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

Role of electron-phonon coupling in excitonic insulator candidate Ta2NiSe5

Cheng Chen These authors contributed equally to this work Department of Physics, University of Oxford, Oxford, OX1 3PU, United Kingdom Department of Applied Physics, Yale University, New Haven, Connecticut 06511, USA    Xiang Chen These authors contributed equally to this work Physics Department, University of California, Berkeley, California 94720, USA Materials Science Division, Lawrence Berkeley National Lab, Berkeley, California 94720, USA    Weichen Tang Physics Department, University of California, Berkeley, California 94720, USA Materials Science Division, Lawrence Berkeley National Lab, Berkeley, California 94720, USA    Zhenglu Li Physics Department, University of California, Berkeley, California 94720, USA Materials Science Division, Lawrence Berkeley National Lab, Berkeley, California 94720, USA    Siqi Wang Department of Applied Physics, Yale University, New Haven, Connecticut 06511, USA    Shuhan Ding Department of Physics and Astronomy, Clemson University, Clemson, South Carolina 29631, USA    Zhibo Kang Department of Applied Physics, Yale University, New Haven, Connecticut 06511, USA    Chris Jozwiak Advanced Light Source, Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA    Aaron Bostwick Advanced Light Source, Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA    Eli Rotenberg Advanced Light Source, Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA    Makoto Hashimoto Stanford Synchrotron Radiation Lightsource, SLAC National Accelerator Laboratory, Menlo Park, CA 94025, USA    Donghui Lu Stanford Synchrotron Radiation Lightsource, SLAC National Accelerator Laboratory, Menlo Park, CA 94025, USA    Jacob P.C. Ruff Cornell High Energy Synchrotron Source, Cornell University, Ithaca, New York 14853, USA    Steven G. Louie Physics Department, University of California, Berkeley, California 94720, USA Materials Science Division, Lawrence Berkeley National Lab, Berkeley, California 94720, USA    Robert J. Birgeneau Physics Department, University of California, Berkeley, California 94720, USA Materials Science Division, Lawrence Berkeley National Lab, Berkeley, California 94720, USA Department of Materials Science and Engineering, University of California, Berkeley, California 94720, USA    Yulin Chen Department of Physics, University of Oxford, Oxford, OX1 3PU, United Kingdom    Yao Wang [email protected] Department of Physics and Astronomy, Clemson University, Clemson, South Carolina 29631, USA    Yu He [email protected] Department of Applied Physics, Yale University, New Haven, Connecticut 06511, USA
Abstract

Electron-hole bound pairs, or excitons, are common excitations in semiconductors. They can spontaneously form and “condense” into a new insulating ground state – the so-called excitonic insulator – when the energy of electron-hole Coulomb attraction exceeds the band gap. In the presence of electron-phonon coupling, a periodic lattice distortion often concomitantly occurs with this exciton condensation. However, similar structural transition can also be induced by electron-phonon coupling itself, therefore hindering the clean identification of bulk excitonic insulators based on reductionistic reasoning (e.g. which instability is the “driving force” of the phase transition). Using high-resolution synchrotron x-ray diffraction and angle-resolved photoemission spectroscopy techniques, we identify key electron-phonon coupling effects in a leading excitonic insulator candidate Ta2NiSe5. These include an extensive unidirectional lattice fluctuation and an electronic pseudogap in the normal state, as well as a negative electronic compressibility in the charge-doped broken-symmetry state. In combination with first principles and model calculations, we determine a minimal lattice model and the corresponding interaction parameters that capture the experimental observations. More importantly, we show how the Coulomb and electron-phonon coupling effects can be separated on the level of lattice model, and demonstrate a general framework beyond the reductionist approach in the investigation of correlated systems with intertwined orders.

I INTRODUCTION

Correlated electron systems often contain multiple instabilities and intertwined orders, where numerous degrees of freedom jointly determine the material properties [1]. A notable example is the iron-based superconductors, where the electronic, orbital and spin channels all contribute to the rotational symmetry-breaking ground state [2]. In these systems, order parameters of the same symmetry often couple to each other, and the common reductionistic approach of ascribing the transition to a single dominant channel becomes ambivalent [3]. Another more recent example is the renewed search for bulk excitonic insulators, where electron-hole pairs spontaneously form in the ground state and cause charge insulation [4, 5]. In indirect band gap semiconductors such as 1T-TiSe2 [6], both the plasmon and phonons soften to zero energy below the phase transition temperature, indicating contributions from both the lattice and electronic degrees of freedom. Indeed, it has been extensively discussed that the presence of electron-phonon coupling, regardless of whether there is pre-existing electron-hole Coulomb attraction, can also lead to a structural distortion with an ordering wave vector matching what connects the conduction band bottom and the valence band top [7, 8]. Such a “chicken-or-egg” dilemma originates from the pursuit of a reductive description, which is not always possible in correlated materials where relevant interaction energy scales are not well-separated. The lack of a clear energy hierarchy severely limits the construction and application of reductive frameworks. One common manifestation of this issue is the often ambivalent interpretations of observations made via different experimental techniques on the same system. A promising approach to avoid this dilemma is to include all relevant interactions on equal footing in the theoretical model, determine the interaction parameters by comparing the simulation with experimental observables through various tuning methods, and subsequently make further predictions for model validation and for material design guidance [9].

Refer to caption
Figure 1: Structural phase transition in Ta2NiSe5. (a) Top view of the lattice structures across the phase transition: orthorhombic (above TsT_{s}) and monoclinic (below TsT_{s}). (b) Temperature dependence of the ±\mathrm{\pm} 1 3 4, 2 0 0 and ±\mathrm{\pm} 1 3 7 Bragg peaks from the single crystal x-ray diffraction experiments. (c) Peak positions, widths, and normalized intensities of the ±\mathrm{\pm} 1 3 4 and 2 0 0 Bragg peaks across the phase transition at TsT_{s}\sim 329 K.

In this work, we will adopt such a framework to address the ongoing search for bulk excitonic insulators in low-dimensional materials. Recent studies revealed evidence of exciton formation and strong electron-phonon coupling (EPC) in the quasi-one-dimensional (quasi-1D) ternary chalcogenide Ta2NiSe5[10, 11, 12, 13, 14, 15]. Upon warming, a q=0q=0 monoclinic-to-orthorhombic structural transition happens at TsT_{s}\sim 329 K [see Fig. 1(a)], above which a semi-metallic electronic structure is supposedly enforced by the mirror symmetry of the crystal and opposite parities of the low-energy orbitals near the Fermi momentum [16]. However, an insulating behavior is observed to persist up to 550K [10]. This is in striking contrast to archetypal metal-to-insulator transition (MIT) systems, such as the perovskite nickelates [17] and the chain compound TTF-TCNQ [18], where the higher-symmetry structure occurs concurrently with the metallic electronic state (usually called the normal state). The nature of the high-temperature electronic state in Ta2NiSe5 remains controversial [19, 20, 21, 22, 23, 24, 25, 26], including suggestions of either a regular gapless semimetal or a pseudogapped state with “preformed excitons” [25, 20, 26]. In addition, contentions remain regarding whether the system’s insulating ground state hosts an exciton condensate [11, 12, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28].

