Pairing by disorder of dopants in a magnetic spin ladder

K. Knakkergaard Nielsen Max-Planck Institute for Quantum Optics, Hans-Kopfermann-Str. 1, D-85748 Garching, Germany
(July 1, 2024)
Abstract

I demonstrate a pairing mechanism of dopants in a magnetic lattice, emerging from the underlying high-temperature disorder of the spins. The effect is demonstrated in a mixed-dimensional model, where dopants travel along a two-leg ladder, while the spins feature Ising interactions along the rungs and the legs of the ladder. The dynamics following the sudden immersion of two nearest neighbor dopants in an infinite temperature spin lattice shows that thermal spin disorder frustrates the relative motion of the dopants and enforces them to co-propagate. The predictions are shown to be realistically testable in quantum simulation experiments.

Understanding pairing of dopants in magnetic lattices is an immense challenge in modern quantum many-body theory. The main goal is to gain a deeper understanding of strongly correlated systems, arising e.g. in the notoriously difficult high-temperature superconductors Schrieffer and Brooks (2007); Trugman (1988); Dagotto (1994). While solid state experiments continue to pursue such inquiries after decades of research O’Mahony et al. (2022), a highly interesting and promising pathway has appeared in quantum simulation experiments with ultracold atoms in optical lattices Gross and Bloch (2017); Ji et al. (2021); Wang et al. (2021); Prichard et al. (2023); Koepsell et al. (2019, 2021); Lebrat et al. (2024); Hirthe et al. (2023). Single-site resolution capabilities Bakr et al. (2009); Sherson et al. (2010) makes it possible to measure low-temperature spatial correlations Koepsell et al. (2019, 2021); Prichard et al. (2023); Lebrat et al. (2024), as well as track them over time by repeated realizations of the same initial state. In particular, single hole dynamics consistent with magnetic polaron formation Nielsen et al. (2022) in quantum antiferromagnets has been measured Ji et al. (2021). Moreover, in such experiments one can tinker with the simulated models, otherwise fixed in the solid state. This has enabled the observation of dopant pairing in a mixed-dimensional magnetic lattice Hirthe et al. (2023). This has also been seen to be a plausible explanation Schlömer et al. (2023) for high-temperature superconductivity in a specific nickelate compound Sun et al. (2023). More broadly speaking, the appearance of pairing is conventionally tied to the distortion of some underlying order. In the aforementioned example, it is the distortion of the underlying AFM order that gives rise to the induced attraction Bohrdt et al. (2022); Nielsen (2023), which is then further enhanced by restricting motion to one dimension. In conventional superconductors, electrons distort the underlying crystal lattice, inducing an attraction between them Fröhlich (1950); Bardeen et al. (1957); Bruus and Flensberg (2016). In materials supporting high-temperature superconductivity, there may be an important counterexample to this generic scenario, in which preformed Cooper pairs Geshkenbein et al. (1997); Randeria et al. (1992) appearing above the superconducting phase transition might explain Li et al. (2010); Seo et al. (2019) the pseudogap phase Stanescu and Phillips (2003).

Refer to caption
Figure 1: Pairing dynamics. (a) Setup: holes move in a magnetic (red: \!\!\uparrow, blue: \!\!\downarrow) two-leg ladder with Ising interactions J,Jsubscript𝐽perpendicular-tosubscript𝐽parallel-toJ_{\perp},J_{\parallel}italic_J start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT , italic_J start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT via nearest-neighbor hop** t𝑡titalic_t along the ladder. (b) Hole-hole correlator Chh(x1,x2;τ)=n^h(x1)n^h(x2)τsubscript𝐶hhsubscript𝑥1subscript𝑥2𝜏subscriptexpectationsubscript^𝑛subscript𝑥1subscript^𝑛subscript𝑥2𝜏C_{\rm hh}(x_{1},x_{2};\tau)=\braket{\hat{n}_{h}(x_{1})\hat{n}_{h}(x_{2})}_{\tau}italic_C start_POSTSUBSCRIPT roman_hh end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ; italic_τ ) = ⟨ start_ARG over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) end_ARG ⟩ start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT for J=0subscript𝐽parallel-to0J_{\parallel}=0italic_J start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT = 0 at indicated times normalized to its maximum after the holes are initialized on the same rung, (l,x1)=(1,0),(l,x2)=(2,0)formulae-sequence𝑙subscript𝑥110𝑙subscript𝑥220(l,x_{1})=(1,0),(l,x_{2})=(2,0)( italic_l , italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) = ( 1 , 0 ) , ( italic_l , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) = ( 2 , 0 ). (c) Dynamics of the root-mean-square dynamics of the center-of-mass X=(x1+x2)/2𝑋subscript𝑥1subscript𝑥22X=(x_{1}+x_{2})/2italic_X = ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) / 2 and relative Δ=(x1x2)/2Δsubscript𝑥1subscript𝑥22\Delta=(x_{1}-x_{2})/2roman_Δ = ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) / 2 coordinates, (d) correlation coefficient corr(x1,x2)=x1x2τ/[σ(x1)σ(x2)]corrsubscript𝑥1subscript𝑥2subscriptexpectationsubscript𝑥1subscript𝑥2𝜏delimited-[]𝜎subscript𝑥1𝜎subscript𝑥2{\rm corr}(x_{1},x_{2})=\braket{x_{1}x_{2}}_{\tau}/[\sigma(x_{1})\sigma(x_{2})]roman_corr ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) = ⟨ start_ARG italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ⟩ start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT / [ italic_σ ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) italic_σ ( italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) ], and (e) relative correlation function for Δ=0Δ0\Delta=0roman_Δ = 0. Steady state values are shown in short dashed lines. In all: J=5tsubscript𝐽perpendicular-to5𝑡J_{\perp}=5titalic_J start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = 5 italic_t.

In this Letter, we demonstrate a pairing mechanism that does not rely on being in nor in the vicinity of a phase transition to an ordered phase. In particular, the discovered mechanism remains at infinite temperatures and is a direct consequence of the thermally disordered spin lattice. I analyze the effect in a simplistic setup of a mixed-dimensional model [Fig. 1(a)], where two holes are placed in two legs of a spin ladder. The spin degree of freedom is assumed to be at infinite temperature and completely disordered. Conventional wisdom, therefore, suggests that the holes should not be able to pair. Nevertheless, when the holes are initialized on the same rung in the infinite temperature spin lattice, the strong thermal disorder of the system frustrates the relative motion of the dopants. Indeed, tracking the hole-hole correlator in time [Fig. 1(b)], shows a significant correlation along the diagonal x1=x2subscript𝑥1subscript𝑥2x_{1}=x_{2}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, where the holes are on the same rung of the ladder. One may equivalently observe that the inter-hole distance is always smaller than the traversed distance of their center-of-mass [Fig. 1(c)]. This results in a strictly positive correlation between the holes’ position, signaling an effective attraction between them [Fig. 1(d)]. Finally, we observe that the probability of finding the holes on the same rung is easily increased by a factor of 4444 as compared to what one expects from their marginal distributions, described by the relative hole-hole correlator in Fig. 1(e). This phenomenon is closely tied to the recent discovery of thermally induced localization in such systems Nielsen (2024a, b). In particular, the holes experience an effective induced interaction that has a disordered character, encouraging the holes to co-propagate. I emphasize that the discovered effect would appear in a wide range of systems, though the robustness of the pairing in more generic setups is presently unclear.

