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

**institutetext: Center for the Fundamental Laws of Nature, Harvard University,
17 Oxford Street, Cambridge, MA 02138, USA

Accelerating Black Hole Thermodynamics with Boost Time

Adam Ball *    Noah Miller [email protected] [email protected]
Abstract

We derive a thermodynamic first law for the electrically charged C-metric with vanishing cosmological constant. This spacetime describes a pair of identical accelerating black holes each pulled by a cosmic string. Treating the “boost time” of this spacetime as the canonical time, we find a thermodynamic first law in which every term has an unambiguous physical meaning. We then show how this first law can be derived using Noetherian methods in the covariant phase space formalism. We argue that the area of the acceleration horizon contributes to the entropy and that the appropriate notion of energy of this spacetime is a “boost mass”, which vanishes identically. The recovery of the Reissner-Nordstrom first law in the limit of small string tension is also demonstrated. Finally, we compute the action of the Euclidean section of the C-metric and show it agrees with the thermodynamic grand potential, providing an independent confirmation of the validity of our first law. We also briefly speculate on the significance of firewalls in this spacetime.

1 Introduction

The C-metric describes a pair of accelerating oppositely charged black holes pulled to infinity by cosmic strings kinwalk . The event horizons of these black holes are connected by a non-traversable wormhole through which the cosmic string threads hawkross . The spacetime has a “boost” symmetry, much like the boost symmetry of Minkowski space. The C-metric and its close cousin, the Ernst metric, have been used in multiple contexts to explore quantum gravity in spacetimes with a vanishing cosmological constant swwp ; ggsErnst ; Dowker:1993bt ; Dowker:1994up ; hhrErnst ; hawkross ; maldacena2013cool . Recently it has been proposed that they hold information about non-perturbative aspects of a putative celestial CFT living on the null infinity of asymptotically locally flat spacetime stromzhib .111The string makes the spacetime not asymptotically flat, but only asymptotically locally flat AshtekarDray .

For any of these applications, it would be beneficial to develop an understanding of the thermodynamics of the C-metric in analogy with the usual laws of black hole thermodynamics. This task is complicated by certain peculiarities of the C-metric. These peculiarities include the lack of a global timelike Killing vector, the existence of a non-compact acceleration horizon, and the presence of the cosmic string.

Our starting point in studying C-metric thermodynamics is to use “boost time” as the canonical time. That is, our first law will be from the perspective of a static observer.222The observer cannot see past the horizons, so their perceived universe is only a patch of the full C-metric spacetime, and in particular contains only one of the black holes. While the boost Killing vector is not globally timelike, adopting this position has proven to be fruitful. After a review of the C-metric and its corresponding background spacetime we discuss how the first law can be derived using the covariant phase space formalism, along the way addressing the subtle issue of the difference between “global” and “local” first laws. We then show how the first law of the Reissner-Nordstrom black hole is recovered in the small string tension limit. Following that, we show that when the temperature of the black hole matches the temperature of the acceleration horizon, the thermodynamic partition function agrees with the semiclassical approximation of the spacetime’s partition function defined through the Euclidean path integral. When the temperatures do not match, we speculate on the existence of firewalls on one or both horizons.

Jumping ahead, our C-metric first law is

0=κacc8πδΔ𝒜acc+κbh8πδ𝒜bh+ΦδQ+Δδμ.0=\frac{\kappa_{\rm acc}}{8\pi}\delta\Delta{\mathcal{A}}_{\rm acc}+\frac{\kappa_{\rm bh}}{8\pi}\delta{\mathcal{A}}_{\rm bh}+\Phi\delta Q+\Delta\ell\,\delta\mu. (1.1)

Every term in this equation has a physical meaning, including zero! κacc\kappa_{\rm acc} and κbh\kappa_{\rm bh} are the surface gravities of the acceleration horizon and black hole horizon, respectively. Δ𝒜acc\Delta{\mathcal{A}}_{\rm acc} is the difference in the area of the acceleration horizon between the C-metric and a related cosmic string spacetime (to be defined later). 𝒜bh{\mathcal{A}}_{\rm bh} is the black hole horizon area. QQ is the physical charge of the black hole and Φ\Phi is the difference in electric potential between the acceleration horizon and the black hole horizon. Δ\Delta\ell is the difference in the “thermodynamic length” Appels:2017xoe of the cosmic string between the C-metric and cosmic string spacetimes, where our thermodynamic length is the rate at which the cosmic string’s worldsheet sweeps out an area in spacetime. μ\mu is the tension of the string. Finally, the left hand side of the equation can be understood as δΔMboost\delta\Delta M_{\rm boost} where ΔMboost=0\Delta M_{\rm boost}=0 is the difference in the boost mass between the C-metric and cosmic string spacetimes. Throughout this paper we work with natural units, c=G==kB=1c=G=\hbar=k_{B}=1.

Many individual aspects of our first law (1.1) have appeared in previous works on accelerating black holes. Some authors have studied first laws in asymptotically locally AdS cases with no acceleration horizon Appels:2016uha ; Appels:2017xoe ; Anabalon:2018ydc ; Anabalon . Both of umass ; hawkross defined their notions of boost mass for the C-metric using a boost charge integral over a large hemisphere near spatial infinity, which we do as well. A quantity appearing conjugate to string tension in a first law is generally referred to as a thermodynamic length Appels:2017xoe . In Herdeiro:2009vd ; KrtousZel a thermodynamic length essentially identical to ours, with the same worldsheet area interpretation, appears in the first law for two non-accelerating charged black holes connected by a “strut”. The work hawkross studied the Euclidean C-metric and its action as a means of calculating the probability for a cosmic string to break via the nucleation of black holes through quantum tunneling. In Astorino (see also acovErnst ; AstReg ), the author analyzed C-metric thermodynamics with similar tools to those of this paper, but had a rather different perspective. Their first law was local to the black hole and rather than using boost time, a more general notion of time was used depending on several free functions. A first law for the C-metric was also proposed in Anabalon by adapting methods the authors had previously used in the AdS case, but the authors warned that its thermodynamic interpretation was uncertain, and it did not incorporate the acceleration horizon as we do here. The notion of a first law applying to a particular horizon, as opposed to the patch between two horizons, is discussed in SdS in the context of spinning de Sitter black holes. However, the mass term in those first laws comes from an integral at infinity and so they are not local in our sense.

The work we most directly build on is umass . Using the canonical Hamiltonian formalism, its authors derived a perturbative first law based on boost time for any asymptotically locally flat spacetime with an acceleration horizon. However they did not allow the string tension to vary, their prescription for boost mass contained an ambiguity (discussed further in section 3), and their handling of the normalization of boost time in the case of the C-metric unnecessarily fixed a metric parameter.

2 The C-metric

In our coordinates, based on those of Hong:2003gx ; stromzhib , the C-metric is

ds(C)2=1A2(xy)2(G(y)dt2dy2G(y)+dx2G(x)+α2G(x)dϕ2).ds_{(C)}^{2}=\frac{1}{A^{2}(x-y)^{2}}\Big{(}G(y)dt^{2}-\frac{dy^{2}}{G(y)}+\frac{dx^{2}}{G(x)}+\alpha^{2}G(x)d\phi^{2}\Big{)}. (2.1)

It solves the Einstein equations everywhere (with a delta function stress tensor on the string), and away from the cosmic string it is an electrovacuum solution (i.e. Tμν=TμνEMT_{\mu\nu}=T_{\mu\nu}^{\rm EM}) with gauge field

Aμdxμ=q(y+1)dt.A_{\mu}dx^{\mu}=q(y+1)dt. (2.2)

Here G(ζ)G(\zeta) is the quartic polynomial

G(ζ)=(1ζ2)(1ζ/ζ1)(1ζ/ζ2)G(\zeta)=(1-\zeta^{2})(1-\zeta/\zeta_{1})(1-\zeta/\zeta_{2}) (2.3)

with roots given by

ζ1=1Ar,ζ2=1Ar+,ζ3=1,ζ4=1\zeta_{1}=-\frac{1}{Ar_{-}},\hskip 14.22636pt\zeta_{2}=-\frac{1}{Ar_{+}},\hskip 14.22636pt\zeta_{3}=-1,\hskip 14.22636pt\zeta_{4}=1 (2.4)

where

r±=m±m2q2.r_{\pm}=m\pm\sqrt{m^{2}-q^{2}}. (2.5)

We require |q|<m|q|<m and Ar+<1Ar_{+}<1 so that the roots satisfy

ζ1<ζ2<ζ3<ζ4.\zeta_{1}<\zeta_{2}<\zeta_{3}<\zeta_{4}. (2.6)

Finally, we define

α=2|G(1)|=ζ1ζ2(1ζ1)(1ζ2).\alpha=\frac{2}{|G^{\prime}(1)|}=\frac{\zeta_{1}\zeta_{2}}{(1-\zeta_{1})(1-\zeta_{2})}. (2.7)

We can see the C-metric depends on three parameters, mm, qq, and AA. Roughly speaking, mm is the mass of a black hole, qq is its charge, and AA is its acceleration. It should be mentioned, however, that for many calculational purposes it is simpler to instead take ζ1\zeta_{1}, ζ2\zeta_{2}, and AA as the three independent parameters.