Combining high-energy x-ray diffraction (XRD) and high-resolution angle-resolved photoemission spectroscopy (ARPES) techniques, we first acquire high-resolution experimental results that will be used to test minimal many-body models. The experimental observations mainly focus on normal state properties, where the symmetry is not yet broken. These include an electronic pseudogap state over a wide temperature range above TsT_{s} in Ta2NiSe5, which occurs concomitantly with a strong unidirectional fluctuation of the lattice; a similarly pseudogapped electronic structure during the insulator-to-semimetal transition upon electron injection below TsT_{s}; and concomitant negative electronic compressibility (NEC), which suggests the thermodynamic inevitability of EPC’s contribution. We then combine experimental results with both first-principles and many-body simulations to estimate the microscopic interaction parameters, including the kinetic energy of the conduction and valence electrons, the strength of the Coulomb interaction, and the EPC vertex. Considering both interactions on an equal footing, we discuss the roles of Coulomb interaction and EPC by comparing the experimental data to the numerically simulated spectra with both, either, and none of the interaction terms. Informed by such a comparison, we attribute the pseudogap state and NEC in Ta2NiSe5 primarily to the interband electron-hole excitation facilitated by a fluctuating lattice. On the contrary, direct Coulomb attraction between electrons and holes alone is insufficient to account for the insulating gap of the system.

Refer to caption
Figure 2: Strong dynamic lattice fluctuation in Ta2NiSe5 above TsT_{s}. (a) Temperature dependence of the diffuse scattering signal along the c*-direction between the 2 0 6 and 2 0 8 Bragg peaks. All figures are plotted under the same log color scale to highlight the evolution. (b) Integrated intensity of the diffuse signal from the blue box in (a). Inset: line cuts [black dashed line in (a)] of the diffuse signal at TT = 305 K, 329 K, and 360 K, respectively.

The organization of this paper is as follows. We report the lattice fluctuations evidenced from XRD and the pseudogap state measured by ARPES in the high-temperature phase in Secs. II and III respectively. Next, we discuss the non-thermal tuning in the low-temperature phase, the emergence of NEC, and bound the Coulomb interaction strength in Sec. IV and Sec. V. We then construct a minimal lattice model, estimate the EPC vertex strength, and explain the experimental spectra using the first-principles and many-body simulations in Sec. VI. The combined impacts of Coulomb interactions and electron-phonon couplings are discussed in Sec. VII. Finally, we conclude by discussing the implication of this framework to other quantum materials with intertwined orders, and make a few predictions for the Ta2NiSe5 system in Sec. VIII.

II High-Temperature Lattice Fluctuations

High-quality Ta2NiSe5 single crystals were grown via the chemical vapor transfer method with iodine (I2) as the transport agent [10, 12, 24]. Starting materials, composed of Ta powder (99.99%), Ni powder (99.99%), Se powder (99.99%) with a nominal molar ratio 2:1:5, were fully ground and mixed inside the glovebox. An additional 50 mg of iodine was then added to the mixed powder before it was vacuumed, backfilled with 1/3 Argon, and sealed inside a quartz tube (inner diameter 8 mm, outer diameter 12 mm, and a length of 120 mm). The sealed quartz tube was placed horizontally inside a muffle furnace during the growth. The hot end reaction temperature was set to 950 C and the cold end was left in air with the temperature stabilized at 850 C. Long and thin single crystals were harvested by quenching the furnace in air after one week of reaction. Residue iodine on the surface of the crystals was removed with ethanol.

Ta2NiSe5 crystallizes in a layered structure stacked via van der Waals interactions [Fig. 1(a)]. Within each layer, the Ta and Ni atoms form a chain structure along the a-axis of the crystal. Our XRD data reveals that the system undergoes a structural phase transition from a high-temperature orthorhombic Cmcm phase to a low-temperature monoclinic C2/c phase at TsT_{s}, in line with the previous report [10]. During the transition, the Ta atoms slightly shear along the chain direction, resulting in an increase of the β\beta angle from 9090^{\circ} to 90.5390.53^{\circ} [exaggerated in Fig. 1(a)]. The second-order nature of this transition is indicated by the continuous separation of the ±\mathrm{\pm} 1 3 4 and ±\mathrm{\pm} 1 3 7 nuclear Bragg peaks, which are identical above TsT_{s} [Fig. 1(b)]. Here, the quantity |β90||\beta-90^{\circ}|, or equivalently the average displacement xx of the Ta atoms, is used as the structural order parameter of this phase transition. We further compare the width of the Bragg peaks below and above TsT_{s} [Fig. 1(c)]. The peak width is nearly identical across the structural transition, ruling out any vestigial static monoclinic phase in the high-temperature normal state.

In addition to the immediate nullification of the lattice order parameter above TsT_{s}, we also observe a pronounced decrease of the 2 0 0 nuclear Bragg peak intensity [Fig. 1(c)], indicating the presence of strong lattice fluctuations. Specifically, this fluctuation originates from the squared average of the atomic displacement xx, and is further corroborated by the dramatically enhanced c*-direction diffuse scattering signal over a broad temperature range above TsT_{s} [Figs. 2(a)-(b)]. Such a diffuse scattering signal can occur when the atomic position deviates from perfect periodicity along the c*-direction. Since static lattice disorder is ruled out, a strong dynamic lattice fluctuation in the form of inter-chain sheering is the most plausible scenario. This aligns with previous reports of the soft transverse acoustic phonon [29] and the electronically coupled optical phonons in Ta2NiSe5[13, 14, 15].

III High-Temperature Pseudogap

The structural phase transition at TsT_{s} could potentially drive an MIT transition in the electronic channel, taking the route of band folding in mean-field theory, which is discussed in many charge density wave (CDW) materials [30, 31]. This synchronization of structural symmetry breaking and electronic MIT is also expected according to the first principles calculation of the electronic structure of Ta2NiSe5  based on density functional theory (DFT). As shown in Fig. 3(a), the relaxed ground-state structure exhibits a lattice order parameter β\beta comparable to the experimental data (detailed in Sec. VI), and the calculated band structure exhibits an energy gap. When the structure is constrained to maintain orthorhombicity, the system is a gapless semimetal. However, previous transport studies have indicated only a small change of the resistivity slope across the structural transition at TsT_{s}, with an insulating behavior persisting up to 550 K, more than 60% above TsT_{s}[10].

Refer to caption
Figure 3: Temperature dependence of the single-particle gap in Ta2NiSe5. (a) DFT calculations of the single-particle bands below and above TsT_{s} using a Fermi-Dirac smearing. States from Ta dx2y2d_{x^{2}-y^{2}} and Ni dxyd_{xy} orbitals are emphasized, corresponding to the linear horizontal (LH, blue) and linear vertical (LV, red) channels in the ARPES experiments, respectively. (b) Photoemission spectra along X-Γ\Gamma-X direction in both photon polarization channels at select temperatures. (c) Temperature-dependent integrated energy distribution curves (EDCs) from the spectra in (b) for both photon polarization channels. (d) Temperature dependence of the integrated spectral weight around EF±E_{F}\mathrm{\pm} 30 meV. The vertical purple dashed line marks the TsT_{s}. (e) Photoemission spectra at TT = 380 K, with intensity normalized along the energy axis to highlight the underlying dispersion.

To reveal the origin of this asynchronized electronic MIT and structural symmetry breaking, we investigate the single-particle spectral function of Ta2NiSe5 with high-resolution ARPES. Figure 3(b) presents the APRES data along the Γ\Gamma-X high symmetry direction [geometry illustrated in Fig. S1] for various temperatures below and above TsT_{s}. Here, the high statistics and energy resolution enable spectra restoration up to \sim150 meV above the Fermi level (EFE_{F}) after dividing the resolution-convolved Fermi-Dirac function [32]. As illustrated in Fig. 3(b), the low-temperature spectra (T<TsT<T_{s}) show a pronounced single-particle gap, where the dispersion and orbital compositions are consistent with the DFT calculations in Fig. 3(a). Therefore, the electronic state below TsT_{s} aligns with the mean-field theory for a second-order phase transition.

