Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Sounds waves and fluctuations in one-dimensional supersolids

L. M. Platt , D. Baillie , and P. B. Blakie


1
Dodd-Walls Centre for Photonic and Quantum Technologies, Dunedin 9054, New Zealand
2
Department of Physics, University of Otago, Dunedin 9016, New Zealand
(Dated: March 29, 2024)
We examine the low-energy excitations of a dilute supersolid state of matter with a one-dimensional crystal
structure. A hydrodynamic description is developed based on a Lagrangian, incorporating generalized elastic
parameters derived from ground state calculations. The predictions of the hydrodynamic theory are validated
against solutions of the Bogoliubov-de Gennes equations, by comparing the speeds of sound, density fluctu-
arXiv:2403.19151v1 [cond-mat.quant-gas] 28 Mar 2024

ations, and phase fluctuations of the two gapless bands. Our results are presented for two distinct supersolid
models: a dipolar Bose-Einstein condensate in an infinite tube and a dilute Bose gas of atoms with soft-core in-
teractions. Characteristic energy scales are identified, highlighting that these two models approximately realize
the bulk incompressible and rigid lattice supersolid limits.

I. INTRODUCTION superfluid fraction in a dipolar supersolid [26].

A supersolid is a state of matter in which superfluid and


crystalline properties coexist as a consequence of the simul-
taneous breaking of phase and translational symmetries. The
properties of supersolids have been discussed in theoretical
works for almost 50 years, prior to their realizations in ultra-
cold atomic systems [1–5]. Experiments with these super-
solids have investigated their basic phase diagram, excita-
tions and dynamics. The excitation spectrum is of great inter-
est, because the broken translational symmetry results in new
gapless excitation branches called Nambu-Goldstone modes
[6]. For the case of supersolids with one-dimensional (1D) FIG. 1. Schematic of systems we consider: (a) A dipolar BEC con-
crystalline structure, two gapless excitation bands have been fined in an infinite tube potential with magnetic dipole moments po-
found. These branches can be classified by the character of larized along z, and (b) a BEC of particles with soft-core interactions.
fluctuations they cause [5, 7, 8], with the lower energy branch Both systems have unconfined motion along x and in the supersolid
transition translational invariance is broken along this axis.
being associated with phase fluctuations and particle tunnel-
ing between lattice sites, and the upper branch being more
strongly associated with density fluctuations arising from the Main motivation of this work is to provide the first quanti-
crystal motion [5, 9–13]. Quantitative descriptions of the ex- tative test of a hydrodynamic description of a supersolids by
citations involve large-scale numerical calculations and do not comparing its predictions to the direct numerical calculation
yield much insight into the underlying physics. However, hy- of the excitations for specific models. The hydrodynamic the-
drodynamic theories offer the possibility to understand the ory depends a set of generalized elastic parameters including
low-energy and long-wavelength aspects of the supersolids, the superfluid fraction, compressibility and lattice elastic pa-
e.g. the speeds of sounds and nature of the fluctuations for rameters, and we show how to determine these from ground
each branch, in terms of fundamental properties of the system. state calculations. In order to illustrate our results we con-
Work on such theories for supersolids dates back to the semi- sider two contrasting systems [see Fig. 1]. First, a dipolar
nal 1969 paper by Andreev and Lifshitz [14], and extended by Bose-Einstein condensate (BEC) in a tube shaped potential
Saslow [15] and Liu [16]. More recently there has been work where transition to a 1D supersolid occurs as the s-wave scat-
on deriving an effective Lagrangian for supersolids by various tering length is changed. This system is the thermodynamic
means. Son [17] has utilized the invariances and broken sym- limit corresponding to experiments that have produced super-
metries of the supersolid state to provide a universal descrip- solids in cigar shaped potentials. Second, a 1D soft-core Bose
tion also see [18, 19] and [20]. Josserand et al. [21, 22] have gas, which has been studied as a basic model of supersolidity
obtained an effective Lagrangian by applying the technique [27–29] and its properties contrast those of the dipolar case.
of homogenisation to a nonlocal Gross-Pitaevskii theory that There are proposals for producing a BEC with soft-core in-
predicts crystallization (also see Refs. [23] and [24]). In recent teractions using atoms that are weakly coupled to a highly
work Šindik et al. [25] have presented a practical proposal for excited Rydberg state [30]. Our results for these two mod-
experiments to investigate the hydrodynamic properties and els show that they have markedly different properties. The
the superfluid fraction using a dipolar supersolid in a toroidal dipolar supersolid arises from a mechanical instability with
trap by using a long-wavelength density-coupled perturbation. the condensate partially collapsing to higher density incom-
We also note a recent experiment that has used an optical lat- pressible state. In contrast, the purely repulsive interactions
tice to excite a self-induced Josephson effect and quantify the of the soft-core model the supersolid arising from a cluster-
2

ing transition, leading to a relatively rigid lattice. These two occurs depends on ω and ρ, and the transition can be contin-
cases are conveniently distinguished by comparing the energy uous or discontinuous. In the results we present here are for
164
scales associated with the isothermal compressibility at con- Dy atoms with add = 130.8 a0 and ω/2π = 150 Hz.
stant strain and the longitudinal elastic modulus of the lattice.
The outline of the paper is as follows. In Sec. II we briefly
introduce the dipolar and soft-core systems we use in this B. 1D soft-core Bose gas
work. The supersolid Lagrangian is introduced in Sec. III
and we discuss the how to obtained the elastic parameters that Our second system is a BEC of atoms free to move in 1D,
appear in the Lagrangian. In Sec. IV we derive the hydro- but interacting via a soft-core interaction potential Usc (x) =
dynamic behavior of the excitations from the Lagrangian and U0 θ(asc − |x|), where asc is the core radius and U0 is the
use this to determine the speeds of sound. We then make a potential strength [see Fig. 1(b)]. It conventional to define the
detailed comparison to calculations of the two models, and dimensionless interaction parameter
explore the limiting behavior of the predictions. The density
fluctuations are discussed in Sec. V. Here a focus is on the 2ma3sc U0 ρ
decomposition of the fluctuations into components related to Λ= , (3)
ℏ2
tunneling of atoms between sites (defect density fluctuations)
and due movement of the crystal lattice (lattice strain fluctu- and continuous transition occurs from a uniform to a modu-
ations). Finally, in Sec. VI we discuss the phase fluctuations lated state at the critical value Λc = 21.05. Further details
before concluding in Sec. VII. about this model are given in Appendix A 2.

