-
Let us begin by briefly reviewing the static BTZ-like spacetime in the Einstein-bumblebee gravity obtained in [9]. The action in this theoretical model with a negative cosmological constant
$ \Lambda = -\frac{1}{l^2} $ is given by [39−41]$ \begin{aligned}[b] {\cal{S}} =& \int d^3x\sqrt{-g}\Big[\frac{R-2\Lambda}{2\kappa}+\frac{\varrho}{2\kappa} B^{\mu}B^{\nu}R_{\mu\nu}-\frac{1}{4}B^{\mu\nu}B_{\mu\nu} \\& -V(B_\mu B^{\mu}\pm b^2)+{\cal{L}}_M\Big]. \end{aligned}$
(1) Here R describes the Ricci scalar, and κ is a constant that has a relationship with the three-dimensional Newton's constant G, given by
$ \kappa = 8\pi G $ .$ B_{\mu\nu} = \partial_{\mu}B_{\nu}-\partial_{\nu}B_{\mu} $ is the strength of bumblebee field$ B_\mu $ , and$ \varrho $ is a coupling constant with the dimension of$ M^{-1} $ in this model. In addition, we note that the potential V has a minimum at$ B^{\mu}B_{\mu}\pm b^2 = 0 $ , where b is a real positive constant and the signs$ \pm $ determine whether the field$ b_\mu $ is timelike or spacelike. This minimum yields a nonzero vacuum value$ \langle B_{\mu}\rangle = b_\mu $ with$ b_\mu b^\mu = \mp b^2 $ , which leads to the breaking of the$ U(1) $ symmetry. It should be noted that the nonzero vector background$ b_\mu $ leads to the violation of Lorentz symmetry [5, 39−41].From the action (1), one can get the static BTZ-like black hole solution in the Einstein-bumblebee gravity [9]
$ ds^{2} = -f(r)dt^{2}+\frac{1+s}{f(r)} dr^{2}+r^{2} d\varphi^{2}, $
(2) with
$ f(r) = \frac{r^2}{l^2}-M, $
(3) where M denotes the mass of the black hole, and
$ s = \xi b^2 $ represents the spontaneous breaking of Lorentz symmetry due to the Einstein-bumblebee vector field with the form$ b_{\mu} = (0,b\xi, 0, 0) $ . Considering the determinant of the metric$ g = -(1+s)r^2 $ , we find that the metric becomes degenerate at$ s = -1 $ . In the following analysis, we will consider the constraint$ s>-1 $ and compare with the results for the case of$ s = 0 $ , i.e., without the Lorentz symmetry breaking. The Hawking temperature of this BTZ-like black hole is$ T_H = \frac{r_{H}}{8\pi l^2\sqrt{1+s}}, $
(4) with the horizon
$ r_{H} = \sqrt{M}l $ . For convenience, we will scale$ l = 1 $ in the numerical calculation.In the static BTZ-like spacetime, the massive scalar field evolves according to
$ \left(\nabla_{\mu}\nabla^{\mu}-\mu^2\right)\Psi = 0, $
(5) where μ is the mass of the scalar field. Assuming that the massive scalar perturbation Ψ has the form
$ \Psi = e^{-i\omega t+im\varphi}\psi(r), $
(6) we can obtain the radial equation
$ \begin{aligned}[b] & \frac{d^2\psi(r)}{dr^2}+\bigg[\frac{1}{r}+\frac{f'(r)}{f(r)}\bigg]\frac{d\psi(r)}{dr}\\&+(1+s) \bigg[\frac{\omega^2}{f(r)^2} -\frac{m^2}{r^2f(r)}-\frac{\mu^2}{f(r)}\bigg]\psi(r) = 0, \end{aligned}$
(7) which can be solved in terms of the hypergeometric functions [42]. We introduce a variable
$ z = \dfrac{r^2-r_H^2}{r^2} $ and then the radial equation (7) can be rewritten as$ \begin{aligned}[b] &z(1-z)\frac{d^2\psi(z)}{dz^2}+(1-z)\frac{d\psi(z)}{dz}\\&+\left(\frac{A}{z}-\frac{B}{1-z}-C\right)\psi(z) = 0, \end{aligned}$
(8) where A, B and C have forms
$ A = \frac{\omega^2l^4(1+s)}{4r_H^2},\quad B = \frac{\mu^2l^2(1+s)}{4},\quad C = \frac{m^2l^2(1+s)}{4r_H^2}. $
