This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Optimizing persistent currents in a ring-shaped Bose-Einstein condensate using machine learning

Simeon Simjanovski ARC Centre of Excellence for Engineered Quantum Systems, School of Mathematics and Physics, University of Queensland, St Lucia, QLD 4072, Australia    Guillaume Gauthier ARC Centre of Excellence for Engineered Quantum Systems, School of Mathematics and Physics, University of Queensland, St Lucia, QLD 4072, Australia    Matthew J. Davis ARC Centre of Excellence for Engineered Quantum Systems, School of Mathematics and Physics, University of Queensland, St Lucia, QLD 4072, Australia    Halina Rubinsztein-Dunlop ARC Centre of Excellence for Engineered Quantum Systems, School of Mathematics and Physics, University of Queensland, St Lucia, QLD 4072, Australia    Tyler W. Neely ARC Centre of Excellence for Engineered Quantum Systems, School of Mathematics and Physics, University of Queensland, St Lucia, QLD 4072, Australia
Abstract

We demonstrate a method for generating persistent currents in Bose-Einstein condensates by using a Gaussian process learner to experimentally control the stirring of the superfluid. The learner optimizes four different outcomes of the stirring process: (O.I) targeting and (O.II) maximization of the persistent current winding number; and (O.III) targeting and (O.IV) maximization with time constraints. The learner optimizations are determined based on the achieved winding number and the number of spurious vortices introduced by stirring. We find that the learner is successful in optimizing the stirring protocols, although the optimal stirring profiles vary significantly depending strongly on the choice of cost function and scenario. These results suggest that stirring is robust and persistent currents can be reliably generated through a variety of stirring approaches.

I Introduction and Background

In a Bose-Einstein condensate (BEC) superfluid, a persistent current (PC) is the quantized circulation of the bulk fluid in a multiply connected geometry, such as a ring [1, 2]. As the phase singularity associated with the PC is located at the center of the ring, where there is no density, the flow is stabilized by topology against dissipation [1, 3]. While PCs are of fundamental interest in BEC research, they are also relevant to atomtronics [4, 5], where they can form the basis of compact matter-wave interferometers [6, 7, 8]. PCs are also the basis of fluxon systems [9] that have potential applications in atomtronics [10], for example as qubits which can be physically relocated [11].

Multiple methods for experimentally generating PCs have been developed. One can imprint the desired currents directly to the wave-function phase using an optical potential [12, 13], inducing superfluid flow proportional to the phase gradient [1]. However, in doing so the phase profile of the condensate acquires a sharp 2Nπ2N\pi phase jump for a charge-NN current at some point along the ring leading to significant density excitations [13, 14]. Alternatively, orbital angular momentum can be directly transferred from Laguerre-Gaussian laser modes [15, 16], to the condensate, although the typical efficiency is limited by mode matching to around 50% [17]. A more straightforward method is to stir the condensate using an optical barrier, analogous to stirring a classical fluid to generate flow [18, 19]. In this paper, we explore the parameter space of a stirring method, with an aim to determine optimal stirring profiles with a machine learning approach. Determining the best-case stirring scenarios will highlight the advantages and disadvantages of this approach compared to the other schemes.

Machine learning has increasingly been used for the experimental control of quantum systems [20, 21, 22]. The key advantage of this approach is the lack of prior knowledge a machine learner has on the system of interest, allowing for unique or counter-intuitive solutions to a given problem to be found [21, 22, 23, 24]. These solutions can be found even when there is an incomplete physical model of the system available, since the learner requires only a restricted set of data points [25]. Furthermore, computers have the ability to handle large amounts of data efficiently as compared to human operators. These factors have established machine learning as a potentially superior method of experimental control [26].

Refer to caption
Figure 1: (a) A simplified diagram of the main apparatus used to generate and control BECs. The light from the digital micromirror device (DMD) is shown by the green beam along the projection (Obj. 1) and imaging (Obj. 2) objectives. The light used for Faraday imaging propagates along the same direction. (b–e) Experimental images demonstrating the stirring process using the DMD projected light. The ring trap shown here has a larger width than the subsequent optimizations for illustrative clarity. (f) A typical time of flight experiment using stirring as described in Sec. II, with red dots indicating spurious vortices.

The stirring of a BEC to generate a PC is complicated by a variety of experimental realities that can be challenging or inefficient to model. For example, there may be density variations due to trap imperfections or roughness that can cause the PC to degrade or decay [27]. The BEC also has a finite lifetime within which stirring, system dynamics, and data extraction must occur, introducing time restrictions to the stirring process [28]. Aggressive stirring of the system will likely result in the generation of undesirable excitations such as quantum vortices and phonons [29]. These aspects suggest the stirring can be a complex process covering a large parameter space that may be challenging to accurately describe and control. Thus, under a set of desired constraints, optimal stirring parameters for this complex system are not a priori obvious. For example, various approaches have moved the barrier at constant speed [29, 30] or accelerating [31] speed profiles, but it is unclear which approach is superior. The stirring barrier must also be removed after inducing flow at some optimal rate. Finally, one might be interested in creating the current in the shortest time possible, to minimize the effect of atom loss due to a finite BEC lifetime. This large parameter space motivates the use of machine learning approaches to optimize the stirring process, subject to user-chosen sets of constraints. Recently, machine learning optimization of persistent currents in a ring lattice has been explored in numerical simulations [32], but such approaches have not yet been applied to experiments.

Here, we use reinforcement learning through the open-source Machine-Learning Online Optimization Package (M-LOOP) [33] to explore the stirring parameter space. Using the Gaussian process learner in M-LOOP and applying it to four separate optimization conditions, we find that the learner is able to successfully optimize the stirring process, generating the desired PCs while minimizing spurious vortices. However, we find that the optimized parameters vary widely depending on the chosen cost function and target parameters. The results suggest that stirring is robust against variations in the particular details of the stirring profile.

II Experimental Implementation

Refer to caption
Figure 2: Process flow chart showing the general stirring parameter optimization process using M-LOOP over 50 iterations. Each iteration is in reality composed of five experimental runs, taken with the same parameter settings. The cost of each iteration is computed as the average cost for these five runs. This results in a mean cost and uncertainty (standard error) returned to M-LOOP for each parameter setting.

Our experiment implements configured optical potentials with blue-detuned repulsive light to both trap a 87Rb BEC in a ring and simultaneously stir it, as shown in Fig. 1. The optical potential for the ring geometry and stirring is generated using a digital micromirror device (DMD). The DMD is an array composed of 2560 ×\times 1600 mirrors that can be individually switched “on” and “off” corresponding to ±12\pm 12^{\circ} tilts, respectively. This enables the creation of a “binary” hard-wall trapping potential [34]. A detailed description of the process for creating and configuring BECs using this apparatus can be found in our prior work [35, 34].

