Next Article in Journal
Kilonova Emission and Heavy Element Nucleosynthesis
Next Article in Special Issue
De Sitter Entropy in Higher Derivative Theories of Gravity
Previous Article in Journal
Cosmological Constant from Boundary Condition and Its Implications beyond the Standard Model
Previous Article in Special Issue
Analytical Approximate Solutions for Scalarized AdS Black Holes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Holographic p-Wave Superconductor with Excited States in 4D Einstein–Gauss–Bonnet Gravity

1
Key Laboratory of Low Dimensional Quantum Structures and Quantum Control of Ministry of Education, Synergetic Innovation Center for Quantum Effects and Applications, Department of Physics, Hunan Normal University, Changsha 410081, China
2
Institute of Interdisciplinary Studies, Hunan Normal University, Changsha 410081, China
*
Author to whom correspondence should be addressed.
Universe 2023, 9(2), 104; https://doi.org/10.3390/universe9020104
Submission received: 19 January 2023 / Revised: 12 February 2023 / Accepted: 15 February 2023 / Published: 17 February 2023
(This article belongs to the Special Issue Black Holes in Einstein–Gauss–Bonnet Theories)

Abstract

:
We construct a holographic p-wave superconductor with excited states in the 4D Einstein–Gauss–Bonnet gravity using the Maxwell complex vector field model. In the probe limit, we observe that, the higher curvature correction or the higher excited state can hinder the vector condensate to be formed in the full parameter space, which is different from the holographic s-wave superconductor. Regardless of the choice of the vector mass by selecting the value of m 2 L 2 or m 2 L eff 2 , we note that the critical chemical potential becomes evenly spaced for the number of nodes and that the difference of the critical chemical potential between the consecutive states depends on the curvature correction. Moreover, we find that the higher curvature correction or the higher excited state will alter the universal relation of the gap frequency, and the pole and delta function of the conductivity for the excited states can be broadened into the peaks with the finite width as the curvature correction increases.