Model.-

I consider spins in a two-leg ladder with Ising interactions and hop** along the ladder,

H^=𝐢,𝐣J𝐢𝐣2S^𝐢(z)S^𝐣(z)t𝐢,𝐣,σ[c~σ𝐢c~σ𝐣+c~σ𝐣c~σ𝐢],^𝐻subscript𝐢𝐣subscript𝐽𝐢𝐣2subscriptsuperscript^𝑆𝑧𝐢subscriptsuperscript^𝑆𝑧𝐣𝑡subscriptsubscriptexpectation𝐢𝐣parallel-to𝜎delimited-[]subscriptsuperscript~𝑐𝜎𝐢subscript~𝑐𝜎𝐣subscriptsuperscript~𝑐𝜎𝐣subscript~𝑐𝜎𝐢\displaystyle{\hat{H}}=\sum_{{\bf i},{\bf j}}\!\frac{J_{{\bf i}-{\bf j}}}{2}% \hat{S}^{(z)}_{\bf i}\hat{S}^{(z)}_{\bf j}-t\!\sum_{\braket{{\bf i},{\bf j}}_{% \parallel},\sigma}\!\!\!\left[\tilde{c}^{\dagger}_{\sigma{\bf i}}\tilde{c}_{% \sigma{\bf j}}+\tilde{c}^{\dagger}_{\sigma{\bf j}}\tilde{c}_{\sigma{\bf i}}% \right],over^ start_ARG italic_H end_ARG = ∑ start_POSTSUBSCRIPT bold_i , bold_j end_POSTSUBSCRIPT divide start_ARG italic_J start_POSTSUBSCRIPT bold_i - bold_j end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT ( italic_z ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT ( italic_z ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_j end_POSTSUBSCRIPT - italic_t ∑ start_POSTSUBSCRIPT ⟨ start_ARG bold_i , bold_j end_ARG ⟩ start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT , italic_σ end_POSTSUBSCRIPT [ over~ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_σ bold_i end_POSTSUBSCRIPT over~ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_σ bold_j end_POSTSUBSCRIPT + over~ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_σ bold_j end_POSTSUBSCRIPT over~ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_σ bold_i end_POSTSUBSCRIPT ] , (1)

i.e. a mixed-dimensional t𝑡titalic_t-Jzsubscript𝐽𝑧J_{z}italic_J start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT model Grusdt and Pollet (2020). I focus on the case of nearest-neighbor couplings along, J=J±𝐞xsubscript𝐽parallel-tosubscript𝐽plus-or-minussubscript𝐞𝑥J_{\parallel}=J_{\pm{\bf e}_{x}}italic_J start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT = italic_J start_POSTSUBSCRIPT ± bold_e start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUBSCRIPT, and across, J=J±𝐞ysubscript𝐽perpendicular-tosubscript𝐽plus-or-minussubscript𝐞𝑦J_{\perp}=J_{\pm{\bf e}_{y}}italic_J start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = italic_J start_POSTSUBSCRIPT ± bold_e start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT end_POSTSUBSCRIPT, the ladder, with S^𝐢(z)subscriptsuperscript^𝑆𝑧𝐢\hat{S}^{(z)}_{\bf i}over^ start_ARG italic_S end_ARG start_POSTSUPERSCRIPT ( italic_z ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT the z𝑧zitalic_z-projection of the spin operator 𝐒^𝐢subscript^𝐒𝐢\hat{\bf S}_{\bf i}over^ start_ARG bold_S end_ARG start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT. The hop**, H^tsubscript^𝐻𝑡{\hat{H}}_{t}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT, is only allowed along the ladder and only to vacant sites, enforced by the constrained operators c~σ𝐢=c^σ𝐢[1n^𝐢]subscriptsuperscript~𝑐𝜎𝐢subscriptsuperscript^𝑐𝜎𝐢delimited-[]1subscript^𝑛𝐢\tilde{c}^{\dagger}_{\sigma{\bf i}}=\hat{c}^{\dagger}_{\sigma{\bf i}}[1-\hat{n% }_{\bf i}]over~ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_σ bold_i end_POSTSUBSCRIPT = over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_σ bold_i end_POSTSUBSCRIPT [ 1 - over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT ], where n^𝐢=σn^σ𝐢=σc^σ𝐢c^σ𝐢subscript^𝑛𝐢subscript𝜎subscript^𝑛𝜎𝐢subscript𝜎subscriptsuperscript^𝑐𝜎𝐢subscript^𝑐𝜎𝐢\hat{n}_{\bf i}=\sum_{\sigma}\hat{n}_{\sigma{\bf i}}=\sum_{\sigma}\hat{c}^{% \dagger}_{\sigma{\bf i}}\hat{c}_{\sigma{\bf i}}over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_σ bold_i end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_σ bold_i end_POSTSUBSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_σ bold_i end_POSTSUBSCRIPT is the local density operator. The effective spin degree of freedom generally describes an internal degree of freedom with the Ising-type coupling in H^Jsubscript^𝐻𝐽{\hat{H}}_{J}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT, the first term in Eq. (1), and since there are no exchange processes of the dopants, both fermions and bosons can realize the model. To have an efficient description of the hole and spin degrees of freedom, we perform a Holstein-Primakoff transformation on top of the ferromagnetic ground state |FM=|\ket{{\rm FM}}=\ket{\cdots\uparrow\uparrow\cdots}| start_ARG roman_FM end_ARG ⟩ = | start_ARG ⋯ ↑ ↑ ⋯ end_ARG ⟩, with all spins in the |ket\ket{\uparrow}| start_ARG ↑ end_ARG ⟩ state. The spin Hamiltonian

H^J=𝐢,𝐣J𝐢𝐣2[12s^𝐢s^𝐢][12s^𝐣s^𝐣][1h^𝐢h^𝐢][1h^𝐣h^𝐣],subscript^𝐻𝐽subscript𝐢𝐣subscript𝐽𝐢𝐣2delimited-[]12subscriptsuperscript^𝑠𝐢subscript^𝑠𝐢delimited-[]12subscriptsuperscript^𝑠𝐣subscript^𝑠𝐣delimited-[]1superscriptsubscript^𝐢subscript^𝐢delimited-[]1superscriptsubscript^𝐣subscript^𝐣\displaystyle\!\!\!\hat{H}_{J}\!=\!\sum_{{\bf i},{\bf j}}\!\frac{J_{{\bf i}-{% \bf j}}}{2}\Big{[}\frac{1}{2}\!-\!\hat{s}^{\dagger}_{{\bf i}}\hat{s}_{{\bf i}}% \Big{]}\Big{[}\frac{1}{2}\!-\!\hat{s}^{\dagger}_{{\bf j}}\hat{s}_{{\bf j}}\Big% {]}\Big{[}1\!-\!\hat{h}_{{\bf i}}^{\dagger}\hat{h}_{{\bf i}}\Big{]}\Big{[}1\!-% \!\hat{h}_{{\bf j}}^{\dagger}\hat{h}_{{\bf j}}\Big{]},\!\!over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT bold_i , bold_j end_POSTSUBSCRIPT divide start_ARG italic_J start_POSTSUBSCRIPT bold_i - bold_j end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG [ divide start_ARG 1 end_ARG start_ARG 2 end_ARG - over^ start_ARG italic_s end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT over^ start_ARG italic_s end_ARG start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT ] [ divide start_ARG 1 end_ARG start_ARG 2 end_ARG - over^ start_ARG italic_s end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_j end_POSTSUBSCRIPT over^ start_ARG italic_s end_ARG start_POSTSUBSCRIPT bold_j end_POSTSUBSCRIPT ] [ 1 - over^ start_ARG italic_h end_ARG start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over^ start_ARG italic_h end_ARG start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT ] [ 1 - over^ start_ARG italic_h end_ARG start_POSTSUBSCRIPT bold_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over^ start_ARG italic_h end_ARG start_POSTSUBSCRIPT bold_j end_POSTSUBSCRIPT ] , (2)

hereby, describes the interaction of bosonic spin flip excitations s^𝐢subscriptsuperscript^𝑠𝐢\hat{s}^{\dagger}_{\bf i}over^ start_ARG italic_s end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT, corresponding to the creation of a spin-\downarrow on site 𝐢𝐢{\bf i}bold_i, and holes created by the operator h^𝐢subscriptsuperscript^𝐢\hat{h}^{\dagger}_{\bf i}over^ start_ARG italic_h end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT, which inherit the statistics of the underlying particles, be it fermionic or bosonic Nielsen (2023). Equivalently, the hop** Hamiltonian

