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

Observing Black Hole Phase Transitions in Extended Phase Space and Holographic Thermodynamics Approaches from Optical Features

Chatchai Promsiri chatchaipromsiri@gmail.com Quantum Computing and Information Research Centre (QX), Faculty of Science, King Mongkut’s University of Technology Thonburi, Bangkok 10140, Thailand    Weerawit Horinouchi wee.hori@gmail.com High Energy Physics Research Unit, Department of Physics, Faculty of Science, Chulalongkorn University, Bangkok 10330, Thailand    Ekapong Hirunsirisawat ekapong.hir@kmutt.ac.th Quantum Computing and Information Research Centre (QX), Faculty of Science, King Mongkut’s University of Technology Thonburi, Bangkok 10140, Thailand Learning Institute, King Mongkut’s University of Technology Thonburi, Bangkok 10140, Thailand
Abstract

The phase transitions of charged Anti-de Sitter (AdS) black holes are characterized by studying null geodesics in the vicinity of the critical curve of photon trajectories around black holes as well as their optical appearance as the black hole images. In the present work, the critical parameters including the orbital half-period τ\tau, the angular Lyapunov exponent λL\lambda_{L}, and the temporal Lyapunov exponent γL\gamma_{L} are employed to characterize black hole phase transitions within both the extended phase space and holographic thermodynamics frameworks. Under certain conditions, we observe multi-valued function behaviors of these parameters as functions of bulk pressure and temperature in the respective approaches. We propose that τ\tau, λL\lambda_{L}, and γL\gamma_{L} can serve as order parameters due to their discontinuous changes at first-order phase transitions. To validate this, we provide detailed analytical calculations demonstrating that these optical critical parameters follow scaling behavior near the critical phase transition point. Notably, the critical exponents for these parameters are found to be 1/21/2, consistent with those of the van der Waals fluid. Our findings suggest that static and distant observers can study black hole thermodynamics by analyzing the images of regions around the black holes.

I Introduction

Thermodynamics of black holes (BHs) has emerged as an active research area that offers deep insights into the interplay among gravity, thermodynamics, quantum mechanics, and information theory. Initiated by Bekenstein [1] and Hawking [2], the energy EE, entropy SS and temperature TT of a BH can be written in the geometric description as follows

E=M,S=A4GN,T=κ2π,\displaystyle E=M,\ \ \ S=\frac{A}{4G_{N}},\ \ \ T=\frac{\kappa}{2\pi}, (1)

where M,AM,A, and κ\kappa are the mass of BH, the surface area of event horizon, and surface gravity at the event horizon, respectively. With such geometric identification, the laws of BH mechanics [3] and the laws of thermodynamics exhibit a surprising mathematical analogy, leading us to consider BHs as thermal objects.

The discussion of BH thermodynamics has been significantly enriched by the discovery of holographic descriptions. Given that the symmetry group of AdSn+1\text{AdS}_{n+1} and the conformal group in nn-dimensional spacetime are both isomorphic to SO(n,2)\text{SO}(n,2), the AdS/CFT correspondence emerged as a duality between the gravitational system in the bulk AdS space and the conformal field theory (CFT) on the boundary [4, 5, 6, 7]. Specifically, thermal radiation and BHs within an (n+1)(n+1)-dimensional AdS space respectively correspond to the nn-dimensional confining phase at zero temperature and thermal states within the deconfining phase of the large-NN gauge theory. Consequently, the confinement-deconfinement phase transition in gauge theory has the gravitational description in the bulk through the Hawking-Page phase transition of AdS-BHs [8].

Despite significant advancements, several issues in BH thermodynamics remain unresolved. The first issue involves the Smarr formula [9], which expresses the relation between different geometric quantities of a BH in the same fashion as the Euler equation in conventional thermodynamics. However, there is an additional term in the Smarr formula when considering BHs with a nonzero cosmological constant Λ\Lambda. The Smarr formula for AdS-BH is given by:

M=n1n2TS+n1n2ΩJ+ΦQ1n2ΘΛ4πGN,\displaystyle M=\frac{n-1}{n-2}TS+\frac{n-1}{n-2}\Omega J+\Phi Q-\frac{1}{n-2}\frac{\Theta\Lambda}{4\pi G_{N}}, (2)

where Ω\Omega is the angular velocity, JJ is the angular momentum, Φ\Phi is the electric potential and QQ is the electric charge of BH. Note that Θ\Theta has been suggested to be defined as the proper volume weighted locally by a Killing vector [10]. Despite this formulation, there continues to be significant debate over the appropriate thermodynamic interpretation of Λ\Lambda and its conjugate variable Θ\Theta. The second issue arises from the observation that the first law of BH thermodynamics does not include the pd𝒱pd\mathcal{V} term. As a result, one may wonder whether the holographic description can provide a gravitational description of the mechanical work term dual to that of the conformal matter.

Remarkably, there are two recent progresses that can be applied to resolve these important issues as follows.

  • Black hole chemistry or extended phase space approach is an extension of thermodynamic phase space for AdS-BHs by identifying the negative cosmological constant Λ\Lambda and its conjugate variable as a bulk pressure PP and thermodynamic volume VV via

    P=Λ8πGNandV=Θ,\displaystyle P=-\frac{\Lambda}{8\pi G_{N}}\ \ \ \text{and}\ \ \ V=-\Theta, (3)

    where V=43πr+3V=\frac{4}{3}\pi r_{+}^{3} for spherical symmetric BHs [11, 12, 13, 14], for good reviews see [15, 16] and reference therein. In this way, the first law of BH thermodynamics has VdPVdP additional work term, so the BH’s mass should be identified as an enthalpy rather than the internal energy of the BH. Notably, the extended phase space approach gives a more consistent gravitational analog of the first law of thermodynamics in conventional matter. For example, the Hawking-Page phase transition in Schwarzschild AdS can be interpreted as solid-liquid phase transition [17], isotherm curves in PVP-V plane of charged AdS-BH in canonical ensemble undergoes a family of Small-Large BHs first order phase transition ending at a critical point of second order phase transition corresponds to the liquid-gas phase transition in the van der Waals (vdW) fluid [18, 19, 20] and the Small-Large-Small BHs phase transition occurs in rotating and nonlinear electrodynamics AdS-BHs analogous to the reentrant phase transition of multicomponent liquids [21, 22]. Moreover, this framework introduces what can be referred to as the black hole’s molecules for black hole thermodynamics [23, 24, 25]. These entities can serve similarly to the microscopic constituents in statistical mechanics, providing a means to describe thermodynamic behavior.

  • Holographic thermodynamics has been proposed to address concerns about the consistency of the extended phase space approach with the AdS/CFT correspondence. Although the extended phase space approach provides a plausible interpretation of Λ\Lambda and its conjugate Θ\Theta for describing the thermodynamics of AdS-BHs, its consistency with the AdS/CFT correspondence remains debated by some researchers [26, 27, 28, 29, 30]. Before discussing progress in holographic thermodynamics aimed at resolving these consistency issues, let us first clarify two important points within the extended phase space approach that continue to be questioned. They are as follows:

    1. (i)

      The equation of state for conformal fluid with energy EE, pressure pp and volume 𝒱\mathcal{V} is described by

      E=(n1)p𝒱.\displaystyle E=(n-1)p\mathcal{V}. (4)

      According to the AdS/CFT dictionary, EE of nn-dimensional conformal fluid is holographically dual to the mass MM of BH in n+1n+1-dimensional spacetime. It is evident that upon replacing pp and 𝒱\mathcal{V} in the above equation with PP and VV in Eq. (3) respectively, the result indicates that EE is not equal to MM of the BH. This implies that the bulk pressure PP and the volume VV for BH are not equivalent to the fluid’s pressure pp and volume 𝒱\mathcal{V}. In other words, the Smarr relation of BHs in the bulk does not correspond to the Euler equation of large-NN gauge theories on the boundary.

    2. (ii)

      In fact, the AdS/CFT correspondence suggests that Λ\Lambda should be dual to the number of colors NcN_{c} in the large-NcN_{c} gauge theory via the relation [30]

      kLn116πGN=Nc2=𝒞,\displaystyle k\frac{L^{n-1}}{16\pi G_{N}}=N_{c}^{2}=\mathcal{C}, (5)

      where 𝒞\mathcal{C} is the central charge and kk is the constant depending on the details of a particular system. These relations suggest that the variation of LL implies a variation of NcN_{c}, which is the rank of the gauge group. Therefore, varying Λ\Lambda of the bulk leads to changing from one field theory to another. Moreover, the variation of LL also corresponds to variation in the volume 𝒱\mathcal{V} of the dual gauge theory since the geometry of dual field theory depends on LL of the bulk.

    Visser [31] proposed an approach to overcome this issue, demonstrating that the Euler equation of the dual field theory requires the μ𝒞𝒞\mu_{\mathcal{C}}\mathcal{C} term but lacks the p𝒱p\mathcal{V} term, while the pd𝒱pd\mathcal{V} term appears in the first law of thermodynamics. Here, 𝒞\mathcal{C} and μ𝒞\mu_{\mathcal{C}} are the central charge and its chemical potential, respectively. On the gravity side, introducing 𝒞\mathcal{C} and μ𝒞\mu_{\mathcal{C}} as thermodynamic variables in the boundary theory is equivalent to allowing for variations in both the Newton constant GNG_{N} and the AdS radius LL in the bulk theory. Using Eq. (5), it becomes possible to vary both LL and GNG_{N} while keeping 𝒞\mathcal{C} fixed at the boundary, ensuring the field theory remains unchanged to another. Thus, one can more appropriately study the variations in the volume 𝒱\mathcal{V} of the dual field theory as LL changes. As suggested in [31], the terms Λ\Lambda and Θ\Theta in Eq. (2) can be rewritten in terms of 𝒞\mathcal{C} and μ𝒞\mu_{\mathcal{C}}. This leads to consistency between the Smarr formula and the Euler equation within the framework known as holographic thermodynamics. Recently, the thermodynamic behaviors resulting from holographic thermodynamics have been extensively studied [32, 33, 34, 35, 36, 37, 38]. For a comprehensive recent review, see [39].

While BH thermodynamics has been developed with intriguing arguments as discussed above, determining the most valid approach remains challenging. It is crucial to support theoretical claims with observations or at least establish a connection between them. It makes sense to argue that the thermal properties of BHs should manifest in specific observational signatures. A natural question that arises is how we can obtain proper observational signatures of a BH to identify its thermodynamic properties and phase transitions. Unfortunately, the observational confirmation of BH thermodynamics remains difficult, although recent advancements that have provided evidence of BH existence through the gravitational wave (GW) signal emitted by a binary black hole (BBH) [40] and the images of two supermassive BHs, i.e., M87* and SgrA* [41, *ETH2, *ETH3, *ETH4, *ETH5, *ETH6, *EventHorizonTelescope:2022wkp, *EventHorizonTelescope:2022apq, *EventHorizonTelescope:2022wok, *EventHorizonTelescope:2022exc, *EventHorizonTelescope:2022urf, *EventHorizonTelescope:2022xqj]. More specifically, the thermodynamic quantities of BHs are typically defined via their event horizons, which are difficult to detect directly from observers at asymptotic infinity.

Although the AdS-BH seems to be interested only in the theoretical aspect, it is valuable to find ways to connect its thermodynamic behaviors with observational signatures. Numerous studies on AdS-BH have attempted to relate the BH phase transition to some signals, which can be observed at asymptotic infinity. Namely, the quasinormal modes (QNMs), which can be characterized by the signature in the GWs emitted from a BH during the ringdown stage [53, 54, 55, 56] and null geodesics of test particles moving near the BH [57, 58, 59, 60, 61]. Recently, the geodesic instability of test particles, specifically focusing on the temporal-Lyapunov exponents of both massless and massive particles, serving as an order parameter to investigate phase transitions in numerous BH solutions in asymptotically AdS spacetime [62, 63, 64, 65, 66].

Strong gravitation near a BH can cause photons traveling close to the critical curve to orbit the BH multiple times before reaching a distant observer, forming a narrow band on the observer’s screen known as the photon ring. This ring can be characterized by three critical parameters: the orbital half-period (τ\tau), the angular-Lyapunov exponent (λL\lambda_{L}), and the temporal-Lyapunov exponent (γL\gamma_{L}), which govern the dynamics of unstable null geodesics and reveal universal properties of the photon ring independent of the distance between the light source and observer from the BH [67, 68] as well as the feature of the emitting light source. In this study, we focus on these observable parameters to decode information about the phase transitions of BHs, particularly charged AdS-BHs, within the framework of extended phase space and holographic thermodynamics. Moreover, we propose that the differences in these three critical parameters at first-order phase transitions could serve as an order parameter for studying scaling behavior near the critical point, and we also provide a mathematical derivation for the corresponding scaling law.

This paper is organized as follows: In section II, we provide a comprehensive review of BH thermodynamics, focusing on three approaches: standard phase space, extended phase space, and holographic thermodynamics. This section sets the stage for our analysis by detailing the theoretical foundations and the distinctions between these approaches. In section III, we introduce three critical parameters of the photon ring region. We discuss the orbital half-period τ\tau, the angular Lyapunov exponent λL\lambda_{L}, and the temporal Lyapunov exponent γL\gamma_{L}, explaining their significance in probing BH phase transitions with horizon-scale observations. Section IV examines the phase transitions of BHs through optical features in the extended phase space and holographic thermodynamics approaches. We analyze how the critical parameters vary with changes in the pressure and temperature, and demonstrate their potential as order parameters indicating phase transitions. In section V, we investigate the scaling behavior of the optical parameters near the critical point within both the extended phase space and holographic thermodynamics approaches. We explore the critical exponents and their implications for understanding BH phase transitions. The results of this investigation support the use of these optical critical parameters as order parameters for characterizing BH phase transitions. Finally, section VI concludes the paper by summarizing our findings and discussing the broader implications of our study for BH thermodynamics and observational astrophysics.

II Many facets of black hole thermodynamics

In this section, we provide a comprehensive review of the thermodynamics of charged AdS-BHs using three approaches: standard phase space, extended phase space, and holographic thermodynamics. The insights gained here will inform our analysis of null geodesics near the critical curve, which may reflect the phase transitions of black holes within each thermodynamic framework discussed in section IV.