The experimental sequence is shown in Fig. 1. The atoms are initially trapped in a 100100-μ\mum diameter disk. A separation barrier is then inserted, growing from zero to 77-μ\mum width over 100100-ms symmetrically about a mean radius of 3434-μ\mum. This insertion method was observed to be sufficiently slow to avoid density excitations. This process results in two separated condensates: an outer ring BEC where the PC is created, and an inner reference BEC that has a uniform phase, and can be interfered with the outer BEC resulting in a pattern from which the PC winding number can be inferred [36]. Simultaneously, a 33-μ\mum thick stirring barrier is inserted into the ring BEC by linearly increasing the barrier thickness from 0-μ\mum. This barrier is used for the stirring of the outer ring, subject to machine learning controlled parameters (see Section III) which aim to generate the desired persistent current. After the stirring process is completed, the stirring barrier is removed by linearly decreasing its thickness over the removal time TRT_{R} while maintaining the same angular velocity. The stirring region and reference BEC are then interfered by releasing the system from the trap, and imaged in time of flight (TOF) [37, 19, 13].

The number of spiral fringes resulting from the interference of the ring and reference condensates identifies the phase winding number of the ring, which is directly proportional to the PC flow speed [37]. Spuriously generated quantum vortices can form due to the stirring and are resolvable as density dips in TOF, as seen in Fig. 1. A single TOF image can thus be used to determine both the winding and vortex number. A detailed description of the image analysis is given in Appendix A.

III Machine Learning Implementation

III.1 Stirrer parametrization

Machine learning control of the stirring is implemented with Gaussian process regression via the freely available package M-LOOP [33, 38]. As a reinforcement machine learning method, M-LOOP requires an external measure of the error produced by the parameter model chosen by the learner [25, 39]. This error can be quantified using a cost function. The choice of cost function can vary specific to the type of optimization required, but there is one underlying property across all choices: the cost should be minimized when the parameter set is optimized [24, 20].

We begin by specifying the angular stirring function of the barrier:

θ(t)=α(tA)P,\displaystyle\theta(t)=\alpha\left(\frac{t}{A}\right)^{P}, (1)

where the constant A=400A=400-ms is introduced to ensure the stirring is slow enough to prevent destroying the BEC. The parameters controlled by the machine learner are as follows: (1) Stirring time of the barrier after insertion, TT; (2) Removal time of the barrier after the stirring time, TRT_{R}; (3) Stirring exponent, PP; and (4) Angular coefficient, α\alpha. Each parameter is given a lower and upper bound: (1) 100100 ms T1500\leq T\leq 1500 ms, (2) 1010 ms TR450\leq T_{R}\leq 450 ms, (3) 1P4.51\leq P\leq 4.5 [dimensionless], and (4) 1.75×1021.75\times 10^{-2}-rad ααmax\leq\alpha\leq\alpha_{\textrm{max}}, where αmax=3.49\alpha_{\textrm{max}}=3.49-rad in the case of optimizations considered in Sec. III and αmax=10.5\alpha_{\textrm{max}}=10.5-rad for those considered in Sec. V. This was done to restrict the parameter space initially (Sec. IV) for a simpler optimizations while increasing complexity with a larger parameter space in the later optimizations (Sec. V) to test the limitations of the learner employed. The lower bound on TT is selected so that even for the minimum stirring time sufficient fringes are generated for the detection algorithm to operate accurately 111With very few fringes present, the fringes themselves wrap around the entire interference region. The detection algorithm effectively counts periodicity of density in these regions meaning that such wrapping obscures the measurement. (see also Appendix A). The upper bound is imposed due to the frame-rate constraints on the DMD, ensuring that the sequence projected onto the BEC is smooth when the stirring occurs, avoiding the stirring barrier moving around the ring in large, discrete jumps. Bounds on TRT_{R} are also chosen such that the removal time is comparable to the insertion time. The ranges for α\alpha and PP were chosen by observation to ensure parameters that both were sufficient to consistently generate observable spiral fringes and did not result in aggressive stirring such that the BEC was destroyed.

III.2 Cost functions

Two main optimizations are considered: targeting a specific winding number, with the cost function

𝒞T\displaystyle\mathcal{C}_{\textrm{T}} =|NTNW|+NV+10TTmax,\displaystyle=\left|N_{T}-N_{W}\right|+N_{V}+10\frac{T}{T_{\text{max}}}, (2)

and maximizing the winding number, with the cost function

𝒞M\displaystyle\mathcal{C}_{\textrm{M}} =NW+NV+10TTmax.\displaystyle=-N_{W}+N_{V}+10\frac{T}{T_{\text{max}}}. (3)

Here, the measured winding number NWN_{W} and the measured spurious vortex number NVN_{V} are obtained from the image analysis [Appendix A]. The target winding number NTN_{T} is chosen prior to optimization. The maximum allowed stirring time TmaxT_{\text{max}} is introduced for optimizations O.III and O.IV where we impose a time restriction for the stirring process (Sec. V). This time restriction on stirring is made so that any experiments which require both stirring and observation of subsequent dynamics of the BEC system can be made within the BEC lifetime, \approx 20 s for our system. By minimizing the stirring time, we make more time available for these subsequent dynamics to be observed. Otherwise, we set TmaxT_{\text{max}}\to\infty such that the last term in the cost functions vanishes. Cost is computed using an average measure of cost for five instances of the experiment. The standard error of this cost measurement is used as the uncertainty estimate for the cost. This uncertainty is not incorporated into the cost itself, but is fed directly to M-LOOP alongside the cost. M-LOOP is then able to estimate errors on the parameter landscapes [e.g., Fig. 3(b)] using these input uncertainties.

While vortices are the primary spurious effects we are interested in minimizing, density excitations (sound) may also result depending on the stirring profile [41]. This information is not readily extracted from the TOF images, although it may affect the regularity of the fringe pattern. Minimizing sound excitation may be an additional refinement in future implementations of this approach.

Equation (2) is minimized when the winding number is exactly NTN_{T} and no vortices are present in the system, resulting in optimization to the target. Equation (3) is minimized for the largest possible winding number given a minimized vortex number, thus maximizing the PC. The weights on each term were chosen such that any spurious vortices are weighted equally to the error in the winding for Eq. (2) or to how large the winding becomes in Eq. (3). This ensures that the best “score” is assigned not only in achieving the target or maximizing the current, but also when spurious vortices are removed. When including the stirring time penalty, the factor of 1010 ensures that that longer stirring times are heavily punished, to the same level of having ten spurious vortices for every stirring time TT near the upper limit TmaxT_{\text{max}}.

IV Optimizations O.I and O.II: Variable stirring time

We begin our investigation by fixing the stirring exponent P=2P=2 (constant acceleration) and considering TmaxT_{\text{max}}\to\infty to remove the stirring time penalty from Eqs. (2) and (3), meaning the stirring time is limited only by its upper bound.

