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

The First Law for the Kerr-NUT Spacetime

Ernesto Frodden [email protected] Departamento de Física, Universidad de Santiago de Chile, Avenida Ecuador 3493, Santiago, Chile.    Diego Hidalgo [email protected] Centro de Estudios Científicos (CECs), Av. Arturo Prat 514, Valdivia, Chile
Departamento de Física, Universidad de Concepción, Casilla 160-C, Concepción, Chile,
Instituto de Ciencias Físicas y Matemáticas, Universidad Austral de Chile, Casilla 567, Valdivia, Chile.
Abstract

We show that the first law for the Kerr-NUT is straightforwardly established with the surface charge method. The entropy is explicitly found as a charge and its value is not proportional to the horizon area. We conclude that there are unavoidable contributions from the Misner strings to the charges, still the mass and angular momentum gets standard values. However, there are no independent charges associated with the Misner strings.

Surface Charges, Misner Strings

I Introduction

Taub-NUT (TN) spacetime is a vacuum solution of General Relativity Taub ; NUT . For a long time, the physical interpretation of TN spacetime has been elusive. This spacetime contains unusual properties that make it attractive for study and constitute a typical counter-example for weakly established statements in the field (think about the area law in black hole physics or standardized asymptotic behavior for spacetimes). The characterizing feature of TN spacetime is that the additional parameter, the NUT parameter nn, causes string-like defects Misner:1963fr , nowadays called Misner strings. These strings are interpreted as the gravitational analogue of the Dirac strings appearing for magnetic monopoles in the electromagnetic context. Thus, the NUT property is further interpreted as a gravitomagnetic charge, i.e. as a charge dual to the mass in the same vein of electromagnetic duality.

There are at least two remarkable interpretations of the TN solution. The first one, presented by Misner Misner:1963fr , argued that the string-like defect could be made invisible if the time coordinate is made periodic. The second interpretation, employed by Bonnor bonnor1969 , keeps the strings as a certain “massless source of angular momentum.”

Most of the studies on TN spacetimes is done in the Euclidean regime Chamblin:1998pz ; Ghezelbash:2008zz ; Emparan:1999pm ; Mann:1999pc ; Mann:1999bt ; Garfinkle:2000ms ; Johnson:2014xza ; Johnson:2014pwa ; Clarkson:2002uj ; Astefanesei:2004kn ; Kalamakis:2020aaj ; Ciambelli:2020qny ; Arratia:2020hoy ; Hawking:1998jf ; Hawking:1998ct because gravitational instantons Hawking:1976jb might play a role in the path integral formulation of gravity. However, the Lorentzian version of TN spacetime was critically revisited Clement:2015cxa ; Clement:2015aka . The authors considered the geodesics motion in the presence of Misner strings. The main result is that TN spacetime turns out to be geodesically complete: i.e Free-falling observers do not see strings as they cross them. Guided by this last analysis, renewed interest in TN spacetime emerged in the community as there is a window for the TN solutions to have physical relevance Kubiznak:2019yiu ; Gonzalez:2017sfq ; Cano:2021qzp ; BallonBordo:2019vrn ; Bordo:2019slw ; Durka:2019ajz ; Bordo:2019rhu ; Bordo:2019tyh ; Zhang:2021pvx .

In this work, we pursue a consistent formulation of thermodynamics for the Kerr-Nut spacetime, also known as rotating TN spacetime in the Lorentzian regime. This solution is also referred to as the Kerr NUT solution because it is a generalization of the Kerr black hole solution including the NUT parameter. The final hope is to settle down a thermodynamics interpretation for the TN spacetime to increase the knowledge about this maybe-physical solution and may help to have a new arena where quantum gravity ideas might be explored. As we will see, the spacetime entropy for this solution is different to the usual black holes.

Recently, Kubiznak:2019yiu , it has been claimed that the first law for Lorentzian rotating TN spacetimes, Kerr-NUT, can be formulated assuming: 1) An entropy given by the one-four of the area of the horizon (area law), and 2) the existence of independently NUT parameters that control the effect of Misner strings at the first law.

In this work, we review how the Misner strings enter in the formulation of the first law. Using the surface charge method Frodden:2019ylc . The TN charges: Mass and angular momentum, are derived with a suitable spacetime integration. With these conserved quantities, the variation of the entropy is naturally obtained as a charge too. This procedure, plus integrability considerations, lead us to restrict the parameter space and find an entropy that differs from the area law, in accordance with Hawking:1998jf .

Our paper is organized as follows: In Section II, we briefly review basic facts about the Kerr-NUT. The surface charge method improved to deal with string-like defects is presented in Section III. Section IV is devoted to constructing the first law of thermodynamic for TN solution and find the appropriate entropy that fits on it. As a consistency check of the results, in Subsection IV.1, we show that all exposed quantities and the first law can be framed in the grand canonical ensemble. Summary and final discussion are presented in Section V. For completeness we include an appendix exhibiting intermediate steps of the computations.

