<<

PHYSICAL REVIEW X 9, 021012 (2019) Featured in Physics

Long-Lived and Transient Supersolid Behaviors in Dipolar Quantum

L. Chomaz,1 D. Petter,1 P. Ilzhöfer,2 G. Natale,1 A. Trautmann,2 C. Politi,2 G. Durastante,1,2 R. M. W. van Bijnen,2 A. Patscheider,1 M. Sohmen,1,2 M. J. Mark,1,2 and F. Ferlaino1,2,* 1Institut für Experimentalphysik, Universität Innsbruck, Technikerstraße 25, 6020 Innsbruck, Austria 2Institut für Quantenoptik und Quanteninformation, Österreichische Akademie der Wissenschaften, Technikerstraße 21a, 6020 Innsbruck, Austria

(Received 11 March 2019; published 19 April 2019; corrected 25 April 2019)

By combining theory and experiments, we demonstrate that dipolar quantum gases of both 166Er and 164Dy support a state with supersolid properties, where a spontaneous density modulation and a global coherence coexist. This paradoxical state occurs in a well-defined parameter range, separating the phases of a regular Bose-Einstein condensate and of an insulating droplet array, and is rooted in the mode softening, on the one side, and in the stabilization driven by quantum fluctuations, on the other side. Here, we identify the parameter regime for each of the three phases. In the experiment, we rely on a detailed analysis of the interference patterns resulting from the free expansion of the , quantifying both its density modulation and its global phase coherence. Reaching the phases via a slow interaction tuning, starting from a stable condensate, we observe that 166Er and 164Dy exhibit a striking difference in the lifetime of the supersolid properties, due to the different atom loss rates in the two systems. Indeed, while in 166Er the supersolid behavior survives only a few tens of milliseconds, we observe coherent density modulations for more than 150 ms in 164Dy. Building on this long lifetime, we demonstrate an alternative path to reach the supersolid regime, relying solely on evaporative cooling starting from a thermal gas.

DOI: 10.1103/PhysRevX.9.021012 Subject Areas: Atomic and Molecular Physics, Condensed Physics, Quantum Physics

I. INTRODUCTION and Rydberg particles [9,10], cold atoms with a soft- core potential [11], or lattice-confined systems [7]. Supersolidity is a paradoxical quantum phase of matter Breakthrough experiments with Bose-Einstein condensates where both crystalline and superfluid order coexist [1–3]. (BECs) coupled to light have recently demonstrated a state Such a counterintuitive phase, featuring rather antithetic with supersolid properties [12,13]. While in these systems properties, has been originally considered for quantum indeed two continuous symmetries are broken, the with mobile bosonic vacancies, the latter being periodicity is set by the laser wavelength, making the responsible for the superfluid order. 4He has long supersolid incompressible. been considered a prime system to observe such a phe- Another key notion concerns the close relation between a nomenon [4,5]. However, after decades of theoretical and possible transition to a supersolid ground state and the experimental efforts, an unambiguous proof of superso- 4 existence of a local energy minimum at large momentum lidity in solid He is still missing [6,7]. in the excitation spectrum of a nonmodulated superfluid, In search of more favorable and controllable systems, known as the roton mode [14]. Since excitations corre- ultracold atoms emerged as a very promising candidate, sponding to a periodic density modulation at the roton thanks to their highly tunable interactions. Theoretical wavelength are energetically favored, the existence of this works point to the existence of a supersolid ground state mode indicates the system’s tendency to crystallize [15] in different cold-atom settings, including dipolar [8] and it is predicted to favor a transition to a supersolid ground state [4,5,9]. *Corresponding author. Remarkably, BECs of highly magnetic atoms, in which [email protected] the particles interact through the long-range and anisotropic dipole-dipole interaction (DDI), appear to gather several Published by the American Physical Society under the terms of key ingredients for realizing a supersolid phase. First, the Creative Commons Attribution 4.0 International license. Further distribution of this work must maintain attribution to as predicted more than 15 years ago [16,17] and recently the author(s) and the published article’s title, journal citation, demonstrated in experiments [18,19], the partial attraction and DOI. in momentum space due to the DDI gives rise to a roton

2160-3308=19=9(2)=021012(12) 021012-1 Published by the American Physical Society L. CHOMAZ et al. PHYS. REV. X 9, 021012 (2019) minimum. The corresponding excitation energy, i.e., the caused by three-body collisions [35,36]. These realizations roton gap, can be tuned in the experiments down to raise the crucial question of whether a long-lived or vanishing values. Here, the excitation spectrum softens stationary supersolid state can be created despite the at the roton momentum and the system becomes unstable. usually non-negligble atom losses and the crossing of a Second, there is a nontrivial interplay between the trap discontinuous , which inherently creates geometry and the phase diagram of a dipolar BEC. For excitations in the system. instance, our recent observations have pointed out the In this work, we study both experimentally and theo- advantage of axially elongated trap geometries (i.e., cigar retically the phase diagram of degenerate gases of highly shaped) compared to the typically considered cylindrically magnetic atoms beyond the roton softening. Our inves- symmetric ones (i.e., pancake shaped) in enhancing the tigations are carried out using two different experimental visibility of the roton excitation in experiments. Last but setups producing BECs of 166Er [22,37] and of 164Dy [38] not least, while the concept of a fully softened mode is and rely on a fine-tuning of the contact-interaction strength typically related to instabilities and disruption of a coherent in both systems. In the regime of interest, these two atomic quantum phase, groundbreaking works in the quantum-gas species have different contact-interaction scattering lengths community have demonstrated that quantum fluctuations as, whose precise dependence on the magnetic field is can play a crucial role in stabilizing a dipolar BEC [20–26]. known only for Er [18,22,39], and different three-body-loss Such a stabilization mechanism enables the existence, rate coefficients. Moreover, Er and Dy possess different beyond the mean-field instability, of a variety of stable magnetic moments μ and masses m, yielding the dipolar a μ μ2m=12πℏ2 65 5a 131a ground states, from a single macrodroplet [22,24,27] to lengths, dd ¼ 0 ,of . 0 and 0, respec- striped phases [28], and droplet crystals [29]; see also tively. Here, μ0 is the vacuum permeability, ℏ ¼ h=2π the related works [30–33]. For multidroplet ground states, reduced Planck constant, and a0 the Bohr radius. For both efforts have been devoted to understanding if a phase systems, we find states showing hallmarks of supersolidity, coherence among ground-state droplets could be estab- namely, the coexistence of density modulation and global lished [28,29]. However, previous experiments with 164Dy phase coherence. For such states, we quantify the extent of have shown the absence of phase coherence across the the as parameter range for their existence and study their 166 droplets [28], probably due to the limited atom numbers. lifetime. For Er, we find results very similar to the one 162 Droplet ground states, quantum stabilization, and dipolar recently reported for Dy [35,36], both systems being have caused a huge amount of excitement with very limited by strong three-body losses, which destroy the recent advancements adding key pieces of information to supersolid properties in about half of a trap period. 164 the supersolid scenario. The quench experiments in an However, for Dy, we have identified an advantageous 166Er BEC at the roton instability have revealed out-of- magnetic-field region where losses are very low and large equilibrium modulated states with an early-time phase BECs can be created. In this condition, we observe that the coherence over a timescale shorter than a quarter of the supersolid properties persist over a remarkably long time, oscillation period along the weak-trap axis [18]. In the same well exceeding the trap period. Based on such a high work, it has been suggested that the roton softening stability, we finally demonstrate a novel route to reach the combined with the quantum stabilization mechanism supersolid state, based on evaporative cooling from a may open a promising route towards a supersolid ground thermal gas. state. A first confirmation came from a recent theoretical work [34], considering an Er BEC in an infinite elongated II. THEORETICAL DESCRIPTION trap with periodic boundary conditions and tight transverse As a first step in our study of the supersolid phase in confinement. The supersolid phase appears to exist within a dipolar BECs, we compute the ground-state phase diagram narrow region in interaction strength, separating a roton for both 166Er and 164Dy quantum gases. The gases are excitation with a vanishing energy and an incoherent confined in a cigar-shaped harmonic trap, as illustrated in assembly of insulating droplets. Almost simultaneously, 162 Fig. 1(a). Our theory is based on numerical calculations of experiments with Dy BECs in a shallow elongated trap, the extended Gross-Pitaevskii equation [40], which performing a slow tuning of the contact interaction, includes our anisotropic trapping potential, the short-range reported on the production of stripe states with phase contact and long-range dipolar interactions at a mean-field coherence persisting up to half of the weak trapping period level, as well as the first-order beyond-mean-field correc- [35]. More recently, such observations have been con- tion in the form of a Lee-Huang-Yang (LHY) term 162 firmed in another Dy experiment [36]. Here, theoretical [18,22–24,27]. We note that, while both the exact strength calculations showed the existence of a phase-coherent of the LHY term and its dependence on the gas character- droplet ground state, linking the experimental findings to istics are under debate [18,19,25,31,41], the importance of the realization of a state with supersolid properties. The such a term, scaling with a higher power in density, is results on 162Dy show, however, transient supersolid prop- essential for stabilizing states beyond the mean-field erties whose lifetime is limited by fast inelastic losses instability [18,25,41]; see also Refs. [8,42–44].