II. SYSTEMS C. General solution properties

We consider two zero-temperature systems that undergo a By solving the (generalized) Gross-Pitaevskii equations for
transition to a 1D crystalline state. The models are briefly in- these models we determine the energy minimising wavefunc-
troduced in the following two subsections. Additional details tion ψ0 . These solutions can be found in a single unit cell of
of the models and numerical solution method are given in Ap- length a along x. If the state is crystalline, then a is the lattice
pendix A. Apart from the microscopic parameters particular constant (for uniform states the energy is independent of a).
to each model, we consider both systems in a thermodynamic Take E to be the energy per unit length of this state for den-
limit with a specified density per unit length ρ. sity ρ and lattice constant a (as defined in Appendix A). In the
thermodynamic limit where the system is free to choose the
microscopic length a to minimise the energy, the ground state
A. Dipolar Bose gas in a tube potential is

Our first system is a Bose gas of magnetic atoms confined E0 (ρ) = min E(ρ, a), (4)
a
in a tube shaped potential [see Fig. 1(a)]. The single particle
Hamiltonian is which can be used to implicitly determine a(ρ) as the value of
a that minimises E at density ρ.
ℏ2 ∇ 2 1
Hsp = − + mω 2 (y 2 + z 2 ), (1)
2m 2
III. SUPERSOLID LAGRANGIAN
where ω is the transverse harmonic confinement angular fre-
quency, and the atoms are free to move along the x-axis. The
atomic dipole moments are polarized along z by a bias field A. Quadratic Lagrangian density
and the interactions are described by the potential
Here our interest is in a Lagrangian description of the low-
4πas ℏ2 3add ℏ2 z2
 
energy, long-wavelength features of a system where 1D crys-
U (r) = δ(r) + 1−3 2 , (2)
m mr3 r talline order can occur along the x-direction. We work in a
thermodynamic limit where the ground state is determined by
where as is the s-wave scattering length, add = the average linear number density ρ in addition to the inter-
mµ0 µ2m /12πℏ2 is the dipole length, and µm is the atomic action and relevant transverse confinement parameters. Long-
magnetic moment. The ratio ϵdd = add /as characterises wavelength perturbations of the ground state can be character-
the relative strength of the dipole-dipole to s-wave interac- ized by three fields: {δρ(x, t), u(x, t), ϕ(x, t)}. Here δρ(x, t)
tions. When this parameter is sufficiently large the ground denotes variations in the average density, u denotes the defor-
state undergoes a transition to a crystalline state with mod- mation field of the crystal lattice along the x direction (relative
ulation along x. Quantum fluctuation effects are necessary to the unstrained ground state), and ϕ denotes the superfluid
to stabilize this state and further details about this model are phase field. For treating small departures from equilibrium
given in Appendix A 1. The value of ϵdd where the transition the Lagrangian density can be expanded to quadratic order in
3

these fields as dimensional crystals the αuu term generalizes to the elastic
αρρ αuu tensor. For the 1D crystal αuu can also be identified as the
L = − ℏ δρ ∂t ϕ − (δρ)2 − (∂x u)2 − αρu δρ ∂x u layer-compressibility used to characterize smectic materials
2 2
1  ℏ 2 ℏ2 [19, 35]. Finally, we define the density-strain coupling pa-
+ mρn ∂t u − ∂x ϕ − ρ (∂x ϕ)2 , (5) rameter by the mixed partial derivative
2 m 2m
 2 
where ρn is the normal density, with ρ = ρs + ρn , and ρs be- ∂ E
αρu = a . (11)
ing the superfluid density. Similar forms of Lagrangian den- ∂ρ∂a
sities have been obtained by other authors (e.g. see [18–22]),
although here we follow the approach of Yoo and Dorsey [18].
A discussion about the relationship between these treatments IV. EXCITATIONS AND SPEEDS OF SOUND
is given in Refs. [18, 19]. The phase gradient relates to the
ℏ In this section we consider the collective excitations of the
superfluid velocity as vs = m ∂x ϕ and the normal velocity
relates to the time-derivative of the lattice deformation field gapless energy bands. We begin by deriving hydrodynamic
vn = ∂t u [17, 18]. results for the collective modes, and the associated speeds of
sound. We then compare these to results for our two models,
and consider the limiting behavior.
B. Generalized elastic parameters

In addition to the superfluid density, which determines the A. Hydrodynamic equations of motion and collective modes
rigidity of the state to twists in the phase ϕ [31], three other
generalized elastic parameters of the system {αρρ , αρu , αuu } The Euler-Lagrange equations obtained from Eq. (5) are
also appear in Eq. (5). Here we discuss how these can be
obtained from ground state calculations. ℏ∂t ϕ = −αρρ δρ − αρu ∂x u, (12)
For 1D crystals the superfluid fraction fs = ρs /ρ has ρn (m∂t2 u − ℏ∂tx ϕ) = αuu ∂x2 u+ αρu ∂x δρ, (13)
−1 ℏ
∂t δρ = −ρn ∂tx u − ρs ∂x2 ϕ.
Z
a dx (14)
fs+ = R , (6) m
ρ uc dy dz |ψ0 |2
Z Z −1 These describe the hydrodynamic evolution for small depar-
a dx
fs− = dy dz 2
, (7) tures from equilibrium.
ρ uc |ψ0 | We look for normal mode solutions of Eqs. (12) - (14) of
as upper and lower bounds, respectively [32, 33]. Hereon, the form X(x, t) = Xw ei(qx−ωt) , where X denotes our three
we neglect the transverse coordinates (y, z) when applying fields appearing in the Lagrangian and q is a quasimomentum.
results to the soft-core case. Both bounds are identical for The resulting linear system is
the soft-core case, and provide the exact superfluid fraction 
δρw

(e.g. see [27]). For the dipolar case the bounds are close to M  ϕw  = 0, (15)
each other and taking fs = 12 (fs+ + fs− ) provides an accurate uw
estimate of the superfluid fraction (see [34]).
The other generalized elastic parameters can be determined where
from the energy density E(ρ, a), for the ground state under  
the constraint of lattice constant being a and the mean density αρρ −iℏω iqαρu
being ρ. The first elastic parameter is defined by M =  iℏω ℏ2 q 2 ρs /m −ℏωqρn . (16)
2 2
 2  iqαρu ℏωqρn ω ρn m − q αuu
∂ E
αρρ = , (8) Nontrivial solutions require the matrix M to be singular, i.e. its
∂ρ2 a
determinant ∆M must be zero, where
and relates to the isothermal compressibility at constant strain:
∆M = mℏρn (ω 2 − c2+ q 2 )(ω 2 − c2− q 2 ). (17)
1
κ̃ = 2 . (9) Here we have introduced
ρ αρρ
1
q
The second parameter describes the effects of straining a site mc2± = (a∆ ± a2∆ − 4b∆ ), (18)
at constant density 2
 2  where
2 ∂ E
αuu = a . (10) αuu
∂a2 ρ a∆ = ραρρ − 2αρu + , (19)
ρn
Noting that lattice strain, given by ∂x u, relates to changes ρs 2
b∆ = (αρρ αuu − αρu ). (20)
in the lattice constant according to δa/a = ∂x u. In higher ρn
4

Thus the hydrodynamic excitation frequencies of the two so- 1


lutions are gapless and given by (a)

0.5
ω± = c± q, (21)

with c± being the speeds of sound. 0


1.4
(b)
B. Results for the speeds of sound 1.2

1
We compare the speeds of sound determined from direct
calculation of the excitations to the hydrodynamic result of 0.8
Eq. (18) in Figs. 2 and 3 for the dipolar and soft-core systems,
respectively. The direct calculation of the excitations involves 0.6

solving the Bogoliubov-de Gennes (BdG) equations (e.g. see


0.4
[7–9, 36, 37]). We make this comparison across the uniform
superfluid to supersolid transition. In Figs. 2(a) and 3(a) we 0.2
show the density contrast C defined as
0
max ϱ0 − min ϱ0
C= , (22) 0.06 (c)
max ϱ0 + min ϱ0
0.04
where ϱ0 (x) = dydz|ψ0 (x)|2 is the line density along x.
R
The contrast reveals the appearance of density modulations, 0.02
and hence acts as an order parameter for the formation of
crystalline order. In the dipolar system, at the densities we 0
consider here [37, 38], the crystalline order develops continu-
ously as ϵdd increases past the critical value of ϵdd,c . For the -0.02
soft-core system the transition is also continuous, and occurs
(d)
when the interaction parameter Λ exceeds Λc . In both sys- 0.6
tems, the superfluid fraction decreases with increasing density
contrast.
0.4
The numerical solution of the BdG equations [37] yields
the energies ℏωνq and amplitude functions {uνq , vνq }, where
ν is the band index and q is the quasimomentum along x. We 0.2
then obtain the speeds of sound cν of gapless branches from
the BdG results by fitting ωνq to cν q, for small values of q. In 0
the uniform superfluid phase only a single branch is gapless, 1.38 1.4 1.42 1.44 1.46 1.48 1.5 1.52

and we identify the speed of sound of this branch as ν = +.