II.1 Thermodynamics of charged AdS-BH within standard phase space

The action for Einstein-Maxwell gravity in (n+1)(n+1)-dimensional AdS spacetime is given by

S=116πGNdn+1xg[R2+n(n1)L2],\displaystyle S=\frac{1}{16\pi G_{N}}\int d^{n+1}x\sqrt{-g}\left[R-\mathcal{F}^{2}+\frac{n(n-1)}{L^{2}}\right], (6)

where FF is the U(1)U(1) gauge field strength tensor and Λ=n(n1)2L2\Lambda=-\frac{n(n-1)}{2L^{2}} is the negative cosmological constant with the length scale of the AdS space LL. A spherical symmetric BH solution from this action is

ds2\displaystyle ds^{2} =\displaystyle= f(r)dt2+dr2f(r)+r2dωn12,\displaystyle-f(r)dt^{2}+\frac{dr^{2}}{f(r)}+r^{2}d\omega^{2}_{n-1}, (7)

where

f(r)\displaystyle f(r) =\displaystyle= 1+r2L2mrn2+q2r2n4.\displaystyle 1+\frac{r^{2}}{L^{2}}-\frac{m}{r^{n-2}}+\frac{q^{2}}{r^{2n-4}}. (8)

Note that dωn12d\omega^{2}_{n-1} is the metric of unit n1n-1 sphere. The parameter mm is related to the mass MM of BH as follows

M=(n1)ωn116πGNm,\displaystyle M=\frac{(n-1)\omega_{n-1}}{16\pi G_{N}}m, (9)

where ωn1\omega_{n-1} is the volume of the unit n1n-1 sphere. The electric charge of BH QQ can be written in the form of the parameter qq as

Q=(n1)ωn18πGNηq,\displaystyle Q=\frac{(n-1)\omega_{n-1}}{8\pi G_{N}}\eta q, (10)

where η\eta is given by

η=2(n2)n1.\displaystyle\eta=\sqrt{\frac{2(n-2)}{n-1}}. (11)

For spherical symmetric and static charged BH, one can choose the gauge potential as

𝒜=(1ηqrn2+Φ)dt.\displaystyle\mathcal{A}=\left(-\frac{1}{\eta}\frac{q}{r^{n-2}}+\Phi\right)dt. (12)

Here, 𝒜\mathcal{A} is fixed and set to vanish at the horizon surface r+r_{+}, so that the electric potential Φ\Phi becomes

Φ=1ηqr+n2.\displaystyle\Phi=\frac{1}{\eta}\frac{q}{r_{+}^{n-2}}. (13)

The Hawking temperature of BH and its corresponding entropy are associated with its geometrical quantities as

T\displaystyle T =\displaystyle= κ2π=n24πr+(1+nn2r+2L2q2r+2n4),\displaystyle\frac{\kappa}{2\pi}=\frac{n-2}{4\pi r_{+}}\left(1+\frac{n}{n-2}\frac{r_{+}^{2}}{L^{2}}-\frac{q^{2}}{r_{+}^{2n-4}}\right), (14)
S\displaystyle S =\displaystyle= A4GN=ωn1r+n14GN,\displaystyle\frac{A}{4G_{N}}=\frac{\omega_{n-1}r_{+}^{n-1}}{4G_{N}}, (15)

where κ\kappa and AA denote the surface gravity and horizon area, respectively.

Recall that the partition function in the grand canonical ensemble 𝒵\mathcal{Z} is the Laplace transform of the density of states g(E,N)g(E,N) as follows

𝒵=g(E,N)eβ(EμN)𝑑E𝑑N,\displaystyle\mathcal{Z}=\int g(E,N)e^{-\beta(E-\mu N)}dEdN, (16)

where β,E,μ\beta,E,\mu and NN denote the inverse temperature, internal energy, chemical potential and number of particles in the system, respectively. To obtain the density of states, we apply an inverse Laplace transform to 𝒵\mathcal{Z} and then use the steepest descent method. Thus, we have

g(E,N)𝒵eβ(EμN).\displaystyle g(E,N)\sim\mathcal{Z}e^{\beta(E-\mu N)}. (17)

As the thermal entropy SS defined as the logarithm of the number of states, the above equation can give the grand potential, ΩTln𝒵\Omega\equiv-T\ln\mathcal{Z}, in the form of thermodynamic variables as

Ω=ETSμN,\displaystyle\Omega=E-TS-\mu N, (18)

The first law of thermodynamics satisfies

dΩ=SdTpd𝒱Ndμ.\displaystyle d\Omega=-SdT-pd\mathcal{V}-Nd\mu. (19)

Thus, we treat T,𝒱T,\mathcal{V} and μ\mu as independent thermodynamic variables so that we can express the grand potential as the function of these three variables, i.e., Ω=Ω(T,𝒱,μ)\Omega=\Omega(T,\mathcal{V},\mu). Applying the Legendre transformation to obtain the Helmholtz free energy of the canonical ensemble, F(T,𝒱,N)=Ω(T,𝒱,μ)+μNF(T,\mathcal{V},N)=\Omega(T,\mathcal{V},\mu)+\mu N, we have

F=ETS.\displaystyle F=E-TS. (20)

With the first law of thermodynamics, the infinitesimal change in FF reads

dF=SdTpd𝒱μdN.\displaystyle dF=-SdT-pd\mathcal{V}-\mu dN. (21)

In gravitational physics, the partition function of BHs can be approximated from the on-shell Euclidean action SES_{E} as follows [69]

𝒵eSE.\displaystyle\mathcal{Z}\sim e^{-S_{E}}. (22)

To evaluate the Einstein-Maxwell action in Eq. (6) with the potential AtA_{t} fixed at infinity, the Gibbons-Hawking boundary term vanishes due to the strength tensor \mathcal{F} becoming zero. Consequently, the partition function 𝒵\mathcal{Z} can be simply calculated from the bulk action without any additional term. Using the subtraction method, the resulting thermodynamic potential is in the form [70, 71, 72]

Tln𝒵Ω=MTSΦQ.\displaystyle-T\ln\mathcal{Z}\equiv\Omega=M-TS-\Phi Q. (23)

Comparing Eqs. (18) with (23), one may identify EE, μ\mu and NN of the thermal systems with MM, Φ\Phi and QQ of the charged BH. It is important to note that the thermodynamic potential corresponding to the grand potential Ω\Omega, in the above equation, is derived by calculating the on-shell action with potential AtA_{t} fixed at the boundary of spacetime. Recall that the potential Φ\Phi is given by taking rr\to\infty into Eq. (12). Alternatively, the thermodynamic potential corresponding to the Helmholtz free energy FF can be obtained by fixing the charge QQ at the spacetime boundary instead. In this approach, the Gibbons-Hawking term no longer vanishes but becomes significant and contributes to the on-shell action SES_{E}. The resulting Euclidean action calculation with fixed QQ at the boundary yields the Helmholtz free energy FF as the corresponding thermodynamic potential:

Tln𝒵F=MTS.\displaystyle-T\ln\mathcal{Z}\equiv F=M-TS. (24)

II.2 Extended phase space approach

As discussed in the previous subsection, it is indeed possible to associate the thermodynamic variables of BH with the geometric properties of its spacetime. However, the Euler’s theorem for a homogeneous function suggests that the Smarr formula in Eq. (2) for BHs in the spacetime with non-zero Λ\Lambda become inconsistent with the first law of black hole thermodynamics unless variations in Λ\Lambda are taken into account. Using Eqs. (2) and (3) and allowing for variations in the cosmological constant Λ\Lambda, we obtain the Smarr formula and the variation of mass in the form

M\displaystyle M =\displaystyle= n1n2TS+ΦQ2n2PV,\displaystyle\frac{n-1}{n-2}TS+\Phi Q-\frac{2}{n-2}PV, (25)
dM\displaystyle dM =\displaystyle= TdS+ΦdQ+VdP,\displaystyle TdS+\Phi dQ+VdP, (26)

respectively. Note that the computations concerning the Smarr formula and its associated first law derived from the scaling argument are detailed in Appendix A of the manuscript.

As presented in Eq. (26), the BH mass MM is a function of S,QS,Q and PP, thus it should be interpreted as the enthalpy H(S,Q,P)H(S,Q,P) rather than the internal energy E(S,Q,V)E(S,Q,V). Using the Legendre transformation E=HPVE=H-PV, treating MM to be enthalpy leads to the first law of black hole thermodynamics in the extended phase space of the form

dE=TdS+ΦdQPdV.\displaystyle dE=TdS+\Phi dQ-PdV. (27)

Significantly, the first law can now describe the change in internal energy during processes associated with changes in volume within a black hole’s event horizon when it absorbs energy or emits Hawking radiation through the PdVPdV term, akin to conventional thermodynamics.

Let us consider black hole thermodynamics in the canonical ensemble within the extended phase space approach. Eliminating LL in Eq. (14) by identifying P=38πGNL2\displaystyle P=\frac{3}{8\pi G_{N}L^{2}}, we obtain the equation of state for a 44-dimensional charged AdS-BH as follows [18]

P=T2r+18πr+2+q28πr+4.\displaystyle P=\frac{T}{2r_{+}}-\frac{1}{8\pi r_{+}^{2}}+\frac{q^{2}}{8\pi r_{+}^{4}}. (28)

By comparing the above equation with the equation of state for vdW fluids, one can find that qq relates to the gas’s constant in the vdW equation of state [18]. Using Eq. (28), we plot the isotherm curves for different temperatures, as shown in the left panel in Fig. 1. This depiction presents the critical behavior in the Pr+P-r_{+} plane. Note that the critical values rc,Tcr_{c},T_{c} and PcP_{c} of this system satisfy the following conditions:

Pr+=0,and2Pr+2=0.\displaystyle\frac{\partial P}{\partial r_{+}}=0,\ \ \ \text{and}\ \ \ \frac{\partial^{2}P}{\partial r_{+}^{2}}=0. (29)

Solving these two conditions, we obtain

rc=6q,Tc=618πq,Pc=196πq2.\displaystyle r_{c}=\sqrt{6}q,\ \ \ T_{c}=\frac{\sqrt{6}}{18\pi q},\ \ \ P_{c}=\frac{1}{96\pi q^{2}}. (30)

Since PP should have positive values for r+>0r_{+}>0, there exists a particular value of temperature T0T_{0} which is the lower bound of temperature where P=0P=0 at a particular value of horizon radius r0r_{0}. Namely,

T0=318πq,r0=3q.\displaystyle T_{0}=\frac{\sqrt{3}}{18\pi q},\ \ \ r_{0}=\sqrt{3}q. (31)

At TT below TcT_{c}, the isotherm curves in the Pr+P-r_{+} plane show the appearance of a local minimum PminP_{\text{min}} and a local maximum PmaxP_{\text{max}} at the horizon radii rminr_{\text{min}} and rmaxr_{\text{max}}, respectively. Using Eq. (30) and fixing q=1q=1, the critical parameters are given by rc=2.45r_{c}=2.45, Tc=0.0433T_{c}=0.0433, and Pc=0.0033P_{c}=0.0033. To illustrate the characteristics of the system, we present the curves for TT less than, equal to, and greater than TcT_{c}. As shown in Fig. 1, the curves for T=0.0310T=0.0310 (below the critical point), T=TcT=T_{c}, and T=0.0500T=0.0500 (above the critical point) are depicted. For T=0.0310T=0.0310, we find that Pmin=0.0001P_{\text{min}}=0.0001 and Pmax=0.0016P_{\text{max}}=0.0016, corresponding to rmin=1.7391r_{\text{min}}=1.7391 and rmax=4.6615r_{\text{max}}=4.6615, respectively.

Refer to caption
Refer to caption
Figure 1: Left: Plots of isotherm curves in the Pr+P-r_{+} plane for fixed q=1q=1. The temperature increases from bottom to top as T=0.0310,0.0433T=0.0310,0.0433 and 0.05000.0500, where Tc=0.0433T_{c}=0.0433 is the critical temperature. Right: The plots of isothermal compressibility κT\kappa_{T} as a function of r+r_{+} correspond to those of the isotherm curves shown in the left panel.

According to the Le Chatelier principle [73, 74, 75, 76], the response function associated to the isotherm curves in Pr+P-r_{+} plane is the isothermal compressibility κT\kappa_{T}, which is defined as

κT=1VVP|T=12πr+42πTr+3r+2+2q2,\displaystyle\kappa_{T}=-\frac{1}{V}\frac{\partial V}{\partial P}\Big{|}_{T}=\frac{12\pi r_{+}^{4}}{2\pi Tr_{+}^{3}-r_{+}^{2}+2q^{2}}, (32)

where V=43πr+3V=\frac{4}{3}\pi r_{+}^{3} is thermodynamic volume. The right panel in Fig. 1 displays r+r_{+} dependence of κT\kappa_{T} corresponding to isotherm curves in Pr+P-r_{+} plane in the left panel. The results reveal three branches of BHs below TcT_{c}: Small BH (red solid curve), Intermediate BH (green solid curve) and Large BH (blue solid curve) correspond to the ranges of r+<rmin,rmin<r+<rmaxr_{+}<r_{\text{min}},r_{\text{min}}<r_{+}<r_{\text{max}} and r+>rmaxr_{+}>r_{\text{max}}, respectively. Here rminr_{\text{min}} and rmaxr_{\text{max}} is the event horizon radii that κT\kappa_{T} diverge. Note that these values of r+r_{+} can be obtained by setting the denominator in Eq. (32) to be zero, namely 2πTr+3r+2+2q2=02\pi Tr_{+}^{3}-r_{+}^{2}+2q^{2}=0, and finding its roots. The Small BH and Large BH (both have κT>0\kappa_{T}>0) are mechanically stable against microscopic fluctuation while the Intermediate BH (κT<0\kappa_{T}<0) is unstable. The authors in [77] have shown that the unstable Intermediate BH in the PVP-V plane can be replaced by a straight line at P=PfP=P_{f} obeying the Maxwell equal area law, where the Small-Large BHs first-order phase transition occurs. Note that PfP_{f} can be called as the Hawking-Page pressure. Defining the reduced variables as

p=PPc,t=TTcandv=VVc,\displaystyle p=\frac{P}{P_{c}},\ \ \ t=\frac{T}{T_{c}}\ \ \ \text{and}\ \ \ v=\frac{V}{V_{c}}, (33)

an analytic expression for Hawking-Page pressure is given by [77]

p=[12cos(arccos(1t2)+π3)]2,\displaystyle p^{*}=\left[1-2\cos\left(\frac{\arccos\left(1-t^{2}\right)+\pi}{3}\right)\right]^{2}, (34)

where p=Pf/Pcp^{*}=P_{f}/P_{c} denotes the reduced Hawking-Page pressure. Considering the isotherm curve with T=0.0310T=0.0310, for example, which is less than TcT_{c} as shown in Fig. 1, the corresponding reduced temperature t=0.7159t=0.7159 can be put into the above equation such that the reduced pressure at which the first-order phase transition occurs is at p=0.4384p^{*}=0.4384, corresponding to Pf=0.0014P_{f}=0.0014. Moreover, the horizon radii for Small and Large BHs at the first-order phase transition are respectively expressed as follows

rSrc\displaystyle\frac{r_{S}}{r_{c}} =\displaystyle= 2tcos2ϕ4t2cos4ϕ2tcosϕ,\displaystyle\frac{2}{t}\cos^{2}\phi-\sqrt{\frac{4}{t^{2}}\cos^{4}\phi-\frac{\sqrt{2}}{t}\cos\phi}, (35)
rLrc\displaystyle\frac{r_{L}}{r_{c}} =\displaystyle= 2tcos2ϕ+4t2cos4ϕ2tcosϕ,\displaystyle\frac{2}{t}\cos^{2}\phi+\sqrt{\frac{4}{t^{2}}\cos^{4}\phi-\frac{\sqrt{2}}{t}\cos\phi}, (36)

where ϕ=πθ3\phi=\frac{\pi-\theta}{3} and cosθ=22t\cos{\theta}=\frac{\sqrt{2}}{2}t.

