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

Helical superfluid in a frustrated honeycomb Bose-Hubbard model

Tzu-Chi Hsieh Department of Physics and Center for Theory of Quantum Matter, University of Colorado, Boulder, Colorado 80309, USA    Han Ma Perimeter Institute for Theoretical Physics, Waterloo, Ontario N2L 2Y5, Canada Department of Physics and Center for Theory of Quantum Matter, University of Colorado, Boulder, Colorado 80309, USA    Leo Radzihovsky Department of Physics and Center for Theory of Quantum Matter, University of Colorado, Boulder, Colorado 80309, USA
Abstract

We study a “helical” superfluid, a nonzero-momentum condensate in a frustrated bosonic model. At mean-field Bogoliubov level, such a novel state exhibits “smectic” fluctuation that are qualitatively stronger than that of a conventional superfluid. We develop a phase diagram and compute a variety of its physical properties, including the spectrum, structure factor, condensate depletion, momentum distribution, all of which are qualitatively distinct from that of a conventional superfluid. Interplay of fluctuations, interaction and lattice effects gives rise to the phenomenon of order-by-disorder, leading to a crossover from the smectic superfluid regime to the anisotropic XY superfluid phase. We complement the microscopic lattice analysis with a field theoretic description for such a helical superfluid, which we derive from microscopics and justify on general symmetry grounds, reassuringly finding full consistency. Possible experimental realizations are discussed.

I Introduction

Frustrated systems exhibit novel emergent phenomena as they often host extensive degeneracy that is sensitive to even weak perturbations Moessner and Chalker (1998); Balents (2010). This manifests itself in rich low-temperature phenomenology, as in the frustrated magnetism, which commonly remains disordered down to temperatures low compared to the Curie-Weiss temperature set by the exchange interaction Ramirez et al. (1990); Gaulin et al. (1992); Ramirez (1994); Uemura et al. (1994); Dunsiger et al. (1996). At even lower temperature, generically order develops through the so-called “order-by-disorder” process Villain et al. (1980); Henley (1989), where fluctuations (e.g., entropically) select ordered states among many competing nearly free-energetically degenerate phases. This contrasts with an even more exotic possibility of a putative quantum liquid, that fails to order down to zero temperature, as believed to be exemplified by some 2d Kagome magnetic materials with a flat-band dispersion Sachdev (1992); Hastings (2000); Ran et al. (2007); Hermele et al. (2008); Evenbly and Vidal (2010); Yan et al. (2011); Depenbrock et al. (2012); Jiang et al. (2012); Iqbal et al. (2013, 2014); He et al. (2014); Gong et al. (2015); Norman (2016); Mei et al. (2017); He et al. (2017); Zhu et al. (2018).

Another rich class of “codimension-one” frustrated systems Bergman et al. (2007); Lee and Balents (2008); Mulder et al. (2010); Matsuda et al. (2010); Gao et al. (2017); Biswas and Damle (2018); Niggemann et al. (2019); Shimokawa and Kawamura (2019); Gao et al. (2020); Liu et al. (2020); Balla et al. (2019, 2020); Yao et al. (2021); Bordelon et al. (2021); Gao et al. (2022), characterized by a bare dispersion that is nearly degenerate along d1d-1 dimensions (dd the spatial dimension), develop even in the bipartite lattice materials, frustrated by competing interactions. Examples include spinel materials such as MnSc2S4\textrm{MnSc}_{2}\textrm{S}_{4} on 3d diamond lattice Bergman et al. (2007); Lee and Balents (2008); Gao et al. (2017); Bordelon et al. (2021) and FeCl3\textrm{FeCl}_{3} on layered honeycomb lattice Gao et al. (2022). Theoretically, they are well described by classical spin models with nearest-neighbor and next-nearest-neighbor exchange interactions. Within certain regime of frustration, the dispersion minima form a degenerate manifold with codimension-one in the reciprocal space. However, at nonzero temperature a set of wavevectors is selected entropically based on their lowest free energy.

The phenomenon of order-by-disorder, however, is elusive in magnetic systems, often obscured by further-neighbor interactions, which instead can lead to a low-temperature multi-step ordering sequence of phases Gao et al. (2017). Alternatively, the codimension-one bare dispersion and the associate phenomenology can be realized in bosonic systems with frustrated hoppings Varney et al. (2012); Sedrakyan et al. (2014) or Rashba spin-orbital coupling Sedrakyan et al. (2012), engineered in a controlled way in cold atom experiments Wintersperger et al. (2020); Zhai (2015). More exotic constructions involve ring exchange interactions that induce an emergent Bose metal Paramekanti et al. (2002); Tay and Motrunich (2011).

It is thus of interest to elucidate the nature of a state that emerges from a nonzero density of interacting frustrated bosons condensed at nonzero momentum on a dispersion minimum contour, and in particular to explore its stability to a putative quantum liquid state Sur and Yang (2019); Lake et al. (2021). To this end, we take two complementary approaches, a paradigmatic two-dimensional Bose-Hubbard model on honeycomb lattice and a continuum field theory that we derive from it. The model is frustrated by competing nearest-neighbor and next-nearest-neighbor hoppings, t1t_{1} and t2t_{2}, and thereby (for a broad range of parameters) exhibits a bare dispersion with a minimum on a closed contour at nonzero momentum k0k_{0} set by the hopping matrix elements. We thus explore in detail the rich phenomenology of the resulting helical superfluid state. In particular, as summarized in detail in Sec. II, we analyze the helical superfluid within the lattice Bogoliubov approximation, computing its dispersion, condensate depletion, momentum distribution, structure factor, and equation of state, all of which different qualitatively from that of a conventional superfluid. We complement this honeycomb lattice analysis by deriving from it (guided by generalized dipolar symmetry Gromov (2019); Radzihovsky (2020); Yuan et al. (2020); Gorantla et al. (2022); Lake et al. (2022)) a superfluid smectic field theory Radzihovsky (2011, 2020) to more generally explore the properties of the helical condensate. By going beyond the Bogoliubov approximation, we analyze the quantum and thermal order-by-disorder phenomenon that sets in at low energies and leads to a crossover to a more conventional but highly anisotropic XY superfluid.

The outline of this paper is as follows. In Sec II, we give a summary of our primary results. In Sec. III, we study a model of bosons hopping on a frustrated honeycomb lattice and its superfluidity within the Bogoliubov approximation. In Sec. IV, we construct a smectic field theory for the helical superfluid, and use it to study its stability to quantum and thermal fluctuations and its properties in the isotropic continuum, thereby complementing the microscopic lattice model analysis in Sec III. Sec. V is devoted to the quantum and thermal order-by-disorder phenomenon, that reduces the degeneracy of the dispersion minimum contour down to a discrete set of six minima, and thereby stabilizes the helical superfluid state. We conclude in Sec. VI with a discussion of results in the contexts of current experiments and remaining open questions.

II Summary of results

Refer to caption
Figure 1: (a) Schematic phase diagram of the frustrated honeycomb Bose-Hubbard model. With the inclusion of lattice order-by-disorder effects, the finite-temperature phase boundaries from a conventional (SFk=0\text{SF}_{k=0}) or a helical (SFHelical\text{SF}_{\text{Helical}}) superfluid to a normal state (N) are of Berezinskii-Kosterlitz-Thouless universality class. To leading order, the phase transition temperature is given by TKTn0BBT_{\text{KT}}\sim n_{0}\sqrt{BB_{\perp}}, where BB and BB_{\perp} are the superfluid stiffnesses defined in the Lagrangian (6). For t2/t1<1/6t_{2}/t_{1}<1/6, BBt1B\sim B_{\perp}\sim t_{1}, giving TKTt1T_{\text{KT}}\propto t_{1}. In contrast, for t2/t1>1/6t_{2}/t_{1}>1/6, BB_{\perp} is induced perturbatively via interaction together with lattice effects, giving TKTU5/8t13/8T_{\text{KT}}\propto U^{5/8}t_{1}^{3/8} in the weakly-interacting limit. In the absence of lattice effects, SFHelical\text{SF}_{\text{Helical}} is always destabilized to the Normal state by divergent thermal fluctuations. (b) Phase diagram of the helical condensate in the regime 1/6<t2/t1<1/21/6<t_{2}/t_{1}<1/2. Bosons can condense at the Vertex (BECV\text{BEC}_{\text{V}}) or the Edge (BECE\text{BEC}_{\text{E}}) of the dispersion minimum contour as a result of the order-by-disorder, obtained by numerically evaluating the one-loop correction to the thermodynamic potential (V.1). The black (red-dashed) curves are zero temperature (T=0.1UnT=0.1Un) phase boundaries, across which there is a first order transition.
Refer to caption
Figure 2: (a) Minima of the thermodynamic potential as a function of the condensate momentum 𝐤0{\bf k}_{0} for intermediate frustration 1/6<t2/t1<1/21/6<t_{2}/t_{1}<1/2. The order-by-disorder effects, which manifest at long scales r>λr_{\perp}>\lambda_{\perp}, break the degenerate ground state manifold from a (schematic) hexagon down to six points. The helical condensate (red dot) spontaneously breaks the U(1) (C6\text{C}_{\text{6}}) symmetry in the left (right) panel, resulting in a low-energy smectic-like (conventional, XY) Goldstone mode (GM). (b) Two-point correlation function C(𝐫)=[ϕ(𝐫)ϕ(0)]2C({\bf r})=\langle[\phi({\bf r})-\phi(0)]^{2}\rangle of (6) along rr_{\perp} with r=0r_{\parallel}=0. For general cases (black curve), the helical superfluid exhibits quantum (r<rTr_{\perp}<r_{\perp T}), thermal (r>rTr_{\perp}>r_{\perp T}), smectic (SFSm\text{SF}_{\text{Sm}}, r<λr_{\perp}<\lambda_{\perp}), and XY (SFXY\text{SF}_{\text{XY}}, r>λr_{\perp}>\lambda_{\perp}) fluctuations, separated by two crossover scales, rTr_{\perp T} and λ\lambda_{\perp}. In the absence of the order-by-disorder (λ\lambda_{\perp}\to\infty, red-dashed curve), the helical superfluid is characterized by linear in rr_{\perp} smectic fluctuations for r>rTr_{\perp}>r_{\perp T}. In the zero temperature limit (rTr_{\perp T}\to\infty, blue-dotted curve), the superfluid exhibits a long-range order with C(0,r)constC(0,r_{\perp}\to\infty)\to\text{const}. Inset: same plot with logarithmic scale in the rr_{\perp} axis, where rr_{\perp} is plotted to longer scale for clarity of the crossover.

Before turning to a detailed analysis, we summarize the key results of this work. We study a Bose-Hubbard model with frustrated nearest-neighbor (NN) and next-nearest-neighbor (NNN) hoppings, t1t_{1} and t2t_{2}. For intermediate frustration 1/6<ρt2/t1<1/21/6<\rho\equiv t_{2}/t_{1}<1/2, its band structure features a closed contour minimum centered around the Γ\Gamma point (𝐤=0{\bf k}=0). We then consider a bosonic condensate with a macroscopic wavefunction Φ=n0+πei𝐤0𝐫+iϕ\Phi=\sqrt{n_{0}+\pi}e^{i{\bf k}_{0}\cdot{\bf r}+i\phi} at a single momentum 𝐤0{\bf k}_{0} on the contour minimum – a “helical condensate”, where n0n_{0} is the condensate density, ϕ\phi and π\pi are the phase and density fluctuations, respectively. As illustrated in Fig. 1, the helical superfluid is a stable phase on the high symmetry points of the contour with its transition temperature to the normal state (below that of a conventional superfluid in the unfrustrated regime ρ<1/6\rho<1/6) strongly suppressed by quantum and thermal fluctuations. Quantum Lifshitz transitions at ρ=1/6,1/2\rho=1/6,1/2111The quantum Lifshitz transition at ρ=1/2\rho=1/2 only happens at the vertices of the hexagonal contour (see Fig. 5), which is energetically stable at one-loop order as shown in Fig. 1(b). Higher-order fluctuations may shift the positions of the Lifshitz transitions, or make the transition at ρ=1/2\rho=1/2 split into two Lifshitz points characterized by smectic-like Goldstone modes with Lagrangian (44b), as allowed by the C6\text{C}_{\text{6}} lattice symmetry. are characterized by “soft” Goldstone modes that exhibit an anisotropic quartic dispersion, which suppresses the transition temperature to zero.

As summarized in Fig. 2, at zero temperature (T=0T=0) and on scale shorter than quantum order-by-disorder length λQ\lambda_{\perp Q}, the highly anisotropic superfluid exhibits an unconventional smectic-like Goldstone mode (GM) with a low-energy dispersion [for full dispersion, see Eq. (31)]

E𝐪Bq2+Kq4,\displaystyle E_{\bf q}\sim\sqrt{Bq_{\parallel}^{2}+Kq_{\perp}^{4}}, (1)

where the subscript \parallel (\perp) denotes the direction that is parallel (perpendicular) to condensate momentum 𝐤0{\bf k}_{0}, located at the high symmetry points of the dispersion contour. At temperatures higher than the interaction UU, the crossover scale is modified to λT\lambda_{\perp T}. On longer scales the so-called order-by-disorder sets in, driven by quantum, thermal and lattice effects, and the system crossovers to a more conventional superfluid with highly anisotropic linear dispersion. In the weakly-interacting limit, the order-by-disorder length is long, given by

λ={λQU5/8,TUn0.λTU1/8T1/2,TUn0.\displaystyle\lambda_{\perp}=\left\{\begin{array}[]{cc}\lambda_{\perp Q}\sim U^{-5/8},\quad T\ll Un_{0}.\\ \lambda_{\perp T}\sim U^{-1/8}T^{-1/2},\quad T\gg Un_{0}.\end{array}\right. (4)

To further explain the order-by-disorder phenomenology (discussed in detail in Sec. V), we first note that the thermodynamic potential of the helical state near the condensate momentum 𝐤0{\bf k}_{0} [see the right panel of Fig. 2(a)] exhibits a property

Ω(𝐤0+𝐪)Ω(𝐤0)N0[bq2+bq2+b4q4],\displaystyle\Omega({\bf k}_{0}+{\bf q})-\Omega({\bf k}_{0})\approx N_{0}\left[bq_{\parallel}^{2}+b_{\perp}q_{\perp}^{2}+b_{4}q_{\perp}^{4}\right], (5)

where the coefficients b(𝐤0)b({\bf k}_{0}) and b4(𝐤0)b_{4}({\bf k}_{0}) are dominated by the bare dispersion (see Appendix B), while b(𝐤0)b_{\perp}({\bf k}_{0}) is generated perturbatively by the lattice order-by-disorder via interaction UU. In the above, N0N_{0} is the condensate number. The corresponding low-energy Goldstone mode Lagrangian density is given by

=n0[Bτ(τϕ)2+B(ϕ)2+B(ϕ)2+K(2ϕ)2],\displaystyle\mathcal{L}=n_{0}\left[B_{\tau}(\partial_{\tau}\phi)^{2}+B(\partial_{\parallel}\phi)^{2}+B_{\perp}(\partial_{\perp}\phi)^{2}+K(\partial_{\perp}^{2}\phi)^{2}\right], (6)

which when treated at a quantum level leads to the dispersion (1) for aq1λ=K/Ba\ll q_{\perp}^{-1}\ll\lambda_{\perp}=\sqrt{K/B_{\perp}} (aa the lattice constant). We observe that a difference choice of the helical condensation, 𝐤0𝐤0+𝐪{\bf k}_{0}\to{\bf k}_{0}+{\bf q}, equivalently corresponds to a linear in 𝐫{\bf r} phase rotation, ϕϕ+𝐪𝐫\phi\to\phi+{\bf q}\cdot{\bf r}. By applying the two transformations to the thermodynamic potential and the Goldstone mode Hamiltonian respectively, and requiring the shifted energies to be identical, we obtain the following Ward identities,

b=B,b=B.\displaystyle b=B,\quad b_{\perp}=B_{\perp}. (7)

We expect the relations (7) to hold at all order of a perturbation theory (see Sec. V). In particular, at the zeroth order, the thermodynamic potential [see the left panel of Fig. 2(a)] is featured by a degenerate contour minimum, and thus b=B=0b_{\perp}=B_{\perp}=0. At one-loop (first) order, the correction to the thermodynamic potential is given by the zero-point energy of the Bogoliubov quasiparticles together with entropic contributions, which in turn gives a bU5/4b_{\perp}\propto U^{5/4} (bU1/4Tb_{\perp}\propto U^{1/4}T) for TUn0T\ll Un_{0} (TUn0T\gg Un_{0}) in the weakly-interacting limit. We also perform a complementary calculation of BB_{\perp} verifying (7). In contrast, while b4=Kb_{4}=K at zeroth-order, the above symmetry does not constrain it to hold generically.

The Goldstone mode theory (6) predicts the low-energy properties of the helical superfluid. In particular, the off-diagonal order of the superfluid Ψ(𝐫)Ψ(0)eC(𝐫)\langle\Psi^{\ast}({\bf r})\Psi(0)\rangle\propto e^{-C({\bf r})}, where C(𝐫)=[ϕ(𝐫)ϕ(0)]2C({\bf r})=\langle[\phi({\bf r})-\phi(0)]^{2}\rangle. As shown in Fig. 2(b), the two-point correlation function C(𝐫)C({\bf r}) exhibits several qualitatively distinct limits. Here we focus on the behavior along the \perp direction, leaving detailed discussions in Sec. IV.1. At zero temperature (see the blue-dotted curve), for either B=0B_{\perp}=0 or B0B_{\perp}\neq 0, C(r=0,r)constC(r_{\parallel}=0,r_{\perp}\to\infty)\to\text{const}, indicating a long-range order of the helical superfluid. At nonzero temperature, as illustrated by the black (red-dashed) curve, C(0,r)log(r)C(0,r_{\perp}\to\infty)\sim\log(r_{\perp}) (r\sim r_{\perp}) for B0B_{\perp}\neq 0 (B=0B_{\perp}=0), which signifies a quasi-long-range (short-range) order as a consequence of the XY (smectic) GM fluctuations. For a general case T0T\neq 0 and B0B_{\perp}\neq 0, there are two important crossover scales, rTr_{\perp T} and λ\lambda_{\perp}. The former separates the low-temperature quantum and higher-temperature classical regimes, while the latter, as discussed above, separates the short-distance smectic and long-distance XY regimes.

Observables 2d Helical SF 2d conventional SF
Static structure factor, SS n0ξ2q2+ξ2λ2q4\sim n_{0}\sqrt{\xi^{2}q_{\parallel}^{2}+\xi^{2}\lambda^{2}q_{\perp}^{4}} n0ξ~q\sim n_{0}\tilde{\xi}q
Condensate depletion, nd/nn_{d}/n U3/4n1/4\sim U^{3/4}n^{-1/4} U\sim U
Momentum distribution, n𝐪n_{\bf q} 1ξqfn(λq2q)\frac{1}{\xi q_{\parallel}}f_{n}\biggl{(}\lambda\frac{q_{\perp}^{2}}{q_{\parallel}}\biggr{)}, when qξ1q_{\parallel}\xi\ll 1 1(ξq)4fn(λq2q)\frac{1}{(\xi q_{\parallel})^{4}}f_{n}\biggl{(}\lambda\frac{q_{\perp}^{2}}{q_{\parallel}}\biggr{)}, when qξ1q_{\parallel}\xi\gg 1 1/q1/q, when qξ~1q\tilde{\xi}\ll 1 C/q4C/q^{4}, when qξ~1q\tilde{\xi}\gg 1
Superfluid stiffness, (ρs)ij(\rho_{s})_{ij} 8Jnk0ik0j8Jnk_{0i}k_{0j} δijn/m\delta_{ij}n/m
Chemical potential correction, μ(1)\mu^{(1)} Un𝒞(U3nB2K)1/4-Un\mathcal{C}\Big{(}\frac{U^{3}}{nB^{2}K}\Big{)}^{1/4} logarithmic correction Popov (1972); Salasnich and Toigo (2016)
Table 1: Comparison of the helical superfluid (within the order-by-disorder crossover scale) and a conventional superfluid at zero temperature in 2d. Note 𝐪=𝐤𝐤0{\bf q}={\bf k}-{\bf k}_{0} (𝐪=𝐤{\bf q}={\bf k}) for the helical (conventional) superfluid. For helical superfluid with low-energy Lagrangian (6), the coherence length ξ=B/2Un\xi=\sqrt{B/2Un}, the anisotropy length scale λ=K/B\lambda=\sqrt{K/B}, the condensate momentum 𝐤𝟎{\bf k_{0}}, JJ and 𝒞\mathcal{C} are constants, and fnf_{n} is a scaling function given in (IV.2.3). For conventional superfluid, CC is the Tan’s contact and ξ~=1/4mUn\tilde{\xi}=1/\sqrt{4mUn} is the coherence length of interacting Bose gas with mm, UU and nn being the mass, interaction strength and particle density respectively.

Consequently, the helical state is characterized by an anisotropic XY superfluidity at the longest scale. However, within a large (for small UU) order-by-disorder crossover scale, (r,r)=(λ2/ξ,λ)(r_{\parallel}^{*},r_{\perp}^{*})=(\lambda_{\perp}^{2}/\xi,\lambda_{\perp}) (ξ\xi the coherence length), the system exhibits soft (B0B_{\perp}\approx 0) quantum and thermal fluctuations, with its physical properties qualitatively distinct from that of a conventional superfluid. Abstracting from the Bose-Hubbard lattice model, we develop a field theory that captures the key qualitative features of the helical superfluid within this regime, and calculate a number of its physical observables, with the zero-temperature results summarized in Table 1. At nonzero temperature, thermal fluctuations set in at scales beyond (rT,rT)=(Un0Tξ,Un0Tλξ)(r_{\parallel T},r_{\perp T})=(\frac{Un_{0}}{T}\xi,\sqrt{\frac{Un_{0}}{T}}\sqrt{\lambda\xi}) (λ=K/B\lambda=\sqrt{K/B}), as illustrated in Fig. 2(b). Such quantum-to-classical crossover is observable in real space- or momentum-resolved quantities, as exemplified by the structure factor in Sec. IV.2.1. We now turn to detailed analysis that leads to the results enumerated above.

III Frustrated Bose-Hubbard model on honeycomb lattice

To explore the phenomenon of nonzero momentum helical superfluidity, we study a frustrated Bose-Hubbard model on honeycomb lattice with Hamiltonian H=H0+HintH=H_{0}+H_{\text{int}}, where

Hint=U2is=1,2ni,s(ni,s1)\displaystyle H_{\text{int}}=\frac{U}{2}\sum_{i}\sum_{s=1,2}n_{i,s}(n_{i,s}-1) (8)

is the usual on-site interactions with the subscripts ii and ss labeling the Bravais lattice sites and the two sublattices respectively. ni,s=ai,sai,sn_{i,s}=a^{\dagger}_{i,s}a_{i,s} is the boson number operator with ai,sa^{\dagger}_{i,s} and ai,sa_{i,s} the corresponding creation and annihilation operators. The kinetic part of the Hamiltonian is given by

H0=t1ijai,1aj,2+t2ij(ai,1aj,1+ai,2aj,2)+h.c.,H_{0}=-t_{1}\sum_{\langle ij\rangle}a^{\dagger}_{i,1}a_{j,2}+t_{2}\sum_{\langle\langle ij\rangle\rangle}(a^{\dagger}_{i,1}a_{j,1}+a^{\dagger}_{i,2}a_{j,2})+h.c., (9)

where t1t_{1} and t2t_{2} are the nearest-neighbor (NN) and the next-nearest-neighbor (NNN) hopping parameters, respectively. We focus on the case of t1>0t_{1}>0, taking the advantage of the symmetry of the model: t1t1t_{1}\rightarrow-t_{1}, a2a2a_{2}\rightarrow-a_{2}. In contrast to the NN hopping, we take the NNN hopping, t2-t_{2}, to be “frustrated antiferromagnetic”, t2<0-t_{2}<0. This has been engineered in atomic systems in an optical lattice through Floquet techniques Wintersperger et al. (2020). Throughout our analysis below, we use dimensionless parameter ρ=t2/t1>0\rho=t_{2}/t_{1}>0 to quantify the frustration in this model.

III.1 Band structure of non-interacting Hamiltonian

Below we first review the band structure of H0H_{0}, identifying the condition when a degenerate minimum along a closed contour develops, and then study a Bose condensate on the this dispersion minimum contour via Bogoliubov approximation. We start with specifying the NN and NNN vectors as [see Fig. 3(a)]

𝐞1\displaystyle{\bf e}_{1} =(0,1),𝐞2=(32,12),𝐞3=(32,12),\displaystyle=(0,1),\quad{\bf e}_{2}=(-\frac{\sqrt{3}}{2},-\frac{1}{2}),\quad{\bf e}_{3}=(\frac{\sqrt{3}}{2},-\frac{1}{2}),
𝐯1\displaystyle{\bf v}_{1} =(3,0),𝐯2=(32,32),𝐯3=(32,32)\displaystyle=(\sqrt{3},0),\quad{\bf v}_{2}=(-\frac{\sqrt{3}}{2},\frac{3}{2}),\quad{\bf v}_{3}=(-\frac{\sqrt{3}}{2},-\frac{3}{2}) (10)

with the lattice vectors 𝐯i{\bf v}_{i} spanning the Bravais triangular lattice and the unit vectors 𝐞i{\bf e}_{i} spanning the honeycomb lattice (the lattice constant a=1a=1). The corresponding reciprocal vectors are [see Fig. 3(b)]

𝐆1=(0,4π3),𝐆2=(2π3,2π3),𝐆3=(2π3,2π3).\displaystyle{\bf G}_{1}=(0,\frac{4\pi}{3}),\ \ {\bf G}_{2}=(-\frac{2\pi}{\sqrt{3}},-\frac{2\pi}{3}),\ \ {\bf G}_{3}=(\frac{2\pi}{\sqrt{3}},-\frac{2\pi}{3}). (11)
Refer to caption
Figure 3: (a) Honeycomb lattice with NN (black) and NNN (red) vectors. The two sublattices are connected by 𝜹=𝐞1{\bm{\delta}}={\bf e}_{1}. The lattice spacing is set to a=1{\rm a}=1 throughout this paper. (b) First Brillouin zone, high symmetry points 333The high symmetry points are given by Γ=(0,0)\Gamma=(0,0), K=(4π33,0)K=(\frac{4\pi}{3\sqrt{3}},0), K=(4π33,0)K^{\prime}=(-\frac{4\pi}{3\sqrt{3}},0), M=(π3,π3)M=(\frac{\pi}{\sqrt{3}},\frac{\pi}{3}), M=(0,2π3)M^{\prime}=(0,-\frac{2\pi}{3}), and M′′=(π3,π3)M^{\prime\prime}=(-\frac{\pi}{\sqrt{3}},\frac{\pi}{3})., and high symmetry (red-dashed) lines. Equivalent high symmetry points are connected by the reciprocal vectors (11).

The Hamiltonian (9) is straightforwardly diagonalized in momentum space (see Appendix A)

H0=\displaystyle H_{0}= t1𝐤(Γ𝐪a𝐤,1a𝐤,2+Γ𝐪a𝐤,2a𝐤,1)\displaystyle\ -t_{1}\sum_{\bf k}(\Gamma_{\bf q}a^{\dagger}_{{\bf k},1}a_{{\bf k},2}+\Gamma_{\bf q}^{*}a^{\dagger}_{{\bf k},2}a_{{\bf k},1})
+t2𝐤ϵ𝐤(a𝐤,1a𝐤,1+a𝐤,2a𝐤,2)\displaystyle\ +t_{2}\sum_{\bf k}\epsilon_{\bf k}(a^{\dagger}_{{\bf k},1}a_{{\bf k},1}+a^{\dagger}_{{\bf k},2}a_{{\bf k},2})
=\displaystyle= 𝐤ϵ𝐤d𝐤,d𝐤,+𝐤ϵ𝐤+d𝐤,+d𝐤,+,\displaystyle\ \sum_{\bf k}\epsilon^{-}_{\bf k}d_{{\bf k},-}^{\dagger}d_{{\bf k},-}+\sum_{\bf k}\epsilon^{+}_{\bf k}d_{{\bf k},+}^{\dagger}d_{{\bf k},+}, (12)

where

d𝐤,±=\displaystyle d_{{\bf k},\pm}= 12(eθ𝐤2a𝐤,1eθ𝐤2a𝐤,2)\displaystyle\ \frac{1}{\sqrt{2}}(e^{-\frac{\theta_{\bf k}}{2}}a_{{\bf k},1}\mp e^{\frac{\theta_{\bf k}}{2}}a_{{\bf k},2}) (13)

create bosonic excitations in the two bands,

ϵ𝐤=t2ϵ𝐤t1|Γ𝐤|,ϵ𝐤+=t2ϵ𝐤+t1|Γ𝐤|,\epsilon^{-}_{\bf k}=t_{2}\epsilon_{{\bf k}}-t_{1}|\Gamma_{{\bf k}}|,\quad\epsilon^{+}_{{\bf k}}=t_{2}\epsilon_{{\bf k}}+t_{1}|\Gamma_{{\bf k}}|, (14)

exhibiting Dirac nodes at the KK points (the corners of the first Brillouin zone), made famous in graphene but irrelevant here for bosonic system dominated by condensation at the bottom of the band. In the above,

ϵ𝐤=2icos𝐤𝐯i,Γ𝐤=iexp(i𝐤𝐞i)\epsilon_{{\bf k}}=2\sum_{i}\cos{\bf k}\cdot{\bf v}_{i},\quad\Gamma_{{\bf k}}=\sum_{i}\exp\left(-i{\bf k}\cdot{\bf e}_{i}\right) (15)

with |Γ𝐤|=3+ϵ𝐤|\Gamma_{{\bf k}}|=\sqrt{3+\epsilon_{{\bf k}}}. The dependence of the band structure on ρ=t2/t1\rho=t_{2}/t_{1} is depicted in Fig. 4.

