Department of Applied Physics, Nagoya University, Nagoya 464-8603, Japan
Received September 13, 2020; Accepted November 27, 2020; Published February 3, 2021
The Nambu–Goldstone (NG) mode in a Bose–Einstein condensate (BEC) transmits a potential barrier with probability 1 in the zero-energy limit, which is known as the anomalous tunneling. In this paper, we investigate the tunneling properties of quasiparticles in a dynamically unstable BEC. We prepare a multicomponent BEC (binary and spin-1 BECs) in a dynamically unstable state and solve the tunneling problem of the spin-wave excitation from the condensate. We find that the perfect transmission occurs even when the BEC is dynamically unstable if the spin-wave mode is the NG mode. Here, the mode that exhibits the perfect transmission is the dynamically unstable spin-wave, which is a pure-imaginary-eigenvalue solution of the Bogoliubov–de Gennes equation. Hence, we should take the zero-energy limit not along the real axis but along the imaginary axis. We also demonstrate the existence of the perfect reflection of a dynamically unstable mode at the point where the imaginary part of the eigenvalue takes its maximum. In this case, the incident and reflected waves destructively interfere, and the amplitude of the quasiparticle wave function is strongly suppressed. We numerically confirm that the perfect reflection is a generic nature of dynamically unstable modes and not related to the NG mode. A possible experimental scheme to observe the tunneling properties is discussed on the basis of real-time dynamics obtained by solving the time-dependent Gross–Pitaevskii equation.
The concept of elementary excitations plays a key role in understanding the fundamental properties of quantum many-body systems, ranging from ground-state properties to non-equilibrium transport phenomena.1) In particular, in a system with spontaneous symmetry breaking, the nature of the system at a low energy is dominated by the gapless Nambu–Goldstone (NG) mode associated with the broken symmetry. Examples include phonons in a Bose–Einstein condensate (BEC) and magnons in a ferromagnet, which are associated with the breaking of the U(1) gauge symmetry and the SO(3) spin-rotational symmetry, respectively.
Anomalous tunneling is one of the salient features of elementary excitations in a BEC, which is the phenomenon that a quasiparticle in a BEC transmits a potential barrier with probability 1 in the low-energy limit2,3) (see Fig. 1). This behavior is the opposite of the tunneling problem of non-interacting particles described by the Schrödinger equation, where the transmission probability becomes zero in the low-energy limit. Thus far, the anomalous tunneling has been studied in BECs with supercurrent flow,4–6) at finite temperatures,7) and with spin degrees of freedom.8–10) The relations of the anomalous tunneling with Josephson current,11,12) impedance matching,13) and scattering at magnetic domain walls and impurities14,15) were also discussed. The above works revealed that the anomalous tunneling occurs in the zero-energy limit of the NG mode: An NG boson in the zero-energy limit is identical to condensed particles; hence, its wave function extends over both sides of the barrier potential, resulting in the perfect transmission. Apart from the zero-energy limit of the NG mode, the Fano resonance between the NG and Higgs modes and the perfect transmission of the Higgs mode via antibound states have recently been predicted.16,17)
Figure 1. (Color online) Schematic of the tunneling problem of a Bogoliubov quasiparticle in a BEC under a potential barrier \(U(\boldsymbol{x})\). A quasiparticle with energy E injected from the left is reflected to the left or transmitted to the right. Whereas a non-interacting particle is reflected with probability 1 in the low-energy limit \(E\to 0\), the Bogoliubov quasiparticle associated with the spontaneous symmetry breaking exhibits the perfect transmission at \(E\to 0\). This perfect transmission is called the anomalous tunneling.2,3)
In this study, we are motivated by the results in Refs. 4–6 showing that the perfect transmission does not occur in a scalar BEC flowing with the critical current at the onset of the Landau instability. There are two types of instability in a BEC: the Landau instability and the dynamical instability. The former is the energetical instability characterized by a negative excitation energy, whereas the latter is the instability against the exponential growth of zero-energy excitations that are characterized by complex eigenfrequencies.18–21) It is then natural to ask how the anomalous tunneling occurs in a dynamically unstable system, which is yet to be investigated.
In this paper, we investigate the tunneling properties of quasiparticles in the presence of dynamically unstable modes. When a prepared BEC is dynamically unstable, the Bogoliubov–de Gennes (BdG) equation, which describes the excitation spectrum from the condensate, has complex eigenvalues. The tunneling problem is well defined even in such a case: We solve the BdG equation with a given eigenvalue E (which can be a complex value) and divide the obtained wave function into the incident, reflected, and transmitted waves, obtaining reflection and transmission probabilities.
Below, we consider two situations: a completely mixed binary (pseudo-spin-1/2) BEC and a spin-1 polar BEC. Dynamical instabilities in such systems are experimentally observed in, e.g., Refs. 20, 22–24. Here, we choose the spin configuration of the condensate such that the system can be stable or dynamically unstable depending on the spin-dependent interaction parameter. The BdG equations for spin waves in these systems are the same except for the quadratic Zeeman energy (\(q_{z}\)) term that appears only in the case of a polar BEC. Since the \(q_{z}\) term breaks the spin rotational symmetry, the spin-wave mode in the polar BEC is not the NG mode, whereas it is the NG mode in a binary BEC. We find that the perfect transmission occurs in the case of a binary BEC even when the system is dynamically unstable. However, the zero-energy limit should be taken not along the real E axis but along the imaginary E axis so that the quasiparticle wave function coincides with the condensate one in the \(E\to 0\) limit. On the other hand, the perfect transmission does not occur for any parameters in a spin-1 polar BEC with \(q_{z}\neq 0\) because the spin-wave mode is not the NG mode. Instead, when the kinetic energy of the incident wave matches the energy in the long-wavelength limit, the transmission probability resonantly increases. In the case of a spin-1 polar BEC, the eigenvalue E remains nonzero in the long-wavelength limit owing to the quadratic Zeeman energy. We find that the transmission probability resonantly increases when the energy of the incident quasiparticle matches this energy.
We also find that the perfect reflection occurs when \(\mathop{\text{Im}}\nolimits E\neq 0\) and \(d\mathop{\text{Im}}\nolimits E/dk=0\), where k is the momentum of the incident wave. This is understood as the consequence of the disappearance of the linearly independent plane-wave solution for the given E. The fact that the perfect reflection occurs in both systems of a binary BEC and a spin-1 polar BEC indicates that the origin of the perfect reflection is not related to the NG mode.
To find a signature of the tunneling properties observable in experiments, we numerically simulate the real-time dynamics of a binary BEC using the time-dependent Gross–Pitaevskii (GP) equation. We add a dynamically unstable fluctuation with a certain momentum only on one side of the potential barrier and see the growth of the fluctuations on each side of the potential. When the initial fluctuation is close to the perfect-transmission mode, its growth is approximated by \(\exp(\mathop{\text{Im}}\nolimits E t)\) for a long period and hardly affected by the potential barrier. At the same time, the fluctuation penetrates into the other side of the potential. However, because the growth is very slow owing to \(E\sim 0\), the initial fluctuation will be smeared out by the growth of more unstable modes. Hence, it would be difficult to observe the perfect transmission experimentally. As for the perfect reflection, although the origin of the perfect reflection is the destructive interference of the incident and reflected waves, the initial fluctuation grows as \(\exp(\mathop{\text{Im}}\nolimits E t)\) once the fluctuation is uniformly added on one side. We find that the fluctuation does not penetrate into the other side, which is the signature of the perfect reflection. Since the perfect reflection occurs for the most unstable (i.e., fastest-growing) mode, the signature will be detectable.
This paper is organized as follows. In Sect. 2, we investigate the tunneling properties of the spin-wave mode in a binary BEC. We introduce the system in Sect. 2.1 and derive solutions of the BdG equation in the absence of the barrier potential in Sect. 2.2. In Sect. 2.3, we solve the BdG equation under a barrier potential and discuss the tunneling properties. In Sect. 3, we discuss the tunneling problem in a spin-1 polar BEC in the same manner as in Sect. 2. In Sect. 4, we investigate the real-time dynamics of the tunneling phenomenon in a binary BEC. Section 5 concludes the paper.
2. Binary BEC
Model
We consider a binary BEC under a barrier potential at \(x=0\) (Fig. 1) and examine the tunneling properties of the quasiparticles through the barrier. The energy functional of the system is given by \begin{align} \mathcal{E}_{\text{binary}} &= \int d\boldsymbol{x}\sum_{m}\left[\frac{\hbar^{2}}{2M}\left|\frac{\partial}{\partial x}\Psi_{m}\right|^{2}{}+U(x)|\Psi_{m}|^{2}\right]\notag\\ &\quad + \int d\boldsymbol{x}\left[\sum_{m}\frac{g}{2}|\Psi_{m}|^{4}+ g'|\Psi_{1}|^{2}|\Psi_{2}|^{2}\right], \end{align} (1) where \(\Psi_{m}\) (\(m=1,2\)) is the condensate wave function, M is the atomic mass, \(U(x)\) is the barrier potential, and \(g>0\) and \(g'>0\) are the intra- and interspecies interaction strengths, respectively. For the sake of simplicity, we assume that the condensate wave function and the barrier potential depend only on x in a three-dimensional system so that the problem becomes essentially one-dimensional. We also assume that the strengths of the intraspecies interactions for the \(m=1\) and 2 components are the same. The two components are miscible (immiscible) for \(g>g'\) (\(g<g'\)).25,26) The potential \(U(x)\) becomes zero and the condensate density \(n(x)=\sum_{m}|\Psi_{m}|^{2}\) converges to a constant value \(n_{0}\) at \(x\to\pm\infty\).
