1 Overview of the powder sample and the pump-probe setup. a, Crystal structure of the tetragonal LLTO in equilibrium. Graphics rendered by VESTA . b, Schematic of the experimental setup for synchrotron-based time-resolved hard X-ray micro-diffraction (see (2024)

Hidden correlations in stochastic photoinduced dynamics of a solid-state electrolyte

JacksonMcClellan,1, 2, 3,123{}^{1,\,2,\,3,\,{\color[rgb]{0,0,0.80}\ast}}start_FLOATSUPERSCRIPT 1 , 2 , 3 , ∗ end_FLOATSUPERSCRIPT AlfredZong,1, 3,,13{}^{1,\,3,\,{\color[rgb]{0,0,0.80}\ast,\,\text{{\char 12\relax}}}}start_FLOATSUPERSCRIPT 1 , 3 , ∗ , ✉ end_FLOATSUPERSCRIPT KimH.Pham,4 HanzheLiu,4 ZacheryW.B.Iton,4 BurakGuzelturk,5 DonaldA.Walko,5 HaidanWen,5, 6 ScottK.Cushing,4,4{}^{4,\,{\color[rgb]{0,0,0.80}\text{{\char 12\relax}}}}start_FLOATSUPERSCRIPT 4 , ✉ end_FLOATSUPERSCRIPT MichaelW.Zuerch,1, 3,13{}^{1,\,3,\,{\color[rgb]{0,0,0.80}\text{{\char 12\relax}}}}start_FLOATSUPERSCRIPT 1 , 3 , ✉ end_FLOATSUPERSCRIPT

(Dated: June 10, 2024)

1Department of Chemistry, University of California, Berkeley, CA 94720, USA.

2Department of Chemistry and Biochemistry, The Ohio State University, Columbus, OH 43210, USA.

3Materials Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, CA 94720, USA.

4Division of Chemistry and Chemical Engineering, California Institute of Technology, Pasadena, CA 91125, USA.

5Advanced Photon Source, Argonne National Laboratory, 9700 South Cass Avenue, Argonne, IL 60439, USA.

6Materials Science Division, Argonne National Laboratory, 9700 South Cass Avenue, Argonne, IL 60439, USA.

{}^{\color[rgb]{0,0,0.80}\ast\,}start_FLOATSUPERSCRIPT ∗ end_FLOATSUPERSCRIPTThese authors contributed equally to this work: JacksonMcClellan and AlfredZong.

{}^{\color[rgb]{0,0,0.80}\text{{\char 12\relax}}\,}start_FLOATSUPERSCRIPT ✉ end_FLOATSUPERSCRIPTCorrespondence to: A.Z. (alfredz@berkeley.edu), S.K.C. (scushing@caltech.edu), and M.W.Z. (mwz@berkeley.edu).

Abstract: Photoexcitation by ultrashort laser pulses plays a crucial role in controlling reaction pathways, creating nonequilibrium material properties, and offering a microscopic view of complex dynamics at the molecular level. The photo response following a laser pulse is, in general, non-identical between multiple exposures due to spatiotemporal fluctuations in a material or the stochastic nature of dynamical pathways. However, most ultrafast experiments using a stroboscopic pump-probe scheme struggle to distinguish intrinsic sample fluctuations from extrinsic apparatus noise, often missing seemingly random deviations from the averaged shot-to-shot response. Leveraging the stability and high photon-flux of time-resolved X-ray micro-diffraction at a synchrotron, we developed a method to quantitatively characterize the shot-to-shot variation of the photoinduced dynamics in a solid-state electrolyte. By analyzing temporal evolutions of the lattice parameter of a single grain in a powder ensemble, we found that the sample responses after different shots contain random fluctuations that are, however, not independent. Instead, there is a correlation between the nonequilibrium lattice trajectories following adjacent laser shots with a characteristic “correlation length” of approximately 1,500 shots, which represents an energy barrier of 0.38eV for switching the photoinduced pathway, a value interestingly commensurate with the activation energy of lithium ion diffusion. Not only does our nonequilibrium noise correlation spectroscopy provide a new strategy for studying fluctuations that are central to phase transitions in both condensed matter and molecular systems, it also paves the way for discovering hidden correlations and novel metastable states buried in oft-presumed random, uncorrelated fluctuating dynamics.

1 Overview of the powder sample and the pump-probe setup. a, Crystal structure of the tetragonal LLTO in equilibrium. Graphics rendered by VESTA . b, Schematic of the experimental setup for synchrotron-based time-resolved hard X-ray micro-diffraction (see Methods). c, Scanning electron micrograph of the powder LLTO sample pressed into a pellet, which was used in the time-resolved X-ray experiment. d, Raw diffraction image of the (004) Bragg peak of a particular crystalline grain recorded before pump pulse arrival. 𝐪_⟂ is along the 𝑐-axis of the grain while 𝐪_∥ lies in the 𝑎-𝑏 plane of the grain. (1)

Even though rapid advances in laser technology have enabled the detection of ultrafast dynamics in biological systems 2, quantum materials 3, 4, and chemical reactions 5 down to the attosecond regime, probing stochastic processes that are intrinsic in these systems remains a conundrum. In cases where random fluctuations and nondeterministic quasiparticle motions dominate the dynamics of interest, such as near the critical point of a phase transition or in a disordered medium riddled with topological defects, a conventional stroboscopic pump-probe scheme is ineffective as it averages out shot-to-shot variations in the pump-induced response. Even if the probe signal achieves a sufficiently high signal-to-noise ratio in a non-stroboscopic single-shot pump-probe measurement 6, 7, 8, typically only the data point at one pump-probe delay is collected after a single pump shot, rendering it impossible to reconstruct the stochastic dynamics over time. Alternatively, a pump pulse may be followed by a train of probe pulselets via echelon-based optics 9, 10, 11 to characterize the entire pump-induced evolution. However, the probe signal spread over individual pulselets can be too weak to yield useful information about the dynamics after only one pump shot, and extensive averaging over multiple pump shots are often necessary. Fundamentally, there is an upper limit on how strong the probe pulse (or pulselets) can be so that it does not significantly alter the dynamics under investigation. This limit hence imposes a trade-off between the signal-to-noise ratio and how many delay times can be measured, preventing access to the full, stochastic dynamics that differ from one pump shot to another.

Here, we tackle this challenge in studying stochastic ultrafast dynamics via a new statistical approach that is capable of uncovering hidden correlations between individual pump-induced events. A key to this method relies on the low noise level associated with the data collection process that can be plagued by instrumental instability or low photon counts in a table-top ultrafast laser-based measurements. In our experiment, we overcome these problems by utilizing the stability and high photon flux of a synchrotron-based time-resolved X-ray setup such that the experimentally observed signal variation is dominated by the intrinsic sample fluctuations in the photoinduced response. Unlike previous approaches that necessitate a single-shot setup 12, 13, our method is deployed to a typical stroboscopic pump-probe scheme and only necessitates repeated data acquisitions of the same pump-probe delay with a stable light source and high measurement statistics. We term this method nonequilibrium noise correlation spectroscopy, which can have broad applicability to a wide range of time-resolved experiments. We expect this approach to yield a deeper understanding of the key role of disordered fluctuations not just in lattice but also in charge, spin, and orbital degrees of freedom that are inherent in the nonequilibrium photoinduced trajectories 14, 15, 16, 17.