1. Introduction

The holographic correspondence, which is expected to be a version of the anti-de Sitter/conformal field theory (AdS/CFT) correspondence if the bulk is an AdS space [1], shows that the perturbation theory in the bulk and in a theory with one extra dimension can be used to investigate the field theory living on the boundary of AdS with strong coupling [2,3,4]. In recent years, this correspondence has been extended to explore the phenomena of high-temperature superconductivity [5,6,7,8], which is one of the unsolved mysteries in modern condensed matter physics. Neglecting the backreaction of matter fields on the spacetime metric, Hartnoll et al. first built the holographic s-wave superconductor [9] by spontaneously breaking an Abelian gauge symmetry of a charged scalar field in the Schwarzschild-AdS spacetime [10] and successfully reproduced the important properties of a ( 2 + 1 )-dimensional superconductor. Gubser et al. constructed a holographic p-wave superconductor model by introducing an S U ( 2 ) Yang–Mills field into the bulk [11], and the authors of [12,13] presented a holographic realization of d-wave superconductivity by introducing a charged massive spin two field propagating in the bulk. Considering a charged vector field into an Einstein–Maxwell theory, Cai et al. introduced a new holographic p-wave superconductor [14] and pointed out that this model is a generalization of the S U ( 2 ) model with a general mass and gyromagnetic ratio [15].
It should be noted that the aforementioned holographic superconductor models focus on the ground state that corresponds to the minimum free energy state. More recently, Wang et al. constructed the holographic s-wave superconductor with the excited states in the probe limit [16] and beyond the probe limit [17], which shows that, as in the evidence of the existence of excited states, the conductivity of each excited state has an additional pole in the imaginary part and a delta function in the real part arising at the low temperature inside the gap. To go a step further, Li et al. studied the non-equilibrium process of the excited states for the holographic s-wave superconductor in detail [18], and we developed a general analytic technique to investigate the excited states of the holographic s-wave and p-wave superconductors [19]. Owing to the importance of the excited states in condensed matter physics, there has been an increasing interest in the excited states regarding the studies of holographic superconductors in the AdS black hole spacetime [20,21,22,23] and soliton spacetime [24].
In this work, in order to systematically understand the influences of the 1 / N or 1 / λ ( λ is the ’t Hooft coupling) corrections on the holographic dual models, we will generalize the investigation on the excited states in the holographic p-wave superconductor to the ( 2 + 1 )-dimensional Gauss–Bonnet model inspired by the work of Glavan and Lin [25], where a 4D Einstein–Gauss–Bonnet gravity was introduced by adopting the Gauss–Bonnet coupling parameter α to scale as 1 / ( D 4 ) in the limit D 4 . This novel 4D gravity preserves the number of graviton degrees of freedom, thus being free from Ostrogradsky instability, and the Gauss–Bonnet coupling will produce the nontrivial dynamics. However, some independent groups pointed out quite serious doubts regarding the credibility of this 4D Gauss–Bonnet bulk construction since the original claim of [25] explicitly contradicts with the established theorem, i.e., the Lovelock theorem [26,27,28,29,30,31,32]. Thus, the researchers further proposed the regularized 4D Einstein–Gauss–Bonnet gravity [33,34] and the consistent D 4 Einstein–Gauss–Bonnet gravity [35]. In particular, the consistent theory of D 4 Einstein–Gauss–Bonnet gravity, which is also called the 4D Aoki–Gorji–Mukohyama theory of gravity, can be served as a consistent realization of D 4 Einstein–Gauss–Bonnet gravity with the rescaled Gauss–Bonnet coupling constant by taking into account all of the above issues [36,37], and recover the black hole solutions provided in [25] and the well-known Friedmann–Lemaître–Robertson–Walker (FLRW) metric. These 4D Einstein–Gauss–Bonnet gravity theories exhibit a number of interesting phenomena in different areas, for reviews, see [38] and the references therein. Interestingly, an exact solution of the same form, named the 4D planar Gauss–Bonnet-AdS black hole, can be obtained via these gravity theories [33,34,35,36,37,39,40,41]:
d s 2 = f ( r ) d t 2 + d r 2 f ( r ) + r 2 ( d x 2 + d y 2 ) ,
with the metric coefficient
f = f ( r ) = r 2 2 α 1 1 4 α L 2 1 r + 3 r 3 ,
where α is the Gauss–Bonnet coupling and r + is the black hole horizon related to the Hawking temperature T H = 3 r + / ( 4 π L 2 ) . Since f r 2 ( 1 1 4 α / L 2 ) / ( 2 α ) , we can define the so-called effective asymptotic AdS scale by [42]
L eff 2 = 2 α 1 1 4 α L 2 ,
where the upper bound α = L 2 / 4 is the Chern–Simons limit [43]. It should be noted that, since we are interested in the application of the Mermin–Wagner theorem to the holographic superconductors, we prefer the Einstein–Gauss–Bonnet gravity theory, where specific combinations of the curvature tensors are used, over other modified theories, such as the f ( R ) gravity, where powers of the Ricci scalar only are used [44,45,46].
As a matter of fact, for the s-wave superconductor, we have constructed the holographic dual model with the excited states away from the probe limit in the 4D Aoki–Gorji–Mukohyama theory of gravity, and observed that the combination of the Gauss–Bonnet gravity and excited state provides richer physics in the scalar condensates and conductivity in the ( 2 + 1 )-dimensional superconductor [47], which exhibits a very interesting and different feature when compared to the higher-dimensional Gauss–Bonnet superconductor [44,45,46]. Bao et al. studied the excited states of the holographic s-wave superconductor with the scalar field coupled to the Gauss–Bonnet invariance and found that the Gauss–Bonnet coupling has interesting prints on the critical temperature, condensation strength, and optical conductivity for the ground state and excited states [48]. However, the study on the ( 2 + 1 )-dimensional Gauss–Bonnet p-wave superconductor is called for. Thus, we will build the holographic p-wave superconductor in the 4D Einstein–Gauss–Bonnet gravity via the Maxwell complex vector field model and investigate the effects of the curvature corrections on the excited states of the ( 2 + 1 )-dimensional p-wave superconductor. Since the probe limit can extract the main physics of the holographic dual system and avoid the complex computation, we will concentrate on this limit in the following.
This work is organized as follows. In Section 2, we will construct the holographic p-wave superconductor model in the 4D Gauss–Bonnet-AdS black hole background. In Section 3, we will investigate the condensate of the vector field and the phase transition in the ( 2 + 1 )-dimensional Gauss–Bonnet p-wave superconductor both for the ground state and excited states. In Section 4, we will calculate the electrical conductivity of our holographic p-wave model with the excited states. We will conclude, in the last section, with our main results.

2. Description of the p-Wave Model

We will build the holographic p-wave superconductor by using the Maxwell complex vector field model [14,15]:
S = d 4 x g 1 4 F μ ν F μ ν 1 2 ρ μ ν ρ μ ν m 2 ρ μ ρ μ + i q γ 0 ρ μ ρ ν F μ ν ,
with the Maxwell field strength F μ ν = μ A ν ν A μ and the tensor ρ μ ν = D μ ρ ν D ν ρ μ . Here, ρ μ is the complex vector field with mass m and charge q, D μ = μ i q A μ denotes the covariant derivative, and γ 0 represents the interaction between the vector field ρ μ and the gauge field A μ .
Since we work in a probe limit where the backreaction of matter fields on the spacetime metric is neglected, the gravitational background will be fixed and the matter fields can be treated as the perturbations. Varying the action (4) with respect to the vector field ρ μ and gauge field A μ , we have
D ν ρ ν μ m 2 ρ μ + i q γ 0 ρ ν F ν μ = 0 ,
ν F ν μ i q ( ρ ν ρ ν μ ρ ν ρ ν μ ) i q γ 0 ν ( ρ ν ρ μ ρ ν ρ μ ) = 0 .
As in [14,15], we also adopt the ansatz for the matter fields ρ μ d x μ = ρ x ( r ) d x and A μ d x μ = A t ( r ) d t . Thus, under the metric (1), the above equations of motion become
ρ x + f f ρ x 1 f m 2 q 2 A t 2 f ρ x = 0 ,
A t + 2 r A t 2 q 2 ρ x 2 r 2 f A t = 0 ,
where the prime denotes the differentiation in r.
In order to investigate the excited states of the holographic p-wave superconductor in the 4D Einstein–Gauss–Bonnet gravity, we have to numerically solve Equations (7) and (8) by using the shooting method [9]. For the boundary condition at the horizon r = r + , the field ρ x is regular but A t obeys A t ( r + ) = 0 . Near the AdS boundary r , we obtain the asymptotic behaviors
ρ x = ρ x r Δ + ρ x + r Δ + , A t = μ ρ r ,
with the characteristic exponents defined by Δ ± = 1 ± 1 + 4 m 2 L eff 2 / 2 . From the AdS/CFT correspondence, μ and ρ are interpreted as the chemical potential and charge density, and ρ x and ρ x + correspond to the source and vacuum expectation value of the vector operator O x in the dual field theory, respectively. In this work, we will impose the boundary condition ρ x = 0 and set Δ = Δ + for clarity.

