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

η\eta Pairing of Light-Emitting Fermions:
Nonequilibrium Pairing Mechanism at High Temperatures

Masaya Nakagawa [email protected] Department of Physics, University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-0033, Japan    Naoto Tsuji Department of Physics, University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-0033, Japan RIKEN Center for Emergent Matter Science (CEMS), Wako, Saitama 351-0198, Japan    Norio Kawakami Department of Physics, Kyoto University, Kyoto 606-8502, Japan    Masahito Ueda Department of Physics, University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-0033, Japan RIKEN Center for Emergent Matter Science (CEMS), Wako, Saitama 351-0198, Japan Institute for Physics of Intelligence, University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-0033, Japan
Abstract

Strongly interacting fermionic atoms are shown to develop η\eta-pairing superfluid correlations in a nonequilibrium steady state in the presence of spontaneous emission of light from atoms. On the basis of the Hubbard model subject to spontaneous decay between internal spin states, we show that prohibition of radiative decay due to the Pauli exclusion principle and destructive interference between doublon-decay processes lead to nonequilibrium η\eta pairing. Because of the non-thermal nature of the steady state, pair correlations arise even from a completely uncorrelated infinite-temperature initial state, allowing coherent atom pairs to be formed at high temperatures. Experimental implementation with fermionic atoms in an optical lattice is discussed.

pacs:

In a seminal paper Yang (1989), Yang showed that exact eigenstates of the Fermi-Hubbard model can be constructed through the η\eta-pairing mechanism. Although the Hubbard model in two and three dimensions is not exactly solvable due to the competing nature of hopping and interaction, a hidden η\eta symmetry of the model permits η\eta-pairing eigenstates Yang and Zhang (1990); Pernici (1990). The η\eta-pairing states exhibit an off-diagonal long-range order Yang (1962), giving a rare example of fermionic superfluidity that is an exact eigenstate. However, since they cannot be the ground state of the Hubbard model except for the limiting case of an infinite attractive interaction Yang (1989); Singh and Scalettar (1991), η\eta pairing requires nonequilibrium situations such as adiabatic state preparation Rosch et al. (2008); Kantian et al. (2010), laser irradiation Kaneko et al. (2019, 2020); Werner et al. (2019); Li et al. (2020), periodic driving Kitamura and Aoki (2016); Peronaci et al. (2020); Cook and Clark (2020); Tindall et al. (2021a, b), and tailored dissipative processes Diehl et al. (2008); Kraus et al. (2008); Bernier et al. (2013); Tindall et al. (2019); Zhang and Song (2020a, b).

Control of superfluidity with light lies at the forefront of nonequilibrium quantum many-body physics Fausti et al. (2011); Kaiser et al. (2014); Hu et al. (2014); Mitrano et al. (2016); Cantaluppi et al. (2018); Suzuki et al. (2019); Buzzi et al. (2020); Budden et al. (2020). Since irradiation of light generally heats up a sample, how to create quantum coherence of a superfluid by the light-matter coupling is a highly nontrivial question with several interesting possibilities Kaneko et al. (2019, 2020); Werner et al. (2019); Li et al. (2020); Fausti et al. (2011); Kaiser et al. (2014); Hu et al. (2014); Mitrano et al. (2016); Cantaluppi et al. (2018); Suzuki et al. (2019); Buzzi et al. (2020); Budden et al. (2020). Importantly, a pairing mechanism in nonequilibrium quantum matter may be fundamentally different from that of the BCS theory Bardeen et al. (1957) because of its nonequilibrium nature.

Refer to caption
Figure 1: (a) Schematic illustration of the model. Spin-1/21/2 fermionic atoms in an optical lattice are described by the Hubbard model with the hopping amplitude tt and the on-site interaction UU, where spontaneous emission of light leads to an on-site decay process from spin-up to spin-down states with rate γ\gamma. When two atoms with opposite spins occupy the same site, the Pauli blocking prevents atoms from radiative decay. (b) Lambda-type level structure for the implementation of spontaneous decay in the Hubbard model. The spin-up state |g,\ket{g,\uparrow} can decay to the spin-down state |g,\ket{g,\downarrow} through an off-resonant coherent transition (shown by a double arrow) to an excited state |e\ket{e} and a subsequent spontaneous decay (shown by wavy arrows) from it.

In this Letter, we propose a nonequilibrium η\eta-pairing mechanism induced by spontaneous emission of light from strongly correlated fermionic atoms. We consider a cold-atom Fermi-Hubbard system in an optical lattice in which a spin-up state of an atom can decay to a spin-down state via a coupling to an excited state that decays by spontaneously emitting a photon, as illustrated in Fig. 1. We show that the nonequilibrium steady states of this light-matter-coupled system are given by the η\eta-pairing states. Here, the η\eta-pairing states are stabilized due to double protection of fermion pairs from possible decay processes. First, as shown in Fig. 1(a), the Pauli exclusion principle prohibits on-site pairs (i.e., doublons) from radiative decay Sandner et al. (2011). Second, complete destructive interference between doublon-decay processes prevents doublons from breaking up into single particles, thereby maintaining the η\eta-pairing states as a coherent steady state. This pairing mechanism is inherently of nonequilibrium nature and makes a marked contrast to the BCS mechanism Bardeen et al. (1957) in which an attractive interaction between fermions and the Fermi-surface instability play essential roles. Remarkably, unlike previous proposals Bernier et al. (2013); Tindall et al. (2019), the η\eta-pairing steady state is observed even when the dynamics starts from an infinite-temperature initial state, because the entropy of the system can be decreased by dissipative processes of spontaneous emission which is described by non-Hermitian Lindblad operators. This indicates that one can create coherent atom pairs at an initial temperature far above the BCS critical temperature. We propose an experimental implementation of the model with alkaline-earth-like fermionic atoms in an optical lattice and discuss how to detect the η\eta-pairing state through measurement of a momentum distribution of doublons.

Model.– We generalize a quantum optical master equation Breuer and Petruccione (2007), which has been used to describe the dynamics of noninteracting atoms subject to spontaneous emission of photons, to interacting fermionic atoms Lehmberg (1970); Pichler et al. (2010); Sarkar et al. (2014). We consider spin-1/21/2 fermionic atoms in a dd-dimensional cubic optical lattice. The Hamiltonian of our system is described by the Hubbard model HH when the lattice potential is sufficiently deep and the single-band approximation is valid Esslinger (2010):

H=\displaystyle H= ti,jσ=,(ciσcjσ+H.c.)+Ujnjnj\displaystyle-t\sum_{\langle i,j\rangle}\sum_{\sigma=\uparrow,\downarrow}(c_{i\sigma}^{\dagger}c_{j\sigma}+\mathrm{H.c.})+U\sum_{j}n_{j\uparrow}n_{j\downarrow}
+δj(njnj),\displaystyle+\delta\sum_{j}(n_{j\uparrow}-n_{j\downarrow}), (1)

where cjσc_{j\sigma} is the annihilation operator of a spin-σ\sigma fermion at a lattice position 𝑹j{0,,L1}d\bm{R}_{j}\in\{0,\cdots,L-1\}^{d}, njσcjσcjσn_{j\sigma}\equiv c_{j\sigma}^{\dagger}c_{j\sigma} is the density operator, tt is the tunneling amplitude, and UU is the on-site interaction strength. Here δ\delta is the energy difference between the internal spin states, which can be set to zero without loss of generality sup . We assume that atoms tunnel only between nearest-neighbor sites. We impose the periodic boundary condition to ensure the translational invariance of the model, but the following results can be generalized to the case of the open boundary condition.

We consider a situation in which an atom in a spin-up state can decay to a spin-down state via spontaneous emission of a photon [see Fig. 1(a)]. Such a decay process can be induced by weakly hybridizing the spin-up state with an excited level of the atom [see Fig. 1(b)]. A detailed discussion will be given later and in Supplemental Material sup with a concrete experimental implementation. Within the Born-Markov approximation, we can trace out the photon degrees of freedom and obtain the Lindblad-type quantum master equation that describes the time evolution of the density matrix ρ\rho of the atomic gas as Breuer and Petruccione (2007)

dρdτ=ρ=i[H,ρ]+j(LjρLj12{LjLj,ρ}),\frac{d\rho}{d\tau}=\mathcal{L}\rho=-i[H,\rho]+\sum_{j}(L_{j}\rho L_{j}^{\dagger}-\frac{1}{2}\{L_{j}^{\dagger}L_{j},\rho\}), (2)

where τ\tau is the time, \mathcal{L} is the Liouvillian superoperator, and the Lindblad operator Lj=γcjcjL_{j}=\sqrt{\gamma}c_{j\downarrow}^{\dagger}c_{j\uparrow} describes a spontaneous emission process with rate γ>0\gamma>0 at site jj.

η\eta-pairing steady state.– The Hubbard Hamiltonian (1) possesses the η\eta SU(2) symmetry generated by ηx=(η++η)/2,ηy=(η+η)/2i\eta^{x}=(\eta^{+}+\eta^{-})/2,\eta^{y}=(\eta^{+}-\eta^{-})/2i, and ηz=12j(nj+nj1)\eta^{z}=\frac{1}{2}\sum_{j}(n_{j\uparrow}+n_{j\downarrow}-1), where η+=jei𝑸𝑹jcjcj\eta^{+}=\sum_{j}e^{i\bm{Q}\cdot\bm{R}_{j}}c_{j\uparrow}^{\dagger}c_{j\downarrow}^{\dagger} creates a doublon with crystal momentum 𝑸=(π,,π)\bm{Q}=(\pi,\cdots,\pi), and η=(η+)\eta^{-}=(\eta^{+})^{\dagger} Yang (1989); Yang and Zhang (1990). Thus, if a state |Ψ\ket{\Psi} is an eigenstate of HH, so is η+|Ψ\eta^{+}\ket{\Psi}. For example, Yang’s η\eta-pairing state |ψN(η+)N/2|0\ket{\psi_{N}}\equiv(\eta^{+})^{N/2}\ket{0}, where |0\ket{0} is the vacuum of fermions, is an NN-particle eigenstate of HH Yang (1989). The η\eta-pairing state can be regarded as a condensate of doublons at momentum 𝑸\bm{Q} and exhibits an off-diagonal long-range order Yang (1989, 1962). We also introduce a family of eigenstates |ψ2Np;𝒌1,,𝒌Nex(η+)Npc𝒌1c𝒌Nex|0\ket{\psi_{2N_{p}};\bm{k}_{1},\cdots,\bm{k}_{N_{\mathrm{ex}}}}\equiv(\eta^{+})^{N_{p}}c_{\bm{k}_{1}\downarrow}^{\dagger}\cdots c_{\bm{k}_{N_{\mathrm{ex}}}\downarrow}^{\dagger}\ket{0}, where c𝒌σ=1Nsjcjσei𝒌𝑹jc_{\bm{k}\sigma}=\frac{1}{\sqrt{N_{s}}}\sum_{j}c_{j\sigma}e^{-i\bm{k}\cdot\bm{R}_{j}} is the annihilation operator of a single-particle eigenstate with crystal momentum 𝒌\bm{k}, and NsN_{s} denotes the number of lattice sites. This state also shows the η\eta-pairing correlations but contains excess spin-down particles that do not form pairs.

Since Lj|ψ=0L_{j}\ket{\psi}=0 for |ψ=|ψN\ket{\psi}=\ket{\psi_{N}} and |ψ=|ψ2Np;𝒌1,,𝒌Nex\ket{\psi}=\ket{\psi_{2N_{p}};\bm{k}_{1},\cdots,\bm{k}_{N_{\mathrm{ex}}}}, we find that the η\eta-pairing states ρ=|ψψ|\rho=\ket{\psi}\bra{\psi} are steady states of the quantum master equation (2): |ψψ|=0\mathcal{L}\ket{\psi}\bra{\psi}=0. Equivalently, no photon is emitted from the η\eta-pairing states, implying that they are many-body dark states Diehl et al. (2008); Kraus et al. (2008). The steadiness of the η\eta-pairing states is physically understood as follows. First, on-site fermion pairs (doublons) are stabilized since a radiative decay of a spin-up particle is Pauli-blocked by a spin-down particle sitting at the same site [see Fig. 1(a)] Sandner et al. (2011). Second, although doublons may have varying center-of-mass momenta and decay via hopping processes, η\eta pairs never decay since they constitute eigenstates of the Hubbard Hamiltonian. Physically, this is understood as complete destructive interference between different decay processes of doublons, as can be seen from ti,jσ(ciσcjσ+H.c.)|ψN=0-t\sum_{\langle i,j\rangle}\sum_{\sigma}(c_{i\sigma}^{\dagger}c_{j\sigma}+\mathrm{H.c.})\ket{\psi_{N}}=0 sup . The above pairing mechanism is clearly distinct from the BCS mechanism which arises from the Fermi-surface instability due to an attractive interaction between fermions.