II Kerr-NUT Solution

The Kerr-NUT, or rotating TN spacetime, generalizes the Kerr metric, with an extra NUT parameter nn as

ds2=ΔΣ(dt+(2ncosθ+2Cnasin2θ)dϕ)2+ΣΔdr2\displaystyle ds^{2}=-\frac{\Delta}{\Sigma}\left(dt+(2n\cos\theta+2Cn-a\sin^{2}\theta)d\phi\right)^{2}+\frac{\Sigma}{\Delta}dr^{2}
+sin2θΣ(adt(r2+a2+n22anC)dϕ)2+Σdθ2,\displaystyle+\,\frac{\sin^{2}\theta}{\Sigma}\left(adt-(r^{2}+a^{2}+n^{2}-2anC)d\phi\right)^{2}+\Sigma\,d\theta^{2}\,, (1)

with

Δ\displaystyle\Delta =\displaystyle= r2+a22mrn2,\displaystyle r^{2}+a^{2}-2mr-n^{2}\,, (2)
Σ\displaystyle\Sigma =\displaystyle= r2+(n+acosθ)2,\displaystyle r^{2}+(n+a\cos\theta)^{2}\,, (3)

where mm is the integration constant associated with the mass, nn the NUT parameter, and aa the rotation parameter. The solution with C=0C=0 is a particular Demianski-Newman vacuum spacetime metriccite . The parameter C=±1,0C=\pm 1,0 was first introduced in Manko:2006bf . The solution presents Misner strings as (t)2(\nabla t)^{2} becomes ill-defined along the axis Misner:1963fr . The discrete parameter CC specifies different configurations for the strings Manko:2006bf : For C=+1C=+1 one string is at the north pole axis while the south pole axis is regular, for C=1C=-1 is the opposite situation, and for C=0C=0 there are two strings symmetrically located.

The solution has a Killing horizon BallonBordo:2019vrn ; podolskybook located at the largest root of Δ=0\Delta=0 and it is generated by the Killing vector ξ=t+Ωϕ\xi=\partial_{t}+\Omega\,\partial_{\phi}, with Ω\Omega the angular velocity of a ZAMO (zero angular momentum observer) at the horizon as seen by a static observer at infinity.

III Surface charges for spacetimes with strings

To characterize TN spacetime we can compute charges associated with isometries. General Relativity is a gauge theory and a naive use of the First Noether Theorem produces misleading charges. An improved method to extract charges is required. We use the standard surface charge method Lee:1990nz ; Iyer:1994ys ; Wald:1999wa ; for isometries it is equivalent to the symplectic method explained in Barnich:2001jy (both methods were revisited in Frodden:2019ylc ).

An exact symmetry (here equivalent to a Killing symmetry) that is generated by a Killing vector field ξ\xi produces an on-shell conservation law equation: d𝒌ξ0d\boldsymbol{k}_{\xi}\approx 0. The integration of this equation over a spacetime volume Σ\Sigma, with two disconnected boundaries Σ=SS\partial\Sigma=S\cup S^{\prime}, defines a conserved surface charge variation as

δ̸Qξ=S𝒌ξ,\not{\delta}Q_{\xi}=\oint_{S}\boldsymbol{k}_{\xi}\,, (4)

the conservation is understood as the fact that for any other closed two-surface SS^{\prime}, obtained from a continuous deformation of SS; the previous expression produces the same δ̸Qξ\not{\delta}Q_{\xi}. In particular, this is true for deformations in time and radial directions. In’ (4) we use bold notation to indicate spacetime differential forms, namely the surface charge density is the two-form 𝒌ξ=14kξμνεμνρσdxρdxσ\boldsymbol{k}_{\xi}=\frac{1}{4}k^{\mu\nu}_{\xi}\varepsilon_{\mu\nu\rho\sigma}dx^{\rho}\wedge dx^{\sigma}. See Frodden:2019ylc for more details and the explicit formula of kξμνk^{\mu\nu}_{\xi} in terms of a general metric or tetrad variables.

Here we stress two aspects of (4). First, the slashed delta symbol in δ̸Qξ\not{\delta}Q_{\xi} means we compute a one-form in phase space which is not necessarily exact. Exactness in phase space, namely δ(δ̸Qξ)=δ2Qξ=0\delta(\not{\delta}Q_{\xi})=\delta^{2}Q_{\xi}=0, is an extra condition required for the existence of the charge QξQ_{\xi}, in the usual parlance this is the integrability condition. A second aspect is that integration is over any closed smooth surface SS where the conservation law holds (not necessarily at infinity). As we check in a moment, Misner strings violate this requirement, and the integration region needs to be corrected accordingly.

As in Kerr spacetime, for the Kerr-NUT, there are also two exact symmetries present: Time translations generated by t\partial_{t} and axial rotations generated by ϕ\partial_{\phi}. A direct computation of (4) fails as it produces charges dependent on the surface SS (e.g. dependent on the radius of a given sphere): Conservation is broken! This is a problem the reader may be tempted to solve by taking a sphere at infinity. But, as explained before, conservation here means precisely that one can compute charges not necessarily at the asymptotic region. The solution to the problem comes by excluding the region with the Misner strings. More precisely, we integrate the conservation law equation in a volume without the strings. As a consequence, the result does not decouple into the usual two integrals, one over each closed surface SS^{\prime} and SS (which is what makes of (4) a seed to compute a charge). Instead, extra contributions to (4) emerge.

Refer to caption
Figure 1: Integration region: A three-volume, Σ\Sigma, delimited by two spheres and two cones. The blue lines stand for the Misner strings symmetrically positioned (case C=0C=0).

Let us consider the surface charge density 𝒌ξ\boldsymbol{k}_{\xi}, with ξ\xi a Killing vector field, either t\partial_{t} or ϕ\partial_{\phi}. We integrate the three-form d𝒌ξd\boldsymbol{k}_{\xi} on a volume Σϵ\Sigma_{\epsilon} limited by two spherical shells, say S1S_{1} and S2S_{2}, with radius r1<r2r_{1}<r_{2}, and two tubes TST_{S} and TNT_{N} (actually cones) excluding the strings from the volume, see Figure 1. The south cone TST_{S} has a small opening angle, θ=πϵ\theta=\pi-\epsilon, towards the south pole, and the north cone TNT_{N} has a small opening angle, θ=0+ϵ\theta=0+\epsilon, towards the north pole. We integrate d𝒌ξ0d\boldsymbol{k}_{\xi}\approx 0 over Σϵ\Sigma_{\epsilon}, thus