3. Condensate and Phase Transition

It is well known that a phase transition is a change in state from one phase to another; it can be driven by many parameters, such as the temperature, pressure, etc. Choosing the temperature as the driving parameter, one finds that the high-temperature phase is almost always more disordered, i.e., has a higher symmetry than the low-temperature phase. In the holographic superconductor model, a charged condensate forms below a critical temperature T c , which is the symmetry breaking phase transition to a superconducting phase. Now, we are in a position to generalize the investigation on the vector condensate and phase transition for the ground state of the holographic p-wave superconductor in the 4D Einstein–Gauss–Bonnet gravity [49] to those for the excited states. When performing numerical calculations, we will use the following scaling symmetries
r λ r , ( t , x , y ) λ 1 ( t , x , y ) , q q , ( ρ x , A t ) λ ( ρ x , A t ) , μ λ μ , ρ λ 2 ρ , ρ x + λ 1 + Δ ρ x + ,
to build the invariant and dimensionless quantities, where λ is a real positive number. For simplicity, we set r + = 1 and q = 1 in the following calculation. On the other hand, from [49], we observe that, regardless of the choice of the mass of the vector field, by choosing the value of m 2 L 2 or m 2 L eff 2 , the critical temperature T c decreases as α increases for all vector field masses in the ground state. We will check it in the excited states and expect this tendency to be the same. For concreteness, we choose the Gauss–Bonnet parameter in the range L 2 / 10 α L 2 / 4 as in [49] and fix the masses of the vector field m 2 L 2 = 3 / 4 and m 2 L eff 2 = 3 / 4 in our calculation. It should be noted that the other choices of the vector mass will not qualitatively modify our results.
In Figure 1, we present the condensates of the vector operator O x as a function of temperature for various Gauss–Bonnet parameters α from the ground state ( n = 0 ) to the second excited state ( n = 2 ) with the fixed vector masses m 2 L 2 = 3 / 4 and m 2 L eff 2 = 3 / 4 . Obviously, there exists the critical behavior of the second-order phase transition O x ( 1 T / T c ) 1 / 2 for the fixed α and n, and each curve is in agreement with that of the holographic p-wave superconducting phase in the literature, where the condensate goes to a constant at zero temperature. For the fixed number of nodes n, we observe that this constant, which corresponds to the condensation gap, increases as α increases, which suggests that, regardless of the ground state or the excited states, the higher curvature correction causes the vector hair to be more difficult to be developed in the ( 2 + 1 )-dimensional p-wave superconductor. This result is independent of the choice of the vector mass by selecting the value of m 2 L 2 or m 2 L eff 2 , which can be used to back up the numerical finding as shown in [49] for the ground state of the ( 2 + 1 )-dimensional Gauss–Bonnet superconductor. For the fixed Gauss–Bonnet parameter α , we see that the condensation gap increases as n increases, which indicates that the higher excited state causes it to be harder for the condensate of the vector operator to form. Thus, the p-wave superconductor still exists even when we consider the Gauss–Bonnet corrections to the standard ( 2 + 1 ) -dimensional holographic superconductor model with the excited states.
In order to further support the observation obtained in Figure 1, we provide the critical temperature T c as a function of the Gauss–Bonnet parameter from the ground state ( n = 0 ) to the third excited state ( n = 3 ) for the fixed vector masses m 2 L 2 = 3 / 4 and m 2 L eff 2 = 3 / 4 in Figure 2. It is shown that, regardless of the choice of the mass of the vector field by choosing the value of m 2 L 2 or m 2 L eff 2 , we find that the increase in the Gauss–Bonnet parameter α or the number of nodes n results in the decrease in the critical temperature T c , which can be used to back up the numerical finding as shown in Figure 1 that the higher curvature correction or the higher excited state can hinder the condensate to be formed in the full parameter space. This behavior is reminiscent of that seen for the s-wave case where the curvature correction has a more subtle effect, i.e., the critical temperature first decreases then increases as the curvature correction tends toward the Chern–Simons value. Increasing n will weaken this subtle effect of the Gauss–Bonnet parameter on the critical temperature [47], so we conclude that the curvature correction has completely different effects on the critical temperature T c for the s-wave and p-wave superconductors with the excited states in ( 2 + 1 )-dimensions.
As a matter of fact, we can also realize the phase transition of the holographic superconductor around the critical chemical potential μ c , above which the vector field begins to condensate. Therefore, we present the critical chemical potential μ c of the vector operator O x with the masses of the vector field m 2 L 2 = 3 / 4 and m 2 L eff 2 = 3 / 4 from the ground state to the 20-th excited state for different values of the Gauss–Bonnet parameter, i.e., α = 0.10 , 0, 0.10 , and 0.25 in Table 1. It is found that the critical chemical potential μ c increases with the increase in α or n, which agrees well with the finding of the critical temperature T c and indicates that the higher curvature correction or the higher excited state causes the condensate of the vector operator O x to be harder. Obviously, the critical chemical potential μ c of m 2 L 2 = 3 / 4 (left column) is much closer to that of m 2 L eff 2 = 3 / 4 (right column) as the number of nodes n increases for each α , for example, in the case of n = 20 here, which means that, regardless of the choice of the mass by selecting the value of m 2 L 2 or m 2 L eff 2 , the critical chemical potentials are the same for a sufficiently large n. Fitting the relation between μ c and n by using the results in Table 1, we have
μ c 4.820 n + 5.042 , for α = 0.10 , 5.189 n + 5.345 , for α = 0.00 , 5.714 n + 5.772 , for α = 0.10 , 7.931 n + 7.391 , for α = 0.25 ,
for the vector mass m 2 L 2 = 3 / 4 and
μ c 4.821 n + 4.941 , for α = 0.10 , 5.189 n + 5.345 , for α = 0.00 , 5.712 n + 5.927 , for α = 0.10 , 7.919 n + 8.479 , for α = 0.25 ,
for the vector mass m 2 L eff 2 = 3 / 4 , which shows that the critical chemical potential μ c becomes evenly spaced for the number of nodes n and that the difference of μ c between the consecutive states does not depend on the choice of the vector mass but does depend on the Gauss–Bonnet parameter, i.e., Δ μ c = 4.8 for α = 0.10 , Δ μ c = 5.2 for α = 0.00 , Δ μ c = 5.7 for α = 0.10 , and Δ μ c = 7.9 for α = 0.25 . Interestingly, increasing the Gauss–Bonnet parameter α increases this difference of the critical chemical potential between the consecutive states.

