Welcome to visit Communications in Theoretical Physics,
Condensed Matter Theory

Kármán vortex street in Bose–Einstein condensate with PT symmetric potential

  • Kaihua Shao 1 ,
  • Baolong Xi 1 ,
  • Zhonghong Xi 1, 2 ,
  • Pu Tu 1, 3 ,
  • Jinping Ma 1 ,
  • Xi Zhao , 1 ,
  • Hongjuan Meng 1, ,
  • Yuren Shi , 1,
Expand
  • 1College of Physics and Electronic Engineering, Northwest Normal University, Lanzhou, 730070, China
  • 2College of Physics and Hydropower Engineering, Gansu Normal College For Nationalities, Hezuo 747000, China
  • 3College of Intelligent Manufacturing, Sichuan University of Arts and Science, DaZhou, 635000, China

Authors to whom any correspondence should be addressed.

Received date: 2024-06-07

  Revised date: 2024-09-30

  Accepted date: 2024-10-16

  Online published: 2024-12-04

Copyright

© 2024 Institute of Theoretical Physics CAS, Chinese Physical Society and IOP Publishing. All rights, including for text and data mining, AI training, and similar technologies, are reserved.

Abstract

Kármán vortex street not only exists in nature, but also widely appears in engineering practice, which is of great significance for understanding superfluid. Parity-time (PT) symmetric potential provides a good platform for the study of Kármán vortex streets. In this paper, different patterns of vortex shedding formed behind PT symmetric potential in Bose–Einstein condensate (BEC) are simulated numerically. Kármán vortex streets and others are discovered to emerge in the wake of a moving obstacle with appropriate parameters. Compared with BEC without PT symmetric potential, the frequency and amplitude of the drag force are more complex. The parametric regions of the combined modes are scattered around the Kármán vortex street. Numerical simulations indicate that the imaginary part of the PT symmetric potential affects the vortex structure patterns. Finally, we proposed an experimental protocol that may observe a Kármán vortex street.

Cite this article

Kaihua Shao , Baolong Xi , Zhonghong Xi , Pu Tu , Jinping Ma , Xi Zhao , Hongjuan Meng , Yuren Shi . Kármán vortex street in Bose–Einstein condensate with PT symmetric potential[J]. Communications in Theoretical Physics, 2025 , 77(3) : 035701 . DOI: 10.1088/1572-9494/ad873e

1. Introduction

Bose–Einstein condensate (BEC) is an interesting and important phenomenon that occurs in a certain system of most atoms at ultra-low temperatures. It is a unique state of matter that was first predicted by Bose and Einstein [1]. Wieman and Cornell observed BEC in 87Rb atomic gases using laser cooling and evaporative cooling techniques in magnetic traps [2]. Later, Ketterle’s group also achieved BEC of 23Na in an experiment [3]. A team led by Jin realized the condensation of 40K in 2004 [4]. Then BEC provides an ideal platform for studying superfluid phenomena [57].
A Kármán vortex street is a well-known phenomenon in classical fluid systems. When the fluid flows past an obstacle, vortex pairs with opposite rotating directions and regular arrangement are formed, and this wake is called a Kármán vortex street [811]. As a long-lived topological excitation, it has a vital effect on understanding the properties of superfluids [12]. The phenomenon of Kármán vortex streets has also been found in quantum fluids [13]. In 2010, Sasaki et al studied a Kármán vortex street formed by moving obstacle potential in the scalar BEC [14]. Saito et al found that the vibration of the obstacle potential plays a modulating role in the formation of vortex streets in 2013 [15]. In the experiment, Kwon et al observed different wake patterns in oblate atomic BEC under various velocities of the obstacle [16]. Furthermore, the Kármán vortex street in a two-component and spin-orbit coupled BEC is investigated as well [17, 18].
Parity-time (PT) symmetry was initially proposed by Bender and Boettcher in 1998, and it has garnered significant attention [19]. Their result shows that the energy spectrum associated with non-Hermitian operators maintains a fully real character [2022]. Usually, an open quantum system takes into account the the imaginary part of the external potential to contribute a gain and loss model [23]. The atoms loss can be induced by a focused electron beam [24] while atoms are added by letting them fall into the condensate from a second condensate [25]. For satisfying the necessary but not sufficient condition of a PT symmetric potential. Its real part exhibits even symmetry, whereas the imaginary part is an odd one [26]. In addition, the similarity between the Schr $\ddot{{\rm{o}}}$ dinger equation in quantum mechanics and the mean-field Gross–Pitaevskii (GP) equation sparked study into BEC under the influence of PT symmetric potentials [27, 28]. In 2017, Lukas Schwarz et al investigated vortex excitations in a dilute BEC with the complex PT symmetric potential [29]. Haag et al studied nonlinear quantum dynamics in a double well potential satisfying PT-symmetry, where particles are injected into one well and removed from the other [30, 31]. Recently, researchers have delved deeply into the study of vortex and solitons excitations caused by PT symmetric potentials in BEC systems [3234]. However, the Kármán vortex street is rarely studied in a system with PT symmetric potential. Thus, it is worth investigating the generation of the Kármán vortex street and exploring the impact of gain-loss strength on its stability in system.
In this paper, we numerically found that the Kármán vortex street can be generated when the width and moving velocity of the obstacle potential are appropriate. The numerical simulation results also include typical combined modes. When the Kármán vortex street is generated, the amplitude of the drag force acting on the obstacle potential increases first and then decreases. By analyzing the influence of the imaginary part of the PT symmetric potential on the motion trajectory of vortex pairs behind the obstacle potential, it is found that the motion trajectory of the vortex pairs gradually changes from V-shaped to the Kármán vortex street. Consequently, it is worth investigating the generation of the Kármán vortex street and exploring the impact of gain-loss strength.