H^t=t𝐢,𝐣subscript^𝐻𝑡𝑡subscriptsubscriptexpectation𝐢𝐣parallel-to\displaystyle{\hat{H}}_{t}=t\sum_{\braket{{\bf i},{\bf j}}_{\parallel}}\!over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = italic_t ∑ start_POSTSUBSCRIPT ⟨ start_ARG bold_i , bold_j end_ARG ⟩ start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_POSTSUBSCRIPT [h^𝐣F(h^𝐢,s^𝐢)F(h^𝐣,s^𝐣)h^𝐢\displaystyle\Big{[}\hat{h}^{\dagger}_{{\bf j}}F(\hat{h}_{{\bf i}},\hat{s}_{{% \bf i}})F(\hat{h}_{{\bf j}},\hat{s}_{{\bf j}})\hat{h}_{{\bf i}}[ over^ start_ARG italic_h end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_j end_POSTSUBSCRIPT italic_F ( over^ start_ARG italic_h end_ARG start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT , over^ start_ARG italic_s end_ARG start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT ) italic_F ( over^ start_ARG italic_h end_ARG start_POSTSUBSCRIPT bold_j end_POSTSUBSCRIPT , over^ start_ARG italic_s end_ARG start_POSTSUBSCRIPT bold_j end_POSTSUBSCRIPT ) over^ start_ARG italic_h end_ARG start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT
+h^𝐣s^𝐢F(h^𝐢,s^𝐢)F(h^𝐣,s^𝐣)s^𝐣h^𝐢]+H.c.\displaystyle+\hat{h}^{\dagger}_{{\bf j}}\hat{s}^{\dagger}_{\bf i}F(\hat{h}_{{% \bf i}},\hat{s}_{{\bf i}})F(\hat{h}_{{\bf j}},\hat{s}_{{\bf j}})\hat{s}_{\bf j% }\hat{h}_{{\bf i}}\Big{]}+{\rm H.c.}+ over^ start_ARG italic_h end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_j end_POSTSUBSCRIPT over^ start_ARG italic_s end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT italic_F ( over^ start_ARG italic_h end_ARG start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT , over^ start_ARG italic_s end_ARG start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT ) italic_F ( over^ start_ARG italic_h end_ARG start_POSTSUBSCRIPT bold_j end_POSTSUBSCRIPT , over^ start_ARG italic_s end_ARG start_POSTSUBSCRIPT bold_j end_POSTSUBSCRIPT ) over^ start_ARG italic_s end_ARG start_POSTSUBSCRIPT bold_j end_POSTSUBSCRIPT over^ start_ARG italic_h end_ARG start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT ] + roman_H . roman_c . (3)

describes two distinct ways in which holes may hop: (1) a hole hop** from site 𝐢𝐢{\bf i}bold_i to 𝐣𝐣{\bf j}bold_j in the absence of a spin flip on site 𝐣𝐣{\bf j}bold_j, and (2) hop** in the presence of a spin flip, whereby the hole and spin flip swap places. The constraint function F(h^,s^)=1s^s^h^h^𝐹^^𝑠1superscript^𝑠^𝑠superscript^^F(\hat{h},\hat{s})=\sqrt{1-\hat{s}^{\dagger}\hat{s}-\hat{h}^{\dagger}\hat{h}}italic_F ( over^ start_ARG italic_h end_ARG , over^ start_ARG italic_s end_ARG ) = square-root start_ARG 1 - over^ start_ARG italic_s end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over^ start_ARG italic_s end_ARG - over^ start_ARG italic_h end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over^ start_ARG italic_h end_ARG end_ARG ensures at most one spin per site. The initial state at time τ=0𝜏0\tau=0italic_τ = 0 of two holes on the same rung of the ladder in a spin background 𝝈𝝈{\boldsymbol{\sigma}}bold_italic_σ is then

|Ψ𝝈(τ=0)=l=12h^(l,0)jS𝝈ls^(l,j)|FM,ketsubscriptΨ𝝈𝜏0superscriptsubscriptproduct𝑙12subscriptsuperscript^𝑙0subscriptproduct𝑗superscriptsubscript𝑆𝝈𝑙subscriptsuperscript^𝑠𝑙𝑗ketFM\displaystyle\ket{\Psi_{\boldsymbol{\sigma}}(\tau=0)}=\prod_{l=1}^{2}\hat{h}^{% \dagger}_{(l,0)}\prod_{j\in S_{{\boldsymbol{\sigma}}}^{l}}\hat{s}^{\dagger}_{(% l,j)}\ket{{\rm FM}},| start_ARG roman_Ψ start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT ( italic_τ = 0 ) end_ARG ⟩ = ∏ start_POSTSUBSCRIPT italic_l = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over^ start_ARG italic_h end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ( italic_l , 0 ) end_POSTSUBSCRIPT ∏ start_POSTSUBSCRIPT italic_j ∈ italic_S start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_l end_POSTSUPERSCRIPT end_POSTSUBSCRIPT over^ start_ARG italic_s end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ( italic_l , italic_j ) end_POSTSUBSCRIPT | start_ARG roman_FM end_ARG ⟩ , (4)

in which S𝝈lsuperscriptsubscript𝑆𝝈𝑙S_{{\boldsymbol{\sigma}}}^{l}italic_S start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_l end_POSTSUPERSCRIPT is the subset of spins in leg l𝑙litalic_l, which are initially in the |ket\ket{\downarrow}| start_ARG ↓ end_ARG ⟩ state. Here, we write 𝐢=(l,j)𝐢𝑙𝑗{\bf i}=(l,j)bold_i = ( italic_l , italic_j ), with leg index l=1,2𝑙12l=1,2italic_l = 1 , 2 and position j𝑗jitalic_j along the leg. Since the spins can only exchange places with the holes along the ladder, the state at any later stage is

|Ψ𝝈(τ)=ketsubscriptΨ𝝈𝜏absent\displaystyle\ket{\Psi_{\boldsymbol{\sigma}}(\tau)}=| start_ARG roman_Ψ start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT ( italic_τ ) end_ARG ⟩ = x1,x2a𝝈(x1,x2;τ)l=12h^(l,xl)subscriptsubscript𝑥1subscript𝑥2subscript𝑎𝝈subscript𝑥1subscript𝑥2𝜏superscriptsubscriptproduct𝑙12subscriptsuperscript^𝑙subscript𝑥𝑙\displaystyle\sum_{x_{1},x_{2}}\!a_{\boldsymbol{\sigma}}(x_{1},x_{2};\tau)% \prod_{l=1}^{2}\hat{h}^{\dagger}_{(l,x_{l})}∑ start_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ; italic_τ ) ∏ start_POSTSUBSCRIPT italic_l = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over^ start_ARG italic_h end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ( italic_l , italic_x start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) end_POSTSUBSCRIPT
×jS𝝈lj[sgn(xl),xl]s^(l,jsgn(xl))jS𝝈lj[sgn(xl),xl]s^(l,j)|FM.\displaystyle\times\!\!\!\!\!\!\!\!\prod_{\begin{subarray}{c}j\in S_{{% \boldsymbol{\sigma}}}^{l}\\ j\in[{\rm sgn}(x_{l}),x_{l}]\end{subarray}}\!\!\!\!\!\!\!\hat{s}^{\dagger}_{(l% ,j-{\rm sgn}(x_{l}))}\!\!\!\!\!\prod_{\begin{subarray}{c}j\in S_{{\boldsymbol{% \sigma}}}^{l}\\ j\notin[{\rm sgn}(x_{l}),x_{l}]\end{subarray}}\!\!\!\!\!\hat{s}^{\dagger}_{(l,% j)}\!\ket{{\rm FM}}.× ∏ start_POSTSUBSCRIPT start_ARG start_ROW start_CELL italic_j ∈ italic_S start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_l end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL italic_j ∈ [ roman_sgn ( italic_x start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) , italic_x start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ] end_CELL end_ROW end_ARG end_POSTSUBSCRIPT over^ start_ARG italic_s end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ( italic_l , italic_j - roman_sgn ( italic_x start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) ) end_POSTSUBSCRIPT ∏ start_POSTSUBSCRIPT start_ARG start_ROW start_CELL italic_j ∈ italic_S start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_l end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL italic_j ∉ [ roman_sgn ( italic_x start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) , italic_x start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ] end_CELL end_ROW end_ARG end_POSTSUBSCRIPT over^ start_ARG italic_s end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ( italic_l , italic_j ) end_POSTSUBSCRIPT | start_ARG roman_FM end_ARG ⟩ . (5)