To demonstrate this capability of unraveling dynamical correlations in the transient stochastic response, we chose to study Li0.5La0.5TiO3 (LLTO), a highly conductive solid-state electrolyte candidate for lithium-ion batteries 18, 19, 20, 21. LLTO has a perovskite crystal structure (Fig.1a) with alternating vacancy-rich and poor layers 22, where the low diffusion barrier leads to high lithium ion conductivity 23. The mobility of lithium ions via LLTO is highly dependent on the crystal structure 24, which changes substantially upon lithium ion insertion and extraction as observed in X-ray diffraction 23. Conversely, if the crystal structure changes upon photoexcitation due to nonthermal phonon population 25, 26 and transient heat deposition 27, 28, one also expects varying dynamics from shot to shot due to the stochastic location of the lithium ions in the crystal, which are free to migrate either due to thermal activation 29 or photoexcitation 30. Therefore, the highly mobile lithium ions in LLTO offer the ideal platform to investigate stochastic fluctuations in the lattice dynamics following photoexcitation, which can in turn yield insight into the strong coupling between lithium ion diffusion and the crystalline lattice 24, 31 for designing next-generation solid-state electrolyte.

To this end, we monitored the evolution of lattice parameters of LLTO following above-bandgap photoexcitation using synchrotron-based X-ray diffraction (see Methods for details). An above-bandgap excitation in LLTO is expected to cause heat-induced initial lattice expansion and subsequent relaxation that modulate lithium site occupations because it is known that the activation energy of ionic transport is closely tied to lattice distortions 31. Importantly, we uncovered a hidden correlation between lattice motions following neighboring pump shots with a characteristic “correlation length” of approximately 1,500 laser shots, which correspond to a similar-to\sim0.38eV energy barrier. The energy barrier for lithium ion conduction in LLTO is similar in magnitude, so the stochastic lattice trajectories may give insight into the photoswitched ion diffusion mechanism. The novel framework of nonequilibrium noise correlation spectroscopy introduced in this work provides new understanding in the potential role of photo-induced structural change in the lithium-ion conduction process, opening the avenue for harnessing tailored laser pulses for manipulating ionic conduction in solids.

1 Overview of the powder sample and the pump-probe setup. a, Crystal structure of the tetragonal LLTO in equilibrium. Graphics rendered by VESTA . b, Schematic of the experimental setup for synchrotron-based time-resolved hard X-ray micro-diffraction (see Methods). c, Scanning electron micrograph of the powder LLTO sample pressed into a pellet, which was used in the time-resolved X-ray experiment. d, Raw diffraction image of the (004) Bragg peak of a particular crystalline grain recorded before pump pulse arrival. 𝐪_⟂ is along the 𝑐-axis of the grain while 𝐪_∥ lies in the 𝑎-𝑏 plane of the grain. (2)

I Time-resolved X-ray micro-diffraction

A schematic of the setup is shown in Fig.1b (see Methods for a more detailed description). As industrial-scale solid-state ionic conductors are often synthesized as sintered, pressed powder pellets 32, the LLTO sample under investigation was prepared in a similar polycrystalline form, whose typical grain size can be up to a few micrometers (Fig.1c). The high momentum-resolution of the setup as well as the comparable X-ray beam spot size and LLTO grain size make it possible to observe individual Bragg peaks instead of Debye-Scherrer rings (Fig.1d), enabling us to focus on the stochastic lattice dynamics of a single grain without an averaging effect within the powder ensemble.

To understand how the experiment is sensitive to shot-to-shot variation of the lattice response despite a conventional stroboscopic pump-probe scheme, it is worth revisiting the measurement protocol, summarized in Fig.2a. The repetition rate of the ultraviolet pump laser was 1kHz, and the X-ray diffraction peak was measured at each pump-probe delay time ti=1,,59subscript𝑡𝑖159t_{i=1,\dots,59}italic_t start_POSTSUBSCRIPT italic_i = 1 , … , 59 end_POSTSUBSCRIPT sequentially, where diffraction intensities of n𝑛nitalic_n pump-probe pulse pairs were averaged for each tisubscript𝑡𝑖t_{i}italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT during the delay time scan. Here, the dwell time at each delay was 1s, leading to n1,000𝑛1000n\approx 1,000italic_n ≈ 1 , 000. Upon the completion of one full pump-probe delay scan, we repeated the procedure for a total of N𝑁Nitalic_N scans (N=10𝑁10N=10italic_N = 10) with nearly zero waiting time in between successive scans. In general, the sample response after every single pump laser pulse can follow different trajectories due to fluctuations (Fig.2b, upper panel). Even though we only captured a single point out of the entire photoinduced evolution at each delay time tisubscript𝑡𝑖t_{i}italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT (solid dots in the upper panel of Fig.2b), as illustrated in the lower panel of Fig.2b, we can still detect the stochastic variation between certain shots based on abrupt discontinuities in the recorded time trace (highlighted by the black arrow), especially if such discontinuities occur long after the region of pump-probe temporal overlap and if they are much larger than the noise level of the measurements (see ref.33 for noise estimates).

These discontinuities are clearly observed in our measurements. Figure2c shows the c𝑐citalic_c-axis lattice parameter as a function of pump-probe delay time for ten scans, two of which are shown in Fig.2d (see ref.33 for the extraction of c𝑐citalic_c from diffraction images). Upon photoexcitation, on average, the lattice exhibits a sudden c𝑐citalic_c-axis expansion of more than 0.1% followed by a slow recovery over tens of microseconds (Fig.3a). For individual scans, however, the c𝑐citalic_c value experiences seemingly random discontinuities in the time trace after pump pulse arrival. By contrast, the variation of c𝑐citalic_c before photoexcitation is much smaller (e.g., 25-times smaller in scan7 in Fig.2d; see ref.33 for noise estimates). This dichotomy of the noise level before and after t=0𝑡0t=0italic_t = 0 excludes extrinsic measurement uncertainties due to factors such as instability of the X-ray beam, which is expected to yield a similar noise level at all time delays. These discontinuities hence suggest that indeed the sample response after each pump shot varies significantly, an observation further substantiated by the large contrast of the transient c𝑐citalic_c-axis parameters at a fixed time delay across different scans (Fig.2e).

1 Overview of the powder sample and the pump-probe setup. a, Crystal structure of the tetragonal LLTO in equilibrium. Graphics rendered by VESTA . b, Schematic of the experimental setup for synchrotron-based time-resolved hard X-ray micro-diffraction (see Methods). c, Scanning electron micrograph of the powder LLTO sample pressed into a pellet, which was used in the time-resolved X-ray experiment. d, Raw diffraction image of the (004) Bragg peak of a particular crystalline grain recorded before pump pulse arrival. 𝐪_⟂ is along the 𝑐-axis of the grain while 𝐪_∥ lies in the 𝑎-𝑏 plane of the grain. (3)

Upon closer scrutiny, two features of the stochastic response reflected in these discontinuities stand out in Fig.2c–e. First, the relative magnitude of a shift in c𝑐citalic_c is larger right after t=0𝑡0t=0italic_t = 0 compared to later pump-probe delays. This feature is further echoed by the larger scan-to-scan variation at t=3μ𝑡3μt=3~{}\upmuitalic_t = 3 roman_μs (blue triangles) compared to t=20μ𝑡20μt=20~{}\upmuitalic_t = 20 roman_μs and 32μμ\upmuroman_μs (red and green triangles) in Fig.2e. Second, the discontinuities in the time traces are often preceded or succeeded by a streak of similar values, as highlighted in dashed boxes in Fig.2c,d. The presence of these streaks are unexpected if the stochastic lattice dynamics following each pump pulse are independent from one another. The streaks hence hint at some correlated photoinduced dynamics in LLTO. We will quantify both features of the stochastic response in the statistical analyses below.