In contrast, the high-temperature electronic structure (T>TsT>T_{s}) exhibits anomalous behavior beyond the DFT prediction. We observe a pronounced spectral weight depletion around EFE_{F} (±\mathrm{\pm} 100 meV), despite the disappearance of the low-temperature flat valence band top [Fig. 3(b)]. Such a strong intensity depletion cannot be addressed in the generic band theory: drastic orbital character change alone cannot account for the missing spectral weight in both Ta and Ni states probed in both photon polarization channels. The dipole transition matrix-element effect is also unlikely, given the robust spectral weight depletion seen across a broad range of photon energies and Brillouin zones [Fig. S2].

The relative evolution of this spectral depletion to the structural phase transition can be better illustrated from the integrated energy distribution curves (EDCs) [Fig. 3(c)]. The structure-related band reconstruction at high binding energy (marked by the shifting of the valence band from -0.11 eV to -0.17 eV) happens within a relatively small temperature range below TsT_{s}  in line with the second-order phase transition. In contrast, the low-energy spectral weight (EF±30E_{F}\pm 30 meV), drops continuously from 380 K to 120 K with little sensitivity to the structural phase transition [Fig. 3(d)]. Such persistent single-particle spectral depletion at EFE_{F} naturally accounts for the insulating behavior in resistivity and optical conductivity above TsT_{s}[10, 28].

Despite this spectral weight depletion, however, the band dispersions above TsT_{s} do not suffer from apparent gap opening or band hybridization. This can be seen from the band dispersions extracted from the ARPES spectra, normalized along the energy axis to remove the intensity depletion effect. As shown in Fig. 3(e), we identify uninterrupted conduction and valence band dispersions (recovered up to 5kBTs\mathrm{\sim}5k_{B}T_{s} above EFE_{F}) consistent with the DFT calculation. Such a single-particle “gapped” spectrum on top of a metallic dispersion, without a global symmetry breaking of the system, is analogous to the pseudogap state found in high-TcT_{c} cuprates [33, 34, 35], and is recently shown to also relate to electron-hole excitations [36].

Summing up the experimental results of Ta2NiSe5 upon thermal tuning, an electronic gapped state is found below TsT_{s} in line with the second-order structure phase transition and the DFT calculation. Meanwhile, we observe concomitant evolutions of strong lattice fluctuation and electronic pseudogap state above TsT_{s}, without any global symmetry breaking, suggesting potential involvement of strong EPC behind the MIT transition in Ta2NiSe5.

Refer to caption
Figure 4: Evidence of pseudogap and negative electronic compressibility (NEC) with potassium dosing. (a) Photoemission spectra along X-Γ\Gamma-X directions at discrete potassium dosing cycles (lower-left indices) at 80K. Distinct regions are shaded in different colors, marking the evolution stages of the band structure. (b) EDCs at Γ\Gamma for both Ta 4f4f corelevel and low energy bands. The peak shifts with potassium dosing are illustrated with dashed lines. CB: Conduction Band. VB: Valence Band. (c) Carrier density and chemical potential estimated from the peak area increment of the K 3p3p corelevel (Fig. S3) and the peak position shift of the Ta 4f4f core-level in (c), respectively. (d) Electronic compressibility calculated from the parameters deduced in (d). Red line is calculated from the polynomial fit of the scattered data points in (d).

IV Low-Temperature Pseudogap and Negative Electronic Compressibility

While the concomitant evolution of the high-temperature electronic pseudogap and strong lattice fluctuation suggest a connected microscopic mechanism, one cannot quantitatively delineate the contributions of EPC and direct electron-hole Coulomb interaction (excitonic instability) to the broken-symmetry ground state. Non-thermal tuning methods offer more possibilities to separate the above two channels. While pump-probe studies could access excited states beyond the temperature range in equilibrium ARPES measurements, the non-equilibrium dynamics involve non-equilibrated electronic and lattice temperatures, as well as other non-thermal effects [37]. Therefore, contentions remain as to which interaction term determines the eventual restoration of metallicity in Ta2NiSe5[21, 22, 23, 24].

Tuning the carrier density proves to be another efficient non-thermal tuning method in Ta2NiSe5, which can restore the gapless state at low temperatures [38, 39]. Here, through the combined use of a micron-spot synchrotron ARPES and in-situ potassium (K) dosing in the low-temperature broken-symmetry state, we achieve a quantitative measure of both the charge doping and chemical potential shift [see Fig. 4]. As will be discussed below, this measurement is necessary to examine the thermodynamic stability of the electronic subsystem. The relative potassium dosage is accessed through the core-level x-ray photoemission spectroscopy (XPS) of the K 3p3p orbital, and the electron doping per unit cell is determined by the Luttinger volume at different dosing levels (see Appendix C for details).

Enabled by these improvements, our experiments reveal a three-stage isothermal spectral evolution across the dosing-tuned MIT [Fig. 4(a) and Fig. S3]. This evolution takes a striking resemblance to the thermal phase transition (insulator-pseudogap-semimetal). Starting from a low-temperature gapped state, the valence band rapidly downshifts upon dosing (shaded in blue). A pseudogap concurrently appears at EFE_{F}, similar to the spectral features right above TsT_{s}. In the second stage (shaded in green), the energy position of the conduction band barely changes while the missing spectral weight in the pseudogap is gradually replenished, resembling the pseudogap filling at higher temperatures. Finally, a semi-metallic state is restored with the full conduction and valence bands intersecting each other (shaded in orange). This semi-metallic state achieved with heavy potassium dosing is qualitatively identical to the normal state of Ta2NiSe5 in the high-temperature orthorhombic phase, as suggested by the pump-probe experiments [23] and the DFT calculation [Fig. 3(a)]. Such resemblance indicates that upon electron doping, the system may have a similar tendency to naturally recover the orthorhombic lattice structure, which is indeed confirmed by our first-principles calculation as will be detailed in Sec. VI.

Meanwhile, thermodynamic inevitability of the lattice’s participation is evidenced by the anomalous chemical potential evolution during the electron injection process. Here, the chemical potential change is evaluated from the energy shift of the fully filled Ta 4f4f core-level and cross-referenced with the Se valence band shift [Fig. 4(b)], to rule out extrinsic contributions [40]. The dosage-independent linewidth indicates insignificant inelastic scattering from potassium disorder. Both the Ta 4f4f core-level and the Se valence band are found to evolve non-monotonically with charge doping, in sharp contrast to the non-interacting scenario where a rigid shift of bands towards higher binding energy is expected. Such a non-monotonic behavior of the chemical potential can be described in terms of electronic compressibility. NEC, where the energy of the electronic subsystem decreases despite the addition of electrons, appears when the pseudogap is being replenished in the second stage of charge doping [the green region in Fig. 4(c-d)]. Considering that purely electronic systems can only repel added electrons, such an NEC behavior requires the involvement of additional degrees of freedom, such as the lattice excitations, in order to maintain the thermodynamic stability of the full system.

