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

Present address: ]Department of Physics, Harvard University, Cambridge, Massachusetts 02138, USA

Doping-dependence of the electron-phonon coupling in two families of bilayer superconducting cuprates

Yingying Peng [email protected] International Center for Quantum Materials, School of Physics, Peking University, Beijing CN-100871, China    Leonardo Martinelli [email protected] Dipartimento di Fisica, Politecnico di Milano, piazza Leonardo da Vinci 32, I-20133 Milano, Italy    Qizhi Li International Center for Quantum Materials, School of Physics, Peking University, Beijing CN-100871, China    Matteo Rossi Dipartimento di Fisica, Politecnico di Milano, piazza Leonardo da Vinci 32, I-20133 Milano, Italy    Matteo Mitrano [ Department of Physics and Materials Research Laboratory, University of Illinois, Urbana, IL 61801, USA    Riccardo Arpaia Dipartimento di Fisica, Politecnico di Milano, piazza Leonardo da Vinci 32, I-20133 Milano, Italy Quantum Device Physics Laboratory, Department of Microtechnology and Nanoscience, Chalmers University of Technology, SE-41296 Göteborg, Sweden    Marco Moretti Sala Dipartimento di Fisica, Politecnico di Milano, piazza Leonardo da Vinci 32, I-20133 Milano, Italy    Qiang Gao Beijing National Laboratory for Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, Beijing CN-100190, China    Xuefei Guo Department of Physics and Materials Research Laboratory, University of Illinois, Urbana, IL 61801, USA    Gabriella Maria De Luca Dipartimento di Fisica Ettore Pancini Università di Napoli ”Federico II”, Complesso Monte-Santangelo via Cinthia, I-80126 Napoli, Italy CNR-SPIN Complesso Monte-Santangelo via Cinthia, I-80126 Napoli, Italy    Andrew Walters Diamond Light Source, Harwell Campus, Didcot OX11 0DE, United Kingdom    Abhishek Nag Diamond Light Source, Harwell Campus, Didcot OX11 0DE, United Kingdom    Andi Barbour National Synchrotron Light Source II, Brookhaven National Laboratory, Upton, NY 11973, USA    Genda Gu Condensed Matter Physics & Materials Science Department, Brookhaven National Laboratory, Utpton, NY 11973, USA    Jonathan Pelliciari National Synchrotron Light Source II, Brookhaven National Laboratory, Upton, NY 11973, USA    Nicholas B. Brookes ESRF—The European Synchrotron, 71 Avenue des Martyrs, CS 40220, F-38043 Grenoble, France    Peter Abbamonte Department of Physics and Materials Research Laboratory, University of Illinois, Urbana, IL 61801, USA    Marco Salluzzo CNR-SPIN Complesso Monte-Santangelo via Cinthia, I-80126 Napoli, Italy    Xingjiang Zhou Beijing National Laboratory for Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, Beijing CN-100190, China    Ke-Jin Zhou Diamond Light Source, Harwell Campus, Didcot OX11 0DE, United Kingdom    Valentina Bisogni National Synchrotron Light Source II, Brookhaven National Laboratory, Upton, NY 11973, USA    Lucio Braicovich Dipartimento di Fisica, Politecnico di Milano, piazza Leonardo da Vinci 32, I-20133 Milano, Italy ESRF—The European Synchrotron, 71 Avenue des Martyrs, CS 40220, F-38043 Grenoble, France    Steven Johnston [email protected] Department of Physics and Astronomy, The University of Tennessee, Knoxville, TN 37996, USA Institute for Advanced Materials and Manufacturing, University of Tennessee, Knoxville, TN 37996, USA    Giacomo Ghiringhelli [email protected] Dipartimento di Fisica, Politecnico di Milano, piazza Leonardo da Vinci 32, I-20133 Milano, Italy CNR-SPIN, Dipartimento di Fisica, Politecnico di Milano, piazza Leonardo da Vinci 32, I-20133 Milano, Italy
Abstract

While electron-phonon coupling (EPC) is crucial for Cooper pairing in conventional superconductors, its role in high-TcT_{c} superconducting cuprates is debated. Using resonant inelastic x-ray scattering at the oxygen KK-edge, we studied the EPC in Bi2Sr2CaCu2O8+δ (Bi2212) and Nd1+xBa2-xCu3O7-δ (NBCO) at different doping levels ranging from heavily underdoped (p=0.07p=0.07) to overdoped (p=0.21p=0.21). We analyze the data with a localized Lang-Firsov model that allows for the coherent excitations of two phonon modes. While electronic band dispersion effects are non-negligible, we are able to perform a study of the relative values of EPC matrix elements in these cuprate families. In the case of NBCO, the choice of the excitation energy allows us to disentangle modes related to the CuO3 chains and the CuO2 planes. Combining the results from the two families, we find the EPC strength decreases with doping at 𝐪=(0.25,0)\mathbf{q_{\parallel}}=(-0.25,0) r.l.u., but has a non-monotonic trend as a function of doping at smaller momenta. This behavior is attributed to the screening effect of charge carriers. We also find that the phonon intensity is enhanced in the vicinity of the charge-density-wave (CDW) excitations while the extracted EPC strength appears to be less sensitive to their proximity. By performing a comparative study of two cuprate families, we are able to identify general trends in the EPC for the cuprates and provide experimental input to theories invoking a synergistic role for this interaction in dd-wave pairing.

preprint: phonons in O K edge RIXS

I Introduction

The pairing interaction in conventional superconductors is mediated by phonons and critically depends on the electron-phonon coupling (EPC) Bardeen et al. (1957); McMillan and Rowell (1965). The role played by EPC in copper-based high critical temperature superconductors (HTSs), however, is still debated. On the one hand, experimental probes like angle-resolved photoemission spectroscopy (ARPES) Lanzara et al. (2001); Cuk et al. (2004); Devereaux et al. (2004); Cuk et al. (2005); Graf et al. (2008); He et al. (2018); Johnston et al. (2012) and scanning tunneling microscopy (STM) Lee et al. (2006); Zhao (2009, 2007) have found evidence for the interplay between lattice vibrations and superconductivity. On the other hand, spin fluctuations are widely believed to provide the dominant contribution to pairing in the cuprates Bulut and Scalapino (1996); Maier et al. (2008); Scalapino (2012), since the superconducting order parameter is of dx2y2d_{x^{2}-y^{2}} symmetry  Tsuei and Kirtley (2000) and superconductivity appears in proximity to antiferromagnetism. This observation, however, does not explicitly rule out a potential role for phonons in pairing. For example, while a complete microscopic picture is currently absent, the interplay of the electron-phonon and electron-electron interactions may lead to the enhanced critical temperature (TcT_{c}) in HTSs. Indeed, theoretical work has shown that the materials with the largest coupling to the out-of-phase bond-buckling oxygen phonon branch (often referred to as the “B1gB_{1g}” modes) also have the largest Tc,maxT_{c,\mathrm{max}} Johnston et al. (2010), suggesting that these modes are relevant to superconductivity.

