqq. 2.86

Moreover, the components of the energy-momentum tensor satisfy

p(Ω4r2Tpp)+Ω2r2ηq(Ω2r2+2ηTpq)=0, 2.87
p(Ω2r2+2ηTpq)+r2ηΩ2q(Ω4r2Tqq)=0, 2.88

where the energy-momentum tensor is given by the formulas

Tpp=(1+ε)ρ(up)2,Tqq=(1+ε)ρ(uq)2,Tpq=(1-ε)ρΩ-2, 2.89
TAB=ερr-2γAB,TpA=TqA=0,A,B=2,3. 2.90

Moreover, its components are related through the algebraic relation

TppTqq=(1+η)2(Tpq)2. 2.91
Proof

The proof is a straightforward calculation and is given in Appendix C.

The Characteristic Cauchy Problem and Asymptotic Flattening

The idea is to choose a point (p0,q0) in the exterior region (recall Figure 5) and solve the Einstein-Euler system in an infinite semi-rectangular domain (in the (p,q)-plane) D depicted in Figure 6. We normalise the choice of the double-null coordinates by making the outgoing curve B1 in the RLP-spacetime (see Figures 45) correspond to the {p=0} level set, and the ingoing curve N (see Figure 5) to the {q=0} level set. We have the freedom to prescribe the data along the ingoing boundary C_={(p,q)|p[p0,0],q=q0} and the outgoing boundary C={(p,q)|p=p0,qq0}. On C_ we demand that data be given by the restriction of the selfsimilar RLP solution to C_, and on the outgoing piece we make the data exactly selfsimilar on a subinterval q[q0,q0+A0] for some A0>0. On the remaining part of the outgoing boundary q[q0+A0,), we prescribe asymptotically flat data.

Fig. 6.

Fig. 6

The grey shaded area is the region D, where the truncation of the selfsimilar profile takes place. Data are prescribed on the two characteristic surfaces C and C_

The key result of Section 9 is Theorem 9.4, which states that the above described PDE is well-posed on D, if we choose |p0|=δ sufficiently small. We are not aware of such a well-posedness result for the Einstein-Euler system in the double-null gauge in the literature, and we therefore carefully develop the necessary theory. Precise statements are provided in Section 9.2. The idea is standard and relies on the method of characteristics. However, to make it work we rely on an effective “diagonalisation" of the Euler equations (2.87)–(2.88), which replaces these equations by two transport equations for the new unknowns f+ and f-, see Lemma 9.1. This change of variables highlights the role of the acoustic cone and allows us to track the acoustic domain of dependence by following the characteristics, see Figure 6. The analysis of the fluid characteristics is presented in Section 9.3. Various a priori bounds are given in Section 9.4. In Section 9.5 we finally introduce an iteration procedure and prove Theorem 9.4.

Formation of Naked Singularities: Proof of Theorem 1.2

Since the data on C_ and the portion of C with q0qq0+A0 agrees with the RLP-solution, the solution must, by the finite-speed of propagation, coincide with the exact RLP solution in the region DA0 depicted in Figure 7. By uniqueness the solution extends smoothly (in fact analytically) to the exact RLP-solution in the past of C_, all the way to the regular centre {r=R=0}.

Fig. 7.

Fig. 7

The light grey shaded area is the region DA0, where the truncated solution from Theorem 9.4 coincides with the exact selfsimilar solution

The future boundary of the maximal development is precisely the surface {p=0}, the Cauchy horizon for the new spacetime. The boundary of the null-cone corresponding to {p=p0} is complete, and we show in Section 9.6 that for any sequence of points (p0,qn) with qn approaching infinity, the affine length of maximal future-oriented ingoing geodesics launched from (p0,qn) and normalised so that the tangent vector corresponds to p, is bounded by a constant, uniformly-in-n. This shows that the future null-infinity is incomplete in the sense of [32, Definition 1.1], thus completing the proof of Theorem 1.2, see the Penrose diagram, Figure 8.

Fig. 8.

Fig. 8

Penrose diagram of our spacetime with an incomplete future null-infinity

Radially Symmetric Einstein-Euler System in Comoving Coordinates

Selfsimilar Comoving Coordinates

We plug in (2.19)–(2.24) into (2.12)–(2.15) and obtain the selfsimilar formulation of the Einstein-Euler system:

2Σ+yΣ(y)+(1+ε)ΣVr~+2Vr~eμ=0, 3.92
yλ(y)=eμVr~, 3.93
e-μyV(y)+11+εr~(y)e-2λ(y)Σ(y)Σ(y)+r~13G~+2εΣ=0, 3.94
r~2e-2λ=1+εV2-23εG~r~2. 3.95

Remark 3.1

Recall that the radial velocity V satisfies V=e-μτr and therefore

V=e-μ(-r~+yr~(y)). 3.96

From (3.92), (3.93), and (3.96) we obtain

0=2+yΣ(y)Σ(y)+(1+ε)yλ(y)+2yr~(y)r~(y)-1

and therefore

λ(y)=-11+εΣΣ+2y-2r~r~-1y. 3.97

For smooth solutions with strictly positive density Σ, an explicit integration of the identity (3.97) gives the formula

eλ=αΣ-11+εy2ε1+εr~-2, 3.98

for some constant α>0.

Remark 3.2

(Gauge normalisation) From (2.16) we also obtain

μ(y)=-ε1+εΣ(y)Σ(y). 3.99

For any given 0<ε<1 we finally fix the remaining freedom in the problem by setting

c=e2μ(0)Σ(0)2εε+1=1(1+ε)2. 3.100

Upon integrating (3.99) we obtain the identity

eμ(y)=11+εΣ(y)-ε1+ε. 3.101

Before we reduce the selfsimilar formulation (3.92)–(3.95) to a suitable 2×2 nonautonomous ODE system, we first prove an auxiliary lemma.

Lemma 3.3

Let (Σ,V,λ,μ) be a smooth solution to (3.92)–(3.95).

(a)
The following important relationship holds:
w+ε1+ε=yr~r~, 3.102
with w defined by (2.27).
(b)
(Local expression for the mean density) The selfsimilar mean density G~ defined by
G~(y)=6r~30yΣr~2r~dy~ 3.103
satisfies the relation
G~=6Σw. 3.104
(c)
The expression K defined by
K:=-w1+ε+e2μ16(1+ε)G~w+ε+ε(1+ε)Σw+ε. 3.105
satisfies the relation
K=11+εd-w, 3.106
with d defined by (2.26).
(d)
The metric coefficient e2λ satisfies the formula
e2λ=r~2(w+ε)2(1+ε)2y21+εΣ2ε1+εr~2(w-1)2-4εΣwr~2. 3.107

Proof

Proof of part (a). This is a trivial consequence of  (3.96) and (2.27).

Proof of part (b). After multiplying (3.92) by r~2r~ we obtain

0=2Σr~2r~+yr~=r~+Veμr~2Σ+(1+ε)Σ(r~2V+2r~r~V)eμ 3.108
=2Σr~2r~+Σr~3+eμΣr~2V+(1+ε)eμΣ(r~2V) 3.109
=2Σr~2r~+Σr~3+(1+ε)(eμΣr~2V) 3.110

where in the second line we used yr~=r~+Veμ (which follows from (3.96)) and in the last line the identity

(Σeμ)=Σeμ+Σeμ(-ε1+εΣΣ)=11+εΣeμ,

which uses the field equation (3.99). After integrating over the interval [0, y] and integrating-by-parts, we obtain

0yΣr~2r~ds=Σr~2(r~+(1+ε)eμV)=Σr~3w 3.111

where we have used (2.27). Hence dividing by r~3 we obtain (3.104).

Proof of part (c). From (3.105) and (3.104) we immediately have

K=(1+ε)e2μΣw1+ε=11+εΣ1-ε1+ε-w, 3.112

where we have used (3.101).

Proof of part (d). The formula is a simple consequence of (3.95), (3.96), and (3.104).

Proposition 3.4

(The ODE system in comoving selfsimilar coordinates) Let (Σ,V,λ,μ) be a smooth solution to (3.92)–(3.95). Then the pair (d,w) solves the system (2.29)–(2.30), where d,w are defined in (2.26)–(2.27).

Proof

A routine calculation starting from (3.92) and (3.94) (these can be thought of as the continuity and the momentum equation respectively) gives:

y+εyw+ε(w-1)Σ+(1+ε)Σyw+εw-(1+3ε)Σ+3(w+ε)Σ=0, 3.113
e-2μyΣw+11+εεe-2μy(w-1)+(w+ε)e-2λyΣ+e-2μΣ(w+ε)(w-1)1+ε+(1+ε)13ΣG~+2εΣ2=0. 3.114

We may rewrite (3.113)–(3.114) in the form

Σ=2y(1+ε)Σy2-e2μ-2λ-w1+ε+e2μ16(1+ε)G~w+ε+ε(1+ε)Σw+ε, 3.115
w1+ε=(w+ε)1-3w(1+ε)2y-2ywy2-e2μ-2λ-w1+ε+e2μ16(1+ε)G~w+ε+ε(1+ε)Σw+ε. 3.116

With (3.105) in mind, equations (3.115)–(3.116) take the form

Σ=-2(1+ε)ΣKy(e2μ-2λy-2-1), 3.117
w=(w+ε)(1-3w)(1+ε)y+2(1+ε)wKy(e2μ-2λy-2-1), 3.118

where K is given by (3.105) and G~ by (3.103). Equations (2.29)–(2.30) now follow from (2.26) and parts (b) and (c) of Lemma 3.3.

Selfsimilar Schwarzschild Coordinates

We recall the selfsimilar Schwarzschild coordinate introduced in (2.32).

Lemma 3.5

(Selfsimilar Schwarzschild formulation) Let (Σ,w) be a smooth solution to (3.117)–(3.118). Then the variables (D,W) defined by (2.34) solve the system (2.35)–(2.36), with B given by (2.37) and η=η(ε)=2ε1-ε.

Proof

With the above notation and Lemma 3.3 we have K=11+εD-W. It is straightforward to see that the system (3.117)–(3.118) transforms into

D=-2(1-ε)xD(W+ε)(D-W)x2(W+ε)2(e2μ-2λy-2-1), 3.119
W=1-3Wx+2(1+ε)xW(W+ε)(D-W)x2(W+ε)2(e2μ-2λy-2-1). 3.120

From (2.33), the constraint equation (3.95), and (2.27)

(W+ε)2(1+ε)2x2e2μ-2λy-2=r~2e2μ-2λ=e2μ(1+εV2-23εG~r~2)=e2μ1+εe-2μx2(W-1)2(1+ε)2-4εΣWx2, 3.121

where we have slightly abused notation by letting Σ(x)=Σ(y). Therefore

(W+ε)2(1+ε)2x2e2μ-2λy-2-(W+ε)2(1+ε)2x2=Σ-2ε1+ε(1+ε)2-(W+ε)2(1+ε)2-ε(W-1)2(1+ε)2+4ε(1+ε)2Σ1-ε1+εWx2, 3.122

where we have used (3.101). Plugging this back into (3.119)–(3.119), the claim follows.

Lemma 3.6

(Algebraic structure of the sonic denominator B) Consider the denominator B[x;D,W] defined in (2.37). We may factorise B in the form

B[x;D,W]=(1-ε)(J[x;D]-xW(x))(H[x;D]+xW(x)), 3.123

where

J[x;D]=J:=-2ε1-ε(1+D)x+4ε2(1-ε)2(1+D)2x2+εx2+D-η1-ε, 3.124
H[x;D]=H:=J[x;D]+4ε1-ε(1+D)x. 3.125

Moreover,

(1-ε)JH=D-η+ε(1-ε)x2, 3.126

and

J=-2εJ(1+D)+ε(1-ε)x-(2εxJ+ε1-εD-η-1)D(1-ε)J+2εx(1+D). 3.127

Proof

From (2.37) it is clear that we may view B[x;D,W] as a quadratic polynomial in W:

W2+4ε1-ε(1+D)W-ε-x-2D-η1-ε=0.

Solving for W, we obtain two roots, which when multiplied by x give

xW±=-2εx(1+D)±4ε2x2(1+D)2+(1-ε)ε(1-ε)+D-η1-ε. 3.128

We now observe from (3.124)–(3.125) that the positive solution corresponds to J[x;D] and the negative one to -H[x;D]. This in turn immediately gives (3.123). Property (3.126) is obvious from (3.124)–(3.125). Finally, to show (3.127) observe that J solves

(1-ε)J2+4εx(1+D)J-ε(1-ε)x2-D-η=0.

We differentiate the above equality, regroup terms and obtain (3.127).

The Monotonicity Lemma

Controlling the sonic denominator B from below will be one of the central technical challenges in our analysis. From Lemma 3.6 it is clear that this can be accomplished by tracking the quantity J[x;D]-xW(x). A related quantity, of fundamental importance in our analysis is given by

f(x):=J[x;D]-xD. 3.129

The goal of the next lemma is to derive a first order ODE satisfied by f, assuming that we have a smooth solution to the system (2.35)–(2.36). This lemma will play a central role in our analysis.

Lemma 3.7

Let (D,W) be a smooth solution to the selfsimilar Einstein-Euler system (2.35)–(2.36) on some interval I(0,), and let f be given by (3.129). Then, the function f satisfies the ODE

f(x)+a[x;D,W]f(x)=b[x;D,W],xI, 3.130

where

a[x;D,W]=a1[x;D,W]+a2[x;D,W],b[x;D,W]=b1[x;D,W]+εb2[x;D,W]

and

a1[x;D,W]=2εxJ+ε1-εD-η-1(1-ε)J+2εx1+D+x2(1-ε)DW+εB-2ε11-εD-ηx+(1-ε)x+(D+ε)fZ-1; 3.131
a2[x;D,W]=2εJ-xWD-1+2f+4εx+xD5+εZ-1; 3.132
b1[x;D,W]=DH+xW(xW-J)+ε(xW-J)Z-1x2D2-2D+1-ε+2(1-ε)xD-η; 3.133
b2[x,D,W]=-2x2{DD2+(3+ε)D-(1-ε)+ε41-εD3+2(1+ε)1-εD(1+D)+5-ε1-εD2+3D-2+ε}Z-1; 3.134
Z[x;D,W]=(1-ε)J+2εx1+DH+xW. 3.135

Proof

Since f=J-xD-D, the goal is to find the desirable form (3.130) by using the dynamics of J and D (3.127) and (2.35). The factorisation of the denominator B in terms of J and H given in (3.123) will be importantly used in the derivation.

Using (3.123) and xD-xW=xD-J+J-xW, we first rewrite D as

D=-2x(1-ε)D(W+ε)(D-W)B=2(1-ε)D(W+ε)Bf-2D(W+ε)H+xW.

Using further (3.127), it leads to

f=-(2εxJ+ε1-εD-η-1(1-ε)J+2εx(1+D)+x)(2(1-ε)D(W+ε)B)f 3.136
-2εJ(1+D)-ε(1-ε)x-(2εxJ+ε1-εD-η-1)2D(W+ε)H+xW(1-ε)J+2εx(1+D) 3.137
--x2D(W+ε)H+xW+D. 3.138

(3.136) is the form of -af, and it corresponds to the first term of a1 in (3.131).

We next examine (3.138). Using (3.125), we see that

-(3.138)=DH+xW-2x(W+ε)+H+xW=DH+xW-2x(W+ε)+J+4ε1-ε(1+D)x+xW=DH+xWJ-xW+DH+xW2ε1+ε1-εx+4ε1-εDx=:I1+I2. 3.139

Then the first term I1 corresponds to the first term of b1 in (3.133). The second term I2 will be combined together with the second line (3.137).

Let us rewrite I2-(3.137) as

I2-(3.137)=I3Z 3.140

where Z is given in (3.135) and the numerator I3 reads as

I3=D((1-ε)J+2εx(1+D))(2ε1+ε1-εx+4ε1-εDx)+2εJ(1+D)(H+xW) 3.141
-2(W+ε)(2εxDJ+ε1-εD-η)-ε(1-ε)x(H+xW) 3.142
=D((1-ε)J+2εx(1+D))(2ε1+ε1-εx+4ε1-εDx)+2εJ(1+D)(H+J) 3.143
-2(Jx+ε)(2εxDJ+ε1-εD-η)-ε(1-ε)x(H+J) 3.144
-2(W-Jx)(2εxDJ+ε1-εD-η)-ε(1-ε)x(xW-J)+2εJ(1+D)(xW-J). 3.145

We first observe that (3.145) can be written into the form of af-b:

(3.145)=2ε(J-xW)(D-1)f 3.146
+ε(J-xW)x2D2-2D+1-ε+21-εD-ηx 3.147

where the first line corresponds to the first term of a2 in (3.132) and the second line corresponds to the second term of b1 in (3.133).

We next examine (3.143) and (3.144). Using (3.125) to replace H, and (3.126) to replace D-η, and writing J=f+xD to replace J, we arrive at

(3.143)+(3.144)=a~f 3.148
+D((1-ε)xD+2εx(1+D))(2ε1+ε1-εx+4ε1-εDx) 3.149
+4εx2D2(1+D)+8ε21-εx2D(1+D)2 3.150
+2ε2x2(D+ε) 3.151
-2(D+ε)(2εx2D2+εx2D2+4ε21-ε(1+D)x2D) 3.152
-ε(1-ε)x(2xD+4ε1-ε(1+D)x), 3.153

where

a~=D(1-ε)(2ε1+ε1-εx+4ε1-εDx) 3.154
+2ε(1+D)2f+2xD+(2xD+4ε1-ε(1+D)x) 3.155
-21x2εxDf+2εx2D2+ε1-εD-η 3.156
-2(D+ε)ε2xD+f+(2xD+4ε1-ε(1+D)x) 3.157
-2ε(1-ε)x 3.158
=-2ε1(1-ε)xD-η+(1-ε)x+(D+ε)f 3.159
+2ε2f+xD(5+ε)+4εx. 3.160

Now (3.159) corresponds to the second term of a1 in (3.131) and (3.160) corresponds to the last three terms of a2 in (3.132). It remains to check the formula for b2. To this end, we now group (3.149)–(3.153) into ε term and ε2 terms:

(3.149)++(3.153) 3.161
=[2ε(1+ε)x2D2+4εD3x2+4ε21+ε1-εx2D(1+D)+8ε21-εx2D2(1+D) 3.162
+4εx2D2(1+D)+8ε21-εx2D(1+D)2+2ε2x2(D+ε)] 3.163
-[6εx2D3+6ε2x2D2+8ε21-εx2D2(1+D)+8ε31-εx2D(1+D) 3.164
+2ε(1-ε)x2D+4ε2x2(1+D)] 3.165
=[8εx2D3+2εx2D2(3+ε)]-[6εx2D3+2ε(1-ε)x2D]+[4ε21+ε1-εx2D(1+D)+8ε21-εx2D3+16ε21-εx2D2+2ε2(41-ε+1)x2D+2ε3x2]-[(6ε2+8ε31-ε)x2D2+(8ε31-ε+4ε2)x2D+4ε2x2]=2εx2DD2+(3+ε)D-(1-ε) 3.166
+2ε2x241-εD3+2(1+ε)1-εD(1+D)+5-ε1-εD2+3D-2+ε. 3.167

This completes the proof.

Corollary 3.8

Let (D,W) be a smooth solution to the selfsimilar Einstein-Euler system (2.35)–(2.36) on some interval I(0,). Then for any x1<x, x1,xI, we have the formula

f(x)=f(x1)e-x1xa[z;D,W]dz+e-x1xa[z;D,W]dzx1xb1[z;D,W]+εb2[z;D,W]ex1za[s;D,W]dsdz. 3.168

Proof

The proof follows by applying the integrating factor method to (3.130).

Lemma 3.9

(Sign properties of b) Assume that J-xW>0 and B[x;D,W]>0. Then there exists an ε0>0 sufficiently small such that for all 0<εε0 the following statements hold:

(a)
For any D>0
b1[x;D,W]<0. 3.169
(b)
Furthermore,
b2[x;D,W]<0for allD>13. 3.170

Proof

The assumptions of the lemma and the decomposition (3.123) imply that Z>0 and H+xW>0.

Proof of part (a). The negativity of b1 is obvious from (3.133) and the obvious bound 2D2-2D+1-ε>0 for ε sufficiently small.

Proof of part (b). Let φ0 be the larger of the two roots of the quadratic polynomial DD2+(3+ε)D-(1-ε), which is given by

φ0:=-(3+ε)+(3+ε)2+4(1-ε)2.

It is easily checked that there exists an ε0>0 such that 0<φ0<13 for all 0<εε0 and in particular

DD2+(3+ε)D-(1-ε)>0

for D>13(>φ0). On the other hand,

41-εD3+2(1+ε)1-εD(1+D)+5-ε1-εD2+3D-2+ε>4D3+6D2+5D-2>0

for D>13 and the claim follows from (3.134).

The Sonic Point Analysis

It turns out that for purposes of homogeneity, it is more convenient to work with rescaled unknowns where the sonic point is pulled-back to a fixed value 1. Namely, we introduce the change of variables:

z=xx,W(z)=W(x),R(z)=D(x). 4.171

so that the sonic point x is mapped to z=1. It is then easily checked from (2.35)–(2.36) that (R,W) solves

dRdz=-2x2z(1-ε)R(W+ε)(R-W)B, 4.172
dWdz=(1-3W)z+2x2z(1+ε)W(W+ε)(R-W)B, 4.173

where

B=R-η-(W+ε)2-ε(W-1)2+4εRWx2z2. 4.174

We introduce

δz:=z-1. 4.175

we look for solutions R,W to (4.172)-(4.173) of the form

R=N=0RN(δz)N,W=N=0WN(δz)N. 4.176

We observe that there is a simple relation between the formal Taylor coefficients of (D,W) and (R,W):

Rj=Djxj,Wj=Wjxj,j=0,1,2,. 4.177

Sonic Point Conditions

Lemma 4.1

(Sonic conditions) There exists a small ε0>0 such that for all |ε|ε0 and all x[32,72] there exists a continuously differentiable curve [-ε0,ε0]ε(R0(ε),W0(ε))=(R0(ε;x),W0(ε;x)) such that

  1. R0(ε)=W0(ε)>0.

  2. B[R0(ε),W0(ε)]=0.

  3. R0(0)=W0(0)=1x.

  4. -<εR0(0)=εW0(0)<0.

Proof

Fix an x[32,72]. Consider a small neighbourhood of (ε,W)=(0,1x), open rectangle (ε,W)(-l,l)×(w1,w2), and a continuously differentiable function h:(-l,l)×(w1,w2)R defined by

h(ε,W):=(1+3ε)W2+4εW+ε(ε-1)Wη-1x2, 4.178

where we recall η=η(ε)=2ε1-ε. Then the sonic point conditions R=W and B=0 with z=1 reduce to h(ε,W)=0 and moreover we have h(0,1x)=0. Clearly h is continuously differentiable in all arguments. Observe that

hW=2(1+3ε)W2+4εW+η(1+3ε)W2+4εW+ε(ε-1)Wη-1 4.179

from which we have

hW|(ε,W)=(0,1x)=2x>0. 4.180

Therefore, by the implicit function theorem, we deduce that there exists an open interval (-l0,l0) of ε=0 and a unique continuously differential function g:(-l0,l0)(w1,w2) such that g(0)=1x and h(ε,g(ε))=0 for all ε(-l0,l0). Moreover, we have

gε=-hεhW,ε(-l0,l0) 4.181

where

hε=3W2+4W+2ε-1+(1+3ε)W2+4εW+ε(ε-1)2lnW(1-ε)2Wη. 4.182

When ε=0,

gε(0)=-hε|(0,1x)hW|(0,1x)=-3x2+4x-1-2lnxx22x=x2-4x+2logx-32x. 4.183

The derivative of the map xx2-4x+2logx-3 is 2(x-1)2x and the function is therefore strictly increasing for x1. It is easy to check that the value at x=72 is negative, and therefore there exists a constant κ>0 such that gε(0)<-κ for all x[32,72]. In particular, there exists a 0<ε01 sufficiently small and a constant c>0 such that

1x-cε<g(ε;x)<1x,ε(0,ε0],x32,72.

We let

R0:=R(ε)=g(ε),W0:=W0(ε)=g(ε). 4.184

Remark 4.2

(The map xW0(ε;x) is decreasing) In order to examine the behaviour of W0(ε;x)=g(ε;x) as a function of x for any fixed ε, we rewrite the relation h(ε,W0(ε))=0 in the form

h(ε,g(ε;x);x)=0.

Upon taking the x derivative of the above, we easily see that xg(ε;x)<0. In fact, we have xg=-xhWh where xh=2x3>0 from (4.178), and Wh>0 is given in (4.179).

For a given function f, we write (f)M, MN, to denote the M-th Taylor coefficient in the expansion of f around the sonic point z=1. In particular,

(R(W+ε)(R-W))M=l+m+n=MRl(Wm+εδm0)(Rn-Wn),(W(W+ε)(R-W))M=l+m+n=MWl(Wm+εδm0)(Rn-Wn),(W2)M=l+m=MWlWm,(RW)M=l+m=MRlWm.

We set (f)M=0 for M<0.

Formula of Faa Di Bruno. Given two functions fg with formula power series expansions

f(x)=n=0fnxn,g(x)=n=1gnxn, 4.185

we can compute the formal Taylor series expansion of the composition h=fg via

h(x)=n=0hnxn 4.186

where

hn=m=1nπ(n,m)m!λ1!λn!fmg1λ1gnλn,h0=f0 4.187

and

π(n,m)=(λ1,,λn):λiZ0,i=1nλi=m,i=1niλi=n. 4.188

An element of π(n,m) encodes the partitions of the first n numbers into λi classes of cardinality i for i{1,,m}. Observe that by necessity

λj=0forn-m+2jn.

To see this, suppose λj=p1 for some n-m+2jn. Then m-p=ijλiijiλi=n-jpn-(n-m+2)p, which leads to (n-m+1)pn-m. But this is impossible if p1.

Now

R-η=R0-η1+i=1RiR0(δz)i-η=R0-η+j=1(R-η)j(δz)j 4.189

where

(R-η)j=R0-ηm=1j1R0mπ(j,m)(-η)m1λ1!λj!R1λ1Rjλj,j1. 4.190

Here (-η)m=(-η)(-η-1)(-η-m+1). Then we may write B as

B=R-η-(1-ε)W2+4εW+4εRW+ε2-εx2z2 4.191
=:R-η-x2Hz2 4.192
=l=0(R-η)l(δz)l-x2l=0Hl(δz)l(1+2δz+(δz)2) 4.193

so that

Hl=(1-ε)(W2)l+4εWl+4ε(RW)l+(ε2-ε)δl0. 4.194

Lemma 4.3

For any N0, the following formulas hold:

l+m=N(m+1)Rm+1(R-η)l-x2l+m=N(m+1)Rm+1Hl+2l+m=N-1(m+1)Rm+1Hl+l+m=N-2(m+1)Rm+1Hl+2x2(1-ε)(R(W+ε)(R-W))N+(R(W+ε)(R-W))N-1=0 4.195

and

l+m=N(m+1)Wm+1(R-η)l-x2l+m=N(m+1)Wm+1Hl+2l+m=N-1(m+1)Wm+1Hl+l+m=N-2(m+1)Wm+1Hl-l+m=N(R-η)l(-1)m+3l+m+n=NWn(R-η)l(-1)m+x2l+m=NHl(-1)m+2l+m=N-1Hl(-1)m+l+m=N-2Hl(-1)m-3x2l+m+n=NWnHl(-1)m+2l+m+n=N-1WnHl(-1)m+l+m+n=N-2WnHl(-1)m-2x2(1+ε)(W(W+ε)(R-W))N+(W(W+ε)(R-W))N-1=0 4.196

where we recall (f)M=0 for M<0.

Proof

Proof of (4.195). We now plug (4.176) into

BR+2x2(1-ε)(1+δz)R(W+ε)(R-W)=0, 4.197
BW-(1-3W)B1+δz-2x2(1+ε)(1+δz)W(W+ε)(R-W)=0. 4.198

(4.197) reads as

0=l=0(R-η)l(δz)l-x2l=0Hl(δz)l(1+2δz+(δz)2)m=0(m+1)Rm+1(δz)m+2x2(1-ε)l=0(R(W+ε)(R-W))l(δz)l(1+δz)=N=0l+m=N(m+1)Rm+1(R-η)l(δz)N-x2N=0l+m=N(m+1)Rm+1Hl+2l+m=N-1(m+1)Rm+1Hl+l+m=N-2(m+1)Rm+1Hl(δz)N+2x2(1-ε)N=0(R(W+ε)(R-W))N+(R(W+ε)(R-W))N-1(δz)N.

Comparing the coefficients, we obtain (4.195).

Proof of (4.196). Since 11+δz=m=0(-1)m(δz)m, we can expand (1-3W)B1+δz as

(1-3W)B1+δz=1-3n=0Wn(δz)nl=0(R-η)l(δz)l-x2l=0Hl(δz)l(1+2δz+(δz)2)m=0(-1)m(δz)m=N=0l+m=N(R-η)l(-1)m(δz)N-3N=0l+m+n=NWn(R-η)l(-1)m(δz)N-x2N=0l+m=NHl(-1)m+2l+m=N-1Hl(-1)m+l+m=N-2Hl(-1)m(δz)N+3x2N=0l+m+n=NWnHl(-1)m+2l+m+n=N-1WnHl(-1)m+l+m+n=N-2WnHl(-1)m(δz)N.

We plug (2.43) into (4.198)

0=l=0(R-η)l(δz)l-x2l=0Hl(δz)l(1+2δz+(δz)2)m=0(m+1)Wm+1(δz)m-N=0(1-3W)B1+δzN(δz)N-2x2(1+ε)l=0(W(W+ε)(R-W))l(δz)l(1+δz)=N=0l+m=N(m+1)Wm+1(R-η)l(δz)N-x2N=0l+m=N(m+1)Wm+1Hl+2l+m=N-1(m+1)Wm+1Hl+l+m=N-2(m+1)Wm+1Hl(δz)N-N=0(1-3W)B1+δzN(δz)N-2x2(1+ε)N=0(W(W+ε)(R-W))N+(W(W+ε)(R-W))N-1(δz)N

which leads to (4.196).

Remark 4.4

(R0,W0) obtained in Lemma 4.1 satisfy the sonic conditions:

R0=W0,R0-η=x2H0 4.199

and it is easy to verify that for such choice of (R0,W0), the relations (4.195)N=0 and (4.196)N=0 trivially hold. Moreover, with (4.199), there are no (RN+1,WN+1) in (4.195) and (4.196).

Remark 4.5

To determine R1,W1, we record (4.195)N=1 and (4.196)N=1:

(R-η)1R1-x2H1R1-2x2H0R1+2x2(1-ε)R0(W0+ε)(R1-W1)=0 4.200

and

(R-η)1W1-x2(W1H1+2W1H0)-[(R-η)1-R0-η]+3[W1R0-η+W0(R-η)1-W0R0-η]+x2[H1+H0]-3x2[W1H0+W0H1+W0H0]-2x2(1+ε)W0(W0+ε)(R1-W1)=0, 4.201

where we have used the sonic conditions (4.199). Recalling from (4.190) and (4.194)

(R-η)1=-ηR0-η-1R1, 4.202
H1=2(1+ε)W0+2εW1+4εW0R1, 4.203

we see that for general nonzero ε, (4.200) is linear in W1 and quadratic in R1 and (4.201) is linear in R1 and quadratic in W1. On the other hand, when ε=0, since (R-η)1=0, H1=2W0W1, H0=W02, (4.200) becomes

-2x2W0W1R1-2x2W02R1+2x2R0W0(R1-W1)=0,

and (4.201) becomes

-x2(2W0W12+2W1W02)+1+3[W1-W0]+x2[2W0W1+W02]-3x2[W1W02+2W02W1+W03]-2x2W02(R1-W1)=0.

Thus by using R0|ε=0=W0|ε=0=1x, we see that (4.200) and (4.201) are reduced to

-2W1xR1+1=0, 4.204
-2xW12-xW1+3W1+R1+3x-1=0, 4.205

which are the the sonic point conditions satisfied in the Newtonian limit, see [15]. In general, there are two pairs of solutions to (4.204)-(4.205), one is of Larson-Penston type given by (R1,W1)=(-1x,1-2x) and the other one is of Hunter type given by (R1,W1)=(1-3x,0).

In the following, we will show that there exists a continuously differentiable curve (R1(ε),W1(ε)) satisfying (4.200)–(4.201), which at ε=0 agrees precisely with the values of (R1,W1) associated with the (Newtonian) Larson-Penston solution, namely (R1(0),W1(0))=(-1x,1-2x) for x>2.

Lemma 4.6

(RLP conditions) Let x(2,72) be fixed and let (R0,W0) be as obtained in Lemma 4.1. Then there exists an ε1>0 such that there exists a continuously differentiable curve (-ε1,ε1)ε(R1(ε), W1(ε)) such that

  1. The relations (4.195)N=1 and (4.196)N=1 hold.

  2. R1(0)=-1x and W1(0)=1-2x.

Proof

As in Lemma 4.1 we will use the implicit function theorem. To this end, consider a small neighborhood (-l,l)×(r1,r2)×(w1,w2) of (0,-1x,1-2x) and introduce F:(-l,l)×(r1,r2)×(w1,w2)R2 as

F(ε,R1,W1)=F1(ε,R1,W1),F2(ε,R1,W1)T 4.206

where

graphic file with name 40818_2022_144_Equ207_HTML.gif 4.207

so that (4.200) and (4.201) are equivalent to F(ε,R1,W1)=0. It is clear that F is continuously differentiable in all arguments and F(0,-1x,1-2x)=0. We next compute the Jacobi matrix F[R1,W1]

F[R1,W1]=F1R1F1W1F2R1F2W1 4.208

where

F1R1=-2ηR0-η-1R1-x2H1-x2R1H1R1-2x2H0+2x2(1-ε)R0(W0+ε),F1W1=-x2R1H1W1-2x2(1-ε)R0(W0+ε),F2R1=-ηR0-η-1W1-x2W1H1R1+ηR0-η-1-3ηW0R0-η-1+x2H1R1-3x2W0H1R1-2x2(1+ε)W0(W0+ε),F2W1=-ηR0-η-1R1-x2(H1+2H0)-x2W1H1W1+3R0-η+x2H1W1-3x2H0-3x2W0H1W1+2x2(1+ε)W0(W0+ε),

and

H1R1=4εW0,H1W1=2[(1+ε)W0+2ε].

Using H1(0)=2W0(0)W1(0), H1R1(0)=0, H1W1(0)=2W0(0), H0(0)=W02(0), W0(0)=R0(0)=1x, W1(0)=x-2x and R1(0)=-1x, we evaluate the above at (ε,R1,W1)=(0,-1x,1-2x)

F1R1|(k,R1,W1)=(0,-1x,1-2x)=-x2H1(0)-2x2H0(0)+2x2R0(0)W0(0)=-2(x-2),F1W1|(k,R1,W1)=(0,-1x,1-2x)=-2x2R1(0)W0(0)-2x2R0(0)W0(0)=0,F2R1|(k,R1,W1)=(0,-1x,1-2x)=-2x2W02(0)=-2,F2W1|(k,R1,W1)=(0,-1x,1-2x)=-x2(H1(0)+2H0(0))-2x2W1(0)W0(0)+3+2x2W0(0)-3x2H0(0)-6x2W02(0)+2x2W02(0)=-4x2W0(0)W1(0)+2x-6=-4(x-2)+2x-6=-2(x-1).

The Jacobian at (ε,R1,W1)=(0,-1x,1-2x) is

F[R1,W1](ε,R1,W1)=(0,-1x,1-2x)=4(x-2)(x-1)>0ifx>2 4.209

and thus, F[R1,W1] is invertible in sufficiently small neighborhood of (0,-1x,1-2x) for any fixed x>2. Now by the implicit function theorem, we deduce that there exists an open interval (-l0,l0) of ε=0 and unique continuously differential functions g1:(-l0,l0)(r1,r2), g2:(-l0,l0)(w1,w2) such that g1(0)=-1x, g2(0)=1-2x and F(ε,g1(ε),g2(ε))=0 for all ε(-l0,l0).

It is evident that R1 and W1 become degenerate at x=2. In particular, when ε=0, the corresponding R1 and W1 for the LP-type and the Hunter-type solutions coincide precisely at x=2, while they are well-defined for further values of x below 2, see [15]. Interestingly this feature does not persist for ε>0. In fact, we will show that R1 and W1 cease to exist as real numbers before x reaches 2 from above. To this end, we first derive the algebraic equation satisfied by W1.

Lemma 4.7

(The cubic equation for W1) The Taylor coefficient W1 is a root of the cubic polynomial

P(X):=X3+a2a3X2+a1a3X+a0a3, 4.210

where aj, j=0,1,2,3 are real-valued continuous functions of W0 given in (4.221)–(4.224) below. Moreover, there exists a constant c>0 and 0<ε01 such that for all 0<εε0 and any 13<W0<1, we have c<a3<1c, c<a0ε(3W0-1)<1c. In particular any root of XP(X) is not zero for 0<εε0.

Proof

Note that (4.201) can be rearranged as

-x2H1[W1+3W0-1]-x2H0[ηR1W0+2][W1+3W0-1]-2x2(1+ε)W0(W0+ε)(R1-W1)=0. 4.211

Rewrite (4.200) as

(1-ε)W0(W0+ε)-H0-[η2H0W0+2εW0]R1R1-(1-ε)W0(W0+ε)+[(1+ε)W0+2ε]R1W1=0, 4.212

and (4.211) as

R1=-[(1+ε)W0+2ε]W12+[(3+5ε)W02-(1-8ε+ε2)W0-(3ε-ε2)]W1+(3W0-1)H0(1+ε)W0(W0+ε)+[η2H0W0+2εW0](3W0-1)+[η2H0W0+2εW0]W1. 4.213

We would like to derive an algebraic equation for W1. To this end, write (4.213) and (4.212) as

R1=-dW12+eW1+cH0h+bW1, 4.214

where

d:=(1+ε)W0+2ε,e:=(3+5ε)W02-(1-8ε+ε2)W0-(3ε-ε2),c:=3W0-1, 4.215
h:=(1+ε)W0(W0+ε)+[η2H0W0+2εW0](3W0-1),b:=η2H0W0+2εW0. 4.216

We observe that we may write h=(1+ε)a+bc, where

a:=W0(W0+ε). 4.217

If we let

f:=(1-ε)a-H0, 4.218

then (4.212) can be rewritten in the form

[f-bR1]R1-[(1-ε)a+dR1]W1=0, 4.219

with f, .bad as above. Now replace R1 in (4.219) using the relation  (4.214):

-f+bdW12+eW1+cH0h+bW1dW12+eW1+cH0h+bW1-(1-ε)a-ddW12+eW1+cH0h+bW1W1=0.