4. Conductivity

In order to calculate the conductivity of the ( 2 + 1 )-dimensional Gauss–Bonnet p-wave superconductor with the excited states, we assume that the perturbed Maxwell field has a form δ A y = e i ω t A y ( r ) d y , just as in [14]. Thus, the equation of motion of the perturbation A y ( r ) is
A y + f f A y + ω 2 f 2 2 q 2 ρ x 2 r 2 f A y = 0 .
At the horizon, we impose the ingoing wave boundary condition A y ( r ) f ( r ) i ω / ( 4 π T H ) . At the asymptotic boundary ( r ), we obtain
A y = A ( 0 ) + A ( 1 ) r ,
which leads to the conductivity of the dual superconductor
σ = i A ( 1 ) ω A ( 0 ) .
Considering that the different choices of the vector mass do not qualitatively modify the effect of the Gauss–Bonnet correction on the behavior of the condensates for the vector operator O x , we will fix the mass of the vector field by m 2 L eff 2 = 3 / 4 in this section.
In Figure 3, we plot the conductivity of the ( 2 + 1 )-dimensional p-wave superconductor for the vector mass m 2 L eff 2 = 3 / 4 with different values of the Gauss–Bonnet parameter from the ground state to the third excited state at temperature T / T c 0.20 . It is shown that, similar to the ground state [49], there also exists a gap in the conductivity of the excited state and the gap frequency ω g increases as the Gauss–Bonnet parameter α increases, where we define the gap frequency ω g as the frequency minimizing Im[ σ ( ω ) ] [50]. Obviously, the increasing Gauss–Bonnet parameter α or number of nodes n is to increase ω g / T c , which implies that the higher curvature correction or the higher excited state will alter the universal relation ω g / T c 8 [50] for the ( 2 + 1 )-dimensional p-wave superconductor. Moreover, we find that, as evidence of the existence of excited states, there exist additional poles in Im[ σ ( ω ) ] and delta functions in Re[ σ ( ω ) ] arising at low temperatures for the excited states, which is similar to the holographic s-wave superconductor with the excited states [16,17]. Interestingly, as α increases, we clearly see that the pole and delta function can be broadened into the peaks with the finite width, which is consistent with the finding in the ( 2 + 1 )-dimensional Gauss–Bonnet s-wave superconductor with the excited states [47]. This may be a quite general feature for the holographic superconductor with the excited states in the 4D Einstein–Gauss–Bonnet gravity.