HTSs have complex phase diagrams that include competing spin-, charge- and pair-density-waves, a pseudogap, and strange metallic behavior Fradkin et al. (2015); Keimer et al. (2015); Comin and Damascelli (2016); Agterberg et al. (2020); Arpaia et al. (2018); Arpaia and Ghiringhelli (2021); Wahlberg et al. (2021). The pivotal parameter that determines the nature of the ground state at a given temperature is the doping pp, which gives the number of holes introduced in the CuO2 planes. In the underdoped region of the phase diagram, the phonon branches are also renormalized in the presence of charge-density-wave (CDW) correlations. This behavior is reminiscent of what happens in conventional metals, where the EPC is the driving force behind the formation of charge modulations Zhu et al. (2015). Recent resonant inelastic X-ray scattering (RIXS) studies have indeed observed anomalous softening and enhanced phonon excitations in the vicinity of charge order Chaix et al. (2017); Miao et al. (2018); Lin et al. (2020); Li et al. (2020); Peng et al. (2020).

In the underdoped region of the phase diagram the coupling between phonon and electrons is also affected by the onset of the pseudogap phase. The pseudogap in HTSs is characterized by a temperature TT^{*}, below which a partial gap opens in the normal state Timusk and Statt (1999). In general, TT^{*} falls to zero at a critical doping pc=0.19p_{c}=0.19, where Hall coefficient measurements show a change of carrier density from pp to 1+p1+p Badoux et al. (2016). Interestingly, a recent ARPES study of Bi2Sr2CaCu2O8+δ (Bi2212) revealed a rapid change of bosonic coupling strength to the bond-buckling B1gB_{1g} phonon branch (37\sim 37 meV) across the pseudogap boundary He et al. (2018). To better understand these observations, it is necessary to investigate how EPC changes with doping across the cuprate phase diagram, spanning from the underdoped region to the overdoped Fermi liquid region found for p>pcp>p_{\text{c}}.

Measuring the full EPC vertex M(𝐤,𝐪)M({\bf k},{\bf q}) is challenging, as it can depend on both the fermionic (𝐤{\bf k}) and bosonic (𝐪{\bf q}) momenta. Techniques like ARPES probe the 𝐤{\bf k} dependence of the EPC via the electron self-energy, which averages over the phonon momentum 𝐪{\bf q}. Similarly, scattering probes like inelastic x-ray and neutron scattering tend to probe the 𝐪{\bf q} dependence of M(𝐤,𝐪)M({\bf k},{\bf q}) via the phonon line width, which averages over the fermion momentum 𝐤{\bf k}. Recently, RIXS has been demonstrated as an effective tool to determine the momentum dependence of the EPC with element specificity, due to the interaction between the excited electrons in the intermediate state and phonons Ament et al. (2011). As a scattering technique, RIXS more directly measures the 𝐪{\bf q}-dependence of the vertex and integrates over the 𝐤{\bf k}-dependence. For example, models have shown that the 𝐤{\bf k}-integration can be weighting by other factors like the electron occupations and orbital character of the bands if the system hosts itinerant electrons Ament et al. (2011); Devereaux et al. (2016); Braicovich et al. (2020). In cases where the electron mobility can be neglected (e.g. due to a strong core-hole potential), the EPC strength can be extracted either from the ratio between the intensity of the phonon and of its overtones in RIXS spectra Ament et al. (2011); Bieniasz et al. (2021). Alternatively, one can measure the phonon intensity as the incident beam detuned from the resonance, as successfully done recently at the Cu L3L_{3} in cuprates Braicovich et al. (2020); Rossi et al. (2019).

Prior RIXS studies of EPC in the 2D cuprates have focused largely on the Cu L3L_{3}-edge Rossi et al. (2019); Braicovich et al. (2020); Peng et al. (2020). However, it has been established that the bands that cross the Fermi level in doped cuprates have significant oxygen 2px,y2p_{x,y} orbital character Chen et al. (1991). Moreover, the phonons believed to be most relevant to pairing in the cuprates are the optical oxygen modes Cuk et al. (2005); Johnston et al. (2010). It is, therefore, desirable to also study the EPC at the O KK-edge in these materials. The RIXS process at this edge involves the 1s2p1s\rightarrow 2p transition, so that in the intermediate state the charge perturbation mostly involves the electron in a 2p2p orbital (plus a core hole, on whose influence we will discuss in the following). Moreover, RIXS experiments at the O KK-edge can benefit from a better energy resolution and longer-lived core-hole compared to experiments conducted at the Cu L3L_{3}-edge Brookes et al. (2018).

Here, we have studied the coupling between electrons and the oxygen-related phonons in bi-layer compounds Bi2212 and NBCO, using O KK-edge RIXS with an energy resolution of about 33 meV. We have extracted the EPC strength from the data using a localized model which is based on a Lang-Firsov transformation (we refer to the model as Lang-Firsov model in the following) Lang and Firsov (1962); Geondzhian and Gilmore (2020); Geondzhian et al. (2020) that accounts for both bond-buckling and bond-stretching phonons. Combining the results of the two cuprate families, we find that EPC strength decreases with doping at 𝐪=(0.25,0)\mathbf{q_{\parallel}}=(-0.25,0) r.l.u. while this trend becomes non-monotonic at smaller momenta. The observed changes in the EPC with doping agree with the nonuniform screening effect discussed in theoretical calculations Johnston et al. (2010). We have also studied the effect of CDW correlations on EPC. Our results provide experimental insights on the general trends of EPC and allow us to better understand their role in high-TcT_{c} cuprate superconductors.

II Methods

II.1 RIXS experiments

We have performed O KK-edge RIXS studies on NBCO and Bi2212 samples with different doping levels. The NBCO samples were 100 nm films grown on SrTiO3, with two different doping levels: a heavily underdoped sample with Tc=38T_{\mathrm{c}}=38 K (UD38, p0.07p\simeq 0.07) and an optimally-doped sample with Tc=92T_{c}=92 K (OP92, p0.16p\simeq 0.16). Lattice constants were a=b=3.9Åa=b=3.9~{}\textrm{\AA}, while c=11.714Åc=11.714~{}\textrm{\AA} in the UD38 sample and c=11.74Åc=11.74~{}\textrm{\AA} in the OP one. RIXS spectra were collected at the I21 beam line of the Diamond Light Source. Energy resolution was determined to be 3434 meV by measuring the width of the elastic line on a carbon tape. We used a fixed temperature of 2020 K.