By taking the functional derivative of the energy functional (1) with respect to \(\Psi^{*}_{m}\), we obtain the time-dependent GP equation: \begin{align} i\hbar \frac{\partial\Psi_{1}}{\partial t} &= \left[-\frac{\hbar^{2}}{2M}\frac{\partial^{2}}{\partial x^{2}}+U(x)+g|\Psi_{1}|^{2}+g'|\Psi_{2}|^{2}\right]\Psi_{1}, \end{align} (2a) \begin{align} i\hbar \frac{\partial\Psi_{2}}{\partial t} &= \left[-\frac{\hbar^{2}}{2M}\frac{\partial^{2}}{\partial x^{2}}+U(x)+g|\Psi_{2}|^{2}+g'|\Psi_{1}|^{2}\right]\Psi_{2}. \end{align} (2b) Below, we consider an equal-population mixture of the two components and prepare a completely overlapped state: \(\Psi_{1}(x)=\Psi_{2}(x)\equiv\Phi(x)/\sqrt{2}\), where \(\Phi(x)\) satisfies \begin{equation} \mu \Phi = \left[-\frac{\hbar^{2}}{2M}\frac{d^{2}}{dx^{2}}+U(x)+\frac{g+g'}{2}|\Phi|^{2}\right]\Phi \end{equation} (3) with μ being the chemical potential. The initial state is a stationary solution of Eq. (2) regardless of whether the binary mixture is miscible or not. However, its stability markedly changes at \(g=g'\): When \(g>g'\) (miscible), the initial state is the ground state of the system; hence, the quasiparticle eigenfrequencies are all positive real numbers. When \(g<g'\) (immiscible), the initial state is dynamically unstable against phase separation, and complex eigenfrequencies appear.
Since \(U(x)\to 0\) and \(|\Phi|^{2}\to n_{0}\) (= const.) at \(x\to\pm\infty\), Eq. (3) leads to \(\mu=(g+g')n_{0}/2\). We describe the equations with rescaling the energy, length, and time scales by μ, \(\hbar/\sqrt{M\mu}\), and \(\mu/\hbar\), respectively. By introducing the interaction parameter β (\(-1\leq\beta\leq 1\)) as \begin{equation} g = (g+g')\frac{1+\beta}{2},\quad g' = (g+g')\frac{1-\beta}{2} \end{equation} (4) and rewriting the wave functions as \(\Psi_{1,2}(x)=\sqrt{n_{0}}\psi_{1,2}(x)\) and \(\Phi(x)=\sqrt{n_{0}}\phi(x)\), we obtain the dimensionless forms of Eqs. (2) and (3) as \begin{align} i\frac{\partial\psi_{1}}{\partial t} &= [\mathcal{L}_{0}+(1+\beta)|\psi_{1}|^{2}+(1-\beta)|\psi_{2}|^{2}]\psi_{1}, \end{align} (5a) \begin{align} i\frac{\partial\psi_{2}}{\partial t} &= [\mathcal{L}_{0}+(1+\beta)|\psi_{2}|^{2}+(1-\beta)|\psi_{1}|^{2}]\psi_{2}, \end{align} (5b) and \begin{equation} (\mathcal{L}_{0}-1+|\phi|^{2})\phi = 0, \end{equation} (6) where \begin{equation} \mathcal{L}_{0} = - \frac{1}{2}\frac{d^{2}}{dx^{2}}+U(x). \end{equation} (7)
The dynamics of quasiparticle excitations from a condensate at a low temperature is well described by the Bogoliubov theory. The BdG equation for a binary BEC is obtained by substituting \begin{equation} \begin{pmatrix} \psi_{1}\\ \psi_{2} \end{pmatrix}= e^{-it} \left[\frac{\phi}{\sqrt{2}} \begin{pmatrix} 1\\ 1 \end{pmatrix}+ \boldsymbol{u}e^{-iE t}+\boldsymbol{v}^{*}e^{+iE^{*} t}\right] \end{equation} (8) into Eq. (5) and linearizing the equation with respect to \(\boldsymbol{u}\) and \(\boldsymbol{v}\), where \(\boldsymbol{u}=(u_{1}(x),u_{2}(x))^{\text{T}}\) and \(\boldsymbol{v}=(v_{1}(x),v_{2}(x))^{\text{T}}\) are two-component spinors. The resulting eigenvalue equation is given by \begin{equation} \begin{pmatrix} (\mathcal{L}_{0}-1)\mathbf{1}+ H_{1} &H_{2}\\ -H^{*}_{2} &-[(\mathcal{L}_{0}-1)\mathbf{1}+ H^{*}_{1}] \end{pmatrix}\begin{pmatrix} \boldsymbol{u}\\ \boldsymbol{v}\end{pmatrix}= E \begin{pmatrix} \boldsymbol{u}\\ \boldsymbol{v}\end{pmatrix}, \end{equation} (9) where \(\mathbf{1}\) is the \(2\times 2\) identity matrix, and \begin{align} H_{1} &= \frac{|\phi|^{2}}{2} \begin{pmatrix} 3+\beta &1-\beta\\ 1-\beta &3+\beta \end{pmatrix}, \end{align} (10) \begin{align} H_{2} &= \frac{\phi^{2}}{2} \begin{pmatrix} 1+\beta &1-\beta\\ 1-\beta &1+\beta \end{pmatrix}. \end{align} (11)
When the two components completely overlap, the density- and spin-wave modes are decoupled in the BdG equation. Indeed, by defining the density-wave (phonon) mode \(u^{\text{d}}=u_{1}+u_{2}\), \(v^{\text{d}}=v_{1}+v_{2}\) and the spin-wave (magnon) mode \(u^{\text{s}}=u_{1}-u_{2}\), \(v^{\text{s}}=v_{1}-v_{2}\), BdG Eq. (9) is divided into the following two equations: \begin{align} &\begin{pmatrix} \mathcal{L}_{0} -1 + 2|\phi|^{2} &\phi^{2}\\ -(\phi^{*})^{2} &-(\mathcal{L}_{0}-1 +2|\phi|^{2}) \end{pmatrix}\begin{pmatrix} u^{\text{d}}\\ v^{\text{d}} \end{pmatrix}\notag\\ &\quad= E^{\text{d}} \begin{pmatrix} u^{\text{d}}\\ v^{\text{d}} \end{pmatrix}, \end{align} (12) \begin{align} &\begin{pmatrix} \mathcal{L}_{0} -1 + (1+\beta)|\phi|^{2} &\beta\phi^{2}\\ -\beta(\phi^{*})^{2} &-[\mathcal{L}_{0} -1 + (1+\beta)|\phi|^{2}] \end{pmatrix}\begin{pmatrix} u^{\text{s}}\\ v^{\text{s}} \end{pmatrix}\notag\\ &\quad= E^{\text{s}} \begin{pmatrix} u^{\text{s}}\\ v^{\text{s}} \end{pmatrix}. \end{align} (13) When \(U(x)=0\), the spectra of these equations exhibit gapless linear dispersions, corresponding to the NG phonon and NG magnon modes associated with the breaking of the U(1) gauge symmetry and the SO(2) spin rotation symmetry, respectively. (The spin here means the pseudo-spin-1/2 of the binary system.) In the case of \(U(x)\neq 0\), the existence of the NG modes is confirmed by the fact that \(u^{\text{d,s}}=-(v^{\text{d,s}})^{*}=\phi\) are the eigen solutions of Eqs. (12) and (13) with \(E^{\text{d,s}}=0\).
We note that GP Eq. (6) and BdG Eq. (12) for the phonon mode are identical to those for a scalar BEC. Hence, the tunneling problem is the same as that in the case of a scalar BEC discussed in Ref. 3. We therefore discuss the tunneling properties of the spin-wave mode in the rest of this section.
Bogoliubov spectrum in a uniform system
Before solving the tunneling problem, we analytically derive asymptotic forms of the quasiparticle wave function at \(x\to\pm\infty\). In the case of \(U(x)=0\), the stationary solution of Eq. (6) is \(\phi=1\). By substituting \((u^{\text{s}},v^{\text{s}})^{\text{T}}=(ue^{ikx},ve^{ikx})^{\text{T}}\) and \(\phi=1\) to Eq. (13), we obtain the BdG equation for spin-wave modes at \(U(x)=0\) as \begin{equation} \begin{pmatrix} \epsilon_{k}+ \beta &\beta\\ -\beta &-(\epsilon_{k} + \beta) \end{pmatrix}\begin{pmatrix} u\\ v \end{pmatrix}= E^{\text{s}} \begin{pmatrix} u\\ v \end{pmatrix}, \end{equation} (14) which has the eigenvalue \begin{equation} E^{\text{s}} = \sqrt{\epsilon_{k}(\epsilon_{k}+2\beta)}, \end{equation} (15) where \(\epsilon_{k}\equiv k^{2}/2\). Note that when \(\beta<0\), \(E^{\text{s}}\) becomes a pure imaginary value for small k and the system becomes dynamically unstable. This instability is caused by the immiscible components [\(g<g'\), which is equivalent to \(\beta<0\), see Eq. (4)] that are mixed in the initial state. Our interest in this paper is to clarify how such dynamically unstable modes are reflected or transmitted by the barrier potential. In the case of \(\beta\geq 0\), for which the system is dynamically stable, almost the same situation has been discussed in the previous works,9,10) where the tunneling problem of spin waves in a spin-1 polar state has been investigated and the perfect transmission was observed in the low-energy limit. The BdG Eq. (12) for density-wave modes at \(U(x)=0\) has the eigenvalue \(E^{\text{d}}=\sqrt{\epsilon_{k}(\epsilon_{k}+2)}\), which is a positive real value for \(^{\forall} k\).
In the tunneling problem, we will solve the quasiparticle wave function injected from \(x=-\infty\) with an energy E. Here, we therefore calculate normalized eigenvectors of Eq. (14) as a function of an incident eigenvalue E. We should be careful in calculating the normalization constant for the BdG equation for a bosonic system since the BdG equation is a non-Hermitian matrix equation. For the case of Eq. (14), the eigenvector is normalized as \(|u|^{2}-|v|^{2}=1\) or −1 (\(u^{2}-v^{2}=1\) or −1) for real (pure imaginary) E. See Appendix for the details.
Real positive eigenvalue state
For a real positive E, there are two propagating modes and two damping/growing modes [Figs. 2(a) and 2(c)] respectively given by \begin{equation} \begin{pmatrix} a_{r}\\ -b_{r} \end{pmatrix}e^{\pm i k_{1} x}, \end{equation} (16) and \begin{equation} \begin{pmatrix} b_{r}\\ a_{r} \end{pmatrix}e^{\mp q_{2} x}, \end{equation} (17) where \begin{align} a_{r} &= \mathop{\text{sgn}}\nolimits(\beta)\sqrt{\frac{\sqrt{\beta^{2} + E^{2}}}{2E}+\frac{1}{2}}, \end{align} (18) \begin{align} b_{r} &= \sqrt{\frac{\sqrt{\beta^{2} + E^{2}}}{2E}-\frac{1}{2}}, \end{align} (19) \begin{align} k_{1} &= \sqrt{2}\sqrt{\sqrt{\beta^{2}+E^{2}} - \beta}, \end{align} (20) \begin{align} q_{2} &= \sqrt{2}\sqrt{\sqrt{\beta^{2}+E^{2}} + \beta}, \end{align} (21) and \(a_{r}\) and \(b_{r}\) satisfy \(|a_{r}|^{2}-|b_{r}|^{2}=1\). These are eigensolutions in the entire region of β: \(-1\leq\beta\leq 1\).