2. Theoretical method

The Hamiltonian of the BEC with PT symmetric potential can be written as [3537]
$\begin{eqnarray}H=\int {\rm{d}}{\boldsymbol{r}}{\psi }^{\dagger }\left(-\displaystyle \frac{{{\hslash }}^{2}}{2m}{{\rm{\nabla }}}^{2}+V+\displaystyle \frac{1}{2}g{\psi }^{\dagger }\psi \right)\psi ,\end{eqnarray}$
where ψ = ψ(r, t) denotes the wave function of the condensate. r = (x, y, z) represents the Cartesian coordinate, t indicates time. ∇2 is a three-dimensional Laplacian operator. The PT symmetric potential V = VR + iVI, where VR and VI are the real and imaginary parts of the external potential. VI can influence the atomic distribution in the BEC. Total particles are N = ∫∣ψ2dr. g = 4π2as/m is the interaction strength between atoms, as stands for the s-wave scattering length [3840].
In the experiments, BEC is usually confined within a harmonic potential ${V}_{\mathrm{har}}=\tfrac{1}{2}m\left({\omega }_{x}^{2}{x}^{2}+{\omega }_{y}^{2}{y}^{2}+{\omega }_{z}^{2}{z}^{2}\right)$, here ωx, ωy and ωz denote the respective frequencies of the potential along x, y and z directions. When confinement is stronger in the z-direction than in the other two directions, the energy scale of ωz is much larger than ωx and ωy [4145]. Thus the ‘disk-shaped’ condensate is formed in the xy plane. Under which, the wave function is reasonably represented as $\psi ({\boldsymbol{r}},t)=\varphi (x,y,t){\phi }_{0}(z){{\rm{e}}}^{-{\rm{i}}{\omega }_{z}t/2}$, indicating that the motion of atoms along the z-direction is confined to the ground state which is expressed as ${\phi }_{0}{(z)=(m{\omega }_{z}/\pi {\hslash })}^{1/4}{{\rm{e}}}^{-m{\omega }_{z}{z}^{2}/2{\hslash }}$.
In the following, we introduce the characteristic time ${t}_{0}={\omega }_{0}^{-1}$ and characteristic length ah0 to facilitate the research. The corresponding variables in equation (1) are normalized by $\tilde{t}={\omega }_{0}t$, $\tilde{x}=x/{a}_{{\rm{h}}0}$, $\tilde{y}=y/{a}_{{\rm{h}}0}$, $\tilde{\psi }\,=\psi /\sqrt{{n}_{0}/{a}_{{\rm{h}}0}^{2}}$, $\tilde{g}=4\pi {n}_{0}{a}_{s}/\sqrt{2\pi }{l}_{z}$, $\tilde{V}=V/{\hslash }{\omega }_{0}$ with ${\omega }_{0}\,=\min ({\omega }_{x},{\omega }_{y})$, ${l}_{z}=\sqrt{{\hslash }/m{\omega }_{z}}$ and ${a}_{{\rm{h}}0}=\sqrt{{\hslash }/m{\omega }_{0}}$. n0 represents the characteristic particle number density. For simplicity, The tilde ‘∼’ has been omitted from all the variables. Therefore, the dimensionless quasi-two-dimensional GP equation is derived as
$\begin{eqnarray}{\rm{i}}\displaystyle \frac{\partial \psi }{\partial t}=\left[-\displaystyle \frac{1}{2}{{\rm{\nabla }}}^{2}+V+g| \psi {| }^{2}\right]\psi .\end{eqnarray}$
The ground state is obtained by adopting the initial state through the use of the imaginary-time propagation method. The initial guess is assumed to be ψ = 1 and Γ is set to be zero. Without loss of generality, we take g = 1. Otherwise, it can be reduced to 1 by appropriate mathematical transformation. The symmetry of the system is intentionally disturbed by introducing a small stochastic perturbation. The time-splitting Fourier pseudo-spectral method is employed to numerically solve equation (2) under periodic boundary conditions [46]. The spatial domain is truncated to [−256, 256] ⨂[−64, 64] and discretized uniformly into 2048 × 512 grids.