Refer to caption
Figure 3: M-LOOP output graphics for optimization to a target of NT=20N_{T}=20 windings (optimization O.I – cost function given by Eq. (2)). (a) Cost vs run number over 50 runs. The initial 20 random runs are shown before the vertical dashed line. Filled (red) points represent the randomly chosen parameter runs, and open (blue) points represent the runs with machine learner chosen parameters. Filled (red) points following the training can also be observed, which also represent random walks designed to drive the learner out of potential local minima while testing the optimum parameters. The solid (blue) line tracks the minimum cost achieved as a function of run number. (b) The predicted landscape for each of the three parameters computed using the costs. The dashed (red) line represents the stirring time parameter TT, the solid (green) line represents the removal time TRT_{R}, and the dotted (blue) line represents the angular coefficient α\alpha. Parameter values are normalized to lower and upper (PminP_{\text{min}} and PmaxP_{\text{max}}) bounds using (PPmin)(P-P_{\text{min}})/(PmaxPmin)(P_{\text{max}}-P_{\text{min}}). Uncertainties, shown by the shaded regions for each curve, are computed through M-LOOP by feeding the learner an error estimate for each run. (c) The 5-ms TOF image resulting from using the optimum parameters in stirring. This particular image contains 19 fringes.

IV.1 Optimization O.I: Targeting

Using Eq. (2) as a cost function aims to realize a target winding number. This case is desirable for experiments investigating various initial conditions of a superfluid shear layer such as with the superfluid Kelvin-Helmholtz instability [42, 43] or when fluxons [9, 10] are experimentally initialized by stirring. A target winding number of NT=20N_{T}=20 is chosen. This choice was made since a winding of 2020 was already known to be achievable in the experiment though manually optimized stirring. The general process for optimization is described in Fig. 2. After 50 iterations, the learner was terminated, and the results of the procedure can be seen in Figure 3.

The cost vs run number is shown in Fig. 3(a) and illustrates how the learner evolved in choosing parameters and their effectiveness in experimentally acquiring a circulation of NT=20N_{T}=20. We note that in this scheme it is difficult to discern between large cost values due to the measured winding number being vastly different to the target or whether it is due to many spurious vortices. Generally, spurious vortices vary greatly shot to shot and so large cost values with larger uncertainties can be associated with large spurious vortex numbers. In contrast, the winding will be relatively stable shot to shot and data points associated with these defects will have lower uncertainty. The blue curve superimposed on the data tracks the running minimum cost. As can be seen, convergence is quickly achieved, justifying the cutoff run number of 5050 since we see no significant improvement in the cost or parameters. With these parameters, the cost landscape forms a four-dimensional region. Landscape curves, shown in Fig. 3(b), are computed as the cross sections of the cost landscape for each of the three parameters, chosen such that the other two parameters coincide with experimentally minimal costs. The optimum stirring parameters can be read directly from the global minima of the curves.

Figure 3(c) shows the results of stirring under the learner-proposed optimum parameters. For this particular image, no spurious vortices were observed while the winding count is NW=19N_{W}=19. Averaging over five runs under these optimum parameters, it was found that the mean winding was N¯W=19.6±0.3\overline{N}_{W}=19.6\pm 0.3 with N¯V=0.2±0.2\overline{N}_{V}=0.2\pm 0.2 spurious vortices. The optimal stirring time TT is 760.6760.6-ms, near the middle of its bound range. The optimal removal time TR=176.2T_{R}=176.2-ms is relatively short, near its lower bound. The coefficient parameter α\alpha exhibits multiple minima but has one clear global minimum at 0.7860.786-rad. Using these, it is possible to also deduce the final speed of the barrier from Eq. 1 to be vB=8.9×103v_{B}=8.9\times 10^{-3}-rad/ms or equivalently vB=0.38v_{B}=0.38-mm/s at the inner radius of 32.2532.25-μ\mum. The optimum parameters, along with the parameters for the subsequent optimizations, are summarized in Table 1.

IV.2 Optimization O.II: Maximization

Optimization O.II seeks to maximize the winding number while suppressing spurious vortices. This can be achieved by running M-LOOP with the cost function in Eq. (3). Figure 4 shows the outcome of this optimization. There is immediately a clear and steady decrease in cost after the final training run, leading to a plateau in the cost for the final runs indicated by the minimum cost curve in Fig. 4(a). This behavior suggests convergence to an optimal parameter set.

The parameter landscapes show the optimized parameter choices as clear global minima of the curves. Unlike the targeting optimization, all curves contain a single minima. The optimal stirring protocol applies aggressive movement of the barrier, indicated by the large α=2.36\alpha=2.36-rad, but for a relatively short stirring time at T=335.3T=335.3-ms and removal time TR=15.2T_{R}=15.2-ms. These parameters correspond to a final barrier speed of vB=1.0×102v_{B}=1.0\times 10^{-2}-rad/ms or vB=0.44v_{B}=0.44-mm/s at a radius of 32.2532.25-μ\mum. An example TOF image using the optimized parameters is shown in Fig. 4(c), which contains 2525 windings. The parameters result in N¯W=25.8±0.4\overline{N}_{W}=25.8\pm 0.4, with N¯V=0.2±0.2\overline{N}_{V}=0.2\pm 0.2 over five experimental runs.

Refer to caption
Figure 4: M-LOOP output graphics for maximization of the winding number (O.II – cost function given by Eq. (3)). (a) The cost vs the run number over 50 runs. (b) The predicted landscape showing the optimum parameter choices. The dashed (red) line represents the stirring time parameter TT, the solid (green) line represents the removal time TRT_{R}, and the dotted (blue) line represents the angular coefficient α\alpha. Uncertainties are indicated by the shaded regions. (c) 55-ms TOF image under optimized parameters. This particular image shows 25 fringes.

IV.3 Remarks on the optimizations

Considering first Fig. 3(a) and Fig. 4(a), we observe that during training runs the computed cost and uncertainty vary significantly from run to run. The concept behind this random training is to walk through as large a portion of the parameter space as possible before the optimization effectively begins. Doing so ensures that the Gaussian process regression is more accurately able to produce meaningful landscape curves [33]. Following the training, we observe a rapid stagnation of the cost in both optimization cases. This is indicated by the minimum cost curves superimposed on Fig. 3(a) and Fig. 4(a), which show a drop after training to a sudden plateau suggesting convergence.