Multiply (h+bW1)2:

-f(h+bW1)+b(dW12+eW1+cH0)(dW12+eW1+cH0)-(1-ε)a(h+bW1)-d(dW12+eW1+cH0)(h+bW1)W1=0.

Note that the highest order term W14 is not present. Rearrange similar terms to conclude the identity

a3W13+a2W12+a1W1+a0=0, 4.220

where

a3:=d(dh-be)-b(fd+(1-ε)ab), 4.221
a2:=-[bcdH0-e(dh-be)+h(fd+(1-ε)ab)+b(fe+(1-ε)ah)], 4.222
a1:=-[bceH0-cH0(dh-be)+bcfH0+h(fe+(1-ε)ah)], 4.223
a0:=-cH0(bcH0+fh). 4.224

By (4.224), the sign of a0 is the same as the sign of bcH0+fh. We now use

bcH0+fh=η2H0W0+2εW03W0-1H0+(1-ε)W0(W0+ε)-H0(1+ε)W0(W0+ε)+[η2H0W0+2εW0](3W0-1)=(1-ε)W0(W0+ε)-H0(1+ε)W0(W0+ε)+(1-ε)W0(W0+ε)[η2H0W0+2εW0](3W0-1)=W0(W0+ε){(1+ε)(1-ε)W0(W0+ε)-H0(I)+ε[H0W0+2(1-ε)W0](II)(3W0-1)}(III),

where we have used η2=ε1-ε in the last line. Now

(I)=(1-ε)W0(W0+ε)-(1+3ε)W02-4εW0+ε-ε2=ε-4W02-(3+ε)W0+1-ε

and

(II)=(1+3ε)W0+4ε-ε-ε2W0+2(1-ε)W0=(3+ε)W0+4ε-ε-ε2W0.

Hence

(III)=ε{-4(1+ε)W02-(3+ε)(1+ε)W0+1-ε2+[(3+ε)W0+4ε-ε-ε2W0](3W0-1)}=ε{(5-ε)W02-(6-7ε+ε2)W0+(1-7ε+2ε2)+ε-ε2W0}=ε{5W02-6W0+1+O(ε)}.

Since 5W02-6W0+1=(5W0-1)(W0-1)<0 for 15<W0<1,

a0=-cH0(bcH0+fh)=-εcH0W0(W0+ε){5W02-6W0+1+O(ε)}>0for13<W0<1and0<ε1. 4.225

From (4.215)–(4.217) it is easy to see that

a=W02+O(ε),b=O(ε),c=3W0-1,d=W0+O(ε), 4.226
e=3W02-W0+O(ε),f=(1-ε)W0(W0+ε)-H0=O(ε), 4.227
h=(1+ε)W0(W0+ε)+[ε1-εH0W0+2εW0](3W0-1)=W02+O(ε). 4.228

This then implies

a3=W04+O(ε),a2=W04(3W0-1)+O(ε),a1=W05(2W0-1)+O(ε),a0ε(3W0-1)1 4.229

which shows the uniform positivity and boundedness of a3 for sufficiently small ε. Upon dividing (4.220) by a3 we finally conclude the proof of the lemma.

Remark 4.8

In the formal Newtonian limit ε=0, the cubic equation P(X)=0 reduces to

X3+(3W0-1)X2+(2W0-1)W0X=X(X+W0)(X+2W0-1)=0. 4.230

The root X=0 corresponds to the Hunter-type, X=1-2W0 is of the LP-type, and X=-W0 is the Newtonian ghost solution, see [15].

Relativistic Larson-Penston-Type Solutions

By Lemma 4.7 we know that there are in general three complex roots of (4.210) giving possible values for W1. One of those roots is a “ghost root" and is discarded as unphysical solution of (4.210). More precisely, in the Newtonian limit ε0 such a spurious root corresponds to the value W1(0)=-1x. Our goal is to first mod out such a solution by using the implicit function theorem to construct a curve εW1gh(ε) of spurious solutions agreeing with -1x when ε=0.

Lemma 4.9

(Ghost root) The cubic polynomial P introduced in Lemma 4.7 can be factorised in the form

P(X)=(X-W1gh(ε))Q(X), 4.231

where W1gh(ε) is a real valued function of W0(ε) and Q is a quadratic polynomial given by

Q(X)=X2+a2a3+W1ghX-a0a3W1gh. 4.232

Proof

We differentiate (4.210) to obtain

PX|(ε,X)=(0,-1x)=3X2+2a2a3X+a1a3|(ε,X)=(0,-1x)=31x2-23-xx1x+2-xx2=x-1x2>0. 4.233

The implicit function theorem implies the existence of a unique ε-parametrised curve

W1gh(ε)=-1x+O(ε) 4.234

satisfying P(W1gh(ε))=0. Identity (4.231) can now be checked directly, keeping in mind (4.232).

The discriminant of Q is given by

ΔQ=a2a3+W1gh2+4a0a3W1gh.

Lemma 4.10

(Definition of xcrit(ε)) There exists a continuous curve (0,ε0]εxcrit(ε)(12,72) such that

ΔQ(xcrit(ε))=0, 4.235
ΔQ(x)>0,x(xcrit(ε),xmax(ε)], 4.236
xcrit(ε)=2+O(ε)>2. 4.237

Moreover, for any x(xcrit(ε),72] there exists a constant C=C(x)>0 such that for all ε(0,ε0]

W1RLP(ε;x)-x-2xCε. 4.238

When x=xcrit(ε) the rate of convergence changes and there exists a constant C¯>0 such that

W1RLP(ε;xcrit(ε))-xcrit(ε)-2xcrit(ε)C¯ε. 4.239

Proof

Step 1. Existence of xcrit. From the asymptotic formulas (4.229) and (4.234) it is clear that for all ε(0,ε0] we have

-a2a3+W1gh=x-2x+O(ε), 4.240

and therefore

ΔQ(ε,x)=x-2x+O(ε)2+4a0a3W1gh.

However, by Lemma 4.7 we know that 4a0a3W1gh-α(3W0-1)ε for some positive constant α>0. Choosing x<2 such that 2-x=O(ε), we see that ΔQ(ε,x)<0. On the other hand, for any x2+δ for some small, but fixed δ>0, we see that ΔQ(ε,x)>0. Therefore, by the intermediate value property, there exists an x(2-O(ε),2+δ) such that ΔQ(x)=0. Let xcrit be the largest such x - it is clear from the construction that the properties (4.235)–(4.236) are satisfied.

Step 2. Asymptotic behaviour of xcrit(ε). From (4.240), the identity

-a2a3+W1gh|x=xcrit(ε)=-4a0a3W1gh|x=xcrit(ε), 4.241

and Lemma 4.7 we easily conclude (4.237).

Step 3. Continuity properties of the map εW1RLP(ε,x). Fix an x(xcrit,xmax]. Then

W1RLP(ε;x)-x-2x=-a2a3+W1gh+(a2a3+W1gh)2+4a0a3W1gh2-x-2x=-a2a3+W1gh-x-2x2+(a2a3+W1gh)2+4a0a3W1gh-x-2x2=:(i)+(ii). 4.242

By (4.240) we have (i)=O(ε). For (ii),

(ii)=(a2a3+W1gh)2-(x-2x)2+4a0a3W1gh2(a2a3+W1gh)2+4a0a3W1gh+x-2x=O(ε)(x-2)2+O(ε)+(x-2)asε0+, 4.243

by (4.229) and (4.240). Therefore, the bound (4.238) follows.

If we now let x=xcrit(ε), since by definition ΔQ(xcrit(ε))=0, we have

W1RLP(ε;xcrit(ε))-xcrit(ε)-2xcrit(ε)=-a2a3+W1gh2-xcrit(ε)-2xcrit(ε)=--a0a3W1gh-xcrit(ε)-2xcrit(ε)=Oε0+(ε), 4.244

where we have used (4.237) and Lemma 4.7. This proves (4.239).

Of special importance in our analysis is the Friedmann solution, which has the property that for any ε>0 W0=R0=13. By Lemma 4.1 there exists a continuously differentiable curve εxmax(ε) defined through the property

W0(ε;xmax(ε))=13,ε[0,ε0]. 4.245

Since W0(0;3)=13, we conclude from Lemma 4.1 that

xmax(ε)<3and3-xmax(ε)=O(ε),ε[0,ε0]. 4.246

With above preparations in place, we are now ready to define what we mean by a solution of the relativistic Larson-Penston type.

Definition 4.11

(Solutions of RLP-type) Let 0<ε01 be given by Lemma 4.7 and let x[xcrit,xmax]. We say that the sequence (RN,WN)NN is of relativistic Larson-Penston (RLP)-type if for all ε(0,ε0]

R0=W0, 4.247

the coefficients (RN,WN)NN satisfy the recursive relations (4.195)–(4.196), W1 is the root of the quadratic polynomial Q (4.232) given by

W1=W1RLP=-a2a3+W1gh+(a2a3+W1gh)2+4a0a3W1gh2, 4.248

and R1 is given as a function of W0 and W1 via (4.213).

Remark 4.12

For all ε(0,ε0] it is clear from the proof of the lemma that the following bound holds in the interval x[xcrit(ε),xmax(ε)]:

W1RLP(ε;x)-x-2x=O(ε)(x-2)2+O(ε)+(x-2). 4.249

High-Order Taylor Coefficients

We now consider (4.195)N2 and (4.196)N2.

Lemma 4.13

Let N2. Then the following holds

AN(W0,W1,R1)RNWN=SNVN. 4.250

Here

AN(W0,W1,R1)=A11A12A21A22 4.251

where

A11=-2x2H0N+2x2(1-ε)R0(W0+ε)-x2H1N-ηR0-η-1R1(N+1)-4εx2W0R1,A12=-x2R1(2(1-ε)W0+4ε+4εR0)-2x2(1-ε)R0(W0+ε),A21=-2x2(1+ε)W0(W0+ε)-(ηR0-η-1+4εW0x2)(W1+3W0-1),A22=-2x2H0N-3x2H0+3R0-η+2x2(1+ε)W0(W0+ε)-x2(W1+3W0-1)(2(1-ε)W0+4ε+4εR0)-x2H1N-ηR0-η-1R1N,

and

SN=SN[R0,W0;R1,W1;,RN-1,WN-1],VN=VN[R0,W0;R1,W1;,RN-1,WN-1],

are given in (4.255) and (4.258).

Proof

We first observe that there are no terms involving (RN+1,WN+1) in (4.195) and (4.196) due to the sonic conditions (4.199) (cf. Remark 4.4 and see the cancelation below). To prove the lemma, we isolate all the coefficients in (R-η)N and HN that contain the top order coefficients RN,WN. For (R-η)N, from (4.190) we have

(R-η)N=-ηR0-η-1RN+R0-ηm=2N1R0mπ(N,m)(-η)m1λ1!λN!R1λ1RNλN 4.252

where we recall λj=0 for N-m+2jN, in particular λN=0 and therefore there are no terms involving RN in the summation term. For HN, from (4.194), we have

HN=2(1-ε)W0WN+4εWN+4ε(R0WN+RNW0)+(1-ε)+m=N1mN-1WWm++m=N1mN-14εRWm. 4.253

Using (4.252) and (4.253), we can isolate all the coefficients in (4.195) that contain contributions from (RN,WN) as follows:

graphic file with name 40818_2022_144_Equ254_HTML.gif 4.254

where

SN=-R1R0-ηm=2N1R0mπ(N,m)(-η)m1λ1!λN!R1λ1RNλN+x2R1[(1-ε)+m=N1mN-1WWm++m=N1mN-14εRWm]+S~N, 4.255

and

S~N=-+m=N1mN-2(m+1)Rm+1(R-η)+x2+m=N1mN-2(m+1)Rm+1H+2+m=N-1mN-2(m+1)Rm+1H++m=N-2(m+1)Rm+1H-2x2(1-ε)+m+n=N1nN-1R(W+ε)m(R-W)n+(R(W+ε)(R-W))N-1. 4.256

Here we recall the definitions of (R-η) in (4.190) and H1 in (4.203) as well as the sonic conditions in (4.199). Note that we have also used (4.202) above.

Following the same procedure, we now isolate all the coefficients in (4.196) that contain contributions from (RN,WN).

graphic file with name 40818_2022_144_Equ257_HTML.gif 4.257

where

VN=-(W1-1+3W0)R0-ηm=2N1R0mπ(N,m)(-η)m1λ1!λN!R1λ1RNλN+x2(W1-1+3W0)[(1-ε)+m=N1mN-1WWm++m=N1mN-14εRWm]+V~N 4.258

and

V~N=-+m=N1mN-2(m+1)Wm+1(R-η)+x2+m=N1mN-2(m+1)Wm+1H+2+m=N-1mN-2(m+1)Wm+1H++m=N-2(m+1)Wm+1H++m=N1mN(R-η)(-1)m-3+m+n=N1mNWn(R-η)(-1)m-x2+m=N1mNH(-1)m+2+m=N-1H(-1)m++m=N-2H(-1)m+3x2+n=N1nN-1WnH++m+n=N1mNWnH(-1)m+2+m+n=N-1WnH(-1)m++m+n=N-2WnH(-1)m+2x2(1+ε)+m+n=N1nN-1W(W+ε)m(R-W)n+(W(W+ε)(R-W))N-1. 4.259

This completes the proof.

Let κ>0 be a sufficiently small number independent of ε to be fixed later in Section 4.5.

Lemma 4.14

Let x[xcrit+κ,xmax] be given. Then the components Aij of the matrix AN satisfy the following

A11=-2Nx-1+O(ε)+2+O(ε), 4.260
A12=O(ε), 4.261
A21=-2+O(ε), 4.262
A22=-2Nx-1+O(ε)+O(ε). 4.263

In particular, the matrix AN is invertible for all sufficiently small ε and detAN=O(N2). Moreover, the Taylor coefficients (RN,WN), N2, satisfy the recursive relationship

RN=A22detANSN-A12detANVN, 4.264
WN=A11detANVN-A21detANSN, 4.265

where the source terms VN,SN, N2, are given by Lemma 4.13.

Proof

We rearrange terms in Aij of Lemma 4.13 as

A11=-2Nx2H0+12x2H1+η2R0-η-1R1+2x2R0W0+2εx2R0(1-ε-W0)-ηR0-η-1R1-4εx2W0R1,A12=-2x2R1W0-2x2R0W0-2εx2R1(2+2R0-W0)-2εx2R0(1-ε-W0),A21=-2x2W02-2εx2W0(W0+1+ε)-2ε(11-εR0-η-1+2W0x2)(W1+3W0-1),A22=-2Nx2H0+12x2H1+η2R0-η-1R1-3x2H0+3R0-η+2x2W02-2x2W0(W1+3W0-1)+2εx2W0(W0+1+ε)-2εx2(W1+3W0-1)(2+2R0-W0).

From (4.199), equations (4.202)–(4.203) and Lemmas 4.14.6, we have the relations

H0=1x2+O(ε),H1=2W0(0)W1(0)+O(ε)=2x1-2x+O(ε),

and therefore

A11=-2N1+W1(0)W0(0)+O(ε)+2+O(ε),A12=-2R1(0)W0(0)-2+O(ε),A21=-2+O(ε),A22=-2N1+W1(0)W0(0)+O(ε)-4+21W0(0)-2W1(0)W0(0)+O(ε).

Since R0(0)=W0(0)=1x, W1(0)W0(0)=x-2 and R1(0)=-1x, the claimed behavior of Aij follows.

Since N2 and xxcrit+κ>2, the determinant of AN has a lower bound

14detAN=N(x-1+O(ε))+O(ε)N(x-1+O(ε))-1+O(ε)+O(ε)N(x-1-Cε0))-Cε0N(x-1-Cε0)-1-Cε0-Cε0

for some universal constant C>0. For a sufficiently small ε0>0 we have N(x-1-Cε0))-Cε032 for N2 and xxcrit+κ>2. We see that detAN>0 and hence AN is invertible for all 0<εε0 with ε0 chosen sufficiently small. It immediately follows that detAN=O(N2). Since AN is invertible, relations (4.264)–(4.265) follow by multiplying (4.250) by AN-1 from the left.

Remark 4.15

A simple consequence of the previous lemma is the existence of a universal constant β0>0 such that for any x[xcrit+κ,xmax] and sufficiently small 0<εε0 the following bounds hold:

|RN|β0N|SN|+εN|VN|, 4.266
|WN|β0N|VN|+1N|SN|. 4.267

Remark 4.16

It is a routine to check

R2=-x2+6x-72x(2x-3)+O(ε), 4.268
W2=-5x2+19x-172x(2x-3)+O(ε), 4.269

for any x[xcr+κ,xmax].

Remark 4.17

In Lemma 4.14, the lower bound xcrit+κ for the x-interval has been chosen for convenience to ensure O(ε) disturbance of the coefficients to the corresponding ones to LP type solutions. It can be relaxed to xcrit by replacing O(ε) by O(ε). See the change of the distance of W1 near xcrit in Lemma 4.10.

Series Convergence and Local Existence Around a Sonic Point

Theorem 4.18

(Local existence around the sonic point) There exist an ε0>0 and r>0 sufficiently small such that for all 0<εε0 and all x[xcrit(ε)+κ,xmax(ε)] the sequence {RN,WN}NZ0 of RLP-type (see Definition 4.11) has the following property: the formal power series

R(z):=N=0RN(δz)N,W(z):=N=0WN(δz)N 4.270

converge for all z(1-r,1+r) and functions zR(z) and zW(z) are real analytic inside |z-1|<r. We can differentiate the infinite sums term by term, the pair z(R(z),W(z)) solves (4.172)–(4.173) for |z-1|<r, and R(z) is strictly positive for |z-1|<r.

Proof

The proof is analogous to Theorem 2.10 of [15]. By Lemma A.9, using α(1,2), there exists C>1 such that

N=1|RN||δz|N+N=1|WN||δz|N2N=1|Cδz|NCN3<

when |δz|<1C=:r. The claim follows by the comparison test. The real analyticity and differentiability statements are clear. Recalling (4.193), we may rewrite B as

B=N=0(R-η)N(δz)N-x2N=0HN(δz)N(1+2δz+(δz)2)=(R-η)1-x2(H1+2H0)δz+N=2(R-η)N-x2(HN+2HN-1+HN-2)(δz)N. 4.271

By (4.202)–(4.203)

(R-η)1-x2(H1+2H0)=-ηR0-η-1R1-x2(2(1+ε)W0+2εW1+4εW0R1+2(1-ε)W02+4εW0+4εR0W0+ε2-ε)=-2x2W0(W1+W0)-2εP0, 4.272

for all sufficiently small ε, where

P=11-εR0-η-1R1+x2W0W1+2W1+2W0R1+4W0+3W02+ε-1.

Therefore, it is now easy to see that for all sufficiently small ε and for r>0 sufficiently small, the function B0 for all |z-1|<r and z1. As a consequence, R(z) and W(z) are indeed the solutions as can be seen by plugging the infinite series (4.270) into (4.197) and (4.198).

Lemma 4.19

There exist an ε0>0 and r>0 sufficiently small such that for any α(1,2) and for all x[xcrit(ε)+κ,xmax(ε)] the sequence {RN,WN}NN0 of RLP-type (see Definition 4.11) has the following property: There exists a constant C=C(x,α)>0 such that for all 0<εε0 the bounds |xR0|, |xW0|, |xR1|, |xW1|C hold and

|xRN|CN-αN3,N2, 4.273
|xWN|CN-αN3,N2. 4.274

In particular, the formal power series

N=0xRN(δz)N,N=0xWN(δz)N

converge for all z satisfying |z-1|<r. Moreover, the function x(xcrit+κ,xmax)(R(z;x),W(z;x)) is C1 and the derivatives xR and xW are given by the infinite series above.

Proof

From Lemma 4.1 and Lemma 4.6, xR0(0)=xW0(0)=-1x2 and xR1(0)=1x2, xW1(0)=2x2, and it is clear that |xR0|, |xW0|, |xR1|, |xW1|C for x[xcrit+κ,xmax] and for all sufficiently small ε. For N2, xRN and xWN are recursively given by differentiating the expression in (4.264) and (4.265):

xRN=xA22detANSN+A22detANxSN-xA12detANVN-A12detANxVN, 4.275
xWN=xA11detANVN+A11detANxVN-xA21detANSN-A21detANxSN. 4.276

When N=2, the claim immediately follows. For N3, we will apply the same induction argument used for RN,WN bounds. To this end, we first observe that from Lemma 4.13 and Lemma 4.14

xA11=-2N(1+O(ε))+O(ε),xA12=O(ε), 4.277
xA21=O(ε),xA22=-2N(1+O(ε))+O(ε), 4.278
xdetAN=4N(1+O(ε))(2N((x-1)+O(ε))-1)+O(ε), 4.279

leading to

xA22detAN,xA11detAN1N,xA12detANεN2,xA21detAN1N2.

Hence using Lemmas A.8 and A.9,

xRN1N|SN|+1N|xSN|+εN2|VN|+εN2|xVN|CN-αN3+1N|xSN|+εN2|xVN|,xWN1N|VN|+1N|xVN|+1N2|SN|+1N2|xSN|CN-αN3+1N|xVN|+1N2|xSN|. 4.280

We now recall that SN and VN consist of sum and product of polynomials in R0,W0,,RN-1,WN-1 and power functions of R0. When we differentiate with respect to x, at most one term indexed by Ri or Wi, 0iN-1 is differentiated. In particular, the same combinatorial structure in the problem is maintained and the same inductive proof relying on the already established bounds (1.828) and (1.829) gives (4.273) and (4.274). The remaining conclusions now follow easily.

The Sonic Window and xmin

The goal of this subsection is to define xmin and to define the sonic window, which serves as the basic interval in our shooting method in the next section. We begin with the following lemma.

Lemma 4.20

Consider the RLP type solution constructed in Theorem 4.18. There exist a small constant δ0>0 independent of ε and 0<z0=z0(δ0)<1 such that for x=2+δ0>xcrit

W(z0)>12-2η

for all sufficiently small ε>0. Here we recall η=2ε1-ε.

Proof

For any x(xcrit,xmax), we Taylor-expand W in z around z=1:

W(z)=W(1)+W(1)(z-1)+W(z~)2(z-1)2 4.281

for some z~[z,1]. We note that W(1)=W0 and W(1)=W1. Let δ>0 be a small constant independent of ε to be fixed and set x=2+δ. Then we have

W0=12+δ+O(ε),W1=δ2+δ+O(ε), 4.282

see the end of Remark 4.5. On the other hand, from (4.269) we have

W(1)=2W2=1-δ-5δ2(2+δ)(1+2δ)+O(ε). 4.283

Now let

z0(δ,ε)=min{z1:W(z)14for allz[z1,1]}.

Observe that z0<1 by (4.283) for ε,δ sufficiently small. We now claim that there exists a small enough δ0>0 such that z0(δ0,ε)1-δ01/4 for all sufficiently small ε. Suppose not. Then for all δ>0 and for some ε0>0, z0(δ,ε0)>1-δ1/4. Thus there exists 1-δ1/4<z1=z1(δ,ε0)<1 so that W(z1)<14, but this is impossible because of (4.283) and the continuity of W. Now the Taylor expansion (4.281) at z=1-δ01/4 gives rise to

W(1-δ01/4)=12+δ0-δ02+δ0δ01/4+W(z~)2δ01/2+O(ε)12+δ0-δ02+δ0δ01/4+18δ01/2+O(ε) 4.284

from which we deduce W(1-δ01/4)>12-2η for all sufficiently small ε.

We now define

xmin:=2+δ0 4.285

where δ0 is given in Lemma 4.20. We observe that by construction xmin is independent of ε. We are ready to introduce the sonic window:

Definition 4.21

(The sonic window) For any 0εε0 we refer to the interval [xmin,xmax(ε)] as the sonic window. We often drop the ε-dependence when the ε is fixed.

Remark 4.22

Observe that by construction, the sonic window [xmin,xmax(ε)] is a strict subset of the interval [2, 3], while the interval [xcrit(ε),xmax(ε)] coincides with the interval [2, 3] when ε=0. The latter is precisely the range of possible sonic points within which we found the Newtonian Larson-Penston solution in [15] and our new sonic window [xmin,xmax(ε)] shows that the lower bound x=2 can be improved to 2+δ0 even for the Newtonian problem.

For future use in Sections 5 and 6, we analyse the behaviour of J-xW and J-xD near the sonic point x.

Lemma 4.23

(Initialisation) Let ε(0,ε0], where ε0>0 is a sufficiently small constant given by Theorem 4.18. There exist a δ>0 and c0>0 such that c0ε0<δ and for any x[xmin,xmax], the unique local RLP-type solution associated to x given by Theorem 4.18, satisfies the bounds

(a)
J[x;D]<xW,x(x,x+δ), 4.286
J[x;D]>xW,x(x-δ,x). 4.287
(b)
J[x;D]>xD,x(x-2δ,x-c0ε0), 4.288
J[x;D]>xD,x(x+c0ε0,x+2δ). 4.289
(c)
Moreover, the following bound holds
x(1-W(x))|x=x+δ12,for allε(0,ε0]. 4.290

Proof

Proof of part (a). Let g(x):=xW-J. Since g is a smooth function of D and W, by Theorem 4.18, g is smooth near the sonic point. Note that g(x)=0 from Lemma 4.1. Since g=W+xW-J, using Lemma 4.1, Lemma 4.6 and (3.127), we deduce that

g(x)=1-1x+O(ε)>13for allx[xmin,xmax]

for all sufficiently small ε>0. Therefore, g is locally strictly increasing and (4.286) and (4.287) hold for some δ>0.

Proof of part (b). Since J-xD=0 at x=x, we use this and the formula (3.127) to conclude

(J-(xD))|x=x=ε-2(1+D0)xD0+(1-ε)x-2x2D0D1+11-εD0-η-1D1(1-ε)xD0+2εx(1+D0)-xD1-D0, 4.291

where we recall (2.43). By Remark 4.5 and (4.177), we have D0=W0=1x+O(ε) and D1=-1x2+O(ε). Plugging this into the above expression, we conclude that

f(x)=(J-(xD))|x=x=O(ε).

Similarly, f=J-2D-xD and thus using D(x)=2x2R2=2x2-x2+6x-72x(2x-3)+O(ε) (see (4.268)) and J(x)=O(ε), we have

f(x)=2x2-x2x2-x2+6x-72x(2x-3)+O(ε)=x2-2x+1x2(2x-3)+O(ε)

and hence f(x)>19 for all x[xmin,xmax]. Since f is uniformly continuous, there exists a δ>0 such that

f(x)>118forx(x-δ,x+δ) 4.292

for all x[xmin,xmax].

We now Taylor-expand f at x=x to obtain

f(x)=O(ε)(x-x)+f(x~)2(x-x)2,x(x-δ,x+δ)

for some x~ between x and x. For x>x and for some c1>0 we have

f(x)>-c1ε(x-x)+136(x-x)2=136(x-x)x-x-36c1ε.

Therefore we deduce (4.289) with c0=36c1. An analogous argument gives (4.288).

Proof of part (c). Bound (4.290) follows trivially from Theorem 4.18 and the asymptotic behaviour as ε0 in Lemma 4.1.

Singularity at the Origin x=0

By analogy to the previous section, we Taylor-expand the solution at the origin z=0 in order to prove a local existence theorem starting from the origin to the right. An immediate consistency condition follows from the presence of 1-3Wz in (4.173): W(0)=13 and R(0)=R~0>0 is a free parameter. Denote the solution from the left by R- and W- and assume that locally around z=0

R-(z;R~0):=N=0R~NzN,W-(z;R~0):=N=0W~NzN 4.293

where R~0>0 is a free parameter and W~0=13. The following theorem asserts that the formal power series converge and hence (R-,W-) are real analytic in a small neighbourhood of z=0.

Theorem 4.24

Let R~0>0 be given. There exists an 0<r~<1 such that the formal power series (4.293) converge for all z[0,r~). In particular, R- and W- are real analytic on [0,r~). We can differentiate the infinite sums term by term and the functions R-(z,R~0) and W-(z,R~0) solve (4.172) and (4.173) with the initial conditions R-(0;R~0)=R~0, W-(0;R~0)=13.

Proof

Around the origin z=0 we write out the formal expansion of B:

B=R-η-x2Hz2=j=0(R-η)jzj-x2j=0Hjzj+2. 4.294

By the Faa di Bruno formula (4.186)–(4.188)

(R-η)j=R~0-ηm=1j1R~0mπ(j,m)(-η)m1λ1!λj!R~1λ1R~jλj,j1 4.295

and (R-η)0=R~0-η. Plugging (4.293) into (4.172), we obtain the formal relation

0==0(R-η)z-x2=0Hz+2m=0(m+1)R~m+1zm+2x2(1-ε)=0(R(W+ε)(R-W))z+1=N=0+m=N(m+1)R~m+1(R-η)zN-x2N=0+m=N-2(m+1)R~m+1HzN+2x2(1-ε)N=0(R(W+ε)(R-W))N-1zN. 4.296

Comparing the coefficients, we obtain

+m=N(m+1)R~m+1(R-η)-x2+m=N-2(m+1)R~m+1H+2x2(1-ε)(R(W+ε)(R-W))N-1=0. 4.297

Similarly, plugging (4.293) into (4.173), we obtain

0==0(R-η)z-x2=0Hz+2m=0(m+1)W~m+1zm+3m=0W~m+1zm-2x2(1+ε)=0(W(W+ε)(R-W))z+1=N=0+m=N(m+4)W~m+1(R-η)zN-x2N=0+m=N-2(m+4)W~m+1HzN-2x2(1+ε)N=0(W(W+ε)(R-W))N-1zN. 4.298

Comparing the coefficients, we obtain

+m=N(m+4)W~m+1(R-η)-x2+m=N-2(m+4)W~m+1H-2x2(1+ε)(R(W+ε)(R-W))N-1=0. 4.299

Identities (4.297) and (4.299) give the recursive relationships

R~N+1=1N+1S~N+1,N0, 4.300
W~N+1=1N+4V~N+1,N0, 4.301

where

S~N+1=R~0η[-+m=NmN-1(m+1)R~m+1(R-η)+x2+m=N-2(m+1)R~m+1H 4.302
-2x2(1-ε)(R(W+ε)(R-W))N-1],V~N+1=R~0η[-+m=NmN-1(m+4)W~m+1(R-η)+x2+m=N-2(m+4)W~m+1H+2x2(1+ε)(W(W+ε)(R-W))N-1], 4.303

where S~N+1 and V~N+1 depend only on (R~i,W~i) for 0iN. The rest of the proof is now entirely analogous to the proof of Theorem 4.18 and we omit the details.

Remark 4.25

We may repeat the same procedure as in Lemma 4.19 to deduce that R~0R-(z;R~0) and R~0W-(z;R~0) have the convergent power series near the origin and the function R~0(0,)(R-(z;R~0),W-(z;R~0)) is C1. And the derivatives R~0R- and R~0W- satisfy the system of ODEs obtained by the differentiating (4.172) and (4.173) with initial conditions R~0R-(0;R~0)=1 and R~0W-(0;R~0)=0.

The Friedmann Connection

Sonic Time, A Priori Bounds, and the Key Continuity Properties of the Flow

We denote xr by r~, where r is the analyticity radius given by Theorem 4.18.

Definition 5.1

(Sonic time) For any x[xmin,xmax] consider the unique local solution on the interval [x-r~,x+r~] given by Theorem 4.18. The sonic time s(x) is given by

s(x):=infx(0,x]{(D(·,x),W(·;x))is a solution to (2.35)--(2.36) on[x,x)andB(x;D,W)>0for allx[x,x)}, 5.304

where we recall the definition (3.123) of B.

Lemma 5.2

Let r~>0 be as above. Then

13<J[x;13]xfor allx[0,xmax-r~]. 5.305

Proof

By (3.124), bound (5.305) is equivalent to showing

19+23η(1+13)<ε+(13)-η(1-ε)x2. 5.306

Since xxmax-r~=3+O(ε)-r~ the right-hand side above is larger than ε+(13)-η(1-ε)(3+O(ε)-r)2, which converges to 1(3-r)2 as ε0+. The left-hand side on the other hand converges to 19 and thus the claim follows.

Lemma 5.3

For any x[xmin,xmax] consider the unique local solution on the interval [x-r~,x+r~] given by Theorem 4.18. Then for any x(s(x),x] the following bounds hold:

W(x)<J(x)x, 5.307
|W(x)|<J(x)x+2η(1+D), 5.308
0<D(x), 5.309

where J is defined in (3.124). Moreover, for any x(s(x),x-δ] such that D(x)13 we have the upper bound

D(x)<J(x)x, 5.310

where 0<δ1 is an ε-independent constant from Lemma 4.23.

Proof

Let x[xmin,xmax] and let x˚(s(x),x). By Definition 5.1 there exists a κ>0 such that B(x)>κ for all x[x˚,x-r~), which according to (3.123) is equivalent to the bound

J-xWJ+2ηx(1+D)+xW>κ1-ε.

If W(x)>0 then from the strict positivity of J and the above bound we immediately have W(x)<J(x)x. If on the other hand W(x)0 then J-xW>0 and therefore from the above bound again |xW|=-xW<J+2ηx(1+D). The two bounds together imply

|W(x)|J(x)x+2η(1+D),x(s(x),x), 5.311

which shows (5.308). The strict positivity of D on (s(x),x] follows by rewriting the equation (2.35) in the form

ddxlogD=-2x(1-ε)(W+ε)(D-W)B. 5.312

Finally, to prove (5.310), we observe that it suffices to show f>0 where we recall the formula f=J-xD. If D(x)=13, by Lemma 5.2, we are done. If D(x)>13, we consider two cases. First suppose D>13 on [x,x-δ]. Then b<0 on [x,x-δ] by Lemma 3.9 and (5.307), and f(x-δ)>0 by Lemma 4.23. Hence, by using Corollary 3.8, we have

f(x)>f(x-δ)exx-δa[z;D,W]dz>0.

If D13 on [x,x-δ], there should exist x1(x,x-δ] such that D>13 on [x,x1) and D(x1)=13. Note that b<0 on [x,x1) and f(x1)>0 by Lemma 5.2. By using Corollary 3.8 again, we obtain

f(x)>f(x1)exx1a[z;D,W]dz>0, 5.313

which proves the claim.

Lemma 5.4

Let ε(0,ε0], where ε0>0 is a small constant given by Theorem 4.18. For any x[xmin,xmax] consider the unique local RLP-type solution associated to x given by Theorem 4.18. If D(x)13 for some x(s(x),x-δ), then

D(x)<1x1-ε. 5.314

Proof

Since D(x)13, by Lemma 5.3, we have DJ[x;D]<1x. Using the definition of J[·;D] (3.124) it is easy to see that the inequality DJ[x;D]<1x is equivalent to

η2(1+D)2x2+εx2+D-η1-ε<εx2+D-η1-εDx-η(1+D)x.

This in turn implies

εx2+D-η1-ε<(εx2+D-η1-ε)2D2x2-2η(1+D)D(εx2+D-η1-ε)1<(εx2+D-η1-ε)D2x2-2η(1+D)D(1+3ε)D2+4εD-ε(1-ε)Dη<1x2.

Now we note that

(1+3ε)D2+4εD-ε(1-ε)=D2+ε3D2+4D-(1-ε)D2

where we have used D13. This implies (5.314).

Remark 5.5

It is a priori possible that the solution blows-up at a point at which B remains strictly positive, for example through blow-up of W. It is trivial to see that this cannot happen in the Newtonian setting, but in the relativistic case it requires a careful argument, which is given in the next lemma.

Lemma 5.6

(No blow up before the sonic point) For any x[xmin,xmax] consider the unique RLP-type solution on the interval (s(x),x]. If s(x)>0, then

lim infxs(x)+B(x)=0. 5.315

Proof

Assume the opposite. In that case there exists a constant κ>0 such that B(x)κ for all x(s(x),x-r~]. Our goal is to show that |D(x)|+|W(x)|< on (s(x),x], which would lead to the contradiction.

Step 1. Boundedness of D. If D(x)13 the bound (5.310) gives D+η(1+D)<η2(1+D)2+ε+D-η(1-ε)x2, which upon taking a square and using 2ηD(1+D)>0 leads to

D2<ε+D-η(1-ε)x2<ε+D-η(1-ε)s(x)2.

Since 0<η1, this gives a uniform upper bound on D on (s(x),x-r~]

D(x)M,x(s(x),x]. 5.316

Step 2. Boundedness of |W|. It follows from (2.35)–(2.36) that there exists a sufficiently large value N=N(s(x),x)>0 such that W<0, D>0 if W>N and W>0, D>0 if W<-N, where we use the already shown upper bound on D. In both cases, the two regions are dynamically trapped and we denote the union of the two regions by IN. For any xIN we multiply (2.35) by 1+ε(1-ε)D, (2.36) by 1W, and sum them to obtain

logD2(1+η)W2=2x1W-3<0for|W|>N, 5.317

with N sufficiently large. In particular, for any s(x)<x1<x2<x, where x1,x2 both belong to the invariant region IN above, we obtain

D(x1)2(1+η)W(x1)2>D(x2)2(1+η)W(x2)2 5.318

On the other hand, since Bκ, |W|>N, and D<M there exists a universal constant constant C=C(s(x),x) such that for a sufficiently large choice of N, rom (2.37) we have

D-η>CW2,xIN. 5.319

We apply this to (5.318) to conclude that

1CD(x1)2+η>D(x2)2(1+η)W(x2)2,for anyx1(s(x),x2).

This gives a lower bound for D and therefore an upper bound for |W| via (5.319) in the region IN. In particular lim supxs(x)+|W(x)|< and the claim follows.

