Thermodynamic Geometry of Bardeen Black Holes

Yang Guo∗ and Yan-Gang Miao† School of Physics, Nankai University, Tianjin 300071, China

Thermodynamic geometry is a powerful tool to probe the microstructure of black holes. Based on the Hessian matrix of black hole mass, we introduce a thermodynamic metric and give its scalar curvature in Weinhold’s geometry. The conformal relation between Weinhold’s geometry and Ruppeiner’s geometry will be changed due to the modified first law of for regular black holes. We revisit the thermodynamics of the Bardeen-AdS black hole and find that its obeys the area law. We also investigate the critical behaviour in an extended phase space, which shows the coincidence with that of liquid-gas systems. In the critical phenomena of thermodynamic curvatures, we find that the Bardeen-AdS black hole as a regular black hole has a remarkable similarity to the van der Waals fluid.

I. INTRODUCTION

In the past decades, a complete statistical mechanics that describes the microstructure of black holes was still lacking although there was a wealth of work in the thermodynamics of black holes. A geometric description of thermodynamic systems may provide an alternative attempt to probe the microstructure of black holes. Weinhold developed [1] the mass (energy) representation by constructing a linear vector space with a full metric structure, where the vector space can be constructed directly from empirical laws of thermodynamics. For each extensity Xµ of a , its conjugate field variable vµ is defined by µ vµ ≡ ∂M/∂X , (1) which constitutes a thermodynamic phase space together with Xµ that is required to have the properties of a vector space. Thus, an abstract space having the full properties of inner product on the thermodynamic phase space is given, which implies that the thermodynamic laws can show the underlying structure of a geometric object. Based on the fluctuation theory of thermodynamics, on the other hand, Ruppeiner introduced [2] the entropy representation, which shows that thermodynamic systems can be described by Riemannian manifolds. Quite interesting is that there exists [3] a conformal equivalence between Weinhold’geometry and Ruppeiner’s geometry. In information geometry, a parameterized statistical model is considered as a Riemannian manifold in which the corresponding probability distributions can be defined. In this view, Ruppeiner’s geometry is one of information theories with the probability distribution [4],

p − 1 gR XµXν P(X) ∝ gR e 2 µν , (2) R R where gµν is the Ruppeiner metric and g is the determinant of the metric. Thermodynamic geometries have been extensively studied and applied [5–12] to numerous systems. In particular, the interpretations of thermodynamic curvatures are linked [13] to the interactions in the microstructure of ideal quantum gases. In recent years, Ruppeiner’s geometry has promoted substantially the connection between microstructure and thermodynamics of black holes, furthermore, such a connection has been shown [14–17] to be analogous to that in the van der Waals fluid. These interpretations are new attempts to extract information from the thermodynamic arXiv:2107.01866v2 [gr-qc] 14 Jul 2021 geometry of black holes. The thermodynamic curvature does not depend [18] on the thermodynamic metric since it is an invariant for a given thermodynamic state, which has recently been verified [19, 20] again in two new phase spaces. As is known, there is only an attractive interaction in the van der Waals fluid. However, there exists [21] a weak repulsive interaction in the microstructure of charged AdS black holes. We may argue whether the existence of a weak repulsive interaction has something to do with the singularity of charged AdS black hole spacetimes. On the other hand, the current thermodynamic geometries are generally based on the first law of black hole mechanics associated with linear electrodynamics. They cannot be applied directly to interpret the thermodynamic geometry of regular black holes that have no singularity. We need to construct the thermodynamic geometry in the context of the nonlinear electrodynamics, which is able to give an interpretation of thermodynamic geometry for regular black holes. It is challenging to generalize thermodynamic geometry to regular black holes due to the intrinsic complications in the

[email protected] † Corresponding author: [email protected] 2 configurations of the first law and the Lagrangian coupled with nonlinear electrodynamics. In this paper, we study the Weinhold geometry of regular black holes in the context of nonlinear electrodynamics. Our main task is to construct the Weinhold geometry by using the normalized scalar curvature [21] and reveal the behaviour of thermodynamic curvatures through macroscopic properties. The outline of this paper is as follows. In Sec.II we review briefly the properties of thermodynamic geometry and then give the general form of Weinhold scalar curvatures. Next, in Sec.III we derive the entropy of Bardeen-AdS black holes in terms of the modified integration approach, and demonstrate the critical behaviour of thermodynamics in Bardeen-AdS black holes and the analogous behaviour in liquid-gas systems. In Sec.IV we compute the thermody- namic curvatures of Bardeen-AdS black holes in terms of Weinhold’s geometry and Ruppeiner’s geometry and make a comparison between our results and the thermodynamic curvature of the van der Waals system. Finally, we give our conclusions in Sec.V.