For maximization O.II, the optimal parameters can be interpreted as those which reach the upper limit for the winding (flow speed) of the persistent current before the shedding of vortices from the stirring barrier cannot be avoided, i.e., the critical velocity has been reached [44]. Applying Eq. (5) [Appendix B] using the final winding numbers for both the targeting and maximization case gives a flow speed of vT(R)347v_{\text{T}}(R)\approx 347-μ\mum/s and vM(R)457v_{\text{M}}(R)\approx 457-μ\mum/s respectively (at the inner ring radius R=32.25μR=32.25~{}\mum). Equivalently, this gives the critical velocity of the experimental system to be vM0.35csv_{\text{M}}\approx 0.35c_{s} where cs1.3c_{s}\approx 1.3-mm/s is the two-dimensional (2D) speed of sound for the system [see Appendix B, Eq. (6)]. This critical velocity measurement is consistent with previous estimates of vortex shedding due to flow past an obstacle [45, 46, 47], suggesting the current is limited by features or roughness in the optical potential and that the machine learner has optimized the stirring process to this limit. We also note that the final barrier and fluid speeds for O.I do not coincide, as shown in Table 1, indicating that the fluid is out of equilibrium with the barrier for this stirring protocol. This behavior is discussed in more detail in Sec. VI.

Overall, the machine learner is able to optimize two different stirring protocols for different final PC conditions. Given the constant angular acceleration used in the stirring profile here, an experimenter might opt to maximize the current via stirring using the same profile as O.I for longer times, given that this is the most straightforward approach. Remarkably, the machine learner finds a different and more efficient approach to stirring: aggressively stirring for a shorter period of time.

Refer to caption
Figure 5: M-LOOP output graphics for time-limited targeting a winding number of NT=20N_{T}=20 windings (O.III – cost function given by Eq. (2)). (a) The cost vs the run number over 50 runs. The color convention is consistent with Fig. 3. (b) The predicted landscape showing the optimum parameter choices. The dashed (red) line represents the stirring time TT landscape while the dotted (blue) line represents the acceleration coefficient α\alpha. Here, the solid (green) line represents the exponent parameter PP. Uncertainties are indicated by the shaded regions. (c) 55 ms TOF image under optimized parameters. This particular image shows 18 fringes.

V Optimizations O.III and O.IV: Minimum stirring time

Motivated by the above results, we next repeat the previous optimizations while attempting to minimize the stirring time. The learner now has to satisfy the desired winding conditions with a cost penalty for long stirring times TT. We no longer hold P=2P=2, in Eqs. (3) and (2), instead allowing it to be a parameter optimized by the learner. This extra flexibility is allowed in order to compensate for the added complexity in the cost function, giving the learner more control over the barrier motion. Tmax=1500T_{\text{max}}=1500-ms is now the maximum allowed stirring time, which is consistent with the upper bound on the stirring time TT. Since the cost was observed to vary weakly with the barrier removal time (see Figs. 3 and 4) the removal time is fixed to TR=100T_{R}=100-ms, simplifying the cost functions and aiding convergence.

V.1 Optimization O.III: Targeting

Figure 5 shows the outcomes from target optimization subject to a time cost imposed by Eq. (2). The stirring time curve in Fig. 5(b) shows a global minimum near 240240-ms, which is close to its lower boundary of 100100-ms, indicating that only short stirring times are required to achieve the desired winding target. The power parameter also has its global minimum at its lower bound value P=1P=1 and the associated coefficient α=2.64\alpha=2.64-rad. This results in an instantaneous and constant angular velocity of 6.6×1036.6\times 10^{-3}-rad/ms, corresponding to 0.280.28-mm/s at the inner radius of 32.2532.25-μ\mum. Spurious vortices were more prevalent in this optimization, as shown in the example image, Fig. 5(c). Here we also observe 18 windings. Averaging over five separate images, a winding number of NW=18.4±0.4N_{W}=18.4\pm 0.4 is observed with NV=1.0±0.3N_{V}=1.0\pm 0.3 spurious vortices.

V.2 Optimization O.IV: Time-limited maximization

Figure 6 shows the outcome of trying to maximize the winding by employing Eq. (3) with a time cost for M-LOOP. The learner was able to find optimized parameters as indicated by the clear minima in the landscape curves of Fig. 6(b). All global minima lie away from the boundaries of the parameters in this optimization. The stirring is highly non-linear, with a value of P=3.5P=3.5. The stirring time was optimized to T=163.5T=163.5-ms suggesting that the learner correctly identified the cost function and time restriction. Accompanying the short, non-linear stirring is a large, α=8.22\alpha=8.22-rad, coefficient contributing to the highly aggressive stirring. In this case, the final barrier speed is vB=2.5×102v_{B}=2.5\times 10^{-2}-rad/ms or 1.11.1-mm/s at a radius of 32.2532.25-μ\mum. Under this proposed optimal stirring, the resulting BEC looks similar to Fig. 6(c) which contains five spurious vortices but only 18 windings. An average winding number of N¯W=18.4±0.4\overline{N}_{W}=18.4\pm 0.4 with N¯V=4.5±0.5\overline{N}_{V}=4.5\pm 0.5 spurious vortices is observed overall. Likely this smaller winding, relative to O.II, is due to the trade-off placed in the cost function weights between the winding and the stirring time, which are necessary to enable the fast creation of the PC.

Refer to caption
Figure 6: M-LOOP output graphics for time-limited maximization of the windings number (Optimization IV – cost function given by Eq. (3)). (a) The cost vs the run number over 50 runs. (b) The predicted landscape shows the optimum parameter choices. The dashed (red) line represents the stirring time TT landscape while the dotted (blue) line represents the acceleration coefficient α\alpha. Here the solid (green) line represents the exponent parameter PP. Uncertainties are indicated by the shaded regions. (c) 55 ms TOF image under optimized parameters. This particular image shows 18 fringes.

V.3 Remarks on the time-limited optimizations

The time series of the cost vs run number for O.III in Fig. 5(a) shows a small dip followed by stagnation within ten machine learner controlled runs. Both cases exhibit the characteristic random cost distribution during training, and the effectiveness of said training is highlighted by the smoothness of the parameter landscape curves in Figs. 5(b) and 6(b). In both cases, the parameter landscape curves have minima near the lower boundary of the stirring time TT. This suggests that the learner correctly includes the added restriction on the stirring time.

Observing the time series for both O.III and O.IV, Figs. 5(a) and 6(a) respectively, we see the cost vs run number in both cases stagnates rapidly, suggesting fast convergence. However, the convergence in winding number is exactly the same between the two cases. Since both cases share NW=18.4±0.4N_{W}=18.4\pm 0.4, both have a flow speed of v(R)v(R)\approx-μ\mum/s or v.25csv\approx.25c_{s} (at R=32.25R=32.25-μ\mum). For O.III, this is close to the target. For O.IV however, this flow rate is far from the critical velocity measured in O.II. The inability of the learner to optimize the stirring as desired arises from the fact that the cost function is designed with a new weighted punishment term for the stirring time, introducing a trade-off between maximum winding and shortest stirring.