(9) Rewriting the radial function
$ \psi(z) $ as the form$ \psi(z) = z^{i\tfrac{l^2\sqrt{1+s}}{2r_H}\omega} (1-z)^{\tfrac{1+\sqrt{1+\mu^2l^2(1+s)}}{2}}F(z), $
(10) we find that the function
$ F(z) $ satisfies the standard hypergeometric equation$ z(1-z)\frac{d^2F(z)}{dz^2}+[c-(1+a+b)z]\frac{dF(z)}{dz}+ab F(z) = 0, $
(11) with
$ a = \frac{1+\sqrt{1+\mu^2l^2(1+s)}}{2}-i\frac{l\sqrt{1+s}}{2r_H}(\omega l-m), $
(12) $ b = \frac{1+\sqrt{1+\mu^2l^2(1+s)}}{2}-i\frac{l\sqrt{1+s}}{2r_H}(\omega l+m), $
(13) $ c = 1-i\frac{l^2\sqrt{1+s}}{r_H}\omega, $
(14) where the scalar field mass must obey
$ \mu^2\geqslant\mu^2_{BF} $ with the Breitenlohner-Freedman (BF) bound$ \mu_{BF}^2 = -\dfrac{1}{l^2(1+s)} $ . Thus, the general solution of the radial equation (8) can be given by a linear combination of hypergeometric functions F, i.e.,$ \begin{aligned}[b] \psi(z) = &z^{i\tfrac{l^2\sqrt{1+s}}{2r_H}\omega} (1-z)^{\tfrac{1+\sqrt{1+\mu^2l^2(1+s)}}{2}} \Bigg[A F(a,b,c;z)\\&+B z^{i\tfrac{\omega l^2 \sqrt{1+s}}{r_H}}F(a-c+1,b-c+1,2-c;z)\Bigg], \end{aligned}$
(15) where A and B are two constants of integration. By imposing the ingoing-wave condition at the black hole horizon, we obtain
$ B = 0 $ and then the solution (15) has a more simple form$ \psi(z) = A z^{i\tfrac{l^2\sqrt{1+s}}{2r_H}\omega} (1-z)^{\tfrac{1+\sqrt{1+\mu^2l^2(1+s)}}{2}} F(a,b,c;z). $
(16) To probe the asymptotic behavior of
$ \psi(z) $ at infinity as$ r\rightarrow\infty $ (i.e.,$ z\rightarrow 1 $ ), we can expand the hypergeometric functions at$ 1-z $ [42], i.e.,$ \begin{aligned}[b] {l} F(a,b,c;z) =& \dfrac{\Gamma(c)\Gamma(c-a-b)}{\Gamma(c-a)\Gamma(c-b)}F(a, b, a+b-c+1; 1-z)\\& +(1-z)^{c-a-b}\dfrac{\Gamma(c)\Gamma(a+b-c)}{\Gamma(a)\Gamma(b)}\\&F(c-a, c-b, c-a-b+1; 1-z). \end{aligned} $
(17) Thus, the general solution (16) can be expressed as
$ \psi(r) = A_{\rm{I}}\psi^{(D)}(r)+A_{\rm{II}}\psi^{(N)}(r), $
(18) where the constants
$ A_{\rm{I}} $ and$ A_{\rm{II}} $ are$ A_{\rm{I}} = A\dfrac{\Gamma(c)\Gamma(c-a-b)}{\Gamma(c-a)\Gamma(c-b)},\quad\quad\quad A_{\rm{II}} = A\dfrac{\Gamma(c)\Gamma(a+b-c)}{\Gamma(a)\Gamma(b)}. $
(19) The functions
$ \psi^{(D)}(r) $ and$ \psi^{(N)}(r) $ respectively denote the forms of the solution (16) satisfied the Dirichlet boundary condition and Neumann boundary condition at infinity [4]. The forms of$ \psi^{(D)}(r) $ and$ \psi^{(N)}(r) $ are$ \begin{aligned}[b]\psi^{(D)}(r) =& \left(1-\dfrac{r_H^2}{r^2}\right)^{i\tfrac{l^2\sqrt{1+s}}{2r_H}\omega}\left(\dfrac{r_H}{r}\right)^{1+\sqrt{1+\mu^2l^2(1+s)}}\\& F\left(a, b, a+b-c+1; \dfrac{r_H^2}{r^2}\right),\end{aligned} $
(20) $ \begin{aligned}[b]\psi^{(N)}(r) =& \left(1-\dfrac{r_H^2}{r^2}\right)^{i\tfrac{l^2\sqrt{1+s}}{2r_H}\omega}\left(\dfrac{r_H}{r}\right)^{1-\sqrt{1+\mu^2l^2(1+s)}} \\ &F\left(c-a, c-b, c-a-b+1; \dfrac{r_H^2}{r^2}\right).\end{aligned} $
(21) With the Dirichlet boundary condition, we find
$ a = -n $ or$ b = -n $ , and then obtain the quasinormal frequencies for the massive scalar perturbation$ \omega^{(D)} = \pm\frac{m}{l}-i\frac{2r_H}{l^2\sqrt{1+s}}\bigg[n+\frac{1}{2}+\frac{1}{2}\sqrt{1+\mu^2l^2(1+s)}\bigg]. $
(22) The signs
