This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

aainstitutetext: Faculty of Science, Theoretical Physics, Memorial University of Newfoundland, St. John’s, NL A1C 5S7, Canada bbinstitutetext: Departament de Física Quàntica i Astrofísica, Institut de Ciències del Cosmos, Universitat de Barcelona, Martí i Franquès 1, E-08028 Barcelona, Spain ccinstitutetext: Department of Mathematic and Statistics and Department of Physics and Astronomy, McMaster University, Hamilton, ON, Canada, L8S 4M1 ddinstitutetext: Department of Physics and Astronomy, University of Waterloo, Waterloo, ON, Canada, N2L 3G1

Phase Transitions and Stability of Eguchi-Hanson-AdS Solitons

Turkuler Durgut b    Robie A. Hennigar c    Hari K. Kunduri d    and Robert B. Mann [email protected] [email protected] [email protected] [email protected]
Abstract

The Eguchi-Hanson-AdS5 family of spacetimes are a class of static, geodesically complete asymptotically locally AdS5 soliton solutions of the vacuum Einstein equations with negative cosmological constant. They have negative mass and are parameterized by an integer p3p\geq 3 with a conformal boundary with spatial topology L(p,1)L(p,1). We investigate mode solutions of the scalar wave equation on this background and show that, similar to AdS5, the geometry admits a normal mode spectrum (i.e. solutions that neither grow or decay in time). In addition, we also discuss other geometric properties of these soliton spacetimes, including the behaviour of causal geodesics and their thermodynamic properties. We also point out a surprising connection with the AdS soliton.

1 Introduction

A classic theorem of Wang MR2156971 states that Anti-de Sitter (AdS) spacetime is the unique conformally compact (globally) static solution to the vacuum Einstein equations Gab+Λgab=0G_{ab}+\Lambda g_{ab}=0 where Λ<0\Lambda<0 and with spherical conformal spatial boundary. The theorem holds in situations where the positive mass theorem for asymptotically hyperbolic manifolds is valid (see also MR771446 ; Qing_2004 ). This result lends support to the expectation that AdS spacetime is the appropriate ‘ground state’ amongst the class of all solutions with the same asymptotic behaviour. A putative ground state would, in turn, be expected to be dynamically stable, and indeed studies of the wave equation on a fixed AdS background are consistent with this intuition. However, the remarkable result of Bizon et al. Bizon:2011gg demonstrates that, in fact, AdS is nonlinearly unstable under arbitrarily small perturbations whose endpoint to the formation of black holes - that is, small amounts of energy tend to concentrate at shorter and shorter scales, rather than dissipating as in Minkowski spacetime Bizon:2011gg ; Bizon:2015pfa .

A natural question is whether the vacuum Einstein equations with negative cosmological constant admit other static solutions that are asymptotically locally AdS but with a different conformal boundary. The one-parameter family of AdS soliton spacetimes provides such an example with toroidal conformal spatial boundary Galloway:2001uv ; Galloway:2002ai . Clarkson and Mann considered the problem of finding static solutions asymptotic to a freely acting discrete quotient of AdS Clarkson:2006zk ; Clarkson:2005qx . They succeeded in constructing solutions in odd dimensions (referred to as Eguchi-Hanson-AdS spacetimes) that are asymptotic to AdS/d+1p{}_{d+1}/\mathbb{Z}_{p}. These spacetimes have negative energy relative to that of pure AdS. In five spacetime dimensions, p3p\geq 3 (see below) and spatial cross sections of the conformal boundary are lens spaces L(p,1)L(p,1) equipped with the standard round metric. Clarkson and Mann conjectured that these metrics are the states of lowest energy in their asymptotic class Clarkson:2006zk .

The Eguchi-Hanson-AdS5 geometry, in addition to being static, has a local SU(2)×U(1)SU(2)\times U(1) isometry group, which acts with three-dimensional orbits. Hence its spatial sections belong to the biaxial Bianchi IX class of geometries. Dold exploited this symmetry to study the evolution of initial data within this symmetry class Dold:2017hwr . In addition to showing that the resulting system of equations forms a well-posed initial-boundary value problem (with the fields satisfying an appropriate Dirichlet condition at conformal infinity), he rigorously proved that the maximal development of this restricted class of initial data sufficiently close to Eguchi-Hanson-AdS5 data cannot form a horizon in the future. Assuming that Eguchi-Hanson-AdS5 is indeed the only static solution within its conformal class, this implies that the endpoint of the evolution must generically be a spacetime containing a naked singularity.

In the present article, we will take a different perspective and study mode solutions to the massless Klein-Gordon equation

gΦ=0\Box_{g}\Phi=0 (1)

on the fixed Eguchi-Hanson-AdS5 background. One advantage of this approach is that we do not need to make any special symmetry restrictions on Φ\Phi. It easy to see that (1) is separable and so it is relatively straightforward to reduce the problem to a single radial Schödinger-type equation. Since the background is static, it is straightforward to show that there is a conserved energy and hence a uniform bound for the energy associated with the field Φ\Phi in terms of its initial energy. This kind of bound, however, does not tell us if the field is being concentrated within a compact region as a result of some geometric mechanism (e.g. trapping).

For simplicity we will study quasinormal mode solutions of (1). We will show that, similar to AdS spacetime, Eguchi-Hanson-AdS5 admits normal mode solutions (i.e. they neither grow nor decay in time). Our results will be based on a robust numerical approach as well as analytic methods.

In addition to the Klein-Gordon test field, we consider many other aspects of these solutions that have not been addressed to date in the literature. We begin with an analysis of the mechanical properties of these solitons. We compute their mass and show that these solutions have a non-trivial thermodynamic volume of topological origin Kastor:2009wy ; Andrews:2019hvq . Examining the thermodynamics in the canonical ensemble, we show that there is an analog of the Hawking-Page phase transition Hawking:1982dh . The two relevant states in the phase transition are the Eguchi-Hanson-AdS5 soliton and the black hole resulting from performing p\mathbb{Z}_{p} identifications to the spherical AdS black hole. We then study the geodesics in the spacetime, which is relevant for two reasons. Firstly, we seek to find whether or not there exists stable trapping of null geodesics (the confinement of null geodesics to a compact subregion of space) in the Eguchi-Hanson-AdS5 spacetime. Stable trapping presents an obstruction to proving strong decay statements for solutions of the wave equation Holzegel:2013kna ; Benomio:2018ivy . For example, decay might be no faster than inversely logarithmic in time, rather than inverse polynomial Keir:2016azt . The latter is expected to be necessary if there is any hope of demonstrating nonlinear stability. In particular, stable trapping has been shown in several examples of families of horizonless soliton spacetimes, which typically have some nontrivial spatial topology (e.g., two-cycles or ‘bubbles’) Keir:2016azt ; Andrews:2019hvq ; Gunasekaran:2020pue or ultracompact objects Keir:2014oka . However, we find that stable trapping is absent in the Eguchi-Hanson-AdS5 spacetime. Secondly, we investigate the light-crossing time of the geometry and find that it turns out to be relevant to understanding the spacing between overtones for the normal mode solutions of the Klein-Gordon equation.

A recurring theme throughout each aspect of our work is a connection between the Eguchi-Hanson-AdS5 soliton and the AdS soliton that has not been pointed out in the literature. Namely, starting with the former with spatial boundary metric (the round L(p,1)L(p,1) lens space), we find that formally taking the pp\to\infty limit gives the AdS soliton as the limiting solution. As such, we show how the relevant quantities of the AdS soliton govern the asymptotics of the corresponding quantities for the Eguchi-Hanson soliton.

The outline of our paper is as follows. In section 2, we review the basic structure of the Eguchi-Hanson soliton and show its limit is AdS soliton as pp\to\infty. In section 3, we analyze the mass and thermodynamic behaviour of the soliton, and in section 4, we consider the behaviour of timelike and null geodesics in this spacetime. We find that both massive and null particles oscillate between the edge of the soliton and infinity, with no stable trapping regions. In section 5, we proceed with the main purpose of our paper, that of analyzing the scalar wave equation in the Eguchi-Hanson soliton spacetime. As an analytic solution is apparently intractable, we solve the equation numerically, checking our results against various approximations in certain limits. Amongst our most intriguing results is that the normal modes interpolate between those of a scalar wave on the orbifold AdS/52{}_{5}/\mathbb{Z}_{2} and on the AdS soliton as the parameter pp is varied. We close our paper with some concluding remarks in 6. Several appendices contain details showing how we arrived at our results.

2 Metric and Structure

Clarkson and Mann Clarkson:2006zk obtained a solution of the D=5D=5 Einstein equations with negative cosmological constant, namely

Rab=42gab,R_{ab}=-\frac{4}{\ell^{2}}g_{ab}\;, (2)

given by

ds2\displaystyle ds^{2} =g(r)dt2+r2f(r)[dψ+cos(θ)2dϕ]2+dr2f(r)g(r)+r24dΩ22\displaystyle=-g(r)dt^{2}+r^{2}f(r)\left[d\psi+\frac{\cos(\theta)}{2}d\phi\right]^{2}+\frac{dr^{2}}{f(r)g(r)}+\frac{r^{2}}{4}d\Omega^{2}_{2} (3)
g(r)\displaystyle g(r) =1+r22andf(r)=1a4r4.\displaystyle=1+\frac{r^{2}}{\ell^{2}}\quad\quad\text{and}\quad\quad f(r)=1-\frac{a^{4}}{r^{4}}. (4)

This is a cohomogeneity-one metric with local isometry group ×SU(2)×U(1)\mathbb{R}\times SU(2)\times U(1). When a=0a=0 this reduces to the AdS5 metric with spherical boundary when tt\in\mathbb{R}, r>0r>0, ψ(0,2π)\psi\in(0,2\pi), θ(0,π)\theta\in(0,\pi) and ϕ(0,2π)\phi\in(0,2\pi) with a standard apparent singularity at r=0r=0, where the (ψ,θ,ϕ)(\psi,\theta,\phi) part of the metric degenerates, representing the origin of coordinates. However, for a0a\neq 0 (we fix a>0a>0 without loss of generality) the metric extends globally to a manifold with non-trivial topology, provided that certain regularity conditions are satisfied. The Killing vector field /ψ\partial/\partial\psi becomes degenerate at r=ar=a; examining the (r,ψ)(r,\psi) sector of the geometry, absence of conical singularities requires the identification

ψψ+2π2g(a).\psi\sim\psi+\frac{2\pi}{2\sqrt{g(a)}}\,. (5)

The ensures that the geometry smoothly ‘pinches off’ leaving a round S2S^{2} of radius a/2a/2. This condition must be combined with the independent condition, arising from regularity of the constant (t,r)(t,r) surfaces, that demands

ψψ+2πp,\psi\sim\psi+\frac{2\pi}{p}\,, (6)

where pp\in\mathbb{Z} (this ensures the geometry is that of L(p,1)L(p,1)). Satisfying both conditions requires that

a2=(p241)2.a^{2}=\left(\frac{p^{2}}{4}-1\right)\ell^{2}\,. (7)

Thus regularity requires that we must have p3p\geq 3. This means that asymptotically the boundary metric is a lens space L(p,1)L(p,1) (p=1p=1 would be S3S^{3}). Thus, we have a gravitational soliton (a geodesically complete, strictly stationary solution) that has a 2-cycle in the interior region and has a lens space as its boundary. The above metric is often referred to as the ‘Eguchi-Hanson-AdS5’ as constant time hypersurfaces generalize the well-known four-dimensional Eguchi-Hanson gravitational instanton metric which must have p=2p=2 Eguchi:1978xp .

2.1 The Large pp Limit of the Metric: AdS Soliton

As we have just seen, the Eguchi-Hanson-AdS5 soliton is characterized by a single integer pp. For several reasons, it will be fruitful to consider these solutions for large values of pp.

Beginning with the Eguchi-Hanson-AdS5 metric (3), we perform the following transformations

t\displaystyle t =2τp,r=az,θ=4ρp,φ=ψp2ϕ,\displaystyle=\frac{2\tau}{p}\,,\quad r=az\,,\quad\theta=\frac{4\rho}{p}\,,\quad\varphi=\psi-\frac{p}{2}\phi\,, (8)

and then take the limit pp\to\infty. The result is

ds2=z2dτ2+2dz2z2f(z)+2z2f(z)4dφ2+2z2[dρ2+ρ2dϕ2]+𝒪(1p)ds^{2}=-z^{2}d\tau^{2}+\frac{\ell^{2}dz^{2}}{z^{2}f(z)}+\frac{\ell^{2}z^{2}f(z)}{4}d\varphi^{2}+\ell^{2}z^{2}\left[d\rho^{2}+\rho^{2}d\phi^{2}\right]+\mathcal{O}\left(\frac{1}{p}\right) (9)

where

f(z)=11z4.f(z)=1-\frac{1}{z^{4}}\,. (10)

One can then convert the polar coordinates on the 2\mathbb{R}^{2} to standard Cartesian coordinates, giving

ds2=z2dτ2+2dz2z2f(z)+2z2f(z)4dφ2+2z2[dx2+dy2]+𝒪(1p).ds^{2}=-z^{2}d\tau^{2}+\frac{\ell^{2}dz^{2}}{z^{2}f(z)}+\frac{\ell^{2}z^{2}f(z)}{4}d\varphi^{2}+\ell^{2}z^{2}\left[dx^{2}+dy^{2}\right]+\mathcal{O}\left(\frac{1}{p}\right)\,. (11)

In the strict pp\to\infty limit, this is an AdS5 soliton belonging to the class first reported in Horowitz:1998ha , in coordinates such that the location of the bubble is at z=1z=1. One can easily check that the coordinate φ\varphi is periodic with period 2π2\pi. Strictly speaking, there are several topologically distinct solitons that can be obtained from the same local metric (11) depending on identifications performed on the auxiliary flat directions (x,y)(x,y) — see, e.g., Page:2002qc . In this case, the soliton corresponding to the large pp limit of Eguchi-Hanson-AdS has no identifications on these coordinates, i.e. the spatial part of the boundary metric is 𝕊1×2\mathbb{S}^{1}\times\mathbb{R}^{2}. This is the same configuration first considered in Horowitz:1998ha , and henceforth we will refer to this case as “the” AdS soliton. To the best of our knowledge, this connection between the Eguchi-Hanson-AdS5 soliton and the AdS soliton has not been previously reported on111The connection between the Eguchi-Hanson and AdS soliton geometries could be inferred from the results for lensed CFT partition functions Shaghoulian:2016gol . We are grateful to Edgar Shaghoulian who, after this work was completed, brought this reference to our attention. .

Regularity of the solution requires that p3p\geq 3 is an integer. Therefore, the pp\to\infty limit may be most cautiously considered as a ‘formal’ limit. Nonetheless, it is difficult to overstate the utility of this result. As we will see in the subsequent sections, many of the quantities of interest cannot be evaluated exactly for the Eguchi-Hanson-AdS5 soliton, but the asymptotics of these quantities can be effectively captured by the corresponding quantities for the AdS soliton. Said another way, Eguchi-Hanson-AdS5 solitons for large values of pp behave in a manner similar to the AdS soliton.

3 Soliton Mechanics

3.1 Smarr Relation & First Law

While it is well-known that black holes satisfy a first law and Smarr relation, similar relationships can be found for solitons and soliton-black hole configurations. This was rigorously demonstrated in the asymptotically flat case in Kunduri:2013vka , and extended to a particular example of an asymptotically globally AdS soliton in Andrews:2019hvq . Here we apply these considerations to the Eguchi-Hanson-AdS5 soliton. While the mass was calculated in the original manuscript Clarkson:2006zk , the notion of thermodynamic volume Kastor:2009wy — which proves crucial for deriving the Smarr relation and first law in this case — was at that point not developed. We note also that considerations of extended thermodynamics have been previously carried out for Eguchi-Hanson-dS soliton in Mbarek:2016mep .