VI Conclusion and Outlook

The first aim of this paper was to realize optimal stirring parameters for creating PCs with a stirring barrier. The machine learning algorithm was successful in this endeavor, creating PCs for each of the optimizations explored, while minimizing spurious vortices 222We note that the lower winding numbers in O.III and O.IV result from the strong weighting on the stirring time punishment term in the cost functions. This observation highlights the trade-offs introduced in the cost functions which are likely the limiting factors preventing more complete optimization in more complicated scenarios.. While our initial hypothesis was that there would be a unique approach to optimal stirring, the results depended greatly on the choice of cost function and the constraints chosen. In the case of variable stirring time without penalty (O.I and O.II), the learner chose dramatically different stirring times, stirring gently for a long time (O.I) and stirring aggressively for a short time (O.II). For the variable time optimizations (O.III and O.IV), the choice of parameters is even more distinct. O.III found the best protocol was to instantaneously accelerate the barrier to move with constant angular velocity, while O.IV found that aggressive stirring with a highly non-linear profile, P=3.5P=3.5, was optimal.

Optimization TT (ms) TRT_{R} (ms) α\alpha (rad) PP vv/csc_{s} vBv_{B}/csc_{s}
O.I 721 176 0.786 2 0.27 0.38
O.II 335 15.2 2.36 2 0.35 0.34
O.III 240 100 2.64 1 0.25 0.22
O.IV 164 100 8.22 3.5 0.25 0.84
Table 1: Summary of optimized parameter values for the optimization cases O.I (targeting), O.II (maximization), O.III (time-limited targeting), and O.IV (time-limited maximization). vv is the PC flow speed in the middle of the ring, while cs1300c_{s}\approx 1300-μ\mum/s is the speed of sound. The final speed of the barrier is indicated by vBv_{B}.

These results suggest that while the learner can successfully control stirring for desired results and minimum cost, the problem of determining an optimal stirring profile is under-constrained by the chosen parameters. We take this result to imply that stirring to create a PC is a robust process, and can be successfully achieved with a variety of stirring profiles. From a practical perspective, the relatively counter-intuitive approach of aggressive stirring for a short time at constant speed (O.III) may be the best approach, since this minimizes atom loss due to the finite lifetime of the BEC. Another interesting feature of these optimizations is that there are some out-of-equilibrium dynamics between the fluid and barrier. This is seen through the difference in final barrier speeds relative to the superfluid velocity. For example, O.IV resulted in a final barrier speed of vB=0.84csv_{B}=0.84c_{s} for a final superfluid speed of v=0.25csv=0.25c_{s}, but relied on a short stirring time of 164164-ms. These results suggest counter-adiabatic stirring protocols may be advantageous, provided the barrier beam is removed at the right time. Previous work has explored the possibility of spatial rapid adiabatic passage for engineering angular momentum states in a short time [49, 50]. Future extension of similar approaches to stirring of persistent currents may be a fruitful area of both theoretical and experimental future work.

Overall, the successful optimizations demonstrate the suitability of using machine learners for the creation and control of PCs through stirring, albeit with no universal stirring protocol. These results suggest useful future applications of the algorithm. Since PCs establish a basis for atom interferometry [51, 52, 53], having high levels of control over PCs using machine learning can be desirable for engineering atom interferometers [54, 55]. Beyond interferometry, it is possible to apply this algorithm to prepare PC states with specified winding numbers for studies in quantum turbulence, such as determining the dynamics of the Kelvin-Helmholtz instability [42, 43]. Atom inertial sensors have used PCs as a source of synthetic rotation for calibration [56], and this method presents a precise way of creating these references. The general structure of the feedback loop described could also be adapted to optimize alternative methods for generating PCs, such as phase imprinting [13].

VII Acknowledgements

We thank M. Reeves and M. Christenhusz for useful discussions and help with computational work. This research was supported by the Australian Research Council (ARC) Centre of Excellence for Engineered Quantum Systems (EQUS, CE170100009). S. S acknowledges the support of an Australian Government Research and Training Program Scholarship. G. G. acknowledges the support of ARC Discovery Projects Grant No. DP200102239. T. W. N. acknowledges the support of Australian Research Council Future Fellowship No. FT190100306.

Appendix A Image Processing

Images of the BEC are produced using Faraday imaging [57, 58]. Automatic image processing is used to extract the winding number and vortex number from the Faraday images. Vortices become directly resolvable in TOF as clusters of dark pixels, corresponding to density dips, which means that detection of such sites can be treated analogously to blob detection in general image processing [59, 60, 61]. The use of this algorithm ensures successful detection of vortices outside of the fringe region but tends to count dark fringes as vortex objects as well. To avoid this issue, the algorithm is restricted to operate only outside the fringe region through masking while the fringe region is analyzed using a Fourier signal method discussed in more detail below. Specifically, we restrict the vortex detection to the outer ring where stirring occurs.

Vortices may still be present in the fringe region and can be observed as either a warping or a fork in a fringe which separates into two fringes, most clearly observable in Fig. 7(a). This results in an increased standard deviation when estimating the winding number as described below. The fringe number standard deviation is simply added to the vortex number NVN_{V} when computing the average cost for the set of five runs. Although this potentially overestimates the number of vortices in the fringe region by mixing the standard deviation in achieved winding number into NVN_{V}, we find that this approach successfully minimizes spurious vortices in both the masked and interference fringe regions. Note that any spurious vortices in the interior of the reference condensate are ignored by the algorithm since they are not caused by the stirring process.

A Fourier signal algorithm can be used for the counting winding number and its standard deviation for the aforementioned estimate of NVN_{V}. We begin by considering the fringe region as indicated by the solid green line in Fig. 8(b). The azimuthal Fourier transform is determined and the mode with the largest Fourier amplitude is identified as corresponding to the winding number for the selected region. A cutoff mode of n=3n=3 is introduced since the signal of these modes is generally large, leading to erroneous detection of winding numbers that do not correspond to the observed spiral pattern. We found this cutoff value sufficient for reducing this effect.

Refer to caption
Figure 7: Processing of a TOF image using the Gaussian blob algorithm. (a) The original image. (b) Masked image to remove the fringe region. The processed image is also smoothed using a Gaussian filter. (c) Laplacian of the masked and filtered image. Thresholding of the image is now possible and the sharp ring edges can be masked out. (d) Detected vortex positions (green solid circles) superimposed on the original image. Additional details can be found in Refs. [59, 60, 61]. Note that due to the masking of the fringe region, vortices in this region are not detected despite being clearly present as evident in the warping and forking of the fringes.
Refer to caption
Figure 8: Example of the processing of a TOF image using the winding counting algorithm. (a) Fourier amplitude for mode numbers 1–50. The largest amplitude mode is taken to be the winding number. Mode numbers at or under the cutoff mode of n=3n=3 are indicated by the black dashed vertical line and are highlighted in red. This cutoff is introduced since these modes have a large amplitude but do not reflect the fringe number. The investigation does not require modes close to n=3n=3 to be detected. The largest amplitude has a mode number of 2424, indicated by the orange, dot-dashed line. (b) Color coded image highlighting the region of interest (green ring) considered in the example, superimposed on the original image. Note that only one small region of interest is shown here, while in practice 1 similar rings of varying radii are considered. These 1 rings overlap slightly since their inner radii are separated by 0.250.25-μ\mum while their thickness remains fixed at 11-μ\mum. Over the 1 rings for the example image the mean fringe number is 24.8±0.424.8\pm 0.4 where the uncertainty is estimated as the standard error of the 1 ring measurements. The observed count in this particular image is 2424 fringes.