Figure 2. (Color online) Spin-wave spectrum [Eq. (15)] for a binary BEC. Shown are \(E^{2}\) as a function of \(\epsilon_{k}=k^{2}/2\) for \(\beta\geq 0\) (a, b) and \(\beta<0\) (c, d). We solve k for a given \(E^{2}\), where \(E^{2}>0\) (\(E^{2}<0\)) means a real (pure imaginary) eigenvalue E. The solutions with \(\epsilon_{k}>0\) (\(\epsilon_{k}<0\)) are propagating (localized) modes. The points indicated by (i)–(iv) correspond to the solutions of the incident wave for the cases of (i)–(iv) with the asymptotic forms given by Eqs. (27)–(29).
Pure imaginary eigenvalue state
We rewrite the pure imaginary eigenvalue E as \(E=i\Delta\), where \(\Delta\in\mathbb{R}\) and \(|\Delta|<|\beta|\). For \(\beta < 0\), there are four propagating solutions [Fig. 2(d)]: \begin{equation} \begin{pmatrix} b_{c}\\ a_{c} \end{pmatrix}e^{\pm ik_{3} x}\quad \text{and}\quad \begin{pmatrix} a_{c}\\ -b_{c} \end{pmatrix}e^{\pm ik_{4} x}, \end{equation} (22) which change to four damping/growing solutions for \(\beta\geq 0\) [Fig. 2(b)]: \begin{equation} \begin{pmatrix} b_{c}\\ a_{c} \end{pmatrix}e^{\mp q_{3} x}\quad \text{and}\quad \begin{pmatrix} a_{c}\\ -b_{c} \end{pmatrix}e^{\mp q_{4} x}, \end{equation} (23) where \begin{equation} a_{c} = \mathop{\text{sgn}}\nolimits(\beta)\frac{e^{-i\theta/2}}{\sqrt{-2i\sin\theta}},\quad b_{c} = - \frac{e^{i\theta/2}}{\sqrt{-2i\sin\theta}} \end{equation} (24) with \(\theta\equiv\sin^{-1}(\Delta/|\beta|)\), and \begin{align} k_{3} &= iq_{3} = \sqrt{2}\sqrt{+\sqrt{\beta^{2}-\Delta^{2}} - \beta}, \end{align} (25a) \begin{align} k_{4} &= iq_{4} = \sqrt{2}\sqrt{-\sqrt{\beta^{2}-\Delta^{2}} - \beta}. \end{align} (25b) Here, \(a_{c}\) and \(b_{c}\) are normalized as \(a_{c}^{2}-b_{c}^{2}=1\). We define \(k_{3,4}\) and \(q_{3,4}\) to be positive real values and do not use them when they are imaginary values.
Tunneling properties of the spin-wave modes
We now solve BdG Eq. (13) and calculate the reflection and transmission probabilities of quasiparticles with energy E and momentum \(k_{\text{in}}\) injected from the left. As for the barrier potential \(U(x)\), we use the Gaussian potential given by \begin{equation} U(x) = U_{0}e^{-x^{2}/(2\sigma^{2})}. \end{equation} (26) We numerically solve BdG Eq. (13) by the finite element method with the asymptotic form of the wave function imposed at \(x\to\pm\infty\).
In the case of \(\beta\geq 0\), a quasiparticle with \(k_{\text{in}}=k_{1}\) injected from the left (\(e^{ik_{1} x}\)) is reflected to the left (\(e^{-ik_{1} x}\)) or transmitted to the right (\(e^{ik_{1} x}\)). In addition, localized modes at the potential barrier appear (\(e^{\pm q_{2} x}\)). The asymptotic form of the wave function at \(x\to\pm \infty\) is then given by
Here, R and T are the reflection and transmission coefficients, respectively, and A and B are the coefficients for the localized modes. The current conservation law requires \(|R|^{2}+|T|^{2}=1\).
In the case of \(\beta<0\), there are three possibilities: (ii) real positive \(E\in\mathbb{R}\) with \(k_{\text{in}}=k_{1}\), (iii) pure imaginary \(E=i\Delta\in i\mathbb{R}\) (\(\Delta>0\)) with \(k_{\text{in}}=-k_{4}\), and (iv) pure imaginary \(E=i\Delta\in i\mathbb{R}\) with \(k_{\text{in}}=k_{3}\). The results for \(\Delta<0\) are obtained by taking the complex conjugate of the wave functions obtained for \(\Delta>0\). The points corresponding to the incident wave are depicted in the dispersion relation in Fig. 2. Note that although the group velocity of a quasiparticle with a pure imaginary eigenvalue is zero, the causality is satisfied when we assume \(-d|\Delta|/dk\) to be a group velocity and choose the momentum of the incident, reflected, and transmitted waves. [The origin of the causality becomes clearer when the spectrum has a gap at \(k=0\), as in Eq. (43). We shall revisit this issue in Sect. 3.3.] Hence, the asymptotic forms of the wave function at \(x\to\pm \infty\) are given as follows:
(ii)
\(\beta<0\), \(E\in\mathbb{R}\) with \(k_{\text{in}}=k_{1}\):
Here, we further introduce the scaled coefficients \begin{equation} \tilde{A} = \sqrt{-\frac{k_{\text{sc}}}{k_{\text{in}}}} A,\quad \tilde{B} = \sqrt{-\frac{k_{\text{sc}}}{k_{\text{in}}}} B, \end{equation} (30) where \(k_{\text{sc}}\) is the real part of the momentum of the A term. For example, \(k_{\text{sc}}=0\), \(-k_{3}\), and \(k_{4}\) for the cases (ii), (iii), and (iv), respectively. Using \(\tilde{A}\) and \(\tilde{B}\), we write the current conservation law as \begin{equation} |T|^{2}+|R|^{2}+|\tilde{A}|^{2}+|\tilde{B}|^{2} = 1 \end{equation} (31) for all cases (i)–(iv).
In Fig. 3, we show the numerical results for the tunneling coefficients for the cases (i)–(iv). The results for the quasiparticle wave functions are depicted in Fig. 4. Figure 3(a) shows the transmission probability for real E (\(>0\)) with various values of β, obtained by imposing the asymptotic form of Eq. (27) [cases (i) and (ii)]. One can see that the perfect transmission occurs at the low-energy limit for \(\beta\geq 0\), i.e., \(|T|^{2}\) goes to unity as \(E\to 0\), which is consistent with the previous works.9,10) On the other hand, for \(\beta<0\), the transmission probability becomes smaller than unity. This difference is understood as follows. The perfect transmission occurs when the quasiparticle wave function coincides with the condensate wave function: \(u(x)=-v^{*}(x)=\phi(x)\). In the present case, the incident momentum \(k_{1}\) goes to zero (nonzero) as \(E\to 0\) when \(\beta\geq 0\) (\(\beta<0\)); hence \(u(x)=-v^{*}(x)=\phi(x)\) is (is not) satisfied at \(E\to 0\). In Figs. 4(a) and 4(b), we show \(|u|^{2}-|v|^{2}\) at \(\beta=0.2\) and −0.2, respectively. [The integral \(\int (|u|^{2}-|v|^{2})\,dx\) gives the norm of the quasiparticle wave function for a real E, whereas it should vanish for a complex E. We therefore plot \(|u|^{2}+|v|^{2}\) for \(\beta<0\) as shown in Figs. 4(c) and 4(d). See also Appendix.] One can see from these figures that \(|u|^{2}-|v|^{2}\) for \(\beta=0.2\) becomes close to \(|\phi|^{2}\) as \(E\to 0\), whereas for \(\beta=-0.2\) at \(E\to 0\), it has a completely different x dependence from \(|\phi|^{2}\), which is consistent with the above discussion. The result in the previous works4–6) that the phonon mode in a scalar BEC moving with the critical current does not exhibit the perfect transmission is due to the same reason, where u and \(v^{*}\) deviate from ϕ owing to local enhancement of density fluctuations around the potential barrier.6)
Figure 3. (Color online) Tunneling property of spin-wave modes in a binary BEC. (a) Transmission probability \(|T|^{2}\) for real E, obtained by imposing the asymptotic forms of Eq. (27) [cases (i) and (ii)]. The perfect transmission occurs in the low-energy limit, i.e., \(|T|^{2}\rightarrow 1\) as \(E\rightarrow 0\), for \(\beta\geq 0\), whereas \(|T|^{2}\) at \(E\to 0\) for \(\beta<0\) is smaller than unity. (b) E dependence of the tunneling coefficients \(|T|^{2}\), \(|R|^{2}\), \(|\tilde{A}|^{2}\), and \(|\tilde{B}|^{2}\) in the presence of dynamical instability at \(\beta=-0.2\). The horizontal axis is chosen such that \(\epsilon_{k_{\text{in}}}=0\) at the left edge of the panel and \(\epsilon_{k_{\text{in}}}\) increases as one goes to the right. Correspondingly, along the horizontal axis, the value of E starts from 0, \(\mathop{\text{Im}}\nolimits E\) first increases [case (iii)], takes a maximum value (0.2), and decreases to 0 [case (iv)], and then E changes to a real value [case (ii)]. The region with pure imaginary E is shaded in gray. The perfect transmission (\(|T|^{2}=1\)) occurs at \(E=0\) at the left edge of the panel, and the perfect reflection (\(|\tilde{A}|^{2}=1\)) occurs at \(E=0.2i\). We use the barrier potential with \(U_{0}=2\) and \(\sigma=0.5\) for both panels. The perfect transmission and the perfect reflection occur independently from the values of \(U_{0}\) and σ.