V Determination of the electron-hole Coulomb attraction

Refer to caption
Figure 5: Comparison of conduction and valence band overlap in the semi-metallic normal state of Ta2NiSe5, achieved under different dielectric environments. (a) heavily potassium-dosed sample. (enhanced nen_{e}) (b) high-temperature spectrum of the intrinsic sample, normalized along the energy axis. (intrinsic ne+npn_{e}+n_{p}) (c) photon-excited sample (enhanced ne+npn_{e}+n_{p}, reproduced from previous work[23]).
Refer to caption
Figure 6: Evolution of band structure with carrier doping by DFT calculation. (a) Evolution of the calculated band structure (using the r2SCAN functional) with carrier doping starting from the low-temperature monoclinic fully gapped state. (b) Evolution of beta angle (marking the low-temperature monoclinic structure when deviated from 9090^{\circ}) evaluated with different DFT functionals.

An important consequence of accessing the semi-metallic normal state via different tuning methods is that it provides a reference to determine the Coulomb interaction strength, which are hard to estimate in the broken symmetry phase without bias. Here, we quantitatively compare the valence-conduction band overlaps EgE_{g} in the semi-metallic normal state reached via heavy potassium dosing with the other two normal states achieved under dramatically different dielectric environments - equilibrium heating and neutral photodoping [see Fig. 5]. To study the influence of direct inter-band Coulomb interaction on this EgE_{g}, we consider a three-orbital model when lattice effects are not included. The inter-band Coulomb interaction term reads as

int=Vi,σ,α,σ(niασc+ni+1ασc)niσf\mathcal{H}_{\mathrm{int}}=V\!\sum_{i,\sigma,\alpha,\sigma^{\prime}}\left(n^{c}_{i\alpha\sigma}+n^{c}_{i+1\alpha\sigma}\right)n^{f}_{i\sigma^{\prime}} (1)

where VV is the Coulomb interaction strength. nicn^{c}_{i}, ni+1cn^{c}_{i+1}, and nifn^{f}_{i} are the density operators of two degenerate conduction bands (Ta 5d5d) and the valence band (Ni 3d3d). If we compare normal states only, the system is resistant to exciton condensation (detailed in Sec. VI) and each band has conserved electron numbers (for specific global doping). We denote their total electron numbers (considering the conduction-band degeneracy) as NcN_{c} and NfN_{f}, respectively. Thus, the effective Hamiltonian can be mapped into

HintcVi,σ,α,σ(niασc+ni+1ασc)niσf=2VNfNi,α,σniασcH^{c}_{\mathrm{int}}\approx V\sum_{i,\sigma,\alpha,\sigma^{\prime}}\left(n^{c}_{i\alpha\sigma}+n^{c}_{i+1\alpha\sigma}\right)\left\langle n^{f}_{i\sigma^{\prime}}\right\rangle=\frac{2VN_{f}}{N}\sum_{i,\alpha,\sigma}n^{c}_{i\alpha\sigma} (2)

for the conduction band and

HintfVi,σ,α,σniασc+ni+1ασcniσf=2VNcNi,σniσfH^{f}_{\mathrm{int}}\approx\ V\sum_{i,\sigma,\alpha,\sigma^{\prime}}\langle n^{c}_{i\alpha\sigma}+n^{c}_{i+1\alpha\sigma}\rangle n^{f}_{i\sigma^{\prime}}=\frac{2VN_{c}}{N}\sum_{i,\sigma}n^{f}_{i\sigma} (3)

for the valence band, where NN is the number of unit cells. As a result, in addition to excitonic instability, which disappears in those normal states, the Hartree part of the Coulomb interaction can be directly read out from the relative shift of both conduction and valence band site energies. The doping-induced change of the band overlap EgE_{g} corresponds to Δμ=2V(ΔNcΔNf)/N\mathrm{\Delta}\mu={2V(\mathrm{\Delta}N_{c}-{\mathrm{\Delta}N}_{f})}/{N}. Therefore, the EgE_{g} is found to be linearly reduced with increasing VV [see Fig. S4]. Here (ΔNcΔNf)/N(\mathrm{\Delta}N_{c}-{\mathrm{\Delta}N}_{f})/N can be estimated from the potassium dosed ARPES spectrum. According to the Luttinger theorem [Appendix C], the doped electron density for the 15th-dosing [Fig. 4(a)] reaches \sim0.4 per unit cell, indicating (ΔNc+ΔNf)/N=0.46(\mathrm{\Delta}N_{c}+{\mathrm{\Delta}N}_{f})/N\ =0.46. As the Fermi velocity is comparable in the valence and two conduction bands, it leads to (ΔNcΔNf)/N0.152(\mathrm{\Delta}N_{c}-{\mathrm{\Delta}N}_{f})/N\ \approx 0.152. On the other side, Δμ\mathrm{\Delta}\mu can be estimated by comparing the change of band overlap EgE_{g} in the normal state [Fig. 5]. The heavily K-dosed spectra (Eg280E_{g}\cong 280 meV, heavily screened situation with enhanced nen_{e}), the high-temperature spectra (Eg270E_{g}\cong 270 meV, intrinsic nen_{e} + npn_{p}), and the pump-probe photodoped spectra [23] (Eg290E_{g}\cong 290 meV, photodoping enhanced nen_{e} + npn_{p}) give an upper bound to the change of EgE_{g} (Δμ\mathrm{\Delta}\mu) of 20\sim 20meV. Therefore, we can stringently place an upper bound of direct Coulomb interaction strength V70V\sim 70 meV in the system. This Coulomb interaction, in stark contrast to the \sim eV scale strength required in previous Coulomb-only mean field calculations [16, 20], is not nearly enough to account for the experimentally observed low-temperature gap by itself [further discussed in Fig. 9].

Summing up our results on low-temperature charge doping, a semi-metallic electronic structure is recovered with heavy electron injection through a pseudogap stage, resembling the normal state reached at sufficiently high temperatures. Moreover, the thermodynamic inevitability of the lattice’s participation is revealed through the appearance of the pseudogap state and the NEC, while the direct Coulomb interaction strength is estimated to be insufficient to account for the insulating state by itself.

VI Theoretical Simulations and Interpretations

The stringent upper bound for the allowed Coulomb interaction and the indispensable contribution from the lattice, revealed by the above experimental data, point to the need to reexamine the proposal of excitonic condensate formation in Ta2NiSe5. Exciton condensation was first proposed as a viable route toward MIT soon after the success of the BCS theory of superconductivity [41]. In Ta2NiSe5, one leading contention is whether it is the electron-phonon coupling or direct electron-hole Coulomb attraction that “dominates” the transition. Based on the new experimental evidence discussed in Sec. II-IV, we quantify the EPC and Coulomb terms, and systematically examine the roles of each term using first-principles and many-body methods in this section.

We first focus on the mechanism of the semi-metallic normal state restoration by charge doping in Fig. 4(a). To describe this process, we simulate the system at different doping levels (see Appendix B for details), using first-principles DFT method with the numerically stable r2SCAN functional [42]. This functional includes the kinetic energy density contribution, whose accuracy in correlated materials is beyond the widely used standard functionals within local density approximation (LDA) or generalized gradient approximation (GGA) [43]. To reveal the ground-state structures, we perform lattice relaxation at each doping level to achieve the most energetically favorable lattice structure. As illustrated in Fig. 6(a), starting from the low-temperature monoclinic gapped state, the system gradually restores semi-metallicity with a transition at the doping level of xdoping0.15x_{\text{doping}}\sim 0.15 (extra electrons per unit cell). Such a dramatic change in the electronic bands is mainly caused by the change of the lattice structure as shown in Fig. 6(b). We follow the definition in Sec. II and use the β\beta angle to depict the structural transition. Cross-checked with different DFT functionals, β\beta decreases as additional electrons dope into the system and eventually reaches 90 at xdoping0.20.3x_{\text{doping}}\sim 0.2-0.3. This simulated doping evolution reflects that the lattice structure of Ta2NiSe5 has a natural tendency to reshape into the high-symmetry orthorhombic phase (the high-temperature lattice structure in Sec. II) upon charge doping. It is worth noting that DFT is a ground-state theory and the functionals we adopted do not include excitonic effects. But the impact of phonons has been included on the level of Born-Oppenheimer approximation through the structural optimization at each doping level. Therefore, the experimentally observed restoration of the high-symmetry lattice structure upon charge doping (see Sec. IV), can be explained as a concomitant effect caused by changes in the lattice structure, without invoking the influence of additional excitonic instability.