021012-2 LONG-LIVED AND TRANSIENT SUPERSOLID BEHAVIORS … PHYS. REV. X 9, 021012 (2019)

(a) z 0 0.1 0.2 0.3 0.4 0.5 B y 4 -1 S x n (10 µm ) 10

(f) (g) ) 1

(b) 4 - 0.6

-5 µm 0.4 8 ID SSP BEC 4 0 2 0.2 (10 (µm) n

) ID

y * 0

5 4 -10 0 10 6 0 y (µm) ( 10

44 46 48 50 52 54 N SSP aS (a0) 4

0.8 BEC 2 (c) (d)0.4 (e)

m ) 2 µ 0.4 166Er 164Dy 4 -1 1 0.2

(10 45 50 55 90 95 100 n 0 0 0 -10 0 10 -10 0 10 -10 0 10 aS (a0) aS (a0) y (µm) y (µm) y (µm)

FIG. 1. Phase diagram of an 166Er and a 164Dy dipolar BEC in a cigar-shaped trap. (a) Illustration of the trap geometry with atomic 166 4 dipoles oriented along z. (b) Integrated density profile as a function of as for an Er ground state of N ¼ 5 × 10 . In the color bar, the density scale is upper limited to 4 × 104 μm−1 in order to enhance the visibility in the supersolid regime. (c)–(e) Exemplary density profiles for an insulating droplet state (ID) at as ¼ 49a0, for a state with supersolid properties (SSP) at 51a0, and for a BEC at 52a0, 166 164 respectively. (f),(g) Phase diagrams for Er and Dy for trap frequencies ωx;y;z ¼ 2π × ð227; 31.5; 151Þ and 2π × ð225; 37; 135Þ Hz, respectively. The gray color identifies ground states with a single peak in nðyÞ of large Gaussian width, σy > 2ly. The dark blue region in (f) shows the region where nðyÞ exhibits a single sharp peak, σy ≤ 2ly, and no density modulation. The red-to-blue color map shows S in the case of a density-modulated nðyÞ. In (g) the color map is upper limited to use the same color code as in (f) and to enhance visibility 164 4 in the low-N regime. The inset in (g) shows the calculated density profile for Dy at N ¼ 7 × 10 and as ¼ 91a0.

Our theoretical results are summarized in Fig. 1.By [27,28,45]; see the ID region in Figs. 1(f) and 1(g). For low varying the condensed-atom number N and as, the phase N, we find a single droplet of high peak density, as in diagram shows three very distinct phases. To illustrate Refs. [24,27]; see dark blue region in Fig. 1(f). Generally them, we first describe the evolution of the integrated in situ speaking, our calculations show that the number of droplets density profile nðyÞ with fixed N for varying as, Fig. 1(b). in the array decreases with lowering as or N. The existence The first phase, appearing at large as, resembles a regular of these three phases (BEC, SSP, ID) is consistent with dilute BEC. It corresponds to a nonmodulated density recent calculations considering an infinitely elongated 162 profile of low peak density and large axial size σy exceed- Er BEC [34] and a cigar-shaped Dy BEC [36], illustrat- ing several timespffiffiffiffiffiffiffiffiffiffiffiffiffiffi the corresponding harmonic oscillator ing the generality of this behavior in dipolar gases. length (ly ¼ ℏ=mωy); see Fig. 1(e) and the region To study the supersolid character of the density-modu- denoted BEC in Figs. 1(f) and 1(g). The second phase lated phases, we compute the average of the wave function S appears when decreasing as down to a certain critical value, overlap between neighboring droplets . As an ansatz to S as. Here, the system undergoes an abrupt transition to a extract , we use a Gaussian function to describe the wave periodic density-modulated ground state, consisting of an function of each individual droplet. This is found to be an array of overlapping narrow droplets, each of high peak appropriate description from an analysis of the density density. Because the droplets are coupled to each other via a profiles of Figs. 1(b)–1(d); see also Ref. [46]. For two density overlap, later quantified in terms of the link strength droplets at a distance d and of identical Gaussian widths σy 2 2 S, particles can tunnel from one droplet to a neighboring along the array direction, S is simply S ¼ expð−d =4σyÞ. one, establishing a global phase coherence across the cloud; Here, we generalize the computation of the wave function see Fig. 1(d). Such a phase, in which periodic density overlap to account for the difference in widths and modulation and phase coherence coexist, is identified as amplitudes among neighboring droplets. This analysis the supersolid (SSP) one [10,34]; see the SSP region in allows us to distinguish between the two types of modu- Figs. 1(f) and 1(g). When further decreasing as, we observe lated ground states, SSP and ID in Figs. 1(f) and 1(g). a fast reduction of the density overlap, which eventually Within the Josephson-junction picture [47–49], the tunnel- vanishes; see Fig. 1(c). Here, the droplets become fully ing rate of atoms between neighboring droplets depends on separated. Under realistic experimental conditions, it is the wave function overlap, and an estimate for the single- expected that the phase relation between such droplets particle tunneling rate can be derived within the Gaussian cannot be maintained; see later discussion. We identify this approximation [46]; see also Ref. [40]. The ID phase third phase as the one of an insulating droplet (ID) array corresponds to vanishingly small values of S, yielding