Figure 4. (Color online) Quasiparticle wave function of spin-wave modes in a binary BEC. Shown are \(|u|^{2}-|v|^{2}\) for real eigenvalue modes (a, b), and \(|u|^{2}+|v|^{2}\) for pure imaginary eigenvalue modes (c, d). The panels (a) and (b) are the results for the zero-energy limit at \(\beta=0.2\) and −0.2, respectively, obtained for the asymptotic form of Eq. (27) [cases (i) and (ii)]. As E approaches 0, \(|u|^{2}-|v|^{2}\) approaches \(|\phi|^{2}\) in (a), whereas its x dependence has a distinctive difference from \(|\phi|^{2}\) in (b), which is consistent with the presence (a) and absence (b) of the perfect transmission (see text). Panels (c) and (d) are the results for \(\beta=-0.2\) with the asymptotic form of Eq. (28) [case (iii)] in the limits of \(E\to 0\) and \(E\to i|\beta|\), respectively. As E approaches 0, \(|u|^{2}+|v|^{2}\) becomes proportional to \(|\phi|^{2}\) in (c), which is consistent with the conditions for the perfect transmission. In (d), the quasiparticle wave function becomes smaller as E approaches \(i|\beta|\). This is due to the destructive interference between the incident and reflected waves. For all panels, we use the barrier potential with \(U_{0}=2\) and \(\sigma=0.5\).
Note that \(k_{4}\) goes to zero and \(u(x)=-v^{*}(x)=\phi(x)\) is satisfied when E goes to zero along the imaginary axis. Hence, the perfect transmission at \(E\to 0\) occurs in this case. Figure 3(b) shows the E dependences of \(|T|^{2}\), \(|R|^{2}\), \(|\tilde{A}|^{2}\), and \(|\tilde{B}|^{2}\) at \(\beta=-0.2\). Here, we set the horizontal axis of Fig. 3(b) such that \(\epsilon_{k_{\text{in}}}\) is zero at the left end of the figure and \(\epsilon_{k_{\text{in}}}\) increases as one goes to the right. The region where E is pure imaginary is shaded in gray. We have numerically confirmed the current conservation law given by Eq. (31). In Fig. 3(b), the transmission probability \(|T|^{2}\) goes to unity as \(\mathop{\text{Im}}\nolimits E\to 0\) at the left end of the figure, indicating that the perfect transmission occurs even for the dynamically unstable modes. We also confirm that \(|u|^{2}+|v|^{2}\) shown in Fig. 4(c) becomes proportional to \(|\phi|^{2}\) as \(\mathop{\text{Im}}\nolimits E\to 0\), which is consistent with \(u(x)=-v^{*}(x)=\phi(x)\). On the other hand, at the point where E changes from pure imaginary values to real values [the point \(E=0\) and \(\epsilon_{k}>0\) in Figs. 2(c) and 2(d), which corresponds to the boundary between the gray shaded region and the white region in Fig. 3(b)], the incident momentum \(k_{3}\) remains nonzero in the limit of \(E\to 0\), which means \(u(x)=-v^{*}(x)=\phi(x)\) is not satisfied in this limit, resulting in the absence of the perfect transmission.
We also find from Fig. 3(b) that \(|\tilde{A}|^{2}\) goes to unity at \(E=i|\beta|\). At this point, \(|{\mathop{\text{Im}}\nolimits E}|\) takes its maximum value, i.e., \(E\in i\mathbb{R}\) and \(d\mathop{\text{Im}}\nolimits E/dk=0\). When E is pure imaginary, the A terms in Eqs. (28) and (29) represent the reflected wave with the momentum different from the incident one. Hence, \(|\tilde{A}|^{2}=1\) means the occurrence of perfect reflection. The origin of the perfect reflection is that the number of linearly independent solutions given in Eq. (22) decreases at \(k_{3}=k_{4}\), and the solution that satisfies the asymptotic forms of Eqs. (28) and (29) disappears. Thus, the incident wave and reflected wave (the A term) destructively interfere. The destructive interference can be confirmed in the quasiparticle wave function shown in Fig. 4(d), in which we plot \(|u|^{2}+|v|^{2}\) obtained for \(\beta=-0.2\) with the asymptotic form of Eq. (28). The wave function goes to zero as \(E\to i|\beta|\) even in the \(x<0\) region. Since we fix the amplitude of the incident wave, the reduction in \(|u|^{2}+|v|^{2}\) means the destructive interference. Perfect reflection was also predicted to occur for the long-wavelength limit of the spin-wave mode in a ferromagnetic BEC.9,10) The perfect reflection we find in this study is different from that in the previous works in that the origin of the perfect reflection in the latter case is essentially the same as that of the ordinary quantum-mechanical tunneling of a free particle.
3. Spin-1 Polar BEC
Model
Next, we consider the tunneling problem in a spin-1 polar BEC. The BdG equation for the spin-wave mode in this system is the same as that in the previous section except for the quadratic Zeeman energy \(q_{z}\) term. Since the quadratic Zeeman effect breaks the spin rotational symmetry, the spin-wave spectrum has a nonzero eigenvalue at \(k=0\). Here, the eigenvalue at \(k=0\) can be real or pure imaginary depending on the value of \(q_{z}\). We therefore focus on the \(q_{z}\) dependence of the tunneling properties. Below, we redefine the variables used in the previous section, so that the resulting BdG equation has the same form as that in the previous section except for the \(q_{z}\) term.
The energy functional of a spin-1 system is given by \begin{align} \mathcal{E}_{\text{spin-1}} &= \int d\boldsymbol{x}\sum_{m}\left\{\frac{\hbar^{2}}{2M}\left|\frac{\partial \Psi_{m}}{\partial x}\right|^{2}{}+[U(x)+q_{z} m^{2}]|\Psi_{m}|^{2}\right\}\notag\\ &\quad + \frac{1}{2}\int d\boldsymbol{x}[c_{0}n^{2}+c_{1}|\boldsymbol{F}|^{2}], \end{align} (32) where \(\Psi_{m}\) is the condensate wave function of the atoms in the magnetic sublevel \(m=1\), 0, and −1, M is the atomic mass, \(U(x)\) is the barrier potential, \(n(x)=\sum_{m}|\psi_{m}(x)|^{2}\) is the condensate density, and \(\boldsymbol{F}(x) =(F_{x}(x),F_{y}(x),F_{z}(x))\) is the spin density vector defined by \(\boldsymbol{F}(x)=\sum_{mm'}\Psi_{m}^{*}(x)\boldsymbol{S}_{mm'}\Psi_{m'}(x)\) with \(\boldsymbol{S}=(S_{x},S_{y},S_{z})\) being the spin-1 matrices. The interaction coefficients are given by \(c_{0}=4\pi\hbar^{2}(2a_{2}+a_{0})/(3M)\) and \(c_{1}=4\pi\hbar^{2}(a_{2}-a_{0})/(3M)\), where \(a_{\mathcal{F}}\) is the s-wave scattering length for the total spin \(\mathcal{F}=0,2\) channel.
The ground-state phase of this system is determined by the values of \(c_{1}\) and \(q_{z}\). (The phase diagram is given in, e.g., Ref. 27.) Below, we consider a condensate in the \(m=0\) state, i.e., a polar BEC. The polar BEC is the ground state of the system when \(q_{z}>\max(0,-2c_{1}n)\), and the system becomes dynamically unstable when \((c_{1},q_{z})\) is outside of this region.
The GP and BdG equations are obtained by the same manner as in the previous section. In the case of a polar BEC, the stationary solution \((\Psi_{1},\Psi_{0},\Psi_{-1})=(0,\Phi(x),0)\) satisfies \begin{equation} \mu \Phi = \left[-\frac{\hbar^{2}}{2M}\frac{d^{2}}{dx^{2}}+U(x)+c_{0}|\Phi|^{2}\right]\Phi, \end{equation} (33) where μ is the chemical potential. Assuming \(U(x)\rightarrow 0\) and \(|\Phi|^{2}\rightarrow n_{0}\) (= const.) at \(x\rightarrow\pm\infty\), we obtain \(\mu=c_{0}n_{0}\). We use this μ to scale the dimensionful variables, i.e., we rescale the energy, length, and time scales by μ, \(\hbar/\sqrt{M\mu}\), and \(\mu/\hbar\), respectively. By rewriting \(\Phi(x)=\sqrt{n_{0}}\phi(x)\), we reduce Eq. (33) to \begin{equation} (\mathcal{L}_{0}-1+|\phi|^{2})\phi = 0, \end{equation} (34) where \(\mathcal{L}_{0}\) is defined in Eq. (7). Equation (34) is identical to Eq. (6).