To investigate the global stability for charged AdS-BH associated with the isotherm curves in the Pr+P-r_{+} plane, one needs to consider the free energy as a function of bulk pressure PP with the temperature TT held fixed. In the fixed charge ensemble, treating MM as the internal energy results in the thermodynamic potential being the Helmholtz free energy FF, as is typical in standard phase space. However, in this section, where MM is treated as the enthalpy, the thermodynamic potential is instead the Gibbs free energy GG. We express GG in term of r+r_{+} with fixed TT in the following form

Tln𝒵G\displaystyle-T\ln\mathcal{Z}\equiv G =\displaystyle= E+PVTS\displaystyle E+PV-TS (37)
=\displaystyle= MTS\displaystyle M-TS
=\displaystyle= 2q2πr+3T+r+23r+.\displaystyle\frac{2q^{2}-\pi r_{+}^{3}T+r_{+}^{2}}{3r_{+}}.

By using dMdM from Eq. (26) with the charge fixed, the infinitesimal change of the Gibbs free energy is

dG=dMTdSSdT,\displaystyle dG=dM-TdS-SdT, (38)

can be changed to be in the form

dG=SdT+VdP.\displaystyle dG=-SdT+VdP. (39)

This relation suggests that G=G(T,P)G=G(T,P) leading to the expression for entropy and thermodynamic volume given by

(GT)P=S,(GP)T=V.\displaystyle\left(\frac{\partial G}{\partial T}\right)_{P}=-S,\ \ \ \left(\frac{\partial G}{\partial P}\right)_{T}=V. (40)
Refer to caption
Refer to caption
Refer to caption
Figure 2: The isotherm curves in the GPG-P plane for fixed q=1q=1 are plotted at different temperatures, namely the temperature 0.03100.0310 (left), 0.04330.0433 (middle), and 0.05000.0500 (right), which represent the behaviors of the systems at the temperature smaller than, equal to, and larger than TcT_{c}, respectively.

We plot GG against PP for a fixed TT in Fig. 2. It is important to note that the point r+r_{+} where κT\kappa_{T} becomes divergent corresponds to a cusp in the G(P)G(P) graph. This feature reflects the second-order phase transition resulting from κT=1V(2GP2)T\kappa_{T}=-\frac{1}{V}\left(\frac{\partial^{2}G}{\partial P^{2}}\right)_{T}. As shown in the left panel in Fig. 2, the free energy in the case of T<TcT<T_{c}, exhibits the swallowtail behavior with two cusps at PminP_{\text{min}} and PmaxP_{\text{max}} corresponding to the horizon radii that κT\kappa_{T} diverges rminr_{\text{min}} and rmaxr_{\text{max}}, respectively. Considering the bulk pressure PP running from a low value where only Large-BH exists, Small and Intermediate BHs both emerge at P=PminP=P_{\text{min}} with free energy larger than the Large BH. Obviously, Large BH is thermodynamically preferred in the region of P<PfP<P_{f}. It turns out that the system encounters the first-order phase transition at PfP_{f} such that Small BH phase becomes thermodynamically preferred when P>PfP>P_{f}. The middle panel in Fig. 2 shows the case of T=TcT=T_{c} where the Small-Large BHs transition becomes the second-order type of phase transition at P=PcP=P_{c}. The right panel in Fig. 2 shows that only one phase of BH exists at high temperature T>TcT>T_{c}, here at T=0.0500T=0.0500, where κT>0\kappa_{T}>0 at all values of PP. Remarkably, by identifying Small, Intermediate, and Large BHs as liquid, metastable, and gas phases, respectively, the behaviors of the BH phase transition is in a similar way as what occurs in the vdW fluid.

II.3 Holographic thermodynamics approach

As mentioned in section I, Introduction, the pressure PP and thermodynamic volume VV characterizing BHs in the bulk through the extended phase space approach do not directly correspond to the pressure pp and volume 𝒱\mathcal{V} of the dual field theory on the boundary within the AdS/CFT correspondence. Additionally, the Smarr formula, which relates the mechanical quantities of BHs to thermodynamic variables, depends on the number of spacetime dimensions, this feature is typically absent in the Euler equation for ordinary matter. Recently, the holographic thermodynamics approach has provided insights into effectively addressing these two issues in associating the black hole thermodynamic properties with its dual field theory counterpart. In this section, we examine the thermodynamics of CFT, which is holographically dual to the charged AdS-BH, within the holographic thermodynamics approach.

Introducing the central charge 𝒞\mathcal{C} and its chemical potential μ𝒞\mu_{\mathcal{C}} as a new pair of thermodynamic variable in the large-NN gauge theory, the author in [31] suggests that the scaling behavior of its internal energy EE may be written as

E(αS,α0𝒱,αBi,α𝒞)=αE(S,𝒱,Bi,𝒞),\displaystyle E(\alpha S,\alpha^{0}\mathcal{V},\alpha B_{i},\alpha\mathcal{C})=\alpha E(S,\mathcal{V},B_{i},\mathcal{C}), (41)

where BiB_{i} denote the conserved quantities, such as charge and angular momentum. The Euler equation and its corresponding first law due to the Euler scaling argument can be obtained by using Eqs. (125) and (126) in Appendix A as

E\displaystyle E =\displaystyle= TS+νiBi+μ𝒞𝒞,\displaystyle TS+\nu^{i}B_{i}+\mu_{\mathcal{C}}\mathcal{C}, (42)
dE\displaystyle dE =\displaystyle= TdSpd𝒱+νidBi+μ𝒞d𝒞,\displaystyle TdS-pd\mathcal{V}+\nu^{i}dB_{i}+\mu_{\mathcal{C}}d\mathcal{C}, (43)

where

T=(ES)𝒱,Bi,𝒞,p=(E𝒱)S,Bi,𝒞,νi=(EBi)S,𝒱,𝒞,μ𝒞=(E𝒞)S,𝒱,Bi.\displaystyle T=\left(\frac{\partial E}{\partial S}\right)_{\mathcal{V},B_{i},\mathcal{C}},p=-\left(\frac{\partial E}{\partial\mathcal{V}}\right)_{S,B_{i},\mathcal{C}},\nu^{i}=\left(\frac{\partial E}{\partial B_{i}}\right)_{S,\mathcal{V},\mathcal{C}},\mu_{\mathcal{C}}=\left(\frac{\partial E}{\partial\mathcal{C}}\right)_{S,\mathcal{V},B_{i}}. (44)

Note that the p𝒱p\mathcal{V} term is absent in the Euler equation but the pd𝒱pd\mathcal{V} term appear in the first law. Since the variation of EE in Eq. (42) should be equal to the first law in Eq (43), i.e., TdS+SdT+νidBi+Bidνi+μ𝒞d𝒞+𝒞dμ𝒞=TdSpd𝒱+νidBi+μd𝒞TdS+SdT+\nu^{i}dB_{i}+B_{i}d\nu^{i}+\mu_{\mathcal{C}}d\mathcal{C}+\mathcal{C}d\mu_{\mathcal{C}}=TdS-pd\mathcal{V}+\nu^{i}dB_{i}+\mu d\mathcal{C}, this implies the Gibbs-Duhem equation as follows

SdT+Bidνi+𝒞dμ𝒞=pd𝒱.\displaystyle SdT+B_{i}d\nu^{i}+\mathcal{C}d\mu_{\mathcal{C}}=-pd\mathcal{V}. (45)

The grand potential is given by

Ω\displaystyle\Omega =\displaystyle= ETSνiBi,\displaystyle E-TS-\nu^{i}B_{i}, (46)
=\displaystyle= (TS+νiBi+μ𝒞𝒞)TSνiBi,\displaystyle(TS+\nu^{i}B_{i}+\mu_{\mathcal{C}}\mathcal{C})-TS-\nu^{i}B_{i},
=\displaystyle= μ𝒞𝒞,\displaystyle\mu_{\mathcal{C}}\mathcal{C},

where we have substituted EE by using Eq. (42).

Since 𝒞\mathcal{C} depends on both LL and GNG_{N} due to the holographic dictionary as shown in Eq. (5), thus varying 𝒞\mathcal{C} in the dual field theory should be equivalent to variation of Λ\Lambda and GNG_{N} in the gravity side. By including Λ\Lambda and GNG_{N} into the mass formula of AdS-BH in Eq. (127), an infinitesimal change of MM can be obtained from Eq. (126) as

dM=κ8πGNdA+ΦdQ+Θ8πGNdΛ(MΦQ)dGNGN,\displaystyle dM=\frac{\kappa}{8\pi G_{N}}dA+\Phi dQ+\frac{\Theta}{8\pi G_{N}}d\Lambda-(M-\Phi Q)\frac{dG_{N}}{G_{N}}, (47)

where an additional partial derivative of MM with respect to GNG_{N} is MGN=(MΦQ)GN\frac{\partial M}{\partial G_{N}}=-\frac{(M-\Phi Q)}{G_{N}}. Substituting Θ\Theta in term of other variables by using the Smarr formula in Eq. (2) with Λ\Lambda expressed in term of LL, the above equation becomes

dM=κ2πd(A4GN)+ΦLd(QL)Mn1dLn1Ln1+(MκA8πGNΦQ)d(Ln1/GN)Ln1/GN.\displaystyle dM=\frac{\kappa}{2\pi}d\left(\frac{A}{4G_{N}}\right)+\frac{\Phi}{L}d(QL)-\frac{M}{n-1}\frac{dL^{n-1}}{L^{n-1}}+\left(M-\frac{\kappa A}{8\pi G_{N}}-\Phi Q\right)\frac{d(L^{n-1}/G_{N})}{L^{n-1}/G_{N}}.
(48)

Thermodynamic variables of dual field theory can be identified with geometric quantities of AdS-BH as

E=M,Φ~=ΦL,Q~=QL,𝒱Ln1,𝒞Ln1GN.\displaystyle E=M,\ \ \ \tilde{\Phi}=\frac{\Phi}{L},\ \ \ \tilde{Q}=QL,\ \ \ \mathcal{V}\sim L^{n-1},\ \ \ \mathcal{C}\sim\frac{L^{n-1}}{G_{N}}. (49)

Substituting them into Eq. (48) and comparing with Eq. (43), we obtain

dE=TdS+Φ~dQ~pd𝒱+μ𝒞d𝒞,\displaystyle dE=TdS+\tilde{\Phi}d\tilde{Q}-pd\mathcal{V}+\mu_{\mathcal{C}}d\mathcal{C}, (50)

where

p=E(n1)𝒱,μ𝒞=1𝒞(ETSΦ~Q~).\displaystyle p=\frac{E}{(n-1)\mathcal{V}},\ \ \ \mu_{\mathcal{C}}=\frac{1}{\mathcal{C}}\left(E-TS-\tilde{\Phi}\tilde{Q}\right). (51)

Notably, the former relation in (51) expresses the pressure pp, in the field theory side, satisfies the equation of state in Eq. (4), while the latter gives the Euler equation corresponding to Eq.(42) with identifying ν=Φ~\nu=\tilde{\Phi} and B=Q~B=\tilde{Q}. It is interesting to note that the CFT in the field theory side from the approach of holographic thermodynamics can live in the curved spacetime with the curvature radius RR, which has a value not necessarily equal to the bulk AdS radius LL. This can be seen by redefining the thermodynamic variables as [31, 32]

E=MLR,T=κ2πLR,S=A4GN,Φ~=ΦLLR,Q~=QL.\displaystyle E=M\frac{L}{R},\ \ \ T=\frac{\kappa}{2\pi}\frac{L}{R},\ \ \ S=\frac{A}{4G_{N}},\ \ \ \tilde{\Phi}=\frac{\Phi}{L}\frac{L}{R},\ \ \ \tilde{Q}=QL. (52)

From the resulting first law, we have 𝒱Rn1\mathcal{V}\sim R^{n-1} and 𝒞Ln1GN\displaystyle\mathcal{C}\sim\frac{L^{n-1}}{G_{N}}. Thus, the volume 𝒱\mathcal{V} and central charge 𝒞\mathcal{C} of the dual CFT are now independent due to the choice of rescaling as shown in Eq. (52).

It is worthwhile to review here some interesting results about the thermodynamics of CFT that are holographically dual to charged AdS-BH within the novel holographic thermodynamics approach. For simplicity, we define the dimensionless quantities:

x=r+L,y=qLn2.\displaystyle x=\frac{r_{+}}{L},\ \ \ y=\frac{q}{L^{n-2}}. (53)

The thermodynamic quantities of the dual CFT, which is in nn-dimensional spacetime of curvature radius RR, can be written in parametric equations of xx and yy as follows [32]

E\displaystyle E =\displaystyle= (n1)𝒞xn2R(1+x2+y2x2n4),\displaystyle\frac{(n-1)\mathcal{C}x^{n-2}}{R}\left(1+x^{2}+\frac{y^{2}}{x^{2n-4}}\right), (54)
T\displaystyle T =\displaystyle= n24πRx(1+nn2x2q2x2n4,),\displaystyle\frac{n-2}{4\pi Rx}\left(1+\frac{n}{n-2}x^{2}-\frac{q^{2}}{x^{2n-4}},\right), (55)
S\displaystyle S =\displaystyle= 4π𝒞xn1,\displaystyle 4\pi\mathcal{C}x^{n-1}, (56)
Q~\displaystyle\tilde{Q} =\displaystyle= 2η(n1)𝒞y,\displaystyle 2\eta(n-1)\mathcal{C}y, (57)
Φ~\displaystyle\tilde{\Phi} =\displaystyle= 1ηRyxn2,\displaystyle\frac{1}{\eta R}\frac{y}{x^{n-2}}, (58)
μ𝒞\displaystyle\mu_{\mathcal{C}} =\displaystyle= xn2R(1x2y2x2n4).\displaystyle\frac{x^{n-2}}{R}\left(1-x^{2}-\frac{y^{2}}{x^{2n-4}}\right). (59)

Here, the field theory under consideration lives in the spacetime with n=3n=3 dimensions. The CFT is in the ensemble with fixed (Q~,𝒱,𝒞)(\tilde{Q},\mathcal{V},\mathcal{C}), corresponding to the canonical ensemble. The temperature and heat capacity of the dual CFT are given by

T\displaystyle T =\displaystyle= 14πRx(1+3x2+Q216𝒞2x2),\displaystyle\frac{1}{4\pi Rx}\left(1+3x^{2}+\frac{Q^{2}}{16\mathcal{C}^{2}x^{2}}\right), (60)
CQ~,𝒱,𝒞\displaystyle C_{\tilde{Q},\mathcal{V},\mathcal{C}} =\displaystyle= T(ST)Q~,𝒱,𝒞=8π𝒞x2(16𝒞2x2(3x2+1)Q~2)16𝒞2x2(3x21)+3Q~2.\displaystyle T\left(\frac{\partial S}{\partial T}\right)_{\tilde{Q},\mathcal{V},\mathcal{C}}=\frac{8\pi\mathcal{C}x^{2}\left(16\mathcal{C}^{2}x^{2}(3x^{2}+1)-\tilde{Q}^{2}\right)}{16\mathcal{C}^{2}x^{2}(3x^{2}-1)+3\tilde{Q}^{2}}. (61)