The range of xx is 1x1-1\leq x\leq 1. ϕ\phi is periodically identified ϕϕ+2π\phi\sim\phi+2\pi. α\alpha has been chosen to avoid a conical singularity at x=1x=1 between the black holes, although one remains at x=1x=-1 which is the location of the cosmic string. The angular deficit around the cosmic string is

δdef=2π(1|G(1)G(1)|).\delta_{\rm def}=2\pi\left(1-\Big{|}\frac{G^{\prime}(-1)}{G^{\prime}(1)}\Big{|}\right). (2.8)

The boost time tt can be any real number. Different ranges of yy correspond to different patches of the spacetime. yy\to-\infty is the black hole singularity, y=ζ1y=\zeta_{1} is the black hole inner horizon, y=ζ2y=\zeta_{2} is the black hole event horizon, y=1y=-1 is the acceleration horizon, and conformal infinity is approached as yxy\to x. If y<1<xy<-1<x and xy0x-y\to 0, spatial infinity is reached. If 1<y<x-1<y<x and yxy\to x, null infinity is reached.

Refer to caption
Figure 1: Lines of constant yy on the t=const.t=\rm const., ϕ=0,π\phi=0,\pi slice.
Refer to caption
Figure 2: Lines of constant xx on the t=const.t=\rm const., ϕ=0,π\phi=0,\pi slice.

The C-metric has two Killing vectors, ϕ\partial_{\phi} and t\partial_{t}. ϕ\partial_{\phi} generates rotations around the string while t\partial_{t} generates “boosts” parallel to the string. t\partial_{t} is timelike between the outer black hole horizon and acceleration horizon, but is spacelike between the acceleration horizon and null infinity, as well as between the outer and inner black hole horizons.

To get an understanding for the coordinates, in figures 2 and 2 we have depicted the t=const.t=\rm const., ϕ=0,π\phi=0,\pi slice of the C-metric in the ζ2y1\zeta_{2}\leq y\leq-1 patch. It has the geometry of a wormhole which we have drawn cut in half. We can see that on a t=const.t=\rm const. slice, surfaces of constant yy are two-spheres. yy can be thought of as 1/Ar-1/Ar, where rr is a radial coordinate from the singularity. The variable xx can be thought of as cosη-\cos\eta, where η\eta is the polar angle in bi-spherical coordinates. Note that xx acts as a radial coordinate on the acceleration horizon.

Different coordinates may be employed in order to study the “worldlines” the black holes trace out in spacetime. In gkp , the authors do this for the uncharged C-metric and confirm that the black holes travel on roughly hyperbolic Rindler-like trajectories.

3 The cosmic string background spacetime

The C-metric is closely related to the flat cosmic string spacetime. This spacetime consists of nothing but a straight string in flat space whose stress-energy density sources a conical singularity. In our first law, Δ𝒜acc\Delta{\mathcal{A}}_{\rm acc} and Δ\Delta\ell are defined as differences between analogous quantities computed in the C-metric and the cosmic string “background,” so it is worth reviewing their relationship carefully.

Refer to caption
Figure 3: For ε0\varepsilon\to 0, the surface of constant ε\varepsilon is a large hemisphere with polar angle θ\theta.

It will be helpful to find coordinates that are naturally adapted to studying the spatial infinity of the C-metric. We can do this by parameterizing yy and xx by ε\varepsilon and θ\theta as follows:

y\displaystyle y =1εcos2θ,\displaystyle=-1-\varepsilon\cos^{2}\theta, (3.1)
x\displaystyle x =1+εsin2θ.\displaystyle=-1+\varepsilon\sin^{2}\theta. (3.2)

The surface of constant ε\varepsilon where ε0\varepsilon\to 0 corresponds to a large hemisphere approaching spatial infinity. The coordinate θ\theta is the polar angle of this sphere. See figure 3.

Using the (t,ε,θ,ϕ(t,\varepsilon,\theta,\phi) coordinates we will define ds(B)2ds^{2}_{(B)}, the exact metric of the cosmic string spacetime, to be the leading order part of the C-metric ds(C)2ds^{2}_{(C)} in the small ε\varepsilon expansion. Here we must be sure to treat dεd\varepsilon as 𝒪(ε){\mathcal{O}}(\varepsilon), which renders the dεdθd\varepsilon d\theta component subleading compared to the dε2d\varepsilon^{2} and dθ2d\theta^{2} components. We find

ds(C)24A2G(1)ε(14G(1)2cos2θdt2+dε24ε2+dθ2+sin2θG(1)2G(1)2dϕ2)ds(B)2.\begin{split}ds_{(C)}^{2}&\approx\frac{4}{A^{2}G^{\prime}(-1)\varepsilon}\Big{(}-\frac{1}{4}G^{\prime}(-1)^{2}\cos^{2}\theta dt^{2}+\frac{d\varepsilon^{2}}{4\varepsilon^{2}}+d\theta^{2}+\sin^{2}\theta\frac{G^{\prime}(-1)^{2}}{G^{\prime}(1)^{2}}d\phi^{2}\Big{)}\\ &\equiv ds^{2}_{(B)}.\end{split} (3.3)

One way to see that ds(B)2ds^{2}_{(B)} really is the metric of a cosmic string is to substitute

R2\displaystyle R^{2} =4A2G(1)ε,\displaystyle=\frac{4}{A^{2}G^{\prime}(-1)\varepsilon}, (3.4)
t~\displaystyle\tilde{t} =G(1)2t,\displaystyle=\frac{G^{\prime}(-1)}{2}t, (3.5)

at which point the metric, in (t~,R,θ,ϕ)(\tilde{t},R,\theta,\phi) coordinates, becomes

ds(B)2=R2cos2θdt~ 2+dR2+R2dθ2+G(1)2G(1)2R2sin2θdϕ2.ds^{2}_{(B)}=-R^{2}\cos^{2}\theta d\tilde{t}^{\,2}+dR^{2}+R^{2}d\theta^{2}+\frac{G^{\prime}(-1)^{2}}{G^{\prime}(1)^{2}}R^{2}\sin^{2}\theta d\phi^{2}. (3.6)

This is the cosmic string spacetime in accelerated spherical coordinates with boost time t~\tilde{t}. We can see that it has the same angular deficit as the C-metric we constructed it from, given in (2.8).

Using the parameterizations of (3.1) and (3.2), we have shown (by construction) that the metric components of the C-metric and cosmic string spacetimes asymptotically agree to leading order in ε\varepsilon. However, we can do better, and doing so is crucial for later calculations.

As discussed in hawkhor , a proper subtraction method should match the induced metric on the boundary exactly between the spacetime and its background. The problem with (3.1) and (3.2) is that the induced metrics on an ε=const.\varepsilon=\rm const. surface in the C-metric and cosmic string spacetimes only match at leading order in ε\varepsilon. The first subleading order can also affect the quantities we will compute, so we need to add a correction to the coordinate transformations (3.1) and (3.2) that is subleading in ε\varepsilon. An improved coordinate transformation is given by

y\displaystyle y =1εcos2θ(1ε(3cos2θ1)G′′(1)4G(1)),\displaystyle=-1-\varepsilon\cos^{2}\theta\Big{(}1-\varepsilon\frac{(3\cos 2\theta-1)G^{\prime\prime}(-1)}{4G^{\prime}(-1)}\Big{)}, (3.7)
x\displaystyle x =1+εsin2θ(1ε(3cos2θ+1)G′′(1)4G(1)).\displaystyle=-1+\varepsilon\sin^{2}\theta\Big{(}1-\varepsilon\frac{(3\cos 2\theta+1)G^{\prime\prime}(-1)}{4G^{\prime}(-1)}\Big{)}. (3.8)

Surprisingly, the result of matching the induced metric components on the surfaces of small constant ε\varepsilon ends up giving a better match of every component of the metrics ds(C)2ds^{2}_{(C)} and ds(B)2ds^{2}_{(B)} to one order higher in ε\varepsilon. Namely, in these new coordinates the nonzero C-metric components satisfy

gtt(C)\displaystyle g^{(C)}_{tt} =gtt(B)(1+𝒪(ε2)),\displaystyle=g^{(B)}_{tt}(1+\mathcal{O}(\varepsilon^{2})), gεε(C)\displaystyle g^{(C)}_{\varepsilon\varepsilon} =gεε(B)(1+𝒪(ε2)),\displaystyle=g^{(B)}_{\varepsilon\varepsilon}(1+\mathcal{O}(\varepsilon^{2})),
gθθ(C)\displaystyle g^{(C)}_{\theta\theta} =gθθ(B)(1+𝒪(ε2)),\displaystyle=g^{(B)}_{\theta\theta}(1+\mathcal{O}(\varepsilon^{2})), gϕϕ(C)\displaystyle g^{(C)}_{\phi\phi} =gϕϕ(B)(1+𝒪(ε2)),\displaystyle=g^{(B)}_{\phi\phi}(1+\mathcal{O}(\varepsilon^{2})), (3.9)
gεθ(C)\displaystyle g^{(C)}_{\varepsilon\theta} =𝒪(ε0).\displaystyle={\mathcal{O}}(\varepsilon^{0}).

In an orthonormal frame these would differ from the cosmic string only at 𝒪(ε2){\mathcal{O}}(\varepsilon^{2}). Happily, this 𝒪(ε2){\mathcal{O}}(\varepsilon^{2}) difference will have no effect on the physical quantities we compute in the ε0\varepsilon\to 0 limit.

The calculation of boost mass in umass , besides using ADM charges, differed from ours by relating the C-metric to the cosmic string using coordinates like our (3.1), (3.2) as opposed to our (3.7), (3.8). This resulted in a different value for the boost mass.