In the crystalline state there are two gapless branches, and we
assign the ν of the lower (upper) branches as − (+). Our
FIG. 2. Speeds of sound for a dipolar gas. (a) Density contrast and
results for the speeds of sound are shown in Figs. 2(b) and 3(b)
superfluid fraction. (b) Speeds of sound across the superfluid to su-
for the dipolar and soft-core systems, respectively. There are persolid transition. BdG results (solid lines) and analytic result (18)
significant differences in behavior revealed by these results, similarly colored dots. For reference cκ (thick light blue line) and
particularly in the nature of the discontinuity of the speeds of cu (thick light red line) are shown. (c) The difference of Eq. (18)
sound at the transition point. For the dipolar system c+ jumps from the BdG speeds of sound (solid lines) using same colors as in
downwards as ϵdd crosses the transition1 . In contrast, for the (a). Also shown is this difference setting αρu = 0 (dashed lines) in
soft-core system the c+ speed of sound jumps upwards at the Eq. (18). (d) The generalized elastic parameters compared to mc2κ
transition. Other aspects of the behavior near the transition and mc2u . Other parameters: 164 Dy atoms with ρ = 2500/µm and
will be discussed further in Sec. IV C. radially symmetric transverse confinement of ωy,z /2π = 150 Hz.
Since the BdG speeds of sound and the hydrodynamic re-
sults are in good agreement, in Figs. 2(c) and 3(c) we show
the difference between the results. The relative error is ≲ 1% model, except very close to the transition. The error for the
for the dipolar gas and significantly smaller for the soft-core dipolar model is close to the accuracy that we can determine
the speeds of sound from the BdG calculations. We also show
the difference arising from neglecting the density-strain term
(i.e. setting αρu = 0, which makes our theory equivalent to
1 For the ρ considered the jump is small and hard to see. Ref. [37] shows that in Refs. [19, 25]), noting that over the parameter range
results at other densities where the discontinuity is more clearly revealed. considered this causes a ∼ 5% relative upward-shift in the c+
5

1 ters are shown in Figs. 2(d) and 3(d). These results reveal that
(a) the crystal-dependent elastic terms αρu and αuu both vanish
as we approach the transition from the modulated side. For
0.5
both systems we find that ραρu is typically smaller than αuu ,
suggesting that neglecting this parameter can be a reasonable
0 first approximation. The parameter αρρ is non-zero in both
16
(b) phases, but changes discontinuously at the transition (also see
[37]).
14

12
C. Supersolid sound limiting behavior
10

8 From the hydrodynamic results for the speed of sound in


6 Eq. (18) we derive limiting results to describe their behavior
approaching the transition and deep into the crystal regime. It
4
is useful to introduce two energy scales as
2
2
αρρ αuu − αρu 1
0 mc2κ ≡ ρ = , (23)
1 αuu ρκ
(c) αuu
0.8 mc2u ≡ , (24)
ρn
0.6

0.4
being the bulk compressibility and lattice compressibility
energies, respectively, defining the associated characteristic
0.2
speeds cκ and cu . Here
0
αuu
-0.2 κ= , (25)
ρ2 (αρρ αuu − αρu
2 )

200 is the isothermal compressibility. Because αρu is relatively


small in both our models κ ≈ κ̃ [see Eq. (9), Figs. 2(d) and
150 (d) 3(d)]. The lattice compressibility is only non-zero in the mod-
ulated state. While both αuu and ρn go to zero as we approach
100
the transition from the modulated side, mc2u has a non-zero
value [see Figs. 2(d) and 3(d)].
50

10 20 30 40 50 60 70 80 90 100 1. Uniform superfluid

In the uniform superfluid fs = 1, αuu = αρu = 0 and κ →


1/ρ2 αρρ = κ̃. In this case Eq. (18) yields the single nontrivial
FIG. 3. Speeds of sound for a soft-core gas. (a) Density contrast
and superfluid fraction. (b) Speeds of sound across the superfluid speed of sound c+ = cκ . This situation is well-known for
to supersolid transition. BdG results (solid lines) and analytic result BECs, with the speed of sound being directly related to the
(18) similarly colored dots. For reference cκ (thick light blue line) isothermal compressibility [39].
and cu (thick light red line) are shown. (c) The difference of Eq. (18)
from the BdG speeds of sound (solid lines) using same colors as in
(a). Also shown is this difference setting αρu = 0 (dashed lines) in 2. Approaching the transition
Eq. (18). (d) The generalized elastic parameters compared to mc2κ
and mc2u . Approaching the transition from the modulated side, the
density contrast vanishes and the superfluid fraction ap-
proaches unity. Here the speeds of sound (18) behavior has
speed of sound and a smaller change in c− . Furthermore, the two cases:
role of the density-strain term vanishes as we approach the 
transition. c+ → cκ ,
for cκ > cu , (26)
The hydrodynamic results presented here are determined c− → cu ,
by generalized elastic parameters (i.e. {fs , αρρ , αρu , αuu }),
which are determined from the ground state calculations as 
discussed in Sec. III B. The superfluid fraction results are c+ → cu ,
for cκ < cu . (27)
shown in Figs. 2(a) and 3(a), and the other elastic parame- c− → cκ ,
6

For the dipolar case cκ > cu , while the soft-core case instead 1.5
has cκ < cu [see Figs. 2(d) and 3(d)]. (a)
These results emphasize an important difference between
the two models we consider. The crystalline phase of
the dipolar system is dominated by the compressive energy 1
(i.e. mc2κ > mc2u ), while the soft-core system is dominated by
the lattice compressive energy.

3. Deep in the modulated regime 0.5

Deep in the modulated phase the density contrast ap-


proaches unity and the superfluid fraction vanishes. This is
known as the isolated droplet or classical crystal regime, be- 0
1.38 1.4 1.42 1.44 1.46 1.48 1.5 1.52
cause the ability for atoms to tunnel between unit cells van-
ishes. In this limit we have that
c+ → csp , (28) 16
(b)
c− → 0, (29) 14

where we have introduced the characteristic speed of sound 12


αuu
mc2sp = ραρρ − 2αρu + , (30) 10
ρ
8
which we discuss further in Sec. IV D. The results in Figs. 4(a)
and (b) show that this estimate works well as the superfluid 6
fraction vanishes, although it provides a reasonable estimate
4
of c+ even close to the transition for the dipolar case.
2