3. Numerical results

To satisfy the condition that the real part of the PT symmetric potential is an even function and the imaginary part is an odd one, we choose the real part to be the Gaussian obstacle potential, which is given by [14]
$\begin{eqnarray}{V}_{{\rm{R}}}={V}_{0}\left[{{\rm{e}}}^{-\tfrac{{\left(x-{x}_{0r}+\upsilon t\right)}^{2}+{\left(y+{y}_{0r}\right)}^{2}}{{d}^{2}}}+{{\rm{e}}}^{-\tfrac{{\left(x+{x}_{0r}-\upsilon t\right)}^{2}+{\left(y-{y}_{0r}\right)}^{2}}{{d}^{2}}}\right],\end{eqnarray}$
where V0 is the peak strength. (x0r, y0r) represents the center position of VR at t = 0. The imaginary part of the PT symmetric potential is adopted as [29]
$\begin{eqnarray}{V}_{{\rm{I}}}={\rm{\Gamma }}\left[{{\rm{e}}}^{-\tfrac{{\left(x-{x}_{0i}\right)}^{2}+{\left(y+{y}_{0i}\right)}^{2}}{{d}^{2}}}-{{\rm{e}}}^{-\tfrac{{\left(x+{x}_{0i}\right)}^{2}+{\left(y-{y}_{0i}\right)}^{2}}{{d}^{2}}}\right].\end{eqnarray}$
Here, Γ is the gain-loss strength. (±x0i, ± y0i) expresses the center position that remains stationary during the dynamical evolution. d denotes the width of VI and VR. The harmonic potential is no longer considered here.
In practice, the Gaussian obstacle potential VR centered at (x0r, y0r) = (204.8, 25) begins to move along −x direction with $v$, while another obstacle moves along the x direction. We investigate various wake patterns by adjusting d and $v$ of the obstacle potential. Figure 1 denotes the wake pattern of the drifting vortex dipoles and V-shaped vortex pairs. The direction of the moving obstacle potential is indicated by the black arrows. Vortex pairs will periodically shed behind the obstacle when $v$ exceeds a critical value [47]. The width of the obstacle and the interaction between atoms affects the critical velocity. Vortex pair shedding reduces the primary velocity to below the critical velocity. After a brief period, the primary velocity field is larger than the critical velocity again, causing a new vortex pair to be shed. When d = 0.5 and $v$ = 0.56, the drifting vortex dipoles are generated behind the obstacle potential (see in figure 1(a)). This is a new wake pattern in the system. It is specified that + represents the clockwise rotation and − symbolizes the counterclockwise rotation. Two vortices exhibit opposite circulation ±2π/m, which is consistent with dimensionless circulation ±2π. The vortex dipoles undergo drift and it is evident that their centers are not distributed along a single line. This indicates that the vortex pairs are not stable and the local density distribution of the particle is unequal in the system. In figure 1(b), vortex-antivortex pairs are emitted alternately behind the obstacle when $v$ increases to 0.65. The distance between two point vortices keeps almost unvaried in the process of dynamic evolution. The vortex-antivortex pairs are unstable and form the V-shaped vortex pair.
Figure 1. Density (left panel) and phase (right panel) distribution of the condensate at different velocity $v$ for d = 0.5. (a) $v$ = 0.56. (b) $v$ = 0.65. The direction of the moving obstacle potential is indicated by the black arrows. The + indicates clockwise circulation of the vortex and − represents counterclockwise circulation.
When the velocity of the obstacle potential is 0.49 and the width increases from 1.8 to 2.1, Kármán vortex street and irregular turbulence appear behind the obstacle potential, as shown in figure 2. Although there is gain and loss of atoms in the system, vortex pairs are alternately and periodically released from the obstacle potential. The upper and lower vortex pairs have opposite circulations ±2π. The wake eventually becomes the Kármán vortex street (see figure 2(a)). The angular frequency of the vortex is 2/(mD2) [48], where D refers to the distance between two vortices center. The average distance between two rows of vortex pairs is b ≈ 29.25, and between adjacent vortex pairs in one row is l ≈ 135.50. So the ratio is b/l ≈ 0.22. In [49], the stability condition for generating the Kármán vortex street in classical fluids satisfies 0.12 < b/l < 0.4. Therefore, our result falls into this range and agrees well with the conclusion. If d increases to 2.1, while $v$ remains at 0.49, vortex pairs form irregular turbulence (see figure 2(b)). It is found that the wake tends to be irregular turbulence when the width of the obstacle potential increases.
Figure 2. The density (left panel) and phase (right panel) distribution of the condensate at different d for $v$ = 0.49. (a) d = 1.8. (b) d = 2.1. The black straight arrows show the direction in which the obstacle is moving. The black curved arrows represent the direction of circulations of vortex pairs.