Here, sgn(x)sgn𝑥{\rm sgn}(x)roman_sgn ( italic_x ) is the sign function. Despite the rather formal description, the physics is simple and clear: if the hole in leg l𝑙litalic_l moves |xl|subscript𝑥𝑙|x_{l}|| italic_x start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT | times to the right, it surpasses |xl|subscript𝑥𝑙|x_{l}|| italic_x start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT | spins that all move one step to the left. If the hole moves to the left, the surpassed spins move one step to the right. The corresponding probability amplitudes a𝝈(x1,x2;τ)subscript𝑎𝝈subscript𝑥1subscript𝑥2𝜏a_{{\boldsymbol{\sigma}}}(x_{1},x_{2};\tau)italic_a start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ; italic_τ ) of finding the holes at positions x1,x2subscript𝑥1subscript𝑥2x_{1},x_{2}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT at time τ𝜏\tauitalic_τ, follow the equations of motion

iτa𝝈(x1,x2;τ)𝑖subscript𝜏subscript𝑎𝝈subscript𝑥1subscript𝑥2𝜏\displaystyle i\partial_{\tau}a_{\boldsymbol{\sigma}}(x_{1},x_{2};\tau)italic_i ∂ start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ; italic_τ ) =V𝝈(x1,x2)a𝝈(x1,x2;τ)absentsubscript𝑉𝝈subscript𝑥1subscript𝑥2subscript𝑎𝝈subscript𝑥1subscript𝑥2𝜏\displaystyle=V_{\boldsymbol{\sigma}}(x_{1},x_{2})a_{{\boldsymbol{\sigma}}}(x_% {1},x_{2};\tau)= italic_V start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) italic_a start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ; italic_τ )
+tδ1,δ2|δ1|+|δ2|=1a𝝈(x1+δ1,x2+δ2;τ),𝑡subscriptsubscript𝛿1subscript𝛿2subscript𝛿1subscript𝛿21subscript𝑎𝝈subscript𝑥1subscript𝛿1subscript𝑥2subscript𝛿2𝜏\displaystyle+t\!\!\!\!\!\!\!\sum_{\begin{subarray}{c}\delta_{1},\delta_{2}\\ |\delta_{1}|+|\delta_{2}|=1\end{subarray}}\!\!\!\!\!\!\!a_{{\boldsymbol{\sigma% }}}(x_{1}+\delta_{1},x_{2}+\delta_{2};\tau),+ italic_t ∑ start_POSTSUBSCRIPT start_ARG start_ROW start_CELL italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL | italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT | + | italic_δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT | = 1 end_CELL end_ROW end_ARG end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ; italic_τ ) , (6)

derived from the Schrödinger equation, iτ|Ψ𝝈(τ)=H^|Ψ𝝈(τ)𝑖subscript𝜏ketsubscriptΨ𝝈𝜏^𝐻ketsubscriptΨ𝝈𝜏i\partial_{\tau}\ket{\Psi_{\boldsymbol{\sigma}}(\tau)}=\hat{H}\ket{\Psi_{% \boldsymbol{\sigma}}(\tau)}italic_i ∂ start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT | start_ARG roman_Ψ start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT ( italic_τ ) end_ARG ⟩ = over^ start_ARG italic_H end_ARG | start_ARG roman_Ψ start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT ( italic_τ ) end_ARG ⟩, with the initial condition that holes are on the same rung: a𝝈(x1=0,x2=0;τ=0)=1subscript𝑎𝝈formulae-sequencesubscript𝑥10formulae-sequencesubscript𝑥20𝜏01a_{\boldsymbol{\sigma}}(x_{1}=0,x_{2}=0;\tau=0)=1italic_a start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0 , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0 ; italic_τ = 0 ) = 1. The sum extends over configurations reached in a single hop. The induced interaction V𝝈(x1,x2)subscript𝑉𝝈subscript𝑥1subscript𝑥2V_{\boldsymbol{\sigma}}(x_{1},x_{2})italic_V start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) emerges, because the motion of the holes alter the spin bonds in the ladder [Fig. 2(a)]. For a random spin background, the induced interaction is also random [Fig. 2(b)], characterized in the relative and center-of-mass coordinates Δ=(x1x2)/2Δsubscript𝑥1subscript𝑥22\Delta=(x_{1}-x_{2})/2roman_Δ = ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) / 2 and X=(x1+x2)/2𝑋subscript𝑥1subscript𝑥22X=(x_{1}+x_{2})/2italic_X = ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) / 2.

Refer to caption
Figure 2: Pairing mechanism. (a) As the holes move (green shaded region), spin bonds in the initial configuration are broken across (dark blue) and within (light blue) the ladder, and new bonds are established (dark and light red, respectively), resulting in a random induced interaction between the holes. (b) Induced interaction for vanishing center-of-mass, X=0𝑋0X=0italic_X = 0, as a function of the relative coordinate ΔΔ\Deltaroman_Δ for 50505050 realizations of the spin background (grey and colored lines), performing a random walk with standard deviation |J||Δ|similar-toabsentsubscript𝐽perpendicular-toΔ\sim|J_{\perp}|\sqrt{|\Delta|}∼ | italic_J start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT | square-root start_ARG | roman_Δ | end_ARG (black lines). (c) Induced interaction at Δ=0Δ0\Delta=0roman_Δ = 0, as a function of X𝑋Xitalic_X. Here, J=Jsubscript𝐽parallel-tosubscript𝐽perpendicular-toJ_{\parallel}=J_{\perp}italic_J start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT = italic_J start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT is used.