0=Σϵ𝑑𝒌ξ=Σϵ𝒌ξ=S1𝒌ξS2𝒌ξ+TS𝒌ξ+TN𝒌ξ.0=\int_{\Sigma_{\epsilon}}\!\!d\boldsymbol{k}_{\xi}=\int_{\partial\Sigma_{\epsilon}}\!\!\boldsymbol{k}_{\xi}=\int_{S_{1}}\!\!\boldsymbol{k}_{\xi}-\int_{S_{2}}\!\!\boldsymbol{k}_{\xi}+\int_{T_{S}}\!\!\!\boldsymbol{k}_{\xi}+\int_{T_{N}}\!\!\!\boldsymbol{k}_{\xi}\,. (5)

When there are no strings, the third and fourth term vanishes in the shrinking limit ϵ0\epsilon\!\to\!0 and we recover the usual expression for the charge (4). However, for TN spacetime, strings contribute in the limit of a vanishing ϵ\epsilon.

In general, the radial integration over each tube decouples into two terms associated with the circles found as the intersection of the tube with the shells

TS𝒌ξ\displaystyle\int_{T_{S}}\boldsymbol{k}_{\xi} =\displaystyle= Kξ|r=r1;θ=πϵKξ|r=r2;θ=πϵ,\displaystyle K_{\xi}\big{|}_{r=r_{1};\,\theta=\pi-\epsilon}-K_{\xi}\big{|}_{r=r_{2};\,\theta=\pi-\epsilon}\,, (6a)
TN𝒌ξ\displaystyle\int_{T_{N}}\boldsymbol{k}_{\xi} =\displaystyle= Kξ|r=r1;θ=ϵKξ|r=r2;θ=ϵ.\displaystyle K_{\xi}\big{|}_{r=r_{1};\,\theta=\epsilon}-K_{\xi}\big{|}_{r=r_{2};\,\theta=\epsilon}\,. (6b)

Then, we take the limit ϵ0\epsilon\to 0 and reorder (5) with terms associated with each shell to get an expression for the surface charge that is defined exclusively at a sphere S=S1S=S_{1}. This quantity is truly conserved, i.e. its value does not depend on the specific surface SS

δ̸Qξ=S𝒌ξ+Kξ|south+Kξ|north,\not{\delta}Q_{\xi}=\oint_{S}\boldsymbol{k}_{\xi}+K_{\xi}\big{|}_{south}+K_{\xi}\big{|}_{north}\,, (7)

where we refer to θ=π\theta=\pi as south, and θ=0\theta=0 as north. Although we formally use a specific parametrization to refer to geometrical objects, the result does not depend on them. Before shrinking the latter around the strings, we can deform either the shells or the tubes and still have the same resulting surface charge. In particular, (7) does not depend on the specific radius we compute it.

Therefore, we have shown how the KK contributions solve the problem of a naive computation with (4) when string-like defects cross the surface SS. For spacetimes without strings the KK terms automatically vanish. Still, we note that we can introduce artificial string-like defects by changing coordinates even for regular spacetimes. This can break down the conservation of (4) unless terms similar to the KK are taken into account. An example of this is the defect appearing in the axis for some versions of the Kerr-(a)dS metric solution which can be avoided by a ϕ\phi escalation.

We emphasize that to get (7), we rely on the correct integration of the local conservation law. This requires identifying the tubes to exclude the defects: The regions where d𝒌ξ0d\boldsymbol{k}_{\xi}\approx 0 does not hold.

III.1 Mass and Angular Momentum

The two Killing symmetries produce two independent varied surface charges. We check that our corrected formula delivers the expected result.

The solution space is spanned by three parameters mm, aa, and nn. For time translations exact symmetry, we expect the mass/energy, thus we rename δMδ̸Qt\delta M\equiv\not{\delta}Q_{\partial_{t}}, and (7) produces (for details check the appendix)

δM=δm,\delta M\,=\,\delta m\,, (8)

which is of course integrable as δ2M=0\delta^{2}M=0. We set the integration constant equal to zero to obtain the finite charge M=mM=m, which means that for Kerr-NUT spacetime the parameter mm is still identified with the total mass/energy of spacetime.

For the axial symmetry generator we expect the angular momentum, thus we rename δJδ̸Qϕ\delta J\equiv\not{\delta}Q_{-\partial_{\phi}}, and we get

δJ=(a3Cn)δm+mδa3Cmδn.\delta J\,=\,(a-3Cn)\delta m+m\delta a-3Cm\delta n\,. (9)

which again is an integrable expression, δ2J=0\delta^{2}J=0, and produces the angular momentum J=(a3Cn)mJ=(a-3Cn)m. Note that, as expected, the TN spacetime with a=0a=0 still carries angular momentum but only when strings are not symmetrically distributed along the rotation axis.

Let us conclude here with two remarks. First, the charges do not present divergences (e.g. rr-divergences when evaluated in asymptotic regions). While each integral in (7) is rr-dependent, therefore susceptible to divergences, it is only the sum of the three that produces rr-independent charges. Then, the KK-es are essentials and the Misner strings contribute to the charges though them. The second remark is that no asymptotic region is required. We do not demand a particular location for the sphere SS. In that sense, the charges computed above are quasi-local.

IV Entropy and The First Law

In analogy with black holes thermodynamics, we now treat the Kerr-NUT spacetime as a thermodynamic system. We constructively present the first law to determine the entropy of this spacetime following the black hole prescription, i.e. we introduce a temperature and an angular momentum related to the surface r=r+r=r_{+}, the larger root of Δ=0\Delta=0. For later convenience, we focus on the case with C=0C=0.

