Increasing superconducting Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT by layering in the attractive Hubbard model

Rodrigo A. Fontenele    Natanael C. Costa    Thereza Paiva    Raimundo R. dos Santos Instituto de Física, Universidade Federal do Rio de Janeiro Cx.P. 68.528, 21941-972 Rio de Janeiro RJ, Brazil
(August 30, 2024)
Abstract

The attractive Hubbard model has become a model readily realizable with ultracold atoms on optical lattices. However, the superconducting (superfluid) critical temperatures, Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT’s, are still somewhat smaller than the lowest temperatures achieved in experiments. Here we consider two possible routes, generically called layering, to increase Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT: a bilayer and a simple cubic lattice, both with tunable hopping, tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, between attractive Hubbard planes. We have performed minus-sign–free determinant quantum Monte Carlo simulations to calculate response functions such as pairing correlation functions, uniform spin susceptibility, and double occupancy, through which we map out some physical properties. We have found that by a judicious choice of fillings and intensity of on-site attraction, a bilayer can exhibit Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT’s between 1.5 and 1.7 times those of the single layer; for the simple-cubic lattice the enhancement can be 30% larger than the maximum for the single layer. We also check the accuracy of both a BCS-like estimate for Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT in the attractive Hubbard model, as well as of an upper bound for Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT based on the superfluid density.

I Introduction

The attractive Hubbard model (AHM) describes the dynamics of fermions moving in a single band (nearest neighbor hopping integral t𝑡titalic_t) subject to an on-site interaction, U<0𝑈0U<0italic_U < 0, which favors the formation of local pairs Micnas et al. (1990). This model has been useful to study several phenomenological aspects of superconductivity. For instance, it contemplates the existence of a temperature scale, Tp≳Tcgreater-than-or-equivalent-tosubscript𝑇𝑝subscript𝑇𝑐T_{p}\gtrsim T_{c}italic_T start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ≳ italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, with Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT being the critical temperature for superconductivity, associated with a suppression of the uniform spin susceptibility, χssubscript𝜒𝑠\chi_{s}italic_χ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, signalling the opening of a spin gap Randeria et al. (1992); Bucher et al. (1993); dos Santos (1994); Paiva et al. (2010); Fontenele et al. (2022). This connects directly with pseudo-gap phenomena in high-temperature cuprate superconductors Wilson (2001). Another important feature of the AHM is the possibility of interpolating between the weak coupling limit, displaying BCS behavior, with long coherence length pairs, and the strong coupling regime, with tightly-bound pairs and short coherence length undergoing Bose-Einstein condensation Micnas et al. (1990); Randeria (1995); Chen et al. (2005); Randeria and Taylor (2014); Fontenele et al. (2022).

With the recent advances in optical lattice experiments (OLE) involving ultracold fermionic atoms Chin et al. (2010); Jaksch and Zoller (2005); Bloch et al. (2008); Esslinger (2010); McKay and DeMarco (2011); Mitra et al. (2018); Jördens et al. (2010), the AHM has found an immediate experimental realization, in which there is a fine control of the interaction strength through an external magnetic field. The advent of the quantum gas microscope Bakr et al. (2009) paved the way to visualise the atomic distribution on the lattice and draw quantitative conclusions, such as correlation functions for the AHM on a square optical lattice Mitra et al. (2018); Gall et al. (2020); Chan et al. (2020); Hartke et al. (2023). Although OLE constitute an important setup for studying strongly correlated phenomena, the current lowest temperatures achieved are still higher than theoretical predictions for Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. More specifically, recent extensive Quantum Monte Carlo (QMC) simulations Fontenele et al. (2022) mapped the square lattice phase diagram, Tc⁹(⟹n⟩,U)subscript𝑇𝑐delimited-âŸšâŸ©đ‘›đ‘ˆT_{c}(\langle n\rangle,U)italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( ⟹ italic_n ⟩ , italic_U ), with ⟹n⟩delimited-âŸšâŸ©đ‘›\langle n\rangle⟹ italic_n ⟩ being the band-filling: a maximum Tc≈0.15⁹t/kBsubscript𝑇𝑐0.15𝑡subscriptđ‘˜đ”T_{c}\approx 0.15t/k_{B}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≈ 0.15 italic_t / italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT occurs around ⟹n⟩≈0.87delimited-âŸšâŸ©đ‘›0.87\langle n\rangle\approx 0.87⟹ italic_n ⟩ ≈ 0.87 Fontenele et al. (2022), where kBsubscriptđ‘˜đ”k_{B}italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT is the Boltzmann constant, which we set to unity from now on. This corresponds to a few tens of nanokelvin, which is still about a third of the current lowest temperatures experimentally accessible in OLE Hartke et al. (2023); Mitra et al. (2018). Hence, it is important to investigate scenarios which could lead to higher Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT’s, closer to an experimentally accessible range.

As a rough guide to this quest, we recall that the BCS behavior of the AHM at weak coupling allows us to expect Micnas et al. (1990)

TcBCS=W~⁹exp⁥(−1/D⁹(0)⁹|U|),superscriptsubscript𝑇𝑐BCS~𝑊1đ·0𝑈T_{c}^{\text{BCS}}=\widetilde{W}\exp{\left(-1/D(0)|U|\right)}~{},italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT BCS end_POSTSUPERSCRIPT = over~ start_ARG italic_W end_ARG roman_exp ( - 1 / italic_D ( 0 ) | italic_U | ) , (1)

where W~~𝑊\widetilde{W}over~ start_ARG italic_W end_ARG is proportional to the bandwidth, W𝑊Witalic_W, and D⁹(0)đ·0D(0)italic_D ( 0 ) is the density of states (DOS) at the Fermi level; throughout this paper we take W~=W~𝑊𝑊\widetilde{W}=Wover~ start_ARG italic_W end_ARG = italic_W since the proportionality constant is immaterial to pinpoint the maxima critical temperatures. Therefore, one possibility to enhance Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is to increase D⁹(0)đ·0D(0)italic_D ( 0 ), while keeping fixed both the fermionic density and U𝑈Uitalic_U. For two-dimensional systems this may be achieved by adding a second layer and adequately tuning the hopping, tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, between the layers; see below. Note, however, that in two dimensions this is only applicable to the doped case, since the degeneracy between charge-density wave and superconductivity suppresses Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT to zero at half-filling, by virtue of the Mermin-Wagner theorem. Physically, the presence of a second layer offers another hopping channel, thus allowing pre-formed pairs to be less scattered throughout the lattice. The possibility of a bilayer inducing a behavior different from the monolayer has been considered in the repulsive Fermi-Hubbard model: it was found that interplane antiferromagnetic correlations tend to enhance d𝑑ditalic_d-wave pairing Scalettar et al. (1994); dos Santos (1995).

Similar arguments apply to a simple cubic lattice, where a tunable tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT could benefit from the fact that Tc≠0subscript𝑇𝑐0T_{c}\neq 0italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≠ 0 for tz=tsubscript𝑡𝑧𝑡t_{z}=titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = italic_t even at half filling dos Santos (1994). By varying tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT between 0 and 1, we follow a dimensional crossover between two- and three dimensions. Recent OLE on Bose-Einstein condensates with Rubidium-87 atoms allowed the study of dynamical properties crossing over between 2D and 3D behavior Zheng et al. (2023); this system can be modeled by a Bose-Hubbard model with anisotropic hopping. Again, QMC simulations on the three-dimensional anisotropic repulsive Hubbard model at half-filling have found that a finite hopping between layers can enhance the magnetic structure factor relative to the 2222D case, and produce a slight increase in the critical temperature Ibarra-García-Padilla et al. (2020). In addition, a perturbative cluster approach to the AHM suggests that the order parameter is enhanced relative to the two-dimensional case Ogawa et al. (2015).

In view of these possibilities, a systematic study of Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT for the anisotropic AHM both in the bilayer and in three dimensions is of interest. With this in mind, here we use Determinant Quantum Monte Carlo (DQMC) simulations to investigate the AHM on these geometries.

The layout of the paper is as follows. In Section II we discuss the model and highlight the main aspects of DQMC, including the quantities used to probe the physical properties of the system. In Section III we present estimates for the critical temperature as a function of interlayer hopping on a square bilayer, followed by a discussion on the pairing temperature scale. In Section IV, we consider the simple cubic lattice with variable interlayer hopping, tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT. In both cases we analyze the superconducting properties at different band fillings and strengths of the on-site attraction. Finally, in Section V we present our conclusions.

II Model and Methodology

Refer to caption
Figure 1: (Color online) Fermionic atoms trapped in a bilayer. Each atom can be in either an |↑⟩ket↑|\!\uparrow\rangle| ↑ ⟩ state (blue) or a |↓⟩ket↓|\!\downarrow\rangle| ↓ ⟩ state (red). An atom may hop to its neighboring site at a rate t𝑡titalic_t within a plane, and tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT between planes. When two atoms in opposite spin states occupy the same site, the energy is lowered by |U|𝑈|U|| italic_U |.

We write the anisotropic attractive Hubbard Hamiltonian as,

ℋℋ\displaystyle\mathcal{H}caligraphic_H =−tⁱ∑⟹𝐱,đŁâŸ©âˆ„,σ(cđąâąÏƒâ€ âącđŁâąÏƒ+h.c.)−tzⁱ∑⟹𝐱,đŁâŸ©âŸ‚,σ(cđąâąÏƒâ€ âącđŁâąÏƒ+h.c.)absent𝑡subscriptsubscript𝐱𝐣parallel-to𝜎superscriptsubscript𝑐𝐱𝜎†superscriptsubscript𝑐𝐣𝜎absenth.c.subscript𝑡𝑧subscriptsubscript𝐱𝐣perpendicular-to𝜎superscriptsubscript𝑐𝐱𝜎†superscriptsubscript𝑐𝐣𝜎absenth.c.\displaystyle=-t\sum_{\langle\mathbf{i},\mathbf{j}\rangle_{\parallel},\sigma}% \left(c_{\mathbf{i}\sigma}^{\dagger}c_{\mathbf{j}\sigma}^{\phantom{\dagger}}+% \mbox{h.c.}\right)-t_{z}\sum_{\langle\mathbf{i},\mathbf{j}\rangle_{\perp},% \sigma}\left(c_{\mathbf{i}\sigma}^{\dagger}c_{\mathbf{j}\sigma}^{\phantom{% \dagger}}+\mbox{h.c.}\right)= - italic_t ∑ start_POSTSUBSCRIPT ⟹ bold_i , bold_j ⟩ start_POSTSUBSCRIPT ∄ end_POSTSUBSCRIPT , italic_σ end_POSTSUBSCRIPT ( italic_c start_POSTSUBSCRIPT bold_i italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT bold_j italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT + h.c. ) - italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT ⟹ bold_i , bold_j ⟩ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT , italic_σ end_POSTSUBSCRIPT ( italic_c start_POSTSUBSCRIPT bold_i italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT bold_j italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT + h.c. )
âˆ’ÎŒâąâˆ‘đąâąÏƒnđąâąÏƒâˆ’|U|ⁱ∑𝐱(n𝐱↑−1/2)⁹(n𝐱↓−1/2),𝜇subscript𝐱𝜎subscript𝑛𝐱𝜎𝑈subscript𝐱subscript𝑛↑𝐱absent12subscript𝑛↓𝐱absent12\displaystyle~{}~{}-\mu\sum_{\mathbf{i}\sigma}n_{\mathbf{i}\sigma}-|U|\sum_{% \mathbf{i}}\left(n_{\mathbf{i}\uparrow}-1/2\right)\left(n_{\mathbf{i}% \downarrow}-1/2\right),- italic_ÎŒ ∑ start_POSTSUBSCRIPT bold_i italic_σ end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT bold_i italic_σ end_POSTSUBSCRIPT - | italic_U | ∑ start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT ( italic_n start_POSTSUBSCRIPT bold_i ↑ end_POSTSUBSCRIPT - 1 / 2 ) ( italic_n start_POSTSUBSCRIPT bold_i ↓ end_POSTSUBSCRIPT - 1 / 2 ) , (2)

where ⟹𝐱,đŁâŸ©âˆ„subscript𝐱𝐣parallel-to\langle\mathbf{i},\mathbf{j}\rangle_{\parallel}⟹ bold_i , bold_j ⟩ start_POSTSUBSCRIPT ∄ end_POSTSUBSCRIPT represents nearest-neighbor sites within a layer, and t𝑡titalic_t is the intralayer hopping amplitude; tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT is the hopping energy between neighboring sites in adjacent layers ⟹𝐱,đŁâŸ©âŸ‚subscript𝐱𝐣perpendicular-to\langle\mathbf{i},\mathbf{j}\rangle_{\perp}⟹ bold_i , bold_j ⟩ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT (see Fig. 1); ÎŒđœ‡\muitalic_ÎŒ is the chemical potential controlling the band filling, and |U|𝑈|U|| italic_U | is the on-site attractive interaction between fermions of opposite spin. We consider unit lattice spacing, and t𝑡titalic_t sets the energy scale.

The physical quantities of interest are obtained through DQMC simulations Blankenbecler et al. (1981); Hirsch (1983, 1985); White et al. (1989); dos Santos (2003). This is an unbiased numerical approach based on an auxiliary-field decomposition of the interaction, which maps onto a quadratic form of free fermions coupled to bosonic degrees of freedom, 𝒼ⁱ(𝐱,τ)=±1𝒼𝐱𝜏plus-or-minus1\mathcal{S}(\mathbf{i},\tau)=\pm 1caligraphic_S ( bold_i , italic_τ ) = ± 1, in both spatial and imaginary time coordinates. This method is based on a separation of the non-commuting parts of the Hamiltonian by means of the Trotter-Suzuki decomposition, i.e.

đ’”đ’”\displaystyle\mathcal{Z}caligraphic_Z =Tr⁹e−ÎČⁱℋ^=Tr⁹[(eâˆ’Î”âąÏ„âą(ℋ^0+ℋ^U))M]≈absentTrsuperscriptđ‘’đ›œ^ℋTrdelimited-[]superscriptsuperscript𝑒Δ𝜏subscript^ℋ0subscript^ℋU𝑀absent\displaystyle=\mathrm{Tr}\,e^{-\beta\widehat{\mathcal{H}}}=\mathrm{Tr}\,[(e^{-% \Delta\tau(\widehat{\mathcal{H}}_{0}+\widehat{\mathcal{H}}_{\rm U})})^{M}]\thickapprox= roman_Tr italic_e start_POSTSUPERSCRIPT - italic_ÎČ over^ start_ARG caligraphic_H end_ARG end_POSTSUPERSCRIPT = roman_Tr [ ( italic_e start_POSTSUPERSCRIPT - roman_Δ italic_τ ( over^ start_ARG caligraphic_H end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + over^ start_ARG caligraphic_H end_ARG start_POSTSUBSCRIPT roman_U end_POSTSUBSCRIPT ) end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT italic_M end_POSTSUPERSCRIPT ] ≈
≈Tr⁹[eâˆ’Î”âąÏ„âąâ„‹^0⁹eâˆ’Î”âąÏ„âąâ„‹^U⁹eâˆ’Î”âąÏ„âąâ„‹^0⁹eâˆ’Î”âąÏ„âąâ„‹^Uⁱ⋯],absentTrdelimited-[]superscript𝑒Δ𝜏subscript^ℋ0superscript𝑒Δ𝜏subscript^ℋUsuperscript𝑒Δ𝜏subscript^ℋ0superscript𝑒Δ𝜏subscript^ℋU⋯\displaystyle\thickapprox\mathrm{Tr}\,[e^{-\Delta\tau\widehat{\mathcal{H}}_{0}% }e^{-\Delta\tau\widehat{\mathcal{H}}_{\rm U}}e^{-\Delta\tau\widehat{\mathcal{H% }}_{0}}e^{-\Delta\tau\widehat{\mathcal{H}}_{\rm U}}\cdots],≈ roman_Tr [ italic_e start_POSTSUPERSCRIPT - roman_Δ italic_τ over^ start_ARG caligraphic_H end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - roman_Δ italic_τ over^ start_ARG caligraphic_H end_ARG start_POSTSUBSCRIPT roman_U end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - roman_Δ italic_τ over^ start_ARG caligraphic_H end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - roman_Δ italic_τ over^ start_ARG caligraphic_H end_ARG start_POSTSUBSCRIPT roman_U end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ⋯ ] , (3)

where ℋ^0subscript^ℋ0\widehat{\mathcal{H}}_{0}over^ start_ARG caligraphic_H end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT contains the terms quadratic in fermion creation and destruction operators, and ℋ^Usubscript^ℋU\widehat{\mathcal{H}}_{\rm U}over^ start_ARG caligraphic_H end_ARG start_POSTSUBSCRIPT roman_U end_POSTSUBSCRIPT contains the quartic terms. We take ÎČ=MâąÎ”âąÏ„đ›œđ‘€Î”đœ\beta=M\Delta\tauitalic_ÎČ = italic_M roman_Δ italic_τ , with Î”âąÏ„Î”đœ\Delta\tauroman_Δ italic_τ being the grid of the imaginary-time coordinate axis. This decomposition leads to an error proportional to (Î”âąÏ„)2superscriptΔ𝜏2(\Delta\tau)^{2}( roman_Δ italic_τ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, which can be systematically reduced as Î”âąÏ„â†’0→Δ𝜏0\Delta\tau\to 0roman_Δ italic_τ → 0. Throughout this work, we choose Î”âąÏ„â‰€0.1Δ𝜏0.1\Delta\tau\leq 0.1roman_Δ italic_τ ≀ 0.1 (depending on the temperature), which is small enough to lead to systematic errors smaller than the statistical ones (from the Monte Carlo sampling). Finally, for the AHM the discrete Hubbard-Stratonovich transformation dealing with the quartic terms in ℋ^Usubscript^ℋU\widehat{\mathcal{H}}_{\rm U}over^ start_ARG caligraphic_H end_ARG start_POSTSUBSCRIPT roman_U end_POSTSUBSCRIPT leads to sign-free simulations Hirsch (1983, 1985); White et al. (1989); dos Santos (2003).

In order to probe the emergence of superconductivity, we analyze the s𝑠sitalic_s-wave pair correlation functions,

C𝐱𝐣Δ≡12ⁱ⟹b𝐱†ⁱb𝐣+H.c.⟩,superscriptsubscriptđ¶đąđŁÎ”12delimited-⟚⟩superscriptsubscript𝑏𝐱†superscriptsubscript𝑏𝐣absentH.c.C_{\mathbf{i}\mathbf{j}}^{\Delta}\equiv\frac{1}{2}\langle b_{\mathbf{i}}^{% \dagger}b_{\mathbf{j}}^{\phantom{\dagger}}+\text{H.c.}\rangle,italic_C start_POSTSUBSCRIPT bold_ij end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ end_POSTSUPERSCRIPT ≡ divide start_ARG 1 end_ARG start_ARG 2 end_ARG ⟹ italic_b start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT bold_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT + H.c. ⟩ , (4)

with b𝐱†≡c𝐱↑†ⁱc𝐱↓†superscriptsubscript𝑏𝐱†superscriptsubscript𝑐↑𝐱absent†superscriptsubscript𝑐↓𝐱absent†b_{\mathbf{i}}^{\dagger}\equiv c_{\mathbf{i}\uparrow}^{\dagger}c_{\mathbf{i}% \downarrow}^{\dagger}italic_b start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ≡ italic_c start_POSTSUBSCRIPT bold_i ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT bold_i ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT (b𝐱≡c𝐱↓ⁱc𝐱↑superscriptsubscript𝑏𝐱absentsuperscriptsubscript𝑐↓𝐱absentabsentsuperscriptsubscript𝑐↑𝐱absentabsentb_{\mathbf{i}}^{\phantom{\dagger}}\equiv c_{\mathbf{i}\downarrow}^{\phantom{% \dagger}}c_{\mathbf{i}\uparrow}^{\phantom{\dagger}}italic_b start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ≡ italic_c start_POSTSUBSCRIPT bold_i ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT bold_i ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT) corresponding to creation (annihilation) of a pair of fermions at a given site 𝐱𝐱\mathbf{i}bold_i. Further, the Fourier transform of C𝐱𝐣Δsuperscriptsubscriptđ¶đąđŁÎ”C_{\mathbf{i}\mathbf{j}}^{\Delta}italic_C start_POSTSUBSCRIPT bold_ij end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ end_POSTSUPERSCRIPT at đȘ=0đȘ0\mathbf{q}=0bold_q = 0 defines the s-wave pair structure factor,

Ps⁹(đȘ=0)=1Nⁱ∑𝐱,𝐣C𝐱𝐣Δsubscript𝑃𝑠đȘ01𝑁subscript𝐱𝐣superscriptsubscriptđ¶đąđŁÎ”P_{s}(\mathbf{q}=0)=\frac{1}{N}\sum_{\mathbf{i},\mathbf{j}}C_{\mathbf{i}% \mathbf{j}}^{\Delta}italic_P start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( bold_q = 0 ) = divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT bold_i , bold_j end_POSTSUBSCRIPT italic_C start_POSTSUBSCRIPT bold_ij end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ end_POSTSUPERSCRIPT (5)

with N=L×L×Lz𝑁𝐿𝐿subscript𝐿𝑧N=L\times L\times L_{z}italic_N = italic_L × italic_L × italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT being the number of sites of the lattice, where L𝐿Litalic_L is the linear dimension of the 2222D layers. For the bilayer, Lz=2subscript𝐿𝑧2L_{z}=2italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 2, and for the cubic lattice, Lz=Lsubscript𝐿𝑧𝐿L_{z}=Litalic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = italic_L; periodic boundary conditions (PBC) are assumed in both cases. One should keep in mind that in Eq. (5) the dependence of Pssubscript𝑃𝑠P_{s}italic_P start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT on L𝐿Litalic_L, U𝑈Uitalic_U, tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, ⟹n⟩delimited-âŸšâŸ©đ‘›\langle n\rangle⟹ italic_n ⟩, and ÎČđ›œ\betaitalic_ÎČ has been omitted, to simplify the notation.

We estimate the inverse critical temperature, ÎČc≡1/Tcsubscriptđ›œđ‘1subscript𝑇𝑐\beta_{c}\equiv 1/T_{c}italic_ÎČ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≡ 1 / italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, through the correlation ratio,

Rc⁹(L,ÎČ)=1−Ps⁹(đȘ+Ύ⁹đȘ)Ps⁹(đȘ),subscriptđ‘…đ‘đżđ›œ1subscript𝑃𝑠đȘ𝛿đȘsubscript𝑃𝑠đȘ\displaystyle R_{c}(L,\beta)=1-\frac{P_{s}(\mathbf{q}+\delta\mathbf{q})}{P_{s}% (\mathbf{q})}~{},italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_L , italic_ÎČ ) = 1 - divide start_ARG italic_P start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( bold_q + italic_ÎŽ bold_q ) end_ARG start_ARG italic_P start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( bold_q ) end_ARG , (6)

where đȘ=0đȘ0\mathbf{q}=0bold_q = 0, and Ύ⁹đȘ=2âąÏ€/L𝛿đȘ2𝜋𝐿\delta\mathbf{q}=2\pi/Litalic_ÎŽ bold_q = 2 italic_π / italic_L. This quantity is a renormalisation-group invariant at the critical point Kaul (2015); Sato et al. (2018); Darmawan et al. (2018); that is, crossings of curves for Rc⁹(L,ÎČ)subscriptđ‘…đ‘đżđ›œR_{c}(L,\beta)italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_L , italic_ÎČ ) for different system sizes provide estimates for ÎČcsubscriptđ›œđ‘\beta_{c}italic_ÎČ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, which improve as L𝐿Litalic_L increases.

For our purposes here, the magnetic properties are probed by the uniform susceptibility. This is obtained through the fluctuation-dissipation theorem, which leads to χs=ÎČⁱ⟹S2⟩subscriptđœ’đ‘ đ›œdelimited-⟚⟩superscript𝑆2\chi_{s}=\beta\langle S^{2}\rangleitalic_χ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = italic_ÎČ âŸš italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩, with ⟹S2⟩delimited-⟚⟩superscript𝑆2\langle S^{2}\rangle⟹ italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ being the uniform spin structure factor,

⟹S2⟩=1Nâąâˆ‘đąđŁâŸšđ’đąâ‹…đ’đŁâŸ©,delimited-⟚⟩superscript𝑆21𝑁subscript𝐱𝐣delimited-⟚⟩⋅subscript𝐒𝐱subscript𝐒𝐣\langle S^{2}\rangle=\frac{1}{N}\sum_{\mathbf{i}\mathbf{j}}\langle\mathbf{S}_{% \mathbf{i}}\cdot\mathbf{S}_{\mathbf{j}}\rangle,⟹ italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ = divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT bold_ij end_POSTSUBSCRIPT ⟹ bold_S start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT ⋅ bold_S start_POSTSUBSCRIPT bold_j end_POSTSUBSCRIPT ⟩ , (7)

where 𝐒𝐱≡(1/2)ⁱ𝐩𝐱subscript𝐒𝐱12subscript𝐩𝐱\mathbf{S}_{\mathbf{i}}\equiv(1/2)\mathbf{m}_{\mathbf{i}}bold_S start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT ≡ ( 1 / 2 ) bold_m start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT, with the components of the magnetization operator being

m𝐱xsuperscriptsubscriptđ‘šđąđ‘„\displaystyle m_{\mathbf{i}}^{x}italic_m start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT ≡\displaystyle\equiv≡ c𝐱↑†ⁱc𝐱↓+c𝐱↓†ⁱc𝐱↑,superscriptsubscript𝑐↑𝐱absent†superscriptsubscript𝑐↓𝐱absentabsentsuperscriptsubscript𝑐↓𝐱absent†superscriptsubscript𝑐↑𝐱absentabsent\displaystyle c_{\mathbf{i}\uparrow}^{\dagger}c_{\mathbf{i}\downarrow}^{% \phantom{\dagger}}+c_{\mathbf{i}\downarrow}^{\dagger}c_{\mathbf{i}\uparrow}^{% \phantom{\dagger}},italic_c start_POSTSUBSCRIPT bold_i ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT bold_i ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT + italic_c start_POSTSUBSCRIPT bold_i ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT bold_i ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT , (8a)
m𝐱ysuperscriptsubscript𝑚𝐱𝑩\displaystyle m_{\mathbf{i}}^{y}italic_m start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_y end_POSTSUPERSCRIPT ≡\displaystyle\equiv≡ −iⁱ(c𝐱↑†ⁱc𝐱↓−c𝐱↓†ⁱc𝐱↑),𝑖superscriptsubscript𝑐↑𝐱absent†superscriptsubscript𝑐↓𝐱absentabsentsuperscriptsubscript𝑐↓𝐱absent†superscriptsubscript𝑐↑𝐱absentabsent\displaystyle-i\left(c_{\mathbf{i}\uparrow}^{\dagger}c_{\mathbf{i}\downarrow}^% {\phantom{\dagger}}-c_{\mathbf{i}\downarrow}^{\dagger}c_{\mathbf{i}\uparrow}^{% \phantom{\dagger}}\right),- italic_i ( italic_c start_POSTSUBSCRIPT bold_i ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT bold_i ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT - italic_c start_POSTSUBSCRIPT bold_i ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT bold_i ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ) , (8b)
m𝐱zsuperscriptsubscript𝑚𝐱𝑧\displaystyle m_{\mathbf{i}}^{z}italic_m start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT ≡\displaystyle\equiv≡ n𝐱↑−n𝐱↓.subscript𝑛↑𝐱absentsubscript𝑛↓𝐱absent\displaystyle n_{\mathbf{i}\uparrow}-n_{\mathbf{i}\downarrow}.italic_n start_POSTSUBSCRIPT bold_i ↑ end_POSTSUBSCRIPT - italic_n start_POSTSUBSCRIPT bold_i ↓ end_POSTSUBSCRIPT . (8c)
Refer to caption
Figure 2: Density of states (DOS) for the non-interacting bilayer as a function of energy, ω=Î”âˆ’ÎŒâą(tz)𝜔𝜀𝜇subscript𝑡𝑧\omega=\varepsilon-\mu(t_{z})italic_ω = italic_Δ - italic_ÎŒ ( italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ), and interlayer hopping, tz/tsubscript𝑡𝑧𝑡t_{z}/titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT / italic_t; the Fermi energy for a given electronic density is located at the origin, for (a) ⟹n⟩=0.87delimited-âŸšâŸ©đ‘›0.87\langle n\rangle=0.87⟹ italic_n ⟩ = 0.87 and (b) ⟹n⟩=0.5delimited-âŸšâŸ©đ‘›0.5\langle n\rangle=0.5⟹ italic_n ⟩ = 0.5. The cyan stars in the main panels mark D⁹(0)đ·0D(0)italic_D ( 0 ), as plotted in the insets as functions of tz/tsubscript𝑡𝑧𝑡t_{z}/titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT / italic_t.

In what follows, we separate the discussion in two parts. First, we consider a bilayer, in which case our simulations were carried out for L≀16𝐿16L\leq 16italic_L ≀ 16. Subsequently, we study the actual dimensional crossover in cubic systems with L≀10𝐿10L\leq 10italic_L ≀ 10. Typically our data have been obtained after 2−4×10424superscript1042-4\times 10^{4}2 - 4 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT warming-up steps, followed by 8−16×104816superscript1048-16\times 10^{4}8 - 16 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT sweeps for measurements, depending on the temperature, interaction strength, and electronic density. We have performed 10101010 to 20202020 realizations for each set of parameters to ensure the best accuracy for our results.

III The bilayer

III.1 Critical temperature

We start the discussion with the non-interacting bilayer. The single-particle energies are given by

Î”đ€=−2⁹t⁹[cos⁥kx+cos⁥ky]±tz,subscriptđœ€đ€plus-or-minus2𝑡delimited-[]subscriptđ‘˜đ‘„subscript𝑘𝑩subscript𝑡𝑧\varepsilon_{\mathbf{k}}=-2t\left[\cos k_{x}+\cos k_{y}\right]\pm t_{z}~{},italic_Δ start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT = - 2 italic_t [ roman_cos italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT + roman_cos italic_k start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ] ± italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , (9)

whose DOS’s are displayed in Fig. 2, for different densities and several values of tz/tsubscript𝑡𝑧𝑡t_{z}/titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT / italic_t. For tz≠0subscript𝑡𝑧0t_{z}\neq 0italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ≠ 0 the DOS may be thought of two square lattice densities of states, displaced from each other by 2⁹tz2subscript𝑡𝑧2t_{z}2 italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, as it can be seen from the positions of the van Hove singularities (VHS’s). For instance, the insets of Fig. 2 show the DOS as a function tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, for ⟹n⟩=0.87delimited-âŸšâŸ©đ‘›0.87\langle n\rangle=0.87⟹ italic_n ⟩ = 0.87 and 0.5, respectively: a substantial increase is found in some cases. The fine tuning of the VHS’s with a single parameter opens the possibility of increasing critical temperatures, as suggested by Eq. (1). Therefore, the effects of the DOS on pairing (i.e. whether Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT increases or not) are investigated below.

Refer to caption
Figure 3: Correlation ratio for the AHM on a bilayer, as a function of the inverse temperature, ÎČđ›œ\betaitalic_ÎČ, for isotropic hopping and ⟹n⟩=0.87delimited-âŸšâŸ©đ‘›0.87\langle n\rangle=0.87⟹ italic_n ⟩ = 0.87. Different curves are labelled by their corresponding linear lattice sizes, L𝐿Litalic_L.

Recalling that the global maximum Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT for the single layer lies near ⟹n⟩=0.87delimited-âŸšâŸ©đ‘›0.87\langle n\rangle=0.87⟹ italic_n ⟩ = 0.87 Fontenele et al. (2022), we compare the effect of tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT on Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT both for this density and for ⟹n⟩=0.5delimited-âŸšâŸ©đ‘›0.5\langle n\rangle=0.5⟹ italic_n ⟩ = 0.5. Similarly, the maximum Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT over a wide range of densities occurs at U=−5⁹t𝑈5𝑡U=-5titalic_U = - 5 italic_t Fontenele et al. (2022), so that most of our simulations will be in the range |U|/t=3−5𝑈𝑡35|U|/t=3-5| italic_U | / italic_t = 3 - 5. In this way, starting from the global maximum Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT one can investigate possible paths in parameter space leading to larger slopes.

A typical behavior of Rc⁹(L,ÎČ)subscriptđ‘…đ‘đżđ›œR_{c}(L,\beta)italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_L , italic_ÎČ ) is shown in Fig. 3 for U=−5⁹t𝑈5𝑡U=-5titalic_U = - 5 italic_t, ⟹n⟩=0.87delimited-âŸšâŸ©đ‘›0.87\langle n\rangle=0.87⟹ italic_n ⟩ = 0.87 and tz=tsubscript𝑡𝑧𝑡t_{z}=titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = italic_t. We clearly distinguish two regimes of ÎČđ›œ\betaitalic_ÎČ: one in which Rcsubscript𝑅𝑐R_{c}italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT decreases with increasing L𝐿Litalic_L, and another in which Rcsubscript𝑅𝑐R_{c}italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT increases. The inversion takes place at the common crossing point, ÎČc⁹t≈4.5±0.1subscriptđ›œđ‘đ‘Ąplus-or-minus4.50.1\beta_{c}t\approx 4.5\pm 0.1italic_ÎČ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_t ≈ 4.5 ± 0.1, or Tc/t=0.217±0.005subscript𝑇𝑐𝑡plus-or-minus0.2170.005T_{c}/t=0.217\pm 0.005italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_t = 0.217 ± 0.005; for ⟹n⟩=0.5delimited-âŸšâŸ©đ‘›0.5\langle n\rangle=0.5⟹ italic_n ⟩ = 0.5, we get Tc/t≈0.217±0.009subscript𝑇𝑐𝑡plus-or-minus0.2170.009T_{c}/t\approx 0.217\pm 0.009italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_t ≈ 0.217 ± 0.009. These values should be compared with those for the single layer, Tc/t=0.147±0.004subscript𝑇𝑐𝑡plus-or-minus0.1470.004T_{c}/t=0.147\pm 0.004italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_t = 0.147 ± 0.004 and Tc/t=0.126±0.003subscript𝑇𝑐𝑡plus-or-minus0.1260.003T_{c}/t=0.126\pm 0.003italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_t = 0.126 ± 0.003, respectively, with corresponding increase of 48% and 72%. Simulations for tz=0.4⁹tsubscript𝑡𝑧0.4𝑡t_{z}=0.4titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0.4 italic_t yield Tc/t=0.185±0.003subscript𝑇𝑐𝑡plus-or-minus0.1850.003T_{c}/t=0.185\pm 0.003italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_t = 0.185 ± 0.003 and 0.200±0.008plus-or-minus0.2000.0080.200\pm 0.0080.200 ± 0.008, respectively for ⟹n⟩=0.87delimited-âŸšâŸ©đ‘›0.87\langle n\rangle=0.87⟹ italic_n ⟩ = 0.87 and 0.5, while for tz=1.6⁹tsubscript𝑡𝑧1.6𝑡t_{z}=1.6titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 1.6 italic_t, one gets Tc/t=0.189±0.007subscript𝑇𝑐𝑡plus-or-minus0.1890.007T_{c}/t=0.189\pm 0.007italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_t = 0.189 ± 0.007 and 0.208±0.008plus-or-minus0.2080.0080.208\pm 0.0080.208 ± 0.008. Figure 4 shows these data, and we see that the maximum Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT occurs at tz=tsubscript𝑡𝑧𝑡t_{z}=titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = italic_t for both fillings. Note that we have allowed interlayer hoppings to be greater than intralayer ones, i.e. tz>tsubscript𝑡𝑧𝑡t_{z}>titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT > italic_t, to broaden the search for maximum Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT; while in OLE this would simply demand adjustments in the relative intensity or frequency of the laser beams, in materials this could be achieved by changing the pressure from hydrostatic to uniaxial. The decrease in Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT beyond the isotropic case may be attributed to an effective dimensional reduction, since hopping along the z𝑧zitalic_z direction becomes more favorable than hopping within the planes, thus restricting particle movement. This, in turn, may increase pair-scattering effects by lattice sites, thereby lowering the critical temperature. Note that a similar effect has been observed in a bilayer made up from a superconducting layer and a metallic layer; see, e.g. Ref. Zujev et al. (2014).

Refer to caption
Figure 4: Critical temperature, Tc/tsubscript𝑇𝑐𝑡T_{c}/titalic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_t, for the AHM on a bilayer (red diamonds), upper bound for Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT (up blue triangles) and the BCS estimate (green circles), Eq. (1) as a function of interlayer hopping, tz/tsubscript𝑡𝑧𝑡t_{z}/titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT / italic_t, for different fermionic densities, ⟹n⟩delimited-âŸšâŸ©đ‘›\langle n\rangle⟹ italic_n ⟩, and fixed U𝑈Uitalic_U. We have arbitrarily divided TcBCSsuperscriptsubscript𝑇𝑐BCST_{c}^{\text{BCS}}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT BCS end_POSTSUPERSCRIPT [Eq. (1)] by 24 just to lie closer to the other curves. All lines are guides to the eye.

It is also instructive to compare Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT both with the BCS estimate, Eq. (1), and with a suggested upper bound Hazra et al. (2019); see below. At first sight, Fig. 4 seems to indicate that TcBCSsuperscriptsubscript𝑇𝑐BCST_{c}^{\text{BCS}}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT BCS end_POSTSUPERSCRIPT tracks Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. However, a closer look reveals that TcBCSsuperscriptsubscript𝑇𝑐BCST_{c}^{\text{BCS}}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT BCS end_POSTSUPERSCRIPT does not capture the position of maximum Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT for either densities. In addition, given that the (non-interacting) bandwidth varies with tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, if one rescales TcBCSsuperscriptsubscript𝑇𝑐BCST_{c}^{\text{BCS}}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT BCS end_POSTSUPERSCRIPT by W=8ⁱt+2ⁱtz𝑊8𝑡2subscript𝑡𝑧W=8t+2t_{z}italic_W = 8 italic_t + 2 italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, the discrepancies with respect to the location of maximum Tc/tsubscript𝑇𝑐𝑡T_{c}/titalic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_t are significantly enhanced. Therefore, one cannot take TcBCSsuperscriptsubscript𝑇𝑐BCST_{c}^{\text{BCS}}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT BCS end_POSTSUPERSCRIPT at face value to pinpoint actual maxima.

Refer to caption
Figure 5: Temperature behavior of the upper bound for the superfluid density, ρ~ssubscript~𝜌𝑠\widetilde{\rho}_{s}over~ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, at fixed ⟹n⟩=0.87delimited-âŸšâŸ©đ‘›0.87\langle n\rangle=0.87⟹ italic_n ⟩ = 0.87, U/t=−5𝑈𝑡5U/t=-5italic_U / italic_t = - 5 and tz/t=0.4subscript𝑡𝑧𝑡0.4t_{z}/t=0.4italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT / italic_t = 0.4, and for different lattice sizes. The intercept with the (dotted) straight line 2⁹T/tâąÏ€2𝑇𝑡𝜋2T/t\pi2 italic_T / italic_t italic_π provides an estimate for Tcboundsuperscriptsubscript𝑇𝑐boundT_{c}^{\text{bound}}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT bound end_POSTSUPERSCRIPT; see text.

In order to further understand the enhancement of Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT presented in Fig. 4, it is worth analyzing its upper bounds, Tcboundsuperscriptsubscript𝑇𝑐boundT_{c}^{\mbox{\tiny bound}}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT bound end_POSTSUPERSCRIPT, as proposed in Ref. Hazra et al. (2019). We start by recalling that the superfluid density (or helicity modulus) may be obtained through Scalapino et al. (1992, 1993)

ρssubscript𝜌𝑠\displaystyle\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT =Ds4âąÏ€âąe2=14⁹[ΛL−ΛT],absentsubscriptđ·đ‘ 4𝜋superscript𝑒214delimited-[]superscriptΛ𝐿superscriptΛ𝑇\displaystyle=\frac{D_{s}}{4\pi e^{2}}=\frac{1}{4}\left[\Lambda^{L}-\Lambda^{T% }\right],= divide start_ARG italic_D start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_ARG start_ARG 4 italic_π italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG = divide start_ARG 1 end_ARG start_ARG 4 end_ARG [ roman_Λ start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT - roman_Λ start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT ] , (10)

which is proportional to the superfluid stiffness Dssubscriptđ·đ‘ D_{s}italic_D start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. The limiting longitudinal and transverse responses are

ΛL≡limqx→0Λx⁹x(qx,qy=0,ωn=0)\Lambda^{L}\equiv\lim_{q_{x}\to 0}\Lambda_{xx}(q_{x},q_{y}=0,\omega_{n}=0)roman_Λ start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT ≡ roman_lim start_POSTSUBSCRIPT italic_q start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT → 0 end_POSTSUBSCRIPT roman_Λ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT ( italic_q start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT , italic_q start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT = 0 , italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = 0 ) (11)

and

ΛT≡limqy→0Λx⁹x⁹(qx=0,qy,ωn=0),superscriptΛ𝑇subscript→subscript𝑞𝑩0subscriptÎ›đ‘„đ‘„formulae-sequencesubscriptđ‘žđ‘„0subscript𝑞𝑩subscript𝜔𝑛0\Lambda^{T}\equiv\lim_{q_{y}\to 0}\Lambda_{xx}(q_{x}=0,q_{y},\omega_{n}=0),roman_Λ start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT ≡ roman_lim start_POSTSUBSCRIPT italic_q start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT → 0 end_POSTSUBSCRIPT roman_Λ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT ( italic_q start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT = 0 , italic_q start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT , italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = 0 ) , (12)

with

Λx⁹x⁹(đȘ,ωn)subscriptÎ›đ‘„đ‘„đȘsubscript𝜔𝑛\displaystyle\Lambda_{xx}(\mathbf{q},\omega_{n})roman_Λ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT ( bold_q , italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) =∑ℓ∫0ÎČđ‘‘Ï„âąei⁹đȘ⋅ℓⁱeiâąÏ‰nâąÏ„âąÎ›x⁹x⁹(ℓ,τ),absentsubscriptbold-ℓsuperscriptsubscript0đ›œdifferential-d𝜏superscript𝑒⋅𝑖đȘbold-ℓsuperscript𝑒𝑖subscript𝜔𝑛𝜏subscriptÎ›đ‘„đ‘„bold-ℓ𝜏\displaystyle=\sum_{\boldsymbol{\ell}}\int_{0}^{\beta}d\tau\,e^{i\mathbf{q}% \cdot\boldsymbol{\ell}}e^{i\omega_{n}\tau}\Lambda_{xx}(\boldsymbol{\ell},\tau),= ∑ start_POSTSUBSCRIPT bold_ℓ end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ÎČ end_POSTSUPERSCRIPT italic_d italic_τ italic_e start_POSTSUPERSCRIPT italic_i bold_q ⋅ bold_ℓ end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_τ end_POSTSUPERSCRIPT roman_Λ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT ( bold_ℓ , italic_τ ) , (13)

where ωn=2⁹nâąÏ€âąTsubscript𝜔𝑛2𝑛𝜋𝑇\omega_{n}=2n\pi Titalic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = 2 italic_n italic_π italic_T is the Matsubara frequency, and

Λx⁹x⁹(ℓ,τ)=⟹jx⁹(ℓ,τ)⁹jx⁹(0,0)⟩,subscriptÎ›đ‘„đ‘„bold-ℓ𝜏delimited-⟚⟩subscriptđ‘—đ‘„bold-ℓ𝜏subscriptđ‘—đ‘„00\displaystyle\Lambda_{xx}(\boldsymbol{\ell},\tau)=\langle j_{x}(\boldsymbol{% \ell},\tau)j_{x}(0,0)\rangle,roman_Λ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT ( bold_ℓ , italic_τ ) = ⟹ italic_j start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ( bold_ℓ , italic_τ ) italic_j start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ( 0 , 0 ) ⟩ , (14)

where

jx⁹(ℓ,τ)=eâ„‹âąÏ„âą[i⁹tâąâˆ‘Ïƒ(cℓ+đ±^,Ïƒâ€ âącℓ,σ−cℓ,Ïƒâ€ âącℓ+đ±^,σ)]⁹eâˆ’â„‹âąÏ„subscriptđ‘—đ‘„bold-ℓ𝜏superscript𝑒ℋ𝜏delimited-[]𝑖𝑡subscript𝜎superscriptsubscript𝑐bold-ℓ^đ±đœŽâ€ subscript𝑐bold-ℓ𝜎superscriptsubscript𝑐bold-ℓ𝜎†subscript𝑐bold-ℓ^đ±đœŽsuperscript𝑒ℋ𝜏j_{x}(\boldsymbol{\ell},\tau)=e^{{\cal H}\tau}\left[it\sum_{\sigma}\left(c_{% \boldsymbol{\ell}+\mathbf{\hat{x}},\sigma}^{{}^{\dagger}}c_{\boldsymbol{\ell},% \sigma}-c_{\boldsymbol{\ell},\sigma}^{{}^{\dagger}}c_{\boldsymbol{\ell}+% \mathbf{\hat{x}},\sigma}\right)\right]e^{-{\cal H}\tau}italic_j start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ( bold_ℓ , italic_τ ) = italic_e start_POSTSUPERSCRIPT caligraphic_H italic_τ end_POSTSUPERSCRIPT [ italic_i italic_t ∑ start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT ( italic_c start_POSTSUBSCRIPT bold_ℓ + over^ start_ARG bold_x end_ARG , italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT † end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT bold_ℓ , italic_σ end_POSTSUBSCRIPT - italic_c start_POSTSUBSCRIPT bold_ℓ , italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT † end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT bold_ℓ + over^ start_ARG bold_x end_ARG , italic_σ end_POSTSUBSCRIPT ) ] italic_e start_POSTSUPERSCRIPT - caligraphic_H italic_τ end_POSTSUPERSCRIPT (15)

is the xđ‘„xitalic_x-component of the current density operator; see Ref. Scalapino et al. (1992) for details.

Further, for a Berezinskii-Kosterlitz-Thouless (BKT) transition, a universal-jump relation involving the helicity modulus holds Nelson and Kosterlitz (1977),

Tc=π2âąÏs−subscript𝑇𝑐𝜋2superscriptsubscript𝜌𝑠\displaystyle T_{c}=\frac{\pi}{2}\rho_{s}^{-}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = divide start_ARG italic_π end_ARG start_ARG 2 end_ARG italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT (16)

where ρs−superscriptsubscript𝜌𝑠\rho_{s}^{-}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT is the value of the helicity modulus just below the critical temperature. Thus, by plotting ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT as a function of temperature, the intercept with 2⁹T/π2𝑇𝜋2T/\pi2 italic_T / italic_π provides an estimate for Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, as discussed in Ref. Paiva et al. (2004). By the same token, a rigorous upper bound for the superfluid density (or, equivalently, for the superfluid stiffness) is given by Hazra et al. (2019)

ρ~s=D~s4âąÏ€âąe2=14ⁱΛL,subscript~𝜌𝑠subscript~đ·đ‘ 4𝜋superscript𝑒214superscriptΛ𝐿\tilde{\rho}_{s}=\frac{\tilde{D}_{s}}{4\pi e^{2}}=\frac{1}{4}\Lambda^{L},over~ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = divide start_ARG over~ start_ARG italic_D end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_ARG start_ARG 4 italic_π italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG = divide start_ARG 1 end_ARG start_ARG 4 end_ARG roman_Λ start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT , (17)

given that ΛTsuperscriptΛ𝑇\Lambda^{T}roman_Λ start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT is a positive quantity Scalapino et al. (1993).

Refer to caption
Figure 6: Critical temperature as a function of the magnitude of the attractive interaction at fixed fermionic density for both the monolayer and the isotropic (i.e., tz=tsubscript𝑡𝑧𝑡t_{z}=titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = italic_t) bilayer. Lines are guides to the eye.

Since a bilayer is topologically a two-dimensional system, the above calculational strategy used for a single layer is equally applicable to the present case. Figure 5 shows typical data for ρ~s⁹(T)subscript~𝜌𝑠𝑇\tilde{\rho}_{s}(T)over~ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_T ) obtained from our DQMC simulations, for fixed ⟹n⟩delimited-âŸšâŸ©đ‘›\langle n\rangle⟹ italic_n ⟩ and U𝑈Uitalic_U, and different system sizes. Since ρ~s≄ρssubscript~𝜌𝑠subscript𝜌𝑠\tilde{\rho}_{s}\geq\rho_{s}over~ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ≄ italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, and the superfluid density is a monotonically decreasing function of the temperature, at least near Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT (see Ref. Paiva et al. (2004)), the intercept of ρ~s⁹(T)subscript~𝜌𝑠𝑇\tilde{\rho}_{s}(T)over~ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_T ) with 2⁹T/π2𝑇𝜋2T/\pi2 italic_T / italic_π will take place at some Tcbound≳Tcgreater-than-or-equivalent-tosuperscriptsubscript𝑇𝑐boundsubscript𝑇𝑐T_{c}^{\text{bound}}\gtrsim T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT bound end_POSTSUPERSCRIPT ≳ italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. Figure 4 shows our estimates for Tcboundsuperscriptsubscript𝑇𝑐boundT_{c}^{\text{bound}}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT bound end_POSTSUPERSCRIPT based on this procedure, from which we see that indeed there is not much room for Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT to increase any further, in particular for tz≳tgreater-than-or-equivalent-tosubscript𝑡𝑧𝑡t_{z}\gtrsim titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ≳ italic_t.

Let us now compare how |U|𝑈|U|| italic_U | influences the increase in Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT in both a monolayer and an isotropic bilayer. Figure 6 shows Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT as a function of |U|𝑈|U|| italic_U |, for a fixed ⟹n⟩=0.87delimited-âŸšâŸ©đ‘›0.87\langle n\rangle=0.87⟹ italic_n ⟩ = 0.87, where one may notice that the critical temperature increases faster for the bilayer (tz=1subscript𝑡𝑧1t_{z}=1italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 1) than for the monolayer (tz=0subscript𝑡𝑧0t_{z}=0italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0). That is, these two curves act as lower and upper bounds for Tc⁹(U)subscript𝑇𝑐𝑈T_{c}(U)italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_U ) when 0<tz/t<10subscript𝑡𝑧𝑡10<t_{z}/t<10 < italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT / italic_t < 1. It is interesting to note that for |U|/t=3𝑈𝑡3|U|/t=3| italic_U | / italic_t = 3, Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is hardly influenced by the presence of a second layer.

Refer to caption
Figure 7: Layer-resolved contributions to the s𝑠sitalic_s-wave pair structure factor as functions of the inverse temperature, ÎČ⁹tđ›œđ‘Ą\beta titalic_ÎČ italic_t, for different values of tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT and on an L=16𝐿16L=16italic_L = 16 bilayer. Lines are guides to the eye: full and dashed lines respectively refer to intralayer and interlayer contributions.
Refer to caption
Figure 8: (a) Finite-size-scaling plot for Pssubscript𝑃𝑠P_{s}italic_P start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, Eq. (19), for lattices with linear sizes L=12,14𝐿1214L=12,14italic_L = 12 , 14 and 16161616. The full and open symbols represent the intra- and inter-layer contributions to Pssubscript𝑃𝑠P_{s}italic_P start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. (b) The intra- (full black symbols) and inter-layer (open red symbols) contributions to the squared zero temperature gap, |Δ0|2superscriptsubscriptΔ02|\Delta_{0}|^{2}| roman_Δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, as functions of tz/tsubscript𝑡𝑧𝑡t_{z}/titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT / italic_t. In addition, the total |Δ0|2superscriptsubscriptΔ02|\Delta_{0}|^{2}| roman_Δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, as a function of tz/tsubscript𝑡𝑧𝑡t_{z}/titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT / italic_t is show (blue diamonds).

At this point, one may wonder which is the division of labor between intra- and interlayer superconducting correlations. In order to shed light into this issue, we restrict the sums in the pair structure factor, Eq. (5), to sites within the same layer, call it PsAAsuperscriptsubscript𝑃𝑠AAP_{s}^{\text{AA}}italic_P start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT AA end_POSTSUPERSCRIPT, and to sites joining the two layers along the z𝑧zitalic_z direction, PsABsuperscriptsubscript𝑃𝑠ABP_{s}^{\text{AB}}italic_P start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT AB end_POSTSUPERSCRIPT, in such way that

Ps=2ⁱPsAA+PsAB.subscript𝑃𝑠2superscriptsubscript𝑃𝑠AAsuperscriptsubscript𝑃𝑠ABP_{s}=2P_{s}^{\text{AA}}+P_{s}^{\text{AB}}.italic_P start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 2 italic_P start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT AA end_POSTSUPERSCRIPT + italic_P start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT AB end_POSTSUPERSCRIPT . (18)

Typical data for PsAAsuperscriptsubscript𝑃𝑠AAP_{s}^{\text{AA}}italic_P start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT AA end_POSTSUPERSCRIPT and PsABsuperscriptsubscript𝑃𝑠ABP_{s}^{\text{AB}}italic_P start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT AB end_POSTSUPERSCRIPT as functions of the inverse temperature are shown in Fig. 7. By extracting similar data for other system sizes, we can plug their limiting (ÎČâ†’âˆžâ†’đ›œ\beta\to\inftyitalic_ÎČ â†’ ∞) values into the Huse scaling Huse (1988),

PsL2=|Δ0|2+BL,L≫1,T→0,formulae-sequencesubscript𝑃𝑠superscript𝐿2superscriptsubscriptΔ02đ”đżformulae-sequencemuch-greater-than𝐿1→𝑇0\frac{P_{s}}{L^{2}}=|\Delta_{0}|^{2}+\frac{B}{L},\ \ \ L\gg 1,\ T\to 0,divide start_ARG italic_P start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_ARG start_ARG italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG = | roman_Δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + divide start_ARG italic_B end_ARG start_ARG italic_L end_ARG , italic_L ≫ 1 , italic_T → 0 , (19)

to obtain the separate contributions to the superconducting gap at zero temperature Moreo and Scalapino (1991); Paiva et al. (2004); Bđ”Bitalic_B is a non-universal constant.

Figure 8(a) illustrates the extrapolation to L→∞→𝐿L\to\inftyitalic_L → ∞ in a scaling plot. Interestingly, when we plot the separate gaps as functions of tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT in Fig. 8(b) for fixed U𝑈Uitalic_U and ⟹n⟩delimited-âŸšâŸ©đ‘›\langle n\rangle⟹ italic_n ⟩, we see that the intralayer contribution initially decreases before stabilizing in a value smaller than that for a monolayer. By contrast, the interlayer contribution rises from zero and stabilizes at roughly the same value as the one for the monolayer. Equation (19) allows us to add the contributions to the total |Δ0|2superscriptsubscriptΔ02|\Delta_{0}|^{2}| roman_Δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, and Fig. 8(b) shows that the latter displays a maximum near the isotropic limit, tz=tsubscript𝑡𝑧𝑡t_{z}=titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = italic_t, similarly to Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. Thus, the key point is that the correlations between layers lead to a stronger order parameter than in a monolayer, which, in turn, results in higher critical temperatures.

III.2 Pairing temperature scale

Refer to caption
Figure 9: The uniform spin susceptibility for the bilayer, as a function of temperature for ⟹n⟩=0.87delimited-âŸšâŸ©đ‘›0.87\langle n\rangle=0.87⟹ italic_n ⟩ = 0.87, |U|/t=5.0𝑈𝑡5.0|U|/t=5.0| italic_U | / italic_t = 5.0, and different values of the interplane hopping tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT; the linear lattice size is L=16𝐿16L=16italic_L = 16. Dashed lines are guides to the eye, and the vertical lines indicate the temperatures at which the downturn in χssubscript𝜒𝑠\chi_{s}italic_χ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT occurs.
Refer to caption
Figure 10: Critical (Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT) and pairing (Tpsubscript𝑇𝑝T_{p}italic_T start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT) temperatures (in units of t𝑡titalic_t) as functions of tz/tsubscript𝑡𝑧𝑡t_{z}/titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT / italic_t, obtained from our DQMC simulations for a bilayer with linear size L=16𝐿16L=16italic_L = 16, and (a) ⟹n⟩=0.87delimited-âŸšâŸ©đ‘›0.87\langle n\rangle=0.87⟹ italic_n ⟩ = 0.87; (b)⟹n⟩=0.5delimited-âŸšâŸ©đ‘›0.5\langle n\rangle=0.5⟹ italic_n ⟩ = 0.5. Lines are guides to the eye.
Refer to caption
Figure 11: Snapshots of the double occupancy, d𝐱subscript𝑑𝐱d_{\mathbf{i}}italic_d start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT, throughout one of the layers. Labels of horizontal and vertical axes denote site coordinates on an L=8𝐿8L=8italic_L = 8 layer, for ⟹n⟩=0.87delimited-âŸšâŸ©đ‘›0.87\langle n\rangle=0.87⟹ italic_n ⟩ = 0.87, U/t=−5𝑈𝑡5U/t=-5italic_U / italic_t = - 5 and tz=tsubscript𝑡𝑧𝑡t_{z}=titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = italic_t. The rows show typical distributions at different temperatures. The calculated double occupancy for each realization appears on top of each panel, and the color map on the right hand side shows the scale for d𝐱subscript𝑑𝐱d_{\mathbf{i}}italic_d start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT.
Refer to caption
Figure 12: Same as Fig. 2, but for a simple cubic lattice: (a) ⟹n⟩=1.0delimited-âŸšâŸ©đ‘›1.0\langle n\rangle=1.0⟹ italic_n ⟩ = 1.0 and (b) ⟹n⟩=0.87delimited-âŸšâŸ©đ‘›0.87\langle n\rangle=0.87⟹ italic_n ⟩ = 0.87.
Refer to caption
Figure 13: Log-linear plot of the pairing correlation function vs. distance along the diagonal of an 8×8×88888\times 8\times 88 × 8 × 8 cubic lattice, for different inverse temperatures, ÎČ⁹tđ›œđ‘Ą\beta titalic_ÎČ italic_t, for a half-filled band, isotropic hopping tz/t=1subscript𝑡𝑧𝑡1t_{z}/t=1italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT / italic_t = 1, and U/t=−8.0𝑈𝑡8.0U/t=-8.0italic_U / italic_t = - 8.0. Due to periodic boundary conditions, the farthest distance is 4⁹a⁹34𝑎34a\sqrt{3}4 italic_a square-root start_ARG 3 end_ARG, with a𝑎aitalic_a being the lattice spacing.
Refer to caption
Figure 14: Correlation ratio as a function of the inverse temperature, ÎČ⁹tđ›œđ‘Ą\beta titalic_ÎČ italic_t, for different linear cubic lattice sizes.
Refer to caption
Figure 15: DQMC estimates for the 3D critical temperature, Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, as a function of tz/tsubscript𝑡𝑧𝑡t_{z}/titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT / italic_t: (red) diamonds and (black) pentagons correspond to U=−5ⁱt𝑈5𝑡U=-5titalic_U = - 5 italic_t and U=−8ⁱt𝑈8𝑡U=-8titalic_U = - 8 italic_t, respectively; scaled data for TcBCSsuperscriptsubscript𝑇𝑐BCST_{c}^{\text{BCS}}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT BCS end_POSTSUPERSCRIPT, Eq. (1), for U=−5ⁱt𝑈5𝑡U=-5titalic_U = - 5 italic_t, are shown as (green) circles. We have arbitrarily divided TcBCSsuperscriptsubscript𝑇𝑐BCST_{c}^{\text{BCS}}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT BCS end_POSTSUPERSCRIPT by 20 just to lie closer to the other curves. The panels correspond to the densities shown. All lines are guides to the eye.
Refer to caption
Figure 16: Critical, Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, and pairing, Tpsubscript𝑇𝑝T_{p}italic_T start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT, temperatures (in units of t𝑡titalic_t) as functions of tz/tsubscript𝑡𝑧𝑡t_{z}/titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT / italic_t, obtained from our DQMC simulations for a cubic lattice with linear lattice size L=8𝐿8L=8italic_L = 8, at (a) half filling and (b) for ⟹n⟩=0.87delimited-âŸšâŸ©đ‘›0.87\langle n\rangle=0.87⟹ italic_n ⟩ = 0.87, for |U|/t=8.0𝑈𝑡8.0|U|/t=8.0| italic_U | / italic_t = 8.0. Dashed lines are guides to the eye.

As mentioned in the Introduction, one of the interesting features of the AHM is the emergence of a pairing temperature scale above Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT which signals the formation of pairs, but not their coherence. Our purpose here is to investigate how the addition of another layer affects the pairing temperature. One manifestation of pair formation is through a gap opening for spin excitations, hence as a downturn in the uniform magnetic susceptibility χssubscript𝜒𝑠\chi_{s}italic_χ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT as the temperature is lowered Micnas et al. (1990); Randeria et al. (1992); dos Santos (1994); WlazƂowski et al. (2013); Tajima et al. (2014); Fontenele et al. (2022).

Figure 9 shows our typical DQMC data for the temperature dependence of the uniform susceptibility, χssubscript𝜒𝑠\chi_{s}italic_χ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT [see Eq. (7)], for a bilayer with |U|/t=5.0𝑈𝑡5.0|U|/t=5.0| italic_U | / italic_t = 5.0 and ⟹n⟩=0.87delimited-âŸšâŸ©đ‘›0.87\langle n\rangle=0.87⟹ italic_n ⟩ = 0.87, for different interplane hoppings; the linear lattice size is kept fixed at L=16𝐿16L=16italic_L = 16. We note that, for fixed U𝑈Uitalic_U, χssubscript𝜒𝑠\chi_{s}italic_χ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT drops steadily below some temperature, whose location depends on tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT. Following the idea that this drop signals the formation of local pairs within some temperature scale Randeria et al. (1992); dos Santos (1994); WlazƂowski et al. (2013); Tajima et al. (2014), we adopt the position of the maximum in χssubscript𝜒𝑠\chi_{s}italic_χ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT as the pairing scale, Tpsubscript𝑇𝑝T_{p}italic_T start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT. As Fig. 9 shows, for strong couplings the maxima can be quite broad, so that their position is determined by inspection of the actual numerical output for χssubscript𝜒𝑠\chi_{s}italic_χ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, taking into account its error bars; this yields a range of temperatures within which the maximum lies. This somewhat flexible definition is in line with the fact that we are dealing with a crossover, hence a temperature scale, not with a sharp transition.

Estimates for Tpsubscript𝑇𝑝T_{p}italic_T start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT as a function of tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT are shown in Fig. 10 for two different densities; for comparison, data for Tcⁱ(tz)subscript𝑇𝑐subscript𝑡𝑧T_{c}(t_{z})italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) are also shown. We see that the overall tendency of Tpsubscript𝑇𝑝T_{p}italic_T start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT is to decrease very slowly with tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, while, as mentioned before, Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT displays a broad maximum around tz=tsubscript𝑡𝑧𝑡t_{z}=titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = italic_t. From this we may conclude that, to some extent, interlayer hopping favors pair breaking if they are not part of a condensate, but suppresses pair breaking if in a condensate.

We have also studied the distribution of doubly occupied sites, through

d𝐱≡⟹n𝐱↑ⁱnđąâ†“âŸ©,subscript𝑑𝐱delimited-⟚⟩subscript𝑛↑𝐱absentsubscript𝑛↓𝐱absentd_{\mathbf{i}}\equiv\langle n_{\mathbf{i}\uparrow}n_{\mathbf{i}\downarrow}\rangle,italic_d start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT ≡ ⟹ italic_n start_POSTSUBSCRIPT bold_i ↑ end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT bold_i ↓ end_POSTSUBSCRIPT ⟩ , (20)

and its average over the lattice sites of the bilayer,

⟹dâŸ©â‰ĄâŸšn↑ⁱnâ†“âŸ©â‰Ą1Nsⁱ∑𝐱d𝐱,delimited-âŸšâŸ©đ‘‘delimited-⟚⟩subscript𝑛↑subscript𝑛↓1subscript𝑁𝑠subscript𝐱subscript𝑑𝐱\langle d\rangle\equiv\langle n_{\uparrow}n_{\downarrow}\rangle\equiv\frac{1}{% N_{s}}\sum_{\mathbf{i}}d_{\mathbf{i}},⟹ italic_d ⟩ ≡ ⟹ italic_n start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT ⟩ ≡ divide start_ARG 1 end_ARG start_ARG italic_N start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT italic_d start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT , (21)

with Ns=2⁹L2subscript𝑁𝑠2superscript𝐿2N_{s}=2L^{2}italic_N start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 2 italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. Figure 11 displays snapshots of d𝐱subscript𝑑𝐱d_{\mathbf{i}}italic_d start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT for ⟹n⟩=0.87delimited-âŸšâŸ©đ‘›0.87\langle n\rangle=0.87⟹ italic_n ⟩ = 0.87 and |U|/t=5.0𝑈𝑡5.0|U|/t=5.0| italic_U | / italic_t = 5.0 for an L=8𝐿8L=8italic_L = 8 bilayer with tz=tsubscript𝑡𝑧𝑡t_{z}=titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = italic_t within three ranges of temperature, namely the metallic phase, T>Tp𝑇subscript𝑇𝑝T>T_{p}italic_T > italic_T start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT, the pseudogap regime, Tc<T<Tpsubscript𝑇𝑐𝑇subscript𝑇𝑝T_{c}<T<T_{p}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT < italic_T < italic_T start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT, and the superconducting phase, T<Tc𝑇subscript𝑇𝑐T<T_{c}italic_T < italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. As expected, we see that ⟹d⟩delimited-âŸšâŸ©đ‘‘\langle d\rangle⟹ italic_d ⟩ typically increases as the temperature is lowered, with the boost being more pronounced as the pairs condense into the superconducting phase. We also see that there is no significant change in the range of values of ⟹d⟩delimited-âŸšâŸ©đ‘‘\langle d\rangle⟹ italic_d ⟩. By contrast, the distribution of doublons seems to follow a more pronounced checkerboard pattern in the latter case. In the superconducting phase the checkerboard pattern is even more pronounced than in the pseudogap phase, in the sense that one finds sites with large ⟹dđąâŸ©delimited-⟚⟩subscript𝑑𝐱\langle d_{\mathbf{i}}\rangle⟹ italic_d start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT ⟩ surrounded by sites with small ⟹dđąâŸ©delimited-⟚⟩subscript𝑑𝐱\langle d_{\mathbf{i}}\rangle⟹ italic_d start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT ⟩. One may attribute this to the presence of more tightly bound pairs.

IV The Simple Cubic lattice

Having established the quantitative influence of hopping to another layer on the critical temperature, it is instructive to compare with the behavior of the AHM on a simple cubic lattice, but with a variable hopping along one of the directions, say the z𝑧zitalic_z direction, as in the bilayer. Recalling that now Tc≠0subscript𝑇𝑐0T_{c}\neq 0italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≠ 0 for tz≠0subscript𝑡𝑧0t_{z}\neq 0italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ≠ 0 at half filling, we will discuss this case, in addition to ⟹n⟩=0.87delimited-âŸšâŸ©đ‘›0.87\langle n\rangle=0.87⟹ italic_n ⟩ = 0.87, which is where the maximum Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT occurs for the two-dimensional case.

Starting with the non-interacting case, Fig. 12 shows the density of states for different values of tz/tsubscript𝑡𝑧𝑡t_{z}/titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT / italic_t, and for the two fillings mentioned above. We see how the van Hove singularity at the two-dimensional particle-hole symmetry point evolves to the typical three-dimensional DOS. As a consequence, the DOS at the corresponding Fermi energies monotonically decrease with tz/tsubscript𝑡𝑧𝑡t_{z}/titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT / italic_t; see the insets in each case.

Figure 13 shows the decay of pairing correlations with the distance for the cubic lattice at half filling. It is interesting to note that the length scale of decay increases by two orders of magnitude as the temperature is lowered within the interval shown, signalling the onset of superconducting long-range order. In order to estimate Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT through our QMC simulations, we adopt a procedure similar to that for the bilayer, namely the crossing of the correlation ratio, Eq. (6). Typical results are shown in Fig. 14, and by drawing similar plots for other values of U𝑈Uitalic_U, tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, and ⟹n⟩delimited-âŸšâŸ©đ‘›\langle n\rangle⟹ italic_n ⟩ one determines Tc⁹(tz/t)subscript𝑇𝑐subscript𝑡𝑧𝑡T_{c}(t_{z}/t)italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT / italic_t ) as displayed in Fig. 15. We see that for U=−5⁹t𝑈5𝑡U=-5titalic_U = - 5 italic_t the maximum Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT occurs at tz=tsubscript𝑡𝑧𝑡t_{z}=titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = italic_t for both fillings; therefore, in this case varying tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT away from the isotropic limit does not lead to higher Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT’s. By contrast, for U=−8⁹t𝑈8𝑡U=-8titalic_U = - 8 italic_t the maximum Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT occurs at larger values of tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT as the filling is decreased: there is an increase of Tc/t=0.34±0.01subscript𝑇𝑐𝑡plus-or-minus0.340.01T_{c}/t=0.34\pm 0.01italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_t = 0.34 ± 0.01, for tz=tsubscript𝑡𝑧𝑡t_{z}=titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = italic_t, from Tc/t=0.40±0.02subscript𝑇𝑐𝑡plus-or-minus0.400.02T_{c}/t=0.40\pm 0.02italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_t = 0.40 ± 0.02 for tz=2⁹tsubscript𝑡𝑧2𝑡t_{z}=2titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 2 italic_t, which is about 16% over the value for the isotropic case. This may be attributed to the fact that for stronger couplings the pairs are more strongly bound, hence more resistant to the suppressing effects of channelling along the z𝑧zitalic_z-direction.

Figure 15 shows that the maximum increase in Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT as |U|𝑈|U|| italic_U | goes from 5⁹t5𝑡5t5 italic_t to 8⁹t8𝑡8t8 italic_t is of about 50%percent5050\%50 %, reachable when tz>1.5⁹tsubscript𝑡𝑧1.5𝑡t_{z}>1.5titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT > 1.5 italic_t. In addition, Fig. 15 also shows an interesting feature: a crossing of the Tc⁹(tz/t)subscript𝑇𝑐subscript𝑡𝑧𝑡T_{c}(t_{z}/t)italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT / italic_t ) curves for small values of tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT. In order to understand this, let us focus on ⟹n⟩=0.87delimited-âŸšâŸ©đ‘›0.87\langle n\rangle=0.87⟹ italic_n ⟩ = 0.87. For the square lattice the maximum Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT occurs for U=−5⁹t𝑈5𝑡U=-5titalic_U = - 5 italic_t Fontenele et al. (2022), namely Tc≈0.22⁹tsubscript𝑇𝑐0.22𝑡T_{c}\approx 0.22titalic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≈ 0.22 italic_t, while for U=−8⁹t𝑈8𝑡U=-8titalic_U = - 8 italic_t, Tc≈0.16⁹tsubscript𝑇𝑐0.16𝑡T_{c}\approx 0.16titalic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≈ 0.16 italic_t; see data for tz=0subscript𝑡𝑧0t_{z}=0italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0 in Fig. 15(b). On the other hand, for the isotropic simple cubic lattice the maximum Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT occurs for |U|/t≳8greater-than-or-equivalent-to𝑈𝑡8|U|/t\gtrsim 8| italic_U | / italic_t ≳ 8, as it can be inferred from the data for ⟹n⟩=0.8delimited-âŸšâŸ©đ‘›0.8\langle n\rangle=0.8⟹ italic_n ⟩ = 0.8 in Ref. dos Santos (1994). The crossing is therefore a consequence of the dimensional crossover driven by tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT. In other words, the crossing at tz≈0.25⁹tsubscript𝑡𝑧0.25𝑡t_{z}\approx 0.25titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ≈ 0.25 italic_t for ⟹n⟩=0.87delimited-âŸšâŸ©đ‘›0.87\langle n\rangle=0.87⟹ italic_n ⟩ = 0.87 seems to separate two regimes: one in which interplane pairing correlations are significantly less relevant than the intraplane ones. Similar arguments should also account for the crossing at half filling, despite the fact that Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is suppressed to zero for tz=0subscript𝑡𝑧0t_{z}=0italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0.

Similarly to what we did for the two-dimensional geometries, we now probe the effectiveness of the BCS estimate for the anisotropic simple cubic lattice. Figure 15 compares the dependence of Tc/tsubscript𝑇𝑐𝑡T_{c}/titalic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_t and TcBCS/Wsuperscriptsubscript𝑇𝑐BCS𝑊T_{c}^{\text{BCS}}/Witalic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT BCS end_POSTSUPERSCRIPT / italic_W with tz/tsubscript𝑡𝑧𝑡t_{z}/titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT / italic_t, for fixed U=−5ⁱt𝑈5𝑡U=-5titalic_U = - 5 italic_t, and for two densities; W=8ⁱt+4ⁱtz𝑊8𝑡4subscript𝑡𝑧W=8t+4t_{z}italic_W = 8 italic_t + 4 italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT is the bandwidth. For half filling, TcBCSsuperscriptsubscript𝑇𝑐BCST_{c}^{\text{BCS}}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT BCS end_POSTSUPERSCRIPT does not satisfy the Mermin-Wagner theorem for tz=0subscript𝑡𝑧0t_{z}=0italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0, as expected. In addition, one sees that TcBCSsuperscriptsubscript𝑇𝑐BCST_{c}^{\text{BCS}}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT BCS end_POSTSUPERSCRIPT does not track our DQMC estimates for Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT for both fillings.

Let us now discuss how the pairing temperature is affected by a variable hopping along the z𝑧zitalic_z-direction. Our DQMC data are shown in Fig. 16 for fixed U=−8⁹t𝑈8𝑡U=-8titalic_U = - 8 italic_t, and for different fillings; Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is also plotted for comparison. We see that Tpsubscript𝑇𝑝T_{p}italic_T start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT decreases much more steadily with tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT than for the bilayer; this leads one to conclude that confinement along the z𝑧zitalic_z-direction favors pair formation, but not necessarily pair condensation. The data also show that the behavior of Tpsubscript𝑇𝑝T_{p}italic_T start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT with tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT is not too dependent on ⟹n⟩delimited-âŸšâŸ©đ‘›\langle n\rangle⟹ italic_n ⟩ within the range considered.

And, finally, Fig. 17 compares the effect of anisotropy on Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT for both the bilayer and the simple cubic lattice. We see that away from half filling Tⁱc𝑇𝑐Tcitalic_T italic_c is about 30% higher for the 3D lattice.

Refer to caption
Figure 17: Comparison of our DQMC data for Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT on the bilayer and on the simple cubic lattice.

V Conclusions

We have examined layering as a possible route to increase the superconducting critical temperature in the attractive Hubbard model.

In two dimensions, layering consists of stacking two planes with hopping tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT between them; intra-plane hoppings and on-site attraction are the same in both planes. We have found that by a judicious choice of fillings and on-site attraction, a bilayer can exhibit critical temperatures between 1.5 and 1.7 times those of the single layer. We have also established that double occupancy can be used to distinguish between normal (metallic), pre-formed pairs (pseudogap), and superconducting phases through quantum gas microscope measurements.

For a simple cubic lattice, layering consists of allowing tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT to vary with respect to the isotropic case, tz=tsubscript𝑡𝑧𝑡t_{z}=titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = italic_t, but the number of planes along the z𝑧zitalic_z direction, Lzsubscript𝐿𝑧L_{z}italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, is the same as the number of sites, Lx=Lysubscriptđżđ‘„subscript𝐿𝑩L_{x}=L_{y}italic_L start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT = italic_L start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT, along the planar directions. We have found a smaller enhancement of Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT relative to the isotropic case, namely of 1.3 times larger, for some specific choices of tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, U𝑈Uitalic_U and ⟹n⟩delimited-âŸšâŸ©đ‘›\langle n\rangle⟹ italic_n ⟩.

For both geometries, we have found that the BCS estimate roughly tracks the DQMC Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT’s, although the positions of Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT maxima are not reliable; less serious is the multiplicative factor of order 20 for both geometries, since the unknown proportionality factor in Eq. (1) could account for it. In addition, since TcBCSsuperscriptsubscript𝑇𝑐BCST_{c}^{\text{BCS}}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT BCS end_POSTSUPERSCRIPT does not follow the Mermin-Wagner theorem this discrepancy at half-filling in 3D is even more serious for tzâ‰Č0.5⁹tless-than-or-similar-tosubscript𝑡𝑧0.5𝑡t_{z}\lesssim 0.5titalic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT â‰Č 0.5 italic_t. Also, care must be taken since the BCS estimate for Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT scales with the bandwidth, which, in turn, depends on tzsubscript𝑡𝑧t_{z}italic_t start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT; this leads to more severe discrepancies in bilayer than in 3D.

We hope our results will stimulate optical lattice experiments to explore this route of layering to further increase Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT.

ACKNOWLEDGMENTS

The authors are grateful to the Brazilian Agencies Conselho Nacional de Desenvolvimento CientĂ­fico e TecnolĂłgico (CNPq), Coordenação de Aperfeiçoamento de Pessoal de Ensino Superior (CAPES), and Instituto Nacional de CiĂȘncia e Tecnologia de Informação QuĂąntica (INCT-IQ) for funding this project. We also gratefully acknowledge support from Fundação Carlos Chagas de Apoio Ă  Pesquisa (FAPERJ), through the grants E-26/200.258/2023 (N.C.C.), E-26/210.974/2024 (R.R.d.S.), E-26/200.959/2022 (T.P.), and E-26/210.100/2023 (T.P.). N.C.C. also acknowledges CNPq, Grant No. 313065/2021-7.