To quantify the impact of phonons, we further evaluate the electron-phonon coupling strength near the Fermi level. Following Ref. 44, the coupling matrix element among Kohn-Sham orbitals is

gmn(ν)(𝒌,𝒒)=um𝒌+𝒒|Δ𝒒(ν)VKS|un𝒌,g_{mn}^{(\nu)}(\boldsymbol{k},\boldsymbol{q})=\langle u_{m\boldsymbol{k+q}}|\Delta^{(\nu)}_{\boldsymbol{q}}V^{\rm KS}|u_{n\boldsymbol{k}}\rangle\,, (4)

where |un𝒌|u_{n\boldsymbol{k}}\rangle is the Bloch periodic part of the Kohn-Sham wavefunction with momentum 𝒌\boldsymbol{k} of the nn-th band; the Δ𝒒(ν)VKS\Delta^{(\nu)}_{\boldsymbol{q}}V^{\rm KS} is the variation of Kohn-Sham potential induced by the ν\nu-th phonon mode at momentum 𝒒\boldsymbol{q}. The definition of the dimensionless operator Δ𝒒(ν)\Delta^{(\nu)}_{\boldsymbol{q}} and details of the simulation are described in Appendix B.

The orthorhombic phase of Ta2NiSe5 is unstable at low temperature [see Figs. 3(a-b) and Fig. 6(a)]. Forcing the lattice structure to stay in a high-symmetry orthorhombic phase, our first-principles simulation identifies two unstable phonon modes responsible for the orthorhombic-monoclinic structural transition. However, one cannot directly evaluate the electron-phonon coupling matrix elements within the DFT framework for these unstable phonons. Instead, we employ the reference system Ta2NiS5 to estimate these couplings, whose orthorhombic phase is stable at low temperature. In Ta2NiS5, we identify two phonon modes, labeled by their irreducible representation B2gB_{2g} and B1gB_{1g}, that have the maximum eigenvector overlap with the two unstable modes in Ta2NiSe5, respectively. In Ta2NiS5, these two phonon modes couple to three Kohn-Sham bands near Fermi surface: the top valence band (v1v_{1}), the lowest conduction band (c1c_{1}), and the second lowest conduction band (c2c_{2}). The coupling matrix elements (at the long-wavelength phonon limit q=0q=0) in this three-orbital subspace are

|gv1c1(B2g)(𝒌=Γ,𝒒=0)|\displaystyle\left|g_{v_{1}c_{1}}^{(B_{2g})}(\boldsymbol{k}=\Gamma,\boldsymbol{q}=0)\right| =\displaystyle= 52.4meV\displaystyle 52.4\,\text{meV}
|gv1c2(B1g)(𝒌=Γ,𝒒=0)|\displaystyle\left|g_{v_{1}c_{2}}^{(B_{1g})}(\boldsymbol{k}=\Gamma,\boldsymbol{q}=0)\right| =\displaystyle= 50.8meV.\displaystyle 50.8\,\text{meV}\,. (5)

The calculated phonon frequencies at 𝒒=0\boldsymbol{q}=0 are ωB2g=1.91\omega_{B_{2g}}=1.91 THz and ωB1g=1.54\omega_{B_{1g}}=1.54 THz, consistent with those identified in Ta2NiSe5  by Raman spectra [13, 14, 15]. Therefore, the phonon parameters extracted here will be used for further analysis in Ta2NiSe5.

Refer to caption
Figure 7: Lattice fluctuation induced pseudogap and the temperature-doping phase diagram of Ta2NiSe5. (a) Simulated band structure from the two-band model with only interband electron-phonon coupling (g=70g=70 meV): the broken-symmetry state caused by structural band hybridization (blue), the pseudogap state caused by lattice fluctuation (green), and the semi-metallic normal state (orange). (b) Temperature-doping phase diagram of Ta2NiSe5. The crystal structures within different regions are illustrated on the side.

As we will show later, simply considering electron-phonon coupling at the mean-field phonon level (Born-Oppenheimer approximation) cannot address the asynchronization between the structural phase transition and electronic spectral evolution along the temperature axis, discussed in Sec. III. Previous theoretical investigations have concluded that an eV-scale Coulomb interaction is required to break the symmetry [16, 20], also without considering the non-mean-field phonon effects. With the parameters estimated from the experimental data and the first-principles simulations, we consider a many-body model explicitly including both electronic interactions and electron-phonon coupling and all orders of fluctuations. Here, we employ a model Hamiltonian in a similar form to Ref. 45:

\displaystyle\mathcal{H}\!\! =\displaystyle= 0+k,q,σgqN[(aq+aq)ck+q,σfkσ+H.c.]\displaystyle\!\!\mathcal{H}_{0}+\sum_{k,q,\sigma}\frac{g_{q}}{\sqrt{N}}\left[\left(a_{q}+a^{\dagger}_{-q}\right)c^{\dagger}_{k+q,\sigma}f_{k\sigma}+H.c.\right] (6)
+qωqaqaq+Vi,σ,σ(niσc+ni+1σc)niσf,\displaystyle+\sum_{q}{{\omega}_{q}\ a^{\dagger}_{q}a_{q}}+V\sum_{i,\sigma,\sigma^{\prime}}\left(n^{c}_{i\sigma}+n^{c}_{i+1\sigma}\right)n^{f}_{i\sigma^{\prime}}\,,
0\displaystyle\mathcal{H}_{0}\!\! =\displaystyle= k,σεkcckσckσ+k,σεkvfkσfkσ,\displaystyle\!\!\sum_{k,\sigma}{\varepsilon^{c}_{k}c^{\dagger}_{k\sigma}c_{k\sigma}}+\sum_{k,\sigma}{\varepsilon^{v}_{k}f^{\dagger}_{k\sigma}f_{k\sigma}}\,, (7)

but the band and interaction parameters are chosen to match the experimental data in this paper and the phonon parameters are chosen according to our first-principles simulations. We write the Hamiltonian in momentum space here: ckσc^{\dagger}_{k\sigma} (ckσc_{k\sigma}) creates (annihilates) an electron at the conduction band (primarily Ta 5d5d) for momentum kk and spin σ\sigma, with dispersion given by εkc\varepsilon^{c}_{k}, while the fkσf^{\dagger}_{k\sigma} (fkσf_{k\sigma}) creates (annihilates) an electron at the valence band (primarily Ni 3d3d), with dispersion given by εkv\varepsilon^{v}_{k}. The niσcn^{c}_{i\sigma} and niσfn^{f}_{i\sigma} are the density operators for the conduction and valence band, respectively. The fitted band structures via experimental results [Fig. 3(e)] read as