Essentially a consequence of the previous lemma and a standard ODE argument is the statement that as long as the sonic denominator B is bounded below by some constant δ>0 for all xx¯>s(x), we can extend the solution to the left to some interval [x¯-t,x], where t>0 depends only on δ and x¯. The statement and the proof are analogous to Lemma 4.3 in [15] and we state it without proof.

Lemma 5.7

Let x[xmin,xmax] be given and consider the unique RLP-type solution (D(·;x),W(·;x)) to the left of x=x, given by Theorem 4.18. Assume that for some x¯(0,x-r~) and δ>0 we have x¯>s(x) and the conditions

B(x)>δ,D(x)>0,x[x¯,x-r~], 5.320

hold. Then there exists a t=t(δ,x¯)>0 such that the solution can be continued to the interval [x¯-t,x] so that

B(x)>0,D(x)>0,x[x¯-t,x-r~].

Proposition 5.8

Let x[xmin,xmax] be given and consider the unique RLP-type solution (D(·;x),W(·;x)) to the left of x=x.

(a)
(Upper semi-continuity of the sonic time). Then
lim supx~xs(x~)s(x),
i.e. the map xs(x) is upper semi-continuous. In particular, if s(x)=0 then the map s(·) is continuous at x.
(b)
([Continuity of the flow away from the sonic time]) Let {xn}nN[xmin,xmax] and x[xmin,xmax] satisfy limnxn=x. Let x-r~>z>max{s(x),supnNs(xn)}. Then
limnW(x;xn)=W(x;x),limnD(x;xn)=D(x;x).
(c)
Let {xn}nN[xmin,xmax] and x[xmin,xmax] satisfy limnxn=x. Assume that there exist 0<x^<x-r~ and κ>0 such that s(xn)<x^ for all nN and the following uniform bound holds:
B[x;xn,W,D]>κ,nN,x[x^,x-r~]. 5.321
Then there exists a T=T(κ,x^)>0 such that
s(x)<x^-T,s(xn)<x^-T,nN. 5.322

The Friedmann Shooting Argument

The basic idea of this section is inspired by a related proof in our earlier work on the existence of Newtonian Larson-Penston solutions [15]. We start by recalling Definition 2.2.

Lemma 5.9

(X>13 and X13 are nonempty) There exists an 0<ε01 sufficiently small so that the following statements hold for any 0<εε0:

(a)

There exists a κ=κ(ε)>0 such that (xmax(ε)-κ,xmax(ε)]XX13.

(b)

Moreover, xminX>13.

Proof

Proof of part (a). We use the mean value theorem to write

W(x;x)=W0+W(x¯;x)(x-x),x(s(x),x)

for some x¯(x,x). Note that W(x;x)=1xW1=1x1-2x+O(ε)=1x-2x2+O(ε). By (4.245), it follows that W(xmax(ε);xmax(ε))=19+O(ε). By Theorem 4.18 there exist small enough r>0 and δ1>0 such that W(x;x)>118 for all x(x-r,x] and x[xmax(ε)-δ1,xmax(ε)]. For such x and x, we have

W(x;x)1x+118(x-x),

where we have used W0=W0<1x for ε(0,ε0], see Lemma 4.1. Note 1x+118(x-x)=13 when x=x~(x)=x-6(3-x)x. Therefore for all x(xmax(ε)-κ,xmax(ε)](3-κ,3) with κ=min{δ1,xmax(ε)-3+3r6+r}, there exists an xx~(x) such that W(x;x)=13, which shows the claim. Here we have used (4.246) and the smallness of 0<ε01.

Proof of part (b). We rewrite (2.36) in the form

xW=1-2W-WB-2x2(1+ε)(W+ε)(D-W)B. 5.323

We now recall Lemma 3.6 and express the numerator above in the form

B-2x2(1+ε)(W+ε)(D-W)=(1-ε)(J[D]-xW)(H[D]+xW)-2(1+ε)x(W+ε)(xD-J[D]+J[D]-xW)=2(1+ε)x(W+ε)f+(J[D]-xW)(1-ε)(H[D]+xW)-2(1+ε)x(W+ε).

where we recall f(x)=J[x;D]-xD. Using (3.125),

(1-ε)(H[D]+xW)-2(1+ε)x(W+ε)=(1-ε)(J[D]+2η(1+D)x+xW)-2(1+ε)x(W+ε)=(1-ε)(J[D]-xW)+4ε(xD-xW)+2ε(1-ε)x=(1+3ε)(J[D]-xW)-4εf+2ε(1-ε)x.

Putting these together, and by (3.123) and W>0, (5.323) reads as

xW=1-2W+2ηfH[D]+xWW-W(1+3ε)(J[D]-xW)2+2(1+ε)x(W+ε)(J[D]-xD)+2ε(1-ε)(J[D]-xW)xB1-(2-2η)W, 5.324

where we have used 0<f=J[D]-xD<H[D]+xW and B>0 for x(s(x),x). Hence the set

x(s(x),x))|W(x;x)>12-2η 5.325

is an invariant set.

We now use Lemma 4.20 and the definition of xmin to obtain W(x0;xmin)>12-2η for some x0(s(x),x)). Together with the invariance of (5.325) we conclude that xminX>13.

Remark 5.10

A corollary of Definition 2.2 and Lemma 5.9 is that the fundamental set X is a simply connected subset of (xmin,xmax] which contains the point xmax. The set X is in particular an interval.

Of fundamental importance in our analysis are the following two quantities.

Definition 5.11

(Critical point and Friedmann time) Let ε0>0 be a small constant given by Theorem 4.18. For any ε(0,ε0] we introduce the critical point

x¯:=infxXx, 5.326

and the Friedmann time

xF=xF(x):=infx(s(x),x)|W(τ;x)>13forτ(x,x). 5.327

We will show later, the unique local RLP-type solution associated with the sonic point x¯ is in fact global and extends all the way to the origin x=0. The Friedmann time introduced above plays a crucial role in the proof.

The sets X13 and X<13 enjoy several important properties which we prove in the next lemma.

Lemma 5.12

Let ε0>0 be a small constant given by Theorem 4.18. Then for any ε(0,ε0] the following statements hold.

(a)
For any xX13X<13 we have
W(x;x)<D(x;x),x(s(x),x), 5.328
W(x;y)<13,x(s(x),xF(x)), 5.329
where (5.329) is considered trivially true in the case s(x)=xF(x).
(b)

For any xX13 we have W(xF(x);x)>0. Moreover, the set X13 is relatively open in [xmin,xmax].

Proof

Proof of Part (a). Let xX13X<13. Then from Lemma 4.6 we know that W(x)<D(x) for all x[(1-r¯)x,x) for some r¯r where r is given by Theorem 4.18. We first prove (5.328). By way of contradiction, assume that there exists xc(s(x),x) such that

W(xc)=D(xc)andW(x)<D(x),x(xc,x).

We distinguish three cases.

Case 1: xc(xF,x). In this case, from (2.35) and (2.36), we have W(xc)<0 and D(xc)=0. In particular, (D-W)(xc)>0 and locally strictly to the left of xc we have

W<0,W-D>0,D>0,W>13.

We observe that these conditions are dynamically trapped and since W<0, we deduce that W stays strictly bounded away from 13 from above for all x(xF,x), which is a contradiction to the assumption xX13X<13.

Case 2: xc=xF. In this case, xX13 necessarily and W(xc)=D(xc)=13. On the other hand, since D-W>0 for x(xF,x), it follows D<0 on (xF,x) from (2.35). Hence, D(xc)>D(x)=W013 since x[xmin,xmax]. This is a contradiction.

Case 3: xc(s(x),xF). In this case, xX13 necessarily. Since xc<xF we know that D-W>0 locally around xF. Therefore, we have

W>0,D-W>0,W<13on(xF-δ,xF)

for a sufficiently small δ>0. These conditions are dynamically trapped and we conclude that D-W>0 on (s(x),xF). This is a contradiction, and hence completing the proof of (5.328).

Inequality (5.329) follows by a similar argument, since the property W>0,D-W>0,W<13 is dynamically preserved and all three conditions are easily checked to hold locally to the left of xF(x).

Proof of Part (b). For any xX13 by part (a) and (2.36) we have W(xF(x);x)>0. Therefore there exists a sufficiently small δ>0 so that W(x;x)<13 for all x(xF(x)-δ,xF(x)). By Proposition 5.8, there exists a small neighborhood of x such that W(x;x)<13 for some x(xF(x)-δ,xF(x)). Therefore X13 is open.

Lemma 5.13

Let ε0>0 be a small constant given by Theorem 4.18 and for any x[xmin,xmax] consider the unique local RLP-type solution given by Theorem 4.18. If xF(x)=s(x)>0 then necessarily

W(xF(x);x)>13. 5.330

In particular if xX, then necessarily s(x)<xF(x).

Proof

Assume the claim is not true, in other words W(xF(x);x)=13. Since D is decreasing on (xF,x-δ] by Lemma 5.12 and (2.35) and since D(x)1/3 it follows that limxxFD(x)>13. In particular, since JxD on (s(x),x-δ) by (5.310), it follows from (3.123) that B(x;x)(1-ε)x(D-W)(H(x)+xW). Therefore

lim infxxFB(x)>0,

a contradiction to the assumption xF(x)=s(x) due to Lemma 5.6. The second claim is a consequence of the just proven claim and the definition of the set X.

Lemma 5.14

Let ε0>0 be a small constant given by Theorem 4.18 and for any xX>13[xmin,xmax] consider the unique local RLP-type solution given by Theorem 4.18. Then there exists a monotonically increasing continuous function

m:(0,x-r~](0,),

such that for any X(0,x-r~]

B(x;x)>m(X)>0for allx(max(X,s(x)),x-r~]. 5.331

Proof

Since xX13, by Lemma 5.12J-xWJ-xD=f on (s(x),x-r~). Since H+xWJ-xWf we conclude from (3.123) that B(1-ε)f2. Since W(x;x)13 for all x[xF(x),x-r~], it follows from Lemma 5.12 that D>13 on the same interval. Therefore, by Lemma 3.9 we conclude that b[x;D,W]<0 on [xF(x),x-r~]. By Corollary 3.8 we then conclude

f(x)f(x-r~)exx-r~a[z;D,W]dz,x[xF(x),x-r~]. 5.332

From (3.132) it is clear that on [xF(x),x-r~],

a2=2εJ-xWD-1+2f+4εx+xD5+εZ-1=2εJ-xWD+1+2xW+4εx+xD3+εZ-1>0.

Similarly, the first line of (3.131) is strictly positive on the same interval, and we conclude from (5.332) and (3.131) that

f(x)f(x-r~)e-2εxx-r~11-εD-ηz+(1-ε)z+(D+ε)fZ-1dz,x[xF(x),x-r~]. 5.333

From (2.46), W13 on [xF(x),x-r~], and HJ we have Zmax{23εz2,13Jz} on [xF(x),x-r~]. Similarly,

2ε11-εD-ηz+(1-ε)z+(D+ε)fZ-12ε3η1-ε1z+(1-ε)x32εz2+2ε(M+ε)J13JzC11z3+1z2+1zC21z3,

for and some C1,C2>0 where we recall the upper bound M of D in (5.316). Combining the two bounds and plugging it back into (5.333) we get

f(x)f(x-r~)e-C2xx-r~1z3dzf(x-r~)e-C22x-2. 5.334

Together with the established lower bound B(1-ε)f2 we conclude the proof.

Proposition 5.15

The unique RLP-type solution associated with x¯(xmin,xmax) exists globally to the left, i.e. s(x¯)=0.

Proof

Case 1. We assume xF(x¯)=0. Since 0s(x¯)xF(x¯) we are done.

Case 2. We assume xF(x¯)>s(x¯)>0. In this case x¯X which is impossible, since X is relatively open in [xmin,xmax] by Lemma 5.12 and xminX by Lemma 5.9.

Case 3. We now assume that xF(x¯)=s(x¯)>0. By Lemma 5.13 this in particular implies that

x¯[xmin,xmax]\X13 5.335

and

W(xF(x¯);x¯)>13.

Consider now a sequence {xn}nNX such that limnxn=x¯. We consider

x¯F:=lim supnxF(xn)

and after possibly passing to a subsequence, we assume without loss that limnxF(xn)=x¯F. We now consider two possibilities.

Case 3 a). Assume that x¯F>0. Since {xn}nNX, by Lemma 5.13 necessarily s(xn)<xF(xn), nN. Upon possibly passing to a further subsequence, we can ascertain that there exists some δ>0 such that x¯F>xF(xn)>x¯F-δ>0 for all nN. By Lemma 5.14 we conclude in particular that

B(x;xn)>m(x¯F-δ)>0for allx[xF(xn),x-r~],nN. 5.336

Therefore, by part (c) of Proposition 5.8, there exists T=T(m(x¯F-δ),x¯F) such that

s(x¯)<x¯F-T,s(xn)<x¯F-T,nN. 5.337

Fix an x(x¯F-T,x¯F) and observe that

W(x;x¯)=limnW(x,xn)13, 5.338

where we have used Lemma 5.12. This implies x¯X13, a contradiction to (5.335).

Case 3 b). Assume that x¯F=0. For any fixed x^>0 we can apply the argument from Case 3 a) to conclude that s(x¯)<x^. Therefore s(x¯)=0 in this case.

Lemma 5.16

(Continuity of X13xxF(x)) The map

X13xxF(x)

is continuous and

limXxx¯xF(x)=0=xF(x¯). 5.339

Proof

The proof is nearly identical to the proof of Lemma 4.13 in [15].

The Solution from the Origin x=0 to the Right

In this section we consider the solutions (D-,W-) to (2.35)–(2.36) generated by the data imposed at x=0 and satisfying W-(0)=13. Upon specifying the value D0=D-(0)>0, the problem is well-posed by Theorem 4.24 on some interval [0, r].

Definition 5.17

We introduce the sonic time from the left

s-(D0):=supx0{x|J[x;D-(x;D0)]-xW-(x;D0)>0}. 5.340

Lemma 5.18

Let D-(0)=D0>13, W-(0)=13 and x[xmin,xmax]. The solution (D-(x;D0),W-(x;D0)) to (2.35)–(2.36) with the initial data

D-(0,D0)=D013,W-(0;D0)=13, 5.341

exists on the interval [0,s-(D0)) and satisfies the following bounds: for x(0,s-(D0))

W->13, 5.342
D-1-ε+W-1+ε<D01-ε+13(1+ε), 5.343
D-1+εW-1-ε<D-1+ε31-ε, 5.344
D->W-, 5.345
D-<0. 5.346

Proof

The proof is analogous to the proof of Lemma 4.14 from [15].

In the following lemma we identify a spatial scale x01D01+O(ε) over which we obtain quantitative lower bounds on the density D- over [0,x0].

Lemma 5.19

(Quantitative lower bounds on D-) Let D0>13 and x[xmin,xmax] be given and consider the unique solution (D-(·;D0),W-(·;D0)) to the initial-value problem (2.35)–(2.36), (5.341). For any D0>13 let

x0=x0(D0):=31-ε2212(1+8ε)12D011-ε,D0>1;31-ε2212(1+8ε)12,13<D01. 5.347

Then s-(D0)>x0 for all D0>13 and

D-(x;D0)D0exp-D0-1+ε,D0>1;D0exp-1,13<D01,x[0,x0]. 5.348

Moreover, for all sufficiently small ε>0, there exists a D^>1 such that for all D0>D^ we have

D-(x0;D0)>1x01-ε. 5.349

Proof

Equation (2.35) is equivalent to

D-(x)=D0exp-0x2τ(1-ε)(W-+ε)(D--W-)Bdτ. 5.350

Since W-=(W-1+εW-1-ε)12<(D-1+εW-1-ε)12 and D-W-=(D-1+εW-1-ε)11+εW-2ε1+ε, by Lemma 5.18, we have the following bounds on the interval (0,s-(D0))

W-<D-<D0, 5.351
13<W-<D01+ε31-ε12, 5.352
D-W-<D01+ε31-ε. 5.353

Now from the definition (2.37) of B, for any 0xx0 using (5.351)-(5.353) and the definition of x0 in (5.347), we have

D-2ε1-εB=1-D-2ε1-ε(1-ε)W-2+4εD-W-+4εW-+ε2-εx21-D02ε1-ε(1+3ε)D01+ε31-ε+4εD01+ε31-ε12+ε2-εx21-D02ε1-ε(1+8ε)D01+ε31-εx21-12D01-ε12,D0>1;1-D01+ε+2ε1-ε212,13<D01. 5.354

Note from the second line to the third line we have used the upper bound of [·] term

D01+ε31-ε(1+3ε)D01+ε31-ε+4εD01+ε31-ε12+ε2-ε(1+8ε)D01+ε31-ε, 5.355

which holds true for D0>13 and sufficiently small 0<ε. Therefore, s-(D0)>x0 for all D0>13.

From (5.351),  (5.353), and (5.354) for any x[0,x0] we obtain

0x2τ(1-ε)(W-+ε)(D--W-)Bdτ4(1-ε)D02ε1-ε(D01+ε31-ε+εD0)0xτdτ2x02(1-ε)D02ε1-ε(D01+ε31-ε+εD0)=1D01-ε(1-ε)(D01+ε31-ε+εD0)(1+8ε)D01+ε31-ε,D0>1;D0(1-ε)D02ε1-ε(D0ε+ε31-ε)(1+8ε),13<D01.

Now comparing the denominator and numerator of the last fractions, we see that

(1+8ε)D01+ε31-ε-(1-ε)(D01+ε31-ε+εD0)=ε9D01+ε31-ε-(1-ε)D00for allD0>1

and

1+8ε-(1-ε)D02ε1-ε(D0ε+ε31-ε)1+8ε-(1-ε)(1+ε31-ε)=ε(9-(1-ε)31-ε)0for all13<D01

and hence we deduce that for any x[0,x0]

0x2τ(1-ε)(W-+ε)(D--W-)Bdτ1D01-ε,D0>1;1,13<D01.

Plugging this bound in (5.350) we obtain (5.348).

To show (5.349), we first rewrite (5.347) for D0>1 as

D0=1x01-ε31-ε2(1+8ε)1-ε2. 5.356

Using this in (5.348)

D-(x0;D0)D0e-1D01-ε=1x01-ε31-ε2(1+8ε)1-ε2e-1D01-ε. 5.357

The bound (5.349) follows if 31-ε2(1+8ε)1-ε2e-1D01-ε>1 which is clearly true for sufficiently large D0 and sufficiently small ε.

Remark 5.20

The specific choice of x0 in (5.347) has been made to ensure the lower bound in (5.349) to be compared with the bound (5.314) satisfied by the solutions emanating from the sonic point. In fact, better bounds are obtained by choosing different x0. For instance, if x0=O(1D01+α) for ε1-ε+ε2+12<1+α11-ε in (5.347), then we may deduce that for such x0, D-(x0;D0)>1x011+α1x01-ε with the equality being valid for α=ε1-ε for all sufficiently large D0 and small ε.

Remark 5.21

Since the mapping D0x0(D0) from (5.347) is nonincreasing, it follows that for any fixed D0>13 we have the uniform bound on the sonic time:

s-(D~0)>x0(D~0)x0(D0),for all13<D~0D0.

The following lemma shows the crucial monotonicity property of D-(·;D0) with respect to D0 on a time-scale of order D0-(34+O(ε)).

Lemma 5.22

Let x[xmin,xmax]. There exists a sufficiently small κ>0 such that for all D013

D0D-(x;D0)>0for allx[0,κD0-b],whereb=3+ε+2η4

Proof

We introduce the short-hand notation D-=D0D- and W-=D0W-, where we note that the map D0(D-,W-) is C1 by Remark 4.25. It is then easy to check that (D-,W-) solve

W-=-3xW--2x(1+ε)W-(W-+ε)BW-+2x(1+ε)(D--W-)B2W-+ε-DWBBW-(W-+ε)W- 5.358
+2x(1+ε)W-(W-+ε)B1-DDBB(D--W-)D-,D-=-2x(1-ε)(W-+ε)B2D--W--DDBBD-(D--W-)D--2x(1-ε)D-BD--2W--ε-DWBB(W-+ε)(D--W-)W-, 5.359

where

DWB=-[2(1-ε)W+4ε(1+D)]x2,DDB=-2ε1-εD-2ε1-ε-1-4εWx2.

At x=0 we have the initial values

D-(0)=1,W-(0)=0. 5.360

We multiply (5.358) by W- and integrate over the region [0, x]. By (5.360) we obtain

12W-2(x)+0x3τ+2τ(1+ε)W-(W-+ε)BW-2dτ=0x2τ(1+ε)(D--W-)B2W-+ε-DWBBW-(W-+ε)W-2dτ+0x2τ(1+ε)W-(W-+ε)B1-DDBB(D--W-)D-W-dτ. 5.361

Just like in (5.354), by using (5.351)–(5.353), and (5.355) we have

D-2ε1-εB1-D02ε1-ε(1+3ε)D01+ε31-ε+4εD01+ε31-ε12+ε2-ετ21-(1+8ε)D02ε1-εD01+ε31-ετ2.

Therefore

D-2ε1-εB12,for anyτ[0,(8D01+ε+2ε1-ε)-12] 5.362

where we have used 2(1+8ε)3ε-1<8 for all xxmax and sufficiently small ε.

Using the bounds xxmax, (5.351)–(5.353), (5.362), we obtain from (5.361)

12W-2(x)+0x3τ+2τ(1+ε)W-(W-+ε)BW-2dτC0xD01+ε+2ε1-ετW-2dτ+CD01+ε+2ε1-ε0xτ|D-||W-|dτ,x(8D01+ε+2ε1-ε)-12 5.363

for all sufficiently small ε. Here we have used -DDBB(D--W-)>0 and DDBBW-<0, and

0<-DDBB(D--W-)<-DDBBD-=-DDBD-ηBD-1+ηC1ε(1+D-1+ηW-τ2)C1ε(1+D01+ε+η31-ετ2)C2ε

for any τ[0,(8D01+ε+2ε1-ε)-12] so that 1-DDBB(D--W-) is bounded by a constant.

Let X^=κD0-b where b=3+ε+4ε1-ε4 with a sufficiently small κ>0 to be specified later. Note that X^<(8D01+ε+2ε1-ε)-12 for all D013 and κ chosen sufficiently small and independent of D0. For any τ[0,X^] we have D0κ1bτ-1b. Therefore D01+ε+2ε1-ετκ1b(1+ε+2ε1-ε)τ-1b+1. From these estimates and (5.363) we conclude for x[0,X^],

12W-2(x)+0x3τ+2τ(1+ε)W-(W-+ε)BW-2dτC0xκ1b(1+ε+2ε1-ε)τ-1b+1W-2dτ+C3D01+ε+η0x3τW-2dτ120xτ3D-2dτ12C0xκ1b(1+ε+2ε1-ε)τ-1b+1W-2dτ+120x3τW-2dτ+C26D02+2ε+2ηD-20xτ3dτ. 5.364

Since -1b+1>-1, with κ chosen sufficiently small, but independent of D0, we can absorb the first two integrals on the right-most side into the term 0x3τW-2dτ on the left-hand side. We then conclude

W-(x)CD01+ε+ηx2D-,x[0,X^]. 5.365

We now integrate (5.359), use (5.351)–(5.353), (5.362), and conclude from (5.360)

D-(x)-1CD01+ε+2ε1-εD-0xτdτ+C0xD02+ητ+D02+2ε+2η+εD03+ε+2ητ3W-(τ)dτCD01+ε+ηx2+D03+ε+2ηx4+(D03+3ε+3η+εD04+2ε+3η)x6D-Cκ2D-,x[0,X^], 5.366

for all sufficiently small ε, where we have used (5.365) and 0xκD0-b. Therefore,

D-1+Cκ2D-

and thus, for κ sufficiently small so that Cκ2<13, we have D-32. From here we infer

D-(x)1-32Cκ2>12>0,x[0,X^].

Upper and Lower Solutions

Definition 5.23

(Upper and lower solution) For any x[xmin,xmax] we say that (D(·;x),W(·;x)) is an upper (resp. lower) solution at x0(0,x) if there exists D0>0 such that

D(x0;x)=D-(x0;D0)

and

W(x0;x)>(resp.<)W-(x0;D0).

Lemma 5.24

(Existence of a lower solution) There exists a κ>0 such that for any x0<κ there exists an x[x¯,xmax] such that (D(·;x),W(·;x)) is a lower solution at x0. Moreover, there exists a universal constant C such that D1<Cx01-ε, where D-(x0;D1)=D(x0;x).

Proof

For any xX we consider the function

S(x):=supx~[x¯,x]xF(x~). 5.367

The function xS(x) is clearly increasing, continuous, and by Lemma 5.16, limxx¯S(x)=0. Therefore, the range of S is of the form [0,κ] for some κ>0. For any xX, by Lemma 5.16, the supremum in (5.367) is attained, i.e. there exists x[x¯,x] such that S(x)=xF(x)=:x0. Therefore, for any x¯<x<x we have

s(x)<xF(x)x0.

By Lemma 5.12 we have the bound D(x0;x)>W(x0;x)=13. By Lemma 5.19 choosing D0=D0(x0)=1x01-ε31-ε2(1+8ε)1-ε2 and using Lemmas 5.4 and 5.3 we have

D-(x0;D0)>1x01-ε>J[x0;D]x0>D(x0;x),

where we have assumed x0 to be sufficiently small. On the other hand D-(x0;13)=13<D(x0;x) (where we recall that D-(·;13) is the Friedmann solution, see (2.40)). Using Remark 5.21 and the Intermediate Value Theorem, there exists a D1(13,D0) such that

D(x0;x)=D-(x0;D1).

By (5.342) W-(x0;D1)>13=W(x0;x) and therefore (D(·,x),W(·;x)) is a lower solution at x0. The upper bound on D1 follows from our choice of D0 above.

Remark 5.25

The proof of Lemma 5.24 follows closely the analogous proof in the Newtonian case (Lemma 4.20 in [15]).

Lemma 5.26

Let xX>13 (see (2.44) for the definition of X>13) and assume that s(x)=0. Then

  1. D(x;x)>W(x;x),x(0,x);
  2. lim supx0x1-εW(x;x)>0.

Proof

Proof of Part (a). If not let

xc:=supx(0,x)D(τ;x)-W(τ;x)>0,τ(x,x),D(x;x)=W(x;x)>0.

At xc we have from (2.35)–(2.36) W(xc;x)=1-3Wxc<0 since W>13 for xX>13, and D(xc;x)=0. Therefore there exists a neighbourhood strictly to the left of xc such that W<0, D<W, and D>0. It is easily checked that this property is dynamically trapped and we conclude

W(x;x)1-3W(x;x)x,xxc. 5.368

Integrating the above equation over [x,xc] we conclude

W(x;x)x3ω(xc;x)xc3-13xc3=W(xc;x)-13xc3=:c>0. 5.369

We now recall (5.317). From (5.369), this implies that logD2(1+η)W2<0 since xX>13. In particular we obtain the lower bound

D(x)2(1+η)W(x)2>D(xc)2(1+η)W(xc)2=W(xc)4+2η>3-(4+2η),xxc.

It follows that

D-η<3η(2+η)1+ηWη1+η,xxc. 5.370

On the other hand, bound (5.369) implies

x2>c23W-23,xxc. 5.371

From (2.37), (5.370)–(5.371) we therefore have

B<3η(2+η)1+ηWη1+η-CW43,xxc. 5.372

for some universal constant C>0. Since W grows to infinity as x0, this implies that the right-hand side above necessarily becomes negative, i.e. s(x)>0. A contradiction.

Proof of Part (b). Since D>W by part (a) and W>13 (since xX>13), from Lemmas 5.3 and 5.4

D1x1-ε. 5.373

By way of contradiction we assume that lim supx0x1-εW(x;x)=0. For any ε~>0 choose δ>0 so small that

x1-εW(x;x)<ε~,x(0,δ). 5.374

Note that in particular, for any x(0,δ) we have

x2WD1+η=x1-εWx1+εD1+ηε~,x(0,δ), 5.375

where we have used (5.373) and (5.374). Similarly,

x2WDη=x1-εWx2εDηx1-εε~δ1-ε,x(0,δ). 5.376

We next claim that there exists a universal constant C¯ such that

2(1+ε)x2(W+ε)(D-W)C¯B. 5.377

Keeping in mind (2.37), this is equivalent to the estimate

x2W2DηC¯(1-ε)-2(1+ε)+x2WDη-2ε(1+ε)+2(1+ε)D+4εC¯(1+D)+x2Dη2ε(1+ε)D+C¯(ε2-ε)C¯. 5.378

For any 0<C¯<2 the first term on the left-hand side of (5.378) is strictly negative, and so are -2ε(1+ε)x2WDη and C¯(ε2-ε)x2Dη. On the other hand

x2WDη2(1+ε)D+4εC¯(1+D)=x2WD1+η2(1+ε)+4εC¯+4εCx¯2WDηε~2(1+ε)+4εC¯+4εC¯ε~δ1-ε, 5.379

where we have used (5.375) and (5.376). Similarly,

2ε(1+ε)x2D1+η2ε(1+ε)x1-ε2ε(1+ε)δ1-ε, 5.380

where we have used (5.373). It is thus clear that we can choose C¯ of the form C¯=C~(ε~+δ1-ε) for some universal constant C~>0 and ε~,δ sufficiently small, so that (5.377) is true. It then follows from (2.36) that

W1-3Wx+C~(ε~+δ1-ε)Wx=1-(3-C~(ε~+δ1-ε))Wx. 5.381

As a consequence of (5.381), for sufficiently small ε~,δ>0 and x(0,δ) we have

W-2Wx,

which in turn implies W(x;x)Cx-2 for some C>0 and sufficiently small x, a contradiction.

Lemma 5.27

(Existence of an upper solution) If

limx0W(x;x¯)13,

then there exists a universal constant C>0 and an arbitrarily small x0>0 such that (D(·;x¯),W(·;x¯)) is an upper solution at x0 and D1<Cx0, where D-(x0;D1)=D(x0;x¯).

Proof

It is clear that lim infx0W(x;x¯)13 as otherwise we would have x¯X, a contradiction to the definition (5.326) of x¯ and the openness of X. We distinguish three cases.

Case 1.

lim infx0W(x;x¯)>13.

In this case x¯X>13 and by Proposition 5.15 we have s(x¯)=0. By part (b) of Lemma 5.26 there exists a constant C>0 and a sequence {xn}nN(0,1) such that limnxn=0 and

W(xn;x¯)>Cxn1-ε 5.382

where C is independent of n. For any such xn we have by part (a) of Lemma 5.26 and Lemma 5.4

D(xn;x¯)>W(xn;x¯)>Cxn1-ε>CD(xn;x¯) 5.383

where we have used the assumption x¯X>13 and part (a) of Lemma 5.26 to conclude D(·;x¯)>13, which is necessary for the application of Lemma 5.4.

For any 0<xn1 sufficiently small consider (D-(·;D0,n),W-(·;D0,n)) with D0,n=D0(xn)=1xn1-ε31-ε2(1+8ε)1-ε2>1 for n large enough. By Lemmas 5.19 and 5.26

D-(xn;D0,n)>1xn1-ε>D(xn;x¯)>W(xn;x¯)>13.

On the other hand,

D(xn;x¯)>13=D-(xn;13),

where we recall that D-(·;13)13 is the Friedmann solution. Moreover, by Remark 5.21[0,xn][0,s-(D~0)) for all D~0[13,D0,n]. By the continuity of the map [13,D0,n]D~0D-(xn;D~0) the Intermediate Value Theorem implies that there exists Dn,1(13,D0,n) such that

D-(xn;D1n)=D(xn;x¯)for all sufficiently largenN. 5.384

Let

c1,n:=exp-Dn,1-1+ε,ifDn,1>1;exp-1,if13<Dn,11.

Clearly c1,ne-1=:c1 for all nN. By Lemma 5.19D-(xn;Dn,1)c1Dn,1. Using (5.344)–(5.345) we also have W-(xn;Dn,1)<Dn,11+ε23-1-ε2. Together with (5.383) and (5.384) we conclude that for all n sufficiently large

W-(xn;Dn,1)<Dn,11+ε23-1-ε2<D-(xn;Dn,1)c11+ε23-1-ε2=D(xn;x¯)c11+ε23-1-ε2<3-1-ε2(c1C)-1+ε2W(xn;x¯)1+ε2. 5.385

By (5.383) W(xn;x¯) grows to positive infinity as xn approaches zero. Therefore, we may choose a sufficiently large NN and set x0=xN1, D0=D0,N, D1=D1,N so that the right-hand side of (5.385) is bounded from above by W(x0;x¯). This gives

W-(x0;D1)<W(x0;x¯).

We conclude that (D(·;x¯),W(·;x¯)) is an upper solution (see Definition 5.23) at x0 and the upper bound on D1 follows from our choice of D0.

Case 2.

13<lim supx0W(x;x¯)<,lim infx0W(x;x¯)=13. 5.386

In particular x¯X<13 (see (2.46)) and by Lemma 5.12D(x;x¯)>W(x;x¯). Assumption (5.386) also implies that there exists a constant c>0 independent of x such that

W(x;x¯)<c,x(0,x¯]. 5.387

We now claim that there exists a constant C such that

D-CDxε, 5.388

for all sufficiently small x. To prove (5.388) we note that by (2.35) this is equivalent to the bound x1+ε(W+ε)(D-W)CB, which, by (2.37) is equivalent to the bound

x2W2Dη-1x1-ε+C(1-ε)+x2WD1+η1x1-ε+4εC)+x2WDη-εx1-ε+4εC+x2Dηεx1-ε+C(ε2-ε)C. 5.389

Using (5.387) and Lemma 5.4, we have

x2W2Dηc2x2-2ε,x2WD1+ηcx1-ε,x2WDηcx2-2ε,x2Dηx2-2ε.

Choosing ε>0 sufficiently small, x so small that -1x1-ε+(1-ε)<0, and C>1 sufficiently large, but independent of ε, we use the above bounds to conclude the claim.

Bound (5.388) gives logD+C1-εx1-ε0, which in turn gives the bound

D(x)D(x¯)eC1-εx¯1-ε-C1-εx1-ε,x(0,x¯], 5.390

for some x¯ sufficiently small. This immediately implies the uniform boundedness of D(·;x¯), i.e.

D(x;x¯)<c,x(0,x], 5.391

where we have (possibly) enlarged c so that (5.387) and (5.391) are both true. There exists an δ>0 and a sequence xnnN such that limnxn=0 and

13+δ<W(xn;x¯),andlimnW(xn;x¯)=lim supx0W(x;x¯).

Since {D(xn;x¯)}nN is bounded, by Lemma 5.19 we can choose an D0>1 such that D-(xn;D0)>D(xn;x¯) for all nN. On the other hand D(xn;x¯)>13=D-(xn;13). By the intermediate value theorem there exists a sequence {D0,n}nN(13,D0) such that

D-(xn;D0,n)=D(xn;x¯).

Since W-(x;D0,n)2D0,n1+ε31-ε<D01+ε31-ε and D-(x;D0,n)<D0,n<D0 (Lemma 5.18) we conclude from (2.35)–(2.36) and Theorem 4.24 that D-(xn;D0,n) and W-(xn;D0,n) are bounded uniformly-in-n, by some constant C~>0. Therefore

W-(xn;D0,n)13+C~xn.

We thus conclude that for a fixed n sufficiently large W-(xn;D0,n)13+δ<W(xn;x¯). Therefore, W(·;x¯) is an upper solution (see Definition 5.23) at x0:=xn with D1=D0,n. The claimed upper bound on D1 is clear.

Case 3.

13<lim supx0W(x;x¯)=,lim infx0W(x;x¯)=13.

As W(·;x¯) must oscillate between 13 and we can use the mean value theorem to conclude that there exists a sequence xnnN such that limnxn=0 and

W(xn;x¯)>n,andW(xn;x¯)=0. 5.392

We claim that there exist N0>0 and 0<η1 such that

W(xn;x¯)12xn1-ε,nN0. 5.393

To prove this, assume that (5.393) is not true. Then there exists a subsequence of {xn}nN such that W(xn;x¯)<12xn1-ε. We now rewrite (2.36) in the form

W=Wx1W-(3-γ)+-γB+2(1+ε)x2(W+ε)(D-W)B, 5.394

where γ(0,3) is a control parameter to be chosen below. We use (2.37) to evaluate

-γB+2x2(1+ε)(W+ε)(D-W)=x2γ(W+ε)2-ε(W-1)2+4εDW-2(1+ε)(W+ε)(W-D)-γD-η=x2{γ(1-ε)-2(1+ε)W2+4γε-2(1+ε)ε+4γε+2(1+ε)DW+γ(ε2-ε)+2ε(1+ε)D}-γD-η. 5.395

Since D13, by Lemma 5.4 we also have the bound D(x)1x1-ε. We may therefore estimate

Dη-γB+2x2(1+ε)(W+ε)(D-W)x2Dηγ(1-ε)-2(1+ε)W2+4γεW+4γε+2(1+ε)x2WD1+η+2ε(1+ε)x2D1+η-γ. 5.396

By (5.392), for any γ(0,32) we have γ(1-ε)-2(1+ε)W(xn;x¯)2+4γεW(xn;x¯)<0 for all sufficiently large n. We use the upper bounds on (D,W) along the sequence {xn}nN to obtain

x2WD1+η12andx2D1+ηx1-ε,

and therefore

Dη-γB+2x2(1+ε)(W+ε)(D-W)|x=xn2γε+1+ε+2ε(1+ε)xn1-ε-γ<0 5.397

for some γ(1,32) and all sufficiently small ε. Feeding this back into (5.394), we conclude

W(xn;x¯)W(xn;x¯)xn1W(xn;x¯)-(3-γ)<0, 5.398

for all sufficiently large n, where we use the first bound in (5.392), which contradicts the second claim in (5.392). We can therefore repeat the same argument following (5.382) to conclude that W(·;x¯) is an upper solution at x0:=xn, for some n sufficiently large. The upper bound on D1 follows in the same way.

We next show that W(·;x¯) takes on value 13 at x=0.

Proposition 5.28

The limit limx0W(x;x¯) exists and

limx0W(x;x¯)=13.