4 Definition of physical quantities

In this section we provide formulas for all physical quantities appearing in our first law (1.1). We note that every integration performed herein occurs on a t=const.t=\rm const. slice. We begin with the physical charge and black hole area,

Q=14πF=14πg(C)Ftydxdϕ|y=const.=αq=ζ1ζ2A(1ζ1)(1ζ2),Q=\frac{1}{4\pi}\int\star F=\frac{1}{4\pi}\int\sqrt{-g^{(C)}}F^{ty}dxd\phi\Bigg{\rvert}_{y=\rm const.}=\alpha q=\frac{\sqrt{\zeta_{1}\zeta_{2}}}{A(1-\zeta_{1})(1-\zeta_{2})}, (4.1)
𝒜bh=gxx(C)gϕϕ(C)dxdϕ|y=ζ2=4πζ1ζ2A2(ζ11)(ζ21)2(ζ2+1).{\mathcal{A}}_{\rm bh}=\int\sqrt{g^{(C)}_{xx}g^{(C)}_{\phi\phi}}dxd\phi\Bigg{\rvert}_{y=\zeta_{2}}=\frac{4\pi\zeta_{1}\zeta_{2}}{A^{2}(\zeta_{1}-1)(\zeta_{2}-1)^{2}(\zeta_{2}+1)}. (4.2)

The string tension μ\mu is related to the angular deficit δdef\delta_{\rm def} (given in (2.8)) by δdef=8πGμ\delta_{\rm def}=8\pi G\mu. Here we set G=1G=1, giving

μ=αAm=ζ1ζ22(1ζ1)(1ζ2).\mu=\alpha Am=\frac{-\zeta_{1}-\zeta_{2}}{2(1-\zeta_{1})(1-\zeta_{2})}. (4.3)

To compute the change in the area of the acceleration horizon, Δ𝒜acc\Delta{\mathcal{A}}_{\rm acc}, we must compute its area in the C-metric, 𝒜acc(C){\mathcal{A}}_{\rm acc}^{(C)}, and its area in the background, 𝒜acc(B){\mathcal{A}}_{\rm acc}^{(B)}. We find

𝒜acc(C)=gxx(C)gϕϕ(C)dxdϕ|y=1=2παA21xmindx(x+1)2=2παA2(1xmin+112),{\mathcal{A}}_{\rm acc}^{(C)}=\int\sqrt{g^{(C)}_{xx}g^{(C)}_{\phi\phi}}dxd\phi\Bigg{\rvert}_{y=-1}=\frac{2\pi\alpha}{A^{2}}\int^{1}_{x_{\rm min}}\frac{dx}{(x+1)^{2}}=\frac{2\pi\alpha}{A^{2}}\Big{(}\frac{1}{x_{\rm min}+1}-\frac{1}{2}\Big{)}, (4.4)
𝒜acc(B)=gεε(B)gϕϕ(B)dεdϕ|θ=π/2=2παA2εmindεε2=2παA2εmin.{\mathcal{A}}^{(B)}_{\rm acc}=\int\sqrt{g^{(B)}_{\varepsilon\varepsilon}g^{(B)}_{\phi\phi}}d\varepsilon d\phi\Bigg{\rvert}_{\theta=\pi/2}=\frac{2\pi\alpha}{A^{2}}\int^{\infty}_{\varepsilon_{\rm min}}\frac{d\varepsilon}{\varepsilon^{2}}=\frac{2\pi\alpha}{A^{2}\varepsilon_{\rm min}}. (4.5)

As we integrate over the whole acceleration horizon, xmin1x_{\rm min}\to-1 and εmin0\varepsilon_{\rm min}\to 0. Using (3.8) at θ=π/2\theta=\pi/2, we have

xmin=1+εmin(1+εminG′′(1)2G(1)).x_{\rm min}=-1+\varepsilon_{\rm min}\Big{(}1+\varepsilon_{\rm min}\frac{G^{\prime\prime}(-1)}{2G^{\prime}(-1)}\Big{)}. (4.6)

In the εmin0\varepsilon_{\rm min}\to 0 limit, we get

Δ𝒜acc=𝒜acc(C)𝒜acc(B)=2πζ1ζ2(ζ1+ζ2+2)A2(ζ121)(ζ221).\Delta{\mathcal{A}}_{\rm acc}={\mathcal{A}}_{\rm acc}^{(C)}-{\mathcal{A}}^{(B)}_{\rm acc}=\frac{2\pi\zeta_{1}\zeta_{2}(\zeta_{1}+\zeta_{2}+2)}{A^{2}\left(\zeta_{1}^{2}-1\right)\left(\zeta_{2}^{2}-1\right)}. (4.7)

The remaining quantities κacc\kappa_{\rm acc}, κbh\kappa_{\rm bh}, Φ\Phi, and Δ\Delta\ell all depend on the normalization of our time coordinate. As it is not immediately clear what the most natural normalization of tt is, we will define the Killing vector

ξ=ξμμ=Nt\xi=\xi^{\mu}\partial_{\mu}=N\partial_{t} (4.8)

with some constant normalization NN which we leave unspecified for now. Eventually we will see that different choices of NN are suitable in different contexts.

We now compute the surface gravities κacc\kappa_{\rm acc} and κbh\kappa_{\rm bh}. It is interesting to note that ξ\xi becomes null on both the acceleration horizon and the black hole horizon, meaning we can use the same vector ξ\xi to calculate the surface gravity of both horizons. Recall that surface gravity is given by

κ=μVμV,whereV=ξμξμ,\kappa=\sqrt{\nabla_{\mu}V\nabla^{\mu}V},\hskip 14.22636pt\text{where}\hskip 14.22636ptV=\sqrt{-\xi^{\mu}\xi_{\mu}}, (4.9)

evaluated on the horizon in question. This gives

κacc=N2G(1)=N(1+ζ11)(1+ζ21),\kappa_{\rm acc}=\frac{N}{2}G^{\prime}(-1)=N\left(1+\zeta_{1}^{-1}\right)\left(1+\zeta_{2}^{-1}\right), (4.10)
κbh=N2|G(ζ2)|=N(ζ221)(ζ2ζ1)2ζ1ζ2.\kappa_{\rm bh}=\frac{N}{2}|G^{\prime}(\zeta_{2})|=N\frac{\left(\zeta_{2}^{2}-1\right)(\zeta_{2}-\zeta_{1})}{2\zeta_{1}\zeta_{2}}. (4.11)

Next we define the values of the electric potential at the black hole horizon, the acceleration horizon, and spatial infinity:

Φbh\displaystyle\Phi_{\rm bh} =(ξμAμ)|y=ζ2=Nq(ζ2+1)=Nζ2+1Aζ1ζ2,\displaystyle=(\xi^{\mu}A_{\mu})\Big{\rvert}_{y=\zeta_{2}}=Nq(\zeta_{2}+1)=N\frac{\zeta_{2}+1}{A\sqrt{\zeta_{1}\zeta_{2}}}, (4.12)
Φacc\displaystyle\Phi_{\rm acc} =Φ=(ξμAμ)|y=1=0.\displaystyle=\Phi_{\infty}=(\xi^{\mu}A_{\mu})\Big{\rvert}_{y=-1}=0. (4.13)

Despite the vanishing of the latter two with our choice of gauge potential in (2.2), we find it instructive to keep them explicit in some contexts. We also define the gauge-invariant quantity

ΦΦaccΦbh.\Phi\equiv\Phi_{\rm acc}-\Phi_{\rm bh}. (4.14)

Finally, we must discuss the physical interpretation of the “thermodynamic length” Appels:2017xoe . As the string evolves along the timelike Killing vector ξ\xi, it sweeps out a worldsheet in spacetime. We take our thermodynamic length as the rate at which the area of the worldsheet increases as the Killing time progresses. Δ\Delta\ell is the difference between the thermodynamic length of the C-metric and the cosmic string background.

We begin by computing the thermodynamic lengths of the C-metric and cosmic string background separately. We find

(C)=Nζ2ymaxdygtt(C)gyy(C)|x=1=NA2(11+ζ211+ymax),\ell^{(C)}=N\int^{y_{\rm max}}_{\zeta_{2}}dy\sqrt{-g^{(C)}_{tt}g^{(C)}_{yy}}\Bigg{\rvert}_{x=-1}=\frac{N}{A^{2}}\Big{(}\frac{1}{1+\zeta_{2}}-\frac{1}{1+y_{\rm max}}\Big{)}, (4.15)
(B)=Nεmindεgtt(B)gεε(B)|θ=0=NA2εmin.\ell^{(B)}=N\int^{\infty}_{\varepsilon_{\rm min}}d\varepsilon\sqrt{-g^{(B)}_{tt}g^{(B)}_{\varepsilon\varepsilon}}\Bigg{\rvert}_{\theta=0}=\frac{N}{A^{2}\varepsilon_{\rm min}}. (4.16)

Using (3.7) at θ=0\theta=0, we find

ymax=1εmin(1εminG′′(1)2G(1)).y_{\rm max}=-1-\varepsilon_{\rm min}\Big{(}1-\varepsilon_{\rm min}\frac{G^{\prime\prime}(-1)}{2G^{\prime}(-1)}\Big{)}. (4.17)

In the εmin0\varepsilon_{\rm min}\to 0 limit, we then find

Δ=(C)(B)=N2A23+ζ11+ζ1.\Delta\ell=\ell^{(C)}-\ell^{(B)}=-\frac{N}{2A^{2}}\frac{3+\zeta_{1}}{1+\zeta_{1}}. (4.18)

This quantity may seem somewhat obscure, but it is closely related to the Nambu-Goto action of the string and has appeared in a similar context in Herdeiro:2009vd ; KrtousZel .