εkc=3.11.8cos(k)0.9cos(2k)0.6cos(3k)\displaystyle\varepsilon^{c}_{k}=3.1-1.8\,{{\mathrm{cos}}}(k)-0.9\,{{\mathrm{cos}}}(2k)-0.6\,{{\mathrm{cos}}}(3k) (8)
εkv=1.8+1.5cos(k)+0.3cos(2k)+0.1cos(3k)\displaystyle\varepsilon^{v}_{k}=-1.8+1.5\,{{\mathrm{cos}}}(k)+0.3\,{{\mathrm{cos}}}(2k)+0.1\,{{\mathrm{cos}}}(3k)\, (9)

The direct hybridization between the conduction and valence bands around Γ\Gamma is forbidden in the orthorhombic phase by the inversion symmetry [16]. However, this hybridization is enabled by an B2gB_{2g} lattice distortion, parametrized as a momentum-dependent displacement xq(B2g)x_{q}^{(B2g)} and quantized as the phonon mode xq(B2g)=aq+aqx_{q}^{(B2g)}=a_{q}+a^{\dagger}_{-q}. As the Fermi momentum kFk_{F} is much smaller than 2π/a02\pi/a_{0}, we further restrict the e-ph coupling to the zone center gq=gδqg_{q}=g{\delta}_{q}. The inter-orbital Coulomb interaction term takes the same nearest-neighbor approximation as Eq. (1). Ideally, the conduction band should contain two sub-bands due to the two Ta chains per unit cell. However, we consider only the B2gB_{2g} coupling corresponding to the lower conduction band [45], where the phonon mode reflects the strongest Raman anomaly across the transition[13, 14, 15].

This model allows beyond-Born-Oppenheimer lattice fluctuations that couple to the electrons, and captures both the thermal and non-thermal evolution of the spectral function. We simulate single-particle spectral functions of the model in Eq. 6 using the exact diagonalization (ED), which accurately describes all correlation effects in a finite system (see Appendix B for details). Here, gg is chosen as 7070 meV, slightly larger than the number from the Ta2NiS5 simulation in Eq. (VI) to compensate the ignored electronic and lattice degrees of freedom, and we first exclude the Coulomb interaction by setting V=0V=0 (see Sec. VII for the discussion about the impact of Coulomb interaction). Enhanced by the low dimensionality, strong lattice fluctuation above TsT_{s} readily enables the conduction-valence band hybridization, and results in an electronic pseudogap without causing a global symmetry breaking [Fig. 7(a), green]. Further reducing temperature results in a divergence of the phonon number, which signifies the phonon condensation and drives the transition to the monoclinic phase. Subsequently, a hard hybridization gap forms below TsT_{s} [Fig. 7(a), blue]. On the other hand, increasing electron carrier density lifts the Fermi level from the charge-neutral point and reduces the interband phonon dressing. Thus, doping ultimately drives the monoclinic phase into the orthorhombic phase even at low temperatures, which in turn reflects the first-principles results in Fig. 6. The observed NEC also reflects strong electron-phonon coupling [46, 47] [see Appendix D], where additional electronic energy from the added carriers is absorbed into the lattice degree of freedom. Therefore, the phase diagram of Ta2NiSe5  along both the temperature and electron-doping axes can be depicted in Fig. 7(b), where the region shaded in green represents the lattice fluctuation induced pseudogap state above TsT_{s}. Clearly, both the thermal and quantum fluctuation effects are strong in this state.

When the phonon fluctuations are sufficient to drive the symmetry breaking through EPC, the many-body model in Eq. (6) can reproduce the sharp band gap of DFT simulations within the Born-Oppenheimer approximation. To show this, we simulate the spectral function by projecting the phonon fluctuations to a classical displacement (so-called the frozen phonon), which is equivalent to a coherent-state phonon wavefunction. Specifically, we assume that the ground state collapses to a broken-symmetry state characterized by a finite XX. Treating this XX as a classical variable, we obtain the frozen-phonon equation for electrons

FP(X)=0+g0Xk,σ[ckσfkσ+H.c.].\displaystyle\mathcal{H}_{\rm FP}\left(X\right)=\mathcal{H}_{0}+g_{0}X\sum_{k,\sigma}{\left[c^{\dagger}_{k\sigma}f_{k\sigma}+H.c.\right]}\ . (10)

Using the (xq=0(B2g))2\langle(x_{q=0}^{\rm(B2g)})^{2}\rangle obtained by ED simulations for various temperatures, we estimate the frozen-phonon displacement by XED=(xq=0(B2g))21/2X_{\rm ED}=\langle(x_{q=0}^{\rm(B2g)})^{2}\rangle^{1/2} and evaluate the spectral functions for the corresponding broken-symmetry states. As shown in the left panel of Fig. 7(a), this mean-field treatment for phonons reproduces the gapped ARPES spectra with well-defined single-particle band folding.

VII Discussion

We emphasize that our discussion of the EPC effects does not exclude the presence of electronic Coulomb interactions, which naturally exist in all materials. Instead, the quantification of a minimal lattice model provides a way to delineate the contributions from both interactions, avoiding the “chicken-or-egg” dilemma. We discuss the combined impact of EPC and electronic Coulomb interactions in this section. Figure 8 shows the single-particle spectra with the EPC gg and interband Coulomb interaction VV, whose upper bound is set by normal-state experimental results in Sec. IV. Comparing Fig. 7(a) with Fig. 8, we find that the Coulomb interaction within the allowed regime does not obviously influence the electronic structure, both in the broken-symmetry and pseudogap states. In contrast, when the EPC is ignored, the 100\leq 100 meV Coulomb interaction can only repel the conduction and valence bands in their entirety at the long-wavelength limit [Fig. 8 dashed red box]. Such a relative band shift is not helpful – and can even be destructive – to interband hybridizations.

Refer to caption
Figure 8: Impact of Coulomb interactions. Blue and green: Simulated band structure with both EPC (g=70g=70 meV) and direct interband Coulomb interaction (V=100V=100 meV). Red: Simulated band structure with only interband Coulomb interaction (V=100V=100 meV).
Refer to caption
Figure 9: Mean-field analysis of the Ta2NiSe5 model with the combined impact of EPC and interband Coulomb interaction. (a) The dependence of the excitonic order parameter Δext=iσfiσciσ/2N\Delta_{\rm ext}=\sum_{i\sigma}\langle f^{\dagger}_{i\sigma}c_{i\sigma}\rangle/2N on the EPC gg and Coulomb interaction VV, obtained by the mean-field solution of Eq. (6). (b) The same as (a) but for the lattice order parameter XMFTX_{\rm MFT}. (c, d) The dependence of the direct gap size on the two interactions for the (c) full mean-field solution and (d) solution without the Hartree term. The shaded regime indicates the opening of the band gap due to the relative shift of two bands, signaled by Δext=0\Delta_{\rm ext}=0. (e) The strength of interband hopping integral tTaNit_{\rm TaNi} as a function of momentum in the Ta2NiSe5. (f) the critical Coulomb interaction strength VcV_{c} needed to generate 300 and 200 meV hybridization gaps ΔEhyb\Delta E_{\rm hyb} at given tTaNit_{\rm TaNi}. Solid line: the result of mean-field simulation. Dashed line: the result excluding the contribution from the Hartree part of the mean-field decomposition. The region shaded in blue: experimentally determined Coulomb interaction strength.