The spiraling nature of the fringes requires some care in considering how to best extract the winding number from the discrete Fourier transform. The solution is to subdivide the region into 1 overlapping rings of 1-μ\mum thickness. These smaller rings are set apart via their inner and outer radii being translated by 0.25-μ\mum. This choice allows the entire fringe region to be spanned by a series of thin rings over which the accuracy of the detection is improved. The standard deviation of the winding number across the rings is used to estimate the number of vortices in the fringe region. Finally, in order to extract a useful measure of winding number, the mean of the mode number with the largest amplitude is selected which need not be an integer. Figure 8(a) shows the Fourier amplitudes for the considered modes for the small ring region considered in the image in Fig. 8(b).

Appendix B Estimating Flow Speed from Winding Numbers

It is possible to estimate the speed of the PCs generated by these different stirring processes via the velocity-phase relationship in a supercurrent:

v(r,t)=mϕ(r,t).\vec{v}(\vec{r},t)=\frac{\hbar}{m}\nabla\phi(\vec{r},t). (4)

Assuming that the winding is in the azimuthal direction, θ^\hat{\theta}, due to the ring geometry, any radial derivatives of the phase are zero. The PC should not decay making the speed independent of time. Therefore, the flow speed can be computed via:

v(r)=m1rNWθ^.\vec{v}(r)=\frac{\hbar}{m}\frac{1}{r}N_{W}\hat{\theta}. (5)

The speed of sound in the quasi-2D BEC system can be estimated for comparison using the relation:

cs=μ2Dm,c_{s}=\sqrt{\frac{\mu_{\text{2D}}}{m}}, (6)

where-μ2D\mu_{\text{2D}} is the effective 2D chemical potential of the persistent current ring region.-μ2D\mu_{\text{2D}} can be deduced from the condensed atom number in the ring Nc9.6×105N_{c}\approx 9.6\times 10^{5} along with the inner and outer radii of the ring, rI=36r_{\text{I}}=36-μ\mum and rO=50r_{\text{O}}=50-μ\mum respectively, using the following relation [62]:

μ2D=25(32gNcmωz2π(rO2rI2))2/3.\mu_{\text{2D}}=\frac{2}{5}\left(\frac{3}{2}\frac{gN_{c}\sqrt{m\omega_{z}^{2}}}{\pi(r_{\text{O}}^{2}-r_{\text{I}}^{2})}\right)^{2/3}. (7)

Using a vertical trapping frequency of ωz=2π×108\omega_{z}=2\pi\times 108 Hz and the ss-wave scattering length as=98.13a0a_{s}=98.13a_{0} to compute the effective 2D interaction parameter g=4π2as/mg=4\pi\hbar^{2}a_{s}/m, we compute that-μ2D2.5×1031\mu_{\text{2D}}\approx 2.5\times 10^{-31} J/kg and therefore cs1.3c_{s}\approx 1.3-mm/s.