$ + $ and$ - $ respectively denote the left-moving and the right-moving modes. As the Lorentz symmetry breaking parameter s increases, the absolute value of the imaginary part of$ \omega^{(D)} $ decreases for different overtone number n [10], which is also shown in Fig. 1(a). Moreover, we also note that all the imaginary parts of$ \omega^{(D)} $ are negative, which means that the black hole is stable under the scalar perturbation with the Dirichlet boundary condition. And the corresponding wave-function$ \Psi^{(D)} $ at infinity has the asymptotic behaviorFigure 1. (color online)The relationship between the imaginary part of the QNM frequencies and the Lorentz symmetry breaking parameter s with different overtone numbers n in different boundary conditions, where we have taken
$ r_H = 1 $ and$ \mu^2 = -0.75 $ .$ \begin{aligned}[b] \Psi^{(D)}|_{r\rightarrow\infty} =&A_{\rm{I}} \left(\dfrac{r_H}{r}\right)^{1+\sqrt{1+\mu^2l^2(1+s)}} \\&\left[1+{\cal{O}}(1/r^2)\right]e^{-i\omega^{(D)}t+im\varphi}. \end{aligned}$
(23) Similarly, for the Neumann boundary condition, one has
$ c-a = -n $ or$ c-b = -n $ , and then the quasinormal frequencies for the scalar field are$ \omega^{(N)} = \pm \frac{m}{l}-i\frac{2r_H}{l^2\sqrt{1+s}}\bigg[n+\frac{1}{2}-\frac{1}{2}\sqrt{1+\mu^2l^2(1+s)}\bigg]. $
(24) As shown in Fig. 1(b), similar to the case of the Dirichlet boundary condition, the imaginary part of
$ \omega^{(N)} $ increases for different overtone numbers n with the increase of the parameter s under the Neumann boundary condition, except the fundamental mode$ n = 0 $ , i.e., the scalar field without nodes, where the imaginary part of$ \omega^{(N)} $ decreases slowly with the increase of s. It should be noted that the asymptotic behavior of the wave-function$ \Psi^{(N)} $ at infinity becomes$ \begin{aligned}[b] \Psi^{(N)}|_{r\rightarrow\infty} =& A_{\rm{II}} \left(\frac{r_H}{r}\right)^{1-\sqrt{1+\mu^2l^2(1+s)}} \\&\left[1+{\cal{O}}(1/r^2)\right]e^{-i\omega^{(N)}t+im\varphi}.\end{aligned} $
(25) -
For the d-dimensional spacetime
$ (M,g_{\mu\nu}) $ with the symmetry under the conformal transformation$ g_{\mu\nu}\rightarrow g'_{\mu\nu} = e^{2Q}g_{\mu\nu} $ and$ Q = \frac{1}{d}\nabla_\sigma\zeta^\sigma $ , there exists a conformal Killing vector (CKV)$ \zeta^\mu $ satisfied the conformal Killing equation$ \nabla_{\mu}\zeta_\nu+\nabla_\nu\zeta_\mu = \frac{2}{d}\nabla_\sigma\zeta^\sigma g_{\mu\nu}. $
(26) Since the closed conformal Killing vector (CCKV)
$ \zeta^{\mu} $ meets an extra condition$ \nabla_{[\mu}\zeta_{\nu]} = 0, $
(27) it satisfies the reduced Killing equation
$ \nabla_{\mu}\zeta_\nu = \frac{1}{d}\nabla_\sigma\zeta^\sigma g_{\mu\nu}. $
(28) Recent investigations show a CCKV
$ \zeta^{\mu} $ is an eigenvector of Ricci tensor with a constant eigenvalue χ [1, 2], i.e.,$ R^\mu_{\enspace\nu}\zeta^{\nu} = \chi(d-1)\zeta^{\mu}. $
(29) One can introduce a one-parameter family of operators
$ D_{k}: = {\cal{L}}_{\zeta}-\frac{k}{d}\nabla_{\mu}\zeta^\mu, $
(30) where
$ {\cal{L}}_{\zeta} $ denotes the Lie derivative with respect to$ \zeta^\mu $ and k is a parameter. The commutation relation between the operator$ D_{k} $ and the d'Alembertian is given by$ \begin{aligned}[b] [\nabla_\mu \nabla^\mu,D_k] =& \chi(2k+d-2)D_k\\&+2Q(\nabla_\mu \nabla^\mu+\chi k(k+d-1)).\end{aligned} $
(31) Acting the commutator on a scalar field ϕ, one can obtain