We introduce the interaction parameter β as \begin{equation} \beta \equiv \frac{c_{1}}{c_{0}}. \end{equation} (35) By rewriting the wave function as \(\Psi_{m}(x)=\sqrt{n_{0}}\psi_{m}(x)\), we obtain the time-dependent GP equation in the dimensionless form as \begin{align} i\frac{\partial\psi_{\pm 1}}{\partial t} &= [\mathcal{L}_{0}+1\pm\beta f_{z}+q_{z}]\psi_{\pm 1}+\frac{\beta}{\sqrt{2}}f_{\mp}\psi_{0}, \end{align} (36a) \begin{align} i\frac{\partial\psi_{0}}{\partial t} &= [\mathcal{L}_{0}+1]\psi_{0} +\frac{\beta}{\sqrt{2}}(f_{+}\psi_{+1}+f_{-}\psi_{-1}), \end{align} (36b) where \(f_{+}=f_{-}^{*}=\sqrt{2}(\psi_{1}^{*}\psi_{0}+\psi_{0}^{*}\psi_{-1})\) and \(f_{z}=|\psi_{1}|^{2}-|\psi_{-1}|^{2}\). By substituting \begin{equation} \begin{pmatrix} \psi_{+1}\\ \psi_{0}\\ \psi_{-1} \end{pmatrix}= e^{-it}\left[ \begin{pmatrix} 0\\ \phi\\ 0 \end{pmatrix}+\boldsymbol{u}e^{-iEt}+\boldsymbol{v}^{*}e^{+iE^{*}t}\right] \end{equation} (37) into Eq. (36) and linearizing the equation with respect to \(\boldsymbol{u}\) and \(\boldsymbol{v}\), with \(\boldsymbol{u}=(u_{+1},u_{0},u_{-1})^{\text{T}}\) and \(\boldsymbol{v}=(v_{+1},v_{0},v_{-1})^{\text{T}}\) being three-component spinors, we obtain the BdG equation for a polar BEC as \begin{equation} \begin{pmatrix} H_{0} + H_{1} &H_{2}\\ -H^{*}_{2} &-[H_{0} + H^{*}_{1}] \end{pmatrix}\begin{pmatrix} \boldsymbol{u}\\ \boldsymbol{v}\end{pmatrix}= E \begin{pmatrix} \boldsymbol{u}\\ \boldsymbol{v}\end{pmatrix}, \end{equation} (38) where \begin{align} H_{0} &= \begin{pmatrix} \mathcal{L}_{0}-1+q_{z} &0 &0\\ 0 &\mathcal{L}_{0}-1 &0\\ 0 &0 &\mathcal{L}_{0}-1+q_{z} \end{pmatrix}, \end{align} (39a) \begin{align} H_{1} &= |\phi|^{2} \begin{pmatrix} 1+\beta &0 &0\\ 0 &2 &0\\ 0 &0 &1+\beta \end{pmatrix}, \end{align} (39b) \begin{align} H_{2} &= \phi^{2} \begin{pmatrix} 0 &0 &\beta\\ 0 &1 &0\\ \beta &0 &0 \end{pmatrix}. \end{align} (39c) This \(6\times 6\) eigenvalue equation can be divided into three \(2\times 2\) equations. By defining the density-wave mode \(u^{\text{d}}=u_{0}\), \(v^{\text{d}}=v_{0}\) and spin-wave modes \(u_{\pm}^{\text{s}}=u_{\pm 1}\), \(v_{\pm}^{\text{s}}=v_{\mp 1}\), we reduce BdG Eq. (38) to \begin{align} \begin{pmatrix} \mathcal{L}_{0} -1 + 2|\phi|^{2} &\phi^{2}\\ -(\phi^{*})^{2} &-(\mathcal{L}_{0}-1 +2|\phi|^{2}) \end{pmatrix}\begin{pmatrix} u^{\text{d}}\\ v^{\text{d}} \end{pmatrix}&= E^{\text{d}} \begin{pmatrix} u^{\text{d}}\\ v^{\text{d}} \end{pmatrix}, \end{align} (40) \begin{align} \begin{pmatrix} \mathcal{L}_{0} -1 +q_{z}+ (1+\beta)|\phi|^{2} &\beta\phi^{2}\\ -\beta(\phi^{*})^{2} &-(\mathcal{L}_{0} -1 +q_{z}+(1+\beta)|\phi|^{2}) \end{pmatrix}\begin{pmatrix} u_{\pm}^{\text{s}}\\ v_{\pm}^{\text{s}} \end{pmatrix}&= E^{\text{s}} \begin{pmatrix} u_{\pm}^{\text{s}}\\ v_{\pm}^{\text{s}} \end{pmatrix}. \end{align} (41) Equation (40) has a zero-energy solution \(u^{\text{d}}=-(v^{\text{d}})^{*}=\phi\), indicating that the density-wave is the NG phonon associated with the spontaneous breaking of the U(1) gauge symmetry. On the other hand, the spin-wave mode is not the NG mode for \(q_{z}\neq 0\), since the SO(3) spin rotational symmetry is broken in the presence of the quadratic Zeeman energy. The previous works investigated the tunneling properties for BdG Eq. (41) with \(q_{z}=0\) and \(\beta>0\) and showed that the perfect transmission occurs in the low energy limit.9,10)
Note that Eq. (41) at \(q_{z}=0\) is identical to Eq. (13). The role of \(q_{z}\) is that it effectively shifts the chemical potential from 1 to \(1-q_{z}\), opening an energy gap at \(k=0\) in a uniform system (see below). This is possible because the condensation occurs in a different internal state from the quasiparticles.
Bogoliubov spectrum in a uniform system
We analytically solve BdG Eq. (41) for \(U(x)=0\) and obtain propagating and damping/growing solutions. By substituting \((u^{\text{s}},v^{\text{s}})^{\text{T}}=(ue^{ikx},ve^{ikx})^{\text{T}}\) and \(\phi=1\) to Eq. (41), we obtain \begin{equation} \begin{pmatrix} \epsilon_{k} + q_{z} + \beta &\beta\\ -\beta &-(\epsilon_{k} + q_{z} + \beta) \end{pmatrix}\begin{pmatrix} u\\ v \end{pmatrix}= E^{\text{s}} \begin{pmatrix} u\\ v \end{pmatrix}, \end{equation} (42) which has the eigenvalue \begin{equation} E^{\text{s}} = \sqrt{(\epsilon_{k}+q_{z})(\epsilon_{k}+q_{z}+2\beta)}. \end{equation} (43) It follows that when \(q_{z}<\max(0,-2\beta)\), \(E^{\text{s}}\) becomes pure imaginary for a certain region of k and the system becomes dynamically unstable. The condition for the dynamical stability \(q_{z}>\max(0,-2\beta)\) agrees with the region for which the polar state is the ground state.
Real positive eigenvalue state
For a real positive E, the four linearly independent solutions are given by \begin{equation} \begin{pmatrix} a_{r}\\ -b_{r} \end{pmatrix}e^{\pm i k_{1} x}\quad \text{and}\quad \begin{pmatrix} b_{r}\\ a_{r} \end{pmatrix}e^{\mp ik_{2} x}, \end{equation} (44) where \(a_{r}\) and \(b_{r}\) are defined in Eqs. (18) and (19), respectively, and \(k_{1}\) and \(k_{2}\) are respectively given by \begin{align} k_{1} &= \sqrt{2}\sqrt{+\sqrt{\beta^{2}+E^{2}}-(\beta+q_{z})}, \end{align} (45) \begin{align} k_{2} &= \sqrt{2}\sqrt{-\sqrt{\beta^{2}+E^{2}}-(\beta+q_{z})}. \end{align} (46) The right-hand sides of Eqs. (45) and (46) can be real or pure imaginary. When \(k_{1,2}\) is pure imaginary, we rewrite it as \(k_{1,2}=iq_{1,2}\) and use real-valued \(q_{1,2}\) (\(>0\)). The corresponding solutions in Eq. (44) express both propagating modes (\(e^{\pm i k_{1,2}x}\)) and growing/damping modes (\(e^{\mp q_{1,2}x}\)).
Pure imaginary eigenvalue state
By rewriting the eigenvalue as \(E=i\Delta\) with \(\Delta\in\mathbb{R}\) and \(|\Delta|<|\beta|\), we obtain the four solutions given by \begin{equation} \begin{pmatrix} b_{c}\\ a_{c} \end{pmatrix}e^{\pm ik_{3} x}\quad \text{and}\quad \begin{pmatrix} a_{c}\\ -b_{c} \end{pmatrix}e^{\pm ik_{4} x}, \end{equation} (47) where \(a_{c}\) and \(b_{c}\) are defined in Eq. (24) and \begin{align} k_{3} &= \sqrt{2}\sqrt{+\sqrt{\beta^{2}-\Delta^{2}} - (\beta+q_{z})}, \end{align} (48) \begin{align} k_{4} &= \sqrt{2}\sqrt{-\sqrt{\beta^{2}-\Delta^{2}} - (\beta+q_{z})}. \end{align} (49) The right-hand sides of Eqs. (48) and (49) can be real or pure imaginary. When \(k_{3,4}\) is pure imaginary, we rewrite it as \(k_{3,4}=iq_{3,4}\) and use real-valued \(q_{3,4}\) (\(>0\)). The corresponding solutions in Eq. (44) express both propagating modes (\(e^{\pm i k_{3,4}x}\)) and growing/damping modes (\(e^{\mp q_{3,4}x}\)).
Tunneling properties of the spin-wave modes
We now investigate the tunneling properties of the spin-wave modes in a polar BEC. In the same way as in the previous section, for given energy E and momentum \(k_{\text{in}}\) of the incident wave, we construct the asymptotic form of the quasiparticle wave function at \(x\to\pm\infty\) and numerically solve BdG Eq. (41) by the finite element method. In the presence of the quadratic Zeeman effect, the energy dispersion is categorized into four types as shown in Fig. 5: (a) \(q_{z}>0\) and \(q_{z}+2\beta>0\), (b) \(q_{z}(q_{z}+2\beta)\leq 0\) and \(q_{z}+\beta>0\), (c) \(q_{z}(q_{z}+2\beta)\leq 0\) and \(q_{z}+\beta\geq 0\), and (d) \(q_{z}<0\) and \(q_{z}+2\beta<0\).
Figure 5. (Color online) Spin-wave spectrum [Eq. (43)] for a polar BEC. Shown are \(E^{2}\) as the function \(\epsilon_{k}\) for (a) \(q_{z}>0\) and \(q_{z}+2\beta>0\), (b) \(q_{z}(q_{z}+2\beta)\leq 0\) and \(q_{z}+\beta>0\), (c) \(q_{z}(q_{z}+2\beta)\leq 0\) and \(q_{z}+\beta\geq 0\), and (d) \(q_{z}<0\) and \(q_{z}+2\beta<0\). We solve k for a given \(E^{2}\), where \(E^{2}>0\) (\(E^{2}<0\)) means a real (pure imaginary) eigenvalue E. Solid circles in (c) and (d) indicate the points \(E =\sqrt{q_{z}(q_{z}+2\beta)}\), where the transmission probability resonantly increases.
It is instructive to discuss the causality of the pure-imaginary-eigenvalue mode mentioned in Sect. 2.3 on the basis of the spectrum shown in Fig. 5(d). For \(0<E<\sqrt{q_{z}(q_{z}+2\beta)}\), \(k_{1}\) and \(k_{2}\) in Eqs. (45) and (46) are both real positive and satisfy \(k_{2}<k_{1}\). Since the group velocity at \(k=k_{1}\) (\(k_{2}\)) is positive (negative), \(e^{ik_{1}x}\) (\(e^{-ik_{2}x}\)) describes a right-going wave. When we continuously change E to pure imaginary, \(k_{1}\) and \(k_{2}\) are analytically continued to \(k_{3}\) and \(k_{4}\), respectively: Equations (48) and (49) are obtained from Eqs. (45) and (46), respectively, by extending the domain of E to a complex value. Hence, it is natural to regard \(e^{ik_{3}x}\) and \(e^{-ik_{4}x}\) having the same causality, obtaining Eqs. (28) and (29), which are the \(q_{z}\to 0\) limit of the present case.