Symmetries, conserved quantity, and exact Liouvillian eigenmodes.– Since the system has many dark states, a general steady state is a statistical mixture of the η\eta-pairing states |ψN\ket{\psi_{N}} and |ψ2Np;𝒌1,,𝒌Nex\ket{\psi_{2N_{p}};\bm{k}_{1},\cdots,\bm{k}_{N_{\mathrm{ex}}}}. The steady-state distribution of the η\eta-pairing states depends on the initial state and the model parameters. To seek an optimal condition, we exploit the fact that the Lindblad operators commute with the generators of the η\eta SU(2) symmetry: [Lj,η+]=[Lj,η]=0[L_{j},\eta^{+}]=[L_{j},\eta^{-}]=0. This property and the η\eta SU(2) symmetry of the Hamiltonian lead to the conservation of

C\displaystyle C\equiv η+η\displaystyle\langle\eta^{+}\eta^{-}\rangle
=\displaystyle= jnjnj+i,j(ij)ei𝑸(𝑹i𝑹j)cicicjcj,\displaystyle\sum_{j}\langle n_{j\uparrow}n_{j\downarrow}\rangle+\sum_{i,j\ (i\neq j)}e^{i\bm{Q}\cdot(\bm{R}_{i}-\bm{R}_{j})}\langle c_{i\uparrow}^{\dagger}c_{i\downarrow}^{\dagger}c_{j\downarrow}c_{j\uparrow}\rangle, (3)

where 𝒪=Tr[𝒪ρ]\langle\mathcal{O}\rangle=\mathrm{Tr}[\mathcal{O}\rho] Buča and Prosen (2012); Albert and Jiang (2014); Tindall et al. (2019). Thus, to achieve large η\eta-pairing correlations in a steady state, we should prepare an initial state having sufficiently large CC. This implies that high-energy or high-temperature states rather than low-energy ones of the repulsive Hubbard model are appropriate for an initial state, since the former involve a large number of doublons which give a large jnjnj\sum_{j}\langle n_{j\uparrow}n_{j\downarrow}\rangle in Eq. (3), whereas the latter do not. The conservation of CC is similar to the situation in Ref. Tindall et al. (2019); however, we emphasize that the physical mechanism of the η\eta pairing is directly linked to the Fermi statistics and the destructive interference, but not to heating discussed in Ref. Tindall et al. (2019).

The time scale of the formation dynamics of the η\eta-pairing state is governed by the eigenvalues of the Liouvillian \mathcal{L}. Remarkably, some eigenvalues and the corresponding eigenmodes of the Liouvillian can exactly be obtained from the procedure in Refs. Torres (2014); Nakagawa et al. (2021). To see this, we decompose the Liouvillian as =𝒦+𝒥\mathcal{L}=\mathcal{K}+\mathcal{J}, where 𝒦ρi(HeffρρHeff)\mathcal{K}\rho\equiv-i(H_{\mathrm{eff}}\rho-\rho H_{\mathrm{eff}}^{\dagger}) describes an effective non-Hermitian Hamiltonian dynamics with HeffH(i/2)jLjLjH_{\mathrm{eff}}\equiv H-(i/2)\sum_{j}L_{j}^{\dagger}L_{j}, and 𝒥ρjLjρLj\mathcal{J}\rho\equiv\sum_{j}L_{j}\rho L_{j}^{\dagger} is the quantum-jump term. The non-Hermitian Hamiltonian HeffH_{\mathrm{eff}} commutes with the total magnetization operator Sz=j(njnj)/2S_{z}=\sum_{j}(n_{j\uparrow}-n_{j\downarrow})/2, while the quantum-jump term decreases the magnetization by one. It follows that in the basis that diagonalizes 𝒦\mathcal{K} the Liouvillian \mathcal{L} is a triangular-matrix form in which the off-diagonal elements are determined from 𝒥\mathcal{J}. Thus, any Liouvillian eigenvalue λ\lambda is the same as that of 𝒦\mathcal{K} and can be calculated from a pair of eigenvalues EaE_{a} and EbE_{b} of the non-Hermitian Hubbard model HeffH_{\mathrm{eff}} as λ=i(EaEb)\lambda=-i(E_{a}-E_{b}^{*}) sup . Such a feature does not exist in models with other Lindblad operators used in Refs. Diehl et al. (2008); Kraus et al. (2008); Bernier et al. (2013); Tindall et al. (2019). In particular, for the one-dimensional (d=1d=1) case, the present model is exactly solvable by the Bethe ansatz method Nakagawa et al. (2021); Essler et al. (2010).

The Liouvillian excitation spectrum of the η\eta-pairing steady states involves single-particle and collective excitations as in conventional BCS superfluids. For example, single-particle excitations used to show the metastability of the η\eta-pairing state Yang (1989) provide exact Liouvillian eigenmodes with degenerate eigenvalues λ=γ\lambda=-\gamma in any spatial dimensions sup . Such single-particle modes decay on the time scale of 1/γ1/\gamma. The late-stage dynamics near the steady state is governed by collective modes of the η\eta SU(2) pseudospins. In fact, the Bethe ansatz solution shows the dispersion relation of the collective modes as λ(P)=2Im16t2cos2(P/2)+(U+iγ/2)2γ2Im[J](1+cosP)\lambda(P)=2\mathrm{Im}\sqrt{16t^{2}\cos^{2}(P/2)+(U+i\gamma/2)^{2}}-\gamma\simeq 2\mathrm{Im}[J](1+\cos P) for PπP\simeq\pi, where PP is the momentum of the excitation and J=4t2/(U+iγ/2)J=4t^{2}/(U+i\gamma/2) is an effective exchange coupling between pseudospins, giving a gapless Liouvillian spectrum that implies a power-law tail in the relaxation dynamics in the thermodynamic limit Cai and Barthel (2013); sup . Note that the dispersion relation of the collective modes near P=πP=\pi is quadratic rather than linear in contrast to the Nambu-Goldstone mode of a conventional charge-neutral superfluid. This property highlights the difference between η\eta-pairing and BCS superfluids; the relevant symmetry of the former is SU(2), while that of the latter is U(1).

Moreover, by the Shiba transformation Shiba (1972) VcjV=cj,VcjV=ei𝑸𝑹jcjVc_{j\uparrow}V^{\dagger}=c_{j\uparrow},Vc_{j\downarrow}V^{\dagger}=e^{i\bm{Q}\cdot\bm{R}_{j}}c_{j\downarrow}^{\dagger}, the model discussed in this Letter can be mapped to a dissipative Hubbard model subject to two-body particle loss Nakagawa et al. (2020, 2021) described by a Lindblad operator ei𝑸𝑹jVLjV=γcjcje^{i\bm{Q}\cdot\bm{R}_{j}}VL_{j}V^{\dagger}=\sqrt{\gamma}c_{j\downarrow}c_{j\uparrow}. Thus, the Liouvillian spectrum of the model described in Eqs. (1) and (2) is equivalent to that of a dissipative Hubbard model subject to two-body particle loss after changing the sign of the interaction strength. The η\eta-pairing states of the model correspond to ferromagnetic steady states discussed in Refs. Nakagawa et al. (2020, 2021) through this mapping, while the physical mechanisms of their stabilization are different. This correspondence extends the Shiba symmetry between the repulsive and attractive Hubbard models to the symmetry between spontaneous emission and particle loss, making a nontrivial connection between different dissipative processes of many-body systems.

Refer to caption
Figure 2: Dynamics of the pair correlation function Re[C(j,0;τ)]\mathrm{Re}[C(j,0;\tau)] [(a), (c)] and the density ndn_{d} of doublons [(b), (d)] for the one-dimensional system with 8 sites and 8 particles. The initial state is the doublon-density-wave state for (a) and (b), and the infinite-temperature state for (c) and (d). The model parameters are set to t=1,U=10,δ=0t=1,U=10,\delta=0, and γ=10\gamma=10. The unit of time is τh=1/t\tau_{h}=1/t.

Dynamics of pair correlations.– We numerically solve the quantum master equation (2) in the one-dimensional case by the quantum-trajectory method Dalibard et al. (1992); Carmichael (1993); Dum et al. (1992); Daley (2014). We take 1000 quantum trajectories, where the time evolution is calculated by exact diagonalization of the non-Hermitian Hamiltonian HeffH_{\mathrm{eff}}. We consider an initial state ρ(0)=|ψ0ψ0|\rho(0)=\ket{\psi_{0}}\bra{\psi_{0}} of an 8-site system, where |ψ0=j=0,2,4,6cjcj|0\ket{\psi_{0}}=\prod_{j=0,2,4,6}c_{j\uparrow}^{\dagger}c_{j\downarrow}^{\dagger}\ket{0} is a density-wave state of doublons. Figure 2 (a) shows the dynamics of the pair correlation function C(j,0;τ)Tr[cjcjc0c0ρ(τ)]C(j,0;\tau)\equiv\mathrm{Tr}[c_{j\uparrow}^{\dagger}c_{j\downarrow}^{\dagger}c_{0\downarrow}c_{0\uparrow}\rho(\tau)]. In the steady state, the amplitude of the pair correlation function does not depend on sites and its sign alternates from site to site, indicating the formation of η\eta pairs with the center-of-mass momentum 𝑸\bm{Q}. The imaginary part of the pair correlation function is negligible in the steady state, while the real part does not significantly depend on the model parameters sup . Figure 2(b) shows the dynamics of the density of doublons nd(1/Ns)jnjnjn_{d}\equiv(1/N_{s})\sum_{j}\langle n_{j\uparrow}n_{j\downarrow}\rangle. After a long time, the density of doublons reaches a nonzero steady-state value. Since the number of doublons in the initial state is 𝒪(Ns)\mathcal{O}(N_{s}) and C0C\geq 0, the amplitude of the pair correlations in the steady states is estimated to be 𝒪(1/Ns)\mathcal{O}(1/N_{s}), which can be observed as a long-range order in finite-size systems such as trapped atoms Tindall et al. (2021a); Mazurenko et al. (2017). The dynamics in Fig. 2 exhibits a two-step relaxation. Since the decay rate γ\gamma is sufficiently larger than the hopping amplitude, the pair correlations initially increase over the time scale of hopping which breaks the doublons in the initial state, and the slow relaxation at a late stage near the steady state is governed by the collective modes with an exchange coupling Im[J]=0.16\mathrm{Im}[J]=-0.16.

We also study the dynamics from an infinite-temperature initial state ρ(0)=I/dH\rho(0)=I/d_{\mathrm{H}}, where II and dHd_{\mathrm{H}} are the identity operator and the dimension of the Hilbert space of the 8-particle sector, respectively. To numerically simulate the dynamics with the quantum-trajectory method, we sample a random pure initial state in each quantum trajectory. Figures 2 (c) and (d) show the dynamics of the pair correlations and that of the density of doublons, respectively. Notably, the η\eta-pairing correlations develop even if the dynamics starts from the infinite-temperature initial state. This behavior is in marked contrast to the heating-induced η\eta-pairing mechanism in Ref. Tindall et al. (2019), where the η\eta paring cannot develop from the infinite-temperature state because of the Hermiticity of Lindblad operators.

Doublon momentum distribution.– The formation of an η\eta-pairing state can be detected from the momentum distribution of doublons, which can be measured with time-of-flight techniques after doublons are converted into molecules by a Feshbach resonance Regal et al. (2004); Zwierlein et al. (2004); Stöferle et al. (2006); Altman and Vishwanath (2005); Perali et al. (2005); Matyjaśkiewicz et al. (2008). The momentum distribution of doublons is defined as R𝒌d𝒌d𝒌/𝒌d𝒌d𝒌R_{\bm{k}}\equiv\langle d_{\bm{k}}^{\dagger}d_{\bm{k}}\rangle/\sum_{\bm{k}}\langle d_{\bm{k}}^{\dagger}d_{\bm{k}}\rangle, where d𝒌(1/Ns)jcjcjei𝒌𝑹jd_{\bm{k}}\equiv(1/\sqrt{N_{s}})\sum_{j}c_{j\downarrow}c_{j\uparrow}e^{-i\bm{k}\cdot\bm{R}_{j}} is the annihilation operator of a doublon with momentum 𝒌\bm{k}. Here, we have normalized the momentum distribution by the total number of doublons such that 0R𝒌10\leq R_{\bm{k}}\leq 1 is satisfied, since the number of doublons is not conserved during the dynamics. Figure 3 shows the dynamics of the doublon momentum distribution for the two initial states as in Fig. 2. The build-up dynamics of the distribution at 𝒌=±𝑸\bm{k}=\pm\bm{Q} clearly signals the formation of η\eta pairing.

In Fig. 3, one can see that the doublons for 𝒌±𝑸\bm{k}\neq\pm\bm{Q} are still populated after a long time. The residual doublon distribution is not due to imperfection of the η\eta-pairing formation but arises from a fundamental constraint due to the Fermi statistics. In fact, the doublon momentum distribution can be bounded from above as d𝒌d𝒌Λ2/2\langle d_{\bm{k}}^{\dagger}d_{\bm{k}}\rangle\leq\Lambda_{2}/2 sup , where Λ2\Lambda_{2} is the maximum eigenvalue of the two-particle reduced density matrix (ρ2)iσ1jσ2;kσ3lσ4cjσ2ciσ1ckσ3clσ4(\rho_{2})_{i\sigma_{1}j\sigma_{2};k\sigma_{3}l\sigma_{4}}\equiv\langle c_{j\sigma_{2}}^{\dagger}c_{i\sigma_{1}}^{\dagger}c_{k\sigma_{3}}c_{l\sigma_{4}}\rangle Yang (1962). Using the bound of Λ2\Lambda_{2} from the Pauli exclusion principle Yang (1962), we obtain

d𝒌d𝒌N(2NsN+2)4Ns,\langle d_{\bm{k}}^{\dagger}d_{\bm{k}}\rangle\leq\frac{N(2N_{s}-N+2)}{4N_{s}}, (4)

where the bound is saturated by Yang’s η\eta-pairing state |ψN\ket{\psi_{N}} for 𝒌=𝑸\bm{k}=\bm{Q} sup . Moreover, provided that the model does not have a dark state other than |ψN\ket{\psi_{N}} and |ψNp;𝒌1,,𝒌Nex\ket{\psi_{N_{p}};\bm{k}_{1},\cdots,\bm{k}_{N_{\mathrm{ex}}}}, we can derive a bound on R𝑸R_{\bm{Q}} in the steady state as sup

R𝑸1N2Ns+1Ns.R_{\bm{Q}}\leq 1-\frac{N}{2N_{s}}+\frac{1}{N_{s}}. (5)

The bound (5) is consistent with the numerical results in Fig. 3. If we take the thermodynamic limit NsN_{s}\to\infty while keeping the density nN/(2Ns)n\equiv N/(2N_{s}) constant, the inequality (5) reduces to R𝑸1nR_{\bm{Q}}\leq 1-n. Thus, the normalized doublon momentum distribution at momentum ±𝑸\pm\bm{Q} cannot reach unity for a system with nonzero density because of the Pauli exclusion principle, indicating a fundamental distinction between doublons and bosons sup .