021012-3 L. CHOMAZ et al. PHYS. REV. X 9, 021012 (2019) tunneling times extremely long compared to any other Ref. [40]. Similarly to Er, the Dy BEC is also confined in a relevant timescale. In contrast, the supersolid phase is cigar-shaped optical dipole trap and a homogeneous identified by a substantial value of S, with a correspond- magnetic field B sets the quantization axis along z and ingly short tunneling time. the value of as. For Dy, we will discuss our results in As shown in Figs. 1(f) and 1(g), a comparative analysis terms of magnetic field B, since the as-to-B conversion is of the phase diagram for 166Er and 164Dy reveals similarities not well known in the magnetic-field range considered between the two species (see also Ref. [36]). A supersolid [25,40,41,52]. In a first set of measurements, we first phase is found for sufficiently high N, in a narrow region produce a stable BEC of about N ¼ 3.5 × 104 condensed of as, upper bounded by the critical value as ðNÞ.For atoms at a magnetic field of B ¼ 2.5 G and then probe the intermediate N, as increases with increasing N. We note phase diagram by tuning as. Here, before ramping the that, for low N, the nonmodulated BEC evolves directly magnetic field to access the interesting as regions, we into a single droplet state for decreasing as [50]. In this slowly increase the power of the trapping beams within case, no supersolid phase is found in between; see also 200 ms. The final trap frequencies are ωx;y;z ¼ 2π × Refs. [24,27]. Despite the general similarities, we see that ð300; 16; 222Þ Hz. After preparing a stable BEC, we ramp the supersolid phase for 164Dy appears for lower atom B to the desired value within 20 ms and hold the sample for number than for Er and has a larger extension in as. This is th [40]. In a second set of measurements, we study a a mainly due to the different dd and strength of the LHY completely different approach to reach the supersolid state. term. We note that, at large N and for decreasing as,Dy As discussed later, here we first prepare a thermal sample at exhibits ground states with a density modulation appearing a B value where supersolid properties are observed and then first in the wings, which then progresses inwards until a further cool the sample until a transition to a coherent substantial modulation over the whole cloud is established droplet-array state is reached. In both cases, at the end [51]; see inset of Fig. 1(g). In this regime, we also observe of the experimental sequence, we perform absorption that as decreases with increasing N. These types of states imaging after typically 27 ms of time-of-flight (TOF) have not been previously reported and, although challeng- expansion. The imaging beam propagates horizontally ing to access in experiments because of the large N, they under an angle α of ≈45° with respect to the weak axis deserve further theoretical investigations. of the trap (y). From the TOF images, we thus extract nðkY;kzÞ with kY ¼ cosðαÞky þ sinðαÞkx. 164 III. EXPERIMENTAL SEQUENCE A special property of Dy is that its background 166 164 a FOR Er AND Dy scattering length is smaller than dd. This allows us to enter the supersolid regime without the need of setting B To experimentally access the above-discussed physics, we 166 162 166 164 close to a FR, as is done for Er and Dy, which produce dipolar BECs of either Er or Dy atoms. These typically causes severe atom losses due to increased three- two systems are created in different setups and below we body-loss coefficients. In contrast, in the case of 164Dy, the summarize the main experimental steps; see also Ref. [40]. supersolid regime is reached by ramping B away from the — 166 Erbium. We prepare a stable Er BEC following FR pole used to produce the stable BEC via evaporative the scheme of Ref. [18]. At the end of the preparation, cooling, as the as range of Fig. 1(g) lies close to the the Er BEC contains about N 8 × 104 atoms at ¼ background as reported in Ref. [52]; see also Ref. [40].At a 64 5a s ¼ . 0. The sample is confined in a cigar-shaped the background level, three-body-loss coefficients below ω optical dipole trap with harmonic frequencies x;y;z ¼ 1.3 × 10−41 m6 s−1 have been reported for 164Dy [25]. 2π × ð227; 31.5; 151Þ Hz. A homogeneous magnetic field B polarizes the sample along z and controls the value of as via a magnetic Feshbach resonance (FR) [18,22,40].Our IV. DENSITY MODULATION AND PHASE COHERENCE measurements start by linearly ramping down as within 20 ms and waiting an additional 15 ms so that as reaches its The coexistence of density modulation and phase coher- target value [40]. We note that ramping times between 20 ence is the key feature that characterizes the supersolid and 60 ms have been tested in the experiment and we do not phase and allows us to discriminate it from the BEC and ID record a significant difference in the system’s behavior. cases. To experimentally probe this aspect in our dipolar After the 15-ms stabilization time, we then hold the sample quantum gases, we record their density distribution after a t for a variable time h before switching off the trap. Finally, TOF expansion for various values of as across the phase we let the cloud expand for 30 ms and perform absorption diagram. As for a BEC in a weak optical lattice [53] or for imaging along the z (vertical) direction, from which we an array of BECs [54–56], the appearance of interference extract the density distribution of the cloud in momentum patterns in the TOF images is associated with a density space, nðkx;kyÞ. modulation of the in situ atomic distribution. Moreover, the Dysprosium.—The experimental procedure to create a shot-to-shot reproducibility of the patterns (in amplitude 164Dy BEC follows the one described in Ref. [38]; see also and position) and the persistence of fringes in averaged

021012-4 LONG-LIVED AND TRANSIENT SUPERSOLID BEHAVIORS … PHYS. REV. X 9, 021012 (2019)

n (103 µm2) 3 4 4 25 x10 x10 x10 -6 (a) (d) 6 (g)5 (j) 5 (m) -4 20 5 ) 4 4 1 -2 - 4 15 m

0 (µm) 3 3 (µ 3 10 n y 2 2 k 2 2 4 5 1 1 1 6 0 0 0 -6 (b) (e) 14 (h)5 (k)5 (n) -4 4 12 ) -2 4 4 -1 3 10 0 8 3 3 (µm) (µm

2 n y 2 6 k 2 2 4 1 4 1 1 2 6 0 0 0 0 -6 (c) (f) 2.5 12 (i) (l)5 (o) -4 5 2 10 ) -2 4 4 m) -1 1.5 8 0 (µ 3 3 n

(µm 6

y 2 1 2 k 4 2 4 0.5 2 1 1 6 0 0 0 0 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 -4 -2 0 2 4 -6 -4 -2 0 2 4 6 -1 -1 k (µm-1) k (µm-1) k (µm-1) k (µm-1) y (µm) k (µm ) k (µm ) x x x x y y

FIG. 2. Coherence in the interference patterns: measurement and toy model. (a)–(c) Examples of single TOF absorption images at 166 th ¼ 5 ms for Er at as ¼f54.7ð2Þ; 53.8ð2Þ; 53.3ð2Þga0, respectively. Corresponding average pictures for 100 images obtained under the same experimental conditions (d)–(f) and their Fourier transform (FT) profiles (g)–(i). The gray lines show the FT norm jF½nðyÞj of the individual profiles. The averages, nM (blue squares) and nΦ (red dots), are fitted to three-Gaussian functions (blue solid line and brown dashed line, respectively). The dotted lines show the components of the total fitted function corresponding to the two side peaks in nΦ. (j)–(l) Interference patterns from the toy-model realizations with 100 independent draws using ND ¼ 4, d ¼ 2.8 μm, σy ¼ 0.56 μm (see text) and for different ϕi distributions: (j) ϕi ¼ 0, (k) ϕi normally distributed around 0 with 0.2π standard deviation, (l) ϕi uniformly distributed between 0 and 2π. (m)–(o) Corresponding FT profiles for the toy model, same color code as (g)–(i). pictures, obtained from many repeated images taken under coherence. We identify this behavior as the one of the same experimental conditions, reveals the presence of ID states. phase coherence across the sample [56]. Toy model—To get an intuitive understanding of the Figure 2 exemplifies snapshots of the TOF distributions interplay between density modulation and phase coherence for Er, measured at three different as values; see and to estimate the role of the different sources of Figs. 2(a)–2(c). Even if very close in scattering length, fluctuations in our experiment, we here develop a simple the recorded nðkx;kyÞ shows a dramatic change in behavior. toy model, which is inspired by Ref. [56]; see also For as ¼ 54.7ð2Þa0, we observe a nonmodulated distribu- Ref. [40]. In our model, the initial state is an array of tion with a density profile characteristic of a dilute BEC. ND droplets containing in total N atoms. Each droplet is When lowering as to 53.8ð2Þa0, we observe the appearance described by a one-dimensional Gaussian wave function of an interference pattern in the density distribution, ψ iðyÞ of amplitude αi, phase ϕi, width σi, and center yi.To consisting of a high central peak and two almost symmetric account for fluctuations in the experiments, we allow αi, low-density side peaks [57]. Remarkably, the observed di ¼ yi − yi−1, and σi to vary by 10% around their expect- pattern is very reproducible with a high shot-to-shot ation values. The spread of the phases ϕi among the stability, as shown in the repeated single snapshots and droplets is treated specially as it controls the global phase in the average image [Figs. 2(b) and 2(e)]. This behavior coherence of the array. By fixing ϕi ¼ 0 for each droplet or indicates a coexistence of density modulation and global by setting a random distribution of ϕi, we range from full phase coherence in the in situ state, as expected in the phase coherence to the incoherent cases. Therefore, the supersolid phase. This observation is consistent with degree of phase incoherence can be varied by changing the our previous quench experiments [18] and with the recent standard deviation of the distribution of ϕi. 162 Dy experiments [35,36]. When further lowering as to To mimic our experiment, we compute the free evolution ψ 53.3ð2Þa0, complicated patterns develop with fringes of each individual i over 30P ms, and then compute the 2 varying from shot to shot in number, position, and axial distribution nðy; tÞ¼j iψ iðy; tÞj , from which we amplitude, signaling the persistence of in situ density extract the momentum distribution nðkyÞ, also accounting modulation. However, the interference pattern is com- for the finite imaging resolution [40]. For each computation pletely washed out in the averaged density profiles run, we randomly draw ND values for ϕi, as well as of σi, [Fig. 2(f)], pointing to the absence of a global phase di, and αi, and extract nðkyÞ. We then collect a set of nðkyÞ