$ \begin{aligned}[b]&D_{k-2}\left[\nabla_\mu \nabla^\mu+\chi k(k+d-1)\right]\phi\\ =\;& \left[\nabla_\mu \nabla^\mu+\chi (k-1)(k+d-2)\right]D_k\phi. \end{aligned}$
(32) Obviously, the operator
$ D_k $ maps a scalar field ϕ with mass squared$ \mu^2 = -\chi k(k+d-1) $ to another scalar field$ D_k\phi $ with a new mass squared$ \mu'^2 = -\chi(k-1)(k+d-2) $ , so$ D_k $ can be regarded as a mass ladder operator, which shifts the parameter k to$ k-1 $ in the mass squared term. Thus, such kind of operators is called as the mass lowering operator$ D_{k_-} $ . Similarly, one can introduce the mass raising operator$ D_{k_+} $ , which shifts the parameter k to$ k+1 $ in the mass squared term.For the BTZ-like spacetime in the Einstein-bumblebee gravity (2), we obtain four independent CCKVs
$ \zeta_{0} = e^{\tfrac{r_H}{l^2\sqrt{1+s}}t}\left(\frac{1}{\sqrt{r^2-{r_H}^2}}\partial_t-\frac{r\sqrt{r^2-{r_H}^2}}{r_H l^2 \sqrt{1+s}}\partial_r\right), $
(33) $ \zeta_{1} = e^{-\tfrac{r_H}{l^2\sqrt{1+s}}t}\left(\frac{1}{\sqrt{r^2-{r_H}^2}}\partial_t+\frac{r\sqrt{r^2-{r_H}^2}}{r_H l^2 \sqrt{1+s}}\partial_r\right), $
(34) $ \zeta_{2} = e^{\tfrac{r_H}{l\sqrt{1+s}}\varphi}\left(\frac{r^2-{r_H}^2}{r_H l\sqrt{1+s}}\partial_r+\frac{1}{r}\partial_\varphi\right), $
(35) $ \zeta_{3} = e^{-\tfrac{r_H}{l\sqrt{1+s}}\varphi}\left(-\frac{r^2-{r_H}^2}{r_H l\sqrt{1+s}}\partial_r+\frac{1}{r}\partial_\varphi\right), $
(36) and the corresponding four mass ladder operators
$ \begin{aligned}[b] D_{0,k} =\;& e^{\tfrac{r_H}{l^2\sqrt{1+s}}t}\Bigg(\frac{1}{\sqrt{r^2-{r_H}^2}}\partial_t-\frac{r\sqrt{r^2-{r_H}^2}}{r_Hl^2 \sqrt{1+s} }\partial_r\\&+k\frac{\sqrt{r^2-{r_H}^2}}{r_H l^2 \sqrt{1+s}}\Bigg), \end{aligned}$
(37) $ \begin{aligned}[b]D_{1,k} =& e^{-\tfrac{r_H}{l^2\sqrt{1+s}}t}\Bigg(\frac{1}{\sqrt{r^2-{r_H}^2}}\partial_t\\&+\frac{r\sqrt{r^2-{r_H}^2}}{r_H l^2\sqrt{1+s} }\partial_r-k\frac{\sqrt{r^2-{r_H}^2}}{r_H l^2 \sqrt{1+s}}\Bigg),\end{aligned} $
(38) $ D_{2,k} = e^{\tfrac{r_H}{l\sqrt{1+s}}\varphi}\Bigg(\frac{r^2-{r_H}^2}{r_H l\sqrt{1+s}}\partial_r+ \frac{1}{r}\partial_\varphi-k\frac{r}{r_H l\sqrt{1+s}}\Bigg), $
(39) $ D_{3,k} = e^{-\tfrac{r_H}{l\sqrt{1+s}}\varphi}\left(-\frac{r^2-{r_H}^2}{r_H l\sqrt{1+s}}\partial_r+\frac{1}{r}\partial_\varphi+k\frac{r}{r_H l\sqrt{1+s}}\right). $
(40) Obviously, the mass ladder operators depend on the Lorentz symmetry breaking parameter s in the Einstein-bumblebee gravity. In an AdS
$ _d $ -like spacetime, the BF-bound$ \mu^2_{BF} $ for a scalar field can be expressed as [1, 4]$ \mu^2_{BF} = \frac{(d-1)^2}{4}\chi,\quad \chi<0. $
(41) For the BTZ-like black hole spacetime in the Einstein-bumblebee gravity (2), we find that χ is exactly equal to the BF-bound
$ \mu^2_{BF} = -\dfrac{1}{l^2(1+s)} $ , and then the commutation relation (31) with the d'Alembertian becomes$ \left[\nabla_\mu\nabla^\mu, D_{i,k}\right] = -\frac{2k+1}{l^2(1+s)}D_{i,k} +\frac{2}{3}(\nabla_\mu \zeta_i^\mu)\left[\nabla_\mu\nabla^\mu -\frac{k (k+2)}{l^2(1+s)}\right]. $
(42) This means that Eq. (32) can be further simplified as
$ D_{i,k-2}\left[\nabla_\mu\nabla^\mu-\frac{k(k+2)}{l^2(1+s)}\right]\Psi = \left[\nabla_\mu\nabla^\mu -\frac{(k-1)(k+1)}{l^2(1+s)}\right]D_{i,k}\Psi. $
(43) As in [1, 2],