Refer to caption
Figure 3: Dynamics of the normalized momentum distribution RkR_{k} of doublons for (a) the doublon-density-wave initial state and (b) the infinite-temperature initial state. The model parameters are the same as those in Fig. 2. The unit of time is τh=1/t\tau_{h}=1/t.

Experimental implementation.– The validity of the Hubbard model for the description of ultracold fermionic atoms in an optical lattice has been tested by a number of experiments Esslinger (2010). In particular, it has been experimentally verified in Ref. Gall et al. (2020) that a cold-atom quantum simulator of the Hubbard model possesses the Shiba symmetry Shiba (1972); Ho et al. (2009) between the repulsive and attractive Hubbard models, which leads to the η\eta SU(2) symmetry when combined with the spin SU(2) symmetry Yang and Zhang (1990). The Lindblad operator LjL_{j} for spontaneous decay from a spin-up state to a spin-down state can be implemented by weakly coupling the spin-up state to an excited state that can decay to the spin-down state [see Fig. 1(b)] Sandner et al. (2011); sup . With large detuning of the laser that couples the spin-up state and the excited state, the latter can be adiabatically eliminated Pichler et al. (2010); Sarkar et al. (2014); Reiter and Sørensen (2012), and the model (2) is obtained sup . To avoid recoil of atoms that induces transitions to neighboring sites or higher bands, a sufficiently deep optical lattice should be used sup ; Pichler et al. (2010); Sarkar et al. (2014).

In this implementation, a spontaneous decay from the excited state to the spin-up state leads to an additional Lindblad operator Lj(d)=γdnjL_{j}^{(d)}=\sqrt{\gamma_{d}}n_{j\uparrow}, which causes dephasing of atoms. Since the inverse of the dephasing rate γd\gamma_{d} determines the lifetime of the η\eta-pairing states as confirmed numerically sup , the condition γdγ,t\gamma_{d}\ll\gamma,t needs to be met. A possible candidate that achieves this condition is alkaline-earth-like fermionic atoms such as 171Yb, 173Yb, and 87Sr sup . We can implement the Fermi-Hubbard model by trapping spin-polarized atoms in the ground S01{}^{1}S_{0} state and in the metastable P03{}^{3}P_{0} state in a magic-wavelength optical lattice Gorshkov et al. (2010). We can also realize an effective spontaneous decay by coupling the P03{}^{3}P_{0} state to a particular hyperfine state in the P11{}^{1}P_{1} state that can quickly decay to the S01{}^{1}S_{0} state Sandner et al. (2011). Since the spontaneous decay from the P11{}^{1}P_{1} state to the P03{}^{3}P_{0} state is negligible in the experimental time scale, the condition γdγ,t\gamma_{d}\ll\gamma,t can well be satisfied.

Summary.– In this Letter, we have proposed a nonequilibrium pairing mechanism for fermionic atoms using spontaneous emission of light. The nonequilibrium η\eta pairing has been shown to be formed through a non-BCS mechanism, where the Pauli blocking of radiative decay of on-site fermion pairs and the destructive interference between doublon-decay processes play vital roles. We have shown that η\eta-pairing correlations in the steady state are supported by high density of doublons in the initial state, which indicates that coherent η\eta pairs can be created at temperatures comparable to or even higher than the Fermi temperature, as long as the single-band approximation is valid. Since charge-neutral atoms do not have the dynamical instability of η\eta-pairing states due to coupling with dynamical electromagnetic fields as shown recently Tsuji et al. (2021), the scheme proposed in this Letter for ultracold atoms serves as a promising candidate for observation of the η\eta-pairing state well above the BCS critical temperature.

Acknowledgements.
This work was supported by KAKENHI (Grant Nos. JP18H01140, JP18H01145, JP19H01838, JP20K03811, and JP20K14383) and a Grant-in-Aid for Scientific Research on Innovative Areas (KAKENHI Grant No. JP15H05855) from the Japan Society for the Promotion of Science.

References

Supplemental Material for
η\eta Pairing of Light-Emitting Fermions: Nonequilibrium Pairing Mechanism at High Temperatures”

Appendix A Eliminating the energy difference between spin states

We show that the energy difference δ\delta between internal spin states in the model (1) does not affect the dynamics of the system. We decompose the Hamiltonian as

H\displaystyle H =H0+Hδ,\displaystyle=H_{0}+H_{\delta}, (S1)
H0\displaystyle H_{0} ti,jσ(ciσcjσ+H.c.)+Ujnjnj,\displaystyle\equiv-t\sum_{\langle i,j\rangle}\sum_{\sigma}(c_{i\sigma}^{\dagger}c_{j\sigma}+\mathrm{H.c.})+U\sum_{j}n_{j\uparrow}n_{j\downarrow}, (S2)
Hδ\displaystyle H_{\delta} δj(njnj),\displaystyle\equiv\delta\sum_{j}(n_{j\uparrow}-n_{j\downarrow}), (S3)

and introduce the interaction representation

ρ~(τ)eiHδτρ(τ)eiHδτ.\tilde{\rho}(\tau)\equiv e^{iH_{\delta}\tau}\rho(\tau)e^{-iH_{\delta}\tau}. (S4)

Substituting Eqs. (S1)-(S4) into the quantum master equation (2), we obtain

dρ~dτ=i[H0,ρ~]+j(Ljρ~Lj12{LjLj,ρ~}).\frac{d\tilde{\rho}}{d\tau}=-i[H_{0},\tilde{\rho}]+\sum_{j}(L_{j}\tilde{\rho}L_{j}^{\dagger}-\frac{1}{2}\{L_{j}^{\dagger}L_{j},\tilde{\rho}\}). (S5)

For a physical quantity 𝒪\mathcal{O} that commutes with the magnetization operator SzS_{z}, we have

Tr[𝒪ρ~(τ)]=Tr[𝒪ρ(τ)].\mathrm{Tr}[\mathcal{O}\tilde{\rho}(\tau)]=\mathrm{Tr}[\mathcal{O}\rho(\tau)]. (S6)

Thus, the expectation value of an observable is independent of δ\delta and therefore we may set δ=0\delta=0 without loss of generality.

Appendix B Destructive interference between doublon-decay processes

Here we show that the probability amplitude of decay processes of doublons vanishes for the η\eta-pairing state. We define a hopping operator for spin-σ\sigma particles by Tijσt(ciσcjσ+H.c.)T_{ij\sigma}\equiv-t(c_{i\sigma}^{\dagger}c_{j\sigma}+\mathrm{H.c.}), where ii and jj denote a pair of nearest-neighbor sites. Then, the commutation relations of the hopping operator and the η\eta operator are calculated as

[Tij,η+]=tei𝑸𝑹j(cicj+ei𝑸(𝑹i𝑹j)cjci),[T_{ij\uparrow},\eta^{+}]=-te^{i\bm{Q}\cdot\bm{R}_{j}}(c_{i\uparrow}^{\dagger}c_{j\downarrow}^{\dagger}+e^{i\bm{Q}\cdot(\bm{R}_{i}-\bm{R}_{j})}c_{j\uparrow}^{\dagger}c_{i\downarrow}^{\dagger}), (S7)

and

[Tij,η+]=tei𝑸𝑹j(ei𝑸(𝑹i𝑹j)cicj+cjci).[T_{ij\downarrow},\eta^{+}]=-te^{i\bm{Q}\cdot\bm{R}_{j}}(e^{i\bm{Q}\cdot(\bm{R}_{i}-\bm{R}_{j})}c_{i\uparrow}^{\dagger}c_{j\downarrow}^{\dagger}+c_{j\uparrow}^{\dagger}c_{i\downarrow}^{\dagger}). (S8)

Since the right-hand sides of Eqs. (S7) and (S8) both commute with η+\eta^{+}, we have

[Tijσ,(η+)n]=n(η+)n1[Tijσ,η+],[T_{ij\sigma},(\eta^{+})^{n}]=n(\eta^{+})^{n-1}[T_{ij\sigma},\eta^{+}], (S9)

which leads to

Tijσ|ψN=N2(η+)N/21[Tijσ,η+]|0,T_{ij\sigma}\ket{\psi_{N}}=\frac{N}{2}(\eta^{+})^{N/2-1}[T_{ij\sigma},\eta^{+}]\ket{0}, (S10)

where |ψN=(η+)N/2|0\ket{\psi_{N}}=(\eta^{+})^{N/2}\ket{0} is Yang’s η\eta-pairing state. Here, we invoke the fact that the center-of-mass momentum 𝑸=(π,,π)\bm{Q}=(\pi,\cdots,\pi) leads to ei𝑸(𝑹i𝑹j)=1e^{i\bm{Q}\cdot(\bm{R}_{i}-\bm{R}_{j})}=-1 for any pair of nearest-neighbor sites ii and jj. Thus, from Eqs. (S7), (S8), and (S10), we obtain

Tij|ψN=T\ij|ψN,T_{ij\uparrow}\ket{\psi_{N}}=-T_{\ij\downarrow}\ket{\psi_{N}}, (S11)

which indicates a vanishing probability amplitude of the doublon-decay processes, i.e., tσ(ciσcjσ+H.c.)|ψN=0-t\sum_{\sigma}(c_{i\sigma}^{\dagger}c_{j\sigma}+\mathrm{H.c.})\ket{\psi_{N}}=0. As clearly seen from Eq. (S11), the physical origin of this vanishing probability amplitude is that the minus sign [ei𝑸(𝑹i𝑹j)=1e^{i\bm{Q}\cdot(\bm{R}_{i}-\bm{R}_{j})}=-1] arising from the η\eta-pair wave function leads to complete destructive interference between two paths of doublon-decay processes described by TijT_{ij\uparrow} and TijT_{ij\downarrow}.

Appendix C Exact Liouvillian eigenmodes

The Liouvillian superoperator and its conjugate are defined as ρ=i[H,ρ]+j(LjρLj12{LjLj,ρ})\mathcal{L}\rho=-i[H,\rho]+\sum_{j}(L_{j}\rho L_{j}^{\dagger}-\frac{1}{2}\{L_{j}^{\dagger}L_{j},\rho\}) and ρ=i[H,ρ]+j(LjρLj12{LjLj,ρ})\mathcal{L}^{\dagger}\rho=i[H,\rho]+\sum_{j}(L_{j}^{\dagger}\rho L_{j}-\frac{1}{2}\{L_{j}^{\dagger}L_{j},\rho\}), respectively. Suppose that the Liouvillian is diagonalized as ρnR=λnρnR\mathcal{L}\rho_{n}^{\mathrm{R}}=\lambda_{n}\rho_{n}^{\mathrm{R}} and ρnL=λnρnL\mathcal{L}^{\dagger}\rho_{n}^{\mathrm{L}}=\lambda_{n}^{*}\rho_{n}^{\mathrm{L}}, where ρnR\rho_{n}^{\mathrm{R}} (ρnL\rho_{n}^{\mathrm{L}}) denotes the nnth right (left) eigenmode. Then, the time evolution of a density matrix can be written as

ρ(τ)=ncneλnτρnR,\rho(\tau)=\sum_{n}c_{n}e^{\lambda_{n}\tau}\rho_{n}^{\mathrm{R}}, (S12)

where the coefficient reads cn=Tr[(ρnL)ρ(0)]c_{n}=\mathrm{Tr}[(\rho_{n}^{\mathrm{L}})^{\dagger}\rho(0)] if the biorthonormal condition Tr[(ρmL)ρnR]=δm,n\mathrm{Tr}[(\rho_{m}^{\mathrm{L}})^{\dagger}\rho_{n}^{\mathrm{R}}]=\delta_{m,n} is imposed. As mentioned in the main text, the Liouvillian =𝒦+𝒥\mathcal{L}=\mathcal{K}+\mathcal{J} can be made to take a triangular-matrix form whose diagonal (off-diagonal) components arise from 𝒦\mathcal{K} (𝒥\mathcal{J}). Since eigenvalues of a triangular matrix are given by its diagonal elements, the Liouvillian eigenvalues coincide with those of 𝒦\mathcal{K}. Thus, eigenvalues λ\lambda of the Liouvillian are given as λ=i(EaEb)\lambda=-i(E_{a}-E_{b}^{*}), where EaE_{a} and EbE_{b} are eigenvalues of the non-Hermitian Hamiltonian

Heff=\displaystyle H_{\mathrm{eff}}= Hi2jLjLj\displaystyle H-\frac{i}{2}\sum_{j}L_{j}^{\dagger}L_{j}
=\displaystyle= ti,jσ(ciσcjσ+H.c.)+(U+iγ2)jnjnj+(δiγ2)jnjδjnj,\displaystyle-t\sum_{\langle i,j\rangle}\sum_{\sigma}(c_{i\sigma}^{\dagger}c_{j\sigma}+\mathrm{H.c.})+\left(U+\frac{i\gamma}{2}\right)\sum_{j}n_{j\uparrow}n_{j\downarrow}+\left(\delta-\frac{i\gamma}{2}\right)\sum_{j}n_{j\uparrow}-\delta\sum_{j}n_{j\downarrow}, (S13)

and the corresponding eigenmodes of the Liouvillian can be obtained from the eigenstates of HeffH_{\mathrm{eff}} and the matrix elements of LjL_{j} between them Torres (2014); Nakagawa et al. (2021).