1 Overview of the powder sample and the pump-probe setup. a, Crystal structure of the tetragonal LLTO in equilibrium. Graphics rendered by VESTA . b, Schematic of the experimental setup for synchrotron-based time-resolved hard X-ray micro-diffraction (see Methods). c, Scanning electron micrograph of the powder LLTO sample pressed into a pellet, which was used in the time-resolved X-ray experiment. d, Raw diffraction image of the (004) Bragg peak of a particular crystalline grain recorded before pump pulse arrival. 𝐪_⟂ is along the 𝑐-axis of the grain while 𝐪_∥ lies in the 𝑎-𝑏 plane of the grain. (4)

II Statistical analysis of lattice dynamics

We first examine the amplitude of the scan-to-scan c𝑐citalic_c value variation that appears to decay as a function of delay time. Figure3b confirms this observation, where the standard deviation csdsubscript𝑐sdc_{\text{sd}}italic_c start_POSTSUBSCRIPT sd end_POSTSUBSCRIPT computed across scans exhibits a remarkably similar temporal evolution as that of the scan-averaged dynamics (cavgsubscript𝑐avgc_{\text{avg}}italic_c start_POSTSUBSCRIPT avg end_POSTSUBSCRIPT) in Fig.3a. Both cavgsubscript𝑐avgc_{\text{avg}}italic_c start_POSTSUBSCRIPT avg end_POSTSUBSCRIPT and csdsubscript𝑐sdc_{\text{sd}}italic_c start_POSTSUBSCRIPT sd end_POSTSUBSCRIPT can be well fitted to a phenomenological model (black curves in Fig.3a,b), which captures essential features of a generic photoinduced response that approximately follows first-order kinetics during long-term recovery (with time constants τavgsubscript𝜏avg\tau_{\text{avg}}italic_τ start_POSTSUBSCRIPT avg end_POSTSUBSCRIPT and τsdsubscript𝜏sd\tau_{\text{sd}}italic_τ start_POSTSUBSCRIPT sd end_POSTSUBSCRIPT)34, 35, 36,

f(t)=fequil+[12(1+erf(2ln2(tt0)w))\displaystyle f(t)=f_{\text{equil}}+\bigg{[}\frac{1}{2}\bigg{(}1+\text{erf}%\bigg{(}\frac{2\sqrt{\ln 2}(t-t_{0})}{w}\bigg{)}\bigg{)}italic_f ( italic_t ) = italic_f start_POSTSUBSCRIPT equil end_POSTSUBSCRIPT + [ divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( 1 + erf ( divide start_ARG 2 square-root start_ARG roman_ln 2 end_ARG ( italic_t - italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) end_ARG start_ARG italic_w end_ARG ) )
(I+I0e(tt0)/τ)],\displaystyle\cdot\big{(}I_{\infty}+I_{0}e^{-(t-t_{0})/\tau}\big{)}\bigg{]},⋅ ( italic_I start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT + italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - ( italic_t - italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) / italic_τ end_POSTSUPERSCRIPT ) ] ,(1)

where f(t)𝑓𝑡f(t)italic_f ( italic_t ) is the observable of interest that depends on the pump-probe delay time t𝑡titalic_t obtained from the electronic delay signal, and fequilsubscript𝑓equilf_{\text{equil}}italic_f start_POSTSUBSCRIPT equil end_POSTSUBSCRIPT is its equilibrium value prior to photoexcitation. w𝑤witalic_w characterizes the initial system response time, t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT determines the pump pulse arrival time, I0subscript𝐼0I_{0}italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT denotes the change right after photoexcitation while Isubscript𝐼I_{\infty}italic_I start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT denotes the value after the system partially relaxes, a process characterized by a time constant τ𝜏\tauitalic_τ.

On the phenomenological level, the sudden increase of csd(t)subscript𝑐sd𝑡c_{\text{sd}}(t)italic_c start_POSTSUBSCRIPT sd end_POSTSUBSCRIPT ( italic_t ) after t=0𝑡0t=0italic_t = 0 is indicative of the variation of the extent of the photoinduced lattice expansion, captured by I0subscript𝐼0I_{0}italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT in Eq.(1). The physical origin of such variations in I0subscript𝐼0I_{0}italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT can be the randomness of the strain environment 37 following the mechanical relaxation of neighboring grains from the previous laser shot, the changing absorbed fluence due to varying light attenuation through a grainy medium where different grain orientations in the vicinity lead to different degrees of scattering, or a combination of such factors. However, a careful comparison between the dynamics of cavg(t)subscript𝑐avg𝑡c_{\text{avg}}(t)italic_c start_POSTSUBSCRIPT avg end_POSTSUBSCRIPT ( italic_t ) and csd(t)subscript𝑐sd𝑡c_{\text{sd}}(t)italic_c start_POSTSUBSCRIPT sd end_POSTSUBSCRIPT ( italic_t ) in Fig.3a,b indicates that I0subscript𝐼0I_{0}italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT cannot be the only stochastic element that varies from shot to shot. Specifically, the relaxation time τavgsubscript𝜏avg\tau_{\text{avg}}italic_τ start_POSTSUBSCRIPT avg end_POSTSUBSCRIPT is more than twice of τsdsubscript𝜏sd\tau_{\text{sd}}italic_τ start_POSTSUBSCRIPT sd end_POSTSUBSCRIPT. This difference in τ𝜏\tauitalic_τ means that the c𝑐citalic_c-axis parameter becomes more consistent in value at larger delay times, suggesting a correlation between I0subscript𝐼0I_{0}italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and τ𝜏\tauitalic_τ for individual pump shot-induced response. Indeed, when we simulate a large number of lattice responses by drawing I0subscript𝐼0I_{0}italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT randomly from a Gaussian distribution, we can reproduce the τavg>τsdsubscript𝜏avgsubscript𝜏sd\tau_{\text{avg}}>\tau_{\text{sd}}italic_τ start_POSTSUBSCRIPT avg end_POSTSUBSCRIPT > italic_τ start_POSTSUBSCRIPT sd end_POSTSUBSCRIPT relation (Fig.3c,d) if we impose a negative correlation between I0subscript𝐼0I_{0}italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and τ𝜏\tauitalic_τ, where the exact functional form of the negative correlation is not critical (see ref.33 for more details of the simulation).

Next, we address the origin of the streaks in Fig.2c,d, which suggest that variations in the lattice dynamics after each pump laser shot are not independent. To obtain a quantitative measure of the hidden dynamical correlation, we examine how the c𝑐citalic_c-axis parameter within a scan at one particular delay time tisubscript𝑡𝑖t_{i}italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is correlated with its value at another delay time tjsubscript𝑡𝑗t_{j}italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT. Even though we use delay times ti,jsubscript𝑡𝑖𝑗t_{i,j}italic_t start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT as a convenient notation, the correlation we are interested in is defined in between two lattice responses following two pump shots that arrive at different lab times; under our experimental condition, a delay difference |titj|subscript𝑡𝑖subscript𝑡𝑗|t_{i}-t_{j}|| italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT | of 1μμ\upmuroman_μs corresponds to approximately 1,000 pump shots elapsed (see measurement scheme in Fig.2a). To the lowest order, we assume that a linear correlation can capture the relation, if any, between c(ti)𝑐subscript𝑡𝑖c(t_{i})italic_c ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) and c(tj)𝑐subscript𝑡𝑗c(t_{j})italic_c ( italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) across different pump shots. This assumption is expected to hold because a larger value of c(ti)𝑐subscript𝑡𝑖c(t_{i})italic_c ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) at an early time delay indicates a larger initial lattice expansion, which in turn leads to a larger value of c(tj)𝑐subscript𝑡𝑗c(t_{j})italic_c ( italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) during its microsecond relaxation period, where no coherent oscillatory dynamics were observed.