4. Rigid lattice limit 0


10 20 30 40 50 60 70 80 90 100

When the lattice compressibility is much larger than the


bulk modulus, i.e. mc2u ≫ mc2κ , the lattice dynamics are
heavily suppressed. Neglecting the αρu term, we have FIG. 4. Comparison of limiting results for the speeds of sound for
p the (a) dipolar and (b) soft-core gases. BdG results (solid lines) is
c− → f s cκ , (31) compared to the limiting results (labelled). Other parameters as in
p Figs. 2 and 3.
c+ → c2u + fn c2κ , (32)
where fn = 1 − fs . This limit is approximately applicable to
the linear-chain model, in which a 1D solid is approximated
the soft-core gas and the results for c± are shown in Fig. 4(b).
by a set of masses joined by springs, yielding a simple the-
ory for the phonons [40]. This idea was recently explored
5. Bulk incompressible limit
and found to be a good description of the dynamics of few
droplets of a dipolar BEC confined by a three-dimensional
cigar-shaped harmonic trap [41].
On the other hand for mc2u ≪ mc2κ , we have To apply this model to the systems we consider here, we
p
c− → f s cu , (33) denote the center-of mass position of the droplet at the j-th
p site by the coordinate xj , with the equilibrium position being
c+ → c2κ + fn c2u . (34) x0j = ja. In the linear-chain model the equation of motion for
the droplet positions is
This limit is approximately applicable to the dipolar gas and
the results for c± are shown in Fig. 4(a). mNuc ẍj = −ksp (2xj − xj−1 − xj+1 ), (35)
where Nuc = ρa is the number of atoms in each unit cell
D. Relationship to linear-chain model (localized droplet) and ksp is the spring constant. We identify
the spring constant from how the energy of the droplet2 E
For a system deep in the modulated regime the ground state
consists of atoms relatively localized in each unit cell (see in-
2 Taking the droplet energy to be the energy in the unit cell, i.e. E = Ea.
sets to Fig. 5). This regime should be well-approximated by
7

140 the frequency of the upper excitation band for dipolar (soft-
core) model. These results confirm that the excitations in this
120
band are closely related to the classical crystal phonon modes.
100

80 V. DENSITY FLUCTUATIONS
4

60 2
We are interested in the density fluctuations caused by the
40 0 quasiparticles of the gapless bands. To do this we examine the
-4 -2 0 2 4
effect of adding an excitation with band index ν and quasimo-
20 mentum q to the condensate with a small coherent amplitude
cνq , i.e.
0
25
ψ(x) = ψ0 (x) + cνq uνq (x) − c∗νq vνq

(x). (38)
20 In particular, we will consider the changes in the linear density
Z
15 ϱ(x) = dy dz |ψ(x)|2 , (39)
4

10 2 relative to the unperturbed case ϱ0 (x). Below we introduce


the various density fluctuation measures and then consider the
0
5 -2 0 2 hydrodynamic limit of these results.

0
0 0.1 0.2 0.3 0.4 0.5 A. Total density fluctuation

The total density fluctuation


FIG. 5. Comparison of the spring dispersion relation (dashed line)  Z 
to the BdG results (solid lines) for (a) the dipolar model with as = δϱ(x) = 2Re cνq dy dz ψ0 (x)[uνq (x) − vνq (x)] , (40)
86 a0 and (b) soft-core model with Λ = 80. The insets show the
linear density of the ground states for reference. Other parameters as
in Figs. 2 and 3. is the leading order expression in cνq for ϱ(x) − ϱ0 (x), and
we have taken ψ0 to be real. In the modulated phase the den-
sity varies rapidly with x, with a dominant periodicity set by
varies with the inter droplet distance a: the lattice constant a. Here we will focus on the density fluc-
 2  tuations arising on length scales much larger than the lattice
∂ E constant, i.e. with wave vectors in the first Brillouin zone. For
ksp = , (36)
∂a2 Nuc a single quasiparticle (cνq = 1) the only non-zero fluctuation
is at wavevector q and −q, with
notably with Nuc held constant. This linear-chain model has
a well-known phonon dispersion relation of the form
Z
δ ρ̃νq = dx e−iqx ψ0 (x)[uνq (x) − vνq (x)]. (41)
2csp  qa 
ωsp (q) = sin , (37)
a 2 We show an example of the density change to a modulated
p
where csp = a ksp /mNuc is the speed of sound and k is the ground state with the addition of a quasiparticle in Fig. 6(a),
wavevector. Evaluating this result in terms of the generalized and the associated long-wavelength total density fluctuation is
elastic parameters yields Eq. (30) for csp . visualized in Fig. 6(c) [also see Appendix B].
In Fig. 5 we compare the dispersion relations for the low-
est two gapless bands obtained from the BdG calculations to
the linear-chain dispersion relation (37), noting that the lat- B. Lattice strain and associated density fluctuation
ter only depends on the elastic parameters ksp obtained from
ground state solutions. The cases considered for these results With crystalline order we can also consider the effect of the
are chosen to be deep in the crystal regime with fs ≈ 8% excitation (38) on the crystal structure. This is characterized
(3%) for the dipolar (soft-core) case. The results show that by the displacement field u(x) which describes the distance
the low q behavior of the upper gapless band is reasonably a point originally at x displaces. We identify the lattice site
well predicted by ωsp (q) [also see cs and c+ in Figs. 2(b) and locations as the local maxima of the linear density. The un-
3(b)]. At higher q values, approaching the band edge, the lin- perturbed locations being x0j = ja (i.e. maxima of ϱ0 ), and
ear chain dispersion slightly overestimates (underestimates) the locations for the perturbed case being xj = x0j + u(x0j )
8

5000

4000

3000

2000

1000

0
0.2

0.1
4500
0
4000
-0.1 3500

-0.2 3000
5.3 5.4 5.5 5.6 5.7 5.8 5.9 6 6.1 6.2

100

-100

-10 -5 0 5 10

0.3 1 1
200

150 0.2

100 0.5 0.5


0.1
50

0 0 0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4

FIG. 6. Effect of an excitation on the density fluctuations of a dipolar supersolid with ϵdd = 1.43. (a) The ground state density ϱ0 (x) (black)
and density ϱ(x) (red) when perturbed by a ν = 1 excitation with q = 2π/9a and cνq = 5. The purely-deformed density profile ϱu (x)
(blue) using the deformation field determined from the quasiparticle. Vertical lines indicate the lattice locations for the ground state (x0j ,
grey lines) and the perturbed state (xj , light red lines). (b) The deformation field u(x) associated with the perturbed state from (a). Small
circle markers indicate the deformation of the density maxima of ϱ(x). Inset: close up of a density peak showing the identification of u(x).
(c) A comparison of the total, lattice deformation, and defect density fluctuations projected to the first Brillouin zone (see Appendix B). (d)
Excitations frequencies for the lowest two bands comparing BdG (solid) and hydrodynamic [i.e. ωνq = cν q] (dashed) results. Marker identifies
the excitation used in subplots (a)-(c). (e)-(f) The density fluctuations comparing the BdG (solid) and hydrodynamic (dashed) results.

(i.e. maxima of ϱ). The displacement field can be expressed on a supersolid state with equilibrium density profile ϱ0 (x).
as Assuming that the atom number in each cell remains constant
then the purely-deformed density profile is given by
u(x) = 2Re{cνq uνq eiqx }. (42)
where the amplitude uνq is a measure of how the quasiparticle ϱu (x) = ϱ0 (x − u(x))(1 − ∂x u(x)), (43)
displaces the site at the origin3 . In Fig. 6(b) we show (42) for
where the first factor describes the displaced density profile
the perturbed density of Fig. 6(a), and give an example of its
and second factor adjusts the envelope of density profile to
relationship to the site displacements [inset to Fig. 6(b)].
keep the atom number per cell constant. The change in cell
To understand the effect of lattice deformation on the den-
width, i.e. strain, leads to a change in density across the sys-
sity we consider a weak displacement field of the form (42)
tem, giving rise to the lattice strain density fluctuation. Fourier
transforming δϱu (x) ≡ ϱu (x) − ϱ0 (x) for a single quasipar-
ticle yields [cf. Eq. (41)]
3 In calculations we evaluate uνq = −δρ′νq (0)/ϱ′′ (0), where the prime is
R 0
derivative with respect to x and δρνq (x) = dy dz ψ0 [uνq − vνq ] δ ρ̃u,νq = −iqρuνq L, (44)
9

and δ ρ̃u,ν−q = δ ρ̃∗u,νq as the nonzero contributions in the first band (ν = +, red) has weak defect density fluctuations, and
Brillouin zone. In Eq. (44) L is the length of the x region the strain density fluctuation makes the dominant contribution
integrated over. to the total density fluctuation. Hence the second band can be
We illustrate the lattice strain density fluctuation construc- considered a crystal-like excitation, consistent with the con-
tion in Fig. 6. Using the identified deformation field u(x) clusions from the linear-chain model presented in Sec. IV D.
[Fig. 6(b)], in Fig. 6(a) we show the purely-deformed den-
sity profile ϱu (x), with associated density fluctuation shown
in Fig. 6(c). 10 1