Let us introduce a temperature (proportional to the surface gravity at r+r_{+} redshifted to infinity)

T=m2a2+n24π(m(m+m2a2+n2)+n2).T=\frac{\sqrt{m^{2}-a^{2}+n^{2}}}{4\pi\left(m\left(m+\sqrt{m^{2}-a^{2}+n^{2}}\right)+n^{2}\right)}.\, (10)

The angular velocity, of the surface r=r+r=r_{+} as measured from infinity, is

Ω=gtϕgϕϕ|r+=a2(m(m+m2a2+n2)+n2).\Omega=\left.\frac{g_{t\phi}}{g_{\phi\phi}}\right|_{r_{+}}=-\frac{a}{2\left(m\left(m+\sqrt{m^{2}-a^{2}+n^{2}}\right)+n^{2}\right)}\,. (11)

The expectation is that in the first law the temperature multiplies the entropy variation, we call it δ̸S\not{\delta}S, and the angular velocity multiplies the angular momentum variation (9). Then, we can isolate δ̸S\not{\delta}S from the first law to get

δ̸S=1T(δMΩδJ)=2πm2a2+n2×((2m(m+m2a2+n2)a2+2n2)δmmaδa).\not{\delta}S=\frac{1}{T}\left(\delta M-\Omega\delta J\right)=\frac{2\pi}{\sqrt{m^{2}-a^{2}+n^{2}}}\times\\ \left(\big{(}2m(m+\sqrt{m^{2}-a^{2}+n^{2}})-a^{2}+2n^{2}\big{)}\delta m-ma\delta a\right)\,. (12)

Note that for n=0n=0 the previous expression is integrable, and as expected, it is the variation of the usual one-fourth of the area black hole entropy of Kerr spacetime. Even for non-rotating TN spacetime, n0n\neq 0, the expression is not integrable anymore, δ2S0\delta^{2}S\neq 0, which is a problem to have a well-defined entropy charge.

Still the entropy variation is conserved in the sense that it can be directly computed from (7) with the combination of Killing vectors ξ=1T(t+Ωϕ)\xi=\frac{1}{T}\left(\partial_{t}+\Omega\,\partial_{\phi}\right). That means it is a proper varied surface charge as it does not depend on the surface one chooses to compute it. Therefore, this expression should be a good seed to compute an entropy as a Noether charge as we already did with the mass or the angular momentum (we keep the Noether charge name, understanding that surface charge method to compute charges generalizes the Noether procedure Frodden:2019ylc ).

In the literature, it has been explored the possibility of adding extra terms to the first law associated with the Misner strings. At this level, this strategy is equivalent to the splitting of (12) in different terms and providing interpretation to each one. Here we choose another path. We use the surface charge method as a guide and keep the possibility of having an entropy that is indeed conserved: A Noether charge. Therefore the expression in (12) must not be split; otherwise, we can not write the entropy variation as (7) with an arbitrary closed surface.

The alternative path: We have checked the suggestive alternative of splitting (12) into surface and tubes terms associated to the Misner strings as in (7). As explained before, separately surfaces and tubes terms are not conserved quantities and neither they are integrable. One may be tempted to fix the closed surface at r=r+r=r_{+}. In this case, the surface integral in (7) indeed becomes the variation of one-fourth of the area. However, we are not computing a conserved charge if we have to fix the surface where it is defined. Furthermore, insisting in this approach, the terms associated with the tubes are not integrable, and another temperatures should be introduced to achieve their integrability. This path would be similar to the one presented in Bordo:2019rhu (still, we have a different angular momentum stemming from the other methods used to compute charges).

A way to have, simultaneously, the entropy as conserved and integrable is to reduce the parameter space. Let us assume that mm, aa, and nn are dependent quantities and solve this dependence such that the integrability condition is satisfied. Consider m=m(n,a)m=m(n,a). Because of the way these parameters enter in the metric components, it is easy to see that all the tree parameters scale with the same factor under scaling of the metric. Then, the allowed relation among the parameters is simpler: m=nh(a/n)m=n\,h(a/n), with h(a/n)h(a/n) a function to be determined by the integrability condition δ(δ̸S)=0\delta(\not{\delta}S)=0. This produces a differential equation solved simply by h(a/n)=αh(a/n)=\alpha, with α\alpha an integration constant. Thus

m=αn,m=\alpha\,n\,, (13)

ensures integrability of the entropy, and it reduces the parameter space of our solution from three to two parameters. Replacing this in the equation for the varied charge (12), we can integrate it to obtain an entropy

S=2παn2(α+1+α2a2n2),S=2\pi\alpha\,n^{2}\left(\alpha+\sqrt{1+\alpha^{2}-\frac{a^{2}}{n^{2}}}\right)\,, (14)

where we have set to zero the integration constant. Our method does not fix α\alpha and in what follows we keep it free.111We can fix α\alpha if we require the entropy to be compatible with the non-rotating TN entropy obtained through other methods. For instance, Hawking and Hunter in Hawking:1998jf found for the Euclidean TN-bolt counterpart SHH=πn2S_{HH}=\pi n^{2}, this fixes α=122\alpha=\frac{1}{2\sqrt{2}}. This entropy fulfills a standard first law

δM=TδS+ΩδJ,\delta M=T\delta S+\Omega\delta J\,, (15)

where M=αnM=\alpha\,n and J=αanJ=\alpha\,a\,n. In contrast with the usual black hole solutions, the entropy is not one-fourth of the area, first noted in Hawking:1998jf for non-rotating TN and using Euclidean methods. The first law presented is minimal because it does not require additional terms not justified in terms of charges. It may be seen as non-standard as we had reduced the parameter space to achieve integrability. But as explained before, this is required to have MM, JJ, and SS as well-defined charges to compose the first law. Therefore, we conclude that there is a consistent thermodynamic description only for a subset of parameter space of the Kerr-NUT spacetime. If TN spacetimes will ever have a physical meaning this is a prediction of our approach.