Proof

Assume that the claim is not true. By Lemmas 5.24 and 5.27 we can find a 0<x01 and xX so that (D(·;x),W(·;x)) and (D(·;x¯),W(·;x¯)) are respectively a lower and an upper solution at x0. Without loss of generality let

A:=D(x0;x)<D(x0;x¯)=:B.

By Lemmas 5.24 and 5.27 there exist DA,DB>13 such that A=D-(x0;DA), B=D-(x0;DB), and DA,DB(13,D0), where D01 and x0C1D0. Therefore by Lemma 5.22, D0D-(x0;D~0)>0 for all D~0[13,D0], since D0-3+ε+2η4D0-1 for D0 large and all εε0 with ε0 sufficiently small. By the inverse function theorem, there exists a continuous function τg(τ) such that

D-(x0;g(τ))=τ,τ[A,B],g(DA)=A.

By strict monotonicity of D~0D-(x0;D~0) on (0,D0] the inverse g is in fact injective and therefore g(DB)=B. We consider the map

[x¯,x]xW(x0;x)-W-(x0;g(D(x0;x)))=:h(x).

By the above discussion h is continuous, h(A)<0 and h(B)>0. Therefore, by the Intermediate Value Theorem there exists an xs(x¯,x) such that h(xs)=0. The solution (D(·;xs),W(·;xs)) exists on [0,xs], satisfies W(0)=13 and belongs to X. This is a contradiction to the definition of X.

In the next proposition we will prove that the solution (D(·;x),W(·;x)) is analytic in a left neighbourhood of x=0, by showing that it coincides with a solution emanating from the origin.

Proposition 5.29

There exists a constant C>0 so that

D(x;x¯)+W(x;x¯)+W(x;x¯)-13x2C,x(0,x¯].

The solution D(·;x¯):(0,x¯]R>0 extends continuously to x=0 and D:=D(0;x¯)<. Moreover, the solution (D(·;x¯),W(·;x¯)) coincides with (D-(·;D),W-(·;D)) and it is therefore analytic at x=0 by Theorem 4.24.

Proof

By Proposition 5.28 it is clear that W(·;x¯) is bounded on [0,x¯]. Moreover, DW13 on [0,x¯] and therefore by Lemma 5.4D1x1-ε. We may now apply the exact same argument as in the proof of (5.391) to conclude that D is uniformly bounded on [0,x¯]. Since B>0 and D>W on [0,x¯], by (2.35) D0 and therefore the limit D:=limx0D(x;x¯) exists and it is finite. It is also clear that there exists a constant c0>0 such that

B(x)>c0,x[0,x¯-r~]. 5.399

Let

ζ=W-130.

It is then easy to check from (2.36)

ζx3=x32(1+ε)xW(W+ε)(D-W)B

and therefore, by (5.399) and the uniform boundedness of W and D, for any 0<x1<x we obtain

ζ(x)x3-ζ(x1)x13Cx1xτ4dτ=C5x5-x15,x(0,x¯-r~].

We now let x10+ and conclude

ζ(x)x2. 5.400

Consider now D¯(x):=D(x;x¯)-D-(x;D) and W¯(x):=W(x;x¯)-W-(x;D), both are defined in a (right) neighbourhood of x=0 and satisfy

D¯=O(1)D¯+O(1)W¯,W¯=-3W¯x+O(1)D¯+O(1)W¯,

where D¯(0)=W¯(0)=0. Here we have used the already proven boundedness of (D(·;x¯),W(·;x¯)) and the boundedness of (D-(·;D),W-(·;D)), see Lemma 5.18. We multiply the first equation by D¯, the second by W¯, integrate over [0, x] and use Cauchy-Schwarz to get

D¯(x)2+W¯(x)2+30xW¯(τ)2τdτC0xD¯(τ)2+W¯(τ)2dτ. 5.401

We note that 0xW¯(τ)2τdτ is well-defined, since W¯=ζ-ζ-, where ζ-=W--13; we use (5.400) and observe ζ-x2 in the vicinity of x=0 by the analyticity of W-, see Theorem 4.24. Therefore D¯(x)2+W¯(x)2=0 by (5.401). The analyticity claim now follows from Theorem 4.24.

Remark 5.30

Propositions 5.28 and 5.29 follow closely the arguments in the Newtonian case (Propositions 4.22 and 4.23 in [15]). However, an important difference is the use of (5.391) in the proof of Proposition 5.29.

Proof of Theorem 2.3. Theorem 2.3 is now a simple corollary of Propositions 5.155.28, and 5.29.

The Far Field Connection

The goal of this section is to construct a global-to-the-right solution of the dynamical system (2.35)–(2.36) for all x in the sonic window [xmin,xmax]. We refer to Section 2.4 for a detailed overview.

A Priori Bounds

Remark 6.1

The condition J>xD is equivalent to the inequality

D-η>x2(1-ε)D2-ε(1-ε)+4εD(1+D), 6.402

which can be seen directly from the definition (3.124) of J[x;D]. Similarly, the inequality xW>J is equivalent to

D-η<x2(1-ε)W2-ε(1-ε)+4εW(1+D). 6.403

Lemma 6.2

(A priori bounds to the right) Let x[xmin,xmax]. Let X>x be the maximal time of existence to the right of the associated RLP-type solution on which we have

xW(x)>J[x;D],x(x+δ,X), 6.404
J[x;D]>xD(x),x(x+δ,X), 6.405

where 0<δ1 is an ε-independent constant from Lemma 4.23. Then the following claims hold

(a)
D<0,x(x+δ,X), 6.406
W>16,x(x+δ,X), 6.407
D(x)W(x)>Dδ2x+δx3,x(x+δ,X), 6.408
where
Dδ:=D(x+δ).
(b)
There exists a constant C~ independent of X and ε0>0 such that for all 0<εε0
W(x)C~,x(x+δ,X). 6.409
(c)
There exist constants C>0, 0<ε01 such that for all ε(0,ε0)
DxD-γ,x(x+δ,X),γ:=1-Cε>0 6.410
and therefore
D(x)Dδxx+δ-γ. 6.411
In particular, with 0<ε01 sufficiently small, we may choose γ=1-Cε0 uniform for all ε(0,ε0).

Proof

We note that by Lemma 4.23 there exists a δ>0 such that X>x+2δ.

Proof of part (a). Notice that by assumptions (6.404)–(6.405) we have

W>D,x(x+δ,X). 6.412

Bound (6.406) follows from (2.35), (3.123), (6.404), and (6.412).

By (2.36) and (6.404)–(6.405) we conclude W1-3Wx or equivalently (Wx3)x2. We conclude that for any x(x+δ,X) we have

W(x)Wδx+δx3+131-x+δx316, 6.413

since Wδ:=W(x+δ)=W0(ε,x)-O(δ)W0(ε;xmax(ε))-O(δ)=13-O(δ), for 0<δ1 sufficiently small and independent of ε. Here we have also used Remark 4.2 and the definition of xmax in (4.245).

From (2.35)–(2.36) we have

x1-βWDxβ=DWx+DWx+βDW=D+(β-3)DW+4εx2DW(W+ε)(D-W)B. 6.414

We may let β=3 and conclude from (6.404), (6.412) that ddxWDxβ>0 on (x+δ,X). In particular

D(x)W(x)>Dδ2x+δx3,x(x+δ,X), 6.415

which is (6.408).

Proof of part (b). Let now γ(0,3) and rewrite (2.36) in the following way:

W=1-(3-γ)Wx+WxB-γB+2x2(1+ε)(W+ε)(D-W). 6.416

We use (2.37) to evaluate

-γB+2x2(1+ε)(W+ε)(D-W)=x2γ(W+ε)2-ε(W-1)2+4εDW-2(1+ε)(W+ε)(W-D)-γD-η=x2{γ(1-ε)-2(1+ε)W2+4γε-2(1+ε)ε+4γε+2(1+ε)DW+γ(ε2-ε)+2ε(1+ε)D}-γD-η=:Gγ[x;D,W]. 6.417

By (6.408) we know that for any x(x+δ,X), D(x)>Dδ2(x+δ)3x-3W(x)-1, and therefore

D-η<Dδ-2η(x+δ)-3ηx3ηW(x)η,x(x+δ,X). 6.418

Fix now any γ(2(1+ε)1-ε,3), which is possible since ε can be chosen small. The expression Gγ[x;D,W] in (6.417) can be bounded from below with the help of (6.418) by

g[x,W]:=x2(1-ε)γ-2(1+ε)1-εW2+x2χ1[D]W+x2χ2[D]-C0(x)x3ηW(x)η, 6.419

where

χ1[D]:=4γε-2(1+ε)ε+4γε+2(1+ε)D,χ2[D]:=γ(ε2-ε)+2ε(1+ε)D,C0(x):=γDδ-2η(x+δ)-3η.

By the already proven bound (6.406) it follows that D(x)Dδ. Therefore, from the above formulas, there exists a universal constant C such that

|χ1[D]|+|χ2[D]|+|C0(x)|C,x(x+δ,X),x[xmin,xmax].

Since x>x+δ1 and for sufficiently small ε we have 3η<2 (recall (2.38)), it follows that there exists a constant C¯=C¯(γ) independent of X and ε such that

g[x,W]>1,ifW(x)>C¯. 6.420

Let now C~=max{C¯,23-γ}. It then follows that both summands on the right-hand side of (6.416) are strictly negative when WC~; the first term is negative due to 1-(3-γ)C<0 and the second one due to (6.420) and the negativity of B. Therefore the region WC~ is dynamically inaccessible. This completes the proof of (6.409).

Proof of part (c). For any γR it follows from (2.35) that

DxD+γ=γB-2(1-ε)x2(W+ε)(D-W)B. 6.421

We focus on the numerator of the right-hand side above. By (2.37) we have

γB-2(1-ε)x2(W+ε)(D-W)=γD-η-γ(W+ε)2-ε(W-1)2+4εDWx2-2(1-ε)x2(W+ε)(D-W)γx2(1-ε)D2-ε(1-ε)+4εD(1+D)-γ(W+ε)2-ε(W-1)2+4εDWx2-2(1-ε)x2(W+ε)(D-W)=x2(W-D)γ(W-D)+(2-2γ)W-ε3γD+W(2-γ)+4γ+2ε-2. 6.422

In the third line above we crucially used the bound (6.402), which by Remark 6.1 is equivalent to the assumption (6.405). Since W>D, C~>W16, and 0<D<Dδ on (x+δ,X), we may choose γ=1-Cε with C sufficiently large, but ε-independent, so that the above expression is strictly positive, since in this case (2-2γ)WCε3. Therefore, for sufficiently small ε and γ=1-Cε, the right-hand side of (6.421) is negative, since B<0 on (x,X). This proves (6.410), which after an integration gives (6.411).

Monotonicity and Global Existence

Lemma 6.3

(Monotonicity properties of the flow) Let x[xmin,xmax]. Let X>x be the maximal time of existence to the right of the associated RLP-type solution on which we have

xW(x)>J[x;D],x(x+δ,X). 6.423

Then there exists an 0<ε01 such that for all ε(0,ε0] the associated RLP-type solution satisfies,

J[x;D]>xD(x),x(x+δ,X). 6.424

If in addition X<, then there exists a constant κ¯=κ¯(X)>0 such that f(x)=J[x;D]-xD(x)>κ¯ for all x(x+δ,X).

Proof

We observe that b1[x;D,W] is strictly positive on (x+δ,X) by our assumption. Moreover, from the definition of (3.134), it is clear that b2[x;D,W]b~2[x;D,W], where

b~2[x;D,W]=-2x2D{D2+(3+ε)D-(1-ε)+ε41-εD2+2(1+ε)1-ε(1+D)+5-ε1-εD+3}Z-1; 6.425

It is clear that for any D14 and ε0 sufficiently small the expression b~2[x;D,W] is strictly positive for all 0<εε0. If there exists an x~>x+δ such that D(x~)14, then by the monotonicity of xD(x), D(x)14 for all xx~. Thus b~2[x;D,W] and therefore b[x;D,W] remain positive. It follows that if f(x~)>0, then by Corollary 3.8f(x)>0 for all x[x~,X).

We next show that for a sufficiently small ε0 the value of D always drops below 14 at some x~ for all ε(0,ε0] so that f remains strictly positive on (x+δ,x~]. By the bound (6.411) and Dδ1 it follows that D(x)14 as long as xx41γ. Since γ=1+O(ε) and x3, then with

x~=3×411-12=48 6.426

it follows that if X>x~ and f(x)>0 on (x+δ,x~], then f(x)>0 for all x(x+δ,X), with ε0 chosen sufficiently small.

It remains to show that for all ε(0,ε0] with ε0 chosen sufficiently small, we indeed have f>0 on (x+δ,min{x~,X}). Let (x+δ,X1)(x+δ,X) be the maximal subinterval of (x+δ,X), such that f>0 on (x+δ,X1). By Lemma 4.23 we have X1x+2δ. Notice that on the interval (x+δ,X1), by the strict negativity of B the factor a1[x;D,W] (3.131) is strictly negative. Since b1 is positive and b2b~2 we have from Lemma 3.7

f+a2fεb~2.

For any x+δ<x1<x<X1 this immediately yields

f(x)f(x1)e-x1xa2[z;D,W]dz+εe-x1xa2[z;D,W]dzx1xb~2[z;D,W]ex1za2[s;D,W]dsdz. 6.427

Observe that

Z-11x2DW6x2D, 6.428

where we have used H+xWxW, (1-ε)J+2εx1+D(1-ε)xD+2εx1+DxD, and finally W16. Therefore

b~2[x;D,W]2x2DZ(D2+(3+ε)D-(1-ε)+2ε41-εD2+2(1+ε)1-ε(1+D)+5-ε1-εD+3)C, 6.429

for some universal constant C, which follows from (6.428) and the a priori bound DDδ1.

We note that by (2.46) and (3.125),

2εJ-xWD-1Z-1=2εJ-xWD-1(1-ε)J+2εx1+DH+xW=2εJ-xWD-1(1-ε)J+2εx1+DJ[D]+4ε1-ε(1+D)x+xW2εJ+xW1-Dδ(1-ε)J+2εx1+DJ+4ε1-ε(1+D)x+xW1-Dδx1, 6.430

where we have used [(1-ε)J+2εx1+D2εx, J+4ε1-ε(1+D)x+xWJ+xW, and xx1. Similarly,

2εfZJ+xD(1-ε)J+2εx1+DH+xW2εJ+xW(1-ε)xD+2εx1+DH+xW2ε12εx=1x1, 6.431

where we have used f=J-xDJ+xD in the first line, DW and J>xD in the second, and finally J+xWH+xW, (1-ε)xD+2εx1+D2εx, and xx1 in the last line. Also, since H+xWxW16x and (1-ε)xD+2εx1+D2εx, we have the following simple bound

Z-13εx2. 6.432

Since DDδ1, we can use (6.430), (6.431), (6.428), (6.432), and the definition (3.132) of a2[x;D,W] to conclude that

a2[x;D,W]2εJ-xWD-1Z-1+4εfZ+2εZ-14εx+xD5+εC. 6.433

We now feed the bounds (6.429) and (6.433) into (6.427) to conclude

eC(x-x1)f(x)f(x)ex1xa2[z;D,W]dzf(x1)-εCeC(x-x1)(x-x1),x(x1,X),x1[x+δ,X1). 6.434

Let x1=x+δ so that by Lemma 4.23f(x1)>c¯ for some c¯>0 and all ε(0,ε0] and all x[xmin,xmax]. From (6.434) we conclude that we can choose ε0 so that f(x)>c>0 for all x(x+δ,min{x~,X1}), where we recall x~=48. By the definition of X1, this concludes the proof of (6.424).

To prove the remaining claim, by (6.434) we need only to address the case when X>x~=48. In this case b~2 is strictly positive for x>x~ by the argument following (6.425) and therefore by (6.427) and (6.433)

f(x)f(x~)e-x~xa2[z;D,W]dzf(x~)e-C(x-x~), 6.435

and thus the claim follows.

We are now ready to prove Theorem 2.4, which asserts global-to-the-right existence of RLP-type solutions.

Proof of Theorem 2.4. Let [x,X) be the maximal time of existence on which

xW>J,x(x,X). 6.436

By Lemma 6.3 we conclude that J>xD on (x+δ,X) for all εε0 sufficiently small and therefore all the conclusions of Lemma 6.2 apply. We argue by contradiction and assume that X<. By Lemma 6.3 we know that there exists a constant κ¯=κ¯(X) such that f(x)>κ¯ for all x(x+δ,X). In particular lim supxX-(W(x)-D(x))>0. It follows that

lim supxX-(xW)-J0. 6.437

By (2.36) and (3.127), we have

(xW)-J=1-2W(x)+2x2(1+ε)W(W+ε)(D-W)B+2εJ(1+D)-ε(1-ε)x(1-ε)J+2εx(1+D)+(2εxJ+ε1-εD-η-1)D(1-ε)J+2εx(1+D)=1-2W(x)+2εJ(1+D)-ε(1-ε)x(1-ε)J+2εx(1+D)+-x(1+ε)W(1-ε)D+ε2xJ+11-εD-η-1(1-ε)J+2εx(1+D)D. 6.438

By the uniform bounds (6.407) and (6.409) we have 16WC~. By (6.406) and (6.408) we have

DδD(x)Dδ2(x+δ)3x3C~Dδ2(x+δ)3X3C~. 6.439

Therefore the first line of the right-most side of (6.438) is bounded. Since limxX-xW-J=0 it follows that limxX-B=0 and therefore lim supxX-D(x)=- by (2.35). On the other hand, the second line of the right-most side of (6.438) is of the form g(x)D, where

g(x)=-x(1+ε)W(1-ε)D+ε2xJ+11-εD-η-1(1-ε)J+2εx(1+D)=-xW(1+ε)(1-ε)J+2εx(1+D)+2εxJ+ε1-εD-η-1(1-ε)D(1-ε)D(1-ε)J+2εx(1+D)=-(1-ε2)xWJ-2ε(1+ε)x2W(1+D)+2ε(1-ε)xDJ+εD-η(1-ε)D(1-ε)J+2εx(1+D). 6.440

We now use (6.403), which for any x<X allows us to estimate the numerator of (6.440) from above:

-(1-ε2)xWJ-2ε(1+ε)x2W(1+D)+2ε(1-ε)xDJ+εD-η<-(1-ε2)xWJ-2ε(1+ε)x2W(1+D)+2ε(1-ε)xDJ+εx2(1-ε)W2-ε(1-ε)+4εW(1+D)=x2-(1-ε2)WJx+2ε(1-ε)DJx+ε(1-ε)W2-(2ε-2ε2)x2W(1+D)-x2ε2(1-ε). 6.441

As xX, Jx approaches W and therefore the above expression is negative for ε sufficiently small (and independent of X). Therefore by (6.440) g(x)<0 as xX-, which implies that the right-most side of (6.438) blows up to + as xX-, which is a contradiction to (6.437). This concludes the proof of the theorem.

Asymptotic Behaviour as x

We introduce the unknown Π(x)=Σ(y), where Σ is the energy density introduced in (2.19). Recalling (2.34) and (2.26), we have

Π=D1+η=D1+ε1-ε. 6.442

It is straightforward to check that the system (2.35)–(2.36) can be reformulated as follows

Π=-2x(1+ε)Π(W+ε)(D-W)B, 6.443
W=(1-3W)x+2x(1+ε)W(W+ε)(D-W)B. 6.444

Lemma 6.4

There exist an x0>x, a constant C and ε0>0 sufficiently small so that for all x>x0

Π(x)xΠ(x)+2=εα(x)(W-1)+β(x),x(x,), 6.445

where

|α(x)|Cxx0, 6.446
|β(x)|Cx-32,xx0. 6.447

Proof

From (6.443) and the definition (3.123) of B it follows that

ΠxΠ+2=2B-2x2(1+ε)(W+ε)(D-W)B=2D-η-x2(W+ε)2-ε(W-1)2+4εDW+(1+ε)(W+ε)(D-W)D-η-x2(W+ε)2-ε(W-1)2+4εDW. 6.448

We rewrite the rectangular brackets in the numerator above in the form

-(W+ε)2-ε(W-1)2+4εDW+(1+ε)(W+ε)(D-W)=ε(W-1)2W+ε-1-D5εW+W+ε(1+ε). 6.449

We feed this back into (6.448), divide the numerator and the denominator by x2, and obtain

ΠxΠ+2=2ε(W-1)2W+ε-1-D5εW+W+ε(1+ε)+D-ηx-2-(W+ε)2-ε(W-1)2+4εDW+D-ηx-2=εα(x)(W-1)+β(x), 6.450

where

α(x):=22W+ε-1-(W+ε)2-ε(W-1)2+4εDW+D-ηx-2, 6.451
β(x):=-2D5εW+W+ε(1+ε)+2D-ηx-2-(W+ε)2-ε(W-1)2+4εDW+D-ηx-2. 6.452

By (6.408) and WC~, it follows that D(x)c1x-3 for some universal constant c1>0. Together with (6.411) we conclude

c1x-3D(x)c2x-γ, 6.453

where γ=1-Cε0 for some ε01. We conclude that D-ηx-2c1-ηx3η-2. Since 16WC~ by Lemma 6.2, for εε0 sufficiently small we conclude that there exist a constant C and x0>x such that for all x>x0

1C-(W+ε)2-ε(W-1)2+4εDW+D-ηx-2C.

This immediately yields (6.446). Note that the leading order behaviour in the rectangular brackets in the very last line of (6.422) is of the form γ(W-D)+(2-2γ)W+O(ε). We use the decay bound (6.411), namely D(x)x-1+Cε, and we see that the right-most side of (6.422) becomes positive with γ=32, with x sufficiently large and ε sufficiently small. The identity (6.421) yields (6.447).

Lemma 6.5

Let x[xmin,xmax] and let (W,D) be the unique RLP-type solution defined globally to the right for all ε(0,ε0]. Then for any ε(0,ε0] there exists a constant d2>0 such that

limxW(x)=1, 6.454
limxΠ(x)x2=d2. 6.455

Claim (6.455) equivalently reads

limxD(x)x21-ε1+ε=d21-ε1+ε. 6.456

Proof

Note that by (6.444)

(W-1)x=Wx+W-1=-2W+2x2(1+ε)W(W+ε)(D-W)B=-W2B-2x2(1+ε)(W+ε)(D-W)B. 6.457

From this and the first line of (6.448) we conclude

(W-1)x=-WΠxΠ+2. 6.458

Let now

ϑ(x):=(1-W)x. 6.459

Equation (6.458) and Lemma 6.4 now give

ϑ(x)=-εW(x)α(x)xϑ(x)+W(x)β(x). 6.460

We now use the integrating factor to solve for ϑ. For any x0>x we have

ϑ(x)eεx0xW(s)α(s)sds=ϑ(x0)+x0xW(s)β(s)eεx0sW(s)α(s)sdsds. 6.461

By Lemma 6.4 and the bound 16WC~ from Lemma 6.2 we have |αW|C and therefore

xx0-Cεeεx0xW(s)α(s)sdsxx0Cε. 6.462

We use this bound in (6.461) and together with (6.447) we conclude that

xx0-Cε|ϑ(x)||ϑ(x0)|+Cx0xs-32+CεdsC. 6.463

We now recall the definition (6.459) of ϑ and conclude from the above bound that

|1-W|Cx-1+Cε,for allε(0,ε0]. 6.464

where we recall (6.410). The claim (6.454) follows.

From Lemma 6.4 we conclude

log(Π(x)x2)=log(Π(x0)x02)+x0xεα(s)s(W(s)-1)+β(s)sds, 6.465

and thus

Π(x)x2=Π(x0)x02ex0xεα(s)s(W(s)-1)+β(s)sds. 6.466

We now use (6.464) and Lemma 6.4 to conclude that εα(s)s(W(s)-1)+β(s)ss-2+O(ε) and is therefore integrable. Letting x we conclude (6.455). That (6.456) is equivalent to (6.455) follows from (6.442).

Remark 6.6

Stronger versions of the claims (6.454) and (6.455) are contained in (6.464) and (6.466) respectively. Formula (6.466) can be used to give a quantitative decay bound for

Π(x)x2-d2,asx.

In the following lemma we establish the strict upper bound W<1 for the RLP-type solutions and provide the crucial ε-independent bounds at spacelike hypersurface x=+, which will play an important role in the construction of the maximal selfsimilar extension of the RLP spacetime.

Lemma 6.7

(W stays below 1) Let x[xmin,xmax] and let (W,D) be the unique RLP-type solution defined globally to the right for all ε(0,ε0]. Then there exist a sufficiently small 0<ε01 and constants c2>c1>0 such that

W(x)<1,x[x,),for allε(0,ε0], 6.467

and

1+c2d21+c1for allε(0,ε0], 6.468

where d2=limxΠ(x)W(x)x2=limxΠ(x)x2.

Proof

We first observe that by (6.443)–(6.444)

(ΠWx2)=Πx(1-W)=Πϑ, 6.469

where we have used (6.459) in the second equality. By way of contradiction, assume that there exists an x¯(x,) such that W(x¯)=1 and W(x)<1 for all x(x,x¯). Integrating (6.469) over [x,x¯] we conclude that

Π(x¯)x¯2=Π(x¯)W(x¯)x¯2=Π(x)W(x)x2+xx¯Π(s)ϑ(s)ds=W02+ηx2+xx¯Π(s)ϑ(s)ds, 6.470

where we recall (6.442). On the other hand, clearly W(x¯)0 and therefore, by (6.444), at x¯ we have

0W(x¯)=-2x¯+2x¯(1+ε)2(D-1)D-η-(1+ε)2+4εDx¯2=-2D-η-(1+ε)2+4εDx¯2+2x¯2(1+ε)2(D-1)x¯D-η-(1+ε)2+4εDx¯2=-2D-η+2Dx¯24ε+(1+ε)2x¯D-η-(1+ε)2+4εDx¯2=-2D-η1-Πx¯24ε+(1+ε)2x¯D-η-(1+ε)2+4εDx¯2. 6.471

Since the denominator is strictly negative, it follows that

Π(x¯)x¯214ε+(1+ε)2. 6.472

Combining (6.470) and (6.472) we conclude that

14ε+(1+ε)2W02+ηx2+xx¯Π(s)ϑ(s)ds. 6.473

Our goal is to provide a lower bound for ϑ(x) which is uniform-in-ε. Using (6.461),(6.462), and (6.447), we conclude that for any x¯>x>x0>x

xx0Cε|ϑ(x)|ϑ(x0)-Cx0xs-32+Cεds. 6.474

Therefore

|ϑ(x)|ϑ(x0)xx0-Cε-Cx0xs-32+Cεdsxx0-Cε. 6.475

Now choose x0=x+δ with δ as in Lemma 4.23 so that, by Lemma 4.23 we can ascertain the uniform-in-ε bound

ϑ(x0)12,for allε(0,ε0]. 6.476

This bound together with the lower bound (6.475) yields

|ϑ(x)|12xx0-Cε-Cx0xs-32+Cεdsxx0-Cε,for allε(0,ε0]. 6.477

The right-hand side of (6.477) is a continuous function in x which converges to 12 as xx0-. Therefore, there exists an ε-independent real number x~>x0 such that

ϑ(x)14,for allx[x0,x~]and allε(0,ε0]. 6.478

Observe that the bound (6.478) also ascertains that x~<x¯, see (6.459).

It is easily checked from (6.443)–(6.444) that (ΠWx3)=Πx20. Since |W|C~ by Lemma 6.2 it then follows that

Π(x)Π(x)W(x)x3C~x3=W02+ηx3C~x3Cx-3,x>x, 6.479

where, as usual, the constant C is independent of ε.

We now use the bounds (6.478) and (6.479) in (6.473) to conclude

14ε+(1+ε)2W02+ηx2+xx¯Π(s)ϑ(s)dsW02+ηx2+x0x~Π(s)ϑ(s)dsW02+ηx2+Cx0x~s-3dsW02+ηx2+C1x02-1x~2. 6.480

Observe that we have used the non-negativity of ϑ on [x,x¯] in the second bound above.

By the ε-asymptotic behaviour of W0 from Lemma 4.1 we have

W02+ηx2=1+O(ε),

and clearly 14ε+(1+ε)2=1+O(ε). Since however the term C1x02-1x~2 is bounded from below by some ε-independent positive real number δ~>0, we obtain contradiction by choosing ε0 sufficiently small. This completes the proof of (6.467).

By Lemma 6.5 we may integrate (6.469) over the whole interval [x,) to conclude

limxΠ(x)W(x)x2=Π(x)W(x)x2+xΠ(s)ϑ(s)dsΠ(x)W(x)x2+x0x~Π(s)ϑ(s)dsW02+ηx2+C1x02-1x~21+O(ε)+δ~>1, 6.481

for εε0 sufficiently small. This proves the lower bound stated in (6.468).

To prove the upper bound on Π(x)W(x)x2, we rewrite (6.469) in the form log(ΠWx2)=1-WxW and integrate to obtain the identity

log(Π(x)W(x)x2)=log(Π(x0)W(x0)x02)+x0x1-W(s)sW(s)ds,xx0<x. 6.482

Using (6.464) and (6.407) we conclude from (6.482) that

log(Π(x)W(x)x2)log(Π(x0)W(x0)x02)+6Cx0xs-2+Cεds, 6.483

where the constant C does not depend on ε. We let x0=x, so that for sufficiently small εε0 the ε-dependent quantity Π(x)W(x) is bounded uniformly-in-ε. On the other hand

6Cxxs-2+Cεds6Cxs-2+Cεds=6C1-Cεx-1+Cε,

which, since x[xmin,xmax], is also bounded uniformly-in-ε. This completes the proof of the lemma.

In the following proposition, we establish the sharp asymptotic behaviour of the variable W as x - this will play an important role in constructing the unique extension of the solution in Section 7.

Proposition 6.8

(Precise asymptotic behaviour for W) Let x[xmin,xmax] and let (W,D) be the unique RLP-type solution defined globally to the right for all ε(0,ε0]. After choosing a possibly smaller ε0>0, there exists a constant c~ such that for any ε(0,ε0] there exists a constant w1<0 such that

1-W(x)=-w1x-1-ε1+ε+ox(x1-ε1+ε),xx, 6.484

where w1<-c~<0 for all ε(0,ε0].

Proof

We may now derive more precise asymptotics for the functions α(x) and β(x) from (6.451)–(6.452). Using (6.456) and (6.454) we see that

limxβ(x)D(x)=2(1+ε)21+ε(6+ε)-limxD-1+ε1-εx-2=2(1+ε)21+ε(6+ε)-1c, 6.485

where 1+c2>c=limxD1+ε1-εx2>1+c1 for all ε(0,ε0] by Lemma 6.7. We conclude that there exists a constant c~>0 such that for all ε(0,ε0],

1c~>limxβ(x)x-21-ε1+ε>c~. 6.486

Using (6.454) and (6.456), it is easy to see that

α(x)x-21+ε. 6.487

Recall now ϑ from (6.459). Letting W=1+(W-1)=1-ϑx, we may rewrite (6.460) in the form

ϑ(x)=2ε1+εϑ(x)x+εϑ(x)2x2α(x)-εα~(x)ϑ(x)x+W(x)β(x),α=-21+ε+α~, 6.488

with limxα~(x)=0. This yields the identity

ϑx-2ε1+ε=x-2ε1+εεϑ(x)2x2α(x)-εα~(x)ϑ(x)x+W(x)β(x). 6.489

By the bound (6.464) and by (6.486)–(6.487), it follows that the right-hand side of of the above identity is integrable on [x,) and therefore (6.484) holds for some v10 (note that v1 cannot be positive by Lemma 6.7). To prove the ε-uniform upper bound on v1, we must first estimate the rate of decay of α~(x) as x. From (6.488) and (6.451), we directly check

α~(x)=-2(1-ε)(1-W)2+4εDW-D-ηx-2(1+ε)D-ηx-2-(W+ε)2-ε(W-1)2+4εDW. 6.490

Therefore, for sufficiently large x1 we have the upper bound

|α~(x)|Kx-2+O(ε), 6.491

for some K>0 independent of ε. Here we have used (6.464), the bound (6.456), and Lemma 6.5. Keeping in mind (6.459), the bound 1-W>0 (from Lemma 6.7) and (6.486) we conclude that for any xx sufficiently large we have

-w1=limx(1-W(x))x1-ε1+ε=limxϑ(x)x-2ε1+ε=1-W(x0)x01-ε1+ε+x0εx-2ε1+εϑ(x)2x2α(x)-α~(x)ϑ(x)x+x-2ε1+εW(x)β(x)dxεx0x-2ε1+εϑ(x)2x2α(x)-α~(x)ϑ(x)xdx+C1x0x-2ε1+εx-21-ε1+εdxC1x0x-21+εdx-C2εx0x-3+Cεdx, 6.492

for some constants C1,C2>0 and independent of ε. Note that we have used W16 in the third line and (6.491) in the last. This implies the claim for ε sufficiently small.

Maximal Analytic Extension

The main results of this section are the description of the local and the global extension of the flow across the coordinate singularity at y=, see Theorems 7.4 and 2.5 respectively. For a detailed overview, we refer to Section 2.5.

Adapted Comoving Coordinates

As explained in Section 2.5 the metric g given by (2.7) becomes singular as y. To show that this is merely a coordinate singularity, we introduced the adapted comoving chart (2.54), and the new selfsimilar variable Y given by (2.55)–(2.56). A simple manipulation of (2.54) gives the relation

τ=τ(τ~,R)=Yητ~,τ<0,R>0. 7.493

Recall the new unknowns χ(Y), d(Y), and w(Y) from (2.57). In the new chart the spacetime metric (2.7) by (2.19)–(2.24), (7.493), and (2.57) takes the form

g=-e2μ~(Y)dτ~2-4ε1+εYe2μ~(Y)dτ~dR+e2λ~(Y)-4ε(1+ε)2Y2e2μ~(Y)dR2+R2χ(Y)2γ,

where we recall (2.68)–(2.69). Note that by (2.52) and (2.68), we have

e2μ~(Y)=1(1-ε)2dY-2-η,Y>0. 7.494

It is clear that in the limit y, i.e. Y0+ the metric coefficient e2μ~(Y) approaches a positive constant due to (2.53) and (2.56).

Lemma 7.1

Let the triple (r~,d,w) be a smooth solution to (3.117)–(3.118). Then the new unknowns (χ,d,w) defined in (2.57) solve the system (2.58)–(2.61).

Proof

The Lagrangian system (3.117)–(3.118) now takes the form

d=-w+ε1+ε2(1-ε)r~2yd(w+ε)(d-w)B, 7.495
w=w+ε1+ε1-3wy+2(1+ε)r~2yw(w+ε)(d-w)B, 7.496
r~=r~yw+ε1+ε, 7.497

where

B=d-η-r~2(w+ε)2-ε(w-1)2+4εwd, 7.498

and refers to differentiation with respect to y. Using (2.57) it is then straightforward to check that (2.58)–(2.61) hold.

Remark 7.2

From (7.497) and the leading order behaviour of w(y)=w(Y) at y= it is easy to see that

r~(y)=ay+oy+(y)=a1Y1+η+oY0+(1Y1+η),Y>0, 7.499
limY0+χ(Y)=a, 7.500

for some constant a>0. Here the constant a corresponds to the labelling gauge freedom and we set without loss of generality

a=1. 7.501

We note that the unknown d corresponds to the modified density d defined in (2.57). By Lemma 6.5 and Proposition 6.8, and the asymptotic behaviour (7.499) with a=1, we have the leading order asymptotic behaviour

d(Y)=d2Y2+oY0+Y2, 7.502
w(Y)=1+w1Y+oY0+Y, 7.503

where 1+c2>d2>1+c1 and

w1<-c~<0 7.504

by Lemma 6.7 and Proposition 6.8. We note that

limY0+χ(Y)2Y2+2ηw(Y)d1+η(Y)=limxΠ(x)W(x)x2=d2>1 7.505

by Lemma 6.7.

Local Extension

To prove the existence of a solution to (2.58)–(2.60) with suitable boundary conditions (see (7.528)–(7.530)), we formally Taylor-expand the unknowns d,w,χ around Y=0 and prove the convergence of the series. Assume that the following expansions hold

d(Y)=d2Y2+N=3dNYN, 7.506
w(Y)=1+N=1wNYN, 7.507
χ(Y)=χ0+N=1χNYN, 7.508

where by (7.501) χ0=1. From this and (2.61) we have the formal expansion

C=N=0CNYN, 7.509

where for any N0

CN=(dY-2)-ηN-2-k+=N(χ2)k(1-ε)(w2)+4εw+(ε2-ε)δ0+4ε(wd). 7.510

Here we employ the convention that (f)j=0 if j<0. We may single out the leading order contribution on the right-hand side of (7.510):

CN=-4εχ02dN-2(1+ε)χ02wN-2(1+ε)2χ0χN+CN,1, 7.511

where

CN,1:=(dY-2)-ηN-2-k+=Nk,N-1(χ2)k(1-ε)(w2)+4εw+(ε2-ε)δ0+4ε(wd)-χ02(1-ε)k+=Nk,N-1wkw+4εk+=NN-1wkd. 7.512

Lemma 7.3

For any N1, NN, the Taylor coefficients (dN,wN,χN) satisfy the following recursive relations

(N-2)χ02(1+ε)2dN=UN, 7.513
(1-ε)(1+ε)2(N-1)χ02wN-2(1+ε)(1+ε)2+4εχ02dN=VN, 7.514
χ0wN+(1-ε)NχN=-k+=Nk,N-1χkw, 7.515

where

UN=k+j=NkN-1kdkCj-2k++m+n=NN-1(χ2)kd(dm-wm)(w2)n+2εwn+ε2δ0n,N3, 7.516
VN=(1-ε)k+j=NkN-1kwkCj-k+=Nk,N-13(w2)k-(1-3ε)wkC+2(1+ε)CN,1+3(1+ε)2χ02k+=Nk,N-1wkw+2(1+ε)k+m+n=Nk,m,nN-1(χ2)k(dm-wm)(w3)n+2ε(w2)n+ε2wn-2(1+ε)χ02k++n=Nk,,nN-1wkwwn+2εk+=Nk,N-1wkw-2(1+ε)3k+=Nk,N-1χkχ. 7.517

Proof

Proof of (7.513). We multiply (2.58) by YC and plug in the formal expansions (7.506)–(7.510). We formally obtain that the N-th Taylor coefficient in the expansion of dYC is given by

k+j=NkdkCj. 7.518

On the other hand, the N-th Taylor coefficient in the expansion of 2χ2d(w+ε)2(d-w)=2χ2d(w2+2εw+ε2)(d-w) is easily checked to be

2k++m+n=N(χ2)kd(dm-wm)(w2)n+2εwn+ε2δ0n. 7.519

We now extract the leading order terms in (7.518) and (7.519) - these are the factors containing either dN or wN. We see that

k+j=NkdkCj=NdNC0+k+j=NkN-1kdkCj=-Nχ02(1+ε)2dN+k+j=NkN-1kdkCj, 7.520

where we have used C0=-χ02(1+ε)2. Note that the term CN does not contribute a copy of dN on the right-hand side above since k=0 when j=N. Similarly, the expression (7.519) can be split into

-2χ02(1+ε)2dN+2k++m+n=NN-1(χ2)kd(dm-wm)(w2)n+2εwn+ε2δ0n. 7.521

Note that we have repeatedly used the assumption d0=d1=0. Equating (7.520) and (7.521) we obtain the recursive relation (7.513).