Based on the values of c(ti)𝑐subscript𝑡𝑖c(t_{i})italic_c ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) and c(tj)𝑐subscript𝑡𝑗c(t_{j})italic_c ( italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) in Fig.2c, we computed the Pearson correlation coefficients 38 ρ(ti,tj)𝜌subscript𝑡𝑖subscript𝑡𝑗\rho(t_{i},t_{j})italic_ρ ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) as a simple measure of their linear correlation: ρ(ti,tj)=cov[c(ti)c(tj)]/(σ[c(ti)]σ[c(tj)])𝜌subscript𝑡𝑖subscript𝑡𝑗covdelimited-[]𝑐subscript𝑡𝑖𝑐subscript𝑡𝑗𝜎delimited-[]𝑐subscript𝑡𝑖𝜎delimited-[]𝑐subscript𝑡𝑗\rho(t_{i},t_{j})=\text{cov}[c(t_{i})c(t_{j})]/(\sigma[c(t_{i})]\sigma[c(t_{j}%)])italic_ρ ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) = cov [ italic_c ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) italic_c ( italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) ] / ( italic_σ [ italic_c ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ] italic_σ [ italic_c ( italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) ] ), where the covariance (cov) and the standard deviation (σ𝜎\sigmaitalic_σ) are computed across 10scans. The symmetric correlation matrix is shown for its lower-half in Fig.4b for all non-negative pump-probe time delays, where each pixel corresponds to a linear correlation coefficient derived from a scatter plot between c(ti)𝑐subscript𝑡𝑖c(t_{i})italic_c ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) and c(tj)𝑐subscript𝑡𝑗c(t_{j})italic_c ( italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) (Fig.4a). Besides the ρ(ti,tj=i)1𝜌subscript𝑡𝑖subscript𝑡𝑗𝑖1\rho(t_{i},t_{j=i})\equiv 1italic_ρ ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT italic_j = italic_i end_POSTSUBSCRIPT ) ≡ 1 entries along the 45 diagonal, the correlation matrix in Fig.4b appears to be populated with mostly random values around ρ=0𝜌0\rho=0italic_ρ = 0. This randomness is reflected in the histogram of all non-diagonal entries in the correlation matrix (Fig.4c), which can be fitted to a Gaussian distribution where we notice a small offset of the histogram towards positive ρ𝜌\rhoitalic_ρ, which is best seen from the unbalanced tails of the histogram highlighted by red and blue bars. The excess positive values of ρ𝜌\rhoitalic_ρ stem from the red features in the correlation matrix in the neighborhood of the 45 diagonal, indicating a high degree of correlation of the transient c𝑐citalic_c-axis parameter if titjsubscript𝑡𝑖subscript𝑡𝑗t_{i}\approx t_{j}italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≈ italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT; namely, pump-induced lattice trajectories are not independent if there are relatively few photoexcitation events elapsed between the corresponding pump pulses. To quantify the correlation length ξ𝜉\xiitalic_ξ in terms of the elapsed pump shots during the measurement, we inspect how fast the Pearson correlation coefficient decays from 1 as one moves away from the 45 diagonal in the correlation matrix. In practice, we compute an averaged correlation coefficient ρavg(Δt)subscript𝜌avgΔ𝑡\rho_{\text{avg}}(\Delta t)italic_ρ start_POSTSUBSCRIPT avg end_POSTSUBSCRIPT ( roman_Δ italic_t ) by going through all ρ(ti,tj)𝜌subscript𝑡𝑖subscript𝑡𝑗\rho(t_{i},t_{j})italic_ρ ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) that satisfies Δt=|titj|Δ𝑡subscript𝑡𝑖subscript𝑡𝑗\Delta t=|t_{i}-t_{j}|roman_Δ italic_t = | italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT |, shown in Fig.4d. The resulting curve ρavgsubscript𝜌avg\rho_{\text{avg}}italic_ρ start_POSTSUBSCRIPT avg end_POSTSUBSCRIPT shows a clear exponentially decaying trend towards zero, yielding a correlation length of ξ=1,500±300𝜉1plus-or-minus500300\xi=1,500\pm 300italic_ξ = 1 , 500 ± 300 shots, which are comparable to the number of pump shots received (n1,000𝑛1000n\approx 1,000italic_n ≈ 1 , 000shots) at one specific delay during one scan in our data acquisition scheme. This statistical analysis hence gives a quantitative measure of how fast the dynamical correlation is lost as more photoexcitation events occur under repeated pump shots.

III Simulation and discussion

To understand the physical origin of the correlation in the stochastic lattice expansion and relaxation, we simulated the measurements following the exact data acquisition scheme in the experiment (Fig.2a). Inspired by the excellent fit of Eq.(1) to the averaged response in Fig.3a, we employed the same phenomenological model to capture the system evolution after individual photoexcitation event. In the simulation, stochastic responses are introduced by drawing the initial lattice response [I0subscript𝐼0I_{0}italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT in Eq.(1)] randomly from a Gaussian distribution. However, if I0subscript𝐼0I_{0}italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is both random and independent after every pump shot, there is no dynamical correlation and the resulting ρ(ti,tj)𝜌subscript𝑡𝑖subscript𝑡𝑗\rho(t_{i},t_{j})italic_ρ ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) matrix is truly randomly populated with no enhanced coefficients when titjsubscript𝑡𝑖subscript𝑡𝑗t_{i}\approx t_{j}italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≈ italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT (see ref.33 and Fig.S5). To model the highly-correlated dynamics between neighboring shots, we introduce the following protocol: with probability p01much-less-thansubscript𝑝01p_{0}\ll 1italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≪ 1, I0subscript𝐼0I_{0}italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT will be randomly selected for the next pump shot; otherwise, the photoinduced dynamics for the next shot will be identical to the dynamics following the current shot.

Despite the simplicity of our model, it recreated the key statistical properties observed in the experiment such as the Pearson correlation matrix and its histogram distribution (Fig.4f,g), which look nearly indistinguishable from the experimental data (Fig.4b,c). In individual time traces, important characteristics such as the discontinuities in between a stretch of continuous streaks are also clearly discernible (Fig.S7). Similar to the experimental correlation matrix in Fig.4d, a high and positive correlation value is observed near the 45 diagonal where titjsubscript𝑡𝑖subscript𝑡𝑗t_{i}\approx t_{j}italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≈ italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT, which decays to zero for large |titj|subscript𝑡𝑖subscript𝑡𝑗|t_{i}-t_{j}|| italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT | (Fig.4h). In the simulation, the value of p0subscript𝑝0p_{0}italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT was adjusted to yield the same correlation length of ξ=1,500±300𝜉1plus-or-minus500300\xi=1,500\pm 300italic_ξ = 1 , 500 ± 300 shots, leading to p0=(0.09±0.02)%subscript𝑝0percentplus-or-minus0.090.02p_{0}=(0.09\pm 0.02)\%italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = ( 0.09 ± 0.02 ) %, where the error bar of p0subscript𝑝0p_{0}italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is computed to correspond to the lower and upper bounds of the correlation length ξ𝜉\xiitalic_ξ.