I indeed show SM that if all spin backgrounds are equally likely – as in an infinite temperature spin state – V𝝈(x1,x2)subscript𝑉𝝈subscript𝑥1subscript𝑥2V_{\boldsymbol{\sigma}}(x_{1},x_{2})italic_V start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) has zero mean and a variance (x1,x20subscript𝑥1subscript𝑥20x_{1},x_{2}\neq 0italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ≠ 0)

Var[V𝝈(x1,x2)]=J24[|Δ|14]+J24Vardelimited-[]subscript𝑉𝝈subscript𝑥1subscript𝑥2superscriptsubscript𝐽perpendicular-to24delimited-[]Δ14superscriptsubscript𝐽parallel-to24\displaystyle{\rm Var}[V_{{\boldsymbol{\sigma}}}(x_{1},x_{2})]=\frac{J_{\perp}% ^{2}}{4}\left[|\Delta|-\frac{1}{4}\right]+\frac{J_{\parallel}^{2}}{4}roman_Var [ italic_V start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) ] = divide start_ARG italic_J start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 end_ARG [ | roman_Δ | - divide start_ARG 1 end_ARG start_ARG 4 end_ARG ] + divide start_ARG italic_J start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 end_ARG (7)

linear in the relative distance 2|Δ|2Δ2|\Delta|2 | roman_Δ | between the holes, describing a classical random walk in the values of the potential [Fig. 2(b)]. I can focus on the evolution in specific spin backgrounds by resolving the dynamics of the density matrix in terms of these states. Starting from the initial density matrix

ρ^(τ=0)=l=1,2,σ(l,0)c^(l,0)σ(l,0)ρ^Jc^(l,0)σ(l,0),^𝜌𝜏0subscript𝑙12subscript𝜎𝑙0subscript^𝑐𝑙0𝜎𝑙0subscript^𝜌𝐽subscriptsuperscript^𝑐𝑙0𝜎𝑙0\hat{\rho}(\tau=0)=\!\!\!\sum_{l=1,2,\sigma_{(l,0)}}\!\!\!\hat{c}_{(l,0)\sigma% {(l,0)}}\hat{\rho}_{J}\hat{c}^{\dagger}_{(l,0)\sigma{(l,0)}},over^ start_ARG italic_ρ end_ARG ( italic_τ = 0 ) = ∑ start_POSTSUBSCRIPT italic_l = 1 , 2 , italic_σ start_POSTSUBSCRIPT ( italic_l , 0 ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT ( italic_l , 0 ) italic_σ ( italic_l , 0 ) end_POSTSUBSCRIPT over^ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ( italic_l , 0 ) italic_σ ( italic_l , 0 ) end_POSTSUBSCRIPT , (8)

where ρ^J=eβH^J/ZJsubscript^𝜌𝐽superscript𝑒𝛽subscript^𝐻𝐽subscript𝑍𝐽\hat{\rho}_{J}=e^{-\beta{\hat{H}}_{J}}/Z_{J}over^ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT = italic_e start_POSTSUPERSCRIPT - italic_β over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT end_POSTSUPERSCRIPT / italic_Z start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT is the Gibbs state of the spins at inverse temperature β=1/(kBT)𝛽1subscript𝑘𝐵𝑇\beta=1/(k_{B}T)italic_β = 1 / ( italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T ) and σ(l,0)=,subscript𝜎𝑙0\sigma_{(l,0)}=\uparrow,\downarrowitalic_σ start_POSTSUBSCRIPT ( italic_l , 0 ) end_POSTSUBSCRIPT = ↑ , ↓ designates the spin configuration at the origin of each leg, l𝑙litalic_l, the ensuing dynamics ρ^(τ)=eiH^τρ^(τ=0)e+iH^τ^𝜌𝜏superscript𝑒𝑖^𝐻𝜏^𝜌𝜏0superscript𝑒𝑖^𝐻𝜏\hat{\rho}(\tau)=e^{-i\hat{H}\tau}\hat{\rho}(\tau=0)e^{+i\hat{H}\tau}over^ start_ARG italic_ρ end_ARG ( italic_τ ) = italic_e start_POSTSUPERSCRIPT - italic_i over^ start_ARG italic_H end_ARG italic_τ end_POSTSUPERSCRIPT over^ start_ARG italic_ρ end_ARG ( italic_τ = 0 ) italic_e start_POSTSUPERSCRIPT + italic_i over^ start_ARG italic_H end_ARG italic_τ end_POSTSUPERSCRIPT – provided that the system is assumed to be closed – is the Boltzmann-weighted sum Nielsen (2024a, b)

ρ^(τ)=𝝈eβEJ(𝝈)ZJ|Ψ𝝈(τ)Ψ𝝈(τ)|,^𝜌𝜏subscript𝝈superscript𝑒𝛽subscript𝐸𝐽𝝈subscript𝑍𝐽ketsubscriptΨ𝝈𝜏brasubscriptΨ𝝈𝜏\hat{\rho}(\tau)=\sum_{{\boldsymbol{\sigma}}}\frac{e^{-\beta E_{J}({% \boldsymbol{\sigma}})}}{Z_{J}}\ket{\Psi_{{\boldsymbol{\sigma}}}(\tau)}\bra{% \Psi_{{\boldsymbol{\sigma}}}(\tau)},over^ start_ARG italic_ρ end_ARG ( italic_τ ) = ∑ start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT divide start_ARG italic_e start_POSTSUPERSCRIPT - italic_β italic_E start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT ( bold_italic_σ ) end_POSTSUPERSCRIPT end_ARG start_ARG italic_Z start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT end_ARG | start_ARG roman_Ψ start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT ( italic_τ ) end_ARG ⟩ ⟨ start_ARG roman_Ψ start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT ( italic_τ ) end_ARG | , (9)

where EJ(𝝈)subscript𝐸𝐽𝝈E_{J}({\boldsymbol{\sigma}})italic_E start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT ( bold_italic_σ ) is the magnetic energy of the spin realization 𝝈𝝈{\boldsymbol{\sigma}}bold_italic_σ before the spins σ(1,0),σ(2,0)subscript𝜎10subscript𝜎20\sigma_{(1,0)},\sigma_{(2,0)}italic_σ start_POSTSUBSCRIPT ( 1 , 0 ) end_POSTSUBSCRIPT , italic_σ start_POSTSUBSCRIPT ( 2 , 0 ) end_POSTSUBSCRIPT are removed. In particular, at infinite temperatures all pure state evolutions |Ψ𝝈(τ)ketsubscriptΨ𝝈𝜏\ket{\Psi_{{\boldsymbol{\sigma}}}(\tau)}| start_ARG roman_Ψ start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT ( italic_τ ) end_ARG ⟩ are sampled with equal probability, eβEJ(𝝈)/ZJ=2Nsuperscript𝑒𝛽subscript𝐸𝐽𝝈subscript𝑍𝐽superscript2𝑁e^{-\beta E_{J}({\boldsymbol{\sigma}})}/Z_{J}=2^{-N}italic_e start_POSTSUPERSCRIPT - italic_β italic_E start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT ( bold_italic_σ ) end_POSTSUPERSCRIPT / italic_Z start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT = 2 start_POSTSUPERSCRIPT - italic_N end_POSTSUPERSCRIPT, for system size N𝑁Nitalic_N. In this high-temperature limit, the dynamics is exactly the same for antiferromagnetic (J,J>0subscript𝐽perpendicular-tosubscript𝐽parallel-to0J_{\perp},J_{\parallel}>0italic_J start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT , italic_J start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT > 0) and ferromagnetic couplings (J,J<0subscript𝐽perpendicular-tosubscript𝐽parallel-to0J_{\perp},J_{\parallel}<0italic_J start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT , italic_J start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT < 0). The calculation of the hole dynamics then follows the recipe: (1) generate a sample S𝑆Sitalic_S of NS=400subscript𝑁𝑆400N_{S}=400italic_N start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT = 400 random spin configurations 𝝈𝝈{\boldsymbol{\sigma}}bold_italic_σ and calculate V𝝈(x1,x2)subscript𝑉𝝈subscript𝑥1subscript𝑥2V_{{\boldsymbol{\sigma}}}(x_{1},x_{2})italic_V start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ), (2) solve the equations of motion in Eq. (6) using exact diagonalization for N=161×2𝑁1612N=161\times 2italic_N = 161 × 2, and (3) compute the hole-hole correlator