II. GENERAL PROPERTIES OF THERMODYNAMIC GEOMETRY

A. Thermodynamic metrics

We start with reviewing some general properties of thermodynamic geometry theories. The thermodynamic ge- ometry links the statistical mechanics to thermodynamics, in which an appropriate line element is crucial in the equilibrium state space of a thermodynamic system. In the Ruppeiner theory [2, 22], since the thermodynamic fluc- tuation theory originates from statistical mechanics, certain properties of thermodynamics and statistical mechanics for a thermodynamic system are encoded into a single thermodynamic manifold equipped with the line element, 2 R µ ν dsR = gµν dX dX , (3) where the independent variables Xµ with the Greek index, µ = 0, 1, 2, ··· , denote the extensive quantities of a R thermodynamic system and gµν is the so-called Ruppeiner metric defined by the Hessian matrix of thermodynamic entropy, ∂2S(X) gR = − . (4) µν ∂Xµ∂Xν R Here S(X) is the entropy of the thermodynamic system with Boltzmann’s constant, kB = 1, and gµν must be a positive definite matrix1 in order to ensure thermodynamic stability. Furthermore, we take note of a related metric which is defined by the Hessian matrix of black hole mass, known as the Weinhold metric [1], ∂2M(X) gW = , (5) µν ∂Xµ∂Xν which can be obtained from the Ruppeiner metric via a transformation of fluctuation coordinates. Technically, the two metrics that describe Ruppeiner’s and Weinhold’s geometries, respectively, are conformally related [3] to each other, 1 ds2 = ds2 , (6) R T W where T is the temperature of the thermodynamic system under consideration.

B. First law of black hole thermodynamics with nonlinear electrodynamics

For the stationary black holes in nonlinear electrodynamics, the first law of black hole thermodynamics has been established [25, 26] via a covariant approach,

dM = T dS + ΦdQe + ΨdQm, (7) where Qe and Qm denote the electric charge and magnetic charge, respectively. It was claimed [27] that the first law of black hole thermodynamics holds for the black holes with nonlinear electrodynamics, but the Smarr formula seems

1 Thermodynamic metrics must be positive definite since the entropy has [18] a maximum value in equilibrium. This is the condition to ensure thermodynamic stability. However, the positive definiteness may fail [19] due to the non-independence between entropy and volume. It is required [23, 24] to impose Sylvetser’s criterion for a global thermodynamic stability. 3 to be violated. It is easy to check that Eq. (7) is satisfied for Born-Infeld black holes but not for Bardeen black holes. The reason why Eq. (7) breaks down for Bardeen black holes has been clarified [28], where the Lagrangian should include the contribution of additional terms besides being a function of electromagnetic invariants when one tries to derive the first law of black hole thermodynamics. Differing from Rasheed’s formula, the first law of Bardeen black hole thermodynamics takes the form, dM = T dS + Ψdq + Kdq + ΠdM, (8) which contains the extra terms,2 Kdq and ΠdM, due to the variation with respect to mass and charge. In an AdS spacetime, the cosmological constant is interpreted as thermodynamic pressure, P = −Λ/(8π). When P dV term is included, the first law can be rewritten to be

T X Yi dM = dS + dXi, (9) 1 − Π 1 − Π i

i µ where the Latin index takes 1, 2, 3 ··· , X = (V, ··· ) and Yi = (−P, ··· ). When the notations, X = (S, V ··· ) and Yµ = (T/(1 − Π),Yi/(1 − Π)), are taken, Eq. (9) becomes a compact form, µ dM = YµdX . (10) We note that the Einstein notation has been adopted in Eq. (10) and will be used in the following contexts.