Physically, p0subscript𝑝0p_{0}italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT represents the probability of a micro-grain of LLTO to randomly change its nonequilibrium lattice expansion and relaxation trajectories after the next pump shot. From an energetics perspective, the value of p0=exp[E0/(kBT)]subscript𝑝0subscript𝐸0subscript𝑘𝐵𝑇p_{0}=\exp{[-E_{0}/(k_{B}T)]}italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = roman_exp [ - italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / ( italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T ) ] implies an energy barrier of E0=0.38±0.01subscript𝐸0plus-or-minus0.380.01E_{0}=0.38\pm 0.01italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.38 ± 0.01eV, where kBsubscript𝑘𝐵k_{B}italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT is the Boltzmann constant and T620𝑇620T\approx 620italic_T ≈ 620K is the lattice temperature right after photoexcitation (see ref.33 for an estimate of the photoinduced lattice temperature rise). In our experiments, E0subscript𝐸0E_{0}italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the activation energy for the micro-grain to change its photoinduced lattice dynamics in the powder ensemble, where local strain, grain orientation, and non-uniform thermal gradients can all contribute to this barrier. To understand what can cause such a change of the micro-grain environment, we note that the value of E0subscript𝐸0E_{0}italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is close to both theoretical and experimental values of the energy barrier for lithium ion migration in LLTO (Eb0.3subscript𝐸𝑏0.3E_{b}\approx 0.3italic_E start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ≈ 0.3–0.5eV) 39, 23, 24. Hence, one plausible scenario is that lithium ion displacement after each pump pulse 40 may result in a slight variation in the metastable lattice structure, where the hopping lithium ion follows a different photoinduced trajectory due to the modified strain environment surrounding the crystalline grain of interest.

The photo-assisted lithium ion diffusion also offers an explanation to the negative correlation between the initial photoinduced lattice expansion [I0subscript𝐼0I_{0}italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT in Eq.(1)] and the corresponding relaxation time τ𝜏\tauitalic_τ. Such a relation typically shows up in optical pump-probe studies of semiconductors 41 and superconductors 42, where in certain regimes, bimolecular recombination of electron-hole pairs or electron-electron pairs leads to a faster decay with a larger number of initially excited free carriers. However, such mechanisms are not expected to apply to the present measurements, where relaxation over tens of microseconds is dictated by heat diffusion, strain relaxation, and macroscopic grain motion 43, 44, 45. In the context of photoinduced structural change — especially through a nonequilibrium transition — a positive correlation between I0subscript𝐼0I_{0}italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and τ𝜏\tauitalic_τ is typically observed 46, 47, 48, contrary to our observation of LLTO. The positive I0subscript𝐼0I_{0}italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT-τ𝜏\tauitalic_τ correlation in those other experiments 46, 47, 48 is indicative of a soft mode as the system enters a flat potential energy landscape near the critical point. By contrast, in LLTO, lithium ion migration during the course of a photoinduced lattice evolution allows the lattice to enter a transient structure that is energetically more favorable than the one without any lithium ion movement. This process hence finds a local minimum by going away from the flat region in the potential energy surface, leading to a stiffer instead of softer lattice and hence accounting for the faster lattice relaxation.

Our results introduce a general statistical framework to extract stochastic fluctuations in photoinduced dynamics that can be applied to a variety of pump-probe experiments, provided that the noise in the data is dominated by intrinsic sample dynamics instead of extrinsic instrumental uncertainties. In the context of X-ray sciences, the nonequilibrium noise correlation spectroscopy demonstrated in this work presents a streamlined, complementary approach to photon correlation spectroscopy in the study of structural correlations at an ultrashort timescale while avoiding the technical complexity of implementing split X-ray pulses in a free-electron laser 49, 50, 51. Our findings hold great potential in the active pursuit of modeling and designing spatially heterogeneous or temporally fluctuating systems that underpin most materials of both fundamental and technological interest, such as unconventional superconductors 52, 53, self-assembled nanostructures 54, and in operando catalysts 55. In intrinsically nonequilibrium context such as biomolecular processes, our approach may help identify new structural dynamics and energy flow that are critical in understanding and orchestrating the reaction pathway 56. This work hence provides a powerful tool to uncover hidden correlations in otherwise random dynamics that occur at the intrinsic timescale and lengthscale of electrons, ions, and molecules.

IV Methods

Time-resolved X-ray micro-diffraction: Experiments were performed at the 7ID-C beamline at the Advanced Photon Source (APS) in Argonne National Laboratory, and the setup configurations were detailed in refs.57, 58. The geometry of the experiment is depicted in Figs.1b and S1. The samples were photoexcited above the bandgap of LLTO using a 3.55eV (349nm) femtosecond laser source, which was produced as the second harmonic of the output of an optical parametric amplifier (OPerA Solo, Coherent Inc.) using the output of an amplified Ti:sapphire laser (Legend, Coherent Inc.). The pump laser operates at 1kHz repetition rate, which was locked to an integer division of the synchrotron repetition rate. The probing X-ray beam from the synchrotron had a pulse duration of 100ps operating at 6.5MHz repetition rate. The X-ray beam was then monochromatized to an energy of 8keV (1.5498Å) and focused by a pair of Kirkpatrick-Baez mirrors to a cross-sectional spot size of 12μm×15μ12μm15μ12~{}\upmu\text{m}\times 15~{}\upmu12 roman_μ m × 15 roman_μm (full-width at half maximum, FWHM), which was an order of magnitude smaller than the pump beam spot size at the sample position to ensure a near-uniform photoexcitation condition in the probed area. The powder pellet was mounted at the center of rotation of a six-circle diffractometer (Huber GmbH). The X-ray diffraction signals were collected by an area detector (Pilatus 100K, DECTRIS Ltd.) that was gated to selectively record the X-ray pulse that was paired with the excitation laser pulse, rendering the overall repetition rate of the experiment to 1kHz. The time delay between the laser pump pulse and X-ray probe pulse was varied electronically, allowing us to access a timescale up to tens of microseconds, which is relevant for the slow relaxation process pertinent to heat dissipation through a grainy pellet 59 as well as macroscopic grain motion as a result of laser-induced lattice parameter change.

Sample preparation and characterization:Li0.5La0.5TiO3 was synthesized according to the procedures in ref.60. A stoichiometric amount of La2O3, Li2CO3, and TiO2 were mixed in an agate mortar and pressed into pellets under 100MPa of pressure. The pellets were placed on a bed of sacrificial powder and calcined at 800C for 4hr then at 1200C for 12hr at a ramp rate of 1C/min. After the crystal structure of LLTO was confirmed with X-ray diffraction, the resulting powder was pressed into a pellet with a diameter of 6mm and a thickness of 0.5–1mm under 620MPa of pressure. The pellet was subsequently annealed at 1100C for 6hr at a ramp rate of 2C/min over a bed of its mother powder. To characterize the grain morphology, scanning electron microscopy (SEM) was performed using the SE2 detector of a ZEISS 1550VP field emission SEM with an acceleration voltage of 10kV at 10k×\times× magnification. As shown in Fig.1c, the grain size of LLTO is comparable to the X-ray beam spot size for the time-resolved X-ray micro-diffraction measurements.

V Additional Information

Acknowledgements:We thank fruitful discussions with X.Xu and A.Kogar, and we appreciate the support in sample synthesis and characterization from K.See. A.Z. acknowledges support from the Miller Institute for Basic Research in Science. J.M. acknowledges funding by the National Science Foundation (NSF-REU EEC-1852537). M.Z. acknowledges funding by the Department of Energy (DE-SC0024123). This research used resources of the Advanced Photon Source, a U.S. Department of Energy (DOE) Office of Science user facility operated for the DOE Office of Science by Argonne National Laboratory under Contract No.DE-AC02-06CH11357.