021012-5 L. CHOMAZ et al. PHYS. REV. X 9, 021012 (2019) by drawing these values multiple times using the same the toy model [Figs. 2(j)–2(l)], respectively. As expected, statistical parameters and compute the expectation value, for the BEC case, both nM and nΦ show a single peak at hnðkyÞi; see Figs. 2(j)–2(l). The angled brackets denote the zero spatial frequency, y ¼ 0, characterizing the absence of ensemble average. density modulation, Fig. 2(g). In the case of phase-coherent The results of our toy model show large similarity with droplets, Fig. 2(e), we observe that nM and nΦ are the observed behavior in the experiment. In particular, superimposed and both show two symmetric side peaks while for each single realization one can clearly distinguish at finite y, in addition to a dominant peak at y ¼ 0; see multipeak structures regardless of the degree of phase Fig. 2(h). In the incoherent droplet case, we find that, while coherence between the droplets, the visibility of the nM still shows side peaks at finite y, the ones in nΦ wash interference pattern in the averaged nðkyÞ survives only out from the averaging, Figs. 2(f), 2(i), 2(l), and 2(o).For if the standard deviation of the phase fluctuations between both coherent and incoherent droplet arrays, the toy-model droplets is small (roughly, below 0.3π). In the incoherent results show behaviors matching the above description, case, we note that the shape of the patterns strongly varies providing a further justification of our FT analysis scheme; from shot to shot. Interestingly, the toy model also shows see Figs. 2(j)–2(o). Our toy model additionally proves two that the visibility of the coherent peaks in the average interesting features. First, it shows that the equality images is robust against the typical shot-to-shot fluctua- nM ¼ nΦ, revealing the global phase coherence of a tions in droplet size, amplitude, and distance that occur in density-modulated state, is remarkably robust to noise in the experiments; see Figs. 2(j) and 2(k). the structure of the droplet arrays; see Figs. 2(j) and 2(m). Probing density modulation and phase coherence.—To Second, our toy model, however, shows that phase fluc- separate and quantify the information on the in situ density tuations across the droplet array on the order of 0.2π modulation and its phase coherence, we analyze the measured standard deviation are already sufficient to make nΦ and interference patterns in Fourier space [36,58–60].Here,we nM deviate from each other; see Figs. 2(k) and 2(n). The extract two distinct averaged density profiles, nM and nΦ. incoherent behavior is also associated with strong varia- Their structures at finite y spatial frequency (i.e., in Fourier tions in the side peak amplitude of the individual realiza- space) quantify the two abovementioned properties. tions of jF½nj, connecting, e.g., to the observations More precisely, we perform a Fourier transform (FT) of of Ref. [36]. the integrated momentum distributions nðkyÞ denoted Finally, to quantify the density modulation and the phase coherence, we fit a three-Gaussian function to both F½nðyÞ. Generally speaking, modulations in nðkyÞ induce n y n y y y Mð Þ and Φð Þ and extract the amplitudes of the peaks at finite spatial frequency, ¼ , in the FT norm, A A jF½nðyÞj; see Figs. 2(g)–2(i) and 2(m)–2(o). Following the finite-spatial-frequency peaks, M and Φ, for both dis- tributions, respectively. Note that for a BEC, which is a above discussion (see also Refs. [56,61]), such peaks in an A individual realization hence reveal a density modulation of phase-coherent state, Φ will be zero since it probes only finite-spatial-frequency peaks; see Figs. 2(g)–2(i) the corresponding in situ state, with a wavelength roughly – equal to y. Consequently, we consider the average of the and 2(m) 2(o). FT norm of the individual images, nMðyÞ¼hjF½nðyÞji,as V. CHARACTERIZATION OF THE the first profile of interest. The peaks of nM at finite y then indicate the mere existence of an in situ density modulation SUPERSOLID STATE of roughly constant spacing within the different realiza- We are now in the position to study two key aspects, tions. As the second profile of interest, we use the FT namely, (i) the evolution of the density modulation and norm of the average profile hnðkyÞi, nΦðyÞ¼jF½hniðyÞj. phase coherence across the BEC-supersolid-ID phases and Connecting to our previous discussion, the peaks of nΦ at (ii) the lifetime of the coherent density-modulated state in finite y point to the persistence of a modulation in the the supersolid regime. average hnðkyÞi, which we identified as a hallmark for a Evolution of the supersolid properties across the phase global phase coherence within the density-modulated state. diagram.—The first type of investigation is conducted with In particular, we point out that a perfect phase coherence, 166Er since, contrary to 164Dy, its scattering length and implying identical interference patterns in all the individual dependence on the magnetic field has been precisely realizations, yields nM ¼ nΦ and, thus, identical peaks characterized [18,22]. After preparing the sample, we ramp at finite y in both profiles. We note that, by linearity, nΦ as to the desired value and study the density patterns as well also matches the norm of the average of the full FT as their phase coherence by probing the amplitudes AM of the individual images, i.e., nΦðyÞ¼jhF½nðyÞij; see and AΦ as a function of as after th ¼ 5 ms. As shown in also Ref. [40]. Fig. 3(a), in the BEC region (i.e., for large as), we observe Figures 2(g)–2(i) and 2(m)–2(o) demonstrate the sig- that both AM and AΦ are almost zero, evidencing the nificance of our FT analysis scheme by applying it expected absence of a density modulation in the system. As to the momentum distributions from the experiment soon as as reaches a critical value as , the system’s behavior [Figs. 2(d)–2(f)] and the momentum distributions from dramatically changes with a sharp and simultaneous

021012-6 LONG-LIVED AND TRANSIENT SUPERSOLID BEHAVIORS … PHYS. REV. X 9, 021012 (2019)

3 x10 the system to establish or maintain a global phase coher- (a) 1 S 0

M ence. In this plot, we set ¼ in the case where no A

4 / modulation is found in the ground state. At the BEC-to- A

M 0 supersolid transition, i.e., at as ¼ as, a density modulation A -2 -1 0 1 abruptly appears in the system’s ground state with S taking A a - a* (a ) 2 s s 0 a finite value. Here, S is maximal, corresponding to a density modulation of minimal amplitude. Below the transition, we observe a progressive decrease of S with 0 a 0.8 lowering s, pointing to the gradual reduction of the (b) tunneling rate in the droplet arrays. Close to the transition, 0.6 we estimate a large tunneling compared to all other relevant timescales. However, we expect this rate to become vanish-