C. Defect density fluctuations 10 0

In Fig. 6(a) we observe that while the maxima and minima


10 -1
of the density ϱ(x) (i.e. lattice distortion) coincide with those
of ϱu (x), these functions are not identical. The difference
arises from the tunnelling of particles between sites, leading 10 -2
to a change in the atom number per site. This is generally re-
ferred to as the defect density [14, 17, 18], and can be defined
10 -3
as ϱ∆ (x) = ϱ(x) − ϱu (x), or equivalently
10 1
δ ρ̃∆,νq = δ ϱ̃νq − δ ϱ̃u,νq . (45)

We illustrate the defect density fluctuation in Fig. 6(c). 10 0

D. Density fluctuations of lowest bands 10 -1

The decomposition of the total density fluctuation into 10 -2


strain and defect contributions is compared for an example
case in Fig. 6(c). Here we see that the strain and defect den-
sities have a relatively large amplitude and are out of phase, 10 -3
1.38 1.4 1.42 1.44 1.46 1.48 1.5
such that the total density fluctuation is small. This behav-
ior of the lattice straining in the opposite sense to the flow
of atoms in the lowest gapless band reveals a general prop-
erty of the lowest energy band we have used as an example FIG. 7. Strengths of the hydrodynamic regime density fluctuations
in Fig. 6(a)-(c). Similar behavior has been observed in exper- for a dipolar supersolid. (a) ξo+ results (solid lines) compared to
iments with dipolar supersolids as a low frequency Nambu- limq→0 2So,+ (q)/q (markers) for the various density fluctuations.
(b) ξo− results (solid lines) compared against limq→0 2So,− (q)/q
Goldstone excitation [5, 10, 11].
(markers). In both subplots we show ξκ for reference. In (b) the
It is useful to define the quantities results from (a) are repeated as light dash-dot lines. Other parame-
ters as in Fig. 2
|δ ρ̃νq |2
Sρ,ν (q) = , (46)
N
|δ ρ̃u,νq |2 The hydrodynamic description (5) also allows the calcula-
Su,ν (q) = , (47) tion of the density correlation functions (see [18]). In particu-
N
|δ ρ̃∆,νq |2 lar, our interest is in the imaginary part of the density response
S∆,ν (q) = , (48) function, which has the form
N
to characterise the strength of the total density, strain den- 1 X
sity and defect density fluctuations, arising from the excitation χ′′o (q, ω) = N qπ ξoν [δ(ω − cν q) − δ(ω + cν q)], (49)
2
P numbers ν and q, where N = Lρ. We note that
with quantum ν
Sρ (q) = ν Sρ,ν (q) is the usual static structure factor, and
that aspects of the static and dynamical structure factors for where the subscript o can take the values {ρ, u, ∆} to rep-
the dipolar supersolid have been discussed in Refs. [5, 12, 37]. resent the total density, strain density and defect density, re-
In Fig. 6(c) we show the dispersion relations, and in spectively. This result is limited to the lowest two bands
Fig. 6(e)-(g) the related fluctuation strengths introduced in (i.e. ν = {−, +}) in the modulated regime and to a single
Eq. (46)-(48), for the lowest two gapless bands of a dipolar band (i.e. ν = {+}) in the uniform regime. The hydrody-
supersolid. The lowest band (ν = −, blue) has strong strain namic results for the speeds of sound cν have already been
and defect density fluctuations as q → 0, but these are out-of- introduced, and ξoν is a correlations length that represents the
phase, leading to a small total density fluctuation. The second contribution of band ν to the response function. For the mod-
10

In Figs. 6(e)-(g) we compare the So,ν (q) functions obtained


10 0
from the BdG calculations Eq. (46)-(48) to the hydrodynamic
result for a dipolar supersolid. All the density fluctuations
10 -1 vanish proportional to q in the q → 0 limit, and the hydro-
dynamic result is seen to agree with the BdG results in this
10
-2 limit. A survey over a broader parameter regime is shown
in Figs. 7 and 8 for the dipolar and soft-core systems, respec-
tively. In these plots we only consider the hydrodynamic char-
10 -3
acter of the fluctuations and compare ξoν to the q → 0 limit of
2So,ν (q)/q calculated from the excitations4 .
10 -4 Our results show that strain and defect density fluctuations
are reasonably strong in the ν = − band of the dipolar super-
10 0
solid, but these contributions cancel out to suppress the total
density functions [Fig. 7(a)]. In contrast for the same band of
the soft-core supersolid the lattice strain density fluctuations
10 -1 are heavily suppressed [Fig. 8(a)]. As noted in Sec. IV, these
two systems differ in the relative size of their characteristic
10 -2 energies mc2κ and mc2u . For the dipolar system the bulk com-
pressibility dominates and it is favorable to reduce the total
10 -3
density fluctuations. In contrast for the soft-core system the
lattice compressibility dominates and it is favourable to re-
duce lattice motion. Interestingly ν = + band dominates the
10 -4 lower band for total density fluctuations in the dipolar super-
10 20 30 40 50 60 70 80
solid, whereas for the soft-core case the ν = − band domi-
nates close to the transition.
For the ν = + band the role of defect fluctuations in the
modulated state is more important for the soft-core system
FIG. 8. Strengths of the hydrodynamic regime density fluctua- close to the transition. However, for both models deep into the
tions for a soft-core supersolid. (a) ξo+ results (solid lines) com- crystalline regime, the defect contribution is negligible and we
pared against limq→0 2So,+ (q)/q (markers) for the various den- find ξu+ ≈ ξρ+ . This is consistent with this band becoming a
sity fluctuations. (b) ξo− results (solid lines) compared against purely crystal excitation.
limq→0 2So,− (q)/q (markers). In both subplots we show ξκ for ref- Finally, we note that the response functions relate to impor-
erence. In (b) the results from (a) are repeated as light dash-dot lines.
tant sum rules. Notably, the compressibility sum rule
Other parameters as in Fig. 3.
dω χ′′ρ (q, ω)
Z ∞
N αuu
= 2
= N ℏρκ, (54)
−∞ π ω ρ(α ρρ αuu − αρu )
ulated case these are given by
2 ρs [cf. Eq.(25)] and the f -sum rule
ℏ ρmc± − ρn αuu
ξρ± =
Z ∞
, (50) dω ′′ ℏq 2
mρ mc± (c2± − c2∓ ) ωχρ (q, ω) = N . (55)
−∞ π m
ℏρ mc2± − ρs αρρ
ξu± = , (51)
mρn mc± (c2± − c2∓ )
± ℏρs ρmc2± − (αuu − 2ραρu + ρ2 αρρ ) VI. PHASE FLUCTUATIONS
ξ∆ = . (52)
mρρn mc± (c2± − c2∓ )
We can also consider the effect of a quasiparticle excita-
In the uniform case the only non-zero quantity is ξρ+ → ξκ ,
tion on the phase fluctuations. The phase is undefined as the
with ξκ = ℏ/mcκ being the usual expression for the healing
density vanishes, so we avoid integrating over transverse coor-
length of a Bose-Einstein condensate.
dinates and only consider the phase on the x axis (i.e. hereon
At zero temperature and for ω > 0, the response func-
we take y = z = 0 in the dipolar model). For a quasipar-
tion relates to the dynamic structure factor as χ′′o (q, ω) =
ticle perturbation of the form (38) that phase fluctuation is
πSo (q, ω) (e.g. see [39]), and from R this we can obtain the ϕ(x) = arg {ψ(x)} (since ψ0 is taken to be real). To lead-
static structure factors N So (q) = dωSo (q, ω).PThe static
ing order in cνq this is
structure factors can be decomposed as So (q) = ν So,ν (q)
where So,ν (q) is the contribution from band ν, which were Im{cνq [uνq (x) + vνq (x)]}
ϕ(x) = (56)
defined in Eqs. (46)-(48). From hydrodynamic result (49) we ψ0 (x)
obtain
Z ∞
1 1
lim So,ν (q) = dω χ′′o (q, ω) = qξoν . (53)
q→0 πN 0 2 4 We use a BdG excitation with q ∼ 10−2 /a to extract the limiting behavior.
11