IV.1 Grand Canonical Ensemble

Our result is compatible with the Euclidean action to describe the thermodynamics. To show this, we use the grand canonical ensemble function, GG, already computed in the Appendix of Bordo:2019rhu ; Emparan:1999pm as the flat limit of a regularized Euclidean (a)dS action divided by β=1/T\beta=1/T. The result is simply G=m/2G=m/2. We express it as a function of the intensive thermodynamic variables at play to get

G(T,Ω)=18πT(1+α1+Ω24π2T2).G(T,\Omega)=\frac{1}{8\pi T\left(1+\alpha^{\prime}\sqrt{1+\displaystyle\frac{\Omega^{2}}{4\pi^{2}T^{2}}}\right)}\,. (16)

With α=1+1/α2\alpha^{\prime}=\sqrt{1+1/\alpha^{2}}. The expression is consequence of a simple algebraic manipulation of T(a,n)T(a,n), Ω(a,n)\Omega(a,n), and m=αnm=\alpha\,n. Note its simplicity. Now, the first law for the Kerr-NUT spacetimes can be reproduced in the standard way. As expected, the entropy (14), the angular momentum, and the mass/energy are: S=GTS=-\frac{\partial G}{\partial T}, J=GΩJ=\frac{\partial G}{\partial\Omega} and M=GJΩ+STM=G-J\Omega+ST, respectively.

V Conclusions and Outlook

In this note, we reported a procedure to use the covariant surface charge method for spacetimes with string-like defects and immediately applied it to the Kerr-NUT spacetime. The critical step was to note that the equation, d𝒌ξ0d{\boldsymbol{k}}_{\xi}\approx 0, is not well behaved over the Misner strings; therefore, the three-volume where it is integrated can not contain them. The procedure is general and can be easily applied in particular for accelerating black holes.

For the Kerr-NUT spacetime, besides the explicit computation of the natural charges, mass and angular momentum, the thermodynamic picture was completed by writing down the first law and computing the entropy as an integrated charge. We achieve the integrability of the entropy without introducing additional terms in the first law. As a general rule, the introduction of new terms in the first law should be done with caution, we try to avoid it. Instead, we impose a reduction in the parameter space of the solution. Thus, the principle that charges appearing in the first law (MM, JJ, and SS) are Noether charges, all at the same footing, is respected. The entropy found is not proportional to the area, which is something already expected for TN spacetimes.

Let us finish with a comment on our attempt to extend the results to the (a)dS Kerr-NUT family. The strategy to compute mass and angular momentum charges is equally successful there, however we could not solve the differential equation for the integrability of the entropy. Then, we leave for a future project to extend the thermodynamic picture for that spacetime family or in full generality for the Plebanski-Demianski family Griffiths:2005qp .

Acknowledgements.
The authors would like thank to Remigiusz Durka for a valuable discussion on the topic. The Centro de Estudios Científicos (CECs) is funded by the Chilean Government through the Centers of Excellence Base Financing Program of ANID.

Appendix A Charge Computations

The surface charge for GR is given by Lee:1990nz ; Iyer:1994ys ; Wald:1999wa ; Barnich:2001jy ; Frodden:2019ylc

kξμν\displaystyle k_{\xi}^{\mu\nu} =\displaystyle= gκ(ξ[νσδgμ]σξ[νμ]δg+ξσ[μδgν]σ\displaystyle\sqrt{-g}\kappa\left(\xi^{[\nu}\nabla_{\sigma}\delta g^{\mu]\sigma}-\xi^{[\nu}\nabla^{\mu]}\delta g+\xi_{\sigma}\nabla^{[\mu}\delta g^{\nu]\sigma}\right. (17)
12δg[νξμ]+δgσ[νσξμ]),\displaystyle~{}~{}~{}~{}~{}~{}~{}~{}~{}\left.-\frac{1}{2}\delta g\nabla^{[\nu}\xi^{\mu]}+\delta g^{\sigma[\nu}\nabla_{\sigma}\xi^{\mu]}\right)\,,

where κ=c4/(8πG)\kappa=c^{4}/(8\pi G), cc the speed of light, GG the Newton’s constant, gg stands for the determinant of the metric gμνg_{\mu\nu}, and δggμνδgμν\delta g\equiv g_{\mu\nu}\delta g^{\mu\nu}.222We work in units where c=G=1c=G=1. Notice that the last three terms of the equation corresponds to the variation of the Komar’s integrand commonly used to compute charges.

The integral of kξμνk_{\xi}^{\mu\nu} over a closed surface splits in four regions: Two concentric spheres with holes and two cones (see Fig. 1). Over a sphere at constant rr the required component is kξtrk^{tr}_{\xi} while for the cones it is kξθtk^{\theta t}_{\xi}. The explicit expressions for them in the Kerr-Taub-NUT spacetime are quite large, thus, for the sake of space with do not present them here (still it is a straightforward computation using a software). To give a taste we exhibit just the coefficient accompanying the δm\delta m term of kttrk(t)trk_{\partial_{t}}^{tr}\equiv k^{tr}_{(t)} associated to a Killing vector field ξ=t\xi=\partial_{t}. That is, we decompose

k(t)tr=[k(t)tr]δmδm+[k(t)tr]δaδa+[k(t)tr]δnδn,k^{tr}_{(t)}=\left[k^{tr}_{(t)}\right]_{\delta m}\delta m+\left[k^{tr}_{(t)}\right]_{\delta a}\delta a+\left[k^{tr}_{(t)}\right]_{\delta n}\delta n\,,

and the first term is explicitly

[k(t)tr]δm=sinθ8π(1+(r2+a22anC+n2)(r2a2cos2θ)(r2+acosθ(acosθ+2n)+n2)2).\left[k^{tr}_{(t)}\right]_{\delta m}=\\ \frac{\sin\theta}{8\pi}\left(1+\frac{\left(r^{2}+a^{2}-2anC+n^{2}\right)(r^{2}-a^{2}\cos^{2}\theta)}{\left(r^{2}+a\cos\theta(a\cos\theta+2n)+n^{2}\right)^{2}}\right)\,. (18)

Note that it is a function of the coordinates. Here we can already identify the first term that will contribute to half of the final charge mm. We can also check why, for this particular charge, the asymptotic analysis works as in the limit rr\to\infty the second term in the parenthesis reduces to 1.

