Quantum fluctuations lead to glassy electron dynamics in the good metal regime of electron doped KTaO3

Shashank Kumar Ojha [email protected] Department of Physics, Indian Institute of Science, Bengaluru 560012, India    Sankalpa Hazra Contributed equally Department of Physics, Indian Institute of Science, Bengaluru 560012, India Department of Materials Science and Engineering, The Pennsylvania State University, University Park, PA 16802, USA    Surajit Bera Contributed equally Department of Physics, Indian Institute of Science, Bengaluru 560012, India    Sanat Kumar Gogoi Department of Physics, Indian Institute of Science, Bengaluru 560012, India Department of Physics, Digboi College, Digboi 786171, India    Prithwijit Mandal Department of Physics, Indian Institute of Science, Bengaluru 560012, India    Jyotirmay Maity Department of Physics, Indian Institute of Science, Bengaluru 560012, India    A. Gloskovskii Deutsches Elektronen-Synchrotron DESY, 22607 Hamburg, Germany    C. Schlueter Deutsches Elektronen-Synchrotron DESY, 22607 Hamburg, Germany    Smarajit Karmakar Tata Institute of Fundamental Research, 36/P, Gopanpally Village, Serilingampally Mandal, Ranga Reddy District, Hyderabad, 500107, India    Manish Jain Department of Physics, Indian Institute of Science, Bengaluru 560012, India    Sumilan Banerjee [email protected] Department of Physics, Indian Institute of Science, Bengaluru 560012, India    Venkatraman Gopalan Department of Materials Science and Engineering, The Pennsylvania State University, University Park, PA 16802, USA    Srimanta Middey [email protected] Department of Physics, Indian Institute of Science, Bengaluru 560012, India
Abstract

Abstract: One of the central challenges in condensed matter physics is to comprehend systems that have strong disorder and strong interactions. In the strongly localized regime, their subtle competition leads to glassy electron dynamics which ceases to exist well before the insulator-to-metal transition is approached as a function of do**. Here, we report on the discovery of glassy electron dynamics deep inside the good metal regime of an electron-doped quantum paraelectric system: KTaO3. We reveal that upon excitation of electrons from defect states to the conduction band, the excess injected carriers in the conduction band relax in a stretched exponential manner with a large relaxation time, and the system evinces simple aging phenomena - a telltale sign of glassy dynamics. Most significantly, we observe a critical slowing down of carrier dynamics below 35 K, concomitant with the onset of quantum paraelectricity in the undoped KTaO3. Our combined investigation using second harmonic generation technique, density functional theory and phenomenological modeling demonstrates quantum fluctuation-stabilized soft polar modes as the impetus for the glassy behavior. This study addresses one of the most fundamental questions regarding the potential promotion of glassiness by quantum fluctuations and opens a route for exploring glassy dynamics of electrons in a well-delocalized regime.

Introduction: The notion of glassy dynamics associated with the electronic degree of freedom in condensed matter systems was first envisaged by Davies, Lee and Rice in 1982 [1]. Building heavily on Anderson’s seminal work on the localization of wave functions in random disordered media [2], it was predicted that, in a disordered insulator with highly localized electronic states, the interplay between disorder and long-range Coulomb interaction (Fig. 1a) should precipitate electronic frustration in real space. Such a scenario would result in a rugged energy landscape with numerous metastable states, leading to the emergence of electron glass [3, 4, 5]. This hypothesis was soon tested on various strongly localized electronic systems, including granular metals, crystalline and amorphous oxides, and later, it was also extended to doped semiconductors and two-dimensional electron gases [6, 3, 4, 5, 7]. The typical relaxation time in these systems ranges from a few seconds to several hours, making them an obvious choice for studying glassy physics in laboratory timescales. What makes these systems even more intriguing is the array of perturbations that one can use to effectively drive them away from equilibrium [4]. Furthermore, due to the light mass of electrons, electron glasses are highly susceptible to quantum fluctuations. This aspect introduces additional complexities in understanding the behavior of electron glasses [8].

In the antithetical regime of highly delocalized electrons i.e. in a metal, the screening effect significantly reduces the strength of the electron-electron and electron-impurity interactions. Consequently, such system generally possesses a non-degenerate ground state with a well-defined Fermi surface. As a result, the dynamics of glassy behavior, which involves the existence of multiple, competing ground states, is incompatible with the behavior of metals. In fact, the manifestation of glassiness fades away considerably prior to the transition from insulator to metal, and there is an absolute lack of any substantiated indication of the presence of glassiness within a good metal regime [3] where quasi-particle mean free path is larger than the electron’s wavelength.

In this work, we report on the discovery of glassy dynamics of conduction electrons in an electron-doped quantum paraelectric system, namely KTaO3, in a good metal regime. Even more surprising observation is that glassiness is found to appear in a regime where quantum fluctuations are inherently present in the system. In pristine KTaO3, the quantum fluctuations associated with the zero point motion of the atoms prohibits the onset of ferroelectric order below 35 K (Fig. 1b), and the system’s properties are typically governed by the presence of an associated low energy transverse optical phonon (Fig. 1c) popularly known as soft polar mode [14]. Our combined transport and optical second harmonic generation measurements find that properties associated with the soft polar mode are preserved even deep inside the metallic regime. Most importantly, such soft modes are found to be directly responsible for emergent glassy dynamics at low temperatures, which is further corroborated by our theoretical calculations. Our observation is one of the rarest examples where quantum fluctuation, which is generally considered as a bottleneck for electron glass formation [8], is ultimately accountable for the appearance of glassy dynamics in a good metallic phase.

Results:

Demonstration of good metallic behavior: Due to the remarkable applications of KTO in the field of spintronics and prospects of studying emergent physics close to the ferroelectric quantum critical point, several successful attempts have been made in recent times to introduce free carriers in the bulk as well as at the surface or interface of KTO and a wide range of phenomena ranging from topology to 2D superconductivity have been reported  [15, 16, 17, 18]. However, so far there has been no report about glassy dynamics in electron doped KTO. For the current investigation, metallic samples have been prepared by introducing oxygen vacancies in pristine single crystalline (001) oriented KTaO3 substrate (for details see methods section and references  [2, 19]). These samples were found to exhibit quantum oscillations below 10 K [2], which is signature of a good metal with well-defined Fermi surface. To further testify this, we have also computed temperature-dependent mean free path of electrons (lesubscript𝑙𝑒l_{e}italic_l start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT) within Drude–Boltzmann picture. Fig. 1d shows the corresponding plots for degenerate and non-degenerate case. A dotted vertical line marks the temperature above which lesubscript𝑙𝑒l_{e}italic_l start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT becomes shorter than the inverse of Fermi wave-vector (kFsubscript𝑘𝐹k_{F}italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT) and the sample crosses over from good metal to a bad metal phase [6]. In the current work, observation of glassy dynamics is inherently constrained to temperatures which is much lower than crossover temperature to bad metal phase and hence for all practical purposes our electron doped KTaO3 system can be considered as a good metal with well-defined scattering.

Demonstration of glassy dynamics: Oxygen vacancy creation in KTaO3 not only adds free electrons but also leads to the formation of highly localized defect states. To determine the exact position of the defect states, valence band spectrum has been mapped out by using hard X-ray photoelectron spectroscopy (HAXPES) at P22 beamline of PETRA III, DESY (see methods for more details). Apart from the well-defined quasi-particle peak, mid-gap states centered at 1.6-1.8 eV is observed (Fig. 2a), which arises due to the clustering of oxygen vacancies [19].

In this work, we utilize these defect states to perturb the system by selective excitation of trapped electrons to the conduction band via sub-bandgap light illumination. Subsequently, the system’s response is studied by monitoring the temporal evolution of the electrical resistance at a fixed temperature (see inset of Fig. 2a for transport measurement set-up). For each measurement, the sample was first cooled down to the desired temperature. Once the temperature stabilizes, the system was driven out of equilibrium by shining light for half an hour, thereafter resistance relaxation was observed for the next 1.5 hours in dark condition. For the next measurement, the sample was heated to room temperature where the original resistance is recovered. Fig. 2b shows one set of data recorded with green light (λ𝜆\lambdaitalic_λ=527 nm) at several fixed temperatures ranging from 15 K to 150 K.

At first glance, Fig. 2b reveals a striking temperature dependence of the photo-do** effect (also see Supplementary Note 2 and Supplementary Figure 2) and the way the system relaxes after turning off the light is also found to be strongly temperature dependent. Further analysis reveals that in the off-stage, the resistance relaxes in a stretched exponential manner [exp(-(t𝑡titalic_t/τ𝜏\tauitalic_τ)β) where τ𝜏\tauitalic_τ is the relaxation time and β𝛽\betaitalic_β (stretching exponent) <<< 1 (see Supplementary Figure 3 for fitting of few representative data)]. Such stretched exponential relaxation is very often considered as a signature of glassy dynamics and has been observed in variety of glassy systems [4, 22, 23]. In glass physics, it is commonly accepted that such stretching is due to a distribution of relaxation times arising from the disorder-induced heterogeneity [24, 25]. In the present case, the distribution of relaxation times would correspond to multiple relaxation channels for the electron-hole recombination [26]. The microscopic origin behind the spatial separation of electron-hole pair and resultant non-exponential relaxation will be discussed later. The temperature evolution of β𝛽\betaitalic_β and τ𝜏\tauitalic_τ obtained from the fitting further reveals a substantial increase in relaxation time below 50 K with a power law behavior (τT2.8similar-to𝜏superscript𝑇2.8\tau\sim T^{-2.8}italic_τ ∼ italic_T start_POSTSUPERSCRIPT - 2.8 end_POSTSUPERSCRIPT, see Supplementary Note 4 and Supplementary Figure 4) followed by constant β𝛽absent\beta\approxitalic_β ≈ 0.5 below 35 K (Fig. 2c). This observation is remarkable given the fact that this crossover temperature roughly coincides with the onset of quantum fluctuation in undoped KTaO3.

One critical test to confirm glassiness is the observation of the aging phenomenon wherein the system’s response depends on its age [27]. More precisely, older systems are found to relax more slowly than younger ones. To examine this, we prepared the system of desired age by tuning the duration of light illumination (t𝑡titalic_tillill{}_{\text{ill}}start_FLOATSUBSCRIPT ill end_FLOATSUBSCRIPT), and measurements similar to that shown in Fig. 2b were performed at 15 K (Fig. 3a). For further analysis, we only focus on relaxation after turning off the light. In Fig. 3b we plot the change in resistance in the off-stage (ΔΔ\Deltaroman_ΔR𝑅Ritalic_Roffoff{}_{\text{off}}start_FLOATSUBSCRIPT off end_FLOATSUBSCRIPT) normalized with a total drop in resistance (ΔΔ\Deltaroman_ΔR𝑅Ritalic_Rillill{}_{\text{ill}}start_FLOATSUBSCRIPT ill end_FLOATSUBSCRIPT) at the end of illumination. Evidently, with increasing t𝑡titalic_tillill{}_{\text{ill}}start_FLOATSUBSCRIPT ill end_FLOATSUBSCRIPT system’s response becomes more and more sluggish. More precisely, τ𝜏\tauitalic_τ obtained from the fitting is found to scale linearly with t𝑡titalic_tillill{}_{\text{ill}}start_FLOATSUBSCRIPT ill end_FLOATSUBSCRIPT (inset of Fig. 3c). This is the defining criteria for simple or full aging [27] which is more clear in Fig. 3c where all the curves can be collapsed to a universal curve by normalizing the abscissa by t𝑡titalic_tillill{}_{\text{ill}}start_FLOATSUBSCRIPT ill end_FLOATSUBSCRIPT. A careful look at Fig. 3c reveals that for larger values of t𝑡titalic_tillill{}_{\text{ill}}start_FLOATSUBSCRIPT ill end_FLOATSUBSCRIPT, curves start to deviate from the universal scaling. This is much clear in Fig. 3d which contains a similar set of scaling data for another sample with a lower carrier concentration. Such an observation is consistent with the criteria for aging that t𝑡titalic_tillill{}_{\text{ill}}start_FLOATSUBSCRIPT ill end_FLOATSUBSCRIPT should be much less than the time required to reach the new equilibrium under perturbation. As evident from Fig. 3a and inset of Fig. 3d, above a critical value of t𝑡titalic_tillill{}_{\text{ill}}start_FLOATSUBSCRIPT ill end_FLOATSUBSCRIPT there is little change in resistance upon shining the light any further. This signifies that the system is closer to its new equilibrium and hence aging ceases to hold at a higher t𝑡titalic_tillill{}_{\text{ill}}start_FLOATSUBSCRIPT ill end_FLOATSUBSCRIPT.

Presence of polar nano regions & importance of quantum fluctuations: As mentioned earlier, the glassy behavior of electrons in conventional electron glasses results from the competition between disorder and Coulomb interactions and is only applicable in strongly localized regime [3, 4, 5]. As glassiness is observed within a good metal regime (kFsubscript𝑘𝐹k_{F}italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPTlesubscript𝑙𝑒l_{e}italic_l start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT>>>1) in our oxygen-deficient KTaO3 samples, we need to find other processes responsible for the electron-hole separation and complex glassy relaxations in the present case. In the context of electron do** in another well-studied quantum paraelectric SrTiO3, a large lattice relaxation (LLR) model involving deep trap levels [10] has been associated as a dominant cause for prohibiting electron-hole recombination [29, 30]. However, our analysis of τ𝜏\tauitalic_τ vs. T𝑇Titalic_T (Supplementary Note 7 and Supplementary Figure 7) does not support the applicability of the LLR model in the present case. In sharp contrast, we will conclusively demonstrate here that the effective charge separation in such systems directly correlates with the appearance of polar nano regions (PNRs), which arise as a direct consequence of the defect dipoles present in a highly polarizable lattice of quantum paraelectric [31].

In an ordinary dielectric host, an electric dipole can polarize the lattice only in its immediate vicinity and hence the correlation length (rcsubscript𝑟𝑐r_{c}italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT) is generally of the order of unit cell length which further remains independent of temperature [17]. However the situation is drastically different in highly polarizable hosts such as KTaO3 where the magnitude of rcsubscript𝑟𝑐r_{c}italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is controlled by the polarizability of the lattice which is inversely proportional to the soft mode frequency (ω𝜔\omegaitalic_ωs). Since ω𝜔\omegaitalic_ωs decreases with decreasing temperature, rcsubscript𝑟𝑐r_{c}italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT becomes large at lower temperatures. As a result, PNRs spanning several unit cells are formed around the defect dipole which is randomly distributed in the lattice (Fig. 4a).

While PNRs are quite well established in the insulating regime (Supplementary Note 8 and Supplementary Figure 8), they are expected to vanish in the metals due to screening effects from free electrons. Surprisingly, several recent experiments have reported that PNRs can exist even in the metallic regime [33, 34, 35, 36, 37]. Motivated by these results, we have carried out temperature-dependent optical second harmonic generation (SHG) measurement (Fig. 4b) which is a powerful technique to probe PNRs [25]. Fig. 4c shows the temperature evolution of SHG intensity which is directly proportional to the volume density of PNRs. As evident, no appreciable SHG signal is observed at room temperature, however, a strong signal enhancement is observed below 150 K, signifying the appearance of PNR below 150 K in our metallic sample. Since this onset temperature exactly coincides with the temperature below which an appreciable photo-do** effect is observed in our transport measurements (Supplementary Note 9 and Supplementary Figure 9), we believe that the internal electric field generated around such PNRs is the major cause behind driving apart the photo-generated electron-hole pairs in real space. Another notable observation is that the SHG intensity is independent of the temperature below 35 K. This immediately reminds of the regime of quantum fluctuation which enforces a constant value of ωssubscript𝜔𝑠\omega_{s}italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT below 35 K [14]. Since rcωs1proportional-tosubscript𝑟𝑐superscriptsubscript𝜔𝑠1r_{c}\propto\omega_{s}^{-1}italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ∝ italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT [14], our SHG measurement conclusively establishes that quantum fluctuation stabilized soft polar mode is retained even in metallic KTaO3 [39] (also see Supplementary Note 10).

We have also conducted an investigation into the primary defect dipoles responsible for the creation of PNRs in our samples. Considering that our samples were prepared through high-temperature annealing within an evacuated sealed quartz tube [19], there is a possibility of K vacancies due to its high volatility. The presence of a certain degree of K vacancy has been indeed observed in our HAXPES measurements (Supplementary Note 11 and Supplementary Figure 11). It is widely established that, in AB𝐴𝐵ABitalic_A italic_BO3 systems, the off-centering of substitute B𝐵Bitalic_B atom (B𝐵Bitalic_B antisite-like defect) in the presence of A𝐴Aitalic_A atom vacancy leads to a macroscopic polarization even in a non-polar matrix  [40, 41, 25, 42, 43]. Our density functional theory calculations (details are in the method section) considering Ta antisite-like defect has found significant Ta off-centering along [100] and [110] in the presence of K vacancy (Fig. 4d). Further, we have also computed induced polarization in the system following the modern theory of polarization where the change in macroscopic electric polarization is represented by a Berry phase [28, 31, 46, 29, 48] and the non-zero Berry curvature i.e., Berry phase per unit area, is taken as a signature of finite polarization in the material. In Fig.4 e, f, we have plotted the Berry curvature in the plane kz=0subscript𝑘𝑧0k_{z}=0italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0 (Ωxy(𝐤)subscriptΩ𝑥𝑦𝐤\Omega_{xy}(\mathbf{k})roman_Ω start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPT ( bold_k )) for the Ta off-centering along [110] and [100], respectively. From the figure, it is clear that the large contributions to the Berry curvature are due to the avoided crossings of bands at the Fermi surface (see inset of Fig. 4f) which are induced by spin-orbit coupling.

Phenomenological model to understand glassy dynamics: We now discuss a possible mechanism for the emergent glassy dynamics in a metal where conduction electrons coexist with PNRs. Since the glassiness in the present case is observed in the dynamical relaxation of excess injected carriers in the conduction band, it is necessary to have a thorough understanding of the relaxation processes happening against the backdrop of randomly oriented PNRs. In indirect-band semiconductors like KTaO3 which have strong electron-lattice and defect-lattice coupling, the relaxation should be predominantly nonradiative and manifest itself as the emission of several low-energy phonons [49]. Further, since inter-band electron-phonon matrix element due to acoustic phonons are negligible, the electron-hole recombination may primarily involve soft polar modes at low temperatures, although it is not the lowest energy phonon [50]. While such a multi-phonon inter-band transition could lead to multichannel relaxation with large time scales [49], it can never give rise to collective glassy behavior.

Instead, we suggest the following scenario for the observed glassiness. As was previously mentioned, there is clear evidence that the internal electric field around PNRs has a significant impact on electron-hole recombination in our sample. It has been demonstrated previously [31] that the random interactions between PNRs (in the limit of dilute defect dipoles) leads to a dipole glass at low temperatures in KTaO3. Recently, long-lived glass-like relaxations in SHG and Kerr signals were observed in pristine KTO at temperatures below 50 K and was attributed to dipolar correlations among PNRs, further highlighting the potential role of PNRs in influencing relaxation properties [51, 52]. Electrons and holes being charged particles would immediately couple to the complex electric field from the dipole glass and hence there is a chance that the glassy background of PNRs can induce glassiness to the free carriers in the system.