Let ξ=t\xi=\partial_{t} be the stationary Killing field. It has zero divergence so dξ=0d\star\xi=0, where \star is the Hodge dual. It follows that one can write the closed four-form ξ=dϖ\star\xi=-d\star\varpi for some locally defined 2-form ϖ\varpi, or equivalently

ξ=dϖ.\xi=\star d\star\varpi. (12)

On the other hand a basic identity is

ddξ=2Ric(ξ)=82dϖd\star d\xi=2\star\text{Ric}(\xi)=\frac{8}{\ell^{2}}d\star\varpi (13)

where the second equality follows from the Einstein equation Rab=42gabR_{ab}=-\tfrac{4}{\ell^{2}}g_{ab} and (12). This means we have the conservation equation

d[dξ82ϖ]=0d\star\left[d\xi-\frac{8}{\ell^{2}}\varpi\right]=0 (14)

which we will integrate over a spatial hypersurface t=t=constant. If we introduce the basis

e0\displaystyle e^{0} =gdt,e1=drfg,e2=rf(dψ+cosθ2dϕ),\displaystyle=\sqrt{g}dt,\qquad e^{1}=\frac{dr}{\sqrt{fg}},\qquad e^{2}=r\sqrt{f}\left(d\psi+\frac{\cos\theta}{2}d\phi\right), (15)
e3\displaystyle e^{3} =r2dθ,e4=r2sinθdϕ\displaystyle=\frac{r}{2}d\theta,\qquad e^{4}=\frac{r}{2}\sin\theta d\phi

and assume ϖ\varpi takes the form

ϖ=A(r)e0e1+B(r)e1e2\varpi=A(r)e^{0}\wedge e^{1}+B(r)e^{1}\wedge e^{2} (16)

then a calculation gives

A(r)=1f(r4+C1r3),B(r)=C2r2gA(r)=\frac{1}{\sqrt{f}}\left(\frac{r}{4}+\frac{C_{1}}{r^{3}}\right),\qquad B(r)=\frac{C_{2}}{r^{2}\sqrt{g}} (17)

so that

ϖ=1f(r4+C1r3)dtdr+C2rgdr(dψ+cosθ2dϕ).\varpi=\frac{1}{f}\left(\frac{r}{4}+\frac{C_{1}}{r^{3}}\right)dt\wedge dr+\frac{C_{2}}{rg}dr\wedge\left(d\psi+\frac{\cos\theta}{2}d\phi\right). (18)

whose Hodge dual is

ϖ=14(r44+C1)sinθdψdθdϕ+C24sinθdtdθdϕ.\star\varpi=-\frac{1}{4}\left(\frac{r^{4}}{4}+C_{1}\right)\sin\theta d\psi\wedge d\theta\wedge d\phi+\frac{C_{2}}{4}\sin\theta dt\wedge d\theta\wedge d\phi. (19)

Note that the degeneracy of /ψ\partial/\partial\psi at the ‘centre’ r=ar=a implies that ϖ\star\varpi is not well defined there unless C1C_{1} is chosen to be C1=a4/4C_{1}=-a^{4}/4. However, as is typical — and as we will see below — regularity is not the correct prescription for fixing the parameter C1C_{1}.

Next, we integrate the closed form defined by (14) over a hypersurface Σ\Sigma defined as a surface of constant time, t=t= constant, over the region R0rR_{0}\leq r\leq\infty with R0>aR_{0}>a. On this region ϖ\varpi is well-defined and we can apply Stokes’ theorem. The identity gives

0=Σd[(dξ82ϖ)]=Σ(dξ82ϖ)ΣR0(dξ82ϖ)0=\int_{\Sigma}d\left[\star\left(d\xi-\frac{8}{\ell^{2}}\varpi\right)\right]=\int_{\partial_{\infty}\Sigma}\star\left(d\xi-\frac{8}{\ell^{2}}\varpi\right)-\int_{\partial\Sigma_{R_{0}}}\star\left(d\xi-\frac{8}{\ell^{2}}\varpi\right) (20)

where Σ\partial_{\infty}\Sigma and ΣR0\partial\Sigma_{R_{0}} represent the asymptotic and inner boundaries respectively.

Let us first focus on the contribution at conformal infinity. A calculation shows that as rr\to\infty,

dξ=(r422+a422+O(1/r2))sinθdψdθdϕ,\star d\xi=\left(-\frac{r^{4}}{2\ell^{2}}+\frac{a^{4}}{2\ell^{2}}+O(1/r^{2})\right)\sin\theta d\psi\wedge d\theta\wedge d\phi\,, (21)

and we note that the divergent term is precisely cancelled by the corresponding divergent term of ϖ\star\varpi when the two terms are combined as in (20). We identify a renormalized Komar mass as Kastor:2009wy

MKomar:=332πΣ(dξ82ϖ)=3π8p2(a4+4C1).M_{\rm Komar}:=-\frac{3}{32\pi}\int_{\partial_{\infty}\Sigma}\star\left(d\xi-\frac{8}{\ell^{2}}\varpi\right)=-\frac{3\pi}{8p\ell^{2}}\left(a^{4}+4C_{1}\right)\,. (22)

The fact that the free parameter C1C_{1} appears in the Komar mass can be understood as an ambiguity in the ground state energy. We will now fix this ambiguity.

We can ensure the integral over the asymptotic boundary evaluates to the mass by choosing C1C_{1} appropriately. To calculate this we use the Ashtekar-Magnon procedure Ashtekar:1984zz which is well-defined in this setting. The relevant component of the Weyl tensor is

Crtrt=a4(r2+2)(r4a4)=a4r6+O(r8)C^{t}_{~{}rtr}=-\frac{a^{4}}{(r^{2}+\ell^{2})(r^{4}-a^{4})}=-\frac{a^{4}}{r^{6}}+O(r^{-8}) (23)

as rr\to\infty. Setting Ω=/r\Omega=\ell/r and defining the conformal metric g¯ab=Ω2gab\bar{g}_{ab}=\Omega^{2}g_{ab} with Ω=0\Omega=0 as rr\to\infty, the Ashtekar-Magnon mass is then defined as

Q[t]=16πM¯ba(t)bdStQ[\partial_{t}]=\frac{\ell}{16\pi}\int_{\partial M}\bar{\mathcal{E}}^{a}_{~{}b}(\partial_{t})^{b}dS_{t} (24)

where dStdS_{t} is a constant time slice of the conformal boundary, which has a round lens space metric of radius \ell. The quantity ¯ba\bar{\mathcal{E}}^{a}_{~{}b} is the electric part of the Weyl tensor

¯ba=2Ω2g¯cdg¯efndnfCcbea,\bar{\mathcal{E}}^{a}_{~{}b}=\frac{\ell^{2}}{\Omega^{2}}\bar{g}^{cd}\bar{g}^{ef}n_{d}n_{f}C^{a}_{~{}cbe}, (25)

with unit spacelike normal n=dΩn=d\Omega.

Noting that g¯rr=r42\bar{g}^{rr}=r^{4}\ell^{2}, it is then a straightforward matter to obtain the relevant component

¯tt=r66Crtrt=a46\bar{\mathcal{E}}^{t}_{~{}t}=\frac{r^{6}}{\ell^{6}}C^{t}_{rtr}=-\frac{a^{4}}{\ell^{6}} (26)

yielding

M:=Q[t]=16πM(a46)3sinθ4𝑑ψ𝑑θ𝑑ϕ=πa482pM:=Q[\partial_{t}]=\frac{\ell}{16\pi}\int_{\partial M}\left(-\frac{a^{4}}{\ell^{6}}\right)\ell^{3}\frac{\sin\theta}{4}d\psi d\theta d\phi=-\frac{\pi a^{4}}{8\ell^{2}p} (27)

so the mass is negative, a fact already observed in Clarkson:2006zk — c. f. eq.(10) of that work. Comparing the above with the result of the Komar integration implies C1=a4/6C_{1}=-a^{4}/6.

Once the Komar mass has been computed, the thermodynamic volume is identified by evaluating the integral of the Killing potential over the inner boundary, and taking the limit R0aR_{0}\to a. This gives:

V=ΣRϖ=π22p(a4+4C1)=π2a46p.V=-\int_{\partial\Sigma_{R}}\star\varpi=\frac{\pi^{2}}{2p}\left(a^{4}+4C_{1}\right)=\frac{\pi^{2}a^{4}}{6p}\,. (28)

using the choice C1=a4/6C_{1}=-a^{4}/6. For some black hole solutions the thermodynamic volume can be interpreted as the volume of a Euclidean ball of radius aa that is removed from the spatial hypersurface Kubiznak:2016qmn . This interpretation is not available in this case because there is no ‘ball’ in Euclidean space for which a lens space is its boundary. In this case, the thermodynamic volume VV has a topological origin: it arises not due to the presence of an horizon, but instead because the choice of constant C1C_{1} that leads to the correct mass leads to a Killing potential that is not regular at the location of the bubble.

Noting that thermodynamic pressure is Kubiznak:2016qmn

P:=Λ8π=34π2P:=-\frac{\Lambda}{8\pi}=\frac{3}{4\pi\ell^{2}} (29)

then we have

M=PVM=-PV (30)

which is the Smarr formula for this system. Using the regularity condition (6) and the mass (27) we have

dM=π4p(p241)2ddM=-\frac{\pi}{4p}\left(\frac{p^{2}}{4}-1\right)^{2}\ell d\ell (31)

Alternatively, using (28), (29), and the regularity condition (6), we have

VdP=π4p(p241)2dVdP=-\frac{\pi}{4p}\left(\frac{p^{2}}{4}-1\right)^{2}\ell d\ell (32)

and consequently

dM=VdPdM=VdP (33)

which is the expression of the first law for the Eguchi-Hanson-AdS5 soliton.

The fact that the regularity condition is required for the validity of the first law is consistent with previous studies of (extended) mechanics of smooth geometries Kunduri:2013vka ; Gunasekaran:2016nep ; Andrews:2019hvq . It is worth remarking that in some cases, e.g., for accelerating black holes or spacetimes containing Misner strings, it is possible to formulate the Smarr relation and first law without requiring the regularity condition to hold Appels:2016uha ; Bordo:2019tyh . It may be interesting to better understand when, exactly, regularity of the geometry is crucial for formulating a sensible first law and Smarr relation.

3.2 Euclidean action

By sending tiτt\to i\tau the Eguchi-Hanson-AdS5 solution (3) may be analytically continued to produce a Riemannian (positive signature) Einstein metric:

ds2=g(r)dτ2+r2f(r)[dψ+cosθ2dϕ]2+dr2f(r)g(r)+r24dΩ22.ds^{2}=g(r)d\tau^{2}+r^{2}f(r)\left[d\psi+\frac{\cos\theta}{2}d\phi\right]^{2}+\frac{dr^{2}}{f(r)g(r)}+\frac{r^{2}}{4}d\Omega^{2}_{2}. (34)

The geometry is smooth and complete with an S2S^{2}-bolt at r=ar=a provided the regularity condition (7) is imposed. We now periodically identify the τ\tau-coordinate as ττ+β\tau\sim\tau+\beta so that it parameterizes an S1S^{1}. The vector field τ\partial_{\tau} is nowhere vanishing since g(r)>0g(r)>0 and therefore this S1S^{1} does not degenerate. In particular there is no condition on β\beta. The underlying manifold will therefore be S1×TS2S^{1}\times T^{*}S^{2} (the latter factor being the cotangent bundle of S2S^{2}). The metric is conformally compact with conformal boundary S1×L(p,1)S^{1}\times L(p,1) equipped with the conformal boundary metric

γ=dτ2+2[(dψ+cosθ2dϕ)2+dΩ224].\gamma=d\tau^{2}+\ell^{2}\left[\left(d\psi+\frac{\cos\theta}{2}d\phi\right)^{2}+\frac{d\Omega_{2}^{2}}{4}\right]. (35)

The metric on the boundary L(p,1)L(p,1) is the round metric.

We may easily produce a Riemannian Einstein metric with the same conformal boundary by taking appropriate angular identifications of the Euclidean Schwarzschild-AdS5 metric to obtain an Einstein metric on 2×L(p,1)\mathbb{R}^{2}\times L(p,1):

ds2\displaystyle ds^{2} =U(r)dτ2+U(r)1dr2+r2[(dψ+cosθ2dϕ)2+dΩ224]\displaystyle=U(r)d\tau^{2}+U(r)^{-1}dr^{2}+r^{2}\left[\left(d\psi+\frac{\cos\theta}{2}d\phi\right)^{2}+\frac{d\Omega_{2}^{2}}{4}\right] (36)

where U(r)=1μ/r2+r2/2U(r)=1-\mu/r^{2}+r^{2}/\ell^{2}. We take, as above, ψψ+2π/p\psi\sim\psi+2\pi/p and θ(0,π)\theta\in(0,\pi), ϕ(0,2π)\phi\in(0,2\pi) and r>r+r>r_{+} where r+r_{+} is the largest root of U(r)U(r). As is well known, regularity at r=r+r=r_{+}, the largest root of U(r)U(r), requires that the angle τ\tau must be identified as ττ+β\tau\sim\tau+\beta with

β=2π2r+2r+2+2.\beta=\frac{2\pi\ell^{2}r_{+}}{2r_{+}^{2}+\ell^{2}}. (37)

Thus for fixed temperature T=β1T=\beta^{-1} there are two possible black holes

r+=π2±π222β22βr_{+}=\frac{\pi\ell^{2}\pm\ell\sqrt{\pi^{2}\ell^{2}-2\beta^{2}}}{2\beta} (38)

provided T>TminT>T_{min} where

Tmin=2π.T_{min}=\frac{\sqrt{2}}{\pi\ell}. (39)

Note that rather than closing smoothly to an S1×S2S^{1}\times S^{2} ‘bolt’ as in Euclidean Eguchi-Hanson-AdS5, the above space has a L(p,1)L(p,1) bolt.

We now follow the standard procedure Emparan:1999pm ; Balasubramanian:1999re to compare the finite Euclidean on-shell actions for these two possible infilling metrics for fixed temperature T=β1T=\beta^{-1}. The renormalized Euclidean action is

I=116πG[M(Rg+122)𝑑Vol(g)+2M(TrK3Rh4)𝑑Vol(h)]I=-\frac{1}{16\pi G}\left[\int_{M}\left(R_{g}+\frac{12}{\ell^{2}}\right)d\text{Vol}(g)+2\int_{\partial M}\left(\text{Tr}K-\frac{3}{\ell}-\frac{\ell R_{h}}{4}\right)d\text{Vol}(h)\right] (40)

where hh is the metric induced on a hypersurface r=Rr=R and Rg,RhR_{g},R_{h} are the respective scalar curvatures of the metrics gg and hh.

For Eguchi-Hanson-AdS5 we have

IEH=β2Vol(L(p,1))16πG[34(p241)2]I_{EH}=\frac{\beta\ell^{2}\text{Vol}(L(p,1))}{16\pi G}\left[\frac{3}{4}-\left(\frac{p^{2}}{4}-1\right)^{2}\right] (41)

where Vol(L(p,1))=2π2/p\text{Vol}(L(p,1))=2\pi^{2}/p is the volume of the boundary L(p,1)L(p,1). For the Euclidean black hole metric (36) a computation gives

IBH=β2Vol(L(p,1))16πG[34+r+22(1r+22)].I_{BH}=\frac{\beta\ell^{2}\text{Vol}(L(p,1))}{16\pi G}\left[\frac{3}{4}+\frac{r_{+}^{2}}{\ell^{2}}\left(1-\frac{r_{+}^{2}}{\ell^{2}}\right)\right]. (42)

With the actions at hand, simple calculations reveal that the situation is analogous to the Hawking-Page transition Hawking:1982dh . At low temperature the Eguchi-Hanson-AdS5 soliton has the least action and dominates the canonical ensemble, whereas at sufficiently large temperatures, it is a large black hole that dominates. When the actions are equal there is a transition analogous to the Hawking-Page transition. The temperature at which the phase transition occurs can be shown to be

TEHB=12π4+p48p2+202+p48p2+20.T_{\rm EHB}=\frac{1}{2\pi\ell}\frac{4+\sqrt{p^{4}-8p^{2}+20}}{\sqrt{2+\sqrt{p^{4}-8p^{2}+20}}}\,. (43)

At large values of pp this has the asymptotic form

TEHB=p2π[1+1p2+52p4+].T_{\rm EHB}=\frac{p}{2\pi\ell}\left[1+\frac{1}{p^{2}}+\frac{5}{2p^{4}}+\cdots\right]\,. (44)

In the strict pp\to\infty limit, this phase transition is related to that which occurs between the toroidal AdS black hole and the AdS soliton Surya:2001vj .