C. Thermodynamic curvatures

Given the first law of black hole thermodynamics Eq. (10), the thermodynamic line elements of Weinhold’s geometry can be written as 1 1 ds2 = dT dS + dY dXi. (11) W 1 − Π 1 − Π i Comparing with the line element of Ruppeiner’s geometry [15], 1 1 ds2 = dT dS + dY dXi, (12) R T T i we can see that there exists an conformal factor (1−Π)/T between the Weinhold line element Eq. (11) and Ruppeiner line element Eq. (12). The conformal factor in Eq. (6) is 1/T , now it changes to (1 − Π)/T due to the introduction of extra terms in the first law of black hole thermodynamics. When considering the phase space coordinates (T,Xi), we have  ∂S   ∂S  dS = dT + dXi, (13) ∂T ∂Xi

∂Y   ∂Y  dY = j dT + j dXi. (14) j ∂T ∂Xi Substituting Eqs. (13) and (14) into Eq. (11) and using the relation, ∂S ∂Y = − i , (15) ∂Xi ∂T we obtain the line element of Weinhold’s geometry,   2 1 CXi 2 1 ∂Yi i j dsW = dT + j dX dX , (16) 1 − Π T 1 − Π ∂X T

2 The extra terms originate from the partial derivative of action with respect to additional parameters that are introduced to give the correct Smarr’s formula (see Ref. [28] for the specific expression of extra terms). For Bardeen black holes, Π takes the form of 3 2 2 −3/2 1 − rh(q + rh) which is obviously a dimensionless constant with the value less than one when the charge is set to be a constant. For more details on the first law and Smarr’s formula of black hole thermodynamics in the context of nonlinear electrodynamics coupled to Einstein gravity, please refer to Refs. [25, 26, 28–31]. 4

i i where CXi ≡ T (∂S/∂T )Xi is the at constant X . Since X can take anyone of extensive quantities, the metric Eq. (16) is two dimensional. In the coordinates (T,X), we obtain the general form of a scalar curvature by using its definition in the ,     1 − Π 2 2 RW = 2 2 CX (∂X CX )(∂X,X Y ) − T (∂T,X Y ) + (∂X Y ) (∂X CX ) 2CX (∂X Y )   −T (∂T CX )(∂T,X Y ) + CX − 2(∂X,X CX ) + ∂T,X Y + 2T (∂T,T,X Y ) , (17)

2 where ∂T,X Y ≡ ∂ Y/(∂T ∂X) as known in General Relativity. We note that Eq. (17) shares the same divergent points at CX = 0 or ∂X Y = 0 with the scalar curvature of Ruppeiner’s geometry given [15] by     1 2 2 RR = 2 2 T (∂X Y ) (∂X CX ) + (∂T CX )(∂X Y − T (∂T,X Y )) + CX (∂X Y ) 2CX (∂X Y )  2 +T [(∂X CX )(∂X,X Y ) − T (∂T,X Y ) ] + 2T (∂X Y )(−(∂X,X CX ) + T (∂T,T,X Y )) . (18)

Following Refs. [15, 21] for the treatment to the divergence of Ruppeiner scalar curvature, we introduce the nor- malized scalar curvature3 defined by

R = R CX , (19) and then reduce the Weinhold scalar curvature Eq. (17) and Ruppeiner scalar curvature Eq. (18) to be

2 (1 − Π)[∂T,X Y − T (∂T,X Y ) + 2T (∂T,T,X Y )] RW = 2 , (20) 2(∂X Y )

2 2 2 2 (∂X Y ) − T ((∂T,X Y ) ) + 2T (∂X Y )(∂T,T,X Y ) RR = 2 , (21) 2(∂X Y ) which are so-called normalized thermodynamic scalar curvatures because their divergent behaviour4 at the critical point of phase transitions has been removed.

III. THERMODYNAMICS AND PHASE TRANSITION OF BARDEEN-ADS BLACK HOLES

In this section, we start with a brief review on the thermodynamic properties of Bardeen-AdS black holes and investigate the phase transition via P − V criticality in an extended phase space. The action of Bardeen-AdS black holes with the cosmological constant Λ in the four-dimensional spacetime reads [32, 33] 1 Z √ S = d4x −g (R − 2Λ − 4L(F )) , (22) 16π where g is the determinant of the metric tensor, R is the Ricci scalar, Λ is related to the AdS radius l via the relation Λ = −3/l2, and the function of electromagnetic invariant F , L(F ) is given by !5/2 M p4q2F L = , (23) q3 1 + p4q2F where M and q denote mass and charge of Bardeen-AdS black holes, respectively. In four dimensions, the line element of spherically symmetric Bardeen-AdS black holes takes the form, dr2 ds2 = −f(r)dt2 + + r2dΩ2, (24) f(r)