S 0.4 ingly small, on the sub-Hertz level [40], when decreasing as 1–2a0 below as. Our observation also hints at the smooth 0.2 character of the transition from a supersolid to an ID phase. The general trend of S, including the extension in as 0 where it takes nonvanishing values, is similar to the as -2 -1 0 1 behavior of AM and AΦ observed in the experiments [62]. a - a* (a ) s s 0 We observe in the experiments that the as dependence at FIG. 3. Supersolid behavior across the phase diagram. Mea- the BEC-to-supersolid transition appears sharper than at sured side peak amplitudes, AΦ (circles) and AM (squares), with the supersolid-to-ID interface, potentially suggesting a their ratio in inset (a), and calculated link strength S (b) as a different nature of the two transitions. However, more 166 function of as − as for Er. For nonmodulated states, we set investigations are needed since atom losses, finite temper- S ¼ 0 in theory and AΦ=AM ¼ 0 in experiment (crosses in inset). ature, and finite-size effects can affect, and in particular In the inset, open and closed symbols correspond to AΦ=AM > smoothen, the observed behavior [64–66]. Moreover, 0.8 and ≤ 0.8, respectively. In the experiments, we probe the dynamical effects, induced by, e.g., excitations created at t 5 system at a fixed h ¼ ms. Horizontal error bars are derived the crossing of the phase transitions or atom losses during B from our experimental uncertainty in , vertical error bars the time evolution, can also play a substantial role in the corresponding to the statistical uncertainty from the fit are smaller experimental observations, complicating a direct compari- than the data points. The measured and calculated critical son with the ground-state calculations. The time dynamics scattering lengths are as ¼ 54.9ð2Þa0 and 51.15a0, respectively [62]. The numerical results are obtained for the experimental trap as well as a different scheme to achieve a state with frequencies and for a constant N ¼ 5 × 104 [63]. supersolid properties is the focus of the remainder of the paper. Lifetime of the supersolid properties.—Having identified a increase of both AM and AΦ. While the strength of AM the s range in which our dipolar quantum gas exhibits and AΦ varies with decreasing as—first increasing then supersolid properties, the next central question concerns the decreasing—we observe that their ratio AΦ=AM remains stability and lifetime of such a fascinating state. Recent 162 constant and close to unity over a narrow as range below as experiments on Dy have shown the transient character of of ≳1a0 width; see the inset of Fig. 3(a). This behavior the supersolid properties, whose lifetime is limited by pinpoints the coexistence in the system of phase coherence three-body losses [35,36]. In these experiments, the phase and density modulation, as predicted to occur in the coherence is found to survive up to 20 ms after the density supersolid regime. For ðas − as Þ < −1a0, we observe that modulation has formed. This time corresponds to about half the two amplitudes depart from each other. Here, while the of the weak-trap period. Stability is a key issue in the density modulation still survives with AM saturating to a supersolid regime, especially since the tuning of as, used to lower finite value, the global phase coherence is lost with enter this regime, has a twofold consequence on the AΦ=AM < 1, as expected in the insulating droplet phase. inelastic loss rate. First, it gives rise to an increase in Note that we also study the evolution of AΦ and AM in the peak density [see Figs. 1(b)–1(d)] and, second, it may 164Dy, but as a function of B, and find a qualitatively similar lead to an enhancement of the three-body-loss coefficient. behavior. We address this question by conducting comparative 166 164 To get a deeper insight on how our observations compare studies on Er and Dy gases. These two species allow to the phase-diagram predictions (see Fig. 1), we study the us to tackle two substantially different scattering scenarios. 166 link strength S as a function of as; see Fig. 3(b). Since S Indeed, the background value of as for Er (as well as for 162 a quantifies the density overlap between neighboring drop- Dy) is larger than dd. Thus, reaching the supersolid a =a ≈ 1 2–1 4 lets and is related to the tunneling rate of atoms across the regime, which occurs at dd s . . in our geometry, droplet array, it thus provides information on the ability of requires us to tune B close to the pole of a FR. This tuning

021012-7 L. CHOMAZ et al. PHYS. REV. X 9, 021012 (2019) also causes an increase of the three-body-loss rate. In Interestingly, AΦ and AM remain approximately equal contrast, 164Dy realizes the opposite case with the back- during the whole time dynamics; see insets of Figs. 4(a) a ground scattering length smaller than dd. This feature and 4(b). This behavior indicates that it is the strength of the brings the important advantage of requiring tuning B away density modulation itself and not the phase coherence from the FR pole to reach the supersolid regime. As we among droplets that decays over time. Similar results have describe below, this important difference in scattering been found theoretically in Ref. [67]. We connect this properties leads to a strikingly longer lifetime of the decay mainly to three-body losses, especially detrimental 166 164Dy supersolid properties with respect to 166Er and to for Er, and possible excitations created while crossing the recently observed behavior in 162Dy [35,36]. the BEC-to-supersolid phase transition [40]. For compari- The measurements proceed as follows. For both 166Er son, the inset of Fig. 4(a) shows also the behavior in the ID 166 and 164Dy, we first prepare the quantum gas in the stable regime for Er, where AΦ=AM < 1 already at short th and BEC regime and then ramp as to a fixed value in the remains so during the time evolution [40]. supersolid regime for which the system exhibits a state of To get a quantitative estimate of the survival time of coherent droplets (i.e., AΦ=AM ≈ 1); see previous discus- the phase-coherent and density-modulated state, we fit a A t sion. Finally, we record the TOF images after a variable th simple exponential function to Φ and extract Φ, defined 1=10 166 and we extract the time evolution of both AΦ and AM. as the lifetime; see Fig. 4.For Er, we extract The study of these two amplitudes will allow us to answer tΦ ¼ 38ð6Þ ms. For th >tΦ, the interference patterns the question of whether the droplet structure—i.e., the become undetectable in our experiment and we recover density modulation in space—persists in time whereas a signal similar to the one of a nonmodulated BEC state [as the coherence among droplets is lost (AM >AΦ → 0)or in Figs. 2(a) and 2(d)]. These results are consistent with if the density structures themselves vanish in time recent observations of transient supersolid properties in 162 164 (AM ≈ AΦ → 0). Dy [35].For Dy, we observe that the coherent density- As shown in Fig. 4, for both species, we observe that AΦ modulated state is remarkably long-lived. Here, we find t 152 13 and AM decay almost synchronously with a remarkably Φ ¼ ð Þ ms. longer lifetime for 164Dy [Fig. 4(b)] than 166Er [Fig. 4(a)]. The striking difference in the lifetime and robustness of the supersolid properties between 166Er and 164Dy becomes even more visible when studying tΦ as a function of as x103 (B for Dy). As shown in Fig. 5, tΦ for Er remains 5 (a) 1 comparatively low in the investigated supersolid regime

M and slightly varies between 20 and 40 ms. Similarly to the 4 A / recent studies with 162Dy, this finding reveals the transient A M 3 A character of the state and opens the question of whether a 0 A 2 0481216 stationary supersolid state can be reached with these t (ms) h 164 t 1 species. On the contrary, for Dy we observe that Φ first increases with B in the range from 1.8 G to about 0 166Er 1.98 G. Then, for B>1.98 G, tΦ acquires a remarkably x102 large and almost constant value of about 150 ms over a 12 wide B range. This shows the long-lived character of the (b) 1 supersolid properties in our 164Dy quantum gas. We note M

A that over the investigated range, as is expected to monoto- 8 / M A nously increase with B [40]. Such a large value of tΦ A 0 exceeds not only the estimated tunneling time across A 0 60 120 t (ms) 4 h neighboring droplets but also the weak-axis trap period, which together set the typical timescale to achieve global equilibrium and to study collective excitations. 164Dy 0 0 40 80 120 VI. CREATION OF STATES WITH SUPERSOLID th (ms) PROPERTIES BY EVAPORATIVE COOLING FIG. 4. Time evolution of the supersolid properties. Amplitudes The long-lived supersolid properties in 164Dy motivate us AΦ (circles) and AM (squares) in the supersolid regime as a 166 to explore an alternative route to cross the supersolid phase function of the holding time in trap for (a) Er at 54.2ð2Þa0 and for (b) 164Dy at 2.04 G. The solid lines are exponential fits to the transition, namely, by evaporative cooling instead of data. The insets show the time evolution of AΦ=AM for the above interaction tuning. For this set of experiments, we have cases (filled triangles), and, for comparison, in the ID regime modified the waists of our trapping beams in order to (empty triangles) for Er at as ¼ 53.1ð2Þa0 (a). achieve quantum degeneracy in tighter traps with respect to