4 Geodesics

In this section we consider the behaviour of null and timelike geodesics in the Eguchi-Hanson-AdS5 geometry. Our primary interest is to investigate instabilities in this horizonless spacetime. As is now well established, there is a close connection between the geometrically induced stable trapping of null geodesics and instabilities due to the clumping of wave energy. In the geometric optics approximation, the propagation of solutions to the wave equation in a fixed background can be described by the trajectory of null geodesics. Stable trapping is a phenomena by which properties of the geometry create obstacles forcing null geodesics to be confined in a spatially compact region.

The prototypical example of null trapping is the photon sphere at r=3Mr=3M in the Schwarzschild spacetime; null geodesics exist in circular orbits at fixed radius. Such trapping, however is unstable because any small perturbation of the orbit will cause the null geodesics to either fall towards the horizon or escape to infinity. By contrast, stable trapping occurs when small perturbations of the orbits remain small so the trapping region is ‘attractive’. Stable trapping has been shown to lead to inverse-logarithmic time decay of wave energy in both asymptotically flat Benomio:2018ivy ; Keir:2016azt ; Gunasekaran:2020pue and asymptotically AdS spacetimes Holzegel:2013kna . As discussed in the introduction, this is suggestive of a non-linear instability, as linear stability typically requires decay that is an inverse polynomial function of time.

To exploit the local ×SU(2)×U(1)\mathbb{R}\times SU(2)\times U(1) isometry of the Eguchi-Hanson-AdS5 spacetime, it is convenient to (3) in the form

ds2\displaystyle ds^{2} =g(r)dt2+dr2f(r)g(r)+r2f(r)4[dψ¯+cos(θ)dϕ]2+r24dΩ22\displaystyle=-g(r)dt^{2}+\frac{dr^{2}}{f(r)g(r)}+\frac{r^{2}f(r)}{4}[d\bar{\psi}+\cos(\theta)d\phi]^{2}+\frac{r^{2}}{4}d\Omega^{2}_{2} (45)

where we have defined a new coordinate ψ¯=2ψ\bar{\psi}=2\psi. Then (ψ¯,θ,ϕ)(\bar{\psi},\theta,\phi) become the familiar Euler angles, with θ(0,π),ψ¯(0,4π/p),ϕ(0,2π)\theta\in(0,\pi),\bar{\psi}\in(0,4\pi/p),\phi\in(0,2\pi), and the metric can be expressed as

ds2=g(r)dt2+dr2f(r)g(r)+r24(σ12+σ22)+r2f(r)4σ32ds^{2}=-g(r)dt^{2}+\frac{dr^{2}}{f(r)g(r)}+\frac{r^{2}}{4}\left(\sigma_{1}^{2}+\sigma_{2}^{2}\right)+\frac{r^{2}f(r)}{4}\sigma_{3}^{2} (46)

where σi\sigma_{i} are left-invariant one-forms on SU(2)SU(2) defined by

σ1=\displaystyle\sigma_{1}= sinψ¯dθ+cosψ¯sinθdϕ,\displaystyle-\sin\bar{\psi}d\theta+\cos\bar{\psi}\sin\theta d\phi\,, (47)
σ2=\displaystyle\sigma_{2}= cosψ¯dθ+sinψ¯sinθdϕ,\displaystyle\cos\bar{\psi}d\theta+\sin\bar{\psi}\sin\theta d\phi\,, (48)
σ3=\displaystyle\sigma_{3}= dψ¯+cosθdϕ.\displaystyle d\bar{\psi}+\cos\theta d\phi\,. (49)

Explicitly, the spatial Killing vector fields are given by

R1\displaystyle R_{1} =cotθcosϕϕ+sinϕθcosϕsinθψ\displaystyle=\cot\theta\cos\phi\partial_{\phi}+\sin\phi\partial_{\theta}-\frac{\cos\phi}{\sin\theta}\partial_{\psi} (50)
R2\displaystyle R_{2} =cotθsinϕϕ+cosϕθ+sinϕsinθψ\displaystyle=-\cot\theta\sin\phi\partial_{\phi}+\cos\phi\partial_{\theta}+\frac{\sin\phi}{\sin\theta}\partial_{\psi} (51)
R3\displaystyle R_{3} =ϕ,L3=ψ.\displaystyle=\partial_{\phi},\qquad L_{3}=\partial_{\psi}. (52)

The trajectories of geodesics of mass MM are easily found using this symmetry and the Hamilton-Jacobi method as outlined in Kunduri:2005zg . In particular the Hamiltonian for the motion of uncharged particles is H=gabpapbH=g^{ab}p_{a}p_{b} where pap_{a} are the canonical momenta. The Hamiltonian system is Liouville integrable as there are five Poisson commuting functions associated with the local isometries (there is an additional conserved quantity associated with a reducible Killing tensor). The Hamilton-Jacobi equation is

Sλ+gabSxaSxb=0\frac{\partial S}{\partial\lambda}+g^{ab}\frac{\partial S}{\partial x^{a}}\frac{\partial S}{\partial x^{b}}=0 (53)

where λ\lambda is an affine curve parameter and it is clear one can express the Hamilton-Jacobi function SS in the separable form

S=M2λEt+pψψ+pϕϕ+Θ(θ)+R(r)S=M^{2}\lambda-Et+p_{\psi}\psi+p_{\phi}\phi+\Theta(\theta)+R(r) (54)

where (E,pψ,pϕ)(E,p_{\psi},p_{\phi}) correspond to conserved energy and angular momenta along particle trajectories, with pa=aSp_{a}=\partial_{a}S. Omitting details, we simply present the resulting curve equations for xa(λ)x^{a}(\lambda):

t˙\displaystyle\dot{t} =2Eg(r)ψ¯˙=8cotθr2[pϕsinθcotθpψ¯]\displaystyle=\frac{2E}{g(r)}\qquad\dot{\bar{\psi}}=-\frac{8\cot\theta}{r^{2}}\left[\frac{p_{\phi}}{\sin\theta}-\cot\theta p_{\bar{\psi}}\right]
ϕ˙\displaystyle\dot{\phi} =8r2sinθ[pϕsinθcotθpψ¯]\displaystyle=\frac{8}{r^{2}\sin\theta}\left[\frac{p_{\phi}}{\sin\theta}-\cot\theta p_{\bar{\psi}}\right] (55)

and

r˙2\displaystyle\dot{r}^{2} =4E2f(r)16g(r)r2pψ¯24M2f(r)g(r)16Cr2f(r)g(r)\displaystyle=4E^{2}f(r)-\frac{16g(r)}{r^{2}}p_{\bar{\psi}}^{2}-4M^{2}f(r)g(r)-\frac{16C}{r^{2}}f(r)g(r) (56)
θ˙2\displaystyle\dot{\theta}^{2} =64r2[C(cotθpψ¯pϕsinθ)2]\displaystyle=\frac{64}{r^{2}}\left[C-\left(\cot\theta p_{\bar{\psi}}-\frac{p_{\phi}}{\sin\theta}\right)^{2}\right] (57)

where CC is another constant of the motion associated with the existence of the reducible Killing tensor.

We will now perform more detailed studies of the geodesics.

4.1 Time-like Geodesics: Negative Mass Repulsion

Let us begin with a consideration of time-like geodesics, i.e. M0M\neq 0. To illustrate some similarities/differences with time-like geodesics in AdS, we will restrict attention here to radial time-like geodesics. After defining =E/M\mathcal{E}=E/M and rescaling the affine parameter accordingly, we arrive at the equation

r˙2=f(r)[2g(r)].\dot{r}^{2}=f(r)\left[\mathcal{E}^{2}-g(r)\right]\,. (58)

We immediately see that the larger-r turning point of the motion is exactly the same as it is for AdS222Of course, the radial coordinate for the soliton is not the same as the radial coordinate for pure AdS. These differences disappear at sufficiently large rr, as can be confirmed by putting the metric into Fefferman-Graham form. What we mean here is that the functional form of rmaxr_{\rm max} is identical.:

rmax=21.r_{\rm max}=\ell\sqrt{\mathcal{E}^{2}-1}\,. (59)

This result is sensible — the space is asymptotically locally AdS and so at large enough distances the radial motion should approach that of AdS. There is a further constraint to consider since the only physically relevant cases are those for which rmaxar_{\rm max}\geq a. This in turn enforces that the energy must be larger than a given threshold,

min=p2,wherea=p241.\mathcal{E}\geq\mathcal{E}_{\rm min}=\frac{p}{2}\,,\quad\text{where}\quad\frac{a}{\ell}=\sqrt{\frac{p^{2}}{4}-1}\,. (60)

A second turning point arises due to the presence of the bubble, this is at r=ar=a, where f(r)=0f(r)=0. The motion of a massive particle is therefore oscillatory, bouncing back and forth on the interval arrmaxa\leq r\leq r_{\rm max}.

The presence of the bubble has implications for the motion at smaller values of rr. To highlight this, consider the acceleration

r¨=r2a4r32+2a4(21)r5\ddot{r}=-\frac{r}{\ell^{2}}-\frac{a^{4}}{r^{3}\ell^{2}}+\frac{2a^{4}(\mathcal{E}^{2}-1)}{r^{5}} (61)

where the first term on the right-hand side is the acceleration term present in pure AdS. This makes manifest the well-known fact that the motion of time-like geodesics in AdS is periodic with period 2π2\pi\ell. Further, note that the sign of the acceleration in pure AdS is always negative — a stone tossed in AdS will always be returned to sender.

The additional acceleration terms for a0a\neq 0 have interesting consequences. The last term in the above always makes a positive contribution, since min>1\mathcal{E}\geq\mathcal{E}_{\rm min}>1. Close to the bubble this positive term actually dominates, leading to a region where the acceleration is positive, indicating a repulsion.

It is a simple matter to prove this. First, note that sufficiently large rr, the leading AdS term will always dominate, meaning that r¨<0\ddot{r}<0 at large rr. Next, let us consider the possibility of solutions to the equation r¨=0\ddot{r}=0. Define a new variable r=(x+α)r=\ell(x+\alpha) where α=a/\alpha=a/\ell. Then, the equation r¨=0\ddot{r}=0 becomes equivalent to the polynomial equation333While an explicit solution for xx in this equation can be obtained, it is sufficiently complicatedly that it is not beneficial to present it here.

x6+6αx5+15α2x4+20α3x3+16α4x2+8α5x+2α4(12+α2)=0.x^{6}+6\alpha x^{5}+15\alpha^{2}x^{4}+20\alpha^{3}x^{3}+16\alpha^{4}x^{2}+8\alpha^{5}x+2\alpha^{4}\left(1-\mathcal{E}^{2}+\alpha^{2}\right)=0\,. (62)

Every term in this polynomial is manifestly positive except for the very last one. It will vanish for

=min=p2.\mathcal{E}=\mathcal{E}_{\rm min}=\frac{p}{2}\,. (63)

For >min\mathcal{E}>\mathcal{E}_{\rm min} the last term is negative. In this case, applying Descartes’ rule of signs tells us that there will be a single positive value of xx where the above polynomial has a zero. There are no zeros for positive xx under other circumstances. Undoing our substitutions, x>0x>0 implies r>ar>a. Thus, we have concluded that there is exactly one zero for the acceleration for r>ar>a. Since we know from the above analysis that r¨<0\ddot{r}<0 for sufficiently large rr, we then conclude that in a neighbourhood of the bubble the acceleration is positive for particles with >p/2\mathcal{E}>p/2. Since this corresponds to the minimum possible energy, it follows that all massive particles feel a repulsion in the vicinity of the bubble, except those for which \mathcal{E} is exactly min\mathcal{E}_{\rm min} as these particles just sit at the bubble without motion.

The extent of this repulsion is bounded and approaches a constant, as can be seen by expanding the acceleration in the vicinity of the bubble:

r¨=(42p2)p24+𝒪(ra).\ddot{r}=\frac{(4\mathcal{E}^{2}-p^{2})}{\ell\sqrt{p^{2}-4}}+\mathcal{O}(r-a)\,. (64)

The origin of this effect is the negative mass of the bubble. To see this, it is instructive to re-write the expression for the acceleration in terms of the mass given in (27). It is a simple matter to show that it then takes the form,

r¨=r28pMπ(2rmax2r2)r5,\ddot{r}=-\frac{r}{\ell^{2}}-\frac{8pM}{\pi}\frac{\left(2r_{\rm max}^{2}-r^{2}\right)}{r^{5}}\,, (65)

where rmaxr_{\rm max} was introduced in eq. (59). Since rrmaxr\leq r_{\rm max} we see directly that if it were possible to have positive mass, then the acceleration would always be attractive. However, since the mass is necessarily negative, there is a competition between the confining potential of AdS and the negative mass repulsion of the bubble. This leads to a thin layer in the vicinity of the bubble where massive particles find themselves accelerated away from the bubble.

Finally, it is important to emphasize that the repulsion does not result in a situation where a (positive energy) particle ‘hovers’ at some fixed position r>ar>a. When >min\mathcal{E}>\mathcal{E}_{\rm min} the velocity is necessarily non-zero at the point where the acceleration vanishes. The only case when it is possible for r˙=r¨=0\dot{r}=\ddot{r}=0 simultaneously is when =min\mathcal{E}=\mathcal{E}_{\rm min}, corresponding to a particle at the location of the bubble.

The (radial) motion of massive particles in Eguchi-Hanson-AdS5 is therefore qualitatively similar to the motion of massive particles in AdS. The motion is forever oscillatory, with particles confined in some layer surrounding the bubble, the thickness of which depends on the energy of the particle. Unsurprisingly, this is suggestive that Eguchi-Hanson-AdS5 may suffer from a similar non-linear instability as global AdS, however, here without the possibility of forming horizons Dold:2017hwr .

4.2 Null Geodesics: Absence of Stable Trapping

Consider, now, null geodesics. These are obtained by setting M=0M=0 in (56), yielding

r˙2=4f(r)16g(r)η2r216C^r2f(r)g(r)\dot{r}^{2}=4f(r)-\frac{16g(r)\eta^{2}}{r^{2}}-\frac{16\hat{C}}{r^{2}}f(r)g(r) (66)

where η:=pψ/E\eta:=p_{\psi}/E and C^=C/E2\hat{C}=C/E^{2} and we have performed an appropriate rescaling of the affine parameter. For convenience, we will work with dimensionless parameters by scaling out \ell (this is equivalent to fixing =1\ell=1). Doing so produces

r˙2=V(x)=4x3P(x),P(x)=c3x3+c2x2+c1x+c0.\dot{r}^{2}=V(x)=\frac{4}{x^{3}}P(x),\qquad P(x)=c_{3}x^{3}+c_{2}x^{2}+c_{1}x+c_{0}. (67)

where x=r2x=r^{2} and

c3=14C^4η2,c2=4(C^+η2),c1=a4(14C^),c0=4a4C^.c_{3}=1-4\hat{C}-4\eta^{2},\quad c_{2}=-4(\hat{C}+\eta^{2}),\quad c_{1}=-a^{4}(1-4\hat{C}),\quad c_{0}=4a^{4}\hat{C}. (68)

Trajectories are only allowed in regions where the effective potential V(r)>0V(r)>0 with turning points at the zeroes, whereas regions with V(r)<0V(r)<0 are forbidden. Stable trapping will occur if there exist x1,x2x_{1},x_{2} such that 0<a2x1<x20<a^{2}\leq x_{1}<x_{2} with V(xi)=0V(x_{i})=0, V(x)>0V(x)>0 for x(x1,x2)x\in(x_{1},x_{2}), and V(x)<0V(x)<0 in neighbourhoods to the left of x1x_{1} and right of x2x_{2} (see (Gunasekaran:2020pue, , Fig 2)). This translates into similar conditions on the cubic P(x)P(x). First note that C^0\hat{C}\geq 0 by definition. Observe that

P(0)=4a4C^0,P(a2)=4a4(1+a2)η20.P(0)=4a^{4}\hat{C}\geq 0,\qquad P(a^{2})=-4a^{4}(1+a^{2})\eta^{2}\leq 0. (69)

We now consider several distinct cases.