3 Since the thermodynamic scalar curvature R diverges at CX = 0, CX is treated as a constant with the value of infinitely close to zero, + namely, CX → 0 . Such a normalization just removes the divergence of R at CX = 0, see Ref. [15] for the details, but not the divergence at ∂X Y = 0, which will be shown in Fig.2 and Fig.3. 4 There exists [15, 22] a relation between the divergence of correlation length and the divergence of thermodynamic curvature occurring at the critical point of phase transitions. We shall discuss this issue in detail in sectionIVC. 5 with the shape function, 2Mr2 r2 f(r) = 1 − + , (25) (q2 + r2)3/2 l2 where dΩ2 is the line element of the unit two sphere. The Hawking temperature of Bardeen-AdS black holes can be calculated, 2 2 4 2 −2q + rh + 3rh/l T = 2 2 , (26) 4πrh(q + rh) where rh stands for the horizon radius which is the solution of the equation f(rh) = 0. Note that we have to use the modified first law, Eq. (8), to compute the entropy of the Bardeen black hole, which leads to Z 1 − Π S = dM = πr2 , (27) T h where the form of Π has been given in footnote 2. We see that the Bardeen black hole obeys the area law. Moreover, one can derive the thermodynamic volume conjugated to the thermodynamic pressure,   3  2 3/2 ∂M 4πrh q V = = 1 + 2 . (28) ∂P S,q 3 rh In the above considerations, the black hole mass has been identified with the enthalpy throughout the thermody- namic process in the extended phase space that includes (P,V ). With the clear representations of the pressure and volume, the heat capacity measuring the thermodynamic stability can be obtained,  ∂S  CV = T = 0, (29) ∂T V   2 2 5/2 2 2 4 ∂S 2π(q + rh) (−2q + rh + 8P πrh) CP = T = 4 2 3 2 5 7 . (30) ∂T P 2q rh + 7q rh + (−1 + 24P πq )rh + 8P πrh Unlike the van der Waals fluid, the heat capacity of Bardeen-AdS black holes at constant volume is zero, and the heat capacity at constant pressure is finite. For Bardeen-AdS black holes, the equation of state P = (T,V ) can be derived from the Hawking temperature Eq. (26),  q  2 6V 2/3 2 6V 2/3 12q + π −1 + 2πT −4q + π P = . (31)  2 2 6V 2/3 2π −4q + π The behaviour of isotherms in P − V diagram is shown in Fig.1. Obviously, there exist oscillating parts (unstable regions) in the isotherms for the case of T < Tc, which is similar to that of the van der Waals fluid [34, 35]. For the Bardeen-AdS black hole, the horizontal segments of the isotherms5 shorten when the temperature is increasing, similarly the specific volumes of gas and liquid phases approach to each other with the increasing of temperature for the van der Waals fluid. Moreover, the left and right ends of the horizontal segments merge when the temperature reaches a certain limit of temperature, i.e., the critical temperature Tc whose corresponding pressure is the critical pressure Pc. This critical point is located at the inflection point of the isotherm of T = Tc, which is determined by  ∂P   ∂2P  = 0, 2 = 0. (32) ∂V T ∂V T We then have the critical thermodynamic quantities as follows, √ √  q 1  273 − 17 2 273 + 15 T = − , (33) c 24πq √ √ 2   3 Vc = 19 13 + 15 21 πq , (34) 3 √ 273 + 27 Pc = √ 2 . (35) 12 273 + 15 πq2

5 These horizontal segments are determined by Maxwell’s equal area law, see [36, 37] for the details. They represent the co-existence curve of small and large black hole phases. 6

FIG. 1. P − V diagram of Bardeen-AdS black holes. Isotherms descend from top to bottom when the temperature is decreasing, where the black dashed curve corresponds to T > Tc, the red curve to the critical temperature T = Tc, and the two blue curves to T < Tc.

The small and large black hole phases merge [36, 37] at the critical point. The equation of state can be rewritten to the reduced form,

√   √ p √ q √   219 + 13 273 −18 + 51 + 3 273 − 4 15 + 273T˜ −2 + 273 + 17 V˜ 2/3 V˜ 2/3 P˜ = − , (36)  √ 2 19 −2 + 273 + 17 V˜ 2/3 where the dimensionless variables are defined by P V T P˜ ≡ , V˜ ≡ , T˜ ≡ . (37) Pc Vc Tc We note that the charge does not appear manifestly in the reduced equation of state Eq. (36). It is crucial to give such an equation of state for us to compute the thermodynamic curvatures of Bardeen-AdS black holes in sectionIV.

IV. THERMODYNAMIC GEOMETRY OF BARDEEN-ADS BLACK HOLES

Now we turn to the thermodynamic geometry of Bardeen-AdS black holes which is a powerful tool to probe [38–42] the microstructure of black holes in an extended phase space. We shall present the results in terms of the Ruppeiner’s geometry and Weinhold’s geometry which are constructed in the thermodynamic state space with the temperature and volume fluctuation coordinates, (T,V ).

A. Ruppeiner’s geometry

We consider the line element of in fluctuation coordinates (T,V ) [15],   2 CV 2 1 ∂P 2 dsR = 2 dT − dV , (38) T T ∂V T 7

where the heat capacity vanishes at constant volume, i.e., CV = T (∂S/∂T )V = 0, which leads to a degenerate metric. Nonetheless, we can deal with this issue by introducing [15] the normalized thermodynamic curvature. Using Eq. (21), we can obtain the normalized Ruppeiner thermodynamic scalar curvature,

  q√  √ √ √ 2/3 4/3 2 RR = − 6 2 273 + 15T˜ − 7(17 273 + 281)V˜ − 4(285 273 + 4709)V˜ + (4777 273 + 78929)V˜ √ q √ √ √ +10( 273 + 17) −2 + ( 273 + 17)V˜ 2/3 − 105(285 273 + 4709)V˜ 2/3 + 33(4777 273 + 78929)V˜ 4/3 √ √  q √  √ −3(80069 273 + 1322957)V˜ 2 + 75(17 273 + 281) 3 ( 273 + 17)V˜ 2/3 − 2 (17 273 + 281)V˜ 2/3

√  q√  √ √  2 −5( 273 + 17) − 2 273 + 15T˜ ( 273 + 17)V˜ 2/3 + (17 273 + 281)V˜ 4/3 − 4 ,

(39) which is obviously independent of the charge of black holes when the reduced thermodynamic coordinates are adopted. However, we note that it is a lengthy and cumbersome formula. This is due in large part to the intrinsic complications of the Bardeen black hole in the contexts of its coupling with nonlinear electrodynamics and its rich thermodynamic phase structures [26]. We plot the behaviour of normalized Ruppeiner thermodynamic curvature in Fig.2. We note that the normalized Ruppeiner curvature changes sign at V = 0.16Vc (round to the nearest hundredth), namely, it is positive for V < 0.16Vc but negative for V > 0.16Vc. In general, we point out that the normalized Ruppeiner curvature of Bardeen-AdS black holes has a very similar microstructure to that of van der Waals fluids [15].

FIG. 2. Ruppeiner thermodynamic curvature of Bardeen-AdS black holes. This picture shows the characteristic behaviour of the Ruppeiner thermodynamic curvature as a function of the reduced temperature and volume. 8

B. Weinhold’s geometry

In (T,V ) fluctuation coordinates, the line element Eq. (16) of Weinhold’s geometry reads   2 1 CV 2 1 ∂P 2 dsW = dT + dV , (40) 1 − Π T 1 − Π ∂V T where CV = 0 still leads to a degenerate metric which can be cured by the normalization [15]. We note that Π satisfies 0 < Π < 1 and so it changes only the quantity rather than the sign of thermodynamic curvature. This means that a repulsive interaction cannot be reversed to an attractive interaction by the presence of Π, and vice versa. Without loss of generality, we set the dimensionless constant Π = 0. Using formula Eq. (20), we obtain the normalized Weinhold thermodynamic scalar curvature straightforwardly, √  √ 7 √ √  p p3 ˜ 2/3 2/3 4/3 −57 273 + 15 V RW = ( 273 + 17)V˜ − 2 ( 273 + 17)V˜ + (17 273 + 281)V˜ − 4 √ [( 273 + 17)V˜ 2/3 − 2]7/2 √ √ √ √ 8( 273 + 15)(13 273 + 219)T˜[( 273 + 17)V˜ 2/3 + (17 273 + 281)V˜ 4/3 − 4] √ − √ 4(13 273 + 219) [( 273 + 17)V˜ 2/3 − 2]7  q √ √ √ q√ √ × 3 ( 273 + 17)V˜ 2/3 − 2[(17 273 + 281)V˜ 2/3 − 5( 273 + 17)] − 2 273 + 15T [( 273 + 17)V˜ 2/3

√ 2 +(17 273 + 281)V˜ 4/3 − 4] .

(41) We plot the behaviour of normalized Weinhold thermodynamic curvature in Fig.3.

FIG. 3. Weinhold thermodynamic curvature of Bardeen-AdS black holes. This picture shows the characteristic behaviour of the Weinhold thermodynamic curvature as a function of the reduced temperature and volume.

From the two expressions in Eqs. (39) and (41), it is obvious that the normalized thermodynamic scalar curvatures in the two geometry theories are different. There is, however, a qualitative similarity between the structures of 9

Ruppeiner’s geometry and Weinhold’s geometry as displayed in Fig.2 and Fig.3. The normalized thermodynamic curvatures diverge to negative infinity in very low temperatures. In other words, for both the Weinhold’s geometry and Ruppeiner’s geometry there exists a strongly attractive interaction in the microstructure of Bardeen-AdS black holes in this region of temperature, which can be explicitly seen in the two figures. Furthermore, it is worth noting that the normalized Weinhold curvature is negative and does not change sign, which is due to the existence of conformal relation between Weinhold’s and Ruppeiner’s geometries. According to the microscopic interpretation of interactions [13], the positive and negative curvatures imply the repulsive and attractive interactions, respectively. Hence, for Weinhold’s geometry there are only attractive interactions but no repulsive interactions in the microstructure of Bardeen-AdS black holes, which coincides with the interacting behaviour in the van der Waals fluid.

C. Critical phenomena

In order to get a deep understanding of the normalized thermodynamic curvatures of Bardeen-AdS black holes, we shall focus on the critical phenomena in which the Bardeen-AdS black hole has a markedly different behaviour from that of the charged AdS black holes with spacetime singularities, such as the Reissner-Nordström AdS and Gauss-Bonnet AdS black holes [14, 16, 43]. The normalized Ruppeiner thermodynamic curvature Eq. (39) diverges and vanishes along the following curves on the (V/Vc, T/Tc) plane, respectively, q √  √ √  3 273 + 17 V˜ 2/3 − 2 17 273 + 281 V˜ 2/3 − 5 273 + 17 T˜| = , (42) RR→∞ p√  √ √  2 273 + 15 273 + 17 V˜ 2/3 + 17 273 + 281 V˜ 4/3 − 4  √ √ √ ˜ ˜ 2/3 ˜ 4/3 ˜ 2 T |RR=0 = 3 35(285 273 + 4709)V − 11(4777 273 + 78929)V + (80069 273 + 1322957)V √  q√ q √  √ −25(17 273 + 281) 2 273 + 15 ( 273 + 17)V˜ 2/3 − 2 − 7(17 273 + 281)V˜ 2/3 √ √ √  −4(285 273 + 4709)V˜ 4/3 + (4777 273 + 78929)V˜ 2 + 10( 273 + 17) , (43) which is plotted in Fig.4.

FIG. 4. Critical behaviour of thermodynamic curvatures. The upper (red) curve stands for the diverging case of RR and RW . The lower (blue) curve corresponds to the vanishing case of RR only. 10

From Eq. (41) we find that the Weinhold curvature shares the same diverging curve as the Ruppeiner curvature, i.e., ˜ ˜ T |RW →∞ = T |RR→∞. This happens due to the existence of conformal relation between the Weinhold’s geometry and Ruppeiner’s geometry. However, the Weinhold curvature does not vanish at any finite positive temperature, which is quite different from the Ruppeiner curvature. For Weinhold’s geometry and Ruppeiner’s geometry the normalized thermodynamic curvatures will diverge at the critical point of a phase transition as shown in Fig.4. We can easily see that this divergent point is indeed located at the maximum of the red curve, where the temperature and volume just take their critical values, i.e., T = Tc and V = Vc. This phenomenon was also noticed [44] for the charged AdS black holes with spacetime singularities, where the divergent points of specific heat match those of thermodynamic curvature. Furthermore, a positive normalized thermodynamic curvature was obtained [15, 16] in the small volume phase of the charged AdS black holes with spacetime singularities, namely, a weak repulsive interaction appears in the microstructure, while it does not in the van der Waals fluid. Interestingly, for the Bardeen-AdS black hole that has no spacetime singularities, the temperature converges to zero as the volume decreases, see Fig.4. Such a behaviour has been observed [15] in the van der Waals fluid. Hence, the Bardeen-AdS black hole has the critical phenomena similar to those of van der Waals fluid.

V. CONCLUSIONS

In the present work, we revisit the thermodynamics and investigate the thermodynamic geometry for Bardeen black holes in the anti-de Sitter spacetime. We find that the entropy of Bardeen-AdS black holes still obeys the area law if the modified first law of thermodynamics is considered. We derive the normalized scalar curvature in the Weinhold’s geometry and find that it has the same divergent behaviour as the Ruppeiner scalar curvature at the critical point of a phase transition, but it has no vanishing behaviour at any finite positive temperature while the Ruppeiner scalar curvature has. Moreover, we notice that the additional parameter Π appeared in the conformal factor of Weinhold’s geometry does not reverse the interaction in the microstructure of Bardeen-AdS black holes. When compared to the charged AdS black holes with spacetime singularities, the Bardeen-AdS black hole as a regular black hole has a rather different behaviour, that is, it has no weak repulsive interaction in the microstructure. When compared to the van der Waals fluid, the Bardeen-AdS black hole has a critical behaviour resembling the liquid-gas phase transition and both its Weinhold scalar curvature and Ruppeiner scalar curvature have a very similar microstructure to that of the van der Waals fluid. Our results seem to justify the mathematical analogy of the Bardeen-AdS black hole with the van der Waals fluid. Whether the vanishing of a weak repulsive interaction in the Bardeen-AdS black hole is universal or not in the other regular black holes, we shall deal with this issue by analyzing their thermodynamic geometries in detail.

ACKNOWLEDGEMENTS

The authors would like to thank R.-G Cai and T. Vetsov for their helpful comments and disscussions. This work was supported in part by the National Natural Science Foundation of China under Grant No. 11675081.

[1] F. Weinhold, J. Chem Phys. 63, 2479 (1975). [2] G. Ruppeiner, Phys. Rev. A 20, 1608 (1979). [3] P. Salamon, J. Nulton, and E. Ihrig, J. Chem. Phys. 80, 436 (1984). [4] J. E. Aman, I. Bengtsson, and N. Pidokrajt, Gen. Rel. Grav. 35, 1733 (2003), arXiv:gr-qc/0304015. [5] J. E. Aman, J. Bedford, D. Grumiller, and et al, J. Phys. Conf. Ser. 66, 012007 (2007), arXiv:gr-qc/0611119. [6] J. E. Aman and N. Pidokrajt, Phys. Rev. D 73, 024017 (2006), arXiv:hep-th/0510139. [7] J.-Y. Shen, R.-G. Cai, B. Wang, and R.-K. Su, Int. J. Mod. Phys. A 22, 11 (2007), arXiv:gr-qc/0512035. [8] R.-G. Cai and J.-H. Cho, Phys. Rev. D 60, 067502 (1999), arXiv:hep-th/9803261. [9] Z.-M. Xu, Phys. Lett. B 807, 135529 (2020), arXiv:2006.00665 [gr-qc]. [10] S.-W. Wei and Y.-X. Liu, (2021), arXiv:2106.06704 [gr-qc]. [11] S. A. H. Mansoori, B. Mirza, and M. Fazel, JHEP 04, 115 (2015), arXiv:1411.2582 [gr-qc]. [12] S. A. Hosseini Mansoori and B. Mirza, Phys. Lett. B 799, 135040 (2019), arXiv:1905.01733 [gr-qc]. [13] H. Oshima, T. Obata, and H. Hara, Journal of Physics A Mathematical General 32, 6373 (1999). [14] C. Niu, Y. Tian, and X.-N. Wu, Phys. Rev. D 85, 024017 (2012), arXiv:1104.3066 [hep-th]. [15] S.-W. Wei, Y.-X. Liu, and R. B. Mann, Phys. Rev. D 100, 124033 (2019), arXiv:1909.03887 [gr-qc]. [16] A. Dehyadegari, A. Sheykhi, and S.-W. Wei, Phys. Rev. D 102, 104013 (2020), arXiv:2006.12265 [gr-qc]. 11