Proof of (7.514). We multiply (2.59) by (1-ε)YC and expand the two sides of the equation by analogy to the above. We obtain formally

(1-ε)wYC=(1-ε)N=0k+j=NkwkCjYN. 7.522

Upon extracting the leading order term in the N-th Taylor coefficient on the right-hand side of (7.522), by analogy to (7.520) we obtain

(1-ε)k+j=NkwkCj=-(1-ε)Nχ02(1+ε)2wN+(1-ε)k+j=NkN-1kwkCj. 7.523

We now turn our attention to the right-hand side. Observe that formally the N-th Taylor coefficient in the expansion of -(w+ε)(1-3w)C=3w2-(1-3ε)w-εC equals

k+=N3(w2)kC-(1-3ε)wkC-εδ0kC. 7.524

To single out the leading order contribution we use (7.511) and thus (7.524) can be rewritten in the form

2(1+ε)CN+3(w2)N-(1-3ε)wNC0+k+=Nk,N-13(w2)k-(1-3ε)wkC=-8ε(1+ε)χ02dN-3(1+ε)2(3+ε)χ02wN-4(1+ε)3χ0χN+k+=Nk,N-13(w2)k-(1-3ε)wkC+2(1+ε)CN,1-3(1+ε)2χ02k+=Nk,N-1wkw. 7.525

By analogy to (7.519) and (7.521) we also check that the N-the Taylor coefficient in the expansion of -2(1+ε)χ2w(w+ε)2(d-w) formally corresponds to

-2(1+ε)k+m+n=N(χ2)k(dm-wm)(w3)n+2ε(w2)n+ε2wn. 7.526

After isolating the leading order coefficients, the above expression can be rewritten in the form

-2(1+ε)3χ02dN+4(1+ε)2(2+ε)χ02wN+4(1+ε)3χ0χN-2(1+ε)k+m+n=Nk,m,nN-1(χ2)k(dm-wm)(w3)n+2ε(w2)n+ε2wn+2(1+ε)χ02k++n=Nk,,nN-1wkwwn+2εk+=Nk,N-1wkw+2(1+ε)3k+=Nk,N-1χkχ. 7.527

Claim (7.514) now follows from (7.523) (7.525), and (7.527).

Proof of (7.515). The claim follows by substituting the formal expansions for χ and w into (2.60), comparing the coefficients and using w0=1.

Theorem 7.4

(Local extension) There exists and 0<ε01 and a Y0<0 such that for any ε(0,ε0] there exists a unique analytic-in-Y solution to (2.58)–(2.60) on the interval (-|Y0|,|Y0|) such that

limY0d(Y)Y2=d2, 7.528
w(0)=1,w(0)=w1, 7.529
χ(0)=1, 7.530
C(Y)<0,Y(-|Y0|,|Y0|), 7.531

where d2,w1 are constants given by Lemma 6.7 and Proposition 6.8. Moreover,

(a)
There exists a δ>0 such that
d(Y0)>δ,χ(Y0)>δ,andw(Y0)<1δfor allε(0,ε0]. 7.532
(b)
There exists a constant c0>0 such that
χ2w(dY-2)1+η>1+c0,for allY[Y0,0]and allε(0,ε0]. 7.533
(c)
There exists a constant δ~>0 such that
1100>dw|Y=Y0>δ~,ε(0,ε0]. 7.534
(d)
There exists a constant cw>0 such that
w(Y0)>1+cw,for allε(0,ε0]. 7.535

Proof

The proof of existence of a local real-analytic solution is analogous to the proof of Theorem 4.18. We observe that the occurrence of the factors (N-2) and (N-1) in (7.513) and (7.514) respectively reflects the fact that d2 and w1 must be prescribed in order to consistently solve the ensuing recursive relations between the higher-order coefficients. We leave out the details, as they are similar to the ideas of the proof of Theorem 4.18.

Proof of part (a). Estimates (7.532) are a trivial consequence of the leading order Taylor expansions from Remark 7.2 the uniform-in-ε positivity of d2 and negativity of w1.

Proof of part (b). Since there exists a constant c~0 such that limY0χ2w(dY-2)1+η>1+c~0 for all ε(0,ε0], by Taylor expansion around Y=0 we can ascertain that there is a constant c0>0 and an interval [-|Y0|,|Y0|], Y0<0, such that (7.533) holds.

Proof of part (c). Observe that dw|Y=0=0. Since d is locally bounded from below and w locally increases to the left, the bound follows from continuity.

Proof of part (d). By the Taylor expansion (7.507) around Y=0 and the uniform-in-ε bound (7.504) we can guarantee that w<0 on [Y0,0) and that cw can be chosen independently of ε.

Remark 7.5

Note that Y0 is ε-independent, which plays an important role in our proof of the existence of outgoing null-geodesics from the scaling origin O. The constants d2 and w1 act as initial conditions for the system (2.58)–(2.60) to extend the solution to the left.

It is a priori possible that the solution constructed by Theorem 7.4 does not coincide with the RLP solution constructed for x(0,). The overlapping region (0,|Y0|) expressed in the x-coordinate is given by (X0,), where X0=r~|Y0|-(1+η)-11+η. As it is less clear how to apply the standard uniqueness theorem for the problem phrased in the Y-coordinate, we shall revert to the coordinate

X=x-11+η,x=X-1-η, 7.536

and reduce the question of uniqueness to the standard ODE theory.

Proposition 7.6

(Uniqueness of the RLP-extension) In the region Y>0, the solution constructed in Theorem 7.4 coincides with the RLP-type solution emanating from the sonic point x¯ constructed in Section 6.

Proof

Introduce the unknowns

D(x)=X2D¯(X),W(x)=1+XW¯(X). 7.537

A direct calculation gives

B(x)=X-2-2η-(1+ε)2-2(1+ε)XW¯+X2K1(X;D¯,W¯), 7.538
K1(X;D,W):=D¯-η-(1-ε)W¯2-4εD¯-4εXD¯W¯. 7.539

Using (7.536) and (2.35)–(2.36), we may further compute

ddXD=DdxdX=2X(1+ε)D¯1+ε+XW¯X2D¯-1-XW¯-(1+ε)2-2(1+ε)XW¯+X2K1(X;D¯,W¯)=:2XD¯+X2K2(X,D¯,W¯), 7.540

where it is easy to check that the function K2(X,D¯,W¯) is Lipschitz. An analogous calculation based on (2.36) then gives

ddXW=WdxdX=(1+η)2X+3W¯+2X(1+ε)1+XW¯1+ε+XW¯(X2D¯-1-XW¯)1+21+εXW¯-X2(1+ε)2K1=:(1+η)2X+3W¯-2X+11+ηW¯+XK3(X;D¯,W¯)=W¯+(1+η)XK3(X;D¯,W¯), 7.541

where it is easy to check that the function K3(X,D¯,W¯) is Lipschitz. Recalling (7.537), we conclude that

ddXD¯=K2(X;D¯,W¯),ddXW¯=(1+η)K3(X;D¯,W¯). 7.542

Therefore, the dynamical system (7.542) is regular at X=0 and must coincide with the RLP-solution emanating from the sonic point X=x-11+η. Since the mapping XY is smooth and invertible locally around X=0, the claimed uniqueness statement follows.

Maximal Extension

By analogy to Lemma 3.6, we factorise the denominator C into

C=(1-ε)J[Y;d,χ]-χwH[Y;d,χ]+χw, 7.543

where

J[Y;d,χ]=J:=-η(1+d)χ+η2(1+d)2χ2+εχ2+dY-2-ηY21-ε, 7.544
H[Y;d,χ]=H:=J+2η(1+d)χ. 7.545

Clearly, for any fixed Y and d , J[Y,d] is the solution of the equation C=0 viewed as a quadratic equation in χw. Just like in the proof of Lemma 3.6, it can be checked that

J=-4εχJ+η(dY-2)-η-12(1-ε)J+4εχ(1+d)d+-4εχ(1+d)J-2(ε2-ε)χχ+2ηY(dY-2)-η-1d+2(dY-2)-ηY2(1-ε)J+4εχ(1+d). 7.546

Our next goal is to prove a global extension result, which is shown later in Theorem 2.5. To that end define

Yms:=infY<0|a smooth solution to (2.58)--(2.60) andw(Y)>1,C(Y)<0,χ(Y)>0. 7.547

Lemma 7.7

Let (χ,w,d) be a local-in-Y solution to (2.58)–(2.60).

(a)
Then for all Y(Yms,0] the following identities hold:
(d1+ηw)=-d1+η(w+ε)(1-3w)(1-ε)Y, 7.548
(χ2d1+ηw)=χ2d1+η(1-ε)Yw2+w(1+3ε)-ε. 7.549
(b)
For all Y(Yms,0) we have the bounds
0<d<w. 7.550

Proof

Proof of part (a). Dividing (2.58) by d and (2.59) by w and the summing the (1+η)-multiple of the first equation with the second, we obtain (7.548). Using (7.548) and (2.60) we then obtain

(χ2d1+ηw)=χ2d1+η(1-ε)Y2w(1-w)-(w+ε)(1-3w)=χ2d1+η(1-ε)Yw2+w(1+3ε)-ε.

Proof of part (b). The strict positivity of d in a small open left neighbourhood of Y=0 follows from Theorem 7.4. The global positivity then follows by integrating (2.58), which can be rewritten as

logd=2χ2Y(w+ε)2(d-w)C.

The upper bound d<w clearly holds at Y=0 and in its small neighbourhood due to (7.502)–(7.503). Assume now by contradiction that there exists Yms<Y<0 so that Y is the infimum over all values of Y(Yms,0] such that d(Y)<w(Y). By continuity obviously d(Y)=w(Y). However, by (2.58)–(2.59) d(Y)-w(Y)=(w+ε)(1-3w)(1-ε)Y>0, since w(Y)>1 by (7.547). This is a contradiction, as this means that d-w decays locally going to the left of Y.

Lemma 7.8

Let (χ,w,d) be a local-in-Y solution to (2.58)–(2.60) and let

Γδ(Y):=(Y2)1+η-(1-δ)χ2wd1+η,δ>0. 7.551

Then there exist a 0<δ<1 and 0<ε01 sufficiently small such that for all ε(0,ε0] we have

Γδ(Y)<0,Y(Yms,0). 7.552

Proof

We observe that by (7.549),

Γδ(Y)=(2+2η)(Y2)ηY-(1-δ)(χ2wd1+η)=(2+2η)(Y2)ηY-(1-δ)χ2d1+η(1-ε)Yw2+w(1+3ε)-ε=1(1-ε)Y2(1+ε)Γδ(Y)+(1-δ)χ2d1+η-w2+w-εw+ε=1(1-ε)Y2(1+ε)Γδ(Y)-(1-δ)χ2d1+η(w-1)(w+ε). 7.553

Since w>1 by our assumptions, it follows that

(1-δ)χ2d1+η(w-1)(w+ε)>0.

Since Γδ(Y)=(Y2)1+η1-(1-δ)χ2w(dY-2)1+η it follows from Theorem 7.4, inequality (7.533) that we can choose δ=δ(c0)>0 such that Γδ<0 on [Y0,0). Using the standard integrating factor argument it then follows from (7.553) that

Γδ(Y)<0,Y(Yms,0). 7.554

Lemma 7.9

Let (χ,w,d) be a local-in-Y solution to (2.58)–(2.60) with the radius of analyticity (Y0,-Y0) given by Theorem 7.4. Then there exists a constant δ~>0 such that

dw>δ~for allYms<YY0,0<εε0. 7.555

Proof

By (2.58)–(2.59) we obtain

dw-dw=2(2+η)χ2dw(w+ε)2(d-w)YC+d(w+ε)(1-3w)(1-ε)Y=d(1-ε)YC{4χ2w(w+ε)2(d-w)+(w+ε)(1-3w)dY-2-ηY2-χ2(w+ε)2-ε(w-1)2+4εwd}=:d(1-ε)YCA¯, 7.556

where we used 2(2+η)=41-ε. We now rewrite A¯ to obtain

A¯=-4χ2w2(w+ε)2-(w+ε)(1-3w)χ2(w+ε)2+4χ2wd(w+ε)2-4εχ2wd(w+ε)(1-3w)+(w+ε)(1-3w)dY-2-ηY2+εχ2(w+ε)(1-3w)(w-1)2=χ2(w+ε)2-w2-w-ε+3εw+4χ2w2d(w+ε)(1+3ε)+(w+ε)(1-3w)dY-2-ηY2+εχ2(w+ε)(1-3w)(w-1)2=χ2(w+ε)4w2d(1+3ε)-(w+ε)w2+w+ε(1-3w)+(w+ε)(1-3w)dY-2-ηY2+εχ2(w+ε)(1-3w)(w-1)2=χ2(w+ε)w24(1+3ε)d-w-ε-(w+ε)w+ε(1-3w)+(w+ε)(1-3w)dY-2-ηY2+εχ2(w+ε)(1-3w)(w-1)2<χ2(w+ε)w24(1+3ε)d-w, 7.557

where we have used w>1, which in turn gives the bounds -(w+ε)w+ε(1-3w)<0, (w+ε)(1-3w)dY-2-ηY2<0, εχ2(w+ε)(1-3w)(w-1)2<0. Recall that dw|Y=0=0 by (7.528)–(7.529). From (7.556) and (7.557) it then follows that

dw<0ifdw<14(1+3ε). 7.558

Therefore by Theorem 7.4, inequality (7.534) and (7.558), the ratio dw increases to the left of Y=Y0 as long as dw<14(1+3ε). If the ratio ever exceeds 14(1+3ε) going to the left then it must stay larger than 14(1+3ε) as seen by a contradiction argument using (7.557). Therefore the claim follows.

The next lemma establishes the monotonicity of the function χw, and as a consequence the strict lower bound w>1+c on (Yms,0] for some c>0. The former is a preparatory step to prove that the flow remains supersonic in Lemma 7.11.

Lemma 7.10

Let (χ,w,d) be a local-in-Y solution to (2.58)–(2.60) analytic in (Y0,-Y0) given by Theorem 7.4. Then there exists an 0<ε01 sufficiently small so that

(χw)<0,Yms<YY0. 7.559

Moreover, there exists a constant c>0 such that

w(Y)>1+c,Yms<YY0,0<εε0. 7.560

In particular, inf(Yms,0]w(Y)>1.

Proof

By (2.59)–(2.60) we have

(χw)=(1-w)χw(1-ε)Y-χ(w+ε)(1-3w)(1-ε)Y-2(1+η)χ3w(w+ε)2(d-w)YC=2χw2-εχ+3εwχ(1-ε)Y-2(1+η)χ3w(w+ε)2(d-w)YC=χ(1-ε)YC2w2-ε+3εwC-2(1+ε)χ2w(w+ε)2(d-w)=:χ(1-ε)YCA. 7.561

From (2.61) we have

A=2w2-ε+3εwdY-2-ηY2-χ2(w+ε)2-ε(w-1)2+4εwd-2(1+ε)χ2wd(w+ε)2+2(1+ε)χ2w2(w+ε)2=χ2(w+ε)2-2w2-3εw+ε+2(1+ε)w2+εχ22w2-ε+3εw(w-1)2-4εχ2wd2w2-ε+3εw-2(1+ε)χ2wd(w+ε)2+2w2-ε+3εwdY-2-ηY2=εχ2(w+ε)2(w-1)(2w-1)+εχ22w2-ε+3εw(w-1)2-4εχ2wd2w2-ε+3εw-2(1+ε)χ2wd(w+ε)2+2w2-ε+3εwdY-2-ηY2 7.562

By Lemma 7.8 there exists a δ>0 such that dY-2-ηY2<(1-δ)χ2wd. Therefore

-4εχ2wd2w2-ε+3εw-2(1+ε)χ2wd(w+ε)2+2w2-ε+3εwdY-2-ηY2<χ2wd-4ε2w2-ε+3εw-2(1+ε)(w+ε)2+2w2-ε+3εw(1-δ)=χ2wd(-2δ-10ε)w2+-12ε2-4ε(1+ε)+3ε(1-δ)w+4ε2-2ε2(1+ε)-ε(1-δ)<-2δχ2w3d, 7.563

for εε0 sufficiently small. Therefore

A<εχ2(w+ε)2(w-1)(2w-1)+εχ22w2-ε+3εw(w-1)2-2δχ2w3d=εχ2(w-1)4w3+(-3+7ε)w2+(2ε2-6ε)w+ε-ε2-2δχ2w3d<4εχ2w4-2δχ2w3d=2χ2w3-δd+2εw, 7.564

where we have used the bounds w-1<w and (-3+7ε)w2+(2ε2-6ε)w+ε-ε2<0, where the latter follows from the assumption w>1 on (Yms,0]. Plugging (7.564) into (7.561) we obtain the estimate

(χw)<2χ3w3(1-ε)YC-δd+2εw=2χ3w4(1-ε)YC-δdw+2ε<2χ3w4(1-ε)YC-δδ~+2ε<-δδ~χ3w4(1-ε)YC<0, 7.565

for 0<εε0 sufficiently small. Here we have crucially used Lemma 7.9.

To prove (7.560) we note that by (7.565)

χ(Y)w(Y)>χ(Y0)w(Y0),Yms<Y<Y0 7.566

and therefore

w(Y)>χ(Y0)χ(Y)w(Y0)w(Y0)>1+cw, 7.567

where we have used part (d) of Theorem 7.4 in the last inequality. We have also used the bound χ(Y)χ(Y0) for YY0 which follows from χ>0 on (Yms,Y0), which in turn follows from the bound w>1 and (2.60).

Lemma 7.11

Let (χ,w,d) be a local-in-Y solution to (2.58)–(2.60) with the radius of analyticity (Y0,-Y0) given by Theorem 7.4. Then

lim supY(Yms)-C(Y)<0,

in other words - the flow remains supersonic.

Proof

Using (7.546) and (7.559), and the bounds

-4εχJ+η(dY-2)-η-12(1-ε)J+4εχ(1+d)d>0,-2(ε2-ε)χχ>0,

we get

J-(χw)>δδ~χ3w4(1-ε)YC+-4εχ(1+d)J+(2η+2)(dY-2)-ηY2(1-ε)J+4εχ(1+d). 7.568

Since (dY-2)-ηY2=(1-ε)J2+4εJ(1+d)χ-ε(1-ε)χ2, it follows that

(dY-2)-ηY2<(1-ε)J2+4εJ(1+d)χ<J2(1-ε)J+4ε(1+d)χ.

Therefore

(2η+2)(dY-2)-ηY2(1-ε)J+4εχ(1+d)CJ|Y|. 7.569

From (2.60) and the bound w>1 we have the rough bound |χ|Cwχ|Y|. Therefore

-4εχ(1+d)J2(1-ε)J+4εχ(1+d)Cεwχ(1+d)J|Y|(2(1-ε)J+4εχ(1+d))CwJ|Y|. 7.570

Using (7.569)–(7.570) in (7.568) we obtain

J-(χw)>δδ~χ3w4(1-ε)YC-CJ(1+w)|Y|=χw2δδ~χ2w2(1-ε)YC-CJχw(1w+1)|Y|. 7.571

Assume now that limY(Yms)-(J-χw)=0. In that case limY(Yms)-C=0 and it is clear from (7.571) that J-(χw) is strictly positive in some right neighbourhood of Yms. Here we use the uniform positivity of χw on (Yms,0], which follows from Lemma 7.10. A contradiction.

We have shown that the flow remains supersonic to the left of Y=0 and therefore the only obstruction to the global existence of the solution is the finite-time blow-up of the unknowns. We shall show that this is precisely the case. The intuition is that right-hand side of (2.59) will be, in a suitable sense, dominated by the first term on the right-hand side, which will lead to the blow-up of w through a Riccati-type argument.

Proof of Theorem 2.5. Let α>0 be a positive constant to be specified later. We rewrite (2.59) in the form

w=-(w+ε)(1-(3-2α)w)(1-ε)Y+2(w+ε)(1-ε)YCαwC-(1+ε)χ2w(w+ε)(d-w). 7.572

We focus on the term

E:=αwC-(1+ε)χ2w(w+ε)(d-w).

By (2.61) we have

E=αwd-η(Y2)1+η-αw(1-ε)χ2w2-4εαw2χ2+αwχ2(ε-ε2)-4εαχ2w2d-(1+ε)χ2wd(w+ε)+(1+ε)χ2w2(w+ε). 7.573

We let

α=1+δ¯1-ε 7.574

where δ¯>0 is a constant to be specified later. After regrouping terms in E above we obtain

E=χ2w2-(1+δ¯)w-2η(1+δ¯)+(1+ε)(w+ε)+ε(1+δ¯)w+1+δ¯1-εwd-η(Y2)1+η-2η(1+δ¯)χ2w2d-(1+ε)χ2wd(w+ε)<χ2w2-(δ¯-ε)w-2η(1+δ¯)+ε+ε2+ε(1+δ¯)w+χ2w2d1+δ¯1-ε(1-δ)-2η(1+δ¯)-(1+ε), 7.575

where we have used Lemma 7.8 in the last inequality. It is now clear that with the choice

δ¯12δ, 7.576

there exists an 0<ε01 sufficiently small, so that both expressions on the right-most side of (7.575) are strictly negative for all 0<εε0. Here we use w>1.

We conclude therefore from (7.572) and (7.574) that

w-(w+ε)(1-1-3ε-2δ¯1-εw)(1-ε)Y=-1-3ε-2δ¯1-ε(w+ε)(1-ε1-3ε-2δ¯-w)(1-ε)Y. 7.577

Now choose δ¯ sufficiently small (but independent of ε) so that w(Y)>1-ε1-3ε-2δ¯ for all Y(Yms,Y0), which is possible due to (7.560). Let now

K:=1-3ε-2δ¯1-ε.

We conclude from (7.577) that

w-C(w+ε)(Kw-1)|Y|-aw(Kw-1)|Y|,Yms<Y<Y0, 7.578

for some universal constant a>0. Inequality (7.578) is a Riccati-type differential inequality and leads to finite Y blow-up of w. We provide here the standard argument for the sake of completeness. Upon multiplying w by K and then redefining a>0, we may assume without loss of generality that K=1. Divide (7.578) (with K=1) by w(w-1) and express both sides as an exact derivative to conclude

log1-1w|Y|-a0,YY0. 7.579

Upon integration this yields the bound

w11-C|Y|a,Y<Y0, 7.580

and therefore w necessarily blows up as YY~+ for some -<Y~<Y0<0. By Lemma 7.9 this also implies that d blows up as YY~+.

We next want to show that limYY~+χ(Y)=0 and therefore Y~=Yms. We note that for any Y(Y~,0] we necessarily have χ(Y)>0, which follows from (2.60). Since χ>0 by (2.60) and the bound w>1 it follows that χ decreases to the left of Y=0 and the limit

χ:=limYY~+χ(Y)0 7.581

exists. Assume by the way of contradiction that χ>0. We look more closely at the leading order behaviour of the right-hand side of (2.59) on approach to the blow-up point Y~. Since

0<δ~<dw<1 7.582

by Lemmas 7.7 and 7.9, and the assumption χ>0 it is easily seen that

w=3w2(1-ε)Y+2(1+ε)w3(d-w)Y(1-ε)w2+4εwd+R, 7.583

where R has the property

limYY~+R(Y)2(1+ε)w3(d-w)Y(1-ε)w2+4εwd=0.

From (7.582) and (7.583) it is now clear that there exist constants 0<κ1<κ2 such that

κ1Y<ww2<κ2Y,Y(Y~,Y~+β), 7.584

for some constant β>0. Integrating the above differential inequalities over an interval [Y0,Y](Y~,Y~+β) and letting Y0Y~, we conclude

1κ2|Y~-Y|+O(|Y~-Y|2)w(Y)1κ1|Y~-Y|+O(|Y~-Y|2) 7.585

in a possibly smaller open right neighbourhood of Y~. By (2.60) we have (logχ)=1-wY, which together with (7.585) shows that there exist some positive constants 0<κ~1<κ~2 such that

κ~1<logχ|Y~-Y|<κ~2 7.586

in a small open right neighbourhood of Y~. Integrating (7.586) we conclude that

limYY~+χ(Y)=0, 7.587

which contradicts the assumption χ>0. It follows that in particular Y~=Yms.

Corollary 7.12

(Uniformity-in-ε) There exist a constant A>0 and 0<ε01 such that for all ε(0,ε0] we have the uniform bounds

1A<|Yms|<A, 7.588

where -<Yms<0 is the maximal existence interval to the left from Theorem 2.5.

Proof

Both constants C and a in (7.580) can be chosen to be ε-independent for ε0 sufficiently small and so we obtain a uniform upper bound on the maximal time Yms. Since by the construction Yms<Y0<0, where Y0 is the ε-independent constant from Theorem 7.4, we conclude the proof.

The Massive Singularity

Definition 7.13

(The massive singularity) The hypersurface MSε defined through

MSε=(τ~,R)|R=ε|Yms|τ~\{(0,0)}

is called the massive singularity.

In this section we compute the precise blow up rates of the RLP-solution at the massive singularity. To this end, it is convenient to introduce the quantity

C¯:=-χ-2w-2C=-χ-2w-2dY-2-ηY2+(1+εw)2-ε(1-1w)2+4εdw. 7.589

Lemma 7.14

The limits Q0:=limYYmsd(Y)w(Y) and limYYmsC¯(Y) exist and are finite. Moreover 0<Q01 and

limYYmsC¯(Y)=1-ε+4εQ0.

Proof

We let

Q:=dw. 7.590

We use (7.556)–(7.557) to derive a differential equation for Q:

Q=dw(w+ε)(1-ε)|Y|w2C¯(1+3ε)(4Q-1)-1wα(Y)+d1-η(w+ε)(1-3w)|Y|2+2η(1-ε)|Y|χ2w4C¯=:α1(Y)+α2(Y)d1-η, 7.591

where the function α(Y) is given by

α(Y)=1-9ε+7ε-3ε2w-ε-ε2w2. 7.592

By the invariance of the flow, we know that (1+3ε)(4Q-1)1wα(Y) and in particular the function α1 is nonnegative on (Yms,Y0]. Integrating (7.591) we conclude that

Q(Y)-Q(Yms)=YmsYα1(s)ds+YmsYd1-ηα2(s)ds. 7.593

We observe that α2 is a bounded function on (Yms,Y0] and d1-η is bounded from above by κ1-η(δY)-1+η and therefore YmsYd1-ηα2(s)ds1ε|Y-Yms|η. Since Q is bounded, it follows from (7.593) that α1L1((Yms,Y0]). For any Y1,Y2(Yms,Y0] we conclude that

Q(Y1)-Q(Y2)=Y1Y2α1(s)ds+Y1Y2d1-ηα2(s)dsY1Y2α1(s)ds+1ε|Y-Yms|η

In particular, Q is uniformly continuous on (Yms,Y0] and therefore Q0=limYYmsQ(Y) exists as claimed. Since 0<dw1 by Lemma 7.9 we have Q0(0,1]. To show that limYYmsC¯(Y) exists it follows from (7.585) and the continuity of Q that we only need to show the continuity of χ-2w-2dY-2-ηY2 at Yms. However, by Lemma 7.10 the quantity χ-2w-2 is uniformly bounded from above on approach to Yms and by (7.585) dY-2-ηY2|Y-Yms|η as YYms, so that χ-2w-2dY-2-ηY2 necessarily converges to 0 as YYms. Therefore YC¯(Y) is continuous and it converges to 1-ε+4εQ0 at Yms.

Lemma 7.15

There exist positive constants w^,d^>0 such that

w(Y)=w^|Y-Yms|+oYYms(|Y-Yms|-1), 7.594
d(Y)=d^|Y-Yms|+oYYms(|Y-Yms|-1). 7.595

Moreover

d^=|Yms|6,w^=2|Yms|3. 7.596

Proof

Dividing (2.58) by d2 and using (7.589) we obtain

-dd2=2|Y|(1+εw)2(wd-1)C¯. 7.597

The right-hand side has a strictly positive limit 1d^ by Lemma 7.14, where d^ equals

d^=|Yms|2Q0(1+ε4Q0-1)1-Q0, 7.598

and we may write the limit in the form 1d^+oYYms(|Y-Yms|). We may now integrate (7.597) to conclude (7.595). Analogously, we divide (2.59) by w2 and obtain

-ww2=1+εw1w-3(1-ε)Y-2(1+η)Y(1+εw)2(dw-1)C¯. 7.599

By Lemma 7.14, the right-hand side converges to a limit denoted by 1w^, given by

1w^=3(1-ε)|Yms|-2(1+η)|Yms|1-Q01+ε(4Q0-1). 7.600

We may now integrate (7.599) to conclude (7.594). Since by (7.594) and (7.595) Q0=d^w^, it follows by multiplying (7.598) and (7.600) that

Q0=|Yms|2Q0(1+ε4Q0-1)1-Q03(1-ε)|Yms|-2(1+η)|Yms|1-Q01+ε(4Q0-1). 7.601

Since Q0>0 by Lemma 7.14, we may divide by Q0 above and reduce the problem to the linear equation (4Q0-1)1+ε=0, hence d^w^=Q0=14. From (7.598) and (7.600) we now conclude (7.596).

Proposition 7.16

(Massive singularity) Let (w^,d^) be given by (7.596). There exists a χ^>0 such that on approach to the massive singularity MSε the solution (d,w,χ) of (2.58)–(2.60) obeys the following asymptotic behaviour:

w(Y)=w^|Y-Yms|1+OYYms(|Y-Yms|η), 7.602
d(Y)=d^|Y-Yms|1+OYYms(|Y-Yms|η), 7.603
χ(Y)=χ^|Y-Yms|23(1-ε)1+OYYms(|Y-Yms|η). 7.604

Moreover, the quantities μ~, λ~, defined by the extension of (2.68) and (3.107) to Y(Yms,0) respectively, satisfy

e2μ~YYms(Y-Yms)2ε1-ε,e2λ~YYms(Y-Yms)-23(1-ε). 7.605

The star density ρ and the Ricci scalar R blow up on approach to MSε.

Proof

We introduce q:=Q-14=dw-14, we can rewrite (7.591) in the form

q=AY-Ymsq+B, 7.606

where

A:=4(1+3ε)d|Y-Yms|w(w+ε)(1-ε)|Y|w2C¯,B:=-Q(w+ε)(1-ε)|Y|wC¯α+d1-η(w+ε)(1-3w)|Y|2+2η(1-ε)|Y|χ2w4C¯.

From (7.585) and Lemma 7.15 we immediately have

limYYmsA(Y)=4(1+3ε)d^(1-ε)|Yms|=23(1+2η)=A0, 7.607
B(Y)=OYYms(|Y-Yms|-1+η). 7.608

We now consider Yms<Y<Y1 for some fixed Y1 and integrate (7.606). We obtain

q(Y)=q(Y1)eY1YA(τ)τ-Ymsdτ+Y1YesYA(τ)τ-YmsdτB(s)ds. 7.609

For any 0<δ1 there exists a Y1>Yms such that |A-A0|<δ. It is then easy to see that

eY1YA(τ)τ-Ymsdτ=e-YY1A(τ)τ-Ymsdτe-YY1A0-δτ-Ymsdτ=|Y-Yms|A0-δ|Y1-Yms|A0-δ.

We use the bound in (7.609) and conclude

q(Y)q(Y1)|Y-Yms|A0-δ|Y1-Yms|A0-δ+Y1Y|Y-Yms|A0-δ|s-Yms|A0-δ|s-Yms|-1+ηds|Y-Yms|A0-δ+|Y-Yms|η|Y-Yms|η. 7.610

Plugging this back into (7.597) and (7.599) allows us to obtain the (suboptimal) rates (7.602)–(7.603). Upon dividing (2.60) by χ and integrating, using (7.602), for any Yms<Y<Y1 we obtain

χ(Y)=χ(Y1)e11-εYY1w^τ(τ-Yms)1+OτYms(|τ-Yms|η)dτ=χ(Y1)e23(1-ε)YY1|Yms|τ(τ-Yms)1+OYYms(|Y-Yms|η)dτ=χ(Y1)e23(1-ε)YY11τ-1τ-Yms1+OYYms(|Y-Yms|η)dτ=O(1)|Y-Yms|23(1-ε)1+OYYms(|Y-Yms|η), 7.611

which proves (7.604).

The asymptotics for e2μ~ in (7.605) follows directly from (7.603) and (7.494). The asymptotics for e2λ~ in (7.605) can be read off from (7.603)–(7.604), (2.69), and the identity (3.98), which in the (d,w,χ) variables reads

d11-εeλ~χ2=αY2+η, 7.612

for some constant α>0. It then follows from (7.612) that as δY0+, e2λ~(δY)-23(1-ε).

From  (2.57), we conclude that Σ(Y)=d(Y)1+ε1-εYYmsδY-1+ε1-ε, where we slightly abuse notation by continuing to denote Σ(Y) to signify the selfsimilar energy density (see (2.19)). Therefore by (2.19), for any τ~>0

ρ(τ~,R)=1τ(τ~,R)2Σ(Y)=R2ητ~21+ε1-εεηΣ(Y)=1τ~2Y2Σ(Y)YYmsδY-1+ε1-ε. 7.613

The fluid density therefore implodes along MSε and as a consequence of (2.75), the Ricci scalar blows up at MSε.

Remark 7.17

We can now use the asymptotic rates from from the previous lemma to replace the rough upper bound χ-2w-21, by the sharp upper bound rate |Y-Yms|2-43(1-ε). This can then be bootstrapped to obtain the near optimal next order correction in the rates (7.603)–(7.602), where η can be replaced by 23+O(ε). We do not pursue this here, as it will not be needed in the rest of the paper.

Causal Structure of the RLP Family of Solutions

As explained in Section 2.6, Theorems 2.32.4, and 2.5 imply the existence of a maximally selfsimilarly extended RLP spacetime, recall Definition 2.7. The standard ODE-theory implies that the solution is indeed real-analytic in Y on (Yms,). We next show that the RLP-spacetimes are not asymptotically flat and compute the associated mass-aspect function as r.

Lemma 8.1

((MRLP,ε,gRLP,ε) is not asymptotically flat) The RLP spacetimes (MRLP,ε,gRLP,ε) are not asymptotically flat, i.e. for any τ<0, limRm(τ,R)=. More precisely the mass aspect function 2m(τ,r)r satisfies

limr2m(τ,r)r=limR2m(τ,R)r(τ,R)=4εd2,

where d2>1 is the ε-independent constant introduced in Lemma 6.5.

Proof

Fix a τ<0. As R increases to we have y=R-ετ. Recalling (2.10) we have

limRm(τ,R)R=limR4π0R12πτ2Σ(y)ετ2r~2(y)r~(y)dR¯R=2εlimy0yΣ(z)r~(z)2r~(z)dzy=2εlimyΣ(y)r~(y)2r~(y)=2εd2, 8.614

where we have changed variables and used y=R-ετ in the second equality, the l’Hospital rule in the third, and (6.454)–(6.455), (2.33), Remark 7.2 in the last. Note that limRrR=limrrR=1 by (7.499)–(7.501).

As stated in [28], it is clear that along any line of the form (τ,ατ) the density ρ(τ,ατ)=12πτ2Σ(-αε) diverges as τ0-. A similar statement applies when we approach the scaling origin (0, 0) along the lines of the form (τ~,ατ~), τ~>0. In particular, by (2.75) the Ricci scalar blows up at (0, 0) and this is a geometric singularity. We now proceed to study the radial null-geodesics that “emanate" from this singularity. We shall henceforth use the abbreviation RNG for the radial null geodesics.

Lemma 8.2

(a)
In the (τ,R)-plane the outgoing/ingoing RNG-s respectively satisfy the equations
dRdτ=±eμ(y)-λ(y),y=R(τ)-ετ, 8.615
whenever the right-hand side is well-defined.
(b)
Similarly, in the (τ~,R)-plane the outgoing/ingoing RNG-s respectively satisfy the equations
dRdτ~=±1eλ~(Y)-μ~(Y)2ε1+εY,Y=-ετ~R(τ~), 8.616
whenever the right-hand side is well-defined.
(c)
If we let Y=-εσ, then the curve (στ~,τ~) is a simple outgoing/ingoingRNG (see Definition 2.11) if and only if
G±(Y)=0, 8.617
where
G±(Y):=εeλ~(Y)-μ~(Y)±1-ε1+εY. 8.618

Proof

Proof of part (a). Equation (8.615) is just the condition that the radial geodesic has null length in the local coordinates (2.70).

Proof of part (b). In the local coordinates (2.65) the RNG-s satisfy

-e2μ~(Y)τ~˙(s)2-4ε1+εYe2μ~R˙(s)τ~˙(s)+e2λ~(Y)-4ε(1+ε)2Y2e2μ~(Y)R˙(s)2=0, 8.619

where it is understood that Y=-ετ~(s)R(s) in the above expression. Reparametrising by τ~ the above ODE, we obtain

1=e2λ~(Y)-2μ~(Y)-4ε(1+ε)2Y2dRdτ~2-4ε1+εYdRdτ~, 8.620

or equivalently

dRdτ~2e2λ~(Y)-2μ~(Y)=1+2ε1+εYdRdτ~2. 8.621

Upon taking the square root, solutions to the equation (8.621) are given by the solutions of (8.616).