From Eq. (60), the horizon radius of extremal BH (T=0)(T=0) can be written as a function of Q~\tilde{Q} and 𝒞\mathcal{C} as

xext2=2𝒞2+4𝒞4+3𝒞2Q~212𝒞2.\displaystyle x_{\text{ext}}^{2}=\frac{-2\mathcal{C}^{2}+\sqrt{4\mathcal{C}^{4}+3\mathcal{C}^{2}\tilde{Q}^{2}}}{12\mathcal{C}^{2}}. (62)

Due to the TxT-x criticality of charged AdS-BH, thermodynamics of dual CFT also critically change in a similar way as the ratio Q~/𝒞\tilde{Q}/\mathcal{C} crosses the critical point, which can be determined from the point of inflection as follows

(Tx)=0,and(2Tx2)=0.\displaystyle\left(\frac{\partial T}{\partial x}\right)=0,\ \ \ \text{and}\ \ \ \left(\frac{\partial^{2}T}{\partial x^{2}}\right)=0. (63)

Solving these two conditions, we obtain

xc=16,Tc=23π,Sc=2π𝒞3,Q~c𝒞c=23.\displaystyle x_{c}=\frac{1}{\sqrt{6}},\ \ \ T_{c}=\frac{\sqrt{2}}{\sqrt{3}\pi},\ \ \ S_{c}=\frac{2\pi\mathcal{C}}{3},\ \ \ \frac{\tilde{Q}_{c}}{\mathcal{C}_{c}}=\frac{2}{3}. (64)
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 3: (a) Plots of isocharge curves in the TxT-x plane for Q~=0.2,2/3,1\tilde{Q}=0.2,2/3,1 with fixed 𝒞=1\mathcal{C}=1. (b) Plots of the heat capacity CQ~,𝒱,𝒞C_{\tilde{Q},\mathcal{V},\mathcal{C}} as a function of xx that corresponds to figure (a). (c) The dependence of TT on parameter xx, while keeping Q~=1\tilde{Q}=1 fixed with varying 𝒞=0.5,3/2,3\mathcal{C}=0.5,3/2,3. (d) Plots of heat capacity CQ~,𝒱,𝒞C_{\tilde{Q},\mathcal{V},\mathcal{C}} versus xx that corresponds to figure (c). Note that the critical values of electric charge and central charge are Q~c=2/3\tilde{Q}_{c}=2/3 and 𝒞c=3/2\mathcal{C}_{c}=3/2, respectively.

We examine thermal phase structures of dual CFT in two cases: (I) fixed 𝒞\mathcal{C} with different values of Q~\tilde{Q} and (II) fixed Q~\tilde{Q} with different values of 𝒞\mathcal{C}. We illustrate the behaviors of TT and CQ~,𝒱,𝒞C_{\tilde{Q},\mathcal{V},\mathcal{C}} as functions of xx for the former and latter cases in the first and second rows of Fig. 3, respectively. Note that xmaxx_{\text{max}} and xminx_{\text{min}} are given by

xmax,min2=2𝒞24𝒞49𝒞2Q~212𝒞2.\displaystyle x_{\text{max,min}}^{2}=\frac{2\mathcal{C}^{2}\mp\sqrt{4\mathcal{C}^{4}-9\mathcal{C}^{2}\tilde{Q}^{2}}}{12\mathcal{C}^{2}}. (65)

Two radii xmaxx_{\text{max}} and xminx_{\text{min}} indicate positions of local maximum and minimum of Hawking temperature as shown in Fig. 3 (a) and (c), namely Tmax=T(xmax)T_{\text{max}}=T(x_{\text{max}}) and Tmin=T(xmin)T_{\text{min}}=T(x_{\text{min}}). At the critical value of Q~/𝒞=2/3\tilde{Q}/\mathcal{C}=2/3, these two radii coincide at xcx_{c}, which corresponds to the critical temperature TcT_{c}. Remarkably, there exist three thermal states of dual CFT for Q~<Q~c\tilde{Q}<\tilde{Q}_{c} in case (I) and 𝒞>𝒞c\mathcal{C}>\mathcal{C}_{c} in case (II). These three states consist of pCFT1 (red), nCFT (green), and pCFT2 (blue), which refers to the states of positive heat capacity when xext<x<xmaxx_{\text{ext}}<x<x_{\text{max}}, negative heat capacity when xmax<x<xminx_{\text{max}}<x<x_{\text{min}}, and positive heat capacity when x>xminx>x_{\text{min}}, respectively. It is important to note that our notation might cause some confusion when xmaxx_{\text{max}} is less than xminx_{\text{min}}. As defined above, xmaxx_{\text{max}} and xminx_{\text{min}} in this paper refer to the points at which the temperature TT is at its maximum and minimum, respectively. Please do not be confused by these terms.

Table 1: The numerical values of parameters that characterize the phase transition of charged AdS-BH in cases I and II from the holographic thermodynamics.
Case I Case II
Q~=0.200\tilde{Q}=0.200 Q~c=0.667\tilde{Q}_{c}=0.667 Q~=1.00\tilde{Q}=1.00 𝒞=0.500\mathcal{C}=0.500 𝒞c=1.50\mathcal{C}_{c}=1.50 𝒞=3.00\mathcal{C}=3.00
xextx_{\text{ext}} 0.0498 0.161 0.232 0.408 0.161 0.0825
xminx_{\text{min}} 0.571 - - - - 0.558
xmaxx_{\text{max}} 0.0876 - - - - 0.149
xcx_{c} - 0.408 - - 0.408 -
TminT_{\text{min}} 0.275 - - - - 0.273
TmaxT_{\text{max}} 0.633 - - - - 0.403
TcT_{c} - 0.260 - - 0.260 -
TfT_{f} 0.302 - - - - 0.291

Using the Maxwell’s equal area law, the curve associated with the unstable nCFT phase in the TxT-x plane in cases (I) and (II) can be replaced by a horizontal line, which indicates the first-order phase transition between these pCFT1 and pCFT2. In Appendix B, we elaborate on the detailed calculations using the method employed by [77] to obtain the Hawking-Page phase transition temperature TfT_{f} within the holographic thermodynamics framework. The temperatures for cases I and II are as follows:

t=3+3(q3)(q1)q23+3(q3)(q1)2q,\displaystyle t^{*}=\frac{3+\sqrt{3(q-3)(q-1)}-q}{2\sqrt{3+\sqrt{3(q-3)(q-1)}-2q}}, (66)

and

t\displaystyle t^{*} =\displaystyle= 3c+3(13c)(1c)123c2+c3(13c)(1c)2c,\displaystyle\frac{3c+\sqrt{3(1-3c)(1-c)}-1}{2\sqrt{3c^{2}+c\sqrt{3(1-3c)(1-c)}-2c}}, (67)

respectively, where t=Tf/Tc,q=Q~/Q~ct^{*}=T_{f}/T_{c},q=\tilde{Q}/\tilde{Q}_{c} and c=𝒞/𝒞cc=\mathcal{C}/\mathcal{C}_{c}. The horizon radii for Small and Large BHs of case I and II are expressed respectively as follows

xSxc=q3+3(q3)(q1)2q,\displaystyle\frac{x_{S}}{x_{c}}=\frac{q}{\sqrt{3+\sqrt{3(q-3)(q-1)}-2q}}, (68)
xLxc=3+3(q3)(q1)2q,\displaystyle\frac{x_{L}}{x_{c}}=\sqrt{3+\sqrt{3(q-3)(q-1)}-2q}, (69)

and

xSxc=13c2+c3(13c)(1c)2c,\displaystyle\frac{x_{S}}{x_{c}}=\frac{1}{\sqrt{3c^{2}+c\sqrt{3(1-3c)(1-c)}-2c}}, (70)
xLxc=1c3c2+c3(13c)(1c)2c.\displaystyle\frac{x_{L}}{x_{c}}=\frac{1}{c}\sqrt{3c^{2}+c\sqrt{3(1-3c)(1-c)}-2c}. (71)

Note that we summarize our numerical results of some important parameters in our study in Table 1.

The thermodynamic potential associated with fixed (Q~,𝒱,𝒞)(\tilde{Q},\mathcal{V},\mathcal{C}) ensemble is the Helmholtz free energy

FETS=𝒞xn2R(1x2+(2n3)4η2(n1)2𝒞2Q~2x2n4).\displaystyle F\equiv E-TS=\frac{\mathcal{C}x^{n-2}}{R}\left(1-x^{2}+\frac{(2n-3)}{4\eta^{2}(n-1)^{2}\mathcal{C}^{2}}\frac{\tilde{Q}^{2}}{x^{2n-4}}\right). (72)

The variation of FF reads

dF\displaystyle dF =\displaystyle= dETdSSdT,\displaystyle dE-TdS-SdT, (73)
=\displaystyle= (TdS+Φ~dQ~pd𝒱+μ𝒞d𝒞)TdSSdT,\displaystyle(TdS+\tilde{\Phi}d\tilde{Q}-pd\mathcal{V}+\mu_{\mathcal{C}}d\mathcal{C})-TdS-SdT,
=\displaystyle= SdT+Φ~dQ~pd𝒱+μ𝒞d𝒞,\displaystyle-SdT+\tilde{\Phi}d\tilde{Q}-pd\mathcal{V}+\mu_{\mathcal{C}}d\mathcal{C},

where we have used Eq. (50) for dEdE, this yields

(FT)Q~,𝒱,𝒞=S,(FQ~)T,𝒱,𝒞=Φ~,(F𝒱)T,Q~,𝒞=p,(F𝒞)T,Q~,𝒱=μ𝒞.\displaystyle\left(\frac{\partial F}{\partial T}\right)_{\tilde{Q},\mathcal{V},\mathcal{C}}=-S,\ \ \ \left(\frac{\partial F}{\partial\tilde{Q}}\right)_{T,\mathcal{V},\mathcal{C}}=\tilde{\Phi},\ \ \ \left(\frac{\partial F}{\partial\mathcal{V}}\right)_{T,\tilde{Q},\mathcal{C}}=-p,\ \ \ \left(\frac{\partial F}{\partial\mathcal{C}}\right)_{T,\tilde{Q},\mathcal{V}}=\mu_{\mathcal{C}}. (74)

As pointed out in [32], there is no critical point found in the p𝒱p-\mathcal{V} plane in the dual CFT. This result from the holographic thermodynamics is different from the PVP-V criticality behavior from the extended phase space approach as discussed in last subsection. Therefore, we will consider FF as a function of TT instead of pp with constant Q~\tilde{Q} or 𝒞\mathcal{C} to investigate the global stability of dual CFT. By considering xx as a parameter, we parametrically plot between FF in Eq.(72) and TT in Eq.(55) for dual CFT in the case (I) and (II) as shown in the left and right panels of Fig 4, respectively. There are cusp points in F(T)F(T) curve where CQ~,𝒱,𝒞C_{\tilde{Q},\mathcal{V},\mathcal{C}} diverge. At these points, there are the second-order phase transitions occuring, manifested from the fact that (2FT2)Q~,𝒱,𝒞=CQ~,𝒱,𝒞T\left(\frac{\partial^{2}F}{\partial T^{2}}\right)_{\tilde{Q},\mathcal{V},\mathcal{C}}=-\frac{C_{\tilde{Q},\mathcal{V},\mathcal{C}}}{T}, which can be derived via the first relation in Eq. (74).

Refer to caption
Refer to caption
Figure 4: Plots of FF versus TT of 33-dimensional CFT which holographically dual to 44-dimensional charged AdS-BH. Note that the radius of curvature R=1R=1 in these plots. Left: We fix 𝒞=1\mathcal{C}=1 with different values of Q~=0.2,0.4,2/3,1\tilde{Q}=0.2,0.4,2/3,1 (pink, dark-green, dark-blue and purple). Right: We fix Q~=1\tilde{Q}=1 and vary 𝒞=1,3/2,2,3\mathcal{C}=1,3/2,2,3 (pink, dark-green, dark-blue and purple). For Q~<Q~c\tilde{Q}<\tilde{Q}_{c} (𝒞>𝒞c)(\mathcal{C}>\mathcal{C}_{c}), the free energy shows the swallowtail behavior and first-order phase transition occur.
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 5: The behaviors of F(T)F(T) curves in the case (I) and (II) are shown in the first row and second rows, respectively. The swallowtail behavior of F(T)F(T) occurs when Q~<Q~c\tilde{Q}<\tilde{Q}_{c} and 𝒞>𝒞c\mathcal{C}>\mathcal{C}_{c} for case (I) and (II) are shown in figures (a) and (f), respectively, where have three thermal states, namely pCFT1, nCFT and pCFT2, represented in red, green and blue curves, respectively.

We will discuss about the phase transition of dual CFT via its F(T)F(T) curve in more details as follows. First, let us consider in the case (I). We depicted the F(T)F(T) curves when Q~<Q~c\tilde{Q}<\tilde{Q}_{c}, Q~=Q~c\tilde{Q}=\tilde{Q}_{c} and Q~>Q~c\tilde{Q}>\tilde{Q}_{c} in figures (a), (b) and (c) of Fig 5. The free energy shows the swallowtail shape for Q~<Q~c\tilde{Q}<\tilde{Q}_{c}. In this case, the pCFT1 phase is thermodynamically preferred in low temperatures, i.e. both in the region of T<TminT<T_{\text{min}} and TminT<TfT_{\text{min}}\leq T<T_{f}. There is a first-order phase transition from pCFT1 to pCFT2 occurring at TfT_{f}. At slightly higher than TfT_{f}, the pCFT2 phase becomes the most thermodynamically preferred since it has the lowest free energy. The swallowtail shape of F(T)F(T) curve with Q~=Q~c\tilde{Q}=\tilde{Q}_{c} shrink to appear as a kink at TcT_{c}, where the transition between pCFT1 and pCFT2 phases become the second-order type of phase transition. For Q~>Q~c\tilde{Q}>\tilde{Q}_{c}, there is only one phase with continuous curves with positive heat capacity for any TT, thus no phase transition occurs.

We plot FF versus TT in the case (II) in Fig 5 (d), (e) and (f) from small to large values of 𝒞\mathcal{C}. The results indicate that phase structures of dual CFT in case (II) are qualitatively similar to the case (I). Namely, F(T)F(T) displays the swallowtail shape of three phases for 𝒞>𝒞c\mathcal{C}>\mathcal{C}_{c}, the pCFT1-pCFT2 first-order phase transition occurs at T=TfT=T_{f}. At the critical value of 𝒞\mathcal{C}, the nCFT disappears and the transition between pCFT1 and pCFT2 becomes the second-order phase transition at TcT_{c}. For the small values of 𝒞\mathcal{C}, there is only one thermal state of CFT with positive heat capacity at any value of TT.