1 3 where we have introduced the phase correlation length


2.5
0.8
ℏ mc± (c2± − c2∓ )
0.6
2
ξϕ± = . (60)
1.5 ρ αρρ mc± − ρ1n (αρρ αuu − αρu
2 2 )
0.4
1

0.2 This shows that the phase response diverges as q → 0. From


0.5
this result we can derive the hydrodynamic result for the phase
0 0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5 fluctuations
10
2 0.4 1
lim Sϕ,ν (q) = . (61)
10 1
0.35
q→0 2ξϕ± q
0.3

0.25
10 0
0.2
In Figs. 9(a) and (b) we show examples of Sϕ,ν (q) com-
10 -1 0.15
puted from the BdG results for the dipolar and soft-core super-
0.1 solid states, respectively. We also confirm that the hydrody-
10 -2
0.05 namic results agree in the q → 0 limit. In Figs. 9(c)-(f) we ex-
10 -3 0 amine the hydrodynamic limiting behavior of the fluctuations
1.5
[i.e. comparing ξϕν to limq→0 q/2Sϕ,ν (q)] crossing the transi-
10 1
tion for both models. In the uniform superfluid state ξϕ+ → ξκ ,
1 being identical to the length scale for the total density fluctua-
10 0
tions in the same regime. In the dipolar supersolid the ν = −
band exhibits stronger phase fluctuations close to the transi-
0.5 10 -1
tion, but as ϵdd increases the phase fluctuations of the ν = +
band eventually dominates. For the soft-core supersolid the
10 -2
0
1.4 1.45 1.5 20 40 60 80
ν = + band dominates the phase fluctuations everywhere.
We note that the Fourier component of the superfluid ve-
locity field is revealed from the phase fluctuations as vνq =
FIG. 9. Phase fluctuations. (a), (b) Comparison of the phase fluctu- iℏqϕq /m, and thus our results are trivially extendible to de-
ations as a function of q obtained from the BdG calculations (solid) scribe the superfluid velocity fluctuations (also see [39]). The
to the hydrodynamic limit ξϕ± (dashed) for the lowest two bands (a) character of quasiparticle effect on the phase or velocity has
dipolar and (b) soft-core supersolids. (c)-(f) Strengths of the hydro- been employed to characterise the bands and excitations of
dynamic regime total density ξρ± (black line) and phase ξϕ± (green dipolar supersolids (e.g. see [5, 42]). However, these char-
line) correlation lengths for a dipolar supersolid, compared to the acterizations involve quantifying the phase variations distin-
relevant limits limq→0 1/[2qSϕ,ν (q)] (markers). Results for the (c) guishing between the high and low density regions of the su-
ν = −1 and (e) ν = +1 bands of a dipolar supersolid, with other persolid within the unit cell, and are thus beyond the hydro-
parameters as in Fig. 2. In both subplots we show ξκ for reference. dynamic description.
Results for (d) ν = −1 and (f) ν = +1 bands of a soft-core super-
solid, with other parameters as in Fig. 3. In (e) and (f) the ξϕ+ and
ξρ+ are identical for the uniform phase, and are indicated by dashed
VII. OUTLOOK AND CONCLUSIONS
lines. Also, the modulated results from (c) and (d) are repeated as
light dash-dot lines.
In this study, we have demonstrated the effectiveness of the
supersolid Lagrangian formalism in providing a quantitative
For a single quasiparticle, the only non-zero fluctuations are description of the hydrodynamic characteristics of dipolar and
at wavevectors q and −q, with [cf. Eq. (41)] soft-core supersolids. The Lagrangian description depends on
Z a number of elastic parameters and we have discussed how
uνq (x) + vνq (x) these can be extracted from ground state calculations. Our
ϕ̃νq = dx e−iqx . (57)
2iψ0 (x) findings include a comparative analysis of the Lagrangian the-
ory to the BdG excitations for the speeds of sound, density and
The strength of the phase fluctuations are given by the dimen- phase fluctuations. The excellent agreement shows that the
sionless quantity hydrodynamic behaviour is well-described by the Lagrangian.
ρ Also, these results underscore the distinct behaviors exhibited
Sϕ,ν (q) = |ϕ̃νq |2 . (58) by dipolar and soft-core models.
L
To unravel these distinctions, we have considered various
Similar to the treatment of density functions, from the limiting properties of the Lagrangian theory and how these
quadratic Lagrangian we obtain the phase response function apply to the two models. This has brought to light the signif-
1 X 1 icance of two energy scales, mc2κ and mc2u , which delineate
χ′′ϕ (q, ω) = Nπ [δ(ω − cν q) − δ(ω + cν q)], (59) the relative impacts of bulk and lattice compressibility, respec-
2 ξϕν q
ν tively. The incompressible limit (mc2κ ≫ mc2u ) applies to the
12

dipolar supersolid, while the rigid lattice limit (mc2κ ≪ mc2u ) where
applies to the soft-core supersolid. Our results illuminate how Z
these distinct limits manifest in the excitation spectrum, den- Φ(x) = dx′ U (x − x′ )|ψ(x′ )|2 , (A2)
sity, and phase fluctuations.
It is interesting to consider how our predictions can be ver- and the effects of quantum fluctuations p are described by
ified in experiments. Recently, the work of Šindik et al. [25] the term with coefficient γQF = 32 3 gs a3s /πQ5 (ϵdd ), with
proposed a scheme for measuring the excitations and density R1
Q5 (x) = ℜ{ 0 du[1+x(3u2 −1)]5/2 } [49]. The energy func-
response of a dipolar supersolid with minimal finite size ef- tional is evaluated in a single unit cell of length a along x (and
fects. Their scheme involves the use of a toroidal potential integrated over all space in the transverse directions). The ef-
for confinement with a well-defined mean density, and an az- fective potential arising from interactions, Φ(x), is evaluated
imuthally varying perturbing potential to excite q → 0 excita- using a truncated interaction kernel (see Ref. [34] for details).
tions. Observing the density oscillations following the sudden The energy minimising wavefunction for (A1) is determined
removal of the potential can be used to determine the speeds using the gradient flow (or imaginary time) evolution [50, 51],
of sound and the amplitude of the density response. Deter- noting that the average density condition enforces the follow-
mining the lattice strain and defect density fluctuations may ing normalization constrain on the wavefunction
be possible with high-resolution in situ imaging which have Z
already been used in experiments to characterize excitations
[11] (also see [43, 44]). dx |ψ(x)|2 = ρa. (A3)
uc
Our results describe the zero temperature fluctuations and
an important extension of this work is to include finite tem- As a result we obtain E(ρ, a), i.e. the minimising energy as
perature effects (see Refs. [18, 19]). Another area of inter- a function of density and cell size. We discuss the energy
est is higher dimensional supersolids where an elastic tensor density function and identifying the ground state in Sec. A 3.
describes the crystal, and the shear modulus will become rele-
vant. Of interest is the two-dimensional supersolid that has re-
cently been produced in experiments with a dipolar BEC [45]. 2. 1D soft-core formalism
In the thermodynamic limit the ground state phase diagram
for this case has been determined [46, 47], yet the excitations The stationary states are determining by minimising the
remain largely unexplored. This system is expected to have meanfield functional of the state
three gapless branches [6] (cf. [7]) with the emergence of a Z 
ℏ2 d2 1