The Bi2212 samples were single crystals with three doping levels: optimally doped with Tc=91T_{c}=91 K (OP91, p0.16p\simeq 0.16), overdoped with Tc=82T_{c}=82 K (OD82, p0.19p\simeq 0.19), and overdoped with Tc=66T_{c}=66 K (OD66, p0.21p\simeq 0.21). The RIXS measurements on Bi2212 were performed using the soft x-ray inelastic beamline (SIX) 22-ID at the National Synchrotron Light Source II, Brookhaven National Laboratory Jarrige et al. (2018). The total experimental energy resolution was 33\sim 33 meV. We measured at 3535 K to avoid sample damage. We define our reciprocal lattice units (r.l.u.) in terms of the tetragonal unit cell with a=b=3.8Åa=b=3.8~{}\textrm{\AA} and c=30.89Åc=30.89~{}\textrm{\AA}.

The RIXS spectra were collected with σ\sigma incident polarization (perpendicular to the scattering plane) to maximize the charge signal. The scattering angle was fixed at 150, and we used a grazing-in geometry, which by convention means negative parallel transferred momentum 𝐪\mathbf{q_{\parallel}}, as shown in Fig. 1(g).

II.2 The Lang-Firsov Localized Models

The first model developed to describe the EPC in RIXS experiments was the single-site model introduced in Ref. Ament et al. (2011) to describe lattice excitations generated at transition metal LL-edges. The model was derived starting from the local Hamiltonian

=ϵ0nd+ω0bb+Mσnd(b+b),{\cal H}=\epsilon_{0}n_{d}+\omega_{0}b^{\dagger}b+M\sum_{\sigma}n_{d}(b+b^{\dagger}), (1)

where dσd^{\dagger}_{\sigma} and bb^{\dagger} represent the creation operators of a spin σ\sigma electron and local phonon mode, respectively, nd=σdσdσn_{d}=\sum_{\sigma}d^{\dagger}_{\sigma}d^{\phantom{\dagger}}_{\sigma}, and MM is the EPC constant, which is related to a dimensionless coupling parameter g=(M/ω0)2g=(M/\omega_{0})^{2}. This model is based on the simplifying assumption that both the electrons and phonon are local in the sense that they cannot travel through the lattice at any stage of the RIXS process. In this limit, the RIXS intensity can be computed exactly by use of a Lang-Firsov transformation, yielding

I(q,Ω)=n|An(Δ+iΓ)|2δ(ω0nΩ),I(q,\Omega)=\sum_{n}|A_{n}(\Delta+\mathrm{i}\Gamma)|^{2}\delta(\omega_{0}n-\Omega), (2)

where Δ\Delta is the detuning of the incident energy from the resonant peak, Γ\Gamma is the half-lifetime of the intermediate state of RIXS process, and

An(z)=mBn,m(g)Bm,0(g)zω0(mg)A_{n}(z)=\sum_{m}\frac{B_{n,m}(g)B_{m,0}(g)}{z-\omega_{0}(m-g)} (3)

with

Bn,m=(1)negn!m!l=0m(g)lgnm2(ml)!l!(nm+l)!B_{n,m}=(-1)^{n}\sqrt{e^{-g}n!m!}\sum_{l=0}^{m}\frac{(-g)^{l}g^{\frac{n-m}{2}}}{(m-l)!l!(n-m+l)!} (4)

being the standard Franck-Condon factor (Bn,mB_{n,m} is understood to mean Bmax(n,m),min(n,m)B_{\mathrm{max}(n,m),\mathrm{min}(n,m)}).

Reference Geondzhian and Gilmore (2020) recently generalized the single-site model to the case where the localized electron couples to more than one type of phonon. This generalization allows each phonon to be coherently excited and is relevant to the cuprates where we naturally expect coupling to different phonon branches. Here, we consider the case with two types of phonon modes (e.g. the bond-buckling and bond-stretching modes), again described by the local Hamiltonian Geondzhian and Gilmore (2020)

=\displaystyle{\cal H}=~{} ϵ0nd+λ=1,2[ωλbλbλ+Mλnd(bλ+bλ)].\displaystyle\epsilon_{0}n_{d}+\sum_{\lambda=1,2}\left[\omega^{\phantom{\dagger}}_{\lambda}b_{\lambda}^{\dagger}b^{\phantom{\dagger}}_{\lambda}+M_{\lambda}n_{d}(b^{\phantom{\dagger}}_{\lambda}+b_{\lambda}^{\dagger})\right]. (5)

The exact RIXS intensity can still be computed, yielding

I(q,Ω)=Cscalen1,n2|An1,n2(Δ+iΓ)|2δ(λ=1,2nλωλΩ),I(q,\Omega)=C_{\mathrm{scale}}\sum_{n_{1},n_{2}}|A_{n_{1},n_{2}}(\Delta+\mathrm{i}\Gamma)|^{2}\delta\big{(}\sum_{\lambda=1,2}n_{\lambda}\omega_{\lambda}-\Omega\big{)}, (6)

where

An1n2=m1m2Dn1m1n2m2(g1,g2)Dm10m20(g1,g2)Δ+iΓλ=1,2ωλ(mλgλ).\displaystyle{A_{n_{1}n_{2}}}=\sum_{m_{1}m_{2}}\frac{D_{n_{1}m_{1}}^{n_{2}m_{2}}(g_{1},g_{2})D_{m_{1}0}^{m_{2}0}(g_{1},g_{2})}{\Delta+i\Gamma-\sum_{\lambda=1,2}\omega_{\lambda}(m_{\lambda}-g_{\lambda})}. (7)

Here, CscaleC_{\mathrm{scale}} is an overall amplitude factor, independent of momentum and detuning and applies to all phonon branches, which we use to scale overall intensity to match experiment. The coefficients DD in Eq. (7) are given by a product of Franck-Condon factors

Dn1,m1n2,m2(g1,g2)=Bn1,m1(g1)Bn2,m2(g2),D_{n_{1},m_{1}}^{n_{2},m_{2}}(g_{1},g_{2})=B_{n_{1},m_{1}}(g_{1})B_{n_{2},m_{2}}(g_{2}), (8)

where gλ=(Mλ/ωλ)2g_{\lambda}=(M_{\lambda}/\omega_{\lambda})^{2}.

We stress that both models neglect all electron mobility involved in the RIXS scattering process. Nevertheless, relatively simple expressions for the RIXS intensity [Eqs. (2) and (6)] are obtained from these approximations that are useful for fitting experimental data. The parameters setting the intensity of phonon excitations in both models are the EPC strength of the different branches gλg_{\lambda}, the detuning energy Δ\Delta, and the core-hole lifetime parameter Γ\Gamma. Throughout, we treat gλg_{\lambda} as a fitting parameter and set Δ\Delta based on experimental conditions. We will also fix Γ=0.15\Gamma=0.15 eV Lee et al. (2013); Johnston et al. (2016); Braicovich et al. (2020) for all the samples investigated, since the intrinsic lifetime of the state is mostly determined by Auger recombination rate and so independent of the material studied.