4.2.1 Case 1: η,C^0\eta,\hat{C}\neq 0

From above, we have P(0)>0P(0)>0 and P(a2)<0P(a^{2})<0. Thus there must be at least 1 root x0x_{0} with 0<x0<a20<x_{0}<a^{2}. For stable trapping we will need two further positive roots x1,x2x_{1},x_{2} each strictly greater than a2a^{2}. As xx\to\infty, the sign of PP is controlled by c3c_{3}. If c3>0c_{3}>0 then it is clear there cannot be a 2nd root x2x_{2}. This occurs if

0<C^<14η2<140<\hat{C}<\frac{1}{4}-\eta^{2}<\frac{1}{4} (70)

Thus we find there is no stable trapping in this case. Now, suppose that c3<0c_{3}<0. then

C^>14η2\hat{C}>\frac{1}{4}-\eta^{2} (71)

Now consider Descartes’ rule of signs. We are assuming c3<0c_{3}<0, and c2<0c_{2}<0, whereas c0>0c_{0}>0 automatically. Thus there is only one sign flip between adjacent coefficients: if c1>0c_{1}>0 there is a sign flip between the x2x^{2} and xx coefficients, and if c1<0c_{1}<0 there is a sign flip between the xx and x0x^{0} coefficients. Thus the rule of signs indicates there can be only one positive root, but we already know one exists in (0,a2)(0,a^{2}). Thus stable trapping cannot occur in this case either.

Finally suppose c3=0c_{3}=0 so that

C^=14η2>0\hat{C}=\frac{1}{4}-\eta^{2}>0 (72)

in which case we must have η2<1/4\eta^{2}<1/4. Writing η2=1/4ϵ\eta^{2}=1/4-\epsilon for 0<ϵ<1/40<\epsilon<1/4, we have

P(x)=x2+a4(4ϵ1)x+4a4ϵP(x)=-x^{2}+a^{4}(4\epsilon-1)x+4a^{4}\epsilon (73)

which is a downward-pointing parabola. It is easy to see that the only turning point has to be for x<0x<0; in particular there cannot be two roots to the right of x=a2x=a^{2}. Thus there is no stable trapping here either.

4.2.2 Case 2: C^=0\hat{C}=0

In this case we get

P(x)=x((14η2)x24η2xa4).P(x)=x((1-4\eta^{2})x^{2}-4\eta^{2}x-a^{4}). (74)

Then x=0x=0 is automatically a root. Suppose that η2=0\eta^{2}=0. Then P=x(xa2)(x+a2)P=x(x-a^{2})(x+a^{2}) for which we easily read off the roots and find stable trapping cannot occur. Hence assume η2>0\eta^{2}>0. To get stable trapping we need the quadratic Q(x)=(14η2)x24η2xa4Q(x)=(1-4\eta^{2})x^{2}-4\eta^{2}x-a^{4} to have two roots that are greater than a2a^{2}. From Descartes’ rule of signs we see that there can be at most 1 sign flip between adjacent coefficients, and hence only one positive root. Thus there is no stable trapping here either.

4.2.3 Case 3: η=0\eta=0

We have

P(x)=(14C^)(xa2)(x+a2)(x4C^14C^)P(x)=(1-4\hat{C})(x-a^{2})(x+a^{2})\left(x-\frac{4\hat{C}}{1-4\hat{C}}\right) (75)

If C=1/4C=1/4 then this is just

P=(xa2)(x+a2)P=-(x-a^{2})(x+a^{2}) (76)

which is a downward pointing parabola and there is no stable trapping here. Hence assume C^1/4\hat{C}\neq 1/4. There is a root x=a2x_{-}=-a^{2} and x1=a2x_{1}=a^{2}. Then the only way we have the final root x2x_{2} to the right of x=a2x=a^{2} is if

4C^14C^>a2\frac{4\hat{C}}{1-4\hat{C}}>a^{2} (77)

which, since C^>0\hat{C}>0 implies we must have 14C^>01-4\hat{C}>0. But then P(x2)>0P^{\prime}(x_{2})>0, which cannot occur for stable trapping.

The above cases exhaust all possibilities, establishing that stable trapping does not occur.

4.3 Null Geodesics: Light-Crossing Time

In AdS, a light ray sent from a given point completes a round-trip to infinity and back in finite time coordinate time. Taking this point to be the orign, we have

TAdS=2r=0r=t˙r˙=201g(r)=π.T_{\rm AdS}=2\int_{r=0}^{r=\infty}\frac{\dot{t}}{\dot{r}}=2\int_{0}^{\infty}\frac{1}{g(r)}=\pi\ell\,. (78)

This in turn defines a fundamental frequency naturally associated with AdS:

ωAdS=2πTAdS=2.\omega_{\rm AdS}\ell=\frac{2\pi\ell}{T_{\rm AdS}}=2\,. (79)

This is relevant because the fundamental frequency matches the spacing for the overtones of scalar normal modes in AdS. Since we will study scalar normal modes in the next section, it is relevant to understand if a similar effect occurs for the soliton.

The computation for the light-crossing time in the soliton proceeds along the same lines. The integral to evaluate is now

Tp=2a1g(r)f(r)𝑑r,T_{\rm p}=2\int_{a}^{\infty}\frac{1}{g(r)\sqrt{f(r)}}dr\,, (80)

where here the subscript pp refers to the integer defining regularity of the geometry. This integral actually has a closed form expression,

Tp\displaystyle\frac{T_{p}}{\ell} =1482π(p24)3/2{48(p24)2Γ(34)2[F12(14,1,14,16(p24)2)1]\displaystyle=\frac{1}{48\sqrt{2\pi}(p^{2}-4)^{3/2}}\left\{48(p^{2}-4)^{2}\Gamma\left(\frac{3}{4}\right)^{2}\left[{}_{2}F_{1}\left(-\frac{1}{4},1,\frac{1}{4},\frac{16}{(p^{2}-4)^{2}}\right)-1\right]\right.
+Γ(14)2[48(p24)+256pp28F12(12,34,74,16(p24)2)]}.\displaystyle\left.\,+\Gamma\left(\frac{1}{4}\right)^{2}\left[48(p^{2}-4)+\frac{256}{p\sqrt{p^{2}-8}}{}_{2}F_{1}\left(\frac{1}{2},\frac{3}{4},\frac{7}{4},\frac{16}{(p^{2}-4)^{2}}\right)\right]\right\}\,. (81)

Expanding this in different limits is more useful. First, for p2p\to 2, the expansion is

Tp=π2πΓ[34]2p24πΓ[1/4]+32Γ[3/4]324πΓ[3/4](p2)3/2+π2(p2)2+𝒪(p2)5/2.\frac{T_{p}}{\ell}=\pi-\sqrt{\frac{2}{\pi}}\Gamma\left[\frac{3}{4}\right]^{2}\sqrt{p-2}-\frac{4\pi\Gamma\left[1/4\right]+3\sqrt{2}\Gamma[3/4]^{3}}{24\sqrt{\pi}\Gamma[3/4]}(p-2)^{3/2}+\frac{\pi}{2}(p-2)^{2}+\mathcal{O}(p-2)^{5/2}\,. (82)

The other limit of interest where this can be expanded is pp\to\infty. In this case the expansion reads,

Tp=Γ[1/4]22πp+2πΓ[1/4]28Γ[3/4]2p3+3(17Γ[3/4]2256Γ[3/4]2)82πp5+𝒪(p6).\frac{T_{p}}{\ell}=\frac{\Gamma[1/4]^{2}}{\sqrt{2\pi}p}+\sqrt{\frac{2}{\pi}}\frac{\Gamma[1/4]^{2}-8\Gamma[3/4]^{2}}{p^{3}}+\frac{3\left(17\Gamma[-3/4]^{2}-256\Gamma[3/4]^{2}\right)}{8\sqrt{2\pi}p^{5}}+\mathcal{O}(p^{-6})\,. (83)
Refer to caption
Figure 1: A plot of the light-crossing time TpT_{p} as a function of pp (black curve) compared with the asymptotic approximations as p2p\to 2 (blue curve) and as pp\to\infty (red curve). In each case we include the first three terms of the approximation. We see that the large pp approximation is quite accurate for all physical values of p3p\geq 3.

We show in Figure 1 a plot of the exact evaluation of the light crossing time, compared with the approximate forms described above. Here we have plotted the result treating pp as a continuous parameter. It must be kept in mind that regular geometries exist only for integer p3p\geq 3. First, we note that the light-crossing time for the soliton is always less than that of pure AdS. Regarding the approximations, for p2p\sim 2 the approximation is accurate in the close vicinity of p=2p=2, but is a poor approximation for larger, physical values of pp. The large pp approximation is much better suited for all cases p3p\geq 3. Finally, in the limit pp\to\infty, we recover the light-crossing time of the AdS soliton, τAdSS=Γ[1/4]2/(22π)\tau_{\rm AdSS}=\Gamma[1/4]^{2}/(2\sqrt{2\pi}).444See, for example, Myers:2017sxr . This is observed after performing the shift of the time coordinate necessary for that limit τ=pt/2\tau=pt/2 — see Section 2.1.

5 Wave Equation on Eguchi-Hanson Solitons

A massive Klein-Gordon field Φ\Phi obeys the equation

μμΦ=M2Φ\nabla^{\mu}\nabla_{\mu}\Phi=M^{2}\Phi (84)

in the spacetime background metric (3). We use the separation ansatz

Φ=eiωteimψY(θ,ϕ)R(r)\Phi=e^{-i\omega t}e^{im\psi}Y(\theta,\phi)R(r) (85)

where Y(θ,ϕ)Y(\theta,\phi) is an eigenfunction of the charged scalar Laplacian on CP1CP^{1}, satisfying

D2Y(θ,ϕ)=μY(θ,ϕ),.D^{2}Y(\theta,\phi)=-\mu Y(\theta,\phi),. (86)

The spectrum of this operator, and the associated eigenfunctions have been studied Wu:1976ge ; Warner:1982fs ; Hoxha:2000jf , and the spectrum is

μ=\displaystyle\mu= l(l+2)m2,\displaystyle\,l\left(l+2\right)-m^{2}\,, (87)
l=\displaystyle l=  2k+|m|,\displaystyle\,2k+|m|\,, (88)
k=\displaystyle k=  0,1,2,.\displaystyle\,0,1,2,\dots\,. (89)

Requiring Φ\Phi to be smooth implies m=npm=np, as ψ\psi is identified with period 2π/p2\pi/p. With this observation, the problem reduces to that of a single radial equation, which reads

1rddr(f(r)g(r)r3dR(r)dr)+[ω2r2g(r)m2f(r)M2r2μ]R(r)=0.\frac{1}{r}\frac{d}{dr}\left(f(r)g(r)r^{3}\frac{dR(r)}{dr}\right)+\left[\frac{\omega^{2}r^{2}}{g(r)}-\frac{m^{2}}{f(r)}-M^{2}r^{2}-\mu\right]R(r)=0\,. (90)

The radial equation (90) can be cast into Schrödinger equation form. To see this, we introduce a new independent variable xx

dxdr=1gf,x(a)=0\frac{dx}{dr}=\frac{1}{g\sqrt{f}},\qquad x(a)=0 (91)

and a dependent variable Ψ\Psi

Ψ=f1/4r3/2R.\Psi=f^{1/4}r^{3/2}R\,. (92)

This puts the radial equation into a formally self-adjoint form

d2Ψdx2+V(r(x))Ψ=ω2Ψ-\frac{d^{2}\Psi}{dx^{2}}+V(r(x))\Psi=\omega^{2}\Psi (93)

with potential

V(r)=gf1/4r3/2ddr[r3fgddr(f1/4r3/2)]+gr2(m2f+μ+M2r2).V(r)=-\frac{gf^{1/4}}{r^{3/2}}\frac{d}{dr}\left[r^{3}fg\frac{d}{dr}(f^{-1/4}r^{-3/2})\right]+\frac{g}{r^{2}}\left(\frac{m^{2}}{f}+\mu+M^{2}r^{2}\right)\,. (94)

We now seek solutions of this equation. We have not been able to find an analytic solution, and so we proceed with a combination of numerical and approximate techniques.

5.1 Approximate Solution: WKB Analysis

In the limit of large eigenvalues the differential equation can be solved approximately using, for example, the WKB method bender78:AMM . For the purposes of this analysis, it is useful to rewrite the differential equation (90) by defining

r=za,r=za\,, (95)

which maps the domain to z[1,)z\in[1,\infty). After this transformation, the resulting equation is given by

0=\displaystyle 0= z(z41)(4+(p24)z2)R′′(z)+(4(p24)z2+12z4+5(p24)z6)R(z)\displaystyle z(z^{4}-1)\left(4+(p^{2}-4)z^{2}\right)R^{\prime\prime}(z)+\left(4-(p^{2}-4)z^{2}+12z^{4}+5(p^{2}-4)z^{6}\right)R^{\prime}(z)
+4z3[4k22np4k(1+np)+n2p2z41z4+(p24)z22ω24+(p24)z2]R(z).\displaystyle+4z^{3}\left[-4k^{2}-2np-4k(1+np)+\frac{n^{2}p^{2}z^{4}}{1-z^{4}}+\frac{(p^{2}-4)z^{2}\ell^{2}\omega^{2}}{4+(p^{2}-4)z^{2}}\right]R(z)\,. (96)

We work here in an approximation where the overtone number NN is larger than the other fixed quantum numbers characterizing the problem, i.e. Nn,kN\gg n,k. In this case, by defining a new function and variable according to

z=1+ex,z=1+e^{x}\,, (97)

and

Ψ=z(z+1)(z2+1)(4+z2(p24))R(z),\Psi=\sqrt{\frac{z}{(z+1)(z^{2}+1)(4+z^{2}(p^{2}-4))}}R(z)\,, (98)

the differential equation takes the form of a Schrödinger equation,

1(ω)2d2Ψdx2+VΨ=0,-\frac{1}{(\omega\ell)^{2}}\frac{d^{2}\Psi}{dx^{2}}+V\Psi=0\,, (99)

with potential

V=4z4(z1)(p24)(4+z2(p24))2(z+1)(z2+1)V=\frac{4z^{4}(z-1)(p^{2}-4)}{\left(4+z^{2}(p^{2}-4)\right)^{2}(z+1)(z^{2}+1)} (100)

in the limit of large (ω)(\omega\ell). The transformation (97) has mapped the problem to the interval (,)(-\infty,\infty), which allows for direct comparison with the analysis of bender78:AMM . Examining the potential as a function of xx we see that it vanishes at the boundaries x±x\to\pm\infty, and there are no turning points on the interior domain. We can then apply directly the results of (bender78:AMM, , Sec. 10.5) for the quantization condition of normal mode solutions in the geometric optics approximation:

ωV(x)𝑑x=(N+12)π+𝒪(1ω),\omega\ell\int_{-\infty}^{\infty}\sqrt{V(x)}dx=\left(N+\frac{1}{2}\right)\pi+\mathcal{O}\left(\frac{1}{\omega}\right)\,, (101)

where NN is a non-negative integer. Written as an integral over zz, the left-hand side of the above becomes

V(x)𝑑x=12z2(4+z2(p24))p24z41𝑑z=Tp2,\int_{-\infty}^{\infty}\sqrt{V(x)}dx=\int_{1}^{\infty}\frac{2z^{2}}{\left(4+z^{2}(p^{2}-4)\right)}\sqrt{\frac{p^{2}-4}{z^{4}-1}}dz=\frac{T_{p}}{2\ell}\,, (102)

where in the last equality we recognize the integral as being identical to that defining the light crossing time of the soliton geometry given in (80). Therefore, applying the quantization condition as above, we have in the geometric optics approximation the following result for the frequencies:

ω=(N+12)2πTp=(N+12)ωfun,\omega=\left(N+\frac{1}{2}\right)\frac{2\pi}{T_{p}}=\left(N+\frac{1}{2}\right)\omega_{\rm fun}\,, (103)

where the fundamental frequency ωfun=2π/Tp\omega_{\rm fun}=2\pi/T_{p} for the Eguchi-Hanson-AdS5 soliton. We expect this relationship to be accurate in the limit of large overtone number, since we have assumed that ω\ell\omega is large. Later, in our numerical results, we will verify this is remarkably accurate even at small overtone number.