A variety of combined modes are also present in the system with PT symmetric potentials. Figure 3 plots V-shaped vortex pairs and irregular turbulence, which are combined with the Kármán vortex street, respectively. Figure 3(a) represents the combined mode of the Kármán vortex street and V-shaped vortex pairs. Unlike the case of the V-shaped vortex pairs (see figure 1(b)), the distance between two adjacent vortex pairs becomes larger. We note that the Kármán vortex street has been formed when the obstacle moves along −x direction. In figure 3(b), we find that irregular turbulence exists when the obstacle moves in the x direction. In figure 3, the ratio is b/l ≈ 0.20 with b ≈ 28.75 and l ≈ 146.63 on average, which is within the range of stability condition. It is worth noting that the combined modes have not occurred in the scalar BEC system which is without PT symmetric potential. The combined mode is asymmetric because of the broken local symmetry caused by gain and loss of atoms. One can come to the conclusion that the density distribution of particles has important effects on the local symmetry of the system.
Figure 3. Density (left panel) and phase (right panel) distribution of combined modes. (a) (d, $v$) = (1.7, 0.45), (b) (d, $v$) = (1.7, 0.51).
The drag force experienced by the moving obstacle potential is defined as f = (fx, fy) = i∂tψ*ψdxdy. Figure 4 indicates a normalized instantaneous drag force when the vortex is released from the obstacle. Figure 4(a) illustrates the drag force corresponding to the drifting vortex dipoles. fx is the shedding frequency of vortices. fy represents the frequency at which the vortex pair moves in an upward or downward direction. It is noticed that the drag force oscillates periodically and the amplitudes are not equal. As the vortex pairs behind the obstacle do not shed simultaneously, and the frequencies of the motion of the vortex pairs along the y direction are inconsistent. Figure 4(b) describes the drag force of the V-shaped vortex pairs plotted in figure 1(b). When the vortex pairs shed, the oscillating trajectory of the drag force has a certain periodicity. fx exhibits periodic oscillations within the range of approximately 388.18 to 514.84 for t. It indicates that the vortex pairs, which are generated by the moving obstacle potentials along both the x direction and the −x direction, simultaneously shed off in the same direction. Figure 4(c) corresponds to the Kármán vortex street. fy oscillates periodically, and it is noteworthy that the amplitude of fy tends to increase then decrease. When the shedding of vortices is irregular turbulence (see figure 2(b)), the drag force has no obvious periodicity, as shown in figure 4(d). Figures 4(e) and (f) correspond to combined modes of the Kármán vortex street and V-shaped vortex pair (see figure 3(a)), and the Kármán vortex street and irregular turbulence (see figure 3(b)). It is found that fy has obvious periodicity, indicating that there exists periodic shedding vortex pairs. One can observe a clear difference in the frequency and magnitude between this system and the scalar BEC lacking PT-symmetry potential.
Figure 4. The variation of the instantaneous drag force with vortex shedding. The blue dashed line and the orange solid line correspond to fx and fy.
Figure 5 depicts the parametric region of different wake patterns. When the moving velocity of the obstacle potential is low, a laminar flow is generated. Irregular turbulence generally occurs when the velocity is larger. There is a green area indicating the Kármán vortex street, in which the range of width and velocity is 1.1 < d < 3.2 and 0.39 < $v$ < 0.54. As opposed to other vortex patterns, it is noticed that the combined modes (red area in figure 5) are scattered around the range of the Kármán vortex street.
Figure 5. Parameter regions for different wake patterns. The purple area is consistent with the stable laminar flow. The blue, light blue, green, yellow and red regions correspond to the patterns shown in figures 1, 2 and 3.
Figure 6 shows dynamic evolutions of the vortex structures with various gain-loss strengths. For Γ = 0.1, the persistently stable Kármán vortex street appears, as shown in figure 6(a). As Γ increases to 0.3, the vortex street tends to disappear in figure 6(b). It is also observed that irregular vortices are formed when the wake is close to the obstacle potential. In figure 6(c), the distances between two point vortices in partial vortex pairs increase at Γ = 0.5 and the vortices become irregular. As the gain-loss strength Γ increases, the vortex structure pattern changes.