Before continuing with the calculation, let us define some functions that will allow us to compactly write our next expressions

h±\displaystyle h_{\pm} =\displaystyle= (a±n)2+r2,\displaystyle(a\pm n)^{2}+r^{2}\,, (19)
fδa±\displaystyle f_{\delta a_{\pm}} =\displaystyle= n(m(a±n)2±2n(a±n)r+mr2),\displaystyle n(-m(a\pm n)^{2}\pm 2n(a\pm n)r+mr^{2})\,, (20)
fδn±\displaystyle f_{\delta n_{\pm}} =\displaystyle= m(a±n)2(2a±n)n(a±n)(3a±n)r\displaystyle m(a\pm n)^{2}(2a\pm n)\mp n(a\pm n)(3a\pm n)r
+m(2a±3n)r23nr3.\displaystyle+\,m(2a\pm 3n)r^{2}\mp 3nr^{3}\,. (21)

For the Killing vector field ξ=t\xi=\partial_{t}, the terms from the cones contributing to the charges in (6) are

TS𝒌(t)\displaystyle\int_{T_{S}}\boldsymbol{k}_{(t)} =\displaystyle= r1r2𝑑r02π𝑑ϕk(t)θt|θ=πϵ=K(t)|r=r1,θ=πϵK(t)|r=r2,θ=πϵ,\displaystyle\int_{r_{1}}^{r_{2}}dr\int_{0}^{2\pi}d\phi\,k^{\theta t}_{(t)}\bigg{|}_{\theta=\pi-\epsilon}=K_{(t)}\bigg{|}_{\begin{subarray}{c}r=r_{1},\\ \theta=\pi-\epsilon\end{subarray}}-K_{(t)}\bigg{|}_{\begin{subarray}{c}r=r_{2},\\ \theta=\pi-\epsilon\end{subarray}}\,, (22)
TN𝒌(t)\displaystyle\int_{T_{N}}\boldsymbol{k}_{(t)} =\displaystyle= r1r2𝑑r02π𝑑ϕk(t)θt|θ=ϵ=K(t)|r=r1,θ=ϵK(t)|r=r2,θ=ϵ,\displaystyle\int_{r_{1}}^{r_{2}}dr\int_{0}^{2\pi}d\phi\,k^{\theta t}_{(t)}\bigg{|}_{\theta=\epsilon}=K_{(t)}\bigg{|}_{\begin{subarray}{c}r=r_{1},\\ \theta=\epsilon\end{subarray}}-K_{(t)}\bigg{|}_{\begin{subarray}{c}r=r_{2},\\ \theta=\epsilon\end{subarray}}\,, (23)

where ϵ\epsilon is the small opening angle for the cones. We take ϵ0\epsilon\to 0 after integrating on each surface. Note that we have included an explicit intermediate term, and note that for that term θ\theta names a component of the tensorial density k(t)μνk^{\mu\nu}_{(t)} but it is also a coordinate where the spacetime function k(t)μνk^{\mu\nu}_{(t)} can be evaluated. We concentrate in the region with r2rr_{2}\to r, then the two ending terms of the last equations are, explicitly, given by

K(t)|north\displaystyle K_{(t)}\big{|}_{north} =\displaystyle= limϵ0K(t)|r,θ=ϵ\displaystyle\lim_{\epsilon\to 0}K_{(t)}\big{|}_{\begin{subarray}{c}r,\\ \theta=\epsilon\end{subarray}}
=\displaystyle= C+12h+2(n(a+n)h+δm+fδa+δa+fδn+δn),\displaystyle\frac{C+1}{2h_{+}^{2}}\left(n(a+n)h_{+}\delta m+f_{\delta a_{+}}\delta a+f_{\delta n_{+}}\delta n\right)\,,
K(t)|south\displaystyle K_{(t)}\big{|}_{south} =\displaystyle= limϵ0K(t)|r,θ=πϵ\displaystyle\lim_{\epsilon\to 0}K_{(t)}\big{|}_{\begin{subarray}{c}r,\\ \theta=\pi-\epsilon\end{subarray}}
=\displaystyle= C12h2(n(an)hδm+fδaδa+fδnδn).\displaystyle\frac{C-1}{2h_{-}^{2}}\left(n(a-n)h_{-}\delta m+f_{\delta a_{-}}\delta a+f_{\delta n_{-}}\delta n\right)\,.