As explained below (80), the light crossing time can be evaluated analytically and was presented in Eq. (4.3). While this closed-form solution for the light-crossing time (or, equivalently, the fundamental frequency) is convenient, it is not necessarily illuminating. Therefore in Table 1 we list the numerical values for the fundamental frequency for a few values of pp. Since the asymptotics of TpT_{p} are dominated by a 1/p1/p term, we have factored out an overall multiple of pp in the numerical expressions for ωfun\omega_{\rm fun}. This allows one to see more clearly the limiting behaviour for larger values of pp. The results in this table should be compared with the fundamental frequency for the AdS soliton, which is

ωfunAdSS=42π3/2Γ[1/4]2=limp2ωfunp2.39628.\ell\omega_{\rm fun}^{\rm AdSS}=\frac{4\sqrt{2}\pi^{3/2}}{\Gamma[1/4]^{2}}=\lim_{p\to\infty}\frac{2\ell\omega_{\rm fun}}{p}\approx 2.39628\,. (104)

That is, at large values of pp, the fundamental frequency, which governs the normal modes at large overtone number Nk,nN\gg k,n, asymptotically approaches p/2p/2 times the fundamental frequency of the AdS soliton. This is directly related to the fact that there is a sense in which the AdS soliton is formally the large pp limit of the Eguchi-Hanson-AdS5 soliton, as explained in Section 2.1. The factor of 2/p2/p is precisely the factor required to match the time coordinates between the two geometries.

Numerical values of ωfun\omega_{\rm fun}
pp 3 4 5 6 7 8 8 10
ωfun/p\ell\omega_{\rm fun}/p 1.16822 1.18337 1.18917 1.19207 1.19375 1.19481 1.19553 1.19604
Table 1: A selection of numerically obtained ωfun\omega_{\rm fun} values.

5.2 Numerical Solution: Boundary Conditions & Regularity

To implement our numerical methods, it is essential to understand the behaviour of the solutions to the radial equation in the vicinity of the bubble and asymptotically. It will also be more convenient to compactify the semi-infinite domain to a finite interval. We therefore begin our numerical analysis of the radial equation with an analysis of the asymptotic behaviour of the solutions and boundary conditions.

We begin by introducing a new coordinate

u=2(ra)r+a1,u=\frac{2(r-a)}{r+a}-1\,, (105)

which maps the semi-infinite domain r(a,)r\in(a,\infty) to the interval u[1,1]u\in[-1,1]. To understand the singularity structure, we perform a Frobenius analysis. Near r=ar=a we write,

R(u)=(u+1)si=0ai(u+1)iR(u)=(u+1)^{s}\sum_{i=0}a_{i}(u+1)^{i} (106)

and extract ss by solving the differential equation near u=1u=-1. Provided that m0m\neq 0, the equation allows for two solutions, corresponding to the values

s=±m2p.s=\pm\frac{m}{2p}\,. (107)

We require regularity of the solution as u1u\to-1, and therefore only one of the above solutions is physically acceptable, depending on the sign of mm. In general, for m0m\neq 0 we have the regular behaviour s=|m|/(2p)s=|m|/(2p).

The case m=0m=0 must be treated separately, since the Frobenius method gives a degenerate root in that case. Since the degenerate root corresponds to s=0s=0, from the general theory of Frobenius analysis, we can conclude in this case that the solution must be of the form

R(u)=i=0ai(u+1)i+log(u+1)i=0bi(u+1)i.R(u)=\sum_{i=0}a_{i}(u+1)^{i}+\log(u+1)\sum_{i=0}b_{i}(u+1)^{i}\,. (108)

It can easily be shown that this ansatz leads to a consistent series solution in the vicinity of u=1u=-1. Regularity of the solution there forces us to set b0=0b_{0}=0, and so when m=0m=0, the solution approaches a constant as rar\to a. This behaviour is in fact captured by the result above, s=|m|/(2p)s=|m|/(2p), in the case m=0m=0.

The asymptotic analysis near u=1u=1 (i.e. rr\to\infty) is more standard. We can again proceed via Frobenius analysis. Taking the Frobenius ansatz

R(u)=(u1)s^i=0ai(u1)i,R(u)=(u-1)^{\hat{s}}\sum_{i=0}a_{i}(u-1)^{i}\,, (109)

expanding the differential equation near u1u\to 1 and demanding a solution of the indicial equation, we find

s^=2±2+(M)2.\hat{s}=2\pm\sqrt{2+(M\ell)^{2}}\,. (110)

Requiring the solution to be real gives the well-known Breitenlohner-Freedman bound Breitenlohner:1982jf , (ML)2>2(ML)^{2}>-2. As usual, we proceed by taking the normalizable solution,

s^=2+2+(M)2.\hat{s}=2+\sqrt{2+(M\ell)^{2}}\,. (111)

With the asymptotic behaviours understood, we now recast the differential equation into a form that is more amenable to numerical solution. To this end, we define a new function

R(u)=(1u)2+2+(M)2(1+u)|m|/(2p)h(u).R(u)=(1-u)^{2+\sqrt{2+(M\ell)^{2}}}(1+u)^{|m|/(2p)}h(u)\,. (112)

We then recast (90) in terms of the new function h(u)h(u). The resulting expression is somewhat messy, and so we do not present it here. The prefactors implement the appropriate fall-off conditions in the two relevant limits, and therefore the only requirement on h(u)h(u) is that it should be regular. Demanding this, a series solution near either u1u\to-1 or u+1u\to+1 results in Robin boundary conditions

h(1)\displaystyle h^{\prime}(-1) =[(4+p2)(m2+2|m|p)+4p4+8ω2+2p2(μω2)2p3(|m|+p)]h(1)\displaystyle=\left[\frac{(4+p^{2})(m^{2}+2|m|p)+4p^{4}+8\omega^{2}+2p^{2}(\mu-\omega^{2})}{2p^{3}(|m|+p)}\right]h(-1) (113)
h(1)\displaystyle h^{\prime}(1) =(1+|m|4p)h(1)\displaystyle=\left(1+\frac{|m|}{4p}\right)h(1) (114)

that must be imposed on the function h(u)h(u).

Before moving on, it is worth commenting in a bit more detail about the solution in the near bubble regime rar\to a. In this regime, the radial solution behaves as R(r)(ra)n/2R(r)\sim(r-a)^{n/2}, and so it appears to be continuous but not smooth there. To study this more carefully, we carry out the transformation

ρ=4rapf(a),φ=pψ\rho=\frac{4\sqrt{r-a}}{p\sqrt{f^{\prime}(a)}}\,,\quad\varphi=p\psi (115)

which yields

ds2=dρ2+ρ2dφ2ds^{2}=d\rho^{2}+\rho^{2}d\varphi^{2} (116)

namely the standard polar metric on 2\mathbb{R}^{2} as rar\to a, with φφ+2π\varphi\sim\varphi+2\pi due to the 2π/p2\pi/p periodicity of ψ\psi.

Consider next the solution to the wave equation as rar\to a. Focusing just on the terms with rr and ψ\psi dependence, and transforming these according to the polar coordinates (115) defined above, we obtain

ϕ(r,ψ)ρneinφ.\phi(r,\psi)\sim\rho^{n}e^{in\varphi}\,. (117)

Despite the fact that neither ρ\rho nor φ\varphi are themselves smooth functions — as can be confirmed by transforming these quantities to a Cartesian frame — the combinations as appearing here are indeed smooth. The solutions have the same structure as Bessel functions.

5.3 Numerical Solution: Normal Modes

We can now perform a numerical analysis of the radial equation (90). To solve this equation, we have employed two different numerical schemes. Our primary method has been a pseudospectral method for obtaining the eigenvalues of the differential operator. However, we have also cross-checked and benchmarked this method with a simpler shooting method. In Appendix A we compare the two techniques.

For an in-depth review of the pseudospectral method we refer to various references, e.g. Boyd ; canuto2007spectral ; Dias:2015nua . After the transformations (105) and (112), the radial equation is a problem on the interval [1,1][-1,1]. On this interval, we introduce a grid consisting of 𝒩+1\mathcal{N}+1 grid points uiu_{i}, which we take to be the Gauss-Lobatto points,

ui=cosiπ𝒩for i=0,,𝒩.u_{i}=\cos\frac{i\pi}{\mathcal{N}}\quad\text{for }i=0,\dots,\mathcal{N}\,. (118)

We then discretize the differential equation. The eigenfunction h(u)h(u) becomes a vector defined at the Gauss-Lobatto points, hi:=h(ui)h_{i}:=h(u_{i}), and the derivatives appearing in the differential equation are replaced with the corresponding Chebychev differentiation matrices555We refer to, for example, canuto2007spectral for an explicit form of these objects — c.f. section 2.4.2 therein. Di,jD_{i,j}. The differential equation then reduces to a generalized eigenvalue problem,

Hi,jhj=ω2Vi,jhjH_{i,j}h_{j}=\omega^{2}V_{i,j}h_{j} (119)

where Hi,jH_{i,j} is the discretization of the differential operator and Vi,jV_{i,j} is the matrix discretization of the terms multiplying ω2\omega^{2} in the original differential equation. To implement the Robin boundary conditions, we replace the first row of the above with the discretization of the second equation in (113), and the last row with the discretization of the first equation of (113).

After the differential equation has been discretized, we utilize the built-in eigenvalue solvers of Mathematica to obtain the eigenvalues ω\omega. For a given choice of 𝒩\mathcal{N} we will obtain a set of eigenvalues {ω}𝒩\{\omega\}_{\mathcal{N}} for the generalized eigenvalue problem described above. Of course, for any discretization of the differential equation, the eigenvalues of the corresponding matrix equation will differ from the true eigenvalues of the differential operator. We expect that the error in this discretization will become smaller as 𝒩\mathcal{N} is increased. We monitor convergence in the following way. We set a priori a tolerance that we regard as the minimum acceptable absolute error in the eigenvalue ω\omega. We then choose a number of grid points 𝒩\mathcal{N} and compute the spectrum {ω}𝒩\{\omega\}_{\mathcal{N}} first for 𝒩\mathcal{N} and then the spectrum {ω}𝒩+1\{\omega\}_{\mathcal{N}+1} for a grid with 𝒩+1\mathcal{N}+1 points. For each eigenvalue in {ω}𝒩\{\omega\}_{\mathcal{N}} we assess whether there is a corresponding eigenvalue in {ω}𝒩+1\{\omega\}_{\mathcal{N}+1} that is within the specified tolerance. If it is, then we conclude that this particular eigenvalue has been determined to the specified tolerance. We continue in this way, identifying each eigenvalue in {ω}𝒩\{\omega\}_{\mathcal{N}} that has converged within the specified tolerance. For most of our results the tolerance has been set to 10510^{-5}, meaning the eigenvalues are accurate to at least five decimal places. In most cases, we have repeated the above procedure for several values of 𝒩\mathcal{N} to gain a better understanding of the convergence properties. In Appendix C we present additional details on convergence for particular cases. Furthermore, in Appendix A we compare the results obtained via the pseudospectral method with those obtained via the shooting method, to ensure consistency.

Refer to caption
Figure 2: A comparison of the WKB approximate results for n=k=0n=k=0 (black lines) with the numerically computed eigenvalues for p=3,4,5,6p=3,4,5,6 corresponding to the blue, red, green and orange dots, respectively.
Refer to caption
Figure 3: A plot showing the error in the WKB approximation as a function of overtone number for p=3,4,5,6p=3,4,5,6 corresponding to the blue, red, green and orange dots, respectively. Here we show the logarithm of the relative absolute error, |ωWKBωNum|/ωNum|\omega_{\rm WKB}-\omega_{\rm Num}|/\omega_{\rm Num}.
Refer to caption
Figure 4: A plot showing the difference between the overtone spacing for the numerically computed frequencies ΔωNum=ωN+1ωN\Delta\omega_{\rm Num}=\omega_{N+1}-\omega_{N} and the fundamental frequency ωfun\omega_{\rm fun} on a logarithmic scale. The plot shows the case n=20,k=0n=20,k=0 for p=3,4,5,6p=3,4,5,6 corresponding to the blue, red, green and orange dots, respectively. The results are consistent with the fundamental frequency governing the spacing between successive overtones provided that Nn,kN\gg n,k.

When confusion may arise, when referring to a particular element of {ω}\{\omega\} we will use the notation ωN,k,n(p)\omega^{(p)}_{N,k,n} to indicate the dependence on the various parameters that appear in the equation. Here NN is a non-negative integer indicating the overtone, and n=m/pn=m/p. Since the equations and boundary conditions are invariant under mmm\to-m, we will without loss of generality consider only the case where nn is a non-negative integer. When there is less risk of confusion, we will suppress additional data attached to the spectral element ωN,k,n(p)\omega^{(p)}_{N,k,n} to avoid unnecessary bulky notion.

To begin our discussion of the numerical results, we compare the output of the pseudospectral method with the approximate results obtained via the WKB method. In the case of modes with n=k=0n=k=0, the WKB approximation is rather accurate for all overtone numbers. This is shown in Figure 2, where we show the output of the numerical methods (coloured dots) with the analytic approximation provided by the WKB method (black lines). The plot shows this comparison for overtones up to N=50N=50 for p=3,4,5p=3,4,5 and 66. In all cases, the agreement between the WKB result and the more accurate numerical approach is superb. This is further backed up by Figure 3, which shows the relative error between the WKB approximation and the numerical result for the same set of pp values. This plot shows that the WKB approximation is most accurate for the p=3p=3 case, but in all cases is accurate to about one part in one-thousand for overtone number N10N\geq 10. As expected, the accuracy of the WKB approximation becomes better the larger the overtone number becomes.

When nn and kk are non-vanishing, the WKB result no longer provides particularly good agreement, and we will discuss these cases at greater length below. However, it is important to note that even in these cases the WKB approach yields the correct spacing between overtones, provided the overtone number is sufficiently large compared to the values of nn and kk. We illustrate this in Figure 4, which compares the overtone spacing ωN+1ωN\omega_{N+1}-\omega_{N} as obtained numerically to the fundamental frequency, which governs the overtone spacing in the WKB approximation. The plot shows this comparison as a function of overtone number for the n=20n=20 and k=0k=0 mode for the cases of p=3,4,5p=3,4,5 and 66. We see in all cases that the absolute difference between the numerically determined spacing and the fundamental frequency is a monotonically decreasing function of the overtone number. This result, which we have verified in more examples in our numerical computations, is consistent with the notion that the fundamental frequency universally governs the spacing between overtones in the large overtone limit.

As mentioned, the WKB results provide a good approximation to the eigenvalues when n=k=0n=k=0, and more generally allow us to understand the spacing between overtones in the limit of large overtone number. However, there are a number of instances when the approximate solutions obtained in this way are insufficient. We move on to discuss these cases now. While the Eguchi-Hanson-AdS5 soliton is defined only for integer p3p\geq 3, the radial equation is sensible from a mathematical perspective under more general circumstances. To understand the spectrum {ω}\{\omega\} for the soliton, we have found it useful to analyse the radial equation for real values of p[2,)p\in[2,\infty). The main reason for this is that it is possible, through a combination of numerical and analytical techniques, to obtain simple asymptotic forms for ω\omega as p2p\to 2 and as pp\to\infty. By piecing together these approximate forms in various ways it is possible to get a good handle on the spectra over the full range of parameters.