Author contributions:A.Z., S.C., and M.Z. conceived the project. The time-resolved X-ray measurements were conducted by A.Z., H.L., K.H.P., B.G., D.A.W., and H.W. Beamline 7ID-C at Advanced Photon Source is operated by B.G., D.A.W., and H.W. A.Z. and J.M. analyzed the data and performed model calculation. K.H.P. synthesized and characterized LLTO, and prepared the sample for the beamline experiments. Z.W.B.I. performed the scanning electron microscopy measurements of LLTO. J.M. and A.Z. wrote the manuscript with critical inputs from K.H.P., S.C., M.Z., and all other authors. The research was supervised by S.C. and M.Z.

Competing interests:The authors declare no competing interests.

Data availability:All of the data and calculations supporting the conclusions are available within the article and the Supplementary Information. Additional data are available from the corresponding authors upon reasonable request.

References

  • MommaandIzumi [2011]K.MommaandF.Izumi,VESTA 3 for three-dimensionalvisualization of crystal, volumetric and morphology data,J. Appl. Crystallogr.44,1272 (2011).
  • Moffat [1998]K.Moffat,Ultrafast time-resolvedcrystallography,Nat. Struct. Mol. Biol.5,641 (1998).
  • dela Torreetal. [2021]A.dela Torre, D.M.Kennes, M.Claassen,S.Gerber, J.W.McIver,andM.A.Sentef,Colloquium: Nonthermal pathways to ultrafast control inquantum materials,Rev. Mod. Phys.93,041002 (2021).
  • Zongetal. [2023]A.Zong, B.R.Nebgen,S.-C.Lin, J.A.Spies,andM.Zuerch,Emerging ultrafast techniques for studying quantummaterials,Nat. Rev. Mater.8,224 (2023).
  • Lépineetal. [2014]F.Lépine, M.Y.Ivanov,andM.J.J.Vrakking,Attosecond moleculardynamics: fact or fiction?,Nat. Photon.8,195 (2014).
  • Moetal. [2016]M.Z.Mo, X.Shen, Z.Chen, R.K.Li, M.Dunning, K.Sokolowski-Tinten, Q.Zheng, S.P.Weathersby, A.H.Reid, R.Coffee, I.Makasyuk,S.Edstrom, D.McCormick, K.Jobe, C.Hast, S.H.Glenzer,andX.Wang,Single-shot mega-electronvolt ultrafast electron diffraction for structuredynamic studies of warm dense matter,Rev. Sci. Instrum.87,11D810 (2016).
  • LiangandWang [2018]J.LiangandL.V.Wang,Single-shot ultrafastoptical imaging,Optica5,1113 (2018).
  • Helketal. [2019]T.Helk, M.Zürch,andC.Spielmann,Perspective: Towards single shot time-resolvedmicroscopy using short wavelength table-top light sources,Struct. Dyn.6,010902 (2019).
  • WakehamandNelson [2000]G.P.WakehamandK.A.Nelson,Dual-echelon single-shotfemtosecond spectroscopy,Opt. Lett.25,505 (2000).
  • Shinetal. [2014]T.Shin, J.W.Wolfson,S.W.Teitelbaum,M.Kandyla,andK.A.Nelson,Dual echelon femtosecond single-shot spectroscopy,Rev. Sci. Instrum.85,083115 (2014).
  • Noeetal. [2016]G.T.Noe, I.Katayama,F.Katsutani, J.J.Allred, J.A.Horowitz, D.M.Sullivan, Q.Zhang, F.Sekiguchi, G.L.Woods, M.C.Hoffmann, H.Nojiri,J.Takeda,andJ.Kono,Single-shot terahertz time-domain spectroscopy in pulsedhigh magnetic fields,Opt. Express24,30328 (2016).
  • Kloseetal. [2023]C.Klose, F.Büttner,W.Hu, C.Mazzoli, K.Litzius, R.Battistelli, I.Lemesh, J.M.Bartell, M.Huang, C.M.Günther,M.Schneider, A.Barbour, S.B.Wilkins, G.S.D.Beach, S.Eisebitt,andB.Pfau,Coherent correlation imaging for resolving fluctuating states ofmatter,Nature614,256 (2023).
  • Manguetal. [2024]A.Mangu, V.A.Stoica,H.Zheng, T.Yang, M.Zhang, H.Wang, Q.L.Nguyen, S.Song, S.Das, P.Meisenheimer, E.Donoway, M.Chollet, Y.Sun, J.J.Turner, J.W.Freeland, H.Wen, L.W.Martin,L.Q.Chen, V.Gopalan, D.Zhu, Y.Cao,andA.M.Lindenberg,Hidden domainboundary dynamics towards crystalline perfection (2024),arXiv:2402.04962 .
  • Walletal. [2018]S.Wall, S.Yang, L.Vidas, M.Chollet, J.M.Glownia, M.Kozina, T.Katayama, T.Henighan, M.Jiang, T.A.Miller, D.A.Reis, L.A.Boatner, O.Delaire,andM.Trigo,Ultrafast disordering of vanadium dimers inphotoexcited VO22{}_{\textrm{2}}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT,Science362,572 (2018).
  • Zongetal. [2021]A.Zong, P.E.Dolgirev,A.Kogar, Y.Su, X.Shen, J.A.W.Straquadine, X.Wang, D.Luo, M.E.Kozina,A.H.Reid, R.Li, J.Yang, S.P.Weathersby, S.Park, E.J.Sie, P.Jarillo-Herrero, I.R.Fisher, X.Wang,E.Demler,andN.Gedik,Role of equilibrium fluctuations in light-induced order,Phys. Rev. Lett.127,227401 (2021).
  • Perez-Salinasetal. [2022]D.Perez-Salinas, A.S.Johnson, D.Prabhakaran,andS.Wall,Multi-mode excitation drivesdisorder during the ultrafast melting of a C4-symmetry-broken phase,Nat. Commun.13,238 (2022).
  • Disaetal. [2023]A.S.Disa, J.Curtis,M.Fechner, A.Liu, A.von Hoegen, M.Först, T.F.Nova, P.Narang, A.Maljuk, A.V.Boris, B.Keimer,andA.Cavalleri,Photo-inducedhigh-temperature ferromagnetism in YTiO33{}_{\textrm{3}}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT,Nature617,73 (2023).
  • Stramareetal. [2003]S.Stramare, V.Thangadurai,andW.Weppner,Lithium lanthanumtitanates: a review,Chem. Mater.15,3974–3990 (2003).
  • Belousetal. [2004]A.Belous, O.Yanchevskiy,O.V’yunov, O.Bohnke, C.Bohnke, F.LeBerre,andJ.Fourquet,Peculiarities of Li0.5La0.5TiO3 formation during the synthesisby solid-state reaction or precipitation from solutions,Chem. Mater.16,407 (2004).
  • Nakayamaetal. [2005]M.Nakayama, T.Usui,Y.Uchimoto, M.Wakihara,andM.Yamamoto,Changes in electronic structure upon lithium insertioninto the A𝐴Aitalic_A-site deficient perovskite type oxides (Li,La)TiO3,J. Phys. Chem. B109,4135 (2005).