A recent theoretical study in the single carrier limit Bieniasz et al. (2021) found that the local model for the RIXS cross-section produces stronger phonon features in comparison to a fully itinerant model. Moreover, the discrepancy between the models becomes more pronounced as the strength of the localizing core-hole potential is reduced. This result suggests that fitting a localized model to phonon spectra may underestimate the strength of the EPC. In the single carrier limit, electron mobility in the intermediate state can also introduce a momentum dependence in the intensity of the phonon excitations that is unrelated to the momentum dependence of the EPC constant M(𝐤,𝐪)M({\bf k},{\bf q}). With these caveats in mind, we proceed by fitting the experimental data with the localized models with the assumption that doing so will allow us to extract general trends in the data. For these reasons, the absolute values of the extracted couplings obtained here should not be equated to the value entering a microscopic Hamiltonian.

III Experimental results

Refer to caption
Figure 1: Overview of data on NBCO and Bi22212 at the O KK-edge. (a, b) XAS spectra for different doping values of Bi2212 and NBCO, respectively. The legends indicate the critical temperature of the samples. The relevant resonances are indicated with blue and red arrows. (c, d) Overview of RIXS spectra in Bi2212 OP91 and NBCO OP92, respectively. The excitation energies for the different spectra are highlighted as ticks in the XAS, using the same colour code. (e) Example of RIXS spectrum at Oxygen KK-edge collected on NBCO OP92 at 528.2 eV. Labels indicate the different excitations. (f) Zoom-in of the low-energy regions of RIXS spectra for all the samples acquired at H=0.15H=-0.15 r.l.u. and at the energy corresponding to the plane resonance. The elastic peak has been subtracted for clarity. The spectra have been normalized to the spectral weight between 1 and 7 eV corresponding to the CT excitation range. (g) Sketch of the experimental geometry. The incoming and outgoing x-rays are reported in yellow arrows, and the incident (θ\theta) and scattering (2θ2\theta) angles are highlighted. The component of the transferred momentum parallel to the CuO2 planes is reported in light green.

III.1 RIXS spectra at O KK-edge

Refer to caption
Figure 2: XAS and phonon resonances in Bi2212 OP91 and NBCO OP92. (a, b) XAS spectra (grey shade area), the spectral area of phonon (orange circles) and charge-transfer excitations (green diamonds) for Bi2212 OP91 and NBCO OP92, respectively. The red and blue dashed vertical lines indicate the chain and plane resonant energies in NBCO OP92, respectively. (c) RIXS spectra collected at the chain (red) and plane (blue) resonances in NBCO OP92 after subtracting the elastic peaks for comparison.

The X-ray absorption spectra (XAS) of our Bi2212 and NBCO samples are displayed in Fig. 1(a) and Fig. 1(b), respectively. It is easy to identify the Zhang-Rice singlet (ZRS) pre-peak in the O KK-edge XAS of Bi2212, which originates from the hybridization of the in-plane O 2pσ2p_{\sigma} and Cu 3dx2y23d_{x^{2}-y^{2}} orbitals Hawthorn et al. (2011). Both the energy and amplitude of the ZRS peak change with doping as expected. On the other hand, we observe two different structures in NBCO whose intensity increases with doping: a main peak at 528.2 eV, and another peak at a lower energy (527.6\sim 527.6 eV). The former one is associated with doped holes in the CuO2 planes, while the latter originates from the CuO3 chains, which act as a charge reservoir in NBCO Hawthorn et al. (2011). Our UD38 sample has short chains in both in-plane directions due to the tetragonal structure, while the OP92 sample has fully developed long chains, and due to the twin structure, they exist in both in-plane directions. The comparison between UD38 and OP92 sample also nicely highlights the decrease in spectral weight of the upper Hubbard band (UHB) peak around 529.5 eV.

To investigate the electron-phonon interaction, we have acquired RIXS spectra at different incident X-ray energies across the doping-related resonances. Figs. 1(c) and 1(d) present an overview of the RIXS spectra for Bi2212 OP91 and NBCO OP92, respectively. The incident energy of each scan is indicated by the tick marks with the same color in panels (a) and (b) for the two compounds. As highlighted in Fig. 1(e), we recognize four features in the spectra: the elastic peak, the phonons (and their overtones) below 0.2 eV, the broad charge excitations peaked around 0.8 eV, and some higher-energy excitations above 1.5 eV from dddd and charge-transfer (CT) excitations. Moreover, we also identify some spectral weight around 0.5 eV, which could be due to broad bimagnon excitations Bisogni et al. (2012). To highlight the phonon contribution to the spectra and compare intensities between different samples at resonant energy, we removed the elastic peak and normalized the spectra to their respective dd+CT excitations as summarized in Fig. 1(f). The phonon excitations have a larger spectral weight in NBCO and the exciations in Bi2212 OD82 (pcp_{c}=0.19) have a stronger intensity than those observed in the OP91 and OD66 samples.

To precisely determine the resonant energy of the phonon signal (i.e., the energy at which the phonon intensity is maximum), we have extracted their integrated spectral weight. This value was determined by removing the elastic peak and charge excitations from RIXS spectra, and then integrating the infrared region of the RIXS spectra (03000-300 meV). As a reference, we have also extracted the spectral weight of CT excitations by integrating the spectra in a 3 eV window centered around their maximum value. The integrated spectral weight of the phonon and CT excitations are plotted in Figs. 2(a) and 2(b) for Bi2212 OP91 and NBCO OP92, respectively, normalized to their maximum value and superimposed over the corresponding XAS spectra. We notice that the one-dimensional breathing mode of the chains has a stronger EPC strength than its two-dimensional counterpart, which is evident from the higher intensity of its first overtone.

The intensity of CT excitations follows closely the XAS spectra in both samples, confirming the stability of our incident energy and the independence of the lifetime of the intermediate state. However, the phonon has a large shift of 200250200-250 meV relative to the maximum of XAS, both in Bi2212 and NBCO. The presence and magnitude of this energy shift are puzzling. A small displacement, of the order of the phonon energy (50100\approx 50-100 meV) between the maximum of the XAS and the maximum of phonon intensity has been observed before Feng et al. (2020); Peng et al. (2020) and can be accounted for by the localized Lang-Firsov model with a strong EPC and a long core-hole lifetime. The large value observed here, however, requires a different explanation. As suggested in Ref. Geondzhian and Gilmore (2018), a possible origin could be a change in the curvature of the potential energy surface between the ground and intermediate state of the RIXS process.

III.2 Energy detuning measurements

Refer to caption
Figure 3: Detuning data on Bi2212 OP91 (left panels) and NBCO UD38 (right panels). Incident energies were chosen around the plane resonance for Bi2212 OP91 and of chain resonance for NBCO AF, for the reasons discussed in the text. The momentum transfer was (0.15,0)(-0.15,0) r.l.u. for Bi2212 OP91 and (0.25,0)(-0.25,0) r.l.u. for NBCO AF. Solid lines are the fitting curves, decomposed into the various phonon branches, obtained through the global fitting discussed in the main text. We have highlighted that, in the case of plane resonance of Bi2212, the phonon resonance is shifted with respect to the maximum of the XAS. A failure of the model is evident at large detuning energies.