In order to study such a possibility, we consider a theoretical model with the Hamiltonian,

H𝐻\displaystyle Hitalic_H =Hel+Hgl+Helgl,absentsubscript𝐻𝑒𝑙subscript𝐻𝑔𝑙subscript𝐻𝑒𝑙𝑔𝑙\displaystyle=H_{el}+H_{gl}+H_{el-gl},= italic_H start_POSTSUBSCRIPT italic_e italic_l end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT italic_g italic_l end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT italic_e italic_l - italic_g italic_l end_POSTSUBSCRIPT , (1)

where Hel=ijtijcicjε0αfαfαsubscript𝐻𝑒𝑙subscript𝑖𝑗subscript𝑡𝑖𝑗superscriptsubscript𝑐𝑖subscript𝑐𝑗subscript𝜀0subscript𝛼superscriptsubscript𝑓𝛼subscript𝑓𝛼H_{el}=-\sum_{ij}t_{ij}c_{i}^{\dagger}c_{j}-\varepsilon_{0}\sum_{\alpha}f_{% \alpha}^{\dagger}f_{\alpha}italic_H start_POSTSUBSCRIPT italic_e italic_l end_POSTSUBSCRIPT = - ∑ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_f start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT (ε0>0subscript𝜀00\varepsilon_{0}>0italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT > 0) describes the electronic part in terms of creation and annihilation operators ci,cisuperscriptsubscript𝑐𝑖subscript𝑐𝑖c_{i}^{\dagger},c_{i}italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT , italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT (i=1,,Nc𝑖1subscript𝑁𝑐i=1,\cdots,N_{c}italic_i = 1 , ⋯ , italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT) and fα,fαsuperscriptsubscript𝑓𝛼subscript𝑓𝛼f_{\alpha}^{\dagger},f_{\alpha}italic_f start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT , italic_f start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT (α=1,,Nf𝛼1subscript𝑁𝑓\alpha=1,\cdots,N_{f}italic_α = 1 , ⋯ , italic_N start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT) of the Ncsubscript𝑁𝑐N_{c}italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT conduction and Nfsubscript𝑁𝑓N_{f}italic_N start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT impurity electronic states, respectively. For our calculations, we consider various lattices and corresponding hop** amplitudes tijsubscript𝑡𝑖𝑗t_{ij}italic_t start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT to describe several different energy dispersions for the conduction band, e.g., a band with semicircular DOS g(ϵ)=(1/2π)W2ω2θ(W|ω|)𝑔italic-ϵ12𝜋superscript𝑊2superscript𝜔2𝜃𝑊𝜔g(\epsilon)=(1/2\pi)\sqrt{W^{2}-\omega^{2}}\theta(W-|\omega|)italic_g ( italic_ϵ ) = ( 1 / 2 italic_π ) square-root start_ARG italic_W start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_θ ( italic_W - | italic_ω | ) with band width W𝑊Witalic_W [θ(x)𝜃𝑥\theta(x)italic_θ ( italic_x ) is heaviside step function] (Fig. 5(a)), and a flat band with width W=0𝑊0W=0italic_W = 0, as discussed in the Supplementary Note 13.

We take a flat impurity band at energy ε0subscript𝜀0-\varepsilon_{0}- italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, separated by a gap Δ=ε0W/2Δsubscript𝜀0𝑊2\Delta=\varepsilon_{0}-W/2roman_Δ = italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_W / 2 from the conduction band minimum. We set the chemical potential μ=0𝜇0\mu=0italic_μ = 0, at the centre of conduction band (Fig. 5(a)).

To model the dynamics of the glassy background, which may result from either a single PNR or randomly-distributed coupled assembly of PNRs, we consider a system with Ngsubscript𝑁𝑔N_{g}italic_N start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT degrees of freedom {xμ}subscript𝑥𝜇\{x_{\mu}\}{ italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT } (Supplementary Note 13).

These position-like variables, related to the electric dipoles inside the PNRs, can be thought of as a multi-coordinate generalization of the usual single configuration coordinate  [10, 49, 30] for a defect or impurity in semiconductors. Such single coordinate defect model, though can give rise to very slow relaxation of photoresistivity [10, 49, 30], is unlikely to lead to collective aging phenomena seen in our experiment.

For our model, we thus assume a collective glassy background that gives rise to a two-step relaxation, Cgl(t)=xμ(t)xμ(0)=Aexp(t/τs)+Bexp[(t/τα)β]subscript𝐶𝑔𝑙𝑡delimited-⟨⟩subscript𝑥𝜇𝑡subscript𝑥𝜇0𝐴𝑡subscript𝜏𝑠𝐵superscript𝑡subscript𝜏𝛼𝛽C_{gl}(t)=\langle x_{\mu}(t)x_{\mu}(0)\rangle=A\exp(-t/\tau_{s})+B\exp[-(t/% \tau_{\alpha})^{\beta}]italic_C start_POSTSUBSCRIPT italic_g italic_l end_POSTSUBSCRIPT ( italic_t ) = ⟨ italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ( italic_t ) italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ( 0 ) ⟩ = italic_A roman_exp ( start_ARG - italic_t / italic_τ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_ARG ) + italic_B roman_exp [ - ( italic_t / italic_τ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT ] as a function of time t𝑡titalic_t, for the dynamical correlation of xμsubscript𝑥𝜇x_{\mu}italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT at temperature T𝑇Titalic_T. The relaxation consists of a short-time exponential decay with the time scale τssubscript𝜏𝑠\tau_{s}italic_τ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT and a long-time stretched exponential α𝛼\alphaitalic_α-relaxation with time scale τα(T)subscript𝜏𝛼𝑇\tau_{\alpha}(T)italic_τ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_T ) and stretching exponent β𝛽\betaitalic_β [53, 54]. We assume that τssubscript𝜏𝑠\tau_{s}italic_τ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT is temperature independent, whereas ταsubscript𝜏𝛼\tau_{\alpha}italic_τ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT increases with decreasing temperature. We also vary the coefficients A(T)𝐴𝑇A(T)italic_A ( italic_T ) and B(T)𝐵𝑇B(T)italic_B ( italic_T ) with T𝑇Titalic_T such that the relative strength of the stretched exponential part increases at lower temperatures (Supplementary Note 13). In the above form of Cgl(t)subscript𝐶𝑔𝑙𝑡C_{gl}(t)italic_C start_POSTSUBSCRIPT italic_g italic_l end_POSTSUBSCRIPT ( italic_t ), we neglect the β𝛽\betaitalic_β relaxation  [54] which leads to a power-law decay of Cgl(t)subscript𝐶𝑔𝑙𝑡C_{gl}(t)italic_C start_POSTSUBSCRIPT italic_g italic_l end_POSTSUBSCRIPT ( italic_t ) in the plateau region (τstταmuch-less-thansubscript𝜏𝑠𝑡much-less-thansubscript𝜏𝛼\tau_{s}\ll t\ll\tau_{\alpha}italic_τ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ≪ italic_t ≪ italic_τ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT), after the microscopic relaxation and before the onset of the α𝛼\alphaitalic_α-relaxation. The β𝛽\betaitalic_β relaxation is only expected to modify finer details of electronic relaxation in Fig.5.

The glassy PNRs lead to a transition between the conduction band and impurity state via a coupling Helgl=iαμ(Viαμcifα+h.c)xμH_{el-gl}=\sum_{i\alpha\mu}(V_{i\alpha\mu}c_{i}^{\dagger}f_{\alpha}+\mathrm{h.% c})x_{\mu}italic_H start_POSTSUBSCRIPT italic_e italic_l - italic_g italic_l end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_i italic_α italic_μ end_POSTSUBSCRIPT ( italic_V start_POSTSUBSCRIPT italic_i italic_α italic_μ end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_f start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT + roman_h . roman_c ) italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT, where we take Viαμsubscript𝑉𝑖𝛼𝜇V_{i\alpha\mu}italic_V start_POSTSUBSCRIPT italic_i italic_α italic_μ end_POSTSUBSCRIPT as Gaussian complex random numbers to keep the model solvable. The coupling can arise either due to direct coupling of the electric field of the PNR to the electrons, or indirectly via coupling between PNR and electrons mediated by phonons, e.g., the soft polar optical phonon mode.

To gain an analytical understanding of the electron-hole recombination dynamics we consider a limit where there are no backactions of the conduction electrons on the impurity electrons and the glass (Supplementary Note 13).

The description of the relaxation of the photo-excited electrons requires the consideration of the out-of-equilibrium quantum dynamics, which is beyond the scope of this paper. Instead, we consider an equilibrium dynamical correlation, namely the (connected) density-density correlation function Cel(t)=ni(t)ni(0)ni(0)2subscript𝐶𝑒𝑙𝑡delimited-⟨⟩subscript𝑛𝑖𝑡subscript𝑛𝑖0superscriptdelimited-⟨⟩subscript𝑛𝑖02C_{el}(t)=\langle n_{i}(t)n_{i}(0)\rangle-\langle n_{i}(0)\rangle^{2}italic_C start_POSTSUBSCRIPT italic_e italic_l end_POSTSUBSCRIPT ( italic_t ) = ⟨ italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( 0 ) ⟩ - ⟨ italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( 0 ) ⟩ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT as a function of time t𝑡titalic_t for the density ni=cicisubscript𝑛𝑖superscriptsubscript𝑐𝑖subscript𝑐𝑖n_{i}=c_{i}^{\dagger}c_{i}italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT of the conduction electron. In our experiment, the resistivity decreases due to the increase of the density of conduction electrons via photo-excitations and relaxes through the relaxation of these excess carriers to the impurity states. The correlation function Cel(t)subscript𝐶𝑒𝑙𝑡C_{el}(t)italic_C start_POSTSUBSCRIPT italic_e italic_l end_POSTSUBSCRIPT ( italic_t ) characterizes a similar relaxation process, albeit close to the thermal equilibrium. The correlation function Cel(t)subscript𝐶𝑒𝑙𝑡C_{el}(t)italic_C start_POSTSUBSCRIPT italic_e italic_l end_POSTSUBSCRIPT ( italic_t ) can be computed exactly at a temperature T𝑇Titalic_T in the toy model (Eq. (1)).

In the limit of no backaction of the conduction electrons, the electronic correlation function can be written as a convolution of the spectral function of the glass (Supplementary Note 13). As a result, the glass spectral function, which contains the information of the multiple time-scales and their non-trivial temperature dependence, may directly induce glassiness in the electronic relaxation.

To verify the above scenario, we numerically compute Cel(t)subscript𝐶𝑒𝑙𝑡C_{el}(t)italic_C start_POSTSUBSCRIPT italic_e italic_l end_POSTSUBSCRIPT ( italic_t ) for two-step glass correlation functions Cgl(t)subscript𝐶𝑔𝑙𝑡C_{gl}(t)italic_C start_POSTSUBSCRIPT italic_g italic_l end_POSTSUBSCRIPT ( italic_t ), shown in Fig. 5(b) (upper panel) for three temperatures T=0.5,0.3,0.1Δformulae-sequence𝑇0.50.3much-less-than0.1ΔT=0.5,0.3,0.1\ll\Deltaitalic_T = 0.5 , 0.3 , 0.1 ≪ roman_Δ, where the temperature dependence is parameterized by B(T)𝐵𝑇B(T)italic_B ( italic_T ) (A=1B𝐴1𝐵A=1-Bitalic_A = 1 - italic_B) and τα(T)subscript𝜏𝛼𝑇\tau_{\alpha}(T)italic_τ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_T ) (Fig. 5(b) (lower panel)). We take a temperature independent exponent β=0.5𝛽0.5\beta=0.5italic_β = 0.5 and τα(T)T2.8similar-tosubscript𝜏𝛼𝑇superscript𝑇2.8\tau_{\alpha}(T)\sim T^{-2.8}italic_τ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_T ) ∼ italic_T start_POSTSUPERSCRIPT - 2.8 end_POSTSUPERSCRIPT, consistent with our experimental results (Fig. 2c). Here τs1superscriptsubscript𝜏𝑠1\tau_{s}^{-1}italic_τ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT (τssubscript𝜏𝑠\tau_{s}italic_τ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT) is set as the unit of energy (time) (=1Planck-constant-over-2-pi1\hbar=1roman_ℏ = 1). The calculated Cel(t)subscript𝐶𝑒𝑙𝑡C_{el}(t)italic_C start_POSTSUBSCRIPT italic_e italic_l end_POSTSUBSCRIPT ( italic_t ) is plotted in Fig. 5(c) for a flat conduction band (W=0𝑊0W=0italic_W = 0) (upper panel) and a semicircular conduction band DOS with W=0.01𝑊0.01W=0.01italic_W = 0.01, and electron-glass coupling V=0.3𝑉0.3V=0.3italic_V = 0.3. As shown in Fig. 5(c), we find that for a local glassy bath, whose bandwidth is comparable or larger than the electronic energy scales W𝑊Witalic_W and ΔΔ\Deltaroman_Δ, the complex glassy correlation, namely the two-step relaxation with long and non-trivial temperature-dependent time scale, is also manifested in the electronic relaxation.

Outlook: While we do observe complex relaxations in such a toy model, these are still weak in contrast to the actual experimental results since our calculations only capture glassy two-step electronic relaxation via equilibrium dynamical correlations, whereas the actual experimental electron dynamics take place in a strongly non-equilibrium condition. We expect that a more realistic non-equilibrium theory considering the direct coupling with a real space distribution of PNRs embedded in the sea of conduction electrons would yield a strong glass. This would further demand intricate knowledge about the nature of interactions in the presence of PNRs which is still a subject of debate. Interestingly, these questions form the basis for understanding the nature of conduction in an interesting class of materials known as polar metals [55, 35, 36], and hence we believe that our finding of glassy relaxations in presence of PNRs will be crucial in building the theory of conduction in quantum critical polar metals. Further, the observation of glassy dynamics deep inside the good metallic regime is in sharp contrast with the conventional semiconductors where glassy relaxation ceases to exist just before the insulator- metal transition (IMT) is approached from the insulating side. This raises the question about the envisaged role of glassy freezing of electrons as a precursor to IMT apart from the Anderson and Mott localization [56, 3].

Methods

Sample preparation: Oxygen deficient KTaO3 single crystals were prepared by heating as received 001 oriented pristine KTaO3 substrate (from Princeton Scientific Corp.) in a vacuum sealed quartz tube in presence of titanium wire. For more details we refer to our previous work  [2, 19].

Dielectric measurement: Temperature-dependent dielectric measurement was performed in a close cycle cryostat using an impedance analyzer from Keysight Technology Instruments (Model No. E49908).

Transport measurement: All the transport measurements were carried out in an ARS close cycle cryostat in van der Pauw geometry using a dc delta mode with a Keithley 6221 current source and a Keithley 2182A nanovoltmeter and also using standard low-frequency lock-in technique. Ohmic contacts were realized by ultrasonically bonding aluminum wire or by attaching gold wire with silver paint.

Light set-up: Light of the desired wavelength was passed through the optical window of the close cycle cryostat from ARS. A home-built setup consisting of a diffuser and lens was used to make light fall homogeneously over the sample (see inset of Fig. 2a ). Commercially available light-emitting diodes from Thor Labs were used as a light source. Incident power on the sample was measured with a laser check handheld power meter from coherent (Model No: 54-018).

HAXPES measurement: Near Fermi level and K 2p core level spectra were collected at Hard X-ray Photoelectron Spectroscopy (HAXPES) beamline (P22) of PETRA III, DESY, Hamburg, Germany using a high-resolution Phoibos electron analyzer [57]. Au Fermi level and Au 4f core level spectra collected on a gold foil (mounted on the same sample holder) were used as a reference for making the correction to the measured kinetic energy. The chamber pressure during the measurement was similar-to\sim 10-10 Torr. An open cycle Helium flow cryostat was used to control the sample temperature.

Second harmonic generation measurement: SHG measurements were performed under reflection off the sample at a 45-degree incidence angle. A p𝑝pitalic_p-polarised 800 nm beam from a Spectra-Physics Spirit-NOPA laser was used as the fundamental beam (pulse width: 300 fs, repetition rate: 1 MHz), and was focused on the sample surface. The p𝑝pitalic_p-polarised SHG intensity generated by the sample, was measured by a photo multiplier tube. The sample temperature was controlled by a helium-cooled Janis 300 cryostat installed with a heating element.

Density functional theory: The noncolinear density functional theory calculations were carried out using the QUANTUM ESPRESSO package[58]. In this calculation optimized norm-conserving pseudopotentials [59, 60, 61] were used and for the exchange-correlation functional [62] we have incorporated Perdew, Burke and Ernzerhof generalized gradient approximation (PBE-GGA). For the unit cell, the Brillouin zone was sampled with 8×8×8cross-product8888\crossproduct 8\crossproduct 88 × 8 × 8 k𝑘kitalic_k-points. The wave functions were expanded in plane waves with an energy up to 90 Ry. Since the effect of SOC in KTaO3 is quite remarkable, we have employed full-relativistic pseudopotential for the Ta atom. The structural relaxations were performed until the force on each atom was reduced to 0.07 eV/Åangstrom\mathrm{\SIUnitSymbolAngstrom}roman_Å. The Berry phase calculations are carried out as implemented in the Wannier90 code [32, 64] We have used a 41×41414141\times 4141 × 41 2D k-mesh for the Berry curvature calculations. No significant change in the result is observed on increasing the k-mesh up to 101×101101101101\times 101101 × 101 (see Supplementary Note 8 for further details).

Data availability

The authors declare that the data supporting the findings of this study are available within the main text and its Supplementary Information. Other relevant data are available from the corresponding author upon request.

References

Acknowledgements

The work is funded by SERB, India, by a core research grant CRG/2022/001906 to SM. Portions of this research were carried out at the light source PETRA III DESY, a member of the Helmholtz Association (HGF). We would like to thank Dr. Anuradha Bhogra and Dr. Thiago Peixoto for their assistance at beamline P22. Financial support by the Department of Science & Technology (Government of India) provided within the framework of the India@DESY collaboration is gratefully acknowledged. S.H. and V.G. acknowledge support from the US Department of Energy under grant no. DE-SC-0012375 for temperature-dependent second-harmonic generation measurements. SB acknowledges support from SERB (CRG/2022/001062), DST, India. SKG and MJ gratefully acknowledge Supercomputer Education and Research Centre, IISc for providing computational facilities SAHASRAT and PARAM-PRAVEGA. SKG acknowledges DST-Inspire fellowship (IF170557). SKO acknowledges the wire bonding facility at the Department of Physics, IISc Bangalore and thanks Shivam Nigam for the experimental assistance. SKO and SM thank Professor D. D. Sarma for giving them access to the quartz tube sealing and dielectric measurement setup in his lab.

Author contribution

SM conceived and supervised the project. SKO, SM, and SK came up with all experimental plans to confirm the glassy behavior. SKO, SH, JM carried out transport measurements. PM helped in making the photoconductivity setup. SH performed SHG measurements under the supervision of VG. SKO, PM, AG and CS carried out HAXPES measurements. SKO performed all the analysis of transport measurements and HAXPES. SKG and MJ performed DFT calculations. S Bera, SK and SB provided phenomenological model calculations. SKO, S Bera, SB, and SM wrote the manuscript with inputs from other authors. All authors discussed the results and participated in finalizing the manuscript.

Competing interests

The authors declare no competing interests.