Proof of part (c). The proof follows by direct substitution in (8.616).

Remark 8.3

For the purpose of describing the causal structure of the selfsimilar spacetimes under consideration, it is convenient to introduce the function

Fε(Y):=Y2e2μ~ε(Y)-2λ~ε(Y),Y>Yms, 8.622

where Y=-ετ~R is the selfsimilar coordinate associated with the patch (2.65) and we keep the index ε to emphasise the dependence on ε. It is then straightforward to check from part (c) of Lemma 8.2 that an RNG (στ~,τ~), σ0, is simple if and only if

1εFε(Y)=1+ε1-ε2,Y=-εσ. 8.623

Formulas (8.618) and (8.622) give the obvious factorisation property:

G+(Y)G-(Y)=-ε1-ε1+ε2e2λ~ε(Y)-2μ~ε(Y)1εFε(Y)-1+ε1-ε2. 8.624

Proposition 8.4

Let Fε(·) be the function given by (8.622). There exists an 0<ε01 such that the following statements are true.

(a)
The function Fε satisfies the formula
Fε(Y)=(1+ε)2(1-ε)2(dεY-2)-ηY2+εχε2(wε-1)2-4wεdε(wε+ε)2χε2,Y>Yms. 8.625
(b)
There exists an Y<0 such that Y>Yms(=Yεms) for all ε(0,ε0] and
1εFε(Y)>2for allε(0,ε0]. 8.626
(c)
For any fixed ε(0,ε0] we have
limY0Fε(Y)=0, 8.627
limYYmsFε(Y)=0. 8.628

Proof

Proof of part (a). By (2.68) we have for any Y>0

Fε(Y)=Y2e2μ~ε(Y)-2λ~ε(Y)=(1+ε)2(1-ε)2Y2+2ηe2με(y)-2λε(y)=(1+ε)2(1-ε)2y-2e2με(y)-2λε(y)=(1+ε)2(1-ε)2dε-η+εr~ε2(wε-1)2-4dεwε(wε+ε)2r~ε2=(1+ε)2(1-ε)2(dεY-2)-ηY2+εχε2(wε-1)2-4wεdε(wε+ε)2χε2,y,Y>0, 8.629

where the index ε is added to emphasise the dependence on ε. We used the formula (2.56) in the third equality, and (7.498) to express y-2e2με(y)-2λε(y) in terms of wε, dε, and r~ε in the fourth. Since the right-most side of (8.629) extends analytically to Y(Yms,0] by Theorems 7.4 and 2.5 the claim in part (a) follows.

Proof of part (b). Note that for Y(Yms,0] by (8.629)

1εFε=(1+ε)2ε(1-ε)2(dεY-2)-ηY2(wε+ε)2χε2+(1+ε)2(1-ε)2(wε-1)2-4wεdε(wε+ε)2. 8.630

We now fix Y=Y0, a constant provided by the local extendibility statement of Theorem 7.4. In particular, by part (a) of Theorem 7.4 there exists a δ>0 such that for all ε(0,ε0] we have dε(Y0)>δ, χε(Y0)>δ and dε(Y0)<wε(Y0)<1δ. By letting ε0 in (8.630) we conclude that

1εFε(Y0)>2forεsufficiently small. 8.631

Proof of part (c). Claim (8.627) follows from the formula (8.629) and the asymptotic behaviour (7.499)–(7.503). To prove (8.628) we work directly with (8.622) and use (7.605). This gives

limY(Yms)+Fε(Y)=limY(Yms)+Y2e2με(Y)-2λε(Y)YYmsc(Yms)2lim(Y-Yms)0+(δY)2(1+3ε)3(1-ε)=0. 8.632

We are now ready to prove Theorem 2.12.

Proof of Theorem 2.12. Our first goal is to show that for 0<ε1 there exist at least two solutions to the equation (8.623). By parts (b) and (c) of Proposition 8.4 it is now clear that the function Y1εFε(Y) converges to 0 at Y=Yms and Y=0, but necessarily peaks above 1+ε1-ε2 at Y=Y0, where Y0 is given by Theorem 7.4. Therefore there exist Y1(Y0,0) and Y2(Yms,Y0) such that (8.623) holds with Y=Y1 and Y=Y2. Since Y1,Y2 are strictly negative and necessarily zeroes of G± defined by (8.618), they must in fact be zeroes of G+ and therefore represent outgoing simple RNG-s. Note that the function YFε(Y)-ε1+ε1-ε2 is real analytic on (Yms,0) and therefore the number of zeroes is finite. By slight abuse of notation we enumerate the zeroes as in (2.77).

We recall the relationship (2.56) between the comoving selfsimilar coordinates y and Y, as well as the relation between the comoving coordinate y and the Schwarzschild coordinates x in Subsection 3.2. Recalling the sonic point x¯, the slope of the sonic line in the Y-coordinates is given by

Ysp:=r~-1(x¯)-11+η. 8.633

The next lemma is important for the description of ingoing null-geodesics. It in particular implies that the curve N is the unique simple ingoing RNG.

Lemma 8.5

For any ε(0,ε0], consider the relativistic Larson-Penston spacetime given by Definition 2.7. Then there exists a YN(0,Ysp) such that the curve

N:={(τ~,R)MRLP,ε|-ετ~R=YN}, 8.634

represents a simple ingoing null-geodesics i.e. the boundary of the past light cone of the scaling origin O. Moreover, the curve N is the unique simple ingoing RNG and the following bounds hold:

G-(Y)>0,Y(Yms,YN), 8.635
G-(Y)<0,Y(YN,), 8.636

where G-(·) is defined in (8.618).

Proof

By Theorem 2.12 it suffices to show that there exists a YN(0,Ysp) which solves the equation (8.623). By part (c) of Proposition 8.4 we know that the function Y1εFε(Y) converges to 0 at Y=0. On the other hand, from (8.622) and (2.68) we have

Fε(Y)=(1+ε)2(1-ε)2y-2e2μ(y)-2λ(y)=(1+ε)2(1-ε)2y-2e2μ(y)-2λ(y)-y2+(1+ε)2(1-ε)2. 8.637

Therefore, at the sonic point Ysp we conclude Fε(Ysp)=(1+ε)2(1-ε)2, which implies that 1εFε(Ysp) is larger than 1+ε1-ε2 for all ε(0,1). By continuity, there exists an YN(0,Ysp) such that 1εFε(YN)=1+ε1-ε2. Now observe that by (8.637), the zeroes of the function 1εFε(Y)-1+ε1-ε2 are in 1-1 relationship with the zeroes of the function (0,)ye2λy2-1εe2μ. It is shown in Lemma A.4 that the map yr~(y)2e2λy2-1εe2μ is strictly monotone on (0,) and the uniqueness claim follows. Inequalities (8.635)–(8.636) are a simple consequence of the factorisation (8.624), the above monotonicity, and the positivity of G+ for all Y>0.

Definition 8.6

We refer to the region in the future of the backward null-curve N as the exterior region, and the region in the past of the null-curve N as the interior region, following here the terminology in [6], see Figure 5.

Remark 8.7

For any (R,τ~) in the exterior region, we note that εeμ~-λ~-2ε1+εY is clearly positive for Y(Yms,0]. When Y(0,YN) we may rewrite this expression as G-(Y)+1-3ε1+εY, which is then positive by Lemma 8.5 for sufficiently small ε. On the other hand, the expression εeμ~-λ~+2ε1+εY is clearly positive for Y0. If Y(Y1,0), we may rewrite it as G+(Y)-1-3ε1+εY, which is then necessarily positive. This shows that the right-hand side of the null-geodesic equation appearing in (8.616) is well-defined in the exterior region {Y1<Y<YN}.

Lemma 8.8

(General structure of radial null geodesics)

  1. For any point (R0,τ~0) in the interior region, the future oriented ingoing null curve τ~R(τ~) through (R0,τ~0) remains in the interior region and intersects the surface {R=0} at some τ~<0.

  2. For any point (R0,τ~0) in the exterior region, the future oriented ingoing null curve τ~R(τ~) exits the exterior region by intersecting B1 and also intersects B2 at positive values of the R-coordinate.

  3. For any (R0,τ~0) in the union of the exterior and the interior region, the outgoing null curve τ~R(τ~) through (R0,τ~0) exists globally-in-τ~ and limτ~R(τ~)=. Moreover, no such curve can converge to the scaling origin O to the past.

Proof

Proof of part (a). Assume now that (R0,τ~0) is in the interior region. The future oriented ingoing geodesic must stay in the interior region, as it cannot cross the backward light cone Y=YN by the ODE uniqueness theorem. In the interior region it is more convenient to switch to the original comoving coordinates (R,τ) as they also cover the centre of symmetry surface {(R,τ)|R=0,τ<0}={(r,τ),|r=0,τ<0}. Let (R0,τ0) correspond to (R0,τ~0). The ingoing geodesic equation reads τR=-eμ-λ and the interior region is characterised by the condition eμ(y)-λ(y)>εy. Let now T=-log(-τ) for τ<0. We then have dydT=dydτdτdT=Rτ-ετ+Rετ2(-τ)=-eμ(y)-λ(y)-εyε. In particular dydT<0, the right-hand side is smooth, and there are no fixed points of the above ODE on the interval (0,yN). We wish to show that the time T it takes to reach y=0 is finite. Integrating the above ODE, we see that

T(y)-T0=y0y-εeμ(θ)-λ(θ)-εθdθ,

where T0=-log|τ0|, y0=R0-ετ0. Note that e-λ(y)y0+1r~(y) by (3.95). We next use (2.73) and eμ(0)>0 to conclude that, as y0+, the denominator inside the integral above asymptotes to a constant multiple y-23(1+ε). Since the latter is integrable near y=0, it follows that limy0+T(y)<, as desired.

Proof of part (b). Let (R0,τ~0) belong to the exterior region. We consider the change of variables τ~Y and the particle label R as a function of Y. By (2.55) and the geodesic equations (8.616), along any null-geodesic we have

dYdτ~=ddτ~-ετ~R(τ~)=-εR(τ~)1-τ~1R(τ~)dRdτ~=-εR(τ~)1±1εYeλ~(Y)-μ~(Y)2εY1+ε=-1R(τ~)G±(Y)eλ~(Y)-μ~(Y)2εY1+ε, 8.638

where

G±(Y):=εeλ~(Y)-μ~(Y)±1-ε1+εY. 8.639

We note that by Remark 8.7 all the denominators appearing above are nonzero. Therefore the geodesic equations (8.616) transform into

±1eλ~(Y)-μ~(Y)2εY1+ε=dRdYdYdτ~=-1R(Y)G±(Y)eλ~(Y)-μ~(Y)2εY1+εdRdY. 8.640

For as long as R>0 we can rewrite the above ODE in the form

ddY(logR)=1G±(Y). 8.641

We remark that by Lemma 8.5 the denominator is strictly positive in the exterior region and in the case of outgoing geodesics, function G+ is in fact strictly positive for all Y(Y1,) (positivity of G+ characterises the region “below" B1).

For the ingoing geodesics, the exterior region is invariant by the flow. We integrate (8.641) to conclude that (with Y:=-ετ~0R0)

logR(Y)-logR0=-YY1G-(Z)dZ. 8.642

The map ZG-(Z) is smooth and by Lemma 8.5 it is strictly positive on the interval (Yms,YN). Therefore, R is positive as the ingoing RNG traverses B1 and B2, as the right-hand side of (8.642) is finite for any Y>Yms.

Proof of part (c). Let (R0,τ~0) belong to the exterior region. The outgoing null-geodesic solves the ODE

ddY(logR)=-1G+(Y).

Note that G+ is smooth on (Y1,YN). Since Y1<0 is the largest negative root of ZG+(Z), the right-hand side above is negative. It follows that YR(Y) is decreasing on (Y1,Y), i.e. R(Y) increases as Y approaches Y1 from the right. In particular, the solution exists for all Y(Y1,Y] by the strict negativity of the right-hand side above. We consequently have the formula

R(Y)=R0expYY1G+(Z)dZ,Y>Y1. 8.643

Since the function G+ is positive for Z>Y1, if G(Z)=C(Z-Y1)m(1+O(|Z-Y1|)) is the first term of Taylor expansion of G at Y1, with mN (recall that G is real analytic), then necessarily C>0. Plugging this expansion into (8.643) and integrating we conclude limYY1+R(Y)=, which shows that the outgoing null-geodesic asymptotes to B1. Since Y=-ετ~(Y)R(Y) and limY(Y1)+R(Y)=, it follows that limY(Y1)+τ~(Y)=, and therefore the outgoing geodesic exists globally on [τ~0,) and limτ~R(τ~)=. If (R0,τ~0) belongs to the interior region, a similar analysis in the original comoving coordinates (R,τ) yields the same conclusion.

Since the function G+(Z) is smooth and bounded in a neighbourhood of Z=0, it follows from (8.643) that any outgoing geodesic starting at (R0,τ~0) with τ~0>0 intersects {τ~=0} axis (i.e. Y=0) at a positive value of R to the past. Due to the monotonicity of the flow (8.643), any outgoing geodesic emanating from (R0,τ~0) with τ~00 remains below {τ~=0} axis to the past.

Asymptotic Flattening of the Selfsimilar Profile

The key result of this section is the local well-posedness for the characteristic initial value problem for the Einstein-Euler system, see Theorem 9.4. The idea is to suitably truncate the selfsimilar spacetime as described in Section 2.7. We work with the double-null formulation, see Section 2.7.1, and our starting point is the reformulation of the fluid evolution equations (2.87)–(2.88).

Reformulation of the Fluid Evolution and the Effective Transport Velocity

We introduce the constant

k±:=1±2η+η21+η, 9.644

where we recall η=η(ε) is given by (2.38) and from (9.644) it is clear that k±=1±O(η).

Lemma 9.1

(Reformulation of the Euler equations) Assume that (ρ,uν,r,Ω) is a C1 solution of (2.87)–(2.88). Let

U:=(1+η)Ω2(uq)2, 9.645
f+:=(1-ε)k+1+εr(2+2η)k+-2ρk+-1(uq)2, 9.646
f-:=1+ε(1-ε)k-r2-(2+2η)k-ρ1-k-(uq)2. 9.647

Then the new unknowns f± satisfy

pf++k+Uqf++2k+(2qΩΩ-2ηqrr)Uf+=0, 9.648
pf-+k-Uqf--2k-(2qΩΩ-2ηqrr)Uf-=0. 9.649

Proof

Let U:=Ω4r2Tpp and V:=Ω2r2+2ηTpq. Using (2.91) we rewrite Ω4r2Tqq=(1+η)2Ω4r4ηV2U. Therefore, (2.87)–(2.88) can be rewritten in the form

pU+Ω2r2ηqV=0,pV+(1+η)2r2ηΩ2q(Ω4r4ηV2U)=0. 9.650

For any kR we now compute pVkU and thereby use (9.650):

p(VkU)=kVk-1UpV-VkU2pU=-k(1+η)2Ω2r2ηVUq(VkU)-2k(1+η)2Vk+1U2qΩ2r2η-k(2-k)(1+η)2Ω2r2ηVkU2qV+VkU2Ω2r2ηqV. 9.651

We see that the last line (9.651) vanishes if k is a solution of the quadratic equation

k2-2k+1(1+η)2=0. 9.652

The two distinct roots k± of (9.652) are given in (9.644), and the equation for Vk±U reads

pVk±U+k±(1+η)2Ω2r2ηVUqVk±U+2k±(1+η)2Vk±+1U2qΩ2r2η=0.

When k=k+>1, we keep Vk+U as the unknown. When however k=k-<1, we work with UVk instead, to avoid singularities for small values of ρ. From above we obtain the equation

pUVk-+k-(1+η)2Ω2r2ηVUqUVk--2k-(1+η)2V1-k-qΩ2r2η=0.

Going back to original variables, note that VU=r2η(1+η)Ω4(up)2 and VkU=(1-ε)k1+εr(2+2η)n-2ρk-1Ω4(up)2 so that

(1+η)2Ω2r2ηVU=(1+η)Ω2(up)2=(1+η)Ω2(uq)2=U,Vk+U=(1-ε)k+1+εr(2+2η)k+-2ρk+-1Ω4(up)2=f+,UVk-=1+ε(1-ε)k-Ω4(up)2r2-(2+2η)k-ρ1-k-=f-,

where we recall (9.645)–(9.647).

Fig. 9.

Fig. 9

The grey shaded area in the infinite rectangular region D is a schematic depiction of the region bounded by the backward fluid characteristics emanating from a point SD. The opening angle is of order ε1, which of course is precisely the speed of sound

Remark 9.2

It follows from (9.646)–(9.647) that

ρ=11-ε(f+f-)1k+-k-r2+2η. 9.653

Moreover, from (2.82) and (9.646)–(9.647) we have f+f-=1(1+η)2r4η(uq)4, which leads to the relation

U=(1+η)Ω2(uq)2=(1+η)2Ω2r2ηf+f-12. 9.654

Statement of the Local Well-Posedness Theorem

In order to flatten the the gRLP,ε metric at asymptotic infinity, we shall treat the system (2.83)–(2.84) and (9.648)–(9.649) as the evolutionary part and equations (2.86)–(2.85) as the constraints.

Fixing the Choice of Double-Null Coordinates

We now fix a choice of double-null coordinates which will then be used to dampen the tails of the solution and produce an asymptotically flat spacetimes containing a naked singularity. Let (τ~0,R0)D~RLP,ε be a given point in the exterior region (see Definition 8.6 and (2.66) for the definition of D~RLP,ε). Through (τ~0,R0) we consider the ingoing null-curve which intersects the outgoing null-curve B1 (given by R=ε|Y1|τ~) at some (R1,τ~1) where by Lemma 8.8, R1>0. We then fix the null-coordinate p by demanding that

p=-2(r-r1)along the ingoing null-geodesic through(R0,τ~0),r1:=r~(R1,τ~1); 9.655

and demanding that the level sets of p correspond to outgoing null-geodesics. The choice (9.655) normalises B1 to correspond to the hypersruface {p=0} in the RLP-spacetime. Note that in the RLP spacetime, the ingoing curve through (τ~0,R0) terminates at the massive singularity MSε to the future.

Let now the outgoing null geodesic through (τ~0,R0) intersect the surface N (boundary of the past of the scaling origin O) at (R,τ~). We then fix the null coordinate q by demanding that

q=2(r-r)along the outgoing null-geodesic through(τ~0,R0),r:=r~(R,τ~); 9.656

and demanding that the level sets of q correspond to ingoing null-geodesics. The normalisation (9.656) makes the surface N correspond to the hypersurface {q=0}. A more detailed description of the RLP-spacetime in this double-null gauge is given in Lemma A.1.

Let (p0,q0) be the point (τ~0,R0) in the above double-null gauge. We shall consider the seminfinite rectangular domain

D:={(p,q):p0<p<0,q>q0}, 9.657

where |p0|>0 is sufficiently small, with data prescribed on the set C_C, where

C_:=(p,q)|q=q0,p[p0,0], 9.658
C:=(p,q)|p=p0,q[q0,) 9.659

correspond to the ingoing and the outgoing null-curves emanating from (p0,q0) respectively. See Figure 6.

Norms and Local Well-Posedness

Fix N±>0 so that

N+=N--4η,N->k+-k-2+2η 9.660

and let 0<θ1 be a small fixed constant. For any f±W2,(D¯) and (Ω,rq)W3,(D¯), we define the total norm by

[f±,Ω,r]:=[f±]+[Ω,r] 9.661

where

[f±]:=log(qN+f+)+j=12qjqjlogf+ 9.662
+log(qN-f-)+j=12qjqjlogf-,[Ω,r]=logΩ+j=13qj+θqjlogΩ 9.663
+j=03qj-2qj(r2)+q2r2+qq(r2), 9.664

and the data norm

[f±,Ω,r]|C_C:=[f±]|C_C+[Ω,r]|C_C:=log(qN+f+)|C_C+j=12qjqjlogf+|C_C+log(qN-f-)|C_C+j=12qjqjlogf-|C_C+logΩ|C_C+j=13qj+θqjlogΩ|C_C+j=03qj-2qj(r2)|C_C+q2r2|C_C+qq(r2)|C_C+q-1p(r2)|C_+r21+4prqrΩ2|C_

We choose the data (f^±,r^,Ω^) on C_ to coincide with the corresponding data obtained by restricting the RLP-solution to C_. Let A0>q0 be a real number to be specified later.

  • (i)

    The data on C is chosen so that [f^±,Ω^,r^] coincide with the exact selfsimilar gRLP,ε solution on the segment {(p,q)|p=p0,q[q0,A0]};

  • (ii)

    f^±W2,(CC_), and Ω^,r^qW3,(C_C) with [f±,Ω^,r^]|C_C< and f^±,Ω^,r^>0;

  • (iii)

    the constraint equations (2.85)–(2.86) hold on C and C_ respectively.

Due to the choice of the double-null gauge (9.655)–(9.656), the metric coefficient r^ is determined along C. In order to impose the constraint (2.85) we can solve it for Ω^. By (9.656) we have qr=12 and by (2.89) and Remark 9.2 we may rewrite (2.85) as

12q(Ω-2)=-πr(f+f-)1k+-k-f-f+12Ω-2. 9.665

Moving the copy of Ω-2 to the left-hand side and then integrating, we obtain the formula

Ω(p0,q)=Ω(p0,q0)expq0q2πs+2r(f+f-)1k+-k-f-f+12ds. 9.666
Remark 9.3

(The Hawking mass) We recall the Hawking mass introduced in (2.10). It can be alternatively expressed via the formula

m=r21+4prqrΩ2. 9.667

Using (2.83)–(2.86) one can show that for classical solutions of the Einstein-Euler system we have the identities

pm=2πr2Ω2Tpqpr-Tqqqr, 9.668
qm=2πr2Ω2Tpqqr-Tpppr, 9.669

see for example Section 1.2 of [9]. Using (2.89) we may rewrite the right-hand side of (9.669) as 2πr2(1-ε)ρqr-(1+η)Ω2(up)2pr. Integration along a constant p-slice then gives

m(p,q)=m(p,q0)+2π(1-ε)q0qr2ρqr-(1+η)Ω2(up)2prds. 9.670

We now state the main local existence and uniqueness theorem for the characteristic problem described above.

Theorem 9.4

There exist sufficiently small δ0>0 such that for any δ(0,δ0) and p0=-δ, with initial boundary data satisfying (i)-(iii) above, there exists a unique asymptotically flat solution [f±,Ω,r] to the system (2.83)–(2.84) and (9.648)–(9.649), with f±W2,(D¯) , [Ω,rq]W3,(D¯), and such that [f±,Ω,r]<. Moreover, this solution is a solution of the original system (2.83)–(2.88).

We shall prove Theorem 9.4 by the method of characteristics. Our strategy is to first solve the fluid evolution equations for f± given the effective fluid velocity U and the metric components Ω,r. Then we will feed that back into the wave equations (2.83)–(2.84) for the metric components to obtain the bounds for [r,Ω]. To make this strategy work, in Section 9.3 we carefully look at the characteristics associated with the fluid evolution. After collecting some preparatory a priori bounds in Section 9.4, in Section 9.5 we use an iteration scheme to conclude the proof of Theorem 9.4.

Characteristics for the Fluid Evolution

Let q±(s)=q±(s;p,q) be the backward characteristics associated with the speeds k±U such that

dq±ds(s;p,q)=k±U(s,q±(s;p,q)),s<p, 9.671
q±(p;p,q)=q. 9.672

Since our solutions as well as U reside in the domain D and the boundary, we track the backward characteristics (s,q±(s)) until they leave the domain. In the next lemma, we show the existence and regularity properties of q±(s) and exit time p=p(p,q) and position q=q(p,q).

Lemma 9.5

Let N. Suppose UC(D¯) or W,(D¯) and 1C0<U<C0 on D¯ for some C0>1. Then for any given (p,q)D, there exist a unique exit time p=p(p,q) and position q=q(p,q) such that

  1. A unique solution q±C((p,p];C(D)) or C((p,p];W,(D)) of (9.671) and (9.672) exists so that (s,q±(s;p,q))D.

  2. At s=p, (p,q)C_C where q=q±(p;p,q) and it satisfies
    q-q=ppk±U(s,q±(s;p,q))ds. 9.673
    If p>p0 then q=q0.
  3. (p,q)p(p,q)C(D).

  4. If pp0, pC(D) or W,(D¯).

In particular, if q>q0+k±C0(p-p0), then (p,q)=(p0,q)C.

Proof

We prove the claims for UC as the case of UW, follows in the same way. The local existence and uniqueness of q±C follows from UC via the Picard iteration

q±(s)=q-spk±U(s~,q±(s~;p,q))ds~. 9.674

Thanks to the positive uniform bound of U and Grönwall, the solution can be continued as long as the characteristics belong to the domain (s,q±(s;p,q))D. Since U>1C0, q±(s)<q-k±C0(p-s), (s,q±(s)) will exit the domain D either through the outgoing boundary C or through the ingoing boundary C_. We denote such an exit time by p and the associated value q=q±(p;p,q) where (p,q)C_C. Note that for each given (p,q)D, (p,q) is uniquely determined since the backward characteristics q± are unique. By integrating (9.671) from s=p to s=p, we see that (p,q) satisfies (9.673). We observe that if (p,q)C_, q=q0 and p0p<p, and (9.673) reads as

q-q0=ppk±U(s,q±(s;p,q))ds. 9.675

If (p,q)C, p=p0 and q is given by (9.673).

Clearly p is continuous in p and q. Since higher regularity of p fails in general at the corner of the domain where (p,q)=(p0,q0), we show the regularity when pp0. First let (p~,q~)D be given. Suppose p~=p(p~,q~)>p0. Consider a small neighborhood B of (p~,q~) in D such that I1=p(B), infI1>p0. Recalling (9.675), we define an auxiliary function H:I1×BR by

H(p¯,p,q)=q-q0-p¯pk±U(s,q±(s;p,q))ds.

Then HC since UC, while we have H(p~,p~,q~)=0 and p¯H(p~,p~,q~)=k±U(p~,q~)>0. Therefore, by the implicit function theorem, p=p(p,q)C in a small neighborhood of (p~,q~).

Lastly, since U<C0, if q>q0+k±C0(p-p0), the backward characteristics will intersect the outgoing surface C, in which case p=p0.

A Priori Bounds

In this subsection, we provide estimates for various quantities appearing the iteration scheme in terms of our norms. We will frequently use the following inequality: for any positive function g>0

max{g,g-1}elogg 9.676

which directly follows from g=elogg and g-1=e-logg.

Lemma 9.6

Suppose [Ω,r]<. Then the following holds:

j=02qj+1qjqΩΩ-ηqrrM1([Ω,r]), 9.677
j=12qj+3qjΩ2r3M2([Ω,r]), 9.678

where M1 and M2 are continuous functions of their arguments.

Proof

We start with (9.677). For j=0, it is easy to see that

qqΩΩ-ηqrr1q0θq1+θqlogΩ+η2q2r2q(r2)q 9.679

where the right-hand side is a continuous function of [Ω,r] by (9.676). For j=1, since

qqΩΩ-ηqrr=q2logΩ-η2q2(r2)r2+η2(q(r2))2r4

we have

q2qqΩΩ-ηqrr1q0θq2+θq2logΩ+η2q2r2q2(r2)+η2q2r22q(r2)q2 9.680

which shows (9.677). Lastly, from

q2qΩΩ-ηqrr=q3logΩ-η2q3(r2)r2+3η2q(r2)q2(r2)r4-η(q(r2))3r6

we obtain

q3q2qΩΩ-ηqrr1q0θq3+θq3logΩ+η2q2r2qq3(r2)+3η2q2r22q(r2)q2q2(r2)+η2q2r23q(r2)q3. 9.681

This completes the proof of (9.677). The estimation of (9.678) follows similarly, we omit the details.

Lemma 9.7

(U bounds) Suppose [f±,Ω,r]<. Then the following holds:

U+U-1+j=12qjqjUM3([f±,Ω,r]) 9.682

where M3 is a continuous function of its argument.

Proof

From the first condition of our choice N± in (9.660), we may rewrite U as

U=(1+η)2Ω2qr2ηqN+f+qN-f-12.

Then by (9.676), we see that

U±1(1+η)±2q2r2±ηe2logΩ+12log(qN+f+)+12log(qN-f-) 9.683

which shows (9.682) for U+U-1. Next computing qU as

qU=2qΩΩ-ηqrr+12qlogf+-12qlogf-U 9.684

we obtain

qqU2qqΩΩ-ηqrr+12qqlogf++12qqlogf-U. 9.685

Together with (9.679), it implies (9.682) for j=1. Moreover, since

q2U=2qqΩΩ-ηqrr+12q2logf+-12q2logf-U+(qU)2U 9.686

we have

q2q2U(2q2qqΩΩ-ηqrr+12q2q2logf++12q2q2logf-)U+U-1qqU2. 9.687

Using (9.680) and (9.685), we deduce (9.682), where we recall (9.661).

Remark 9.8

The relation N+=N--4η in (9.660) is importantly used to ensure the boundedness (both upper and lower) of the transport speed U.

We introduce the constant β>0:

β:=2η+N++N-k+-k--1=2η+2N--(k+-k-+4η)k+-k->0, 9.688

where we have used (9.660) in the second equality.

Lemma 9.9

Suppose [f±,Ω,r]<. Then the following bounds hold:

j=02qj+1+βqjΩ2r2η(f+f-)1k+-k-M4([f±,Ω,r]), 9.689
j=02qj+3+βqjΩ2r2+2η(f+f-)1k+-k-M5([f±,Ω,r]), 9.690

where M4 and M5 are continuous functions of their arguments.

Proof

We will first prove (9.689). We start with j=0. Using (9.688), we rewrite H:=Ω2r2η(f+f-)1k+-k- as

q1+βH=q1+βΩ2r2η(f+f-)1k+-k-=Ω2qr2ηqN+f+qN-f-1k+-k-

from which we have

q1+βHΩ2q2r2η(qN+f+qN-f-)1k+-k-. 9.691

By (9.676), the claim immediately follows. We next compute

qH=2qΩΩ-ηqrr+qlogf++qlogf-k+-k-H 9.692

and obtain

q2+βqH2qqΩΩ-ηqrr+qqlogf++qqlogf-k+-k-q1+βH. 9.693

With (9.679) and (9.691), it gives (9.689) for j=1. We next have

q2H=2qqΩΩ-ηqrr+q2logf++q2logf-k+-k-H+2qΩΩ-ηqrr+qlogf++qlogf-k+-k-qH, 9.694

and therefore

q3+βq2H(2q2qqΩΩ-ηqrr+q2q2logf++q2q2logf-k+-k-)q1+βH+2qqΩΩ-ηqrr+qqlogf++qqlogf-k+-k-q2+βqH. 9.695

Hence the claim follows from (9.691), (9.693) and (9.677).

The proof of (9.690) follows easily from (9.689) by applying the product rule for Ω2r2+2η(f+f-)1k+-k-=r-2H or by estimating them directly in the same way as done for (9.689). We omit the details.

In the iteration scheme, we will make use of the Hawking mass m given in (9.667). From (9.669) and (2.82) we see that the Hawking mass satisfies

qm=2πr2(1-ε)ρqr-(1+ε)Ω-2ρ(uq)-2pr=2π(f+f-)1k+-k-1r2ηqr-Ω-2f-f+12pr. 9.696

We observe that p-derivatives are not featured in our function space hierarchy, however pr appears in the above expression. To go around this difficulty, we observe that (2.83) can be formally rewritten in the form

pq(r2)=-Ω22+2πr2Ω4Tpq=-Ω22+2πΩ2r2η(f+f-)1k+-k-, 9.697

where we have used (2.89) and (9.653) in the second equality. Therefore, rather than directly estimating m from (9.667), we slightly abuse notation, and redefine m to be

m(p,q):=m(p,q0)+q0q2π(f+f-)1k+-k-1r2ηqr-Ω-22rf-f+12sdq~ 9.698

where

m(p,q0):=r^21+4pr^qr^Ω^2|(p,q0), 9.699
s(p,q):=p(r^2)|(p,q0)+q0q-Ω22+2πΩ2r2η(f+f-)1k+-k-dq~. 9.700

We observe that pr^ is well-defined at (p,q0) since the data there is given by the exact selfsimilar RLP-solution. We note that s corresponds exactly to p(r2) for a C2-solution of the problem. In the following, we show that m in (9.698) can be estimated by using the norm [f±,Ω,r].

Lemma 9.10

Suppose [f±,Ω,r]<. Then the following holds:

m[Ω,r]|C_C+M6([f±,Ω,r]), 9.701
j=12qj+βqjmM7([f±,Ω,r]), 9.702

where M6 and M7 are continuous functions of its argument.

Proof

We first observe that from (9.700) using q>q0 and q-q0q<1

q-1sp(r^2)q|(p,q0)+Ω22+2πq01+βq1+βΩ2r2η(f+f-)1k+-k- 9.703

and

qsΩ22+2πq01+βq1+βΩ2r2η(f+f-)1k+-k- 9.704

From the definition of the boundary data norm, we have |m(p,q0)|[Ω,r]|C_C. For the integral term in (9.698), using (9.688) and (9.660), we rewrite the integrand as

q-1-β(qN+f+qN-f-)1k+-k-q2ηr2ηqr-Ω-2q2rqN-f-qN+f+12sq 9.705

so that

q0q2π(f+f-)1k+-k-1r2ηqr-Ω-22rf-f+12sdq~1βq0β(qN+f+qN-f-)1k+-k-·12qr2η+1q-1q(r2)+12Ω-12qrqN-f-12(qN+f+)-112q-1s

where we have used q0qq~-1-βdq~1βq0β. Hence, using (9.676), we deduce (9.701). Moreover since

qm=2π(f+f-)1k+-k-1r2ηqr-Ω-22rf-f+12s 9.706

from (9.705), we immediately obtain

q1+βqm2π(qN+f+qN-f-)1k+-k-·12qr2η+1q-1q(r2)+12Ω-12qrqN-f-12(qN+f+)-112q-1s 9.707

which shows (9.702) for j=1. Lastly, we compute q2m as

q2m=π(f+f-)1k+-k-1r2η+1q2(r2)-Ω-2rf-f+12qs+π(f+f-)1k+-k-r2ηq(r2)rqlogf++qlogf-k+-k--2η+12q(r2)r2-πΩ-2(f+f-)1k+-k-f-f+12sr-2qlogΩ+c1qlogf++c2qlogf+-12q(r2)r2 9.708

where c1=1k+-k--12 and c2=1k+-k-+12. Following the same strategy, it is now clear that q2+βq2m is bounded by a continuous function of [f±,Ω,r]. This finishes the proof.

Proof of the Local Well-Posedness

Iteration Scheme

We now set up the iteration scheme. Let [fn±,Ωn,rn] be given so that [fn±,Ωn,rn]<. And let Un and mn be given in (9.654) and (9.698) where [f,Ω,r]=[fn±,Ωn,rn]:

Un=(1+η)2(Ωn)2(rn)2ηfn+fn-12

and

mn=m(p,q0)+q0q2π(fn+fn-)1k+-k-1(rn)2ηqrn-12(Ωn)-2(rn)fn-fn+12sndq~.

Here sn is given by the right-hand side of (9.700) where [f±,Ω,r]=[fn±,Ωn,rn]. We then define [fn+1±,Ωn+1,rn+1] to be the solution of the following system

pfn+1±+k±Unqfn+1±±2k±(2qΩnΩn-2ηqrnrn)Unfn+1±=0, 9.709
pq[(rn+1)2]=-(Ωn)22+2π(Ωn)2(rn)2η(fn+fn-)1k+-k-, 9.710
pqlogΩn+1=(Ωn)22(rn)2mnrn-(1+η)π(Ωn)2(rn)2+2η(fn+fn-)1k+-k-, 9.711

with given characteristic data [fn+1±,Ωn+1,rn+1]=[f^±,Ω^,r^]|C_C satisfying the conditions (i)–(iii) as in Theorem 9.4. We note that for exact solutions, (9.710)-(9.711) are equivalent to (2.83), (2.84).

Our next goal is to prove the solvability of the iterative system (9.709)-(9.711) and derive the uniform bounds of [fn+1±,Ωn+1,rn+1] for sufficiently small δ>0. Let A:=max{4,1+1θq0θ}.

Proposition 9.11

Let the characteristic data [f^±,Ω^,r^] satisfying the assumptions of Theorem 9.4 be given. Suppose [fn±,Ωn,rn]2A[f^±,Ω^,r^]|C_C. Then there exist sufficiently small δ0>0 depending only on [f^±,Ω^,r^]|C_C such that for all δ(0,δ0) there exists a unique solution fn+1±W2,(D¯), [Ωn+1,rn+1q]W3,(D¯) to (9.709)-(9.711) with [fn+1±,Ωn+1,rn+1]=[f^±,Ω^,r^] on C_C satisfying

[fn+1±,Ωn+1,rn+1]2A[f^±,Ω^,r^]|C_C. 9.712

Proposition 9.11 immediately follows from the following two lemmas.

Lemma 9.12

(Solving the Euler part) Assume the same as in Proposition 9.11. Then there exist sufficiently small δ0>0 depending only on [f^±,Ω^,r^]|C_C such that for all δ(0,δ0) there exists a unique solution fn+1±W2,(D¯) to (9.709) with fn+1±=f^± on C_C satisfying

[fn+1±]2[f^±]|C_C+A[f^±,Ω^,r^]|C_C. 9.713
Lemma 9.13

(Solving the metric part) Assume the same as in Proposition 9.11. Then there exist sufficiently small δ0>0 depending only on [f^±,Ω^,r^]|C_C such that for all δ(0,δ0) there exists a unique solution [Ωn+1,rn+1q]W3,(D¯) to (9.710)-(9.711) with [Ωn+1,rn+1]=[Ω^,r^] on C_C satisfying

[Ωn+1,rn+1]A[Ω^,r^]|C_C+A[f^±,Ω^,r^]|C_C. 9.714

In what follows, we prove the above two lemmas. We start with Lemma 9.12.