Figure 6. Density distribution for d = 1.6, $v$ = 0.48. (a) Γ = 0.1, (b) Γ = 0.3, (c) Γ = 0.5.
To better understand the impact of the imaginary part of the PT symmetric potential on the evolution of the Kármán vortex street, figure 7 shows the probability current density vector field behind the obstacle potential. The wake generated by the obstacle potential moving in the x direction is similar to that produced by one moving in the opposite direction. Thus, ignoring the situation where obstacles move along the x direction. At t = 360, the obstacle potential is far from the imaginary part (in figure 7(a)). That is, the vortex pairs are almost unaffected by the gain-loss strength, forming V-shaped vortex pairs. It is evident that the circulation between the vortex pairs is opposite. Figure 7(b) exhibits the wake when the obstacle enters the influence region of the imaginary part. The atomic number is uneven, breaking the symmetry within this range and leading to the probability current density vector field changes. Furthermore, it alters the track of vortices, evolving the V-shaped vortex pair into the Kármán vortex street. As time evolution goes on, the obstacle moves away from the position of the imaginary part, while the gain-loss strength exerts a diminished influence on vortices. The probability current density vector field keeps almost unchanged behind the obstacle. Eventually, the Kármán vortex is generated, as shown in figure 7(c).
Figure 7. The probability current density vector field distribution at different times for d = 1.6, $v$ = 0.48. The obstacle potential moves along the −x direction. The arrows denote the direction of the probability current density vector field.
For a deeper understanding of how the PT symmetric potential affects the system dynamics, figure 8 focuses on the density distribution of the Kármán vortex street generated in systems with or without PT symmetric potential for d = 1.8 and $v$ = 0.49. The black dashed circle outlines the Kármán vortex street formed after the moving obstacle. Figure 8(a) depicts vortex pairs alternately released from the obstacle potential in a scalar BEC without PT symmetric potential (the model is same as in [14]), forming a neatly Kármán vortex street. Figures 8(b) and (c) shows the density distribution for our model with Γ = 0 or Γ = 0.1. As can be seen from figure 8(b), the short segment of Kármán vortex street is formed behind the obstacle potential moving along the x or −x directions, respectively. Figure 8(c) demonstrates the Kármán vortex street generated behind the PT symmetric potential when gain-loss strength Γ = 0.1. During the same evolutionary period, the number of vortex pairs constituting the Kármán vortex street in figure 8(c) is the least compared to figures 8(a) and (b), implying that the imaginary part of the PT symmetric potential is unfavorable for the formation of the Kármán vortex street.
Figure 8. The density distribution of the condensate past the obstacle potential with or without PT symmetric potential. (a) Kármán vortex street formed in a scalar BEC. The model is the same as that in [14]. (b) Γ = 0 for our model. (c) Γ = 0.1 for our model. (d = 1.8, $v$ = 0.49).
In order to observe the Kármán vortex street with PT symmetric potential in experiment, we propose a potential approach for implementation. We consider that the BEC is composed of 87Rb atoms, and its hyperfine state is ∣f = 1, mf = –1⟩ [50] with m ≈ 1.45 × 10−25 kg. The total number of atoms in condensed matter is N ≈ 6.55 × 106(n0 = 100). The condensate is confined in a harmonic potential with (ωx, ωy, ωz) ≈ 2π × (43, 56, 350) Hz [51]. Thus ω0 = 2π × 43 Hz and the oscillator length is ${a}_{\mathrm{ho}}=\sqrt{{\hslash }/m{\omega }_{x}}\approx 3.6\,{\rm{\mu }}$m. We adjust as to be approximately equal to 54.22aB [52]. Here, aB is the Bohr radius. The interaction strength between atoms g ≈ 2.74 × 10−51 J. The repulsive laser beam excites Gaussian obstacle potential along the z-direction [5355]. It moves with $v$ ≈ 0.099 mms−1 from the initial position (x0, y0) = (0.74, 0.09) mm along the −x direction and (x0, y0) = ( − 0.74, − 0.09) mm along the opposite direction, respectively. The peak intensity is V0 ≈ 5.88 × 10−31 J. The width of the PT symmetric potential is d ≈ 6.47μm and the gain-loss strength is Γ ≈ 5.88 × 10−34 J. In the experiment, an atomic laser injects atoms into a condensate to obtain gain [56]. By utilizing a laser beam to excite atoms into excited states and subsequently use photon recoil to eject them from the condensate [25]. Thus, the parameters are approximate figure 2(a) and the Kármán may be achieved.