In the limit p2p\to 2 we find the following behaviour:

ω0,k,n(p)\displaystyle\omega^{(p)}_{0,k,n}\approx  2(2+k+n)+n(p2)\displaystyle\,2\left(2+k+n\right)+n\left(p-2\right)
+[6(2n2nk)(k+n)(1+2k+2n)(1+2k+2n)](p2)2+\displaystyle+\left[\frac{6\left(2n^{2}-n-k\right)}{(k+n)(1+2k+2n)(-1+2k+2n)}\right](p-2)^{2}+\cdots (120)

where the first term has been obtained analytically; the analysis is outlined in Appendix C. Essentially, in the strict p=2p=2 limit, the radial equation reduces to that of a Klein-Gordon field on the orbifold AdS/52{}_{5}/\mathbb{Z}_{2}. That problem is obviously analytically solvable, and leads to the first term in the above. We find excellent numerical agreement between the p2p\to 2 limit of the numerical results and this analytically derived result, as detailed in Appendix C.

The second and third terms in (5.3) have been inferred from the numerical results. For simplicity, we have focused here only on the fundamental mode, and have not worked out the dependence for the overtones due to the complexity. Our procedure to determine these corrections was the following. We have computed, for numerous choices of nn and kk, the spectrum on the interval p(2,3]p\in(2,3]. For a given choice of nn and kk, we then fit the numerical results to a polynomial in (p2)(p-2). Adjusting the order of the polynomial fit, we see that the coefficients converge rapidly to particular values666For example, including terms up to order (p2)10(p-2)^{10} in the fit we find that the coefficient of the linear term (p2)(p-2) converges to five decimal places.. We take this as an indication that a series in (p2)(p-2) captures accurately the behaviour of the spectrum in the close vicinity of p=2p=2 for given values of nn and kk. By comparing the results of this procedure across several different values of nn and kk, we arrive at a collection of values cn,k,ic_{n,k,i} for the coefficients of the (p2)i(p-2)^{i} term in the polynomial fit. We then study the way the different cn,k,ic_{n,k,i} depend on nn and kk, and from this infer their analytical dependence. The analytic dependence is then cross-checked against numerical results for values of nn and kk that were not part of the sample used when deducing the candidate analytic form. This process becomes more involved as the power ii is increased. However, for the linear and quadratic terms, the dependence is simple enough that it can be inferred, giving the results presented above777We have also found that the behaviour of the (p2)3(p-2)^{3} term can be deduced in certain limits. For example, we have found that for n=0n=0 the coefficient of this term is 3/(4k21)-3/(4k^{2}-1) while for k=0k=0 the coefficient of this term is 3/(2n+1)23/(2n+1)^{2}. However, the functional form of the cubic term for general values of nn and kk has eluded us.. The fact that the dependence of these terms on nn and kk appears so simple suggests that it may be possible to determine these corrections analytically, though we shall not pursue this any further.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 5: Here we compare the approximate form of fundamental mode (black line) against the numerically determined spectra (blue dots) over the range p(2,3]p\in(2,3]. The plots show the cases n=1,k=1n=1,k=1 (top left), n=1,k=10n=1,k=10 (top right), n=10,k=1n=10,k=1 (bottom left) and n=10,k=10n=10,k=10 (bottom right). The plots indicate that the approximate form is valid over a larger range of pp values as the value of nn or kk is made larger.

In Figure 5, we compare the analytic approximation (5.3) with the numerical results for different values of nn and kk, over the interval p[2,3)p\in[2,3). While this range of pp does not correspond to regular five-dimensional geometries, it does concisely summarize important information about this approximation. First, we see that the approximate form is quite accurate over this interval. Second, we note that the approximation does better for large values of nn and kk. The reason for this second fact seems to be the following. For large nn or kk, the coefficient of the quadratic term in (5.3) behaves like 1/n1/n or 1/k21/k^{2}, respectively. This suggests that series (5.3) — if it is convergent — converges more rapidly for large values of these quantum numbers, or — if it is an asymptotic series rather than a convergent one — that the series approximates the true function for a large range of pp values in this limit.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 6: Plots of the frequencies vs. pp for n=0n=0 and k=9,11,13k=9,11,13 and 2020 (top left, top right, bottom left, bottom right, respectively). These plots highlight the fact that, for k>2n2nk>2n^{2}-n, it is possible for the eigenvalue to initially decrease as pp increases. In all cases, the light gray line indicates the analytic approximation (5.3), while the blue dots indicate numerical data.

A careful analysis of the analytic approximation shows that the quadratic term is negative whenever k>2n2nk>2n^{2}-n. This leads to some interesting structure in the spectrum, but does not lead to any complex eigenvalues. Namely, we find that in some circumstances the eigenvalues can initially decrease as a function of pp, however this behaviour is ultimately reversed when pp becomes large — for sufficiently large pp, the eigenvalues are monotonically increasing as a function of pp, as we will discuss below. Examples are shown in Figure 6 for n=0n=0 and different choices of kk. For all values of kk that we have explored, the shape of the eigenvalue curve as a function of pp is qualitatively the same. At larger values of kk, the ‘dip’ occurs at larger values of pp. In all cases, the depth of the dip is rather small, and in no cases do the eigenvalues come close to crossing through zero. In practice, we observe this effect for physical values of pp only in the case n=0n=0. For n0n\neq 0, the small p2p\to 2 approximation is less accurate for k>2n2nk>2n^{2}-n. Indeed, for any fixed values of nn and kk, the approximate form derived in the p2p\to 2 limit will ultimately become bad when pp is sufficiently large. Roughly, this occurs when the quadratic piece in the approximation dominates.

When pp becomes sufficiently large compared to the quantum numbers nn and kk, the approximate form derived in the p2p\to 2 limit fails to accurately capture the details of the spectrum. In this regime, we can make progress by understanding the solution of the radial equation in the limit of large pp. Numerical experiments suggest that, when pn,kp\gg n,k the eigenvalues exhibit a linear dependence on pp. This observation can be explained analytically. As described in Section 2.1, in the large pp limit, there is a formal sense in which the geometry limits to the AdS soliton. The normal modes of the AdS soliton then govern the slope of the eigenvalues ωN,k,n\omega_{N,k,n}. Explicitly, we have the following relation that holds for large pp:

dωN,k,n(p)dp=αN,n+asp,\frac{d\omega^{(p)}_{N,k,n}}{dp}=\alpha_{N,n}+\dots\quad\text{as}\quad p\to\infty\,, (121)

where the dots denote terms that are subleading as pp\to\infty. Here αN,n\alpha_{N,n} are the normal modes of the AdS soliton which, unfortunately, cannot be determined analytically888Strictly speaking, αN,n\alpha_{N,n} are the normal modes for the AdS soliton when the time coordinate is rescaled by a factor of 22 from the usual value.. We tabulate in Appendix B the numerically determined values of these normal modes and overtones for several values of nn. We also note that the WKB approximation gives a reasonably accurate approximation when NnN\gg n,

αN,n22π3/2Γ[1/4]2(N+12).\ell\alpha_{N,n}\approx\frac{2\sqrt{2}\pi^{3/2}}{\Gamma[1/4]^{2}}\left(N+\frac{1}{2}\right)\,. (122)

Note that the fact that the quantum number kk is absent in the above expressions is result of neither an assumption nor a typo. There is an ‘emergent degeneracy’ that appears for sufficiently large pp that eliminates any dependence on kk in the leading expressions.

Refer to caption
Refer to caption
Refer to caption
Figure 7: Plot of the fundamental mode for k=0k=0 (top left), k=1k=1 (top right), and k=2k=2 (bottom). In all cases, the curves correspond to n=0,1,2,3n=0,1,2,3 in order of bottom to top (or, in colour order, orange, blue, red, and green). The solid black lines correspond to an analytic approximation to the curves, while the coloured discs correspond to the results of the numerical computation.

While (121) provides a good approximation for the slope, additional details are required to determine the intercept and thereby obtain a good approximation to the eigenvalues themselves, rather than just their slope. One simple — and surprisingly accurate — option is to combine this large pp approximation with the fact that we know as p2p\to 2 the eigenvalues approach 2(2+n+k+N)2(2+n+k+N). Choosing the constant of integration to be such that the eigenvalues have this point as the intercept, we obtain the approximation:

ωN,k,n(p)(p2)αN,n+2(2+n+k+N).\omega^{(p)}_{N,k,n}\approx(p-2)\alpha_{N,n}+2(2+n+k+N)\,. (123)

This approximate form gives very good agreement with the numerically determined eigenvalues over the full range of physical pp values, provided that nn is sufficiently large. In this sense, one may consider (to a first approximation) that the normal modes of Eguchi-Hanson-AdS interpolate between the normal modes of AdS/52{}_{5}/\mathbb{Z}_{2} and the AdS soliton as the parameter pp is varied.

A more accurate approximation can be obtained by carrying out the same series of steps that we did in the p2p\to 2 limit. That is, through a combination of numeric and analytic techniques, we can determine the next-to-leading order terms in the large pp expansion. If instead we do this, we obtain the following:

ωN,k,n(p)pαN,n+n(2k+1)αN,n+𝒪(1/p).\omega^{(p)}_{N,k,n}\approx p\alpha_{N,n}+\frac{n(2k+1)}{\alpha_{N,n}}+\mathcal{O}(1/p)\,. (124)

We show this approximation for a few representative cases in Figure 7 for small values of nn.

Despite the constant term being heuristically derived, the approximation (123) actually does a better job than one might naively expect, at least for the fundamental mode. One reason for this is the following. Though we have not been able to prove this behaviour analytically, we numerically observe that as nn becomes large, α0,nn\alpha_{0,n}\approx n. Therefore, in this limit, the large pp approximation given in (123) approaches

ω0,k,n(p)2(2+n+k)+n(p2)forn.\omega^{(p)}_{0,k,n}\approx 2(2+n+k)+n(p-2)\quad\text{for}\quad n\to\infty\,. (125)

If we compare this result with the large nn\to\infty limit of the p2p\to 2 approximation given in (5.3) we see that the two expressions are exactly the same. Therefore, the approximate forms derived in the pnp\ll n and pnp\gg n limits are identical. This explains why both approximate forms do better than expected for large values of nn.

For the sake of completeness, we tabulate a number of the numerically determined normal modes in Appendix D.

6 Conclusions

Eguchi-Hanson-AdS solitons are conjectured to be the ground states for anti de Sitter gravity with lens space L(p,1)L(p,1) boundary conditions at infinity, analogous to how the AdS soliton is the conjectured ground state for toroidal boundary conditions at infinity. In this manuscript, we have considered various aspects of Eguchi-Hanson-AdS5 solitons. Our primary objective, which we have initiated here, is to understand to what extent these geometries are likely to be stable or not. In this sense, it is known Dold:2017hwr that under certain circumstances perturbations of the geometry may result in the formation of naked singularities. However, the precise mechanism underlying such an instability remains unknown. Since relatively little is understood about these geometries, our study here has focused on other related properties as well.

A key observation we have made concerns the relationship between the Eguchi-Hanson solitons and the AdS soliton. We demonstrated that in the limit where the lens space L(p,1)L(p,1) parameter pp\to\infty, the geometry becomes identically equal to that of the AdS soliton, up to a rescaling of the time coordinate. This observation is more than a curiosity, as we have found here this allows one to obtain approximations for various quantities of interest in terms of those same quantities for the simpler AdS soliton geometry.999After the first version of this manuscript appeared, Edgar Shaghoulian brought to our attention Shaghoulian:2016gol , where the same observation was made for the partition functions of CFTs on lens spaces. Our result for the connection between Eguchi-Hanson-AdS and the AdS soliton could be inferred from these results.

We have studied the conserved quantities of the Eguchi-Hanson solitons and constructed the extended first law of soliton mechanics. In doing so, we observe that these solitons possess a non-trivial thermodynamic volume, despite having vanishing entropy. Commensurate with a previous study of thermdynamic volume for asymptotically globally AdS5 solitons Andrews:2019hvq , the source of this quantity is topological in nature, and has to do with the structure of the Killing potential in the vicinity of the bubble. It would be interesting to extend these observations for other examples of solitons, as it may allow for further elucidation of the role and interpretation of thermodynamic volume in gravitational thermodynamics.

We also considered the thermodynamics of Eguchi-Hanson solitons in the canonical ensemble. After computing the on-shell action, we showed that for low temperatures the soliton dominates the canonical ensemble, while at higher temperature a Schwarzschild-AdS-type black hole with lens space horizon dominates. The phase transition between these two states is of the same type as studied for the toroidal AdS black hole and the AdS soliton first studied in Surya:2001vj .

We examined the geodesics of the Eguchi-Hanson solitons, finding no evidence suggestive of instability such as trapping of null geodesics. However, the time-like geodesics exhibit are oscillatory in nature analogous to geodesics in global AdS. Thus, an instability of the type shown in Bizon:2011gg for pure AdS may be present also for these solitons. Investigating this in more detail is something we hope to return to in the future.

Finally, we studied the separable solutions to the massless Klein-Gordon equation on this background. Via numerical and approximate analytical methods, we find a set of modes that oscillate, never decaying, analogous to AdS. We find no evidence of frequencies with a non-vanishing imaginary component. Using the WKB approximation, we have been able to study the separation between successive overtones, showing that it is well-approximated by the light-crossing time of the geometry by radial null geodesics. We have found it possible to understand the normal mode frequencies by piecing together two approximations. The first involves an analysis of the equation in the vicinity of the value p2p\to 2. While this does not correspond to a physical soliton solution, the radial equation in this limit reduces to that for AdS/52{}_{5}/\mathbb{Z}_{2}, which admits analytic solutions. The second approximation involves the pp\to\infty limit of the radial equation, which reduces to the wave equation on the AdS soliton. Joining the two approximations together gives a quite accurate approximation for the normal modes of the Eguchi-Hanson solitons for any physical value of pp in terms of a single parameter that must be numerically determined: the normal modes for the AdS soliton. In an approximate sense, one can then consider the normal modes of the Eguchi-Hanson solitons as interpolating between those of the orbifold AdS/52{}_{5}/\mathbb{Z}_{2} and the AdS soliton.

There remain a number of areas for future investigation. First among these, along the lines of understanding potential mechanisms for instability, would be an understanding of the gravitational perturbations of the geometry. This task is simplified since the solution is cohomogeneity-one, allowing for the techniques of Murata:2008yx to be used. An understanding of these perturbations may hint toward the existence of ‘resonating’ solutions constructible as non-linear extensions of the normal mode solutions Ishii:2020muv ; Garbiso:2020dys . It would furthermore be interesting to understand the implications of these geometries within the AdS/CFT correspondence, where they may be relevant for understanding aspects of confined phases of CFTs on lens space geometries Constable:1999gb ; Myers:1999psa ; Myers:2017sxr . It would also be interesting to understand the role of higher-curvature corrections for these geometries Wong:2011aa ; Corral:2021xsu ; Corral:2022udb .

Acknowledgements

We would like to thank Roberto Emparan, Edgar Shaghoulian, and Eric Woolgar for helpful discussions, comments, and correspondence. The work of RAH received the support of a fellowship from “la Caixa” Foundation (ID 100010434) and from the European Union’s Horizon 2020 research and innovation programme under the Marie Skłodowska-Curie grant agreement No 847648” under fellowship code LCF/BQ/PI21/11830027. HKK acknowledges the support of the NSERC Grant RGPIN-2018-04887. The work of TD and RBM was supported by the Natural Sciences and Engineering Research Council of Canada.

The lands on which Memorial University’s campuses are situated are in the traditional territories of the Beothuk, Mi’kmaq, Innu, and Inuit of the province of Newfoundland and Labrador. McMaster University is located on the traditional territories of the Mississauga and Haudenosaunee nations and within the lands protected by the “Dish with One Spoon” wampum agreement. University of Waterloo is situated on the traditional territory of the Neutral, Anishinaabeg and Haudenosaunee peoples.

Appendix A Solution of the Radial Equation via the Shooting Method

In this section, we provide independent confirmation for the results obtained via the pseudospectral method. To this end, we consider the solution of the radial equation via the shooting method.

We construct two numerical solutions of the radial equation. One solution begins near the bubble at r=ar=a, using a power series solution of the differential equation in this neighbourhood to construct initial conditions, along with an initial guess for the eigenvalue ω\omega. This solution is integrated outward, toward r=r=\infty, using the standard built-in numerical methods of Mathematica. The second numerical solution proceeds in much the same way, but begins near the asymptotic boundary, and is integrated inward toward r=ar=a.