We initially attempted to use the energy detuning method to extract the value of the EPC for the buckling and breathing branches, which has been described in detail in Ref. Braicovich et al. (2020) and has been successfully applied at the Cu L3L_{3} edge Braicovich et al. (2020); Rossi et al. (2019). This approach is based on the single-site model presented in Sec. II.2 and consists in acquiring scans at different incident energies, at and below the resonant peak in the XAS. The EPC of the different phonon branches can then be extracted by fitting the intensities of the different phonon branches (and of their higher harmonics) with Eq. (6). To reduce the number of parameters, we fit all of the spectra for a given sample at the same time, using a single overall scale factor and fixing Γ=0.15\Gamma=0.15 eV Lee et al. (2013); Johnston et al. (2016). In this way, the only free parameters are the EPC strengths MλM_{\lambda} of the two branches and their energies Ωλ\Omega_{\lambda}. We have also subtracted the elastic peaks and considered only the inelastic parts, since the elastic peaks can carry additional contribution from other types of scattering.

As highlighted in Sec. III.1, one difficulty is that the maximum of the phonon signal is displaced from the maximum of XAS spectra by 200250\sim 200-250 meV. While the employed model can account for a small displacement between the two energies (of the order of ωλ\omega_{\lambda}) in the strong coupling limit, the value of MλM_{\lambda} that is needed is too large and would produce phonon harmonics inconsistent with the RIXS spectra. Therefore, we have decided to artificially put the zero of the detuning in the model at the phonon resonance.

To test the validity of the detuning approach, we have used two different datasets: the chain resonance of NBCO OP92 and the plane resonance of Bi2212 OP91. We chose the chain resonance for NBCO for two reasons. First, it selects the one-dimensional CuO3 chains as opposed to the two-dimesional CuO2 planes of Bi2212, which allows us verify whether there are differences related to dimensionality. Secondly, the chain resonance energy (527.6527.6 eV) is lower than that of the plane (528.2528.2 eV). This difference means that the detuning energies of the chain are not obscured by additional contributions from the plane resonance, while the detuning energies of the plane overlap with the chains.

The spectra with the corresponding fittings are presented in Fig. 3. The solid orange and purple lines are the bond-buckling and bond-stretching phonons including the overtones, respectively, while the green shaded area represents the mixed coherent multi-phonons. However, it is evident from the spectra that the detuning method does not reproduce the measured intensities for all incident photon energies. While the agreement is quite good close to the resonance energy for both the cases, the calculated spectra far from the resonance significantly underestimate (by more than 50\sim 50%) the intensity of the phonon branches. This failure of the detuning method at the ligand KK edge is not completely unexpected given the very different nature of the intermediate state between the Cu L3L_{3}- and the O KK-edges, and of the localized assumptions underpinning the Lang-Firsov models. Indeed, at the Cu L3L_{3}-edge, the excited electron is more strongly bound to the core hole in comparison to the O KK-edge. This means that the excited electron is more free to propagate around the lattice during its lifetime in the latter case. Reference Bieniasz et al. (2021) recently developed a model for an intermediate state consisting of an itinerant electron interacting with phonons and the core hole in an otherwise empty band. One of the main results derived there is that the detuning curves are quantitatively different from the localized model; they have a weaker dependence on the EPC when the core-hole potentials are weak and they generally decay faster as a function of the detuning energy in comparison to the localized model. We observe a small difference between the decay of the intensity of the two phonon branches with detuning, as evident from Fig. 3. This observation suggests that the real situation close to the Fermi energy in cuprates is intermediate between the two approaches.

Refer to caption
Figure 4: (a) Fits of RIXS spectra after subtracting the high-energy excitations for Bi2212 OD82, taken at momenta transfer H=0.2H=-0.2 r.l.u.. Elastic peak is fitted by a Gaussian function determined by the energy resolution. The phonon excitations are fitted with the Frank-Condon model considering two phonons, i.e. the buckling mode (40\sim 40 meV) and stretching mode (70\sim 70 meV). (b) Phonon components are shown after subtracting the elastic peak. (c-e) Doping dependence of phonon amplitude for Bi2212, which shows a maximum at pc=0.19p_{c}=0.19. (f-h) Doping dependence of the electron-phonon coupling strengths MM for both NBCO and Bi2212. The SC phase diagram Fradkin et al. (2010) of cuprates (shaded area) is displayed in (f).

III.3 Effect of doping on EPC

Determining the EPC as a function of doping is crucial for understanding its potential contribution to dd-wave pairing. For this purpose, we fit the RIXS spectrum at the resonant energy using Eq. (6), after subtracting the elastic, charge, and CT excitations from the data, as shown in Fig. 4. As discussed above, the use of Eq. (6) is likely to underestimate the overall strength of the EPC, particularly at the O KK-edge, where electron itinerancy in the RIXS intermediate state can be more pronounced than at the Cu LL-edge Bieniasz et al. (2021). We nevertheless proceed with Eq. (6) in a comparative analysis to identify trends in the data, both as a function of doping and between Bi2212 and NBCO. Presumably, the degree of itinerancy in these systems is comparable for similar doping levels. Thus, the analysis for both samples should be subject to the same systematic error when applying the local model to the data. One should keep in mind that the absolute magnitudes of EPC constants extracted in this analysis cannot be mapped directly onto the Mλ(𝐪)M_{\lambda}({\bf q}) entering in any microscopic Hamiltonian.

To fit the data, we consider the coherent contribution from two phonon modes – one from the bond-buckling branch (ω140\omega_{1}\simeq 40 meV) and one from the bond-stretching branch (ω270\omega_{2}\simeq 70 meV) – while treating the coupling strengths MλM_{\lambda} as fitting parameters. We also introduce a prefactor CscaleC_{\mathrm{scale}}, which multiplies the entire spectrum by a constant amount. We are able to obtain a consistent fit to the doping-dependent RIXS spectra using this approach. The doping evolution of the extracted amplitude CscaleC_{\mathrm{scale}} for the Bi2212 sample are plotted in Figs. 4(c-e) for different momentum transfers, while the corresponding MλM_{\lambda} values are displayed in Figs. 4(f-h). For comparison, we have also overlaid the doping dependence of the extracted MλM_{\lambda} in NBCO at H=0.25H=-0.25 r.l.u. in Fig. 4(f). Notably, the values of the extracted coupling in Bi2212 and NBCO are almost identical at optimal doping, which provides additional post-hoc justification for our comparative approach.