021012-8 LONG-LIVED AND TRANSIENT SUPERSOLID BEHAVIORS … PHYS. REV. X 9, 021012 (2019)

(a) (b) (a)20 (e) 10 ms (i)

µm) 15 150 2

(10 10 n 5

100 0 (ms) (b)20 (f) (j) t 50 ms

15 µm) 2

10 10 50 ( n 5

0 166 164 Er Dy 100 ms 0 (c)20 (g) (k) 53.5 54 54.5 1.8 1.9 2 2.1 15 µm)

a (a ) B (G) 2 S 0 10 10 ( n

FIG. 5. Survival time of the coherent density-modulated state. time (ms) 5 t 166 a 164 B Φ in Er as a function of s (a) and Dy as a function of (b). 0 The error bars refer to the statistical uncertainty from the fit. The (d)20 (h) 300 ms (l) range of investigation corresponds to the supersolid regime for

µm) 15 which phase-coherent density-modulated states are observed. 2

166 10 This range is particularly narrow for Er. ( 10 n 5 10 µm-1 0 -15 -10 -5 0 5 10 15 the one used for in the previous set of 0246 8 10 k (µm-1) measurements. In this way, the interference peaks in the n (102 µm2) Y supersolid region are already visible without the need to apply a further compression of the trap since the side- FIG. 6. Evaporative cooling to a state with supersolid properties. 164Dy absorption images showing the transition to a state with to-central-peak distance in the momentum distribution – 1=l supersolid properties at 2.43 G (a) (d) and to a BEC state at 2.55 G scales roughly as z [18]. Forced evaporative cooling (i)–(l), via different durations of the last evaporation step. These is performed by reducing the power of the trapping beams durations are 10 ms (a),(i), 50 ms (b),(j), 100 ms (c),(k), and piecewise linearly in subsequent evaporation steps until a 300 ms (d),(l). The density profiles (e)–(h) are integrated over the final trap with frequencies 2π × ð225; 37; 134Þ Hz is central regions of the corresponding absorption images (a)–(d). achieved. During the whole evaporation process, which The color map indicates the atomic density in momentum space. has an overall duration of about 3 s, the magnetic field is kept either at B ¼ 2.43 G, where we observe long-lived interference patterns, or at B ¼ 2.55 G, where we produce By increasing the ramp time, the evaporative cooling a stable nonmodulated BEC. We note that these two B becomes more efficient and we observe the appearance of a values are very close without any FR lying in between [40]. bimodal density profile with a narrow and dense peak at the Figure 6 shows the phase transition from a thermal cloud center, which we identify as a regular BEC; see Fig. 6(b). to a final state with supersolid properties by evaporative By further cooling, the BEC fraction increases and the cooling. In particular, we study the phase transition by characteristic pattern of the supersolid state emerges; see varying the duration of the last evaporation ramp, while Figs. 6(c) and 6(d). The observed evaporation process maintaining the initial and final trap-beam power fixed. shows a strikingly different behavior in comparison This procedure effectively changes the atom number and with the corresponding situation at B ¼ 2.55 G, where temperature in the final trap while keeping the trap the usual thermal-to-BEC phase transition is observed; see parameters unchanged, which is important to not alter Figs. 6(i)–6(l). the final ground-state phase diagram of the system. At the We finally probe the lifetime of the supersolid properties end of the evaporation, we let the system equilibrate and by extracting the time evolution of both the amplitudes AΦ thermalize for th ¼ 100 ms, after which we switch off the and AM, as previously discussed. We use the same trap, let the atoms expand for 26.5 ms, and finally perform experimental sequence as the one in Fig. 6(d)—i.e., 300- absorption imaging. We record the TOF images for differ- ms duration of the last evaporation ramp and 100 ms of ent ramp durations, i.e., for different thermalization times. equilibration time—and subsequently hold the sample in For a short ramp, too many atoms are lost such that the the trap for a variable th. As shown in Fig. 7(a), we observe critical atom number for condensation is not reached, and a very long lifetime with both amplitudes staying large and the atomic distribution remains thermal; see Fig. 6(a). almost constant over more than 200 ms. At longer holding

021012-9 L. CHOMAZ et al. PHYS. REV. X 9, 021012 (2019)

300 (a) (b) with supersolid properties by evaporative cooling, demon- 250 strating a powerful alternative approach to interaction 200

tuning on a BEC. This novel technique provides prospects

M 10

, A 150 1 of creating states with supersolid properties while avoiding 8 A M (c) additional excitations and dynamics. The ability to produce 100 6 / A

A 4 long-lived supersolid states paves the way for future 50 0 0200400 2 investigations on quantum fluctuations and many-body 0 5 µm-1 0 50 100 150 200 250 300 350 400 0 correlations, as well as of collective excitations in such t (ms) n (102 µm2) h an intriguing many-body quantum state. A central goal of these future investigations lies in proving the superfluid FIG. 7. Lifetime of the supersolid properties achieved via character of this phase, beyond its global phase coherence A evaporative cooling. Time evolution of the amplitudes Φ (red [7,34,68–70]. circle) and AM (square) after an evaporation time of 300 ms at 2.43 G and an equilibration time of 100 ms. The inset shows ACKNOWLEDGMENTS the time evolution of AΦ=AM.Atth ¼ 0 ms, the atom number N in the phase-coherent density-modulated component is ¼ We thank R. Bisset, B. Blakie, M. Boninsegni, G. 2 2 2 104 . ð Þ × . (b),(c) Averaged absorption images of 25 realiza- Modugno, T. Pfau, and, in particular, L. Santos for many tions after 50 and 300 ms of holding time, respectively. Note stimulating discussions. Part of the computational results that the thermal background has been subtracted from the images. The color map indicates the atomic density in momentum presented have been achieved using the HPC infrastructure space. LEO of the University of Innsbruck. We acknowledge support by the Austrian Science Fund FWF through the DFG/FWF Forschergruppe (FOR 2247/PI2790), by the time, we observe a slow decay of AΦ and AM, following the ERC Consolidator Grant (RARE, No. 681432), and by a decay of the atom number. Moreover, during the dynamics, NFRI Grant (MIRARE, No. ÖAW0600) from the Austrian the ratio AΦ=AM stays constant. The long lifetime of Academy of Science. G. D. and M. S. acknowledge support the phase-coherent density modulation is also directly by the Austrian Science Fund FWF within the DK-ALM visible in the persistence of the interference patterns in (No. W1259-N27). the averaged momentum density profiles [similar to Fig. 2(e)], both at intermediate and long times; see Note added.—Recently, we became aware of related works Figs. 7(b) and 7(c), respectively. For even longer th,we reporting theoretical studies of the ground-state phase cannot resolve anymore interference patterns in the TOF diagram [71,72]. images. Here, we recover a signal consistent with a regular BEC of low N. Achieving the coherent droplet phase via evaporative cooling is a very powerful alternative path to supersolidity. [1] A. F. Andreev and I. M. Lifshitz, Many-Body Problem in We speculate that, for instance, excitations, which might be Quantum Mechanics and Quantum Statistical Mechanics, important when crossing the phase transitions by inter- Sov. Phys. JETP 29, 1107 (1969). action tuning, may be small or removed by evaporation [2] G. V. Chester, Speculations on Bose-Einstein Condensation when reaching this state kinematically. Other interesting and Quantum Crystals, Phys. Rev. A 2, 256 (1970). “ ” questions, open to future investigations, are the nature of [3] A. J. Leggett, Can a Solid Be Superfluid ?, Phys. Rev. the phase transition, the critical atom number, and the role Lett. 25, 1543 (1970). [4] D. A. Kirzhnits and Y. A. Nepomnyashchii, Coherent Crys- of noncondensed atoms. tallization of Quantum , Sov. Phys. JETP 32, 1191 (1971). VII. CONCLUSIONS [5] T. Schneider and C. P. Enz, Theory of the Superfluid-Solid 4 166 164 Transition of He, Phys. Rev. Lett. 27, 1186 (1971). For both Er and Dy dipolar quantum gases, we have [6] S. Balibar, The Enigma of Supersolidity, Nature (London) identified and studied states showing hallmarks of super- 464, 176 (2010). solidity, namely, global phase coherence and spontaneous [7] M. Boninsegni and N. V. Prokof’ev, Colloquium: Super- density modulations. These states exist in a narrow scatter- : What and Where Are They?, Rev. Mod. Phys. 84, ing-length region, lying between a regular BEC phase and a 759 (2012). phase of an insulating droplet array. While for 166Er, [8] Z.-K. Lu, Y. Li, D. S. Petrov, and G. V. Shlyapnikov, Stable Dilute Supersolid of Two-Dimensional Dipolar , similarly to the recently reported 162Dy case [35,36], the Phys. Rev. Lett. 115, 075303 (2015). observed supersolid properties fade out over a compara- 164 [9] N. Henkel, R. Nath, and T. Pohl, Three-Dimensional Roton tively short time because of atom losses, we find that Dy Excitations and Supersolid Formation in Rydberg-Excited exhibits remarkably long-lived supersolid properties. Bose-Einstein Condensates, Phys. Rev. Lett. 104, 195302 Moreover, we are able to directly create stationary states (2010).