  • Qianetal. [2012]D.Qian, B.Xu, H.Cho, T.Hatsukade, K.Carroll,andY.Meng,Lithium lanthanum titanium oxides: a fast ionic conductive coatingfor lithium-ion battery cathodes,Chem. Mater.24,2744–2751 (2012).
  • Okumuraetal. [2011]T.Okumura, T.Ina,Y.Orikasa, H.Arai, Y.Uchimoto,andZ.Ogumi,Effect of average and local structures on lithium ion conductivityin La2/3-xLi3xTiO3,J. Mater. Chem.21,10195 (2011).
  • Zhangetal. [2020]L.Zhang, X.Zhang,G.Tian, Q.Zhang, M.Knapp, H.Ehrenberg, G.Chen, Z.Shen, G.Yang, L.Gu,andF.Du,Lithium lanthanum titanate perovskite as an anode for lithium ionbatteries,Nat. Commun.11,3490 (2020).
  • Woodahletal. [2023]C.Woodahl, S.Jamnuch,A.Amado, C.B.Uzundal, E.Berger, P.Manset, Y.Zhu, Y.Li, D.D.Fong,J.G.Connell, Y.Hirata, Y.Kubota, S.Owada, K.Tono, M.Yabashi, S.G.E.teVelthuis, S.Tepavcevic, I.Matsuda,W.S.Drisdell, C.P.Schwartz, J.W.Freeland, T.A.Pascal, A.Zong,andM.Zuerch,Probing lithium mobility at a solid electrolyte surface,Nat. Mater.22,848 (2023).
  • Renéde Cotretetal. [2019]L.P.Renéde Cotret, J.-H.Pöhls, M.J.Stern,M.R.Otto, M.Sutton,andB.J.Siwick,Time- and momentum-resolved phonon population dynamicswith ultrafast electron diffuse scattering,Phys. Rev. B100,214115 (2019).
  • Chengetal. [2024]Y.Cheng, A.Zong,L.Wu, Q.Meng, W.Xia, F.Qi, P.Zhu, X.Zou, T.Jiang, Y.Guo, J.van Wezel, A.Kogar,M.W.Zuerch, J.Zhang, Y.Zhu,andD.Xiang,Ultrafast formation of topological defects in a two-dimensionalcharge density wave,Nat. Phys.20,54 (2024).
  • Sokolowski-Tintenetal. [1995]K.Sokolowski-Tinten, J.Bialkowski,andD.vonder Linde,Ultrafastlaser-induced order-disorder transitions in semiconductors,Phys. Rev. B51,14186 (1995).
  • Shugaevetal. [2016]M.V.Shugaev, C.Wu,O.Armbruster, A.Naghilou, N.Brouwer, D.S.Ivanov, T.J.-Y.Derrien, N.M.Bulgakova, W.Kautek, B.Rethfeld,andL.V.Zhigilei,Fundamentals of ultrafastlaser–material interaction,MRS Bulletin41,960 (2016).
  • Yashimaetal. [2005]M.Yashima, M.Itoh,Y.Inaguma,andY.Morii,Crystal structure and diffusion path in the fastlithium-ion conductor La0.62Li0.16TiO3,J. Am. Chem. Soc.127,3491 (2005).
  • Kimetal. [2023]T.Kim, S.Park, V.Iyer, B.Shaheen, U.Choudhry, Q.Jiang, G.Eichman, R.Gnabasik, K.Kelley, B.Lawrie, K.Zhu,andB.Liao,Mappingthe pathways of photo-induced ion migration in organic-inorganic hybridhalide perovskites,Nat. Commun.14,1846 (2023).
  • Wakamura [1998]K.Wakamura,Effects of electronicband on activation energy and of effective charge on lattice distortion insuperionic conductors,J. Phys. Chem. Solids59,591 (1998).
  • Rahaman [2017]M.N.Rahaman,Ceramic Processing and Sintering, 2nd ed.(CRC Press,2017).
  • [33]See supplemental materials.
  • Mooreetal. [2016]R.G.Moore, W.S.Lee,P.S.Kirchman, Y.D.Chuang, A.F.Kemper, M.Trigo, L.Patthey, D.H.Lu, O.Krupin, M.Yi, D.A.Reis, D.Doering, P.Denes, W.F.Schlotter, J.J.Turner, G.Hays, P.Hering,T.Benson, J.-H.Chu, T.P.Devereaux, I.R.Fisher, Z.Hussain,andZ.-X.Shen,Ultrafast resonant soft X-ray diffraction dynamics of the chargedensity wave in TbTe33{}_{\textrm{3}}start_FLOATSUBSCRIPT 3 end_FLOATSUBSCRIPT,Phys. Rev. B93,024304 (2016).
  • Zong [2021]A.Zong,Emergent States in Photoinduced Charge-Density-WaveTransitions(Springer,Cham,2021).
  • DemsarandDekorsy [2012]J.DemsarandT.Dekorsy,Carrier dynamics in bulk semiconductors andmetals after ultrashort pulse excitation,inOpticalTechniques for Solid-State Materials Characterization,edited byR.P.PrasankumarandA.J.Taylor(CRC Press,BocaRaton,2012)Chap.8, pp.291–328.
  • Rubio-Marcosetal. [2021]F.Rubio-Marcos, A.DelCampo, J.Ordoñez-Pimentel, M.Venet, R.E.Rojas-Hernandez, D.Páez-Margarit, D.A.Ochoa, J.F.Fernández,andJ.E.García,Photocontrolled strainin polycrystalline ferroelectrics via domain engineering strategy,ACS Appl. Mater. Interfaces13,20858 (2021).
  • Boslaugh [2012]S.Boslaugh,Statistics in aNutshell, 2nd ed.(O’Reilly Media, Inc.,2012).
  • Gengetal. [2009]H.Geng, A.Mei, C.Dong, Y.Lin,andC.Nan,Investigation of structure and electrical properties ofLi0.5La0.5TiO3 ceramics via microwave sintering,J. Alloys Compd.481,555 (2009).
  • Defferriereetal. [2022]T.Defferriere, D.Klotz,J.C.Gonzalez-Rosillo,J.L.M.Rupp,andH.L.Tuller,Photo-enhanced ionic conductivity across grainboundaries in polycrystalline ceramics,Nat. Mater.21,438–444 (2022).
  • ShkrobandCrowell [1998]I.A.ShkrobandR.A.Crowell,Ultrafast chargerecombination in undoped amorphous hydrogenated silicon,Phys. Rev. B57,12207 (1998).
  • Gediketal. [2004]N.Gedik, P.Blake,R.C.Spitzer, J.Orenstein, R.Liang, D.A.Bonn,andW.N.Hardy,Single-quasiparticle stability and quasiparticle-pair decay inYBa2Cu3O6.5,Phys. Rev. B70,014504 (2004).
  • Stoicaetal. [2024]V.A.Stoica, T.Yang,S.Das, Y.Cao, H.Wang, Y.Kubota, C.Dai,H.Padmanabhan, Y.Sato, A.Mangu, Q.L.Nguyen, Z.Zhang, D.Talreja, M.E.Zajac, D.A.Walko, A.D.DiChiara,S.Owada, K.Miyanishi, K.Tamasaku, T.Sato, J.M.Glownia, V.Esposito, S.Nelson, M.C.Hoffmann, R.D.Schaller, A.M.Lindenberg, L.W.Martin, R.Ramesh, I.Matsuda,D.Zhu, L.-Q.Chen, H.Wen, V.Gopalan,andJ.W.Freeland,Non-equilibriumpathways to emergent polar supertextures (2024),arXiv:2402.11634 .
  • Cherukaraetal. [2017]M.J.Cherukara, K.Sasikumar, W.Cha,B.Narayanan, S.J.Leake, E.M.Dufresne, T.Peterka, I.McNulty, H.Wen, S.K. R.S.Sankaranarayanan,andR.J.Harder,Ultrafast three-dimensional X-ray imaging of deformation modes in ZnOnanocrystals,Nano Lett.17,1102 (2017).
  • Chen [2005]G.Chen,Nanoscale Energy Transport and Conversion: AParallel Treatment of Electrons, Molecules, Phonons, and Photons(Oxford University Press,2005).