References

  • Micnas et al. (1990) R. Micnas, J. Ranninger,  and S. Robaszkiewicz, “Superconductivity in narrow-band systems with local nonretarded attractive interactions,” Rev. Mod. Phys. 62, 113–171 (1990).
  • Randeria et al. (1992) Mohit Randeria, Nandini Trivedi, Adriana Moreo,  and Richard T. Scalettar, “Pairing and spin gap in the normal state of short coherence length superconductors,” Phys. Rev. Lett. 69, 2001–2004 (1992).
  • Bucher et al. (1993) B. Bucher, P. Steiner, J. Karpinski, E. Kaldis,  and P. Wachter, “Influence of the spin gap on the normal state transport in YBa2Cu4O8,” Phys. Rev. Lett. 70, 2012–2015 (1993).
  • dos Santos (1994) Raimundo R. dos Santos, “Spin gap and superconductivity in the three-dimensional attractive Hubbard model,” Phys. Rev. B 50, 635–638 (1994).
  • Paiva et al. (2010) Thereza Paiva, Richard Scalettar, Mohit Randeria,  and Nandini Trivedi, “Fermions in 2d optical lattices: Temperature and entropy scales for observing antiferromagnetism and superfluidity,” Phys. Rev. Lett. 104, 066406 (2010).
  • Fontenele et al. (2022) Rodrigo A. Fontenele, Natanael C. Costa, Raimundo R. dos Santos,  and Thereza Paiva, “Two-dimensional attractive Hubbard model and the BCS-BEC crossover,” Phys. Rev. B 105, 184502 (2022).
  • Wilson (2001) John A. Wilson, “Developments in the negative-U modelling of the cuprate HTSC systems,” Journal of Physics: Condensed Matter 13, R945–R977 (2001).
  • Randeria (1995) Mohit Randeria, “Crossover from BCS theory to Bose–Einstein condensation,” in Bose-Einstein Condensation, edited by A. Griffin, D. W. Snoke,  and S.Editors Stringari (Cambridge University Press, 1995) p. 355–392.
  • Chen et al. (2005) Qijin Chen, Jelena Stajic, Shina Tan,  and K. Levin, “Bcs–bec crossover: From high temperature superconductors to ultracold superfluids,” Physics Reports 412, 1–88 (2005).
  • Randeria and Taylor (2014) Mohit Randeria and Edward Taylor, “Crossover from Bardeen-Cooper-Schrieffer to Bose-Einstein condensation and the unitary Fermi gas,” Annual Review of Condensed Matter Physics 5, 209–232 (2014).
  • Chin et al. (2010) Cheng Chin, Rudolf Grimm, Paul Julienne,  and Eite Tiesinga, “Feshbach resonances in ultracold gases,” Rev. Mod. Phys. 82, 1225–1286 (2010).
  • Jaksch and Zoller (2005) D. Jaksch and P. Zoller, “The cold atom Hubbard toolbox,” Ann. Phys. 315, 52 (2005).
  • Bloch et al. (2008) I. Bloch, J. Dalibard,  and W. Zwerger, “Many-body physics with ultracold gases,” Rev. Mod. Phys. 80, 885 (2008).
  • Esslinger (2010) T. Esslinger, “Fermi-Hubbard physics with atoms in an optical lattice,” Annu. Rev. Condens. Matter Phys. 1, 129–152 (2010).
  • McKay and DeMarco (2011) D. C. McKay and B. DeMarco, “Cooling in strongly correlated optical lattices: prospects and challenges,” Rep. Prog. Phys. 74, 054401 (2011).
  • Mitra et al. (2018) Debayan Mitra, Peter T. Brown, Elmer Guardado-Sanchez, Stanimir S. Kondov, Trithep Devakul, David A. Huse, Peter Schauss,  and Waseem S. Bakr, “Quantum gas microscopy of an attractive Fermi-Hubbard system,” Nature Physics 14, 173–177 (2018).
  • Jördens et al. (2010) R. Jördens, L. Tarruell, D. Greif, T. Uehlinger, N. Strohmaier, H. Moritz, T. Esslinger, L. De Leo, C. Kollath, A. Georges, V. Scarola, L. Pollet, E. Burovski, E. Kozik,  and M. Troyer, “Quantitative determination of temperature in the approach to magnetic order of ultracold fermions in an optical lattice,” Phys. Rev. Lett. 104, 180401 (2010).
  • Bakr et al. (2009) Waseem S. Bakr, Jonathon I. Gillen, Amy Peng, Simon Fölling,  and Markus Greiner, “A quantum gas microscope for detecting single atoms in a Hubbard-regime optical lattice,” Nature 462, 74–77 (2009).
  • Gall et al. (2020) M. Gall, C. F. Chan, N. Wurz,  and M. Köhl, “Simulating a mott insulator using attractive interaction,” Phys. Rev. Lett. 124, 010403 (2020).
  • Chan et al. (2020) C. F. Chan, M. Gall, N. Wurz,  and M. Köhl, “Pair correlations in the attractive Hubbard model,” Phys. Rev. Res. 2, 023210 (2020).
  • Hartke et al. (2023) Thomas Hartke, Botond Oreg, Carter Turnbaugh, Ningyuan Jia,  and Martin Zwierlein, “Direct observation of nonlocal fermion pairing in an attractive Fermi-Hubbard gas,” Science 381, 82–86 (2023).
  • Scalettar et al. (1994) Richard T. Scalettar, Joel W. Cannon, Douglas J. Scalapino,  and Robert L. Sugar, “Magnetic and pairing correlations in coupled Hubbard planes,” Phys. Rev. B 50, 13419–13427 (1994).
  • dos Santos (1995) Raimundo R. dos Santos, “Magnetism and pairing in hubbard bilayers,” Phys. Rev. B 51, 15540–15546 (1995).
  • Zheng et al. (2023) Qinpei Zheng, Yuqing Wang, Libo Liang, Qi Huang, Shuai Wang, Wei Xiong, Xiaoji Zhou, Wenlan Chen, Xuzong Chen,  and Jiazhong Hu, “Dimensional crossover of quantum critical dynamics in many-body phase transitions,” Phys. Rev. Res. 5, 013136 (2023).
  • Ibarra-GarcĂ­a-Padilla et al. (2020) Eduardo Ibarra-GarcĂ­a-Padilla, Rick Mukherjee, Randall G. Hulet, Kaden R. A. Hazzard, Thereza Paiva,  and Richard T. Scalettar, “Thermodynamics and magnetism in the two-dimensional to three-dimensional crossover of the Hubbard model,” Phys. Rev. A 102, 033340 (2020).
  • Ogawa et al. (2015) Naoki Ogawa, Tatsuya Kaneko,  and Yukinori Ohta, “Dimensional crossover of the superfluid state in the attractive Hubbard model,” Journal of Physics: Conference Series 592, 012103 (2015).
  • Blankenbecler et al. (1981) R. Blankenbecler, D. J. Scalapino,  and R. L. Sugar, “Monte carlo calculations of coupled boson-fermion systems. I,” Phys. Rev. D 24, 2278–2286 (1981).
  • Hirsch (1983) J. E Hirsch, “Discrete Hubbard-Stratonovich transformation for fermion lattice models,” Phys. Rev. B 28, 4059 (1983).
  • Hirsch (1985) J. E. Hirsch, “Two-dimensional Hubbard model: Numerical simulation study,” Phys. Rev. B 31, 4403–4419 (1985).
  • White et al. (1989) S. R. White, D. J. Scalapino, R. L. Sugar, E. Y. Loh, J. E. Gubernatis,  and R. T. Scalettar, “Numerical study of the two-dimensional Hubbard model,” Phys. Rev. B 40, 506–516 (1989).
  • dos Santos (2003) Raimundo R. dos Santos, “Introduction to quantum Monte Carlo simulations for fermionic systems,” Braz. J. Phys. 33, 36 – 54 (2003).
  • Kaul (2015) Ribhu K. Kaul, “Spin nematics, valence-bond solids, and spin liquids in SO⁹(n)SO𝑛\mathrm{SO}(n)roman_SO ( italic_n ) quantum spin models on the triangular lattice,” Phys. Rev. Lett. 115, 157202 (2015).
  • Sato et al. (2018) Toshihiro Sato, Fakher F. Assaad,  and Tarun Grover, “Quantum Monte Carlo simulation of frustrated Kondo lattice models,” Phys. Rev. Lett. 120, 107201 (2018).
  • Darmawan et al. (2018) Andrew S. Darmawan, Yusuke Nomura, Youhei Yamaji,  and Masatoshi Imada, “Stripe and superconducting order competing in the Hubbard model on a square lattice studied by a combined variational Monte Carlo and tensor network method,” Phys. Rev. B 98, 205132 (2018).
  • Zujev et al. (2014) Aleksander Zujev, Richard T Scalettar, George G Batrouni,  and Pinaki Sengupta, “Pairing correlations in the two-layer attractive Hubbard model,” New Journal of Physics 16, 013004 (2014).
  • Hazra et al. (2019) Tamaghna Hazra, Nishchhal Verma,  and Mohit Randeria, “Bounds on the superconducting transition temperature: Applications to twisted bilayer graphene and cold atoms,” Phys. Rev. X 9, 031049 (2019).
  • Scalapino et al. (1992) D. J. Scalapino, S. R. White,  and S. C. Zhang, “Superfluid density and the Drude weight of the Hubbard model,” Phys. Rev. Lett. 68, 2830–2833 (1992).
  • Scalapino et al. (1993) Douglas J. Scalapino, Steven R. White,  and Shoucheng Zhang, “Insulator, metal, or superconductor: The criteria,” Phys. Rev. B 47, 7995–8007 (1993).
  • Nelson and Kosterlitz (1977) David R. Nelson and J. M. Kosterlitz, “Universal jump in the superfluid density of two-dimensional superfluids,” Phys. Rev. Lett. 39, 1201–1205 (1977).
  • Paiva et al. (2004) Thereza Paiva, Raimundo R. dos Santos, R. T. Scalettar,  and P. J. H. Denteneer, “Critical temperature for the two-dimensional attractive Hubbard model,” Phys. Rev. B 69, 184501 (2004).
  • Huse (1988) David A. Huse, “Ground-state staggered magnetization of two-dimensional quantum heisenberg antiferromagnets,” Phys. Rev. B 37, 2380–2382 (1988).
  • Moreo and Scalapino (1991) A. Moreo and D. J. Scalapino, “Two-dimensional negative-U Hubbard model,” Phys. Rev. Lett. 66, 946–948 (1991).
  • WlazƂowski et al. (2013) Gabriel WlazƂowski, Piotr Magierski, JoaquĂ­n E. Drut, Aurel Bulgac,  and Kenneth J. Roche, “Cooper pairing above the critical temperature in a unitary fermi gas,” Phys. Rev. Lett. 110, 090401 (2013).
  • Tajima et al. (2014) Hiroyuki Tajima, Takashi Kashimura, Ryo Hanai, Ryota Watanabe,  and Yoji Ohashi, “Uniform spin susceptibility and spin-gap phenomenon in the BCS-BEC-crossover regime of an ultracold Fermi gas,” Phys. Rev. A 89, 033617 (2014).