For momentum transfer H=0.25H=-0.25 r.l.u., we find that the EPC of the lower-energy buckling mode is stronger than the EPC of the breathing mode in both samples and for all the measured doping levels. This observation is consistent with the previous Cu LL-edge RIXS measurements on NBCO UD38 sample for small momentum transfers Braicovich et al. (2020) and prior ARPES experiments that infer the strongest coupling to the bond-buckling phonon modes Cuk et al. (2004, 2005); Johnston et al. (2010, 2012). As the hole concentration increases, we find that the extracted MM values for both branches decrease monotonically at H=0.25H=-0.25 r.l.u.. This behavior can be understood as being a consequence of the improved electronic screening as the system becomes more metallic Johnston et al. (2010). The situation is quite different for smaller momenta. For H=0.20H=-0.20 r.l.u., the inferred EPC constants are very weakly dependent on doping, while at H=0.15H=-0.15 r.l.u. the doping dependence even becomes non-monotonic, with both branches becoming more strongly coupled near pc=0.19p_{c}=0.19.

The doping dependence summarized here in both compounds is consistent with the poor screening scenario discussed in Ref. Johnston et al. (2010). There, it was shown that the poor conductivity perpendicular to the CuO2 plane results in a nonuniform screening of the electron-phonon interaction whereby it was well screened for large momentum transfers but poorly screened at small 𝐪{\bf q}. The suppression of large 𝐪{\bf q} coupling relative to small 𝐪{\bf q} will increase the EPC in the dd-wave channel, suggesting that the contribution of these modes to pairing increases as the doping increases and may become largest at optimal doping.

III.4 Influence of CDW on EPC

Both optimally doped NBCO Ghiringhelli et al. (2012); Blanco-Canosa et al. (2014) and Bi2212 Hashimoto et al. (2014); da Silva Neto et al. (2014); Lee et al. (2021) present a CDW ordering with qCDW0.29q_{\text{CDW}}\simeq 0.29 r.l.u., though the CDW signals become weaker with respect to underdoped samples Blanco-Canosa et al. (2014). Previous RIXS studies at Cu L3L_{3}-edge and O KK-edge on doped cuprates have found an anomalous enhancement of phonon intensity near the CDW wavevector, which is ascribed to the presence of dispersive CDW excitations that overlap and interfere with the phonon excitations Chaix et al. (2017); Li et al. (2020). We have, therefore, searched for possible enhancements of EPC of buckling and breathing branches due to the presence of charge order.

Employing again the generalized Lang-Firsov model [Eq. (7)] that accounts for coupling to both modes, we have extracted the evolution of the EPC values and the overall scaling prefactor along the (H,0)(H,0) direction. Figure LABEL:fig:momentumdep shows the momentum dependence of the phonon amplitudes (CscaleC_{\mathrm{scale}}) and MλM_{\lambda} values in optimally doped Bi2212 and NBCO. Evidently, the phonon amplitudes increase monotonically on approaching the qCDWq_{\text{CDW}}. At the same time, the EPC does not show a clear trend with M1M_{1} larger than M2M_{2} for all our measured momenta. These results indicates that while the phonon excitations increase in intensity in proximity of the CDW, there is no clear corresponding enhancement in the relative intensity of the phonon overtones. This behavior can be reconciled if there is an overlapping contribution to the elastic line intensity due to the CDW excitations and no corresponding increase in the EPC close to 𝐪CDW{\bf q}_{\mathrm{CDW}}. One should, therefore, be careful in interpreting directly the phonon intensity as the EPC strength at O KK-edge due to the longer intermediate state. The explanation of this discrepancy remains a topic for further investigation.

IV Conclusions

We have measured the phonon dynamics in two families of cuprates superconductors as a function of incident energy, momentum and doping, ranging from strongly underdoped to overdoped samples. Our experimental results highlight that the localized Lang-Firsov model does not describe phonon spectra when the incident energy is detuned away from the resonance peak. We believe that this discrepancy is a fingerprint of an itineracy of the excited electron and phonon cloud, something which is not captured in the localized model Bieniasz et al. (2021). The detuning method, which has been successfully applied at the Cu LL-edge to extract the EPC, is therefore not applicable at the O KK-edge within the present theoretical framework.

By extracting the momentum and doping dependence of EPC measured at resonant x-ray energy, we have revealed a systematic decrease of EPC with increasing doping from strongly underdoped to overdoped samples at high momentum transfer, while closer to the Γ\Gamma point EPC seems to be less affected by screening, consistently with previous calculations Johnston et al. (2010). We have also detected an anomalous dependence of phonon intensity close to the CDW wave vector. However, we have revealed that momentum dependence of the EPC measured at O KK-edge is different from what measured at Cu LL-edge Braicovich et al. (2020), implying that the RIXS cross section contains additional form factors. The present results show that investigations of the phonon dynamics by O KK-edge RIXS in presence of delocalized electrons has to be made with precautions in the framework of the models developed so far, and new approaches have to be developed to reach a quantitative reliability of results.

Acknowledgements.
Y. Y. P. is grateful for financial support from the Ministry of Science and Technology of China (Grant No. 2019YFA0308401) and the National Natural Science Foundation of China (Grant No. 11974029). P. A. acknowledges support from the Gordon and Betty Moore Foundation, grant no. GMBF-9452. X. J. Z. thank the financial support from the National Natural Science Foundation of China (Grant no. 11888101), the National Key Research and Development Program of China (Grant no. 2016YFA0300300) and the Strategic Priority Research Program (B) of the Chinese Academy of Sciences (Grant no. XDB25000000). S. J. is supported by the National Science Foundation under Grant No. DMR-1842056. L. M., M. M. S. and G. G. acknowledge support by the project PRIN2017 “Quantum-2D” ID 2017Z8TS5B of the Ministry for University and Research (MIUR) of Italy. R.A. acknowledges support by the Swedish Research Council (VR) under the Project 2020-04945. This research used Diamond Light Source beam line I21 under the proposal SP20012. N. B. B thanks Diamond Light Source for hosting him at the time the experiment was carried out at I21. This research used the beam line 2-ID of the National Synchrotron Light Source II, a U.S. Department of Energy (DOE) Office of Science User Facility operated for the DOE Office of Science by Brookhaven National Laboratory under Contract No. DE-SC0012704. The work at BNL was supported by the US Department of Energy, office of Basic Energy Sciences, contract no. DOE-sc0012704.