5. Conclusions

In order to understand the influences of the 1 / N or 1 / λ corrections on the vector condensate, we have built the ( 2 + 1 )-dimensional p-wave superconductor with the excited states in the 4D Einstein–Gauss–Bonnet gravity using the Maxwell complex vector field model. In the probe limit, regardless of the choice of the mass of the vector field by choosing the value of m 2 L 2 or m 2 L eff 2 , the system undergoes a second-order phase transition near the critical temperature T c both for the ground state and excited states. We noted that, different from the holographic s-wave superconductor where the Gauss–Bonnet parameter α and the number of nodes n have a more subtle effect on the critical temperature, in the holographic p-wave case, the critical temperature T c always decreases as α or n increases in the full parameter space. This means that the curvature corrections have completely different effects on the critical temperature T c for the s-wave and p-wave superconductors with the excited states in ( 2 + 1 )-dimensions and implies that the higher curvature correction or the higher excited state will cause the vector hair to be more difficult to develop. We also calculated the critical chemical potential μ c of the excited states, and found that μ c becomes evenly spaced for the number of nodes n and that the difference of μ c between the consecutive states is independent of the choice of the vector mass but dependent of the Gauss–Bonnet parameter α , i.e., increasing α increases the difference of μ c between the consecutive states. Furthermore, we investigated the conductivity of the system and showed that the gap frequency ω g / T c increases with the increase in the Gauss–Bonnet parameter α or number of nodes n, which indicates that the higher curvature correction or the higher excited state will alter the universal relation ω g / T c 8 for the ( 2 + 1 )-dimensional p-wave superconductors. Interestingly, we observed that, as the Gauss–Bonnet parameter α increases, the pole in the imaginary part of the conductivity and delta function in the real part for the excited states can be broadened into the peaks with the finite width, which agrees well with the finding in the ( 2 + 1 )-dimensional Gauss–Bonnet s-wave superconductor with the excited states. This may be a quite general feature for the holographic superconductors with the excited states in the 4D Einstein–Gauss–Bonnet gravity.