$ D_{i,k} $ maps a solution Ψ to the Klein-Gordon equation with the mass squared$ \mu^2 = \dfrac{k (k+2)}{l^2(1+s)} $ into another solution$ D_{i,k}\Psi $ with the mass squared$ \mu^2 = \dfrac{(k-1)(k+1)}{l^2(1+s)} $ , namely, shifts k to$ k-1 $ in the mass squared term.To require that the characteristic exponents with the mass
$ \Delta_{\pm} = \frac{1}{2}[1\pm\sqrt{1+\mu^{2}l^{2}(1+s)}] $ in (23) and (25) are real, we here only focus on the mass squared$ \mu^2\geqslant\mu^2_{BF} $ . As in [4], one can introduce a parameter ν, which obeys$ \mu^2 l^2(1+s) = \nu(\nu+2). $
(44) This means that
$ \nu = -1\pm\sqrt{1+\mu^2l^2(1+s)} $ . And the relationship between$ \mu^2l^2(1+s) $ and ν for different Lorentz symmetry breaking parameters s is presented in Fig. 2, which shows that the function$ \mu^2l^2(1+s) $ is symmetric about$ \nu = -1 $ . Thus, without loss of generality, we only focus on the right part of the parabola where$ \nu = -1+\sqrt{1+\mu^2l^2(1+s)} $ with$ \nu\in[-1,\infty) $ . For a given parameter ν, there are two values of k satisfied$ k(k+2) = \nu(\nu+2) $ , i.e.,Figure 2. (color online)The term
$ \mu^2l^2(1+s) $ as a function of ν for different Lorentz symmetry breaking parameters s.$ k_+ = -2-\nu,\;\; k_- = \nu. $
(45) The corresponding mass ladder operators
$ D_{i,k_+} $ and$ D_{i,k_-} $ respectively shift ν to$ \nu+1 $ and$ \nu-1 $ .Now we are in a position to discuss the mapped solutions resulting from the mass ladder operators acting on the original solution Ψ. Since the operators
$ D_{2,k\pm} $ and$ D_{3,k\pm} $ are non-global operators, they are not globally smooth in the φ direction. As a result, they change the quantum number m of the scalar field. However, the operators$ D_{0,k\pm} $ and$ D_{1,k\pm} $ can directly affect the frequencies of the scalar perturbation. Therefore, as in [4], we only focus on the effects of$ D_{0,k\pm} $ and$ D_{1,k\pm} $ on the general solution of the scalar field under different boundary conditions.The general solution (16) of the scalar field can be rewritten as
$ \Psi = A\left(1-\frac{r^2_H}{r^2}\right)^{-i\tfrac{l^2\sqrt{1+s}}{2r_H}\omega} \left(\frac{r_H}{r}\right)^{2+\nu} F\left(a,b,c;1-\frac{r^2_H}{r^2}\right)e^{-i\omega t+im\varphi}. $
(46) The asymptotic behavior of the general solution near the horizon is
$ \begin{aligned}[b]\Psi|_{r\simeq r_H} =& 2^{-i\tfrac{l^2\sqrt{1+s}}{2r_H}\omega}A\left(\frac{r-r_H}{r}\right)^{-i\tfrac{l^2\sqrt{1+s}}{2r_H}\omega}\\ &\left[1+{\cal{O}}(r-r_H)\right]e^{-i\omega t+im\varphi}. \end{aligned}$
(47) Under the Dirichlet or Neumann boundary conditions at infinity, the above general solution has the asymptotic behavior
$ \Psi|_{r\simeq \infty} = A_{\rm{I}}\left(\frac{r_H}{r}\right)^{2+\nu} \left[1+{\cal{O}}(1/r^2)\right]e^{-i\omega t+im\varphi} \; \; \; \; { ({\rm{DBC}})}, $
(48) $ \Psi|_{r\simeq \infty} = A_{{\rm{II}}}\left(\frac{r_H}{r}\right)^{-\nu}\left[1+{\cal{O}}(1/r^2)\right]e^{-i\omega t+im\varphi}\; \; \; \; {\rm{(NBC)}}. $
(49) Acting the mass ladder operators
$ D_{0,k\pm} $ and$ D_{1,k\pm} $ on the exact solution (46), one can obtain the asymptotic behaviors of the mapped solutions near the horizon$ \begin{aligned}[b] D_{0,k\pm}\Psi|_{r\simeq r_H} =&C_{0,k\pm}\left(\frac{r-r_H}{r}\right)^{-i\tfrac{l^2\sqrt{1+s}}{2r_H}\left(\omega+i\tfrac{r_H}{l^2\sqrt{1+s}}\right)}\\ &\left[1+{\cal{O}}(r-r_H)\right]e^{-i\left(\omega+i\tfrac{r_H}{l^2\sqrt{1+s}}\right)t+im\varphi}, \end{aligned}$