The idea is simple: if ω\omega has been chosen correctly, then the solution must be everywhere regular, and moreover, the two numerical solutions should agree over the domain. However, since the equation is linear, and both R(r)R(r) and cR(r)cR(r) for some constant cc are equally valid solutions, it is not so simple to directly compare the solutions obtained by integrating from either end. (As there is no simple way to ensure consistent normalization from the different starting points.) Instead, we compare a logarithmic derivative,

Rin/outRin/out,\frac{R^{\prime}_{\rm in/out}}{R_{\rm in/out}}\,, (126)

of the numerical solutions at intermediate points. In this way, differences in the normalization of the solution can be eliminated. Matching of the logarithmic derivatives of the two numerically constructed solutions over the domain is taken to mean the correct value of ω\omega has been identified.

As the purpose of this method is to serve as independent confirmation of the pseudospectral results, we will not perform an exhaustive analysis here. Rather, we will present a few instances that demonstrate consistency of the two approaches.

Refer to caption
Refer to caption
Refer to caption
Figure 8: Plots of the logarithmic derivative of the numerical solutions. The blue curves correspond to solutions integrated from u=1u=-1, while the red curves correspond to solutions integrated from u=+1u=+1. In this case, we have set p=3p=3, n=0n=0, k=0k=0. The plots correspond to ω=5.2,5.4,5.2985999\omega=5.2,5.4,5.2985999 in order of left to right.

We show in Figure 8 an example of how the process works. The graphs show the logarithmic derivative of the solution over the domain u[1,1]u\in[-1,1]. The blue curves are those solutions obtained by numerically integrating the equation beginning at u=1u=-1, while the red curves are those solutions obtained by numerically integrating the equation beginning from u=+1u=+1. The three graphs correspond to three different choices of the eigenvalue ω\omega, while we have set p=3p=3, n=0n=0 and k=0k=0 here. The pseudospectral method outputs a value ωPS=5.2985999\omega_{\rm PS}=5.2985999 — this choice of ω\omega appears in the rightmost graph. The leftmost graph corresponds to the choice ω=5.2\omega=5.2, while the center one corresponds to ω=5.4\omega=5.4. For both cases shown where ω\omega is different from ωPS\omega_{\rm PS}, the two solutions are clearly in disagreement. However, for the choice ω=ωPS\omega=\omega_{\rm PS} the curves are visually indistinguishable over much of the domain, except for very near the end points101010Quantitatively, the difference in solutions is on the order of 10610^{-6} over the interval u[1/2,1/2]u\in[-1/2,1/2]. . This is consistent with ω=ωPS\omega=\omega_{\rm PS} being the correct choice for a consistent solution.

It is also insightful to understand the difference in the in/out solutions as a function of ω\omega, to ensure that the implementation of the pseudospectral method is not missing certain overtones, or returning incorrect eigenvalues. To this end, it is useful to introduce an integrated residual. Of course, for any finite resolution, we expect that the numerical solution that begins from u=1u=-1 to blow up sufficiently close to u=+1u=+1, and vice versa for the solution that begins from u=+1u=+1. Therefore, an integrated residual should focus on an intermediate domain where both solutions can be expected to be accurate. For this, we (arbitrarily) choose u[1/2,1/2]u\in[-1/2,1/2]. Another issue is the following. When ω\omega becomes larger than the fundamental frequency, the solution possesses a zero somewhere in the domain. (The number of zeros increases as ω\omega becomes larger than the various overtones). This leads to poles in the logarithmic derivative — see the center plot of Figure 8 for an example. To remedy this, it is convenient to work with an integrated reciprocal residual,

Res(ω)=1212du|ulogRoutulogRin|.{\rm Res}(\omega)=\int_{-\frac{1}{2}}^{\frac{1}{2}}\frac{du}{\left|\partial_{u}\log R_{\rm out}-\partial_{u}\log R_{\rm in}\right|}\,. (127)

Viewed as a function of ω\omega, the residual Res(ω){\rm Res}(\omega) will peak when the difference between the two solutions tends to zero. We can identify those peaks as the values of ω\omega where a consistent solution exists.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 9: Plots of the integrated (reciprocal) residual for n=1,2,3n=1,2,3 (first, second, and third rows, respectively) and k=0,1k=0,1 (first and second columns, respectively). In all cases, p=3p=3 has been chosen. The reciprocal residual peaks at the values of ω\omega for which a consistent solution exists — the absolute heights of the peaks shown in the graphs is meaningless. In all cases, the peaks correspond precisely to those values of ω\omega determined via the pseudospectral method.

We show a few representative examples in Figure 9. In these plots, the peaks correspond to values of ω\omega for which the boundary value problem admits a sensible solution. In all cases, these peaks correspond exactly to those values of ω\omega obtained via the spectral method. No additional values of ω\omega have been obtained, and no errors have be determined.

Appendix B Radial Equation at Large pp: Extracting the Slope of the Eigenvalues

As discussed in the numerical solution of the radial equation, it is fruitful to consider the radial equation for arbitrary values of pp. In particular, we have discussed the numerical observation that the behaviour of the eigenvalue ω\omega rapidly becomes dominated at large pp by a linear dependence. This fact can be explored in more detail, semi-analytically, to give a better understanding of that result.

Consider the radial equation (90) and perform the transformation

r=za,r=za\,, (128)

which maps the domain to z[1,)z\in[1,\infty). After this transformation, the resulting equation is given by

0=\displaystyle 0= z(z41)(4+(p24)z2)Rp′′(z)+(4(p24)z2+12z4+5(p24)z6)Rp(z)\displaystyle z(z^{4}-1)\left(4+(p^{2}-4)z^{2}\right)R^{\prime\prime}_{p}(z)+\left(4-(p^{2}-4)z^{2}+12z^{4}+5(p^{2}-4)z^{6}\right)R^{\prime}_{p}(z)
+4z3[4k22np4k(1+np)+n2p2z41z4+(p24)z22ω24+(p24)z2]Rp(z),\displaystyle+4z^{3}\left[-4k^{2}-2np-4k(1+np)+\frac{n^{2}p^{2}z^{4}}{1-z^{4}}+\frac{(p^{2}-4)z^{2}\ell^{2}\omega^{2}}{4+(p^{2}-4)z^{2}}\right]R_{p}(z)\,, (129)

where we have included the subscript pp to reinforce that, here, we are thinking of this as an arbitrary constant parametrizing the solution.

Consider (B) in the limit where pp is very large. Doing so, it becomes clear that the leading-order behaviour in each term behaves as 𝒪(p2)\mathcal{O}(p^{2}). Thus, a consistent possibility is for the dominant pp-dependence of ω\omega to be such that

ωαp\omega\sim\frac{\alpha}{\ell}p (130)

at large pp.

We can extract the equation that governs this behaviour and determines α\alpha. Rescale ω=αp\omega=\alpha p and peel off the terms in the differential equation that behave as 𝒪(p2)\mathcal{O}(p^{2}) at large pp. For consistency of notation, let us call Rp(z)R_{p}(z) in this limit h0(z)h_{0}(z). Carrying this out, we find the following differential equation:

0=(1+z4)h0′′(z)+(1+5z4)zh0(z)+4(n2z41z4+α2)h0(z).0=(-1+z^{4})h^{\prime\prime}_{0}(z)+\frac{(-1+5z^{4})}{z}h^{\prime}_{0}(z)+4\left(\frac{n^{2}z^{4}}{1-z^{4}}+\alpha^{2}\right)h_{0}(z)\,. (131)

By determining the values of α\alpha for which regular solutions to the above equation exist, we can determine the large pp slope of the eigenvalues ω\omega. Note that it is only the quantum number nn (=m/p)(=m/p) that enters into the equation, as kk has no fixed dependence on pp. Thus, the eigenvalues α=αN,n\alpha=\alpha_{N,n} are characterized by two numbers: the overtone NN and the quantum number nn. In other words, there is an ‘emergent degeneracy’ at large pp: The eigenvalues for all choices of kk are effectively the same, provided that p>kp>k is sufficiently large.

Unfortunately, we have not been able to solve this problem analytically (the equation is of Heun type). However, it is straight-forward to apply the same pseudospectral numerical techniques to this problem. We do not repeat the basic set up of this problem, and proceed directly to the results.

Values of αN,n\alpha_{N,n}
nn NN
0 1 2
0 1.70203 2.93798 4.15256
1 2.42243 3.61436 4.80816
2 3.25439 4.38480 5.54238
3 4.14123 5.21301 6.33085
4 5.05869 6.07883 7.15802
5 5.99504 6.97048 8.01368
6 6.94397 7.88064 8.89089
7 7.90176 8.80456 9.78483
8 8.86607 9.73904 10.69202
9 9.83533 10.68180 11.60991
10 10.80848 11.63122 12.53658
50 50.47989 50.98209 51.54973
100 100.38136 100.78145 101.23488
500 500.22318 500.45767 500.72385
1000 1000.17715 1000.36329 1000.57462
Table 2: A selection of numerically obtained αN,n\alpha_{N,n} values.

Some selected values of αN,n\alpha_{N,n} are presented in Table 2. These have been computed using the pseudospectral method to the accuracy shown in the table. While there is not a clear, discernible pattern for smaller values of nn, it is clear that when nn becomes large we have αN,nn\alpha_{N,n}\approx n. Unfortunately, it becomes increasingly expensive to perform the computations when nn becomes very large. Computing αN,n\alpha_{N,n} to five decimal places requires a Chebychev discretization of more than 1000 points when nn exceeds about 1000.

B.1 Subleading Terms in a Large pp Expansion

In the previous subsection, we have extracted a differential equation whose solution leads to the leading-order slopes of the eigenvalues ω\omega at large values of pp. It is possible to do somewhat better than this, without encountering additional obstructions. To this end, we approach the problem as one in perturbation theory, expanding in the small parameter ρ=1/p\rho=1/p.

We begin by expanding the eigenvalue and the eigenfunction as a perturbative series in ρ\rho:

ρ2Rp(z)\displaystyle\rho^{2}R_{p}(z) =h0(z)+ρh1(z)+ρ2h2(z)+,\displaystyle=h_{0}(z)+\rho h_{1}(z)+\rho^{2}h_{2}(z)+\cdots\,, (132)
ρ2ω2\displaystyle\rho^{2}\omega^{2} =αN,n2+ρβ1+ρ2β2+.\displaystyle=\alpha_{N,n}^{2}+\rho\beta_{1}+\rho^{2}\beta_{2}+\cdots\,. (133)

These expansions are then inserted into the differential equation, and an expansion in powers of ρ\rho is performed. In general, determining the corrections requires knowledge of both the eigenvalues and the eigenfunctions at previous orders. However, a happy accident at 𝒪(ρ)\mathcal{O}(\rho) allows for direct determination of β1\beta_{1}. The differential operator for h1(z)h_{1}(z) is identical to that determining h0(z)h_{0}(z), allowing for all terms involving derivatives to be eliminated. The problem then reduces to an algebraic equation that determines β1\beta_{1}, with the result

β1=2n(1+2k).\beta_{1}=2n(1+2k)\,. (134)

Unfortunately, at higher orders it does not appear to be possible to obtain analytic results for the corrections. While progress could be made numerically, the problem becomes more complicated. Therefore, it is not clear there is any benefit to pursuing this path further.

Appendix C Eguchi-Hanson Solitons as p2p\to 2

Having fruitfully studied the large pp limit of the Eguchi-Hanson soliton in the previous appendix, here we will consider the limit p2p\to 2. The limit should be taken at fixed cosmological scale \ell, so that the theory under consideration is not altered. To this end, we write a=𝒜a=\mathscr{A}\ell and consider the limit p2p\to 2. This is simple, and just results in 𝒜0\mathscr{A}\to 0. The resulting geometry is

ds2\displaystyle ds^{2} =g(r)dt2+dr2g(r)+r24[dψ+cos(θ)dϕ]2+r24dΩ22,g(r)=1+r22,\displaystyle=-g(r)dt^{2}+\frac{dr^{2}}{g(r)}+\frac{r^{2}}{4}\left[d\psi+\cos(\theta)d\phi\right]^{2}+\frac{r^{2}}{4}d\Omega^{2}_{2}\,,\quad g(r)=1+\frac{r^{2}}{\ell^{2}}\,, (135)

with ψ\psi normalized to have period 2π2\pi. The metric on the sections of constant (t,r)(t,r) is that of the projective space 𝕊3/2\mathbb{S}^{3}/\mathbb{Z}_{2}. The analysis of the wave equation on this space is useful for understanding the normal modes of Eguchi-Hanson-AdS for smaller values of pp.

C.1 Solution of the Radial Equation as p2p\to 2

While (regular) Eguchi-Hanson solitons exist only for integer p3p\geq 3, it is sensible from the mathematical point of view to study properties of the radial equation (90) without this restriction imposed on pp. Within this line of thought, the case p=2p=2 is special, since for p=2p=2 the parameter aa vanishes. That is, if we directly substitute p=2p=2 into the radial equation, it reduces to the radial equation for a scalar field on AdS, but with a special choice m=2nm=2n inherited from the fact that the limiting geometry is not globally AdS5, but instead the quotient space AdS5/2\mathbb{Z}_{2}. In this special case, the equation can be solved directly.

Substituting p=2p=2 into the radial equation we obtain,

0=r2(r2+L2)R′′(r)+r(3L2+5r2)R(r)+L2(4(k+n)(1+k+n)+L2r2ω2L2+r2)R(r).0=r^{2}(r^{2}+L^{2})R^{\prime\prime}(r)+r(3L^{2}+5r^{2})R^{\prime}(r)+L^{2}\left(-4(k+n)(1+k+n)+\frac{L^{2}r^{2}\omega^{2}}{L^{2}+r^{2}}\right)R(r)\,. (136)

After setting L=1L=1, the above equation has the following solution in terms of hypergeometric functions

R(r)=\displaystyle R(r)= (1+r2)ω/2[C1r2(1kn)F12(1kn+ω2,1kn+ω2,2k2n,r2)\displaystyle\,(1+r^{2})^{\omega/2}\left[C_{1}r^{2(-1-k-n)}{}_{2}F_{1}\left(-1-k-n+\frac{\omega}{2},1-k-n+\frac{\omega}{2},-2k-2n,-r^{2}\right)\right.
+C2r2(k+n)F12(k+n+ω2,2+k+n+ω2,2(1+k+n),r2)].\displaystyle+\left.C_{2}r^{2(k+n)}{}_{2}F_{1}\left(k+n+\frac{\omega}{2},2+k+n+\frac{\omega}{2},2(1+k+n),-r^{2}\right)\right]\,. (137)

Enforcing the boundary conditions proceeds in exactly the same manner as when studying the wave equation on AdS. To ensure regularity as r0r\to 0 we must set C2=0C_{2}=0. Then expanding near rr\to\infty, the behaviour is

R(r)\displaystyle R(r)\sim C1Γ[2(1+k+n)]Γ[2+k+nω/2]Γ[2+k+n+ω/2]\displaystyle\,\frac{C_{1}\Gamma\left[2(1+k+n)\right]}{\Gamma\left[2+k+n-\omega/2\right]\Gamma\left[2+k+n+\omega/2\right]}
+C1[2(k+n)+ω]Γ[2(1+k+n)]2r2Γ[1+k+nω/2]Γ[2+k+n+ω/2]+𝒪(r4).\displaystyle+\frac{C_{1}\left[2(k+n)+\omega\right]\Gamma\left[2(1+k+n)\right]}{2r^{2}\Gamma\left[1+k+n-\omega/2\right]\Gamma\left[2+k+n+\omega/2\right]}+\mathcal{O}(r^{-4})\,. (138)

To obtain the proper fall off, we must have the first two terms vanish, so that the radial solution decays at 𝒪(r4)\mathcal{O}(r^{-4}). Obviously, we cannot set C1=0C_{1}=0, since this results in the trivial solution. Therefore, we use the property of the gamma function that Γ[N]=\Gamma[-N]=\infty. Then, taking ω\omega to be a positive quantity, we obtain the solution:

ωN,k,n(p=2)=2(2+k+n+N).\omega^{(p=2)}_{N,k,n}=2\left(2+k+n+N\right)\,. (139)

Remarkably, this analytic result matches with numerical results, despite the fact that the limit p2p\to 2 is in a sense singular111111Consider the behaviour in the vicinity of r=ar=a. For p>2p>2, we have the behaviour R(r)(ra)|n|/2R(r)\sim(r-a)^{|n|/2} (140) while for p=2p=2, the behaviour is R(r)r2(k+|n|).R(r)\sim r^{2(k+|n|)}\,. (141) Obviously, there is a singular change in behaviour as p2p\to 2 (i.e. a0a\to 0), provided n,k0n,k\neq 0. . Numerical indications suggest that the eigenvalues ω\omega of the radial equation have a well-defined limit as p2p\to 2 with result the same as that given just above:

ωN,k,np2+=ωN,k,n(p=2)=2(2+k+n+N),\omega_{N,k,n}^{p\to 2^{+}}=\omega^{(p=2)}_{N,k,n}=2\left(2+k+n+N\right)\,, (142)

where NN is the overtone. We initially deduced this form by inspection, computing numerically values of ω\omega as a function of pp in the close vicinity of p=2p=2 for several values of nn and kk.121212For example, it is possible to obtain convergent results with the pseudospectral method for p=2+105p=2+10^{-5} with a Chebychev discretization of about 500 points.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 10: Top left: A plot of the lowest eigenvalue for n=0n=0, k=0k=0 and p=2+103p=2+10^{-3} as a function of the number of points 𝒩\mathcal{N} used in the Chebychev discretization. The even and odd values of NN converge to the final result from opposite directions. Top right: A plot of the difference between the values of ω\omega determined for even and odd 𝒩\mathcal{N} — it is clear that the difference is quite rapidly approaching zero as the number of points used in the discretization is increased, indicating convergence. At 𝒩=300\mathcal{N}=300, the value is ω2=16.000044\omega^{2}=16.000044, with even and odd discretizations differing at order 10610^{-6}. The bottom row shows the same information, but now for the n=1n=1 mode. In this case, convergence is more rapid.