4. Conclusion

We have simulated the dynamics of the vortex released from a PT symmetric potential. By varying the width and velocity of the real part of PT symmetric potential, we observed that the potential moving at a constant velocity exhibits various wake patterns. The ratio b/l for Kármán vortex street is slightly below the stability criterion in classical fluid theory. We also calculate the drag force experienced by the obstacle. In contrast to the scalar BEC without PT symmetric potential, the drag force exhibits different frequencies and magnitudes. We numerically found the parameter regions of different vortex shedding modes. Meanwhile, the velocity field after the obstacle potential is also affected by the imaginary part of the PT symmetric potential. We explored that the imaginary part can introduce additional asymmetry in the system. It alters the shedding frequency and the trajectory of the vortex, thus influencing the generation of the Kármán vortex street. We proposed an experimental solution which is relevant to the parameters in figure 2(a). The research may enhance the exploration of other fluid dynamics issues.

This work is supported by the National Natural Science Foundation of China under Grant Nos. 12065022, 12147213.

1
Bose V 1924 Plancks gesetz und lichtquantenhypothese Z. Angew. Phys. 26 178 181

DOI

2
Anderson M H, Ensher J R, Matthews M R, Wieman C E, Cornell E A 1995 Observation of Bose–Einstein condensation in a dilute atomic vapor Science 269 198 201

DOI

3
Davis K B, Mewes M O, Andrews M R, van Druten N J, Durfee D S, Kurn D M, Ketterle W 1995 Bose–Einstein condensation in a gas of sodium atoms Phys. Rev. Lett. 75 3969 3973

DOI

4
Regal C A, Greiner M, Jin D S 2004 Observation of resonance condensation of fermionic atom pairs Phys. Rev. Lett. 92 040403

DOI

5
Wen X Y, Lin Z, Wang D S 2024 High-order rogue wave and mixed interaction patterns for the three-component Gross–Pitaevskii equations in F = 1 spinor Bose–Einstein condensates Phys. Rev. E 109 044215

DOI

6
Aioi T, Kadokura T, Saito H 2012 Penetration of a vortex dipole across an interface of Bose–Einstein condensates Phys. Rev. A 85 023618

DOI

7
Xing J, Bai W, Xiong B, Zheng J H, Yang T 2023 Structure and dynamics of binary Bose–Einstein condensates with vortex phase imprinting Frontiers of Physics 18 62302

DOI

8
Abo-Shaeer J R, Raman C, Vogels J M, Ketterle W 2001 Observation of vortex lattices in Bose–Einstein condensates Science 292 476 479

DOI

9
Kim I, Wu X L 2015 Unified Strouhal–Reynolds number relationship for laminar vortex streets generated by different-shaped obstacles Phys. Rev. E 92 043011

DOI

10
Ponta F L, Aref H 2004 Strouhal–Reynolds number relationship for vortex street Phys. Rev. Lett. 93 084501

DOI

11
Williamson C H 1996 Vortex dynamics in the cylinder wake Annu. Rev. Fluid. Mech. 28 477 539

DOI

12
Dong S, Triantafyllou G S, Karniadakis G E 2008 Elimination of vortex streets in Bluff-body flows Phys. Rev. Lett. 100 20450