(50) $ \begin{aligned}[b] D_{1,k\pm}\Psi|_{r\simeq r_H} =& C_{1,k\pm}\left(\frac{r-r_H}{r}\right)^{-i\tfrac{l^2\sqrt{1+s}}{2r_H}\left(\omega-i\tfrac{r_H}{l^2\sqrt{1+s}}\right)}\\&\left[1+{\cal{O}}(r-r_H)\right]e^{-i\left(\omega-i\tfrac{r_H}{l^2\sqrt{1+s}}\right)t+im\varphi},\end{aligned} $
(51) where
$ C_{0,k\pm} $ and$ C_{1,k\pm} $ are two constants$ \begin{aligned} C_{0,k\pm} = &\dfrac{iA}{r_Hl^4(1+s)\left(\omega +i\dfrac{r_H}{l^2\sqrt{1+s}}\right)}2^{-1-i\tfrac{l^2\sqrt{1+s}}{2r_H}\left(\omega+i\tfrac{r_H}{l^2\sqrt{1+s}}\right)} \\ &\times\left[r_H^2(k_{\pm}(2+k_{\pm})-\nu(2+\nu))+l^4(1+s)\left(\omega+\tfrac{m}{l}-ik_{\pm}\frac{r_H}{l^2\sqrt{1+s}}\right)\left(\omega-\frac{m}{l}- ik_{\pm}\frac{r_H}{l^2\sqrt{1+s}}\right)\right], \end{aligned} $ (52) $ C_{1,k\pm} = -2^{1-i\tfrac{l^2\sqrt{1+s}}{2r_H}\left(\omega-i\tfrac{r_H}{l^2\sqrt{1+s}}\right)}\frac{i\omega A}{r_H}. $
(53) Under the Dirichlet boundary condition, the asymptotic behaviors of the mapped solutions acted by the mass ladder operators
$ D_{0,k\pm} $ and$ D_{1,k\pm} $ at infinity can be expressed as$ \begin{aligned}[b]D_{0,k+}\Psi^{(D)}|_{r\simeq\infty} =& C_{0,k+}^{(D)}\left(\frac{r_H}{r}\right)^{3+\nu} \\ &\left[1+{\cal{O}}(1/r^2)\right]e^{-i\left(\omega^{(D)}+i\tfrac{r_H}{l^2\sqrt{1+s}} \right)t+im\varphi}, \end{aligned}$
(54) $ \begin{aligned}[b] D_{1,k+}\Psi^{(D)}|_{r\simeq\infty} =& C_{1,k+}^{(D)}\left(\frac{r_H}{r}\right)^{3+\nu}\\ &\left[1+{\cal{O}} (1/r^2)\right]e^{-i\left(\omega^{(D)}-i\tfrac{r_H}{l^2\sqrt{1+s}}\right)t+im\varphi},\end{aligned} $
(55) $ \begin{aligned}[b]D_{0,k-}\Psi^{(D)}|_{r\simeq\infty} =& C_{0,k-}^{(D)}\left(\frac{r_H}{r}\right)^{1+\nu}\\&\left[1+{\cal{O}}(1/r^2)\right]e^{-i\left(\omega^{(D)}+i\tfrac{r_H}{l^2\sqrt{1+s}}\right)t+im\varphi}, \end{aligned}$
(56) $ \begin{aligned}[b] D_{1,k-}\Psi^{(D)}|_{r\simeq\infty} =& C_{1,k-}^{(D)}\left(\frac{r_H}{r}\right)^{1+\nu}\\&\left[1+{\cal{O}}(1/r^2)\right]e^{-i\left(\omega^{(D)}-i\tfrac{r_H}{l^2\sqrt{1+s}}\right)t+im\varphi}, \end{aligned}$
(57) where
$ \begin{aligned}[b] C^{(D)}_{0,k+} =& -\frac{A_{\rm{I}}}{(2+\nu)}\frac{l^2\sqrt{1+s}}{2r_H^2}\left[\omega^{(D)}-\frac{m}{l}+i\frac{r_H}{l^2\sqrt{1+s}}(2+\nu)\right]\\&\left[\omega^{(D)}+\frac{m}{l}+i\frac{r_H}{l^2\sqrt{1+s}}(2+\nu)\right], \end{aligned}$
(58) $ \begin{aligned}[b]C^{(D)}_{1,k+} =& \frac{A_{\rm{I}}}{(2+\nu)}\frac{l^2\sqrt{1+s}}{2r_H^2}\left[\omega^{(D)}-\frac{m}{l}-i\frac{r_H}{l^2\sqrt{1+s}}(2+\nu)\right]\\&\left[\omega^{(D)}+\frac{m}{l}-i\frac{r_H}{l^2\sqrt{1+s}}(2+\nu)\right], \end{aligned}$
(59) $ C^{(D)}_{0,k-} = \frac{2(1+\nu)}{l^2\sqrt{1+s}}A_{\rm{I}},\quad\quad C^{(D)}_{1,k-} = -\frac{2(1+\nu)}{l^2\sqrt{1+s}}A_{\rm{I}}. $
(60) Combining with Eqs. (50), (51) and Eqs. (54)-(57), one can find that
$ D_{0,k\pm} $ and$ D_{1,k\pm} $ respectively shift ω to$ \omega+i\dfrac{r_H}{l^2\sqrt{1+s}} $ and$ \omega-i\dfrac{r_H}{l^2\sqrt{1+s}} $ , while they keep m unchanged. Thus, the quasinormal frequencies of the mapped solutions of the scalar field by the mass ladder operators$ D_{0,k\pm} $ and$ D_{1,k\pm} $ are$ \omega^{(D)}_{0,1} = \omega^{(D)}\pm i\frac{r_H}{l^2\sqrt{1+s}}, $