Chh(x1,x2;τ)=n^h(x1)n^h(x2)τ=𝝈S|a𝝈(x1,x2;τ)|2NS,subscript𝐶hhsubscript𝑥1subscript𝑥2𝜏subscriptexpectationsubscript^𝑛subscript𝑥1subscript^𝑛subscript𝑥2𝜏subscript𝝈𝑆superscriptsubscript𝑎𝝈subscript𝑥1subscript𝑥2𝜏2subscript𝑁𝑆C_{\rm hh}(x_{1},x_{2};\tau)=\braket{\hat{n}_{h}(x_{1})\hat{n}_{h}(x_{2})}_{% \tau}=\sum_{{\boldsymbol{\sigma}}\in S}\frac{|a_{{\boldsymbol{\sigma}}}(x_{1},% x_{2};\tau)|^{2}}{N_{S}},italic_C start_POSTSUBSCRIPT roman_hh end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ; italic_τ ) = ⟨ start_ARG over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) end_ARG ⟩ start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT bold_italic_σ ∈ italic_S end_POSTSUBSCRIPT divide start_ARG | italic_a start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ; italic_τ ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_N start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT end_ARG , (10)

whereby it may be tracked in time [Fig. 1(b)]. This describes the probability of simultaneously finding holes at positions (1,x1),(2,x2)1subscript𝑥12subscript𝑥2(1,x_{1}),(2,x_{2})( 1 , italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) , ( 2 , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ). Here, n^h(xl)=h^(l,xl)h^(l,xl)subscript^𝑛subscript𝑥𝑙subscriptsuperscript^𝑙subscript𝑥𝑙subscript^𝑙subscript𝑥𝑙\hat{n}_{h}(x_{l})=\hat{h}^{\dagger}_{(l,x_{l})}\hat{h}_{(l,x_{l})}over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) = over^ start_ARG italic_h end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ( italic_l , italic_x start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) end_POSTSUBSCRIPT over^ start_ARG italic_h end_ARG start_POSTSUBSCRIPT ( italic_l , italic_x start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) end_POSTSUBSCRIPT is the hole density operator. Figure 1(b) manifestly breaks the individual inversion symmetries x1x1subscript𝑥1subscript𝑥1x_{1}\to-x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT → - italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT or x2x2subscript𝑥2subscript𝑥2x_{2}\to-x_{2}italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT → - italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT of uncorrelated holes, and shows a clear tendency for them to be on the same rung (x1=x2Δ=0subscript𝑥1subscript𝑥2Δ0x_{1}=x_{2}\Leftrightarrow\Delta=0italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⇔ roman_Δ = 0). Their motion is further analyzed by the root-mean-square of the center-of-mass and relative coordinates [Fig. 1(c)],

Arms(τ)subscript𝐴rms𝜏\displaystyle A_{\rm rms}(\tau)italic_A start_POSTSUBSCRIPT roman_rms end_POSTSUBSCRIPT ( italic_τ ) =[X,ΔA2Chh(X+Δ,XΔ;τ)]1/2,absentsuperscriptdelimited-[]subscript𝑋Δsuperscript𝐴2subscript𝐶hh𝑋Δ𝑋Δ𝜏12\displaystyle=\left[\sum_{X,\Delta}A^{2}C_{\rm hh}(X+\Delta,X-\Delta;\tau)% \right]^{1/2},= [ ∑ start_POSTSUBSCRIPT italic_X , roman_Δ end_POSTSUBSCRIPT italic_A start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_C start_POSTSUBSCRIPT roman_hh end_POSTSUBSCRIPT ( italic_X + roman_Δ , italic_X - roman_Δ ; italic_τ ) ] start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT , (11)

for A=X,Δ𝐴𝑋ΔA=X,\Deltaitalic_A = italic_X , roman_Δ. Since Xrms>Δrmssubscript𝑋rmssubscriptΔrmsX_{\rm rms}>\Delta_{\rm rms}italic_X start_POSTSUBSCRIPT roman_rms end_POSTSUBSCRIPT > roman_Δ start_POSTSUBSCRIPT roman_rms end_POSTSUBSCRIPT, the holes are closer together than they have propagated into the lattice. This is further characterized by the correlation coefficient

corr(x1,x2)=x1x2τσ(x1)σ(x2)=Xrms2Δrms2Xrms2+Δrms2,corrsubscript𝑥1subscript𝑥2subscriptexpectationsubscript𝑥1subscript𝑥2𝜏𝜎subscript𝑥1𝜎subscript𝑥2superscriptsubscript𝑋rms2superscriptsubscriptΔrms2superscriptsubscript𝑋rms2superscriptsubscriptΔrms2\displaystyle{\rm corr}(x_{1},x_{2})=\frac{\braket{x_{1}x_{2}}_{\tau}}{\sigma(% x_{1})\sigma(x_{2})}=\frac{X_{\rm rms}^{2}-\Delta_{\rm rms}^{2}}{X_{\rm rms}^{% 2}+\Delta_{\rm rms}^{2}},roman_corr ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) = divide start_ARG ⟨ start_ARG italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ⟩ start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT end_ARG start_ARG italic_σ ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) italic_σ ( italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) end_ARG = divide start_ARG italic_X start_POSTSUBSCRIPT roman_rms end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - roman_Δ start_POSTSUBSCRIPT roman_rms end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_X start_POSTSUBSCRIPT roman_rms end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + roman_Δ start_POSTSUBSCRIPT roman_rms end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (12)

which is always restricted to the interval [1,+1]11[-1,+1][ - 1 , + 1 ]. Positive values of corr(x1,x2)corrsubscript𝑥1subscript𝑥2{\rm corr}(x_{1},x_{2})roman_corr ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ), as seen in Fig. 1(d), correspond to Xrms>Δrmssubscript𝑋rmssubscriptΔrmsX_{\rm rms}>\Delta_{\rm rms}italic_X start_POSTSUBSCRIPT roman_rms end_POSTSUBSCRIPT > roman_Δ start_POSTSUBSCRIPT roman_rms end_POSTSUBSCRIPT, and signify attraction. Likewise, negative values correspond to an effective repulsion, while corr(x1,x2)=0corrsubscript𝑥1subscript𝑥20{\rm corr}(x_{1},x_{2})=0roman_corr ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) = 0 corresponds to no linear correlations. To further substantiate the findings, I analyze the relative hole-hole correlator

Chhrel(Δ,τ)=XChh(X+Δ,XΔ;τ)XChhdisc(X+Δ,XΔ;τ),superscriptsubscript𝐶hhrelΔ𝜏subscript𝑋subscript𝐶hh𝑋Δ𝑋Δ𝜏subscript𝑋superscriptsubscript𝐶hhdisc𝑋Δ𝑋Δ𝜏C_{\rm hh}^{\rm rel}(\Delta,\tau)=\frac{\sum_{X}C_{\rm hh}(X+\Delta,X-\Delta;% \tau)}{\sum_{X}C_{\rm hh}^{\rm disc}(X+\Delta,X-\Delta;\tau)},italic_C start_POSTSUBSCRIPT roman_hh end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_rel end_POSTSUPERSCRIPT ( roman_Δ , italic_τ ) = divide start_ARG ∑ start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT italic_C start_POSTSUBSCRIPT roman_hh end_POSTSUBSCRIPT ( italic_X + roman_Δ , italic_X - roman_Δ ; italic_τ ) end_ARG start_ARG ∑ start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT italic_C start_POSTSUBSCRIPT roman_hh end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_disc end_POSTSUPERSCRIPT ( italic_X + roman_Δ , italic_X - roman_Δ ; italic_τ ) end_ARG , (13)