We calculate the reflection and transmission probabilities for each case of (a)–(d) at various energies and momenta of the incident wave and find that the perfect transmission does not occur except for \(q_{z}=0\). Figures 6(a) and 6(b) show the behavior of the coefficients for the cases in Figs. 5(c) and 5(d), respectively, where we set the horizontal axis in the same manner as that in Fig. 3(b), i.e., \(\epsilon_{k_{\text{in}}}\) is zero at the left end of the figure and \(\epsilon_{k_{\text{in}}}\) increases as one goes right. We find that \(|T|^{2}<1\) for all regions of the figure. This result is consistent with the fact that the perfect transmission occurs when the quasiparticle wave function coincides with the condensate wave function: Since the quadratic Zeeman effect breaks the spin rotational symmetry, the spin-wave mode is not the NG mode of the system. On the other hand, one can see that the perfect reflection occurs at the maximum \(|{\mathop{\text{Im}}\nolimits E}|\) in both figures [at \(E=0.25i\) and \(0.15i\) for Figs. 6(a) and 6(b), respectively], indicating that this is a universal property of dynamically unstable modes.
Figure 6. (Color online) Tunneling property of the spin-wave modes in a polar BEC for (a) \(\beta=-0.25\), \(q_{z}=0.1\), \(U_{0}=3\), and \(\sigma=0.5\), and (b) \(\beta=-0.15\), \(q_{z}=-0.1\), \(U_{0}=2\), and \(\sigma=0.5\). The horizontal axis is taken in the same manner as in Fig. 3(b). The region where E is pure imaginary is shaded with gray. The transmission probability \(|T|^{2}\) is always smaller than unity, and its actual value depends on the detail of the barrier potential. At \(E=i|\beta|\) in both panels, the perfect reflection (\(|\tilde{A}|^{2}=1\)) occurs. The transmission probability resonantly increases at \(E=\sqrt{q_{z}(q_{z}+2\beta)}\), which corresponds to \(E=0.2i\) in (a) and \(E=0.2\) in (b).
Note that the transmission probability resonantly increases at \(E=0.2i\) in Fig. 6(a) and at \(E=0.2\) in Fig. 6(b). These points correspond to the incident energy \(E=\sqrt{q_{z}(q_{z}+2\beta)}\) with nonzero \(k_{\text{in}}\), which are depicted in Figs. 5(c) and 5(d) with filled circles. At these points, the momenta of the A and B terms in the asymptotic form become zero. Namely, the propagating modes for \(\mathop{\text{Im}}\nolimits E>0.2\) [\(\mathop{\text{Re}}\nolimits E<0.2\)] change into localized modes for \(\mathop{\text{Im}}\nolimits E<0.2\) [\(\mathop{\text{Re}}\nolimits E>0.2\)] in Fig. 5(c) [5(d)]. The increase in \(|T|^{2}\) at this point is understood as a resonance with these A and B terms. The peak value of the transmission probability depends on the barrier potential and decreases with increasing barrier potential.
4. Real-time Dynamics Calculated Using the Gross–Pitaevskii Equation
We investigate the time-dependent tunneling properties of the spin-wave modes by solving the time-dependent GP equation [Eq. (5)] of the binary BEC. To see the behavior of the Bogoliubov quasiparticles, we calculate the amplitude of the spin-wave mode \((u^{\text{s}},v^{\text{s}})^{\text{T}}=(u_{k}e^{ikx},v_{k}e^{ikx})^{\text{T}}\) given by21) \begin{equation} \Lambda(t) = \int dx\,e^{-ikx}[\bar{\boldsymbol{u}}_{k}{\boldsymbol{\psi}}(x,t)-\bar{\boldsymbol{v}}_{k}{\boldsymbol{\psi}}^{*}(x,t)], \end{equation} (50) where \({\boldsymbol{\psi}}=(\psi_{1}(x,t),\psi_{2}(x,t))^{\text{T}}\) is the time-dependent order parameter, and \(\bar{\boldsymbol{u}}_{k}=(\bar{u}_{k},-\bar{u}_{k})\) and \(\bar{\boldsymbol{v}}_{k}=(\bar{v}_{k},-\bar{v}_{k})\) are the quasiparticle wavefunctions in a uniform system with \((\bar{u}_{k},\bar{v}_{k})\) being the left eigenvector of Eq. (14) with the same eigenvalue E as the right eigenvector \((u_{k},v_{k})^{\text{T}}\). In the case of a uniform system, the Bogoliubov analysis predicts \(\Lambda(t)\approx\Lambda(0)\exp(iEt)\) when \(|\Lambda(t)|\ll 1\).
To see tunneling properties at the barrier potential at \(x=0\), we add a spin-wave fluctuation with the wavenumber \(k_{\text{in}}\) in the \(x<0\) region and calculate the dynamics starting from the following initial state: \begin{align} {\boldsymbol{\psi}}(x,0) &= \frac{\phi(x)}{\sqrt{2}} \begin{pmatrix} 1\\ 1 \end{pmatrix}+ {\boldsymbol{\delta}}{\boldsymbol{\psi}}(x,0)\Theta(-x), \end{align} (51a) \begin{align} {\boldsymbol{\delta}}{\boldsymbol{\psi}}(x) &= \delta\left[u_{k_{\text{in}}} e^{ik_{\text{in}}x} \begin{pmatrix} 1\\ -1 \end{pmatrix}+ v^{*}_{k_{\text{in}}} e^{-ik_{\text{in}}x} \begin{pmatrix} 1\\ -1 \end{pmatrix}\right], \end{align} (51b) where \(\Theta(x)\) is the Heaviside step function and \(|\delta|\ll 1\). Below, we focus on the cases (iii) and (iv) [Eqs. (28) and (29)] and choose \(k_{\text{in}}=-k_{4}\) and \(k_{3}\) with a pure imaginary E. We then investigate the growth of the incident wave, the reflected wave [the A terms in Eqs. (28) and (29)], and the transmitted waves (the T and B terms in the same equations). For example, in the case of \(k_{\text{in}}=-k_{4}\), the fraction of each mode is defined by \begin{align} \Lambda^{\text{{IN}}}(t) &= \frac{2}{L}\int^{0}_{-L/2} dx\,e^{+ik_{4}x}[\bar{\boldsymbol{u}}_{k_{4}}{\boldsymbol{\psi}}(x,t)-\bar{\boldsymbol{v}}_{k_{4}}{\boldsymbol{\psi}}^{*}(x,t)], \end{align} (52a) \begin{align} \Lambda^{\text{{A}}}(t) &= \frac{2}{L}\int^{0}_{-L/2} dx\,e^{+ik_{3}x}[\bar{\boldsymbol{u}}_{k_{3}}{\boldsymbol{\psi}}(x,t)-\bar{\boldsymbol{v}}_{k_{3}}{\boldsymbol{\psi}}^{*}(x,t)], \end{align} (52b) \begin{align} \Lambda^{\text{{T}}}(t) &= \frac{2}{L}\int^{L/2}_{0} dx\,e^{+ik_{4}x}[\bar{\boldsymbol{u}}_{k_{4}}{\boldsymbol{\psi}}(x,t)-\bar{\boldsymbol{v}}_{k_{4}}{\boldsymbol{\psi}}^{*}(x,t)], \end{align} (52c) \begin{align} \Lambda^{\text{{B}}}(t) &= \frac{2}{L}\int^{L/2}_{0} dx\,e^{-ik_{3}x}[\bar{\boldsymbol{u}}_{k_{3}}{\boldsymbol{\psi}}(x,t)-\bar{\boldsymbol{v}}_{k_{3}}{\boldsymbol{\psi}}^{*}(x,t)], \end{align} (52d) where \((\bar{u}_{k_{3}},\bar{v}_{k_{3}})=(-b_{c},a_{c})\) and \((\bar{u}_{k_{4}},\bar{v}_{k_{4}})=(a_{c},b_{c})\) are the left eigenstates corresponding to the right ones given in Eq. (22), and L is the system size. We note that the amplitude of the reflected wave with momentum \(k_{4}\) [the R term in Eqs. (28)] given by \begin{equation} \Lambda^{\text{{R}}}(t) = \frac{2}{L}\int^{0}_{-L/2} dx\,e^{-ik_{4}x}[\bar{\boldsymbol{u}}_{k_{4}}{\boldsymbol{\psi}}(x,t)-\bar{\boldsymbol{v}}_{k_{4}}{\boldsymbol{\psi}}^{*}(x,t)], \end{equation} (52e) satisfies \(|\Lambda^{\text{R}}|=|\Lambda^{\text{IN}}|\) when E is pure imaginary and the added fluctuation is dynamically unstable. The \(\Lambda^{\alpha}(t)\) (α = IN, A, T, and B) for \(k_{\text{in}}=k_{3}\) are defined in a similar manner.
Figure 7 shows the results for (a) an almost perfect transmission regime with \(k_{\text{in}}=-0.11\) and \(E=0.05i\), and (b) a nonperfect transmission regime with \(k_{\text{in}}=0.88\) and \(E=0.05i\), which are close to the \(E\to 0\) limit of cases (iii) and (iv), respectively. For both cases, we use \(\delta=1.0\times 10^{-3}\) and \(L=1600\). Because the system size is much larger than the coherence length, which is unity in the dimensionless form, the effect of the barrier potential on the initial amplitude is small, resulting in \(|\Lambda^{\text{IN}}(0)|\approx\delta\) and \(|\Lambda^{\text{T,A,B}}(0)|\approx 0\). In the early stages of the dynamics, \(|\Lambda^{\text{IN}}(t)|\) can be well approximated by \(|\Lambda^{\text{IN}}(0)|\exp(\mathop{\text{Im}}\nolimits Et)\). With time however, the deviation becomes larger owing to the reflection at the potential barrier. For the cases in Figs. 7(a) and 7(b), the times when the deviation becomes prominent are \(\mathop{\text{Im}}\nolimits Et\sim 2.1\) and 0.9, respectively, which are evaluated from \([|\Lambda^{\text{IN}}(t)|-|\Lambda^{\text{IN}}(0)|e^{\mathop{\text{Im}}\nolimits Et}]/|\Lambda^{\text{IN}}(t)|\approx 0.1\). The fact that the deviation grows more rapidly in Fig. 7(b) is consistent with the fact that the reflection probability is larger for this case [see Fig. 3(b)]. We also note that the amplitudes of the other modes, \(|\Lambda^{\text{T,A,B}}(t)|\), grow faster than \(\exp(\mathop{\text{Im}}\nolimits Et)\), indicating that these modes are generated at the potential barrier as reflected and transmitted waves.