Refer to caption
Figure 1: Electron glass, quantum paraelectricity and temperature dependence of electron’s mean free path in KTaO3-δ a. A schematic to describe the concept of electron glass in a disordered insulator with highly localized electronic states. Here the solid colored cones represent the random disorder potential and yellow-filled circles represent electrons trapped in them. Disorder tries to make the distribution of doped carriers random, however long-range Coulomb interactions (U1subscript𝑈1U_{1}italic_U start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, U2subscript𝑈2U_{2}italic_U start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT) try to make the distribution homogeneous leading to electronic frustration. b. Temperature dependent dielectric constant (ϵitalic-ϵ\epsilonitalic_ϵ) of pristine (001) oriented KTaO3 single crystal recorded at an AC frequency of 10 kHz, taken from our previous paper [19]. We further note that the value of ε𝜀\varepsilonitalic_ε for our sample appears to be slightly lower than the reported values [10]. We attribute this difference to the difference in the sample preparation process, which may add slight oxygen vacancies, even in the pristine, as-received crystal [11]. Sitting on the verge of quantum critical point, KTaO3 is long known for its peculiar dielectric properties [12] wherein the onset of quantum fluctuation leads to a marked departure from classical paraelectric behavior below 35 K leading to saturation of ε𝜀\varepsilonitalic_ε at low temperature. This is further evident from the modified Curie–Weiss fitting (ε𝜀\varepsilonitalic_ε = ε0subscript𝜀0\varepsilon_{0}italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT+ CWTθCWsubscript𝐶𝑊𝑇subscript𝜃𝐶𝑊\frac{C_{W}}{T-\theta_{CW}}divide start_ARG italic_C start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT end_ARG start_ARG italic_T - italic_θ start_POSTSUBSCRIPT italic_C italic_W end_POSTSUBSCRIPT end_ARG, ε0subscript𝜀0\varepsilon_{0}italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is temperature independent component, CWsubscript𝐶𝑊C_{W}italic_C start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT is Curie-Weiss constant and θCWsubscript𝜃𝐶𝑊\theta_{CW}italic_θ start_POSTSUBSCRIPT italic_C italic_W end_POSTSUBSCRIPT is Curie-Weiss temperature, respectively) shown in the inset of panel b. θCWsubscript𝜃𝐶𝑊\theta_{CW}italic_θ start_POSTSUBSCRIPT italic_C italic_W end_POSTSUBSCRIPT obtained from the fitting is found to be 37 K. c. Unit cell of pristine KTaO3 along with the transverse optical soft mode. d. Temperature-dependent electron’s mean free path lesubscript𝑙𝑒l_{e}italic_l start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT for oxygen-deficient KTO calculated within the Drude-Boltzmann picture. Blue and orange curves correspond to degenerate and non-degenerate cases respectively. ΛΛ\Lambdaroman_Λ denotes the thermal de Broglie wavelength. A dotted vertical line marks the temperature above which lesubscript𝑙𝑒l_{e}italic_l start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT becomes shorter than the inverse of Fermi wave-vector (kFsubscript𝑘𝐹k_{F}italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT) and the sample crosses over from good metal to a bad metal phase. For details of the calculations, we refer to Supplementary Note 1 which significantly overlaps with our earlier work [2]. Source data are provided as a Source Data file.
Refer to caption
Figure 2: HAXPES and transport measurements under light a. Near Fermi level electronic states measured at 40 K using hard X-ray photoelectron spectroscopy on the KTaO3 sample. Binding energy was corrected by taking Au Fermi level as a reference. In the figure, Au Fermi level intensity has been reduced by thousand times and shifted rightward for comparison. Inset shows the setup for transport measurement under light illumination (see methods for more details). Green arrow shows the incoming photon which excites trapped electrons to the conduction band (shown with a slate-gray color). A slate-gray shade has been used to highlight the presence of defect states. b. Temporal evolution of resistance under green light illumination (λ𝜆\lambdaitalic_λ = 527 nm, power = 145 μ𝜇\muitalic_μ watt) for 30 minutes measured at several fixed temperatures. After 30 minutes, resistance relaxation was observed in dark for the next 1.5 hours. For comparative analysis, change in resistance has been converted into relative percentage change (ΔΔ\Deltaroman_ΔR𝑅Ritalic_R/R0subscript𝑅0R_{0}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT)×cross-product\crossproduct×100. Additional measurements with red light have been shown in Supplementary Note 5 and Supplementary Figure 5. c. Temperature dependence of stretching exponent (β𝛽\betaitalic_β) and relaxation time (τ𝜏\tauitalic_τ) obtained from fitting of resistance relaxation in light off stage with a stretched exponential function. A slate-gray shade has been used to highlight the distinct behavior of β𝛽\betaitalic_β and τ𝜏\tauitalic_τ below 35 K. The error bar in β𝛽\betaitalic_β and τ𝜏\tauitalic_τ has been estimated from the variation of corresponding parameters which results in similar fitting. We further emphasize that, while the value β𝛽\betaitalic_β has been obtained by fitting the data for 1.5 hours, the same value is obtained even when the data is fitted for a longer period of up to 24 hours Supplementary Note 6 and Supplementary Figure 6). Source data are provided as a Source Data file.
Refer to caption
Figure 3: Aging phenomena a. Relative percentage change in resistance (ΔΔ\Deltaroman_ΔR𝑅Ritalic_R/R0subscript𝑅0R_{0}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT)×cross-product\crossproduct×100 measured at 15 K for different t𝑡titalic_tillill{}_{\text{ill}}start_FLOATSUBSCRIPT ill end_FLOATSUBSCRIPT ranging from 0.5 min to 30 min. After turning off the light, resistance relaxation was observed for the next 1.5 hours in every case. b. Temporal evolution of resistance relaxation in off-stage for different values of t𝑡titalic_tillill{}_{\text{ill}}start_FLOATSUBSCRIPT ill end_FLOATSUBSCRIPT. Since the amount of do** depends on the duration of light illumination, the change in resistance in the off-stage (ΔΔ\Deltaroman_ΔR𝑅Ritalic_Roffoff{}_{\text{off}}start_FLOATSUBSCRIPT off end_FLOATSUBSCRIPT) has been normalized with a total drop in resistance (ΔΔ\Deltaroman_ΔR𝑅Ritalic_Rillill{}_{\text{ill}}start_FLOATSUBSCRIPT ill end_FLOATSUBSCRIPT) at the end of illumination [21]. c. Upon re-scaling the time axis with t𝑡titalic_tillill{}_{\text{ill}}start_FLOATSUBSCRIPT ill end_FLOATSUBSCRIPT, all the curves in off-stage (except t𝑡titalic_tillill{}_{\text{ill}}start_FLOATSUBSCRIPT ill end_FLOATSUBSCRIPT = 30 min) are found to fall on a universal curve. Inset shows the linear relationship between τ𝜏\tauitalic_τ and t𝑡titalic_tillill{}_{\text{ill}}start_FLOATSUBSCRIPT ill end_FLOATSUBSCRIPT which is the defining criteria for simple/full aging. d. Simple aging observed for another sample at 15 K. Inset shows one representative resistance relaxation data at 15 K. The sheet resistance of this sample is approximately 22 kΩΩ\Omegaroman_Ω/sq. (at room temperature) which is much larger than the one discussed throughout this manuscript which has around 200 ΩΩ\Omegaroman_Ω/sq. emphasizing that the oxygen vacancy concentration for this sample is much lower. Source data are provided as a Source Data file.
Refer to caption
Figure 4: Polar nano regions, second harmonic generation measurement and density functional theory a. A schematic depicting real space random distribution of polar nano regions in a highly polarizable host lattice. b. A schematic of the experimental setup for measuring optical second harmonic generation signal in reflection geometry (for more details see methods section). c. Temperature-dependent second harmonic generation intensity measured on KTaO3 sample. A slate-gray shade has been used to highlight the distinct second harmonic generation signal below 35 K. d. A portion of the relaxed structure of supercell of size 2×cross-product\crossproduct×2×cross-product\crossproduct×2 of KTaO3 with Ta off-centering along the (100) and (110) direction around K vacancy. For more details, see the methods section, Supplementary Note 12, and Supplementary Figure 12. e. Plot of Berry curvature Ωxy(𝐤)subscriptΩ𝑥𝑦𝐤\Omega_{xy}(\mathbf{k})roman_Ω start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPT ( bold_k ) (shown in color-map) for kz=0subscript𝑘𝑧0k_{z}=0italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0 and bands (solid contours) intersecting the Fermi surface for the system having an antisite-like Ta defect along (110) direction. f. Same as e. for the system with an antisite-like Ta defect along (100) direction. The inset shows the spin-orbit induced avoided crossing of bands yielding large contributions to the Berry curvature. We get average polarization of 5.8 μC/cm2𝜇𝐶𝑐superscript𝑚2\mu C/cm^{2}italic_μ italic_C / italic_c italic_m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and 11.1 μC/cm2𝜇𝐶𝑐superscript𝑚2\mu C/cm^{2}italic_μ italic_C / italic_c italic_m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT for Ta antisite-like defect along (110) and (100) respectively for the supercell of size 2×2×2cross-product2222\crossproduct 2\crossproduct 22 × 2 × 2. We also note that, we do not observe any polarization in pristine KTaO3. Source data are provided as a Source Data file.
Refer to caption
Figure 5: Theoretical calculations a. A schematic of band diagram considered for computing inter band transitions when the conduction electron is coupled to glassy background from randomly oriented polar nano regions. b. (Upper panel) The glass correlation function Cgl(t)subscript𝐶𝑔𝑙𝑡C_{gl}(t)italic_C start_POSTSUBSCRIPT italic_g italic_l end_POSTSUBSCRIPT ( italic_t ) vs. t𝑡titalic_t for three temperatures for τs(T)=1subscript𝜏𝑠𝑇1\tau_{s}(T)=1italic_τ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_T ) = 1, b=0.5𝑏0.5b=0.5italic_b = 0.5. (Lower panel) Plot of τα(T)T2.8similar-tosubscript𝜏𝛼𝑇superscript𝑇2.8\tau_{\alpha}(T)\sim T^{-2.8}italic_τ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_T ) ∼ italic_T start_POSTSUPERSCRIPT - 2.8 end_POSTSUPERSCRIPT vs. T𝑇Titalic_T (left y axis, in units of τssubscript𝜏𝑠\tau_{s}italic_τ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT) and B(T)𝐵𝑇B(T)italic_B ( italic_T ) (right y axis) vs. T𝑇Titalic_T (A=1B𝐴1𝐵A=1-Bitalic_A = 1 - italic_B) for three temperatures. c. (upper panel) The density-density correlation function of conduction electron Cel(t)subscript𝐶𝑒𝑙𝑡C_{el}(t)italic_C start_POSTSUBSCRIPT italic_e italic_l end_POSTSUBSCRIPT ( italic_t ) vs. t𝑡titalic_t for three temperatures for flat (W=0𝑊0W=0italic_W = 0) conduction band. (lower panel) Cel(t)subscript𝐶𝑒𝑙𝑡C_{el}(t)italic_C start_POSTSUBSCRIPT italic_e italic_l end_POSTSUBSCRIPT ( italic_t ) vs. t𝑡titalic_t for three temperatures for semicircular conduction band with W=0.01𝑊0.01W=0.01italic_W = 0.01. Source data are provided as a Source Data file.

Supplementary Information

List of contents

Note 1. Calculation of electron’s mean free path.

Note 2. Sheet resistance vs temperature plot.

Note 3. Stretched exponential behavior of resistance relaxation in the light off stage.

Note 4. Power law behavior of persistence photo-resistance relaxation time (τ𝜏\tauitalic_τ).

Note 5. Experiment with red light.

Note 6. Long-time relaxation of persistence photo-resistance.

Note 7. Deviation of persistence photo-resistance relaxation time (τ𝜏\tauitalic_τ) from activated behavior at low temperature.

Note 8. Dielectric loss in pristine KTaO3.

Note 9. Correspondence between the appearance of polar nano regions and photo-do** effect.

Note 10. Raman measurement on a metallic oxygen deficient KTO.

Note 11. Presence of potassium vacancy.

Note 12. Berry phase calculations.

Note 13. Toy model of complex inter-band electronic relaxation mediated by a glassy bath.

Supplementary Note 1: Calculation of electron’s mean free path

The Fermi surface of the electron doped KTO [1] comprises of three ellipsoids of revolution centered at ΓΓ\Gammaroman_Γ point (see Supplementary Fig. 1a) with major axis kF,maxsubscript𝑘𝐹𝑚𝑎𝑥k_{F,max}italic_k start_POSTSUBSCRIPT italic_F , italic_m italic_a italic_x end_POSTSUBSCRIPT and the minor axis kF,minsubscript𝑘𝐹𝑚𝑖𝑛{k_{F,min}}italic_k start_POSTSUBSCRIPT italic_F , italic_m italic_i italic_n end_POSTSUBSCRIPT transverse to it. In the presence of magnetic field along [001] crystallographic axis, electrons traverse around the extremal orbits in the momentum space as shown in the Supplementary Fig. 1b, leading to the observation of SdH oscillations in magnetoresistance. Please refer to our earlier work [2] for SdH oscillation data on the same sample which is being investigated in the current study. These oscillations are periodic in 1/B𝐵Bitalic_B whose frequency F𝐹Fitalic_FSdHSdH{}_{\textnormal{SdH}}start_FLOATSUBSCRIPT SdH end_FLOATSUBSCRIPT is related to the area of extremal orbits Aext as given below [3]

Refer to caption
Supplementary Figure 1: a. Ellipsoidal Fermi surfaces of electron doped KTO. Here kF,maxsubscript𝑘𝐹𝑚𝑎𝑥k_{F,max}italic_k start_POSTSUBSCRIPT italic_F , italic_m italic_a italic_x end_POSTSUBSCRIPT was taken to be 1.541 times kF,minsubscript𝑘𝐹𝑚𝑖𝑛k_{F,min}italic_k start_POSTSUBSCRIPT italic_F , italic_m italic_i italic_n end_POSTSUBSCRIPT for plotting. b. Cross section of the Fermi surface with the plane kZsubscript𝑘𝑍k_{Z}italic_k start_POSTSUBSCRIPT italic_Z end_POSTSUBSCRIPT=0
FSdH=Aext/2πesubscript𝐹SdHPlanck-constant-over-2-pisubscript𝐴𝑒𝑥𝑡2𝜋𝑒F_{\textnormal{SdH}}=\hbar A_{ext}/{2\pi e}italic_F start_POSTSUBSCRIPT SdH end_POSTSUBSCRIPT = roman_ℏ italic_A start_POSTSUBSCRIPT italic_e italic_x italic_t end_POSTSUBSCRIPT / 2 italic_π italic_e (2)

The extremal orbits corresponding to the two ellipsoids directed along kxsubscript𝑘𝑥k_{x}italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT and kysubscript𝑘𝑦k_{y}italic_k start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT axis are ellipses with area equal to π𝜋\piitalic_πkF,minsubscript𝑘𝐹𝑚𝑖𝑛k_{F,min}italic_k start_POSTSUBSCRIPT italic_F , italic_m italic_i italic_n end_POSTSUBSCRIPTkF,maxsubscript𝑘𝐹𝑚𝑎𝑥{k_{F,max}}italic_k start_POSTSUBSCRIPT italic_F , italic_m italic_a italic_x end_POSTSUBSCRIPT, whereas the extremal orbit corresponding to the ellipsoid along the kzsubscript𝑘𝑧k_{z}italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT direction is circular with an area πkF,min2𝜋subscriptsuperscript𝑘2𝐹𝑚𝑖𝑛{\pi k^{2}_{F,min}}italic_π italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_F , italic_m italic_i italic_n end_POSTSUBSCRIPT (see Supplementary Fig. 1b). Further, from the literature it is known that [4, 1]

kF,max/kF,min=1.541subscript𝑘𝐹𝑚𝑎𝑥subscript𝑘𝐹𝑚𝑖𝑛1.541k_{F,max}/k_{F,min}=1.541italic_k start_POSTSUBSCRIPT italic_F , italic_m italic_a italic_x end_POSTSUBSCRIPT / italic_k start_POSTSUBSCRIPT italic_F , italic_m italic_i italic_n end_POSTSUBSCRIPT = 1.541 (3)

for electron doped KTO, which would further imply that there are more electrons around elliptical orbit than those orbiting around circular orbit [5]. From above fact, we assume that the main frequency of SdH oscillation comes from electrons orbiting around the elliptical orbit and therefore Aextsubscript𝐴𝑒𝑥𝑡A_{ext}italic_A start_POSTSUBSCRIPT italic_e italic_x italic_t end_POSTSUBSCRIPT in Supplementary Eq. (1) is taken to be the area of the ellipse π𝜋\piitalic_πkF,minsubscript𝑘𝐹𝑚𝑖𝑛k_{F,min}italic_k start_POSTSUBSCRIPT italic_F , italic_m italic_i italic_n end_POSTSUBSCRIPTkF,maxsubscript𝑘𝐹𝑚𝑎𝑥{k_{F,max}}italic_k start_POSTSUBSCRIPT italic_F , italic_m italic_a italic_x end_POSTSUBSCRIPT. From the SdH analysis, F𝐹Fitalic_FSdHSdH{}_{\textnormal{SdH}}start_FLOATSUBSCRIPT SdH end_FLOATSUBSCRIPT for our oxygen deficient KTO is found to be 12.8 Tesla [2]. Putting this value in Supplementary Eq. (1) and using the Supplementary Eq. (2), we first calculate kF,maxsubscript𝑘𝐹𝑚𝑎𝑥k_{F,max}italic_k start_POSTSUBSCRIPT italic_F , italic_m italic_a italic_x end_POSTSUBSCRIPT and kF,minsubscript𝑘𝐹𝑚𝑖𝑛k_{F,min}italic_k start_POSTSUBSCRIPT italic_F , italic_m italic_i italic_n end_POSTSUBSCRIPT individually and then 3D carrier density is determined. For this we note that the total volume occupied by three ellipsoid is V=k3×(4π/3)k2F,minkF,max{{}_{k}=3\times(4\pi/3){k^{2}}_{F,min}k_{F,max}}start_FLOATSUBSCRIPT italic_k end_FLOATSUBSCRIPT = 3 × ( 4 italic_π / 3 ) italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_F , italic_m italic_i italic_n end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT italic_F , italic_m italic_a italic_x end_POSTSUBSCRIPT from which the carrier concentration n𝑛nitalic_n can be calculated as [5]

n=(k2F,minkF,max)/π2𝑛subscriptsuperscript𝑘2𝐹𝑚𝑖𝑛subscript𝑘𝐹𝑚𝑎𝑥superscript𝜋2n={({k^{2}}_{F,min}k_{F,max})/{\pi^{2}}}italic_n = ( italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_F , italic_m italic_i italic_n end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT italic_F , italic_m italic_a italic_x end_POSTSUBSCRIPT ) / italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (4)

Putting the value of kF,maxsubscript𝑘𝐹𝑚𝑎𝑥k_{F,max}italic_k start_POSTSUBSCRIPT italic_F , italic_m italic_a italic_x end_POSTSUBSCRIPT and kF,minsubscript𝑘𝐹𝑚𝑖𝑛k_{F,min}italic_k start_POSTSUBSCRIPT italic_F , italic_m italic_i italic_n end_POSTSUBSCRIPT in the above Supplementary Eq. (3), the 3D carrier density n𝑛nitalic_n for the oxygen deficient KTO is found to be 5.7×\times×1017 cm-3. Having obtained the 3D carrier density, we next compute the temperature dependent electron’s mean free path following the well established approach described in the papers [6, 7].