References

  • Bardeen et al. (1957) J. Bardeen, L. N. Cooper,  and J. R. Schrieffer, Phys. Rev. 108, 1175 (1957).
  • McMillan and Rowell (1965) W. L. McMillan and J. M. Rowell, Phys. Rev. Lett. 14, 108 (1965).
  • Lanzara et al. (2001) A. Lanzara, P. V. Bogdanov, X. J. Zhou, S. A. Kellar, D. L. Feng, E. D. Lu, T. Yoshida, H. Eisaki, A. Fujimori, K. Kishio, J.-I. Shimoyama, T. Noda, S. Uchida, Z. Hussain,  and Z.-X. Shen, Nature 412, 510 (2001).
  • Cuk et al. (2004) T. Cuk, F. Baumberger, D. H. Lu, N. Ingle, X. J. Zhou, H. Eisaki, N. Kaneko, Z. Hussain, T. P. Devereaux, N. Nagaosa,  and Z.-X. Shen, Phys. Rev. Lett. 93, 117003 (2004).
  • Devereaux et al. (2004) T. P. Devereaux, T. Cuk, Z.-X. Shen,  and N. Nagaosa, Phys. Rev. Lett. 93, 117004 (2004).
  • Cuk et al. (2005) T. Cuk, D. H. Lu, X. J. Zhou, Z.-X. Shen, T. P. Devereaux,  and N. Nagaosa, Phys. Status Solidi B 242, 11 (2005).
  • Graf et al. (2008) J. Graf, M. d’Astuto, C. Jozwiak, D. R. Garcia, N. L. Saini, M. Krisch, K. Ikeuchi, A. Q. R. Baron, H. Eisaki,  and A. Lanzara, Phys. Rev. Lett. 100, 227002 (2008).
  • He et al. (2018) Y. He, M. Hashimoto, D. Song, S.-D. Chen, J. He, I. M. Vishik, B. Moritz, D.-H. Lee, N. Nagaosa, J. Zaanen, T. P. Devereaux, Y. Yoshida, H. Eisaki, D. H. Lu,  and Z.-X. Shen, Science 362, 62 (2018).
  • Johnston et al. (2012) S. Johnston, I. M. Vishik, W. S. Lee, F. Schmitt, S. Uchida, K. Fujita, S. Ishida, N. Nagaosa, Z. X. Shen,  and T. P. Devereaux, Phys. Rev. Lett. 108, 166404 (2012).
  • Lee et al. (2006) J. Lee, K. Fujita, K. McElroy, J. A. Slezak, M. Wang, Y. Aiura, H. Bando, M. Ishikado, T. Masui, J. X. Zhu, A. V. Balatsky, H. Eisaki, S. Uchida,  and J. C. Davis, Nature 442, 546 (2006).
  • Zhao (2009) G. Zhao, Phys. Rev. Lett. 103, 236403 (2009).
  • Zhao (2007) G. Zhao, Phys. Rev. B 75, 214507 (2007).
  • Bulut and Scalapino (1996) N. Bulut and D. J. Scalapino, Phys. Rev. B 54, 14971 (1996).
  • Maier et al. (2008) T. A. Maier, D. Poilblanc,  and D. J. Scalapino, Phys. Rev. Lett. 100, 237001 (2008).
  • Scalapino (2012) D. J. Scalapino, Rev. Mod. Phys. 84, 1383 (2012).
  • Tsuei and Kirtley (2000) C. C. Tsuei and J. R. Kirtley, Rev. Mod. Phys. 72, 969 (2000).
  • Johnston et al. (2010) S. Johnston, F. Vernay, B. Moritz, Z.-X. Shen, N. Nagaosa, J. Zaanen,  and T. P. Devereaux, Phys. Rev. B 82, 064513 (2010).
  • Fradkin et al. (2015) E. Fradkin, S. A. Kivelson,  and J. M. Tranquada, Rev. Mod. Phys. 87, 457 (2015).
  • Keimer et al. (2015) B. Keimer, S. A. Kivelson, M. R. Norman, S. Uchida,  and J. Zaanen, Nature 518, 179 (2015).
  • Comin and Damascelli (2016) R. Comin and A. Damascelli, Annu. Rev. Condens. Matter Phys. 7, 369 (2016).
  • Agterberg et al. (2020) D. F. Agterberg, J. S. Davis, S. D. Edkins, E. Fradkin, D. J. Van Harlingen, S. A. Kivelson, P. A. Lee, L. Radzihovsky, J. M. Tranquada,  and Y. Wang, Annu. Rev. Condens. Matter Phys. 11, 231 (2020).
  • Arpaia et al. (2018) R. Arpaia, E. Andersson, E. Trabaldo, T. Bauch,  and F. Lombardi, Phys. Rev. Materials 2, 024804 (2018).
  • Arpaia and Ghiringhelli (2021) R. Arpaia and G. Ghiringhelli, Journal of the Physical Society of Japan 90, 111005 (2021).
  • Wahlberg et al. (2021) E. Wahlberg, R. Arpaia, G. Seibold, M. Rossi, R. Fumagalli, E. Trabaldo, N. B. Brookes, L. Braicovich, S. Caprara, U. Gran, G. Ghiringhelli, T. Bauch,  and F. Lombardi, Science 373, 1506 (2021).
  • Zhu et al. (2015) X. Zhu, Y. Cao, J. Zhang, E. W. Plummer,  and J. Guo, Proceedings of the National Academy of Sciences 112, 2367 (2015).
  • Chaix et al. (2017) L. Chaix, G. Ghiringhelli, Y. Y. Peng, M. Hashimoto, B. Moritz, K. Kummer, N. B. Brookes, Y. He, S. Chen, S. Ishida, Y. Yoshida, H. Eisaki, M. Salluzzo, L. Braicovich, Z.-X. Shen, T. P. Devereaux,  and W.-S. Lee, Nat. Phys. 13, 952 (2017).
  • Miao et al. (2018) H. Miao, D. Ishikawa, R. Heid, M. Le Tacon, G. Fabbris, D. Meyers, G. D. Gu, A. Q. R. Baron,  and M. P. M. Dean, Phys. Rev. X 8, 011008 (2018).
  • Lin et al. (2020) J. Q. Lin, H. Miao, D. G. Mazzone, G. D. Gu, A. Nag, A. C. Walters, M. García-Fernández, A. Barbour, J. Pelliciari, I. Jarrige, M. Oda, K. Kurosawa, N. Momono, K.-J. Zhou, V. Bisogni, X. Liu,  and M. P. M. Dean, Phys. Rev. Lett. 124, 207005 (2020).
  • Li et al. (2020) J. Li, A. Nag, J. Pelliciari, H. Robarts, A. Walters, M. Garcia-Fernandez, H. Eisaki, D. Song, H. Ding, S. Johnston, R. Comin,  and K.-J. Zhou, Proceedings of the National Academy of Sciences 117, 16219 (2020).
  • Peng et al. (2020) Y. Y. Peng, A. A. Husain, M. Mitrano, S. X.-L. Sun, T. A. Johnson, A. V. Zakrzewski, G. J. MacDougall, A. Barbour, I. Jarrige, V. Bisogni,  and P. Abbamonte, Phys. Rev. Lett. 125, 097002 (2020).
  • Timusk and Statt (1999) T. Timusk and B. Statt, Reports on Progress in Physics 62, 61 (1999).