transverse crystal mode. E[ψ, a] = dx ψ ∗ − + Φ sc (x) ψ, (A4)
uc 2m dx2 2
Note added. As we were about to submit, the preprint
Ref. [48] appeared, which also considers the elastic proper- where the integration is over a unit cell along x of length a,
ties of softcore supersolids. and
Z
Φsc (x) = dx′ Usc (x − x′ )|ψ(x′ )|2 . (A5)
ACKNOWLEDGMENTS
To implement the average density constraint on the stationary
PBB acknowledges useful discussions with J. Hofmann and states ψ has the normalization condition
use of New Zealand eScience Infrastructure (NeSI) high per- Z
formance computing facilities. LMP, DB and PBB, acknowl- dx |ψ(x)|2 = ρa. (A6)
edge support from the Marsden Fund of the Royal Society of uc

New Zealand. Identical to the dipolar case, we obtain the minimum energy
per unit cell E(ρ, a), by applying gradient flow algorithm op-
timise the energy functional Eq. (A4) against the wavefunc-
Appendix A: Additional details on supersolid models tion.

Here we outline some additional details about the two mod-


3. Energy density and ground state
els we consider and how we solve for the energy minimising
states.
For both models are hence able to solve for energy minimis-
ing states with specified average linear density ρ and lattice
1. Dipolar EGPE theory constant a. The ground state ψ0 for density ρ is then deter-
mined by minimising this against a [i.e. Eq. (4)]. For deter-
mining the generalized elastic parameters we used the energy
The eGPE energy functional for this system is density (per unit length) we is given by
Z
E[ψ, a] = dx ψ ∗ Hsp + 21 Φ(x) + 25 γQF |ψ|3 ψ, (A1)
  E(ρ, a)
E(ρ, a) = . (A7)
uc a
13

We only need to obtain E(ρ, a) for values of a close to the is the (constant) mean linear density. We also define the pro-
ground state value at each density, since the elastic param- jected total density fluctuation [shown in Fig. 6(c)]
eters involve derivatives evaluated at the ground state value.
We also note that in situations where the stationary state is δ ϱ̄(x) = PBZ δϱ(x), (B2)
uniform E independent of a.
where for an excitation of quasimomentum q only the ±q
wavevectors contribute following projection, i.e.
Appendix B: Projected density fluctuations
δ ϱ̄(x) = 2Re{cνq δ ρ̃νq eiqx /L}, (B3)
We reveal the long-wavelength character of the density fluc- with δ ρ̃νq as defined in Eq. (41). We can similarly project the
tuations by projecting them to the first Brillouin zone with the density fluctuations δϱu (x) and δϱ∆ (x) to obtain δ ϱ̄u (x) and
operation δ ϱ̄∆ (x), respectively [also shown in Fig. 6(c)]. These satisfy
X 1Z ′
similar relations to (B3) but in terms of δ ρ̃u,νq and δ ρ̃∆,νq ,
PBZ f (x) = dx′ eik(x−x ) f (x′ ), (B1) respectively. For lattice strain density fluctuation this is
L
k∈BZ
δ ϱ̄u (x) = 2Re{−iqρcνq uνq eiqx }, (B4)
where BZ denotes wavevectors in the range [−π/a, π/a]. For
the unperturbed linear density only the constant (k = 0) term where we have used (44). From this we see that δ ϱ̄u (x) =
survives projection ρ = PBZ ϱ0 (x), i.e., the projected density −ρ ∂x u with u given by Eq. (42) [52].

[1] J. Léonard, A. Morales, P. Zupancic, T. Esslinger, and T. Don- [12] D. Petter, A. Patscheider, G. Natale, M. J. Mark, M. A. Baranov,
ner, Supersolid formation in a quantum gas breaking a continu- R. van Bijnen, S. M. Roccuzzo, A. Recati, B. Blakie, D. Bail-
ous translational symmetry, Nature 543, 87 (2017). lie, L. Chomaz, and F. Ferlaino, Bragg scattering of an ultracold
[2] J.-R. Li, J. Lee, W. Huang, S. Burchesky, B. Shteynas, F. Ç. dipolar gas across the phase transition from Bose-Einstein con-
Top, A. O. Jamison, and W. Ketterle, A stripe phase with super- densate to supersolid in the free-particle regime, Phys. Rev. A
solid properties in spin–orbit-coupled Bose–Einstein conden- 104, L011302 (2021).
sates, Nature 543, 91 (2017). [13] T. Ilg and H. P. Büchler, Ground-state stability and excitation
[3] L. Tanzi, E. Lucioni, F. Famà, J. Catani, A. Fioretti, C. Gab- spectrum of a one-dimensional dipolar supersolid, Phys. Rev.
banini, R. N. Bisset, L. Santos, and G. Modugno, Observation A 107, 013314 (2023).
of a dipolar quantum gas with metastable supersolid properties, [14] A. F. Andreev and I. M. Lifshitz, Quantum theory of defects in
Phys. Rev. Lett. 122, 130405 (2019). crystals, Sov. Phys. JETP 29, 1107 (1969).
[4] F. Böttcher, J.-N. Schmidt, M. Wenzel, J. Hertkorn, M. Guo, [15] W. M. Saslow, Microscopic and hydrodynamic theory of super-
T. Langen, and T. Pfau, Transient supersolid properties in an ar- fluidity in periodic solids, Phys. Rev. B 15, 173 (1977).
ray of dipolar quantum droplets, Phys. Rev. X 9, 011051 (2019). [16] M. Liu, Two possible types of superfluidity in crystals, Phys.
[5] G. Natale, R. M. W. van Bijnen, A. Patscheider, D. Petter, M. J. Rev. B 18, 1165 (1978).
Mark, L. Chomaz, and F. Ferlaino, Excitation spectrum of a [17] D. T. Son, Effective lagrangian and topological interactions in
trapped dipolar supersolid and its experimental evidence, Phys. supersolids, Phys. Rev. Lett. 94, 175301 (2005).
Rev. Lett. 123, 050402 (2019). [18] C.-D. Yoo and A. T. Dorsey, Hydrodynamic theory of super-
[6] H. Watanabe and T. Brauner, Spontaneous breaking of continu- solids: variational principle, effective lagrangian, and density-
ous translational invariance, Phys. Rev. D 85, 085010 (2012). density correlation function, Phys. Rev. B 81, 134518 (2010).
[7] T. Macrì, F. Maucher, F. Cinti, and T. Pohl, Elementary ex- [19] J. Hofmann and W. Zwerger, Hydrodynamics of a superfluid
citations of ultracold soft-core bosons across the superfluid- smectic, J. Stat. Mech.: Theory Exp. 2021 (3), 033104.
supersolid phase transition, Phys. Rev. A 87, 061602(R) (2013). [20] C. Bühler, T. Ilg, and H. P. Büchler, Quantum fluctuations in
[8] F. Ancilotto, M. Rossi, and F. Toigo, Supersolid structure and one-dimensional supersolids, Phys. Rev. Res. 5, 033092 (2023).
excitation spectrum of soft-core bosons in three dimensions, [21] C. Josserand, Y. Pomeau, and S. Rica, Coexistence of ordinary
Phys. Rev. A 88, 033618 (2013). elasticity and superfluidity in a model of a defect-free super-
[9] S. M. Roccuzzo and F. Ancilotto, Supersolid behavior of a dipo- solid, Phys. Rev. Lett. 98, 195301 (2007).
lar Bose-Einstein condensate confined in a tube, Phys. Rev. A [22] C. Josserand, Y. Pomeau, and S. Rica, Patterns and supersolids,
99, 041601(R) (2019). Eur. Phys. J.: Spec. 146, 47 (2007).
[10] L. Tanzi, S. M. Roccuzzo, E. Lucioni, F. Famà, A. Fioretti, [23] J. Ye, Elementary excitation in a supersolid, EPL 82, 16001
C. Gabbanini, G. Modugno, A. Recati, and S. Stringari, Su- (2008).
persolid symmetry breaking from compressional oscillations in [24] A. S. Peletminskii, Classical and relativistic dynamics of super-
a dipolar quantum gas, Nature 574, 382 (2019). solids: variational principle, J. Phys. A 42, 045501 (2008).
[11] M. Guo, F. Böttcher, J. Hertkorn, J.-N. Schmidt, M. Wenzel, [25] M. Šindik, T. Zawiślak, A. Recati, and S. Stringari, Sound, su-
H. P. Büchler, T. Langen, and T. Pfau, The low-energy Gold- perfluidity and layer compressibility in a ring dipolar super-
stone mode in a trapped dipolar supersolid, Nature 564, 386 solid, arXiv:2308.05981.
(2019). [26] G. Biagioni, N. Antolini, B. Donelli, L. Pezzè, A. Smerzi,
M. Fattori, A. Fioretti, C. Gabbanini, M. Inguscio, L. Tanzi,
14