In the Drude-Boltzmann picture, 3 dimensional conductivity (σ𝜎\sigmaitalic_σ) of a metal with spherical Fermi surface is given by

σ=ne2τ/m=e2(kF)2le/3π2𝜎𝑛superscript𝑒2𝜏superscript𝑚superscript𝑒2superscriptsubscript𝑘𝐹2subscript𝑙𝑒3superscript𝜋2Planck-constant-over-2-pi\sigma=ne^{2}\tau/m^{*}=e^{2}(k_{F})^{2}l_{e}/3\pi^{2}\hbaritalic_σ = italic_n italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_τ / italic_m start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_l start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT / 3 italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_ℏ (5)

where n𝑛nitalic_n is the 3D carrier density given by n𝑛nitalic_n=kFsubscript𝑘𝐹k_{F}italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT3/3π𝜋\piitalic_π2 (kFsubscript𝑘𝐹k_{F}italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT is the Fermi wavevector), and τ𝜏\tauitalic_τ is scattering time constant given by τ𝜏\tauitalic_τ=lesubscript𝑙𝑒l_{e}italic_l start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPTmsuperscript𝑚m^{*}italic_m start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT/Planck-constant-over-2-pi\hbarroman_ℏkFsubscript𝑘𝐹k_{F}italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT where m𝑚mitalic_m is the effective mass of electrons and lesubscript𝑙𝑒l_{e}italic_l start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT is the electron’s mean free path [8]. Rearranging the Supplementary Eq. (4), the expression for lesubscript𝑙𝑒l_{e}italic_l start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT can be written as

le=3π2/ρ(kF)2e2subscript𝑙𝑒3superscript𝜋2Planck-constant-over-2-pi𝜌superscriptsubscript𝑘𝐹2superscript𝑒2l_{e}=3\pi^{2}\hbar/\rho(k_{F})^{2}e^{2}italic_l start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = 3 italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_ℏ / italic_ρ ( italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (6)

where ρ𝜌\rhoitalic_ρ=1/σ𝜎\sigmaitalic_σ is the resistivity. Since all the measurements in the present study are performed in Van der Pauw geometry, the temperature dependent ρ𝜌\rhoitalic_ρ is obtained from the formula

ρ=(π/ln2)Rt𝜌𝜋𝑙𝑛2𝑅𝑡\rho=(\pi/ln2)Rtitalic_ρ = ( italic_π / italic_l italic_n 2 ) italic_R italic_t (7)

where R𝑅Ritalic_R is the measured temperature dependent resistance and t𝑡titalic_t is the thickness of the conducting region. In the present case, t𝑡titalic_t is calculated by equating the 3D carrier density obtained from the SdH analysis with the sheet carrier density nSsubscript𝑛𝑆n_{S}italic_n start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT obtained from the Hall measurement divided by the thickness [n𝑛nitalic_n=nSsubscript𝑛𝑆n_{S}italic_n start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT/t𝑡titalic_t, see the reference  [3]]. Apart from the ρ𝜌\rhoitalic_ρ, kFsubscript𝑘𝐹k_{F}italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT is the another input parameter in Supplementary Eq. (5) for calculating lesubscript𝑙𝑒l_{e}italic_l start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT. Since the Supplementary Eq. (5) holds only for the spherical Fermi surface, we make the following approximation. Since the Fermi surface of KTO is only moderately anisotropic at such dilute carrier density, for simplification, the value of kFsubscript𝑘𝐹k_{F}italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT has been calculated by effectively map** the ellipsoidal Fermi surface onto a spherical Fermi surface wherein we equate the total volume of the three ellipsoids with a single spherical Fermi surface of radius kFsubscript𝑘𝐹k_{F}italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT. The value of kFsubscript𝑘𝐹k_{F}italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT is then estimated from the relation kFsubscript𝑘𝐹k_{F}italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT=(3π𝜋\piitalic_π2n𝑛nitalic_n)1/3 [8]. Once we obtain kFsubscript𝑘𝐹k_{F}italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT, the temperature dependent ρ𝜌\rhoitalic_ρ obtained from the Supplementary Eq. (6) along with kFsubscript𝑘𝐹k_{F}italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT is plugged in the Supplementary Eq. (5) and temperature dependent electron’s mean free path is calculated.

lesubscript𝑙𝑒l_{e}italic_l start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT calculated using this approach for the oxygen deficient KTO has been shown by a blue curve in the Fig. 1d of the main text. As evident from the plot, above similar-to\sim130 K, lesubscript𝑙𝑒l_{e}italic_l start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT becomes shorter than the inverse of kFsubscript𝑘𝐹k_{F}italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT (kFsubscript𝑘𝐹k_{F}italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPTlesubscript𝑙𝑒l_{e}italic_l start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT >>> 1) and the system enters into a bad metal phase. A further correction to the calculated lesubscript𝑙𝑒l_{e}italic_l start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT is required above the Fermion degeneracy temperature above which the thermal de Broglie wavelength (given by ΛΛ\Lambdaroman_Λ=hhitalic_h/2πmkBT2𝜋superscript𝑚subscript𝑘𝐵𝑇\sqrt{2\pi m^{*}k_{B}T}square-root start_ARG 2 italic_π italic_m start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG) becomes larger than the inter electron separation given by n𝑛nitalic_n-1/3. Above this temperature, electrons become non-degenerate and the modified lesubscript𝑙𝑒l_{e}italic_l start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT is given by lesubscriptsuperscript𝑙𝑒l^{\prime}_{e}italic_l start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPTsimilar-to\sim lesubscript𝑙𝑒l_{e}italic_l start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT n𝑛nitalic_n-1/3/ΛΛ\Lambdaroman_Λ [6, 7]. For calculation of lesubscriptsuperscript𝑙𝑒l^{\prime}_{e}italic_l start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT, one requires an estimation of msuperscript𝑚m^{*}italic_m start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT. In the present work we estimate the msuperscript𝑚m^{*}italic_m start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT by analyzing the temperature dependent SdH oscillations and is found to be 0.56 mesubscript𝑚𝑒m_{e}italic_m start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT (please refer to the section B of the supplemental material of our earlier paper [2] for more details). The orange curve in Fig. 1d of the main text shows the calculated lesubscript𝑙𝑒l_{e}italic_l start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT for the non-degenerate case. As evident, this exercise only shifts the curve to a little higher temperature and crossover to bad metal phase occurs around 145 K, however, the conclusions remain the same. In the current work, observation of glassy dynamics is inherently constrained to the temperatures which is much lower than the crossover temperature to bad metal phase and hence for all practical purposes our electron doped KTO system is in good metal regime with well defined scattering.

Supplementary Note 2: Sheet resistance vs temperature plot.

As depicted in Fig. 2b of the main text, shining light above 150 K does not have any noticeable effect. This observation is further supported by the plot shown in Supplementary Fig. 2, where we compare the RSsubscript𝑅𝑆R_{S}italic_R start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT vs T𝑇Titalic_T curve obtained in the dark (taken in the cooling run) with the measurement conducted during the heating run after exposing the sample to green light (power= 145 μ𝜇\muitalic_μ Watt) for 30 minutes at 10 K. It is evident from the plot that the two curves converge around 150 K, indicating a lack of any significant photo-do** above 150 K.

Refer to caption
Supplementary Figure 2: Temperature dependent sheet resistance (R𝑅Ritalic_RSS{}_{\text{S}}start_FLOATSUBSCRIPT S end_FLOATSUBSCRIPT) of oxygen-deficient KTaO3 sample (n𝑛nitalic_n=5.7×cross-product\crossproduct×1017 cm-3) taken in a cooling run in dark (sky blue curve). The slate-gray color curve shows data taken in a heating run after shining light for 30 minutes at 10 K.

Supplementary Note 3: Stretched exponential behavior of resistance relaxation in the light off stage.

The panels a-f in the Supplementary Fig. 3 show the temporal evolution of resistance before and after turning off the light along with the fitting in light off stage with an stretched exponential function (exp(-(t𝑡titalic_t/τ𝜏\tauitalic_τ)β) where τ𝜏\tauitalic_τ is the relaxation time and β𝛽\betaitalic_β (stretching exponent) <<< 1). As evident, this function provides an excellent fit to experimental data at range of temperatures.

Refer to caption
Supplementary Figure 3: Fitting of resistance relaxation in the light off stage with stretched exponential function at several fixed temperatures.

Supplementary Note 4: Power law behavior of persistence photo-resistance relaxation time (τ𝜏\tauitalic_τ).

Supplementary Fig. 4 shows log𝑙𝑜𝑔logitalic_l italic_o italic_g-log𝑙𝑜𝑔logitalic_l italic_o italic_g plot of relaxation time (τ𝜏\tauitalic_τ) as a function of temperature. As evident, the plot looks linear in wide range of temperatures belo 50 K signifying a power law dependence (τTαsimilar-to𝜏superscript𝑇𝛼\tau\sim T^{-\alpha}italic_τ ∼ italic_T start_POSTSUPERSCRIPT - italic_α end_POSTSUPERSCRIPT). The value of alpha obtained from fitting is found to be 2.8. We also emphasize that such power-law divergence of relaxation time has been also discussed theoretically in context of α𝛼\alphaitalic_α relaxation in glasses [9].

Refer to caption
Supplementary Figure 4: Power law temperature dependence of relaxation time (τ𝜏\tauitalic_τ).

Supplementary Note 5: Experiment with red light.

In the main text, we have presented all the measurements conducted using the green light. In this section, we provide an additional set of data obtained in three consecutive cycles (see Supplementary Fig. 5a) using a red light of wavelength λ𝜆\lambdaitalic_λ = 650 nm, power = 60 μ𝜇\muitalic_μ Watt). Supplementary Fig. 5b and 5c shows the temperature evolution of β𝛽\betaitalic_β and τ𝜏\tauitalic_τ respectively obtained from the fitting. As evident, while the enhancement in τ𝜏\tauitalic_τ below 50 K is consistent with results with green light, β𝛽\betaitalic_β slightly decreases from its peak value at lower temperatures. However, this is within the error bar and the results qualitatively align with the findings from the measurements using green light and suggest that the behavior observed is not specific to a particular wavelength of light.

Refer to caption
Supplementary Figure 5: a. Temporal evolution of resistance under red light illumination (λ𝜆\lambdaitalic_λ = 650 nm, power = 60 μ𝜇\muitalic_μ watt) for 30 minutes measured at several fixed temperatures. After 30 minutes, resistance relaxation was observed in dark for the next one hour. This measurement was repeated for 3 consecutive cycles. For comparative analysis, change in resistance has been converted into relative percentage change (ΔΔ\Deltaroman_ΔR𝑅Ritalic_R/R0subscript𝑅0R_{0}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT)×cross-product\crossproduct×100. b. Temperature dependence of stretching exponent (β𝛽\betaitalic_β) obtained from fitting of resistance relaxation in light off stage with a stretched exponential function for all three cycles. c. Temperature dependence of corresponding relaxation time (τ𝜏\tauitalic_τ) for all three cycles.

Supplementary Note 6: Long-time relaxation of persistence photo-resistance.

In Fig. 2c of the main text and Supplementary Fig. 5b, the value of the stretching exponent was determined by fitting the resistance relaxation during the off-stage, which lasted for 1.5 hours. To ensure that the resistance relaxation was substantial enough to yield a reliable fitting, we conducted an additional measurement where the resistance relaxation was observed for an entire day (see Supplementary Fig. 6). It is important to emphasize that even when the data is fitted for an extended period of up to 24 hours, the same value of β𝛽\betaitalic_β = 0.5 is obtained.

Refer to caption
Supplementary Figure 6: Temporal evolution of resistance under green light illumination (λ𝜆\lambdaitalic_λ = 527 nm, power = 145 μ𝜇\muitalic_μ Watt) for 30 minutes measured at 15 K. After 30 minutes, resistance relaxation was observed in dark for the next 24 hours. For comparative analysis, change in resistance has been converted into relative percentage change (ΔΔ\Deltaroman_ΔR𝑅Ritalic_R/R0subscript𝑅0R_{0}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT)×cross-product\crossproduct×100.

Supplementary Note 7: Deviation of persistence photo-resistance relaxation time (τ𝜏\tauitalic_τ) from activated behavior at low temperature.

In the large lattice relaxation (LLR) model [10], the recombination of electron-hole pairs occurs through the thermal excitation of electrons over an energy barrier. This process is an activated process of the Arrhenius type. To validate this model, we have plotted ln τ𝜏\tauitalic_τ against 1000/T𝑇Titalic_T in Supplementary Fig. 7. It is evident from the plot that the data does not exhibit the expected linear behavior at low temperatures signifying that the LLR model can not account for our experimental observation of glassy dynamics below 35 K.

Refer to caption
Supplementary Figure 7: Arrhenius plot of photo-resistance relaxation time (τ𝜏\tauitalic_τ). As evident, a clear deviation from activation behavior is observed below 35 K.

Supplementary Note 8: Dielectric loss in pristine KTaO3.

In its ideal form, KTaO3 is perfectly centrosymmetric and hence, should not possess any electric dipole moment. However, the presence of impurities and disorder can disrupt the crystal’s inversion symmetry, leading to the development of permanent electric dipoles [11, 12, 13, 14, 15, 16]. In highly polarizable host such as KTaO3, these dipoles polarize the surrounding lattice leading to formation of polar nano regions (PNRs) [17]. Under an applied AC electric field, these PNRs act as a source of dielectric losses, which can be observed as a peak in the loss tangent (tan δ𝛿\deltaitalic_δ = ϵitalic-ϵ\epsilonitalic_ϵ′′′′{}^{{}^{\prime\prime}}start_FLOATSUPERSCRIPT start_FLOATSUPERSCRIPT ′ ′ end_FLOATSUPERSCRIPT end_FLOATSUPERSCRIPT/ϵitalic-ϵ\epsilonitalic_ϵ{}^{{}^{\prime}}start_FLOATSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_FLOATSUPERSCRIPT, where ϵitalic-ϵ\epsilonitalic_ϵ{}^{{}^{\prime}}start_FLOATSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_FLOATSUPERSCRIPT and ϵitalic-ϵ\epsilonitalic_ϵ′′′′{}^{{}^{\prime\prime}}start_FLOATSUPERSCRIPT start_FLOATSUPERSCRIPT ′ ′ end_FLOATSUPERSCRIPT end_FLOATSUPERSCRIPT are the real and imaginary parts of the complex dielectric function, respectively).

Our temperature-dependent measurement of dielectric function indeed reveals the presence of PNRs even in our pristine KTaO3 single crystal (see inset of Supplementary Fig. 8). Further, frequency-dependent measurements reveal that the dielectric loss is an activated process emphasizing that the PNRs in pristine crystals are very dilute and independent.

Refer to caption
Supplementary Figure 8: Arrhenius plot of dielectric relaxation frequency (f𝑓fitalic_f) for pristine KTaO3. Inset shows the dielectric loss tangent (tan δ𝛿\deltaitalic_δ) around 54 K at f𝑓fitalic_f = 50 kHz.

Supplementary Note 9. Correspondence between the appearance of polar nano regions and photo-do** effect.

In order to study the correlation between PNRs and photo-do** effect, we have compared the temperature-dependent total SHG intensity with the fraction of resistance (in terms of relative percentage change (ΔΔ\Deltaroman_ΔR𝑅Ritalic_R/R0subscript𝑅0R_{0}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT)×cross-product\crossproduct×100) which has not been recovered at the end of 1.5 hours after turning off the light. We call this quantity the persistence photo resistance denoted with the symbol δ𝛿\deltaitalic_δ. See the inset of the bottom panel of Supplementary Fig. 9 for the definition of δ𝛿\deltaitalic_δ. As evident, the appearance of finite signal in SHG exactly coincides with the δ𝛿\deltaitalic_δ, strongly signifying the direct role of PNRs behind the effective electron-hole separation of our samples.

Refer to caption
Supplementary Figure 9: (upper panel) Temperature-dependent SHG intensity measured on oxygen deficient KTaO3 sample. (lower panel) Temperature dependence of persistence photo resistance (δ𝛿\deltaitalic_δ) in terms of relative percentage change at the end of 1.5 hours after turning off the light (see the inset of the lower panel for the definition of δ𝛿\deltaitalic_δ).

Supplementary Note 10: Raman measurement on a metallic oxygen deficient KTO.

Supplementary Fig. 10 shows the Raman spectra for one of the metallic oxygen-deficient KTO sample at room temperature. As evident, the soft polar mode (TO1, marked by a red arrow) is preserved even in the metallic sample [18, 19, 20, 21, 22]. This result is consistent with our SHG measurement.

Refer to caption
Supplementary Figure 10: Raman spectra of a metallic oxygen deficient KTO sample at room temperature.

Supplementary Note 11: Presence of potassium vacancy.

To investigate the presence of potassium vacancy in our oxygen-deficient KTaO3 sample, K 2p core levels were collected at the Hard X-ray Photoelectron Spectroscopy (HAXPES) beamline (P22) at PETRA III, DESY. Supplementary Fig. 11a shows one representative experimental data (recorded with an incident photon energy of 5800 eV at room temperature) along with its fitting with the convolution of Lorentzian and Gaussian functions. It is evident from the figure that, in addition to the K 2p 3/2 peak at 292.65 eV arising from the lattice, there is an extra peak appearing at 293.1 eV. It has been previously shown that the presence of an additional peak at a higher binding energy is a characteristic feature of a potassium vacancy in the system [23]. In order to further understand the potassium vacancy profile in our sample, we have carried out measurements with varying photon energy from 3400 eV to 5800 eV (see Table I for the values of mean free path (MFP) and mean escape depth (MED)). As evident from Supplementary Fig. 11b, there are hardly any significant alterations in the line shape of the spectra as the photon energy was increased. This observation indicates that the potassium vacancy profile is homogeneous throughout the bulk of the sample.

Refer to caption
Supplementary Figure 11: a Deconvolution of K 2p core level spectrum (for oxygen-deficient KTaO3 sample) by using the convolution of Lorentzian and Gaussian function. This data was recorded at room temperature with photon energy hhitalic_hν𝜈\nuitalic_ν = 5800 eV. Yellow-filled circles denote the experimental data and a solid black line denotes the simulated spectra. Curves filled with gray and sky blue color correspond to the lattice potassium and potassium vacancy respectively. The dashed black line denotes the Shirley background and the red-filled curve corresponds to the difference between the experimental data and simulated spectra. b K 2p core level spectra recorded with varying incident photon energy. All the spectra were recorded at room temperature.
Photon energy (eV) MFP (nm) MED (nm)
3400 5 15
4600 6.4 19.2
5800 7.7 23.1
Table 1: Table of mean free path (MFP) and mean escape depth (MED) with increasing photon energy. MFP was calculated from the formula m𝑚mitalic_m(K.E.)γ where K.E. is the kinetic energy of the ejected electrons and the values of m𝑚mitalic_m and γ𝛾\gammaitalic_γ were taken to be 0.12 and 0.75 respectively [24]. Further, MED was roughly assumed to be 3 times MFP which would roughly account for 95 percent of the total collected intensity.

Supplementary Note 12: Berry phase calculations.

Refer to caption
Supplementary Figure 12: a. The relaxed structure of KTaO3 with Ta off-centering along the (100) and (110) direction around K vacancy (antisite like Ta defect) for a supercell of size 2×cross-product\crossproduct×2×cross-product\crossproduct×2. b. Band structure plot for (110) direction antisite like Ta defect calculated using density functional theory within PBE-GGA (shown in red solid lines) and Wannier interpolated bands (shown in black dotted lines). The blue solid line is the Fermi energy. c. Plot of total Berry curvature in the plane kz=0subscript𝑘𝑧0k_{z}=0italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0 (in log scale) for the (110) antisitelike Ta defect structure calculated on a 2D k-mesh of sample-size 41×41cross-product414141\crossproduct 4141 × 41 (top panel) and 101×101cross-product101101101\crossproduct 101101 × 101 (bottom panel).