DOI

13
Barenghi C F 2008 Is the Reynolds number infinite in superfluid turbulence Physica D 237 2195 2202

DOI

14
Sasaki K, Suzuki N, Saito H 2010 Bénard-von Kármán vortex street in a Bose–Einstein condensate Phys. Rev. Lett. 104 150404

DOI

15
Saito H, Tazaki K, Aioi T 2013 Modulation of a quantized vortex street with a vibrating obstacle Procedia IUTAM 9 121 128

DOI

16
Kwon W J, Kim J H, Seo S W, Shin Y 2016 Observation of von Kármán vortex street in an atomic superfluid gas Phys. Rev. Lett. 117 245301

DOI

17
Li X L, Yang X Y, Tang N, Song L, Zhou Z K, Zhang J, Shi Y R 2019 Kármán vortex street in a two-component Bose–Einstein condensate New J. Phys. 21 103046

DOI

18
Yang X Y, Li X L, Tang N, Zhou Z K, Song L, Zhang J, Shi Y R 2020 Bénard-von Kármán vortex street in a spin-orbit-coupled Bose–Einstein condensate Phys. Rev. E 102 032217

DOI

19
Bender C M, Boettcher S 1998 Real spectra in non-Hermitian Hamiltonians having PT symmetry Phys. Rev. Lett. 80 5243 5246

DOI

20
Guo P, Gasparian V, Jódar E, Wisehart C 2023 Tunneling time in PT-symmetric systems Phys. Rev. A 107 032210

DOI

21
Suchkov S V, Chekhovskoy I S, Shtyrina O V, Wabnitz S, Fedoruk M P 2023 Nonlinear twisted multicore fibers with PT-symmetry Opt. Commun. 530 129147

DOI

22
Li H, Mekawy A, Krasnok A, Alù A 2020 Virtual parity-time symmetry Phys. Rev. Lett. 124 193901

DOI

23
Single F, Cartarius H, Wunner G, Main J 2014 Coupling approach for the realization of a PT-symmetric potential for a Bose–Einstein condensate in a double well Phys. Rev. A 90 042123

DOI

24
Gericke T, Würtz P, Reitz D, Langen T, Ott H 2008 High-resolution scanning electron microscopy of an ultracold quantum gas Nature Phys. 4 949 953

DOI

25
Li J, Harter A K, Liu J, de Melo L, Joglekar Y N, Luo L 2019 Observation of parity-time symmetry breaking transitions in a dissipative Floquet system of ultracold atoms Nat. Commun. 10 855

DOI

26
Bender C M, Boettcher S, Meisinger P N 1999 PT-symmetric quantum mechanics J. Math. Phys. 40 2201 2229

DOI

27
Konotop V V, Yang J, Zezyulin D A 2016 Nonlinear waves in PT-symmetric systems Rev. Mod. Phys. 88 035002

DOI

28
Lao J Y, Qin Z Y, Zhang J R, Shen Y J 2024 Peakons in spinor F=1 Bose–Einstein condensates with PT-symmetric δ-function potentials Chaos, Solitons Fractals 180 114497

DOI

29
Schwarz L, Cartarius H, Musslimani Z H, Main J, Wunner G 2017 Vortices in Bose–Einstein condensates with PT-symmetric gain and loss Phys. Rev. A 95 053613

DOI

30
Cartarius H, Wunner G 2012 Model of a PT-symmetric Bose–Einstein condensate in a δ-function double-well potential Phys. Rev. A 86 013612

DOI

31
Haag D, Dast D, Löhle A, Cartarius H, Main J, Wunner G 2014 Nonlinear quantum dynamics in a PT-symmetric double well Phys. Rev. A 89 023601

DOI

32
Xu S L, Wu T, Hu H J, He J R, Zhao Y, Fan Z 2024 Vortex solitons in Rydberg-excited Bose–Einstein condensates with rotating PT-symmetric azimuthal potentials Chaos, Solitons Fractals 184 115043

DOI

33
Ma J P, Wang Q Q, Tu P, Shao K H, Zhao Y X, Su R M, Zhao X, Xi B L, Shi Y R 2024 Gap solitons in quasi-1D Bose–Einstein condensate with three-body interactions under PT symmetry Phys. Scr. 99 045251

DOI

34
Zhao Y, Huang Q H, Gong T X, Xu S L, Li Z P, Malomed B A 2024 Three-dimensional solitons supported by the spin-orbit coupling and Rydberg–Rydberg interactions in PT-symmetric potentials Chaos, Solitons Fractals 187 115329