For the sphere contribution we have

S𝒌(t)\displaystyle\oint_{S}\boldsymbol{k}_{(t)} =\displaystyle= limϵ0[ϵπϵ𝑑θ02π𝑑ϕk(t)tr]\displaystyle\lim_{\epsilon\to 0}\left[\int_{\epsilon}^{\pi-\epsilon}d\theta\int_{0}^{2\pi}d\phi\ k^{tr}_{(t)}\right]
=\displaystyle= (1(C1)n(an)2h(C+1)n(a+n)2h+)δm\displaystyle\left(1-\frac{(C-1)n(a-n)}{2h_{-}}-\frac{(C+1)n(a+n)}{2h_{+}}\right)\delta m
((C1)fδa2h2+(C+1)fδa+2h+2)δa\displaystyle-\left(\frac{(C-1)f_{\delta a_{-}}}{2h_{-}^{2}}+\frac{(C+1)f_{\delta a_{+}}}{2h_{+}^{2}}\right)\delta a
((C1)fδn2h2+(C+1)fδn+2h+2)δn.\displaystyle-\left(\frac{(C-1)f_{\delta n_{-}}}{2h_{-}^{2}}+\frac{(C+1)f_{\delta n_{+}}}{2h_{+}^{2}}\right)\delta n\,.

Finally, the variation of the charge reads (see Eq. (7))

δMδ̸Q(t)=S𝒌(t)+K(t)|south+K(t)|north=δm,\delta M\equiv\not{\delta}Q_{(t)}=\oint_{S}\boldsymbol{k}_{(t)}+K_{(t)}\big{|}_{south}+K_{(t)}\big{|}_{north}=\delta m\,, (27)

which happens to be trivially integrable. We see that most of the contributions to the charge nicely cancel such that the final result does not depend on coordinates (here the radius of the sphere). Note how complicated the rr-dependency of each term may be. Still, the whole charge is independent as the method ensures conservation by virtue of μk(t)μν0\partial_{\mu}k^{\mu\nu}_{(t)}\approx 0.
The computation of the (varied) angular momentum δJδ̸Qϕ=δ((a3Cn)m)\delta J\equiv\not{\delta}Q_{-\partial_{\phi}}=\delta\left((a-3Cn)m\right) associated with the Killing vector field ξ=ϕ\xi=-\partial_{\phi} can be carried out following the same steps as the derivation of the mass.