III Critical Parameters in Photon Ring Region

The photon trajectories play a crucial role in determining the images of BH surrounded by emitting matter. Computing such null geodesics has become particularly interesting since the first image of a black hole was published by the Event Horizon Telescope Collaboration [41, *ETH2, *ETH3, *ETH4, *ETH5, *ETH6, *EventHorizonTelescope:2022wkp, *EventHorizonTelescope:2022apq, *EventHorizonTelescope:2022wok, *EventHorizonTelescope:2022exc, *EventHorizonTelescope:2022urf, *EventHorizonTelescope:2022xqj]. In this section, we will explore null geodesics in the spacetime of charged AdS black holes. This analysis aims to provide a clear understanding of three critical parameters in the photon ring region: the orbital half-period τ\tau, the angular-Lyapunov exponent λL\lambda_{L}, and the temporal-Lyapunov exponent γL\gamma_{L}. These parameters play an important role in investigating black hole phase transitions through the BH’s optical characteristics.

The Lagrangian of test particle moving in the curved spacetime of the metric in Eq. (7) is given by

2=f(r)(dtds)2+1f(r)(drds)2+r2(dϕds)2,\displaystyle 2\mathcal{L}=-f(r)\left(\frac{dt}{ds}\right)^{2}+\frac{1}{f(r)}\left(\frac{dr}{ds}\right)^{2}+r^{2}\left(\frac{d\phi}{ds}\right)^{2}, (75)

where ss denotes the affine parameter for a null geodesic. Since the spacetime is static and spherical symmetric, so the motion of test particle is confined in a plane that we can choose θ=π/2\theta=\pi/2 (equatorial plane) without loss of generality. The coordinates tt and ϕ\phi are cyclic coordinates due to the symmetry of the background, resulting in the energy ω\omega and angular momentum \ell of test particles can take the form

ω=f(r)dtds,and=r2dϕds,\displaystyle\omega=f(r)\frac{dt}{ds},\ \ \ \text{and}\ \ \ \ell=r^{2}\frac{d\phi}{ds}, (76)

where they are the constant of motion. The equation of motion in the radial direction reads

(drds)2+Veff=ω2,\displaystyle\left(\frac{dr}{ds}\right)^{2}+V_{\text{eff}}=\omega^{2}, (77)

where the effective potential VeffV_{\text{eff}} is given by

Veff=f(r)(2r2+δ1).\displaystyle V_{\text{eff}}=f(r)\left(\frac{\ell^{2}}{r^{2}}+\delta_{1}\right). (78)

Note that δ1=0\delta_{1}=0 and 11 correspond to null-like and time-like geodesics, respectively. In the following formulas, we only consider the case of photon trajectories, i.e., δ1=0\delta_{1}=0. The unstable circular orbit can be determined by

Veff(ru)=ω2,andVeff(ru)=0,\displaystyle V_{\text{eff}}(r_{u})=\omega^{2},\ \ \ \text{and}\ \ \ V^{\prime}_{\text{eff}}(r_{u})=0, (79)

where rur_{u} is the radius of unstable photon sphere. Using the first condition in (79) together with (77), we obtain

bc2=(ω)2=ru2f(ru),\displaystyle b_{c}^{2}=\left(\frac{\ell}{\omega}\right)^{2}=\frac{r_{u}^{2}}{f(r_{u})}, (80)

where bcb_{c} denotes a critical impact parameter. Solving the second condition in Eq. (79), we obtain

ru\displaystyle r_{u} =\displaystyle= 14(3m±9m232q2),\displaystyle\frac{1}{4}\left(3m\pm\sqrt{9m^{2}-32q^{2}}\right), (81)
=\displaystyle= 14[3r+(1+r+2L2+q2r+2)±9r+2(1+r+2L2+q2r+2)232q2]\displaystyle\frac{1}{4}\left[3r_{+}\left(1+\frac{r_{+}^{2}}{L^{2}}+\frac{q^{2}}{r_{+}^{2}}\right)\pm\sqrt{9r_{+}^{2}\left(1+\frac{r_{+}^{2}}{L^{2}}+\frac{q^{2}}{r_{+}^{2}}\right)^{2}-32q^{2}}\right]

which is independent of angular momentum \ell.

III.1 Orbital half-period

The incoming photons with impact parameter bb larger than and close to bcb_{c} can wind several times around the BH before scattering back to the asymptotic region of spacetime. This effect can cause the light to take different paths around the BH, leading to the formation of multiple images of the source as seen by the distant observer.

Remarkably, the travel time before photons arrive at the observer screen along their null geodesics in different images is not generally equal because of the strong gravity caused by BH. This phenomenon is called the gravitational time delay, which could be measured via the gravitational lensing observation.

It is found that the time delay between successive images in the photon ring region is governed by the time-lapse τ\tau over each half-orbit of a photon sphere [78, 79, 80]. The half-time period τ\tau depends only on the metric spacetime. In other words, τ\tau turns out to be independent of the distance of the light source from observers. This implies that τ\tau is the parameter that is appropriate to encode the important feature about the region of spacetime near the event horizon.

To obtain the orbital half-period τ\tau of the bound photon orbit, we can start by calculating the angular velocity of incoming photons Ω=dϕ/dt\Omega=d\phi/dt measured by a distant observer. Using Eq. (76), we obtain

dϕds\displaystyle\frac{d\phi}{ds} =\displaystyle= r2,\displaystyle\frac{\ell}{r^{2}},
dtdsdϕdt\displaystyle\frac{dt}{ds}\frac{d\phi}{dt} =\displaystyle= r2,\displaystyle\frac{\ell}{r^{2}},
Ω\displaystyle\Omega =\displaystyle= bf(r)r2,\displaystyle b\frac{f(r)}{r^{2}}, (82)

where we have used b=/ωb=\ell/\omega. At the photon sphere radius rur_{u}, the angular velocity Ω\Omega is given by

Ωu=bcf(ru)ru2=f(ru)ru.\displaystyle\Omega_{u}=b_{c}\frac{f(r_{u})}{r_{u}^{2}}=\frac{\sqrt{f(r_{u})}}{r_{u}}. (83)

Since the time lapse over a full orbit is T=2π/ΩT=2\pi/\Omega, therefore the orbital half-period τ\tau reads

τ=T2=πruf(ru).\displaystyle\tau=\frac{T}{2}=\frac{\pi r_{u}}{\sqrt{f(r_{u})}}. (84)

III.2 Angular-Lyapunov exponent

As evident from the fact that Veff(ru)=0V^{\prime}_{\text{eff}}(r_{u})=0 and Veff′′(ru)<0V^{\prime\prime}_{\text{eff}}(r_{u})<0, the orbiting photons at rur_{u} are unstable. The Lyapunov exponents measure the sensitivity of the system to changes in initial conditions, which in this work refers to slight alterations in the photon trajectories near rur_{u} from one to a neighboring one. In this subsection and the following one, we investigate the sub-rings structure of photon trajectories in the region very close to the unstable photon sphere by using two Lyapunov exponents, namely the angular-Lyapunov exponent λL\lambda_{L} and temporal-Lyapunov exponent γL\gamma_{L} [81, 82]. Studying BH’s images of some given emitting light sources indicates that the ratio of photon flux received between adjacent sub-rings is determined by λL\lambda_{L} [67, 83, 82].

To obtain λL\lambda_{L}, we linearize Eq. (77) near the unstable shell of photon sphere in δr=rru\delta r=r-r_{u}. The effective potential becomes Veff(r)=Veff(ru+δr)V_{\text{eff}}(r)=V_{\text{eff}}(r_{u}+\delta r). Consequently, one finds that

dδrds=ω2Veff(ru+δr),\displaystyle\frac{d\delta r}{ds}=\sqrt{\omega^{2}-V_{\text{eff}}(r_{u}+\delta r)}, (85)

Expanding VeffV_{\text{eff}} around rur_{u} and using two conditions in Eq. (79), we obtain

dδrds=12Veff′′(ru)δr.\displaystyle\frac{d\delta r}{ds}=\sqrt{-\frac{1}{2}V^{\prime\prime}_{\text{eff}}(r_{u})}\delta r. (86)

Applying the second relation of Eq. (76) into Eq. (86), which leads to the formula describing the angular ϕ\phi dependence of the deviation δr\delta r from rur_{u} as follows:

πdδrdϕ=λLδr,\displaystyle\pi\frac{d~{}\delta r}{d\phi}=\lambda_{L}~{}\delta r, (87)

where we have defined the angular Lyapunov exponent as

λLπru212Veff′′(ru).\displaystyle\lambda_{L}\equiv\pi r_{u}^{2}\sqrt{-\frac{1}{2}V_{\text{eff}}^{\prime\prime}(r_{u})}. (88)

The angular dependence of the deviation δr(ϕ)\delta r(\phi) is given by

δr(ϕ)=δr0eλLϕ/π,\displaystyle\delta r(\phi)=\delta r_{0}e^{\lambda_{L}\phi/\pi}, (89)

where δr0\delta r_{0} is the initial deviation of a geodesic from the critical circular orbit.

III.3 Temporal-Lyapunov exponent

Gravitational waves emitted by BHs during the ringdown phase are characterized by QNMs. However, determining the QNMs from the field perturbation of BH can be accomplished by examining the unstable circular geodesics of photons. Specifically, the temporal-Lyapunov exponent γL\gamma_{L} is related to the imaginary part of the complex quasinormal frequency. Namely, we have 1/γL1/\gamma_{L} as the timescale for the instability of the ringdown amplitude. Consequently, γL\gamma_{L} associated with the photon ring region serves as an important parameter for studying BHs under various scenarios.

We calculate γL\gamma_{L}, which represents the deviation rate in time tt of photon trajectory from the unstable photon sphere. With the first relation of Eq. (76), we obtain the rate from Eq. (86) as follows:

dδrdt=γLδr,\displaystyle\frac{d\delta r}{dt}=\gamma_{L}\delta r, (90)

where we have defined the temporal-Lyapunov exponent

γLru2f(ru)22Veff′′(ru).\displaystyle\gamma_{L}\equiv\sqrt{-\frac{r_{u}^{2}f(r_{u})}{2\ell^{2}}V^{\prime\prime}_{\text{eff}}(r_{u})}. (91)

Thus, the time evolution of δr(t)\delta r(t) is

δr(t)=δr0eγLt,\displaystyle\delta r(t)=\delta r_{0}e^{\gamma_{L}t}, (92)

where δr0\delta r_{0} is the initial deviation of a geodesic from the photon sphere.

IV Probing Black Hole Phase Transitions through Optical Features

In this section, we will investigate BHs undergoing thermal phase transitions by exploring three critical parameters involving the optical features of BHs introduced in the previous section. As discussed in the Introduction, section I, it is worthwhile to investigate thermal phase transitions through optical probing in the extended phase space approach and holographic thermodynamics. As will be shown later in this section, the results from both approaches indicate that these parameters, treated as functions of either PP or TT, can express discontinuities as well as behaviors of multi-valued functions. This is in a similar way to exploring thermodynamic phase transitions of conventional substances by observing abrupt changes in their free energy and response functions as PP or TT are varied. For convenience, we use 𝒪i\mathcal{O}_{i} where i=1,2,3i=1,2,3 to represent the three critical parameters: 𝒪1\mathcal{O}_{1}, 𝒪2\mathcal{O}_{2}, and 𝒪3\mathcal{O}_{3} represent τ\tau, λL\lambda_{L}, and γL\gamma_{L}, respectively. These three critical parameters can potentially serve as order parameters in the consideration of phase transitions, as we will discuss later in Section V.

IV.1 Probing the phase transition in the extended phase space approach

The thermal profiles of 𝒪i\mathcal{O}_{i} in the extended phase space can be considered as a function of either PP or TT separately, i.e., 𝒪i(P)\mathcal{O}_{i}(P) or 𝒪i(T)\mathcal{O}_{i}(T). This is because rur_{u} in Eq. (81) can be written as ru=ru(r+,P)r_{u}=r_{u}(r_{+},P) and ru=ru(r+,T)r_{u}=r_{u}(r_{+},T) as follows:

ru(r+,P)\displaystyle r_{u}(r_{+},P) =\displaystyle= 14[3(83πPr+3+q2r++r+)+9(83πPr+3+q2r++r+)232q2],\displaystyle\frac{1}{4}\left[3\left(\frac{8}{3}\pi Pr_{+}^{3}+\frac{q^{2}}{r_{+}}+r_{+}\right)+\sqrt{9\left(\frac{8}{3}\pi Pr_{+}^{3}+\frac{q^{2}}{r_{+}}+r_{+}\right)^{2}-32q^{2}}\,\right], (93)
ru(r+,T)\displaystyle r_{u}(r_{+},T) =\displaystyle= q2r++12[r++2πr+2T+r+2(1+2πr+T)2+(8πr+T4)q2+4q4r+2].\displaystyle\frac{q^{2}}{r_{+}}+\frac{1}{2}\left[r_{+}+2\pi r_{+}^{2}T+\sqrt{r_{+}^{2}(1+2\pi r_{+}T)^{2}+(8\pi r_{+}T-4)q^{2}+\frac{4q^{4}}{r_{+}^{2}}}\,\right]. (94)
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 6: Isotherm curves of τ\tau, λL\lambda_{L} and γL\gamma_{L} as a function of PP are shown in the first, second and third rows, respectively. In the left, middle and right columns correspond to the temperature T<TcT<T_{c}, T=TcT=T_{c} and T>TcT>T_{c}, respectively.

In the context of the extended phase space, the present paper focuses on the effect of PP on three critical parameters during BH phase transitions in the isothermal process. Namely, we analyze the discontinuities and multi-value behavior of 𝒪i\mathcal{O}_{i} as functions of PP while keeping TT constant. By substituting ru(r+,T)r_{u}(r_{+},T) as described in Eq. (94) into Eqs. (84), (88) and (91), one can use r+r_{+} as a parameter running in the parametric plot as the isotherm curves, i.e. 𝒪i\mathcal{O}_{i} versus the bulk pressure PP. Note that the bulk pressure PP has been described in Eq. (28). The isotherm curves of τ,λL\tau,\lambda_{L} and γL\gamma_{L}, as a function of PP, are illustrated in the first, second and third rows of Fig. 6, respectively. The left, middle and right columns of Fig. 6 depict the cases with T<Tc,T=TcT<T_{c},T=T_{c} and T>TcT>T_{c}, respectively.