and G. Modugno, Sub-unity superfluid fraction of a supersolid [41] K. Mukherjee and S. M. Reimann, Classical-linear-chain be-
from self-induced Josephson effect, arXiv:2311.04757. havior from dipolar droplets to supersolids, Phys. Rev. A 107,
[27] N. Sepúlveda, C. Josserand, and S. Rica, Nonclassical rota- 043319 (2023).
tional inertia fraction in a one-dimensional model of a super- [42] W. Kirkby, A.-C. Lee, D. Baillie, T. Bland, F. Ferlaino, P. B.
solid, Phys. Rev. B 77, 054513 (2008). Blakie, and R. N. Bisset, Excitations of a binary supersolid,
[28] M. Kunimi, Y. Nagai, and Y. Kato, Josephson effects in one- arXiv:2312.03390.
dimensional supersolids, Phys. Rev. B 84, 094521 (2011). [43] J. Hertkorn, J.-N. Schmidt, F. Böttcher, M. Guo, M. Schmidt,
[29] S. Prestipino, A. Sergi, and E. Bruno, Clusterization of weakly- K. S. H. Ng, S. D. Graham, H. P. Büchler, T. Langen, M. Zwier-
interacting bosons in one dimension: an analytic study at zero lein, and T. Pfau, Density fluctuations across the superfluid-
temperature, J. Phys. A 52, 015002 (2018). supersolid phase transition in a dipolar quantum gas, Phys. Rev.
[30] N. Henkel, R. Nath, and T. Pohl, Three-dimensional roton ex- X 11, 011037 (2021).
citations and supersolid formation in Rydberg-excited Bose- [44] J.-N. Schmidt, J. Hertkorn, M. Guo, F. Böttcher, M. Schmidt,
Einstein condensates, Phys. Rev. Lett. 104, 195302 (2010). K. S. H. Ng, S. D. Graham, T. Langen, M. Zwierlein, and
[31] M. E. Fisher, M. N. Barber, and D. Jasnow, Helicity modulus, T. Pfau, Roton excitations in an oblate dipolar quantum gas,
superfluidity, and scaling in isotropic systems, Phys. Rev. A 8, Phys. Rev. Lett. 126, 193002 (2021).
1111 (1973). [45] M. A. Norcia, C. Politi, L. Klaus, E. Poli, M. Sohmen,
[32] A. J. Leggett, Can a solid be "superfluid"?, Phys. Rev. Lett. 25, M. J. Mark, R. N. Bisset, L. Santos, and F. Ferlaino, Two-
1543 (1970). dimensional supersolidity in a dipolar quantum gas, Nature 596,
[33] A. Leggett, On the superfluid fraction of an arbitrary many- 357 (2021).
body system at T = 0, J. Stat. Phys. 93, 927 (1998). [46] Y.-C. Zhang, F. Maucher, and T. Pohl, Supersolidity around a
[34] J. C. Smith, D. Baillie, and P. B. Blakie, Supersolidity and crys- critical point in dipolar Bose-Einstein condensates, Phys. Rev.
tallization of a dipolar Bose gas in an infinite tube, Phys. Rev. Lett. 123, 015301 (2019).
A 107, 033301 (2023). [47] B. T. E. Ripley, D. Baillie, and P. B. Blakie, Two-dimensional
[35] P. M. Chaikin and T. C. Lubensky, Principles of Con- supersolidity in a planar dipolar Bose gas, Phys. Rev. A 108,
densed Matter Physics (Cambridge University Press, Cam- 053321 (2023).
bridge, 1995). [48] M. Rakic, A. F. Ho, and D. K. K. Lee, Elastic properties and
[36] M. Kunimi and Y. Kato, Mean-field and stability analyses of thermodynamic anomalies of supersolids, arXiv:2403.13727.
two-dimensional flowing soft-core bosons modeling a super- [49] A. R. P. Lima and A. Pelster, Quantum fluctuations in dipolar
solid, Phys. Rev. B 86, 060510(R) (2012). Bose gases, Phys. Rev. A 84, 041604(R) (2011).
[37] P. B. Blakie, L. Chomaz, D. Baillie, and F. Ferlaino, Compress- [50] W. Bao and Q. Du, Computing the ground state solution
ibility and speeds of sound across the superfluid-to-supersolid of Bose–Einstein condensates by a normalized gradient flow,
phase transition of an elongated dipolar gas, Phys. Rev. Res. 5, SIAM J. Sci. Comput. 25, 1674 (2004).
033161 (2023). [51] A.-C. Lee, D. Baillie, and P. B. Blakie, Numerical calculation
[38] P. B. Blakie, D. Baillie, L. Chomaz, and F. Ferlaino, Superso- of dipolar-quantum-droplet stationary states, Phys. Rev. Res. 3,
lidity in an elongated dipolar condensate, Phys. Rev. Res. 2, 013283 (2021).
043318 (2020). [52] A. Zippelius, B. I. Halperin, and D. R. Nelson, Dynamics of
[39] L. Pitaevskii and S. Stringari, Bose-Einstein Condensation and two-dimensional melting, Phys. Rev. B 22, 2514 (1980).
Superfluidity, Vol. 164 (Oxford University Press, 2016).
[40] C. Kittel, Introduction to Solid State Physics, 8th ed. (Wiley,
2004).

You might also like