References

  • Leggett [2001] A. J. Leggett, Bose-Einstein condensation in the alkali gases: Some fundamental concepts, Reviews of Modern Physics 73, 307 (2001).
  • Ryu et al. [2007] C. Ryu, M. F. Andersen, P. Clade, V. Natarajan, K. Helmerson, and W. D. Phillips, Observation of persistent flow of a Bose-Einstein condensate in a toroidal trap, Phys. Rev. Lett. 99, 260401 (2007).
  • Neely et al. [2013] T. W. Neely, A. S. Bradley, E. C. Samson, S. J. Rooney, E. M. Wright, K. J. H. Law, R. Carretero-González, P. G. Kevrekidis, M. J. Davis, and B. P. Anderson, Characteristics of two-dimensional quantum turbulence in a compressible superfluid, Phys. Rev. Lett. 111, 235301 (2013).
  • Amico et al. [2021] L. Amico, M. Boshier, G. Birkl, A. Minguzzi, C. Miniatura, L.-C. Kwek, D. Aghamalyan, V. Ahufinger, D. Anderson, N. Andrei, et al., Roadmap on Atomtronics: State of the art and perspective, AVS Quantum Science 3, 039201 (2021).
  • Amico et al. [2022] L. Amico, D. Anderson, M. Boshier, J.-P. Brantut, L.-C. Kwek, A. Minguzzi, and W. von Klitzing, Colloquium: Atomtronic circuits: From many-body physics to quantum technologies, Reviews of Modern Physics 94 (2022).
  • Ryu et al. [2013] C. Ryu, P. W. Blackburn, A. A. Blinova, and M. G. Boshier, Experimental Realization of Josephson Junctions for an Atom SQUID, Phys. Rev. Lett. 111, 205301 (2013).
  • Halkyard et al. [2010] P. L. Halkyard, M. P. A. Jones, and S. A. Gardiner, Rotational response of two-component Bose-Einstein condensates in ring traps, Phys. Rev. A 81, 061602 (2010).
  • Helm et al. [2018] J. L. Helm, T. P. Billam, A. Rakonjac, S. L. Cornish, and S. A. Gardiner, Spin-orbit-coupled interferometry with ring-trapped Bose-Einstein condensates, Phys. Rev. Lett. 120, 063201 (2018).
  • Oliinyk et al. [2019] A. Oliinyk, A. Yakimenko, and B. Malomed, Tunneling of persistent currents in coupled ring-shaped Bose–Einstein condensates, Journal of Physics B: Atomic, Molecular and Optical Physics 52, 225301 (2019).
  • Oliinyk et al. [2020] A. Oliinyk, B. Malomed, and A. Yakimenko, Nonlinear dynamics of Josephson vortices in merging superfluid rings, Communications in Nonlinear Science and Numerical Simulation 83, 105113 (2020).
  • Kaurov and Kuklov [2005] V. M. Kaurov and A. B. Kuklov, Josephson vortex between two atomic Bose-Einstein condensates, Phys. Rev. A 71, 011601 (2005).
  • Denschlag et al. [2000] J. Denschlag, J. E. Simsarian, D. L. Feder, C. W. Clark, L. A. Collins, J. Cubizolles, L. Deng, E. W. Hagley, K. Helmerson, W. P. Reinhardt, S. L. Rolston, B. I. Schneider, and W. D. Phillips, Generating Solitons by Phase Engineering of a Bose-Einstein Condensate, Science 287, 97 (2000).
  • Del Pace et al. [2022] G. Del Pace, K. Xhani, A. Muzi Falconi, M. Fedrizzi, N. Grani, D. Hernandez Rajkov, M. Inguscio, F. Scazza, W. J. Kwon, and G. Roati, Imprinting persistent currents in tunable fermionic rings, Phys. Rev. X 12, 041037 (2022).
  • Kumar et al. [2018] A. Kumar, R. Dubessy, T. Badr, C. De Rossi, M. de Goër de Herve, L. Longchambon, and H. Perrin, Producing superfluid circulation states using phase imprinting, Phys. Rev. A 97, 043615 (2018).
  • Allen et al. [1992] L. Allen, M. W. Beijersbergen, R. J. C. Spreeuw, and J. P. Woerdman, Orbital angular momentum of light and the transformation of Laguerre-Gaussian laser modes, Phys. Rev. A 45, 8185 (1992).
  • Beattie et al. [2013] S. Beattie, S. Moulder, R. J. Fletcher, and Z. Hadzibabic, Persistent currents in spinor condensates, Phys. Rev. Lett. 110, 025301 (2013).
  • Andersen et al. [2006] M. F. Andersen, C. Ryu, P. Cladé, V. Natarajan, A. Vaziri, K. Helmerson, and W. D. Phillips, Quantized rotation of atoms from photons with orbital angular momentum, Phys. Rev. Lett. 97, 170406 (2006).
  • Wright et al. [2013a] K. C. Wright, R. B. Blakestad, C. J. Lobb, W. D. Phillips, and G. K. Campbell, Driving phase slips in a superfluid atom circuit with a rotating weak link, Phys. Rev. Lett. 110, 025302 (2013a).
  • Eckel et al. [2014a] S. Eckel, J. G. Lee, F. Jendrzejewski, N. Murray, C. W. Clark, C. J. Lobb, W. D. Phillips, M. Edwards, and G. K. Campbell, Hysteresis in a quantized superfluid ‘atomtronic’ circuit, Nature 506, 200 (2014a).
  • Carleo et al. [2019] G. Carleo, I. Cirac, K. Cranmer, L. Daudet, M. Schuld, N. Tishby, L. Vogt-Maranto, and L. Zdeborová, Machine learning and the physical sciences, Reviews of Modern Physics 91, 045002 (2019).
  • Zhou et al. [2018] Q. Zhou, P. Tang, S. Liu, J. Pan, Q. Yan, and S.-C. Zhang, Learning atoms for materials discovery, Proceedings of the National Academy of Sciences 115, E6411–E6417 (2018).
  • Barker et al. [2020] A. J. Barker, H. Style, K. Luksch, S. Sunami, D. Garrick, F. Hill, C. J. Foot, and E. Bentine, Applying machine learning optimization methods to the production of a quantum gas, Machine Learning: Science and Technology 1, 015007 (2020).
  • Mendelson et al. [2003] S. Mendelson, A. J. Smola, and J. van Leeuwen, Advanced Lectures on Machine Learning, Lecture Notes in Computer Science, Vol. 2600 (Springer Berlin / Heidelberg, Berlin, Heidelberg, 2003).
  • Jones et al. [1998] D. Jones, M. Schonlau, and W. Welch, Efficient Global Optimization of Expensive Black-Box Functions, Journal of Global Optimization 13, 455 (1998).
  • Bonaccorso [2018] G. Bonaccorso, Mastering Machine Learning Algorithms, 1st ed. (Packt Publishing, Birmingham, United Kingdom, 2018).
  • Nakamura et al. [2019] I. Nakamura, A. Kanemura, T. Nakaso, R. Yamamoto, and T. Fukuhara, Non-standard trajectories found by machine learning for evaporative cooling of 87Rb atoms, Optics Express 27, 20435 (2019).
  • Bell et al. [2016] T. A. Bell, J. A. P. Glidden, L. Humbert, M. W. J. Bromley, S. A. Haine, M. J. Davis, T. W. Neely, M. A. Baker, and H. Rubinsztein-Dunlop, Bose-Einstein condensation in large time-averaged optical ring potentials, New Journal of Physics 18, 035003 (2016).
  • Gauthier [2019] G. Gauthier, Transport and turbulence in quasi-uniform and versatile Bose-Einstein condensatesPh.D. thesis, The University of Queensland (2019).
  • Wright et al. [2013b] K. C. Wright, R. B. Blakestad, C. J. Lobb, W. D. Phillips, and G. K. Campbell, Threshold for creating excitations in a stirred superfluid ring, Phys. Rev. A 88, 063633 (2013b).
  • Yakimenko et al. [2015] A. I. Yakimenko, K. O. Isaieva, S. I. Vilchinskii, and E. A. Ostrovskaya, Vortex excitation in a stirred toroidal Bose-Einstein condensate, Phys. Rev. A 91, 023607 (2015).
  • Murray et al. [2013] N. Murray, M. Krygier, M. Edwards, K. C. Wright, G. K. Campbell, and C. W. Clark, Probing the circulation of ring-shaped Bose-Einstein condensates, Phys. Rev. A 88, 053615 (2013).