Proof

(Proof of Lemma 9.12 (Solving the Euler part).) For the sake of notational convenience, throughout the proof, we use U=Un, Ω=Ωn, r=rn, f±=fn± so that the variables without the indices refer to the ones from the n-th step.

Uniqueness follows from the uniqueness of characteristics since W2,C1 solutions fn+1±>0 satisfy the ODE along the characteristics

ddplogfn+1±(p,q±)=2k±(2qΩΩ-2ηqrr)U(p,q±). 9.715

The existence follows by the integral representation of (9.715)

fn+1±(p,q)=f^±(p,q)exppp2k±(2qΩΩ-2ηqrr)U(s,q±(s))ds 9.716

where f^±(p,q)=f^±(p,q), and (p,q) is the exit time and position associated with (p,q) constructed in Lemma 9.5. We focus on verifying the desired regularity and estimates.

First of all, clearly fn+1±C(D¯) since p is continuous. To estimate log(qN±fn+1±), we take log of (9.716) and rewrite it as

log(qN±fn+1±)=log(qN±f^±)pp2k±(2qΩΩ-2ηqrr)U(s,q±(s))ds. 9.717

For the first term, we may write it as

log(qN±f^±)=log(qN±f^±)+N±logqq.

Since |log(qN±f^±)|log(qN±f±)|C_C, it suffices to estimate log(qq). To this end, first let p=p0. Then we have

q=q±(p0;p,q)=q-p0pk±U(s)ds>q-k±U(p-p0)q-k±Uδ.

In particular, using qq0, we have q1-k±Uq0-1δq. Hence,

|logqq|-log(1-k±Uq0-1δ)q0-1δk±U 9.718

for sufficiently small δ>0. If p>p0, we have q=q0 and q<q0+k±U(p-p0). Hence in this case, for sufficiently small δ>0,

|logqq|log(1+k±Uq0-1δ)k±Uq0-1δ. 9.719

We next estimate the integral term in (9.717). Using q±(s)q for all psp, and p-pδ,

pp2k±(2qΩΩ-2ηqrr)U(s,q±(s))ds=4k±pp1q±q±(qΩΩ-ηqrr)U(s,q±(s))ds4k±UqqΩΩ-ηqqrrp-pq4k±UqqΩΩ-ηqqrrδq0. 9.720

Therefore by (9.717), (9.718), (9.719), (9.720), we obtain

log(qN±fn+1±)log(qN±f±)|C_C+N±k±Uδq0+4k±UqqΩΩ-ηqqrrδq0. 9.721

Next we show that qfn+1± is continuous. Since pC1 if pp0 by Lemma 9.5, the right-hand side of (9.716) is C1 and thus if pp0, qfn+1± is clearly continuous. Therefore, it suffices to show that qfn+1±(p,q) is continuous when p=p0 and q=q0. To this end, let p(p,q)=p0 and q(p,q)=q0, and take any (p¯,q¯)(p,q) in a small neighborhood of such (p,q). Then we have p(p¯,q¯)p0 and hence qfn+1±(p¯,q¯) is continuous. In fact, for any (p¯,q¯) with p¯:=p(p¯,q¯)p0 we can solve the equations for qlogfn+1±:

p(qlogfn+1±)+k±Uq(qlogfn+1±)+k±qUqlogfn+1±±4k±q(qΩΩ-ηqrr)U=0

along the characteristics and obtain the integral representation

qlogfn+1±(p¯,q¯)=qlogf^±(p¯,q¯)-p¯p¯k±qUqlogfn+1±±4k±q(qΩΩ-ηqrr)U(s,q±(s))ds 9.722

where

qlogf^±(p¯,q¯)=1p¯>p0qlogf^|C_±(p¯,q¯)+δp¯p0qlogf^±(p¯,q¯),

where δp¯p0 is the usual Kronecker delta. Notice that the integral terms are all continuous due to the continuity of p in Lemma 9.5 and therefore fn+1±C1(D¯).

To estimate qqlogfn+1±, as done in (9.718) and (9.719), we first observe that qq11-k±Uq0-1δ. Then for sufficiently small δ>0 we have

qq,q2q21+2k±Uδq0. 9.723

Based on the integral representation (9.722) we proceed to estimate

|qqlogf^±(p¯,q¯)|1+2k±Uδq0qqlogf^±|C_C, 9.724

where we have used the first bound in (9.723).

For the second term in (9.722),

qppk±qUqlogfn+1±±4k±q(qΩΩ-ηqrr)U(s,q±(s))dsk±qqUqqlogfn+1±+4q2q(qΩΩ-ηqrr)Uq(p-p)q2k±qqUqqlogfn+1±+4q2q(qΩΩ-ηqrr)U2δq0

where we have used the upper bound qq2 as it follows from (9.723) for δ>0 sufficiently small. Therefore, we deduce that

1-2k±qqUδq0qqlogfn+1±1+2k±Uδq0qqlogf^±|C_C+8k±δq0q2q(qΩΩ-ηqrr)U. 9.725

From (9.716), since fn+1±W2,, we just need to estimate q2q2logfn+1±. After applying qq to the equation (9.709), we have the following integral representation:

q2logfn+1±(p,q)=q2logf^±,(p,q)-ppk±q2Uqlogfn+1±+2k±qUq2logfn+1±±4k±q2(qΩΩ-ηqrr)U(s,q±(s))ds. 9.726

Analogously to the previous step, we obtain

1-2k±qqUδq0q2q2logfn+1±1+2k±Uδq0q2q2logf^±|C_C+2k±δq0q2q2Uqqlogfn+1±+8k±δq0q3q2(qΩΩ-ηqrr)U. 9.727

We now collect (9.721), (9.725), (9.727) and use (9.677), (9.682) as well as [fn±,Ωn,rn]2A[f^±,Ω^,r^]|C_C to deduce (9.713) for sufficiently small δ>0. This completes the proof.

Proof

(Proof of Lemma 9.13 (Solving the metric part).) As in the previous proof, for the sake of notational convenience, throughout the proof, we use Ω=Ωn, r=rn, f±=fn±, m=mn so that the variables without the indices refer to the ones from the n-th step.

Since (9.710), (9.711) are linear inhomogeneous ODEs, we directly integrate them to solve for rn+1 and Ωn+1. As the existence is clear, we focus on the estimates.

We now integrate (9.710) along an ingoing null curve from p0 to p to obtain

q((rn+1)2)(p,q)=q(r^2)(p0,q)+p0p-Ω22+2πΩ2r2η(f+f-)1k+-k-dp~ 9.728

with

(rn+1)2(p,q)=r^2(p,q0)+q0qq((rn+1)2)(p,q~)dq~. 9.729

We first estimate 1qq((rn+1)2). To this end, we bound the integral term of (9.728) as

1qp0p-Ω22+2πΩ2r2η(f+f-)1k+-k-dp~δq0Ω22+2πq0βq1+βΩ2r2η(f+f-)1k+-k- 9.730

where we have used p-p0qδq0. Therefore, we have

1qq((rn+1)2)1qq(r^2)|C+δq0Ω22+2πq0βq1+βΩ2r2η(f+f-)1k+-k-. 9.731

We also have

1qq((rn+1)2)(p,q)1qq(r^2)(p0,q)-δq0Ω22+2πq0βq1+βΩ2r2η(f+f-)1k+-k-. 9.732

Since 1qq(r^2)(p0,q)1qq(r^2)|C, for sufficiently small δ we obtain

qq((rn+1)2)(p,q)11qq(r^2)|C-δq0Ω22+2πq0βq1+βΩ2r2η(f+f-)1k+-k-qq(r^2)|C+2δq0qq(r^2)|C2Ω22+2πq0βq1+βΩ2r2η(f+f-)1k+-k-. 9.733

On the other hand, from (9.729) and (9.731), the upper bound of (rn+1)2q2 is easily obtained: since q0<q and using (9.731),

(rn+1)2(p,q)q2=r^2(p,q0)q2+1q2q0qq~q((rn+1)2)(p,q~)q~dq~r^2q2|C_C+121qq(r^2)|C+12δq0Ω22+2πq0βq1+βΩ2r2η(f+f-)1k+-k-. 9.734

Next we use (9.732) to first obtain from (9.729)

(rn+1)2(p,q)q2r^2(p,q0)q2+q2-(q0)22q2infC_C1qq(r^2)-12δq0Ω22+2πq0βq1+βΩ2r2η(f+f-)1k+-k-13min1q2r^2|C_C,1qq(r^2)|C_C-12δq0Ω22+2πq0βq1+βΩ2r2η(f+f-)1k+-k-, 9.735

where we have used the following: if q3q0, then r^2(p,q0)q2r^2(p,q0)3(q0)2131q2r^2|C_C, while if q0<q3, q2-(q0)22q2infC_C1qq(r^2)131qq(r^2)|C_C. As done in (9.733), we deduce

q2(rn+1)2(p,q)3maxq2r^2|C_C,qq(r^2)|C_C+9δq0maxq2r^2|C_C,qq(r^2)|C_C2Ω22+2πq0βq1+βΩ2r2η(f+f-)1k+-k-. 9.736

We now differentiate (9.710) with respect to q and integrate:

xq2((rn+1)2)(p,q)=q2(r^2)(p0,q)+p0pq-Ω22+2πΩ2r2η(f+f-)1k+-k-dp~. 9.737

Now the estimation of q2((rn+1)2) follows similarly. Using p-p0qδq0,

q2((rn+1)2)q2(r^2)|C+δq0Ω2q0θq1+θqlogΩ+2πq01+βq2+βqΩ2r2η(f+f-)1k+-k-. 9.738

The third derivative qq3((rn+1)2) can be estimated analogously:

qq3((rn+1)2)qq3(r^2)|C+δq0Ω2q2+θq2logΩq0θ+2q1+θqlogΩ2q02θ+δq02πq01+βq3+βq2Ω2r2η(f+f-)1k+-k-. 9.739

We proceed in the same way for Ωn+1. By integrating (9.711) along an ingoing curve we obtain the expression for qlogΩn+1(p,q)

qlogΩn+1(p,q)=q(logΩ^)(p0,q)+p0pΩ22r2mr-(1+η)πΩ2r2+2η(f+f-)1k+-k-dp~ 9.740

with

Ωn+1(p,q)=Ω^(p,q0)eq0qqlogΩn+1(p,q~)dq~. 9.741

To estimate q1+θqlogΩn+1, we start with the integral term of (9.740). By taking the sup norm, and using p-p0qδq0, we obtain

q1+θp0pΩ22r2mr-(1+η)πΩ2r2+2η(f+f-)1k+-k-dp~δq0121q1-θΩ2qr3m+(1+η)πq1-θ+βq3+βΩ2r2+2η(f+f-)1k+-k- 9.742

and hence

q1+θqlogΩn+1q1+θqlogΩ^|C+δq0121q01-θΩ2qr3m+(1+η)πq01-θ+βq3+βΩ2r2+2η(f+f-)1k+-k-. 9.743

For Ωn+1, we first estimate the integral of (9.741).

q0qqlogΩn+1(p,q~)dq~=q0q1q~1+θq~1+θqlogΩn+1(p,q~)dq~q1+θqlogΩn+11θ1q0θ.

Therefore we deduce

logΩn+1logΩ^|C_C+1θ1(q0)θq1+θqlogΩn+1|C+1θ1q0θδq0121q01-θΩ2qr3m+(1+η)πq01-θ+βq3+βΩ2r2+2η(f+f-)1k+-k-. 9.744

To estimate q2logΩn+1, we first observe that

q2logΩn+1(p,q)=q2(logΩ^)(p0,q)+p0pqΩ22r2mr-(1+η)πΩ2r2+2η(f+f-)1k+-k-dp~, 9.745

wherefrom we deduce

q2+θq2logΩn+1q2+θq2logΩ^|C+δq012Ω2qr3q1+βqmq01-θ+β+q4qΩ2r3mq01-θ+δq0(1+η)πq0q4+βqΩ2r2+2η(f+f-)1k+-k-. 9.746

Similarly one can derive

q3+θq3logΩn+1q3+θq3logΩ^|C+δq012Ω2qr3q2+βq2mq01-θ+β+2q4qΩ2r3q1+βqmq01-θ+β+q5q2Ω2r3mq01-θ+δq0(1+η)πq0q5+βq2Ω2r2+2η(f+f-)1k+-k-. 9.747

We now collect (9.731), (9.733), (9.734), (9.736), (9.738), (9.739), (9.743), (9.744), (9.746), (9.747) and use (9.689), (9.690), (9.678), (9.701), (9.702) as well as [fn±,Ωn,rn]2A[f^±,Ω^,r^]|C_C to deduce (9.713) for sufficiently small δ>0. This concludes the proof.

Convergence of the Iteration Scheme

From Proposition 9.11, the solutions [fn+1±,Ωn+1,rn+1] of (9.709), (9.710), (9.711) have the uniform bounds [fn+1±,Ωn+1,rn+1]2A[f^±,Ω^,r^]|C_C for all n0. In this section, we show the convergence of the sequence of approximations [fn+1±,Ωn+1,rn+1]. We will estimate the difference between [fn+1±,Ωn+1,rn+1] and [fn±,Ωn,rn]. Let

[fn+1±,Ωn+1,rn+1]:=[logfn+1±-logfn±,logΩn+1-logΩn,(rn+1)2-(rn)2] 9.748

for n0. Our next task is to prove the corresponding difference bounds.

Proposition 9.14

Let [fn+1±,Ωn+1,rn+1] be the solution (9.709), (9.710), (9.711) enjoying the uniform bounds [fn+1±,Ωn+1,rn+1]2A[f^±,Ω^,r^]|C_C for all n0. Then the difference norm for [fn+1±,Ωn+1,rn+1] defined in (9.748) satisfies the following recursive inequality

an+1Cδq0anforn1 9.749

where

an+1=fn+1±+q-2rn+1+q-1qrn+1+Ωn+1+q1+θqΩn+1 9.750

and C does not depend on n but only on 2A[f^±,Ω^,r^]|C_C.

The proof relies on C0 estimates of [fn+1±] and C1 estimates of [Ωn+1,rn+1] in the spirit of Lemma 9.12 and Lemma 9.13. Before giving the proof, we present some preliminary estimates. First, we note that [fn+1±,Ωn+1,rn+1] satisfy

pfn+1±+k±Unqfn+1±+k±(Un-Un-1)qlogfn± 9.751
±2k±(2qlogΩn-ηq(rn)2(rn)2)Un-(2qlogΩn-1-ηq(rn-1)2(rn-1)2)Un-1=0,pqrn+1=(Ωn-1)2-(Ωn)22+2π(Ωn)2(rn)2η(fn+fn-)1k+-k--(Ωn-1)2(rn-1)2η(fn-1+fn-1-)1k+-k-, 9.752
pqΩn+1=12(Ωn)2(rn)2mnrn-(Ωn-1)2(rn-1)2mn-1rn-1-(1+η)π(Ωn)2(rn)2+2η(fn+fn-)1k+-k--(Ωn-1)2(rn-1)2+2η(fn-1+fn-1-)1k+-k-, 9.753

with [fn+1±,Ωn+1,rn+1]|C_C=0.

We start with the following inequality, which will allow us to compare the difference of two functions to the logarithm of their ratio.

Lemma 9.15

Let L>0 be given. For any -1<xL, the following holds

|x|(1+L)|log(1+x)|. 9.754
Proof

If x(-1,0), we claim -x-(1+L)log(1+x). Letting h(x)=(1+L)log(1+x)-x, we have h(0)=0 and h(x)=1+L1+x-1>0 for x(-1,0). Hence h<0 for x(-1,0). Now let 0xL. Then since h(x)0 for 0xL, h0 for 0xL, which shows (9.754).

Lemma 9.16

For any b>0 let

Ω:=maxsupn(Ωn)2(Ωn-1)2,supn(Ωn-1)2(Ωn)2,±=maxsupnfn±fn-1±b,supnfn-1±fn±b, 9.755

and

R:=maxsupnqb(rn-1)b(q-1rn)b-(q-1rn-1)b(q-1rn)2-(q-1rn-1)2,supnqb(rn)b(q-1rn)b-(q-1rn-1)b(q-1rn)2-(q-1rn-1)2. 9.756

Note that Ω, ±, R are finite by the uniform bound [fn+1±,Ωn+1,rn+1]2A[f^±,Ω^,r^]|C_C. Then the following bounds hold:

fn±fn-1±b-1,fn-1±fn±b-1b±fn, 9.757
(Ωn)2(Ωn-1)2-1,(Ωn-1)2(Ωn)2-12ΩΩn, 9.758
(rn)b(rn-1)b-1,(rn-1)b(rn)b-1Rq-2rn. 9.759
Proof

(9.757) and (9.758) are direct consequences of (9.754), (9.755) with x+1=(fn±fn-1±)b, (fn-1±fn±)b, (Ωn)2(Ωn-1)2, (Ωn-1)2(Ωn)2. (9.759) directly follows by writing

(rn)b(rn-1)b-1=qb(rn-1)b(q-1rn)b-(q-1rn-1)b(q-1rn)2-(q-1rn-1)2q-2[(rn)2-(rn-1)2].

We are now ready to prove Proposition 9.14.

Proof of Proposition 9.14

We start with fn+1±. We recall the notations from Lemma 9.12. By integrating (9.751) along the characteristics and using the zero data, we have the representation of fn+1±:

fn+1±(p,q)=k±pp(S1+S2)(s,q±(s))ds 9.760

where

S1=-(Un-Un-1)qlogfn±,S2=2(2qlogΩn-ηq(rn)2(rn)2)Un-(2qlogΩn-1-ηq(rn-1)2(rn-1)2)Un-1. 9.761

Using q±(s)q for all psp and p-pδ,

ppS1(s,q±(s))ds=pp1q±(Un-Un-1)q±qlogfn±(s,q±(s))dsUn-1qqlogfn±UnUn-1-1δq0. 9.762

For UnUn-1-1, we rewrite

UnUn-1-1=ΩnΩn-12fn+fn-1+12fn-1-fn-12(rn-1)2(rn)2η-1=ΩnΩn-12-1fn+fn-1+12fn-1-fn-12(rn-1)2(rn)2η+fn+fn-1+12-1fn-1-fn-12(rn-1)2(rn)2η+fn-1-fn-12-1(rn-1)2(rn)2η+(rn-1)2(rn)2η-1. 9.763

Note that factors next to the rectangular brackets are all uniformly bounded. By Lemma 9.16, we deduce that

UnUn-1-1c1(fn+fn+q-2rn) 9.764

where c1>0 is independent of n and hence using the uniform bounds of the uniform bounds of [fn-1±,Ωn-1,rn-1] and [fn±,Ωn,rn], we have

ppS1(s,q±(s))dsc2δq0(fn+Ωn+q-2rn) 9.765

where c2>0 is independent of n. For S2, we first rewrite

S22=2qΩn-ηqrn(rn)2+q(rn-1)2(rn-1)2(rn-1)2(rn)2-1Un+(2qlogΩn-1-ηq(rn-1)2(rn-1)2)Un-1UnUn-1-1. 9.766

Now by using the uniform bounds on [fn-1±,Ωn-1,rn-1] and [fn±,Ωn,rn], Lemma 9.16, (9.764), and p-pqδq0, we deduce

ppS2(s,q±(s))dsc3δq0(q1+θqΩn+q-1qrn+Ωn+fn+q-2rn). 9.767

We now estimate q-2rn+1+q-1qrn+1. We recall the notations used in the proof of Lemma 9.13. We first integrate (9.752) along an ingoing null curve from the initial point p0=p0 to p to obtain

qrn+1(p,q)=p0pS3+S4dp~ 9.768

where

S3=-(Ωn-1)22(Ωn)2(Ωn-1)2-1,S4=2π(Ωn-1)2(rn-1)2η(fn-1+fn-1-)1k+-k-(Ωn)2(Ωn-1)2(rn-1)2η(rn)2η(fn+fn-1+)1k+-k-(fn-fn-1-)1k+-k--1. 9.769

The last factor of S4 can be rewritten via add and subtract trick as done for UnUn-1-1 in (9.763). Then using p-p0qδq0, Lemma 9.16, the uniform bounds, we deduce that

q-1qrn+1c4δq0(fn+Ωn+q-2rn). 9.770

By integrating (9.768) with respect to q, we have rn+1(p,q)=q0qqrn+1(p,q~)dq~. Then by writing q-2rn+1(p,q)=q-2q0qq~q~-1qrn+1(p,q~)dq~, we deduce that

q-2rn+1c42δq0(fn+Ωn+q-2rn). 9.771

The estimates for Ωn+1+q1+θqΩn+1 can be derived in the same fashion. By integrating (9.753),

qΩn+1(p,q)=p0pS5+S6dp~, 9.772

where

S5=12(Ωn)2(rn)2mnrn-(Ωn-1)2(rn-1)2mn-1rn-1,S6=-(1+η)π(Ωn-1)2(rn-1)2+2η(fn-1+fn-1-)1k+-k-(Ωn)2(Ωn-1)2(rn-1)2+2η(rn)2+2η(fn+fn-1+)1k+-k-(fn-fn-1-)1k+-k--1. 9.773

The structure of S6 is similar to S4 in the previous case. Using the uniform bounds for [fn-1±,Ωn-1,rn-1] and [fn±,Ωn,rn] and Lemma 9.16, we have

q1+θp0pS6dp~c5δq0(fn+Ωn+q-2rn) 9.774

for some c5>0 independent of n. For S5, we rewrite it

2S5=(Ωn)2(rn)3(mn-mn-1)+(Ωn)2(Ωn-1)2(rn-1)3(rn)3-1(Ωn-1)2(rn-1)3mn-1. 9.775

The second term is in the form where we can apply Lemma 9.16. For the first term, note that

mn-mn-1=q0q2π(fn+fn-)1k+-k-1(rn)2ηqrn-12(Ωn)-2rnfn-fn+12sn-2π(fn-1+fn-1-)1k+-k-1(rn-1)2ηqrn-1-12(Ωn-1)-2rn-1fn-1-fn-1+12sn-1dq~ 9.776

where we also recall (9.700) for sn. As done for UnUn-1-1 in (9.763), it is evident that the integrand can be written as the sum of the terms that contain one of the forms in Lemma 9.16 or qrn or sn-sn-1 where

sn-sn-1=q0q(Ωn-1)2-(Ωn)22+2π(Ωn)2(rn)2η(fn+fn-)1k+-k--2π(Ωn-1)2(rn-1)2η(fn-1+fn-1-)1k+-k-dq~.

Following the strategy of the proof of (9.701) in Lemma 9.10, using Lemma 9.16 and the uniform bounds of [fn-1±,Ωn-1,rn-1] and [fn±,Ωn,rn], we deduce

mn-mn-1c6δq0(fn+Ωn+q-2rn+q-1qrn). 9.777

Thus the S5 term gives the desired estimates and hence

q1+θqΩn+1c7δq0(fn+Ωn+q-2rn+q-1qrn) 9.778

for some constant c7>0 independent of n. The estimation of Ωn+1 directly follows from Ωn+1(p,q)=qqqΩn+1(p,q~)dq~ and (9.778):

Ωn+1c8δq0(fn+Ωn+q-2rn+q-1qrn) 9.779

for some constant c8>0 independent of n. Collecting (9.765), (9.767), (9.770), (9.771), (9.778), (9.779), we obtain (9.749).

We will now finish the proof of Theorem 9.4.

Proof of Theorem 9.4

Convergence and uniqueness. By Proposition 9.11 and Proposition 9.14, the iterates [fn±,Ωn,rn]nN satisfy the uniform bounds [fn±,Ωn,rn]2A[f^±,Ω^,r^]|C_C for all nN. Moreover fn±+q-2rn+q-1qrn+Ωn+q1+θqΩn0 as n, which in turn implies the strong convergence in C1 for fn± and in C2 for [Ωn,rnq] by standard interpolation. Hence, as n, [Ωn,rnq] converges to [Ω,rq] strongly in C2 and fn± to f± in C1. Using the strong convergence shown above, the fundamental theorem of calculus, and the formulas (9.722), (9.728)–(9.729), (9.737), (9.740)–(9.741), and (9.745), we may pass to the limit as n to conclude that pr(p,·),pΩ(p,·)W2,([q0,)) and pf(p,·)W1,([q0,)). In particular, using (9.697) and (9.710), upon passing to the limit we obtain a classical solution of (2.83). After passing to the limit in (9.709)-(9.711), we see that [f±,Ω,rq] is the desired classical solution to (2.83)–(2.84) and (9.648)–(9.649) in D. Note that in fact ppr(p,·),ppΩ(p,·)W1,([q0,)), which follows easily from (2.84), (9.697) and the bootstrap argument (involving one application of the q-derivative) similarly to above.

Observe that f±W2, and [Ω,r]W3, by standard weak-* convergence arguments, since the iterates are uniformly bounded in the same spaces. Uniqueness easily follows an adaptation of the difference estimates of Proposition 9.14. To show that the solution corresponds to the solution of the original problem (2.83)–(2.88), it only remains to show that the constraint equations (2.86)–(2.85) are satisfied on the interior of the domain D. This is a standard argument, which follows from the observation that for the solutions of (2.83)–(2.84) and (2.87)–(2.88) necessarily pqΩ-2qr+πrΩ2Tpp=0 and qpΩ-2pr+πrΩ2Tqq=0. Since the constraints are satisfied by the characteristic data, this gives the claim. Note that these expressions make sense given the above shown regularity.

Asymptotic flatness. To show the asymptotic flatness we must show that limqm(p,q)< for all p[p0,p0+δ]. By (9.670) this follows if we can show that ρqr and ρΩ2(up)2pr are integrable on [q0,). Note that qrL([q0,))< by the boundedness of our norms. Moreover, by integrating (9.697) with respect to q and using qrq and the boundedness of Ω, we conclude that pr1. From (9.653) we have

ρ=(f+f-)1k+-k-(1-ε)r2+2η=11-εq-N++N-k+-k--2-2η(qr)2+2η(qN+f+qN-f-)1k+-k-11-εq-3-βqr2+2η(qN+f+qN-f-)1k+-k-, 9.780

where we recall β from (9.688). Therefore ρ(p,·)L1([q0,)) since [f±,Ω,r]<. To show that ρΩ2(up)2 is integrable, we observe that

ρΩ2(up)2=(f+f-)1k+-k-(1-ε)r2+2η11+ηr2ηΩ2f-f+12=(f+)1k+-k--12(f-)1k+-k-+12(1+ε)Ω2r2.

Using (2.82) and (9.654), we also have the expression

Ω2(up)2=11+ηr2ηΩ2f-f+12q2η+N+-N-2qN-f-qN+f+121, 9.781

where we have used (9.660), (9.676), and the boundedness of our norms. Therefore, from (9.780) we conclude ρΩ2(up)2L1([q0,)), and the spacetime is therefore asymptotically flat.

Proof of Theorem 1.2: Existence of Naked Singularities

Recall the discussion of naked singularities in the introduction and in Section 2.7.3. Following [32], a spacetime contains a naked singularity if it corresponds to a maximal hyperbolic development of suitably regular data, and the future null-infinity is geodesically incomplete. The latter statement does not actually require the construction of future null-infinity as an idealised boundary attached to a suitable spacetime compactification. Instead, we define it to mean that affine length of a sequence of maximal ingoing null geodesics initiated along a sequence of points (approaching infinity) along an asymptotically flat outgoing null-surface, and suitably normalised, is uniformly bounded by some positive constant.

Proof of Theorem 1.2. For any ε(0,ε0] we consider the associated RLP spacetime (MRLP,ε,gRLP,ε) given in Definition 2.7. We use it to prescribe the data for the characteristic problem in the region D as described in Section 9.2. The associated solution to the Einstein-Euler system exists in region D by Theorem 9.4. Since U±=1+O(ε)>0, and since the data are exactly selfsimilar on C_ and on the finite segment {(p0,q)|q[q0,q0+A0]}C, we conclude that the solution coincides with the selfsimilar RLP-solution in the region DA0={(p,q)|p0p<0,q0q<q0+A0}, see Figure 7. We now consider a new spacetime (Mε,gε) obtained by gluing together the solution in the region D and to the past of the ingoing null-segment C_ inside MRLP,ε. Clearly, the new spacetime is identical to the RLP spacetime (and therefore smooth) in an open neighbourhood across C_. It therefore coincides with the exact selfsimilar RLP-spacetime (MRLP,ε,gRLP,ε) in the past of C_.

The exterior region, viewed as a development of the characteristic problem with data prescribed along the semi-infinite rectangle with outgoing data prescribed on {q0} and ingoing data on p[p0,0) is maximal, as the ingoing null-curve N is incomplete on approach to the singularity. In fact, since along backward null-cone N we have R-ετ=yN, we have

limτ0-(τ,R)Nρ(τ,R)=12πτ2Σ(yN)=, 9.782

since Σ(yN)0, where we recall that the density Σ is in fact strictly positive on [0,). Therefore, by (2.75) the Ricci scalar blows up as the observer approaches the scaling origin O along N.

It remains to show that the future null-infinity is incomplete in the sense of Definition 1.1. from [32]. Consider now a sequence of points (p0,qn) such that limnqn=. For any nN, consider the future oriented ingoing radial null-geodesic emanating from (p0,qn). Let the affine parameter4 be denoted by . Then q()qn, and the angular coordinates are also constant. The p-component satisfies the ODE

0=p¨()+Γppp(p˙())2=p¨()+plog(Ω2)(p˙())2, 9.783

where we have used (3.869). By our assumptions p(0)=p0 and we normalise the tangent vector to be parallel to p so that at (p0,qn), we have -2=g(q,p˙(0)p)=-p˙(0)12Ωn2, where Ωn:=Ω(p0,qn) for all nN. Therefore p˙(0)=4Ωn-2. We may integrate (9.783) once to conclude that dd(-1p˙+log(Ω2))=0 and therefore

-1p˙+log(Ω2)=-Ωn24+log(Ωn2). 9.784

It then follows that

p˙()=1Ωn24-2logΩnΩ(p(),qn). 9.785

However, by the proof of Theorem 9.4 there exists a constant C which depends only on the data [f±,Ω,r]|C_C such that plogΩL(D)C. In particular, by the mean value theorem logΩnΩ(p(),qn)C|p-p0|Cδ, for any p[p0,0) and nN. Note further that by (9.666) and our bounds on the data along C, Ωn2 is bounded from below and above uniformly in n. For δ1 sufficiently small, we conclude that p˙ remains positive and p reaches 0 (i.e. the Cauchy horizon) in finite -time, independent of n.

Remark 9.17

(The Cauchy horizon) By construction, the Cauchy horizon coincides with the null-curve {p=0}.

Remark 9.18

By construction, for any ε(0,ε0] there exists in fact an infinite family of naked singularity solutions. This freedom comes from the essentially arbitrary choice of the truncation in the region D, modulo the size and decay limitations imposed by the local existence theorem, Theorem 9.4. In a neighbourhood of the scaling origin O, our solutions are however exactly selfsimilar.

Acknowledgements

Y. Guo’s research is supported in part by NSF DMS-grant 2106650. M. Hadzic’s research is supported by the EPSRC Early Career Fellowship EP/S02218X/1. J. Jang’s research is supported by the NSF DMS-grant 2009458 and the Simons Fellowship (grant number 616364).

Appendix

Proof of Local Existence Around the Sonic Point: Combinatorial Argument

The main goal of the rest of this section is to show the convergence of the power series N=0RN(δx)N and N=0WN(δx)N. We will do so by induction on the coefficients RN and WN. Before proceeding, we record some technical lemmas. The proofs of Lemmas A.1A.4 below are given in detail in [16] and we therefore only state the lemmas without proof for reader’s convenience.

Lemma A.1

There exists a constant c>0 such that for all NN, the following holds

+m=N,m113m3cN3, 1.786
+m=N,m113m2cN2, 1.787
+m+n=N,m,n113m3n3cN3, 1.788
+m+n=N,m,n113m2n3cN2. 1.789

Lemma A.2

There exists a constant c>0 such that for all N3 and all C2, the following holds

=2N-1C-1qcCN-2Nq,q=2,3. 1.790

For any αR, we let

αj=(α)jj!=α(α-1)(α-j+1)j!forjN,andα0=1.

Lemma A.3

Recall the set π(n,m) defined in (4.188).

  1. For each nN,
    m=1nπ(n,m)(-1)mm!λ1!λn!121λ112nλn=2(n+1)12n+1 1.791
    holds.
  2. There exist universal constants c1,c2>0 such that
    c11n32(-1)n-112nc21n32,nN. 1.792

Lemma A.4

Let p>0 be a given positive number. Let (λ1,,λ)π(,m) where 1m and 2 be given.

  1. If 1m3, there exists a constant c3=c3(p)>0 such that
    n=11nλnpc3p. 1.793
  2. There exist c4=c4(p)>0 and L0=L0(p)>1 such that if LL0, the following holds:
    1Lm-1n=11nλnpc4p,for all1m. 1.794
  3. Let 3. Then there exists c5=c5(p)>0 such that if LL0, the following holds:
    1Lm-2n=11nλnpc5p,for all2m. 1.795

By Lemmas 4.1 and 4.6 there exist constants 0<m<M<1 such that

|R0|,|W0|,|R1|,|W1|M, 1.796
R0>m, 1.797

for all ε(0,ε0], with ε0>0 chosen sufficiently small as in Lemma 4.1.

Lemma A.5

Let x[xcrit+κ,xmax] and α(1,2). Assume that

|Rm|Cm-αm3,2mN-1, 1.798
|Wm|Cm-αm3,2mN-1, 1.799

for some C1 and N3. Then there exists a constant C^=C^(M)>0 such that

|(W2)|+|(RW)|+|(R2)|C^if=0,1,C^C-α3if2N-1, 1.800
|H|C^if=0,1,C^(1+ε)C-α3if2N-1. 1.801

Proof

We first prove the bounds for |(W2)|, 0. The bounds |(W2)0|M2 and |(W2)1|2M2 are obvious from (1.796). Clearly

|(W2)2|2M|W2|+M22MC2-α23+M2(2M+23M2)C2-α23 1.802

where we have used C2-α1. If 3 we then have

|(W2)|m=0|Wm||W-m|2|W0||W|+2|W1||W-1|+m=2-2|Wm||W-m|2MC-α3+2MC-1-α(-1)3+m=2-2C-2αm3(-m)32MC-α13+1(-1)3+12Mm=2-21m3(-m)32MC~C-α3, 1.803

for some constant C~. It is now clear, that the estimates for (RW) and (R2), 0 follow in the same way, as the only estimates we have used are (1.796) and the inductive assumptions (1.798)–(1.798), which both depend only on the index, and are symmetric with respect to R and W. Recalling H from (4.194), the bounds (1.801) now immediately follow from (1.800) and (1.799).

Lemma A.6

Let x[xcrit+κ,xmax] and α(1,2). Assume that (1.798) and (1.799) for N3 and some large enough C>1 satisfying

C>4L0c1 1.804

where c1 and L0=L0(32) are universal constants in (1.792) and Lemma A.4. Then there exists a constant C^=C^(R0) such that

|(R-η)|C^if=1,C^C-α3+C-22if2N-1. 1.805

Proof

When =1, |(R-η)1|=|-ηR0-η-1R1|R0-η-1. When =2 it is easy to see from (4.190) that

(R-η)2=-ηR0-ηR0-1R2-(η+1)R0-2R12. 1.806

In particular, (R-η)2cm-η-2R2+R12cm-η-2C2-α23+M2C^C2-α23+122 for some universal constant C^ and the claim is thus clear for =2. For 3, we rewrite (R-η) in the form

(R-η)=-ηR0-η-1R+R0-ηm=21R0mπ(,m)(-η)m1λ1!λ!R1λ1Rλ. 1.807

Now clearly, by the inductive assumption

-ηR0-η-1RC^C-α2 1.808

for a universal constant C^>0. To bound the second summand on the right-hand side of (1.807), we use (1.798) and Lemma A.3 first to conclude

R1λ1Rλ=n=1Rnλn113λ1C2-α23λ2C-α3λ=C(α-1)λ1+i=1(iλi-αλi)n=11n32λnn=11nλn32C(α-1)mC-αmc1-mn=1(-1)n-112nλnn=11nλn32

where we have used (α-1)λ1(α-1)m in the third line since α>1 and λ1m. Hence, using |(-η)m|m!, we observe that

(-η)m1λ1!λ!R1λ1RλC-mc1-m(-1)(-1)mm!λ1!λ!121λ112λn=11nλn32. 1.809

In turn by recalling (4.190) and using Lemma A.3 and Lemma A.4 with p=32, we see that

R0-ηm=21R0mπ(,m)(-η)m1λ1!λ!R1λ1RλR0-ηC(-1)(c1CR0)2m=2π(,m)1(c1CR0)m-2n=11nλn32(-1)mm!λ1!λ!121λ112λR0-ηC(c1CR0)2c432(-1)2(+1)12+1R0-η-12c2c4c12C-232(+1)12 1.810

where C is large enough so that (1.804) holds. This proves (1.805).

Remark A.7

In (1.804) the constant 4 in the numerator ensures that C>L0c1R0 for all x[xcrit+κ,xmax], since for ε0>0 sufficiently small, there exists a constant 0<δ<1 such that 1R0<3+δ for all ε(0,ε0].

Lemma A.8

Let x[xcrit+κ,xmax] and α(1,2). Then there exists a constant C>0 such that if C>C and for any N3, the following assumptions hold

|Rm|Cm-αm3,2mN-1, 1.811
|Wm|Cm-αm3,2mN-1, 1.812

then we have

|SN|βCN-αN21Cα-1+1C2-α+1C, 1.813
|VN|βCN-αN21Cα-1+1C2-α+1CN, 1.814

for some universal constant β>0.

Proof

We start with (1.813). Recall (4.255). First we show

R1R0-ηm=2N1R0mπ(N,m)(-η)m1λ1!λN!R1λ1RNλNCN-2N2. 1.815

Note that λN=0 and thus (1.815) does not depend on RN. As in (1.809), using (1.811) and Lemma A.3, we have

(-η)m1λ1!λN!R1λ1RNλNCN-mc1-m(-1)N(-1)mm!λ1!λN!121λ112NλNn=1N1nλn32.