We have now given a physical interpretation to every quantity in (1.1) and expressed them in terms of the three independent parameters ζ1\zeta_{1}, ζ2\zeta_{2}, and AA (and the arbitrary normalization NN). It can be explicitly verified that (1.1) holds by writing everything in terms of ζ1,ζ2,A\zeta_{1},\zeta_{2},A and applying the chain rule when computing the variations δΔ𝒜acc,δ𝒜bh,δQ,δμ\delta\Delta{\mathcal{A}}_{\rm acc},\delta{\mathcal{A}}_{\rm bh},\delta Q,\delta\mu.

These quantities also satisfy the Smarr relation

0=κacc8πΔ𝒜acc+κbh8π𝒜bh+12ΦQ0=\frac{\kappa_{\rm acc}}{8\pi}\Delta{\mathcal{A}}_{\rm acc}+\frac{\kappa_{\rm bh}}{8\pi}{\mathcal{A}}_{\rm bh}+\frac{1}{2}\Phi Q (4.19)

which essentially follows from (1.1), Euler’s theorem for homogeneous functions, and the scaling dimensions of the quantities involved. μ\mu does not appear as it is dimensionless.

5 Deriving the first law(s) with the covariant phase space formalism

Noetherian methods in the covariant phase space formalism offer an elegant derivation of the first law for Kerr-Newman black holes wald1993black ; bhmech . Given a Killing vector ξ\xi and field variations δgμν\delta g_{\mu\nu} and δAμ\delta A_{\mu} one can construct a 2-form kξk_{\xi} which is closed, satisfying dkξ=0dk_{\xi}=0, away from matter sources. The integral of kξk_{\xi} over a 2-surface gives the variation of the enclosed ξ\xi-charge. This quantity is invariant under continuous deformations of the surface that do not pass through matter. It should be noted that this charge variation is not necessarily integrable — i.e. it is not necessarily the variation of a well-defined finite charge. Integrability must be checked separately. When ξ\xi is a time translation, one refers to the ξ\xi-charge as the energy (or mass). Similarly when ξ\xi is a boost we will refer to the ξ\xi-charge as the “boost mass” umass .

In Einstein-Maxwell gravity this 2-form has two contributions, one “gravitational” and one “electromagnetic.” Explicitly,

kξ=kξgrav+kξemk_{\xi}=k^{\rm grav}_{\xi}+k^{\rm em}_{\xi} (5.1)

where

kξgrav\displaystyle k^{\rm grav}_{\xi} =δKξgrav+KδξgraviξΘgrav,\displaystyle=-\delta K_{\xi}^{\rm grav}+K_{\delta\xi}^{\rm grav}-i_{\xi}\Theta^{\rm grav}, (5.2)
Kξgrav\displaystyle K_{\xi}^{\rm grav} =116π14εμνρσ(μξννξμ)dxρdxσ,\displaystyle=\frac{1}{16\pi}\cdot\frac{1}{4}\varepsilon_{\mu\nu\rho\sigma}(\nabla^{\mu}\xi^{\nu}-\nabla^{\nu}\xi^{\mu})dx^{\rho}\wedge dx^{\sigma}, (5.3)
iξΘgrav\displaystyle i_{\xi}\Theta^{\rm grav} =116π12εμνρσξν(αδgαμμ(gαβδgαβ))dxρdxσ,\displaystyle=-\frac{1}{16\pi}\cdot\frac{1}{2}\varepsilon_{\mu\nu\rho\sigma}\xi^{\nu}(\nabla_{\alpha}\delta g^{\alpha\mu}-\nabla^{\mu}(g_{\alpha\beta}\delta g^{\alpha\beta}))dx^{\rho}\wedge dx^{\sigma}, (5.4)

and

kξem\displaystyle k^{\rm em}_{\xi} =δKξem+KδξemiξΘem,\displaystyle=-\delta K^{\rm em}_{\xi}+K^{\rm em}_{\delta\xi}-i_{\xi}\Theta^{\rm em}, (5.5)
Kξem\displaystyle K^{\rm em}_{\xi} =116π14εμνρσ(FμνξαAα)dxρdxσ,\displaystyle=\frac{1}{16\pi}\cdot\frac{1}{4}\varepsilon_{\mu\nu\rho\sigma}(F^{\mu\nu}\xi^{\alpha}A_{\alpha})dx^{\rho}\wedge dx^{\sigma}, (5.6)
iξΘem\displaystyle i_{\xi}\Theta^{\rm em} =116π12εμνρσξνFαμδAαdxρdxσ.\displaystyle=\frac{1}{16\pi}\cdot\frac{1}{2}\varepsilon_{\mu\nu\rho\sigma}\xi^{\nu}F^{\alpha\mu}\delta A_{\alpha}dx^{\rho}\wedge dx^{\sigma}. (5.7)

Here εμνρσ\varepsilon_{\mu\nu\rho\sigma} is the totally antisymmetric tensor with ε0123=g\varepsilon_{0123}=\sqrt{-g}. The formalism works for generic variations δgμν,δAμ\delta g_{\mu\nu},\delta A_{\mu}, but often one works within a parametrized family of solutions, for example the Kerr-Newman or C-metric families.

The derivation of the first law for Reissner-Nordstrom black holes using covariant phase space methods is instructive, so we briefly outline it here. ξ\xi is chosen to be the horizon generator of the black hole, normalized so that ξ2=1\xi^{2}=-1 at spatial infinity. The physical motivation for this normalization is that ξ\xi coincides with sense of time (i.e. four-velocity) of a distant static observer. In the standard Reissner-Nordstrom coordinates, ξ=t\xi=\partial_{t}.333In this discussion on Reissner-Nordstrom black holes we temporarily redefine ξ\xi and other quantities. For our surface, we take the union of a large two-sphere at spatial infinity and the black hole bifurcation two-sphere with opposite orientations. As this surface is contractible, the integral of kξk_{\xi} over it must vanish. When kξgrav\int k^{\rm grav}_{\xi} is evaluated on the horizon it reduces to κbh8πδ𝒜bh\frac{\kappa_{\rm bh}}{8\pi}\delta{\mathcal{A}}_{\rm bh} and when kξem\int k_{\xi}^{\rm em} is evaluated on the horizon it gives ΦbhδQ-\Phi_{\rm bh}\delta Q. The variation in energy of the spacetime is given by kξ\int k_{\xi} evaluated at spatial infinity. When we use the standard gauge where Φ=0\Phi_{\infty}=0, this integral is equal to δM\delta M where MM is the usual mass of the black hole. From the contractibility of the total surface, these quantities sum to zero, giving the first law for Reissner-Nordstrom black holes.

One might call this the “global” version of the first law, associated with the full spacetime, as opposed to a version “local” to the horizon. Local first laws, described herein, are appropriate to use in contexts where there are multiple horizons or matter present. For pure Reissner-Nordstrom the difference between the global and local first law is immaterial because the sphere on the horizon can be continuously deformed to the sphere at infinity without passing through matter. But if we take a charged black hole with matter outside it (such that we still have a timelike Killing vector) then this is no longer the case. The difference in the surface integral on the horizon and at spatial infinity would be given by integrals on “bubbles” surrounding the parcels of matter. The surface integral at spatial infinity therefore gives the energy variation of the black hole plus that of the external matter. In contrast, the surface integral on the black hole itself gives only the black hole’s energy variation and can be written in terms of quantities local to the black hole such as its area and charge.

All these general comments apply in particular to the C-metric. It has the string as a matter source as well as two horizons. Our main first law (1.1) is of the global type, but there is also a local first law for each horizon. The local first law of the black hole is precisely the one discussed in Astorino . As we will see later, in the small tension limit the results of Reissner-Nordstrom are most easily reproduced using not our global first law but rather the local law of the black hole.

We now turn to the actual derivation of the C-metric’s first laws. We take our system to be the ζ2y1\zeta_{2}\leq y\leq-1 patch, which is bounded by the acceleration horizon, the black hole horizon, and spatial infinity. We can parametrize our phase space of C-metrics with the three numbers ζ1\zeta_{1}, ζ2\zeta_{2}, and AA. For our Killing vector we take ξ=Nt\xi=N\partial_{t} as in (4.8), which is the generator of both horizons. For now we leave the normalization constant NN to be any function on phase space, N=N(ζ1,ζ2,A)N=N(\zeta_{1},\zeta_{2},A). In order to calculate kξk_{\xi} we must specify what our metric and gauge field variations are, using δ\delta for the differential on phase space. For example

δgμν=gμνζ1δζ1+gμνζ2δζ2+gμνAδA.\delta g_{\mu\nu}=\frac{\partial g_{\mu\nu}}{\partial\zeta_{1}}\delta\zeta_{1}+\frac{\partial g_{\mu\nu}}{\partial\zeta_{2}}\delta\zeta_{2}+\frac{\partial g_{\mu\nu}}{\partial A}\delta A. (5.8)

For the global first law, we must construct a two-dimensional surface that is contractible to a point without passing through the string. This surface is depicted in figure 4. Embedded in a constant tt slice, the surface begins by surrounding most of the black hole horizon before connecting to a thin sheath which encases the string. The sheath then connects to a large hemisphere at spatial infinity, which is capped off by the acceleration horizon. This large hemisphere is the surface of constant ε=εmin\varepsilon=\varepsilon_{\rm min} where εmin0\varepsilon_{\rm min}\to 0.

Refer to caption
Figure 4: This surface can be contracted to a point without passing through the string, which is located in the thin sheath. In this diagram ϕ\phi has been suppressed.