giving the probability of finding the holes at distance 2|Δ|2Δ2|\Delta|2 | roman_Δ | relative to what is expected from their marginals, i.e. the disconnected correlator Chhdisc(Δ,τ)=Xn^h(X+Δ)n^h(XΔ)subscriptsuperscript𝐶dischhΔ𝜏subscript𝑋expectationsubscript^𝑛𝑋Δexpectationsubscript^𝑛𝑋ΔC^{\rm disc}_{\rm hh}(\Delta,\tau)=\sum_{X}\braket{\hat{n}_{h}(X+\Delta)}% \braket{\hat{n}_{h}(X-\Delta)}italic_C start_POSTSUPERSCRIPT roman_disc end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_hh end_POSTSUBSCRIPT ( roman_Δ , italic_τ ) = ∑ start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT ⟨ start_ARG over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_X + roman_Δ ) end_ARG ⟩ ⟨ start_ARG over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_X - roman_Δ ) end_ARG ⟩. This is shown for Δ=0Δ0\Delta=0roman_Δ = 0 in Fig. 1(e), demonstrating a significant increase in the probability of finding the holes on the same rung. On long timescales, the system settles into a steady state, by which XrmslX,ΔrmslΔformulae-sequencesubscript𝑋rmssubscript𝑙𝑋subscriptΔrmssubscript𝑙ΔX_{\rm rms}\to l_{X},\Delta_{\rm rms}\to l_{\Delta}italic_X start_POSTSUBSCRIPT roman_rms end_POSTSUBSCRIPT → italic_l start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT , roman_Δ start_POSTSUBSCRIPT roman_rms end_POSTSUBSCRIPT → italic_l start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT. For J2tsubscript𝐽perpendicular-to2𝑡J_{\perp}\geq 2titalic_J start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ≥ 2 italic_t, there are negligible finite-size corrections for the chosen length of N/2=161𝑁2161N/2=161italic_N / 2 = 161, and I, therefore, confine the analysis to this regime. Figure 3(a), hereby, shows that lX>lΔsubscript𝑙𝑋subscript𝑙Δl_{X}>l_{\Delta}italic_l start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT > italic_l start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT remains true over this entire range of parameters. The associated asymptote of the correlation coefficient [Fig. 3(b)] shows very strong attractive correlations for Jtmuch-greater-thansubscript𝐽perpendicular-to𝑡J_{\perp}\gg titalic_J start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ≫ italic_t and J=0subscript𝐽parallel-to0J_{\parallel}=0italic_J start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT = 0.

Refer to caption
Figure 3: Steady state. (a) The rms distances lXsubscript𝑙𝑋l_{X}italic_l start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT and lΔsubscript𝑙Δl_{\Delta}italic_l start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT of the center-of-mass, X𝑋Xitalic_X, and relative, ΔΔ\Deltaroman_Δ, coordinates of the holes as a function of J/tsubscript𝐽perpendicular-to𝑡J_{\perp}/titalic_J start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT / italic_t for indicated values of Jsubscript𝐽parallel-toJ_{\parallel}italic_J start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT in the steady state. (b) Associated correlation coefficient corr(x1,x2)=x1x2/[σ(x1)σ(x2)]corrsubscript𝑥1subscript𝑥2expectationsubscript𝑥1subscript𝑥2delimited-[]𝜎subscript𝑥1𝜎subscript𝑥2{\rm corr}(x_{1},x_{2})=\braket{x_{1}x_{2}}/[\sigma(x_{1})\sigma(x_{2})]roman_corr ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) = ⟨ start_ARG italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ⟩ / [ italic_σ ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) italic_σ ( italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) ]. (c) Relative hole-hole correlator [Eq. (13)] for holes on the same rung (Δ=0Δ0\Delta=0roman_Δ = 0) vs J/tsubscript𝐽perpendicular-to𝑡J_{\perp}/titalic_J start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT / italic_t. (d) Relative hole-hole correlator vs |Δ|Δ|\Delta|| roman_Δ | for indicated values of J/tsubscript𝐽perpendicular-to𝑡J_{\perp}/titalic_J start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT / italic_t and Jsubscript𝐽parallel-toJ_{\parallel}italic_J start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT.

The dynamical binding – a finite lΔsubscript𝑙Δl_{\Delta}italic_l start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT – has the exact same origin as the thermal localization phenomenon discovered recently for the motion of single holes Nielsen (2024a, b), by which the separation of the holes is avoided, because the induced interaction V𝝈(x1,x2)subscript𝑉𝝈subscript𝑥1subscript𝑥2V_{\boldsymbol{\sigma}}(x_{1},x_{2})italic_V start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) inevitably fluctuates to large values in the relative distance 2|Δ|=|x1x2|2Δsubscript𝑥1subscript𝑥22|\Delta|=|x_{1}-x_{2}|2 | roman_Δ | = | italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT |, as seen in Fig. 2(b). Therefore, at the very least, they back-reflect once they meet their classical turning points, corresponding to an emergent kind of Anderson localization Anderson (1958) in the presence of strong disorder Lagendijk et al. (2009). The retained localization of the holes – a finite lXsubscript𝑙𝑋l_{X}italic_l start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT – is more subtle, because V𝝈(x,x)subscript𝑉𝝈𝑥𝑥V_{{\boldsymbol{\sigma}}}(x,x)italic_V start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT ( italic_x , italic_x ) is restricted to take values between ±|J|plus-or-minussubscript𝐽parallel-to\pm|J_{\parallel}|± | italic_J start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT | [Fig. 2(c)], such that it vanishes for holes on the same rung in the extreme limit of J=0subscript𝐽parallel-to0J_{\parallel}=0italic_J start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT = 0. Even in this case, however, the holes localize [Fig. 3(b)]. The reason is that there are many ways in which the holes can move to sites (1,x)1𝑥(1,x)( 1 , italic_x ) and (2,x)2𝑥(2,x)( 2 , italic_x ). Along each path, the wave builds up a phase directly depending on the randomly fluctuating induced interaction they meet along the way. The ensuing destructive interference of these pathways corresponds to Anderson localization for weak disorder Lagendijk et al. (2009). Finally, the relative hole-hole correlator in Figs. 3(c) and 3(d) show that the probability of finding holes in close proximity can be several times higher than expected from their marginals. Overall, it is clear that at large J/tsubscript𝐽perpendicular-to𝑡J_{\perp}/titalic_J start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT / italic_t the strength of the pairing sensitively depends on Jsubscript𝐽parallel-toJ_{\parallel}italic_J start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT, essentially because the holes localize over very few lattice sites if JJsimilar-tosubscript𝐽parallel-tosubscript𝐽perpendicular-toJ_{\parallel}\sim J_{\perp}italic_J start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ∼ italic_J start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT. However, already at intermediate values of J/tsubscript𝐽perpendicular-to𝑡J_{\perp}/titalic_J start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT / italic_t, the pairing for J=Jsubscript𝐽parallel-tosubscript𝐽perpendicular-toJ_{\parallel}=J_{\perp}italic_J start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT = italic_J start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT is comparable to the case of J=0subscript𝐽parallel-to0J_{\parallel}=0italic_J start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT = 0. I expect this to persist at lower J/tsubscript𝐽perpendicular-to𝑡J_{\perp}/titalic_J start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT / italic_t.

