Paper The following article is Open access

AdS Robin solitons and their stability

, and

Published 24 April 2020 © 2020 The Author(s). Published by IOP Publishing Ltd
, , Citation Piotr Bizoń et al 2020 Class. Quantum Grav. 37 105010 DOI 10.1088/1361-6382/ab7ee4

0264-9381/37/10/105010

Abstract

We consider the four-dimensional Einstein–Klein–Gordon–AdS system with the conformal mass subject to the Robin boundary conditions at infinity. Above a critical value of the Robin parameter, at which the AdS spacetime goes linearly unstable, we prove existence of a family of globally regular static solutions (that we call AdS Robin solitons) and discuss their properties.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

We consider the four-dimensional Einstein–Klein–Gordon–AdS system with mass μ related to the negative cosmological constant Λ through ${\mu }^{2}=\frac{2}{3}{\Lambda}$. For this, and only this, value of mass the system is conformally well-behaved at null and spatial infinity and consequently the initial-boundary value problem is well-posed for a variety of different boundary conditions at infinity [1, 2]. Here, we focus on the one-parameter family of Robin boundary conditions. It has been known that along this family there is a critical parameter value at which the system undergoes a bifurcation: the (zero energy) anti-de Sitter (AdS) spacetime becomes linearly unstable above that critical value [3] and there emerges a pair of (negative energy) globally regular static solutions (henceforth called AdS Robin solitons) [4]. The main goal of this paper is to establish the existence of AdS Robin solitons rigorously and analyze the structure of the bifurcation in more detail. In preparation of future analysis of the role of solitons in dynamics, we also determine their spectrum of linearized perturbations.

2. Setup

The Einstein–Klein–Gordon–AdS system is given by

Equation (1a)

Equation (1b)

where □g = gαβ α β is the wave operator associated with the metric gαβ , μ is the mass of the scalar field, and Λ is a negative constant. We assume spherical symmetry and write the metric in the form

Equation (2)

where $\left(t,x,\omega \right)\in \left(-\infty ,\infty \right){\times}\left[0,\pi /2\right){\times}{\mathbb{S}}^{2}$, dω2 is the round metric on ${\mathbb{S}}^{2}$ and 2 = −3/Λ. The metric functions A, δ and the scalar field ϕ depend on (t, x). We choose units such that = 1 and 4πG = 1 and introduce new variables

Equation (3)

Then the system (1) reduces to

Equation (4a)

Equation (4b)

Equation (4c)

Equation (4d)

where ${\Phi}={\left(\mathrm{cos}x\;{\partial }_{x}f-f\mathrm{sin}x\right)}^{2}+{A}^{-2}\;{\mathrm{e}}^{2\delta }\;{\mathrm{cos}}^{2}x\;{\left({\partial }_{t}f\right)}^{2}$ and

is the polar wave operator associated with the conformal metric ĝαβ = cos2 xgαβ . On the right side of equation (4a) the derivatives of metric functions were eliminated using equations (4b) and (4c).

In the following we set μ2 = −2. For this value of mass the wave equation (4a) is regular at x = π/2 because the first term on the right side (which is the only singular term) vanishes 3 . Thanks to this fact, the initial-boundary value problem for the system (4) is well posed for a variety of boundary conditions at the conformal boundary (both reflective and dissipative) [1, 2]. 4 In this paper we focus our attention on the one-parameter family of Robin boundary conditions

Equation (5)

where b is a constant (hereafter referred to as the Robin parameter). For b = 0 the Robin condition reduces to the Neumann condition ∂x f|x=π/2 = 0.

Assuming (5) and expanding the fields in power series in z = π/2 − x we obtain the following asymptotic behavior near z = 0

Equation (6)

Equation (7)

Equation (8)

where α(t) and δ(t) are free functions 5 and M is a constant. To see the physical meaning of M, let us define the renormalized mass function

Equation (9)

The first term on the right side is the Misner–Sharp mass function defined by mMS = r(1 + r2gμν μ rν r), where r = tan x is the areal radial coordinate. This function diverges as xπ/2 and the purpose of the second term (called the counterterm) is to cancel this divergence. The leading order behavior of the counterterm is determined by the asymptotics (6) and (7) but otherwise can be chosen freely. Using equation (4b) we get

Equation (10)

where

Equation (11)

hence

Equation (12)

This quantity can be interpreted as the bulk energy. From the asymptotic expansions (6) and (7) it follows that

Equation (13)

where the second term on the right side can be viewed as the energy stored on the boundary. Although both the bulk and boundary energies are time dependent, their sum M is conserved. In what follows, we will refer to M as the total energy (mass). The exchange of energy between the bulk and the boundary is a characteristic feature of systems subject to the Robin boundary conditions. Note that some of the bulk energy is 'lost' to the boundary if b < 0 and 'gained' from the boundary if b > 0.