We break the integral over this surface into four parts and evaluate them using (5.1) – (5.7). The explicit results in terms of ζ1\zeta_{1}, ζ2\zeta_{2}, AA, δζ1\delta\zeta_{1}, δζ2\delta\zeta_{2}, and δA\delta A are unenlightening, but can be expressed more briefly in terms of quantities defined in section 4. We find

δ̸Mbh\displaystyle\not{\delta}M_{\rm bh} bhkξ=κbh8πδ𝒜bhΦbhδQ,\displaystyle\equiv\int_{\rm bh}k_{\xi}=\frac{\kappa_{\rm bh}}{8\pi}\delta{\mathcal{A}}_{\rm bh}-\Phi_{\rm bh}\delta Q, (5.9)
δ̸Macc(C)\displaystyle\not{\delta}M^{(C)}_{\rm acc} acckξ=κacc8πδ𝒜acc(C)+ΦaccδQ,\displaystyle\equiv\int_{\rm acc}k_{\xi}=\frac{\kappa_{\rm acc}}{8\pi}\delta{\mathcal{A}}_{\rm acc}^{(C)}+\Phi_{\rm acc}\delta Q, (5.10)
δ̸Mhemi(C)\displaystyle\not{\delta}M^{(C)}_{\rm hemi} hemikξ=κacc8πδ𝒜acc(B)(B)δμ,\displaystyle\equiv\int_{\rm hemi}k_{\xi}=-\frac{\kappa_{\rm acc}}{8\pi}\delta{\mathcal{A}}^{(B)}_{\rm acc}-\ell^{(B)}\delta\mu, (5.11)
δ̸Msheath(C)\displaystyle\not{\delta}M_{\rm sheath}^{(C)} sheathkξ=(C)δμ.\displaystyle\equiv\int_{\rm sheath}k_{\xi}=\ell^{(C)}\delta\mu. (5.12)

Each quantity δ̸M\not{\delta}M is meaningful on its own as a variation in boost mass. The local first law of the black hole is (5.9), as in Astorino . Note that the quantities δ̸Mbh\not{\delta}M_{\rm bh}, κbh\kappa_{\rm bh}, 𝒜bh{\mathcal{A}}_{\rm bh}, Φbh\Phi_{\rm bh}, and QQ can all be computed locally on the black hole horizon. Similarly we identify (5.10) as the local first law of the acceleration horizon and note that all of its quantities can be computed locally as well. We identify δ̸Mhemi(C)-\not{\delta}M^{(C)}_{\rm hemi} as the boost mass of the system (with a sign for orientation) since it is the charge evaluated at spatial infinity, in analogy with Reissner-Nordstrom. Both hawkross and umass define their boost masses using this same surface. Lastly δ̸Msheath(C)\not{\delta}M^{(C)}_{\rm sheath} is simply the boost mass contribution of the string. As these quantities sum to zero due to the contractibility of the surface, it appears we are ready to derive our first law. We have

δ̸Mhemi(C)\displaystyle-\not{\delta}M^{(C)}_{\rm hemi} =δ̸Mbh+δ̸Macc(C)+δ̸Msheath(C)\displaystyle=\not{\delta}M_{\rm bh}+\not{\delta}M_{\rm acc}^{(C)}+\not{\delta}M_{\rm sheath}^{(C)} (5.13)
=κbh8πδ𝒜bh+κacc8πδ𝒜acc(C)+ΦδQ+(C)δμ.\displaystyle=\frac{\kappa_{\rm bh}}{8\pi}\delta{\mathcal{A}}_{\rm bh}+\frac{\kappa_{\rm acc}}{8\pi}\delta{\mathcal{A}}^{(C)}_{\rm acc}+\Phi\delta Q+\ell^{(C)}\delta\mu.

The left hand side is the variation of the system’s boost mass, and the right hand side has the usual form of a first law, being a sum of products of conjugate variables. However, our expression as written is not fully satisfactory since the “bare” quantities δ̸Mhemi(C)\not{\delta}M^{(C)}_{\rm hemi}, 𝒜acc(C){\mathcal{A}}^{(C)}_{\rm acc}, and (C)\ell^{(C)} all diverge as εmin0\varepsilon_{\rm min}\to 0. A little rearrangement by subtracting κacc8πδ𝒜acc(B)\frac{\kappa_{\rm acc}}{8\pi}\delta{\mathcal{A}}^{(B)}_{\rm acc} and (B)δμ\ell^{(B)}\delta\mu from both sides would solve this at the level of equations, reproducing (1.1). However, it is preferable to regularize terms in a principled manner so that each term in the equation retains a physical interpretation and in particular that the left hand side can still be considered a boost mass. The natural way to regularize is by comparing each term with the corresponding quantity in the cosmic string background. To that end, we construct a surface analogous to the one in figure 4 in the cosmic string spacetime. The main difference is that, as there is no black hole in this spacetime, the sheath connects directly to the acceleration horizon. We may then evaluate the surface charges just as before, finding

δ̸Macc(B)\displaystyle\not{\delta}M^{(B)}_{\rm acc} acckξ=κacc8πδ𝒜acc(B),\displaystyle\equiv\int_{\rm acc}k_{\xi}=\frac{\kappa_{\rm acc}}{8\pi}\delta{\mathcal{A}}^{(B)}_{\rm acc}, (5.14)
δ̸Mhemi(B)\displaystyle\not{\delta}M^{(B)}_{\rm hemi} hemikξ=κacc8πδ𝒜acc(B)(B)δμ,\displaystyle\equiv\int_{\rm hemi}k_{\xi}=-\frac{\kappa_{\rm acc}}{8\pi}\delta{\mathcal{A}}^{(B)}_{\rm acc}-\ell^{(B)}\delta\mu, (5.15)
δ̸Msheath(B)\displaystyle\not{\delta}M^{(B)}_{\rm sheath} sheathkξ=(B)δμ.\displaystyle\equiv\int_{\rm sheath}k_{\xi}=\ell^{(B)}\delta\mu. (5.16)

As this analogous surface is also contractible, we have the relation

δ̸Mhemi(B)=δ̸Macc(B)+δ̸Msheath(B).-\not{\delta}M_{\rm hemi}^{(B)}=\not{\delta}M_{\rm acc}^{(B)}+\not{\delta}M_{\rm sheath}^{(B)}. (5.17)

We now define our regularized first law by subtracting (5.17) from (5.13) term by term. Defining

δ̸ΔMhemi\displaystyle\not{\delta}\Delta M_{\rm hemi} δ̸Mhemi(C)δ̸Mhemi(B)=0,\displaystyle\equiv\not{\delta}M_{\rm hemi}^{(C)}-\not{\delta}M_{\rm hemi}^{(B)}=0, (5.18)
δ̸ΔMacc\displaystyle\not{\delta}\Delta M_{\rm acc} δ̸Macc(C)δ̸Macc(B)=κacc8πδΔ𝒜acc+ΦaccδQ,\displaystyle\equiv\not{\delta}M_{\rm acc}^{(C)}-\not{\delta}M_{\rm acc}^{(B)}=\frac{\kappa_{\rm acc}}{8\pi}\delta\Delta{\mathcal{A}}_{\rm acc}+\Phi_{\rm acc}\delta Q, (5.19)
δ̸ΔMsheath\displaystyle\not{\delta}\Delta M_{\rm sheath} δ̸Msheath(C)δ̸Msheath(B)=Δδμ,\displaystyle\equiv\not{\delta}M_{\rm sheath}^{(C)}-\not{\delta}M_{\rm sheath}^{(B)}=\Delta\ell\,\delta\mu, (5.20)

the subtracted first law reads

δ̸ΔMhemi\displaystyle-\not{\delta}\Delta M_{\rm hemi} =δ̸ΔMacc+δ̸Mbh+δ̸ΔMsheath\displaystyle=\not{\delta}\Delta M_{\rm acc}+\not{\delta}M_{\rm bh}+\not{\delta}\Delta M_{\rm sheath} (5.21)
0\displaystyle 0 =κacc8πδΔ𝒜acc+κbh8πδ𝒜bh+ΦδQ+Δδμ.\displaystyle=\frac{\kappa_{\rm acc}}{8\pi}\delta\Delta{\mathcal{A}}_{\rm acc}+\frac{\kappa_{\rm bh}}{8\pi}\delta{\mathcal{A}}_{\rm bh}+\Phi\delta Q+\Delta\ell\,\delta\mu.

This completes the derivation of our regularized first law. We may define the left hand side as the regularized boost mass of the spacetime

δΔMboostδ̸ΔMhemi=0.\delta\Delta M_{\rm boost}\equiv-\not{\delta}\Delta M_{\rm hemi}=0. (5.22)

Note that δΔMboost=0\delta\Delta M_{\rm boost}=0 is trivially integrable, yielding ΔMboost=0\Delta M_{\rm boost}=0. This vanishing of the boost mass is consistent with hawkross . It follows directly from the fact that the C-metric and cosmic string spacetimes match at spatial infinity to subleading order in ε\varepsilon.

We may now inspect our first law and try to assign a temperature to the spacetime using the formula T=κ/2πT=\kappa/2\pi. However, we see the appearance of two possible temperatures, one for the acceleration horizon and one for the black hole horizon. The proper way to assign temperature when there are multiple horizons is still an open question SdS . However when the two temperatures are equal, a direct thermodynamic interpretation is possible. This is discussed in section 8. In section 9 we briefly speculate on how to handle temperature with distinct surface gravities.

One may wonder if δ̸Mbh\not{\delta}M_{\rm bh} and δ̸ΔMacc\not{\delta}\Delta M_{\rm acc} can be integrated, as doing so would allow a finite boost mass to be ascribed to the acceleration horizon and black hole horizon. We can use gauge transformations and choices of NN, the normalization of our horizon generator ξ\xi, to do this.