  • Zongetal. [2019]A.Zong, A.Kogar,Y.-Q.Bie, T.Rohwer, C.Lee, E.Baldini, E.Ergeçen, M.B.Yilmaz, B.Freelon, E.J.Sie,H.Zhou, J.Straquadine, P.Walmsley, P.E.Dolgirev, A.V.Rozhkov, I.R.Fisher, P.Jarillo-Herrero, B.V.Fine,andN.Gedik,Evidence for topological defects in a photoinduced phasetransition,Nat. Phys.15,27 (2019).
  • Huberetal. [2014]T.Huber, S.O.Mariager,A.Ferrer, H.Schäfer, J.A.Johnson, S.Grübel, A.Lübcke, L.Huber, T.Kubacka, C.Dornes, C.Laulhe, S.Ravy, G.Ingold, P.Beaud,J.Demsar,andS.L.Johnson,Coherent structural dynamics of a prototypicalcharge-density-wave-to-metal transition,Phys. Rev. Lett.113,026401 (2014).
  • Beaudetal. [2014]P.Beaud, A.Caviezel,S.O.Mariager, L.Rettig, G.Ingold, C.Dornes, S.-W.Huang, J.A.Johnson, M.Radovic,T.Huber, T.Kubacka, A.Ferrer, H.T.Lemke, M.Chollet, D.Zhu, J.M.Glownia, M.Sikorski, A.Robert,H.Wadati, M.Nakamura, M.Kawasaki, Y.Tokura, S.L.Johnson,andU.Staub,Atime-dependent order parameter for ultrafast photoinduced phasetransitions,Nat. Mater.13,923 (2014).
  • Osakaetal. [2016]T.Osaka, T.Hirano,Y.Sano, Y.Inubushi, S.Matsuyama, K.Tono, T.Ishikawa, K.Yamauchi,andM.Yabashi,Wavelength-tunable split-and-delay optical system for hard X-rayfree-electron lasers,Opt. Express24,9187 (2016).
  • Rosekeretal. [2018]W.Roseker, S.O.Hruszkewycz, F.Lehmkühler, M.Walther, H.Schulte-Schrepping, S.Lee, T.Osaka, L.Strüder, R.Hartmann, M.Sikorski, S.Song, A.Robert, P.H.Fuoss, M.Sutton, G.B.Stephenson,andG.Grübel,Towards ultrafastdynamics with split-pulse X-ray photon correlation spectroscopy at freeelectron laser sources,Nat. Commun.9,1704 (2018),publisher: Nature Publishing Group.
  • Zhuetal. [2017]D.Zhu, Y.Sun, D.W.Schafer, H.Shi, J.H.James, K.L.Gumerlock, T.O.Osier, R.Whitney, L.Zhang,J.Nicolas, B.Smith, A.H.Barada,andA.Robert,Development of a hard x-ray split-delay system at the LinacCoherent Light Source,inAdvancesin X-ray Free-Electron Lasers Instrumentation IV,Vol.10237(SPIE,2017)pp.103–110.
  • Torchinskyetal. [2013]D.H.Torchinsky, F.Mahmood,A.T.Bollinger, I.Božović,andN.Gedik,Fluctuating charge-density waves in a cupratesuperconductor,Nat. Mater.12,387 (2013).
  • Arpaiaetal. [2019]R.Arpaia, S.Caprara,R.Fumagalli, G.D.Vecchi, Y.Y.Peng, E.Andersson, D.Betto, G.M.D.Luca, N.B.Brookes, F.Lombardi, M.Salluzzo,L.Braicovich, C.D.Castro, M.Grilli,andG.Ghiringhelli,Dynamical charge density fluctuations pervading the phasediagram of a Cu-based high-Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT superconductor,Science365,906 (2019).
  • Ekizetal. [2016]M.S.Ekiz, G.Cinar,M.A.Khalily,andM.O.Guler,Self-assembled peptide nanostructures forfunctional materials,Nanotechnology27,402002 (2016).
  • Joháneketal. [2004]V.Johánek, M.Laurin,A.W.Grant, B.Kasemo, C.R.Henry,andJ.Libuda,Fluctuations and bistabilities on catalyst nanoparticles,Science304,1639 (2004).
  • Bhowmicketal. [2023]A.Bhowmick, R.Hussein,I.Bogacz, P.S.Simon, M.Ibrahim, R.Chatterjee, M.D.Doyle, M.H.Cheah, T.Fransson,P.Chernev, I.-S.Kim, H.Makita, M.Dasgupta, C.J.Kaminsky, M.Zhang, J.Gätcke,S.Haupt, I.I.Nangca, S.M.Keable, A.O.Aydin, K.Tono, S.Owada, L.B.Gee, F.D.Fuller, A.Batyuk,R.Alonso-Mori, J.M.Holton, D.W.Paley, N.W.Moriarty, F.Mamedov, P.D.Adams, A.S.Brewster, H.Dobbek, N.K.Sauter,U.Bergmann, A.Zouni, J.Messinger, J.Kern, J.Yano,andV.K.Yachandra,Structuralevidence for intermediates during O2 formation in photosystem II,Nature617,629 (2023).
  • Zhouetal. [2022]F.Zhou, K.Hwangbo,Q.Zhang, C.Wang, L.Shen, J.Zhang, Q.Jiang, A.Zong, Y.Su, M.Zajac, Y.Ahn, D.A.Walko, R.D.Schaller, J.-h.Chu, N.Gedik, X.Xu, D.Xiao,andH.Wen,Dynamical criticality of spin-shearcoupling in van der Waals antiferromagnets,Nat. Commun.13,6598 (2022).
  • Walkoetal. [2016]D.A.Walko, B.W.Adams,G.Doumy, E.M.Dufresne, Y.Li, A.M.March, A.R.Sandy,andJ.Wang.,Developments in time-resolved x-ray research at APS beamline 7ID,AIP Conf. Proc.1741 (2016).
  • Hanetal. [2017]Q.Han, Y.Bai, J.Liu, K.-z.Du, T.Li, D.Ji, Y.Zhou, C.Cao, D.Shin, J.Ding, A.D.Franklin, J.T.Glass, J.Hu, M.J.Therien,J.Liu,andD.B.Mitzi,Additive engineering for high-performanceroom-temperature-processed perovskite absorbers with micron-size grains andmicrosecond-range carrier lifetimes,Energy Environ. Sci.10,2365 (2017).
  • Inagumaetal. [1994]Y.Inaguma, L.Chen,M.Itoh,andT.Nakamura,Candidate compounds with perovskite structure for highlithium ionic conductivity,Solid State Ion.70-71,196 (1994).
1 Overview of the powder sample and the pump-probe setup. a, Crystal structure of the tetragonal LLTO in equilibrium. Graphics rendered by VESTA . b, Schematic of the experimental setup for synchrotron-based time-resolved hard X-ray micro-diffraction (see (2024)

References

Top Articles
Latest Posts
Article information

Author: Geoffrey Lueilwitz

Last Updated:

Views: 5898

Rating: 5 / 5 (80 voted)

Reviews: 87% of readers found this page helpful

Author information

Name: Geoffrey Lueilwitz

Birthday: 1997-03-23

Address: 74183 Thomas Course, Port Micheal, OK 55446-1529

Phone: +13408645881558

Job: Global Representative

Hobby: Sailing, Vehicle restoration, Rowing, Ghost hunting, Scrapbooking, Rugby, Board sports

Introduction: My name is Geoffrey Lueilwitz, I am a zealous, encouraging, sparkling, enchanting, graceful, faithful, nice person who loves writing and wants to share my knowledge and understanding with you.