Author Contributions

Investigation, D.W., X.D., Q.P. and J.J.; writing—original draft preparation, D.W. and X.D.; writing—review and editing, Q.P.; supervision, Q.P. and J.J. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Key Research and Development Program of China (Grant No. 2020YFC2201400), National Natural Science Foundation of China (Grant Nos. 12275079, 12035005 and 11690034) and Postgraduate Scientific Research Innovation Project of Hunan Province (Grant No. CX20210472).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bousso, R. The holographic principle. Rev. Mod. Phys. 2002, 74, 825. [Google Scholar] [CrossRef] [Green Version]
  2. Maldacena, J. The large-N limit of superconformal field theories and supergravity. Int. J. Theor. Phys. 1999, 38, 1113. [Google Scholar] [CrossRef] [Green Version]
  3. Witten, E. Anti-de Sitter space and holography. Adv. Theor. Math. Phys. 1998, 2, 253. [Google Scholar] [CrossRef] [Green Version]
  4. Gubser, S.S.; Klebanov, I.R.; Polyakov, A.M. Gauge theory correlators from non-critical string theory. Phys. Lett. B 1998, 428, 105. [Google Scholar] [CrossRef] [Green Version]
  5. Hartnoll, S.A. Lectures on holographic methods for condensed matter physics. Class. Quant. Grav. 2009, 26, 224002. [Google Scholar] [CrossRef] [Green Version]
  6. Herzog, C.P. Lectures on Holographic Superfluidity and Superconductivity. J. Phys. A 2009, 42, 343001. [Google Scholar] [CrossRef] [Green Version]
  7. Horowitz, G.T. Introduction to Holographic Superconductors. Lect. Notes Phys. 2011, 828, 313. [Google Scholar]
  8. Cai, R.G.; Li, L.; Li, L.F.; Yang, R.Q. Introduction to Holographic Superconductor Models. Sci. China Phys. Mech. Astron. 2015, 58, 060401. [Google Scholar] [CrossRef] [Green Version]
  9. Hartnoll, S.A.; Herzog, C.P.; Horowitz, G.T. Building a Holographic Superconductor. Phys. Rev. Lett. 2008, 101, 031601. [Google Scholar] [CrossRef] [Green Version]
  10. Gubser, S.S. Breaking an Abelian gauge symmetry near a black hole horizon. Phys. Rev. D 2008, 78, 065034. [Google Scholar] [CrossRef] [Green Version]
  11. Gubser, S.S.; Pufu, S.S. The gravity dual of a p-wave superconductor. J. High Energy Phys. 2008, 11, 33. [Google Scholar] [CrossRef]
  12. Chen, J.W.; Kao, Y.J.; Maity, D.; Wen, W.Y.; Yeh, C.P. Towards a holographic model of d-wave superconductors. Phys. Rev. D 2010, 81, 106008. [Google Scholar] [CrossRef] [Green Version]
  13. Benini, F.; Herzog, C.P.; Rahman, R.; Yarom, A. Gauge gravity duality for d-wave superconductors: Prospects and challenges. J. High Energy Phys. 2010, 11, 137. [Google Scholar] [CrossRef] [Green Version]
  14. Cai, R.G.; He, S.; Li, L.; Li, L.F. A holographic study on vector condensate induced by a magnetic field. J. High Energy Phys. 2013, 12, 36. [Google Scholar] [CrossRef] [Green Version]
  15. Cai, R.G.; Li, L.; Li, L.F. A holographic p-wave superconductor model. J. High Energy Phys. 2014, 1, 32. [Google Scholar] [CrossRef] [Green Version]
  16. Wang, Y.Q.; Hu, T.T.; Liu, Y.X.; Yang, J.; Zhao, L. Excited states of holographic superconductors. J. High Energy Phys. 2020, 6, 13. [Google Scholar] [CrossRef]
  17. Wang, Y.Q.; Li, H.B.; Liu, Y.X.; Zhong, Y. Excited states of holographic superconductors with backreaction. Eur. Phys. J. C 2021, 81, 628. [Google Scholar] [CrossRef]
  18. Li, R.; Wang, J.; Wang, Y.Q.; Zhang, H.B. Nonequilibrium dynamical transition process between excited states of holographic superconductors. J. High Energy Phys. 2020, 11, 59. [Google Scholar] [CrossRef]
  19. Qiao, X.Y.; Wang, D.; OuYang, L.; Wang, M.J.; Pan, Q.Y.; Jing, J.L. An analytic study on the excited states of holographic superconductors. Phys. Lett. B 2020, 811, 135864. [Google Scholar] [CrossRef]
  20. Xiang, Q.; Zhao, L.; Wang, Y.Q. Excited states of holographic superconductors from massive gravity. Commun. Theor. Phys. 2022, 74, 115401. [Google Scholar] [CrossRef]
  21. Zhang, S.H.; Zhao, Z.X.; Pan, Q.Y.; Jing, J.L. Excited states of holographic superconductors with hyperscaling violation. Nucl. Phys. B 2022, 976, 115701. [Google Scholar] [CrossRef]
  22. Nguyen, T.T.; Phat, T.H. Asymptotic critical behavior of holographic phase transition at finite topological charge–the spectrum of excited states becomes continuous at T = 0. J. High Energy Phys. 2022, 6, 4. [Google Scholar] [CrossRef]
  23. Xiang, Q.; Zhao, L.; Wang, Y.Q. Spontaneously Translational Symmetry Breaking in the Excited States of Holographic Superconductor. arXiv 2022, arXiv:2207.10593. [Google Scholar]
  24. Ouyang, L.; Wang, D.; Qiao, X.Y.; Wang, M.J.; Pan, Q.Y.; Jing, J.L. Holographic Insulator/Superconductor Phase Transitions with Excited States. Sci. China Phys. Mech. Astron. 2021, 64, 240411. [Google Scholar] [CrossRef]
  25. Glavan, D.; Lin, C.S. Einstein–Gauss–Bonnet Gravity in Four-Dimensional Spacetime. Phys. Rev. Lett. 2020, 124, 081301. [Google Scholar] [CrossRef] [Green Version]