We begin with the acceleration horizon,

δ̸ΔMacc=κacc8πδΔ𝒜acc+ΦaccδQ.\not{\delta}\Delta M_{\rm acc}=\frac{\kappa_{\rm acc}}{8\pi}\delta\Delta{\mathcal{A}}_{\rm acc}+\Phi_{\rm acc}\delta Q. (5.23)

Similar to the case of Reissner-Nordstrom, it is natural to choose the gauge and normalization which would naturally be used by a distant static observer. That is, use the gauge (2.2), for which Φacc=Φ=0\Phi_{\rm acc}=\Phi_{\infty}=0, and normalize time to agree with asymptotic Rindler time, which means κacc=1\kappa_{\rm acc}=1. In this case δ̸ΔMacc\not{\delta}\Delta M_{\rm acc} is trivially integrable as

δ̸ΔMacc=δ(18πΔ𝒜acc).\not{\delta}\Delta M_{\rm acc}=\delta(\frac{1}{8\pi}\Delta{\mathcal{A}}_{\rm acc}). (5.24)

While of questionable physical content, we believe this is the statement that most closely comprises a first law of the acceleration horizon. Note that since the regularized hemisphere charge vanishes, ΔMacc\Delta M_{\rm acc} also equals the boost mass of the black hole plus string subsystem. Since the black hole and string are coupled we cannot generally write it as a sum of two separate boost masses.

We now turn to the black hole horizon, with

δ̸Mbh=κbh8πδ𝒜bhΦbhδQ.\not{\delta}M_{\rm bh}=\frac{\kappa_{\rm bh}}{8\pi}\delta{\mathcal{A}}_{\rm bh}-\Phi_{\rm bh}\delta Q. (5.25)

The perspective of a distant static observer is not integrable here, and the somewhat trivial choice κbh=1\kappa_{\rm bh}=1, Φbh=0\Phi_{\rm bh}=0 just reproduces the area 𝒜bh{\mathcal{A}}_{\rm bh}. There are no other clear natural choices of normalization NN or gauge. If we allow ourselves complete freedom in choosing them, we have the power to not only make δ̸Mbh\not{\delta}M_{\rm bh} integrable, but to make MbhM_{\rm bh} any function of 𝒜bh{\mathcal{A}}_{\rm bh} and QQ. Explicitly, given any Mbh(𝒜bh,Q)M_{\rm bh}({\mathcal{A}}_{\rm bh},Q), all we need to do is choose an NN such that κbh=8πMbh𝒜bh\kappa_{\rm bh}=8\pi\frac{\partial M_{\rm bh}}{\partial{\mathcal{A}}_{\rm bh}} and choose the gauge so that Φbh=MbhQ\Phi_{\rm bh}=-\frac{\partial M_{\rm bh}}{\partial Q}.

If we want our mass to be consistent with the Reissner-Nordstrom black hole, then we can choose MbhM_{\rm bh} to satisfy the Christodolou-Ruffini formula

Mbh(𝒜bh,Q)=𝒜bh/π4(1+4πQ2𝒜bh).M_{\rm bh}({\mathcal{A}}_{\rm bh},Q)=\frac{\sqrt{{\mathcal{A}}_{\rm bh}/\pi}}{4}\left(1+\frac{4\pi Q^{2}}{{\mathcal{A}}_{\rm bh}}\right). (5.26)

This is essentially the point of view taken by Astorino (see also acovErnst ; AstReg ). In our view this result is something one has chosen rather than found unless the choice of normalization and gauge were physically motivated. However, as we will see, when the string tension is small there is a natural choice that recovers the Reissner-Nordstrom mass.

6 The small string tension limit

Recall the charged C-metric is specified by three parameters. Two can be taken dimensionless, such as ζ1,ζ2\zeta_{1},\zeta_{2}, but the third parameter, such as AA, must be dimensionful. Let us now take the two dimensionless parameters to be μ\mu and ρ\rho, where ρ\rho is the “charge to mass ratio”

ρq/m.\rho\equiv q/m. (6.1)

It is possible to express ζ1\zeta_{1} and ζ2\zeta_{2} as functions of μ\mu and ρ\rho.

In this section we explore the limit of small μ\mu when ρ\rho and AA remain finite. Since μ=αAm\mu=\alpha Am and α=1+𝒪(μ)\alpha=1+{\mathcal{O}}(\mu), small μ\mu is equivalent to

Am1.Am\ll 1. (6.2)

Physically this means that the black hole’s length scale mm is small compared to the acceleration length scale A1A^{-1}. Therefore the black hole can be understood as a pointlike Rindler particle from the perspective of a distant observer kinwalk . To an observer close to the black hole, however, the black hole is barely accelerating and resembles a Reissner-Nordstrom black hole at rest. These two notions of scale, one far from the black hole and one intermediate, lead to different notions of the mass of the black hole.444 In fact, when μ0\mu\to 0, the C-metric (2.1) can be shown to reduce to the Minkowski metric and the Reissner-Nordstrom metric using different limiting procedures. If we fix A,ρA,\rho and send μ0\mu\to 0 the resulting metric describes Minkowski space. If we fix m,ρm,\rho and switch to tRNt/At_{RN}\equiv t/A, and send μ0\mu\to 0, the resulting metric describes a Reissner-Nordstrom black hole of mass mm and charge qq.

Refer to caption
Figure 5: a) In the “far region” outside the dotted lines the black hole appears pointlike, tracing out a hyperbolic trajectory in time. b) In the “intermediate region” outside the dotted lines the black hole still appears pointlike, but has negligible acceleration at this scale.

With our choice of gauge in (2.2), where Φacc=0\Phi_{\rm acc}=0, the local first law of the black hole reduces at small μ\mu to

δ̸Mbh=κbh8πδΔ𝒜bhΦbhδQN(μA3δA+1A2δμ).\not{\delta}M_{\rm bh}=\frac{\kappa_{\rm bh}}{8\pi}\delta\Delta{\mathcal{A}}_{\rm bh}-\Phi_{\rm bh}\delta Q\approx N\left(-\frac{\mu}{A^{3}}\delta A+\frac{1}{A^{2}}\delta\mu\right). (6.3)

We will use this equation to study both the point particle and Reissner-Nordstrom limits. First consider the perspective of a distant observer where the black hole appears pointlike. In Minkowski space with coordinates (X0,X1,X2,X3)(X^{0},X^{1},X^{2},X^{3}) the canonical boost generator is

K=X30+X03K=X^{3}\partial_{0}+X^{0}\partial_{3} (6.4)

and a Rindler point particle of mass mptm_{\rm pt} and acceleration AptA_{\rm pt} has boost mass

Mpt=d3xKμTμ0=mptApt.M_{\rm pt}=\int d^{3}x\,K^{\mu}T_{\mu 0}=\frac{m_{\rm pt}}{A_{\rm pt}}. (6.5)

To compare with the black hole we start by normalizing ξ\xi to agree with KK asymptotically. This means κacc=1\kappa_{\rm acc}=1, or N1N\approx 1. Now at leading order in μ\mu we have

δ̸MbhμA3δA+1A2δμ.\not{\delta}M_{\rm bh}\approx-\frac{\mu}{A^{3}}\delta A+\frac{1}{A^{2}}\delta\mu. (6.6)

Although δ̸Mbh\not{\delta}M_{\rm bh} is not integrable, MbhM_{\rm bh} can be perturbatively defined around μ=0\mu=0. For any finite AA, we know μ=0\mu=0 corresponds to Minkowski space and so the (absent) black hole’s mass vanishes. If we increase μ\mu slightly from zero then the boost mass is

MbhμA2mA.M_{\rm bh}\approx\frac{\mu}{A^{2}}\approx\frac{m}{A}. (6.7)

This agrees with the boost mass of the point particle. Note we can reproduce this answer using ΔMacc\Delta M_{\rm acc}, which equals the combined boost mass of the black hole and string. One can check that for small μ\mu,

ΔMaccmA+μΔ\Delta M_{\rm acc}\approx\frac{m}{A}+\mu\Delta\ell (6.8)

where the first term is the boost mass of the black hole and the second term is the boost mass of the string.

Now we turn to the perspective where we have zoomed in on the black hole. This region is characterized by y1y\ll-1, meaning the observer is much closer to the black hole than the acceleration horizon. However, if we also demand y/ζ21y/\zeta_{2}\ll 1 then the observer will also be much further away from the black hole than the length scale mm. The region described by both limits is called the “intermediate region,” and can be thought of as the analogue of the region far from the black hole in the pure Reissner-Nordstrom spacetime. As usual we normalize by matching ξ\xi with the four-velocity of our static observer, which here requires NAN\approx A. This physically motivated choice of NN yields

δ̸MbhμA2δA+1Aδμδ(μA)δm,\not{\delta}M_{\rm bh}\approx-\frac{\mu}{A^{2}}\delta A+\frac{1}{A}\delta\mu\approx\delta\left(\frac{\mu}{A}\right)\approx\delta m, (6.9)
Mbhm\implies M_{\rm bh}\approx m (6.10)

which of course agrees with Reissner-Nordstrom. The black hole surface gravity, with N=AN=A in the small μ\mu limit, equals

κbhm2q2(m+m2q2)2,\kappa_{\rm bh}\approx\frac{\sqrt{m^{2}-q^{2}}}{(m+\sqrt{m^{2}-q^{2}})^{2}}, (6.11)

which also agrees with Reissner-Nordstrom. In fact, Φ\Phi and 𝒜bh{\mathcal{A}}_{\rm bh} also reduce to their Reissner-Nordstrom values in the small μ\mu limit, meaning the whole first law is reproduced.