[17] B. Wu, C. Wang, Z.-M. Xu, and W.-L. Yang, (2020), arXiv:2006.09021 [gr-qc]. [18] G. Ruppeiner, Springer Proc. Phys. 153, 179 (2014), arXiv:1309.0901 [gr-qc]. [19] Z.-M. Xu, B. Wu, and W.-L. Yang, Phys. Rev. D 101, 024018 (2020), arXiv:1910.12182 [gr-qc]. [20] Z.-M. Xu, Front. Phys. (Beijing) 16, 24502 (2021), arXiv:2011.06736 [gr-qc]. [21] S.-W. Wei, Y.-X. Liu, and R. B. Mann, Phys. Rev. Lett. 123, 071103 (2019), arXiv:1906.10840 [gr-qc]. [22] G. Ruppeiner, Rev. Mod. Phys. 67, 605 (1995), [Erratum: Rev.Mod.Phys. 68, 313–313 (1996)]. [23] T. Vetsov, Eur. Phys. J. C 79, 71 (2019), arXiv:1806.05011 [gr-qc]. [24] H. Dimov, R. C. Rashkov, and T. Vetsov, Phys. Rev. D 99, 126007 (2019), arXiv:1902.02433 [hep-th]. [25] D. A. Rasheed, (1997), arXiv:hep-th/9702087. [26] Z.-Y. Fan and X. Wang, Phys. Rev. D 94, 124027 (2016), arXiv:1610.02636 [gr-qc]. [27] N. Bretón, General Relativity and Gravitation 37, 643 (2005), arXiv:gr-qc/0405116 [gr-qc]. [28] Y. Zhang and S. Gao, Classical and Quantum Gravity 35, 145007 (2018), arXiv:1610.01237 [gr-qc]. [29] W.-J. Mo, R.-G. Cai, and R.-K. Su, Commun. Theor. Phys. 46, 453 (2006). [30] L. Gulin and I. Smolić, Class. Quant. Grav. 35, 025015 (2018), arXiv:1710.04660 [gr-qc]. [31] A. Bokulić, T. Jurić, and I. Smolić, Phys. Rev. D 103, 124059 (2021), arXiv:2102.06213 [gr-qc]. [32] A. G. Tzikas, Physics Letters B 788, 219 (2019), arXiv:1811.01104 [gr-qc]. [33] E. Ayón-Beato and A. García, Physics Letters B 493, 149 (2000), arXiv:gr-qc/0009077 [gr-qc]. [34] D. Kubiznak and R. B. Mann, JHEP 07, 033 (2012), arXiv:1205.0559 [hep-th]. [35] S.-W. Wei and Y.-X. Liu, Phys. Rev. D 87, 044014 (2013), arXiv:1209.1707 [gr-qc]. [36] S.-W. Wei and Y.-X. Liu, Phys. Rev. D 91, 044018 (2015), arXiv:1411.5749 [hep-th]. [37] Z.-Y. Fan, Eur. Phys. J. C 77, 266 (2017), arXiv:1609.04489 [hep-th]. [38] S. A. Hosseini Mansoori, M. Rafiee, and S.-W. Wei, Phys. Rev. D 102, 124066 (2020), arXiv:2007.03255 [gr-qc]. [39] S.-W. Wei and Y.-X. Liu, (2021), arXiv:2105.01295 [gr-qc]. [40] A. Ghosh and C. Bhamidipati, Phys. Rev. D 101, 046005 (2020), arXiv:1911.06280 [gr-qc]. [41] S.-W. Wei and Y.-X. Liu, Phys. Lett. B 803, 135287 (2020), arXiv:1910.04528 [gr-qc]. [42] R. Zhou, Y.-X. Liu, and S.-W. Wei, Phys. Rev. D 102, 124015 (2020), arXiv:2008.08301 [gr-qc]. [43] S.-W. Wei and Y.-X. Liu, Phys. Rev. D 101, 104018 (2020), arXiv:2003.14275 [gr-qc]. [44] S. A. H. Mansoori and B. Mirza, Eur. Phys. J. C 74, 2681 (2014), arXiv:1308.1543 [gr-qc].