  26. Gürses, M.; Şişman, T.Ç.; Tekin, B. Is there a novel Einstein–Gauss–Bonnet theory in four dimensions? Eur. Phys. J. C 2020, 80, 647. [Google Scholar] [CrossRef]
  27. Ai, W.Y. A note on the novel 4D Einstein–Gauss–Bonnet gravity. Commun. Theor. Phys. 2020, 72, 095402. [Google Scholar] [CrossRef]
  28. Mahapatra, S. A note on the total action of 4D Gauss–Bonnet theory. Eur. Phys. J. C 2020, 80, 992. [Google Scholar] [CrossRef]
  29. Shu, F.W. Vacua in novel 4D Einstein–Gauss–Bonnet Gravity: Pathology and instability? Phys. Lett. B 2020, 811, 135907. [Google Scholar] [CrossRef]
  30. Tian, S.X.; Zhu, Z.H. Non-full equivalence of the four-dimensional Einstein–Gauss–Bonnet gravity and Horndeksi gravity for Bianchi type I metric. arXiv 2020, arXiv:2004.09954. [Google Scholar]
  31. Arrechea, J.; Delhom, A.; Jiménez-Cano, A. Inconsistencies in four-dimensional Einstein–Gauss–Bonnet gravity. Chin. Phys. C 2021, 45, 013107. [Google Scholar] [CrossRef]
  32. Gürses, M.; Şişman, T.Ç.; Tekin, B. Comment on “Einstein–Gauss–Bonnet Gravity in Four-Dimensional Spacetime”. Phys. Rev. Lett. 2020, 125, 149001. [Google Scholar] [CrossRef] [PubMed]
  33. Lu, H.; Pang, Y. Horndeski Gravity as D→4 Limit of Gauss–Bonnet. Phys. Lett. B 2020, 809, 135717. [Google Scholar] [CrossRef]
  34. Hennigar, R.A.; Kubiznak, D.; Mann, R.B.; Pollack, C. On Taking the D→4 limit of Gauss–Bonnet Gravity: Theory and Solutions. J. High Energy Phys. 2020, 7, 27. [Google Scholar] [CrossRef]
  35. Aoki, K.; Gorji, M.A.; Mukohyama, S. A consistent theory of D→4 Einstein–Gauss–Bonnet gravity. Phys. Lett. B 2020, 810, 135843. [Google Scholar] [CrossRef]
  36. Aoki, K.; Gorji, M.A.; Mukohyama, S. Cosmology and gravitational waves in consistent D→4 Einstein–Gauss–Bonnet gravity. J. Cosmol. Astropart. Phys. 2020, 9, 14. [Google Scholar] [CrossRef]
  37. Aoki, K.; Gorji, M.A.; Mizuno, S.; Mukohyama, S. Inflationary gravitational waves in consistent D→4 Einstein–Gauss–Bonnet gravity. J. Cosmol. Astropart. Phys. 2021, 1, 54. [Google Scholar] [CrossRef]
  38. Fernandes, P.G.S.; Carrilho, P.; Clifton, T.; Mulryne, D.J. The 4D Einstein–Gauss–Bonnet theory of gravity: A review. Class. Quantum Grav. 2022, 39, 063001. [Google Scholar] [CrossRef]
  39. Konoplya, R.A.; Zhidenko, A. Black holes in the four-dimensional Einstein–Lovelock gravity. Phys. Rev. D 2020, 101, 084038. [Google Scholar] [CrossRef] [Green Version]
  40. Fernandes, P.G.S. Charged Black Holes in AdS Spaces in 4D Einstein Gauss–Bonnet Gravity. Phys. Lett. B 2020, 805, 135468. [Google Scholar] [CrossRef]
  41. Konoplya, R.A.; Zhidenko, A. 4D Einstein–Lovelock black holes: Hierarchy of orders in curvature. Phys. Lett. B 2020, 807, 135607. [Google Scholar] [CrossRef]
  42. Cai, R.G. Gauss–Bonnet black holes in AdS spaces. Phys. Rev. D 2002, 65, 084014. [Google Scholar] [CrossRef] [Green Version]
  43. Crisostomo, J.; Troncoso, R.; Zanelli, J. Black hole scan. Phys. Rev. D 2000, 62, 084013. [Google Scholar] [CrossRef] [Green Version]
  44. Gregory, R.; Kanno, S.; Soda, J. Holographic superconductors with higher curvature corrections. J. High Energy Phys. 2009, 10, 10. [Google Scholar] [CrossRef] [Green Version]
  45. Pan, Q.Y.; Wang, B.; Papantonopoulos, E.; Oliveria, J.; Pavan, A.B. Holographic superconductors with various condensates in Einstein–Gauss–Bonnet gravity. Phys. Rev. D 2010, 81, 106007. [Google Scholar] [CrossRef] [Green Version]
  46. Brihaye, Y.; Hartmann, B. Holographic superconductors in 3+1 dimensions away from the probe limit. Phys. Rev. D 2010, 81, 126008. [Google Scholar] [CrossRef] [Green Version]
  47. Pan, J.; Qiao, X.Y.; Wang, D.; Pan, Q.Y.; Nie, Z.Y.; Jing, J.L. Holographic superconductors in 4D Einstein–Gauss–Bonnet gravity with backreactions. Phys. Lett. B 2021, 823, 136755. [Google Scholar] [CrossRef]
  48. Bao, Y.; Guo, H.; Kuang, X.M. Excited states of holographic superconductor with scalar field coupled to Gauss–Bonnet invariance. Phys. Lett. B 2021, 822, 136646. [Google Scholar] [CrossRef]
  49. Qiao, X.Y.; OuYang, L.; Wang, D.; Pan, Q.Y.; Jing, J.L. Holographic superconductors in 4D Einstein–Gauss–Bonnet gravity. J. High Energy Phys. 2020, 12, 192. [Google Scholar] [CrossRef]
  50. Horowitz, G.T.; Roberts, M.M. Holographic superconductors with various condensates. Phys. Rev. D 2008, 78, 126008. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (Color online) The condensates of the vector operator O x as a function of temperature for the masses of the vector field m 2 L 2 = 3 / 4 (left) and m 2 L eff 2 = 3 / 4 (right) with different values of the Gauss–Bonnet parameter from the ground state to the second excited state. In each panel, the four lines correspond to increasing Gauss–Bonnet parameters, i.e., α = 0.1 (red and dashed), 0.0 (blue), 0.1 (green), and 0.25 (black and dashed), respectively.