DOI

35
Kagan Y, Muryshev A E, Shlyapnikov G V 1998 Collapse and Bose–Einstein condensation in a trapped bose gas with negative scattering length Phys. Rev. Lett. 81 933 937

DOI

36
Zhang X F, Kato M, Han W, Zhang S G, Saito H 2017 Spin-orbit-coupled Bose–Einstein condensates held under a toroidal trap Phys. Rev. A 95 033620

DOI

37
Cui X 2022 Quantum fluctuations on top of a PT-symmetric Bose–Einstein condensate Phys. Rev. Res. 4 013047

DOI

38
Akram J, Pelster A 2016 Numerical study of localized impurity in a Bose–Einstein condensate Phys. Rev. A 93 033610

DOI

39
Jeszenszki P, Cherny A Y, Brand J 2018 S-wave scattering length of a Gaussian potential Phys. Rev. A 97 042708

DOI

40
Boniface P, Lebon L, Limat L, Receveur M 2017 Absolute stability of a Bénard-von Kármán vortex street in a confined geometry Europhys. Lett. 117 34001

DOI

41
Wang L X, Dai C Q, Wen L, Liu T, Jiang H F, Saito H, Zhang S G, Zhang X F 2018 Dynamics of vortices followed by the collapse of ring dark solitons in a two-component Bose–Einstein condensate Phys. Rev. A 97 063607

DOI

42
Zhang X F, Du Z J, Tan R B, Dong R F, Chang H, Zhang S G 2014 Vortices in a rotating two-component Bose–Einstein condensate with tunable interactions and harmonic potential Ann. Phys. 346 154 163

DOI

43
Wang D S, Hu X H, Hu J, Liu W M 2010 Quantized quasi-two-dimensional Bose–Einstein condensates with spatially modulated nonlinearity Phys. Rev. A 81 025604

DOI

44
Wang D S, Song S W, Xiong B, Liu W M 2011 Quantized vortices in a rotating Bose–Einstein condensate with spatiotemporally modulated interaction Phys. Rev. A 84 053607

DOI

45
Qiu X, Hu A Y, Cai Y Y, Saito H, Zhang X F, Wen L 2023 Dynamics of spin-nematic bright solitary waves in spin-tensor-momentum coupled Bose–Einstein condensates Phys. Rev. A 107 033308

DOI

46
Bao W, Chern I, Lim F Y 2006 Efficient and spectrally accurate numerical methods for computing ground and first excited states in Bose–Einstein condensates J. Comput. Phys. 219 836 854

DOI

47
Park J W, Ko B, Shin Y 2018 Critical vortex shedding in a strongly interacting fermionic superfluid Phys. Rev. Lett. 121 225301

DOI

48
Fujiyama S, Tsubota M 2009 Drag force on an oscillating object in quantum turbulence Phys. Rev. B 79 094513

DOI

49
Wille R 1960 Kármán vortex streets Adv. Appl. Mech. 6 273 287

DOI

50
Wang F, Xiong D, Li X, Wang D, Tiemann E 2013 Observation of Feshbach resonances between ultracold Na and Rb atoms Phys. Rev. A 87 050702

DOI

51
Sadler L E, Higbie J M, Leslie S R, Vengalattore M, Stamper-Kurn D M 2006 Spontaneous symmetry breaking in a quenched ferromagnetic spinor Bose–Einstein condensate Nature 443 312 315

DOI

52
Wang F D, Li X K, Xiong D Z, Wang D J 2015 A double species 23Na and 87Rb Bose–Einstein condensate with tunable miscibility via an interspecies Feshbach resonance J. Phys. B: At. Mol. Opt. Phys. 49 015302

DOI

53
Neely T W, Samson E C, Bradley A S, Davis M J, Anderson B P 2010 Observation of vortex dipoles in an oblate Bose–Einstein condensate Phys. Rev. Lett. 104 160401

DOI

54
Kwon W J, Moon G, Choi J, Seo S W, Shin Y 2014 Relaxation of superfluid turbulence in highly oblate Bose–Einstein condensates Phys. Rev. A 90 063627

DOI

55
Kwon W J, Moon G, Seo S W, Shin Y 2015 Critical velocity for vortex shedding in a Bose–Einstein condensate Phys. Rev. A 91 053615

DOI

56
Robins N P, Figl C, Jeppesen M, Dennis G R, Close J D 2008 A pumped atom laser Nature Phys. 4 731 736

DOI

Outlines

/