Single-particle excitations from the η\eta-pairing state |ψN\ket{\psi_{N}} were constructed by Yang Yang (1989). We define

|ζN,𝒂(η+)N/21η𝒂+|0,\displaystyle\ket{\zeta_{N,\bm{a}}}\equiv(\eta^{+})^{N/2-1}\eta_{\bm{a}}^{+}\ket{0}, (S14)
η𝒂+𝒌ei𝒌𝒂c𝒌c𝑸𝒌,\displaystyle\eta_{\bm{a}}^{+}\equiv\sum_{\bm{k}}e^{-i\bm{k}\cdot\bm{a}}c_{\bm{k}\uparrow}^{\dagger}c_{\bm{Q}-\bm{k}\downarrow}^{\dagger}, (S15)

for 𝒂{0,,L1}d\bm{a}\in\{0,\cdots,L-1\}^{d} with 𝒂(0,,0)\bm{a}\neq(0,\cdots,0). Here |ζN,𝒂\ket{\zeta_{N,\bm{a}}} is an exact NN-particle eigenstate of HeffH_{\mathrm{eff}} in any spatial dimension:

Heff|ζN,𝒂=[NU2Uiγ2]|ζN,𝒂,H_{\mathrm{eff}}\ket{\zeta_{N,\bm{a}}}=\left[\frac{NU}{2}-U-\frac{i\gamma}{2}\right]\ket{\zeta_{N,\bm{a}}}, (S16)

which indicates that |ζN,𝒂\ket{\zeta_{N,\bm{a}}} is a single-particle excited state with the same momentum as the η\eta-pairing state Yang (1989). The gap Re[Uiγ/2]=U\mathrm{Re}[-U-i\gamma/2]=-U in the excitation energy is related to the metastability of the η\eta-pairing state for U<0U<0 in an isolated system Yang (1989). To promote the eigenstates of the non-Hermitian Hamiltonian to eigenmodes of the Liouvillian, we assume the form of an eigenmode as

ρN,𝒂,𝒂R=CN,𝒂,𝒂|ζN,𝒂ζN,𝒂|+𝒌1,𝒌2,𝒌1,𝒌2D𝒌1,𝒌2,𝒌1,𝒌2(N,𝒂,𝒂)|ψN2;𝒌1,𝒌2ψN2;𝒌1,𝒌2|,\rho_{N,\bm{a},\bm{a}^{\prime}}^{\mathrm{R}}=C_{N,\bm{a},\bm{a}^{\prime}}\ket{\zeta_{N,\bm{a}}}\bra{\zeta_{N,\bm{a}^{\prime}}}+\sum_{\bm{k}_{1},\bm{k}_{2},\bm{k}^{\prime}_{1},\bm{k}^{\prime}_{2}}D_{\bm{k}_{1},\bm{k}_{2},\bm{k}^{\prime}_{1},\bm{k}^{\prime}_{2}}^{(N,\bm{a},\bm{a}^{\prime})}\ket{\psi_{N-2};\bm{k}_{1},\bm{k}_{2}}\bra{\psi_{N-2};\bm{k}^{\prime}_{1},\bm{k}^{\prime}_{2}}, (S17)

which is obtained from the general form of solutions in Refs. Torres (2014); Nakagawa et al. (2021). Note that Lj|ψN2;𝒌1,𝒌2=0L_{j}\ket{\psi_{N-2};\bm{k}_{1},\bm{k}_{2}}=0. Substituting Eq. (S17) into the eigenvalue equation ρN,𝒂,𝒂R=λN,𝒂,𝒂ρN,𝒂,𝒂R\mathcal{L}\rho_{N,\bm{a},\bm{a}^{\prime}}^{\mathrm{R}}=\lambda_{N,\bm{a},\bm{a}^{\prime}}\rho_{N,\bm{a},\bm{a}^{\prime}}^{\mathrm{R}}, we obtain

λN,𝒂,𝒂=γ,\displaystyle\lambda_{N,\bm{a},\bm{a}^{\prime}}=-\gamma, (S18)

and

D𝒌1,𝒌2,𝒌1,𝒌2(N,𝒂,𝒂)=γCN,𝒂,𝒂Ns(λN,𝒂,𝒂λ𝒌1,𝒌2,𝒌1,𝒌2)δ𝒌1+𝒌2,𝒌1+𝒌2ei(𝑸𝒌2)𝒂ei(𝑸𝒌2)𝒂,\displaystyle D_{\bm{k}_{1},\bm{k}_{2},\bm{k}^{\prime}_{1},\bm{k}^{\prime}_{2}}^{(N,\bm{a},\bm{a}^{\prime})}=\frac{\gamma C_{N,\bm{a},\bm{a}^{\prime}}}{N_{s}(\lambda_{N,\bm{a},\bm{a}^{\prime}}-\lambda_{\bm{k}_{1},\bm{k}_{2},\bm{k}^{\prime}_{1},\bm{k}^{\prime}_{2}})}\delta_{\bm{k}_{1}+\bm{k}_{2},\bm{k}^{\prime}_{1}+\bm{k}^{\prime}_{2}}e^{-i(\bm{Q}-\bm{k}_{2})\cdot\bm{a}}e^{i(\bm{Q}-\bm{k}^{\prime}_{2})\cdot\bm{a}^{\prime}}, (S19)

where λ𝒌1,𝒌2,𝒌1,𝒌2=2itμ=1d(cosk1μ+cosk2μcosk1μcosk2μ)\lambda_{\bm{k}_{1},\bm{k}_{2},\bm{k}^{\prime}_{1},\bm{k}^{\prime}_{2}}=2it\sum_{\mu=1}^{d}(\cos k_{1\mu}+\cos k_{2\mu}-\cos k^{\prime}_{1\mu}-\cos k^{\prime}_{2\mu}). The normalization constant is given by

CN,𝒂,𝒂=(ζN,𝒂|ζN,𝒂ζN,𝒂|ζN,𝒂)1/2=(NsN/21)!Ns(N/21)!(Ns2)!.C_{N,\bm{a},\bm{a}^{\prime}}=(\braket{\zeta_{N,\bm{a}}}{\zeta_{N,\bm{a}}}\braket{\zeta_{N,\bm{a}^{\prime}}}{\zeta_{N,\bm{a}^{\prime}}})^{-1/2}=\frac{(N_{s}-N/2-1)!}{N_{s}(N/2-1)!(N_{s}-2)!}. (S20)

Since the eigenvalue (S18) does not depend on NN, 𝒂\bm{a}, and 𝒂\bm{a}^{\prime}, the single-particle excitations decay at a constant rate γ\gamma.

The η\eta-pairing state breaks the η\eta SU(2) symmetry. This implies that the system hosts collective Nambu-Goldstone modes. An exact expression of the dispersion relation of the collective modes can be obtained for a one-dimensional system by the Bethe ansatz method Nakagawa et al. (2021). We consider an NN-particle eigenstate of the form |ϕN(η+)N/21|ϕ2\ket{\phi_{N}}\equiv(\eta^{+})^{N/2-1}\ket{\phi_{2}}, where |ϕ2\ket{\phi_{2}} is a two-particle eigenstate. The Bethe equations for a two-particle system are given by Essler et al. (2010)

eik1L=\displaystyle e^{ik_{1}L}= sink1sink2+2iusink1sink22iu,\displaystyle\frac{\sin k_{1}-\sin k_{2}+2iu}{\sin k_{1}-\sin k_{2}-2iu}, (S21)
eik2L=\displaystyle e^{ik_{2}L}= sink2sink1+2iusink2sink12iu,\displaystyle\frac{\sin k_{2}-\sin k_{1}+2iu}{\sin k_{2}-\sin k_{1}-2iu}, (S22)

where k1k_{1} and k2k_{2} denote quasimomenta, and u(U+iγ/2)/(4t)u\equiv(U+i\gamma/2)/(4t). The collective excitations are obtained from string solutions in which the quasimomenta take complex values. Without loss of generality, we assume that k1=q1iκk_{1}=q_{1}-i\kappa and k2=q2+iκ(q1,q2,κ>0)k_{2}=q_{2}+i\kappa\ (q_{1},q_{2}\in\mathbb{R},\kappa>0). Then, taking the limit of LL\to\infty in Eqs. (S21) and (S22), we have

sink1sink2=2iu,\sin k_{1}-\sin k_{2}=2iu, (S23)

the real and imaginary parts of which read

cosP2sinq2coshκ=\displaystyle\cos\frac{P}{2}\sin\frac{q}{2}\cosh\kappa= Im[u],\displaystyle-\mathrm{Im}[u], (S24)
cosP2cosq2sinhκ=\displaystyle\cos\frac{P}{2}\cos\frac{q}{2}\sinh\kappa= Re[u],\displaystyle-\mathrm{Re}[u], (S25)

where Pq1+q2P\equiv q_{1}+q_{2} is the momentum eigenvalue, and qq1q2q\equiv q_{1}-q_{2}. Using the solution of Eqs. (S24) and (S25), the two-particle eigenstate |ϕ2\ket{\phi_{2}} is given by |ϕ2=x1,x2ψ(x1,x2)cx1cx2|0\ket{\phi_{2}}=\sum_{x_{1},x_{2}}\psi(x_{1},x_{2})c_{x_{1}\uparrow}^{\dagger}c_{x_{2}\downarrow}^{\dagger}\ket{0}, where

ψ(x1,x2)=A[θ(x2x1)eiq1x1+iq2x2+θ(x1x2)eiq1x2+iq2x1]eκ|x1x2|\psi(x_{1},x_{2})=A[\theta(x_{2}-x_{1})e^{iq_{1}x_{1}+iq_{2}x_{2}}+\theta(x_{1}-x_{2})e^{iq_{1}x_{2}+iq_{2}x_{1}}]e^{-\kappa|x_{1}-x_{2}|} (S26)

is the Bethe wavefunction Essler et al. (2010), AA is the normalization constant, and θ(x)\theta(x) is the Heaviside unit-step function. The energy eigenvalue E(P)E(P) of |ϕN\ket{\phi_{N}} is obtained by using that of |ϕ2\ket{\phi_{2}} and the η\eta SU(2) symmetry:

E(P)=\displaystyle E(P)= 2t(cosk1+cosk2)iγ2+(N21)U\displaystyle-2t(\cos k_{1}+\cos k_{2})-\frac{i\gamma}{2}+\left(\frac{N}{2}-1\right)U
=\displaystyle= 4tcosP2cos(q2iκ)iγ2+(N21)U\displaystyle-4t\cos\frac{P}{2}\cos\left(\frac{q}{2}-i\kappa\right)-\frac{i\gamma}{2}+\left(\frac{N}{2}-1\right)U
=\displaystyle= 16t2cos2P2+(U+iγ2)2iγ2+(N21)U,\displaystyle\sqrt{16t^{2}\cos^{2}\frac{P}{2}+\left(U+\frac{i\gamma}{2}\right)^{2}}-\frac{i\gamma}{2}+\left(\frac{N}{2}-1\right)U, (S27)

where we take the branch of the square root so as to make its imaginary part positive. The Liouvillian eigenmode for a collective excitation with momentum PP is given by

ρNR(P)=CN|ϕNϕN|+𝒌1,𝒌2,𝒌1,𝒌2D𝒌1,𝒌2,𝒌1,𝒌2(N)|ψN2;𝒌1,𝒌2ψN2;𝒌1,𝒌2|,\displaystyle\rho_{N}^{\mathrm{R}}(P)=C_{N}\ket{\phi_{N}}\bra{\phi_{N}}+\sum_{\bm{k}_{1},\bm{k}_{2},\bm{k}^{\prime}_{1},\bm{k}^{\prime}_{2}}D_{\bm{k}_{1},\bm{k}_{2},\bm{k}^{\prime}_{1},\bm{k}^{\prime}_{2}}^{(N)}\ket{\psi_{N-2};\bm{k}_{1},\bm{k}_{2}}\bra{\psi_{N-2};\bm{k}^{\prime}_{1},\bm{k}^{\prime}_{2}}, (S28)

and its eigenvalue is

λ(P)=i(E(P)E(P))=2Im16t2cos2P2+(U+iγ2)2γ.\displaystyle\lambda(P)=-i(E(P)-E^{*}(P))=2\mathrm{Im}\sqrt{16t^{2}\cos^{2}\frac{P}{2}+\left(U+\frac{i\gamma}{2}\right)^{2}}-\gamma. (S29)

The coefficients CNC_{N} and D𝒌1,𝒌2,𝒌1,𝒌2(N)D_{\bm{k}_{1},\bm{k}_{2},\bm{k}^{\prime}_{1},\bm{k}^{\prime}_{2}}^{(N)} can be determined from the general procedure in Refs. Torres (2014); Nakagawa et al. (2021). In particular, for PπP\simeq\pi, the Liouvillian eigenvalue reads

λ(P)2Im[J](1+cosP),\lambda(P)\simeq 2\mathrm{Im}[J](1+\cos P), (S30)

where J=4t2/(U+iγ/2)J=4t^{2}/(U+i\gamma/2) is an effective exchange coupling between the η\eta SU(2) pseudospins. The dispersion relation is gapless and quadratic near P=πP=\pi. The collective excitations with the quadratic dispersion are understood as pseudospin waves of the η\eta SU(2) pseudospins, which are similar to magnons of the ferromagnetic Heisenberg model. In fact, by the Shiba transformation, η\eta SU(2) pseudospins are mapped to ordinary SU(2) spins and the collective mode corresponds to a ferromagnetic spin wave in a dissipative Hubbard model subject to two-body loss Nakagawa et al. (2021).

Appendix D Derivation of the bound on the doublon momentum distribution

Here, we derive the bounds [Eqs. (4) and (5) in the main text] on the momentum distribution of doublons. We define a two-particle reduced density matrix ρ2\rho_{2} by

(ρ2)iσ1jσ2,kσ3lσ4Tr[cjσ2ciσ1ckσ3clσ4ρ](\rho_{2})_{i\sigma_{1}j\sigma_{2},k\sigma_{3}l\sigma_{4}}\equiv\mathrm{Tr}[c_{j\sigma_{2}}^{\dagger}c_{i\sigma_{1}}^{\dagger}c_{k\sigma_{3}}c_{l\sigma_{4}}\rho] (S31)

for a density matrix ρ\rho of an NN-particle system Yang (1962). The two-particle reduced density matrix ρ2\rho_{2} is Hermitian and positive semidefinite Yang (1962). Let Λ2\Lambda_{2} and v=(viσ1jσ2)v=(v_{i\sigma_{1}j\sigma_{2}}) be the maximum eigenvalue and the corresponding eigenvector of ρ2\rho_{2} that satisfy

k,lσ3,σ4(ρ2)iσ1jσ2,kσ3lσ4vkσ3lσ4=Λ2viσ1jσ2.\sum_{k,l}\sum_{\sigma_{3},\sigma_{4}}(\rho_{2})_{i\sigma_{1}j\sigma_{2},k\sigma_{3}l\sigma_{4}}v_{k\sigma_{3}l\sigma_{4}}=\Lambda_{2}v_{i\sigma_{1}j\sigma_{2}}. (S32)

Then, the inequality

f,ρ2fi,j,k,lσ1,,σ4fiσ1jσ2(ρ2)iσ1jσ2,kσ3lσ4fkσ3lσ4Λ2\langle f,\rho_{2}f\rangle\equiv\sum_{i,j,k,l}\sum_{\sigma_{1},\cdots,\sigma_{4}}f_{i\sigma_{1}j\sigma_{2}}^{*}(\rho_{2})_{i\sigma_{1}j\sigma_{2},k\sigma_{3}l\sigma_{4}}f_{k\sigma_{3}l\sigma_{4}}\leq\Lambda_{2} (S33)

holds for any vector f=(fiσ1jσ2)f=(f_{i\sigma_{1}j\sigma_{2}}) which is normalized as i,jσ1,σ2|fiσ1jσ2|2=1\sum_{i,j}\sum_{\sigma_{1},\sigma_{2}}|f_{i\sigma_{1}j\sigma_{2}}|^{2}=1. To derive the bound on d𝒌d𝒌\langle d_{\bm{k}}^{\dagger}d_{\bm{k}}\rangle, we set

fiσ1jσ2=12Nsδi,jei𝒌𝑹i(δσ1,δσ2,δσ1,δσ2,),\displaystyle f_{i\sigma_{1}j\sigma_{2}}=\frac{1}{\sqrt{2N_{s}}}\delta_{i,j}e^{-i\bm{k}\cdot\bm{R}_{i}}(\delta_{\sigma_{1},\downarrow}\delta_{\sigma_{2},\uparrow}-\delta_{\sigma_{1},\uparrow}\delta_{\sigma_{2},\downarrow}), (S34)

and define

Fi,jσ1,σ2fiσ1jσ2ciσ1cjσ2=2Nsjei𝒌𝑹jcjcj.\displaystyle F\equiv\sum_{i,j}\sum_{\sigma_{1},\sigma_{2}}f_{i\sigma_{1}j\sigma_{2}}c_{i\sigma_{1}}c_{j\sigma_{2}}=\frac{\sqrt{2}}{\sqrt{N_{s}}}\sum_{j}e^{-i\bm{k}\cdot\bm{R}_{j}}c_{j\downarrow}c_{j\uparrow}. (S35)

Then, we have

d𝒌d𝒌=1Nsi,jei𝒌(𝑹i𝑹j)Tr[cicicjcjρ]=12Tr[FFρ]=12f,ρ2fΛ22.\displaystyle\langle d_{\bm{k}}^{\dagger}d_{\bm{k}}\rangle=\frac{1}{N_{s}}\sum_{i,j}e^{i\bm{k}\cdot(\bm{R}_{i}-\bm{R}_{j})}\mathrm{Tr}[c_{i\uparrow}^{\dagger}c_{i\downarrow}^{\dagger}c_{j\downarrow}c_{j\uparrow}\rho]=\frac{1}{2}\mathrm{Tr}[F^{\dagger}F\rho]=\frac{1}{2}\langle f,\rho_{2}f\rangle\leq\frac{\Lambda_{2}}{2}. (S36)

The bound on Λ2\Lambda_{2} was derived by Yang Yang (1962):

Λ2N(MN+2)M,\Lambda_{2}\leq\frac{N(M-N+2)}{M}, (S37)

where MM denotes the number of single-particle states of fermions. For our case, M=2NsM=2N_{s}, and thus we obtain the bound on the number of doublons at momentum 𝒌\bm{k}:

d𝒌d𝒌N(2NsN+2)4Ns=N2N(N2)4Ns,\langle d_{\bm{k}}^{\dagger}d_{\bm{k}}\rangle\leq\frac{N(2N_{s}-N+2)}{4N_{s}}=\frac{N}{2}-\frac{N(N-2)}{4N_{s}}, (S38)

which indicates that an NN-particle system of N/2(2)N/2(\geq 2) doublons cannot condense all doublons into a single momentum state. This bound (S38) holds for an arbitrary density matrix ρ\rho. The physical origin of this bound can be understood from the Pauli exclusion principle; fermions cannot be condensed without forming pairs and therefore the maximum eigenvalue of the one-particle (two-particle) reduced density matrix cannot be of the order of NN (N2N^{2}) Yang (1962).

For Yang’s η\eta-pairing state |ψN=(η+)N/2|0\ket{\psi_{N}}=(\eta^{+})^{N/2}\ket{0}, the pair correlation function can be calculated as cicicjcj=(N/2)(NsN/2)Ns(Ns1)ei𝑸(𝑹i𝑹j)\langle c_{i\uparrow}^{\dagger}c_{i\downarrow}^{\dagger}c_{j\downarrow}c_{j\uparrow}\rangle=\frac{(N/2)(N_{s}-N/2)}{N_{s}(N_{s}-1)}e^{i\bm{Q}\cdot(\bm{R}_{i}-\bm{R}_{j})} Yang (1989). Thus, we have

d𝑸d𝑸=\displaystyle\langle d_{\bm{Q}}^{\dagger}d_{\bm{Q}}\rangle= 1Nsjnjnj+1Nsijcicicjcjei𝑸(𝑹i𝑹j)\displaystyle\frac{1}{N_{s}}\sum_{j}\langle n_{j\uparrow}n_{j\downarrow}\rangle+\frac{1}{N_{s}}\sum_{i\neq j}\langle c_{i\uparrow}^{\dagger}c_{i\downarrow}^{\dagger}c_{j\downarrow}c_{j\uparrow}\rangle e^{i\bm{Q}\cdot(\bm{R}_{i}-\bm{R}_{j})}
=\displaystyle= N/2Ns+1Ns(N/2)(NsN/2)Ns(Ns1)×Ns(Ns1)\displaystyle\frac{N/2}{N_{s}}+\frac{1}{N_{s}}\frac{(N/2)(N_{s}-N/2)}{N_{s}(N_{s}-1)}\times N_{s}(N_{s}-1)
=\displaystyle= N(2NsN+2)4Ns.\displaystyle\frac{N(2N_{s}-N+2)}{4N_{s}}. (S39)

Thus, Yang’s η\eta-pairing state |ψN\ket{\psi_{N}} achieves the bound (S38) for 𝒌=𝑸\bm{k}=\bm{Q}, and describes the maximal condensation of doublons at momentum 𝑸\bm{Q}. The effect of the Pauli exclusion principle in the momentum distribution can clearly be seen by comparing the commutation relations of annihilation and creation operators of doublons with those of bosons; if b𝒌b_{\bm{k}} is an annihilation operator of a boson that satisfies the canonical commutation relation [b𝒌,b𝒌]=δ𝒌,𝒌[b_{\bm{k}},b_{\bm{k}^{\prime}}^{\dagger}]=\delta_{\bm{k},\bm{k}^{\prime}}, the momentum distribution of bosons for a state (b𝑸)N|0(b_{\bm{Q}}^{\dagger})^{N}\ket{0} satisfies 0|(b𝑸)Nb𝒌b𝒌(b𝑸)N|0=0|(b𝑸)Nb𝒌(b𝑸)Nb𝒌|0=0\bra{0}(b_{\bm{Q}})^{N}b_{\bm{k}}^{\dagger}b_{\bm{k}}(b_{\bm{Q}}^{\dagger})^{N}\ket{0}=\bra{0}(b_{\bm{Q}})^{N}b_{\bm{k}}^{\dagger}(b_{\bm{Q}}^{\dagger})^{N}b_{\bm{k}}\ket{0}=0 for 𝒌𝑸\bm{k}\neq\bm{Q}, and thus all bosons are condensed at momentum 𝑸\bm{Q}. However, for the annihilation operator of a doublon, the commutation relations read

[d𝒌,d𝒌]=[d𝒌,d𝒌]=0,\displaystyle[d_{\bm{k}},d_{\bm{k}^{\prime}}]=[d_{\bm{k}}^{\dagger},d_{\bm{k}^{\prime}}^{\dagger}]=0, (S40)
[d𝒌,d𝒌]=δ𝒌,𝒌1Nsj(nj+nj)ei(𝒌𝒌)𝑹j,\displaystyle[d_{\bm{k}},d_{\bm{k}^{\prime}}^{\dagger}]=\delta_{\bm{k},\bm{k}^{\prime}}-\frac{1}{N_{s}}\sum_{j}(n_{j\uparrow}+n_{j\downarrow})e^{-i(\bm{k}-\bm{k}^{\prime})\cdot\bm{R}_{j}}, (S41)

and do not satisfy the canonical commutation relations for bosons because of the Pauli exclusion principle. For this reason, doublons can be distributed at momentum 𝒌𝑸\bm{k}\neq\bm{Q} in the η\eta-pairing state |ψN=(d𝑸)N/2|0\ket{\psi_{N}}=(d_{\bm{Q}}^{\dagger})^{N/2}\ket{0}.

The momentum distribution of doublons for the η\eta-pairing state |ψ2Np;𝒌1,,𝒌Nex=(η+)Npc𝒌1c𝒌Nex|0\ket{\psi_{2N_{p}};\bm{k}_{1},\cdots,\bm{k}_{N_{\mathrm{ex}}}}=(\eta^{+})^{N_{p}}c_{\bm{k}_{1}\downarrow}^{\dagger}\cdots c_{\bm{k}_{N_{\mathrm{ex}}}\downarrow}^{\dagger}\ket{0} with excess spin-down particles can also be calculated. We note that

𝜼2|ψ2Np;𝒌1,,𝒌Nex=\displaystyle\bm{\eta}^{2}\ket{\psi_{2N_{p}};\bm{k}_{1},\cdots,\bm{k}_{N_{\mathrm{ex}}}}= NsNex2(NsNex2+1)|ψ2Np;𝒌1,,𝒌Nex,\displaystyle\frac{N_{s}-N_{\mathrm{ex}}}{2}\left(\frac{N_{s}-N_{\mathrm{ex}}}{2}+1\right)\ket{\psi_{2N_{p}};\bm{k}_{1},\cdots,\bm{k}_{N_{\mathrm{ex}}}}, (S42)
ηz|ψ2Np;𝒌1,,𝒌Nex=\displaystyle\eta^{z}\ket{\psi_{2N_{p}};\bm{k}_{1},\cdots,\bm{k}_{N_{\mathrm{ex}}}}= 2Np+NexNs2|ψ2Np;𝒌1,,𝒌Nex,\displaystyle\frac{2N_{p}+N_{\mathrm{ex}}-N_{s}}{2}\ket{\psi_{2N_{p}};\bm{k}_{1},\cdots,\bm{k}_{N_{\mathrm{ex}}}}, (S43)

where 𝜼2=(ηx)2+(ηy)2+(ηz)2\bm{\eta}^{2}=(\eta^{x})^{2}+(\eta^{y})^{2}+(\eta^{z})^{2}. Thus, using η+η=𝜼2(ηz)2+ηz\eta^{+}\eta^{-}=\bm{\eta}^{2}-(\eta^{z})^{2}+\eta^{z}, we obtain

d𝑸d𝑸=\displaystyle\langle d_{\bm{Q}}^{\dagger}d_{\bm{Q}}\rangle= 1Nsψ2Np;𝒌1,,𝒌Nex|η+η|ψ2Np;𝒌1,,𝒌Nexψ2Np;𝒌1,,𝒌Nex|ψ2Np;𝒌1,,𝒌Nex\displaystyle\frac{1}{N_{s}}\frac{\bra{\psi_{2N_{p}};\bm{k}_{1},\cdots,\bm{k}_{N_{\mathrm{ex}}}}\eta^{+}\eta^{-}\ket{\psi_{2N_{p}};\bm{k}_{1},\cdots,\bm{k}_{N_{\mathrm{ex}}}}}{\braket{\psi_{2N_{p}};\bm{k}_{1},\cdots,\bm{k}_{N_{\mathrm{ex}}}}{\psi_{2N_{p}};\bm{k}_{1},\cdots,\bm{k}_{N_{\mathrm{ex}}}}}
=\displaystyle= 1NsNsNex2(NsNex2+1)1Ns2Np+NexNs2(2Np+NexNs21)\displaystyle\frac{1}{N_{s}}\frac{N_{s}-N_{\mathrm{ex}}}{2}\left(\frac{N_{s}-N_{\mathrm{ex}}}{2}+1\right)-\frac{1}{N_{s}}\frac{2N_{p}+N_{\mathrm{ex}}-N_{s}}{2}\left(\frac{2N_{p}+N_{\mathrm{ex}}-N_{s}}{2}-1\right)
=\displaystyle= Np(NsNpNex+1)Ns,\displaystyle\frac{N_{p}(N_{s}-N_{p}-N_{\mathrm{ex}}+1)}{N_{s}}, (S44)

which is lower than that for |ψ2Np\ket{\psi_{2N_{p}}} due to the Pauli exclusion between doublons and excess spin-down particles.

In the main text, we have calculated the doublon momentum distribution R𝒌=d𝒌d𝒌/𝒌d𝒌d𝒌R_{\bm{k}}=\langle d_{\bm{k}}^{\dagger}d_{\bm{k}}\rangle/\sum_{\bm{k}}\langle d_{\bm{k}}^{\dagger}d_{\bm{k}}\rangle that is normalized by the total number of doublons. The bound of the normalized momentum distribution can be obtained under the assumption that a steady state ρSS\rho_{\mathrm{SS}} is a statistical mixture of the η\eta-pairing states |ψN\ket{\psi_{N}} and |ψ2Np;𝒌1,,𝒌Nex\ket{\psi_{2N_{p}};\bm{k}_{1},\cdots,\bm{k}_{N_{\mathrm{ex}}}} as

ρSS=Nex=0Np(Nex)ρ(NNex2,Nex),\rho_{\mathrm{SS}}=\sum_{N_{\mathrm{ex}}=0}^{N}p(N_{\mathrm{ex}})\rho\left(\frac{N-N_{\mathrm{ex}}}{2},N_{\mathrm{ex}}\right), (S45)

where

ρ(N/2,0)|ψNψN|ψN|ψN,\displaystyle\rho(N/2,0)\equiv\frac{\ket{\psi_{N}}\bra{\psi_{N}}}{\braket{\psi_{N}}{\psi_{N}}}, (S46)
ρ(Np,Nex)𝒌1,,𝒌Nexp(Np,𝒌1,,𝒌Nex)𝒌1,,𝒌Nexp(Np,𝒌1,,𝒌Nex)|ψ2Np;𝒌1,,𝒌Nexψ2Np;𝒌1,,𝒌Nex|ψ2Np;𝒌1,,𝒌Nex|ψ2Np;𝒌1,,𝒌Nex(forNex1),\displaystyle\rho(N_{p},N_{\mathrm{ex}})\equiv\sum_{\bm{k}_{1},\cdots,\bm{k}_{N_{\mathrm{ex}}}}\frac{p(N_{p},\bm{k}_{1},\cdots,\bm{k}_{N_{\mathrm{ex}}})}{\sum_{\bm{k}_{1},\cdots,\bm{k}_{N_{\mathrm{ex}}}}p(N_{p},\bm{k}_{1},\cdots,\bm{k}_{N_{\mathrm{ex}}})}\frac{\ket{\psi_{2N_{p}};\bm{k}_{1},\cdots,\bm{k}_{N_{\mathrm{ex}}}}\bra{\psi_{2N_{p}};\bm{k}_{1},\cdots,\bm{k}_{N_{\mathrm{ex}}}}}{\braket{\psi_{2N_{p}};\bm{k}_{1},\cdots,\bm{k}_{N_{\mathrm{ex}}}}{\psi_{2N_{p}};\bm{k}_{1},\cdots,\bm{k}_{N_{\mathrm{ex}}}}}\ \ (\mathrm{for}\ N_{\mathrm{ex}}\geq 1), (S47)

and the probability distribution satisfies p(Np,𝒌1,,𝒌Nex)0,p(Nex)0p(N_{p},\bm{k}_{1},\cdots,\bm{k}_{N_{\mathrm{ex}}})\geq 0,p(N_{\mathrm{ex}})\geq 0, and Nex=0Np(Nex)=1\sum_{N_{\mathrm{ex}}=0}^{N}p(N_{\mathrm{ex}})=1. Note that the dynamics under consideration conserves the total number of particles, and NexN_{\mathrm{ex}} should be even if the total number of particles is even. Then, from Eqs. (S39) and (S44), the doublon momentum distribution for the steady state is given by

d𝑸d𝑸=\displaystyle\langle d_{\bm{Q}}^{\dagger}d_{\bm{Q}}\rangle= Tr[d𝑸d𝑸ρSS]\displaystyle\mathrm{Tr}[d_{\bm{Q}}^{\dagger}d_{\bm{Q}}\rho_{\mathrm{SS}}]
=\displaystyle= Nex=0Np(Nex)NNex2Ns(NsN2Nex2+1)\displaystyle\sum_{N_{\mathrm{ex}}=0}^{N}p(N_{\mathrm{ex}})\frac{N-N_{\mathrm{ex}}}{2N_{s}}\left(N_{s}-\frac{N}{2}-\frac{N_{\mathrm{ex}}}{2}+1\right)
=\displaystyle= NsN/2+1NsNex=0Np(Nex)NNex21NsNex=0Np(Nex)NNex2Nex2.\displaystyle\frac{N_{s}-N/2+1}{N_{s}}\sum_{N_{\mathrm{ex}}=0}^{N}p(N_{\mathrm{ex}})\frac{N-N_{\mathrm{ex}}}{2}-\frac{1}{N_{s}}\sum_{N_{\mathrm{ex}}=0}^{N}p(N_{\mathrm{ex}})\frac{N-N_{\mathrm{ex}}}{2}\frac{N_{\mathrm{ex}}}{2}. (S48)

The total number of doublons in the steady state is

𝒌d𝒌d𝒌=jnjnj=jTr[njnjρSS]=Nex=0Np(Nex)NNex2.\displaystyle\sum_{\bm{k}}\langle d_{\bm{k}}^{\dagger}d_{\bm{k}}\rangle=\sum_{j}\langle n_{j\uparrow}n_{j\downarrow}\rangle=\sum_{j}\mathrm{Tr}[n_{j\uparrow}n_{j\downarrow}\rho_{\mathrm{SS}}]=\sum_{N_{\mathrm{ex}}=0}^{N}p(N_{\mathrm{ex}})\frac{N-N_{\mathrm{ex}}}{2}. (S49)

Thus, we obtain

R𝑸=\displaystyle R_{\bm{Q}}= d𝑸d𝑸𝒌d𝒌d𝒌\displaystyle\frac{\langle d_{\bm{Q}}^{\dagger}d_{\bm{Q}}\rangle}{\sum_{\bm{k}}\langle d_{\bm{k}}^{\dagger}d_{\bm{k}}\rangle}
=\displaystyle= NsN/2+1Ns(1/Ns)Nex=0Np(Nex)(NNex)/2Nex/2Nex=0Np(Nex)(NNex)/2\displaystyle\frac{N_{s}-N/2+1}{N_{s}}-\frac{(1/N_{s})\sum_{N_{\mathrm{ex}}=0}^{N}p(N_{\mathrm{ex}})(N-N_{\mathrm{ex}})/2\cdot N_{\mathrm{ex}}/2}{\sum_{N_{\mathrm{ex}}=0}^{N}p(N_{\mathrm{ex}})(N-N_{\mathrm{ex}})/2}
\displaystyle\leq NsN/2+1Ns\displaystyle\frac{N_{s}-N/2+1}{N_{s}}
=\displaystyle= 1N2Ns+1Ns,\displaystyle 1-\frac{N}{2N_{s}}+\frac{1}{N_{s}}, (S50)

which completes the proof of the bound (5) in the main text.

Appendix E Details of the experimental implementation

Here, we describe an experimental situation that leads to the quantum master equation (2) in the main text. We consider fermionic atoms with a lambda-type level structure shown in Fig. 1(b) in the main text. The two low-lying energy levels are denoted by |g,σ(σ=,)\ket{g,\sigma}\ (\sigma=\uparrow,\downarrow), and the highest energy level is denoted by |e\ket{e}. We assume that the |g,\ket{g,\uparrow} state is coupled to the |e\ket{e} state by a coherent laser field, which is described by a classical electromagnetic field, and that the |e\ket{e} state can decay to the |g,σ\ket{g,\sigma} state via spontaneous emission of a photon. Then, the Hamiltonian HtotH_{\mathrm{tot}} of the total system is given by

Htot=\displaystyle H_{\mathrm{tot}}= HA+HI+Hph,\displaystyle H_{\mathrm{A}}+H_{\mathrm{I}}+H_{\mathrm{ph}}, (S51)
HA=\displaystyle H_{\mathrm{A}}= d3𝒙σ=,ψσ(𝒙)(22m+V(𝒙))ψσ(𝒙)+g0d3𝒙ψ(𝒙)ψ(𝒙)ψ(𝒙)ψ(𝒙)\displaystyle\int d^{3}\bm{x}\sum_{\sigma=\uparrow,\downarrow}\psi_{\sigma}^{\dagger}(\bm{x})\left(\frac{-\nabla^{2}}{2m}+V(\bm{x})\right)\psi_{\sigma}(\bm{x})+g_{0}\int d^{3}\bm{x}\psi_{\uparrow}^{\dagger}(\bm{x})\psi_{\downarrow}^{\dagger}(\bm{x})\psi_{\downarrow}(\bm{x})\psi_{\uparrow}(\bm{x})
+δ0d3𝒙(ψ(𝒙)ψ(𝒙)ψ(𝒙)ψ(𝒙))+(Δ+δ0)d3𝒙ψe(𝒙)ψe(𝒙)+d3𝒙(Ω(𝒙)ψe(𝒙)ψ(𝒙)+H.c.),\displaystyle+\delta_{0}\int d^{3}\bm{x}(\psi_{\uparrow}^{\dagger}(\bm{x})\psi_{\uparrow}(\bm{x})-\psi_{\downarrow}^{\dagger}(\bm{x})\psi_{\downarrow}(\bm{x}))+(\Delta+\delta_{0})\int d^{3}\bm{x}\psi_{e}^{\dagger}(\bm{x})\psi_{e}(\bm{x})+\int d^{3}\bm{x}(\Omega(\bm{x})\psi_{e}^{\dagger}(\bm{x})\psi_{\uparrow}(\bm{x})+\mathrm{H.c.}), (S52)
HI=\displaystyle H_{\mathrm{I}}= σ=,d3𝒙(𝒅σ𝑬(+)(𝒙)ψe(𝒙)ψσ(𝒙)+H.c.),\displaystyle-\sum_{\sigma=\uparrow,\downarrow}\int d^{3}\bm{x}(\bm{d}_{\sigma}\cdot\bm{E}^{(+)}(\bm{x})\psi_{e}^{\dagger}(\bm{x})\psi_{\sigma}(\bm{x})+\mathrm{H.c.}), (S53)
Hph=\displaystyle H_{\mathrm{ph}}= 𝒌,λω𝒌a𝒌,λa𝒌,λ,\displaystyle\sum_{\bm{k},\lambda}\omega_{\bm{k}}a_{\bm{k},\lambda}^{\dagger}a_{\bm{k},\lambda}, (S54)

where ψσ(𝒙)\psi_{\sigma}(\bm{x}) (ψe(𝒙)\psi_{e}(\bm{x})) is the field operator of atoms in the |g,σ\ket{g,\sigma} (|e\ket{e}) state, mm is the mass of an atom, V(𝒙)V(\bm{x}) is an optical lattice potential, 2δ02\delta_{0} is the energy difference between the |g,\ket{g,\uparrow} and |g,\ket{g,\downarrow} states, g0g_{0} is the contact interaction strength, Δ\Delta is the detuning of a laser that couples the |g,\ket{g,\uparrow} state to the |e\ket{e} state with the Rabi frequency Ω(𝒙)\Omega(\bm{x}), 𝒅σ\bm{d}_{\sigma} is the electric-dipole matrix element between the states |g,σ\ket{g,\sigma} and |e\ket{e}, a𝒌,λa_{\bm{k},\lambda} is the annihilation operator of a photon field with momentum 𝒌\bm{k}, polarization λ\lambda and frequency ω𝒌\omega_{\bm{k}}, and

𝑬(+)(𝒙)=𝒌,λε𝒌𝒆𝒌,λei𝒌𝒙a𝒌,λ\bm{E}^{(+)}(\bm{x})=\sum_{\bm{k},\lambda}\varepsilon_{\bm{k}}\bm{e}_{\bm{k},\lambda}e^{i\bm{k}\cdot\bm{x}}a_{\bm{k},\lambda} (S55)

is the positive-frequency part of the electric-field operator with the polarization vector 𝒆𝒌,λ\bm{e}_{\bm{k},\lambda}. Here, the rotating-wave approximation has been applied to the light-matter couplings. The kinetic energy of the |e\ket{e} state can be neglected if the decay of this state occurs sufficiently faster than the time scale of the atomic gas. We assume that the relevant optical frequency for the transition between the |g,σ\ket{g,\sigma} and |e\ket{e} states is much higher than the energy scale of the atomic gas, and that the quantized radiation modes are in the vacuum state. Then, we can trace out the photon degrees of freedom using the Born-Markov approximation Breuer and Petruccione (2007), which leads to the quantum master equation for the density matrix ρA\rho_{\mathrm{A}} of the atomic system as Lehmberg (1970); Pichler et al. (2010); Sarkar et al. (2014)

dρAdτ=\displaystyle\frac{d\rho_{\mathrm{A}}}{d\tau}= i[HA+Hdip,ρA]+𝒟ρA,\displaystyle-i[H_{A}+H_{\mathrm{dip}},\rho_{\mathrm{A}}]+\mathcal{D}\rho_{\mathrm{A}}, (S56)

where

Hdip=\displaystyle H_{\mathrm{dip}}= σ=,Γσ2d3𝒙d3𝒚Gσ(kσ(𝒙𝒚))ψe(𝒙)ψσ(𝒙)ψσ(𝒚)ψe(𝒚)\displaystyle\sum_{\sigma=\uparrow,\downarrow}\frac{\Gamma_{\sigma}}{2}\int d^{3}\bm{x}\int d^{3}\bm{y}G_{\sigma}(k_{\sigma}(\bm{x}-\bm{y}))\psi_{e}^{\dagger}(\bm{x})\psi_{\sigma}(\bm{x})\psi_{\sigma}^{\dagger}(\bm{y})\psi_{e}(\bm{y}) (S57)

is an effective dipole-dipole interaction,

𝒟ρA=\displaystyle\mathcal{D}\rho_{\mathrm{A}}= σ=,Γσ2d3𝒙d3𝒚Fσ(kσ(𝒙𝒚))(2ψσ(𝒙)ψe(𝒙)ρAψe(𝒚)ψσ(𝒚)\displaystyle\sum_{\sigma=\uparrow,\downarrow}\frac{\Gamma_{\sigma}}{2}\int d^{3}\bm{x}\int d^{3}\bm{y}F_{\sigma}(k_{\sigma}(\bm{x}-\bm{y}))\Bigl{(}2\psi_{\sigma}^{\dagger}(\bm{x})\psi_{e}(\bm{x})\rho_{\mathrm{A}}\psi_{e}^{\dagger}(\bm{y})\psi_{\sigma}(\bm{y})
ψe(𝒙)ψσ(𝒙)ψσ(𝒚)ψe(𝒚)ρAρAψe(𝒙)ψσ(𝒙)ψσ(𝒚)ψe(𝒚))\displaystyle-\psi_{e}^{\dagger}(\bm{x})\psi_{\sigma}(\bm{x})\psi_{\sigma}^{\dagger}(\bm{y})\psi_{e}(\bm{y})\rho_{\mathrm{A}}-\rho_{\mathrm{A}}\psi_{e}^{\dagger}(\bm{x})\psi_{\sigma}(\bm{x})\psi_{\sigma}^{\dagger}(\bm{y})\psi_{e}(\bm{y})\Bigr{)} (S58)

describes spontaneous-emission processes, Γσ\Gamma_{\sigma} is the rate of spontaneous decay from the |e\ket{e} state to the |g,σ\ket{g,\sigma} state, and ωσ=ckσ\omega_{\sigma}=ck_{\sigma} is the energy difference between those states. Here, the functions G(𝝃)G(\bm{\xi}) and F(𝝃)F(\bm{\xi}) are given from a correlation function of the vacuum radiation modes

0𝑑s(𝒅σ𝑬(+)(𝒙,s))(𝒅σ𝑬(+)(𝒚,0))eiωσs=Γσ2(Fσ(kσ(𝒙𝒚))+iGσ(kσ(𝒙𝒚)))\displaystyle\int_{0}^{\infty}ds\langle(\bm{d}_{\sigma}\cdot\bm{E}^{(+)}(\bm{x},s))(\bm{d}_{\sigma}\cdot\bm{E}^{(+)}(\bm{y},0))^{\dagger}\rangle e^{i\omega_{\sigma}s}=\frac{\Gamma_{\sigma}}{2}\Bigl{(}F_{\sigma}(k_{\sigma}(\bm{x}-\bm{y}))+iG_{\sigma}(k_{\sigma}(\bm{x}-\bm{y}))\Bigr{)} (S59)

as Lehmberg (1970); Pichler et al. (2010); Sarkar et al. (2014)

Fσ(𝝃)=\displaystyle F_{\sigma}(\bm{\xi})= 38π𝑑Ωk(1|𝒌^𝒅^σ|2)ei𝒌^𝝃,\displaystyle\frac{3}{8\pi}\int d\Omega_{k}(1-|\hat{\bm{k}}\cdot\hat{\bm{d}}_{\sigma}|^{2})e^{-i\hat{\bm{k}}\cdot\bm{\xi}}, (S60)
Gσ(𝝃)=\displaystyle G_{\sigma}(\bm{\xi})= 1ξ3𝒫dζ2πζ3ζξFσ(ζ𝝃/ξ),\displaystyle-\frac{1}{\xi^{3}}\mathcal{P}\int\frac{d\zeta}{2\pi}\frac{\zeta^{3}}{\zeta-\xi}F_{\sigma}(\zeta\bm{\xi}/\xi), (S61)

where 𝒌^𝒌/|𝒌|,𝒅^σ𝒅σ/|𝒅σ|\hat{\bm{k}}\equiv\bm{k}/|\bm{k}|,\hat{\bm{d}}_{\sigma}\equiv\bm{d}_{\sigma}/|\bm{d}_{\sigma}|, 𝑑Ωk\int d\Omega_{k} is the integral over the direction of 𝒌^\hat{\bm{k}}, and 𝒫\mathcal{P} denotes the principal value integral. In deriving the above quantum master equation, we have assumed that photons emitted in the |e|g,\ket{e}\to\ket{g,\uparrow} process and those emitted in the |e|g,\ket{e}\to\ket{g,\downarrow} process can be distinguished from their frequencies or polarizations.

If the detuning Δ\Delta is much larger than the other energy scales of the atomic gas, i.e., |Δ|Γσ,ER,|g0|,|δ0|,|Ω(𝒙)||\Delta|\gg\Gamma_{\sigma},E_{R},|g_{0}|,|\delta_{0}|,|\Omega(\bm{x})| (ERE_{R} is the recoil energy), the excited state |e\ket{e} can be adiabatically eliminated Pichler et al. (2010); Sarkar et al. (2014). Using the second-order perturbation theory Reiter and Sørensen (2012), we obtain

dρdτ=\displaystyle\frac{d\rho}{d\tau}= i[Hg+Hg,dip,ρ]+σ=,d3𝒙d3𝒚ΓσΩ(𝒙)Ω(𝒚)2Δ2Fσ(kσ(𝒙𝒚))(2ψσ(𝒙)ψ(𝒙)ρψ(𝒚)ψσ(𝒚)\displaystyle-i[H_{g}+H_{g,\mathrm{dip}},\rho]+\sum_{\sigma=\uparrow,\downarrow}\int d^{3}\bm{x}\int d^{3}\bm{y}\frac{\Gamma_{\sigma}\Omega(\bm{x})\Omega^{*}(\bm{y})}{2\Delta^{2}}F_{\sigma}(k_{\sigma}(\bm{x}-\bm{y}))\Bigl{(}2\psi_{\sigma}^{\dagger}(\bm{x})\psi_{\uparrow}(\bm{x})\rho\psi_{\uparrow}^{\dagger}(\bm{y})\psi_{\sigma}(\bm{y})
ψ(𝒙)ψσ(𝒙)ψσ(𝒚)ψ(𝒚)ρρψ(𝒙)ψσ(𝒙)ψσ(𝒚)ψ(𝒚)),\displaystyle-\psi_{\uparrow}^{\dagger}(\bm{x})\psi_{\sigma}(\bm{x})\psi_{\sigma}^{\dagger}(\bm{y})\psi_{\uparrow}(\bm{y})\rho-\rho\psi_{\uparrow}^{\dagger}(\bm{x})\psi_{\sigma}(\bm{x})\psi_{\sigma}^{\dagger}(\bm{y})\psi_{\uparrow}(\bm{y})\Bigr{)}, (S62)

where ρPgρAPg\rho\equiv P_{g}\rho_{\mathrm{A}}P_{g} is the density matrix of the atomic gas with PgP_{g} being a projection operator onto the subspace that does not contain atoms in the |e\ket{e} state,

Hg=\displaystyle H_{g}= d3𝒙σ=,ψσ(𝒙)(22m+V(𝒙))ψσ(𝒙)+g0d3𝒙ψ(𝒙)ψ(𝒙)ψ(𝒙)ψ(𝒙)\displaystyle\int d^{3}\bm{x}\sum_{\sigma=\uparrow,\downarrow}\psi_{\sigma}^{\dagger}(\bm{x})\left(\frac{-\nabla^{2}}{2m}+V(\bm{x})\right)\psi_{\sigma}(\bm{x})+g_{0}\int d^{3}\bm{x}\psi_{\uparrow}^{\dagger}(\bm{x})\psi_{\downarrow}^{\dagger}(\bm{x})\psi_{\downarrow}(\bm{x})\psi_{\uparrow}(\bm{x})
+δ0d3𝒙(ψ(𝒙)ψ(𝒙)ψ(𝒙)ψ(𝒙))d3𝒙|Ω(𝒙)|2Δψ(𝒙)ψ(𝒙),\displaystyle+\delta_{0}\int d^{3}\bm{x}(\psi_{\uparrow}^{\dagger}(\bm{x})\psi_{\uparrow}(\bm{x})-\psi_{\downarrow}^{\dagger}(\bm{x})\psi_{\downarrow}(\bm{x}))-\int d^{3}\bm{x}\frac{|\Omega(\bm{x})|^{2}}{\Delta}\psi_{\uparrow}^{\dagger}(\bm{x})\psi_{\uparrow}(\bm{x}), (S63)

and

Hg,dip=σ=,d3𝒙d3𝒚ΓσΩ(𝒙)Ω(𝒚)2Δ2Gσ(kσ(𝒙𝒚))ψ(𝒙)ψσ(𝒙)ψσ(𝒚)ψ(𝒚).\displaystyle H_{g,\mathrm{dip}}=\sum_{\sigma=\uparrow,\downarrow}\int d^{3}\bm{x}\int d^{3}\bm{y}\frac{\Gamma_{\sigma}\Omega(\bm{x})\Omega^{*}(\bm{y})}{2\Delta^{2}}G_{\sigma}(k_{\sigma}(\bm{x}-\bm{y}))\psi_{\uparrow}^{\dagger}(\bm{x})\psi_{\sigma}(\bm{x})\psi_{\sigma}^{\dagger}(\bm{y})\psi_{\uparrow}(\bm{y}). (S64)

Then, substituting the expansion ψσ(𝒙)=jnwn(𝒙𝒙j)cjσn\psi_{\sigma}(\bm{x})=\sum_{j}\sum_{n}w_{n}(\bm{x}-\bm{x}_{j})c_{j\sigma n} with the Wannier function wn(𝒙)w_{n}(\bm{x}) of the nnth band of the optical lattice into Eq. (S62), we obtain

dρdτ=\displaystyle\frac{d\rho}{d\tau}= i[Hmulti,ρ]+σ=,j1,,j4n1,,n4γσ;j1,j2,j3,j4(n1,n2,n3,n4)2(2cj1σn1cj2n2ρcj3n3cj4σn4\displaystyle-i[H_{\mathrm{multi}},\rho]+\sum_{\sigma=\uparrow,\downarrow}\sum_{j_{1},\cdots,j_{4}}\sum_{n_{1},\cdots,n_{4}}\frac{\gamma_{\sigma;j_{1},j_{2},j_{3},j_{4}}^{(n_{1},n_{2},n_{3},n_{4})}}{2}(2c_{j_{1}\sigma n_{1}}^{\dagger}c_{j_{2}\uparrow n_{2}}\rho c_{j_{3}\uparrow n_{3}}^{\dagger}c_{j_{4}\sigma n_{4}}
cj1n1cj2σn2cj3σn3cj4n4ρρcj1n1cj2σn2cj3σn3cj4n4),\displaystyle-c_{j_{1}\uparrow n_{1}}^{\dagger}c_{j_{2}\sigma n_{2}}c_{j_{3}\sigma n_{3}}^{\dagger}c_{j_{4}\uparrow n_{4}}\rho-\rho c_{j_{1}\uparrow n_{1}}^{\dagger}c_{j_{2}\sigma n_{2}}c_{j_{3}\sigma n_{3}}^{\dagger}c_{j_{4}\uparrow n_{4}}), (S65)

where

Hmulti=\displaystyle H_{\mathrm{multi}}= i,jσ=,ntij(n)ciσncjσn+j1,j2,j3,j4σn1,n2,n3,n4Uσ;j1,j2,j3,j4(n1,n2,n3,n4)cj1n1cj2σn2cj3σn3cj4n4\displaystyle-\sum_{i,j}\sum_{\sigma=\uparrow,\downarrow}\sum_{n}t_{ij}^{(n)}c_{i\sigma n}^{\dagger}c_{j\sigma n}+\sum_{j_{1},j_{2},j_{3},j_{4}}\sum_{\sigma}\sum_{n_{1},n_{2},n_{3},n_{4}}U_{\sigma;j_{1},j_{2},j_{3},j_{4}}^{(n_{1},n_{2},n_{3},n_{4})}c_{j_{1}\uparrow n_{1}}^{\dagger}c_{j_{2}\sigma n_{2}}^{\dagger}c_{j_{3}\sigma n_{3}}c_{j_{4}\uparrow n_{4}}
+jσnδσ(n)cjσncjσn\displaystyle+\sum_{j}\sum_{\sigma}\sum_{n}\delta^{(n)}_{\sigma}c_{j\sigma n}^{\dagger}c_{j\sigma n} (S66)

is the multiband tight-binding Hamiltonian, and

γσ;j1,j2,j3,j4(n1,n2,n3,n4)=d3𝒙d3𝒚ΓσΩ(𝒙)Ω(𝒚)Δ2Fσ(kσ(𝒙𝒚))wn1(𝒙𝒙j1)wn2(𝒙𝒙j2)wn3(𝒚𝒙j3)wn4(𝒚𝒙j4).\displaystyle\gamma_{\sigma;j_{1},j_{2},j_{3},j_{4}}^{(n_{1},n_{2},n_{3},n_{4})}=\int d^{3}\bm{x}\int d^{3}\bm{y}\frac{\Gamma_{\sigma}\Omega(\bm{x})\Omega^{*}(\bm{y})}{\Delta^{2}}F_{\sigma}(k_{\sigma}(\bm{x}-\bm{y}))w_{n_{1}}^{*}(\bm{x}-\bm{x}_{j_{1}})w_{n_{2}}(\bm{x}-\bm{x}_{j_{2}})w^{*}_{n_{3}}(\bm{y}-\bm{x}_{j_{3}})w_{n_{4}}(\bm{y}-\bm{x}_{j_{4}}). (S67)

When the optical lattice potential is sufficiently deep, the nearest-neighbor hopping and the on-site interaction are dominant and the Hamiltonian HmultiH_{\mathrm{multi}} is well approximated by the single-band Hubbard model HH [Eq. (1) in the main text] Esslinger (2010). Similarly, the spontaneous-emission processes can be neglected except for on-site ones and we may set j1=j2=j3=j4j_{1}=j_{2}=j_{3}=j_{4} in Eq. (S67). In addition, it is shown from Eq. (S67) that the inter-band processes are suppressed by a factor of η2+ησ2\eta_{\uparrow}^{2}+\eta_{\sigma}^{2}, where ησkσa\eta_{\sigma}\equiv k_{\sigma}a is the Lamb-Dicke parameter with aa being the localization length of the Wannier function Pichler et al. (2010); Sarkar et al. (2014). Hence, for a sufficiently deep optical lattice with small η\eta_{\uparrow} and η\eta_{\downarrow}, we obtain

dρdτ=\displaystyle\frac{d\rho}{d\tau}= i[H,ρ]+j(LjρLj12{LjLj,ρ})+j(Lj(d)ρ(Lj(d))12{(Lj(d))Lj(d),ρ}),\displaystyle-i[H,\rho]+\sum_{j}(L_{j}\rho L_{j}^{\dagger}-\frac{1}{2}\{L_{j}^{\dagger}L_{j},\rho\})+\sum_{j}(L_{j}^{(d)}\rho(L_{j}^{(d)})^{\dagger}-\frac{1}{2}\{(L_{j}^{(d)})^{\dagger}L_{j}^{(d)},\rho\}), (S68)

where Lj=γcjcj,γ=γ;j,j,j,j(0,0,0,0),Lj(d)=γdnj,γd=γ;j,j,j,j(0,0,0,0)L_{j}=\sqrt{\gamma}c_{j\downarrow}^{\dagger}c_{j\uparrow},\gamma=\gamma_{\downarrow;j,j,j,j}^{(0,0,0,0)},L_{j}^{(d)}=\sqrt{\gamma_{d}}n_{j\uparrow},\gamma_{d}=\gamma_{\uparrow;j,j,j,j}^{(0,0,0,0)}, and the lowest-band index is denoted by n=0n=0. The Lindblad operator LjL_{j} describes an effective decay process from the |g,\ket{g,\uparrow} state to the |g,\ket{g,\downarrow} state, and Lj(d)L_{j}^{(d)} leads to dephasing of spin-up atoms due to photon scattering. Note that the ratio between the effective decay rate γ\gamma and the tunneling amplitude tt reads

γt=ΓtΩ2Δ2,\frac{\gamma}{t}=\frac{\Gamma_{\downarrow}}{t}\frac{\Omega^{2}}{\Delta^{2}}, (S69)

where Ω2d3𝒙d3𝒚3Ω(𝒙)Ω(𝒚)F(k(𝒙𝒚))|w0(𝒙𝒙j)|2|w0(𝒚𝒙j)|2\Omega^{2}\equiv\int d^{3}\bm{x}\int d^{3}\bm{y}^{3}\Omega(\bm{x})\Omega^{*}(\bm{y})F(k_{\downarrow}(\bm{x}-\bm{y}))|w_{0}(\bm{x}-\bm{x}_{j})|^{2}|w_{0}(\bm{y}-\bm{x}_{j})|^{2}, and this ratio is not necessarily small if Γ/t\Gamma_{\downarrow}/t is large. Thus, if the condition γdγ,t\gamma_{d}\ll\gamma,t is satisfied, the model in the main text and hence the η\eta-pairing states can be realized.

The condition γdγ\gamma_{d}\ll\gamma requires ΓΓ\Gamma_{\uparrow}\ll\Gamma_{\downarrow}. To achieve this, we propose to use alkaline-earth(-like) fermionic atoms such as 171Yb, 173Yb, and 87Sr Sandner et al. (2011). These atoms have the electronic ground state S01{}^{1}S_{0} and a metastable excited state P03{}^{3}P_{0}. First, ground-state atoms in two particular hyperfine states |S01,σ(σ=,)\ket{{}^{1}S_{0},\sigma}\ (\sigma=\uparrow,\downarrow) are prepared in some initial state in an optical lattice. Subsequently, atoms in the |S01,\ket{{}^{1}S_{0},\downarrow} state are excited to the |P03,\ket{{}^{3}P_{0},\uparrow} state by using a π\pi-pulse laser. When an optical lattice is created by a laser at a magic wavelength, the lattice potentials for the S01{}^{1}S_{0} and P03{}^{3}P_{0} states are the same and thus those atoms are described by the Hubbard model with the following identifications: |g,=|S01,\ket{g,\uparrow}=\ket{{}^{1}S_{0},\uparrow} and |g,=|P03,\ket{g,\downarrow}=\ket{{}^{3}P_{0},\uparrow} Gorshkov et al. (2010). Since the lifetime of the P03{}^{3}P_{0} state is sufficiently longer than the typical experimental time scale, spontaneous decay from the P03{}^{3}P_{0} state to the S01{}^{1}S_{0} state is negligible. Then, the effective decay process described above is induced by coupling the |P03,\ket{{}^{3}P_{0},\uparrow} state with the |e=|P11,\ket{e}=\ket{{}^{1}P_{1},\uparrow} state that can quickly decay to the |S01,\ket{{}^{1}S_{0},\uparrow} state Sandner et al. (2011). The nuclear-spin-changing decay processes can be suppressed at a high magnetic field Sandner et al. (2011). Since the magnetic-dipole-allowed transition from the |P03,\ket{{}^{3}P_{0},\uparrow} state to the |P11,\ket{{}^{1}P_{1},\uparrow} state is forbidden for the electric dipole coupling, the spontaneous decay between these states is negligible, and thus ΓΓ\Gamma_{\uparrow}\ll\Gamma_{\downarrow} is satisfied.

Appendix F Dependence of pair correlations on model parameters

In Fig. S1, we show the dependence of the pair correlation function on the model parameters UU and γ\gamma. Here, we take an initial state ρ(0)=|ψ0ψ0|\rho(0)=\ket{\psi_{0}}\bra{\psi_{0}} of an 8-site system, where |ψ0=j=0,2,4,6cjcj|0\ket{\psi_{0}}=\prod_{j=0,2,4,6}c_{j\uparrow}^{\dagger}c_{j\downarrow}^{\dagger}\ket{0}, as in Fig. 2(a) in the main text. The pair correlation C(1,0;τmax)C(1,0;\tau_{\mathrm{max}}) (C(4,0;τmax)C(4,0;\tau_{\mathrm{max}})) between the zeroth site and the first (fourth) site is evaluated at τmax/τh=40\tau_{\mathrm{max}}/\tau_{h}=40. As shown in Fig. S1, the parameter dependence of the pair correlation is weak. The decrease in pair correlations for large UU and small γ\gamma in Fig. S1(a) is attributed to the fact that the system does not reach a steady state at τmax/τh=40\tau_{\mathrm{max}}/\tau_{h}=40 due to the slow relaxation in this parameter regime where |Im[J]|0.01|\mathrm{Im}[J]|\simeq 0.01. This result indicates that the η\eta-pairing state can be realized over a wide range of parameters of the model, if the initial state is appropriately chosen.

Refer to caption
Figure S1: Dependence of the pair correlations on the interaction strength UU and the spontaneous-emission rate γ\gamma. The initial state is the doublon-density-wave state of an 8-site system as in Fig. 2(a) in the main text. The unit of energy is set to t=1t=1. (a) Pair correlation |Re[C(1,0,τmax)]||\mathrm{Re}[C(1,0,\tau_{\mathrm{max}})]| between sites j=1j=1 and j=0j=0 at time τmax/τh=40\tau_{\mathrm{max}}/\tau_{h}=40. (b) Pair correlation |Re[C(4,0,τmax)]||\mathrm{Re}[C(4,0,\tau_{\mathrm{max}})]| between sites j=4j=4 and j=0j=0 at time τmax/τh=40\tau_{\mathrm{max}}/\tau_{h}=40.

Appendix G Effect of dephasing

As described in the experimental implementation of the model, the second-order process involving spontaneous decay from the |e\ket{e} state to the |g,\ket{g,\uparrow} state is described by an additional Lindblad operator Lj(d)L_{j}^{(d)} which leads to dephasing of spin-up particles. In Fig. S2, we show the effect of dephasing on η\eta-pairing formation by numerically solving the quantum master equation (S68) with the quantum-trajectory method. The initial state and the model parameters except for γd\gamma_{d} are the same as those in Figs. 2 (a) and (b). Since the η\eta-pairing states are not steady states in the presence of nonzero γd\gamma_{d}, the pair correlations decrease in a time scale which is roughly given by 1/γd1/\gamma_{d} [see Figs. S2 (a), (e), and (i)]. The imaginary part of each pair correlation function only shows a transient increase and it is negligible after a long time for all parameter regimes including the case of γd=0\gamma_{d}=0 [see Figs. S2 (b), (f), and (j)]. While the density of doublons stays nonvanishing as shown in Fig. 2(b), it slowly decreases in the presence of dephasing [see Figs. S2 (c), (g), and (k)], implying that the steady state under nonzero γd\gamma_{d} is a fully polarized state that does not contain spin-up particles. This behavior is consistent with the fact that the quantity C=η+ηC=\langle\eta^{+}\eta^{-}\rangle [see Eq. (3) in the main text] is not conserved in the presence of nonzero γd\gamma_{d}, as shown in Figs. S2 (d), (h), and (l). Thus, if the condition γdγ,t\gamma_{d}\ll\gamma,t is satisfied, the formation of the η\eta-pairing state can be observed during its lifetime determined by 1/γd1/\gamma_{d}.

Refer to caption
Figure S2: Dynamics of η\eta-pairing formation in the presence of dephasing. (a) (e) (i) Real part of the pair correlation functions C(j,0,τ)C(j,0,\tau). (b) (f) (j) Imaginary part of the pair correlation functions. (c) (g) (k) Density of doublons. (d) (h) (l) η+η/L\langle\eta^{+}\eta^{-}\rangle/L defined in Eq. (3) in the main text. The initial state is the same as in Figs. 2 (a) and (b). The model parameters are set to t=1,U=10,δ=0t=1,U=10,\delta=0, and γ=10\gamma=10, and the rate of dephasing is γd=0.1\gamma_{d}=0.1 for (a)-(d), γd=0.5\gamma_{d}=0.5 for (e)-(h), and γd=1\gamma_{d}=1 for (i)-(l). The unit of time is τh=1/t\tau_{h}=1/t.

In Fig. S3, we also show the momentum distribution of doublons in the presence of dephasing. Although the dephasing suppresses the peaks at k=±πk=\pm\pi, one can still observe the signature of η\eta-pairing formation unless the rate of dephasing is large.

Refer to caption
Figure S3: Dynamics of the normalized momentum distribution RkR_{k} of doublons in the presence of dephasing. The initial state and the model parameters t,U,δ,γt,U,\delta,\gamma are the same as in Fig. 3(a), and the rate of dephasing is (a) γd=0.1\gamma_{d}=0.1, (b) γd=0.5\gamma_{d}=0.5, and (c) γd=1\gamma_{d}=1.