We remark that the expression (13) can be obtained in a systematic way within the diffeomorphism covariant Hamiltonian framework of Wald and Zoupas [7] (see section 2.2 in [8]). Nonetheless, we believe that our hands-on approach, based solely on the analysis of the system (4), is helpful in getting insight into not so widely known physics of the Robin boundary conditions.

3. AdS Robin solitons

For time-independent solutions the system (4) with μ2 = −2 takes the form

Equation (14a)

Equation (14b)

Equation (14c)

where the derivatives of metric functions were eliminated from equation (14a) using equations (14b) and (14c). It is routine to prove that this system has local solutions near x = 0 which behave as follows

Equation (15)

where c is a free parameter.

Lemma. For any c the local solution (15) extends smoothly up to x = π/2 and fulfills the boundary conditions (6)–(8).

Proof. To prove this lemma it is convenient to use the radial coordinate r = tan x and return to the original field variables

Equation (16)

Then, equations (14a) and (14b) become

Equation (17a)

Equation (17b)

The local solutions (15) translate to

Equation (18)

We first observe that the function B = (A − 1)(1 + r2) is monotone increasing. To see this, suppose that B(r) has a maximum at some point r0 > 0. Differentiating equation (17b), substituting B'(r0) = 0, and eliminating ϕ''(r0) and B(r0) using equations (17a) and (17b), respectively, we get after simplifications

Equation (19)

which is manifestly positive, contradicting that the point r0 exists. Since B(r) is positive for small r > 0, this implies that A(r) ⩾ 1 for all r.

Next, we define a function

Equation (20)

Using the system (17) we obtain

Equation (21)

which is manifestly negative, hence H(r) is a monotonically decreasing Lyapunov function. Since A ⩾ 1, it follows that r2 ϕ'2 and ϕ2 remain bounded for all r.

To determine the asymptotic behavior of solutions for r it is convenient to use the logarithmic radial variable τ = log r. In terms of τ the system (17) is asymptotically autonomous for τ and the limiting autonomous system is

Equation (22a)

Equation (22b)

where dot denotes the derivative with respect to τ (by an abuse of notation, we use the same symbols for the original and limiting systems). From the general theory of asymptotically autonomous dynamical system [9, 10] and the existence of the Lyapunov function H, it follows that the asymptotic behavior of solutions of the system (17) for τ is governed by the above limiting system. Elementary analysis gives the attracting fixed point $\phi =0,\;\dot {\phi }=0,\;A=1$ with the leading order behavior

Equation (23)

where ck are free parameters, which are related to the free parameters α, b, and M in the expansions (6) and (7) by

Equation (24)

This completes the proof. □

The above lemma ensures that for each c the solution starting with the initial conditions (15) automatically satisfies the Robin condition f'(π/2) = bf(π/2) for some parameter b (which depends on c). We will refer to these globally regular static solutions as the AdS Robin solitons (or just solitons for short) and denote them by (fs , Bs , δs ). The profiles of solitons can be easily determined numerically by integrating the system (14) with the boundary conditions (15). We note in passing that an analogous reasoning leads to a two-parameter family of hairy black holes (where the second parameter is the horizon radius).

As far as we know, the AdS Robin solitons and hairy black holes were first studied in the literature in the context of so called 'designer gravity' [4, 11], however, to the best of our knowledge, their existence remained unproven.

4. Bifurcation analysis

It is illuminating to look at the solitons from the viewpoint of the local bifurcation theory. To this end, consider the perturbation expansion of solitons for small c

Equation (25)

Inserting this expansion into the system (14) and requiring regularity at x = 0, at the lowest order we get

Equation (26)

Calculations of higher orders are tedious and we relegate them to the appendix. In particular, in (A9) we obtain the explicit expression for the coefficient b2

Equation (27)

The fact that b2 is positive means that at b* we have a supercritical pitchfork bifurcation where the AdS solution bifurcates into a pair of solitons (±fs , Bs , δs ). As usual, this kind of bifurcation is associated with exchange of linear stability and, indeed, in the next section we will show that for b > b* the AdS space becomes linearly unstable whereas the solitons are linearly stable.