(61) with
$ \omega^{(D)} = \pm\frac{m}{l}-i\frac{r_H}{l^2\sqrt{1+s}}(2n+\nu+2). $
(62) In particular, when the mass ladder operator
$ D_{0,k+} $ acts on the fundamental mode, the mapped mode vanishes rather than becomes a “negative overtone mode” due to the restriction imposed by (58) as in [4]. Fig. 3 shows the change of quasinormal frequencies of the mapped modes$ D_{0,k\pm}\Psi^{(D)} $ and$ D_{1,k\pm}\Psi^{(D)} $ with the parameter s, which is similar to that of the original modes$ \Psi^{(D)} $ . Notice that, for all the modes with$ \mu^2 = -0.75 $ and$ l = 1 $ , there exists a threshold value$ s_{c} = \frac{1}{3} $ , corresponding to a vertical dashed line in each panel, which means that all the modes stop increasing their imaginary parts when they reach this vertical line. As a matter of fact, the threshold value$ s_{c} = -\left(1+\dfrac{1}{\mu^2 l^2}\right) $ only depends on the mass of the scalar field, and increases with the increase of the mass squared term$ \mu^2 $ , just as shown in Fig. 4 with$ \mu^2\leqslant0 $ .Figure 3. (color online)The relationship between the imaginary part
$ {\rm{Im}} $ $ [\omega^{(D)}] $ and the Lorentz symmetry breaking parameter s respectively for$ \Psi^{(D)} $ ,$ D_{0,k\pm}\Psi^{(D)} $ and$ D_{1,k\pm}\Psi^{(D)} $ with different overtone numbers n, where we have taken$ r_H = 1 $ and$ \mu^2 = -0.75 $ .Figure 4. (color online) The relationship between the imaginary part
$ {\rm{Im}} $ $ [\omega^{(D)}] $ and the Lorentz symmetry breaking parameter s respectively for$ \Psi^{(D)} $ ,$ D_{0,k_-}\Psi^{(D)} $ and$ D_{1,k\pm}\Psi^{(D)} $ with different masses$ \mu^2 $ , where we have taken$ n = 0 $ and$ r_H = 1 $ .Under the Neumann boundary condition, we can obtain the asymptotic behaviors of the mapped solutions by mass ladder operators
$ D_{0,k\pm} $ and$ D_{0,k\pm} $ at infinity$ \begin{aligned}[b] D_{0,k+}\Psi^{(N)}|_{r\simeq\infty} =& C_{0,k+}^{(N)}\left(\frac{r_H}{r}\right)^{-1-\nu}\\&\left[1+{\cal{O}}(1/r^2)\right]e^{-i\left(\omega^{(N)}+i\tfrac{r_H}{l^2\sqrt{1+s}}\right)t+im\varphi}, \end{aligned}$
(63) $ \begin{aligned}[b] D_{1,k+}\Psi^{(N)}|_{r\simeq\infty} =& C_{1,k+}^{(N)}\left(\frac{r_H}{r}\right)^{-1-\nu}\\&\left[1+{\cal{O}}(1/r^2)\right]e^{-i\left(\omega^{(N)}-i\tfrac{r_H}{l^2\sqrt{1+s}}\right)t+im\varphi}, \end{aligned}$
(64) $ \begin{aligned}[b] D_{0,k-}\Psi^{(N)}|_{r\simeq\infty} =& C_{0,k-}^{(N)}\left(\frac{r_H}{r}\right)^{1-\nu}\\&\left[1+{\cal{O}}(1/r^2)\right]e^{-i\left(\omega^{(N)}+i\tfrac{r_H}{l^2\sqrt{1+s}}\right)t+im\varphi},\end{aligned} $
(65) $\begin{aligned}[b] D_{1,k-}\Psi^{(N)}|_{r\simeq\infty} =& C_{1,k-}^{(N)}\left(\frac{r_H}{r}\right)^{1-\nu}\\&\left[1+{\cal{O}}(1/r^2)\right]e^{-i\left(\omega^{(N)}-i\tfrac{r_H}{l^2\sqrt{1+s}}\right)t+im\varphi},\end{aligned} $
(66) where
$ C^{(N)}_{0,k+} = -\frac{2(1+\nu)}{l^2\sqrt{1+s}}A_{\rm{II}},\quad\quad C^{(N)}_{1,k+} = \frac{2(1+\nu)}{l^2\sqrt{1+s}}A_{\rm{II}}, $
(67) $ \begin{aligned}[b]C^{(N)}_{0,k-} =& \frac{A_{\rm{II}}}{\nu}\frac{l^2\sqrt{1+s}}{2r_H^2}\left[\omega^{(N)} -\frac{m}{l}-i\frac{r_H}{l^2\sqrt{1+s}}\nu\right]\\&\left[\omega^{(N)} +\frac{m}{l}-i\frac{r_H}{l^2\sqrt{1+s}}\nu\right],\end{aligned} $