References

  • (1) A. H. Taub,  Empty space-times admitting a three parameter group of motions, Annals of Mathematics 53 472-490 (1951).
  • (2) E. Newman, L. Tamburino, and T. Unti, Empty-space generalization of the Schwarzschild metric, Journal of Mathematical Physics 4 (1963), no. 7 915–923.
  • (3) C. W. Misner, “The Flatter regions of Newman, Unti and Tamburino’s generalized Schwarzschild space,” J. Math. Phys. 4, 924-938 (1963) doi:10.1063/1.1704019
  • (4) Bonnor, W. B.  1969 Proc. Camb. Phil. Soc. 66 145
  • (5) A. Chamblin, R. Emparan, C. V. Johnson and R. C. Myers, “Large N phases, gravitational instantons and the nuts and bolts of AdS holography,” Phys. Rev. D 59, 064010 (1999) doi:10.1103/PhysRevD.59.064010 [arXiv:hep-th/9808177 [hep-th]]. Copy to ClipboardDownload
  • (6) A. M. Ghezelbash, R. B. Mann and R. D. Sorkin, Can. J. Phys. 86, 579-582 (2008) doi:10.1139/p07-184
  • (7) R. Emparan, C. V. Johnson and R. C. Myers, “Surface terms as counterterms in the AdS / CFT correspondence,” Phys. Rev. D 60, 104001 (1999) doi:10.1103/PhysRevD.60.104001 [arXiv:hep-th/9903238 [hep-th]].
  • (8) R. B. Mann, “Misner string entropy,” Phys. Rev. D 60, 104047 (1999) doi:10.1103/PhysRevD.60.104047 [arXiv:hep-th/9903229 [hep-th]].
  • (9) R. B. Mann, “Entropy of rotating Misner string space-times,” Phys. Rev. D 61, 084013 (2000) doi:10.1103/PhysRevD.61.084013 [arXiv:hep-th/9904148 [hep-th]].
  • (10) D. Garfinkle and R. B. Mann, “Generalized entropy and Noether charge,” Class. Quant. Grav. 17, 3317-3324 (2000) doi:10.1088/0264-9381/17/16/314 [arXiv:gr-qc/0004056 [gr-qc]].
  • (11) C. V. Johnson, “Thermodynamic Volumes for AdS-Taub-NUT and AdS-Taub-Bolt,” Class. Quant. Grav. 31, no.23, 235003 (2014) doi:10.1088/0264-9381/31/23/235003 [arXiv:1405.5941 [hep-th]].
  • (12) C. V. Johnson, “The Extended Thermodynamic Phase Structure of Taub-NUT and Taub-Bolt,” Class. Quant. Grav. 31, 225005 (2014) doi:10.1088/0264-9381/31/22/225005 [arXiv:1406.4533 [hep-th]].
  • (13) R. Clarkson, L. Fatibene and R. B. Mann, “Thermodynamics of (d+1)-dimensional NUT charged AdS space-times,” Nucl. Phys. B 652, 348-382 (2003) doi:10.1016/S0550-3213(02)01143-4 [arXiv:hep-th/0210280 [hep-th]].
  • (14) D. Astefanesei, R. B. Mann and E. Radu, “Nut charged space-times and closed timelike curves on the boundary,” JHEP 01, 049 (2005) doi:10.1088/1126-6708/2005/01/049 [arXiv:hep-th/0407110 [hep-th]].
  • (15) G. Kalamakis, R. G. Leigh and A. C. Petkou, “Aspects of holography of Taub-NUT- AdS4 spacetimes,” Phys. Rev. D 103, no.12, 126012 (2021) doi:10.1103/PhysRevD.103.126012 [arXiv:2009.08022 [hep-th]].
  • (16) L. Ciambelli, C. Corral, J. Figueroa, G. Giribet and R. Olea, “Topological Terms and the Misner String Entropy,” Phys. Rev. D 103, no.2, 024052 (2021) doi:10.1103/PhysRevD.103.024052 [arXiv:2011.11044 [hep-th]].
  • (17) E. Arratia, C. Corral, J. Figueroa and L. Sanhueza, “Hairy Taub-NUT/bolt-AdS solutions in Horndeski theory,” Phys. Rev. D 103, no.6, 064068 (2021) doi:10.1103/PhysRevD.103.064068 [arXiv:2010.02460 [hep-th]].
  • (18) S. W. Hawking and C. J. Hunter, “Gravitational entropy and global structure,” Phys. Rev. D 59, 044025 (1999) doi:10.1103/PhysRevD.59.044025 [arXiv:hep-th/9808085 [hep-th]].
  • (19) S. W. Hawking, C. J. Hunter and D. N. Page, “Nut charge, anti-de Sitter space and entropy,” Phys. Rev. D 59, 044033 (1999) doi:10.1103/PhysRevD.59.044033 [arXiv:hep-th/9809035 [hep-th]].
  • (20) S. W. Hawking, “Gravitational Instantons,” Phys. Lett. A 60, 81 (1977) doi:10.1016/0375-9601(77)90386-3
  • (21) G. Clément, D. Gal’tsov and M. Guenouche, “Rehabilitating space-times with NUTs,” Phys. Lett. B 750, 591-594 (2015) doi:10.1016/j.physletb.2015.09.074 [arXiv:1508.07622 [hep-th]].
  • (22) G. Clément, D. Gal’tsov and M. Guenouche, “NUT wormholes,” Phys. Rev. D 93, no.2, 024048 (2016) doi:10.1103/PhysRevD.93.024048 [arXiv:1509.07854 [hep-th]].
  • (23) R. A. Hennigar, D. Kubizňák and R. B. Mann, “Thermodynamics of Lorentzian Taub-NUT spacetimes,” Phys. Rev. D 100, no.6, 064055 (2019) doi:10.1103/PhysRevD.100.064055 [arXiv:1903.08668 [hep-th]].
  • (24) P. A. Cano and D. Pereñiguez, [arXiv:2101.10652 [hep-th]].
  • (25) H. Gonzalez, D. Grumiller, W. Merbis and R. Wutte, EPJ Web Conf. 168, 01009 (2018) doi:10.1051/epjconf/201816801009 [arXiv:1709.09667 [hep-th]].
  • (26) A. Ballon Bordo, F. Gray, R. A. Hennigar and D. Kubizňák, “The First Law for Rotating NUTs,” Phys. Lett. B 798, 134972 (2019) doi:10.1016/j.physletb.2019.134972 [arXiv:1905.06350 [hep-th]].
  • (27) A. B. Bordo, F. Gray and D. Kubizňák, “Thermodynamics and Phase Transitions of NUTty Dyons,” JHEP 07, 119 (2019) doi:10.1007/JHEP07(2019)119 [arXiv:1904.00030 [hep-th]].
  • (28) R. Durka, “The first law of black hole thermodynamics for Taub-NUT spacetime,” [arXiv:1908.04238 [gr-qc]].
  • (29) A. B. Bordo, F. Gray, R. A. Hennigar and D. Kubizňák, “Misner Gravitational Charges and Variable String Strengths,” Class. Quant. Grav. 36, no.19, 194001 (2019) doi:10.1088/1361-6382/ab3d4d [arXiv:1905.03785 [hep-th]].
  • (30) A. Ballon Bordo, F. Gray, R. A. Hennigar and D. Kubizňák, “The First Law for Rotating NUTs,” Phys. Lett. B 798, 134972 (2019) doi:10.1016/j.physletb.2019.134972 [arXiv:1905.06350 [hep-th]].
  • (31) M. Zhang and J. Jiang, “NUT charges and black hole shadows,” Phys. Lett. B 816, 136213 (2021) doi:10.1016/j.physletb.2021.136213 [arXiv:2103.11416 [gr-qc]].
  • (32) J. Lee and R. M. Wald, J. Math. Phys. 31, 725-743 (1990) doi:10.1063/1.528801
  • (33) V. Iyer and R. M. Wald, Phys. Rev. D 50, 846-864 (1994) doi:10.1103/PhysRevD.50.846 [arXiv:gr-qc/9403028 [gr-qc]].
  • (34) R. M. Wald and A. Zoupas, Phys. Rev. D 61, 084027 (2000) doi:10.1103/PhysRevD.61.084027 [arXiv:gr-qc/9911095 [gr-qc]].
  • (35) G. Barnich and F. Brandt, Nucl. Phys. B 633, 3-82 (2002) doi:10.1016/S0550-3213(02)00251-1 [arXiv:hep-th/0111246 [hep-th]].
  • (36) E. Frodden and D. Hidalgo, “Surface Charges Toolkit for Gravity,” Int. J. Mod. Phys. D 29, no.06, 2050040 (2020) doi:10.1142/S0218271820500406 [arXiv:1911.07264 [hep-th]].
  • (37) Demianski M and Newman E T 1966 Bull. Acad. Polon. Sci. Ser. Math. Astron. Phys. 14 653.
  • (38) V. S. Manko, J. Martin and E. Ruiz, Class. Quant. Grav. 23, 4473-4484 (2006) doi:10.1088/0264-9381/23/13/011 [arXiv:gr-qc/0603002 [gr-qc]].
  • (39) J. B. Griffiths and J. Podolsky, “Exact Space-Times in Einstein’s General Relativity,” doi:10.1017/CBO9780511635397.
  • (40) J. B. Griffiths and J. Podolsky, Int. J. Mod. Phys. D 15, 335-370 (2006) doi:10.1142/S0218271806007742 [arXiv:gr-qc/0511091 [gr-qc]].