Figure 1. (Color online) The condensates of the vector operator O x as a function of temperature for the masses of the vector field m 2 L 2 = 3 / 4 (left) and m 2 L eff 2 = 3 / 4 (right) with different values of the Gauss–Bonnet parameter from the ground state to the second excited state. In each panel, the four lines correspond to increasing Gauss–Bonnet parameters, i.e., α = 0.1 (red and dashed), 0.0 (blue), 0.1 (green), and 0.25 (black and dashed), respectively.
Universe 09 00104 g001
Figure 2. (Color online) The critical temperature T c of the vector operator O x as a function of the Gauss–Bonnet parameter for the fixed masses of the vector field m 2 L 2 = 3 / 4 (left) and m 2 L eff 2 = 3 / 4 (right) from the ground state to the third excited state.
Figure 2. (Color online) The critical temperature T c of the vector operator O x as a function of the Gauss–Bonnet parameter for the fixed masses of the vector field m 2 L 2 = 3 / 4 (left) and m 2 L eff 2 = 3 / 4 (right) from the ground state to the third excited state.
Universe 09 00104 g002
Figure 3. (Color online) The conductivity for the mass of the vector field m 2 L eff 2 = 3 / 4 with different values of the Gauss–Bonnet parameter from the ground state to the third excited state, where the solid line and dashed line represent the real part and imaginary part of the conductivity σ ( ω ) . In each panel, the blue, green, red, and black lines denote the ground ( n = 0 ), first ( n = 1 ), second ( n = 2 ), and third ( n = 3 ) states, respectively.
Figure 3. (Color online) The conductivity for the mass of the vector field m 2 L eff 2 = 3 / 4 with different values of the Gauss–Bonnet parameter from the ground state to the third excited state, where the solid line and dashed line represent the real part and imaginary part of the conductivity σ ( ω ) . In each panel, the blue, green, red, and black lines denote the ground ( n = 0 ), first ( n = 1 ), second ( n = 2 ), and third ( n = 3 ) states, respectively.
Universe 09 00104 g003aUniverse 09 00104 g003b
Table 1. The critical chemical potential μ c of the vector operator O x for the masses of the vector field m 2 L 2 = 3 / 4 (left column) and m 2 L eff 2 = 3 / 4 (right column) with different values of the Gauss–Bonnet parameter from the ground state to the 20-th excited state. It should be noted that the critical chemical potentials for the different choices of the vector mass are the same for the case of α = 0 , i.e., the Schwarzschild-AdS black hole.
Table 1. The critical chemical potential μ c of the vector operator O x for the masses of the vector field m 2 L 2 = 3 / 4 (left column) and m 2 L eff 2 = 3 / 4 (right column) with different values of the Gauss–Bonnet parameter from the ground state to the 20-th excited state. It should be noted that the critical chemical potentials for the different choices of the vector mass are the same for the case of α = 0 , i.e., the Schwarzschild-AdS black hole.
n α = 0.10 α = 0 α = 0.10 α = 0.25
0 5.163 5.047 5.465 5.896 6.073 7.629 8.887
1 9.896 9.790 10.569 11.527 11.688 15.389 16.524
2 14.682 14.581 15.723 17.203 17.356 23.251 24.321
3 19.486 19.389 20.894 22.897 23.045 31.149 32.178
4 24.298 24.203 26.074 28.601 28.745 39.064 40.064
5 29.115 29.021 31.258 34.309 34.451 46.987 47.967
6 33.934 33.842 36.445 40.021 40.161 54.916 55.880
7 38.755 38.664 41.635 45.735 45.873 62.848 63.800
8 43.578 43.488 46.825 51.451 51.588 70.783 71.725
9 48.401 48.312 52.016 57.168 57.304 78.720 79.653
10 53.225 53.136 57.208 62.886 63.021 86.658 87.584
11 58.050 57.961 62.401 68.604 68.738 94.597 95.517
12 62.875 62.787 67.594 74.323 74.457 102.537 103.452
13 67.700 67.612 72.788 80.042 80.175 110.478 111.388
14 72.526 72.438 77.982 85.762 85.895 118.419 119.325
15 77.351 77.264 83.176 91.482 91.615 126.361 127.263
16 82.177 82.090 88.370 97.203 97.334 134.303 135.202
17 87.003 86.916 93.565 102.923 103.054 142.246 143.142
18 91.829 91.743 98.759 108.644 108.774 150.188 151.081
19 96.656 96.569 103.954 114.365 114.495 158.131 159.022
20 101.482 101.396 109.149 120.086 120.216 166.074 166.963
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, D.; Du, X.; Pan, Q.; Jing, J. Holographic p-Wave Superconductor with Excited States in 4D Einstein–Gauss–Bonnet Gravity. Universe 2023, 9, 104. https://doi.org/10.3390/universe9020104

AMA Style

Wang D, Du X, Pan Q, Jing J. Holographic p-Wave Superconductor with Excited States in 4D Einstein–Gauss–Bonnet Gravity. Universe. 2023; 9(2):104. https://doi.org/10.3390/universe9020104

Chicago/Turabian Style

Wang, Dong, Xinyi Du, Qiyuan Pan, and Jiliang Jing. 2023. "Holographic p-Wave Superconductor with Excited States in 4D Einstein–Gauss–Bonnet Gravity" Universe 9, no. 2: 104. https://doi.org/10.3390/universe9020104

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop