Photoinduced phase switching from Mott insulator to metallic state in the quarter-filled Peierls-Hubbard model

Can Shao [email protected] Department of Applied Physics &\&& MIIT Key Laboratory of Semiconductor Microstructure and Quantum Sensing, Nan**g University of Science and Technology, Nan**g 210094, China    Takami Tohyama [email protected] Department of Applied Physics, Tokyo University of Science, Tokyo 125-8585, Japan    Hantao Lu [email protected] School of Physical Science and Technology, Lanzhou University, Lanzhou 730000, China Lanzhou Center for Theoretical Physics &\&& Key Laboratory of Theoretical Physics of Gansu Province, Lanzhou University, Lanzhou 730000, China
(June 3, 2024)
Abstract

Utilizing the exact diagonalization method, we investigate the one-dimensional Peierls-Hubbard model at quarter filling, where it manifests as an antiferromagnetic Mott insulator in units of dimers. By increasing the on-site Coulomb repulsion U𝑈Uitalic_U, we observe a significant suppression of the Drude peak, based on a nonequilibrium linear response theory capable of capturing the zero-frequency (Drude) weight of the optical conductivity under periodic boundary conditions. However, after the ultrafast photoirradiation of this model with large U𝑈Uitalic_U, we detect a distinct enhancement of the Drude peak, signifying the onset of a photoinduced insulator-metal transition. Comparing these dynamics with the half-filled Hubbard model and a noninteracting spinless half-filled Su-Schrieffer-Heeger model (corresponding to the quarter-filled Peierls-Hubbard model with infinite U𝑈Uitalic_U), we propose a novel mechanism for the photoinduced metallic state: the empty-occupied and double-occupied dimers serve as the photoinduced charge carriers, akin to the holons and doublons in Hubbard model.

I Introduction

The exploration of nonequilibrium dynamics in strongly correlated systems opens up new avenues for disentangling the complicated interactions among different degrees of freedom and for realizing or manipulating quantum states, particularly within the realm of ultrafast pump-probe measurements Murakami et al. (2023). Experimental observations have revealed photoinduced transitions from Mott insulators to metallic states in various materials, including the halogen-bridged Ni-chain compound Iwai et al. (2003), the organic charge-transfer compound ET-F2TCNQ Okamoto et al. (2007); Mitrano et al. (2014), and the undoped cuprates Nd2CuO4 and La2CuO4 Okamoto et al. (2010). In the underdoped cuprate compounds, phenomena such as the enhancement of superconductivity caused by microwave Vedeneev et al. (2008) and the emergence of light-induced superconducting-like phases Fausti et al. (2011); Nicoletti et al. (2014); Hu et al. (2014); Kaiser et al. (2014) provide valuable insights into the origin of high-temperature superconductivity.

Theoretical investigations of photoinduced metallic state Oka et al. (2003); Oka and Aoki (2005); Maeshima and Yonemitsu (2005); Takahashi et al. (2008); Eckstein et al. (2010); Eckstein and Werner (2013); Shao et al. (2016); Shinjo and Tohyama (2017); Werner et al. (2019); Rincón and Feiguin (2021) and superconductivity Werner et al. (2018); Wang et al. (2018); Bittner et al. (2019); Werner et al. (2019); Kaneko et al. (2019); Wang et al. (2021); Zhang et al. (2023) in strongly correlated systems, mostly relying on half-filled Hubbard-type models, suggest that photoinduced charge carriers (holons and doublons) play a crucial role in generating the metallic state, while the pairing of these carriers contributes to the emergence of superconductivity. Recently, a photoinduced metallic state is studied in monoclinic VO2 using the nonequilibrium cluster dynamical mean field theory (cDMFT) and exact diagonalization (ED) methods. Their ED results, based on a Hubbard dimer, indicate that the mixed-orbital doublon state generated by laser pulses may contribute to the metallic state Chen et al. (2024).

The one-dimensional (1D) Peierls-Hubbard model at quarter filling has also attracted significant attention because its ground state exhibits an antiferromagnetic Mott insulator in units of dimers Le et al. (2020); Benthien and Jeckelmann (2005). The model can be used to describe a family of the charge-transfer salts Pedron et al. (1994); Nishimoto et al. (2000); Shibata et al. (2001); Tsuchiizu et al. (2001); Penc and Mila (1994); Mila (1995); Favand and Mila (1996); Benthien and Jeckelmann (2005). Photoinduced melting of charge order Yonemitsu and Maeshima (2007); Yonemitsu et al. (2009) and insulator-metal transition Lee (2009) have been reported by incorporating Holstein types of electron-phonon couplings to the Peierls-Hubbard model, aligning with the photoresponse in corresponding materials like (EDO-TTF)2PF6 Chollet et al. (2005). Furthermore, the ultrafast dynamics of the 1D extended Peierls-Hubbard model (including additional nearest-neighbour interactions) has been investigated, with a focus on the photoexcitation of electronic instabilities and polarization-inverted domains Yamaguchi et al. (2019); Rincón et al. (2014).

In this paper, we propose a novel mechanism for the photoinduced metallic state in the 1D quarter-filled Peierls-Hubbard model, based on the ED method. In equilibrium, increasing the interaction strength in a dimerized chain results in a significant suppression of the Drude weight in the optical conductivity, indicating the formation of an insulating state. Upon ultrafast photoirradiation of the system, we observe a substantial enhancement of the Drude peak, suggesting a photoinduced insulator-metal transition. Unlike the scenario in the half-filled Hubbard model where photoinduced charge carriers such as holons and doublons contribute to the appearance of Drude peak, we propose that the empty-occupied and double-occupied dimers serve as the photoinduced charge carriers in the quarter-filled Peierls-Hubbard model, thus facilitating the transition to a metallic state.

II Models and measurements

The 1D Peierls-Hubbard model can be written as H=Hk+HU𝐻subscript𝐻𝑘subscript𝐻𝑈H=H_{k}+H_{U}italic_H = italic_H start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT, where the kinetic part

Hk=i,σ(th(1+δ(1)i)ci,σci+1,σ+H.c.)subscript𝐻𝑘subscript𝑖𝜎subscript𝑡1𝛿superscript1𝑖subscriptsuperscript𝑐𝑖𝜎subscript𝑐𝑖1𝜎H.c.\displaystyle H_{k}=-\sum_{i,\sigma}\left(t_{h}(1+\delta(-1)^{i})c^{\dagger}_{% i,\sigma}c_{i+1,\sigma}+\text{H.c.}\right)italic_H start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = - ∑ start_POSTSUBSCRIPT italic_i , italic_σ end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( 1 + italic_δ ( - 1 ) start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ) italic_c start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i , italic_σ end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_i + 1 , italic_σ end_POSTSUBSCRIPT + H.c. ) (1)

and the interaction term

HU=Uini,ni,.subscript𝐻𝑈𝑈subscript𝑖subscript𝑛𝑖subscript𝑛𝑖\displaystyle H_{U}=U\sum_{i}n_{i,\uparrow}n_{i,\downarrow}.italic_H start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT = italic_U ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_i , ↑ end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_i , ↓ end_POSTSUBSCRIPT . (2)

Here ci,σsubscriptsuperscript𝑐𝑖𝜎c^{\dagger}_{i,\sigma}italic_c start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i , italic_σ end_POSTSUBSCRIPT (ci,σsubscript𝑐𝑖𝜎c_{i,\sigma}italic_c start_POSTSUBSCRIPT italic_i , italic_σ end_POSTSUBSCRIPT) is the creation (annihilation) operator of an electron at site i𝑖iitalic_i with spin σ𝜎\sigmaitalic_σ, and ni,σsubscript𝑛𝑖𝜎n_{i,\sigma}italic_n start_POSTSUBSCRIPT italic_i , italic_σ end_POSTSUBSCRIPT the number operator. th(1+δ(1)i)subscript𝑡1𝛿superscript1𝑖t_{h}(1+\delta(-1)^{i})italic_t start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( 1 + italic_δ ( - 1 ) start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ) denotes the staggered hop** constants and we set δ=0.5𝛿0.5\delta=0.5italic_δ = 0.5 throughout the paper. We consider a periodic chain with L𝐿Litalic_L sites and U𝑈Uitalic_U represents the strength of on-site Coulomb interaction. L𝐿Litalic_L is set to be 12121212 for the 1D Peierls-Hubbard model, and it is important to note that the number of the unit cells is L/2=6𝐿26L/2=6italic_L / 2 = 6. In this paper, we use units with e==c=a0=1𝑒Planck-constant-over-2-pi𝑐subscript𝑎01e=\hbar=c=a_{0}=1italic_e = roman_ℏ = italic_c = italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1, where e𝑒eitalic_e, Planck-constant-over-2-pi\hbarroman_ℏ and c𝑐citalic_c are the elementary charge, the reduced Planck constant and the speed of light, respectively, and 2a02subscript𝑎02a_{0}2 italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the lattice constant. With these units, thsubscript𝑡t_{h}italic_t start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT and th1superscriptsubscript𝑡1{t_{h}}^{-1}italic_t start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT are the unit of energy and time, respectively. Our focus is on zero temperature, with the system evolving from its ground state.

We utilize the time-dependent optical conductivity to characterize the photoinduced metallic behavior. The external electric field of a pum** pulse is included via the Peierls substitution

ci,σci+1,σ+H.c.eiA(t)ci,σci+1,σ+H.c..subscriptsuperscript𝑐𝑖𝜎subscript𝑐𝑖1𝜎H.c.superscript𝑒i𝐴𝑡subscriptsuperscript𝑐𝑖𝜎subscript𝑐𝑖1𝜎H.c.c^{\dagger}_{i,\sigma}c_{i+1,\sigma}+\text{H.c.}\rightarrow e^{\mathrm{i}A(t)}% c^{\dagger}_{i,\sigma}c_{i+1,\sigma}+\text{H.c.}.italic_c start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i , italic_σ end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_i + 1 , italic_σ end_POSTSUBSCRIPT + H.c. → italic_e start_POSTSUPERSCRIPT roman_i italic_A ( italic_t ) end_POSTSUPERSCRIPT italic_c start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i , italic_σ end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_i + 1 , italic_σ end_POSTSUBSCRIPT + H.c. . (3)

A(t)𝐴𝑡A(t)italic_A ( italic_t ) represents the vector potential of the pum** pulse, which can be expressed as

A(t)=A0e(tt0)2/2td2cos[ω0(tt0)].𝐴𝑡subscript𝐴0superscript𝑒superscript𝑡subscript𝑡022superscriptsubscript𝑡𝑑2subscript𝜔0𝑡subscript𝑡0A(t)=A_{0}e^{-\left(t-t_{0}\right)^{2}/2t_{d}^{2}}\cos\left[\omega_{0}\left(t-% t_{0}\right)\right].italic_A ( italic_t ) = italic_A start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - ( italic_t - italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 italic_t start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT roman_cos [ italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_t - italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) ] . (4)

This is a cosine function with a Gaussian envelope centered at t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. The parameters tdsubscript𝑡𝑑t_{d}italic_t start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT and ω0subscript𝜔0\omega_{0}italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT control its width and central frequency, respectively. To obtain the time-dependent wave function, the Lanczos method Prelovsˇˇs\check{\text{s}}overroman_ˇ start_ARG s end_ARGek and Boncˇˇc\check{\text{c}}overroman_ˇ start_ARG c end_ARGa is employed with the key formula

|ψ(t+δt)=eiH(t)δt|ψ(t)l=1Meiϵlδt|ϕlϕl|ψ(t),ket𝜓𝑡𝛿𝑡superscript𝑒i𝐻𝑡𝛿𝑡ket𝜓𝑡similar-to-or-equalssuperscriptsubscript𝑙1𝑀superscript𝑒isubscriptitalic-ϵ𝑙𝛿𝑡ketsubscriptitalic-ϕ𝑙inner-productsubscriptitalic-ϕ𝑙𝜓𝑡{|\psi(t+\delta{t})\rangle}=e^{-\mathrm{i}H(t)\delta t}{|\psi(t)\rangle}\simeq% \sum_{l=1}^{M}{e^{-\mathrm{i}\epsilon_{l}\delta{t}}}{|\phi_{l}\rangle}\langle% \phi_{l}|\psi(t)\rangle,| italic_ψ ( italic_t + italic_δ italic_t ) ⟩ = italic_e start_POSTSUPERSCRIPT - roman_i italic_H ( italic_t ) italic_δ italic_t end_POSTSUPERSCRIPT | italic_ψ ( italic_t ) ⟩ ≃ ∑ start_POSTSUBSCRIPT italic_l = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_M end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - roman_i italic_ϵ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT italic_δ italic_t end_POSTSUPERSCRIPT | italic_ϕ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ⟩ ⟨ italic_ϕ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT | italic_ψ ( italic_t ) ⟩ , (5)

where ϵlsubscriptitalic-ϵ𝑙\epsilon_{l}italic_ϵ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT and |ϕlketsubscriptitalic-ϕ𝑙|\phi_{l}\rangle| italic_ϕ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ⟩ are the Krylov eigenvalues and eigenvectors generated in the Lanczos process, respectively. Note that the initial state is the ground state of Hamiltonian before pump. In our simulations, we choose the time step δt=0.02𝛿𝑡0.02\delta{t}=0.02italic_δ italic_t = 0.02 and M=30𝑀30M=30italic_M = 30 to ensure the convergence of numerical results.

The time-dependent optical conductivity σ(ω,t)𝜎𝜔𝑡\sigma(\omega,t)italic_σ ( italic_ω , italic_t ) can be calculated by the nonequilibrium linear-response theory Lenarčič et al. (2014):

σ(ω,t)=0+σ(t+s,t)ei(ω+iη)sds,𝜎𝜔𝑡superscriptsubscript0𝜎𝑡𝑠𝑡superscript𝑒i𝜔𝑖𝜂𝑠differential-d𝑠\sigma(\omega,t)=\int_{0}^{+\infty}\sigma(t+s,t)e^{\mathrm{i}(\omega+i\eta)s}% \,\mathrm{d}s,italic_σ ( italic_ω , italic_t ) = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + ∞ end_POSTSUPERSCRIPT italic_σ ( italic_t + italic_s , italic_t ) italic_e start_POSTSUPERSCRIPT roman_i ( italic_ω + italic_i italic_η ) italic_s end_POSTSUPERSCRIPT roman_d italic_s , (6)

where the response function σ(t,t)𝜎superscript𝑡𝑡\sigma(t^{\prime},t)italic_σ ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_t ) (with ttsuperscript𝑡𝑡t^{\prime}\geq titalic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ≥ italic_t) reads

σ(t,t)=1L[ψ(t)|τ|ψ(t)+ttχ(t,t′′)dt′′].𝜎superscript𝑡𝑡1𝐿delimited-[]quantum-operator-product𝜓superscript𝑡𝜏𝜓superscript𝑡superscriptsubscript𝑡superscript𝑡𝜒superscript𝑡superscript𝑡′′differential-dsuperscript𝑡′′\sigma(t^{\prime},t)=\frac{1}{L}\left[\langle\psi(t^{\prime})|\tau|\psi(t^{% \prime})\rangle+\int_{t}^{t^{\prime}}\chi(t^{\prime},t^{\prime\prime})\,% \mathrm{d}t^{\prime\prime}\right].italic_σ ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_t ) = divide start_ARG 1 end_ARG start_ARG italic_L end_ARG [ ⟨ italic_ψ ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) | italic_τ | italic_ψ ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ⟩ + ∫ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT italic_χ ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_t start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ) roman_d italic_t start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ] . (7)

The first term of Eq. (7) is referred to as the diamagnetic term, and the 1D stress tensor operator reads

τ=i,σ(theiA(t)(1+δ(1)i)ci,σci+1,σ+H.c.).𝜏subscript𝑖𝜎subscript𝑡superscript𝑒i𝐴𝑡1𝛿superscript1𝑖subscriptsuperscript𝑐𝑖𝜎subscript𝑐𝑖1𝜎H.c.\tau=\sum_{i,\sigma}\left(t_{h}e^{\mathrm{i}A(t)}(1+\delta(-1)^{i})c^{\dagger}% _{i,\sigma}c_{i+1,\sigma}+\text{H.c.}\right).italic_τ = ∑ start_POSTSUBSCRIPT italic_i , italic_σ end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT roman_i italic_A ( italic_t ) end_POSTSUPERSCRIPT ( 1 + italic_δ ( - 1 ) start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ) italic_c start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i , italic_σ end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_i + 1 , italic_σ end_POSTSUBSCRIPT + H.c. ) . (8)

In the second term of Eq. (7), the two-time susceptibility

χ(t,t′′)=iθ(tt′′)ψ(t)|[jI(t),jI(t′′)]|ψ(t),𝜒superscript𝑡superscript𝑡′′i𝜃superscript𝑡superscript𝑡′′quantum-operator-product𝜓𝑡superscript𝑗𝐼superscript𝑡superscript𝑗𝐼superscript𝑡′′𝜓𝑡\chi(t^{\prime},t^{\prime\prime})=-\mathrm{i}\theta(t^{\prime}-t^{\prime\prime% })\langle\psi(t)|[j^{I}(t^{\prime}),j^{I}(t^{\prime\prime})]|\psi(t)\rangle,italic_χ ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_t start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ) = - roman_i italic_θ ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT - italic_t start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ) ⟨ italic_ψ ( italic_t ) | [ italic_j start_POSTSUPERSCRIPT italic_I end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) , italic_j start_POSTSUPERSCRIPT italic_I end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ) ] | italic_ψ ( italic_t ) ⟩ , (9)

where jI(t)=U(t,t)jU(t,t)superscript𝑗𝐼superscript𝑡superscript𝑈superscript𝑡𝑡𝑗𝑈superscript𝑡𝑡j^{I}(t^{\prime})=U^{\dagger}(t^{\prime},t)\,j\,U(t^{\prime},t)italic_j start_POSTSUPERSCRIPT italic_I end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) = italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_t ) italic_j italic_U ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_t ) is the interaction representation of the current operator, with U(t,t)𝑈superscript𝑡𝑡U(t^{\prime},t)italic_U ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_t ) the time-evolution operator and the current operator

j=ii,σ(theiA(t)(1+δ(1)i)ci,σci+1,σH.c.).𝑗isubscript𝑖𝜎subscript𝑡superscript𝑒i𝐴𝑡1𝛿superscript1𝑖subscriptsuperscript𝑐𝑖𝜎subscript𝑐𝑖1𝜎H.c.\displaystyle j=-\mathrm{i}\sum_{i,\sigma}\left(t_{h}e^{\mathrm{i}A(t)}(1+% \delta(-1)^{i})c^{\dagger}_{i,\sigma}c_{i+1,\sigma}-\text{H.c.}\right).italic_j = - roman_i ∑ start_POSTSUBSCRIPT italic_i , italic_σ end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT roman_i italic_A ( italic_t ) end_POSTSUPERSCRIPT ( 1 + italic_δ ( - 1 ) start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ) italic_c start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i , italic_σ end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_i + 1 , italic_σ end_POSTSUBSCRIPT - H.c. ) . (10)

Note that in Eq. (6), the integration cutoff for s𝑠sitalic_s is taken to be one hundred time units, and the broadening factor η=0.1𝜂0.1\eta=0.1italic_η = 0.1.

In what follows, we study the real part of the optical conductivity, i.e., Re σ(ω,t)𝜎𝜔𝑡\sigma(\omega,t)italic_σ ( italic_ω , italic_t ), before and after pump. The detailed derivation of the time-dependent optical conductivity can be found in Ref. Lenarčič et al. (2014), and its validity has been discussed in Ref. Shao et al. (2016).

III Results

Results of the optical conductivity before the pump (in equilibrium), Re σ(ω)𝜎𝜔\sigma(\omega)italic_σ ( italic_ω ), are depicted in Fig. 1 with different values of U𝑈Uitalic_U. Note that the optical conductivity obtained from the nonequilibrium linear-response theory Lenarčič et al. (2014) provides insights into the Drude weight, i.e., the zero-frequency weight of optical conductivity. For U=0𝑈0U=0italic_U = 0 in Fig. 1(a), a prominent Drude peak can be observed along with additional structures within the range of ω[3,4.5]𝜔34.5\omega\in[3,4.5]italic_ω ∈ [ 3 , 4.5 ]. As U𝑈Uitalic_U increases from 00 to 4444, as shown in Figs. 1(a), 1(b) and 1(c), the Drude peak lose its weight, while a side peak around ω=1.0𝜔1.0\omega=1.0italic_ω = 1.0 emerges and grows. It is worth noting that the y𝑦yitalic_y-axis range in Figs. 1(a) and 1(b) is [0,6]06[0,6][ 0 , 6 ], whereas in Figs. 1(c), 1(d), 1(e) and 1(f) it is [0,2]02[0,2][ 0 , 2 ]. This side peak becomes the main peak when U6𝑈6U\geq 6italic_U ≥ 6, and its position shifts to the right to ω=1.76𝜔1.76\omega=1.76italic_ω = 1.76 when U=10𝑈10U=10italic_U = 10, connecting to the left-shifted structures originally located at ω[3,4.5]𝜔34.5\omega\in[3,4.5]italic_ω ∈ [ 3 , 4.5 ] in Fig. 1(a). Despite a small remaining Drude peak in Fig. 1(f) with U=10𝑈10U=10italic_U = 10, attributed to the finite-size effect in PBC even for the insulator state Shao et al. (2016), the nonequilibrium linear-response theory proves useful in characterizing the metallic state in and out of equilibrium by analyzing the relative magnitudes of Drude peak. In addition, our results of equilibrium optical conductivity based on the ED method are similar to the results from the density-matrix renormalization group method Benthien and Jeckelmann (2005), despite slight variations in model parameters.

Refer to caption
Figure 1: (Color online) The optical conductivity Re σ(ω)𝜎𝜔\sigma(\omega)italic_σ ( italic_ω ) in equilibrium for the quarter-filled Peierls-Hubbard with L=12𝐿12L=12italic_L = 12, δ=0.5𝛿0.5\delta=0.5italic_δ = 0.5, and (a) U=0.0𝑈0.0U=0.0italic_U = 0.0, (b) U=2.0𝑈2.0U=2.0italic_U = 2.0, (c) U=4.0𝑈4.0U=4.0italic_U = 4.0, (d) U=6.0𝑈6.0U=6.0italic_U = 6.0, (e) U=8.0𝑈8.0U=8.0italic_U = 8.0, and (f) U=10.0𝑈10.0U=10.0italic_U = 10.0, respectively.
Refer to caption
Figure 2: (Color online) (a) The time-dependent optical conductivity Re σ(ω,Δt)𝜎𝜔Δ𝑡\sigma(\omega,\Delta t)italic_σ ( italic_ω , roman_Δ italic_t ), (b) the time-dependent kinetic energy Ek(Δt)subscript𝐸𝑘Δ𝑡E_{k}(\Delta t)italic_E start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( roman_Δ italic_t ) and interaction energy EU(Δt)subscript𝐸𝑈Δ𝑡E_{U}(\Delta t)italic_E start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT ( roman_Δ italic_t ), (c) the number of the double-occupied dimers NDsubscript𝑁DN_{\rm D}italic_N start_POSTSUBSCRIPT roman_D end_POSTSUBSCRIPT as a function of time, and (d) the time-dependent inter- and intra-dimer kinetic energys. Parameters of the quarter-filled Peierls-Hubbard model: L=12𝐿12L=12italic_L = 12, δ=0.5𝛿0.5\delta=0.5italic_δ = 0.5 and U=10.0𝑈10.0U=10.0italic_U = 10.0. Parameters of the pum** pulse: A0=0.2subscript𝐴00.2A_{0}=0.2italic_A start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.2, td=2.0subscript𝑡𝑑2.0t_{d}=2.0italic_t start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = 2.0, ω0=1.76subscript𝜔01.76\omega_{0}=1.76italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1.76.

In the nonequilibrium scenario, we consider the quarter-filled Peierls-Hubbard model with L=12𝐿12L=12italic_L = 12, δ=0.5𝛿0.5\delta=0.5italic_δ = 0.5 and U=10.0𝑈10.0U=10.0italic_U = 10.0, which is driven by a laser pulse described by Eq. (4). The laser frequency ω0=1.76subscript𝜔01.76\omega_{0}=1.76italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1.76 matches the main peak of the equilibrium optical conductivity. In the following discussions, we set the central time of pum** pulse t0=0subscript𝑡00t_{0}=0italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0 and substitute the variable t𝑡titalic_t in σ(ω,t)𝜎𝜔𝑡\sigma(\omega,t)italic_σ ( italic_ω , italic_t ) with Δt=tt0Δ𝑡𝑡subscript𝑡0\Delta t=t-t_{0}roman_Δ italic_t = italic_t - italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, representing the time difference between the probing time and t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. The real part of the optical conductivity, Re σ(ω,Δt)𝜎𝜔Δ𝑡\sigma(\omega,\Delta t)italic_σ ( italic_ω , roman_Δ italic_t ), for Δt=Δ𝑡\Delta t=-\inftyroman_Δ italic_t = - ∞, 10101010, 20202020 and 30303030 is depicted in Fig. 2(a) with different colors, where Δt=Δ𝑡\Delta t=-\inftyroman_Δ italic_t = - ∞ indicates the equilibrium optical conductivity without the application of the pum** pulse. We observe the emergence of a dominant peak at ω=0𝜔0\omega=0italic_ω = 0 after photoirradiation, signifying a photoinduced metallic state. To analyze the details of the nonequilibrium dynamics, we present the time-dependent kinetic energy Ek(Δt)subscript𝐸𝑘Δ𝑡E_{k}(\Delta t)italic_E start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( roman_Δ italic_t ), which is composed of the intra- and inter-dimer kinetic energies, and the time-dependent interaction energy EU(Δt)subscript𝐸𝑈Δ𝑡E_{U}(\Delta t)italic_E start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT ( roman_Δ italic_t ) in Fig. 2 (b). We find that EUsubscript𝐸𝑈E_{U}italic_E start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT increases from 0.240.240.240.24 to 0.630.630.630.63 after the pum** pulse. Referring to Eq. (2) where EU=Undsubscript𝐸𝑈𝑈subscript𝑛𝑑E_{U}=Un_{d}italic_E start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT = italic_U italic_n start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT, ndsubscript𝑛𝑑n_{d}italic_n start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT represents the total doublon number and U=10𝑈10U=10italic_U = 10 in the nonequilibrium discussions. It is worth noting that the doublon number is counted over sites, rather than dimers. As a result, the increase of doublon number caused by the pum** pulse is approximately 0.040.040.040.04 on the 12121212-site lattice. This suggests that the photoinduced doublons are insufficient to produce a metallic state.

However, the increase of kinetic energy amounts to 1.11.11.11.1, indicating that more than 70707070 percent of the injected energy contributes to the kinetic energy. Furthermore, from Fig. 2 (d), we observe that the intra-dimer kinetic energy increases, while the inter-dimer kinetic energy decreases. It is important to note that in equilibrium (before pump), the intra-dimer kinetic energy is smaller than the inter-dimer kinetic energy due to the formation of the hybridization gap (bonding-antibonding separation). Specifically, every dimer is occupied only by one electron, forming the antiferromagnetic ground state in units of dimers. With the injection of energy from the pump, electrons are facilitated in overcoming the barrier among dimers, leading to the appearance of empty-occupied and double-occupied dimers. The double-occupied dimer with one spin-\uparrow electron and one spin-\downarrow electron has four configurations: (\uparrow,\downarrow), (\downarrow,\uparrow), (\downarrow\uparrow,o) and (o,\downarrow\uparrow). Here o represents an empty site in the dimer. We present the number of the double-occupied dimers, NDsubscript𝑁DN_{\rm D}italic_N start_POSTSUBSCRIPT roman_D end_POSTSUBSCRIPT, as a function of time in Fig. 2 (c), where we observe that its increment is more than 0.60.60.60.6 for our 6666-dimer chain. We thus propose that the empty-occupied and double-occupied dimers serve as the photoinduced carriers contributing to a metallic state. Additionally, both empty-occupied and double-occupied dimers can result in less mobile electrons within dimers and more mobile electrons between dimers. This leads to an increase of the intra-dimer kinetic energy and a decrease of the intra-dimer kinetic energy (since kinetic energy is zero when there is no motion), as shown in Fig. 2 (d).

This phenomenon is reminiscent of the photoinduced carriers, i.e., holons or doublons, in the half-filled Hubbard model. Especially, the low-energy negative spectral weight in the postpump optical conductivity shown in Fig. 2 (a) can also be observed in the half-filled Hubbard model, see Fig. 3(a) which we will discuss later. As also mentioned in Ref. Lu et al. (2015), this negative peak at low frequency originates from an excitation from the optically allowed odd-parity state to an optically forbidden even-parity state. Furthermore, the energy separation between them vanishes when either U𝑈Uitalic_U or L𝐿Litalic_L becomes infinite, indicating that this negative peak can be attributed to the finite size effect. More details on the analysis of the energy separation between the odd- and even-parity states can be found in the Appendix.

Refer to caption
Figure 3: (Color online) (a) The time-dependent optical conductivity Re σ(ω,Δt)𝜎𝜔Δ𝑡\sigma(\omega,\Delta t)italic_σ ( italic_ω , roman_Δ italic_t ), (b) the time-dependent kinetic energy Ek(Δt)subscript𝐸𝑘Δ𝑡E_{k}(\Delta t)italic_E start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( roman_Δ italic_t ) and interaction energy EU(Δt)subscript𝐸𝑈Δ𝑡E_{U}(\Delta t)italic_E start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT ( roman_Δ italic_t ), (c) the time-dependent current j(Δt)𝑗Δ𝑡j(\Delta t)italic_j ( roman_Δ italic_t ), and (d) the time-dependent order parameter of spin-density-wave order 𝒪SDW(Δt)subscript𝒪SDWΔ𝑡\mathcal{O}_{\rm SDW}(\Delta t)caligraphic_O start_POSTSUBSCRIPT roman_SDW end_POSTSUBSCRIPT ( roman_Δ italic_t ). Parameters of the half-filled Hubbard model: L=10𝐿10L=10italic_L = 10 and U=10.0𝑈10.0U=10.0italic_U = 10.0. Parameters of the pum** pulse: A0=0.2subscript𝐴00.2A_{0}=0.2italic_A start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.2, td=2.0subscript𝑡𝑑2.0t_{d}=2.0italic_t start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = 2.0, ω0=7.68subscript𝜔07.68\omega_{0}=7.68italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 7.68.

Now we discuss the photoinduced metallic state in the 1D half-filled Hubbard model in order to compare the results with those of the 1D quarter-filled Peierls-Hubbard model. Here we set the lattice size L=10𝐿10L=10italic_L = 10 and the interaction U=10𝑈10U=10italic_U = 10, respectively. The equilibrium optical conductivity is shown in Fig. 3(a) with a black line, where the primary absorbing peak is observed at ω=7.68𝜔7.68\omega=7.68italic_ω = 7.68. Setting the central frequency of the pum** pulse, ω0subscript𝜔0\omega_{0}italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, to 7.687.687.687.68 and choosing A0=0.2subscript𝐴00.2A_{0}=0.2italic_A start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.2 and td=2.0subscript𝑡𝑑2.0t_{d}=2.0italic_t start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = 2.0 to be consistent with the laser parameters used in the quarter-filled Peierls-Hubbard model, we observe the emergence of the zero-frequency Drude peak and a low-frequency negative peak in the time-dependent optical conductivity, Re σ(ω,Δt)𝜎𝜔Δ𝑡\sigma(\omega,\Delta t)italic_σ ( italic_ω , roman_Δ italic_t ), with Δt=10Δ𝑡10\Delta t=10roman_Δ italic_t = 10, 20202020 and 30303030, as shown in Fig. 3(a) after the pum** pulse. The presence of the Drude peak signifies the onset of a photoinduced metallic state, while the emergence of the low-frequency negative peak is associated with the excitation between the odd- and even-parity states mentioned earlier. With an increment in interaction energy of 13131313, corresponding to an increase in doublon number by 1.31.31.31.3 for a 10101010-site lattice, the kinetic energy decreases from 5.25.2-5.2- 5.2 to 7.27.2-7.2- 7.2. This decrease suggests an increase of the mobility of carriers (doublons), including the formation of η𝜂\etaitalic_η pairing discussed in Ref. Kaneko et al. (2019). It becomes evident that photoinduced carriers and their transport contribute to the formation of photoinduced metallic state. This behavior is very similar to the results observed in the 1D quarter-filled Peierls-Hubbard model, where the photoinduced carriers are empty-occupied and double-occupied dimers instead.

Refer to caption
Figure 4: (Color online) (a) The time-dependent optical conductivity Re σ(ω,Δt)𝜎𝜔Δ𝑡\sigma(\omega,\Delta t)italic_σ ( italic_ω , roman_Δ italic_t ), (b) the time-dependent kinetic energy Ek(Δt)subscript𝐸𝑘Δ𝑡E_{k}(\Delta t)italic_E start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( roman_Δ italic_t ) inter- and intra-dimers. Parameters of the noninteracting hall-filled SSH model: L=12𝐿12L=12italic_L = 12 and δ=0.5𝛿0.5\delta=0.5italic_δ = 0.5. Parameters of the pum** pulse: A0=0.2subscript𝐴00.2A_{0}=0.2italic_A start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.2, td=2.0subscript𝑡𝑑2.0t_{d}=2.0italic_t start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = 2.0, ω0=2.00subscript𝜔02.00\omega_{0}=2.00italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 2.00.

The photoinduced metallic behavior of a noninteracting half-filled Su-Schrieffer-Heeger (SSH) model is also evident in the time-dependent optical conductivity in Fig. 4(a). The SSH model in our paper is based on a 6666-dimer chain with six spinless electrons filled in. Comparatively, the quarter-filled Peierls-Hubbard model is also based on a 6666-dimer chain, but with three spin-\uparrow and three spin-\downarrow electrons. When the interaction strength U𝑈Uitalic_U approaches infinity, the quarter-filled Peierls-Hubbard model transforms into this spinless half-filled SSH model. We examine their different mechanisms for photoinduced metallic state. The central frequency of the pum** pulse, ω0=2.0subscript𝜔02.0\omega_{0}=2.0italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 2.0, aligns with the first absorbing peak of the optical conductivity before pump. We set A0=0.2subscript𝐴00.2A_{0}=0.2italic_A start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.2 and td=2.0subscript𝑡𝑑2.0t_{d}=2.0italic_t start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = 2.0, consistent with the laser parameters applied on the quarter-filled Peierls-Hubbard model. After the pump, the first absorbing peak of the optical conductivity loses its weight and the Drude peak emerges, as shown in the time-dependent optical conductivity, Re σ(ω,Δt)𝜎𝜔Δ𝑡\sigma(\omega,\Delta t)italic_σ ( italic_ω , roman_Δ italic_t ) with Δt=10Δ𝑡10\Delta t=10roman_Δ italic_t = 10, 20202020 and 30303030, in Fig. 4(a). While, the absence of a low-energy negative peak in the time-dependent optical conductivity confirms our results in Appendix, indicating that as U𝑈Uitalic_U becomes infinite in the quarter-filled Peierls-Hubbard model, the energy separation between the odd- and even-parity states vanishes, and negative low-frequency peak will not emerge. On the other hand, the noticeable increase (decrease) observed in the time-dependent kinetic energy intra-(inter-)dimers, as depicted in Fig. 4(b), compared to those shown in Fig. 2 (d) in the quarter-filled Peierls-Hubbard model, can be attributed to the limited mobility of photoinduced double-occupied electrons within dimers, stemming from their spinless fermionic nature.

IV Summary and discussion

To summarize, employing the exact diagonalization method, we presented the equilibrium outcomes of the optical conductivity in the 1D quarter-filled Peierls-Hubbard model. Starting with a purely dimerized chain, we observe that increasing U𝑈Uitalic_U leads to the suppression of the Drude weight in the optical conductivity. This confirms previous studies suggesting that this model manifests an antiferromagnetic insulator, arising from the combined effects of Peierls instability and electron correlations. Under nonequilibrium conditions, we observed the emergence of a photoinduced metallic state following a pum** pulse, as indicated by the appearance of a prominent Drude peak. However, we noted that the increase of photoinduced carriers, i.e., doublon numbers, is very small compared to the photoinduced metallic state in the pure Hubbard chain at half-filling. Instead, we propose that the empty-occupied and double-occupied dimers can serve as the photoinduced carriers contributing to the transition from a Mott insulator to a metallic state.

Acknowledgements.
C. S. acknowledges support from the National Natural Science Foundation of China (NSFC; Grant No. 12104229) and the Fundamental Research Funds for the Central Universities (Grant No. 30922010803). T. T. is partly supported by the Japan Society for the Promotion of Science, KAKENHI (Grant No. 24K00560) from the Ministry of Education, Culture, Sports, Science, and Technology, Japan. H. L. acknowledges support from the National Natural Science Foundation of China (NSFC; Grants No. 11474136 and No. 12247101).

*

Appendix A Appendix
Analysis of the energy separation between the odd- and even-parity states

Refer to caption
Figure 5: The correlation function jjdelimited-⟨⟩𝑗𝑗\langle jj\rangle⟨ italic_j italic_j ⟩ (blue lines) and ττdelimited-⟨⟩𝜏𝜏\langle\tau\tau\rangle⟨ italic_τ italic_τ ⟩ (red lines) with L=12𝐿12L=12italic_L = 12 and U=10𝑈10U=10italic_U = 10 in (a), L=12𝐿12L=12italic_L = 12 and U=100𝑈100U=100italic_U = 100 in (b), and L=20𝐿20L=20italic_L = 20 and U=10𝑈10U=10italic_U = 10 in (c), respectively. (d) The energy separation between the odd- and even-parity states, ΔEΔ𝐸\Delta Eroman_Δ italic_E, as a function of 1/L1𝐿1/L1 / italic_L.

In the main text, after photoirradiation of the one-dimensional (1D) quarter-filled Peierls-Hubbard model, we observe the emergence of a low-energy negative peak at ω0.23𝜔0.23\omega\approx 0.23italic_ω ≈ 0.23 in the postpump optical conductivity. Similar to observations in the 1D Hubbard model Lu et al. (2015), this negative peak is attributed to an excitation from the optically allowed odd-parity state to an optically forbidden even-parity state. To confirm this, we calculate the current-current correlation function

jj=πLa|a|j|0|2δ(ωEa+E0),delimited-⟨⟩𝑗𝑗𝜋𝐿subscript𝑎superscriptquantum-operator-product𝑎𝑗02𝛿𝜔subscript𝐸𝑎subscript𝐸0\langle jj\rangle=\frac{\pi}{L}\sum_{a}|\langle a|j|0\rangle|^{2}\delta(\omega% -E_{a}+E_{0}),⟨ italic_j italic_j ⟩ = divide start_ARG italic_π end_ARG start_ARG italic_L end_ARG ∑ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT | ⟨ italic_a | italic_j | 0 ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_δ ( italic_ω - italic_E start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT + italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) , (11)

and the correlation function replacing the current operator j𝑗jitalic_j with the stress tensor operator τ𝜏\tauitalic_τ

ττ=πLb|b|τ|0|2δ(ωEb+E0).delimited-⟨⟩𝜏𝜏𝜋𝐿subscript𝑏superscriptquantum-operator-product𝑏𝜏02𝛿𝜔subscript𝐸𝑏subscript𝐸0\langle\tau\tau\rangle=\frac{\pi}{L}\sum_{b}|\langle b|\tau|0\rangle|^{2}% \delta(\omega-E_{b}+E_{0}).⟨ italic_τ italic_τ ⟩ = divide start_ARG italic_π end_ARG start_ARG italic_L end_ARG ∑ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT | ⟨ italic_b | italic_τ | 0 ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_δ ( italic_ω - italic_E start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT + italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) . (12)

Considering the symmetries of the operators j𝑗jitalic_j and τ𝜏\tauitalic_τ (see their definition the main text), |aket𝑎|a\rangle| italic_a ⟩ and |bket𝑏|b\rangle| italic_b ⟩ must be the odd- and even-parity states, respectively Takahashi et al. (2002).

We present the results of jjdelimited-⟨⟩𝑗𝑗\langle jj\rangle⟨ italic_j italic_j ⟩ (blue lines) and ττdelimited-⟨⟩𝜏𝜏\langle\tau\tau\rangle⟨ italic_τ italic_τ ⟩ (red lines) as a function of EnE0subscript𝐸𝑛subscript𝐸0E_{n}-E_{0}italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT in Fig. 5 (a) with L=12𝐿12L=12italic_L = 12 and U=10𝑈10U=10italic_U = 10, in Fig. 5 (b) with L=12𝐿12L=12italic_L = 12 and U=100𝑈100U=100italic_U = 100, in Fig. 5 (c) with L=20𝐿20L=20italic_L = 20 and U=10𝑈10U=10italic_U = 10, respectively. Notice that L𝐿Litalic_L and U𝑈Uitalic_U represent the lattice size and the interaction strength, respectively, and the results are normalized by odd and even state we can concerned with. In Fig. 5 (a), the normalized delta peak of jjdelimited-⟨⟩𝑗𝑗\langle jj\rangle⟨ italic_j italic_j ⟩ corresponds to the main peak of the equilibrium optical conductivity, as seen in Fig.1 (f) in the main text. The energy separation between the odd-parity state and its nearest even-parity state is ΔE0.22Δ𝐸0.22\Delta E\approx 0.22roman_Δ italic_E ≈ 0.22, matching the low-energy negative peak at ω0.23𝜔0.23\omega\approx 0.23italic_ω ≈ 0.23 in the postpump optical conductivity, as shown in Fig. 2(a) in the main text. However, ΔEΔ𝐸\Delta Eroman_Δ italic_E becomes very small when U𝑈Uitalic_U is increased to 100100100100, as shown in Fig. 5 (b), and the odd- and even-parity states actually merge together when U=+𝑈U=+\inftyitalic_U = + ∞. In this scenario, the quarter-filled Peierls-Hubbard model transforms into the spinless noninteracting half-filled Su-Schrieffer-Heeger (SSH) model, which explains why we do not observe that low-frequency negative peak in the post-pump optical conductivity of the SSH model, see Fig. 4(a) in the main text.

On the other hand, we find that the energy separation between the odd- and even-parity states may be attributed to the finite-size effect even for finite U𝑈Uitalic_U. For example, we present the results of jjdelimited-⟨⟩𝑗𝑗\langle jj\rangle⟨ italic_j italic_j ⟩ and ττdelimited-⟨⟩𝜏𝜏\langle\tau\tau\rangle⟨ italic_τ italic_τ ⟩ with L=20𝐿20L=20italic_L = 20 and U=10𝑈10U=10italic_U = 10 in Fig. 5(c), where we find that ΔEΔ𝐸\Delta Eroman_Δ italic_E is approximately 0.130.130.130.13. In Fig 5(d), we show the results of ΔEΔ𝐸\Delta Eroman_Δ italic_E as a function of 1/L1𝐿1/L1 / italic_L for L=12𝐿12L=12italic_L = 12 and L=20𝐿20L=20italic_L = 20, from which it can be inferred that ΔEΔ𝐸\Delta Eroman_Δ italic_E may vanish in the thermodynamic limit (L+𝐿L\rightarrow+\inftyitalic_L → + ∞).

We employ the Lanczos method for the calculations of these correlation functions, with the cutoff number M=200𝑀200M=200italic_M = 200 and 300300300300 for L=12𝐿12L=12italic_L = 12 and L=20𝐿20L=20italic_L = 20, respectively.

References