Figure 7. (Color online) Time evolution of the amplitudes of dynamically unstable spin-wave modes [Eq. (52)] in the almost perfect transmission regime (a) and in the nonperfect transmission regime (b), where we start from the initial state given by Eq. (51) with \(k_{\text{in}}=-0.11\) and 0.88, respectively. For both cases, the added fluctuation is an eigenmode of the BdG equation in a uniform system with \(E=0.05i\). The thick dashed curve depicts \(|\Lambda^{\text{IN}}(0)|e^{\mathop{\text{Im}}\nolimits Et}\), which is the growth of \(|\Lambda^{\text{IN}}(t)|\) in the linear regime in the absence of the barrier potential. The effect of the barrier potential is more prominent in (b); the deviations of \(|\Lambda^{\text{IN}}(t)|\) and \(|\Lambda^{\text{IN}}(0)|e^{\mathop{\text{Im}}\nolimits Et}\) grow more rapidly in (b). For all panels, we use \(\beta=-0.2\), \(U_{0}=2\), \(\sigma=0.5\), \(\delta=1.0\times 10^{-3}\), and \(L=1600\).
Next, we investigate the properties of the perfect reflection. Although the origin of the perfect reflection is the destructive interference of the incident and reflected waves, the initial fluctuation exponentially grows once the fluctuation is uniformly added in the \(x<0\) region. We choose \(k_{\text{in}}\) as the wavenumber of the perfect reflection, \(k_{\text{in}}=0.63\) with \(E=0.2i\), and plot the time evolution of \(|\Lambda^{\alpha}(t)|\) (α = IN, A, T, and B) in Fig. 8(a). Since \(k_{3}=k_{4}=k_{\text{in}}\) and the two eigenstates in Eq. (22) are identical to each other, we have \(\Lambda^{\text{IN}}(t)=\Lambda^{\text{A}}(t)\) and \(\Lambda^{\text{T}}(t)=\Lambda^{\text{B}}(t)\). To clarify the consequence of the perfect reflection, we also calculate \(|\Lambda^{\alpha}(t)|\) (α = IN, A, T, and B) for another wavenumber, \(k_{\text{in}}=0.86\) with \(E=0.1i\), whose results are shown in Fig. 8(b). In both Figs. 8(a) and 8(b), \(|\Lambda^{\text{IN}}(t)|\) shows good agreement with \(|\Lambda^{\text{IN}}(0)|\exp(\mathop{\text{Im}}\nolimits Et)\) for a long period. However, there is a significant difference in the growth rate of \(|\Lambda^{\text{T}}(t)|\): \(|\Lambda^{\text{T}}(t)|\) in Fig. 8(a) exhibits an exponential growth that is almost proportional to \(\exp(\mathop{\text{Im}}\nolimits Et)\), whereas in Fig. 8(b), it grows more rapidly than \(\exp(\mathop{\text{Im}}\nolimits Et)\). In the latter case, the amplitudes of the other spin modes, \(|\Lambda^{\text{A}}(t)|\) and \(|\Lambda^{\text{B}}(t)|\), also grow more rapidly than \(\exp(\mathop{\text{Im}}\nolimits Et)\). These results suggest that the fluctuations in the \(x<0\) region are not transmitted to the \(x>0\) region when the perfect reflection occurs. Indeed, the condensate wavefunction near the potential barrier clearly shows this behavior as shown in Figs. 8(c) and 8(d), where we plot the wavefunction \(|\psi_{1,2}(x)|\) at \(\mathop{\text{Im}}\nolimits Et=4\) for the cases in Figs. 8(a) (\(k_{\text{in}}=0.63\)) and 8(b) (\(k_{\text{in}}=0.86\)), respectively. One can see that the initial fluctuation in the \(x>0\) region does not (does) penetrate into the \(x>0\) region in Fig. 8(c) [Fig. 8(d)].
Figure 8. (Color online) (a, b) Time evolution of the amplitudes of the spin-wave modes when the initially added fluctuation is the perfect-reflection mode with \(k_{\text{in}}=0.63\) at \(E=0.2i\) (a) and the ordinary mode with \(k_{\text{in}}=0.86\) at \(E=0.1i\) (b). The growth of \(|\Lambda^{\text{T,B}}(t)|\) in (a) is almost proportional to \(\exp(\mathop{\text{Im}}\nolimits Et)\), whereas that in (b) is more rapid than \(\exp(\mathop{\text{Im}}\nolimits Et)\). This difference is attributed to whether the transmission probability is zero or not (see text). (c, d) Snapshots of condensate wavefunctions around the barrier potential at \(\mathop{\text{Im}}\nolimits Et=4\) for (a) and (b), respectively. In (c), the wavefunction in \(x>0\) is almost not affected by the fluctuation in the \(x<0\) region, whereas in (d), the initial fluctuation in the \(x<0\) region is transmitted to \(x>0\). The other parameters are the same as those used in Fig. 7.
To observe the above results in experiments, a possible initial state instead of Eq. (51) is given by \begin{equation} {\boldsymbol{\psi}}(x,0) = \frac{\phi(x)}{\sqrt{2}} \begin{pmatrix} 1\\ 1 \end{pmatrix}+ \begin{pmatrix} \delta e^{ik_{\text{in}}x}\\ -\delta^{2}/\sqrt{2} \end{pmatrix}\Theta(-x). \end{equation} (53) We can prepare this initial state as follows. First, we prepare a stationary BEC of component 2 under a barrier potential: \({\boldsymbol{\psi}}=\phi(x)(0,1)^{\text{T}}\). Second, we introduce the Rabi coupling between components 1 and 2 for a short period such that their populations become the same: \({\boldsymbol{\psi}}=\phi(x)(1,1)^{\text{T}}/\sqrt{2}\). Third, we introduce the Raman coupling only in the \(x<0\) region for a short period such that a small fraction of atoms of component 2 is transferred to the momentum \(k_{\text{in}}\) state of component 1,28) obtaining the state given by Eq. (53). We have numerically confirmed that the dynamics starting from Eq. (53) is qualitatively the same as those in Figs. 7 and 8. Namely, the nonzero reflection coefficient R makes \(|\Lambda^{\text{IN}}(t)|\) deviate from the exponential growth \(\exp(\mathop{\text{Im}}\nolimits Et)\), and the fluctuation in the \(x<0\) region cannot penetrate into the \(x>0\) region at the perfect reflection. However, it would be difficult to observe the growth of spin waves in the \(E\to 0\) limit because their growth is very slow. The dynamically unstable modes with larger \(\mathop{\text{Im}}\nolimits E\) grow faster, seeded from thermal and quantum fluctuations, and the initially added fluctuation will be smeared out. On the other hand, since the perfect reflection occurs for the mode with the largest \(\mathop{\text{Im}}\nolimits E\), its rapid growth and the absence of the transmitted wave would be observable in experiments.
5. Conclusion
We have studied the tunneling properties of the spin-wave mode in a dynamically unstable BEC and numerically shown that the perfect transmission occurs even in the presence of dynamical instability if the spin-wave mode is the NG mode. When the prepared BEC is dynamically unstable, the eigenvalue E of the BdG equation can be complex. In the models we discussed, E becomes pure imaginary for the incident momentum in a specific region and becomes zero when it changes from real to pure imaginary. We have found that the perfect transmission occurs in the limit that both the eigenvalue E and the momentum of the injected quasiparticle go to zero. This is the condition for the quasiparticle to have the same form as the condensate wave function. When the above condition is satisfied, even a dynamically unstable mode that has a pure imaginary E exhibits the perfect transmission in the limit of \(E\to 0\). On the other hand, even when \(E=0\), the perfect transmission does not occur when the incident momentum is nonzero.
Apart from the perfect transmission at \(E\to 0\), we have also clarified that the perfect reflection occurs at a point where \(|{\mathop{\text{Im}}\nolimits E}|\) becomes maximum. Around the maximum of \(|{\mathop{\text{Im}}\nolimits E}|\), there is a reflected wave that has a different momentum from the incident one. Upon the occurrence of the perfect reflection, the incident wave destructively interferes with this reflected wave, and the quasiparticle wave function is strongly suppressed. We have also found that the transmission probability resonantly increases when the reflected wave changes to a bound state around the potential barrier.
The numerical simulation by solving the time-dependent GP equation revealed that the tunneling properties affect the growth of the spin-wave mode initially added on one side of the barrier potential. We find that the perfect-transmission mode exhibits exponential growth with growth time \((\mathop{\text{Im}}\nolimits E)^{-1}\) for a long period, which is however difficult to observe experimentally owing to the growth of other more unstable modes. On the other hand, the signature of the perfect reflection is that the initial fluctuation on one side cannot penetrate into the other side, which would be observable as it occurs for the most unstable mode.
Acknowledgment
The authors thank Kazuya Fujimoto, Shun Tamura, and Ryoi Ohashi for fruitful discussions. This work was supported by JST-CREST (Grant No. JPMJCR16F2) and JSPS KAKENHI (Grants Nos. JP18K03538 and JP19H01824).