7 The Euclidean C-metric

Comparison of our thermodynamic quantities with those found semiclassically from the Euclidean path integral offers a robust check on both our results and their interpretation. In particular we will compare the thermodynamic partition function with an approximation of the Euclidean path integral partition function. Euclidean solutions of the equations of motion, also known as instantons, represent saddle points in the finite temperature gravitational path integral. Therefore their action approximates the log partition function gibbons1977action :

logZIE.-\log Z\approx I_{E}. (7.1)

Before discussing the partition function via the Euclidean path integral, we must review the geometry of the Euclidean C-metric. Performing a Wick rotation τ=it\tau=it on the Lorentzian C-metric, we obtain the Euclidean C-metric

ds2=1A2(xy)2(G(y)dτ2dy2G(y)+dx2G(x)+α2G(x)dϕ2).ds^{2}=\frac{1}{A^{2}(x-y)^{2}}\Big{(}-G(y)d\tau^{2}-\frac{dy^{2}}{G(y)}+\frac{dx^{2}}{G(x)}+\alpha^{2}G(x)d\phi^{2}\Big{)}. (7.2)

It is electrovacuum almost everywhere, with gauge field

Aμdxμ=iq(y+1)dτ.A_{\mu}dx^{\mu}=-iq(y+1)d\tau. (7.3)

Previously y=ζ2y=\zeta_{2} and y=1y=-1 were horizons, but now they are more like tips of cones. In particular, with ζ2<y<1\zeta_{2}<y<-1 the Euclidean C-metric (7.2) is geodesically complete. Note then G(y)<0G(y)<0, so it is Euclidean everywhere, justifying the name. The topology of the Euclidean C-metric is S2×S2{pt}S^{2}\times S^{2}-\{\mathrm{pt}\}. One S2S^{2} factor is parameterized by (y,τ)(y,\tau) and the other by (x,ϕ)(x,\phi). The removed point “pt\mathrm{pt}” is x=y=1x=y=-1 and corresponds to spatial infinity. The metric retains the conical singularity sourced by the cosmic string at x=1x=-1.555The conical singularity on the string must remain in the Euclidean metric because it is physically sourced by the string. This is necessary for the metric to solve the Euclidean equations of motion and thus be a saddle point of the path integral. See hawkross . However, since the Lorentzian C-metric has no stress-energy source on the horizons (away from the string), it seems most natural to take the stress-energy to vanish at y=ζ2y=\zeta_{2} and y=1y=-1 in the Euclidean solution as well (away from the string). That is, we would like to make τ\tau periodic such that the “cones” are actually disks. We can avoid a conical singularity at y=ζ2y=\zeta_{2} if we give τ\tau a periodicity ττ+β\tau\sim\tau+\beta, where

β=4π|G(ζ2)|=N2πκbh.\beta=\frac{4\pi}{|G^{\prime}(\zeta_{2})|}=N\frac{2\pi}{\kappa_{\rm bh}}. (7.4)

However, a conical singularity will remain at y=1y=-1 unless we have

|G(ζ2)|=G(1).|G^{\prime}(\zeta_{2})|=G^{\prime}(-1). (7.5)

The above equation is called the temperature matching condition, and has been imposed in hawkross ; stromzhib . It is equivalent to

ζ2ζ1=2\zeta_{2}-\zeta_{1}=2 (7.6)

assuming we require ζ21\zeta_{2}\neq-1 so the acceleration horizon and black hole horizon remain distinct. The temperature matching condition fixes a parameter, bringing the number of free parameters of the C-metric down from three to two. It can also be rewritten as

A=m2q2q2A=\frac{\sqrt{m^{2}-q^{2}}}{q^{2}} (7.7)

which shows that, assuming AA and mm are finite, the temperature matching condition can only be satisfied if q2>0q^{2}>0, i.e. if the black holes are charged.

One final way to express the temperature matching condition is

κbh=κacc.\kappa_{\rm bh}=\kappa_{\rm acc}. (7.8)

The physical interpretation of the temperature matching condition is that the rate at which the black hole is fed Unruh radiation exactly equals the rate at which it emits Hawking radiation, thus rendering the C-metric quantum mechanically stable.

The Euclidean C-metric can be pictured by taking the t=0t=0 slice of the Lorentzian C-metric and rotating it around the y=1y=-1 and y=ζ2y=\zeta_{2} axis by the angle τ\tau, as in figure 6.

Refer to caption
Figure 6: The Euclidean C-metric can be obtained by taking a t=0t=0 slice of the Lorentzian C-metric and sweeping it around by the periodic Euclidean time τ\tau. Here we have only drawn the x=1,1x=1,-1 slice of the full three-dimensional t=0t=0 geometry.

We now compute the action of the Euclidean C-metric. The Euclidean action has three components, coming from the gravitational field, the electromagnetic field, and the string:

IE=IEgrav+IEem+IEstring,I_{E}=I_{E}^{\rm grav}+I_{E}^{\rm em}+I_{E}^{\rm string}, (7.9)
IEgrav\displaystyle I_{E}^{\rm grav} =116πd4xgR18πd3xhK,\displaystyle=-\frac{1}{16\pi}\int d^{4}x\sqrt{g}R-\frac{1}{8\pi}\int d^{3}x\sqrt{h}K, (7.10)
IEem\displaystyle I_{E}^{\rm em} =116πd4xgFμνFμν,\displaystyle=-\frac{1}{16\pi}\int d^{4}x\sqrt{g}F^{\mu\nu}F_{\mu\nu}, (7.11)
IEstring\displaystyle I_{E}^{\rm string} =μd2xγ.\displaystyle=\mu\int d^{2}x\sqrt{\gamma}. (7.12)

This action is evaluated out to a large surface approaching spatial infinity. The surface is defined by ε=εmin\varepsilon=\varepsilon_{\rm min} where εmin0\varepsilon_{\rm min}\to 0. hh is the induced metric on this surface and KK is its extrinsic curvature. γ\gamma is the induced metric on the string worldsheet. As usual for a non-compact spacetime, the naïve action diverges so we will actually compute the difference in action relative to a background. In this case we take as our background the Euclidean cosmic string spacetime.

We begin with the gravitational term. For the C-metric, R=0R=0 everywhere except on the string where it has a delta function singularity. When integrating over a conical singularity in a two-dimensional manifold, one may use the Gauss-Bonnet theorem to show that d2xgR=2δdef\int d^{2}x\;\sqrt{g}R=2\delta_{\rm def}, where here δdef\delta_{\rm def} is the deficit angle of the singularity. It may then be shown that the Einstein-Hilbert action of a string-like singularity in four dimensions is given by d4xgR=2𝒜δdef\int d^{4}x\sqrt{g}R=2{\mathcal{A}}\delta_{\rm def}, where 𝒜{\mathcal{A}} is the area that the cusp sweeps out in the transverse directions. For the cosmic string in the Euclidean C-metric, this area can be expressed in terms of the thermodynamic length as

116πd4xgR=βμ1N(C).-\frac{1}{16\pi}\int d^{4}x\sqrt{g}R=-\beta\mu\tfrac{1}{N}\ell^{(C)}. (7.13)

The expression for the Euclidean cosmic string background metric is analogous.

The Gibbons-Hawking-York boundary term is evaluated on the ε=εmin\varepsilon=\varepsilon_{\rm min} surface where εmin0\varepsilon_{\rm min}\to 0. The difference in this boundary term between the C-metric and cosmic string background is zero. Similar to the vanishing of δ̸ΔMhemi\not{\delta}\Delta M_{\rm hemi} in (5.18), this is a consequence of the agreement of the C-metric and the cosmic string background metric at subleading order in ε\varepsilon. The total gravitational portion of the difference in Euclidean action is then

ΔIEgrav=βμ1NΔ.\Delta I_{E}^{\rm grav}=-\beta\mu\tfrac{1}{N}\Delta\ell. (7.14)

Next, the contribution from the Nambu-Goto string action may also be expressed using the thermodynamic length. One finds

ΔIEstring=βμ1NΔ.\Delta I_{E}^{\rm string}=\beta\mu\tfrac{1}{N}\Delta\ell. (7.15)

Note that this cancels out the contribution from the conical singularity sourced by the string in the Einstein-Hilbert action. A similar cancellation was noted in Almheiri:2019qdq in the context of replica wormholes.

Finally we come to the electromagnetic term. Using the equation of motion μFμν=0\nabla_{\mu}F^{\mu\nu}=0, we can write the electromagnetic Lagrangian as a total derivative: FμνFμν=μ(2FμνAν)F^{\mu\nu}F_{\mu\nu}=\nabla_{\mu}(2F^{\mu\nu}A_{\nu}). Therefore the action may be expressed as a boundary integral. However, there is a small issue: the coordinate τ\tau is ill-defined at y=1y=-1 and y=ζ2y=\zeta_{2}. This makes the gauge potential singular at those locations unless it is zero. Our choice of gauge in (7.3) makes the potential vanish at y=1y=-1, meaning the boundary integral at spatial infinity vanishes as well, but the gauge potential remains singular at y=ζ2y=\zeta_{2}. To account for this, we must create an inner boundary of integration on a surface of constant yy and push it up against y=ζ2y=\zeta_{2}. This gives a final result of

IEem=12β1NΦQ=πκbhΦQ.I_{E}^{\rm em}=-\frac{1}{2}\beta\tfrac{1}{N}\Phi Q=-\frac{\pi}{\kappa_{\rm bh}}\Phi Q. (7.16)

There is no contribution to the electromagnetic portion of the action from the background cosmic string action.