When T<TcT<T_{c}, the three critical parameters exhibit a multi-value function within the range PminPPmaxP_{\text{min}}\leq P\leq P_{\text{max}}, as shown in the left column of Fig. 6. In other words, at the values of PP in this range, there are three values for each 𝒪i\mathcal{O}_{i} corresponding with the Small (red), Intermediate (green) and Large-BHs (blue) branches. Outside this range of pressure, there exists only one branch, namely the Large-BH branch for P<PminP<P_{\text{min}} and the Small-BH branch for P>PmaxP>P_{\text{max}}. Interestingly, the isotherm curves of three critical parameters at T<TcT<T_{c} expressed itself as a multi-value function has the boundary at PminP_{\text{min}} and PmaxP_{\text{max}}, which correspond to the positions of the cusps of GG (see Fig. 2) with the divergent value of κT\kappa_{T} (see Fig. 1). Therefore, considering the parameters 𝒪i\mathcal{O}_{i} can exhibit the second-order phase transition of charged AdS-BH from observing the divergence of their derivative with respect to PP. Remarkably, when a Small-Large BHs first-order phase transition occurs at P=PfP=P_{f}, the values of 𝒪i\mathcal{O}_{i} will discontinuously change between Small and Large BH branches due to discontinuity jumping of event horizon radius caused by the Maxwell equal area law. This is an important result that will be applied in our study about the 𝒪iP\mathcal{O}_{i}-P criticality, as discussed in the next section.

As the temperature TT is larger, the pressures PminP_{\text{min}} and PmaxP_{\text{max}} converge and degenerate to be PcP_{c} at T=TcT=T_{c}, namely the Intermediate-BH phase becomes absent, as shown in the middle colums of Fig. 6. The values of three critical parameters at the critical point of the second-order phase transition between Small (red) and Large (blue) BHs can be expressed as follows

τc=3π126+2923,λc=2π2+6,γc=23655332.\displaystyle\tau_{c}=3\pi\sqrt{\frac{12\sqrt{6}+29}{23}},\ \ \ \lambda_{c}=\frac{2\pi\ell}{\sqrt{2+\sqrt{6}}},\ \ \ \gamma_{c}=\frac{2}{3}\sqrt{65-53\sqrt{\frac{3}{2}}}. (95)

Note that we use q=1q=1. For TT is larger than TcT_{c}, it turns out that only one BH solution exists and no phase transition occurs.

IV.2 Probing the phase transition in holographic thermodynamics

In this subsection, we probe phase structures of charged AdS-BH in the holographic thermodynamics approach via three critical parameters, i.e. τ\tau, λ\lambda and γ\gamma. The expression for the photon sphere radius rur_{u} in Eq. (81) can be written in term of the bulk parameters xx and yy, as defined in Eq. (53), in the form

ru=L4[3x(1+x2+y2x2)+9x2(1+x2+y2x2)232y2].\displaystyle r_{u}=\frac{L}{4}\left[3x\left(1+x^{2}+\frac{y^{2}}{x^{2}}\right)+\sqrt{9x^{2}\left(1+x^{2}+\frac{y^{2}}{x^{2}}\right)^{2}-32y^{2}}\right]. (96)

To investigate the influence of Q~\tilde{Q} and 𝒞\mathcal{C} of the boundary CFT on the behaviors of null geodesics around critical photon orbits in bulk spacetime, yy can be substituted using Eq. (57) into the above equation in terms of Q~\tilde{Q} and 𝒞\mathcal{C} as following

ru=L4[3x(1+x2+116𝒞2Q~2x2)+9x2(1+x2+116𝒞2Q~2x2)22Q~2𝒞2].\displaystyle r_{u}=\frac{L}{4}\left[3x\left(1+x^{2}+\frac{1}{16\mathcal{C}^{2}}\frac{\tilde{Q}^{2}}{x^{2}}\right)+\sqrt{9x^{2}\left(1+x^{2}+\frac{1}{16\mathcal{C}^{2}}\frac{\tilde{Q}^{2}}{x^{2}}\right)^{2}-\frac{2\tilde{Q}^{2}}{\mathcal{C}^{2}}}\right]. (97)

Since the p𝒱p-\mathcal{V} criticality is absent in the boundary thermodynamics and no p𝒱p\mathcal{V} term appears in the Smarr formula, we focus on the behavior of 𝒪i\mathcal{O}_{i} as a function of TT instead pp for the holographic thermodynamics. Substituting rur_{u} into Eqs. (84), (88) and (91), one can express τ\tau, λL\lambda_{L} and γL\gamma_{L} versus TT for study the influence of Q~\tilde{Q} and 𝒞\mathcal{C} on 𝒪i(T)\mathcal{O}_{i}(T) in the case I and case II as discussed before, respectively.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 7: Isocharge curves of τ\tau, λL\lambda_{L} and γL\gamma_{L} as a function of TT are shown in the first, second and third rows, respectively. In the left, middle and right columns correspond to the charge Q~<Q~c\tilde{Q}<\tilde{Q}_{c}, Q~=Q~c\tilde{Q}=\tilde{Q}_{c} and Q~>Q~c\tilde{Q}>\tilde{Q}_{c}, respectively.

In the case I, the behaviors of 𝒪i\mathcal{O}_{i} versus TT for Q~<Q~c\tilde{Q}<\tilde{Q}_{c}, Q~=Q~c\tilde{Q}=\tilde{Q}_{c} and Q~>Q~c\tilde{Q}>\tilde{Q}_{c} are illustrated in the left, middle and right columns in Fig. 7, respectively. From the graph of F(T)F(T) with Q~<Q~c\tilde{Q}<\tilde{Q}_{c} for dual CFT (see the figure (a) of Fig 5) reveals that three phases of CFT, namely pCFT1, nCFT and pCFT2 coexist when Tmin<T<TmaxT_{\text{min}}<T<T_{\text{max}} implying that the Small, Intermediate and Large-BH phases also coexist in this range of TT in the gravity picture. We find that the parameters 𝒪i\mathcal{O}_{i} associated with the critical curve in Tmin<T<TmaxT_{\text{min}}<T<T_{\text{max}} exhibits a multi-value function, i.e. three phases of BH have three different values of 𝒪i\mathcal{O}_{i}, as shown in the left column in Fig. 7. Outside this range of TT, there exists only one phase, namely the Small-BH phase for T<TminT<T_{\text{min}} and the Large-BH phase for T>TmaxT>T_{\text{max}}. Moreover, the slope of isocharge curves in the 𝒪iT\mathcal{O}_{i}-T plane diverges at TminT_{\text{min}} and TmaxT_{\text{max}}, which correspond to the positions of cusp of FF with the divergent of CQ~,𝒱,𝒞C_{\tilde{Q},\mathcal{V},\mathcal{C}}. Remarkably, when pCFT1-pCFT2 first-order phase transition occur at T=TfT=T_{f}, the values of 𝒪i\mathcal{O}_{i} will discontinuously change between Small-BH and Large-BH phases due to discontinuos jumping of entropy. This can be exhibited by the Maxwell equal area law as illustrated in Appendix B. This remarkable results from the present study could reveal the behaviors of 𝒪iT\mathcal{O}_{i}-T criticality, as will be discussed later in the next section.

The middle column in Fig. 7 illustrates the case of Q~=Q~c\tilde{Q}=\tilde{Q}_{c}. Notably, 𝒪i\mathcal{O}_{i} becomes a single-valued function, and Intermediate-BH phase disappears. Two distinct configuration of BHs, i.e. Small and Large-BHs, exist without coexistence. Furthermore, Small-Large BHs second-order phase transition takes place at TcT_{c}, where the critical values of τ\tau, λL\lambda_{L} and γL\gamma_{L} can be expressed as follows:

τc=πL29+12692,λc=2π2+6,γc=4L655332.\displaystyle\tau_{c}=\pi L\sqrt{\frac{29+12\sqrt{6}}{92}},\ \ \ \lambda_{c}=\frac{2\pi\ell}{\sqrt{2+\sqrt{6}}},\ \ \ \gamma_{c}=\frac{4}{L}\sqrt{65-53\sqrt{\frac{3}{2}}}. (98)

In the right column of Fig. 7, we examine the scenario where Q~>Q~c\tilde{Q}>\tilde{Q}_{c}. In this case, the BH is characterized by a single phase (purple curve).

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 8: The curves with constant central charge for τ\tau, λL\lambda_{L} and γL\gamma_{L} as a function of TT are shown in the first, second and third rows, respectively. In the left, middle and right columns correspond to the central charge 𝒞<𝒞c\mathcal{C}<\mathcal{C}_{c}, 𝒞=𝒞c\mathcal{C}=\mathcal{C}_{c} and 𝒞>𝒞c\mathcal{C}>\mathcal{C}_{c}, respectively.

By introducing 𝒞\mathcal{C} as a new thermodynamic variable, we investigate its influence on 𝒪i(T)\mathcal{O}_{i}(T) in the case II as shown in Fig. 8. We examine the curves of 𝒪i\mathcal{O}_{i} versus TT in 𝒞<𝒞c\mathcal{C}<\mathcal{C}_{c} scenario in the left column in Fig. 8. Since the BH and its dual CFT on the boundary exist in a single phase without any phase transition, the 𝒪i(T)\mathcal{O}_{i}(T) curves exhibit no multi-value function and discontinuity.

In the case of 𝒞=𝒞c\mathcal{C}=\mathcal{C}_{c}, the curves 𝒪i(T)\mathcal{O}_{i}(T) display a single-value function, whose derivative with respect to TT diverges at TcT_{c}. The BH configurations below and above the critical temperature TcT_{c} correspond to the Small and Large BH phases, respectively. Note that the values of τ\tau, λL\lambda_{L} and γL\gamma_{L} at the critical point of case II is the same as those of case I, as shown in Eq.(98).

For 𝒞>𝒞c\mathcal{C}>\mathcal{C}_{c}, we find that the parameters 𝒪i\mathcal{O}_{i} in Tmin<T<TmaxT_{\text{min}}<T<T_{\text{max}} range exhibits a multi-value function. Namely, three different values of 𝒪i\mathcal{O}_{i} correspond to three phases of BH, as shown in the right column in Fig. 8. Outside this range of TT, there exists only one phase, namely the Small-BH phase for T<TminT<T_{\text{min}} and the Large-BH phase for T>TmaxT>T_{\text{max}}. Moreover, the slope of constant-𝒞\mathcal{C} curves in the 𝒪iT\mathcal{O}_{i}-T plane diverges at TminT_{\text{min}} and TmaxT_{\text{max}}, which correspond to the positions of cusp of FF with the divergent value of CQ~,𝒱,𝒞C_{\tilde{Q},\mathcal{V},\mathcal{C}}.

V Scaling Behavior of the optical parameters

In this section, we investigate the scaling behavior of the BH’s optical appearance parameters OiO_{i}. In phase transition studies, scaling behavior typically emerges near the critical transition point, where different types of matter can exhibit similar scaling laws for thermodynamic quantities. The scaling behavior with identical critical exponents across different types of matter suggests that these matters belong to the same universality class, independent of their particle composition and interactions. Notably, the critical exponents in the scaling law are crucial for understanding the nature of matter near the critical point of phase transition.

To investigate the system near the critical point, we focus on order parameter that vanishes at this point, while maintaining non-zero values prior to reaching it. In examining the dependence of τ\tau, λL\lambda_{L}, and γL\gamma_{L} on PP and TT within the extended phase space approach and holographic thermodynamics, there is a potential that Δ𝒪i=|𝒪iL𝒪iS|\Delta\mathcal{O}_{i}=\left|\mathcal{O}_{iL}-\mathcal{O}_{iS}\right| can serve as an order parameter, where subscripts SS and LL denote Small and Large BHs, respectively. At PfP_{f} and TfT_{f}, where the Small-Large BH phase transition occurs, 𝒪i\mathcal{O}_{i} experiences a discontinuous jump, indicating Δ𝒪i0\Delta\mathcal{O}_{i}\neq 0. As the value of PfP_{f} or TfT_{f} changes along the phase boundary, the values of 𝒪i\mathcal{O}_{i} for Small and Large BHs become increasingly similar, causing Δ𝒪i\Delta\mathcal{O}_{i} to decrease, eventually reaching zero at the critical point.

Here, we will categorize the phase transition of the charged AdS-BH in the extended phase space and holographic thermodynamics description via investigating the scaling law of Δ𝒪i\Delta\mathcal{O}_{i}. In particular, our analysis in this section will show that the scaling law can be written in the following form

𝒪i𝒪icai(1p)αi\displaystyle\frac{\mathcal{O}_{i}}{\mathcal{O}_{ic}}\sim a_{i}(1-p^{*})^{\alpha_{i}} (99)

in the extended phase space approach, whereas in the form

Δ𝒪i𝒪icbi(t1)βi\displaystyle\frac{\Delta\mathcal{O}_{i}}{\mathcal{O}_{ic}}\sim b_{i}(t^{*}-1)^{\beta_{i}} (100)

in holographic thermodynamics approach. Note that 𝒪ic\mathcal{O}_{ic} denotes the value of 𝒪i\mathcal{O}_{i} at the critical point. Moreover, aia_{i} and bib_{i} are the proportionality constants, αi\alpha_{i} and βi\beta_{i} are the critical exponents. Recall that p=Pf/Pcp^{*}=P_{f}/P_{c} and t=Tf/Tct^{*}=T_{f}/T_{c}, as previously defined in section II.