Refer to caption
Figure 4: Band structure (14) of the frustrated bosonic tight-binding model at different ρ\rho. (a) ρ=0.1<1/6\rho=0.1<1/6. In this case, there is only one energy minimum at 𝐤=0{\bf k}=0 (Γ\Gamma point). Bosons form a condensate at zero momentum. (b) 1/6<ρ=0.3<1/21/6<\rho=0.3<1/2. Energy minimum shifts away from Γ\Gamma point to nonzero momentum 𝐤0{\bf k}_{0}, forming a closed contour. This contour is enlarged and deformed as ρ\rho increases from 1/61/6 to 1/21/2 (c) ρ=0.5\rho=0.5. Now the contour is a perfect hexagon with corner at the center of each edge of first Brillouin zone (MM point).

Momenta 𝐤¯0\bar{\bf k}_{0} is the dispersion minimum that satisfies

kϵ𝐤|𝐤=𝐤¯0=0ρ=12|Γ𝐤¯0|.\nabla_{k}\epsilon^{-}_{\bf k}|_{{\bf k}=\bar{\bf k}_{0}}=0\rightarrow\rho=\frac{1}{2|\Gamma_{\bar{\bf k}_{0}}|}. (16)
Refer to caption
Figure 5: Solution of energy minimum, plotted in unit of a1a^{-1}. Blue-dashed lines denote the first Brillouin zone boundaries. (a) This is the case shown in Fig. 4(a), where the energy minimum is a point. (b) This case corresponds to Fig. 4(b) with a dispersion minimum contour. (c) This is the case denoted in Fig. 4(c) with a perfect hexagon contour minimum. (d) When ρ>1/2\rho>1/2, the energy minima form a contour centered at every corner of the first Brillouin zone. We are not focus on this part in this paper.

As shown in Fig. 5, ρ\rho controls the form of the dispersion minimum in the reciprocal space. In detail, for ρ<1/6\rho<1/6, the minimum is at the Γ\Gamma point (𝐤¯0=(0,0)\bar{\bf k}_{0}=(0,0)), while for ρ\rho\rightarrow\infty, the minima are located at the KK points. For intermediate frustration 1/6<ρ<1/21/6<\rho<1/2 (1/2<ρ<1/2<\rho<\infty), the dispersion minima are extended to closed contour(s) centered around the Γ\Gamma (K) point(s). This macroscopic degeneracy makes the system sensitive to perturbations. Below we discuss these three cases respectively.

III.1.1 ρ<1/6\rho<1/6

The simplest is the weakly-frustrated case of ρ<1/6\rho<1/6, for which the bosons condense at the Γ\Gamma point [see Fig. 5(a)], exhibiting conventional zero momentum superfluidity. In the XY spin language, this corresponds to a ferromagnetic spin state.

III.1.2 Intermediate frustration, ρ>12\rho>\frac{1}{2}

In this case, the energy minima form closed contours encircling six corners of the first Brillouin zone, as shown in Fig. 5(d), with contours shrinking as ρ\rho increases to \infty.

As ρ\rho\rightarrow\infty, two sublattices decouple, with bosons on each triangular lattice condensing at a K point of the reciprocal lattice.444In XY spin language, this corresponds to the coplanar 120120^{\circ} state, the ground state of XXZ model on the triangular lattice. Generally, such superfluid is described by two order parameters as,i=csei𝐊s𝐫s,i\langle a_{s,i}\rangle=c_{s}e^{i{\bf K}_{s}\cdot{\bf r}_{s,i}} where cs=1,2c_{s=1,2} are two independent complex condensate amplitudes at 𝐊s{𝐊,𝐊}{\bf K}_{s}\in\{{\bf K},{\bf K^{\prime}}\}. Each of these site-dependent order parameters can be written as a two component XY vector S𝐫=(Rea𝐫,Ima𝐫)\vec{S}_{\bf r}=(\textrm{Re}\langle a_{{\bf r}}\rangle,\textrm{Im}\langle a_{{\bf r}}\rangle). For a particular choice of c1,2c_{1,2}, a configuration of S𝐫\vec{S}_{\bf r} is pictorially illustrated in Fig. 6.

Refer to caption
Figure 6: Spin representation of the bosonic condensates on the two sublattices (black and red), condensed at the first Brillouin zone corner KK or KK^{\prime}. These become independent in the ρ\rho\rightarrow\infty limit, coupled only by the boson interaction.

III.1.3 1/6<ρ<1/21/6<\rho<1/2

The intermediate-frustration regime – the main focus of this paper – exhibits a closed dispersion minimum centered around the Γ\Gamma point, as illustrated in Fig. 5(b). Generically, the contour exhibits a C6\text{C}_{\text{6}} lattice symmetry. The contour degenerates into a circle [hexagon] for ρ16+\rho\to\frac{1}{6}^{+}, see Fig. 5(b) [ρ12\rho\to\frac{1}{2}^{-}, see Fig. 5(c)].

III.2 Helical superfliud in Bogoliubov approximation

The interplay of interactions and the macroscopic degeneracy discussed above presents a rich and challenging problem. However, for weak interactions, a Bose condensate on a finite set of 𝐤0i{\bf k}_{0i} points555This is to be contrasted with an exotic Bose-Luttinger liquid state, where the boson amplitude is ordered with 𝐤0{\bf k}_{0} fluctuating across the entire contour, as recently studied by Sur and Yang Sur and Yang (2019) and Lake et al. Lake et al. (2021) on the minimum of the dispersion contour is at least a locally stable state of matter. For a set of non-collinear 𝐤0i{\bf k}_{0i} points, the ground state is a supersolid that generically breaking the underlying crystal lattice and the U(1) symmetries. A simpler collinear superfluid states are represented by a helical FF-like condensate at a 𝐤0{\bf k}_{0} and an LO-like condensate at {𝐤0,𝐤0}\{{\bf k}_{0},-{\bf k}_{0}\} Radzihovsky (2011); Choi and Radzihovsky (2011). For interacting bosons, finding a ground state even among these set of supersolids (characterized by a set of 𝐤0i{\bf k}_{0i}) is a challenging problem that we do not address here and is best studied numerically.

Here we focus on a helical superfluid corresponding to a condensation at a single 𝐤0{\bf k}_{0} that we find exhibits rich phenomenology that we explore in detail. Such a state can also be described by a helical spin density wave S𝐫=(Rea𝐫,Ima𝐫)=N0(cos𝐤𝟎𝐫,sin𝐤𝟎𝐫)\vec{S}_{\bf r}=(\textrm{Re}\langle a_{\bf r}\rangle,\textrm{Im}\langle a_{\bf r}\rangle)=\sqrt{N_{0}}(\cos{\bf k_{0}}\cdot{\bf r},\sin{\bf k_{0}}\cdot{\bf r}) that breaks the lattice rotational and time-reversal symmetries. However, in this FF helical state, the physical observables remain (lattice-) translationally invariant.

Thus, focusing on such a helical condensate at 𝐤0{\bf k}_{0}, we study interacting bosons on a frustrated honeycomb lattice (9) within the Bogoliubov approximation in a canonical ensemble. In momentum space, the interaction (8) is given by

Hint=\displaystyle H_{\text{int}}= U2Vs=1,2𝐤1,𝐤2,𝐩a𝐤1+𝐩2,sa𝐤2𝐩2,sa𝐤1𝐩2,sa𝐤2+𝐩2,s,\displaystyle\ \frac{U}{2V}\sum_{s=1,2}\sum_{{\bf k}_{1},{\bf k}_{2},{\bf p}}a^{\dagger}_{{\bf k}_{1}+\frac{{\bf p}}{2},s}a^{\dagger}_{{\bf k}_{2}-\frac{{\bf p}}{2},s}a_{{\bf k}_{1}-\frac{{\bf p}}{2},s}a_{{\bf k}_{2}+\frac{{\bf p}}{2},s}, (17)

where VV is the number of the unit cells, which is one-half of the number of the sites, Ns=2VN_{s}=2V. At low energies, neglecting inter-band excitations, we can focus on the lowest band, and express the site boson operators a𝐤,1a_{{\bf k},1} and a𝐤,2a_{{\bf k},2} in terms of the lower-band operator d𝐤,d_{{\bf k},-},

a𝐤,112eiθ𝐤2d𝐤,,a𝐤,212eiθ𝐤2d𝐤,,a_{{\bf k},1}\approx\frac{1}{\sqrt{2}}e^{i\frac{\theta_{{\bf k}}}{2}}d_{{\bf k},-},\quad a_{{\bf k},2}\approx\frac{1}{\sqrt{2}}e^{-i\frac{\theta_{{\bf k}}}{2}}d_{{\bf k},-}, (18)

where θ𝐤=Arg(Γ𝐤)\theta_{\bf k}=\textrm{Arg}(\Gamma_{\bf k}) defined in Eq. (15). Expressing the Hamiltonian in terms of excitations

d𝐤0+𝐪,N0δ𝐤,𝐤0+d𝐤,,d_{{\bf k}_{0}+{\bf q},-}\approx-\sqrt{N_{0}}\delta_{{\bf k},{\bf k}_{0}}+d_{{\bf k},-}, (19)

out of a helical condensate N0N_{0}, to quadratic order we have the Bogoliubov Hamiltonian,

H\displaystyle H\approx ϵ𝐤0N+nNU212𝐪ε𝐪\displaystyle\ \epsilon^{-}_{{\bf k}_{0}}N+\frac{nNU}{2}-\frac{1}{2}\sum_{{\bf q}}\varepsilon^{\prime}_{{\bf q}} (20)
+12𝐪(d𝐤0+𝐪,d𝐤0𝐪,)h0(d𝐤0+𝐪,d𝐤0𝐪,),\displaystyle+\frac{1}{2}\sum_{{\bf q}}\left(\begin{array}[]{cc}d^{\dagger}_{{\bf k}_{0}+{\bf q},-}&d_{{\bf k}_{0}-{\bf q},-}\end{array}\right)h_{0}^{-}\left(\begin{array}[]{c}d_{{\bf k}_{0}+{\bf q},-}\\ d^{\dagger}_{{\bf k}_{0}-{\bf q},-}\end{array}\right), (24)

where

h0=(ε𝐪Uncos(Θ𝐪)Uncos(Θ𝐪)ε𝐪)\displaystyle h_{0}^{-}=\left(\begin{array}[]{cc}\varepsilon^{\prime}_{\bf q}&Un\cos(\Theta_{\bf q})\\ Un\cos(\Theta_{\bf q})&\varepsilon^{\prime}_{\bf-q}\end{array}\right) (27)

and

ε𝐪=\displaystyle\varepsilon^{\prime}_{\bf q}= ϵ𝐤0+𝐪ϵ𝐤0+Un,\displaystyle\ \epsilon^{-}_{{\bf k}_{0}+{\bf q}}-\epsilon^{-}_{{\bf k}_{0}}+Un,
Θ𝐪=\displaystyle\Theta_{\bf q}= θ𝐤0+𝐪+θ𝐤0𝐪2θ𝐤0,\displaystyle\ \frac{\theta_{{\bf k}_{0}+{\bf q}}+\theta_{{\bf k}_{0}-{\bf q}}}{2}-\theta_{{\bf k}_{0}},
n=\displaystyle n= NNs=N2V,\displaystyle\ \frac{N}{N_{s}}=\frac{N}{2V},
N=\displaystyle N= N0+𝐪d𝐤0+𝐪,d𝐤0+𝐪,,\displaystyle\ N_{0}+\sum_{{\bf q}}d_{{\bf k}_{0}+{\bf q},-}^{\dagger}d_{{\bf k}_{0}+{\bf q},-}, (28)

deferring details to Appendix C. Bogoliubov-diagonalizing the Hamiltonian gives

H=\displaystyle H= Egs+𝐪E𝐪α𝐤0+𝐪α𝐤0+𝐪,\displaystyle\ E_{gs}+\sum_{{\bf q}}E_{\bf q}\alpha_{{\bf k}_{0}+{\bf q}}^{\dagger}\alpha_{{\bf k}_{0}+{\bf q}}, (29)

where the ground state energy

Egs\displaystyle E_{gs} =ϵ𝐤0N+nNU2+12𝐪(E𝐪ε𝐪)\displaystyle=\epsilon^{-}_{{\bf k}_{0}}N+\frac{nNU}{2}+\frac{1}{2}\sum_{{\bf q}}(E_{{\bf q}}-\varepsilon^{\prime}_{{\bf q}}) (30)

and the dispersion

E𝐪=\displaystyle E_{\bf q}= 2,𝐪+1,𝐪,\displaystyle\ \mathcal{E}_{2,{\bf q}}+\mathcal{E}_{1,{\bf q}}, (31)

with

1,𝐪=𝐪2+2Un𝐪+U2n2sin2(Θ𝐪)\displaystyle\mathcal{E}_{1,{\bf q}}=\sqrt{\mathcal{E}_{{\bf q}}^{2}+2Un\mathcal{E}_{{\bf q}}+U^{2}n^{2}\sin^{2}(\Theta_{\bf q})} (32)

and

𝐪=\displaystyle\mathcal{E}_{\bf q}= ϵ𝐤0+𝐪+ϵ𝐤0𝐪2ϵ𝐤0,2,𝐪=ϵ𝐤0+𝐪ϵ𝐤0𝐪2.\displaystyle\ \frac{\epsilon^{-}_{{\bf k}_{0}+{\bf q}}+\epsilon^{-}_{{\bf k}_{0}-{\bf q}}}{2}-\epsilon^{-}_{{\bf k}_{0}},\quad\mathcal{E}_{2,{\bf q}}=\frac{\epsilon^{-}_{{\bf k}_{0}+{\bf q}}-\epsilon^{-}_{{\bf k}_{0}-{\bf q}}}{2}. (33)

We postpone the discussion of EgsE_{gs} until Sec. V, where we show that fluctuations together with the lattice effects lead to order-by-disorder selection of the condensation at high symmetry points of the contour. In anticipation of this, here we note that there are two classes (and C6\text{C}_{\text{6}} equivalents) of high symmetry points represented by 𝐤0=(0,k0V){\bf k}_{0}=(0,k_{0V}) “Vertex” BECV\text{BEC}_{\text{V}} and (k0E,0)(k_{0E},0) “Edge” BECE\text{BEC}_{\text{E}}, with (16) giving,

k0V=\displaystyle k_{0V}= 23arccos[116ρ254],\displaystyle\ \frac{2}{3}\textrm{arccos}\left[\frac{1}{16\rho^{2}}-\frac{5}{4}\right],
k0E=\displaystyle k_{0E}= 23arccos[14ρ12].\displaystyle\ \frac{2}{\sqrt{3}}\textrm{arccos}\left[\frac{1}{4\rho}-\frac{1}{2}\right]. (34)

We next examine the details of the Bogoliubov dispersion E𝐪E_{\bf q} that is positive-semi-definite (gapless at 𝐪0{\bf q}\to 0) for all UU and 𝐪{\bf q}, which for U0U\to 0 reduces to ϵ𝐤0+𝐪ϵ𝐤0>0\epsilon^{-}_{{\bf k}_{0}+{\bf q}}-\epsilon^{-}_{{\bf k}_{0}}>0. We thus find that the helical condensate at the Bogoliubov level is locally stable.

Refer to caption
Figure 7: (a) The helical superfluid dispersion E𝐪E_{\bf q} from Eq. (31) with 𝐤0=(0,k0V){\bf k}_{0}=(0,k_{0V}), ρ=0.3\rho=0.3, Un/t1=0.2Un/t_{1}=0.2 and 𝐪{\bf q} in unit of a1a^{-1}. The black-dashed contour indicates the degenerate minimum of the noninteracting band ϵ𝐤\epsilon^{-}_{\bf k} in Eq. (14). In the presence of interaction, the minimum appears at the origin, 𝐪=0{\bf q}=0. The blue and green curves are dispersion cuts along and perpendicular to 𝐤0{\bf k}_{0}, indicating small qq linear and quadratic dispersions, respectively. The corresponding long wavelength description in (b) illustrates the symmetry-expected smectic form (III.2). The inset of (b) shows dispersions along qq_{\parallel} for a set of interaction strengths, from the lowest to the highest given by Un/t1=0.2,0.6,1,1.4Un/t_{1}=0.2,0.6,1,1.4.

In Fig. 7, we plot the quasiparticle dispersion E𝐪E_{\bf q} as a function of 𝐪=𝐪+𝐪{\bf q}={\bf q}_{\parallel}+{\bf q}_{\perp}, where 𝐪=(𝐪𝐤^0)𝐤^0{\bf q}_{\parallel}=({\bf q}\cdot\hat{{\bf k}}_{0})\hat{{\bf k}}_{0} and 𝐪=𝐤^0×𝐪×𝐤^0{\bf q}_{\perp}=\hat{{\bf k}}_{0}\times{\bf q}\times\hat{{\bf k}}_{0}. Without loss of generality, we plot the case 𝐤0=(0,k0V){\bf k}_{0}=(0,k_{0V}), which illustrates general properties of the dispersion (31). We note that it is asymmetric under qqq_{\parallel}\to-q_{\parallel} due to the 2,𝐪\mathcal{E}_{2,{\bf q}} contribution in (33). Physically, this broken inversion symmetry is a consequence of the spontaneously-chosen 𝐤0{\bf k}_{0}. At q=mod(2k0,4π/3)1.5q_{\parallel}=\text{mod}(-2k_{0},4\pi/3)\sim 1.5, the disperion also displays a metastable saddle point inherited from the contour minimum in the noninteracting dispersion ϵ𝐤\epsilon^{-}_{\bf k}, as shown by the blue curve in Fig. 7(b). This feature disappears gradually with increasing interaction [see inset of Fig. 7(b)]. In the small 𝐪{\bf q} limit, as expected on rotational symmetry grounds (here neglecting lattice order-by-disorder effects, but see Sec. V) of a helical superfluid (striped) state, the dispersion is well-described by an anisotropic smectic form,

E𝐪\displaystyle E_{\bf q}\approx 2Un(Bq2+Kq4+Kq4+Kq2q2)\displaystyle\ \sqrt{2Un(Bq_{\parallel}^{2}+K_{\perp}q_{\perp}^{4}+K_{\parallel}q_{\parallel}^{4}+K_{\parallel\perp}q_{\parallel}^{2}q_{\perp}^{2})}
\displaystyle\sim {q2,qλq2,|q|,qλq2,\displaystyle\ \left\{\begin{array}[]{cc}q_{\perp}^{2},\quad q_{\parallel}\ll\lambda q_{\perp}^{2},\\ |q_{\parallel}|,\quad q_{\parallel}\gg\lambda q_{\perp}^{2},\end{array}\right. (37)

with a “soft” \perp direction [see the green curve in Fig. 7(b)], where λ=K/B\lambda=\sqrt{K_{\perp}/B}. For qq2q_{\parallel}\sim q_{\perp}^{2}, the KK_{\parallel} and KK_{\parallel\perp} can be neglected at small 𝐪{\bf q}, while in the rotational invariant limit ρ1/6+\rho\to 1/6^{+}, KKK/2K_{\parallel}\approx K_{\perp}\approx K_{\parallel\perp}/2. The parameters in Eq. (III.2) can be obtained by expanding the free dispersion ϵ𝐤0+𝐪\epsilon^{-}_{{{\bf k}_{0}}+{\bf q}} in 𝐪{\bf q} with the results given in Appendix B. We note that at the Lifshitz transition, B(ρ1/6)0B(\rho\to 1/6)\to 0. Then, the (above neglected) higher-order contribution Kq4K_{\parallel}q_{\parallel}^{4} should be taken into account, with the Goldstone mode exhibiting an anisotropic quartic dispersion in all directions.

IV field theory of helical superfluid on a closed contour minimum

We now turn to a complementary continuum field-theoretic description of the helical superfuild state, characterized by Bose condensation on a closed contour minimum in momentum space. In the isotropic case, neglecting lattice effects, this physics can be encoded through a quartic dispersion

ε𝐤=J(𝐤2k¯02)2+ε0\varepsilon_{{\bf k}}=J({\bf k}^{2}-\bar{k}_{0}^{2})^{2}+\varepsilon_{0} (38)

with minima on a contour 𝐤2=k¯02{\bf k}^{2}=\bar{k}_{0}^{2}. As we demonstrated in Sec. III, such dispersion is indeed realized for bosons on a honeycomb lattice for ρ1/6+\rho\approx 1/6^{+}, with

J\displaystyle J =t1(364+27ρ32),\displaystyle=t_{1}\left(-\frac{3}{64}+\frac{27\rho}{32}\right),
k¯0\displaystyle\bar{k}_{0} =848ρ118ρ,\displaystyle=\sqrt{\frac{8-48\rho}{1-18\rho}},
ε0\displaystyle\varepsilon_{0} =t1[3(16ρ)2118ρ3+6ρ].\displaystyle=t_{1}\left[\frac{3(1-6\rho)^{2}}{1-18\rho}-3+6\rho\right]. (39)

We can encode this physics in a noninteracting Hamiltonian H0=𝐫Φ^𝐫ε^𝐫Φ^𝐫H_{0}=\int_{\bf r}\hat{\Phi}_{\bf r}^{\dagger}\hat{\varepsilon}_{{\bf r}}\hat{\Phi}_{\bf r} with

ε^𝐫=J(2k¯02)2+ε0.\hat{\varepsilon}_{{\bf r}}=J(-\nabla^{2}-\bar{k}_{0}^{2})^{2}+\varepsilon_{0}. (40)

We then extend this to an interacting field theory and study it in a grand canonical ensemble, with the corresponding imaginary-time action S=d2r𝑑τ(0+1)S=\int d^{2}rd\tau(\mathcal{L}_{0}+\mathcal{L}_{1}),

0\displaystyle\mathcal{L}_{0} =\displaystyle= ΦτΦ+J|2Φ|22Jk¯02|Φ|2+(ε~0μ)|Φ|2\displaystyle\Phi^{\ast}\partial_{\tau}\Phi+J|\nabla^{2}\Phi|^{2}-2J\bar{k}_{0}^{2}|\nabla\Phi|^{2}+(\tilde{\varepsilon}_{0}-\mu)|\Phi|^{2}
1\displaystyle\mathcal{L}_{1} =\displaystyle= U2|Φ|4,\displaystyle\frac{U}{2}|\Phi|^{4}, (41)

Φ(𝐫,τ)\Phi({\bf r},\tau) a complex bosonic field and μ\mu the chemical potential. We also defined an energy offset ε~0=Jk¯04+ε0\tilde{\varepsilon}_{0}=J\bar{k}_{0}^{4}+\varepsilon_{0} and interaction UU parameter inherited from the Hubbard model.

As with the lattice model in Sec. III, here we focus on the helical superfluid state – a condensate at a single point 𝐤0{\bf k}_{0} on the dispersion contour minimum – encoded in ε𝐤\varepsilon_{\bf k}, (38), and 0\mathcal{L}_{0} in (41). In the density-phase representation, the state is characterized by

Φ(𝐫)=nei𝐤0𝐫+iϕ=n0+πei𝐤0𝐫+iϕ\Phi({\bf r})=\sqrt{n}e^{i{\bf k}_{0}\cdot{\bf r}+i\phi}=\sqrt{n_{0}+\pi}e^{i{\bf k}_{0}\cdot{\bf r}+i\phi} (42)

with a nonzero condensate density (at mean-field level) n0=(με0)/Un_{0}=(\mu-\varepsilon_{0})/U and momentum k0=k¯0k_{0}=\bar{k}_{0}, for μ>ε0\mu>\varepsilon_{0} (vanishing otherwise), and density and phase fluctuations, π\pi and ϕ\phi, respectively. To quadratic order, the helical superfluid is then described by a Goldstone mode Lagrangian density 666 Each term in Eq. (41) can be rewritten in terms of π\pi and ϕ\phi as τΦ=\displaystyle\partial_{\tau}\Phi= ei𝐤0𝐫+iϕn(12τπ+inτϕ),\displaystyle\frac{e^{i{\bf k}_{0}\cdot{\bf r}+i\phi}}{\sqrt{n}}\biggl{(}\frac{1}{2}\partial_{\tau}\pi+in\partial_{\tau}\phi\biggr{)}, Φ=\displaystyle\nabla\Phi= ei𝐤0𝐫+iϕn[12π+in(𝐤0+ϕ)],\displaystyle\frac{e^{i{\bf k}_{0}\cdot{\bf r}+i\phi}}{\sqrt{n}}\biggl{[}\frac{1}{2}\nabla\pi+in({\bf k}_{0}+\nabla\phi)\biggr{]}, 2Φ=\displaystyle\nabla^{2}\Phi= ei𝐤0𝐫+iϕn3/2[12n2πn2(𝐤0+ϕ)214(π)2\displaystyle\frac{e^{i{\bf k}_{0}\cdot{\bf r}+i\phi}}{n^{3/2}}\biggl{[}\frac{1}{2}n\nabla^{2}\pi-n^{2}({\bf k}_{0}+\nabla\phi)^{2}-\frac{1}{4}(\nabla\pi)^{2} +in(π𝐤0+πϕ+n2ϕ)].\displaystyle+in(\nabla\pi\cdot{\bf k}_{0}+\nabla\pi\cdot\nabla\phi+n\nabla^{2}\phi)\biggr{]}.

0=\displaystyle\mathcal{L}^{\prime}_{0}= iπτϕ+J[k02n0(π)2+14n0(2π)2+4n0k02(ϕ)2\displaystyle i\pi\partial_{\tau}\phi+J\biggl{[}\frac{k_{0}^{2}}{n_{0}}(\partial_{\parallel}\pi)^{2}+\frac{1}{4n_{0}}(\nabla^{2}\pi)^{2}+4n_{0}k_{0}^{2}(\partial_{\parallel}\phi)^{2}
+n0(2ϕ)2+4k0(π)(2ϕ)]+U2π2,\displaystyle+n_{0}(\nabla^{2}\phi)^{2}+4k_{0}(\partial_{\parallel}\pi)(\nabla^{2}\phi)\biggr{]}+\frac{U}{2}\pi^{2}, (43)

where the subscripts \parallel and \perp denote axes parallel and perpendicular to 𝐤0{\bf k}_{0}, respectively. In Sec. V, we will consider higher-order fluctuations that renormalize the parameters n0n_{0}, 𝐤0{\bf k}_{0} and reduce the symmetry of the Lagrangian by C6C_{6} lattice effects. We note that the phase ϕ\phi is compact with 2π2\pi periodicity, and thereby allows vortex configurations – dislocation in the helical pattern of the state. However, because we are primarily interested in the ordered helical superfluid at zero temperature, where these defects are confined, we will neglect them in most of our analysis. They are however important for the finite temperature Kosterlitz-Thouless superfluid-to-normal transition in Fig. 1(a).

Integrating out π\pi field in Eq. (43), we obtain the effective theory of ϕ\phi described by a quantum smectic Lagrangian (consistent with the honeycomb lattice model analysis in Sec. III)

ϕ=\displaystyle\mathcal{L}_{\phi}= n0ϕ[(τ+^2)2^+Bτ1+^]ϕ\displaystyle~{}n_{0}\phi\left[\frac{-(\partial_{\tau}+\hat{\mathcal{E}}_{2})^{2}}{\hat{\mathcal{E}}+B_{\tau}^{-1}}+\hat{\mathcal{E}}\right]\phi (44a)
\displaystyle\approx n0[Bτ(τϕ)2+B(ϕ)2+K(2ϕ)2],\displaystyle~{}n_{0}\left[B_{\tau}(\partial_{\tau}\phi)^{2}+B(\partial_{\parallel}\phi)^{2}+K(\partial_{\perp}^{2}\phi)^{2}\right], (44b)

where Fourier transform of the operators ^\hat{\mathcal{E}} and ^2\hat{\mathcal{E}}_{2} are given by

𝐪=Bq2+Kq4,2,𝐪=2BKqq2,\displaystyle\mathcal{E}_{\bf q}=Bq_{\parallel}^{2}+Kq^{4},\quad\mathcal{E}_{2,{\bf q}}=2\sqrt{BK}q_{\parallel}q^{2}, (45)

with the couplings777At T=0T=0 and weak interaction, we neglect the difference between the boson and condensate densities, nn and n0n_{0}, with (nn0)/nU3/4(n-n_{0})/n\sim U^{3/4}, see Eq. (83).

Bτ=12Un0,B=4Jk02,K=J.B_{\tau}=\frac{1}{2Un_{0}},\quad B=4Jk_{0}^{2},\quad K=J. (46)

The final form in Eq. (44b) corresponds to the asymptotic long wavelength limit. The superfluid mode ϕ\phi exhibits an unusual dispersion given by

E𝐪=\displaystyle E_{\bf q}= 2,𝐪+𝐪2+Bτ1𝐪\displaystyle\ \mathcal{E}_{2,{\bf q}}+\sqrt{\mathcal{E}_{\bf q}^{2}+B_{\tau}^{-1}\mathcal{E}_{\bf q}}
\displaystyle\sim {q2,qλq2,qqc,|q|,qλq2,qqc,\displaystyle\ \left\{\begin{array}[]{cc}q_{\perp}^{2},\quad q_{\parallel}\ll\lambda q_{\perp}^{2},\quad q_{\perp}\ll q_{\perp c},\\ |q_{\parallel}|,\quad q_{\parallel}\gg\lambda q_{\perp}^{2},\quad q_{\parallel}\ll q_{\parallel c},\end{array}\right. (49)

consistent with our lattice analysis, Eq. (31) and the features illustrated in Fig. 7. Two length scales emerge from the zero-sound dispersion (49). The crossover scale ξ=BBτ\xi=\sqrt{BB_{\tau}} is the coherence length along 𝐤0{\bf k}_{0}, beyond which the dispersion is controlled by interaction. The length λ=K/B=1/(2k0)\lambda=\sqrt{K/B}=1/(2k_{0}) characterizes the anisotropy along (\parallel) and transverse (\perp) to 𝐤0{\bf k}_{0}. The corresponding momentum 𝐪c=(qc,qc)(ξ1,(λξ)1/2){\bf q}_{c}=(q_{\parallel c},q_{\perp c})\equiv(\xi^{-1},(\lambda\xi)^{-1/2}) separates the noninteracting regime, E𝐪(Bq+Kq2)2E_{\bf q}\approx\Big{(}\sqrt{B}q_{\parallel}+\sqrt{K}q^{2}\Big{)}^{2}, for 𝐪𝐪c{\bf q}\gg{\bf q}_{c} from the long-wavelength interacting form, E𝐪Bτ1/2Bq2+Kq4E_{\bf q}\approx B_{\tau}^{-1/2}\sqrt{Bq_{\parallel}^{2}+Kq^{4}}, for 𝐪𝐪c{\bf q}\ll{\bf q}_{c}. The latter agrees with the predictions of the lattice model, Eq. (III.2), if the lattice symmetry breaking effects are neglected.

Importantly, we note that, in the absent of lattice effects, (ϕ)2(\nabla_{\perp}\phi)^{2} is absent in the expansion of Eq. (44b) at all orders. This is enforced by the underlying rotational symmetry that is spontaneously broken by the helical superfluid state. More explicitly, an infinitesimal shift of the condensate momentum 𝐤0𝐤0+δ𝐤0{\bf k}_{0}\rightarrow{\bf k}_{0}+\delta{\bf k}_{0} along the dispersion minimum contour is a symmetry888This is closely related to the spontaneous breaking of dipole conservation (in addition to charge) of Bose condensation, which prohibits linear derivative terms, as recently studied in the dipolar Bose-Hubbard model by Lake et al. Lake et al. (2022) Here, the helical superfluid is characterized by an emerged dipole conservation perpendicular to 𝐤0{\bf k}_{0}, which however, only holds infinitesimally.. According to (42), this transforms ϕ𝐫ϕ𝐫+δ𝐤0𝐫\phi_{\bf r}\to\phi_{\bf r}+\delta{\bf k}_{0}\cdot{\bf r}, which then enforces the exact vanishing of the coefficient of (ϕ)2(\nabla_{\perp}\phi)^{2}.

IV.1 Stability

We now turn to the analysis of the stability of the helical condensate at 𝐤0{\bf k}_{0} to quantum and thermal fluctuations, characterized by Goldstone mode fluctuations, ϕ2\langle\phi^{2}\rangle. The divergence of this fluctuations in the thermodynamic limit is a signature of instability. We generalize our analysis to d+1d+1 space-time dimensions, with two “hard” axes, \parallel and τ\tau, and d1d-1 transverse “soft” axes. A complementary generalization to “columnar” Goldstone mode (with dd “hard” and 11 “soft” axes) is relegated to Appendix D. At d=2d=2, both generalizations reduce to the physical case of interest, but corresponding to distinct analytical continuation of the physical problem. Below, we employ the low-energy Goldstone mode theory, (44b), valid within a UV cutoff ΛU\Lambda_{U} defined by 𝐪<Un0\mathcal{E}_{\bf q}<Un_{0}.

IV.1.1 Local fluctuations ϕ2\langle\phi^{2}\rangle

At zero temperature, quantum fluctuations are quantified by

ϕ2Q\displaystyle\langle\phi^{2}\rangle_{Q} =12n0ΛUdωdqdd1q(2π)d+11Bτω2+Kq4+Bq2\displaystyle=\frac{1}{2n_{0}}\int^{\Lambda_{U}}\frac{d\omega dq_{\parallel}d^{d-1}q_{\perp}}{(2\pi)^{d+1}}\frac{1}{B_{\tau}\omega^{2}+Kq_{\perp}^{4}+Bq_{\parallel}^{2}}
=𝒪(1)×n01ξd+12λd12,d>1,\displaystyle=\mathcal{O}(1)\times n_{0}^{-1}\xi^{-\frac{d+1}{2}}\lambda^{-\frac{d-1}{2}},\quad d>1, (50)

and is finite (system size independent) in the thermodynamic limit. This thereby demonstrates the stability of the helical condensate to quantum fluctuations.

At nonzero temperature, thermal fluctuations are dominated by the zeroth Matsubara frequency ωn=0\omega_{n=0}, where ωn=2πn/β\omega_{n}=2\pi n/\beta, with β=1/T\beta=1/T. This gives

ϕ2T\displaystyle\langle\phi^{2}\rangle_{T}\approx T2n0L1ΛUdqdd1q(2π)d1Kq4+Bq2\displaystyle\ \frac{T}{2n_{0}}\int_{L^{-1}}^{\Lambda_{U}}\frac{dq_{\parallel}d^{d-1}q_{\perp}}{(2\pi)^{d}}\frac{1}{Kq_{\perp}^{4}+Bq_{\parallel}^{2}}
\displaystyle\approx Tϕ2QUn0×{min(L^1/2,L^)3d,d<3,ln[min(L^1/2,L^)],d=3,𝒪(1),d>3,\displaystyle\ \frac{T\langle\phi^{2}\rangle_{Q}}{Un_{0}}\times\left\{\begin{array}[]{ccc}\min\left(\hat{L}_{\parallel}^{1/2},\hat{L}_{\perp}\right)^{3-d},&\quad\ d<3,\\ \ln\left[\min\left(\hat{L}_{\parallel}^{1/2},\hat{L}_{\perp}\right)\right],&\quad\ d=3,\\ \mathcal{O}(1),&\quad\ d>3,\end{array}\right. (54)

where the dimensionless parameters L^=L/ξ\hat{L}_{\parallel}=L_{\parallel}/\xi and L^=L/ξλ\hat{L}_{\perp}=L_{\perp}/\sqrt{\xi\lambda}. LL_{\parallel} and LL_{\perp} are the system size along the \parallel and \perp directions, respectively. We thus observe that for d3d\leq 3, (in a striking contrast to a conventional superfluid) thermal fluctuations diverge with system size, and thereby destabilize the helical condensate at any nonzero temperature in the thermodynamic limit. For the physical case of d=2d=2, the length scales for this instability are set by ξT=ξTλUn0Tϕ2Qξλ\xi_{\perp}^{T}=\sqrt{\xi_{\parallel}^{T}\lambda}\sim\frac{Un_{0}}{T\langle\phi^{2}\rangle_{Q}}\sqrt{\xi\lambda}.

IV.1.2 Two-point correlation function of ϕ\phi

Next, we calculate the static two-point correlation function of ϕ(𝐫,τ)\phi({\bf r},\tau), in the physical case given by

C(𝐫)=\displaystyle C({\bf r})= [ϕ(𝐫,0)ϕ(0,0)]2\displaystyle\langle[\phi({\bf r},0)-\phi(0,0)]^{2}\rangle
=\displaystyle= Tn0ωnΛUdqdq(2π)21ei𝐪𝐫Bτωn2+Bq2+Kq4.\displaystyle\frac{T}{n_{0}}\sum_{\omega_{n}}\int^{\Lambda_{U}}\frac{dq_{\parallel}dq_{\perp}}{(2\pi)^{2}}\frac{1-e^{i{\bf q}\cdot{\bf r}}}{B_{\tau}\omega_{n}^{2}+Bq_{\parallel}^{2}+Kq_{\perp}^{4}}. (55)

We are interested in its asymptotics in the high-temperature classical and low-temperature quantum limits. At T=0T=0, the Matsubara converts into an integral over a continuous frequency ω\omega, giving

CQ(𝐫)=\displaystyle C_{Q}({\bf r})= UΛUdqdq(2π)21ei𝐪𝐫2Un0(Bq2+Kq4)\displaystyle\ U\int^{\Lambda_{U}}\frac{dq_{\parallel}dq_{\perp}}{(2\pi)^{2}}\frac{1-e^{i{\bf q}\cdot{\bf r}}}{\sqrt{2Un_{0}(Bq_{\parallel}^{2}+Kq_{\perp}^{4})}}
\displaystyle\approx n01ξ3/2λ1/2,rΛ1,\displaystyle\ n_{0}^{-1}\xi^{-3/2}\lambda^{-1/2},\quad\quad\ r\gg\Lambda^{-1}, (56)

growing quadratically to a finite asymptote ϕ2Q\langle\phi^{2}\rangle_{Q}, (50), at d=2d=2.

In the high-temperature limit, only ωn=0\omega_{n}=0 contributes to the correlation function, giving

CT(𝐫)=\displaystyle C_{T}({\bf r})= 2Uβdqdq(2π)21ei𝐪𝐫2Un0(Bq2+Kq4)\displaystyle\frac{2U}{\beta}\int\frac{dq_{\parallel}dq_{\perp}}{(2\pi)^{2}}\frac{1-e^{i{\bf q}\cdot{\bf r}}}{2Un_{0}(Bq_{\parallel}^{2}+Kq_{\perp}^{4})}
=\displaystyle= T2Un02ξ2[(r4πλ)1/2er24λr+r4λerf(r4λr)]\displaystyle\frac{T}{2Un_{0}^{2}\xi^{2}}\left[\left(\frac{r_{\parallel}}{4\pi\lambda}\right)^{1/2}e^{-\frac{r_{\perp}^{2}}{4\lambda r_{\parallel}}}+\frac{r_{\perp}}{4\lambda}\text{erf}\left(\frac{r_{\perp}}{\sqrt{4\lambda r_{\parallel}}}\right)\right]
\displaystyle\approx {Tξ3/2λ1/24Un02(rπξ)1/2,rr2λ,rΛU1,Tξ3/2λ1/28Un02(r2ξλ)1/2,rr2λ,rΛU1,\displaystyle\left\{\begin{array}[]{cc}\frac{T\xi^{-3/2}\lambda^{-1/2}}{4Un_{0}^{2}}\left(\frac{r_{\parallel}}{\pi\xi}\right)^{1/2},&\quad r_{\parallel}\gg\frac{r_{\perp}^{2}}{\lambda},\ r\gg\Lambda_{U}^{-1},\\ \frac{T\xi^{-3/2}\lambda^{-1/2}}{8Un_{0}^{2}}\left(\frac{r_{\perp}^{2}}{\xi\lambda}\right)^{1/2},&\quad r_{\parallel}\ll\frac{r_{\perp}^{2}}{\lambda},\ r\gg\Lambda_{U}^{-1},\end{array}\right. (59)

where erf(x)\text{erf}(x) is the error function Toner and Nelson (1981); Radzihovsky (2011).

The above analysis indicates a crossover between the low and high temperature limits of C(𝐫)C({\bf r}). To this end, we perform Matsubara summation in Eq. (55) using

1βωneiωnτiωnϵ=eτϵeβϵ1,for τ0,\frac{1}{\beta}\sum_{\omega_{n}}\frac{e^{i\omega_{n}\tau}}{i\omega_{n}-\epsilon}=-\frac{e^{\tau\epsilon}}{e^{\beta\epsilon}-1},\quad\textrm{for }\tau\geq 0, (60)

and define a crossover scale

𝐪T=(qT,qT)=(TUn0qc,TUn0qc),\displaystyle{\bf q}_{T}=(q_{\parallel T},q_{\perp T})=\left(\frac{T}{Un_{0}}q_{\parallel c},\sqrt{\frac{T}{Un_{0}}}q_{\perp c}\right), (61)

separating the quantum (𝐪>𝐪T{\bf q}>{\bf q}_{T}) and classical (𝐪<𝐪T{\bf q}<{\bf q}_{T}) regions of momenta. This gives

C(𝐫)=\displaystyle C({\bf r})= U(qT+qTΛU)1ei𝐪𝐫E𝐪coth(βE𝐪/2)\displaystyle\ U\left(\int^{q_{T}}+\int_{q_{T}}^{\Lambda_{U}}\right)\frac{1-e^{i{\bf q}\cdot{\bf r}}}{E_{\bf q}}\coth(\beta E_{\bf q}/2)
\displaystyle\approx 2UβqT1ei𝐪𝐫E𝐪2+UqTΛU1ei𝐪𝐫E𝐪,\displaystyle\ \frac{2U}{\beta}\int^{q_{T}}\frac{1-e^{i{\bf q}\cdot{\bf r}}}{E_{\bf q}^{2}}+U\int_{q_{T}}^{\Lambda_{U}}\frac{1-e^{i{\bf q}\cdot{\bf r}}}{E_{\bf q}}, (62)

where E𝐪=Bτ1(Bq2+Kq4)E_{\bf q}=\sqrt{B_{\tau}^{-1}(Bq_{\parallel}^{2}+Kq_{\perp}^{4})}. At low temperature, β\beta\rightarrow\infty and qT0q_{T}\rightarrow 0, the quantum, second term dominates, giving C(𝐫)CQ(𝐫)C({\bf r})\approx C_{Q}({\bf r}). While at high temperature, β0\beta\rightarrow 0 and qTΛUq_{T}\rightarrow\Lambda_{U}, C(𝐫)C({\bf r}) is dominated by the classical, first term, giving C(𝐫)CT(𝐫)C({\bf r})\approx C_{T}({\bf r}).

Refer to caption
Figure 8: Numerical plots (rr_{\parallel} and rr_{\perp} in units of ξ\xi and ξλ\sqrt{\xi\lambda}, respectively) of the correlation function (55) (a) along rr_{\perp} with r=0r_{\parallel}=0 and (b) along rr_{\parallel} with r=0r_{\perp}=0. The helical superfluid exhibits a quantum-to-classical (Q-to-C) crossover around r=ξλUn0/Tr_{\perp}=\sqrt{\xi\lambda}\sqrt{Un_{0}/T} and r=ξUn0/Tr_{\parallel}=\xi Un_{0}/T with Un0/T=10Un_{0}/T=10, where the finite-temperature (black-solid) curves deviate from the zero-temperature (blue-dashed) curves. For B0B_{\perp}\neq 0, the correlation function (c) along rr_{\perp} with r=0r_{\parallel}=0 and (d) along rr_{\parallel} with r=0r_{\perp}=0 (black curves) shows a crossover from smectic superfluidity (SFSm\text{SF}_{\text{Sm}}) to conventional XY superfluidity (SFXY\text{SF}_{\text{XY}}) around r=λr_{\perp}=\lambda_{\perp} and r=λ2/ξr_{\parallel}=\lambda_{\perp}^{2}/\xi with λ=10ξ\lambda_{\perp}=10\xi. The red-dashed curves show the case B=0B_{\perp}=0 for comparison. Inset of (c)(d): logarithmic scale in 𝐫{\bf r} axes.

In Fig. 8, we plot C(𝐫)C({\bf r}) in (55), computed numerically along rr_{\perp} and rr_{\parallel}, showing its asymptotic behaviors in the low-temperature (56) and high-temperature (59) regimes. The crossover scale in real space is given by the inverse of the thermal wavevectors,

rT=qT1=Un0Tξ,rT=qT1=Un0Tλξ.\displaystyle r_{\parallel T}=q_{\parallel T}^{-1}=\frac{Un_{0}}{T}\xi,\quad r_{\perp T}=q_{\perp T}^{-1}=\sqrt{\frac{Un_{0}}{T}}\sqrt{\lambda\xi}. (63)

IV.1.3 Order-by-disorder

Although the effective theory introduced in this section is unstable to thermal fluctuations for d3d\leq 3, we anticipate that the presence of the underlying honeycomb lattice explicitly breaks the degenerate contour in Eq. (40) down to six-fold minima through quantum and thermal fluctuations, that probe all momentum scales (detailed in Sec. V). This in turn introduces perpendicular stiffness B(ϕ)2B_{\perp}(\nabla_{\perp}\phi)^{2}, modifying the low-energy Lagrangian to

ϕn0=Bτ(τϕ)2+B(ϕ)2+B(ϕ)2+K(2ϕ)2,\displaystyle\frac{\mathcal{L}_{\phi}}{n_{0}}=B_{\tau}(\partial_{\tau}\phi)^{2}+B(\partial_{\parallel}\phi)^{2}+B_{\perp}(\partial_{\perp}\phi)^{2}+K(\partial_{\perp}^{2}\phi)^{2}, (64)

with suppressed fluctuations. As discussed in Sec. V, BU5/4B_{\perp}\propto U^{5/4} (BU1/4TB_{\perp}\propto U^{1/4}T) for TUn0T\ll Un_{0} (TUn0T\gg Un_{0}) in the weakly-interacting limit Un0t1Un_{0}\ll t_{1}. Importantly, this leads to a crossover length scale λ=K/B\lambda_{\perp}=\sqrt{K/B_{\perp}}. Consequently, as illustrated in Fig. 8(c)(d), the helical superfluid exhibits an extended crossover from a smectic regime to an anisotropic XY regime (but with the condensate at finite momentum), separated by (r,r)=(λ2/ξ,λ)(r_{\parallel}^{*},r_{\perp}^{*})=(\lambda_{\perp}^{2}/\xi,\lambda_{\perp}). Since BB_{\perp} is perturbatively small in UU, λ\lambda_{\perp} can be very large, exhibiting long range of anomalous smectic superfluidity, that we study in detail below.

IV.2 Physical observables in smectic superfluid regime

IV.2.1 Structure factor

To calculate the structure factor, we employ the Lagrangian (43). In Fourier space, it is given by

0=12(ϕ𝐪,ωnπ𝐪,ωn)D1(𝐪,ωn)(ϕ𝐪,ωnπ𝐪,ωn),\mathcal{L}^{\prime}_{0}=\frac{1}{2}\biggl{(}\begin{array}[]{cc}\phi_{-{\bf q},-\omega_{n}}&\pi_{-{\bf q},-\omega_{n}}\end{array}\biggr{)}D^{-1}({\bf q},\omega_{n})\biggl{(}\begin{array}[]{c}\phi_{{\bf q},\omega_{n}}\\ \pi_{{\bf q},\omega_{n}}\end{array}\biggr{)}, (65)

where

D(𝐪,ωn)=\displaystyle D({\bf q},\omega_{n})= 1(ωn+i2,𝐪)2+𝐪2+2Un0𝐪\displaystyle\ \frac{1}{(\omega_{n}+i\mathcal{E}_{2,{\bf q}})^{2}+\mathcal{E}_{\bf q}^{2}+2Un_{0}\mathcal{E}_{\bf q}}
×(12n0𝐪+Uωn+i2,𝐪ωni2,𝐪2n0𝐪)\displaystyle\times\biggl{(}\begin{array}[]{cc}\frac{1}{2n_{0}}\mathcal{E}_{\bf q}+U&\omega_{n}+i\mathcal{E}_{2,{\bf q}}\\ -\omega_{n}-i\mathcal{E}_{2,{\bf q}}&2n_{0}\mathcal{E}_{\bf q}\end{array}\biggr{)} (68)

with 𝐪\mathcal{E}_{\bf q} and 2,𝐪\mathcal{E}_{2,{\bf q}} defined in Eq. (45).

The structure factor is the density-density correlation function, in the Fourier space given by

S(𝐪,ωn)=D22(𝐪,ωn)=2n0𝐪(ωn+i2,𝐪)2+𝐪2+2Un0𝐪.S({\bf q},\omega_{n})=D_{22}({\bf q},\omega_{n})=\frac{2n_{0}\mathcal{E}_{\bf q}}{(\omega_{n}+i\mathcal{E}_{2,{\bf q}})^{2}+\mathcal{E}_{\bf q}^{2}+2Un_{0}\mathcal{E}_{\bf q}}. (69)

This static structure factor is then given by

S(𝐪)\displaystyle S({\bf q}) =1βωnS(𝐪,ωn)\displaystyle=\frac{1}{\beta}\sum_{\omega_{n}}S({\bf q},\omega_{n})
=n0𝐪E𝐪2,𝐪sinh[β(E𝐪2,𝐪)]cosh[β(E𝐪2,𝐪)]cosh(β2,𝐪),\displaystyle=\frac{n_{0}\mathcal{E}_{\bf q}}{E_{\bf q}-\mathcal{E}_{2,{\bf q}}}\frac{\sinh[\beta(E_{\bf q}-\mathcal{E}_{2,{\bf q}})]}{\cosh[\beta(E_{\bf q}-\mathcal{E}_{2,{\bf q}})]-\cosh(\beta\mathcal{E}_{2,{\bf q}})}, (70)

where E𝐪E_{\bf q} is defined in Eq. (49) and we used Eq. (60) to carry out the Matsubara sum.

Refer to caption
Figure 9: Static structure factor (70) (in unit of n0n_{0}) as a function of 𝐪{\bf q} (in the unit of λ1\lambda^{-1}) with λ/ξ=0.3\lambda/\xi=0.3. (a) High-temperature classical limit with qTλ=0.4q_{\parallel T}\lambda=0.4. (b) Low-temperature quantum limit qTλ=0.01q_{\parallel T}\lambda=0.01. Quantum-to-classical crossover (c) along qq_{\parallel} with q=0q_{\perp}=0 and (d) along qq_{\perp} with q=0q_{\parallel}=0. From top to bottom the temperature decreases with qTλ=0.4,0.3,0.2,0.1,0.01q_{\parallel T}\lambda=0.4,0.3,0.2,0.1,0.01. The red double-headed arrows in (a) and (c) indicate the scale of condensate momentum 2k0=λ12k_{0}=\lambda^{-1}.

The structure factor (70) depends on the a number of scales: coherence length ξ=BBτ\xi=\sqrt{BB_{\tau}}, anisotropic length factor λ=K/B\lambda=\sqrt{K/B} and parallel thermal wavevector qT=(2T2/Un0B)1/2q_{\parallel T}=(2T^{2}/Un_{0}B)^{1/2}. Below we first present the results in the low- and high-temperature limits, and then discuss the crossover between them.

In the high-temperature limit, β0\beta\to 0, the structure factor becomes

ST(𝐪)=\displaystyle S_{T}({\bf q})= 2n0T𝐪2Un0𝐪+𝐪22,𝐪2\displaystyle\ \frac{2n_{0}T\mathcal{E}_{\bf q}}{2Un_{0}\mathcal{E}_{\bf q}+\mathcal{E}_{\bf q}^{2}-\mathcal{E}_{2,{\bf q}}^{2}}
\displaystyle\approx {T/U,qξ,qλξ1,0,qξ,qλξ1,\displaystyle\ \biggl{\{}\begin{array}[]{cc}T/U,&\quad q_{\parallel}\xi,\ q_{\perp}\sqrt{\lambda\xi}\ll 1,\\ 0,&\quad q_{\parallel}\xi,\ q_{\perp}\sqrt{\lambda\xi}\gg 1,\end{array} (73)

exhibiting a maximum on a closed contour (inherited from the noninteracting dispersion minimum) with a height T/UT/U and a width increasing with UU, illustrated at high temperature in Fig. 9(a).

In the complementary low-temperature limit, β\beta\to\infty, the structure factor becomes

SQ(𝐪)=\displaystyle S_{Q}({\bf q})= n0𝐪𝐪+2Un0\displaystyle\ n_{0}\sqrt{\frac{\mathcal{E}_{\bf q}}{\mathcal{E}_{\bf q}+2Un_{0}}}
\displaystyle\approx {n0𝐪/2U,qξ,qλξ1,n0,qξ,qλξ1,\displaystyle\ \biggl{\{}\begin{array}[]{cc}\sqrt{n_{0}\mathcal{E}_{\bf q}/2U},&\quad q_{\parallel}\xi,\ q_{\perp}\sqrt{\lambda\xi}\ll 1,\\ n_{0},&\quad q_{\parallel}\xi,\ q_{\perp}\sqrt{\lambda\xi}\gg 1,\end{array} (76)

in contrast exhibiting a single minimum at 𝐪=0{\bf q}=0, which becomes more shallow with increases UU, as shown in Fig. 9(b). It can be written as a scaling form

SQ(𝐪,0)n0(ξq)αS^(λq2q)S_{Q}({\bf q},0)\sim n_{0}(\xi q_{\parallel})^{\alpha}\hat{S}\biggl{(}\lambda\frac{q^{2}}{q_{\parallel}}\biggr{)} (77)

in the limits, ξq1\xi q_{\parallel}\ll 1 and ξq1\xi q_{\parallel}\gg 1, with

S^(x)\displaystyle\hat{S}(x) =1+x2,α=1,\displaystyle=\sqrt{1+x^{2}},\quad\alpha=1,\quad for ξq1,\displaystyle\text{for }\xi q_{\parallel}\ll 1,
S^(x)=1,α=0,\displaystyle\hat{S}(x)=1,\quad\alpha=0,\quad for ξq1.\displaystyle\text{for }\xi q_{\parallel}\gg 1. (78)

The structure factor exhibits a quantum-to-classical crossover for 𝐪{\bf q} separated by the thermal wavevectors 𝐪T{\bf q}_{T} in Eq. (61), as illustrated in Fig. 9(c)(d).

IV.2.2 Condensate depletion

The condensate depletion is given by

nd=\displaystyle n_{d}= nn0=Φ(𝐫,0)Φ(𝐫,η)n0\displaystyle\ n-n_{0}=\langle\Phi({\bf r},0){\Phi}^{\ast}({\bf r},\eta)\rangle-n_{0}
=\displaystyle= 1βd2q(2π)2ωneiωnηG(𝐪,ωn),\displaystyle\ \frac{1}{\beta}\int\frac{d^{2}q}{(2\pi)^{2}}\sum_{\omega_{n}}e^{-i\omega_{n}\eta}G({\bf q},\omega_{n}), (79)

where η=0+\eta=0^{+} is an infinitesimal positive number to encode operator time ordering and with Φ\Phi in Eq. (42) expanded up to quadratic order in π\pi and ϕ\phi,

G(𝐪,ωn)\displaystyle G({\bf q},\omega_{n})\approx 14n0D22(𝐪,ωn)+n0D11(𝐪,ωn)\displaystyle\ \frac{1}{4n_{0}}D_{22}({\bf q},\omega_{n})+n_{0}D_{11}({\bf q},\omega_{n})
+i2D12(𝐪,ωn)i2D21(𝐪,ωn)\displaystyle+\frac{i}{2}D_{12}({\bf q},\omega_{n})-\frac{i}{2}D_{21}({\bf q},\omega_{n})
=\displaystyle= u𝐪2iωn+E𝐪+v𝐪2iωn+E𝐪\displaystyle\ \frac{u_{\bf q}^{2}}{-i\omega_{n}+E_{\bf q}}+\frac{v_{\bf q}^{2}}{i\omega_{n}+E_{-\bf q}} (80)

with matrix DD given in Eq. (68), E𝐪E_{\bf q} given in Eq. (49) and

u𝐪2=\displaystyle u_{\bf q}^{2}= 12(𝐪+Un0𝐪2+2Un0𝐪+1),\displaystyle\frac{1}{2}\biggl{(}\frac{\mathcal{E}_{\bf q}+Un_{0}}{\sqrt{\mathcal{E}_{\bf q}^{2}+2Un_{0}\mathcal{E}_{\bf q}}}+1\biggr{)},
v𝐪2=\displaystyle v_{\bf q}^{2}= 12(𝐪+Un0𝐪2+2Un0𝐪1).\displaystyle\frac{1}{2}\biggl{(}\frac{\mathcal{E}_{\bf q}+Un_{0}}{\sqrt{\mathcal{E}_{\bf q}^{2}+2Un_{0}\mathcal{E}_{\bf q}}}-1\biggr{)}. (81)

After performing the Matsubara summation in Eq. (79) and using the identity (60),

nd=\displaystyle n_{d}= d2q(2π)2(v𝐪2+u𝐪2+v𝐪2eβE𝐪1),\displaystyle\int\frac{d^{2}q}{(2\pi)^{2}}\biggl{(}v_{\bf q}^{2}+\frac{u_{\bf q}^{2}+v_{\bf q}^{2}}{e^{\beta E_{\bf q}}-1}\biggr{)}, (82)

where the first term gives the zero-temperature interaction driven depletion while the second term describes additional depletion due to thermal excitations.

At zero temperature, the depletion (with details relegated to Appendix E) is given by

nd(T=0)=\displaystyle n_{d}(T=0)= d2q(2π)2v𝐪2=𝒞n(U3n0B2K)1/4\displaystyle\ \int\frac{d^{2}q}{(2\pi)^{2}}v_{\bf q}^{2}=\mathcal{C}n\biggl{(}\frac{U^{3}}{n_{0}B^{2}K}\biggr{)}^{1/4}
\displaystyle\sim nU3/4n01/4(Un0)3/4,\displaystyle\ nU^{3/4}n_{0}^{-1/4}\sim(Un_{0})^{3/4},{} (83)

where 𝒞=23/4Iθ|fe(0)|/(96π2)0.035\mathcal{C}=2^{3/4}I_{\theta}|f_{e}(0)|/(96\pi^{2})\approx 0.035 with Iθ10.4882I_{\theta}\approx 10.4882 and fe(0)3.70815f_{e}(0)\approx-3.70815.

88footnotetext: In the weakly interacting limit, we neglect the difference between nn and n0n_{0} as it is of higher order in UU.

At nonzero temperature, the depletion in Eq. (82) diverges in the thermodynamic limit for d3d\leq 3, as analyzed in Appendix E, indicating helical superfluid thermal instability as found in Sec. IV.1.1.

For the microscopic lattice model discussed in Sec. III, the depleting is still given by (82), but with u𝐪u_{\bf q} and v𝐪v_{\bf q} given by

u𝐪2\displaystyle u_{{\bf q}}^{2} =12(𝐪+Un𝐪2+2Un𝐪+U2n2sin2(Θ𝐪)+1),\displaystyle=\frac{1}{2}\left(\frac{\mathcal{E}_{{\bf q}}+Un}{\sqrt{\mathcal{E}^{2}_{{\bf q}}+2Un\mathcal{E}_{{\bf q}}+U^{2}n^{2}\sin^{2}(\Theta_{\bf q})}}+1\right), (84)
v𝐪2\displaystyle v_{{\bf q}}^{2} =12(𝐪+Un𝐪2+2Un𝐪+U2n2sin2(Θ𝐪)1),\displaystyle=\frac{1}{2}\left(\frac{\mathcal{E}_{{\bf q}}+Un}{\sqrt{\mathcal{E}^{2}_{{\bf q}}+2Un\mathcal{E}_{{\bf q}}+U^{2}n^{2}\sin^{2}(\Theta_{\bf q})}}-1\right),

where the lattice effects are encoded in the periodic (in momentum space) forms of Θ𝐪\Theta_{\bf q} and 𝐪\mathcal{E}_{\bf q} defined in Eqs. (28) and (33).

At zero temperature, the depletion as a function of UnUn is computed numerically and presented in Fig. 10. At small UnUn, it is a power-law as found in Eq. (83). As UnUn increases, ndn_{d} deviates from this perfect power-law given by a weakly interacting field theory. At nonzero temperature, thermal fluctuations still destroy the condensate, as discussed above.

Refer to caption
Figure 10: The log-log plot of the zero temperature depletion ndn_{d} (in unit of a2a^{-2}) in Eq. (83) as a function of Un/t1Un/t_{1} for ρ=0.2\rho=0.2, and 𝐤0=(0,k0V){\bf k}_{0}=(0,k_{0V}). The red dots are calculated with the lattice model expression (84). The dashed line gives a fitting in the small UnUn region with the predicted slope =3/4=3/4.

IV.2.3 Momentum distribution

Another important characterization of the helical superfluid is the momentum distribution, n𝐪=d𝐤0+𝐪,d𝐤0+𝐪,n_{\bf q}=\langle d^{\dagger}_{{\bf k}_{0}+{\bf q},-}d_{{\bf k}_{0}+{\bf q},-}\rangle. At T=0T=0, it is given by

n𝐪=|v𝐪|2=12(1+Un0𝐪1+2Un0𝐪1).n_{{\bf q}}=|v_{{\bf q}}|^{2}=\frac{1}{2}\biggl{(}\frac{1+\frac{Un_{0}}{\mathcal{E}_{{\bf q}}}}{\sqrt{1+\frac{2Un_{0}}{\mathcal{E}_{{\bf q}}}}}-1\biggr{)}. (85)

The behavior of n𝐪n_{\bf q} below (ξq1\xi q_{\parallel}\gg 1) and beyond (ξq1\xi q_{\parallel}\ll 1) the coherence length ξ=BBτ\xi=\sqrt{BB_{\tau}} is described by the following scaling form,

n𝐪(ξq)αfn(λq2q),n_{{\bf q}}\sim(\xi q_{\parallel})^{\alpha}f_{n}\biggl{(}\lambda\frac{q_{\perp}^{2}}{q_{\parallel}}\biggr{)}, (86)

with the anisotropy length scale given by λ=K/B\lambda=\sqrt{K/B} and

fn(x)=11+x2,α=\displaystyle f_{n}(x)=\frac{1}{\sqrt{1+x^{2}}},\quad\alpha= 1,for ξq1,\displaystyle\ -1,\quad\text{for }\xi q_{\parallel}\ll 1,
fn(x)=1(1+x2)2,α=\displaystyle f_{n}(x)=\frac{1}{(1+x^{2})^{2}},\quad\alpha= 4,for ξq1.\displaystyle\ -4,\quad\text{for }\xi q_{\parallel}\gg 1. (87)

We note that for λq2/q1\lambda q_{\perp}^{2}/q_{\parallel}\ll 1, n𝐪(ξq)αn_{\bf q}\sim(\xi q_{\parallel})^{\alpha}. In the opposite limit, λq2/q1\lambda q_{\perp}^{2}/q_{\parallel}\gg 1, the momentum distribution is asymptotically given by

n𝐪{1/q2, for qξ1.q3/q8, for qξ1.\displaystyle n_{{\bf q}}\sim\biggl{\{}\begin{array}[]{cc}\vspace{1mm}1/q_{\perp}^{2},&\textrm{ for }q_{\parallel}\xi\ll 1.\\ q_{\parallel}^{3}/q_{\perp}^{8},&\textrm{ for }q_{\parallel}\xi\gg 1.\end{array} (90)

This contrasts qualitatively with that of a conventional superfluid (see Table 1).

Lattice effects beyond this field theoretical treatment are straightforwardly incorporated with n𝐪=v𝐪2n_{\bf q}=v_{\bf q}^{2} at T=0T=0, with the lattice form of v𝐪v_{\bf q} given by Eq. (84). In Fig. 11, we plot nq,q=0n_{q_{\parallel},q_{\perp}=0} as a function of qq_{\parallel} which verifies the scaling behavior nq,q=01/qn_{q_{\parallel},q_{\perp}=0}\sim 1/q_{\parallel} (1/q41/q_{\parallel}^{4}) for qξ1q_{\parallel}\xi\ll 1 (qξ1q_{\parallel}\xi\gg 1). Note, somewhat surprisingly the (T=0T=0) momentum distribution is symmetric under the inversion of momentum n𝐪=n𝐪n_{\bf q}=n_{-{\bf q}}, and its profile has the reflection symmetry, e.g. with respect to q=2π/3q_{\parallel}=2\pi/3, due to the underlying lattice structure, as shown in the inset of Fig. 11. As a result, the 1/q41/q_{\parallel}^{4} tail is only visible for sufficiently small UnUn.

Refer to caption
Figure 11: Momentum distribution nq,q=0n_{q_{\parallel},q_{\perp}=0} vs. qq_{\parallel} (in unit of a1a^{-1}) for ρ=0.2\rho=0.2, Un/t1=104Un/t_{1}=10^{-4}, and 𝐤0=(0,k0V){\bf k}_{0}=(0,k_{0V}). The red dots are calculated from lattice model expression (84). The dashed line gives the asymptotic behavior for qξ1q_{\parallel}\xi\ll 1 and qξ1q_{\parallel}\xi\gg 1. Inset: linear scale of the same plot.

IV.2.4 Superflow

As in a conventional superfluid a spatial gradient of ϕ\phi gives rise to a supercurrent,

𝐣s=n𝐯s\displaystyle{\bf j}_{s}=n{\bf v}_{s} (91)

characterized by the anisotropic superfluid velocity with its form obtained from the Noether’s theorem or, equivalently by gauging (see Appendix F)

𝐯s=4J[(ϕ)2+2𝐤0ϕ](𝐤0+ϕ)2J3ϕ.\displaystyle{\bf v}_{s}=4J\left[(\nabla\phi)^{2}+2{\bf k}_{0}\cdot\nabla\phi\right]({\bf k}_{0}+\nabla\phi)-2J\nabla^{3}\phi. (92)

We emphasize that in 𝐯s{\bf v}_{s} of (92) we included higher-order terms, previously left out in the quadratic Lagrangian, (43). For the simplest uniform gradient configuration ϕ=𝐪𝐫\phi={\bf q}\cdot{\bf r}, we have

𝐯s(ϕ=𝐪𝐫)=\displaystyle{\bf v}_{s}(\phi={\bf q}\cdot{\bf r})= 4J[q2+2𝐤0𝐪](𝐤0+𝐪)\displaystyle 4J\left[q^{2}+2{\bf k}_{0}\cdot{\bf q}\right]({\bf k}_{0}+{\bf q})
=\displaystyle= kε𝐤|𝐤=𝐤0+𝐪\displaystyle\nabla_{k}\varepsilon_{\bf k}|_{{\bf k}={\bf k}_{0}+{\bf q}} (93)

with

ε𝐤=J(k2k02)2+ε0,𝐤=𝐤0+𝐪.\varepsilon_{\bf k}=J(k^{2}-k_{0}^{2})^{2}+\varepsilon_{0},\quad{\bf k}={\bf k}_{0}+{\bf q}. (94)

In the long-wavelength limit to linear order, the supercurrent is given by

(js)i=ϕ0j(ρs)ijjϕ\displaystyle(j_{s})_{i}\underset{\nabla\phi\to 0}{=}\sum_{j}(\rho_{s})_{ij}\nabla_{j}\phi (95)

with the superfluid stiffness (ρs)ij=8Jn(k0)i(k0)j(\rho_{s})_{ij}=8Jn(k_{0})_{i}(k_{0})_{j}, which vanishes transversely to 𝐤0{\bf k}_{0}. This helical condensate thus (in the absent of stabilizing lattice effects) cannot maintain a superflow perpendicular to 𝐤0{\bf k}_{0}, and thus by this criterion is not a superfluid.

IV.2.5 Chemical potential

Refer to caption
Figure 12: μ(1)\mu^{(1)} (in unit of Ua2Ua^{-2}) vs. Un/t1Un/t_{1} for ρ=0.2\rho=0.2, and 𝐤0=(0,k0V){\bf k}_{0}=(0,k_{0V}). The red dots are calculated from (IV.2.5). The dashed line gives a fitting to (98) in the small UnUn region with a slope =3/4=3/4.

We next calculate the equation of state, given by the chemical potential as a function of the boson density nn,

μ=EgsN=μ(0)+μ(1),\displaystyle\mu=\frac{\partial E_{gs}}{\partial N}=\mu^{(0)}+\mu^{(1)}, (96)

where ground state energy EgsE_{gs} is given by Eq. (30). The lowest-order mean-field contribution and the one-loop correction are respectively given by

μ(0)=\displaystyle\mu^{(0)}= ϵ𝐤0+Un,\displaystyle\ \epsilon^{-}_{{\bf k}_{0}}+Un,
μ(1)=\displaystyle\mu^{(1)}= U4𝐪(1𝐪+Un(sinΦ𝐪)2𝐪2+2Un𝐪+(UnsinΦ𝐪)2)\displaystyle\ -\frac{U}{4}\int_{\bf q}\left(1-\frac{\mathcal{E}_{{\bf q}}+Un(\sin\Phi_{\bf q})^{2}}{\sqrt{\mathcal{E}^{2}_{{\bf q}}+2Un\mathcal{E}_{\bf q}+(Un\sin\Phi_{\bf q})^{2}}}\right)
\displaystyle\approx U4𝐪(1𝐪𝐪2+2Un𝐪),\displaystyle\ -\frac{U}{4}\int_{\bf q}\left(1-\frac{\mathcal{E}_{{\bf q}}}{\sqrt{\mathcal{E}^{2}_{{\bf q}}+2Un\mathcal{E}_{\bf q}}}\right), (97)

where in the last line we neglected lattice effects and took 𝐪Bq2+Kq4\mathcal{E}_{{\bf q}}\approx Bq_{\parallel}^{2}+Kq_{\perp}^{4}.

Performing the integration (see Appendix G), we obtain

μ(1)Un𝒞(U3nB2K)1/4,\mu^{(1)}\approx-Un\mathcal{C}\biggl{(}\frac{U^{3}}{nB^{2}K}\biggr{)}^{1/4}, (98)

where constants 𝒞=23/4IθIσ/(32π2)0.069\mathcal{C}=2^{3/4}I_{\theta}I_{\sigma}/(32\pi^{2})\approx 0.069, Iσ1.23605I_{\sigma}\approx 1.23605 and Iθ10.4882I_{\theta}\approx 10.4882. The dimensionless factor in μ(1)\mu^{(1)} expressed in terms of microscopic parameters is given by

(U3nB2K)1/4n1/4U3/4t13/4k¯01,\displaystyle\biggl{(}\frac{U^{3}}{nB^{2}K}\biggr{)}^{1/4}\propto n^{-1/4}U^{3/4}t_{1}^{-3/4}\bar{k}_{0}^{-1}, (99)

where k¯0(ρ)\bar{k}_{0}(\rho) is given by Eq. (39) in the limit of ρ1/6\rho\to 1/6. For a general dd dimensional helical superfluid, this dimensionless factor is given by

𝒬s=(Ud+1nd3B2Kd1)1/4.\mathcal{Q}_{s}=\biggl{(}\frac{U^{d+1}n^{d-3}}{B^{2}K^{d-1}}\biggr{)}^{1/4}. (100)

In Fig. 12, we plot an exact numerical evaluation of the chemical potential in Eq. (IV.2.5) with lattice effects as a function of UnUn. At small UnUn, Eq. (98) fits the data well. We note, as indicated in Eq. (98), μ/U\mu/U decreases (increases) if UU(nn) increases, in qualitative distinction from a conventional superfluid (see Table 1).

V Order-by-disorder

So far, we studied the helical condensation at a single momentum 𝐤0{\bf k}_{0} on the dispersion minimum contour at the Bogoliubov level. At this order, it exhibits smectic fluctuations that are qualitatively larger than that of conventional XY superfluids due to the macroscopic degenerate ground states. In this section, we go beyond the quadratic action, and study the order-by-disorder phenomenon that appears due to the interplay of quantum and thermal fluctuations, and the C6\text{C}_{\text{6}} lattice effects. This generates a nonzero BB_{\perp} that stabilizes the helical superfluid even at nonzero temperature.

Formally, we work with the coherent state path integral of the lattice model introduced in Sec. III, where the partition function is given by Z=𝒟ΦeS[Φ]Z=\int\mathcal{D}\Phi e^{-S\left[\Phi\right]}, with the imaginary-time action S=𝑑τLS=\int d\tau L and the corresponding Lagrangian

L=i,a=1,2Φi,aτΦi,a+H0+Hint.\displaystyle L=\sum_{i,a=1,2}\Phi_{i,a}^{*}\partial_{\tau}\Phi_{i,a}+H_{0}+H_{\text{int}}. (101)

In the above, H0H_{0} and HintH_{\text{int}} are functionals that respectively have the same expressions as the kinetic and the interacting parts of the Hamiltonian introduced in Sec. III with the operators ai,aa_{i,a} and ai,aa_{i,a}^{\dagger} replaced by coherent state fields Φi,a\Phi_{i,a} and Φi,a\Phi_{i,a}^{*}. To study the helical superfluid, we consider the density-phase fluctuations around a mean-field of a helical condensate with density n0n_{0} and momentum 𝐤0{\bf k}_{0}:

Φi,1=\displaystyle\Phi_{i,1}= eiθ𝐤0/2ei𝐤0𝐫i,1+iϕi,1n0+πi,1,\displaystyle\ e^{i\theta_{{\bf k}_{0}}/2}~{}e^{i{\bf k}_{0}\cdot{\bf r}_{i,1}+i\phi_{i,1}}~{}\sqrt{n_{0}+\pi_{i,1}},
Φi,2=\displaystyle\Phi_{i,2}= eiθ𝐤0/2ei𝐤0𝐫i,2+iϕi,2n0+πi,2,\displaystyle\ e^{-i\theta_{{\bf k}_{0}}/2}~{}e^{i{\bf k}_{0}\cdot{\bf r}_{i,2}+i\phi_{i,2}}~{}\sqrt{n_{0}+\pi_{i,2}}, (102)

where 𝐫i,a{\bf r}_{i,a} is the position of the lattice site ii in aa sublattice. Below, we drop the subscripts ii and aa for notation simplicity. Then, the action can be written as S[Φ]=βΩ(0)+τL(0)S[\Phi]=\beta\Omega^{(0)}+\int_{\tau}L^{(0)} with the constant mean-field part Ω(0)(n0,𝐤0)\Omega^{(0)}(n_{0},{\bf k}_{0}) (zeroth-order thermodynamic potential) and the bare Lagrangian, L(0)(π,ϕ)L^{(0)}(\pi,\phi). Such a procedure parallels the field theoretical approach in Sec. IV, but now encodes the underlying C6\text{C}_{\text{6}} lattice effects. Analysis of the quadratic part of the action reproduces the Bogoliubov results, predicted in Sec. III. To study the effects of the higher-order nonlinearalities, various approximation schemes can be employed. Here we safely employ an analysis perturbative in Un0Un_{0}, self-consistently verifying the corrections to BB_{\perp} are finite. We thereby formally obtain

Z=\displaystyle Z= 𝒟π𝒟ϕeβΩ(0)τL(0)=𝒟π𝒟ϕeβΩτL,\displaystyle\ \int\mathcal{D}\pi\mathcal{D}\phi e^{-\beta\Omega^{(0)}-\int_{\tau}L^{(0)}}=\int\mathcal{D}\pi\mathcal{D}\phi e^{-\beta\Omega-\int_{\tau}L}, (103)

where the renormalized thermodynamic potential and Lagrangian are given by perturbation series Ω(n0,𝐤0)=n=0Ω(n)\Omega(n_{0},{\bf k}_{0})=\sum_{n=0}^{\infty}\Omega^{(n)} and L=n=0L(n)L=\sum_{n=0}^{\infty}L^{(n)} [“(n)(n)” stands for the nth loop correction], respectively. Similar to a conventional superfluid, global U(1) symmetry ϕϕ+α\phi\to\phi+\alpha ensures that ϕ\phi is massless, i.e., in the continuum, only gradients of ϕ\phi appears in the Lagrangian. For a helical condensate, another constraint is the equivalence of the transformations 𝐤0𝐤0+𝐪{\bf k}_{0}\to{\bf k}_{0}+{\bf q} and ϕϕ+𝐪𝐫\phi\to\phi+{\bf q}\cdot{\bf r} in the long wavelength limit. The former transforms the free energy,

Ω(𝐤0+𝐪)Ω(𝐤0)=N0lmblmqlqm.\displaystyle\Omega({\bf k}_{0}+{\bf q})-\Omega({\bf k}_{0})=N_{0}\sum_{lm}b_{lm}q_{\parallel}^{l}q_{\perp}^{m}. (104)

The latter, when applied to the continuum Lagrangian density (discussed in detail in Appendix H),

=n0lmBlm(ϕ)l(ϕ)m+,\displaystyle\mathcal{L}=n_{0}\sum_{lm}B_{lm}(\partial_{\parallel}\phi)^{l}(\partial_{\perp}\phi)^{m}+..., (105)

gives an energy shift lmBlmqlqm\sum_{lm}B_{lm}q_{\parallel}^{l}q_{\perp}^{m}. In the above, the ellipsis denotes π\pi and higher derivative ϕ\phi terms. Since 𝐪{\bf q} can be chosen arbitrarily, we conclude that

blm=Blm,\displaystyle b_{lm}=B_{lm}, (106)

for any integers l,m0l,m\geq 0, holding in every nth order of the perturbation theory, blm(n)=Blm(n)b_{lm}^{(n)}=B_{lm}^{(n)}. Below, we study the order-by-disorder phenomenon by performing one-loop calculations of b(1)=b02(1)b_{\perp}^{(1)}=b_{02}^{(1)} and B(1)=B02(1)B_{\perp}^{(1)}=B_{02}^{(1)} for both zero and nonzero temperatures.

V.1 Thermodynamic potential

Refer to caption
Figure 13: bb_{\perp} (b(1)\approx b_{\perp}^{(1)}, in unit of t1a2t_{1}a^{2}) vs. Un/t1Un/t_{1} for ρ=0.3\rho=0.3, T=0T=0, N0=1N_{0}=1. The black-dashed line gives a fitting to (111) in the small UnUn region with a slope =5/4=5/4.

Before working in the more convenient functional integral formalism, we note the ground state energy Egs(𝐤0)=Egs(0)(𝐤0)+Egs(1)(𝐤0)E_{gs}({\bf k}_{0})=E_{gs}^{(0)}({\bf k}_{0})+E_{gs}^{(1)}({\bf k}_{0}) in Eq. (30) consists of two contributions, the mean-field and the zero-point energies, respectively, given by

Egs(0)=ϵ𝐤0N+nNU2,Egs(1)=12𝐪(E𝐪ε𝐪).\displaystyle E_{gs}^{(0)}=\epsilon^{-}_{{\bf k}_{0}}N+\frac{nNU}{2},\quad E_{gs}^{(1)}=\frac{1}{2}\sum_{{\bf q}}(E_{{\bf q}}-\varepsilon^{\prime}_{{\bf q}}). (107)

The former is the dominate contribution that exhibits a sub-extensive degeneracy on the contour minimum, while the latter is the lowest-order in Un0Un_{0} correction, corresponding to the one-loop correction in the field theory treatment below.

In the coherent-state path integral, the zeroth-order thermodynamic potential is equal to the mean-field ground state energy above (with NN0N\to N_{0})

Ω(0)=Egs(0),\displaystyle\Omega^{(0)}=E^{(0)}_{gs}, (108)

and the one-loop correction (neglecting the upper band, valid for Un0/t11Un_{0}/t_{1}\ll 1) is given by

Ω(1)(N0,𝐤0)=\displaystyle\Omega^{(1)}(N_{0},{\bf k}_{0})= 12β𝐪,ωnTrlnG01(𝐪,ωn)\displaystyle\ \frac{1}{2\beta}\sum_{{\bf q},\omega_{n}}\text{Tr}\ln G_{0}^{-1}({\bf q},\omega_{n})
=\displaystyle= 12β𝐪,ωnln[(ωn+i2,𝐪)2+1,𝐪2]\displaystyle\ \frac{1}{2\beta}\sum_{{\bf q},\omega_{n}}\ln\left[(\omega_{n}+i\mathcal{E}_{2,{\bf q}})^{2}+\mathcal{E}_{1,{\bf q}}^{2}\right]
=\displaystyle= 1β𝐪ln[2sinh(βE𝐪/2)],\displaystyle\ \frac{1}{\beta}\sum_{{\bf q}}\ln\left[2\sinh(\beta E_{\bf q}/2)\right], (109)

where G0(𝐪,ωn)G_{0}({\bf q},\omega_{n}) is defined in Eq. (210). In the low-temperature limit β\beta\to\infty, reduces to the zero-point energy of the Bogoliubov quasiparticles Egs(1)=12𝐪(E𝐪ε𝐪)E^{(1)}_{gs}=\frac{1}{2}\sum_{{\bf q}}(E_{\bf q}-\varepsilon^{\prime}_{{\bf q}}) (where we measured related to the zero-point energy of the normal state). In the above, the Matsubara sum was carried out by using

ωnln[(ωn+i2)2+12]=p=±ln[2sinh(1+p22T)],\displaystyle\sum_{\omega_{n}}\ln\left[(\omega_{n}+i\mathcal{E}_{2})^{2}+\mathcal{E}_{1}^{2}\right]=\sum_{p=\pm}\ln\left[2\sinh\left(\frac{\mathcal{E}_{1}+p\mathcal{E}_{2}}{2T}\right)\right], (110)

obtained by integrating (60) over ϵ\epsilon or by Poisson summation formula followed by a contour integral Altland and Simons (2010).

The perturbation theory also gives order-by-order corrections to the condensate momentum, 𝐤0=n𝐤0(n){\bf k}_{0}=\sum_{n}{\bf k}_{0}^{(n)}. This allows us to study the corrections of the thermodynamic potential and its curvature as an expansion at the mean-field minimum: Ω(𝐤0)Ω(0)(𝐤0(0))+Ω(1)(𝐤0(0))\Omega({\bf k}_{0})\approx\Omega^{(0)}({\bf k}_{0}^{(0)})+\Omega^{(1)}({\bf k}_{0}^{(0)}) and 2N0b(𝐤0)2Ω(0)(𝐤0(0))k0(1)+2Ω(1)(𝐤0(0))2N_{0}b_{\perp}({\bf k}_{0})\approx\partial_{\parallel}\partial_{\perp}^{2}\Omega^{(0)}({\bf k}_{0}^{(0)})k_{0}^{(1)}+\partial_{\perp}^{2}\Omega^{(1)}({\bf k}_{0}^{(0)}).999However, the full expression Ω=Ω(0)+Ω(1)\Omega=\Omega^{(0)}+\Omega^{(1)} is not real for any 𝐤0{\bf k}_{0} since the quasiparticle is only guaranteed to be stable when expanding the condensate around the contour minimum 𝐤0(0){\bf k}_{0}^{(0)}. This is an artifact of the perturbation theory arising from b(0)=0b_{\perp}^{(0)}=0, and can be avoided in a self-consistent or renormalization group analysis. Numerical evaluation of the integral (V.1) lifts the degeneracy of the bare dispersion contour, replacing it by six minima with the C6\text{C}_{\text{6}} lattice symmetry. We refer to this as order-by-disorder phenomenon (demonstrated in the context of frustrated magnetism in Bergman et al. (2007); Mulder et al. (2010)). As shown in Fig. 1(b), the helical condensate is stable on the high symmetry points of the contour, the Vertex (BECV\text{BEC}_{\text{V}}) and the Edge (BECE\text{BEC}_{\text{E}}) condensates, depending on the microscopic parameters ρ\rho, Un0Un_{0}, and temperature TT. The phase boundaries are all first-order. We note that there is a reentrant BECV\text{BEC}_{\text{V}} around ρ0.22\rho\approx 0.22, related to the detailed quasiparticle dispersion at small 𝐪{\bf q} characterized by B(ρ)B(\rho) and K(ρ)K(\rho). As detailed in Appendix I, an expansion at the minimum of the thermodynamic potential gives

b(1)=b0fb(Un0/T)\displaystyle b_{\perp}^{(1)}=b_{0\perp}f_{b}(Un_{0}/T) (111)

for small Un0Un_{0}, which gives an effective UVUV cutoff such that 𝐪Bq2+Kq4<Un0\mathcal{E}_{\bf q}\approx Bq_{\parallel}^{2}+Kq_{\perp}^{4}<Un_{0}. In the above,

b0=(Un0)5/4t11/4gb(ρ)\displaystyle b_{0\perp}=(Un_{0})^{5/4}t_{1}^{-1/4}g_{b}(\rho) (112)

with gb(ρ)g_{b}(\rho) a dimensionless 𝒪(1)\mathcal{O}(1) constant and the scaling function

fb(x){1,for x1.1/x,for x1.\displaystyle f_{b}(x)\sim\left\{\begin{array}[]{cc}1,\quad\text{for }x\gg 1.\\ 1/x,\quad\text{for }x\ll 1.\end{array}\right. (115)

With increasing Un0Un_{0}, higher powers terms in 𝐪\mathcal{E}_{\bf q} are important, leading to a deviation from the scaling form (111), see Fig. 13.

As a consequence of the order-by-disorder phenomenon, BB_{\perp} is induced given by B(1)=b(1)B_{\perp}^{(1)}=b_{\perp}^{(1)} as discussed above. As a complementary analysis, below we perform a direct calculation of B(1)B_{\perp}^{(1)}, and show that it reduces to zero in the isotropic, continuum limit.

V.2 BB_{\perp}

To obtain BB_{\perp}, one first needs to calculate the zeroth-order action S(0)[π,ϕ]S^{(0)}[\pi,\phi] based on the lattice model in Sec. III. A shortcut to this is to expand the mean-field ground state energy Egs(0)(𝐤0+𝐪)E_{gs}^{(0)}({\bf k}_{0}+{\bf q}) in 𝐪{\bf q}, and use the relation Blm(0)=blm(0)B_{lm}^{(0)}=b_{lm}^{(0)}. However, this has a shortcoming as it only gives the leading order in 𝐪{\bf q} couplings. Alternatively, one can start with the lattice model, do an expansion in π\pi and ϕ\phi, and then take the continuum limit. The latter approach is more laborious, but gives the full expression that is valid even at large momenta. We relegate the details of this analysis to Appendix H. For simplicity, we consider a long-wavelength form in analogy to (44b) that is valid within the coherence length 𝐪<𝐪c{\bf q}<{\bf q}_{c} in the weakly-interacting limit. By including up to quartic in ϕ\phi terms required for one-loop calculations, and choosing 𝐤0{\bf k}_{0} at one of the C6\text{C}_{\text{6}} symmetric points, the zeroth-order Lagrangian takes the following form (with the superscripts (0)(0) dropped for the simplicity of notation)

ϕn0=\displaystyle\frac{\mathcal{L}_{\phi}}{n_{0}}= Δ[c(ϕ)+c(ϕ)2]+Bτ(τϕ)2+B(ϕ)2+K20(2ϕ)2+K02(2ϕ)2+K11(2ϕ)(2ϕ)\displaystyle\ \Delta\left[c(\partial_{\parallel}\phi)+c^{\perp}(\partial_{\perp}\phi)^{2}\right]+B_{\tau}(\partial_{\tau}\phi)^{2}+B(\partial_{\parallel}\phi)^{2}+K_{20}(\partial^{2}_{\parallel}\phi)^{2}+K_{02}(\partial^{2}_{\perp}\phi)^{2}+K_{11}(\partial_{\parallel}^{2}\phi)(\partial^{2}_{\perp}\phi)
+B30(ϕ)3+B12(ϕ)(ϕ)2+B40(ϕ)4+B04(ϕ)4+B22(ϕ)2(ϕ)2,\displaystyle\ +B_{30}(\partial_{\parallel}\phi)^{3}+B_{12}(\partial_{\parallel}\phi)(\partial_{\perp}\phi)^{2}+B_{40}(\partial_{\parallel}\phi)^{4}+B_{04}(\partial_{\perp}\phi)^{4}+B_{22}(\partial_{\parallel}\phi)^{2}(\partial_{\perp}\phi)^{2}, (116)

where Bτ=1/2Un0B_{\tau}=1/2Un_{0} and Δ𝐤0=1|Γ𝐤¯0|/|Γ𝐤0|\Delta_{{\bf k}_{0}}=1-|\Gamma_{\bar{{\bf k}}_{0}}|/|\Gamma_{{\bf k}_{0}}|, vanishing at k0=k¯0k_{0}=\bar{k}_{0}. All the other parameters are of the form t1g(ρ)t_{1}g(\rho) with g(ρ)g(\rho) a dimensionless 𝒪(1)\mathcal{O}(1) constant while the explicit expressions can be obtained via Blm(0)=blm(0)B_{lm}^{(0)}=b_{lm}^{(0)} along with K20(0)=B40(0)K_{20}^{(0)}=B_{40}^{(0)}, K02(0)=B04(0)K_{02}^{(0)}=B_{04}^{(0)} and K11(0)=B22(0)K_{11}^{(0)}=B_{22}^{(0)}, where the latter relations are found by inspecting the explicit form of the quadratic action, (207), and blm(0)b_{lm}^{(0)} are given in Eqs. (B) and (B) for BECV\text{BEC}_{\text{V}} and BECE\text{BEC}_{\text{E}}, respectively. The linear derivative term (tadpole diagram) ϕ\partial_{\parallel}\phi can be eliminated by choosing the condensate momentum such that k0Ω(𝐤0)=0\nabla_{k_{0}}\Omega({\bf k}_{0})=0, which at mean-field level, gives k0(0)=k¯0k_{0}^{(0)}=\bar{k}_{0}. We note B=0B_{\perp}=0 in the Lagrangian (V.2) even in the presence of C6\text{C}_{\text{6}} lattice (but excluding the zero-point energy) effects. This is consistent with the mean-field ground state energy in Eq. (107) that exhibits a degenerate contour minimum. However, higher-order fluctuations generally generate a nonzero BB_{\perp}. At one-loop order and the weakly-interacting limit, it is given by (see details in Appendix J)

B(1)=\displaystyle B_{\perp}^{(1)}= cB122cq3q2+q2Bτωn2+𝐪+12qB22q2+6B40q2Bτωn2+𝐪B122qq2q2(Bτωn2+𝐪)2\displaystyle\ -\frac{c_{\perp}B_{12}}{2c}\int_{q}\frac{3q_{\parallel}^{2}+q_{\perp}^{2}}{B_{\tau}\omega_{n}^{2}+\mathcal{E}_{\bf q}}+\frac{1}{2}\int_{q}\frac{B_{22}q_{\parallel}^{2}+6B_{40}q_{\perp}^{2}}{B_{\tau}\omega_{n}^{2}+\mathcal{E}_{\bf q}}-B_{12}^{2}\int_{q}\frac{q_{\parallel}^{2}q_{\perp}^{2}}{(B_{\tau}\omega_{n}^{2}+\mathcal{E}_{\bf q})^{2}}
\displaystyle\approx B0fB(Un0/T),\displaystyle\ B_{0\perp}f_{B}(Un_{0}/T), (117)

where the dimensionless scaling function fBf_{B} has the same limits as in Eq. (115) and B0=(Un0)5/4t11/4gB(ρ)B_{0\perp}=(Un_{0})^{5/4}t_{1}^{-1/4}g_{B}(\rho) with gB(ρ)g_{B}(\rho) a dimensionless 𝒪(1)\mathcal{O}(1) constant. In the isotropic limit ρ1/6\rho\to 1/6, K20=K02=K11/2K_{20}=K_{02}=K_{11}/2 and B/k02=B30/k0=B12/k0=4B40=4B04=2B22B/k_{0}^{2}=B_{30}/k_{0}=B_{12}/k_{0}=4B_{40}=4B_{04}=2B_{22}, resulting in B(1)=0B_{\perp}^{(1)}=0. This suggests the robustness of the Lifshitz transition point at ρ=1/6\rho=1/6 at one-loop order and associated relation between harmonic and nonlinear terms enforced by the rotational symmetry of 𝐤0{\bf k}_{0}, also corresponding to the dipole conservation symmetry.

VI Conclusion

In summary, we studied a nonzero-momentum superfluid state as a condensate in a frustrated honeycomb Bose-Hubbard model that features a dispersion minimum contour with a nonzero momentum scale k0k_{0}, set by frustrated hopping. We supplemented the lattice model by a continuum smectic field theory and explored in detail the rich phenomenology that emerges from the helical (single momentum) Bose-condensation on this dispersion minimum contour.

Consistent with generalized Hohenberg-Mermin-Wagner theorem Mermin and Wagner (1966); Hohenberg (1967); Halperin (2019), we found that the helical condensate is stable at zero temperature despite quantum fluctuations that are qualitatively stronger than that of a conventional superfluid. The helical state is spontaneously highly anisotropic in its spectrum of low-energy excitations. We explored in detail the physical properties of such state, such as the equation of state (chemical potential), condensate depletion, momentum distribution, structure factor, of all which are qualitatively distinct from a conventional XY zero-momentum superfluid. Because of the “soft” smectic dispersion, at nonzero temperature, thermal fluctuations that diverge in the absence of lattice effects lead to a vanishing of the helical condensate.

However, somewhat paradoxically, a subtle interplay of lattice effects with quantum and thermal fluctuations leads to stabilization of the helical smectic state, through the phenomenon of order-by-disorder via a crossover to a more conventional stable XY superfluid.

Our study thus predicts a stable helical superfluid state and its rich phenomenology that can emerge in frustrated bosonic systems, characterized by a bare dispersion minimum contour. Such sub-extensive contour degeneracy of the condensate momentum, 𝐤0{\bf k}_{0}, can be realized in cold atom experiments through Floquet engineering of the optical lattice with frustrated hopping Struck et al. (2011); Wintersperger et al. (2020) or synthetic Rashba spin-orbit coupling Zhai (2015). Because the latter approach is not based on frustration, its closed contour minimum dispersion would be spoiled by an optical lattice, reduced to a discrete set of dispersion minima Zhang et al. (2013); Toniolo and Linder (2014); Mæland et al. (2020). It would thereby exhibit a bare BB_{\perp} modulus controlled by the depth of the optical lattice, rather than Feshbach tunable interactions Chin et al. (2010), as considered in our study. Yet, in a shallow optical lattice with a weak BB_{\perp}, within a long crossover scale we expect a nonzero momentum condensate Wu et al. (2011); Barnett et al. (2012); Cui and Zhou (2013) in spin-orbit coupled Bose gases, exhibiting phenomenology predicted in Table 1.

The frustrated bosonic model that we studied is isomorphic to a quantum O(2) easy-plane magnet with frustrated exchange couplings, which is also characterized by a spiral contour as manifold of its classical ground states. The helical superfluid we explored thus maps directly on such a coplanar spin spiral state with the U(1) superfluid phase corresponding to the O(2) direction of the XY spin. Theoretical studies suggest the stability of such spiral states in a quantum easy-plane magnet Liu et al. (2020) and in XY models (for spin-1/2, equivalent to the Bose-Hubbard model at half-filling in the UU\to\infty hard-core limit, in contrast to the weak UU limit considered here) with 𝐤0{\bf k}_{0} at high symmetry points selected by quantum fluctuations (Varney et al., 2011; Di Ciolo et al., 2014; Sedrakyan et al., 2015). This contrasts with the corresponding Heisenberg O(3) model Bergman et al. (2007); Mulder et al. (2010), which at nonzero temperature in 2d is always unstable (even with the usual linear dispersion) to a disordered state due to strongly coupled nonlinear spin wave fluctuations.

Our analysis, however, does not provide the sufficient conditions to realize the helical superfluid discussed here. Other interesting supersolid phases, e.g., Bose condensate at multiple momenta, are also competing ground states that can emerge from the interplay of interactions and the macroscopic degeneracy of the dispersion minimum. Selecting between these is a challenging problem that we have not addressed and likely requires extensive numerical analysis. Cold-atoms systems where our model is realized using Floquet engineering will likely be challenged by heating problem as well as further neighbor hopping and interactions.

Here we neglected the vortices (allowed by the periodicity of the superfluid phase) that can play an important role for the quantum and thermal phase transitions out of the helical superfluid state, sketched at the mean-field level in Fig. 1(a). In 2+1d it is isomorphic to the melting of a quantum smectic (at low energies perturbed by stabilizing modulus BB_{\perp} generated by order-by-disorder) that has been formulated through a dual gauge theory in Ref. Radzihovsky (2020); Zhai and Radzihovsky (2021), but with critical RG analysis remaining as an open problem. We note that, as discussed in Ref. Yan and Reuther (2022) in addition to conventional vortices of the superfluid phase ϕ\phi, thermal excitations of orientational (“momentum”) vortices in 𝐤0{\bf k}_{0} (based on numerical analysis) appear to play an important role. We leave these and other questions discussed above for future theoretical research and hope that the rich phenomenology presented here will stimulate further experimental and theoretical works on frustrated bosonic and magnetic systems.

Acknowledgement

L. R. thanks Arun Paramekanti for discussions. This work is supported by the Simons Investigator Award to L. R. from the James Simons Foundation. L. R. also thanks the KITP for its hospitality as part of the Spring 2022 Pair-Density-Wave Order Rapid Response Workshop, during which part of this work was completed and in part supported by the National Science Foundation under Grant No. NSF PHY-1748958. Research at Perimeter Institute is supported in part by the Government of Canada through the Department of Innovation, Science and Economic Development Canada and by the Province of Ontario through the Ministry of Colleges and Universities.

Appendix A Diagonalization of frustrated tight-binding model on honeycomb lattice

The tight-binding Hamiltonian in Eq. (9) can be written in terms of Pauli matrices H0=𝐤(a𝐤,1,a𝐤,2)h0(a𝐤,1,a𝐤,2)TH_{0}=\sum_{\bf k}(a^{\dagger}_{{\bf k},1},a^{\dagger}_{{\bf k},2})h_{0}(a_{{\bf k},1},a_{{\bf k},2})^{T}, where

h0=t2ϵ𝐤𝟙t1|Γ𝐤|cosθ𝐤σx+t1|Γ𝐤|sinθ𝐤σy,h_{0}=t_{2}\epsilon_{\bf k}\mathbbm{1}-t_{1}|\Gamma_{\bf k}|\cos\theta_{\bf k}~{}\sigma_{x}+t_{1}|\Gamma_{\bf k}|\sin\theta_{\bf k}~{}\sigma_{y}, (118)

with

ϵ𝐤\displaystyle\epsilon_{{\bf k}} =2i(cos𝐤𝐯i),\displaystyle=2\sum_{i}(\cos{\bf k}\cdot{\bf v}_{i}),
Γ𝐤\displaystyle\Gamma_{{\bf k}} =iexp(i𝐤𝐞i),\displaystyle=\sum_{i}\exp(-i{\bf k}\cdot{\bf e}_{i}),
θ𝐤\displaystyle\theta_{\bf k} =Arg(Γ𝐤).\displaystyle=\text{Arg}(\Gamma_{\bf k}). (119)

The Hamiltonian h0h_{0} can be diagonalized by the transformation

U=12(exp(iθ𝐤2)exp(iθ𝐤2)exp(iθ𝐤2)exp(iθ𝐤2)),U=\frac{1}{\sqrt{2}}\biggl{(}\begin{array}[]{cc}\exp(i\frac{\theta_{{\bf k}}}{2})&\exp(i\frac{\theta_{{\bf k}}}{2})\\ -\exp(-i\frac{\theta_{{\bf k}}}{2})&\exp(-i\frac{\theta_{{\bf k}}}{2})\end{array}\biggr{)}, (120)

which gives

Uh0U=t2ϵ𝐤𝟙t1(|Γ𝐤|00|Γ𝐤|),U^{\dagger}h_{0}U=t_{2}\epsilon_{{\bf k}}\mathbbm{1}-t_{1}\biggl{(}\begin{array}[]{cc}-|\Gamma_{{\bf k}}|&0\\ 0&|\Gamma_{{\bf k}}|\end{array}\biggr{)}, (121)

with two bands

ϵ𝐤=t2ϵ𝐤t1|Γ𝐤|,ϵ𝐤+=t2ϵ𝐤+t1|Γ𝐤|.\epsilon^{-}_{{\bf k}}=t_{2}\epsilon_{{\bf k}}-t_{1}|\Gamma_{{\bf k}}|,\quad\quad\epsilon^{+}_{{\bf k}}=t_{2}\epsilon_{{\bf k}}+t_{1}|\Gamma_{{\bf k}}|. (122)

The lower band exhibits a contour minimum for 1/6<ρ=t2/t1<1/21/6<\rho=t_{2}/t_{1}<1/2.

In general, the dispersion minimum contour can appear in a class of Hamiltonians written as

h0=t1T+t2T2,h_{0}=t_{1}T+t_{2}T^{2}, (123)

where t1t_{1} and t2t_{2} are the nearest-neighbor and next-nearest-neighbor hopping amplitudes and TT is a 2×22\times 2 hopping matrix, with the form

T=(0G𝐤G𝐤0)T=\left(\begin{array}[]{cc}0&G_{{\bf k}}\\ G_{{\bf k}}^{\dagger}&0\end{array}\right) (124)

in the sublattice basis. Then the two band are given by

ϵ𝐤±=±|t1||G𝐤|+t2|G𝐤|2\epsilon_{{\bf k}}^{\pm}=\pm|t_{1}||G_{{\bf k}}|+t_{2}|G_{{\bf k}}|^{2} (125)

with the lower energy band ϵ𝐤\epsilon_{{\bf k}}^{-} exhibiting macroscopic ground state degeneracy along the contour defined by |G𝐤|=|t1|/2t2|G_{{\bf k}}|=|t_{1}|/2t_{2}.

Appendix B Expansion of ϵ𝐤\epsilon^{-}_{{\bf k}} around the dispersion minimum contour

In the weakly-interacting regime, the band dispersion around the bose condensate plays an important role for the low-energy properties of the helical superfluid. This is established through the facts: (i) the zeroth-order thermodynamic potential Ω(0)(N0,𝐤0)=N0ϵ𝐤0+N02U/4V\Omega^{(0)}(N_{0},{\bf k}_{0})=N_{0}\epsilon^{-}_{{\bf k}_{0}}+N_{0}^{2}U/4V depending on the free dispersion ϵ𝐤\epsilon^{-}_{{\bf k}} and (ii) The relation between Ω(0)\Omega^{(0)} and the zeroth-order Goldstone mode theory (0)\mathcal{L}^{(0)} as discussed in Sec. V. Expansion on the lower band dispersion (14) up to quartic order gives

ϵ𝐤0+𝐪ϵ𝐤0Δ(cq+cq2)+b(0)q2+b30(0)q3+b12(0)qq2+b40(0)q4+b04(0)q4+b22(0)q2q2,\displaystyle\epsilon^{-}_{{\bf k}_{0}+{\bf q}}-\epsilon^{-}_{{\bf k}_{0}}\approx\Delta(cq_{\parallel}+c_{\perp}q_{\perp}^{2})+b^{(0)}q_{\parallel}^{2}+b_{30}^{(0)}q_{\parallel}^{3}+b_{12}^{(0)}q_{\parallel}q_{\perp}^{2}+b_{40}^{(0)}q_{\parallel}^{4}+b_{04}^{(0)}q_{\perp}^{4}+b_{22}^{(0)}q_{\parallel}^{2}q_{\perp}^{2}, (126)

where for the Vertex condensate (BECV\text{BEC}_{\text{V}}), 𝐤0=(0,k0V){\bf k}_{0}=(0,k_{0V}), |Γ𝐤0|=5+4cos(3k0V/2)|\Gamma_{{\bf k}_{0}}|=\sqrt{5+4\cos(3k_{0V}/2)}, and the coefficients

c=\displaystyle c= 6t2sin(3k0V/2),c=3t2[2+cos(3k0V/2)]2,b(0)=9t2sin2(3k0V/2)|Γ𝐤0V|2,\displaystyle\ 6t_{2}\sin(3k_{0V}/2),\quad c_{\perp}=\frac{3t_{2}[2+\cos(3k_{0V}/2)]}{2},\quad b^{(0)}=\frac{9t_{2}\sin^{2}(3k_{0V}/2)}{|\Gamma_{{\bf k}_{0V}}|^{2}},
b30(0)=\displaystyle b_{30}^{(0)}= 27t2[3+5cos(3k0V/2)+cos(3k0V)]sin(3k0V/2)2|Γ𝐤0|4,b12(0)=9t2[4sin(3k0V/2)+sin(3k0V)]4|Γ𝐤0|2,\displaystyle\ \frac{27t_{2}[3+5\cos(3k_{0V}/2)+\cos(3k_{0V})]\sin(3k_{0V}/2)}{2|\Gamma_{{\bf k}_{0}}|^{4}},\quad b_{12}^{(0)}=\frac{9t_{2}[4\sin(3k_{0V}/2)+\sin(3k_{0V})]}{4|\Gamma_{{\bf k}_{0}}|^{2}},
b40(0)=\displaystyle b_{40}^{(0)}= 27t2[76+190cos(3k0V/2)+163cos(3k0V)+50cos(9k0V/2)+7cos(6k0V)]32|Γ𝐤0|6,b04(0)=9t2[2+cos(3k0V/2)]216|Γ𝐤0|2,\displaystyle\ \frac{27t_{2}[76+190\cos(3k_{0V}/2)+163\cos(3k_{0V})+50\cos(9k_{0V}/2)+7\cos(6k_{0V})]}{32|\Gamma_{{\bf k}_{0}}|^{6}},\quad b_{04}^{(0)}=\frac{9t_{2}[2+\cos(3k_{0V}/2)]^{2}}{16|\Gamma_{{\bf k}_{0}}|^{2}},
b22(0)=\displaystyle b_{22}^{(0)}= 27t2[15+25cos(3k0V/2)+11cos(3k0V)+3cos(9k0V/2)]16|Γ𝐤0|4\displaystyle\ \frac{27t_{2}[15+25\cos(3k_{0V}/2)+11\cos(3k_{0V})+3\cos(9k_{0V}/2)]}{16|\Gamma_{{\bf k}_{0}}|^{4}} (127)

and for the Edge condensate (BECE\text{BEC}_{\text{E}}), 𝐤0=(k0E,0){\bf k}_{0}=(k_{0E},0), |Γ𝐤0|=|1+2cos(3k0E/2)||\Gamma_{{\bf k}_{0}}|=|1+2\cos(\sqrt{3}k_{0E}/2)|, and the coefficients

c=\displaystyle c= 23t2[sin(3k0E/2)+sin(3k0E)],c=9t2cos(3k0E/2)2,b(0)=3t2sin2(3k0E/2),\displaystyle\ 2\sqrt{3}t_{2}[\sin(\sqrt{3}k_{0E}/2)+\sin(\sqrt{3}k_{0E})],\quad c_{\perp}=\frac{9t_{2}\cos(\sqrt{3}k_{0E}/2)}{2},\quad b^{(0)}=3t_{2}\sin^{2}(\sqrt{3}k_{0E}/2),
b30(0)=\displaystyle b_{30}^{(0)}= 33t2sin(3k0E)4,b12(0)=93t2sin(3k0E)4|Γ𝐤0|,b40(0)=3t2[1+7cos(3k0E)]32,\displaystyle\ \frac{3\sqrt{3}t_{2}\sin(\sqrt{3}k_{0E})}{4},\quad b_{12}^{(0)}=\frac{9\sqrt{3}t_{2}\sin(\sqrt{3}k_{0E})}{4|\Gamma_{{\bf k}_{0}}|},\quad b_{40}^{(0)}=\frac{3t_{2}[-1+7\cos(\sqrt{3}k_{0E})]}{32},
b04(0)=\displaystyle b_{04}^{(0)}= 81t2[1+cos(3k0E)]32|Γ𝐤0|2,b22(0)=27t2[1+3cos(3k0E/2)+3cos(3k0E)+cos(33k0E/2)]16|Γ𝐤0|2.\displaystyle\ \frac{81t_{2}[1+\cos(\sqrt{3}k_{0E})]}{32|\Gamma_{{\bf k}_{0}}|^{2}},\quad b_{22}^{(0)}=\frac{27t_{2}[-1+3\cos(\sqrt{3}k_{0E}/2)+3\cos(\sqrt{3}k_{0E})+\cos(3\sqrt{3}k_{0E}/2)]}{16|\Gamma_{{\bf k}_{0}}|^{2}}. (128)

In the above, Δ=1|Γ𝐤¯0|/|Γ𝐤0|\Delta=1-|\Gamma_{\bar{{\bf k}}_{0}}|/|\Gamma_{{\bf k}_{0}}|, which vanishes at 𝐤0=𝐤¯0{\bf k}_{0}=\bar{{\bf k}}_{0}. For 𝐤0𝐤¯0{\bf k}_{0}\neq\bar{{\bf k}}_{0}, there are corrections to all the coefficients above, but here we only give the ones for qq_{\parallel} and q2q_{\perp}^{2} that are important for the one-loop calculations in Sec. V.

Appendix C Bogoliubov approximation to the frustrated Bose-Hubbard model

For a condensate at 𝐤0{\bf k}_{0}, the bosonic operators are given by

a𝐤,1A𝐤0,1δ𝐤,𝐤0+a𝐤0+𝐪,1,a𝐤,2A𝐤0,2δ𝐤,𝐤0+a𝐤0+𝐪,2,a_{{\bf k},1}\approx A_{{\bf k}_{0},1}\delta_{{\bf k},{\bf k}_{0}}+a_{{\bf k}_{0}+{\bf q},1},\ \ \ a_{{\bf k},2}\approx A_{{\bf k}_{0},2}\delta_{{\bf k},{\bf k}_{0}}+a_{{\bf k}_{0}+{\bf q},2}, (129)

where A𝐤0,1=12eiθ𝐤02N0A_{{\bf k}_{0},1}=\frac{1}{\sqrt{2}}e^{i\frac{\theta_{{\bf k}_{0}}}{2}}\sqrt{N_{0}} and A𝐤0,2=12eiθ𝐤02N0A_{{\bf k}_{0},2}=\frac{1}{\sqrt{2}}e^{-i\frac{\theta_{{\bf k}_{0}}}{2}}\sqrt{N_{0}}. We then expand the interacting part of the Hamiltonian, (17), which up to quadratic order of a𝐤0+𝐪a_{{\bf k}_{0}+{\bf q}} and a𝐤0+𝐪a^{\dagger}_{{\bf k}_{0}+{\bf q}} is given by

Hint\displaystyle H_{\text{int}}\approx U2Vs=1,2(|A𝐤0,s|4+4|A𝐤0,s|2𝐪a𝐤0+𝐪,sa𝐤0+𝐪,s+(A𝐤0,s)2𝐪a𝐤0𝐪,sa𝐤0+𝐪,s+A𝐤0,s2𝐪a𝐤0𝐪,sa𝐤0+𝐪,s)\displaystyle\ \frac{U}{2V}\sum_{s=1,2}\left(|A_{{\bf k}_{0},s}|^{4}+4|A_{{\bf k}_{0},s}|^{2}\sum_{{\bf q}}a^{\dagger}_{{\bf k}_{0}+{\bf q},s}a_{{\bf k}_{0}+{\bf q},s}+(A_{{\bf k}_{0},s}^{*})^{2}\sum_{{\bf q}}a_{{\bf k}_{0}-{\bf q},s}a_{{\bf k}_{0}+{\bf q},s}+A_{{\bf k}_{0},s}^{2}\sum_{{\bf q}}a^{\dagger}_{{\bf k}_{0}-{\bf q},s}a^{\dagger}_{{\bf k}_{0}+{\bf q},s}\right)
=\displaystyle= UN024V+UN04V(4𝐪a𝐤0+𝐪,1a𝐤0+𝐪,1+eiθ𝐤0𝐪a𝐤0𝐪,1a𝐤0+𝐪,1+eiθ𝐤0𝐪a𝐤0𝐪,1a𝐤0+𝐪,1)\displaystyle\ \frac{UN_{0}^{2}}{4V}+\frac{UN_{0}}{4V}\biggl{(}4\sum_{{\bf q}}a^{\dagger}_{{\bf k}_{0}+{\bf q},1}a_{{\bf k}_{0}+{\bf q},1}+e^{-i\theta_{{\bf k}_{0}}}\sum_{{\bf q}}a_{{\bf k}_{0}-{\bf q},1}a_{{\bf k}_{0}+{\bf q},1}+e^{i\theta_{{\bf k}_{0}}}\sum_{{\bf q}}a^{\dagger}_{{\bf k}_{0}-{\bf q},1}a^{\dagger}_{{\bf k}_{0}+{\bf q},1}\biggr{)}
+UN04V(4𝐪a𝐤0+𝐪,2a𝐤0+𝐪,2+eiθ𝐤0𝐪a𝐤0𝐪,2a𝐤0+𝐪,2+eiθ𝐤0𝐪a𝐤0𝐪,2a𝐤0+𝐪,2).\displaystyle+\frac{UN_{0}}{4V}\biggl{(}4\sum_{{\bf q}}a^{\dagger}_{{\bf k}_{0}+{\bf q},2}a_{{\bf k}_{0}+{\bf q},2}+e^{i\theta_{{\bf k}_{0}}}\sum_{{\bf q}}a_{{\bf k}_{0}-{\bf q},2}a_{{\bf k}_{0}+{\bf q},2}+e^{-i\theta_{{\bf k}_{0}}}\sum_{{\bf q}}a^{\dagger}_{{\bf k}_{0}-{\bf q},2}a^{\dagger}_{{\bf k}_{0}+{\bf q},2}\biggr{)}. (130)

With a transformation to the band basis

a𝐤,1=12eiθ𝐤2(d𝐤,++d𝐤,),a𝐤,2=12eiθ𝐤2(d𝐤,++d𝐤,),a_{{\bf k},1}=\frac{1}{\sqrt{2}}e^{i\frac{\theta_{{\bf k}}}{2}}(d_{{\bf k},+}+d_{{\bf k},-}),\quad a_{{\bf k},2}=\frac{1}{\sqrt{2}}e^{-i\frac{\theta_{{\bf k}}}{2}}(-d_{{\bf k},+}+d_{{\bf k},-}), (131)

followed by neglecting the upper band contributions, d𝐤,+d_{{\bf k},+}, that only give O(Un0/t1)O(Un_{0}/t_{1}) corrections to the ground state, we obtain the following total Hamiltonian

H=\displaystyle H= ϵ𝐤0N0+UN024V+𝐪ϵ𝐤0+𝐪d𝐤0+𝐪,d𝐤0+𝐪,+UN02V𝐪(d𝐤0+𝐪,d𝐤0+𝐪,+d𝐤0𝐪,d𝐤0𝐪,)\displaystyle\ \epsilon^{-}_{{\bf k}_{0}}N_{0}+\frac{UN_{0}^{2}}{4V}+\sum_{{\bf q}}\epsilon^{-}_{{\bf k}_{0}+{\bf q}}d_{{\bf k}_{0}+{\bf q},-}^{\dagger}d_{{\bf k}_{0}+{\bf q},-}+\frac{UN_{0}}{2V}\sum_{{\bf q}}\left(d_{{\bf k}_{0}+{\bf q},-}^{\dagger}d_{{\bf k}_{0}+{\bf q},-}+d_{{\bf k}_{0}-{\bf q},-}^{\dagger}d_{{\bf k}_{0}-{\bf q},-}\right)
+UN04V𝐪cos(θ𝐤0θ𝐤0+𝐪+θ𝐤0𝐪2)(d𝐤0+𝐪,d𝐤0𝐪,+d𝐤0+𝐪,d𝐤0𝐪,),\displaystyle+\frac{UN_{0}}{4V}\sum_{{\bf q}}\cos\left(\theta_{{\bf k}_{0}}-\frac{\theta_{{\bf k}_{0}+{\bf q}}+\theta_{{\bf k}_{0}-{\bf q}}}{2}\right)\left(d_{{\bf k}_{0}+{\bf q},-}d_{{\bf k}_{0}-{\bf q},-}+d_{{\bf k}_{0}+{\bf q},-}^{\dagger}d_{{\bf k}_{0}-{\bf q},-}^{\dagger}\right), (132)

which gives the Hamiltonian (24) after using the canonical ensemble relation

N=N0+𝐪d𝐤0+𝐪,d𝐤0+𝐪,.N=N_{0}+\sum_{{\bf q}}d_{{\bf k}_{0}+{\bf q},-}^{\dagger}d_{{\bf k}_{0}+{\bf q},-}. (133)

The Hamiltonian (24) then can be diagonalized by the bosonic Bogoliubov transformation

V=(u𝐪v𝐪v𝐪u𝐪),V=\biggl{(}\begin{array}[]{cc}u_{\bf q}&v_{\bf q}\\ v_{-{\bf q}}^{*}&u_{-{\bf q}}^{*}\end{array}\biggr{)}, (134)

where u𝐪u_{\bf q} and v𝐪v_{\bf q} (chosen to be real) are given in Eq. (84) with u𝐪v𝐪=u𝐪2v𝐪2u_{\bf q}v_{\bf q}=-\sqrt{u_{\bf q}^{2}v_{\bf q}^{2}}, and VV is a nonunitary matrix that preserves the bosonic commutation relation after the basis transformation

(d𝐤0+𝐪,d𝐤0𝐪,)=V(α𝐤0+𝐪α𝐤0𝐪).\biggl{(}\begin{array}[]{c}d_{{\bf k}_{0}+{\bf q},-}\\ d^{\dagger}_{{\bf k}_{0}-{\bf q},-}\end{array}\biggr{)}=V\biggl{(}\begin{array}[]{c}\alpha_{{\bf k}_{0}+{\bf q}}\\ \alpha^{\dagger}_{{\bf k}_{0}-{\bf q}}\end{array}\biggr{)}. (135)

The dispersions E𝐪±=E±𝐪E^{\pm}_{\bf q}=E_{\pm\bf q} of α𝐤0±𝐪\alpha_{{\bf k}_{0}\pm{\bf q}} in Eq. (31) can be obtained by solving the determinant equation

|Hλσz|=0|H-\lambda\sigma_{z}|=0 (136)

with the solutions λ=E𝐪+\lambda=E^{+}_{\bf q}, E𝐪-E^{-}_{\bf q}, where E𝐪±>0E^{\pm}_{\bf q}>0.

Appendix D Stability of quantum “smectic” and “columnar” phases in d-dimensions: generalized Hohenberg-Mermin-Wagner theorems

In this appendix, we perform a simple dimensional analysis of the stability of a smectic (columnar) phase in dd spatial dimensions, which exhibits a Goldstone mode with 11 (d1d-1) hard direction(s), \parallel, and the other d1d-1 (11) soft direction(s), \perp. We set U=B=K=2n0=1U=B=K=2n_{0}=1 for simplicity.

For the smectic phase, the quantum fluctuations at T=0T=0 is characterized by

ϕ2Q=dω2πdqdd1q(2π)d1ω2+q2+(q2)2.\langle\phi^{2}\rangle_{Q}=\int\frac{d\omega}{2\pi}\frac{dq_{\parallel}d^{d-1}q_{\perp}}{(2\pi)^{d}}\frac{1}{\omega^{2}+q_{\parallel}^{2}+(q_{\perp}^{2})^{2}}. (137)

With the change of variables q2=yq_{\perp}^{2}=y_{\perp}, the above equation becomes

ϕ2Q\displaystyle\langle\phi^{2}\rangle_{Q} =dωdqdq(2π)d+1qd2ω2+q2+q4\displaystyle=\int\frac{d\omega dq_{\parallel}dq_{\perp}}{(2\pi)^{d+1}}\frac{q_{\perp}^{d-2}}{\omega^{2}+q_{\parallel}^{2}+q_{\perp}^{4}} (138)
dωdqdy(2π)d+1yd32ω2+q2+y2.\displaystyle\propto\int\frac{d\omega dq_{\parallel}dy_{\perp}}{(2\pi)^{d+1}}\frac{y_{\perp}^{\frac{d-3}{2}}}{\omega^{2}+q_{\parallel}^{2}+y_{\perp}^{2}}.

The stability of this state requires the convergence in the IR, which requires

3+d32>2d>1.3+\frac{d-3}{2}>2\Rightarrow d>1. (139)

For the columnar phase, the quantum fluctuations are

ϕ2Q\displaystyle\langle\phi^{2}\rangle_{Q} =dωdd1qdq(2π)d+11ω2+q2+q4\displaystyle=\int\frac{d\omega d^{d-1}q_{\parallel}dq_{\perp}}{(2\pi)^{d+1}}\frac{1}{\omega^{2}+q_{\parallel}^{2}+q_{\perp}^{4}}
dωdqdy(2π)d+1qd2y1/2ω2+q2+y2,\displaystyle\propto\int\frac{d\omega dq_{\parallel}dy_{\perp}}{(2\pi)^{d+1}}\frac{q_{\parallel}^{d-2}y^{-1/2}_{\perp}}{\omega^{2}+q_{\parallel}^{2}+y_{\perp}^{2}}, (140)

which requires

3+(d2)12>2d>323+(d-2)-\frac{1}{2}>2\Rightarrow d>\frac{3}{2} (141)

for the stability. Therefore, for the physical dimension of our interest, d=2d=2, both requirements are satisfied, which suggests the helical superfluid stable under quantum fluctuations.

At nonzero temperature, we consider the dominant classical contributions at ωn=0\omega_{n}=0, which for the smectic are given by

ϕ2T\displaystyle\langle\phi^{2}\rangle_{T} =dqdd1q(2π)d1q2+(q2)2\displaystyle=\int\frac{dq_{\parallel}d^{d-1}q_{\perp}}{(2\pi)^{d}}\frac{1}{q_{\parallel}^{2}+(q_{\perp}^{2})^{2}}
dqdy(2π)d+1yd22y1/2q2+y2.\displaystyle\propto\int\frac{dq_{\parallel}dy_{\perp}}{(2\pi)^{d+1}}\frac{y_{\perp}^{\frac{d-2}{2}}y^{-1/2}_{\perp}}{q_{\parallel}^{2}+y_{\perp}^{2}}. (142)

Its stability requires

2+d32>2d>3.2+\frac{d-3}{2}>2\Rightarrow d>3. (143)

While in the columnar phase,

ϕ2T\displaystyle\langle\phi^{2}\rangle_{T} =dd1qdq(2π)d1q2+q4\displaystyle=\int\frac{d^{d-1}q_{\parallel}dq_{\perp}}{(2\pi)^{d}}\frac{1}{q_{\parallel}^{2}+q_{\perp}^{4}}
dqdy(2π)d+1qd2y1/2q2+y2,\displaystyle\propto\int\frac{dq_{\parallel}dy_{\perp}}{(2\pi)^{d+1}}\frac{q_{\parallel}^{d-2}y^{-1/2}_{\perp}}{q_{\parallel}^{2}+y_{\perp}^{2}}, (144)

which requires

2+d212>2d>522+d-2-\frac{1}{2}>2\Rightarrow d>\frac{5}{2} (145)

Thus, thermal fluctuations make the 2d smectic/columnar phase unstable.

Appendix E Calculation details of depletion

The condensate depletion is given by Eq. (82) at weak interactions as discussed in the main text. At zero temperature, β=\beta=\infty leads to

nd=dqdq(2π)212(𝐪+Un𝐪2+2Un𝐪1),n_{d}=\int_{-\infty}^{\infty}\frac{dq_{\parallel}dq_{\perp}}{(2\pi)^{2}}\frac{1}{2}\biggl{(}\frac{\mathcal{E}_{{\bf q}}+Un}{\sqrt{\mathcal{E}^{2}_{{\bf q}}+2Un\mathcal{E}_{\bf q}}}-1\biggr{)}, (146)

where 𝐪Bq2+Kq4\mathcal{E}_{{\bf q}}\approx Bq_{\parallel}^{2}+Kq_{\perp}^{4} as defined in Eq. (45). In this section, we perform the integral exactly. With the change of variables:

σ=B2Unq,σ=K2Unq2,σ2=σ2+σ2,\displaystyle\sigma_{\parallel}=\sqrt{\frac{B}{2Un}}q_{\parallel},\hskip 2.27621pt\sigma_{\perp}=\sqrt{\frac{K}{2Un}}q_{\perp}^{2},\hskip 2.27621pt\sigma^{2}=\sigma_{\parallel}^{2}+\sigma_{\perp}^{2}, (147)

the integral can be rewritten as

I\displaystyle I =dqdq(2π)212(𝐪+Un𝐪2+2Un𝐪1)\displaystyle=\int_{-\infty}^{\infty}\frac{dq_{\parallel}dq_{\perp}}{(2\pi)^{2}}\frac{1}{2}\biggl{(}\frac{\mathcal{E}_{{\bf q}}+Un}{\sqrt{\mathcal{E}^{2}_{{\bf q}}+2Un\mathcal{E}_{{\bf q}}}}-1\biggr{)}
=(2Un)3/4B1/2K1/4dσdσ16π21|σ|(2+1σ2+σ221+1σ2+σ21)\displaystyle=\frac{(2Un)^{3/4}}{B^{1/2}K^{1/4}}\int_{-\infty}^{\infty}\frac{d\sigma_{\parallel}d\sigma_{\perp}}{16\pi^{2}}\frac{1}{\sqrt{|\sigma_{\perp}|}}\biggl{(}\frac{2+\frac{1}{\sigma_{\parallel}^{2}+\sigma_{\perp}^{2}}}{2\sqrt{1+\frac{1}{\sigma_{\parallel}^{2}+\sigma_{\perp}^{2}}}}-1\biggr{)}
=(2Un)34B12K140σdσ16π202πdθ|sinθ|(2σ2+12σσ2+11)\displaystyle=\frac{(2Un)^{\frac{3}{4}}}{B^{\frac{1}{2}}K^{\frac{1}{4}}}\int_{0}^{\infty}\frac{\sqrt{\sigma}d\sigma}{16\pi^{2}}\int_{0}^{2\pi}\frac{d\theta}{\sqrt{|\sin\theta|}}\biggl{(}\frac{2\sigma^{2}+1}{2\sigma\sqrt{\sigma^{2}+1}}-1\biggr{)}
=(2Un)3/416π2B1/2K1/4IθIσ,\displaystyle=\frac{(2Un)^{3/4}}{16\pi^{2}B^{1/2}K^{1/4}}I_{\theta}I_{\sigma}, (148)

where Iθ=02πdθ|sinθ|10.4882I_{\theta}=\int_{0}^{2\pi}\frac{d\theta}{\sqrt{|\sin\theta|}}\approx 10.4882 and

Iσ\displaystyle I_{\sigma} =0𝑑σ(2σ2+12σ3+σσ)\displaystyle=\int_{0}^{\infty}d\sigma\biggl{(}\frac{2\sigma^{2}+1}{2\sqrt{\sigma^{3}+\sigma}}-\sqrt{\sigma}\biggr{)}
=23σ3/2+23σ+σ3+fe(σ)6|0\displaystyle=\left.-\frac{2}{3}\sigma^{3/2}+\frac{2}{3}\sqrt{\sigma+\sigma^{3}}+\frac{f_{e}(\sigma)}{6}\right|_{0}^{\infty}
=fe(0)6=|fe(0)|6\displaystyle=-\frac{f_{e}(0)}{6}=\frac{|f_{e}(0)|}{6} (149)

with

fe(σ)=2(1)1/4EllipticF[iArcSinh(1)1/4σ,1].f_{e}(\sigma)=2(-1)^{1/4}\textrm{EllipticF}\left[i\textrm{ArcSinh}\frac{(-1)^{1/4}}{\sqrt{\sigma}},-1\right]. (150)

Consequently, the integral is

I=(2Un)3/416π2B1/2K1/4Iθ|fe(0)|6,I=\frac{(2Un)^{3/4}}{16\pi^{2}B^{1/2}K^{1/4}}I_{\theta}\frac{|f_{e}(0)|}{6}, (151)

which gives the depletion

ndn=(U3nB2K)1/423/4Iθ|fe(0)|96π2.\frac{n_{d}}{n}=\biggl{(}\frac{U^{3}}{nB^{2}K}\biggr{)}^{1/4}\frac{2^{3/4}I_{\theta}|f_{e}(0)|}{96\pi^{2}}. (152)

Similarly, we generalize the study of depletion to be in d=l+md=l+m dimensions. The system possesses dispersion hard along ll directions and soft along mm directions. With q();iq_{\parallel(\perp);i} being the momentum along the ii-th hard (soft) direction, the generalized dispersion is

𝐪(l,m)=i=1mBiq;i2+j=1mKjq;j4.\displaystyle\mathcal{E}^{(l,m)}_{\bf q}=\sum_{i=1}^{m}B_{i}q_{\parallel;i}^{2}+\sum_{j=1}^{m}K_{j}q_{\perp;j}^{4}. (153)

In terms of variables

σ;i\displaystyle\sigma_{\parallel;i} =\displaystyle= Bi2Unq;i,σ;i=Ki2Unq;i2,\displaystyle\sqrt{\frac{B_{i}}{2Un}}q_{\parallel;i},\hskip 5.69054pt\sigma_{\perp;i}=\sqrt{\frac{K_{i}}{2Un}}q_{\perp;i}^{2},
σ2\displaystyle\sigma^{2} =\displaystyle= ilσ;i2+jmσ;j2,\displaystyle\sum_{i}^{l}\sigma_{\parallel;i}^{2}+\sum_{j}^{m}\sigma_{\perp;j}^{2}, (154)

the depletion is given by

nd(l,m)\displaystyle n_{d}^{(l,m)} =dlqdmq(2π)212(𝐪(l,m)+Un(𝐪(l,m))2+2Un𝐪(l,m)1).\displaystyle=\int_{-\infty}^{\infty}\frac{d^{l}q_{\parallel}d^{m}q_{\perp}}{(2\pi)^{2}}\frac{1}{2}\biggl{(}\frac{\mathcal{E}_{{\bf q}}^{(l,m)}+Un}{\sqrt{(\mathcal{E}^{(l,m)}_{{\bf q}})^{2}+2Un\mathcal{E}_{\bf q}^{(l,m)}}}-1\biggr{)}. (155)

The integral can be proceeded as

I(l,m)\displaystyle I^{(l,m)} =(2Un)(2l+m)/42(ilBijmKj)1/2dlσdmσ(2π)m+l\displaystyle=\frac{(2Un)^{(2l+m)/4}}{2(\prod_{i}^{l}B_{i}\prod_{j}^{m}K_{j})^{1/2}}\int_{-\infty}^{\infty}\frac{d^{l}\sigma_{\parallel}d^{m}\sigma_{\perp}}{(2\pi)^{m+l}} (156)
×jmKj1/42|σ,j|1/2(2σ2+12σ4+σ21).\displaystyle\times\prod_{j}^{m}\frac{K_{j}^{1/4}}{2|\sigma_{\perp,j}|^{1/2}}\biggl{(}\frac{2\sigma^{2}+1}{2\sqrt{\sigma^{4}+\sigma^{2}}}-1\biggr{)}. (157)

Particularly, when Bi=BB_{i}=B and Kj=KK_{j}=K for any ii and jj,

I(l,m)(Un)(2l+m)/4Bl/2Km/4.I^{(l,m)}\propto\frac{(Un)^{(2l+m)/4}}{B^{l/2}K^{m/4}}. (158)

The depletion is then given by

nd(l,m)n(Un142l+m)(2l+m)/4Bl/2Km/4.\frac{n_{d}^{(l,m)}}{n}\propto\frac{(Un^{1-\frac{4}{2l+m}})^{(2l+m)/4}}{B^{l/2}K^{m/4}}. (159)

Now we discuss the thermal corrections to the depletion, which is given by

δnd(T)=\displaystyle\delta n_{d}(T)= nd(T)nd(0)=ddq(2π)du𝐪2+v𝐪2eβE𝐪1\displaystyle n_{d}(T)-n_{d}(0)=\int_{-\infty}^{\infty}\frac{d^{d}q}{(2\pi)^{d}}\frac{u_{\bf q}^{2}+v_{\bf q}^{2}}{e^{\beta E_{\bf q}}-1} (160)

with E𝐪E_{\bf q} and 1,𝐪\mathcal{E}_{1,\bf q} defined in Eq. (31), and u𝐪2u_{\bf q}^{2} and v𝐪2v_{\bf q}^{2} defined in Eq. (84). In the long-wavelength limit, E𝐪1,𝐪2Un𝐪(l,m)E_{\bf q}\approx\mathcal{E}_{1,\bf q}\approx\sqrt{2Un\mathcal{E}_{\bf q}^{(l,m)}}. Then, with the change of variables σ,i=β2BUn0q,i\sigma_{\parallel,i}=\beta\sqrt{2BUn_{0}}q_{\parallel,i} and σ,j=β2KUn0q,j2\sigma_{\perp,j}=\beta\sqrt{2KUn_{0}}q_{\perp,j}^{2} (set Bi=BB_{i}=B and Kj=KK_{j}=K for simplicity),

δnd(T)=\displaystyle\delta n_{d}(T)= ddq(2π)d𝐪(l,m)+Un1,𝐪1eβE𝐪1\displaystyle\ \int_{-\infty}^{\infty}\frac{d^{d}q}{(2\pi)^{d}}\frac{\mathcal{E}_{\bf q}^{(l,m)}+Un}{\mathcal{E}_{1,\bf q}}\frac{1}{e^{\beta E_{\bf q}}-1}
\displaystyle\approx dlqdmq(2π)l+mUn2Un𝐪(l,m)1eβ2Un𝐪(l,m)1\displaystyle\ \int_{-\infty}^{\infty}\frac{d^{l}q_{\parallel}d^{m}q_{\perp}}{(2\pi)^{l+m}}\frac{Un}{\sqrt{2Un\mathcal{E}_{\bf q}^{(l,m)}}}\frac{1}{e^{\beta\sqrt{2Un\mathcal{E}_{\bf q}^{(l,m)}}}-1}
\displaystyle\propto 1(2π)l+m2mT(T22BUn0)l/2(T22KUn0)m/4\displaystyle\ \frac{1}{(2\pi)^{l+m}2^{m}T}\biggl{(}\frac{T^{2}}{2BUn_{0}}\biggr{)}^{l/2}\biggl{(}\frac{T^{2}}{2KUn_{0}}\biggr{)}^{m/4}
×0dσσl+m/22eσ1,\displaystyle\times\int_{0}^{\infty}d\sigma\frac{\sigma^{l+m/2-2}}{e^{\sigma}-1}, (161)

where the radial integral over σ\sigma is finite if l+m/22>0l+m/2-2>0. This suggests the superfluid is stable when

2l+m>4.2l+m>4. (162)

For a smectic (columnar) phase with l=1l=1 (l=d1l=d-1) and m=d1m=d-1 (m=1m=1), we reproduce the stability condition d>3d>3 (d>5/2d>5/2) as obtained in Appendix D.

Appendix F Superflow

In this section, we derive the expression of the supercurrent. Notice that there involves higher-order derivatives in the Lagrangian 0\mathcal{L}_{0} in Eq. (41), its equation of motion is thus modified as

0Φμ(0(μΦ))+ij(0(ijΦ))=0;\displaystyle\frac{\partial\mathcal{L}_{0}}{\partial\Phi}-\partial_{\mu}\Big{(}\frac{\partial\mathcal{L}_{0}}{\partial(\partial_{\mu}\Phi)}\Big{)}+\partial_{i}\partial_{j}\Big{(}\frac{\partial\mathcal{L}_{0}}{\partial(\partial_{i}\partial_{j}\Phi)}\Big{)}=0; (163)

where μ=(τ,i,j,k,)\mu=(\tau,i,j,k,\dots) and i,j,k,i,j,k,\dots are spatial indices. With the detailed form of 0\mathcal{L}_{0}, We obtain

τΦ+J(4Φ+2k022Φ)=0\displaystyle-\partial_{\tau}{\Phi}^{\ast}+J(\nabla^{4}{\Phi}^{\ast}+2k_{0}^{2}\nabla^{2}{\Phi}^{\ast})=0 (164)

along with a similar equation for Φ{\Phi}. In addition, the Lagrangian 0\mathcal{L}_{0} respects the global U(1) symmetry. Therefore, an infinitesimal transformation ΦeiεΦΦ+iεΦ\Phi\rightarrow e^{i\varepsilon}\Phi\approx\Phi+i\varepsilon\Phi (and ΦeiεΦΦiεΦ{\Phi}^{\ast}\rightarrow e^{-i\varepsilon}{\Phi}^{\ast}\approx{\Phi}^{\ast}-i\varepsilon{\Phi}^{\ast}, where ε1\varepsilon\ll 1) results in 00+εΔ0\mathcal{L}_{0}\rightarrow\mathcal{L}_{0}+\varepsilon\Delta\mathcal{L}_{0} with Δ0\Delta\mathcal{L}_{0} being a total derivative, given by

Δ0=\displaystyle\Delta\mathcal{L}_{0}= 0Φ(iΦ)+(0(μΦ))(iμΦ)\displaystyle\ \frac{\partial\mathcal{L}_{0}}{\partial\Phi}(i\Phi)+\Big{(}\frac{\partial\mathcal{L}_{0}}{\partial(\partial_{\mu}\Phi)}\Big{)}(i\partial_{\mu}\Phi)
+(0(ijΦ))(iijΦ)+ΦΦ\displaystyle+\Big{(}\frac{\partial\mathcal{L}_{0}}{\partial(\partial_{i}\partial_{j}\Phi)}\Big{)}(i\partial_{i}\partial_{j}\Phi)+\Phi\leftrightarrow{\Phi}^{\ast}
=\displaystyle= μ(0(μΦ)(iΦ))+i(0(ijΦ)(ijΦ))\displaystyle\ \partial_{\mu}\Big{(}\frac{\partial\mathcal{L}_{0}}{\partial(\partial_{\mu}\Phi)}(i\Phi)\Big{)}+\partial_{i}\Big{(}\frac{\partial\mathcal{L}_{0}}{\partial(\partial_{i}\partial_{j}\Phi)}(i\partial_{j}\Phi)\Big{)}
+ΦΦ,\displaystyle+\Phi\leftrightarrow{\Phi}^{\ast},

where we applied the equation of motion in Eq. (163) to get the second equality. Then, the Noether current is

j0=\displaystyle j^{0}= 0(τΦ)(iΦ)+0(τΦ)(iΦ)=2iΦΦ\displaystyle\ \frac{\partial\mathcal{L}_{0}}{\partial(\partial_{\tau}\Phi)}(i\Phi)+\frac{\partial\mathcal{L}_{0}}{\partial(\partial_{\tau}{\Phi}^{\ast})}(-i{\Phi}^{\ast})=2i{\Phi}^{\ast}\Phi
ji=\displaystyle j_{i}= 0(iΦ)(iΦ)+0(iΦ)(iΦ)\displaystyle\ \frac{\partial\mathcal{L}_{0}}{\partial(\partial_{i}\Phi)}(i\Phi)+\frac{\partial\mathcal{L}_{0}}{\partial(\partial_{i}{\Phi}^{\ast})}(-i{\Phi}^{\ast})
+0(ijΦ)(ijΦ)+0(ijΦ)(ijΦ)\displaystyle+\frac{\partial\mathcal{L}_{0}}{\partial(\partial_{i}\partial_{j}\Phi)}(i\partial_{j}\Phi)+\frac{\partial\mathcal{L}_{0}}{\partial(\partial_{i}\partial_{j}{\Phi}^{\ast})}(-i\partial_{j}{\Phi}^{\ast})
=\displaystyle= iJΦ(i2Φ+2k02iΦ)+iJΦ(i2Φ+2k¯02iΦ)\displaystyle\ -iJ\Phi(\partial_{i}\nabla^{2}{\Phi}^{\ast}+2k_{0}^{2}\partial_{i}{\Phi}^{\ast})+iJ{\Phi}^{\ast}(\partial_{i}\nabla^{2}\Phi+2\bar{k}_{0}^{2}\partial_{i}\Phi)
+iJ2ΦiΦiJ2ΦiΦ,\displaystyle+iJ\nabla^{2}{\Phi}^{\ast}\partial_{i}\Phi-iJ\nabla^{2}\Phi\partial_{i}{\Phi}^{\ast}, (165)

which satisfy the continuity equation, μjμ=0\partial_{\mu}j^{\mu}=0.

With Φ=nei𝐤0𝐫+iϕ\Phi=\sqrt{n}e^{i{\bf k}_{0}\cdot{\bf r}+i\phi} (and at mean-field level k0=k¯0k_{0}=\bar{k}_{0}), the supercurrent is then a current in space,

𝐣s=\displaystyle{\bf j}_{s}= 4Jn[(ϕ)2+2𝐤0ϕ1n2n+34n2(n)2]\displaystyle\ 4Jn\left[(\nabla\phi)^{2}+2{\bf k}_{0}\cdot\nabla\phi-\frac{1}{n}\nabla^{2}n+\frac{3}{4n^{2}}(\nabla n)^{2}\right]
×(𝐤0+ϕ)2Jn2ϕ2Jn3ϕ\displaystyle\times({\bf k}_{0}+\nabla\phi)-2J\nabla n\nabla^{2}\phi-2Jn\nabla^{3}\phi
\displaystyle\approx 4Jn[(ϕ)2+2𝐤0ϕ](𝐤0+ϕ)2Jn3ϕ,\displaystyle\ 4Jn\left[(\nabla\phi)^{2}+2{\bf k}_{0}\cdot\nabla\phi\right]({\bf k}_{0}+\nabla\phi)-2Jn\nabla^{3}\phi, (166)

generated through the twisting of the superfluid phase ϕ\phi. In the last line, we consider the long-wavelength limit, where the fluctuation of density is small such that n0\nabla n\approx 0.

The supercurrent density can be written as 𝐣s=n𝐯s{\bf j}_{s}=n{\bf v}_{s}, where 𝐯s{\bf v}_{s} the superfluid velocity. For a linear in 𝐫{\bf r} classical phase variation ϕ=𝐪𝐫\phi={\bf q}\cdot{\bf r}, the superfluid velocity reduces to the derivative of the bare dispersion

𝐯s=\displaystyle{\bf v}_{s}= 4J[q2+2𝐤0𝐪](𝐤0+𝐪)\displaystyle\ 4J\left[q^{2}+2{\bf k}_{0}\cdot{\bf q}\right]({\bf k}_{0}+{\bf q})
=\displaystyle= kε𝐤|𝐤=𝐤0+𝐪,\displaystyle\ \nabla_{k}\varepsilon_{\bf k}|_{{\bf k}={\bf k}_{0}+{\bf q}}, (167)

where

ε𝐤=J(k2k02)2+ε0,𝐤=𝐤0+𝐪.\varepsilon_{\bf k}=J(k^{2}-k_{0}^{2})^{2}+\varepsilon_{0},\quad{\bf k}={\bf k}_{0}+{\bf q}. (168)

We emphasize that the existence of the nonlinear ϕ\phi terms in Eq. (166) makes this relation hold.

Alternatively, the supercurrent can be obtained as a response to a background probe gauge field. We first generalize the long-wavelength harmonic Goldstone mode Hamiltonian density [with the corresponding Lagrangian density (44b)],

0ϕ=4Jnk02(ϕ)2+Jn(2ϕ)2,\displaystyle\mathcal{H}_{0\phi}=4Jnk_{0}^{2}(\nabla_{\parallel}\phi)^{2}+Jn(\nabla^{2}\phi)^{2}, (169)

to the quartic order in ϕ\phi by requiring the rotational symmetry of 𝐤0{\bf k}_{0}, giving

ϕ=4Jn[k0ϕ+12(ϕ)2]2+Jn(2ϕ)2.\displaystyle\mathcal{H}_{\phi}=4Jn\left[k_{0}\nabla_{\parallel}\phi+\frac{1}{2}(\nabla\phi)^{2}\right]^{2}+Jn(\nabla^{2}\phi)^{2}. (170)

As the system coupled with a background U(1) gauge field, the Hamiltonian density is modified as ϕ[ϕ]ϕ[ϕ+𝐀]\mathcal{H}_{\phi}[\nabla\phi]\to\mathcal{H}_{\phi}[\nabla\phi+{\bf A}]. The supercurrent density is then obtained by taking derivative with respect to the gauge field

𝐣s=\displaystyle{\bf j}_{s}= Hϕ𝐀|𝐀=0\displaystyle\ \left.\frac{\partial H_{\phi}}{\partial{\bf A}}\right|_{{\bf A}=0}
=\displaystyle= 4Jn[(ϕ)2+2k0ϕ](𝐤0+ϕ)2Jn3ϕ,\displaystyle\ 4Jn\left[(\nabla\phi)^{2}+2k_{0}\nabla_{\parallel}\phi\right]({\bf k}_{0}+\nabla\phi)-2Jn\nabla^{3}\phi, (171)

which reproduces the long-wavelength expression in Eq. (166).

Appendix G Calculation details of the chemical potential

In this appendix, we evaluate the integral in Eq. (IV.2.5). As in Appendix E, we use the dimensionless variables defined in Eq. (147), and written integral as

I\displaystyle I =dqdq(2π)2(1𝐪𝐪2+2Un𝐪)\displaystyle=\int_{-\infty}^{\infty}\frac{dq_{\parallel}dq_{\perp}}{(2\pi)^{2}}\biggl{(}1-\frac{\mathcal{E}_{{\bf q}}}{\sqrt{\mathcal{E}_{{\bf q}}^{2}+2Un\mathcal{E}_{{\bf q}}}}\biggr{)}
=dqdq(2π)2(111+2Un𝐪)\displaystyle=\int_{-\infty}^{\infty}\frac{dq_{\parallel}dq_{\perp}}{(2\pi)^{2}}\biggl{(}1-\frac{1}{\sqrt{1+\frac{2Un}{\mathcal{E}_{{\bf q}}}}}\biggr{)}
=dσdσ8π2(2Un)3/4B1/2K1/41|σ|(1σσ2+1)\displaystyle=\int_{-\infty}^{\infty}\frac{d\sigma_{\parallel}d\sigma_{\perp}}{8\pi^{2}}\frac{(2Un)^{3/4}}{B^{1/2}K^{1/4}}\frac{1}{\sqrt{|\sigma_{\perp}|}}\biggl{(}1-\frac{\sigma}{\sqrt{\sigma^{2}+1}}\biggr{)}
=(2Un)3/48π2B1/2K1/4IθIσ,\displaystyle=\frac{(2Un)^{3/4}}{8\pi^{2}B^{1/2}K^{1/4}}I_{\theta}I^{\prime}_{\sigma}, (172)

where σ2=σ2+σ2\sigma^{2}=\sigma_{\parallel}^{2}+\sigma_{\perp}^{2}, IθI_{\theta} is defined and computed in Eq. (148), and

Iσ\displaystyle I^{\prime}_{\sigma} =\displaystyle= 0σ𝑑σ(1σσ2+1)\displaystyle\int_{0}^{\infty}\sqrt{\sigma}d\sigma\biggl{(}1-\frac{\sigma}{\sqrt{\sigma^{2}+1}}\biggr{)} (173)
=\displaystyle= (2σ3(σ1+σ2)+13fe(σ))|0\displaystyle\left.\biggl{(}\frac{2\sqrt{\sigma}}{3}(\sigma-\sqrt{1+\sigma^{2}})+\frac{1}{3}f_{e}(\sigma)\biggr{)}\right|_{0}^{\infty}
=\displaystyle= 13(fe()fe(0))1.23605,\displaystyle\frac{1}{3}(f_{e}(\infty)-f_{e}(0))\approx 1.23605,

with fe(σ)f_{e}(\sigma) is in Eq. (150). We note that fe()limσfe(σ)=0f_{e}(\infty)\equiv\lim_{\sigma\rightarrow\infty}f_{e}(\sigma)=0 and fe(0)limσ0fe(σ)3.70815f_{e}(0)\equiv\lim_{\sigma\rightarrow 0}f_{e}(\sigma)\approx-3.70815, which lead to the last approximation in Eq. (173).

As a result, the chemical potential for the helical superfluid in 2d2d is

μU=n[123/4IθIσ32π2(U3B2Kn)1/4].\frac{\mu}{U}=n\biggl{[}1-\frac{2^{3/4}I_{\theta}I^{\prime}_{\sigma}}{32\pi^{2}}\biggl{(}\frac{U^{3}}{B^{2}Kn}\biggr{)}^{1/4}\biggr{]}. (174)

In general, for a dd dimensional system with a dispersion 𝐪(l,m)\mathcal{E}_{\bf q}^{(l,m)} defined in Eq. (153), the integral in Eq. (172) can be generalized as

I(l,m)=dlqdmq(2π)m+l(1𝐪(l,m)(𝐪(l,m))2+2Un𝐪(l,m)).I^{(l,m)}=\int_{-\infty}^{\infty}\frac{d^{l}q_{\parallel}d^{m}q_{\perp}}{(2\pi)^{m+l}}\biggl{(}1-\frac{\mathcal{E}^{(l,m)}_{{\bf q}}}{\sqrt{(\mathcal{E}^{(l,m)}_{{\bf q}})^{2}+2Un\mathcal{E}^{(l,m)}_{{\bf q}}}}\biggr{)}. (175)

In terms of the dimensionless variables defined in Eq. (154), the integral becomes

I(l,m)=\displaystyle I^{(l,m)}= dlσdmσ(2π)m+l(2Un)(2l+m)/4(ilBijmKj)1/2\displaystyle\int_{-\infty}^{\infty}\frac{d^{l}\sigma_{\parallel}d^{m}\sigma_{\perp}}{(2\pi)^{m+l}}\frac{(2Un)^{(2l+m)/4}}{(\prod_{i}^{l}B_{i}\prod_{j}^{m}K_{j})^{1/2}}
×jmKj1/42|σ,j|1/2(1σσ2+1).\displaystyle\times\prod_{j}^{m}\frac{K_{j}^{1/4}}{2|\sigma_{\perp,j}|^{1/2}}\biggl{(}1-\frac{\sigma}{\sqrt{\sigma^{2}+1}}\biggr{)}. (176)

For the special case Bi=BB_{i}=B and Kj=KK_{j}=K for any ii and jj, we have

I(l,m)(2Un)(2l+m)/4Bl/2Km/4.I^{(l,m)}\propto\frac{(2Un)^{(2l+m)/4}}{B^{l/2}K^{m/4}}. (177)

Accordingly, in terms of the dimensionless quantity

𝒬l,m=(Un142l+m)(2l+m)/4Bl/2Km/4,\mathcal{Q}^{l,m}=\frac{(Un^{1-\frac{4}{2l+m}})^{(2l+m)/4}}{B^{l/2}K^{m/4}}, (178)

the chemical potential is

μ(l,m)U=n[1l,m𝒬l,m],\frac{\mu^{(l,m)}}{U}=n\left[1-\mathcal{I}^{l,m}\mathcal{Q}^{l,m}\right], (179)

where l,m\mathcal{I}^{l,m} is a dimensionless 𝒪(1)\mathcal{O}(1) constant fully determined by ll and mm. For a dd-dimensional smectic phase with l=1l=1 and m=d1m=d-1, the dimensionless quantity

𝒬smectic1,d1=(Ud+1nd3B2Kd1)1/4.\mathcal{Q}_{\rm smectic}^{1,d-1}=\biggl{(}\frac{U^{d+1}n^{d-3}}{B^{2}K^{d-1}}\biggr{)}^{1/4}. (180)

We can check that with dim[Un]=E\textrm{dim}\left[Un\right]=E, dim[n]=Ld\textrm{dim}[n]=L^{-d}, dim[B]=EL2\textrm{dim}\left[B\right]=EL^{2}, and dim[K]=EL4\textrm{dim}\left[K\right]=EL^{4} (EE, LL with dimensions of energy and length, respectively), 𝒬smectic1,d1\mathcal{Q}_{\rm smectic}^{1,d-1} is in fact dimensionless. On the other hand, in a columnar phase with l=d1l=d-1 and m=1m=1, the dimensionless quantity is given by

𝒬columnard1,1=(U2d1n2d5B2d2K)1/4.\mathcal{Q}_{\rm columnar}^{d-1,1}=\biggl{(}\frac{U^{2d-1}n^{2d-5}}{B^{2d-2}K}\biggr{)}^{1/4}. (181)

Appendix H ϕ\phi theory with C6\text{C}_{\text{6}} lattice effects

We start with the coherent state path integral formalism of the frustrated Bose-Hubbard model introduced in the beginning of Sec. V. In the density-phase representation (V), the Lagrangian (including the constant part) can be written as

L=ii,a=1,2πi,aτϕi,a+H0+Hint,\displaystyle L=i\sum_{i,a=1,2}\pi_{i,a}\partial_{\tau}\phi_{i,a}+H_{0}+H_{\text{int}}, (182)

where the kinetic part

H0=\displaystyle H_{0}= 2t1n0i,j1+πi,1n0+πj,2n0+πi,1πj,2n02cos[θ𝐤0+𝐤0(𝐫i,1𝐫j,2)+ϕi,1ϕj,2]\displaystyle\ -2t_{1}n_{0}\sum_{\langle i,j\rangle}\sqrt{1+\frac{\pi_{i,1}}{n_{0}}+\frac{\pi_{j,2}}{n_{0}}+\frac{\pi_{i,1}\pi_{j,2}}{n_{0}^{2}}}\cos\left[\theta_{{\bf k}_{0}}+{\bf k}_{0}\cdot({\bf r}_{i,1}-{\bf r}_{j,2})+\phi_{i,1}-\phi_{j,2}\right]
+2t2n0a=1,2i,j1+πi,an0+πj,an0+πi,aπj,an02cos[𝐤0(𝐫i,a𝐫j,a)+ϕi,aϕj,a]\displaystyle\ +2t_{2}n_{0}\sum_{a=1,2}\sum_{\langle\langle i,j\rangle\rangle}\sqrt{1+\frac{\pi_{i,a}}{n_{0}}+\frac{\pi_{j,a}}{n_{0}}+\frac{\pi_{i,a}\pi_{j,a}}{n_{0}^{2}}}\cos\left[{\bf k}_{0}\cdot({\bf r}_{i,a}-{\bf r}_{j,a})+\phi_{i,a}-\phi_{j,a}\right] (183)

and the interaction

Hint=U2i[(n0+πi,1)2+(n0+πi,2)2].\displaystyle H_{\text{int}}=\frac{U}{2}\sum_{i}[(n_{0}+\pi_{i,1})^{2}+(n_{0}+\pi_{i,2})^{2}]. (184)

For a bose condensate, we assume the fluctuations of π\pi and ϕ\phi fields are small, and thereby expand in a power series, which up to quadratic order in π\pi and ϕ\phi is given by

H0=\displaystyle H^{\prime}_{0}= t1n0i,jeiθ𝐤0+i𝐤0(𝐫i,1𝐫j,2)[1+πi,12n0+πj,22n0+i(ϕi,1ϕj,2)18n02(πi,1πj,2)212(ϕi,1ϕj,2)2\displaystyle\ -t_{1}n_{0}\sum_{\langle i,j\rangle}e^{i\theta_{{\bf k}_{0}}+i{\bf k}_{0}\cdot({\bf r}_{i,1}-{\bf r}_{j,2})}\left[1+\frac{\pi_{i,1}}{2n_{0}}+\frac{\pi_{j,2}}{2n_{0}}+i(\phi_{i,1}-\phi_{j,2})-\frac{1}{8n_{0}^{2}}(\pi_{i,1}-\pi_{j,2})^{2}-\frac{1}{2}(\phi_{i,1}-\phi_{j,2})^{2}\right.
i(πi,1+πj,2)2n0(ϕi,1ϕj,2)]+t2n0ai,jei𝐤0(𝐫i,a𝐫j,a)[1+πi,a2n0+πj,a2n0+i(ϕi,aϕj,a)\displaystyle\ \left.-i\frac{(\pi_{i,1}+\pi_{j,2})}{2n_{0}}(\phi_{i,1}-\phi_{j,2})\right]+t_{2}n_{0}\sum_{a}\sum_{\langle\langle i,j\rangle\rangle}e^{i{\bf k}_{0}\cdot({\bf r}_{i,a}-{\bf r}_{j,a})}\left[1+\frac{\pi_{i,a}}{2n_{0}}+\frac{\pi_{j,a}}{2n_{0}}+i(\phi_{i,a}-\phi_{j,a})\right.
18n02(πi,aπj,a)212(ϕi,aϕj,a)2i(πi,a+πj,a)2n0(ϕi,aϕj,a)]+U4i,a(n0+πi,a)2+h.c..\displaystyle\ \left.-\frac{1}{8n_{0}^{2}}(\pi_{i,a}-\pi_{j,a})^{2}-\frac{1}{2}(\phi_{i,a}-\phi_{j,a})^{2}-i\frac{(\pi_{i,a}+\pi_{j,a})}{2n_{0}}(\phi_{i,a}-\phi_{j,a})\right]+\frac{U}{4}\sum_{i,a}(n_{0}+\pi_{i,a})^{2}+\text{h.c.}. (185)

To perform one-loop calculations, we include the cubic and quartic terms of ϕ\phi, given by

H1=\displaystyle H^{\prime}_{1}= t1n06i,jeiθ𝐤0+i𝐤0(𝐫i,1𝐫j,2)[i(ϕi,1ϕj,2)3+14(ϕi,1ϕj,2)4]\displaystyle\ -\frac{t_{1}n_{0}}{6}\sum_{\langle i,j\rangle}e^{i\theta_{{\bf k}_{0}}+i{\bf k}_{0}\cdot({\bf r}_{i,1}-{\bf r}_{j,2})}\left[-i(\phi_{i,1}-\phi_{j,2})^{3}+\frac{1}{4}(\phi_{i,1}-\phi_{j,2})^{4}\right]
+t2n06ai,jei𝐤0(𝐫i,a𝐫j,a)[i(ϕi,aϕj,a)3+14(ϕi,aϕj,a)4]++h.c.\displaystyle+\frac{t_{2}n_{0}}{6}\sum_{a}\sum_{\langle\langle i,j\rangle\rangle}e^{i{\bf k}_{0}\cdot({\bf r}_{i,a}-{\bf r}_{j,a})}\left[-i(\phi_{i,a}-\phi_{j,a})^{3}+\frac{1}{4}(\phi_{i,a}-\phi_{j,a})^{4}\right]+...+\text{h.c.} (186)

At this point, one can either directly take the continuum limit in the real space or go to momentum space to get an effective action. The calculation for the former is simpler but misses some contributions even at long length scale, while the later is tedious but valid for all momentum. We discuss both methods below.

H.1 Taking continuum limit in the real space

To take the continuum limit, we consider the fields ϕi,α{\phi}_{i,\alpha} and πi,α{\pi}_{i,\alpha} on lattice sites as slowly-varying functions of continuous spacetime coordinates (𝐫,τ)({\bf r},\tau) compared to the lattice constant. The time coordinate is already continuous. Therefore, the difference of two operators at distinct spatial locations can be expanded as

ϕ(𝐫i,a)ϕ(𝐫j,a)\displaystyle{\phi}({\bf r}_{i,a})-{\phi}({\bf r}_{j,a^{\prime}}) =n=11n!(Δ𝐫)nϕ|𝐫=𝐫j,a,\displaystyle=\sum_{n=1}^{\infty}\frac{1}{n!}(\Delta{\bf r}\cdot\nabla)^{n}{\phi}|_{{\bf r}={\bf r}_{j,a^{\prime}}}, (187)

where Δ𝐫=𝐫i,a𝐫j,a\Delta{\bf r}={\bf r}_{i,a}-{\bf r}_{j,a^{\prime}} with aa and aa^{\prime} label the sublattices. Below we focus on the case 𝐤0=(0,k0){\bf k}_{0}=(0,k_{0}) while other cases can be derived following the same procedure. Including up to the second (first) order for the quadratic (cubic and quartic) term(s), the Hamiltonian density, defined by 233𝐫=H0+H1\frac{2}{3\sqrt{3}}\int_{\bf r}\mathcal{H}^{\prime}=H^{\prime}_{0}+H^{\prime}_{1}, is given by

n0\displaystyle\frac{\mathcal{H}^{\prime}}{n_{0}}\approx Δ[c(ϕ)+c(ϕ)2]+B(ϕ)2+K20(2ϕ)2+K02(2ϕ)2+K11(2ϕ)(2ϕ)\displaystyle\ \Delta\left[c(\partial_{\parallel}\phi)+c_{\perp}(\partial_{\perp}\phi)^{2}\right]+B(\partial_{\parallel}\phi)^{2}+K_{20}(\partial^{2}_{\parallel}\phi)^{2}+K_{02}(\partial^{2}_{\perp}\phi)^{2}+K_{11}(\partial^{2}_{\parallel}\phi)(\partial^{2}_{\perp}\phi)
+B30(ϕ)3+B12(ϕ)(ϕ)2+B40(ϕ)4+B04(ϕ)4+B22(ϕ)2(ϕ)2+U2n0π2,\displaystyle\ +B_{30}(\partial_{\parallel}\phi)^{3}+B_{12}(\partial_{\parallel}\phi)(\partial_{\perp}\phi)^{2}+B_{40}(\partial_{\parallel}\phi)^{4}+B_{04}(\partial_{\perp}\phi)^{4}+B_{22}(\partial_{\parallel}\phi)^{2}(\partial_{\perp}\phi)^{2}+\frac{U}{2n_{0}}\pi^{2}, (188)

where Δ𝐤0=1|Γ𝐤¯0|/|Γ𝐤0|\Delta_{{\bf k}_{0}}=1-|\Gamma_{\bar{{\bf k}}_{0}}|/|\Gamma_{{\bf k}_{0}}| with 𝐤¯0\bar{\bf k}_{0} satisfying Eq. (16),

c=12t2sin(3k0/2)\displaystyle c=12t_{2}\sin(3k_{0}/2) ,c=3t2[2+cos(3k0/2)],B=4t2[1cos(3k0/2)],B30=B12=3t2sin(3k0/2),\displaystyle,\quad c_{\perp}=3t_{2}[2+\cos(3k_{0}/2)],\quad B=4t_{2}[1-\cos(3k_{0}/2)],\quad B_{30}=B_{12}=3t_{2}\sin(3k_{0}/2),
K20=\displaystyle K_{20}= t224[5+32cos(3k0/2)],K02=9t28,K11=3t24[1+4cos(3k0/2)],\displaystyle\ \frac{t_{2}}{24}[-5+32\cos(3k_{0}/2)],\quad K_{02}=\frac{9t_{2}}{8},\quad K_{11}=\frac{3t_{2}}{4}[-1+4\cos(3k_{0}/2)],
B40=\displaystyle B_{40}= t224[5+32cos(3k0/2)],B04=9t28,B22=3t24[1+4cos(3k0/2)],\displaystyle\ \frac{t_{2}}{24}[-5+32\cos(3k_{0}/2)],\quad B_{04}=\frac{9t_{2}}{8},\quad B_{22}=\frac{3t_{2}}{4}[-1+4\cos(3k_{0}/2)], (189)

and the constant and linear in π\pi terms are dropped. In the above, the ρ\rho dependence is hidden in k0(ρ)k_{0}(\rho) and t2=t1ρt_{2}=t_{1}\rho. By integrating over π\pi, we obtain the zeroth-order Lagrangian density, (V.2), with the parameters (H.1), however, not satisfying the relation Blm(0)=blm(0)B_{lm}^{(0)}=b_{lm}^{(0)}. This inconsistency is due to the θ𝐤0\theta_{{\bf k}_{0}} factor in Eq. (H), or more generally the two-band nature of the model, which invalids the simple continuum limit above that neglects the spatial variation between the two sublattice sites within a unit cell, as discussed in more details below.

H.2 Continuous effective theory in the momentum space

Alternatively, We can Fourier transform the Hamiltonian H0{H}^{\prime}_{0} and H1{H}^{\prime}_{1} followed by a change of basis defined as

(π𝐪,1π𝐪,2ϕ𝐪,1ϕ𝐪,2)=\displaystyle\left(\begin{array}[]{cccc}\pi_{{\bf q},1}\\ \pi_{{\bf q},2}\\ \phi_{{\bf q},1}\\ \phi_{{\bf q},2}\end{array}\right)= 12(U𝐪+U𝐪in0(U𝐪U𝐪)i1n0(U𝐪U𝐪)U𝐪+U𝐪)(π𝐪,+π𝐪,ϕ𝐪,+ϕ𝐪,),\displaystyle\ \frac{1}{2}\left(\begin{array}[]{cc}U_{{\bf q}}+U_{-{\bf q}}^{*}&in_{0}(U_{{\bf q}}-U_{-{\bf q}}^{*})\\ -i\frac{1}{n_{0}}(U_{{\bf q}}-U_{-{\bf q}}^{*})&U_{{\bf q}}+U_{-{\bf q}}^{*}\end{array}\right)\left(\begin{array}[]{cccc}\pi_{{\bf q},+}\\ \pi_{{\bf q},-}\\ \phi_{{\bf q},+}\\ \phi_{{\bf q},-}\end{array}\right), (200)

where the subscripts 1,21,2 label the two sublattices and ±\pm denote the two bands as in the main text. The unitary matrix above is given by

U𝐪=\displaystyle U_{\bf q}= 12(exp(iθ𝐤0+𝐪θ𝐤02)exp(iθ𝐤0+𝐪θ𝐤02)exp(iθ𝐤0+𝐪θ𝐤02)exp(iθ𝐤0+𝐪θ𝐤02)).\displaystyle\ \frac{1}{\sqrt{2}}\biggl{(}\begin{array}[]{cc}\exp(i\frac{\theta_{{\bf k}_{0}+{\bf q}}-\theta_{{\bf k}_{0}}}{2})&\exp(i\frac{\theta_{{\bf k}_{0}+{\bf q}}-\theta_{{\bf k}_{0}}}{2})\\ -\exp(-i\frac{\theta_{{\bf k}_{0}+{\bf q}}-\theta_{{\bf k}_{0}}}{2})&\exp(-i\frac{\theta_{{\bf k}_{0}+{\bf q}}-\theta_{{\bf k}_{0}}}{2})\end{array}\biggr{)}. (203)

At quadratic order, the action for the lower band is given by

S0=12q(πq,ϕq,)G01(q,ωn)(πq,ϕq,),\displaystyle S_{0}=\frac{1}{2}\sum_{q}\left(\begin{array}[]{cc}\pi_{-q,-}&\phi_{-q,-}\end{array}\right)G_{0}^{-1}(q,\omega_{n})\left(\begin{array}[]{cc}\pi_{q,-}\\ \phi_{q,-}\end{array}\right), (207)

where

G01(q,ωn)=(𝐪/(2n0)+U1+cosΘ𝐪2ωn+i𝐪,2ωni𝐪,22n0𝐪+4n02U1cosΘ𝐪2).\displaystyle G_{0}^{-1}(q,\omega_{n})=\left(\begin{array}[]{cc}\mathcal{E}_{\bf q}/(2n_{0})+U\frac{1+\cos\Theta_{\bf q}}{2}&\omega_{n}+i\mathcal{E}_{{\bf q},2}\\ -\omega_{n}-i\mathcal{E}_{{\bf q},2}&2n_{0}\mathcal{E}_{\bf q}+4n_{0}^{2}U\frac{1-\cos\Theta_{\bf q}}{2}\end{array}\right). (210)

In the above, we consider the weakly-interacting limit and can thereby drop the upper band contributions which only give O(Un0/t1)O(Un_{0}/t_{1}) corrections to the low energy physics. Before computing higher-order terms, we have two comments in order.

Firstly, we note the quadratic action reproduces the same Bogoliubov quasiparticle spectrum in Eq. (31). Technically, both the density-phase representation (V) and the Bogoliubov approximation (19) share the same classical background field, which in real space is given by n0e±iθ𝐤02+i𝐤0𝐫i,1(2)\sqrt{n_{0}}e^{\pm i\frac{\theta_{{\bf k}_{0}}}{2}+i{\bf k}_{0}\cdot{\bf r}_{i,1(2)}} with the ++ and - factors for the sites in the 11 and 22 sublattices respectively. The fluctuations around the condensate are given by π\pi and ϕ\phi in Eq. (V) or dd_{-} in Eq. (19). An expansion of the former representation gives linear terms of π\pi and ϕ\phi that takes the same form as the real and imaginary parts of dd_{-} in the latter representation, and thus both give the same form of the quadratic action. While the higher-order terms which modifies the S0S_{0} in Eq. (207) do not have a simple relation compared with higher-order terms which corrects Eq. (24), they should describe the same physics.

Secondly, as shown above, an accurate calculation gives BB and KK identical with the ones from Bogoliubov theory in Eq. (III.2). In Appendix H.1, fields on the two sites within a unit cell are considered to be the same, i.e., ϕi,1ϕi,2ϕi\phi_{i,1}\approx\phi_{i,2}\approx\phi_{i}. This is equivalent to approximate the transformation in Eq. (200) with U𝐪U0U_{{\bf q}}\approx U_{0}, which gives ϕ𝐪,1ϕ𝐪,212ϕ𝐪,\phi_{{\bf q},1}\approx\phi_{{\bf q},2}\approx\frac{1}{\sqrt{2}}\phi_{{\bf q},-}. Therefore, the discrepancy between BB and KK in (H.1) and the accurate ones in Eq. (B) comes from the definition of the fields, which by inspection gives a factor of 1/21/2 via the relation ϕ𝐪=12ϕ𝐪,\phi_{{\bf q}}=\frac{1}{\sqrt{2}}\phi_{{\bf q},-}, and the difference between U𝐪U_{{\bf q}} and U0U_{0}. The latter can be neglected in the limit k00k_{0}\to 0, where the difference is of higher-order in 𝐪{\bf q}.

The calculations for the nonlinear terms are quite tedious. Especially, it is hard to simplify the expressions after the transformation in Eq. (200). Here, we employ the same approximation ϕ𝐪,1ϕ𝐪,2ϕ𝐪\phi_{{\bf q},1}\approx\phi_{{\bf q},2}\approx\phi_{{\bf q}} to derive below the cubic and quartic terms in ϕ\phi.

S1\displaystyle S_{1}\approx in0V𝐪,𝐪Γ3(𝐪)ϕ𝐪ϕ𝐪ϕ𝐪𝐪+n012V{𝐪}[4Γ4(𝐪1)+3Γ4(𝐪1+𝐪2)]×ϕ𝐪1ϕ𝐪2ϕ𝐪3ϕ𝐪1𝐪2𝐪3,\displaystyle\ -\frac{in_{0}}{\sqrt{V}}\sum_{{\bf q},{\bf q}^{\prime}}\Gamma_{3}({\bf q})\phi_{{\bf q}}\phi_{{\bf q}^{\prime}}\phi_{-{\bf q}-{\bf q}^{\prime}}+\frac{n_{0}}{12V}\sum_{\{{\bf q}\}}\left[-4\Gamma_{4}({\bf q}_{1})+3\Gamma_{4}({\bf q}_{1}+{\bf q}_{2})\right]\times\phi_{{\bf q}_{1}}\phi_{{\bf q}_{2}}\phi_{{\bf q}_{3}}\phi_{-{\bf q}_{1}-{\bf q}_{2}-{\bf q}_{3}}, (211)

where

Γ3(𝐪)=\displaystyle\Gamma_{3}({\bf q})= t1Re[(Γ𝐤0+𝐪Γ𝐤0𝐪)eiϕ𝐤0+4ραsin(𝐤0𝐯α)sin(𝐪𝐯α)]\displaystyle\ -t_{1}\text{Re}\biggl{[}(\Gamma_{{\bf k}_{0}+{\bf q}}-\Gamma_{{\bf k}_{0}-{\bf q}})e^{-i\phi_{{\bf k}_{0}}}+4\rho\sum_{\alpha}\sin\left({\bf k}_{0}\cdot{\bf v}_{\alpha}\right)\sin({\bf q}\cdot{\bf v}_{\alpha})\biggr{]}
Γ4(𝐪)=\displaystyle\Gamma_{4}({\bf q})= 2t1Re[(Γ𝐤0+𝐪+Γ𝐤0𝐪2Γ𝐤0)eiϕ𝐤02ραcos(𝐤0𝐯α)[1cos(𝐪𝐯α)]],\displaystyle\ -2t_{1}\text{Re}\biggl{[}\biggl{(}\frac{\Gamma_{{\bf k}_{0}+{\bf q}}+\Gamma_{{\bf k}_{0}-{\bf q}}}{2}-\Gamma_{{\bf k}_{0}}\biggr{)}e^{-i\phi_{{\bf k}_{0}}}-2\rho\sum_{\alpha}\cos\left({\bf k}_{0}\cdot{\bf v}_{\alpha}\right)\left[1-\cos\left({\bf q}\cdot{\bf v}_{\alpha}\right)\right]\biggr{]}, (212)

where an expansion of Γ3\Gamma_{3} and Γ4\Gamma_{4} in small 𝐪{\bf q} together with a Fourier transform back to the real space reproduce the nonlinear terms in the field theory (H.1) with the same coefficients obtained earlier (H.1).

Appendix I One-loop calculation of bb_{\perp}

Here we calculate the perpendicular curvature b=12N0k2Ω(𝐤0)b_{\perp}=\frac{1}{2N_{0}}\partial_{k_{\perp}}^{2}\Omega({\bf k}_{0}) at one-loop order. By expanding the condensate momentum 𝐤0=𝐤0(0)+𝐤0(1){\bf k}_{0}={\bf k}_{0}^{(0)}+{\bf k}_{0}^{(1)} around its mean-field value 𝐤0(0){\bf k}_{0}^{(0)} and keeping the leading-order terms, we get

b(1)=12N0kk2Ω(0)(𝐤0(0))k0(1)+12N0k2Ω(1)(𝐤0(0)),k0(1)=kΩ(1)(𝐤0(0))k2Ω(0)(𝐤0(0)),\displaystyle b_{\perp}^{(1)}=\frac{1}{2N_{0}}\partial_{k_{\parallel}}\partial_{k_{\perp}}^{2}\Omega^{(0)}\left({\bf k}_{0}^{(0)}\right)k_{0}^{(1)}+\frac{1}{2N_{0}}\partial_{k_{\perp}}^{2}\Omega^{(1)}\left({\bf k}_{0}^{(0)}\right),\quad k_{0}^{(1)}=-\frac{\partial_{k_{\parallel}}\Omega^{(1)}\left({\bf k}_{0}^{(0)}\right)}{\partial_{k_{\parallel}}^{2}\Omega^{(0)}\left({\bf k}_{0}^{(0)}\right)}, (213)

where k0(1)k_{0}^{(1)} can be determined by 𝐤0Ω(𝐤0)=0\nabla_{{\bf k}_{0}}\Omega({\bf k}_{0})=0 and

kΩ(1)(𝐤0(0))=\displaystyle\partial_{k_{\parallel}}\Omega^{(1)}\left({\bf k}_{0}^{(0)}\right)= 12𝐪coth(βE𝐪/2)kE𝐪\displaystyle\ \frac{1}{2}\sum_{\bf q}\coth\left(\beta E_{\bf q}/2\right)\partial_{k_{\parallel}}E_{\bf q}
k2Ω(1)(𝐤0(0))=\displaystyle\partial_{k_{\perp}}^{2}\Omega^{(1)}\left({\bf k}_{0}^{(0)}\right)= 𝐪[coth(βE𝐪/2)k2E𝐪β2sinh2(βE𝐪/2)(kE𝐪)2].\displaystyle\ \sum_{\bf q}\left[\coth\left(\beta E_{\bf q}/2\right)\partial_{k_{\perp}}^{2}E_{\bf q}-\frac{\beta}{2\sinh^{2}\left(\beta E_{\bf q}/2\right)}(\partial_{k_{\perp}}E_{\bf q})^{2}\right]. (214)

In the absence of interaction, kE𝐪=qE𝐪\partial_{k_{\parallel}}E_{\bf q}=\partial_{q_{\parallel}}E_{\bf q} and k2E𝐪=q2E𝐪\partial_{k_{\perp}}^{2}E_{\bf q}=\partial_{q_{\perp}}^{2}E_{\bf q}. Accordingly, the integrands in kΩ(1)(𝐤0(0))\partial_{k_{\parallel}}\Omega^{(1)}\left({\bf k}_{0}^{(0)}\right) and k2Ω(1)(𝐤0(0))\partial_{k_{\perp}}^{2}\Omega^{(1)}\left({\bf k}_{0}^{(0)}\right) are total derivatives, which makes kΩ(1)(𝐤0(0))=k2Ω(1)(𝐤0(0))=0\partial_{k_{\parallel}}\Omega^{(1)}\left({\bf k}_{0}^{(0)}\right)=\partial_{k_{\perp}}^{2}\Omega^{(1)}\left({\bf k}_{0}^{(0)}\right)=0 and subsequently b(1)=0b_{\perp}^{(1)}=0. In the presence of weak interaction, the dispersion deviates from the noninteracting form within a crossover scale 𝐪c=(qc,qc)=(ξ1,(λξ)1/2){\bf q}_{c}=(q_{\parallel}^{c},q_{\perp}^{c})=(\xi^{-1},(\lambda\xi)^{-1/2}). Therefore, the integrals (I) are dominated by the region 𝐪<𝐪c{\bf q}<{\bf q}_{c}. In this small UU limit, we obtain b(1)=b0fb(Un0/T)b_{\perp}^{(1)}=b_{0\perp}f_{b}(Un_{0}/T) by using change of variables q=ξqq^{\prime}_{\parallel}=\xi q_{\parallel} and q=(λξ)1/2qq^{\prime}_{\perp}=(\lambda\xi)^{1/2}q_{\perp} in Eq. (I), where b0=(Un0)5/4t11/4gb(ρ)b_{0\perp}=(Un_{0})^{5/4}t_{1}^{-1/4}g_{b}(\rho) with gb(ρ)g_{b}(\rho) a dimensionless 𝒪(1)\mathcal{O}(1) constant and the scaling function fbf_{b} is defined in Eq. (115).

Appendix J One-loop calculation of BB_{\perp}

A complete one-loop calculation should give the perpendicular stiffness of ϕ\phi equals to the corresponding curvature in the thermodynamic potential, i.e., B(1)=b(1)B_{\perp}^{(1)}=b_{\perp}^{(1)}, with b(1)b_{\perp}^{(1)} being calculated in Appendix I. Here, we instead perform an approximate calculation of B(1)B_{\perp}^{(1)} in the long-wavelength limit, where the Lagrangian (V.2) is valid. Specifically, we consider the leading derivative terms and thereby only keep Uπ2U\pi^{2} for the π\pi field, which is enabled by an UU-dependent UV cutoff ΛU\Lambda_{U} defined by 𝐪<𝒪(1)×Un0\mathcal{E}_{\bf q}<\mathcal{O}(1)\times Un_{0}, which in the weakly-interacting limit reduces to 𝐪<𝒪(1)×𝐪c{\bf q}<\mathcal{O}(1)\times{\bf q}_{c}. Although such calculation neglects the range of integration 𝐪>𝐪c{\bf q}>{\bf q}_{c}, the contribution to BB_{\perp} in this range is actually very small as seen in the calculation of b(1)b_{\perp}^{(1)} in Appendix I that takes into account all momentum. Therefore, the integral of B(1)B_{\perp}^{(1)} or b(1)b_{\perp}^{(1)} actually self regulates at 𝐪c{\bf q}_{c} in the weakly-interacting limit.

J.1 Long-wavelength limit

We consider the low-energy ϕ\phi-only theory (V.2), where the bare Green function is given by (kB=1k_{B}=1)

G0,ϕ(q)=\displaystyle G_{0,\phi}(q)= ϕqϕq0=12n01Bτωn2+𝐪,\displaystyle\ \langle\phi_{-q}\phi_{q}\rangle_{0}=\frac{1}{2n_{0}}\frac{1}{B_{\tau}\omega_{n}^{2}+\mathcal{E}_{\bf q}}, (215)

where 𝐪=Bq2+K20q4+K02q4+K11q2q2\mathcal{E}_{\bf q}=Bq_{\parallel}^{2}+K_{20}q_{\parallel}^{4}+K_{02}q_{\perp}^{4}+K_{11}q_{\parallel}^{2}q_{\perp}^{2} and Bτ=1/2Un0B_{\tau}=1/2Un_{0}.

We first calculate the coefficient of ϕ\partial_{\parallel}\phi, which corresponds to the slope of the thermodynamic potential at 𝐤0{\bf k}_{0} along the \parallel direction. At one-loop order, it is given by

+=\displaystyle\leavevmode\hbox to6.67pt{\vbox to8.3pt{\pgfpicture\makeatletter\hbox{\hskip 3.33301pt\lower 10.89337pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ } {{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{0.0pt}{14.22638pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{{{}}}{{}}{}{}{}{}{}{}{}{}{}{\pgfsys@moveto{2.12134pt}{17.07182pt}\pgfsys@curveto{2.12134pt}{18.24341pt}{1.17159pt}{19.19316pt}{0.0pt}{19.19316pt}\pgfsys@curveto{-1.17159pt}{19.19316pt}{-2.12134pt}{18.24341pt}{-2.12134pt}{17.07182pt}\pgfsys@curveto{-2.12134pt}{15.90024pt}{-1.17159pt}{14.95049pt}{0.0pt}{14.95049pt}\pgfsys@curveto{1.17159pt}{14.95049pt}{2.12134pt}{15.90024pt}{2.12134pt}{17.07182pt}\pgfsys@closepath\pgfsys@moveto{0.0pt}{17.07182pt}\pgfsys@fill\pgfsys@invoke{ } }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{0.0pt}{17.07182pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{{ {}{}{}}}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}+\leavevmode\hbox to17.47pt{\vbox to19.39pt{\pgfpicture\makeatletter\hbox{\hskip 8.7359pt\lower 14.95049pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ } {{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{0.0pt}{22.76228pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{{{}}}{{}}{}{}{}{}{}{}{}{}{}{\pgfsys@moveto{2.12134pt}{17.07182pt}\pgfsys@curveto{2.12134pt}{18.24341pt}{1.17159pt}{19.19316pt}{0.0pt}{19.19316pt}\pgfsys@curveto{-1.17159pt}{19.19316pt}{-2.12134pt}{18.24341pt}{-2.12134pt}{17.07182pt}\pgfsys@curveto{-2.12134pt}{15.90024pt}{-1.17159pt}{14.95049pt}{0.0pt}{14.95049pt}\pgfsys@curveto{1.17159pt}{14.95049pt}{2.12134pt}{15.90024pt}{2.12134pt}{17.07182pt}\pgfsys@closepath\pgfsys@moveto{0.0pt}{17.07182pt}\pgfsys@fill\pgfsys@invoke{ } }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{0.0pt}{17.07182pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {}{}{{}}{}{} {}{}{} {}{{}{}{}}{}{}{}{{}{}{}}{}{}{}{{}{}{}}{{}{}{}}{}{}\pgfsys@moveto{-8.5359pt}{25.6073pt}\pgfsys@curveto{-8.5359pt}{30.34454pt}{-4.73735pt}{34.14322pt}{0.0pt}{34.14322pt}\pgfsys@curveto{4.73735pt}{34.14322pt}{8.5359pt}{30.34454pt}{8.5359pt}{25.6073pt}\pgfsys@curveto{8.5359pt}{20.87006pt}{4.73735pt}{17.07182pt}{0.0pt}{17.07182pt}\pgfsys@curveto{-4.73735pt}{17.07182pt}{-8.5359pt}{20.87006pt}{-8.5359pt}{25.6073pt}\pgfsys@closepath\pgfsys@stroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{{ {}{}{}}}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}= n0CΔ𝐤0+B122qΛU3q2+q2Bτωn2+𝐪,\displaystyle\ n_{0}C\Delta_{{\bf k}_{0}}+\frac{B_{12}}{2}\int_{q}^{\Lambda_{U}}\frac{3q_{\parallel}^{2}+q_{\perp}^{2}}{B_{\tau}\omega_{n}^{2}+\mathcal{E}_{\bf q}}, (216)

where the first term is the zeroth-order contribution that vanishes at k0=k¯0k_{0}=\bar{k}_{0} and q=1βV𝐪,ωn\int_{q}=\frac{1}{\beta V}\sum_{{\bf q},\omega_{n}}. With the one-loop correction, 𝐤0{\bf k}_{0} needs to be shifted such that the coefficient of ϕ\partial_{\parallel}\phi is zero again, which assures 𝐤0{\bf k}_{0} is located at the local minimum of the loop-corrected thermodynamic potential with vanished first derivative. In the above, the tadpole diagrams are dropped as they cancel out up to all orders when choosing the correct k0k_{0}.

Next we calculate BB_{\perp}, which at one-loop order is given by

B(1)=\displaystyle B_{\perp}^{(1)}= ++\displaystyle\ \leavevmode\hbox to6.67pt{\vbox to8.3pt{\pgfpicture\makeatletter\hbox{\hskip 3.33301pt\lower 10.89337pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ } {{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{0.0pt}{14.22638pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{{{}}}{{}}{}{}{}{}{}{}{}{}{}{\pgfsys@moveto{2.12134pt}{17.07182pt}\pgfsys@curveto{2.12134pt}{18.24341pt}{1.17159pt}{19.19316pt}{0.0pt}{19.19316pt}\pgfsys@curveto{-1.17159pt}{19.19316pt}{-2.12134pt}{18.24341pt}{-2.12134pt}{17.07182pt}\pgfsys@curveto{-2.12134pt}{15.90024pt}{-1.17159pt}{14.95049pt}{0.0pt}{14.95049pt}\pgfsys@curveto{1.17159pt}{14.95049pt}{2.12134pt}{15.90024pt}{2.12134pt}{17.07182pt}\pgfsys@closepath\pgfsys@moveto{0.0pt}{17.07182pt}\pgfsys@fill\pgfsys@invoke{ } }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{0.0pt}{17.07182pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{{ {}{}{}}}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}+\leavevmode\hbox to17.47pt{\vbox to19.39pt{\pgfpicture\makeatletter\hbox{\hskip 8.7359pt\lower 14.95049pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ } {{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{0.0pt}{22.76228pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{{{}}}{{}}{}{}{}{}{}{}{}{}{}{\pgfsys@moveto{2.12134pt}{17.07182pt}\pgfsys@curveto{2.12134pt}{18.24341pt}{1.17159pt}{19.19316pt}{0.0pt}{19.19316pt}\pgfsys@curveto{-1.17159pt}{19.19316pt}{-2.12134pt}{18.24341pt}{-2.12134pt}{17.07182pt}\pgfsys@curveto{-2.12134pt}{15.90024pt}{-1.17159pt}{14.95049pt}{0.0pt}{14.95049pt}\pgfsys@curveto{1.17159pt}{14.95049pt}{2.12134pt}{15.90024pt}{2.12134pt}{17.07182pt}\pgfsys@closepath\pgfsys@moveto{0.0pt}{17.07182pt}\pgfsys@fill\pgfsys@invoke{ } }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{0.0pt}{17.07182pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {}{}{{}}{}{} {}{}{} {}{{}{}{}}{}{}{}{{}{}{}}{}{}{}{{}{}{}}{{}{}{}}{}{}\pgfsys@moveto{-8.5359pt}{25.6073pt}\pgfsys@curveto{-8.5359pt}{30.34454pt}{-4.73735pt}{34.14322pt}{0.0pt}{34.14322pt}\pgfsys@curveto{4.73735pt}{34.14322pt}{8.5359pt}{30.34454pt}{8.5359pt}{25.6073pt}\pgfsys@curveto{8.5359pt}{20.87006pt}{4.73735pt}{17.07182pt}{0.0pt}{17.07182pt}\pgfsys@curveto{-4.73735pt}{17.07182pt}{-8.5359pt}{20.87006pt}{-8.5359pt}{25.6073pt}\pgfsys@closepath\pgfsys@stroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{{ {}{}{}}}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}+\leavevmode\hbox to48.13pt{\vbox to17.47pt{\pgfpicture\makeatletter\hbox{\hskip 3.33301pt\lower-8.7359pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ } {{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{0.0pt}{-2.84544pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{{{}}}{{}}{}{}{}{}{}{}{}{}{}{\pgfsys@moveto{2.12134pt}{0.0pt}\pgfsys@curveto{2.12134pt}{1.17159pt}{1.17159pt}{2.12134pt}{0.0pt}{2.12134pt}\pgfsys@curveto{-1.17159pt}{2.12134pt}{-2.12134pt}{1.17159pt}{-2.12134pt}{0.0pt}\pgfsys@curveto{-2.12134pt}{-1.17159pt}{-1.17159pt}{-2.12134pt}{0.0pt}{-2.12134pt}\pgfsys@curveto{1.17159pt}{-2.12134pt}{2.12134pt}{-1.17159pt}{2.12134pt}{0.0pt}\pgfsys@closepath\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@fill\pgfsys@invoke{ } }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{0.0pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{{{}}}{{}}{}{}{}{}{}{}{}{}{}{\pgfsys@moveto{44.80048pt}{0.0pt}\pgfsys@curveto{44.80048pt}{1.17159pt}{43.85072pt}{2.12134pt}{42.67914pt}{2.12134pt}\pgfsys@curveto{41.50755pt}{2.12134pt}{40.5578pt}{1.17159pt}{40.5578pt}{0.0pt}\pgfsys@curveto{40.5578pt}{-1.17159pt}{41.50755pt}{-2.12134pt}{42.67914pt}{-2.12134pt}\pgfsys@curveto{43.85072pt}{-2.12134pt}{44.80048pt}{-1.17159pt}{44.80048pt}{0.0pt}\pgfsys@closepath\pgfsys@moveto{42.67914pt}{0.0pt}\pgfsys@fill\pgfsys@invoke{ } }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{42.67914pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {}{}{}{}{} {}{{}{}{}}{}{}{}{}{}\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@curveto{0.0pt}{0.0pt}{7.12775pt}{8.5359pt}{21.33957pt}{8.5359pt}\pgfsys@curveto{35.55139pt}{8.5359pt}{42.67914pt}{0.0pt}{42.67914pt}{0.0pt}\pgfsys@stroke\pgfsys@invoke{ } {}{}{}{}{} {}{{}{}{}}{}{}{}{}{}\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@curveto{0.0pt}{0.0pt}{7.12775pt}{-8.5359pt}{21.33957pt}{-8.5359pt}\pgfsys@curveto{35.55139pt}{-8.5359pt}{42.67914pt}{0.0pt}{42.67914pt}{0.0pt}\pgfsys@stroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{{ {}{}{}}}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}
=\displaystyle= cn0Δ𝐤0+18qΛUB22q2+6B04q2Bτωn2+𝐪B122qΛUq2q2(Bτωn2+𝐪)2\displaystyle\ c_{\perp}n_{0}\Delta_{{\bf k}_{0}}+\frac{1}{8}\int_{q}^{\Lambda_{U}}\frac{B_{22}q_{\parallel}^{2}+6B_{04}q_{\perp}^{2}}{B_{\tau}\omega_{n}^{2}+\mathcal{E}_{\bf q}}-B_{12}^{2}\int_{q}^{\Lambda_{U}}\frac{q_{\parallel}^{2}q_{\perp}^{2}}{(B_{\tau}\omega_{n}^{2}+\mathcal{E}_{\bf q})^{2}}
=\displaystyle= cB122cqΛU3q2+q2Bτωn2+𝐪+18qΛUB22q2+6B04q2Bτωn2+𝐪B122qΛUq2q2(Bτωn2+𝐪)2,\displaystyle\ -\frac{c_{\perp}B_{12}}{2c}\int_{q}^{\Lambda_{U}}\frac{3q_{\parallel}^{2}+q_{\perp}^{2}}{B_{\tau}\omega_{n}^{2}+\mathcal{E}_{\bf q}}+\frac{1}{8}\int_{q}^{\Lambda_{U}}\frac{B_{22}q_{\parallel}^{2}+6B_{04}q_{\perp}^{2}}{B_{\tau}\omega_{n}^{2}+\mathcal{E}_{\bf q}}-B_{12}^{2}\int_{q}^{\Lambda_{U}}\frac{q_{\parallel}^{2}q_{\perp}^{2}}{(B_{\tau}\omega_{n}^{2}+\mathcal{E}_{\bf q})^{2}}, (217)

where in the last line we choose 𝐤0{\bf k}_{0} at which the linear term ϕ\partial_{\parallel}\phi vanished. In the above, the arguments of the coefficients are taken to be their mean-field value k0=k¯0k_{0}=\bar{k}_{0} as the difference is of higher order. With a change of variables q=ξqq^{\prime}_{\parallel}=\xi q_{\parallel} and q=(λξ)1/2qq^{\prime}_{\perp}=(\lambda\xi)^{1/2}q_{\perp}, we can rewrite the expression above that in the weakly-interacting limit is given by

B(1)\displaystyle B_{\perp}^{(1)}\approx cB12Bτ4cξ(λξ)3/2qΛq2ωn+2q+2q4+3B04Bτ4ξ(λξ)3/2qΛq2ωn+2q+2q4B122Bτ2ξ3(λξ)3/2qΛqq22(ωn+2q+2q)42\displaystyle\ -\frac{c_{\perp}B_{12}B_{\tau}}{4c\xi(\lambda\xi)^{3/2}}\int_{q^{\prime}}^{\Lambda}\frac{q^{\prime}_{\perp}{}^{2}}{\omega^{\prime}_{n}{}^{2}+q^{\prime}_{\parallel}{}^{2}+q^{\prime}_{\perp}{}^{4}}+\frac{3B_{04}B_{\tau}}{4\xi(\lambda\xi)^{3/2}}\int_{q^{\prime}}^{\Lambda}\frac{q^{\prime}_{\perp}{}^{2}}{\omega^{\prime}_{n}{}^{2}+q^{\prime}_{\parallel}{}^{2}+q^{\prime}_{\perp}{}^{4}}-\frac{B_{12}^{2}B_{\tau}^{2}}{\xi^{3}(\lambda\xi)^{3/2}}\int_{q^{\prime}}^{\Lambda}\frac{q^{\prime}_{\parallel}{}^{2}q^{\prime}_{\perp}{}^{2}}{(\omega^{\prime}_{n}{}^{2}+q^{\prime}_{\parallel}{}^{2}+q^{\prime}_{\perp}{}^{4})^{2}}
=\displaystyle= 3B04cB12/c8ξ(λξ)3/2𝐪Λq2E𝐪coth(Un0E𝐪T)BτB1222ξ3(λξ)3/2𝐪Λqq222E𝐪3[coth(Un0E𝐪T)+Un0E𝐪Tcsch2(Un0E𝐪T)],\displaystyle\ \frac{3B_{04}-c_{\perp}B_{12}/c}{8\xi(\lambda\xi)^{3/2}}\int_{{\bf q}^{\prime}}^{\Lambda}\frac{q^{\prime}_{\perp}{}^{2}}{E^{\prime}_{{\bf q}^{\prime}}}\coth\left(\frac{Un_{0}E^{\prime}_{{\bf q}^{\prime}}}{T}\right)-\frac{B_{\tau}B_{12}^{2}}{2\xi^{3}(\lambda\xi)^{3/2}}\int_{{\bf q}^{\prime}}^{\Lambda}\frac{q^{\prime}_{\parallel}{}^{2}q^{\prime}_{\perp}{}^{2}}{2E^{\prime}_{{\bf q}^{\prime}}{}^{3}}\left[\coth\left(\frac{Un_{0}E^{\prime}_{{\bf q}^{\prime}}}{T}\right)+\frac{Un_{0}E^{\prime}_{{\bf q}^{\prime}}}{T}\text{csch}^{2}\left(\frac{Un_{0}E^{\prime}_{{\bf q}^{\prime}}}{T}\right)\right], (218)

where ωn=Bτωn=πnT/Un0\omega^{\prime}_{n}=B_{\tau}\omega_{n}=\pi nT/Un_{0} (here nn is an integer, not confused with the particle density), E𝐪=q+2q4E^{\prime}_{{\bf q}^{\prime}}=q^{\prime}_{\parallel}{}^{2}+q^{\prime}_{\perp}{}^{4} and the UV cutoff Λ\Lambda indicates 𝐪=(q,q)<(𝒪(1),𝒪(1)){\bf q}^{\prime}=(q^{\prime}_{\parallel},q^{\prime}_{\perp})<(\mathcal{O}(1),\mathcal{O}(1)). In the last line above, we can read B(1)=B0fB(Un0/T)B_{\perp}^{(1)}=B_{0\perp}f_{B}(Un_{0}/T), where the dimensionless scaling function fBf_{B} exhibits the same limits as fbf_{b} in Eq. (115) and B0=(Un0)5/4t11/4gB(ρ)B_{0\perp}=(Un_{0})^{5/4}t_{1}^{-1/4}g_{B}(\rho) with gB(ρ)g_{B}(\rho) a dimensionless 𝒪(1)\mathcal{O}(1) constant.

J.2 Isotropic limit

In the isotropic limit ρ1/6\rho\to 1/6 or k00k_{0}\to 0, the coefficients KK20=K02=K11/2K\equiv K_{20}=K_{02}=K_{11}/2, B/k02=B30/k0=B12/k0=B40=B04=B22/2B/k_{0}^{2}=B_{30}/k_{0}=B_{12}/k_{0}=B_{40}=B_{04}=B_{22}/2 and c=2k0cc=2k_{0}c_{\perp}. Consequently, B(1)B_{\perp}^{(1)} in the last line of Eq. (J.1) vanishes

B(1)=\displaystyle B_{\perp}^{(1)}= B4k02q3q2+q2Bτωn2+Bq2+Kq4+B4k02qq2+3q2Bτωn2+Bq2+Kq4B2k02qq2q2(Bτωn2+Bq2+Kq4)2=0,\displaystyle\ -\frac{B}{4k_{0}^{2}}\int_{q}\frac{3q_{\parallel}^{2}+q_{\perp}^{2}}{B_{\tau}\omega_{n}^{2}+Bq_{\parallel}^{2}+Kq^{4}}+\frac{B}{4k_{0}^{2}}\int_{q}\frac{q_{\parallel}^{2}+3q_{\perp}^{2}}{B_{\tau}\omega_{n}^{2}+Bq_{\parallel}^{2}+Kq^{4}}-\frac{B^{2}}{k_{0}^{2}}\int_{q}\frac{q_{\parallel}^{2}q_{\perp}^{2}}{(B_{\tau}\omega_{n}^{2}+Bq_{\parallel}^{2}+Kq^{4})^{2}}=0, (219)

where we used the following integration by part for the last term

𝐪q2q2(Bτωn2+Bq2+Kq4)2=0𝑑qq02π𝑑θq4cos(θ)2sin(θ)2[Bτωn2+Bq2cos(θ)2+Kq4]2\displaystyle\int_{\bf q}\frac{q_{\parallel}^{2}q_{\perp}^{2}}{(B_{\tau}\omega_{n}^{2}+Bq_{\parallel}^{2}+Kq^{4})^{2}}=\int_{0}^{\infty}dqq\int_{0}^{2\pi}d\theta\frac{q^{4}\cos(\theta)^{2}\sin(\theta)^{2}}{[B_{\tau}\omega_{n}^{2}+Bq^{2}\cos(\theta)^{2}+Kq^{4}]^{2}}
=\displaystyle= 12B0𝑑qq02π𝑑θq2[sin(θ)2cos(θ)2]Bτωn2+Bq2cos(θ)2+Kq4=12B𝐪q2q2Bτωn2+Bq2+Kq4.\displaystyle\ \frac{1}{2B}\int_{0}^{\infty}dqq\int_{0}^{2\pi}d\theta\frac{q^{2}[\sin(\theta)^{2}-\cos(\theta)^{2}]}{B_{\tau}\omega_{n}^{2}+Bq^{2}\cos(\theta)^{2}+Kq^{4}}=\frac{1}{2B}\int_{\bf q}\frac{q_{\perp}^{2}-q_{\parallel}^{2}}{B_{\tau}\omega_{n}^{2}+Bq_{\parallel}^{2}+Kq^{4}}. (220)

References