Hence by using (1.795) of Lemma A.4, the left-hand side of (1.815) is bounded by

LHS of (1.815)R1R0-2k1-kCNm=2Nπ(N,m)(-1)N(c1CR0)mn=1N1nλn32(-1)mm!λ1!λl!121λ112NλNR1R0-2k1-kCN(c1CR0)2c5N32(-1)N2(N+1)12N+1R1R0-2k1-k-22c2c5c12CN-2N32(N+1)12

which shows (1.815). To estimate the second term on the right-hand side of (4.255), we use (1.811), (1.812) and (1.786) to obtain

x2R1[(1-ε)+m=N1mN-1WWm++m=N1mN-14εRWm]x2R1[2(1-ε)W1WN-1+4ε(R1WN-1+RN-1W1)]+x2R1[(1-ε)m=2N-2WN-mWm+m=2N-24εRN-mWm](1+ε)CN-1-αN3+(1+ε)m=2N-2CN-2α(N-m)3m3(1+ε)CN-1-αN3. 1.816

Finally, we estimate the last term on the right-hand side of (4.255) - the expression S~N - given by (4.256). To treat the first line of (4.256), by (1.811) and (1.805), we obtain

+m=N1mN-2(m+1)Rm+1(R-η)+m=N1mN-2Cm+1-α(m+1)2C-α3+C-22CN+1-2α+m=N1mN-21m23+CN-1-αm=1[N2]1m21(N-m)2CN-αN21Cα-1+1C. 1.817

For the first term of the second line of (4.256), by (1.811), (1.801), (1.787)

x2l+m=N1mN-2(m+1)Rm+1Hl+m=N1mN-2Cm+1-α(m+1)2C-α(+1)3CN+1-2αN2. 1.818

The other two terms of the second line can be estimated analogously. For the first term of the third line of (4.256), by (1.811), (1.812), (1.800), (1.786) we obtain

2x2(1-ε)+m+n=N1nN-1R(W+ε)m(R-W)n(R(W+ε))N-1(R1-W1)+(R(W+ε))1(RN-1-WN-1)+n=2N-2(R(W+ε))N-n(R-W)nCN-1-α(N-1)3+CN-1-α(N-1)3+n=2N-2CN-n-α(N-n)3Cn-αn3CN-1-αN3+CN-2αN3. 1.819

The bound on the remaining term in the third line of (4.256) is entirely analogous. Collecting all the bounds (1.815), (1.816), (1.817), (1.818), (1.819), we conclude

|SN|CN-αN21Cα-1+1CN+1CαN+1C, 1.820

which leads to (1.813) since 1<α<2 and C>1.

To prove (1.814), we first recall (4.258). Two first two terms on the right-hand side of (4.258) have the same structure as the first two terms in (4.255) and hence, by using the same uniform bound (1.796), they can be bounded analogously to (1.815) and (1.816). It remains to estimate V~N given in (4.259). The first, second and sixth lines of (4.259) have the same formal structure as the first, second and third lines of (4.256) and hence we obtain the same bounds as in (1.817), (1.818), and (1.819) by using (1.812) in place of (1.811) when necessary. We focus on the third, fourth and fifth lines of (4.259). For the first term of the third line, by using (1.805)

+m=N1mN(R-η)(-1)mR0-η+(R-η)1+=2N-1(R-η)1+C^=2N-1C-α3+C-22CN-1-αN3+CN-3N2 1.821

where we have used N3CN-1-α for all N3 and (1.790) with q=2,3 and C2. For the second term of the third line, we isolate m=N, m=N-1 cases, =0,1 cases, further =N-m and =N-m-1, and use (1.796), (1.805), (1.812)

3+m+n=N1mNWn(R-η)(-1)m1+ε+m=1N-2=0N-mWN-m-(R-η)1+m=1N-2|WN-m|+m=1N-2|WN-m-1|+C^m=1N-2=2N-mWN-m-C-α3+C-221+m=1N-2CN-m-α(N-m)3+m=1N-3CN-m-1-α(N-m-1)3+m=1N-2CN-m-α(N-m)3+CN-m-2(N-m)2+m=1N-2CN-m-1-α(N-m-1)3+CN-m-3(N-m-1)2+m=1N-2=2N-m-2CN-m--α(N-m-)3C-α3+C-22. 1.822

To bound the summations appearing in the third and the fourth line of (1.822), we apply (1.790) six times with =N-m and additionally we use 1CN-1-αN3 to bound them by

S1LCN-1-αN3+CN-3N2. 1.823

To bound the last line of (1.822) we use (1.787) and (1.790) to bound it by

m=1N-2CN-m-α=2N-m-21(N-m-)313+m=1N-2CN-m-2-α=2N-m-21(N-m-)312m=1N-2CN-m-α(N-m)3+m=1N-2CN-m-2-α(N-m)2CN-1-αN3+CN-3-αN2. 1.824

For the fourth line of (4.259) we only present the details for the first term as the other two terms are estimated analogously. We first isolate m=N and m=N-1 and then use (1.801), (1.790), and (1.786) to obtain

x2+m=N1mNH(-1)m1+=2N-1|H|1+=2N-1C-α3CN-1-αN3. 1.825

For the fifth line of (4.259) we only present the detail for the first two terms, as the estimate for the remaining two terms in the fifth line of the right-hand side of (4.259) is analogous and strictly easier. By (1.796), (1.812), (1.801) we have

3x2+n=N1nN-1WnH|WN-1H1|+|W1HN-1|+=2N-2|WN-H|CN-1-α(N-1)3+=2N-2CN--α(N-)3C-α3CN-1-α(N-1)3+CN-2αN3. 1.826

The second term of the fifth line of (4.259) can be estimated in a similar way as in (1.822) by using (1.801) instead of (1.805):

3x2+m+n=N1mNWnH(-1)m1+m=1N-2=0N-m|WN-m-H|1+m=1N-2|WN-m|+m=1N-3|WN-m-1|+m=1N-2=2N-m|WN-m-H|1+m=1N-2CN-m-α(N-m)3+m=1N-3CN-m-1-α(N-m-1)3+m=1N-2CN-m-α(N-m)3+CN-m-1-α(N-m-1)3+m=1N-2=2N-m-2CN-m--α(N-m-)3C-α3CN-1-αN3+CN-1-2αN3 1.827

where we have used (1.790) and (1.786) as before. Combining all the estimates, we deduce the desired bound (1.814).

Lemma A.9

Let x[xcrit+κ,xmax] and α(1,2). Consider (R0,W0), (R1,W1) constructed in Lemma 4.1 and Lemma 4.6, and let (RN,WN) be given recursively by (4.264) and (4.265). There exist a constant C>1 and ε0>0 such that for all 0<ε<ε0 and all x(xmin,xmax)

|RN|CN-αN3, 1.828
|WN|CN-αN3, 1.829

for all N2.

Proof

The proof is based on induction on N. When N=2, it is clear that there exists C0=C0(α)>1 such that for any C>C0 the bound

|R2|,|W2|C2-α23 1.830

holds true for any x[xcrit+κ,xmax]. Fix an N3 and suppose the claim is true for all 2mN-1. Then (1.811) and (1.812) are satisfied and therefore by Lemma A.8 we conclude that (1.813) and (1.814) hold for all C>C. Together with (4.266) and (4.267), those bounds lead to

|RN|β0β1+εN1Cα-1+1CN+εCN-αN3β0β1+ε31Cα-1+13C+εCN-αN3=:c1CN-αN3,|WN|β0β1+1N1Cα-1+1CN+εCN-αN343β0β1Cα-1+13C+εCN-αN3=:c2CN-αN3.

It is now clear that since α>1 we can choose C>C,C0 sufficiently large and ε<ε0 with ε0>0 sufficiently small so that c1,c2<1 and hence (1.828) and (1.829) hold true.

Null-Geodesic Flow in the RLP-Spacetime

We shall carry out the analysis of nonradial null-geodesics (NNG) in the comoving coordinates (τ,R,θ,ϕ). Partial analysis of simple NNG-s in Schwarzschild coordinates was already carried out in [28] and a further analysis of non-spacelike geodesics was done in [19]. A related problem for the selfsimilar dust collapsing clouds was studied in detail in [29].

It is convenient to introduce

z:=ετR=-1y,s:=-logR, 2.831

so that

zN:=-1yN,z1:=-1y1, 2.832

correspond to the boundary of the backward light cone N (yN=YN-1-η) and the “first" outgoing null-geodesic B1 (y1=-|Y1|-1-η) respectively. Then zτ=1εe-s and sτ=-1εze-s and hence dτ=e-sε(dz-zds) and dR=-e-sds. It is straightforward to check that the homothetic Killing vector field ξ=ττ+RR takes the form ξ=-s and the metric g in these coordinates reads

g=e-2s-e2μεdz2+2e2μεzdzds+e2λ-e2μεz2ds2+χ2dϕ2. 2.833

We note that the metric is not regular at z=0. This corresponds to a harmless coordinate singularity which can be easily avoided by introducing a suitable change of variables, see Section 7. We shall nevertheless work with (2.833) to avoid further notational complications and formally limit our analysis to the regions of MRLP,ε satisfying {z<0} and {z>0}.

Let γκ, κ=z,s,θ,ϕ be a null-geodesic and we denote the associated tangent vector by Vκ:=γ˙κ, κ=z,s,θ,ϕ. We let denote the affine parameter. Due to spherical symmetry, we have

γθ()=θ()=π2. 2.834

The remaining geodesic equations read

dVκd=12(κgαβ)VαVβ,κ=z,s,ϕ, 2.835

where Vκ=gκνVν.

Lemma A.1

(Geodesic flow) Let Vκ be the tangent to a null-geodesic as above. Then, there exist constants C,LR such that

-e2με(Vz)2+2e2μεzVzVs+e2λ-e2μεz2(Vs)2+χ2(Vϕ)2=0, 2.836
Vϕ=L, 2.837
Vs=C. 2.838

Moreover, the Vs-component satisfies the quadratic equation

eλ+μz2-εe2λ-2μVse2s2+2Cεeλ-μVse2s-e-λ-μεC2+e-λ+μz2L2χ2=0. 2.839

Proof

Equation (2.836) is just the statement that the 4-vector Vα is null. We now let α=ϕ in (2.835). Since gαβ-terms do not depend on ϕ, from (2.835) we immediately see that dVϕd=0, which implies (2.837). Since Vϕ=gϕϕVϕ, we also obtain

Vϕ=Lχ2e2s. 2.840

Next let κ=s in (2.835). Since sg=-2g (cf. (2.833)) we deduce that dVsd=0 from (2.835) and (2.838) follows.

Since Vs=gssVs+gszVz, we obtain

e2λ-e2μεz2Vse2s+e2μεzVze2s=C, 2.841

which in turn gives

zVze2s=εCe2μ+z2-εe2λ-2μVse2s. 2.842

Using (2.840), we may rewrite (2.836) as

-e2μεVze2s2+2e2μεzVze2sVse2s+e2λ-e2μεz2Vse2s2+L2χ2=0. 2.843

Recall that we work in a patch where z0. Plugging (2.842) into (2.843) to replace Vz, we get

-e2μεεCze2μ-+1zz2-εe2λ-2μVse2s2+2e2μεzεCze2μ+1zz2-εe2λ-2μVse2sVse2s-e2μεz2-εe2λ-2μVse2s2+L2χ2=0

which is easily simplified to give (2.839).

Remark A.2

Relations (2.837)–(2.838) are the Hamiltonian constraints associated with the geodesic flow. Equation (2.837) expresses the conservation of angular momentum associated with the Killing vector field ϕ, while (2.838) is the conservation law generated by the homothetic Killing vector field -s=ττ+RR.

Lemma A.3

Assume that there exists a simple nonradial null-geodesic, i.e. a null-geodesic such that L0 and z()=z¯const for all R. Then z¯ is a critical point of the function

H(z):=e2λ(z)-e2μ(z)εz2χ(z)-2. 2.844

Proof

By our assumption, for any simple geodesic we have Vz=0. By letting κ=z in (2.835), we then obtain the formula

dde2μεe-2sVs=e-2s2ze2λ-e2μεz2(Vs)2+e-2s2z(χ2)(Vϕ)2. 2.845

Since Vz=0 in (2.841), we conclude that Vse-2s is -independent and therefore the left-hand side of (2.845) vanishes. Letting Vz=0 in (2.836) we have (Vϕ)2=-e2λ-e2μεz2(Vs)2χ2. We plug this back into (2.845) and the claim follows.

Lemma A.4

(Monotonicity of H) The function H defined by (2.844) is strictly monotone on (-,0).

Proof

Recalling the notation x=r~(y), χ(y)=xy, and z=-1y, we see that H, written as a function of y can be written in the form

H(y)=x-2e2λy2-1εe2μ. 2.846

Therefore

H(y)=-2x-2xxe2λy2-1εe2μ+x-22e2λy2λ+1y-2εe2μμ=-2x-2w+ε(1+ε)ye2λy2-1εe2μ+2x-2e2λy2-11+εΣΣ+2y-2xx-1y+1y+1εe2με1+εΣΣ=-2x-2w+ε(1+ε)ye2λy2-1εe2μ+2x-21+εe2λy2-ΣΣ+1+3εy-2w+εy+e2μΣΣ=-2x-2w+ε(1+ε)ye2λy2-1εe2μ+2x-21+εΣΣe2μ-e2λy2+e2λy1+ε-2w, 2.847

where we have used the relation xx=w+ε(1+ε)y (see (2.33)), formula (3.97) for λ, and (3.99) to substitute for μ. We now use (3.117) and part (c) of Lemma 3.3 to obtain the formula

Σ(1+ε)Σe2μ-e2λy2=-2ye2λK=-2(d-w)(1+ε)e2λy. 2.848

Using this in (2.847) we obtain

H(y)=-2x-2w+ε(1+ε)ye2λy2-1εe2μ+2x-21+εe2λy1+ε-2d=2x-2(1+ε)ye2λ1ε(w+ε)e2μ-2λ+y21-w-2d. 2.849

We now recall the formula

e2μ-2λ=y2x2(w+ε)2d-η+εx2(w-1)2-4εdwx2, 2.850

see (3.121). We plug it back into (2.849) to conclude that

H(y)=2x-2y(1+ε)e2λ1εd-ηx2(w+ε)+(w-1)2w+ε-4dww+ε+1-w-2d=2x-2y(1+ε)e2λd-ηx21ε1w+ε-d1+ηx22+4ww+ε+(w-1)2w+ε+1-w. 2.851

Recall that 16w1 for all y(0,). Moreover, by (6.469) and the inequality w(y)=W(x)<1 for all y,x0 we know that that the function xD1+ηx2W(x) is increasing. It follows from (6.468) that there exists an ε-independent constant C such that D1+ηx2C for all ε(0,ε0], with ε0 sufficiently small. Since 1w+ε11+ε, 2+4ww+ε2+413+ε, and (w-1)2w+ε+1-w>0, we conclude from (2.851) that for ε>0 sufficiently small that H is strictly positive for y(0,).

Lemma A.5

(L=0: radial null-geodesics) Let Vκ as above. If L=0 the flow takes the form

dsdz=VsVz=1z±εeλ-μ, 2.852

while Vs, Vz are then recovered using (2.842). The behaviour of the solutions is described by Lemma 8.8.

Proof

Equation (2.852) is a simple consequence of (2.843). Equation (2.852) is equivalent to the radial null geodesic equation (8.641) expressed in (Y,logR)-variables.

Lemma A.6

(L0: nonradial null-geodesics) Let Vκ as above and L0. Then any such geodesic that emanates from the union of the exterior and the interior region exists globally in its affine time and does not converge to the scaling origin.

Proof

Step 1. We first consider the case where C=0. Then (2.842) gives the relation

dsdz=VsVz=-zεe2λ-2μ-z2. 2.853

In order for Vs to satisfy (2.839) we must have εe2λ-2μ-z2>0 since L2χ2>0. In particular, the null geodesics are confined to the exterior region zN<z<z1 where zN:=-|YN|1+η<0 and z1:=|Y1|1+η>0. Let s0R and z0(zN,z1) be given initial point in the exterior region. Then by integrating (2.853), we obtain

s=s0-z0z(s)zεe2λ-2μ-z2dz. 2.854

We recall that εe2λ-2μ-z2|z=zN,z1=0 and εe2λ-2μ-z2>0 for zN<z<z1. Since s=-logR, we deduce that R (s-) as z(s)zN or z(s)z1. Therefore a null geodesic in this case asymptotes to the backward light cone z=zN in the past and to the Cauchy horizon z=z1 in the future. In particular, there are no null geodesics emanating from or going into the origin.

Step 2. We now consider the general case C0. Assume without loss of generality C>0. In order for the solution of (2.839) to exist the discriminant must be nonnegative. This amounts to requiring

a(z):=e2μ(z2-εe2λ-2μ)χ2εC2L2. 2.855

Note that limz-a(z)=+, which follows from z=-1y, (2.74), and eμ|y=0>0, while limzzmsa(z)=- by Proposition 7.16. Moreover, we also know a(zN)=0, a(z1)=0 and a(z)<0 for zN<z<z1 from the analysis of radial null geodesics. Therefore, for given εC2L2>0, there exists z¯<zN such that a(z)>εC2L2 for all z(-,z¯) and hence, the non-radial null geodesics cannot enter the part of the interior region z<z¯.

Under the assumption (2.855), the solution to (2.839) is given by

Vse2s=Cε-εeλ-μ±|z|1-e2μ(z2-εe2λ-2μ)χ2L2εC2eλ+μz2-εe2λ-2μ. 2.856

From (2.842) we also have (recall z0)

zVze2s=±Cε|z|eλ+μ1-e2μ(z2-εe2λ-2μ)χ2L2εC2. 2.857

Let

Q(z):=1-e2μ(z2-εe2λ-2μ)χ2L2εC2. 2.858

We then conclude that

dsdz=VsVz=z-εeλ-μ±|z|Q(z)±|z|Q(z)(z2-εe2λ-2μ). 2.859

Step 3. No simple NNG-s. By definition, we have dzds=0 for simple NNG-s. Therefore by (2.859) they must satisfy either z2-εe2λ-2μ=0 or Q(z)=0. The first case corresponds to simple radial null geodesics, while zeros of Q(z)=0 possibly describe the simple non-radial geodesics. We know that z2-εe2λ-eμ<0 for all zN<z<z1 and thus, Q>0 for all zN<z<z1. Therefore we deduce that there is no simple non-radial null geodesics in the exterior (i.e. in the causal past of B1). In the interior region however z¯ cannot correspond to a geodesic, since otherwise by Lemma A.3 we would have H(z¯)=0, but this is impossible by Lemma A.4.

Step 4. NNG-s cannot emanate from or go towards the scaling origin O.

Step 4.1. First letz1>z0>0 and s0R be given in the exterior region above τ=0 (z>0). Then the null geodesics satisfy

dsdz=-εeλ-μ±zQ(z)±Q(z)(z2-εe2λ-2μ). 2.860

We start with (-) part. In this case, the right-hand side of (2.860) is negative and hence the null geodesics are outgoing. We claim that outgoing null geodesics (-) starting from z0>0 and s0R meet z=0 (namely τ=0) for some positive R>0 in the past. Observe that there exist -<M1<M0<0 such that M1<Q(z)(z2-εe2λ-2μ)<M0 for all z[0,z0]. Integrating (2.860), we have

s(0)=s(z0)+z00εeλ-μ+zQ(z)Q(z)(z2-εe2λ-2μ)dz. 2.861

The integral in the right-hand side is finite and using s=-logR, the claim follows. On the other hand, in the future direction, since z2-εe2λ-2μ0 as zz1-, s- as zz1- and thus it asymptotes to the Cauchy horizon z=z1.

We move onto (+) part, the ingoing case. As above, the null geodesics meet τ=0 for some positive R>0 in the past as the corresponding integral stays finite. We claim that the null geodesics meets B1 at positive R>0 in the future. Integrating (2.860), we have

s=s(z0)+z0z(s)-εeλ-μ+zQ(z)Q(z)(z2-εe2λ-2μ)dz 2.862

for z=z(s)<z1. To prove the claim, it suffices to show the integral is finite when z(s)=z1. To this end, we rewrite the integrand as

-εeλ-μ+zQ(z)Q(z)(z2-εe2λ-2μ)=-εeλ-μ+z+z(Q(z)-1)Q(z)(z2-εe2λ-2μ)=1Q(z)(z+εeλ-μ)+z(Q(z)2-1)Q(z)(z2-εe2λ-2μ)(Q(z)+1)=1Q(z)(z+εeλ-μ)-ze2μχ2L2εC2Q(z)(Q(z)+1)

for z0<z<z1, where we have used (2.858) in the last line. It is now clear that both terms are finite and thus the integral is indeed bounded, thereby proving the claim.

Step 4.2. Next we let zN<z0<0 and s0R be given in the exterior region below τ=0 (z<0). The null geodesics satisfy

dsdz=-εeλ-μzQ(z)Q(z)(z2-εe2λ-2μ). 2.863

We start with (+) part. In this case, we claim that the null geodesics meet τ=0 for positive R>0 in the future and asymptotes to the past light cone zN in the past. The geodesic in the future satisfies

s(0)=s(z0)+z00-εeλ-μ+zQ(z)Q(z)(z2-εe2λ-2μ)dz. 2.864

Observe that the integrand is positive and uniformly bounded. Therefore, the claim for the future follows. Now in the past, the right-hand side of (2.863) is positive and the denominator goes to 0 as zzN. Therefore s- and R as zzN, and hence the null geodesics asymptote to z=zN.

We move onto (-) part. Since the denominator is uniformly bounded for z[z0,0], the integral is finite and hence we deduce that the null geodesics meet τ=0 for some positive R>0 in the future. On the other hand, the null geodesics in the past satisfy

s=s(z0)+z0z(s)εeλ-μ+zQ(z)Q(z)(z2-εe2λ-2μ)dz. 2.865

We claim that they meet z=zN for finite R>0. To this end, we rewrite the integrand as

εeλ-μ+zQ(z)Q(z)(z2-εe2λ-2μ)=εeλ-μ+z+z(Q(z)-1)Q(z)(z2-εe2λ-2μ)=1Q(z)(z-εeλ-μ)-ze2μχ2L2εC2Q(z)(Q(z)+1)

for zN<z<z0<0, where we have used (2.858) in the last line. It is now clear that both terms are finite and thus the integral is bounded, thereby proving the claim.

Step 5. NNG-s in the interior region. Using the same strategy as above, one can show that nonradial geodesics (L0) starting in the interior region will asymptote to the past to the boundary of the backward light cone N, and in the future they will asymptote to the Cauchy horizon B1 to the future. In between, they will go through a turning point, which corresponds to the point (s¯,z¯), where z¯<zN is the zero of Q discussed above. We omit the details.

Einstein-Euler System in the Double-Null Gauge

Proof of Lemma 2.13

Proof of Lemma 2.13

Note that

Tpq=14Ω4Tpq,Tpp=14Ω4Tqq,Tqq=14Ω4Tpp. 3.866

We use (2.89) to obtain γABTAB=2ερr2=ηΩ2r2Tpq. Equations (2.83)–(2.86) now follow directly from Section 5 of [11]. Formulas (2.89)-(2.90) follow directly from (1.4) and (2.80). The fluid evolution equations (2.87)–(2.88) are a consequence of the Bianchi identities μTμν=0, ν=p,q,2,3. The Christoffel symbols associated with the purely angular degrees of freedom ΓBCA, A,B,C{2,3} may not vanish, but are not needed in the computations to follow, so we do not compute them. All the remaining non-vanishing Christoffel symbols are given by (Appendix A in [11])

ΓABp=-gpqrqrγAB,ΓABq=-gpqrprγAB, 3.867
ΓBqA=ΓqBA=qrrδBA,ΓBpA=ΓpBA=prrδBA, 3.868
Γppp=plog(Ω2),Γqqq=qlog(Ω2). 3.869

By (2.90) we note that

γABTAB=ερr-2γABγAB=2ερr-2. 3.870

We also observe that by (3.870) and (2.89)

Ω-2γABTAB=2ερΩ-2r-2=2ε1-εr-2Tpq=ηr-2Tpq, 3.871

where we recall the notation η=2ε1-ε. To express the Bianchi identities in coordinates, we use the formula

δTαβ=δTαβ+ΓγδαTγβ+ΓγδβTαγ 3.872

and the formulas (3.867)–(3.869). We start the with the p-equation and obtain from (1.2)

0=μTμp=pTpp+qTqp+ATAp. 3.873

By (3.872) and (3.867)–(3.869) we have

pTpp=pTpp+2ΓpppTpp=p+4pΩΩTpp, 3.874
qTqp=qTpq+ΓqqqTpq=q+2qΩΩTpq, 3.875
ATAp=ATAp+ΓγAATγp+ΓγApTAγ=ΓpAATpp+ΓqAATqp+ΓABpTAB=2prrTpp+2qrrTpq-gpqrqrγABTAB=2prrTpp+2qrrTpq+2Ω-2rqrγABTAB=2prrTpp+(2+2η)qrrTpq, 3.876

where we have used (2.80) in the next-to-last line and (3.871) in the last. Similarly, we compute the q-equation of the Bianchi identities (1.2).

0=μTμq=pTpq+qTqq+ATAq. 3.877

By (3.872) and (3.867)–(3.869) we have just like above

pTpq=p+2pΩΩTpq,qTqq=q+4qΩΩTqq, 3.878
ATAq=ΓpAATpq+ΓqAATqq+ΓABqTAB=2prrTpq+2qrrTqq+2ηprrTpq=2+2ηprrTpq+2qrrTqq. 3.879

Therefore, keeping in mind equations (3.873) and (3.877) can be rewritten in the form

p+plogΩ4r2Tpp+q+qlogΩ2r2+2ηTpq=0, 3.880
p+plogΩ2r2+2ηTpq+q+qlogΩ4r2Tqq=0, 3.881

which is equivalent to (2.87)–(2.88). From (2.89)–(2.90) and (2.82) we have the additional relationship

TppTqq=(1+ε)2ρ2(up)2(uq)2=(1+ε)2ρ2Ω-4=1+ε1-ε2(Tpq)2=(1+η)2(Tpq)2. 3.882

Double-Null Description of the RLP-Spacetime

Lemma A.1

The region to the past of the curve B1 in the RLP spacetime (MRLP,ε,gRLP,ε) is well-defined in the double-null coordinates, with the normalisation conditions (9.655)–(9.656).

Proof

Recalling the expressions (2.7) and (2.79), it is easy to see that

Ω2τpτq=e2μ,Ω2RpRq=-e2λ,τpRp+τqRq=0. 3.883

Using (3.883) we conclude that τpRp=-τqRq=eμ-λ. Recall that the 4-velocity uν in the comoving frame takes the form e-μτ and therefore

V=e-μτr=uppr+uqqr=1Ω2uqpr+uqqr. 3.884

It follows that uq solves the quadratic equation qr(uq)2-Vuq+1Ω2pr=0. Therefore

up=V-V2-4prqrΩ22pr=V-V2+1-2mr2pr, 3.885
uq=V+V2-4prqrΩ22qr=V+V2+1-2mr2qr, 3.886

where the choice of signs in the solutions of the quadratic equations is consistent with the normalisation that e-μτ=upp+uqq is future pointing.

We now let r along the outgoing geodesic through (τ0,R0). By Lemma 8.8 we know that this curve asymptotes to B1 in the (τ~,R) plane, or equivalently, y converges to y1 in the (τ,R)-plane. Recalling (2.21), (3.99)–(3.101), and (2.27), we see that V(τ,R)=εV(y)=εr~(y)w(y)-11+εΣ(y)ε1+εC¯1ε as yy1 for some constant C¯1>0. Recall here that the density Σ(y)=d(y)1-ε1+ε=d(Y)1-ε1+ε for y<0. Observe that along the outgoing geodesic

limq2mr-1=-1+Cε 3.887

for some ε-independent constant C>0 by an argument analogous to (8.614). Using the normalisation condition (9.656), we see that there exists a constant C¯2>0 such that

limq(uq)2=C¯2>0. 3.888

Using the normalisation condition (9.656), equation (2.85) along the outgoing null-geodesic through (τ~0,R0) takes the form

q(Ω-2)=-2π(1+ε)(12q+r)Ω-2ρ(uq)-2, 3.889

where we have used (2.82) and (2.82). Equivalently, after multiplying by Ω2 and integrating, we obtain the formula

qlogΩ=(1+ε)π(12q+r)ρ(uq)-2. 3.890

Note that by (2.19) ρ(τ,R)=εR2y2Σ(y). Since we know by Lemma 8.8 that as q, yy1, and thereby rRχ(y1) it follows by q=2(r-r) that there exists a constant C1 independent of ε such that

ρ(τ,R)1q2qC1ε.

We then use (3.888) to conclude that that the leading order asymptotic behaviour of the right-hand side of (3.890) is Cεq for some ε-independent constant C>0. Therefore Ω(p0,·) is well-defined and ΩqqCε. Moreover, from (9.667) and (3.887), we have

limq4prqrΩ2=limq2mr-1=-1+Cε. 3.891

It follows therefore, that along the outgoing geodesic prqΩ2qq2Cε. We can similarly use the normalisation condition (9.655) to determine Ω along the ingoing geodesic through (τ~0,R0) in the past of B1. Using (2.83)–(2.84), it is straightforward to show that the the solution exists everywhere in the past of B1 with the asymptotic behaviour ΩqqCε, prqq2Cε, qrq1.

Data Availability

All data generated or analysed during this study are included in this published article.

Declarations

Conflict of Interests

The authors have no relevant financial or non-financial interests to disclose.

Footnotes

1
For two non-vanishing functions xA(x), xB(x) we write Axx¯B to mean that there exist ε-independent constants C1,C2>0 such that
C1lim infxx¯A(x)B(x)lim supxx¯A(x)B(x)C2.
2

The LP- and Hunter-type solutions coincide at x=2.

3

The nomenclature is motivated by their Newtonian analogue discovered by Larson [20] and Penston [31] in 1969, see [15] for the rigorous proof of existence of the Newtonian solution.

4

A geodesic γν is parametrised by an affine parameter if γ¨ν()+Γαβνγ˙α()γ˙β()=0.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Contributor Information

Yan Guo, Email: Yan_Guo@brown.edu.

Mahir Hadzic, Email: m.hadzic@ucl.ac.uk.

Juhi Jang, Email: juhijang@usc.edu.

References

  • 1.Bondi H. Spherically symmetrical models in general relativity. Monthly Notices of the Royal Astronomical Society. 1947;107(5–6):410–425. doi: 10.1093/mnras/107.5-6.410. [DOI] [Google Scholar]
  • 2.Carr BJ, Coley AA. Self-similarity in general relativity. Class. Quantum Grav. 1999;16:R31. doi: 10.1088/0264-9381/16/7/201. [DOI] [Google Scholar]
  • 3.Carr BJ, Coley AA, Goliath M, Nilsson US, Uggla C. Critical phenomena and a new class of self-similar spherically symmetric perfect-fluid solutions. Phys. Rev. D. 2000;61(081502):1–5. [Google Scholar]
  • 4.Carr BJ, Gundlach C. Spacetime structure of self-similar spherically symmetric perfect fluid solutions Phys. Rev. D. 2003;67:024035. doi: 10.1103/PhysRevD.67.024035. [DOI] [Google Scholar]
  • 5.Christodoulou D. Violation of cosmic censorship in the gravitational collapse of a dust cloud. Commun. Math. Phys. 1984;93:171–195. doi: 10.1007/BF01223743. [DOI] [Google Scholar]
  • 6.Christodoulou D. Examples of naked singularity formation in the gravitational collapse of a scalar field. Annals of Math. 1994;140:607–653. doi: 10.2307/2118619. [DOI] [Google Scholar]
  • 7.Christodoulou, D., On the global initial value problem and the issue of singularities, Class. Quantum Grav., A 23 (1999)
  • 8.Christodoulou D. The Instability of Naked Singularities in the Gravitational Collapse of a Scalar Field. Annals of Math. 1999;149(1):183–217. doi: 10.2307/121023. [DOI] [Google Scholar]
  • 9.Dafermos MC. Spherically symmetric spacetimes with a trapped surface Class. Quantum Grav. 2005;22(11):2221–2232. doi: 10.1088/0264-9381/22/11/019. [DOI] [Google Scholar]
  • 10.Dafermos, M. C., Luk, J., The interior of dynamical vacuum black holes I: The C0-stability of the Kerr Cauchy horizon. preprint,arXiv:1710.01722v1 (2017)
  • 11.Dafermos MC, Rendall A. An extension principle for the Einstein-Vlasov system under spherical symmetry. Ann. Henri Poincare. 2005;6:1137–1155. doi: 10.1007/s00023-005-0235-7. [DOI] [Google Scholar]
  • 12.Ehlers J, Kind S. Initial-boundary value problem for the spherically symmetric Einstein equations for a perfect fluid. Class. Quant. Grav. 1993;10:2123–2136. doi: 10.1088/0264-9381/10/10/020. [DOI] [Google Scholar]
  • 13.Goliath M, Nilsson US, Uggla C. Timelike self-similar spherically symmetric perfect-fluid models. Class. Quantum Grav. 1998;15(9):2841–2863. doi: 10.1088/0264-9381/15/9/028. [DOI] [Google Scholar]
  • 14.Gundlach, C., Martín-García, J.M., Critical Phenomena in Gravitational Collapse. Living Rev. Relativ.10, 5 (2007) 10.12942/lrr-2007-5 [DOI] [PMC free article] [PubMed]
  • 15.Guo Y, Hadzic M, Jang J. Larson-Penston Self-similar Gravitational Collapse. Comm. Math. Phys. 2021;386:1551–1601. doi: 10.1007/s00220-021-04175-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Guo Y, Hadzic M, Jang J, Schrecker M. Gravitational Collapse for Polytropic Gaseous Stars: Self-similar Solutions. Archive Rat. Mech. Anal. 2022;246:957–1066. doi: 10.1007/s00205-022-01827-8. [DOI] [Google Scholar]
  • 17.Harada T. Final fate of the spherically symmetric collapse of a perfect fluid. Phys. Rev. D. 1998;58:104015. doi: 10.1103/PhysRevD.58.104015. [DOI] [Google Scholar]
  • 18.Harada T, Maeda H. Convergence to a self-similar solution in general relativistic gravitational collapse. Phys. Rev. D. 2001;63:084022. doi: 10.1103/PhysRevD.63.084022. [DOI] [Google Scholar]
  • 19.Joshi PS, Dwivedi IH. The structure of naked singularity in self-similar gravitational collapse. Comm. Math. Phys. 1992;146(2):333–342. doi: 10.1007/BF02102631. [DOI] [Google Scholar]
  • 20.Larson RB. Numerical calculations of the dynamics of a collapsing protostar. Mon. Not. R. Astr. Soc. 1969;145:271–295. doi: 10.1093/mnras/145.3.271. [DOI] [Google Scholar]
  • 21.Lemaître, G., L’Univers en expansion. Ann. Soc. Sci. Bruxelles, A 53, 51 (1933)
  • 22.Merle F, Raphaël P, Rodnianski I, Szeftel J. On the implosion of a compressible fluid I: Smooth self-similar inviscid profiles. Ann. of Math. 2022;196:567–778. doi: 10.4007/annals.2022.196.2.3. [DOI] [Google Scholar]
  • 23.Misner CW, Sharp DH. Relativistic equations for adiabatic, spherically symmetric gravitational collapse. Phys. Rev. 1964;136:B571–B576. doi: 10.1103/PhysRev.136.B571. [DOI] [Google Scholar]
  • 24.Neilsen DW, Choptuik MW. Critical phenomena in perfect fluids. Classical Quantum Gravity. 2000;17:761–782. doi: 10.1088/0264-9381/17/4/303. [DOI] [Google Scholar]
  • 25.Oppenheimer JR, Snyder H. On continued gravitational contraction. Phys. Rev. 1939;56:455–459. doi: 10.1103/PhysRev.56.455. [DOI] [Google Scholar]
  • 26.Ori A, Piran T. Naked singularities in self-similar spherical gravitational collapse. Phys. Rev. Lett. 1987;59:2137. doi: 10.1103/PhysRevLett.59.2137. [DOI] [PubMed] [Google Scholar]
  • 27.Ori A, Piran T. Self-similar spherical gravitational collapse and the cosmic censorship hypothesis. Gen. Relativ. Grav. 1988;20:7. doi: 10.1007/BF00759251. [DOI] [Google Scholar]
  • 28.Ori A, Piran T. Naked singularities and other features of self-similar general-relativistic gravitational collapse. Phys. Rev. D. 1990;42:4. doi: 10.1103/PhysRevD.42.1068. [DOI] [PubMed] [Google Scholar]
  • 29.Ortiz, N., Sarbach, O., Zannias, T., On the behaviour of non-radial null geodesics in self-similar Tolman-Bondi collapse. J. Phys. Conf. Ser., 1208 no.1, 012010 (2019)
  • 30.Penrose R. Gravitational collapse: The role of general relativity. Riv. Nuovo Cim. 1969;1:252–276. [Google Scholar]
  • 31.Penston MV. Dynamics of self-gravitating gaseous spheres III. Mon. Not. R. Astr. Soc. 1969;144:425–448. doi: 10.1093/mnras/144.4.425. [DOI] [Google Scholar]
  • 32.Rodnianski, I., Shlapentokh-Rothman, Y., Naked Singularities for the Einstein Vacuum Equations: The Exterior Solution. preprint, arxiv:1912.08478 (2019)
  • 33.Stanyukovich, K. P., Sharshekeev, O., Gurovich, V. Ts., Automodel motion of a relativistic gas in general relativity if there is point symmetry. Dokl. Akad. Nauk SSSR, 165, 3, 510–513 (1965)
  • 34.Tolman, R.C.: Effect of inhomogeneity on cosmological models. Proc. Nat. Acad. Sci. U. S. 20, 169–176 (1934) [DOI] [PMC free article] [PubMed]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

All data generated or analysed during this study are included in this published article.


Articles from Annals of Pde are provided here courtesy of Springer

RESOURCES