  • Badoux et al. (2016) S. Badoux, W. Tabis, F. Laliberté, G. Grissonnanche, B. Vignolle, D. Vignolles, J. Béard, D. Bonn, W. Hardy, R. Liang, et al.Nature 531, 210 (2016).
  • Ament et al. (2011) L. J. P. Ament, M. van Veenendaal,  and J. van den Brink, EPL 95, 27008 (2011).
  • Devereaux et al. (2016) T. P. Devereaux, A. M. Shvaika, K. Wu, K. Wohlfeld, C. J. Jia, Y. Wang, B. Moritz, L. Chaix, W.-S. Lee, Z.-X. Shen, G. Ghiringhelli,  and L. Braicovich, Phys. Rev. X 6, 041019 (2016).
  • Braicovich et al. (2020) L. Braicovich, M. Rossi, R. Fumagalli, Y. Peng, Y. Wang, R. Arpaia, D. Betto, G. M. De Luca, D. Di Castro, K. Kummer, M. Moretti Sala, M. Pagetti, G. Balestrino, N. B. Brookes, M. Salluzzo, S. Johnston, J. van den Brink,  and G. Ghiringhelli, Phys. Rev. Research 2, 023231 (2020).
  • Bieniasz et al. (2021) K. Bieniasz, S. Johnston,  and M. Berciu, SciPost Phys. 11, 62 (2021).
  • Rossi et al. (2019) M. Rossi, R. Arpaia, R. Fumagalli, M. Moretti Sala, D. Betto, K. Kummer, G. M. De Luca, J. van den Brink, M. Salluzzo, N. B. Brookes, L. Braicovich,  and G. Ghiringhelli, Phys. Rev. Lett. 123, 027001 (2019).
  • Chen et al. (1991) C. T. Chen, F. Sette, Y. Ma, M. S. Hybertsen, E. B. Stechel, W. M. C. Foulkes, M. Schulter, S.-W. Cheong, A. S. Cooper, L. W. Rupp, B. Batlogg, Y. L. Soo, Z. H. Ming, A. Krol,  and Y. H. Kao, Phys. Rev. Lett. 66, 104 (1991).
  • Brookes et al. (2018) N. B. Brookes, F. Yakhou-Harris, K. Kummer, A. Fondacaro, J. Cezar, D. Betto, E. Velez-Fort, A. Amorese, G. Ghiringhelli, L. Braicovich, R. Barrett, G. Berruyer, F. Cianciosi, L. Eybert, P. Marion, P. van der Linden,  and L. Zhang, Nucl. Instr. Meth. Phys. Res. A 903, 175 (2018).
  • Lang and Firsov (1962) I. G. Lang and Y. A. Firsov, Sov. Phys. JETP 16, 1301 (1962).
  • Geondzhian and Gilmore (2020) A. Geondzhian and K. Gilmore, Phys. Rev. B 101, 214307 (2020).
  • Geondzhian et al. (2020) A. Geondzhian, A. Sambri, G. M. De Luca, R. Di Capua, E. Di Gennaro, D. Betto, M. Rossi, Y. Y. Peng, R. Fumagalli, N. B. Brookes, L. Braicovich, K. Gilmore, G. Ghiringhelli,  and M. Salluzzo, Phys. Rev. Lett. 125, 126401 (2020).
  • Jarrige et al. (2018) I. Jarrige, V. Bisogni, Y. Zhu, W. Leonhardt,  and J. Dvorak, Synchrotron Radiation News 31, 7 (2018).
  • Lee et al. (2013) W. S. Lee, S. Johnston, B. Moritz, J. Lee, M. Yi, K. J. Zhou, T. Schmitt, L. Patthey, V. Strocov, K. Kudo, Y. Koike, J. van den Brink, T. P. Devereaux,  and Z. X. Shen, Phys. Rev. Lett. 110, 265502 (2013).
  • Johnston et al. (2016) S. Johnston, C. Monney, V. Bisogni, K.-J. Zhou, R. Kraus, G. Behr, V. N. Strocov, J. Málek, S.-L. Drechsler, J. Geck, T. Schmitt,  and J. van den Brink, Nat. Commun. 7, 10563 (2016).
  • Hawthorn et al. (2011) D. G. Hawthorn, K. M. Shen, J. Geck, D. C. Peets, H. Wadati, J. Okamoto, S.-W. Huang, D. J. Huang, H.-J. Lin, J. D. Denlinger, R. Liang, D. A. Bonn, W. N. Hardy,  and G. A. Sawatzky, Phys. Rev. B 84, 075125 (2011).
  • Bisogni et al. (2012) V. Bisogni, L. Simonelli, L. J. P. Ament, F. Forte, M. Moretti Sala, M. Minola, S. Huotari, J. van den Brink, G. Ghiringhelli, N. B. Brookes,  and L. Braicovich, Phys. Rev. B 85, 214527 (2012).
  • Feng et al. (2020) X. Feng, S. Sallis, Y.-C. Shao, R. Qiao, Y.-S. Liu, L. C. Kao, A. S. Tremsin, Z. Hussain, W. Yang, J. Guo,  and Y.-D. Chuang, Phys. Rev. Lett. 125, 116401 (2020).
  • Geondzhian and Gilmore (2018) A. Geondzhian and K. Gilmore, Phys. Rev. B 98, 214305 (2018).
  • Fradkin et al. (2010) E. Fradkin, S. A. Kivelson, M. J. Lawler, J. P. Eisenstein,  and A. P. Mackenzie, Annual Review of Condensed Matter Physics 1, 153 (2010).
  • Ghiringhelli et al. (2012) G. Ghiringhelli, M. Le Tacon, M. Minola, S. Blanco-Canosa, C. Mazzoli, N. B. Brookes, G. M. De Luca, A. Frano, D. G. Hawthorn, F. He, T. Loew, M. M. Sala, D. C. Peets, M. Salluzzo, E. Schierle, R. Sutarto, G. A. Sawatzky, E. Weschke, B. Keimer,  and L. Braicovich, Science 337, 821 (2012).
  • Blanco-Canosa et al. (2014) S. Blanco-Canosa, A. Frano, E. Schierle, J. Porras, T. Loew, M. Minola, M. Bluschke, E. Weschke, B. Keimer,  and M. Le Tacon, Phys. Rev. B 90, 054513 (2014).
  • Hashimoto et al. (2014) M. Hashimoto, G. Ghiringhelli, W.-S. Lee, G. Dellea, A. Amorese, C. Mazzoli, K. Kummer, N. B. Brookes, B. Moritz, Y. Yoshida, H. Eisaki, Z. Hussain, T. P. Devereaux, Z.-X. Shen,  and L. Braicovich, Phys. Rev. B 89, 220511 (2014).
  • da Silva Neto et al. (2014) E. H. da Silva Neto, P. Aynajian, A. Frano, R. Comin, E. Schierle, E. Weschke, A. Gyenis, J. Wen, J. Schneeloch, Z. Xu, S. Ono, G. Gu, M. Le Tacon,  and A. Yazdani, Science 343, 393 (2014).
  • Lee et al. (2021) W.-S. Lee, K.-J. Zhou, M. Hepting, J. Li, A. Nag, A. Walters, M. Garcia-Fernandez, H. Robarts, M. Hashimoto, H. Lu, et al.Nature Physics 17, 53 (2021).