V.1 Extended phase space approach

Let us consider the system near the critical point, one can apply the power series expansion to express 𝒪iS(rS)\mathcal{O}_{iS}(r_{S}) and 𝒪iL(rL)\mathcal{O}_{iL}(r_{L}) as follows

𝒪iS(rS)=𝒪ic+(𝒪ir+)c(rSrc)+,\displaystyle\mathcal{O}_{iS}(r_{S})=\mathcal{O}_{ic}+\left(\frac{\partial\mathcal{O}_{i}}{\partial r_{+}}\right)_{c}(r_{S}-r_{c})+\dots, (101)
𝒪iL(rL)=𝒪ic+(𝒪ir+)c(rLrc)+,\displaystyle\mathcal{O}_{iL}(r_{L})=\mathcal{O}_{ic}+\left(\frac{\partial\mathcal{O}_{i}}{\partial r_{+}}\right)_{c}(r_{L}-r_{c})+\dots, (102)

where rSr_{S} and rLr_{L} are the horizon radii of Small and Large BHs at Hawking-Page pressure. The order parameter is then

Δ𝒪i𝒪ic=|𝒪iL𝒪iS|𝒪ic=1𝒪ic(𝒪ir+)c(rLrS).\displaystyle\frac{\Delta\mathcal{O}_{i}}{\mathcal{O}_{ic}}=\frac{\left|\mathcal{O}_{iL}-\mathcal{O}_{iS}\right|}{\mathcal{O}_{ic}}=\frac{1}{\mathcal{O}_{ic}}\left(\frac{\partial\mathcal{O}_{i}}{\partial r_{+}}\right)_{c}\left(r_{L}-r_{S}\right). (103)

From Eq. (35) and (36), the leading order of rSr_{S} and rLr_{L} around the critical point is given by

rS\displaystyle r_{S} =\displaystyle= 623t¯+,\displaystyle\sqrt{6}-2\sqrt{3}\sqrt{-\bar{t}}+\cdots,
rL\displaystyle r_{L} =\displaystyle= 6+23t¯+,\displaystyle\sqrt{6}+2\sqrt{3}\sqrt{-\bar{t}}+\cdots,

where

t¯=TfTcTc.\displaystyle\bar{t}=\frac{T_{f}-T_{c}}{T_{c}}. (104)

Thus

rLrS\displaystyle r_{L}-r_{S} =\displaystyle= 43t¯.\displaystyle 4\sqrt{3}\sqrt{-\bar{t}}. (105)

Expanding the reduced Hawking-Page pressure pp^{*} as expressed in Eq. (34) near the critical temperature t=1t=1, we obtain

p\displaystyle p^{*} =\displaystyle= 1+83t¯+.\displaystyle 1+\frac{8}{3}\bar{t}+\cdots. (106)

It is important to emphasize that we are considering the system near the critical point, allowing us to express the above equation up to the first order of t¯\bar{t}, thereby neglecting higher-order terms. Consequently, we can use this first-order approximation to eliminate t¯\bar{t} in Eq. (105). By substituting the result into Eq. (103), we obtain

Δ𝒪i𝒪ic\displaystyle\frac{\Delta\mathcal{O}_{i}}{\mathcal{O}_{ic}} =\displaystyle= ai(1p)1/2,\displaystyle a_{i}\left(1-p^{*}\right)^{1/2}, (107)

where aia_{i} could be expressed as

ai=32𝒪ic(𝒪ir+)c.a_{i}=\frac{3\sqrt{2}}{\mathcal{O}_{ic}}\left(\frac{\partial\mathcal{O}_{i}}{\partial r_{+}}\right)_{c}. (108)

From Eq. (107), we observe that the critical exponent αi=1/2\alpha_{i}=1/2, which suggests that its behaviors are in the same way as the vdW fluid type. The scaling behavior for three optical order parameters are

Δττc\displaystyle\frac{\Delta\tau}{\tau_{c}} \displaystyle\sim 0.7337(1p)1/2,\displaystyle 0.7337\left(1-p^{*}\right)^{1/2}, (109)
ΔλLλc\displaystyle\frac{\Delta\lambda_{L}}{\lambda_{c}} \displaystyle\sim 0.2384(1p)1/2,\displaystyle 0.2384\left(1-p^{*}\right)^{1/2}, (110)
ΔγLγc\displaystyle\frac{\Delta\gamma_{L}}{\gamma_{c}} \displaystyle\sim 0.4953(1p)1/2.\displaystyle 0.4953\left(1-p^{*}\right)^{1/2}. (111)

Using Eqs. (34), (35), (36) and (94), we can determine each pair of the reduced order parameters Δ𝒪i𝒪ic\displaystyle\frac{\Delta\mathcal{O}_{i}}{\mathcal{O}_{ic}} and the reduced Hawking-Page pressure pp^{*} by running the temperature increasing from the minimum value T0T_{0} to TcT_{c}. Recall that we have shown the formula of T0T_{0} in Eq. (31). In other words, by writing in the form of reduced temperature t=T/Tct=T/T_{c}, we run tt from t=t0=T0/Tc=1/2t=t_{0}=T_{0}/T_{c}=1/\sqrt{2} to the critical value t=1t=1, and then we obtain the plot of the reduced order parameters versus pp^{*}, as shown in Fig 9. Notably, we find that three reduced order parameters decrease as pp^{*} increase and become vanish at p=1p^{*}=1. In the zoomed panels, the blue points represent the numerical values of the exact results of the reduced order parameter near the critical point.

Interestingly, the expressions of the scaling laws, as shown in Eqs. (109), (110) and (111), provide the results that fit the blue points remarkably well, as illustrated as the red curves. These indicate that the three order parameters exhibit a discontinuous jump, resulting in non-vanishing Δ𝒪i\Delta\mathcal{O}_{i} within the range p<1p^{*}<1. This behavior signifies the occurrence of a second-order phase transition at p=1p^{*}=1. At this critical point, all order parameters Δ𝒪i\Delta\mathcal{O}_{i} vanish as 𝒪iS\mathcal{O}_{iS} and 𝒪iL\mathcal{O}_{iL} converge to the same value. Beyond the critical point, the order parameters no longer exhibit multiple values, indicating the presence of a single BH phase.

Refer to caption
Refer to caption
Refer to caption
Figure 9: Plots of the reduced order parameters Δ𝒪i/𝒪ic\Delta\mathcal{O}_{i}/\mathcal{O}_{ic} versus pp^{*} at the Small-Large BH first-order phase transition in the extended phase space approach are shown as the blue curve, while they are shown as blue points in the zoomed panel with p1p^{*}\sim 1 in order to compare with the scaling laws obtained from our analysis (red).

V.2 Holographic thermodynamics approach

In the case of the holographic thermodynamics approach, we consider the power series expansion of 𝒪i(x)\mathcal{O}_{i}(x) around the critical point up to the first order as

𝒪iS(x)=𝒪ic+(𝒪ix)c(xSxc)+,\displaystyle\mathcal{O}_{iS}(x)=\mathcal{O}_{ic}+\left(\frac{\partial\mathcal{O}_{i}}{\partial x}\right)_{c}(x_{S}-x_{c})+\dots, (112)
𝒪iL(x)=𝒪ic+(𝒪ix)c(xLxc)+,\displaystyle\mathcal{O}_{iL}(x)=\mathcal{O}_{ic}+\left(\frac{\partial\mathcal{O}_{i}}{\partial x}\right)_{c}(x_{L}-x_{c})+\dots, (113)

where xSx_{S} and xLx_{L} refer to the horizon radii for Small and Large BHs at the first-order phase transition, respectively. The reduced order parameters near the critical point are then

Δ𝒪i𝒪ic=|𝒪iL𝒪iS|𝒪ic=1𝒪ic(𝒪ix)c(xLxS).\displaystyle\frac{\Delta\mathcal{O}_{i}}{\mathcal{O}_{ic}}=\frac{\left|\mathcal{O}_{iL}-\mathcal{O}_{iS}\right|}{\mathcal{O}_{ic}}=\frac{1}{\mathcal{O}_{ic}}\left(\frac{\partial\mathcal{O}_{i}}{\partial x}\right)_{c}\left(x_{L}-x_{S}\right). (114)

Let us consider the phase transition in the case I first. As 𝒞\mathcal{C} is fixed while Q~\tilde{Q} can be varied, we define the parameter q¯\bar{q} as

q¯=Q~Q~cQ~c.\displaystyle\bar{q}=\frac{\tilde{Q}-\tilde{Q}_{c}}{\tilde{Q}_{c}}. (115)

The expression of xSx_{S} and xLx_{L} in Eqs. (68) and (69) can be expanded around xcx_{c} as follows

xS\displaystyle x_{S} =\displaystyle= xc12q¯,\displaystyle x_{c}-\frac{1}{2}\sqrt{-\bar{q}}-\dots,
xL\displaystyle x_{L} =\displaystyle= xc+12q¯.\displaystyle x_{c}+\frac{1}{2}\sqrt{-\bar{q}}-\dots.

By keeping up to only the leading order, the difference between these two horizon radii is given by

xLxS=q¯.\displaystyle x_{L}-x_{S}=\sqrt{-\bar{q}}. (116)

Since we consider the behavior of reduced order parameters whose values are calculated at the Hawking-Page temperature TfT_{f} near the critical temperature TcT_{c}, let us write tt^{*}, Eq. (66), in the form of the series expansion as

t\displaystyle t^{*} =\displaystyle= 114q¯+.\displaystyle 1-\frac{1}{4}\bar{q}+\dots.

Keeping up to only the leading order, we can have

q¯\displaystyle-\bar{q} =\displaystyle= 4(t1).\displaystyle 4(t^{*}-1). (117)

Consequently, using Eqs. (116) and (117), Eqs. (114) can be rewritten in the form of the scaling law of reduced order parameters as

Δ𝒪i𝒪ic=bi(t1)1/2,\displaystyle\frac{\Delta\mathcal{O}_{i}}{\mathcal{O}_{ic}}=b_{i}(t^{*}-1)^{1/2}, (118)

where the proportionality coefficient is

bi=2𝒪ic(𝒪ix)c.\displaystyle b_{i}=\frac{2}{\mathcal{O}_{ic}}\left(\frac{\partial\mathcal{O}_{i}}{\partial x}\right)_{c}. (119)

As obviously shown in Eq. (118), the critical exponents of three order parameters βi=1/2\beta_{i}=1/2 for i=1,2i=1,2 and 33.

Refer to caption
Refer to caption
Refer to caption
Figure 10: Plots of the reduced order parameters Δ𝒪i/𝒪ic\Delta\mathcal{O}_{i}/\mathcal{O}_{ic} versus tt^{*} at the Small-Large BH first-order phase transition in the holographic thermodynamics approach are shown in blue curve, while they are shown as blue points in the zoomed panel with t1t^{*}\sim 1. Note that the red curve follows the scaling laws as obtained from our analysis.

For the case II of phase transition in the holographic thermodynamics approach, we introduce

c¯=𝒞𝒞c𝒞c.\displaystyle\bar{c}=\frac{\mathcal{C}-\mathcal{C}_{c}}{\mathcal{C}_{c}}. (120)

According to the previous approximation, we find that the scaling law for three reduced order parameters near the critical point of case II is the same as case I as follows

Δττc\displaystyle\frac{\Delta\tau}{\tau_{c}} \displaystyle\sim 2.0751(t1)1/2,\displaystyle 2.0751\left(t^{*}-1\right)^{1/2}, (121)
ΔλLλc\displaystyle\frac{\Delta\lambda_{L}}{\lambda_{c}} \displaystyle\sim 0.6742(t1)1/2,\displaystyle 0.6742\left(t^{*}-1\right)^{1/2}, (122)
ΔγLγc\displaystyle\frac{\Delta\gamma_{L}}{\gamma_{c}} \displaystyle\sim 1.4008(t1)1/2.\displaystyle 1.4008\left(t^{*}-1\right)^{1/2}. (123)

Since the scaling behavior near the critical point of reduced order parameters in case II is similar to case I, we will display the plot of three reduced order parameters Δ𝒪i𝒪ic\displaystyle\frac{\Delta\mathcal{O}_{i}}{\mathcal{O}_{ic}} versus the reduced Hawking-Page temperature tt^{*} only in case I, as illustrated in Fig 10. Using Eqs. (66), (68), (69) and (97), Δ𝒪i𝒪ic\displaystyle\frac{\Delta\mathcal{O}_{i}}{\mathcal{O}_{ic}} can be determined at fixed cc versus tt^{*}, with qq running from 1 to 0 in case I, whereas it is determined at fixed qq versus tt^{*} with c1c^{-1} running from 1 to 0 in case II. On the other hand, the figure indicates that three reduced order parameters are more than zero starting at t=1t^{*}=1 and terminating at t1.22t^{*}\approx 1.22, beyond which a Hawking-Page phase transition does not exist. This can also be seen from Fig. 11 (b) in Appendix B. In the zoomed panels, the blue points represent the numerical values of the exact results of the reduced order parameter near the critical point. Interestingly, the expressions of the scaling laws, as shown in Eqs. (121), (122) and (123), provide the results that fit the blue points very well, as illustrated as the red curves. These indicate that the three order parameters exhibit a discontinuous jump, resulting in non-vanishing Δ𝒪i\Delta\mathcal{O}_{i} within the range 1<t<1.221<t^{*}<1.22. This behavior signifies the occurrence of a second-order phase transition at t=1t^{*}=1. At this critical point, all order parameters Δ𝒪i\Delta\mathcal{O}_{i} vanish as 𝒪iS\mathcal{O}_{iS} and 𝒪iL\mathcal{O}_{iL} converge to the same value. For T<TcT<T_{c}, the order parameters no longer exhibit multiple values, indicating the presence of a single BH phase.

It is worth emphasizing that in the extended phase space approach, the scaling law can be found near the critical point for these three optical order parameters as a function of pp^{*}, where all exhibit the critical exponent with the value of 1/21/2. On the other hand, in holographic thermodynamics where the concepts of bulk pressure and volume are absent, the scaling law behavior can also be found near the critical point for these parameters as a function of tt^{*} with the critical exponent equal to 1/21/2. The results in the holographic thermodynamics context are particularly noteworthy because they do not rely on bulk pressure and volume. This suggests that the critical behavior of BHs can be understood only in terms of boundary field theory quantities.

VI Conclusions

In this study, we have thoroughly examined the intricate phase structures of charged AdS-BHs within the extended phase space and holographic thermodynamics approaches by analyzing null geodesics near their critical curves. By utilizing horizon-scale observation of BHs, namely the orbital half-period τ\tau, angular Lyapunov exponent λL\lambda_{L}, and temporal Lyapunov exponent γL\gamma_{L}, we can characterize BH phase transitions in these two frameworks.

In particular, for the extended phase space approach, when T<TcT<T_{c}, these three critical parameters are multi-valued functions of bulk pressure PP, indicating the presence of three coexisting phases of BH. When T>TcT>T_{c}, these parameters display the singled-valued function versus PP corresponding with a single BH phase.

In the holographic thermodynamics approach, the phase transition of charged AdS-BH in the bulk is holographically dual to the CFT phase transition in the boundary with including the central charge 𝒞\mathcal{C} and its conjugate as a new pair of thermodynamic variable. We investigate the phase structures of BH dual to the fixed (Q~,𝒱,𝒞)(\tilde{Q},\mathcal{V},\mathcal{C}) ensemble of CFT. For Q~<Q~c\tilde{Q}<\tilde{Q}_{c} or 𝒞>𝒞c\mathcal{C}>\mathcal{C}_{c}, the three critical parameters exhibit a multi-valued function of temperature TT, indicating the presence of three coexisting BH phases. Conversely, for Q~>Q~c\tilde{Q}>\tilde{Q}_{c} or 𝒞<𝒞c\mathcal{C}<\mathcal{C}_{c}, these parameters display a single-valued function with respect to TT, corresponding to a single BH phase.

In many fields, including condensed matter physics, there are numerous examples of phase transitions that often lack a clearly defined order parameter or the concept may not be entirely appropriate. In this study, we propose the possibility of an order parameter based on the difference between each optical parameter of the Small and that of Large BH phases at the Hawking-Page phase transition. This approach is advantageous because upcoming space-based VLBI missions aim at studying the photon subring structure of BH, thereby providing a practical and observable method to characterize BH phase transitions. Remarkably, the critical exponents for these parameters near the critical point, from both extended phase space and holographic thermodynamics approaches, are found to be 1/2, indicating a significant thermodynamic similarity to van der Waals fluids. It is important to emphasize that, in our work, the scaling law near the critical point can be derived through precise theoretical calculations, avoiding any ambiguous methods.