Rydberg-dressed atoms.-

The predicted results may be feasibly tested in ultracold quantum simulators based on Rydberg-dressed atoms in optical lattices Zeiher et al. (2016). Indeed, such schemes natively realizes spin-specific density-density interactions

H^J=12𝐢𝐣J𝐢𝐣n^𝐢n^𝐣subscript^𝐻𝐽12subscript𝐢𝐣subscript𝐽𝐢𝐣subscript^𝑛absent𝐢subscript^𝑛absent𝐣{\hat{H}}_{J}=\frac{1}{2}\sum_{{\bf i}\neq{\bf j}}J_{{\bf i}-{\bf j}}\hat{n}_{% \uparrow{\bf i}}\hat{n}_{\uparrow{\bf j}}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT bold_i ≠ bold_j end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT bold_i - bold_j end_POSTSUBSCRIPT over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT ↑ bold_i end_POSTSUBSCRIPT over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT ↑ bold_j end_POSTSUBSCRIPT (14)

of the internal atomic state |ket\ket{\uparrow}| start_ARG ↑ end_ARG ⟩, by optically dressing it with a higher-lying Rydberg state. Here, J𝐢𝐣subscript𝐽𝐢𝐣J_{{\bf i}-{\bf j}}italic_J start_POSTSUBSCRIPT bold_i - bold_j end_POSTSUBSCRIPT takes on a soft core shape Henkel et al. (2010); SM . Moreover, such systems were recently demonstrated to enter the itinerant regime Weckesser et al. (2024), by which the desired t𝑡titalic_t-J𝐽Jitalic_J model may be realized. In particular, the spin model in Eq. (14) is obtained by using a Feshbach resonance to enhance the onsite interaction deep into the Mott-insulating phase. This ensures at most one spin per site and negligible spin superexchange. Precise control of the temperature in such systems is difficult, but the hole dynamics may still be investigated in the following manner Nielsen (2024a). First, initialize holes at x1=x2=0subscript𝑥1subscript𝑥20x_{1}=x_{2}=0italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0 by applying a strong repulsive light field to those sites Ji et al. (2021). Second, rotate all spins from the fully polarized spin sample in the non-interacting |ket\ket{\downarrow}| start_ARG ↓ end_ARG ⟩ state into the equal superposition state |Ψ0=𝐢𝟎12[c^𝐢+c^𝐢]|0ketsubscriptΨ0subscriptproduct𝐢012delimited-[]subscriptsuperscript^𝑐𝐢absentsubscriptsuperscript^𝑐𝐢absentket0\ket{\Psi_{0}}=\prod_{{\bf i}\neq{\bf 0}}\frac{1}{\sqrt{2}}\left[\hat{c}^{% \dagger}_{{\bf i}\uparrow}+\hat{c}^{\dagger}_{{\bf i}\downarrow}\right]\ket{0}| start_ARG roman_Ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ⟩ = ∏ start_POSTSUBSCRIPT bold_i ≠ bold_0 end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG [ over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_i ↑ end_POSTSUBSCRIPT + over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_i ↓ end_POSTSUBSCRIPT ] | start_ARG 0 end_ARG ⟩. Third, turn off the focused light fields, freeing the holes to move along the ladder Hirthe et al. (2023). Although this at face value seems distinct from the infinite temperature scenario above, the hole-hole correlator takes on the exact same form

Chh(x1,x2;τ)subscript𝐶hhsubscript𝑥1subscript𝑥2𝜏\displaystyle C_{\rm hh}(x_{1},x_{2};\tau)italic_C start_POSTSUBSCRIPT roman_hh end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ; italic_τ ) =2N𝝈|a𝝈(x1,x2;τ)|2,absentsuperscript2𝑁subscript𝝈superscriptsubscript𝑎𝝈subscript𝑥1subscript𝑥2𝜏2\displaystyle=2^{-N}\sum_{{\boldsymbol{\sigma}}}|a_{\boldsymbol{\sigma}}(x_{1}% ,x_{2};\tau)|^{2},= 2 start_POSTSUPERSCRIPT - italic_N end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT | italic_a start_POSTSUBSCRIPT bold_italic_σ end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ; italic_τ ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (15)

such that the holes move as though the system is at infinite temperature. The polarized Ising form of the interaction is crucial for this exact equivalence, and one should note that spin correlators would presumably differ from their infinite temperature counterpart. The computation now follows the same recipe as before, and the resulting correlation coefficient dynamics [Eq. (12)] and relative correlation function [Eq. (13)] are given in Figs. 4(a) and 4(b), respectively. This is averaged over NS=1000subscript𝑁𝑆1000N_{S}=1000italic_N start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT = 1000 realizations and given in two instances [inset of Fig. 4(a)]: (1) isotropic couplings akin to the J=Jsubscript𝐽parallel-tosubscript𝐽perpendicular-toJ_{\parallel}=J_{\perp}italic_J start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT = italic_J start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT case, and (2) anisotropic couplings where the 30P3/230subscript𝑃3230P_{3/2}30 italic_P start_POSTSUBSCRIPT 3 / 2 end_POSTSUBSCRIPT orbital of the Rydberg state is rotated to lie along the rungs using an external magnetic field Zeiher et al. (2016), suppressing coupling in other directions similar to low values of Jsubscript𝐽parallel-toJ_{\parallel}italic_J start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT SM . Crucially, we observe that the appearing disorder-induced attraction between the holes can be probed for experimentally feasible system sizes and timescales. Indeed, stroboscopic dressing Weckesser et al. (2024) makes 20202020 tunneling events feasible, while longer timescales are inhibited by the intrinsic lifetime of the Rydberg state Weckesser (2024). I note that the rung case gives a much more favorable signal, just as was the case for J=0subscript𝐽parallel-to0J_{\parallel}=0italic_J start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT = 0 vs J=Jsubscript𝐽parallel-tosubscript𝐽perpendicular-toJ_{\parallel}=J_{\perp}italic_J start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT = italic_J start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT.

Refer to caption
Figure 4: Rydberg dressing. (a) Dynamics of the correlation coefficient for isotropic (orange) and strong rung (purple) spin couplings. (b) Associated relative correlator for holes on the same rung, Δ=0Δ0\Delta=0roman_Δ = 0. Dashed lines: steady state values. Nearest neighbor rung interaction: J𝐞y=8tsubscript𝐽subscript𝐞𝑦8𝑡J_{{\bf e}_{y}}=8titalic_J start_POSTSUBSCRIPT bold_e start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT end_POSTSUBSCRIPT = 8 italic_t. Insets: orientation of 30P3/230subscript𝑃3230P_{3/2}30 italic_P start_POSTSUBSCRIPT 3 / 2 end_POSTSUBSCRIPT Rydberg orbitals enforced by shown magnetic fields 𝐁𝐁{\bf B}bold_B. System size: N=21×2𝑁212N=21\times 2italic_N = 21 × 2.

Conclusions and outlook.-

I have demonstrated that thermal disorder of spins can frustrate the relative motion of dopants and lead to a dynamical pairing mechanism. The effect is closely related to the thermally induced localization phenomenon recently discovered for single holes Nielsen (2024a, b). By carefully analyzing its origin and behavior at infinite temperature, it is, furthermore, clear that the effect persists for all values of the spin coupling versus the hop**. For diminishing temperatures, ferromagnetic interaction is expected to lead to a crossover to delocalized and uncorrelated holes. For antiferromagnetic interactions, on the other hand, a crossover from fluctuation induced pairing to confinement of the holes emerges at low temperatures Nielsen (2023). Moreover, it demonstrates that the localization phenomenon Nielsen (2024a, b) is not sensitive to further do**. The results finally show that it is feasible to experimentally test my predictions using currently available techniques with ultracold atoms in optical lattices. One could also engineer similar models using Rydberg arrays Schauss et al. (2015); Labuhn et al. (2016); Bernien et al. (2017); Guardado-Sanchez et al. (2018); Lienhard et al. (2018) or polar molecules Gorshkov et al. (2011). In the future, the generality of the discovered effect may be tested by, e.g., tuning in trans-leg hop** or spin superexchange. The robustness of the phenomenon to such alterations is presently unclear. The nature of the effect is, however, very general – it comes down to a randomly fluctuating field emerging as the dopants separate, frustrating their relative motion.

Acknowledgements.
The author thanks Pascal Weckesser, Johannes Zeiher, Pavel Kos, and J. Ignacio Cirac for helpful input and discussions. A special thanks goes to Pascal Weckesser for computing the soft core potentials. This work has been supported by the Carlsberg Foundation through a Carlsberg Internationalisation Fellowship, Grant No. CF21_0410.

References