We have investigated the probable mechanism for the realization of PNRs theoretically using noncollinear density functional (DFT) theory. Antisite defect in perovskite materials like SrTiO3 and complex perovskite oxides (Ca, Sr)3Mn2O7 is known for causing macroscopic polarization in the system [25, 26]. Due to the similarities in the electronic properties of SrTiO3 and KTaO3, we expect KTaO3 to develop polarization due to antsite defects. We have created Ta antisite defect in a supercell of size 2×2×2cross-product2222\crossproduct 2\crossproduct 22 × 2 × 2 conventional unit cell. In our calculations, the Ta off-centering is considered along (100) and (110) directions. The relaxed structures are shown in Supplementary Fig. 12a. In Supplementary Fig. 12b we have shown the band structure of the system with Ta antisite defect along (110). The red solid lines are the bands calculated using DFT within PBE-GGA[27]. In order to calculate the macroscopic polarization in the system, we use the modern theory of polarization. The change in electronic contribution to the polarization ΔPΔ𝑃\Delta Proman_Δ italic_P is defined as [28, 29]

ΔP=e2πϕΔ𝑃𝑒2𝜋italic-ϕ\Delta P=-\dfrac{e}{2\pi}\phiroman_Δ italic_P = - divide start_ARG italic_e end_ARG start_ARG 2 italic_π end_ARG italic_ϕ (8)

where ϕitalic-ϕ\phiitalic_ϕ is the Berry phase, which is the integral of the Berry curvature over a surface S𝑆Sitalic_S bounded by a closed path in klimit-from𝑘k-italic_k -space, i.e., [30]

ϕ=SΩ(𝐤)𝑑𝐤italic-ϕsubscript𝑆Ω𝐤differential-d𝐤\phi=\int_{S}\Omega(\mathbf{k})d\mathbf{k}italic_ϕ = ∫ start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT roman_Ω ( bold_k ) italic_d bold_k (9)

One can calculate Berry curvature using Bloch states un(𝐤)subscript𝑢𝑛𝐤u_{n}(\mathbf{k})italic_u start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( bold_k ) as

Ωαβ(𝐤)=nfn(𝐤)Ωn,αβ(𝐤)=n2fn(𝐤)Imun(𝐤)kα|un(𝐤)kβsubscriptΩ𝛼𝛽𝐤subscript𝑛subscript𝑓𝑛𝐤subscriptΩ𝑛𝛼𝛽𝐤subscript𝑛2subscript𝑓𝑛𝐤Iminner-productpartial-derivativesubscript𝑘𝛼subscript𝑢𝑛𝐤partial-derivativesubscript𝑘𝛽subscript𝑢𝑛𝐤\Omega_{\alpha\beta}(\mathbf{k})=\sum_{n}f_{n}(\mathbf{k})\Omega_{n,\alpha% \beta}(\mathbf{k})=\sum_{n}-2f_{n}(\mathbf{k})\text{Im}\innerproduct{% \partialderivative{u_{n}(\mathbf{k})}{k_{\alpha}}}{\partialderivative{u_{n}(% \mathbf{k})}{k_{\beta}}}roman_Ω start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT ( bold_k ) = ∑ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( bold_k ) roman_Ω start_POSTSUBSCRIPT italic_n , italic_α italic_β end_POSTSUBSCRIPT ( bold_k ) = ∑ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT - 2 italic_f start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( bold_k ) Im ⟨ start_ARG divide start_ARG ∂ start_ARG italic_u start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( bold_k ) end_ARG end_ARG start_ARG ∂ start_ARG italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT end_ARG end_ARG end_ARG | start_ARG divide start_ARG ∂ start_ARG italic_u start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( bold_k ) end_ARG end_ARG start_ARG ∂ start_ARG italic_k start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT end_ARG end_ARG end_ARG ⟩ (10)

where α,β𝛼𝛽\alpha,\betaitalic_α , italic_β are the cartesian indices and fn(𝐤)subscript𝑓𝑛𝐤f_{n}(\mathbf{k})italic_f start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( bold_k ) is the occupation number of the state n𝑛nitalic_n. Ta antisite defects in KTaO3 have a partially occupied band. As it well known, taking derivatives of un(𝐤)subscript𝑢𝑛𝐤u_{n}(\mathbf{k})italic_u start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( bold_k ) in a finite-difference scheme in the presence of band crossing and avoided crossing becomes difficult. We followed the procedure described by X. Wang et. al. [31] for the calculation of Berry curvature using Wannier functions as implemented in the Wannier90 code [32]. In Supplementary Fig. 12b, we have shown the Wannier interpolated bands in black dotted lines. From the figure, it is clear that both the set of bands calculated using DFT and Wannier interpolation match very well near the Fermi level. In Supplementary Fig. 12c, we have shown the total Berry curvature for the antisite Ta defect along (110) in the plane kz=0subscript𝑘𝑧0k_{z}=0italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0 calculated using 2D k-mesh of size 41×41cross-product414141\crossproduct 4141 × 41 (top panel) and 101×101cross-product101101101\crossproduct 101101 × 101 (bottom panel). While there are small quantitative differences, the qualitative features in both the panels remain the same.

Supplementary Note 13: Toy model of complex inter-band electronic relaxation mediated by a glassy bath.

As discussed in the main text, we consider the following model of coupled electron (el)-glass (gl) system.

H𝐻\displaystyle Hitalic_H =Hel+Hgl+Helglabsentsubscript𝐻𝑒𝑙subscript𝐻𝑔𝑙subscript𝐻𝑒𝑙𝑔𝑙\displaystyle=H_{el}+H_{gl}+H_{el-gl}= italic_H start_POSTSUBSCRIPT italic_e italic_l end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT italic_g italic_l end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT italic_e italic_l - italic_g italic_l end_POSTSUBSCRIPT (11a)
Helsubscript𝐻𝑒𝑙\displaystyle H_{el}italic_H start_POSTSUBSCRIPT italic_e italic_l end_POSTSUBSCRIPT =i,j=1Nctijcicjε0α=1Nffαfα,(ε0>0)absentsuperscriptsubscript𝑖𝑗1subscript𝑁𝑐subscript𝑡𝑖𝑗superscriptsubscript𝑐𝑖subscript𝑐𝑗subscript𝜀0superscriptsubscript𝛼1subscript𝑁𝑓superscriptsubscript𝑓𝛼subscript𝑓𝛼subscript𝜀00\displaystyle=-\sum_{i,j=1}^{N_{c}}t_{ij}c_{i}^{\dagger}c_{j}-\varepsilon_{0}% \sum_{\alpha=1}^{N_{f}}f_{\alpha}^{\dagger}f_{\alpha},~{}~{}~{}~{}(\varepsilon% _{0}>0)= - ∑ start_POSTSUBSCRIPT italic_i , italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_t start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_α = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_f start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_f start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT , ( italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT > 0 ) (11b)
Hglsubscript𝐻𝑔𝑙\displaystyle H_{gl}italic_H start_POSTSUBSCRIPT italic_g italic_l end_POSTSUBSCRIPT =μ=1Ngpμ22m+U({xμ})absentsuperscriptsubscript𝜇1subscript𝑁𝑔superscriptsubscript𝑝𝜇22𝑚𝑈subscript𝑥𝜇\displaystyle=\sum_{\mu=1}^{N_{g}}\frac{p_{\mu}^{2}}{2m}+U(\{x_{\mu}\})= ∑ start_POSTSUBSCRIPT italic_μ = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_POSTSUPERSCRIPT divide start_ARG italic_p start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m end_ARG + italic_U ( { italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT } ) (11c)
Helglsubscript𝐻𝑒𝑙𝑔𝑙\displaystyle H_{el-gl}italic_H start_POSTSUBSCRIPT italic_e italic_l - italic_g italic_l end_POSTSUBSCRIPT =iαμ(Viαμcifα+h.c.)xμ\displaystyle=\sum_{i\alpha\mu}(V_{i\alpha\mu}c_{i}^{\dagger}f_{\alpha}+% \mathrm{h.c.})x_{\mu}= ∑ start_POSTSUBSCRIPT italic_i italic_α italic_μ end_POSTSUBSCRIPT ( italic_V start_POSTSUBSCRIPT italic_i italic_α italic_μ end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_f start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT + roman_h . roman_c . ) italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT (11d)

The electronic part consists of a conduction band and a flat impurity band at energy ε0subscript𝜀0-\varepsilon_{0}- italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. We consider three different energy dispersions for the conduction band, corresponding to different lattices and hop** amplitudes tijsubscript𝑡𝑖𝑗t_{ij}italic_t start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT, namely – (1) a flat band or δ𝛿\deltaitalic_δ function density of states (DOS) with bandwidth W=0𝑊0W=0italic_W = 0, (2) a semicircular DOS, g(ϵ)=(1/2π)W2ω2θ(W|ω|)𝑔italic-ϵ12𝜋superscript𝑊2superscript𝜔2𝜃𝑊𝜔g(\epsilon)=(1/2\pi)\sqrt{W^{2}-\omega^{2}}\theta(W-|\omega|)italic_g ( italic_ϵ ) = ( 1 / 2 italic_π ) square-root start_ARG italic_W start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_θ ( italic_W - | italic_ω | ) with band width W𝑊Witalic_W [θ(x)𝜃𝑥\theta(x)italic_θ ( italic_x ) is heaviside step function], e.g., corresponding to a Bethe lattice, and (3) DOS corresponding to three-dimensional (3d) simple cubic lattice with nearest-neighbour hop** W/12𝑊12W/12italic_W / 12. The band gap between the conduction band minimum and the impurity band is Δ=ε0W/2Δsubscript𝜀0𝑊2\Delta=\varepsilon_{0}-W/2roman_Δ = italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_W / 2. We set the chemical potential at the center of the conduction band, i.e., μ=0𝜇0\mu=0italic_μ = 0, as appropriate for a metallic system (see Fig. 5(a) of main text.

The Hamiltonian Hglsubscript𝐻𝑔𝑙H_{gl}italic_H start_POSTSUBSCRIPT italic_g italic_l end_POSTSUBSCRIPT for a set of particles with positions {xμ}subscript𝑥𝜇\{x_{\mu}\}{ italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT }, and their canonically conjugate momenta pμsubscript𝑝𝜇p_{\mu}italic_p start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ([xμ,pν]=ıδμνsubscript𝑥𝜇subscript𝑝𝜈italic-ısubscript𝛿𝜇𝜈[x_{\mu},p_{\nu}]=\imath\delta_{\mu\nu}[ italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT , italic_p start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ] = italic_ı italic_δ start_POSTSUBSCRIPT italic_μ italic_ν end_POSTSUBSCRIPT with =1Planck-constant-over-2-pi1\hbar=1roman_ℏ = 1) models the dynamics of a local glassy background. The exact form of the inter-particle interaction U({xμ})𝑈subscript𝑥𝜇U(\{x_{\mu}\})italic_U ( { italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT } ) is not crucial for our calculations. However, as an example, we take U({xμ})=μ<ν<γJμνγxμxνxγ𝑈subscript𝑥𝜇subscript𝜇𝜈𝛾subscript𝐽𝜇𝜈𝛾subscript𝑥𝜇subscript𝑥𝜈subscript𝑥𝛾U(\{x_{\mu}\})=\sum_{\mu<\nu<\gamma}J_{\mu\nu\gamma}x_{\mu}x_{\nu}x_{\gamma}italic_U ( { italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT } ) = ∑ start_POSTSUBSCRIPT italic_μ < italic_ν < italic_γ end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT italic_μ italic_ν italic_γ end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT with the spherical constraint μxμ2=Ngsubscript𝜇superscriptsubscript𝑥𝜇2subscript𝑁𝑔\sum_{\mu}x_{\mu}^{2}=N_{g}∑ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = italic_N start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT, corresponding a well-known solvable model for glasses, namely the infinite-range spherical p𝑝pitalic_p-spin glass model with p=3𝑝3p=3italic_p = 3-spin coupling [33]. Here, Jμνγsubscript𝐽𝜇𝜈𝛾J_{\mu\nu\gamma}italic_J start_POSTSUBSCRIPT italic_μ italic_ν italic_γ end_POSTSUBSCRIPT is real Gaussian random number with zero mean and variance Jμνγ2¯=3!J2/2Ng2¯subscriptsuperscript𝐽2𝜇𝜈𝛾3superscript𝐽22superscriptsubscript𝑁𝑔2\overline{J^{2}_{\mu\nu\gamma}}=3!J^{2}/2N_{g}^{2}over¯ start_ARG italic_J start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_μ italic_ν italic_γ end_POSTSUBSCRIPT end_ARG = 3 ! italic_J start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 italic_N start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, where the overline denotes averaging over different realizations of Jμνγsubscript𝐽𝜇𝜈𝛾J_{\mu\nu\gamma}italic_J start_POSTSUBSCRIPT italic_μ italic_ν italic_γ end_POSTSUBSCRIPT. The particular scaling with Ngsubscript𝑁𝑔N_{g}italic_N start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ensures extensive free energy in the thermodynamic limit Ngsubscript𝑁𝑔N_{g}\to\inftyitalic_N start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT → ∞. The p𝑝pitalic_p-spin glass model, with both quantum [33] and classical dissipative [34] dynamics, undergoes a glass transition at temperature TgJsimilar-tosubscript𝑇𝑔𝐽T_{g}\sim Jitalic_T start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ∼ italic_J. For temperature TTggreater-than-or-equivalent-to𝑇subscript𝑇𝑔T\gtrsim T_{g}italic_T ≳ italic_T start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT, the model gives rise to a supercooled liquid regime, like standard structural glasses [35], with complex two-step relaxation for the dynamical correlation function

Cgl(t)subscript𝐶𝑔𝑙𝑡\displaystyle C_{gl}(t)italic_C start_POSTSUBSCRIPT italic_g italic_l end_POSTSUBSCRIPT ( italic_t ) =xμ(t)xμ(0)¯=Ae|t|/τs+Be(|t|/τα)βabsent¯delimited-⟨⟩subscript𝑥𝜇𝑡subscript𝑥𝜇0𝐴superscript𝑒𝑡subscript𝜏𝑠𝐵superscript𝑒superscript𝑡subscript𝜏𝛼𝛽\displaystyle=\overline{\langle x_{\mu}(t)x_{\mu}(0)\rangle}=Ae^{-|t|/\tau_{s}% }+Be^{-(|t|/\tau_{\alpha})^{\beta}}= over¯ start_ARG ⟨ italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ( italic_t ) italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ( 0 ) ⟩ end_ARG = italic_A italic_e start_POSTSUPERSCRIPT - | italic_t | / italic_τ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_POSTSUPERSCRIPT + italic_B italic_e start_POSTSUPERSCRIPT - ( | italic_t | / italic_τ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT (12)

, having a short-time exponential decay, and long-time stretched exponential decay.

To obtain a solvable model for the coupled el-gl system, we also take the coupling Viαμsubscript𝑉𝑖𝛼𝜇V_{i\alpha\mu}italic_V start_POSTSUBSCRIPT italic_i italic_α italic_μ end_POSTSUBSCRIPT between the glass, and the conduction and impurity electronic states, infinite range. Here Viαμsubscript𝑉𝑖𝛼𝜇V_{i\alpha\mu}italic_V start_POSTSUBSCRIPT italic_i italic_α italic_μ end_POSTSUBSCRIPT is a complex Gaussian random number with zero mean and variance |Viαμ|2¯=V/(NcNfNg)1/3¯superscriptsubscript𝑉𝑖𝛼𝜇2𝑉superscriptsubscript𝑁𝑐subscript𝑁𝑓subscript𝑁𝑔13\overline{|V_{i\alpha\mu}|^{2}}=V/(N_{c}N_{f}N_{g})^{1/3}over¯ start_ARG | italic_V start_POSTSUBSCRIPT italic_i italic_α italic_μ end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG = italic_V / ( italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT. The particular scaling ensures a well-defined thermodynamic limit Nc,Nf,Ngsubscript𝑁𝑐subscript𝑁𝑓subscript𝑁𝑔N_{c},N_{f},N_{g}\to\inftyitalic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT , italic_N start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT , italic_N start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT → ∞ with finite ratios pf=Nf/Ncsubscript𝑝𝑓subscript𝑁𝑓subscript𝑁𝑐p_{f}=N_{f}/N_{c}italic_p start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = italic_N start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT / italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT and pg=Ng/Ncsubscript𝑝𝑔subscript𝑁𝑔subscript𝑁𝑐p_{g}=N_{g}/N_{c}italic_p start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT = italic_N start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT / italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. The ratios allow us tune the backaction of one part of the system on the other. For example, for simplicity, and to gain an analytical understanding, as discussed below, we take pf,pg1much-greater-thansubscript𝑝𝑓subscript𝑝𝑔1p_{f},p_{g}\gg 1italic_p start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT , italic_p start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ≫ 1, such that the back actions of the conduction electrons on the impurity electrons and the glass are negligible. We expect our main conclusions to be valid for pf,pg1subscript𝑝𝑓subscript𝑝𝑔1p_{f},p_{g}\approx 1italic_p start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT , italic_p start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ≈ 1, i.e., when the backaction is substantial.

For the above-disordered model (Supplementary Eq. (10)), using the standard replica method [36, 33], we calculate the disorder-averaged connected equilibrium dynamical density-density correlation function for the conduction electrons, Cel(t)=ni(t)ni(0)¯ni(0)2subscript𝐶𝑒𝑙𝑡¯delimited-⟨⟩subscript𝑛𝑖𝑡subscript𝑛𝑖0superscriptdelimited-⟨⟩subscript𝑛𝑖02C_{el}(t)=\overline{\langle n_{i}(t)n_{i}(0)\rangle}-\langle n_{i}(0)\rangle^{2}italic_C start_POSTSUBSCRIPT italic_e italic_l end_POSTSUBSCRIPT ( italic_t ) = over¯ start_ARG ⟨ italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( 0 ) ⟩ end_ARG - ⟨ italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( 0 ) ⟩ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, at temperature TΔmuch-less-than𝑇ΔT\ll\Deltaitalic_T ≪ roman_Δ. Here ni=cicisubscript𝑛𝑖superscriptsubscript𝑐𝑖subscript𝑐𝑖n_{i}=c_{i}^{\dagger}c_{i}italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is the conduction electron density. The correlation function Cel(t)subscript𝐶𝑒𝑙𝑡C_{el}(t)italic_C start_POSTSUBSCRIPT italic_e italic_l end_POSTSUBSCRIPT ( italic_t ) captures the relaxation of the thermally excited electrons in the conduction band at a finite temperature.

We obtain the replicated partition function, Zn=𝒟(c¯,c)𝒟(f¯,f)𝒟xexp(S[c¯,c,f¯,f,x])superscript𝑍𝑛𝒟¯𝑐𝑐𝒟¯𝑓𝑓𝒟𝑥𝑆¯𝑐𝑐¯𝑓𝑓𝑥Z^{n}=\int\mathcal{D}(\bar{c},c)\mathcal{D}(\bar{f},f)\mathcal{D}x\exp(-S[\bar% {c},c,\bar{f},f,x])italic_Z start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT = ∫ caligraphic_D ( over¯ start_ARG italic_c end_ARG , italic_c ) caligraphic_D ( over¯ start_ARG italic_f end_ARG , italic_f ) caligraphic_D italic_x roman_exp ( start_ARG - italic_S [ over¯ start_ARG italic_c end_ARG , italic_c , over¯ start_ARG italic_f end_ARG , italic_f , italic_x ] end_ARG ), as coherent-state imaginary-time path integral over the fermionic Grassmann fields {c¯ia,cia,f¯αa,fαa}subscript¯𝑐𝑖𝑎subscript𝑐𝑖𝑎subscript¯𝑓𝛼𝑎subscript𝑓𝛼𝑎\{\bar{c}_{ia},c_{ia},\bar{f}_{\alpha a},f_{\alpha a}\}{ over¯ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_i italic_a end_POSTSUBSCRIPT , italic_c start_POSTSUBSCRIPT italic_i italic_a end_POSTSUBSCRIPT , over¯ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_α italic_a end_POSTSUBSCRIPT , italic_f start_POSTSUBSCRIPT italic_α italic_a end_POSTSUBSCRIPT } and position variables {xia}subscript𝑥𝑖𝑎\{x_{ia}\}{ italic_x start_POSTSUBSCRIPT italic_i italic_a end_POSTSUBSCRIPT }, with replica index a=1,,n𝑎1𝑛a=1,\cdots,nitalic_a = 1 , ⋯ , italic_n and the action,

Snsubscript𝑆𝑛\displaystyle S_{n}italic_S start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT =01/T𝑑τij,ac¯ia(τ)[(τμ)δijtij]cja(τ)+01/T𝑑τα,af¯αa(τ)(τμε0)fαa(τ)absentsuperscriptsubscript01𝑇differential-d𝜏subscript𝑖𝑗𝑎subscript¯𝑐𝑖𝑎𝜏delimited-[]subscript𝜏𝜇subscript𝛿𝑖𝑗subscript𝑡𝑖𝑗subscript𝑐𝑗𝑎𝜏superscriptsubscript01𝑇differential-d𝜏subscript𝛼𝑎subscript¯𝑓𝛼𝑎𝜏subscript𝜏𝜇subscript𝜀0subscript𝑓𝛼𝑎𝜏\displaystyle=\int_{0}^{1/T}d\tau\sum_{ij,a}\bar{c}_{ia}(\tau)[(\partial_{\tau% }-\mu)\delta_{ij}-t_{ij}]c_{ja}(\tau)+\int_{0}^{1/T}d\tau\sum_{\alpha,a}\bar{f% }_{\alpha a}(\tau)(\partial_{\tau}-\mu-\varepsilon_{0})f_{\alpha a}(\tau)= ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 / italic_T end_POSTSUPERSCRIPT italic_d italic_τ ∑ start_POSTSUBSCRIPT italic_i italic_j , italic_a end_POSTSUBSCRIPT over¯ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_i italic_a end_POSTSUBSCRIPT ( italic_τ ) [ ( ∂ start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT - italic_μ ) italic_δ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT - italic_t start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ] italic_c start_POSTSUBSCRIPT italic_j italic_a end_POSTSUBSCRIPT ( italic_τ ) + ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 / italic_T end_POSTSUPERSCRIPT italic_d italic_τ ∑ start_POSTSUBSCRIPT italic_α , italic_a end_POSTSUBSCRIPT over¯ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_α italic_a end_POSTSUBSCRIPT ( italic_τ ) ( ∂ start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT - italic_μ - italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) italic_f start_POSTSUBSCRIPT italic_α italic_a end_POSTSUBSCRIPT ( italic_τ )
+01/T𝑑τ[12μ,a[m(τxμa)2+zxμa2]+μνγ,aJμνγxμa(τ)xνa(τ)xγa(τ)]superscriptsubscript01𝑇differential-d𝜏delimited-[]12subscript𝜇𝑎delimited-[]𝑚superscriptsubscript𝜏subscript𝑥𝜇𝑎2𝑧superscriptsubscript𝑥𝜇𝑎2subscript𝜇𝜈𝛾𝑎subscript𝐽𝜇𝜈𝛾subscript𝑥𝜇𝑎𝜏subscript𝑥𝜈𝑎𝜏subscript𝑥𝛾𝑎𝜏\displaystyle+\int_{0}^{1/T}d\tau\bigg{[}\frac{1}{2}\sum_{\mu,a}[m({\partial_{% \tau}x_{\mu a}})^{2}+zx_{\mu a}^{2}]+\sum_{\mu\nu\gamma,a}J_{\mu\nu\gamma}x_{% \mu a}(\tau)x_{\nu a}(\tau)x_{\gamma a}(\tau)\bigg{]}+ ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 / italic_T end_POSTSUPERSCRIPT italic_d italic_τ [ divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT italic_μ , italic_a end_POSTSUBSCRIPT [ italic_m ( ∂ start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_μ italic_a end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_z italic_x start_POSTSUBSCRIPT italic_μ italic_a end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] + ∑ start_POSTSUBSCRIPT italic_μ italic_ν italic_γ , italic_a end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT italic_μ italic_ν italic_γ end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_μ italic_a end_POSTSUBSCRIPT ( italic_τ ) italic_x start_POSTSUBSCRIPT italic_ν italic_a end_POSTSUBSCRIPT ( italic_τ ) italic_x start_POSTSUBSCRIPT italic_γ italic_a end_POSTSUBSCRIPT ( italic_τ ) ]
+01/Tdτiαμ,a[Viαμc¯ia(τ)fαa(τ)xμa(τ)+h.c.],\displaystyle+\int_{0}^{1/T}d\tau\sum_{i\alpha\mu,a}[V_{i\alpha\mu}\bar{c}_{ia% }(\tau)f_{\alpha a}(\tau)x_{\mu a}(\tau)+\mathrm{h.c.}],+ ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 / italic_T end_POSTSUPERSCRIPT italic_d italic_τ ∑ start_POSTSUBSCRIPT italic_i italic_α italic_μ , italic_a end_POSTSUBSCRIPT [ italic_V start_POSTSUBSCRIPT italic_i italic_α italic_μ end_POSTSUBSCRIPT over¯ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_i italic_a end_POSTSUBSCRIPT ( italic_τ ) italic_f start_POSTSUBSCRIPT italic_α italic_a end_POSTSUBSCRIPT ( italic_τ ) italic_x start_POSTSUBSCRIPT italic_μ italic_a end_POSTSUBSCRIPT ( italic_τ ) + roman_h . roman_c . ] , (13)

where the Lagrange’s multiplier z(τ)𝑧𝜏z(\tau)italic_z ( italic_τ ) imposes the spherical constraint on xμa(τ)subscript𝑥𝜇𝑎𝜏x_{\mu a}(\tau)italic_x start_POSTSUBSCRIPT italic_μ italic_a end_POSTSUBSCRIPT ( italic_τ ). After averaging over the distributions of Jμνγsubscript𝐽𝜇𝜈𝛾J_{\mu\nu\gamma}italic_J start_POSTSUBSCRIPT italic_μ italic_ν italic_γ end_POSTSUBSCRIPT and Viαμsubscript𝑉𝑖𝛼𝜇V_{i\alpha\mu}italic_V start_POSTSUBSCRIPT italic_i italic_α italic_μ end_POSTSUBSCRIPT, we obtain Zn¯¯superscript𝑍𝑛\overline{Z^{n}}over¯ start_ARG italic_Z start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG and the corresponding action as

S~nsubscript~𝑆𝑛\displaystyle\tilde{S}_{n}over~ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT =01/T𝑑τij,ac¯ia(τ)[(τμ)δijtij]cja(τ)+01/T𝑑τα,af¯αa(τ)(τμε0)fαa(τ)absentsuperscriptsubscript01𝑇differential-d𝜏subscript𝑖𝑗𝑎subscript¯𝑐𝑖𝑎𝜏delimited-[]subscript𝜏𝜇subscript𝛿𝑖𝑗subscript𝑡𝑖𝑗subscript𝑐𝑗𝑎𝜏superscriptsubscript01𝑇differential-d𝜏subscript𝛼𝑎subscript¯𝑓𝛼𝑎𝜏subscript𝜏𝜇subscript𝜀0subscript𝑓𝛼𝑎𝜏\displaystyle=\int_{0}^{1/T}d\tau\sum_{ij,a}\bar{c}_{ia}(\tau)[(\partial_{\tau% }-\mu)\delta_{ij}-t_{ij}]c_{ja}(\tau)+\int_{0}^{1/T}d\tau\sum_{\alpha,a}\bar{f% }_{\alpha a}(\tau)(\partial_{\tau}-\mu-\varepsilon_{0})f_{\alpha a}(\tau)= ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 / italic_T end_POSTSUPERSCRIPT italic_d italic_τ ∑ start_POSTSUBSCRIPT italic_i italic_j , italic_a end_POSTSUBSCRIPT over¯ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_i italic_a end_POSTSUBSCRIPT ( italic_τ ) [ ( ∂ start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT - italic_μ ) italic_δ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT - italic_t start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ] italic_c start_POSTSUBSCRIPT italic_j italic_a end_POSTSUBSCRIPT ( italic_τ ) + ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 / italic_T end_POSTSUPERSCRIPT italic_d italic_τ ∑ start_POSTSUBSCRIPT italic_α , italic_a end_POSTSUBSCRIPT over¯ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_α italic_a end_POSTSUBSCRIPT ( italic_τ ) ( ∂ start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT - italic_μ - italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) italic_f start_POSTSUBSCRIPT italic_α italic_a end_POSTSUBSCRIPT ( italic_τ )
+1201/Tdτμaxμa(τ)(mτ2+z)xμa(τ)+Ncdτdτab[V2(pfpg)1/3Gab(τ,τ)𝒢ba(τ,τ)Qab(τ,τ)\displaystyle+\frac{1}{2}\int^{1/T}_{0}d\tau\sum_{\mu a}x_{\mu a}(\tau)(-m% \partial^{2}_{\tau}+z)x_{\mu a}(\tau)+N_{c}\int d\tau d\tau^{\prime}\sum_{ab}% \bigg{[}V^{2}(p_{f}p_{g})^{1/3}G_{ab}(\tau,\tau^{\prime})\mathcal{G}_{ba}(\tau% ^{\prime},\tau)Q_{ab}(\tau,\tau^{\prime})+ divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∫ start_POSTSUPERSCRIPT 1 / italic_T end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_d italic_τ ∑ start_POSTSUBSCRIPT italic_μ italic_a end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_μ italic_a end_POSTSUBSCRIPT ( italic_τ ) ( - italic_m ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT + italic_z ) italic_x start_POSTSUBSCRIPT italic_μ italic_a end_POSTSUBSCRIPT ( italic_τ ) + italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ∫ italic_d italic_τ italic_d italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT [ italic_V start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_p start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT italic_G start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT ( italic_τ , italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) caligraphic_G start_POSTSUBSCRIPT italic_b italic_a end_POSTSUBSCRIPT ( italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_τ ) italic_Q start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT ( italic_τ , italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT )
J2pg4Qab3(τ,τ)],\displaystyle-\frac{J^{2}p_{g}}{4}Q^{3}_{ab}(\tau,\tau^{\prime})\bigg{]},- divide start_ARG italic_J start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_p start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG start_ARG 4 end_ARG italic_Q start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT ( italic_τ , italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ] , (14)

where we have introduced the large N𝑁Nitalic_N-fields,

Gab(τ,τ)subscript𝐺𝑎𝑏𝜏superscript𝜏\displaystyle G_{ab}(\tau,\tau^{\prime})italic_G start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT ( italic_τ , italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) =1Ncicia(τ)c¯ib(τ)absent1subscript𝑁𝑐subscript𝑖subscript𝑐𝑖𝑎𝜏subscript¯𝑐𝑖𝑏superscript𝜏\displaystyle=-\frac{1}{N_{c}}\sum_{i}c_{ia}(\tau)\bar{c}_{ib}(\tau^{\prime})= - divide start_ARG 1 end_ARG start_ARG italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_i italic_a end_POSTSUBSCRIPT ( italic_τ ) over¯ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_i italic_b end_POSTSUBSCRIPT ( italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) (15a)
𝒢ab(τ,τ)subscript𝒢𝑎𝑏𝜏superscript𝜏\displaystyle\mathcal{G}_{ab}(\tau,\tau^{\prime})caligraphic_G start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT ( italic_τ , italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) =1Nfαfαa(τ)f¯αb(τ)absent1subscript𝑁𝑓subscript𝛼subscript𝑓𝛼𝑎𝜏subscript¯𝑓𝛼𝑏superscript𝜏\displaystyle=-\frac{1}{N_{f}}\sum_{\alpha}f_{\alpha a}(\tau)\bar{f}_{\alpha b% }(\tau^{\prime})= - divide start_ARG 1 end_ARG start_ARG italic_N start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_α italic_a end_POSTSUBSCRIPT ( italic_τ ) over¯ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_α italic_b end_POSTSUBSCRIPT ( italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) (15b)
Qab(τ,τ)subscript𝑄𝑎𝑏𝜏superscript𝜏\displaystyle Q_{ab}(\tau,\tau^{\prime})italic_Q start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT ( italic_τ , italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) =1Ngμxμa(τ)xμb(τ)absent1subscript𝑁𝑔subscript𝜇subscript𝑥𝜇𝑎𝜏subscript𝑥𝜇𝑏superscript𝜏\displaystyle=\frac{1}{N_{g}}\sum_{\mu}x_{\mu a}(\tau)x_{\mu b}(\tau^{\prime})= divide start_ARG 1 end_ARG start_ARG italic_N start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_μ italic_a end_POSTSUBSCRIPT ( italic_τ ) italic_x start_POSTSUBSCRIPT italic_μ italic_b end_POSTSUBSCRIPT ( italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) (15c)

To promote the above as fluctuating dynamical fields, conjugate fields Σba(τ,τ)subscriptΣ𝑏𝑎superscript𝜏𝜏\Sigma_{ba}(\tau^{\prime},\tau)roman_Σ start_POSTSUBSCRIPT italic_b italic_a end_POSTSUBSCRIPT ( italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_τ ), σba(τ,τ)subscript𝜎𝑏𝑎superscript𝜏𝜏\sigma_{ba}(\tau^{\prime},\tau)italic_σ start_POSTSUBSCRIPT italic_b italic_a end_POSTSUBSCRIPT ( italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_τ ) and Πab(τ,τ)subscriptΠ𝑎𝑏𝜏superscript𝜏\Pi_{ab}(\tau,\tau^{\prime})roman_Π start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT ( italic_τ , italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) are introduced for G,𝒢𝐺𝒢G,~{}\mathcal{G}italic_G , caligraphic_G and Q𝑄Qitalic_Q, respectively, e.g., by using the relation

𝒟Gaτ,bτδ(NcGab(τ,τ)+icia(τ)c¯ib(τ))=𝒟G𝒟Σe𝑑τ𝑑τΣba(τ,τ)[NcGab(τ,τ)+icia(τ)c¯ib(τ)]=1𝒟𝐺subscriptproduct𝑎𝜏𝑏superscript𝜏𝛿subscript𝑁𝑐subscript𝐺𝑎𝑏𝜏superscript𝜏subscript𝑖subscript𝑐𝑖𝑎𝜏subscript¯𝑐𝑖𝑏superscript𝜏𝒟𝐺𝒟Σsuperscript𝑒differential-d𝜏differential-dsuperscript𝜏subscriptΣ𝑏𝑎superscript𝜏𝜏delimited-[]subscript𝑁𝑐subscript𝐺𝑎𝑏𝜏superscript𝜏subscript𝑖subscript𝑐𝑖𝑎𝜏subscript¯𝑐𝑖𝑏superscript𝜏1\displaystyle\int\mathcal{D}G\prod_{a\tau,b\tau^{\prime}}\delta(N_{c}G_{ab}(% \tau,\tau^{\prime})+\sum_{i}c_{ia}(\tau)\bar{c}_{ib}(\tau^{\prime}))=\int% \mathcal{D}G\mathcal{D}\Sigma e^{-\int d\tau d\tau^{\prime}\Sigma_{ba}(\tau^{% \prime},\tau)[N_{c}G_{ab}(\tau,\tau^{\prime})+\sum_{i}c_{ia}(\tau)\bar{c}_{ib}% (\tau^{\prime})]}=1∫ caligraphic_D italic_G ∏ start_POSTSUBSCRIPT italic_a italic_τ , italic_b italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_δ ( italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_G start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT ( italic_τ , italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) + ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_i italic_a end_POSTSUBSCRIPT ( italic_τ ) over¯ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_i italic_b end_POSTSUBSCRIPT ( italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ) = ∫ caligraphic_D italic_G caligraphic_D roman_Σ italic_e start_POSTSUPERSCRIPT - ∫ italic_d italic_τ italic_d italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT roman_Σ start_POSTSUBSCRIPT italic_b italic_a end_POSTSUBSCRIPT ( italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_τ ) [ italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_G start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT ( italic_τ , italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) + ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_i italic_a end_POSTSUBSCRIPT ( italic_τ ) over¯ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_i italic_b end_POSTSUBSCRIPT ( italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ] end_POSTSUPERSCRIPT = 1 (16)

As a result, we can now integrate out the fields (c¯,c)¯𝑐𝑐(\bar{c},c)( over¯ start_ARG italic_c end_ARG , italic_c ), (f¯,f)¯𝑓𝑓(\bar{f},f)( over¯ start_ARG italic_f end_ARG , italic_f ), and x𝑥xitalic_x. Assuming replica diagonal ansatz, e.g., Gab(τ,τ)=δabG(τ,τ)subscript𝐺𝑎𝑏𝜏superscript𝜏subscript𝛿𝑎𝑏𝐺𝜏superscript𝜏G_{ab}(\tau,\tau^{\prime})=\delta_{ab}G(\tau,\tau^{\prime})italic_G start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT ( italic_τ , italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) = italic_δ start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT italic_G ( italic_τ , italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ), we obtain Zn¯=𝒟(G,𝒢,Q,Σ,σ,Π)enSeff¯superscript𝑍𝑛𝒟𝐺𝒢𝑄Σ𝜎Πsuperscript𝑒𝑛subscript𝑆𝑒𝑓𝑓\overline{Z^{n}}=\int\mathcal{D}(G,\mathcal{G},Q,\Sigma,\sigma,\Pi)e^{-nS_{eff}}over¯ start_ARG italic_Z start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG = ∫ caligraphic_D ( italic_G , caligraphic_G , italic_Q , roman_Σ , italic_σ , roman_Π ) italic_e start_POSTSUPERSCRIPT - italic_n italic_S start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT end_POSTSUPERSCRIPT and the effective action

Seffsubscript𝑆eff\displaystyle S_{\rm eff}italic_S start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT =Nc𝑑ϵg(ϵ)Trln(τ+μϵΣmissing)NfTrln(τ+μ+ε0σmissing)+Ng2Trln(mτ2+zΠmissing)absentsubscript𝑁𝑐differential-ditalic-ϵ𝑔italic-ϵtracesubscript𝜏𝜇italic-ϵΣmissingsubscript𝑁𝑓tracesubscript𝜏𝜇subscript𝜀0𝜎missingsubscript𝑁𝑔2trace𝑚subscriptsuperscript2𝜏𝑧Πmissing\displaystyle=-N_{c}\int d\epsilon g(\epsilon)\Tr\ln\big(-\partial_{\tau}+\mu-% \epsilon-\Sigma\big{missing})-N_{f}\Tr\ln\big(-\partial_{\tau}+\mu+\varepsilon% _{0}-\sigma\big{missing})+\frac{N_{g}}{2}\Tr\ln\big(-m\partial^{2}_{\tau}+z-% \Pi\big{missing})= - italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ∫ italic_d italic_ϵ italic_g ( italic_ϵ ) roman_Tr roman_ln ( start_ARG - ∂ start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT + italic_μ - italic_ϵ - roman_Σ roman_missing end_ARG ) - italic_N start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT roman_Tr roman_ln ( start_ARG - ∂ start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT + italic_μ + italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_σ roman_missing end_ARG ) + divide start_ARG italic_N start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG roman_Tr roman_ln ( start_ARG - italic_m ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT + italic_z - roman_Π roman_missing end_ARG )
Ncdτdτ[Σ(τ,τ)G(τ,τ)+pfσ(τ,τ)𝒢(τ,τ)pg2Π(τ,τ)Q(τ,τ)+pgJ24Q(τ,τ)3\displaystyle-N_{c}\int d\tau d\tau^{\prime}\big{[}\Sigma(\tau,\tau^{\prime})G% (\tau^{\prime},\tau)+p_{f}\sigma(\tau,\tau^{\prime})\mathcal{G}(\tau^{\prime},% \tau)-\frac{p_{g}}{2}\Pi(\tau,\tau^{\prime})Q(\tau^{\prime},\tau)+\frac{p_{g}J% ^{2}}{4}Q(\tau,\tau^{\prime})^{3}- italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ∫ italic_d italic_τ italic_d italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT [ roman_Σ ( italic_τ , italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) italic_G ( italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_τ ) + italic_p start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT italic_σ ( italic_τ , italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) caligraphic_G ( italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_τ ) - divide start_ARG italic_p start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG roman_Π ( italic_τ , italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) italic_Q ( italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_τ ) + divide start_ARG italic_p start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT italic_J start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 end_ARG italic_Q ( italic_τ , italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT
V2(pfpg)1/3G(τ,τ)𝒢(τ,τ)Q(τ,τ)]\displaystyle-V^{2}(p_{f}p_{g})^{1/3}G(\tau,\tau^{\prime})\mathcal{G}(\tau^{% \prime},\tau)Q(\tau,\tau^{\prime})\big{]}- italic_V start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_p start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT italic_G ( italic_τ , italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) caligraphic_G ( italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_τ ) italic_Q ( italic_τ , italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ] (17)

In the large Ncsubscript𝑁𝑐N_{c}italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT limit, the saddle point solution is obtained by varying the above action with respect to G,Σ,𝒢,σ,Q,Π𝐺Σ𝒢𝜎𝑄ΠG,\Sigma,\mathcal{G},\sigma,Q,\Piitalic_G , roman_Σ , caligraphic_G , italic_σ , italic_Q , roman_Π and setting the variations to zero. Due to time-translation invariance at equilibrium, e.g., G(τ,τ)=G(ττ)𝐺𝜏superscript𝜏𝐺𝜏superscript𝜏G(\tau,\tau^{\prime})=G(\tau-\tau^{\prime})italic_G ( italic_τ , italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) = italic_G ( italic_τ - italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) and z(τ)=z𝑧𝜏𝑧z(\tau)=zitalic_z ( italic_τ ) = italic_z. As a result, we can write the saddle point equations in the following form after performing a Matsubara Fourier transform first, e.g., G(ττ)G(ıωn)𝐺𝜏superscript𝜏𝐺italic-ısubscript𝜔𝑛G(\tau-\tau^{\prime})\to G(\imath\omega_{n})italic_G ( italic_τ - italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) → italic_G ( italic_ı italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) with fermionic Matsubara frequency ωnsubscript𝜔𝑛\omega_{n}italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, and then doing an analytical continuation, G(ıωn)GR(ω+ı0+G(\imath\omega_{n})\to G_{R}(\omega+\imath 0^{+}italic_G ( italic_ı italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) → italic_G start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_ω + italic_ı 0 start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT), to real frequencies, where GRsubscript𝐺𝑅G_{R}italic_G start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT is the retarded Green’s function.

GR(ω)subscript𝐺𝑅𝜔\displaystyle G_{R}(\omega)italic_G start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_ω ) =𝑑ϵg(ϵ)1ω+μϵΣR(ω)absentdifferential-ditalic-ϵ𝑔italic-ϵ1𝜔𝜇italic-ϵsubscriptΣ𝑅𝜔\displaystyle=\int d\epsilon g(\epsilon)\frac{1}{\omega+\mu-\epsilon-\Sigma_{R% }(\omega)}= ∫ italic_d italic_ϵ italic_g ( italic_ϵ ) divide start_ARG 1 end_ARG start_ARG italic_ω + italic_μ - italic_ϵ - roman_Σ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_ω ) end_ARG (18a)
𝒢R1(ω)superscriptsubscript𝒢𝑅1𝜔\displaystyle\mathcal{G}_{R}^{-1}(\omega)caligraphic_G start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( italic_ω ) =ω+μ+ε0σR(ω)absent𝜔𝜇subscript𝜀0subscript𝜎𝑅𝜔\displaystyle=\omega+\mu+\varepsilon_{0}-\sigma_{R}(\omega)= italic_ω + italic_μ + italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_σ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_ω ) (18b)
QR1(ω)superscriptsubscript𝑄𝑅1𝜔\displaystyle Q_{R}^{-1}(\omega)italic_Q start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( italic_ω ) =mω2+zΠR(ω)absent𝑚superscript𝜔2𝑧subscriptΠ𝑅𝜔\displaystyle=-m\omega^{2}+z-\Pi_{R}(\omega)= - italic_m italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_z - roman_Π start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_ω ) (18c)
Σ(τ)Σ𝜏\displaystyle\Sigma(\tau)roman_Σ ( italic_τ ) =V2(pfpg)1/3𝒢(τ)Q(τ)absentsuperscript𝑉2superscriptsubscript𝑝𝑓subscript𝑝𝑔13𝒢𝜏𝑄𝜏\displaystyle=V^{2}(p_{f}p_{g})^{1/3}\mathcal{G}(\tau)Q(\tau)= italic_V start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_p start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT caligraphic_G ( italic_τ ) italic_Q ( italic_τ ) (18d)
σ(τ)𝜎𝜏\displaystyle\sigma(\tau)italic_σ ( italic_τ ) =V2pg1/3pf2/3G(τ)Q(τ)absentsuperscript𝑉2superscriptsubscript𝑝𝑔13superscriptsubscript𝑝𝑓23𝐺𝜏𝑄𝜏\displaystyle=V^{2}p_{g}^{1/3}p_{f}^{-2/3}G(\tau)Q(\tau)= italic_V start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_p start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT italic_p start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 2 / 3 end_POSTSUPERSCRIPT italic_G ( italic_τ ) italic_Q ( italic_τ ) (18e)
Π(τ)Π𝜏\displaystyle\Pi(\tau)roman_Π ( italic_τ ) =3J22Q3(τ)+V2pf1/3pg2/3G(τ)𝒢(τ)absent3superscript𝐽22superscript𝑄3𝜏superscript𝑉2superscriptsubscript𝑝𝑓13superscriptsubscript𝑝𝑔23𝐺𝜏𝒢𝜏\displaystyle=\frac{3J^{2}}{2}Q^{3}(\tau)+V^{2}p_{f}^{1/3}p_{g}^{-2/3}G(\tau)% \mathcal{G}(-\tau)= divide start_ARG 3 italic_J start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG italic_Q start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ( italic_τ ) + italic_V start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_p start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT italic_p start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 2 / 3 end_POSTSUPERSCRIPT italic_G ( italic_τ ) caligraphic_G ( - italic_τ ) (18f)

In the limit pf,pg1much-greater-thansubscript𝑝𝑓subscript𝑝𝑔1p_{f},p_{g}\gg 1italic_p start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT , italic_p start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ≫ 1, we can approximate σ(τ)0𝜎𝜏0\sigma(\tau)\approx 0italic_σ ( italic_τ ) ≈ 0 and Π(τ)(3J2/2)Q3(τ)Π𝜏3superscript𝐽22superscript𝑄3𝜏\Pi(\tau)\approx(3J^{2}/2)Q^{3}(\tau)roman_Π ( italic_τ ) ≈ ( 3 italic_J start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 ) italic_Q start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ( italic_τ ), and neglect the back action of the conduction electrons on the impurity states and the glass.

In this limit, following the numerical procedure similar to that in reference [37, 38], we can numerically solve the above self-consistency equations for various conduction electron DOS g(ϵ)𝑔italic-ϵg(\epsilon)italic_g ( italic_ϵ ) to obtain the retarded functions GR(ω)subscript𝐺𝑅𝜔G_{R}(\omega)italic_G start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_ω ) and QR(ω)subscript𝑄𝑅𝜔Q_{R}(\omega)italic_Q start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_ω ), whereas the retarded Green’s function of the impurity band is given by 𝒢R(ω)=(ω+ı0++μ+ε0)1subscript𝒢𝑅𝜔superscript𝜔italic-ısuperscript0𝜇subscript𝜀01\mathcal{G}_{R}(\omega)=(\omega+\imath 0^{+}+\mu+\varepsilon_{0})^{-1}caligraphic_G start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_ω ) = ( italic_ω + italic_ı 0 start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + italic_μ + italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. However, instead of self-consistently solving for QR(ω)subscript𝑄𝑅𝜔Q_{R}(\omega)italic_Q start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_ω ), we use the Supplementary Eq. (11) to obtain the spectral function of the glass from the fluctuation-dissipation relation

ρgl(ω)subscript𝜌𝑔𝑙𝜔\displaystyle\rho_{gl}(\omega)italic_ρ start_POSTSUBSCRIPT italic_g italic_l end_POSTSUBSCRIPT ( italic_ω ) =1πImQR(ω)=1πtanh(ω2T)Cgl(ω),absent1𝜋Imsubscript𝑄𝑅𝜔1𝜋𝜔2𝑇subscript𝐶𝑔𝑙𝜔\displaystyle=-\frac{1}{\pi}\mathrm{Im}Q_{R}(\omega)=-\frac{1}{\pi}\tanh\left(% \frac{\omega}{2T}\right)C_{gl}(\omega),= - divide start_ARG 1 end_ARG start_ARG italic_π end_ARG roman_Im italic_Q start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_ω ) = - divide start_ARG 1 end_ARG start_ARG italic_π end_ARG roman_tanh ( divide start_ARG italic_ω end_ARG start_ARG 2 italic_T end_ARG ) italic_C start_POSTSUBSCRIPT italic_g italic_l end_POSTSUBSCRIPT ( italic_ω ) , (19)

where Cgl(ω)=𝑑teıωtCgl(t)subscript𝐶𝑔𝑙𝜔superscriptsubscriptdifferential-d𝑡superscript𝑒italic-ı𝜔𝑡subscript𝐶𝑔𝑙𝑡C_{gl}(\omega)=\int_{-\infty}^{\infty}dte^{\imath\omega t}C_{gl}(t)italic_C start_POSTSUBSCRIPT italic_g italic_l end_POSTSUBSCRIPT ( italic_ω ) = ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_d italic_t italic_e start_POSTSUPERSCRIPT italic_ı italic_ω italic_t end_POSTSUPERSCRIPT italic_C start_POSTSUBSCRIPT italic_g italic_l end_POSTSUBSCRIPT ( italic_t ). As a result, the conduction electron self-energy can be obtained as

ΣR(ω)subscriptΣ𝑅𝜔\displaystyle\Sigma_{R}(\omega)roman_Σ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_ω ) =V2𝑑ω1𝑑ω2ρf(ω1)ρgl(ω2)nF(ω1)nB(ω2)nF(ω1)nB(ω2)ωω1ω2+ı0+,absentsuperscript𝑉2differential-dsubscript𝜔1differential-dsubscript𝜔2subscript𝜌𝑓subscript𝜔1subscript𝜌𝑔𝑙subscript𝜔2subscript𝑛𝐹subscript𝜔1subscript𝑛𝐵subscript𝜔2subscript𝑛𝐹subscript𝜔1subscript𝑛𝐵subscript𝜔2𝜔subscript𝜔1subscript𝜔2italic-ısuperscript0\displaystyle=V^{2}\int d\omega_{1}d\omega_{2}\rho_{f}(\omega_{1})\rho_{gl}(% \omega_{2})\frac{n_{F}(\omega_{1})n_{B}(\omega_{2})-n_{F}(-\omega_{1})n_{B}(-% \omega_{2})}{\omega-\omega_{1}-\omega_{2}+\imath 0^{+}},= italic_V start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∫ italic_d italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_d italic_ω start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) italic_ρ start_POSTSUBSCRIPT italic_g italic_l end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) divide start_ARG italic_n start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) italic_n start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) - italic_n start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( - italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) italic_n start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ( - italic_ω start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) end_ARG start_ARG italic_ω - italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_ı 0 start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT end_ARG , (20)

where nF(ω)subscript𝑛𝐹𝜔n_{F}(\omega)italic_n start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_ω ) and nB(ω)subscript𝑛𝐵𝜔n_{B}(\omega)italic_n start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ( italic_ω ) are Fermi and Bose functions, respectively, and ρf(ω)=δ(ω+ε0)subscript𝜌𝑓𝜔𝛿𝜔subscript𝜀0\rho_{f}(\omega)=\delta(\omega+\varepsilon_{0})italic_ρ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ( italic_ω ) = italic_δ ( italic_ω + italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) is the spectral function of the impurity electrons. Here we have redefined V2(pfpg)1/3superscript𝑉2superscriptsubscript𝑝𝑓subscript𝑝𝑔13V^{2}(p_{f}p_{g})^{1/3}italic_V start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_p start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT as V2superscript𝑉2V^{2}italic_V start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. Thus, using ΣR(ω)subscriptΣ𝑅𝜔\Sigma_{R}(\omega)roman_Σ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_ω ) in Supplementary Eq. (17a)), we can obtain the conduction electron Green’s function GR(ω)subscript𝐺𝑅𝜔G_{R}(\omega)italic_G start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_ω ).

.1 Density density correlation function

To capture the manifestation of the glassy relaxation in the electron-hole recombination process we look into the connected density-density correlator Cel(t)=ni(t)ni(0)ni(0)2subscript𝐶𝑒𝑙𝑡delimited-⟨⟩subscript𝑛𝑖𝑡subscript𝑛𝑖0superscriptdelimited-⟨⟩subscript𝑛𝑖02C_{el}(t)=\langle n_{i}(t)n_{i}(0)\rangle-\langle n_{i}(0)\rangle^{2}italic_C start_POSTSUBSCRIPT italic_e italic_l end_POSTSUBSCRIPT ( italic_t ) = ⟨ italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( 0 ) ⟩ - ⟨ italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( 0 ) ⟩ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, which captures the relaxation of the thermally excited carriers at temperature T𝑇Titalic_T. Cel(t)subscript𝐶𝑒𝑙𝑡C_{el}(t)italic_C start_POSTSUBSCRIPT italic_e italic_l end_POSTSUBSCRIPT ( italic_t ) can be obtained from the imaginary-time correlation function Cel(τ)=n(τ)n(0)n(0)2subscript𝐶𝑒𝑙𝜏delimited-⟨⟩𝑛𝜏𝑛0superscriptdelimited-⟨⟩𝑛02C_{el}(\tau)=\langle n(\tau)n(0)\rangle-\langle n(0)\rangle^{2}italic_C start_POSTSUBSCRIPT italic_e italic_l end_POSTSUBSCRIPT ( italic_τ ) = ⟨ italic_n ( italic_τ ) italic_n ( 0 ) ⟩ - ⟨ italic_n ( 0 ) ⟩ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, where n(τ)=(1/Nc)ici(τ)c(τ)𝑛𝜏1subscript𝑁𝑐subscript𝑖subscriptsuperscript𝑐𝑖𝜏𝑐𝜏n(\tau)=(1/N_{c})\sum_{i}c^{\dagger}_{i}(\tau)c(\tau)italic_n ( italic_τ ) = ( 1 / italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_c start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_τ ) italic_c ( italic_τ ). In the large-Ncsubscript𝑁𝑐N_{c}italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT limit, the connected correlator is given by the bubble diagram and can be expressed as Cel(τ)=G(τ)G(τ)subscript𝐶𝑒𝑙𝜏𝐺𝜏𝐺𝜏C_{el}(\tau)=G(\tau)G(-\tau)italic_C start_POSTSUBSCRIPT italic_e italic_l end_POSTSUBSCRIPT ( italic_τ ) = italic_G ( italic_τ ) italic_G ( - italic_τ ). Performing Matsubara Fourier transformation and then analytically continuing iΩnω+i0+𝑖subscriptΩ𝑛𝜔𝑖superscript0i\Omega_{n}\to\omega+i0^{+}italic_i roman_Ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT → italic_ω + italic_i 0 start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, where ΩmsubscriptΩ𝑚\Omega_{m}roman_Ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT is bosonic Matsubara frequency, we can obtain the retarded correlator

Cel,R(ω)=𝑑ω1𝑑ω2ρc(ω1)ρc(ω2)(nF(ω1)nF(ω2)nF(ω1)nF(ω2)ω1ω2ωni0+),subscript𝐶𝑒𝑙𝑅𝜔differential-dsubscript𝜔1differential-dsubscript𝜔2subscript𝜌𝑐subscript𝜔1subscript𝜌𝑐subscript𝜔2subscript𝑛𝐹subscript𝜔1subscript𝑛𝐹subscript𝜔2subscript𝑛𝐹subscript𝜔1subscript𝑛𝐹subscript𝜔2subscript𝜔1subscript𝜔2subscript𝜔𝑛𝑖superscript0\displaystyle C_{el,R}(\omega)=\int d\omega_{1}d\omega_{2}\rho_{c}(\omega_{1})% \rho_{c}(\omega_{2})\bigg{(}\frac{n_{F}(\omega_{1})n_{F}(-\omega_{2})-n_{F}(-% \omega_{1})n_{F}(\omega_{2})}{\omega_{1}-\omega_{2}-\omega_{n}-i0^{+}}\bigg{)},italic_C start_POSTSUBSCRIPT italic_e italic_l , italic_R end_POSTSUBSCRIPT ( italic_ω ) = ∫ italic_d italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_d italic_ω start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) italic_ρ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) ( divide start_ARG italic_n start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) italic_n start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( - italic_ω start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) - italic_n start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( - italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) italic_n start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) end_ARG start_ARG italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT - italic_i 0 start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT end_ARG ) , (21)

where ρc(ω)=(1/π)ImGR(ω)subscript𝜌𝑐𝜔1𝜋Imsubscript𝐺𝑅𝜔\rho_{c}(\omega)=-(1/\pi)\mathrm{Im}G_{R}(\omega)italic_ρ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_ω ) = - ( 1 / italic_π ) roman_Im italic_G start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_ω ) is the conduction electron spectral function. Using the fluctuation-dissipation theorem, we can relate Cel,R(ω)subscript𝐶𝑒𝑙𝑅𝜔C_{el,R}(\omega)italic_C start_POSTSUBSCRIPT italic_e italic_l , italic_R end_POSTSUBSCRIPT ( italic_ω ) to the Fourier transform of the real-time density-density correlation function Cel(t)subscript𝐶𝑒𝑙𝑡C_{el}(t)italic_C start_POSTSUBSCRIPT italic_e italic_l end_POSTSUBSCRIPT ( italic_t ) as Cel(ω)=coth(ω/2T)ImCel,R(ω)subscript𝐶𝑒𝑙𝜔hyperbolic-cotangent𝜔2𝑇Imsubscript𝐶𝑒𝑙𝑅𝜔C_{el}(\omega)=\coth(\omega/2T)\text{Im}C_{el,R}(\omega)italic_C start_POSTSUBSCRIPT italic_e italic_l end_POSTSUBSCRIPT ( italic_ω ) = roman_coth ( start_ARG italic_ω / 2 italic_T end_ARG ) Im italic_C start_POSTSUBSCRIPT italic_e italic_l , italic_R end_POSTSUBSCRIPT ( italic_ω ). Finally, performing the inverse Fourier transform we obtain the real-time density-density correlation function Cel(t)=(dω/2π)eiωtCel(ω)subscript𝐶𝑒𝑙𝑡subscriptsuperscript𝑑𝜔2𝜋superscript𝑒𝑖𝜔𝑡subscript𝐶𝑒𝑙𝜔C_{el}(t)=\int^{\infty}_{-\infty}(d\omega/2\pi)e^{-i\omega t}C_{el}(\omega)italic_C start_POSTSUBSCRIPT italic_e italic_l end_POSTSUBSCRIPT ( italic_t ) = ∫ start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT ( italic_d italic_ω / 2 italic_π ) italic_e start_POSTSUPERSCRIPT - italic_i italic_ω italic_t end_POSTSUPERSCRIPT italic_C start_POSTSUBSCRIPT italic_e italic_l end_POSTSUBSCRIPT ( italic_ω ). We show the results for the numerically computed Cel(t)subscript𝐶𝑒𝑙𝑡C_{el}(t)italic_C start_POSTSUBSCRIPT italic_e italic_l end_POSTSUBSCRIPT ( italic_t ) in Fig. 5(c) of the main text for (1) a flat conduction band (W=0𝑊0W=0italic_W = 0), and (2) a semicircular conduction band DOS, using Cgl(t)subscript𝐶𝑔𝑙𝑡C_{gl}(t)italic_C start_POSTSUBSCRIPT italic_g italic_l end_POSTSUBSCRIPT ( italic_t ) from Supplementary Eq. (11) (Fig. 5(b), main text). For the latter, we take a temperature-independent stretching exponent β=0.5𝛽0.5\beta=0.5italic_β = 0.5 and the α𝛼\alphaitalic_α-relaxation time τα(T)subscript𝜏𝛼𝑇\tau_{\alpha}(T)italic_τ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_T ), which varies as T2.8similar-toabsentsuperscript𝑇2.8\sim T^{-2.8}∼ italic_T start_POSTSUPERSCRIPT - 2.8 end_POSTSUPERSCRIPT, consistent with our experimental results (Fig. 2(c), main text). We also vary the coefficient B(T)1𝐵𝑇1B(T)\leq 1italic_B ( italic_T ) ≤ 1 such that it increases with decreasing temperature, while A(T)=1B(T)𝐴𝑇1𝐵𝑇A(T)=1-B(T)italic_A ( italic_T ) = 1 - italic_B ( italic_T ). This leads to a more dominant long-time stretched exponential part in Cgl(t)subscript𝐶𝑔𝑙𝑡C_{gl}(t)italic_C start_POSTSUBSCRIPT italic_g italic_l end_POSTSUBSCRIPT ( italic_t ) at lower temperatures, compared to the short-time exponential decay in Supplementary Eq. (11). In Supplementary Fig. 13, we show the results for the conduction band energy dispersion corresponding to a nearest-neighbour tight binding model on a simple cubic lattice. We can see that in all the cases the glassy relaxation of the bath influences the relaxation of the conduction electrons leading complex and temperature dependent relaxation profile for Cel(t)subscript𝐶𝑒𝑙𝑡C_{el}(t)italic_C start_POSTSUBSCRIPT italic_e italic_l end_POSTSUBSCRIPT ( italic_t ).

We can obtain a simple analytical understanding of the above results as follows. From Supplementary Eq. (19), we can obtain for μ=0𝜇0\mu=0italic_μ = 0

Refer to caption
Supplementary Figure 13: The density-density correlation function of conduction electron Cel(t)subscript𝐶𝑒𝑙𝑡C_{el}(t)italic_C start_POSTSUBSCRIPT italic_e italic_l end_POSTSUBSCRIPT ( italic_t ) vs. t𝑡titalic_t for three temperatures corresponding to a nearest-neighbour tight binding model on a simple cubic lattice
γ(ω)=ImΣR(ω)𝛾𝜔ImsubscriptΣ𝑅𝜔\displaystyle\gamma(\omega)={\rm Im}\Sigma_{R}(\omega)italic_γ ( italic_ω ) = roman_Im roman_Σ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_ω ) =πV2ρgl(ω+ε0)(nF(ε0)+nB(ε0ω))absent𝜋superscript𝑉2subscript𝜌𝑔𝑙𝜔subscript𝜀0subscript𝑛𝐹subscript𝜀0subscript𝑛𝐵subscript𝜀0𝜔\displaystyle=-\pi V^{2}\rho_{gl}(\omega+\varepsilon_{0})\big{(}n_{F}(-% \varepsilon_{0})+n_{B}(-\varepsilon_{0}-\omega)\big{)}= - italic_π italic_V start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ρ start_POSTSUBSCRIPT italic_g italic_l end_POSTSUBSCRIPT ( italic_ω + italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) ( italic_n start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( - italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) + italic_n start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ( - italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_ω ) ) (22)

Thus, by defining k(ω)=ReΣR(ω)𝑘𝜔ResubscriptΣ𝑅𝜔k(\omega)=\mathrm{Re}\Sigma_{R}(\omega)italic_k ( italic_ω ) = roman_Re roman_Σ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_ω ), we obtain from Supplementary Eq. (17a)

ImGR(ω)Imsubscript𝐺𝑅𝜔\displaystyle{\rm Im}G_{R}(\omega)roman_Im italic_G start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_ω ) =𝑑ϵg(ϵ)γ(ω)(ωϵk(ω))2+γ(ω)2absentdifferential-ditalic-ϵ𝑔italic-ϵ𝛾𝜔superscript𝜔italic-ϵ𝑘𝜔2𝛾superscript𝜔2\displaystyle=\int d\epsilon g(\epsilon)\frac{\gamma(\omega)}{(\omega-\epsilon% -k(\omega))^{2}+\gamma(\omega)^{2}}= ∫ italic_d italic_ϵ italic_g ( italic_ϵ ) divide start_ARG italic_γ ( italic_ω ) end_ARG start_ARG ( italic_ω - italic_ϵ - italic_k ( italic_ω ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_γ ( italic_ω ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG (23)

As a result, from Supplementary Eq. (20) we get

ImCel,R(ω)Imsubscript𝐶𝑒𝑙𝑅𝜔\displaystyle{\rm Im}C_{el,R}(\omega)roman_Im italic_C start_POSTSUBSCRIPT italic_e italic_l , italic_R end_POSTSUBSCRIPT ( italic_ω ) =𝑑ω1ρc(ω1)ρc(ω1ω)(nF(ω1)nF(ω1ω))absentdifferential-dsubscript𝜔1subscript𝜌𝑐subscript𝜔1subscript𝜌𝑐subscript𝜔1𝜔subscript𝑛𝐹subscript𝜔1subscript𝑛𝐹subscript𝜔1𝜔\displaystyle=\int d\omega_{1}\rho_{c}(\omega_{1})\rho_{c}(\omega_{1}-\omega)% \big{(}n_{F}(\omega_{1})-n_{F}(\omega_{1}-\omega)\big{)}= ∫ italic_d italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) italic_ρ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_ω ) ( italic_n start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) - italic_n start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_ω ) ) (24)

The above can be expressed as

ImCR(ω)Imsubscript𝐶𝑅𝜔\displaystyle{\rm Im}C_{R}(\omega)roman_Im italic_C start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( italic_ω ) =1π2dϵ1dϵ2g(ϵ1)g(ϵ2)dω1[γ(ω1)(ω1ϵ1k(ω1))2+γ(ω1)2\displaystyle=\frac{1}{\pi^{2}}\int d\epsilon_{1}d\epsilon_{2}g(\epsilon_{1})g% (\epsilon_{2})\int d\omega_{1}\bigg{[}\frac{\gamma(\omega_{1})}{(\omega_{1}-% \epsilon_{1}-k(\omega_{1}))^{2}+\gamma(\omega_{1})^{2}}= divide start_ARG 1 end_ARG start_ARG italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ∫ italic_d italic_ϵ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_d italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_g ( italic_ϵ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) italic_g ( italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) ∫ italic_d italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT [ divide start_ARG italic_γ ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) end_ARG start_ARG ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_ϵ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_k ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_γ ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG
γ(ω1ω)(ω1ωϵ2k(ω1ω))2+γ(ω1ω)2](nF(ω1)nF(ω1ω))\displaystyle\frac{\gamma(\omega_{1}-\omega)}{(\omega_{1}-\omega-\epsilon_{2}-% k(\omega_{1}-\omega))^{2}+\gamma(\omega_{1}-\omega)^{2}}\bigg{]}\big{(}n_{F}(% \omega_{1})-n_{F}(\omega_{1}-\omega)\big{)}divide start_ARG italic_γ ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_ω ) end_ARG start_ARG ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_ω - italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - italic_k ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_ω ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_γ ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_ω ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ] ( italic_n start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) - italic_n start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_ω ) ) (25)

Thus,

Cel(t)=𝑑ω1𝑑ωeıωtρgl(ω1+ε0)ρgl(ω1+ε0ω)F(ω,ω1)subscript𝐶𝑒𝑙𝑡superscriptsubscriptdifferential-dsubscript𝜔1differential-d𝜔superscript𝑒italic-ı𝜔𝑡subscript𝜌𝑔𝑙subscript𝜔1subscript𝜀0subscript𝜌𝑔𝑙subscript𝜔1subscript𝜀0𝜔𝐹𝜔subscript𝜔1\displaystyle C_{el}(t)=\int_{-\infty}^{\infty}d\omega_{1}d\omega e^{-\imath% \omega t}\rho_{gl}(\omega_{1}+\varepsilon_{0})\rho_{gl}(\omega_{1}+\varepsilon% _{0}-\omega)F(\omega,\omega_{1})italic_C start_POSTSUBSCRIPT italic_e italic_l end_POSTSUBSCRIPT ( italic_t ) = ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_d italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_d italic_ω italic_e start_POSTSUPERSCRIPT - italic_ı italic_ω italic_t end_POSTSUPERSCRIPT italic_ρ start_POSTSUBSCRIPT italic_g italic_l end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) italic_ρ start_POSTSUBSCRIPT italic_g italic_l end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_ω ) italic_F ( italic_ω , italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) (26)

can be written as a convolution over the glass spectral function ρgl(ω)subscript𝜌𝑔𝑙𝜔\rho_{gl}(\omega)italic_ρ start_POSTSUBSCRIPT italic_g italic_l end_POSTSUBSCRIPT ( italic_ω ), where

F(ω,ω1)𝐹𝜔subscript𝜔1\displaystyle F(\omega,\omega_{1})italic_F ( italic_ω , italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) =V42π𝑑ϵ1𝑑ϵ2g(ϵ1)g(ϵ2)coth(ω/2T)[nF(ω1)nF(ω1ω)][(ω1ϵ1k(ω1))2+γ(ω1)2][(ω1ωϵ2k(ω1ω))2+γ(ω1ω)2]absentsuperscript𝑉42𝜋differential-dsubscriptitalic-ϵ1differential-dsubscriptitalic-ϵ2𝑔subscriptitalic-ϵ1𝑔subscriptitalic-ϵ2hyperbolic-cotangent𝜔2𝑇delimited-[]subscript𝑛𝐹subscript𝜔1subscript𝑛𝐹subscript𝜔1𝜔delimited-[]superscriptsubscript𝜔1subscriptitalic-ϵ1𝑘subscript𝜔12𝛾superscriptsubscript𝜔12delimited-[]superscriptsubscript𝜔1𝜔subscriptitalic-ϵ2𝑘subscript𝜔1𝜔2𝛾superscriptsubscript𝜔1𝜔2\displaystyle=\frac{V^{4}}{2\pi}\int d\epsilon_{1}d\epsilon_{2}\frac{g(% \epsilon_{1})g(\epsilon_{2})\coth\left(\omega/2T\right)[n_{F}(\omega_{1})-n_{F% }(\omega_{1}-\omega)]}{[(\omega_{1}-\epsilon_{1}-k(\omega_{1}))^{2}+\gamma(% \omega_{1})^{2}][(\omega_{1}-\omega-\epsilon_{2}-k(\omega_{1}-\omega))^{2}+% \gamma(\omega_{1}-\omega)^{2}]}= divide start_ARG italic_V start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_π end_ARG ∫ italic_d italic_ϵ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_d italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT divide start_ARG italic_g ( italic_ϵ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) italic_g ( italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) roman_coth ( italic_ω / 2 italic_T ) [ italic_n start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) - italic_n start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_ω ) ] end_ARG start_ARG [ ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_ϵ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_k ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_γ ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] [ ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_ω - italic_ϵ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - italic_k ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_ω ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_γ ( italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_ω ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] end_ARG
(nF(ε0)+nB(ε0ω1))(nF(ε0)+nB(ε0ω1+ω))subscript𝑛𝐹subscript𝜀0subscript𝑛𝐵subscript𝜀0subscript𝜔1subscript𝑛𝐹subscript𝜀0subscript𝑛𝐵subscript𝜀0subscript𝜔1𝜔\displaystyle\big{(}n_{F}(-\varepsilon_{0})+n_{B}(-\varepsilon_{0}-\omega_{1})% \big{)}\big{(}n_{F}(-\varepsilon_{0})+n_{B}(-\varepsilon_{0}-\omega_{1}+\omega% )\big{)}( italic_n start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( - italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) + italic_n start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ( - italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) ) ( italic_n start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( - italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) + italic_n start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ( - italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_ω ) ) (27)

The spectral function of the glass contains the information about the multiple time scales and the non-trivial temperature dependence of the glassy relaxation. Thus if appropriate conditions on the electronic energy scales W𝑊Witalic_W and ε0subscript𝜀0\varepsilon_{0}italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (ΔΔ\Deltaroman_Δ) are made relative to the energy scales of the glass, e.g., τs1superscriptsubscript𝜏𝑠1\tau_{s}^{-1}italic_τ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT and τα1superscriptsubscript𝜏𝛼1\tau_{\alpha}^{-1}italic_τ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, then the complex relaxation of the conduction electrons can be obtained. We numerically find that these conditions are met if the glassy bath is broad, i.e., the bandwidth of the glass is comparable or larger than the electronic energy scales.

References

  • [1] Mattheiss, L. F. Energy bands for knif3subscriptf3{\mathrm{f}}_{3}roman_f start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT, srtio3subscripto3{\mathrm{o}}_{3}roman_o start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT, kmoo3subscripto3{\mathrm{o}}_{3}roman_o start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT, and ktao3subscripto3{\mathrm{o}}_{3}roman_o start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT. Phys. Rev. B 6, 4718–4740 (1972). URL https://link.aps.org/doi/10.1103/PhysRevB.6.4718.
  • [2] Ojha, S. K. et al. Oxygen vacancy-induced topological hall effect in a nonmagnetic band insulator. Advanced Quantum Technologies 3, 2000021 (2020). URL https://doi.org/10.1002/qute.202000021.
  • [3] Herranz, G. et al. High mobility in laalo3/srtio3subscriptlaalo3subscriptsrtio3{\mathrm{laalo}}_{3}/{\mathrm{srtio}}_{3}roman_laalo start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT / roman_srtio start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT heterostructures: Origin, dimensionality, and perspectives. Phys. Rev. Lett. 98, 216803 (2007). URL https://link.aps.org/doi/10.1103/PhysRevLett.98.216803.
  • [4] Uwe, H., Kinoshita, J., Yoshihiro, K., Yamanouchi, C. & Sakudo, T. Evidence for light and heavy conduction electrons at the zone center in ktao3subscripto3{\mathrm{o}}_{3}roman_o start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT. Phys. Rev. B 19, 3041–3044 (1979). URL https://link.aps.org/doi/10.1103/PhysRevB.19.3041.
  • [5] Herranz, G. et al. Full oxide heterostructure combining a high-Tcsubscript𝑇c{T}_{\mathrm{c}}italic_T start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT diluted ferromagnet with a high-mobility conductor. Phys. Rev. B 73, 064403 (2006). URL https://link.aps.org/doi/10.1103/PhysRevB.73.064403.
  • [6] Lin, X. et al. Metallicity without quasi-particles in room-temperature strontium titanate. npj Quantum Materials 2, 41 (2017). URL https://doi.org/10.1038/s41535-017-0044-5.
  • [7] Collignon, C., Bourges, P., Fauqué, B. & Behnia, K. Heavy nondegenerate electrons in doped strontium titanate. Phys. Rev. X 10, 031025 (2020). URL https://link.aps.org/doi/10.1103/PhysRevX.10.031025.
  • [8] Ashcroft, N. W. & Mermin, N. D. Solid state physics (Cengage Learning, 2022).
  • [9] Kob, W. Supercooled liquids, the glass transition, and computer simulations. lecture notes” slow relaxation and nonequilibrium dynamics in condensed matter. In Les Houches Session, vol. 77 (2002).
  • [10] Lang, D. V. & Logan, R. A. Large-lattice-relaxation model for persistent photoconductivity in compound semiconductors. Phys. Rev. Lett. 39, 635–639 (1977). URL https://link.aps.org/doi/10.1103/PhysRevLett.39.635.
  • [11] Grenier, P., Bernier, G., Jandl, S., Salce, B. & Boatner, L. A. Fluorescence and ferroelectric microregions in ktao3. Journal of Physics: Condensed Matter 1, 2515 (1989). URL https://dx.doi.org/10.1088/0953-8984/1/14/007.
  • [12] Geifman, I. N. & Golovina, I. S. Discussion about the nature of polarized microregions in ktao3. Ferroelectrics 199, 115–120 (1997). URL https://doi.org/10.1080/00150199708213433.
  • [13] Voigt, P. & Kapphan, S. Experimental study of second harmonic generation by dipolar configurations in pure and li-doped ktao3 and its variation under electric field. Journal of Physics and Chemistry of Solids 55, 853–869 (1994). URL https://www.sciencedirect.com/science/article/pii/0022369794900108.
  • [14] Grenier, P., Jandl, S., Blouin, M. & Boatner, L. A. Study of ferroelectric microdomains due to oxygen vacancies in ktao3. Ferroelectrics 137, 105–111 (1992). URL https://doi.org/10.1080/00150199208015942.
  • [15] Trybuła, Z., Miga, S., Łoś, S., Trybuła, M. & Dec, J. Evidence of polar nanoregions in quantum paraelectric ktao3. Solid State Communications 209-210, 23–26 (2015). URL https://www.sciencedirect.com/science/article/pii/S0038109815000782.
  • [16] Tkach, A., Zlotnik, S. & Vilarinho, P. M. Dielectric response of ktao3 single crystals weakly co-doped with li and mn (2021). URL https://doi.org/10.3390/cryst11101222.
  • [17] Samara, G. A. The relaxational properties of compositionally disordered abo¡sub¿3¡/sub¿perovskites. Journal of Physics: Condensed Matter 15, R367–R411 (2003). URL https://doi.org/10.1088/0953-8984/15/9/202.
  • [18] Shirane, G., Nathans, R. & Minkiewicz, V. J. Temperature dependence of the soft ferroelectric mode in ktao3subscripto3{\mathrm{o}}_{3}roman_o start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT. Phys. Rev. 157, 396–399 (1967). URL https://link.aps.org/doi/10.1103/PhysRev.157.396.
  • [19] Nilsen, W. G. & Skinner, J. G. Raman spectrum of potassium tantalate. The Journal of Chemical Physics 47, 1413–1418 (2004). URL https://doi.org/10.1063/1.1712096.
  • [20] Perry, C. H., Fertel, J. H. & McNelly, T. F. Temperature dependence of the raman spectrum of srtio3 and ktao3. The Journal of Chemical Physics 47, 1619–1625 (2004). URL https://doi.org/10.1063/1.1712142.
  • [21] Uwe, H., Lyons, K. B., Carter, H. L. & Fleury, P. A. Ferroelectric microregions and raman scattering in ktao3subscriptktao3{\mathrm{ktao}}_{3}roman_ktao start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT. Phys. Rev. B 33, 6436–6440 (1986). URL https://link.aps.org/doi/10.1103/PhysRevB.33.6436.
  • [22] Golovina, I. S. et al. Defect driven ferroelectricity and magnetism in nanocrystalline ktao3. Physica B: Condensed Matter 407, 614–623 (2012). URL https://www.sciencedirect.com/science/article/pii/S0921452611011707.
  • [23] Kubacki, J., Molak, A., Rogala, M., Rodenbücher, C. & Szot, K. Metal–insulator transition induced by non-stoichiometry of surface layer and molecular reactions on single crystal ktao3. Surface Science 606, 1252–1262 (2012). URL https://www.sciencedirect.com/science/article/pii/S0039602812001252.
  • [24] Pal, B., Mukherjee, S. & Sarma, D. D. Probing complex heterostructures using hard x-ray photoelectron spectroscopy (haxpes). Journal of Electron Spectroscopy and Related Phenomena 200, 332–339 (2015). URL https://www.sciencedirect.com/science/article/pii/S0368204815001334.
  • [25] Jang, H. W. et al. Ferroelectricity in strain-free srtio3subscriptsrtio3{\mathrm{srtio}}_{3}roman_srtio start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT thin films. Phys. Rev. Lett. 104, 197601 (2010). URL https://link.aps.org/doi/10.1103/PhysRevLett.104.197601.
  • [26] Miao, L. et al. Double-bilayer polar nanoregions and mn antisites in (ca, sr)3mn2o7. Nature Communications 13, 4927 (2022). URL https://doi.org/10.1038/s41467-022-32090-w.
  • [27] Perdew, J. P., Burke, K. & Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865–3868 (1996). URL https://link.aps.org/doi/10.1103/PhysRevLett.77.3865.
  • [28] King-Smith, R. D. & Vanderbilt, D. Theory of polarization of crystalline solids. Phys. Rev. B 47, 1651–1654 (1993). URL https://link.aps.org/doi/10.1103/PhysRevB.47.1651.
  • [29] Resta, R. Macroscopic polarization in crystalline dielectrics: the geometric phase approach. Rev. Mod. Phys. 66, 899–915 (1994). URL https://link.aps.org/doi/10.1103/RevModPhys.66.899.
  • [30] Berry, M. V. Quantal phase factors accompanying adiabatic changes. Proceedings of the Royal Society of London. A. Mathematical and Physical Sciences 392, 45–57 (1984). URL https://royalsocietypublishing.org/doi/abs/10.1098/rspa.1984.0023. eprint https://royalsocietypublishing.org/doi/pdf/10.1098/rspa.1984.0023.
  • [31] Wang, X., Yates, J. R., Souza, I. & Vanderbilt, D. Ab initio calculation of the anomalous hall conductivity by wannier interpolation. Phys. Rev. B 74, 195118 (2006). URL https://link.aps.org/doi/10.1103/PhysRevB.74.195118.
  • [32] Pizzi, G. et al. Wannier90 as a community code: new features and applications. Journal of Physics: Condensed Matter 32, 165902 (2020). URL https://dx.doi.org/10.1088/1361-648X/ab51ff.
  • [33] Cugliandolo, L. F., Grempel, D. R. & da Silva Santos, C. A. Imaginary-time replica formalism study of a quantum spherical p-spin-glass model. Phys. Rev. B 64, 014403 (2001). URL https://link.aps.org/doi/10.1103/PhysRevB.64.014403.
  • [34] Crisanti, A. & Sommers, H. J. The sphericalp-spin interaction spin glass model: the statics. Zeitschrift fur Physik B Condensed Matter 87, 341–354 (1992).
  • [35] Kob, W. The Mode-Coupling Theory of the Glass Transition. arXiv e-prints cond–mat/9702073 (1997). eprint cond-mat/9702073.
  • [36] Marinari, E., Parisi, G. & Ritort, F. Replica field theory for deterministic models. ii. a non-random spin glass with glassy behaviour 27, 7647 (1994). URL https://dx.doi.org/10.1088/0305-4470/27/23/011.
  • [37] Banerjee, S. & Altman, E. Solvable model for a dynamical quantum phase transition from fast to slow scrambling. Phys. Rev. B 95, 134302 (2017). URL https://link.aps.org/doi/10.1103/PhysRevB.95.134302.
  • [38] Bera, S., Venkata Lokesh, K. Y. & Banerjee, S. Quantum-to-classical crossover in many-body chaos and scrambling from relaxation in a glass. Phys. Rev. Lett. 128, 115302 (2022). URL https://link.aps.org/doi/10.1103/PhysRevLett.128.115302.