The total action is then666One finds the same action by borrowing the result from hawkross for the magnetically charged C-metric and adding the term ΦQ/T-\Phi Q/T to it, the necessity of which is explained in HawkRossDuality .

ΔIE=πκbhΦQ.\Delta I_{E}=-\frac{\pi}{\kappa_{\rm bh}}\Phi Q. (7.17)

8 The temperature-matched thermodynamic interpretation

In our first law (1.1) we vary three parameters. We can spend one to set κacc=κbh\kappa_{\rm acc}=\kappa_{\rm bh}, leaving us with two.777This condition holds for black holes produced via quantum tunneling on a cosmic string hawkross . Now that both horizons emit radiation at the same temperature the system should be in equilibrium and we may attempt a full thermodynamic interpretation. That is, we write

δΔMboost=TδΔS+ΦδQ+Δδμ\delta\Delta M_{\rm boost}=T\delta\Delta S+\Phi\delta Q+\Delta\ell\,\delta\mu (8.1)

where

T\displaystyle T =κacc2π=κbh2π,\displaystyle=\frac{\kappa_{\rm acc}}{2\pi}=\frac{\kappa_{\rm bh}}{2\pi}, (8.2)
ΔS\displaystyle\Delta S =14(Δ𝒜acc+𝒜bh),\displaystyle=\frac{1}{4}\left(\Delta{\mathcal{A}}_{\rm acc}+{\mathcal{A}}_{\rm bh}\right), (8.3)
ΔMboost\displaystyle\Delta M_{\rm boost} =0.\displaystyle=0. (8.4)

We still have the Smarr relation

0=TΔS+12ΦQ.0=T\Delta S+\frac{1}{2}\Phi Q. (8.5)

Note that ΔS<0\Delta S<0, but as it is the change in entropy compared to the cosmic string background this is not a problem.

Using these quantities, we can write the standard thermodynamic grand potential of the system, ΔΩ\Delta\Omega, as

ΔΩ\displaystyle\Delta\Omega =ΔMboostTΔSΦQ\displaystyle=\Delta M_{\rm boost}-T\Delta S-\Phi Q (8.6)
=12ΦQ,\displaystyle=-\frac{1}{2}\Phi Q,

where we used the Smarr relation to get the second line. This is simply related to the thermodynamic partition function by

logZthermo=ΔΩ/T.-\log Z_{\rm thermo}=\Delta\Omega/T. (8.7)

We are now in a position to confirm that this thermodynamic partition function agrees with the semiclassical one, (7.1). The check boils down to

ΔIE=?ΔΩ/T.\Delta I_{E}\stackrel{{\scriptstyle?}}{{=}}\Delta\Omega/T. (8.8)

This equality does indeed hold by (7.17), (8.6), and (8.2). This agreement with semiclassical methods offers a strong endorsement for the thermodynamic interpretation of the quantities associated with boost time.

Let us now turn to the question of what the unambiguous temperature of the C-metric is by finally choosing a normalization NN of the boost Killing vector ξ\xi. The most natural choice matches the norm of ξ\xi to that of a boost in the cosmic string background asymptotically. This is accomplished through N=2G(1)N=\frac{2}{G^{\prime}(-1)}, or equivalently κacc=1\kappa_{\rm acc}=1.888The authors of umass similarly required κacc=1\kappa_{\rm acc}=1, but in their C-metric example they took t\partial_{t} as their Killing vector rather than allowing for a general normalization. Accordingly they needed to fix one of their C-metric parameters to impose κacc=1\kappa_{\rm acc}=1, rather than merely fixing the normalization as we do. Therefore, the temperature is always

T=12π.T=\frac{1}{2\pi}. (8.9)

It is dimensionless because boost time is dimensionless. It is somewhat strange to have a first law with a fixed temperature, but this can be viewed as a natural consequence of the non-compactness of the acceleration horizon. It extends far away from the black hole, where the metric is nearly flat, and so its temperature must be consistent with the flat case of Rindler. This point about fixed periodicity was also made in hawkhor .

9 Speculation on firewalls

In this section we speculate on how to treat temperature when the two horizons have different surface gravities. As we mentioned before, this is still an open question SdS . Most attempts to answer it offer a single temperature Shankaranarayanan ; PappasKanti ; TopoTemp . However, here we explore the idea that a continuum of temperatures may be possible. Even for a spacetime with a single horizon, it is not unheard of to have a temperature other than the Hawking temperature TH=κ2πT_{\rm H}=\frac{\kappa}{2\pi}. One example is the Boulware vacuum for the Schwarzschild spacetime, which has zero temperature for static observers Candelas . It is well-known that the stress-energy tensor in the Boulware vacuum diverges near the horizon. This signifies the existence of a firewall, which is related to the lack of entanglement across the horizon. The existence of a firewall can also be seen using the Euclidean section of the spacetime — if we use a different periodicity for Euclidean time, there will be a conical singularity at the horizon representing a delta-function stress-energy tensor from the firewall. If we treat the singularity as a physical membrane, the action can be computed just as it was for the cosmic string.

For the C-metric, if we do not match the surface gravities of the black hole and acceleration horizons then there is necessarily a firewall on at least one horizon. We will consider an arbitrary temperature TT, in which case there are generically firewalls on both horizons. If we introduce action terms for the firewall membranes,

IEfire,acc=σaccd2xγ,IEfire,bh=σbhd2xγ,I_{E}^{\rm fire,\,acc}=\sigma_{\rm acc}\int d^{2}x\sqrt{\gamma},\qquad I_{E}^{\rm fire,\,bh}=\sigma_{\rm bh}\int d^{2}x\sqrt{\gamma}, (9.1)

where σacc\sigma_{\rm acc} and σbh\sigma_{\rm bh} are the firewall energy densities

σacc14(1κacc2πT),σbh14(1κbh2πT),\sigma_{\rm acc}\equiv\frac{1}{4}\left(1-\frac{\kappa_{\rm acc}}{2\pi T}\right),\qquad\sigma_{\rm bh}\equiv\frac{1}{4}\left(1-\frac{\kappa_{\rm bh}}{2\pi T}\right), (9.2)

we find that they exactly cancel the contributions of the horizon conical singularities to the Einstein-Hilbert action. Building on the results of section 7, we therefore have

IEgrav+IEfire,acc+IEfire,bh+IEstring=0.I_{E}^{\rm grav}+I_{E}^{\rm fire,\,acc}+I_{E}^{\rm fire,\,bh}+I_{E}^{\rm string}=0. (9.3)

The electromagnetic action is unaffected by the firewalls, and the temperature simply appears as an overall factor. The total Euclidean action is

IE=IEem=12ΦQ/T.I_{E}=I_{E}^{\rm em}=-\frac{1}{2}\Phi Q/T. (9.4)

The grand potential is then

ΔΩ=TΔIE\displaystyle\Delta\Omega=T\Delta I_{E} =12ΦQ\displaystyle=-\frac{1}{2}\Phi Q (9.5)
=κacc8πΔ𝒜acc+κbh8π𝒜bh\displaystyle=\frac{\kappa_{\rm acc}}{8\pi}\Delta{\mathcal{A}}_{\rm acc}+\frac{\kappa_{\rm bh}}{8\pi}{\mathcal{A}}_{\rm bh}
=TΔSTσaccΔ𝒜accTσbh𝒜bh.\displaystyle=T\Delta S-T\sigma_{\rm acc}\Delta{\mathcal{A}}_{\rm acc}-T\sigma_{\rm bh}{\mathcal{A}}_{\rm bh}.

We assume the entropy is still ΔS=14(Δ𝒜acc+𝒜bh)\Delta S=\frac{1}{4}(\Delta{\mathcal{A}}_{\rm acc}+{\mathcal{A}}_{\rm bh}). Using these definitions, the first law (5.21) can be rewritten as

0=TδΔSTσaccδΔ𝒜accTσbhδ𝒜bh+ΦδQ+Δδμ.0=T\delta\Delta S-T\sigma_{\rm acc}\delta\Delta{\mathcal{A}}_{\rm acc}-T\sigma_{\rm bh}\delta{\mathcal{A}}_{\rm bh}+\Phi\delta Q+\Delta\ell\,\delta\mu. (9.6)

In this form there is a single temperature TT along with energy contributions from the new firewalls.

10 Conclusion

We studied the thermodynamics of the C-metric with respect to canonical charges defined by boost time. As the C-metric has two horizons and a cosmic string, we drew a distinction between global and local first laws. We derived a global first law for the spacetime in terms of a change in boost mass ΔMboost\Delta M_{\rm boost} which we found to vanish. We then showed that the local first law of the black hole horizon reproduced the Reissner-Nordstrom first law in the small string tension limit. Upon matching the temperatures of the acceleration horizon and the black hole horizon, we were able to assign traditional thermodynamic interpretations to the quantities in the global first law. We showed agreement between the thermodynamic and path integral partition functions which served as an independent check on the thermodynamic interpretation. For all of these reasons, we believe that (1.1) can rightfully be called the first law of the C-metric.

Acknowledgements.
We are grateful to Andy Strominger for acquainting us with the C-metric, and for his generous guidance. We also thank Kevin Nguyen, Jakob Salzer, and Rudranil Basu for useful discussions. AB gratefully acknowledges support from NSF grant 1707938 and the Fundamental Laws Initiative. NM gratefully acknowledges support from NSF GRFP grant DGE1745303.

References