021012-10 LONG-LIVED AND TRANSIENT SUPERSOLID BEHAVIORS … PHYS. REV. X 9, 021012 (2019)

[10] F. Cinti, P. Jain, M. Boninsegni, A. Micheli, P. Zoller, and G. [28] M. Wenzel, F. Böttcher, T. Langen, I. Ferrier-Barbut, and T. Pupillo, Supersolid Droplet Crystal in a Dipole-Blockaded Pfau, Striped States in a Many-Body System of Tilted Gas, Phys. Rev. Lett. 105, 135301 (2010). Dipoles, Phys. Rev. A 96, 053630 (2017). [11] M. Boninsegni, Supersolid Phases of Cold Atom Assem- [29] D. Baillie and P. B. Blakie, Droplet Crystal Ground blies, J. Low Temp. Phys. 168, 137 (2012). States of a Dipolar , Phys. Rev. Lett. 121, [12] J. L´eonard, A. Morales, P. Zupancic, T. Esslinger, and T. 195301 (2018). Donner, Supersolid Formation in a Quantum Gas Breaking [30] D. S. Petrov, Quantum Mechanical Stabilization of a a Continuous Translational Symmetry, Nature (London) Collapsing Bose-Bose Mixture, Phys. Rev. Lett. 115, 543, 87 (2017). 155302 (2015). [13] J.-R. Li, J. Lee, W. Huang, S. Burchesky, B. Shteynas, F. Ç. [31] C. R. Cabrera, L. Tanzi, J. Sanz, B. Naylor, P. Thomas, P. Top, A. O. Jamison, and W. Ketterle, A Stripe Phase with Cheiney, and L. Tarruell, Quantum Liquid Droplets in a Supersolid Properties in Spin–Orbit-Coupled Bose– Mixture of Bose-Einstein Condensates, Science 359, 301 Einstein Condensates, Nature (London) 543, 91 (2017). (2018). [14] L. D. Landau, The Theory of of Helium II, [32] G. Semeghini, G. Ferioli, L. Masi, C. Mazzinghi, L. J. Phys. (Moscow) 5, 71 (1941). Wolswijk, F. Minardi, M. Modugno, G. Modugno, M. [15] P. Nozi`eres, Is the Roton in Superfluid 4He the Ghost of a Inguscio, and M. Fattori, Self-Bound Quantum Droplets Bragg Spot?, J. Low Temp. Phys. 137, 45 (2004). of Atomic Mixtures in Free Space, Phys. Rev. Lett. 120, [16] D. H. J. O’Dell, S. Giovanazzi, and G. Kurizki, Rotons in 235301 (2018). Gaseous Bose-Einstein Condensates Irradiated by a Laser, [33] P. Cheiney, C. R. Cabrera, J. Sanz, B. Naylor, L. Tanzi, and Phys. Rev. Lett. 90, 110402 (2003). L. Tarruell, Bright Soliton to Quantum Droplet Transition in [17] L. Santos, G. V. Shlyapnikov, and M. Lewenstein, Roton- a Mixture of Bose-Einstein Condensates, Phys. Rev. Lett. Maxon Spectrum and Stability of Trapped Dipolar Bose- 120, 135301 (2018). Einstein Condensates, Phys. Rev. Lett. 90, 250403 (2003). [34] S. M. Roccuzzo and F. Ancilotto, Supersolid Behaviour of a [18] L. Chomaz, R. M. W. van Bijnen, D. Petter, G. Faraoni, S. Dipolar Bose-Einstein Condensate Confined in a Tube, Baier, J. H. Becher, M. J. Mark, F. Wächtler, L. Santos, and Phys. Rev. A 99, 041601(R) (2019). F. Ferlaino, Observation of Roton Mode Population in a [35] L. Tanzi, E. Lucioni, F. Fama, J. Catani, A. Fioretti, Dipolar Quantum Gas, Nat. Phys. 14, 442 (2018). C. Gabbanini, R. N. Bisset, L. Santos, and G. Modugno, [19] D. Petter, G. Natale, R. M. W. van Bijnen, A. Patscheider, Observation of a Dipolar Quantum Gas with Metastable M. J. Mark, L. Chomaz, and F. Ferlaino, Probing the Roton Supersolid Properties, Phys. Rev. Lett. 122, 130405 (2019). Excitation Spectrum of a Stable Dipolar Bose Gas, arXiv: [36] F. Böttcher, J.-N. Schmidt, M. Wenzel, J. Hertkorn, M. Guo, 1811.12115. T. Langen, and T. Pfau, Transient Supersolid Properties in [20] H. Kadau, M. Schmitt, M. Wenzel, C. Wink, T. Maier, I. an Array of Dipolar Quantum Droplets, Phys. Rev. X 9, Ferrier-Barbut, and T. Pfau, Observing the Rosensweig 011051 (2019). Instability of a Quantum Ferrofluid, Nature (London) [37] K. Aikawa, A. Frisch, M. Mark, S. Baier, A. Rietzler, R. 530, 194 (2016). Grimm, and F. Ferlaino, Bose-Einstein Condensation of [21] I. Ferrier-Barbut, H. Kadau, M. Schmitt, M. Wenzel, and Erbium, Phys. Rev. Lett. 108, 210401 (2012). T. Pfau, Observation of Quantum Droplets in a Strongly [38] A. Trautmann, P. Ilzhöfer, G. Durastante, C. Politi, M. Dipolar Bose Gas, Phys. Rev. Lett. 116, 215301 (2016). Sohmen, M. J. Mark, and F. Ferlaino, Dipolar Quantum [22] L. Chomaz, S. Baier, D. Petter, M. J. Mark, F. Wächtler, Mixtures of Erbium and Dysprosium Atoms, Phys. Rev. Lett. L. Santos, and F. Ferlaino, Quantum-Fluctuation-Driven 121, 213601 (2018). Crossover from a Dilute Bose-Einstein Condensate to a [39] S. Baier, M. J. Mark, D. Petter, K. Aikawa, L. Chomaz, Z. Macrodroplet in a Dipolar Quantum , Phys. Rev. X 6, Cai, M. Baranov, P. Zoller, and F. Ferlaino, Extended Bose- 041039 (2016). Hubbard Models with Ultracold Magnetic Atoms, Science [23] F. Wächtler and L. Santos, Quantum Filaments in Dipolar 352, 201 (2016). Bose-Einstein Condensates, Phys. Rev. A 93, 061603(R) [40] See Supplemental Material at http://link.aps.org/ (2016). supplemental/10.1103/PhysRevX.9.021012 for details on [24] F. Wächtler and L. Santos, Ground-State Properties and the experimental setup, measurement schemes, analysis, Elementary Excitations of Quantum Droplets in Dipolar and theory calculations. Bose-Einstein Condensates, Phys. Rev. A 94, 043618 [41] I. Ferrier-Barbut, M. Wenzel, F. Böttcher, T. Langen, M. (2016). Isoard, S. Stringari, and T. Pfau, Scissors Mode of Dipolar [25] M. Schmitt, M. Wenzel, F. Böttcher, I. Ferrier-Barbut, and T. Quantum Droplets of Dysprosium Atoms, Phys. Rev. Lett. Pfau, Self-Bound Droplets of a Dilute Magnetic Quantum 120, 160402 (2018). Liquid, Nature (London) 539, 259 (2016). [42] A. Gammal, T. Frederico, L. Tomio, and P. Chomaz, Atomic [26] I. Ferrier-Barbut, M. Wenzel, M. Schmitt, F. Böttcher, and Bose-Einstein Condensation with Three-Body Interactions T. Pfau, Onset of a Modulational Instability in Trapped and Collective Excitations, J. Phys. B 33, 4053 (2000). Dipolar Bose-Einstein Condensates, Phys. Rev. A 97, [43] A. Bulgac, Dilute Quantum Droplets, Phys. Rev. Lett. 89, 011604(R) (2018). 050402 (2002). [27] R. N. Bisset, R. M. Wilson, D. Baillie, and P. B. Blakie, [44] D. S. Petrov, Quantum Mechanical Stabilization of a Col- Ground-State Phase Diagram of a Dipolar Condensate with lapsing Bose-Bose Mixture, Phys. Rev. Lett. 115, 155302 Quantum Fluctuations, Phys. Rev. A 94, 033619 (2016). (2015).