Appendix:
Normalization Condition for the Bosonic BdG Equation
The left and right eigenvectors of a non-Hermitian matrix H are defined by \begin{align} H|w_{n}^{\text{R}}\rangle &= E_{n}|w_{n}^{\text{R}}\rangle, \end{align} (A·1a) \begin{align} \langle w_{n}^{\text{L}}|H &= \langle w_{n}^{\text{L}}|E_{n}. \end{align} (A·1b) The second equation can be rewritten as \begin{equation} H^{\dagger}|w_{n}^{\text{L}}\rangle = E_{n}^{*}|w_{n}^{\text{L}}\rangle, \end{equation} (A·1c) where \(|w_{n}^{\text{L}}\rangle\) is the Hermite conjugate of \(\langle w_{n}^{\text{L}}|\). When we multiply \(\langle w_{m}^{\text{L}}|\) to Eq. (A·1a) from the left, we obtain \begin{align} \langle w_{m}^{\text{L}}|H|w_{n}^{\text{R}}\rangle = \langle w_{m}^{\text{L}}|E_{n}|w_{n}^{\text{R}}\rangle &= \langle w_{m}^{\text{L}}|E_{m}|w_{n}^{\text{R}}\rangle, \end{align} (A·2) \begin{align} (E_{m}-E_{n})\langle w_{m}^{\text{L}}|w_{n}^{\text{R}}\rangle &= 0, \end{align} (A·3) from which the orthonormal condition is given by \begin{equation} \langle w_{m}^{\text{L}}|w_{n}^{\text{R}}\rangle = \delta_{mn}. \end{equation} (A·4)
In the case of a bosonic BdG equation, the matrix H satisfies the pseudo-Hermiticity and the particle–hole symmetry: \begin{align} \tau_{z} H \tau_{z} &= H^{\dagger}, \end{align} (A·5) \begin{align} \mathcal{C}^{-1} H \mathcal{C} &= - H, \end{align} (A·6) where \(\mathcal{C}\equiv\tau_{x} K\) is the particle–hole operator with K being the complex-conjugate operator and \(\tau_{x,y,z}\) the Pauli matrices in the Nambu space. From Eqs. (A·1c) and (A·5), we obtain \begin{equation} H\tau_{z}|w_{n}^{\text{L}}\rangle = E_{n}^{*} \tau_{z}|w_{n}^{\text{L}}\rangle, \end{equation} (A·7) which leads to \(|w_{n}^{\text{L}}\rangle\propto \tau_{z}|w_{n}^{\text{R}}\rangle\) for real \(E_{n}\). From Eqs. (A·1a) and (A·6), we obtain \begin{equation} H\mathcal{C}|w_{n}^{\text{R}}\rangle = - \mathcal{C}H|w_{n}^{\text{R}}\rangle = - E_{n}^{*}\mathcal{C}|w_{n}^{\text{R}}\rangle. \end{equation} (A·8)
When \(E_{n}\) is real, \(|w_{n}^{\text{R}}\rangle\) and \(\mathcal{C}|w_{n}^{\text{R}}\rangle\) are a particle–hole pair and satisfy \begin{equation} \langle w_{n}^{\text{R}}|\mathcal{C}\tau_{z} \mathcal{C} |w_{n}^{\text{R}}\rangle = - \langle w_{n}^{\text{R}}|\tau_{z}|w_{n}^{\text{R}}\rangle, \end{equation} (A·9) which implies \(|w_{n}^{\text{L}}\rangle=\tau_{z}|w_{n}^{\text{R}}\rangle\) (\(|w_{n}^{\text{L}}\rangle=-\tau_{z}|w_{n}^{\text{R}}\rangle\)) for a particle (hole) mode. We therefore define the normalization constant for a real-eigenvalue mode as \begin{equation} \langle w_{n}^{\text{R}}|\tau_{z}|w_{n}^{\text{R}}\rangle = 1\ \text{or}\ -1. \end{equation} (A·10)
In the case when \(\mathop{\text{Im}}\nolimits E_{n}\neq 0\), there exists \(n'\) such that \(E_{n'}=E_{n}^{*}\), \(|w_{n'}^{\text{R}}\rangle\propto \tau_{z} |w_{n}^{\text{L}}\rangle\), and \(|w_{n}^{\text{R}}\rangle\propto \tau_{z} |w_{n'}^{\text{L}}\rangle\) [see Eq. (A·7)]. Then, the normalization condition is given by \(|\langle w_{n'}^{\text{R}}|\tau_{z}|w_{n}^{\text{R}}\rangle|=1\). As a special case, when the matrix elements of H are all real, we obtain \(|w_{n'}^{\text{R}}\rangle=(|w_{n}^{\text{R}}\rangle)^{*}\), from which the normalization condition is given by \begin{equation} |(\langle w_{n}^{\text{R}}|)^{*}\tau_{z}|w_{n}^{\text{R}}\rangle| = 1. \end{equation} (A·11)
The BdG equations discussed in this paper have only real (\(\in\mathbb{R}\)) or pure imaginary (\(\in i\mathbb{R}\)) eigenvalues. When the BdG equation is written in the Fourier space as in Eq. (14), the normalization condition is given by \begin{align} |u|^{2}-|v|^{2} &= \pm 1\quad (E\in \mathbb{R}), \end{align} (A·12a) \begin{align} u^{2}-v^{2} &= \pm 1\quad (E\in i\mathbb{R}), \end{align} (A·12b) where the second equation also determines the phase of the eigenvector. When the BdG equation is written in the coordinate space as in Eqs. (13) and (41), Eq. (A·3) means that \begin{equation} \int_{-\infty}^{\infty} [|u(x)|^{2}-|v(x)|^{2}]\,dx \end{equation} (A·13) can be regarded as a norm for \(E\in\mathbb{R}\), whereas it always vanishes for \(E\in\mathbb{C}\).
References
1 P. W. Anderson, Basic Notions of Condensed Matter Physics (Addison-Wesley, Boston, MA, 1983) Vol. 55. Google Scholar
2 D. Kovrizhin, Phys. Lett. A 287, 392 (2001). 10.1016/S0375-9601(01)00503-5 Crossref, Google Scholar
3 Y. Kagan, D. L. Kovrizhin, and L. A. Maksimov, Phys. Rev. Lett. 90, 130402 (2003). 10.1103/PhysRevLett.90.130402 Crossref, Google Scholar
4 I. Danshita, N. Yokoshi, and S. Kurihara, New J. Phys. 8, 44 (2006). 10.1088/1367-2630/8/3/044 Crossref, Google Scholar
5 Y. Ohashi and S. Tsuchiya, Phys. Rev. A 78, 043601 (2008). 10.1103/PhysRevA.78.043601 Crossref, Google Scholar
6 D. Takahashi and Y. Kato, J. Phys. Soc. Jpn. 78, 023001 (2009). 10.1143/JPSJ.78.023001 Link, Google Scholar
7 Y. Kato, H. Nishiwaki, and A. Fujita, J. Phys. Soc. Jpn. 77, 013602 (2008). 10.1143/JPSJ.77.013602 Link, Google Scholar
8 S. Watabe, Y. Kato, and Y. Ohashi, Phys. Rev. A 83, 033627 (2011). 10.1103/PhysRevA.83.033627 Crossref, Google Scholar
9 S. Watabe and Y. Kato, Phys. Rev. A 83, 053624 (2011). 10.1103/PhysRevA.83.053624 Crossref, Google Scholar
10 S. Watabe, Y. Kato, and Y. Ohashi, Phys. Rev. A 84, 013616 (2011). 10.1103/PhysRevA.84.013616 Crossref, Google Scholar
11 S. Tsuchiya and Y. Ohashi, Phys. Rev. A 78, 013628 (2008). 10.1103/PhysRevA.78.013628 Crossref, Google Scholar
12 S. Tsuchiya and Y. Ohashi, Phys. Rev. A 79, 063619 (2009). 10.1103/PhysRevA.79.063619 Crossref, Google Scholar
13 S. Watabe and Y. Kato, Phys. Rev. A 78, 063611 (2008). 10.1103/PhysRevA.78.063611 Crossref, Google Scholar
14 S. Watabe, Y. Kato, and Y. Ohashi, Phys. Rev. A 86, 023622 (2012). 10.1103/PhysRevA.86.023622 Crossref, Google Scholar
15 S. Watabe and Y. Kato, Europhys. Lett. 111, 50008 (2015). 10.1209/0295-5075/111/50008 Crossref, Google Scholar
16 T. Nakayama, I. Danshita, T. Nikuni, and S. Tsuchiya, Phys. Rev. A 92, 043610 (2015). 10.1103/PhysRevA.92.043610 Crossref, Google Scholar
17 T. Nakayama and S. Tsuchiya, Phys. Rev. A 100, 063612 (2019). 10.1103/PhysRevA.100.063612 Crossref, Google Scholar
18 L. Fallani, L. De Sarlo, J. E. Lye, M. Modugno, R. Saers, C. Fort, and M. Inguscio, Phys. Rev. Lett. 93, 140406 (2004). 10.1103/PhysRevLett.93.140406 Crossref, Google Scholar
19 Y. Shin, M. Saba, M. Vengalattore, T. A. Pasquini, C. Sanner, A. E. Leanhardt, M. Prentiss, D. E. Pritchard, and W. Ketterle, Phys. Rev. Lett. 93, 160406 (2004). 10.1103/PhysRevLett.93.160406 Crossref, Google Scholar
20 L. Sadler, J. Higbie, S. Leslie, M. Vengalattore, and D. Stamper-Kurn, Nature 443, 312 (2006). 10.1038/nature05094 Crossref, Google Scholar
21 Y. Kawaguchi and M. Ueda, Phys. Rep. 520, 253 (2012). 10.1016/j.physrep.2012.07.005 Crossref, Google Scholar
22 H.-J. Miesner, D. M. Stamper-Kurn, J. Stenger, S. Inouye, A. P. Chikkatur, and W. Ketterle, Phys. Rev. Lett. 82, 2228 (1999). 10.1103/PhysRevLett.82.2228 Crossref, Google Scholar
23 E. Nicklas, M. Karl, M. Höfer, A. Johnson, W. Muessel, H. Strobel, J. Tomkovič, T. Gasenzer, and M. K. Oberthaler, Phys. Rev. Lett. 115, 245301 (2015). 10.1103/PhysRevLett.115.245301 Crossref, Google Scholar
24 J. H. Kim, S. W. Seo, and Y. Shin, Phys. Rev. Lett. 119, 185302 (2017). 10.1103/PhysRevLett.119.185302 Crossref, Google Scholar
25 C. J. Pethick and H. Smith, Bose–Einstein Condensation in Dilute Gases (Cambridge University Press, Cambridge, U.K., 2008) Vol. 164. Crossref, Google Scholar
26 L. Pitaevskii and S. Stringari, Bose–Einstein Condensation and Superfluidity (Oxford University Press, Oxford, U.K., 2016) Vol. 164. Crossref, Google Scholar
27 A. Tokuno and S. Uchino, Phys. Rev. A 87, 061604 (2013). 10.1103/PhysRevA.87.061604 Crossref, Google Scholar
28 Y.-J. Lin, R. L. Compton, K. Jiménez-García, J. V. Porto, and I. B. Spielman, Nature 462, 628 (2009). 10.1038/nature08609 Crossref, Google Scholar