More intuitions about the impact of EPC and Coulomb interactions can be obtained by analyzing the mean-field order parameters when both interactions are included. Thus, we further conduct a mean-field simulation for Hamiltonian in Eq. (6) with two order parameters: the electronic (excitonic) order parameter is defined as Δext=iσfiσciσ/2N\Delta_{\rm ext}=\sum_{i\sigma}\langle f^{\dagger}_{i\sigma}c_{i\sigma}\rangle/2N, reflecting the hybridization between Ta to Ni orbitals; the lattice order parameter is XMFT=xq=0(B2g)X_{\rm MFT}=\langle x_{q=0}^{\rm(B2g)}\rangle, reflecting the symmetry breaking of the crystal and demonstrating the same information as |β90||\beta-90^{\circ}| in Sec. II[48]. As shown in Figs. 9(a) and (b), these two order parameters follow the same dependence on interactions at the mean-field level. For weak interactions, the EPC (gg) and Coulomb interaction (VV) are cooperative in driving the hybridization fc\langle f^{\dagger}c\rangle and forming excitons. However, since the Hartree part of the Coulomb interaction [i.e. Eqs. (2) and (3)] separates the two bands, it competes with the hybridization or exciton formation. When VV is larger than some critical value (80\sim 80 meV), its suppression on the excitonic or lattice order overwhelms the latter.

To connect the order parameters with the single-particle spectra, we simulate the single-particle band gap within the mean-field framework. Note that these gaps are different from the “pseudogap” from Figs. 7 and 8, the latter of which contains many-body fluctuations beyond the mean field. Figure 9(c) shows the mean-field gap size for different combinations of both interactions. For weak Coulomb interaction, this gap is proportional to the excitonic order parameter Δext\Delta_{\rm ext} (and the lattice order parameter XMFTX_{\rm MFT}). They start to deviate from each other for a relatively strong VV, denoted by the dashed line. In this case, the Hartree part of the Coulomb interaction dominates and a band gap opens, suppressing the hybridization. Obviously, overlapping conduction and valence bands are always observed in ARPES experiments in the normal state of Ta2NiSe5, further supporting the conclusion about limited Coulomb interaction strength. To avoid the impact of this band gap, we further solve an artificial mean-field equation by eliminating the Hartree terms. This treatment leads to a pure hybridization gap [see Fig. 9(d)], depending quadratically on gg and slightly on VV. Detailed discussions about the hybridization gap and the impact of the Hartree terms are presented in Appendix E.

The inefficiency of the interband Coulomb interaction in Ta2NiSe5 originates from the material’s mirror symmetry with respect to the Ni-chain at small momenta, where the exact symmetry is present at k=0k=0[16]. To illustrate this relation, we take the multi-band hopping parameters from Ref. 16 and extract the component corresponding to the Ta 5dx2y25d_{x^{2}-y^{2}} - Ni 3dxy3d_{xy} hopping integral, denoted as tTaNit_{\rm TaNi}. Due to the different parities, tTaNit_{\rm TaNi} is an odd function of the momentum kk and its strength at the kFk_{F} (determined by experiments) is extremely small compared to the bandwidth [see Fig. 9(e)]. Adding this interband hopping tTaNit_{\rm TaNi} in our two-band model, we conduct the same mean-field simulation for the order parameters and single-particle gap. Figure. 9(f) shows the critical Coulomb interaction VcV_{c} necessary to generate a 200-300 meV hybridization gap (ΔEhyb\Delta E_{\rm hyb}) [49], without contributions from phonons. Due to the competition from the Hartree term, a hybridization gap comparable to experiments can not be reached given the experimentally bounded VV. Excluding the contribution from the Hartree part of the mean-field decomposition can relax the requirement on VcV_{c} to a certain extent (dashed line) [20]. However, compared to other excitonic insulator candidates, the required VcV_{c} in Ta2NiSe5 is still far beyond the experimentally identified upper bound (blue shade), due to its uniquely small Fermi momenta (black dashed line).

VIII Conclusion

Our result reveals the important role of EPC in both the normal and broken-symmetry states of Ta2NiSe5, based on the electrons’ coupling to a strongly fluctuating lattice. Combining advanced spectroscopy and computation methods, we determine and quantify a minimal lattice model that captures the pseudogap state, the broken symmetry state, and the negative electronic compressibility. The direct electron-hole Coulomb interaction may not be zero but is not required to capture the above experimental observations. In the pseudogap phase without any global structural symmetry lowering, phonons cannot be treated under the Born-Oppenheimer approximation, under which the lattice’s impact on the electronic structure is only reflected by a nonzero lattice distortion x\langle x\rangle. Instead, phonons act on the electronic structure through a fluctuating state where x=0\langle x\rangle=0 but x2 0{\langle x}^{2}\rangle\ \neq\ 0. In this regard, the pseudogap state may be conceptually likened to the ‘preformed excitons’ proposed in recent studies [26], except that the binding is facilitated by interband electron-phonon coupling. In comparison, fluctuating CDW states in cuprates, K0.3MoO3, ZrTe3, NbSe3, and NbSe2 are usually intertwined with electronic instabilities induced by strong Coulomb interaction [35, 50, 51, 52, 53]. In Ta2NiSe5, however, the major distinction lies in the small momentum low-energy band, which prohibits Coulomb interaction-induced charge transfer to the leading order. Such distinct lattice-symmetry-protected low-energy band crossings can be used to engineer accentuated lattice fluctuation effects in correlated materials.

The strong EPC in the Ta2NiSe5 system also leads to predictions of distinct material properties upon further thermal and nonthermal tuning. First, in the strong-coupling scenario, the phase transition should not be tied to the Fermi surface topology. Recent studies showed S-doping as an effective method to control the conduction and valence band overlap in Ta2NiSe5 [12]. The minimal many-body model shall produce an evolution of the phase transition temperature that is insensitive to the S-doping-tuned semimetal-to-semiconductor transition in Ta2Ni(Se,S)5, much different from the popular expectation of the dome-shaped excitonic insulator phase diagram upon band-gap tuning [41]. Second, given such strong EPC, the phonon self-energy will exhibit anomalies over a broad momentum range extending towards 𝐪0\mathbf{q}\rightarrow 0, rather than a narrow region sharply centered around 𝐪2𝐤F\mathbf{q}\sim 2\mathbf{k}_{F} [54]. This can be verified in the dynamic structural factor accessed with high-resolution inelastic x-ray or neutron scattering experiments.

While we qualitatively explain the observed spectral behaviors using a single phonon mode, further experiments have indicated that multiple phonons are involved in this transition: the diffuse scattering and inelastic x-ray scattering results point to extensive transverse acoustic phonon softening [29]; Raman scattering also suggests strong involvement of electronically coupled optical phonons [13, 14, 15]. Distinct from the acoustic-phonon-driven structural transition in 3D perovskite PrAlO3[55], the strong fluctuations in Ta2NiSe5 seem to suggest additional contributions from the lower system dimension and the optical phonons. The strong lattice fluctuations and electron-phonon coupling in this system offer a knob to realize optically, electrically, and mechanically controlled MIT [56], and a platform to investigate the role of electron-phonon coupling behind phase instabilities in low dimensional systems.