Evidently, the BH optical appearance parameters from both approached are effective in characterizing BH phase transitions in AdS space. It is important to highlight that these critical parameters may be linked to some phenomena in strongly coupled systems through holographic duality. In the context of the AdS/CFT correspondence, a thermal state in the CFT is dual to a BH in the bulk AdS space. A perturbation in the dual field theory, induced by a local operator, corresponds to a perturbation of the BH by turning on the scalar fields propagating in the bulk. The process of returning to thermal equilibrium in the field theory side can be dual to the black hole returning to equilibrium after a perturbation, which can be observed through imaginary part of quasinormal frequencies describing the rate at which a perturbed BH returns to equilibrium. Consequently, the QNM spectrum of a BH can be interpreted holographically as the quantum Ruelle resonances in boundary theory, which are complex frequency modes that describe the late-time decay of perturbations in chaotic systems [84]. Moreover, in the eikonal limit, the decay of QNMs can be governed by the temporal Lyapunov exponent γL\gamma_{L} of unstable null geodesics [81, 85, 86, 87]. In this eikonal regime in the AdS-BHs background, there exist the long-lived modes, which dominate the late-time behavior of BH while returning to equilibrium [88, 89]. The existence of long-lived QNMs in AdS suggests that certain perturbations in the dual CFT decay very slowly, implying that the system takes a long time to return to thermal equilibrium. In our results, we find that at the Small-Large BHs phase transitions, the behavior of γL\gamma_{L} and hence the late-time behavior of long-lived modes change discontinuously due to the Maxwell equal area law. Holographically, this behavior presumably reflects a sudden change in the spectrum of Ruelle resonances across the first-order phase transition on the field theory side. The order parameter we propose could be valuable for extending studies on holographic thermalization in a field theory, particularly concerning phase transitions, in future research.

In particular, we focus on optically probing the phase structures of charged BH in AdS space, as they exhibit rich phase structures and undergo intriguing phase transitions, such as the Hawking-Page phase transition and critical behavior. However, it is also worthwhile to consider applying these methods to study BH phase transitions in more realistic scenarios, such as BHs in asymptotically flat or de Sitter (dS) spacetimes. Additionally, this approach could be extended to investigate BH phase transitions in the context of modified entropies [90, *Promsiri:2020jga, *ElMoumni:2022chi, *Tannukij:2020njz, *Nakarachinda:2021jxd, *Barzi:2024bbj, *MEJRHIT201945, *barrow2020area, *Jawad:2022bmi], providing a broader framework for understanding BH thermodynamics across different theoretical models.

Acknowledgement

EH is supported by Thailand Science Research and Innovation (TSRI) Basic Research Fund: Fiscal year 2023 under project number FRB660073/0164. WH is supported by the National Science and Technology Development Agency under the Junior Science Talent Project (JSTP) scholarship, grant number SCA-CO-2565-16992-TH. This research has received funding support from the NSRF via the Program Management Unit for Human Resources & Institutional Development, Research and Innovation [grant number B13F660066].

Appendix A Eulerian scaling argument

In the following discussions, we review how the first law of thermodynamics is obtained from the Smarr formula by using Euler’s theorem for a homogeneous function as detailed in [15, 16].

A function f(x1,,xn)f(x_{1},\dots,x_{n}) is a homogeneous function of degree qq if

f(αp1x1,αp2x2,,αpnxn)=αqf(x1,x2,,xn).\displaystyle f(\alpha^{p_{1}}x_{1},\alpha^{p_{2}}x_{2},\dots,\alpha^{p_{n}}x_{n})=\alpha^{q}f(x_{1},x_{2},\dots,x_{n}). (124)

The Euler’s theorem states that if f(x1,,xn)f(x_{1},\dots,x_{n}) is a homogeneous function of degree qq, then it satisfy the partial differential equation:

qf(x1,,xn)=p1(fx1)x1+p2(fx2)x2++pn(fxn)xn,\displaystyle qf(x_{1},\dots,x_{n})=p_{1}\left(\frac{\partial f}{\partial x_{1}}\right)x_{1}+p_{2}\left(\frac{\partial f}{\partial x_{2}}\right)x_{2}+\cdots+p_{n}\left(\frac{\partial f}{\partial x_{n}}\right)x_{n}, (125)

then an infinitesimal change of ff is given by

df=(fx1)dx1+(fx2)dx2++(fxn)dxn.\displaystyle df=\left(\frac{\partial f}{\partial x_{1}}\right)dx_{1}+\left(\frac{\partial f}{\partial x_{2}}\right)dx_{2}+\dots+\left(\frac{\partial f}{\partial x_{n}}\right)dx_{n}. (126)

It is noticed that the mass of non-rotating charged AdS-BH can be expressed in terms of the area AA, the electric charge QQ and the cosmological constant Λ\Lambda as follows

M(A,Q,Λ)=(n1)ωn11n116πGNAn2n118πGNωn11n1nΛAnn1+4πGN(n1)ωn11n1η2Q2An2n1,\displaystyle M(A,Q,\Lambda)=\frac{(n-1)\omega_{n-1}^{\frac{1}{n-1}}}{16\pi G_{N}}A^{\frac{n-2}{n-1}}-\frac{1}{8\pi G_{N}\omega_{n-1}^{\frac{1}{n-1}}n}\Lambda A^{\frac{n}{n-1}}+\frac{4\pi G_{N}}{(n-1)\omega_{n-1}^{\frac{1}{n-1}}\eta^{2}}\frac{Q^{2}}{A^{\frac{n-2}{n-1}}}, (127)

which is indeed the homogeneous function degree n2n-2 because M(αn1A,αn2Q,α2Λ)=αn2M(A,Q,Λ)M(\alpha^{n-1}A,\alpha^{n-2}Q,\alpha^{-2}\Lambda)=\alpha^{n-2}M(A,Q,\Lambda). From Eq. (125), we have

(n2)M=(n1)MAA+(n2)MQQ+(2)MΛΛ.\displaystyle(n-2)M=(n-1)\frac{\partial M}{\partial A}A+(n-2)\frac{\partial M}{\partial Q}Q+(-2)\frac{\partial M}{\partial\Lambda}\Lambda. (128)

Using Eq. (127), the partial derivative terms in the above equation can be derived as

MAA=MSS=TS,MQQ=ΦQandMΛΛ=ΘΛ8πGN,\displaystyle\frac{\partial M}{\partial A}A=\frac{\partial M}{\partial S}S=TS,\ \ \ \frac{\partial M}{\partial Q}Q=\Phi Q\ \ \ \text{and}\ \ \ \frac{\partial M}{\partial\Lambda}\Lambda=\frac{\Theta\Lambda}{{8\pi G_{N}}}, (129)

and the Smarr formula is then

M=n1n2κA8πGN+ΦQ1n2ΘΛ4πGN.\displaystyle M=\frac{n-1}{n-2}\frac{\kappa A}{8\pi G_{N}}+\Phi Q-\frac{1}{n-2}\frac{\Theta\Lambda}{4\pi G_{N}}. (130)

The formula in Eq. (126) give the variation of MM as follow

dM=κ8πGNdA+ΦdQ+Θ8πGNdΛ.\displaystyle dM=\frac{\kappa}{8\pi G_{N}}dA+\Phi dQ+\frac{\Theta}{8\pi G_{N}}d\Lambda. (131)

Identifying respectively, Λ\Lambda and its conjugate variable Θ\Theta as the bulk pressure PP and volume VV via the relation in Eq. (3) in Eqs. (130) and (131), we will obtain Eqs. (25) and (26).

Appendix B Holographic Maxwell’s equal area law in TST-S plane

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 11: The curves of the reduced BH temperature tt versus the reduced entropy ss with the Maxwell construction for finding the first-order phase transition tt^{*} are plotted for case I with q=0.7q=0.7 (a) and case II with c=1.5c=1.5 (c). Resulting from the Maxwell construction, we obtain the values of tt^{*} at different values of qq presented as the line terminated at the critical point q=1q=1 (b), whereas tt^{*} versus 1/c1/c presented as the line terminated at the critical point 1/c=11/c=1 (d).

Here, we consider the calculations of pCFT1-pCFT2 phase transition temperature TfT_{f} for CFT, which dual to the Small-Large BHs phase transition of charged AdS-BH in the holographic description. Substituting the horizon radius xx into Eq. (60) and then writing xx in the form of the entropy SS using the area law, we obtain T=T(S,Q~,𝒞)T=T(S,\tilde{Q},\mathcal{C}) as follows

T=12𝒞πS(1+3S4π𝒞πQ~24𝒞S).\displaystyle T=\frac{1}{2}\sqrt{\frac{\mathcal{C}}{\pi S}}\left(1+\frac{3S}{4\pi\mathcal{C}}-\frac{\pi\tilde{Q}^{2}}{4\mathcal{C}S}\right). (132)

We can define the reduced parameters as follows:

t=TTc,s=SSc,q=Q~Q~candc=𝒞𝒞c.\displaystyle t=\frac{T}{Tc},\ \ \ s=\frac{S}{S_{c}},\ \ \ q=\frac{\tilde{Q}}{\tilde{Q}_{c}}\ \ \ \text{and}\ \ \ c=\frac{\mathcal{C}}{\mathcal{C}_{c}}. (133)

Consider case I as introduced in section II, we can express the reduced BH temperature in Eq. (132) in terms of ss and qq as

t=34(1s+s2q26s32).\displaystyle t=\frac{3}{4}\left(\frac{1}{\sqrt{s}}+\frac{\sqrt{s}}{2}-\frac{q^{2}}{6s^{\frac{3}{2}}}\right). (134)

Interestingly, it does not explicitly depend on central charge 𝒞\mathcal{C}. As shown in Fig. 11 (a), the reduced Hawking-Page phase transition temperature t=Tf/Tct^{*}=T_{f}/T_{c} is represented in dashed-black horizontal line. This line is positioned such that the areas above and below the isocharge curve are equal, in accordance with Maxwell’s equal area law. With s1s_{1} and s3s_{3} representing the reduced entropy associated with pCFT1 and pCFT2 at the tt^{*}, respectively, the equal area law is expressed as

t(s3s1)=s1s3t(s)𝑑s.\displaystyle t^{*}(s_{3}-s_{1})=\int_{s_{1}}^{s_{3}}t(s)ds. (135)

Then, we obtain

t=1s1+s3(32q24s1s3+14(s1+s3+s1s3)).\displaystyle t^{*}=\frac{1}{\sqrt{s_{1}}+\sqrt{s_{3}}}\left(\frac{3}{2}-\frac{q^{2}}{4\sqrt{s_{1}s_{3}}}+\frac{1}{4}(s_{1}+s_{3}+\sqrt{s_{1}s_{3}})\right). (136)

Note that the first-order phase transition appear when q<1q<1 (Q~<Q~c\tilde{Q}<\tilde{Q}_{c}). Substituting x=s1x=\sqrt{s_{1}} and y=s3y=\sqrt{s_{3}} into Eq. (134) together with Eq. (136), we obtain the system of equations as follows

t\displaystyle t^{*} =\displaystyle= 34(1x+x2q26x3),\displaystyle\frac{3}{4}\left(\frac{1}{x}+\frac{x}{2}-\frac{q^{2}}{6x^{3}}\right), (137)
t\displaystyle t^{*} =\displaystyle= 34(1y+y2q26y3),\displaystyle\frac{3}{4}\left(\frac{1}{y}+\frac{y}{2}-\frac{q^{2}}{6y^{3}}\right), (138)
t\displaystyle t^{*} =\displaystyle= 1x+y(32q24xy+14(x2+y2+xy)).\displaystyle\frac{1}{x+y}\left(\frac{3}{2}-\frac{q^{2}}{4xy}+\frac{1}{4}(x^{2}+y^{2}+xy)\right). (139)

These equations can be solved analytically by following the procedure as suggested in [77]. First, using the fact the RHS of Eq. (137) equal to Eq. (138), we obtain

x2+y2=6q2z2(1z2)z,\displaystyle x^{2}+y^{2}=\frac{6}{q^{2}}z^{2}\left(1-\frac{z}{2}\right)-z, (140)

where z=xyz=xy. Then, using 2(139)=(137)+(138)2\eqref{t3}=\eqref{t1}+\eqref{t2}, we have

3zq22+z2(x2+y2+z)=34(x2+y2+2z)(1+z2q26z2(x2+y2z)).\displaystyle 3z-\frac{q^{2}}{2}+\frac{z}{2}(x^{2}+y^{2}+z)=\frac{3}{4}(x^{2}+y^{2}+2z)\left(1+\frac{z}{2}-\frac{q^{2}}{6z^{2}}(x^{2}+y^{2}-z)\right). (141)

By eliminating term x2+y2x^{2}+y^{2} via the relation in Eq. (140), we have the quartic equation in variable zz as follows

z42z3+2q2zq4=0.\displaystyle z^{4}-2z^{3}+2q^{2}z-q^{4}=0. (142)

The nontrivial solution is z=xy=qz=xy=q. To obtain xx and yy, we substitute z=qz=q into Eq. (140) and solve following equations

x2+y2\displaystyle x^{2}+y^{2} =\displaystyle= 64q,\displaystyle 6-4q, (143)
xy\displaystyle xy =\displaystyle= q.\displaystyle q. (144)

simultaneously. Since x<yx<y, therefore the solutions for xx and yy are

x\displaystyle x =\displaystyle= q3+3(q3)(q1)2q,\displaystyle\frac{q}{\sqrt{3+\sqrt{3(q-3)(q-1)}-2q}}, (145)
y\displaystyle y =\displaystyle= 3+3(q3)(q1)2q.\displaystyle\sqrt{3+\sqrt{3(q-3)(q-1)}-2q}. (146)

Thus, tt^{*} becomes

t=3+3(q3)(q1)q23+3(q3)(q1)2q.\displaystyle t^{*}=\frac{3+\sqrt{3(q-3)(q-1)}-q}{2\sqrt{3+\sqrt{3(q-3)(q-1)}-2q}}. (147)

For case II, we express the BH temperature in Eq. (132) by substituting S=sScS=sS_{c} and 𝒞=c𝒞c\mathcal{C}=c\mathcal{C}_{c}. The resulting tt does not depend on electric charge Q~\tilde{Q} as following

t=34(1s+s216c2s32).\displaystyle t=\frac{3}{4}\left(\frac{1}{\sqrt{s}}+\frac{\sqrt{s}}{2}-\frac{1}{6c^{2}s^{\frac{3}{2}}}\right). (148)

Note that the first-order phase transition appears when c>1c>1 (𝒞>𝒞c\mathcal{C}>\mathcal{C}_{c}). Since Eqs. (134) and (148) is symmetric under q21/c2q^{2}\rightarrow 1/c^{2}, the solutions of variables x,yx,y and tt^{*} as a function of cc^{*} in this case can be written as follows

x\displaystyle x =\displaystyle= 13c2+c3(13c)(1c)2c,\displaystyle\frac{1}{\sqrt{3c^{2}+c\sqrt{3(1-3c)(1-c)}-2c}}, (149)
y\displaystyle y =\displaystyle= 1c3c2+c3(13c)(1c)2c,\displaystyle\frac{1}{c}\sqrt{3c^{2}+c\sqrt{3(1-3c)(1-c)}-2c}, (150)
t\displaystyle t^{*} =\displaystyle= 3c+3(13c)(1c)123c2+c3(13c)(1c)2c.\displaystyle\frac{3c+\sqrt{3(1-3c)(1-c)}-1}{2\sqrt{3c^{2}+c\sqrt{3(1-3c)(1-c)}-2c}}. (151)

References