  • Haug et al. [2021] T. Haug, R. Dumke, L.-C. Kwek, C. Miniatura, and L. Amico, Machine-learning engineering of quantum currents, Phys. Rev. Research 3, 013034 (2021).
  • Wigley et al. [2016] P. B. Wigley, P. J. Everitt, A. van den Hengel, J. W. Bastian, M. A. Sooriyabandara, G. D. McDonald, K. S. Hardman, C. D. Quinlivan, P. Manju, and C. C. N. Kuhn, Fast machine-learning online optimization of ultra-cold-atom experiments, Scientific Reports 6 (2016).
  • Gauthier et al. [2016] G. Gauthier, I. Lenton, N. M. Parry, M. Baker, M. J. Davis, H. Rubinsztein-Dunlop, and T. W. Neely, Direct imaging of a digital-micromirror device for configurable microscopic optical potentials, Optica 3, 1136 (2016).
  • Gauthier et al. [2020] G. Gauthier, T. A. Bell, M. Baker, H. Rubinsztein-Dunlop, and T. W. Neely, Feedforward Optimisation of Optical Trapping Potentials for Ultracold Atoms, in 2020 Conference on Lasers and Electro-Optics Pacific Rim (CLEO-PR) (2020) pp. 1–2.
  • Andrews et al. [1997] M. R. Andrews, C. G. Townsend, H.-J. Miesner, D. S. Durfee, D. M. Kurn, and W. Ketterle, Observation of Interference Between Two Bose Condensates, Science 275, 637 (1997).
  • Eckel et al. [2014b] S. Eckel, F. Jendrzejewski, A. Kumar, C. J. Lobb, and G. K. Campbell, Interferometric Measurement of the Current-Phase Relationship of a Superfluid Weak Link, Phys. Rev. X 4, 031052 (2014b).
  • Wigley [2017] P. B. Wigley, Generating and observing soliton dynamics in Bose-Einstein Condensates, Ph.D. thesis, Department of Quantum Science, Research School of Physics and Engineering, The Australian National University (2017).
  • Mehta et al. [2019] P. Mehta, M. Bukov, C.-H. Wang, A. G. Day, C. Richardson, C. K. Fisher, and D. J. Schwab, A high-bias, low-variance introduction to machine learning for physicists, Physics Reports 810, 1 (2019).
  • Note [1] With very few fringes present, the fringes themselves wrap around the entire interference region. The detection algorithm effectively counts periodicity of density in these regions meaning that such wrapping obscures the measurement.
  • Kiehn et al. [2022] H. Kiehn, V. P. Singh, and L. Mathey, Superfluidity of a laser-stirred Bose-Einstein condensate, Phys. Rev. A 105, 043317 (2022).
  • Baggaley and Parker [2018] A. W. Baggaley and N. G. Parker, Kelvin-Helmholtz instability in a single-component atomic superfluid, Phys. Rev. A 97, 053608 (2018).
  • Hernandez-Rajkov et al. [2023] D. Hernandez-Rajkov, N. Grani, F. Scazza, G. Del Pace, W. Kwon, M. Inguscio, K. Xhani, C. Fort, M. Modugno, F. Marino, et al., Universality of the superfluid Kelvin-Helmholtz instability by single-vortex tracking, arXiv preprint arXiv:2303.12631  (2023).
  • Reeves et al. [2015] M. T. Reeves, T. P. Billam, B. P. Anderson, and A. S. Bradley, Identifying a Superfluid Reynolds Number via Dynamical Similarity, Phys. Rev. Lett. 114, 155302 (2015).
  • Raman et al. [1999] C. Raman, M. Köhl, R. Onofrio, D. S. Durfee, C. E. Kuklewicz, Z. Hadzibabic, and W. Ketterle, Evidence for a critical velocity in a Bose-Einstein condensed gas, Phys. Rev. Lett. 83, 2502 (1999).
  • Stießberger and Zwerger [2000] J. S. Stießberger and W. Zwerger, Critcal velocity of superfluid flow past large obstacles in Bose-Einstein condensates, Phys. Rev. A 62, 061601 (2000).
  • Kwon et al. [2015] W. Kwon, G. Moon, S. Seo, and Y.-i. Shin, Critical Velocity for Vortex Shedding in a Bose-Einstein Condensate, Phys. Rev. A 91 (2015).
  • Note [2] We note that the lower winding numbers in O.III and O.IV result from the strong weighting on the stirring time punishment term in the cost functions. This observation highlights the trade-offs introduced in the cost functions which are likely the limiting factors preventing more complete optimization in more complicated scenarios.
  • Menchon-Enrich et al. [2016] R. Menchon-Enrich, A. Benseny, V. Ahufinger, A. D. Greentree, T. Busch, and J. Mompart, Spatial adiabatic passage: a review of recent progress, Reports on Progress in Physics 79, 074401 (2016).
  • Guéry-Odelin et al. [2019] D. Guéry-Odelin, A. Ruschhaupt, A. Kiely, E. Torrontegui, S. Martínez-Garaot, and J. G. Muga, Shortcuts to adiabaticity: Concepts, methods, and applications, Reviews of Modern Physics 91, 045001 (2019).
  • Pandey et al. [2019] S. Pandey, H. Mas, G. Drougakis, P. Thekkeppatt, V. Bolpasi, G. Vasilakis, K. Poulios, and W. von Klitzing, Hypersonic Bose–Einstein condensates in accelerator rings, Nature 570, 205 (2019).
  • Edwards [2013] M. Edwards, Atom SQUID, Nature Physics 9, 68 (2013).
  • Gross et al. [2010] C. Gross, T. Zibold, E. Nicklas, J. Estève, and M. K. Oberthaler, Nonlinear atom interferometer surpasses classical precision limit, Nature 464, 1165 (2010).
  • Levy et al. [2007] S. Levy, E. Lahoud, I. Shomroni, and J. Steinhauer, The a.c. and d.c. Josephson effects in a Bose–Einstein condensate, Nature 449, 579 (2007).
  • Packard and Vitale [1992] R. E. Packard and S. Vitale, Principles of superfluid-helium gyroscopes, Phys. Rev. B 46, 3540 (1992).
  • Woffinden et al. [2023] C. W. Woffinden, A. J. Groszek, G. Gauthier, B. J. Mommers, M. W. J. Bromley, S. A. Haine, H. Rubinsztein-Dunlop, M. J. Davis, T. W. Neely, and M. Baker, Viability of rotation sensing using phonon interferometry in Bose-Einstein condensates, SciPost Phys. 15, 128 (2023).
  • Gajdacz et al. [2013] M. Gajdacz, P. L. Pedersen, T. Mørch, A. J. Hilliard, J. Arlt, and J. F. Sherson, Non-destructive Faraday imaging of dynamically controlled ultracold atoms, Review of Scientific Instruments 84, 083105 (2013).
  • Wilson [2015] K. E. Wilson, Developing a Toolkit for Experimental Studies of Two-Dimensional Quantum Turbulence in Bose-Einstein CondensatesPh.D. thesis, The University of Arizona. (2015).
  • Bauckhage and Sifa [2015] C. Bauckhage and R. Sifa, k-Maxoids Clustering., in Proceedings of the LWA 2015 Workshops: KDML, FGWM, IR, and FGDB, Trier, Germany, October 7-9, 2015, CEUR Workshop Proceedings, Vol. 1458, edited by R. Bergmann, S. Görg, and G. Müller (CEUR-WS.org, 2015) pp. 133–144.
  • Rakonjac et al. [2016] A. Rakonjac, A. Marchant, T. Billam, J. Helm, M. Yu, S. Gardiner, and S. Cornish, Measuring the disorder of vortex lattices in a Bose-Einstein condensate, Phys. Rev. A 93, 013607 (2016).
  • Gauthier [2020] G. Gauthier, Creation and Dynamics of Onsager Vortex Clusters, in Transport and Turbulence in Quasi-Uniform and Versatile Bose-Einstein Condensates (Springer International Publishing, 2020) pp. 139–169.
  • Gauthier et al. [2019] G. Gauthier, M. T. Reeves, X. Yu, A. S. Bradley, M. A. Baker, T. A. Bell, H. Rubinsztein-Dunlop, M. J. Davis, and T. W. Neely, Giant vortex clusters in a two-dimensional quantum fluid, Science 364, 1264 (2019).