To highlight the convergence of the numerical scheme in the limit p2p\to 2, we include in Figure 10 relevant plots. The top row depicts the case n=0,k=0n=0,k=0 with p=2+103p=2+10^{-3}, while in the bottom row the case n=1,k=0n=1,k=0 is shown — the results for different parameters are qualitatively similar. The top left plot shows the numerically computed value for the fundamental frequency as a function of the number of points used in the Chebychev discretization. The result converges to (142) from above/below depending on whether the number of points is odd/even. In the top right plot, we show the difference between the numerical results computed with 𝒩\mathcal{N} and 𝒩+1\mathcal{N}+1 points. The plot shows clearly the convergence. The plots included in the bottom row illustrate how, for n0n\neq 0, convergence is much more rapid (this has been a general feature of all our analysis here; it is not particular to the p2p\to 2 limit.)

Refer to caption
Refer to caption
Figure 11: Left: a plot of ω\omega as a function of pp near p=2p=2 for n=1n=1 and k=0,1,2k=0,1,2. The value of kk increases vertically between the different curves. Right: a plot of ω\omega as a function of pp near p=2p=2 for n=2n=2 and k=0,1,2k=0,1,2. The value of kk increases vertically between the different curves. In all cases, the red dots at the far left indicate the value of ω\omega at p=2p=2 determined via analytic calculations.

Finally, in Figure 11, we illustrate the dependence of the eigenvalues on nn and kk as a function of pp in the case when p2p\to 2. The left plot shows n=1n=1 with the three curves corresponding to k=0,1,2k=0,1,2 (bottom to top), while the right plot shows the same values of kk now for n=2n=2. The structure of the curve is rather close to a linear dependence on pp, but this is not as accurate as in the large pp limit.

Appendix D Tabulated Normal Mode Frequencies

Here we list the numerical values of the normal mode frequencies and the first eight overtones for p=3,4p=3,4 and 55 for 0n50\leq n\leq 5 and 0k50\leq k\leq 5.

Selected Normal Modes for Eguchi-Hanson Solitons with p=3p=3
nn kk ωN,n,k\ell\omega_{N,n,k}
0 0 5.29860 8.76887 12.26738 15.77064 19.27494 22.77952 26.28420 29.78891
0 1 6.30468 9.36992 12.69502 16.10285 19.54661 23.00934 26.48335 29.96461
0 2 7.92469 10.47381 13.51160 16.74837 20.07941 23.46252 26.87741 30.31310
0 3 9.82910 11.95092 14.65595 17.67462 20.85428 24.12706 27.45838 30.82877
0 4 11.84344 13.69088 16.06398 18.84279 21.84686 24.98695 28.21527 31.50380
0 5 13.88455 15.60829 17.68099 20.21454 23.03088 26.02396 29.13512 32.32874
1 0 7.91778 11.14589 14.49899 17.90981 21.35104 24.81006 28.28034 31.75819
1 1 9.42922 12.27359 15.38076 18.62934 21.95728 25.33320 28.74012 32.16812
1 2 11.18450 13.69814 16.53884 19.59402 22.78002 26.04869 29.37222 32.73377
1 3 13.06839 15.33579 17.92177 20.77130 23.79787 26.94192 30.16632 33.44760
1 4 15.01696 17.12117 19.48357 22.12918 24.98833 27.99678 31.11066 34.30082
1 5 16.99615 19.00469 21.18562 23.63863 26.32954 29.19684 32.19277 35.28387
2 0 10.70812 13.74673 16.95599 20.26005 23.62075 27.01752 30.43849 33.87643
2 1 12.41615 15.15809 18.13273 21.25980 24.48596 27.77830 31.11641 34.48725
2 2 14.24590 16.75210 19.50481 22.44836 25.52750 28.70191 31.94433 35.23645
2 3 16.14687 18.47772 21.03384 23.79816 26.72551 29.77380 32.91141 36.11579
2 4 18.08915 20.29630 22.68719 25.28394 28.06072 30.97927 34.00632 37.11648
2 5 20.05521 22.17902 24.43771 26.88344 29.51531 32.30412 35.21778 38.22946
3 0 13.57626 16.47128 19.55515 22.75613 26.03311 29.36154 32.72616 36.11706
3 1 15.38154 18.04874 20.92806 23.95921 27.09791 30.31357 33.58533 36.89890
3 2 17.25647 19.74496 22.44133 25.30787 28.30562 31.40240 34.57400 37.80273
3 3 19.17545 21.52829 24.06727 26.78005 29.63908 32.61474 35.68177 38.82033
3 4 21.12238 23.37482 25.78276 28.35616 31.08231 33.93772 36.89830 39.94336
3 5 23.08710 25.26671 27.56875 30.01910 32.62084 35.35928 38.21360 41.16357
4 0 16.48583 19.27040 22.24843 25.35576 28.55222 31.81198 35.11796 38.45854
4 1 18.34645 20.95265 23.75643 26.70864 29.77172 32.91802 36.12749 39.38560
4 2 20.25075 22.71605 25.36763 28.17479 31.10726 34.13883 37.24843 40.41971
4 3 22.18434 24.54022 27.06206 29.73690 32.54451 35.46273 38.47128 41.55312
4 4 24.13770 26.40968 28.82319 31.37983 34.07039 36.87869 39.78682 42.77817
4 5 26.10450 28.31272 30.63743 33.09049 35.67318 38.37654 41.18636 44.08744
5 0 19.41999 22.11767 25.00675 28.03092 31.15262 34.34625 37.59409 40.88356
5 1 21.31516 23.87006 26.61105 29.49631 32.49325 35.57697 38.72861 41.93395
5 2 23.23959 25.68024 28.29302 31.05090 33.92866 36.90416 39.95895 43.07815
5 3 25.18453 27.53459 30.03810 32.68119 35.44702 38.31775 41.27663 44.30901
5 4 27.14397 29.42265 31.83427 34.37546 37.03769 39.80841 42.67359 45.61964
5 5 29.11376 31.33639 33.67164 36.12358 38.69119 41.36761 44.14231 47.00348
Selected Normal Modes for Eguchi-Hanson Solitons with p=4p=4
nn kk ωN,n,k\ell\omega_{N,n,k}
0 0 6.93332 11.72096 16.48465 21.23594 25.98100 30.72257 35.46202 40.20008
0 1 7.60124 12.11154 16.76178 21.45093 26.15667 30.87111 35.59069 40.31358
0 2 8.78461 12.85744 17.30284 21.87462 26.50456 31.16608 35.84667 40.53964
0 3 10.30489 13.90214 18.08440 22.49536 27.01807 31.60343 36.22728 40.87641
0 4 12.03169 15.18497 19.07726 23.29755 27.68810 32.17741 36.72868 41.32121
0 5 13.88412 16.65214 20.25081 24.26344 28.50374 32.88095 37.34608 41.87064
1 0 10.24051 14.73250 19.33705 23.99139 28.67155 33.36672 38.07137 42.78236
1 1 11.51463 15.63997 20.03509 24.55669 29.14584 33.77494 38.42953 43.10133
1 2 13.01805 16.77399 20.92989 25.29093 29.76658 34.31176 38.90201 43.52302
1 3 14.67832 18.09221 21.99768 26.18007 30.52492 34.97130 39.48473 44.04451
1 4 16.44563 19.55760 23.21485 27.20908 31.41103 35.74686 40.17295 44.66234
1 5 18.28645 21.13971 24.55950 28.36294 32.41454 36.63116 40.96143 45.37263
2 0 13.83063 18.05755 22.48254 27.01100 31.59912 36.22446 40.87447 45.54163
2 1 15.36757 19.27010 23.47076 27.84021 32.31142 36.84779 41.42809 46.03928
2 2 17.02663 20.63114 24.60448 28.80398 33.14613 37.58225 42.08290 46.62951
2 3 18.77399 22.11310 25.86469 29.88939 34.09436 38.42154 42.83431 47.30889
2 4 20.58567 23.69313 27.23391 31.08382 35.14699 39.35901 43.67739 48.07369
2 5 22.44462 25.35262 28.69663 32.37531 36.29506 40.38791 44.60699 48.91994
3 0 17.55790 21.56648 25.82210 30.22019 34.70659 39.25103 43.83548 48.44859
3 1 19.23356 22.97093 27.01462 31.24942 35.60858 40.05209 44.55499 49.10106
3 2 20.98243 24.47698 28.31610 32.38580 36.61238 40.94854 45.36342 49.83637
3 3 22.78667 26.06675 29.71220 33.61852 37.70992 41.93430 46.25615 50.65094
3 4 24.63314 27.72554 31.19021 34.93744 38.89332 43.00329 47.22844 51.54105
3 5 26.51208 29.44134 32.73898 36.33320 40.15505 44.14953 48.27554 52.50291
4 0 21.36191 25.19187 29.29583 33.56898 37.95281 42.41246 46.92611 51.47943
4 1 23.12032 26.72433 30.63627 34.75182 39.00691 43.36073 47.78648 52.26597
4 2 24.92639 28.32918 32.05982 36.02062 40.14580 44.39074 48.72473 53.12633
4 3 26.76978 29.99453 33.55583 37.36661 41.36251 45.49695 49.73651 54.05703
4 4 28.64264 31.71054 35.11495 38.78177 42.65040 46.67401 50.81744 55.05453
4 5 30.53890 33.46910 36.72904 40.25880 44.00325 47.91671 51.96326 56.11529
5 0 25.21339 28.89566 32.86589 37.02311 41.30781 45.68286 50.12401 54.61483
5 1 27.02548 30.51774 34.31643 38.32580 42.48500 46.75368 51.10430 55.51758
5 2 28.87050 32.19326 35.83160 39.69815 43.73313 47.89463 52.15280 56.48605
5 3 30.74204 33.91404 37.40346 41.13315 45.04632 49.10086 53.26550 57.51693
5 4 32.63507 35.67325 39.02504 42.62444 46.41905 50.36769 54.43849 58.60695
5 5 34.54561 37.46519 40.69025 44.16628 47.84620 51.69068 55.66798 59.75290
Selected Normal Modes for Eguchi-Hanson Solitons with p=5p=5
nn kk ωN,n,k\ell\omega_{N,n,k}
0 0 8.60470 14.66629 20.66753 26.64481 32.61081 38.57065 44.52678 50.48049
0 1 9.11324 14.96233 20.87733 26.80748 32.74369 38.68298 44.62408 50.56631
0 2 10.05321 15.53752 21.29074 27.12989 33.00784 38.90668 44.81805 50.73752
0 3 11.31716 16.36255 21.89628 27.60647 33.40018 39.23985 45.10745 50.99326
0 4 12.80896 17.40204 22.67861 28.22944 33.91625 39.67973 45.49046 51.33228
0 5 14.45752 18.62026 23.62024 28.98940 34.55055 40.22285 45.96475 51.75293
1 0 12.62102 18.34009 24.16444 30.03559 35.93107 41.84070 47.75927 53.68385
1 1 13.78099 19.15265 24.78554 30.53704 36.35106 42.20180 48.07588 53.96567
1 2 15.14170 20.14934 25.56199 31.16994 36.88404 42.66161 48.47991 54.32587
1 3 16.65357 21.30438 26.48018 31.92650 37.52523 43.21697 48.96923 54.76291
1 4 18.27861 22.59355 27.52596 32.79821 38.26919 43.86428 49.54130 55.27497
1 5 19.98870 23.99529 28.68546 33.77618 39.11009 44.59954 50.19332 55.86000
2 0 17.02736 22.41656 28.02660 33.75054 39.53958 45.36902 51.22514 57.09976
2 1 18.48045 23.54323 28.93701 34.51085 40.19082 45.93785 51.72970 57.55290
2 2 20.04220 24.79302 29.96427 35.37729 40.93753 46.59272 52.31222 58.07710
2 3 21.68881 26.14828 31.09683 36.34228 41.77463 47.33009 52.97014 58.67048
2 4 23.40202 27.59345 32.32364 37.39823 42.69681 48.14618 53.70070 59.33097
2 5 25.16787 29.11511 33.63442 38.53767 43.69872 49.03708 54.50100 60.05635
3 0 21.62601 26.73381 32.13452 37.70106 43.36904 49.10308 54.88238 60.69395
3 1 23.23681 28.06247 33.25231 38.66034 44.20662 49.84506 55.54763 61.29643
3 2 24.91336 29.47653 34.45875 39.70509 45.12438 50.66151 56.28186 61.96287
3 3 26.64301 30.96426 35.74486 40.82877 46.11756 51.54890 57.08243 62.69125
3 4 28.41581 32.51548 37.10236 42.02508 47.18140 52.50364 57.94659 63.47945
3 5 30.22393 34.12150 38.52371 43.28799 48.31124 53.52215 58.87156 64.32526
4 0 26.33349 31.20580 36.41467 41.82709 47.37057 53.00282 58.69778 64.43856
4 1 28.03963 32.67149 37.68495 42.94117 48.35921 53.88953 59.50052 65.17118
4 2 29.78887 34.19866 39.02355 44.12449 49.41520 54.84053 60.36408 65.96115
4 3 31.57388 35.77939 40.42368 45.37162 50.53434 55.85255 61.28592 66.80645
4 4 33.38873 37.40685 41.87916 46.67746 51.71253 56.92233 62.26344 67.70499
4 5 35.22865 39.07512 43.38440 48.03724 52.94584 58.04671 63.29407 68.65470
5 0 31.10896 35.78320 40.81996 46.08697 51.50826 57.03751 62.64499 68.31085
5 1 32.87760 37.34615 42.20524 47.32327 52.62040 58.04577 63.56568 69.15707
5 2 34.67621 38.95485 43.64412 48.61614 53.78933 59.10958 64.53997 70.05460
5 3 36.50024 40.60382 45.13146 49.96119 55.01143 60.22602 65.56548 71.00152
5 4 38.34592 42.28828 46.66260 51.35432 56.28326 61.39222 66.63984 71.99588
5 5 40.21015 44.00411 48.23336 52.79171 57.60152 62.60539 67.76074 73.03575

References