021012-11 L. CHOMAZ et al. PHYS. REV. X 9, 021012 (2019)

[45] F. Wächtler and L. Santos, Ground-State Properties and [59] C. Kohstall, S. Riedl, E. R. S. Guajardo, L. A. Sidorenkov, Elementary Excitations of Quantum Droplets in Dipolar J. H. Denschlag, and R. Grimm, Observation of Interference Bose-Einstein Condensates, Phys. Rev. A 94, 043618 between Two Molecular Bose–Einstein Condensates, New (2016). J. Phys. 13, 065027 (2011). [46] M. Wenzel, F. Böttcher, J.-N. Schmidt, M. Eisenmann, T. [60] L. Chomaz, L. Corman, T. Bienaim´e, R. Desbuquois, C. Langen, T. Pfau, and I. Ferrier-Barbut, Anisotropic Super- Weitenberg, S. Nascimb`ene, J. Beugnon, and J. Dalibard, fluid Behavior of a Dipolar Bose-Einstein Condensate, Emergence of Coherence via Transverse Condensation in a Phys. Rev. Lett. 121, 030401 (2018). Uniform Quasi-Two-Dimensional Bose Gas, Nat. Commun. [47] B. D. Josephson, Possible New Effects in Superconductive 6, 6162 (2015). Tunnelling, Phys. Lett. 1, 251 (1962). [61] S. Hofferberth, I. Lesanovsky, B. Fischer, T. Schumm, and J. [48] J. Javanainen, Oscillatory Exchange of Atoms between Traps Schmiedmayer, Non-Equilibrium Coherence Dynamics in Containing Bose Condensates, Phys. Rev. Lett. 57,3164 One-Dimensional Bose Gases, Nature (London) 449, 324 (1986). (2007). [49] S. Raghavan, A. Smerzi, S. Fantoni, and S. R. Shenoy, [62] The values of as extracted from experiments and from Coherent Oscillations between Two Weakly Coupled Bose- theory disagree by 3.8a0. Such a mismatch is similar to π Einstein Condensates: Josephson Effects, Oscillations, and the one observed in our previous work on the roton Macroscopic Quantum Self-Trapping, Phys. Rev. A 59,620 instability [18,19]. This behavior might be related to the (1999). possibility that beyond-mean-field effects are not pro- [50] A similar behavior is also seen for 164Dy for N<4000 perly accounted for by the conventional LHY correction [out of the scale of Fig. 1(g)]. [51] We note that a similar behavior is also found in our 166Er term; see main text. In addition, other effects, such as phase diagram but for larger N than the one reported here. the dynamics and the atom losses [40], can affect the [52] Y. Tang, A. Sykes, N. Q. Burdick, J. L. Bohn, and B. L. Lev, experimental observations. s-Wave Scattering Lengths of the Strongly Dipolar Bosons [63] This atom number corresponds to the condensed 162Dy and 164Dy, Phys. Rev. A 92, 022703 (2015). atom number measured within the stabilization time of a [53] M. Greiner, O. Mandel, T. Esslinger, T. W. Hänsch, and I. s [40]. Bloch, Quantum Phase Transition from a Superfluid to a [64] M. E. Fisher and A. E. Ferdinand, Interfacial, Boundary, Mott in a Gas of Ultracold Atoms, Nature (London) and Size Effects at Critical Points, Phys. Rev. Lett. 19, 169 415,39(2002). (1967). [54] M. Greiner, I. Bloch, O. Mandel, T. W. Hänsch, and T. [65] Y. Imry and D. Bergman, Critical Points and Scaling Laws Esslinger, Exploring Phase Coherence in a 2D Lattice of for Finite Systems, Phys. Rev. A 3, 1416 (1971). Bose-Einstein Condensates, Phys. Rev. Lett. 87, 160405 [66] Y. Imry, Finite-Size Rounding of a First-Order Phase (2001). Transition, Phys. Rev. B 21, 2042 (1980). [55] B. Paredes, A. Widera, V. Murg, O. Mandel, S. Fölling, I. [67] See the theoretical analysis in the revised version of Cirac, G. V. Shlyapnikov, T. W. Hänsch, and I. Bloch, Ref. [35]. Tonks–Girardeau Gas of Ultracold Atoms in an Optical [68] V. W. Scarola, E. Demler, and S. D. Sarma, Searching for a Lattice, Nature (London) 429, 277 (2004). Supersolid in Cold-Atom Optical Lattices, Phys. Rev. A 73, [56] Z. Hadzibabic, S. Stock, B. Battelier, V. Bretin, and J. 051601(R) (2006). Dalibard, Interference of an Array of Independent Bose- [69] Z.-K. Lu, Y. Li, D. S. Petrov, and G. V. Shlyapnikov, Stable Einstein Condensates, Phys. Rev. Lett. 93, 180403 (2004). Dilute Supersolid of Two-Dimensional Dipolar Bosons, [57] In our absorption imaging a residual asymmetry can be Phys. Rev. Lett. 115, 075303 (2015). [70] F. Cinti and M. Boninsegni, Classical and Quantum observed between ky. This asymmetry, providing a Filaments in the Ground State of Trapped Dipolar Bose sharper peak at ky > 0, stems from optical aberrations in Gases, Phys. Rev. A 96, 013627 (2017). our imperfect optical setup. These imperfections make the [71] Y. Kora and M. Boninsegni, Patterned Supersolids in quantitative analysis of our data more challenging, yet they Dipolar Bose Systems, arXiv:1902.08256. do not affect the quantitative interpretation. In particular, we [72] Y.-C. Zhang, F. Maucher, and T. Pohl, Supersolidity around have checked that the peak structures are also observable a Critical Point in Dipolar Bose Einstein Condensates, along a distinct imaging axis, nonorthogonal to y, using an arXiv:1903.06161. independent imaging setup. [58] M. Takeda, H. Ina, and S. Kobayashi, Fourier-Transform Method of Fringe-Pattern Analysis for Computer-Based Correction: The inadvertent omission of a marker indicating Topography and Interferometry, J. Opt. Soc. Am. 72, 156 “Featured in Physics” has been fixed. (1982).

021012-12