Using (7) and the expansion (25), we get the approximation for the mass

which upon substitution of (A4) and (A8) yields

Equation (28)

It is instructive to rederive this result along the lines of designer gravity [4]. Letting α = f(π/2) and β = f'(π/2), we get from (25) (in this paragraph ' = ' means equality up to order $\mathcal{O}\left({c}^{4}\right)$)

Equation (29)

which can be viewed as the parametric equation of the curve in the (α, β) plane. Eliminating c we get the function

Equation (30)

where the subscript 's' indicates that the function is associated with solitons. Following the approach used in designer gravity we introduce the effective potential

Equation (31)

By construction, critical points of the effective potential correspond to solitons. The key observation, made by Hertog and Horowitz in [4], is that the value of the effective potential at the critical point is equal to the soliton mass. In our case, ${\mathcal{V}}^{\prime }\left(\alpha \right)=0$ for ${\alpha }_{\text{s}}^{2}=\frac{{\pi }^{2}}{4{b}_{2}}\left(b-{b}_{{\ast}}\right)$ (and, of course, for α = 0 corresponding to the AdS space). Substituting this into (31) we get ${M}_{\text{s}}=\mathcal{V}\left({\alpha }_{s}\right)$ which reproduces the formula (28). Note that ${\mathcal{V}}^{{\prime\prime}}\left({\alpha }_{\text{s}}\right){ >}0$.

Further from the bifurcation point, the soliton function βs(α) and the corresponding effective potential $\mathcal{V}\left(\alpha \right)$ can be determined numerically 6 . We find that for each b > b* the effective potential has the shape of a Mexican hat (see figure 1) with exactly three critical points: the local maximum at zero and two global minima at ±αs. This implies that the soliton solution is unique (modulo reflection symmetry) and suggests that it is stable.

Figure 1.

Figure 1. Effective potentials for sample values of b.

Standard image High-resolution image

Remark. It is natural to expect that for any given b > b* the soliton is the ground state, i.e. for any regular initial data satisfying the Robin condition (5) there holds the inequality MMs which saturates if and only if the data correspond to the soliton [4]. Our numerical constructions of initial data corroborate this conjecture but we have not been able to prove it (see [8] for partial results in this direction).

5. Linear stability analysis

Linearizing the system (4) around the AdS solution (f = B = δ = 0) and separating time f(t, x) = eiωt v(x) we get the eigenvalue problem 7

Equation (32)

The operator L (which is just the polar conformal Laplacian on the 3-sphere) is symmetric on the Hilbert space ${L}^{2}\left(\left[0,\pi /2\right],\;{\mathrm{sin}}^{2}\;x\;\;\mathrm{d}x\right)$ and the Robin boundary condition

Equation (33)

provides a one-parameter family of its self-adjoint extensions.

For ω2 > 0 the regular solution of (32) is

Equation (34)

Imposing the Robin condition (33) we obtain the quantization condition for the eigenfrequencies

Equation (35)

From the graphical analysis shown in figure 2(a) we see that for each non-negative integer n there is exactly one eigenfrequency ωn such that

For large n the quantization condition (35) gives the asymptotically resonant spectrum

Equation (36)

Figure 2.

Figure 2. Graphical solutions of the quantization conditions (35) and (38).

Standard image High-resolution image

The lowest eigenvalue ${\omega }_{0}^{2}$ vanishes at b = b* = 2/π; the corresponding eigenfunction is the linearized static solution f1 given in (26). An elementary perturbative calculation gives near b*

Equation (37)

For general b > b* there is an exponentially growing mode ${\mathrm{e}}^{{\lambda }_{0}t}{v}_{0}\left(x\right)$, where the exponent ${\lambda }_{0}=\sqrt{-{\omega }_{0}^{2}}$ is given by the unique positive root of the equation (see figure 2(b))

Equation (38)

and the corresponding eigenfunction is v0(X) = sinh(λ0 x)/sin x.

Next, we look at the linear stability of solitons. Linearizing the system (4) around the soliton and separating time, we get the eigenvalue problem

Equation (39)

where

Equation (40)

and

Equation (41)

For b slightly above b* (i.e. for small c), we have

Equation (42)

where the operator P can be calculated using the expansions (25). To calculate the perturbations of eigenvalues we assume the following ansatz

Equation (43)

where ${{\omega }_{n}^{{\ast}}}^{2}$ and ${v}_{n}^{{\ast}}$ are the eigenvalues and normalized eigenfunctions of the operator L at the bifurcation point and b2 is given in (27). Substituting this ansatz into the Robin boundary condition we get at the first and third order in c

Equation (44)

Substituting the ansatz (43) into (39), we get at the third order in c

Equation (45)

Projecting this equation on ${v}_{n}^{{\ast}}$ and noting, via (44), that

Equation (46)

we obtain the leading order approximation for the eigenvalues

Equation (47)

In particular, for the lowest eigenvalue we obtain

Equation (48)

where

Equation (49)

The positivity of γ0 confirms the expectation that the solitons are linearly stable near the bifurcation point. Solving the eigenvalue problem numerically, we find that the eigenvalues ${\tilde{\omega }}_{n}^{2}$ grow monotonically with c. The numerical values of the first few eigenfrequencies (as measured by the central observer) for a small parameter c = 0.1 (corresponding to b ≈ 0.6467) are displayed in table 1.