(68) $ \begin{aligned}[b]C^{(N)}_{1,k-} =& -\frac{A_{\rm{II}}}{\nu}\frac{l^2\sqrt{1+s}}{2r_H^2}\left[\omega^{(N)} -\frac{m}{l}+i\frac{r_H}{l^2\sqrt{1+s}}\nu\right]\\ &\left[\omega^{(N)} +\frac{m}{l}+i\frac{r_H}{l^2\sqrt{1+s}}\nu\right]. \end{aligned}$
(69) Combining with Eqs. (50), (51) and Eqs. (63)-(66), one can find that
$ D_{0,k\pm} $ and$ D_{1,k\pm} $ respectively shift ω to$ \omega+i\dfrac{r_H}{l^2\sqrt{1+s}} $ and$ \omega-i\dfrac{r_H}{l^2\sqrt{1+s}} $ , while they keep m unchanged. Thus, the quasinormal frequencies of the mapped solutions of the scalar field by the mass ladder operators$ D_{0,k\pm} $ and$ D_{1,k\pm} $ are$ \omega^{(N)}_{0,1} = \omega^{(N)}\pm i\frac{r_H}{l^2\sqrt{1+s}}, $
(70) with
$ \omega^{(N)} = \pm\frac{m}{l}-i\frac{r_H}{l^2\sqrt{1+s}}(2n-\nu). $
(71) Similarly, under the Neumann boundary condition, there is also no “negative overtones” generated from the fundamental modes for
$ D_{0,k_-} $ as in the case with the Dirichlet boundary condition. For the fundamental mode with$ \mu^2<0 $ , the mass ladder operators do not change the BF-bound$ \mu^2_{BF} $ . Furthermore, as in the Dirichlet boundary condition, we find that there also exists the same threshold value$ s_c = \frac{1}{3} $ for the mapped modes and the original modes. The threshold value$ s_c $ for all the mapped modes with the same mass squared term is a constant. Thus, the mass ladder operators keep the threshold value$ s_c $ for the mapped modes.Moreover, as shown in Fig. 6, the threshold value
$ s_c $ increases with the increase of$ \mu^2 $ . It should be noted that the imaginary parts of the quasinormal frequencies of the mapped modes caused by$ D_{0,k_+}\Psi^{(N)} $ are all positive, which indicates that all the fundamental modes of$ D_{0,k_+}\Psi^{(N)} $ are unstable.Figure 6. (color online)The relationship between the imaginary part
$ {\rm{Im}} $ $ [\omega^{(N)}] $ and the Lorentz symmetry breaking parameter s respectively for$ \Psi^{(N)} $ ,$ D_{0,k_+}\Psi^{(N)} $ and$ D_{1,k\pm}\Psi^{(N)} $ with different masses$ \mu^2 $ , where we have taken$ n = 0 $ and$ r_H = 1 $ .Figure 5. (color online)The relationship between the imaginary part
${\rm{Im}}$ $[\omega^{(N)}]$ and the Lorentz symmetry breaking parameter s respectively for$\Psi^{(N)}$ ,$D_{0,k\pm}\Psi^{(N)}$ and$D_{1,k\pm}\Psi^{(N)}$ with different overtone numbers n, where we have taken$r_H=1$ and$\mu^2=-0.75$ .
Mass ladder operators and quasinormal modes of the static BTZ-like black hole in the Einstein-bumblebee gravity
- Received Date: 2024-07-26
- Available Online: 2025-01-01
Abstract: We investigate the mass ladder operators for the static BTZ-like black hole in the Einstein-bumblebee gravity, and probe the quasinormal frequencies of the mapped modes by the mass ladder operators for a scalar perturbation under the Dirichlet and Neumann boundary conditions. It is found that the mass ladder operators depend on the Lorentz symmetry breaking parameter, and the imaginary parts of the frequencies shifted by the mass ladder operators increase with the increase of the Lorentz symmetry breaking parameter under two boundary conditions. It should be noted that, under the Neumann boundary condition, the mapped modes caused by the mass ladder operator