Table 1.  The first six eigenfrequencies of linear perturbations around the soliton for the parameter c = 0.1. In the second row the approximate eigenfrequencies given by (47) are shown for comparison.

n 0 1 2 3 4 5
${\tilde{\omega }}_{n}$ 0.197 35 2.870 65 4.930 28 6.957 14 8.973 63 10.985 49
${\tilde{\omega }}_{n}^{\text{pert}}$ 0.196 42 2.870 62 4.930 25 6.957 11 8.973 60 10.985 46

From the leading order WKB approximation [12] it follows that for large n

Equation (50)

which compares well with numerical results even if n is not very large.

6. Discussion

The Einstein–Klein–Gordon–AdS system with mass ${\mu }^{2}=\frac{2}{3}{\Lambda}{< }0$ is well-behaved at the conformal boundary which makes it a good toy model for studying the role of boundary conditions in dynamics of asymptotically AdS spacetimes [13]. In this paper we focused on the Robin boundary conditions and proved existence of a one-parameter family of solitons for b > b*. We also demonstrated that the linearized perturbations around these solitons have no growing modes. A natural question is: are the AdS Robin solitons nonlinearly stable? Numerical simulations, to be reported in [14], indicate a positive answer and provide evidence for existence of plethora of time-periodic and quasiperiodic solutions, not only in the perturbative regime (which is expected in view of the non-resonant spectrum) but also, somewhat surprisingly, for large perturbations.

Of course, the analogous question of nonlinear stability arises for the AdS spacetime for b < b*. However here, in contrast to the Dirichlet case [6], the numerical simulations are as yet not conclusive and we leave this question to future investigations.

Acknowledgments

We thank Piotr Chruściel, Oleg Evnin, Helmut Friedrich and Arthur Wasserman for helpful remarks. This work was supported in part by the Polish National Science Centre Grant No. 2017/26/A/ST2/00530. PB and MM acknowledge the support of the Alexander von Humboldt Foundation.

: Appendix

Here we give the details of the higher orders of perturbation expansion (25). At the third order equation (14a) becomes

Equation (A1)

Substituting f1 and B2, given in (26), into the right hand side and imposing f3(0) = 0, we find

Equation (A2)

where ζ is the Riemann zeta function and we defined the functions

Equation (A3)

From (A2) we read off

Equation (A4)

Equation (A5)

At the fourth order equation (14b) becomes

Equation (A6)

Substituting (26) and (A2) into the right hand side and requiring regularity at x = 0, we find

Equation (A7)

from which we read off

Equation (A8)

Using (A4) and (A5) and imposing the Robin condition in the expansion (25), we get

Equation (A9)

Footnotes

  • This cancellation is due to the fact that for μ2 = −2 the left sides of equations (1b) and (4a) are asymptotically conformal, that is (□g + 2)ϕ ≈ (cos x)−3(□ĝ − 1)f near x = π/2. However, the constraint equation (4b) has a singularity at x = π/2 so it does not appear possible to extend the solutions 'beyond infinity' (cf [5] where an extension of solutions across the conformal boundary at timelike infinity was analyzed for the system (1) with ${\mu }^{2}=\frac{2}{3}{\Lambda}{ >}0$).

  • This should be contrasted with the widely studied massless case for which only the Dirichlet boundary condition is compatible with the basic requirement of finite total mass [6].

  • We use the normalization δ(t, 0) = 0, hence t is the proper time at the center.

  • For large values of c, there develops a boundary layer near x = π/2 with exponentially shrinking width. Using the method of matched asymptotics one can show that both α and b grow as ${\mathrm{e}}^{{c}^{2}}$ for c which makes the numerics (in compactified variable x) cumbersome.

  • The eigenvalue problem (32) is a particularly simple case of the master eigenvalue problem for linear perturbations of AdS space that was solved by Ishibashi and Wald in full generality using the properties of hypergeometric functions [3]. For the reader's convenience we reproduce their results in our special case using more elementary tools.

Please wait… references are loading.
10.1088/1361-6382/ab7ee4