Last but not least, we demonstrated a general method to unambiguously resolve the typical chick-or-egg problems in correlated material systems. First-principles and many-body numerical simulations can be combined with spectroscopic experiments not only to determine the minimal microscopic lattice model, but also to quantify the microscopic interaction parameters – especially through numerous tuning methods. The minimal many-body model can be considered a faithful description of the material under investigation, and the roles of each interaction – if a reductionist view is desired – may be discussed by performing numerical experiments considering one term at a time. Most importantly, further material property predictions or engineering guidance can then be made by solving the perturbed minimal many-body model containing experimentally determined interaction parameters. This approach is ready to be generalized in low-dimensional correlated materials amid the rapid development of advanced spectroscopy, in situ material tuning, and modern computational methods.

ACKNOWLEDGEMENTS

We thank E. Baldini, S. D. Chen, E. W. Huang, A. Kemper, D. H. Lee, Y. B. Li, D. Y. Qiu, M. Trigo, Z. X. Shen, M. Yi, A. Zong, and W. T. Zhang for helpful discussions. Use of the Stanford Synchrotron Radiation Light Source, SLAC National Accelerator Laboratory, is supported by the US Department of Energy, Office of Science, Office of Basic Energy Sciences under Contract No. DE-AC02-76SF00515. This research used resources of the Advanced Light Source, a US DOE Office of Science User Facility under Contract No. DE-AC02-05CH11231. This work is based upon research conducted at the Center for High Energy X-ray Sciences (CHEXS) which is supported by the National Science Foundation under award DMR-1829070. Work at Lawrence Berkeley National Laboratory was funded by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences, Materials Sciences and Engineering Division under Contract No. DE-AC02-05-CH11231 within the Quantum Materials Program (KC2202) and within the Theory of Materials Program (KC2301) which provided the DFT calculations. S.D. and Y.W. acknowledge support from the National Science Foundation (NSF) award DMR-2038011. The electron-phonon model simulations were performed on the Frontera computing system at the Texas Advanced Computing Center. The DFT calculations were performed using computation resources at the National Energy Research Scientific Computing Center (NERSC). The work at Yale University is partially supported by the National Science Foundation under DMR-2132343.

Appendix A Experimental methods

Electric resistivity and heat capacity measurements were carried out by using a commercial PPMS (Quantum Design). The electric resistivity was measured by the four-probe method with the current applied in the ac-plane of a Ta2NiSe5 single crystal. The specific heat measurement was performed in the temperature range from 200K to 400K where the background signal was recorded in the same temperature range.

To characterize the structural dynamics of Ta2NiSe5 with a changing temperature, we carried out XRD measurement with the X-ray energy of 44 keV at the beamline QM2 of the Cornell High Energy Synchrotron Source (CHESS). Needle-like samples were chosen with a typical lateral dimension of \sim100 microns, which were then mounted with GE Varnish on a rotating pin. A Pilatus 6M 2D detector is used to collect the diffraction pattern with the sample rotated 360 around three different axes at 0.10.1^{\circ} step and 0.1s/frame data rate. The full 3D intensity cube is stacked and indexed with the beamline software package created by Jacob P. C. Ruff.

Synchrotron-based ARPES measurements were performed at beamline BL5-2 of Stanford Synchrotron Radiation Laboratory (SSRL), SLAC, USA, and BL 7.0.2 (MAESTRO) of Advanced Light Source (ALS), USA. The samples were cleaved in situ and measured under the ultra-high vacuum below 3×\mathrm{\times}10-11 Torr. Data was collected by R4000 and DA30L analyzers. The total energy and angle resolutions were 10 meV and 0.2, respectively.

Appendix B First-Principles DFT and Many-Body Simulations

The ab initio calculations presented in Sec. VI are performed using the Quantum ESPRESSO package [57]. The structural relaxation is calculated using the r2SCAN [42] functional with a semiempirical Grimme’s DFT-D2 Van-der-Waals correction [58]. A 30×\mathrm{\times}30×\mathrm{\times}15 k-mesh was used with a 100 Ry wavefunction energy cut-off. The phonon calculations are performed by frozen phonon method using phonopy code [59]. A 3×\mathrm{\times}2×\mathrm{\times}1 supercell was used. The dimensionless operator Δ𝒒(ν)\Delta_{\boldsymbol{q}}^{(\nu)} in Eq. (4) is directly related to the ν\nu-th phonon mode at 𝒒\boldsymbol{q} point

Δ𝒒(ν)=2ω𝒒νκ1Mκ𝐞κ(ν)(𝒒)𝝉κ\Delta_{\boldsymbol{q}}^{(\nu)}=\sqrt{\frac{\hbar}{2\omega_{\boldsymbol{q}\nu}}}\sum_{\kappa}\frac{1}{\sqrt{M_{\kappa}}}\mathbf{e}^{(\nu)}_{\kappa}(\boldsymbol{q})\cdot\frac{\partial}{\partial\boldsymbol{\tau}_{\kappa}} (11)

where κ\kappa labels atoms in the primitive unit cell, MκM_{\kappa} is the mass and 𝝉κ\boldsymbol{\tau}_{\kappa} is the coordinate of the κ\kappa-th atom, 𝐞κ(ν)(𝒒)\mathbf{e}^{(\nu)}_{\kappa}(\boldsymbol{q}) is ν\nu-th eigenvector of the dynamical matrix at 𝒒\boldsymbol{q} point on κ\kappa-th atom and ω\omega is the phonon frequency.

The exact diagonalization (ED) simulations presented in Sec. VI are conducted on an 8-site chain (effectively 16 sites due to the two bands). The diagonalization of the Hamiltonian leads to the ground-state wavefunction |G|G\rangle and all excited states (denoted as |m|m\rangle, with |m=0=|G|m=0\rangle=|G\rangle). The finite-temperature spectral function, for each band, is calculated through a canonical ensemble average

Ac(k,ω)=1πNm,σeEmkBTZImm|ckσ1ω+Em+iδckσ|m\displaystyle A^{c}\!\left(k,\omega\right)=-\frac{1}{\pi N}\!\sum_{m,\sigma}\frac{e^{-\frac{E_{m}}{k_{B}\!T}}}{Z}\mathrm{Im}\langle m|c^{\dagger}_{k\sigma}\frac{1}{\omega\!+\!\mathcal{H}\!-\!E_{m}\!+\!i\delta}c_{k\sigma}|m\rangle
Af(k,ω)=1πNm,σeEmkBTZImm|fkσ1ω+Em+iδfkσ|m\displaystyle A^{f}\!\left(k,\omega\right)=-\frac{1}{\pi N}\!\sum_{m,\sigma}\frac{e^{-\frac{E_{m}}{k_{B}\!T}}}{Z}\mathrm{Im}\langle m|f^{\dagger}_{k\sigma}\frac{1}{\omega\!+\!\mathcal{H}\!-\!E_{m}\!+\!i\delta}f_{k\sigma}|m\rangle

where ZZ is the partition function. The ensemble-averaged equal-time observables, e.g., the average phonon occupation, are defined in a similar manner

O=1ZmeEmkBTm|O|m\displaystyle\langle O\rangle=\ \frac{1}{Z}\sum_{m}{e^{-\frac{E_{m}}{k_{B}T}}\langle m|O|m\rangle\ \ \ } (12)

To capture the high momentum resolution comparable to experiments, we employ the twisted average boundary condition (TABC) in the simulation, with 50 equal-spacing phases for the spectral simulation and 30 phases for the equal-time observables. In addition to achieving the momentum resolution, the TABC is known to reduce the finite-size effects, particularly for the model with q=0q=0 interactions [60]. The ED simulation provides the solution of the full many-body state in 1D and preserves reflection symmetry, i.e. x0\langle x\rangle\equiv 0; however, the fluctuation of displacements x20\langle x^{2}\rangle\neq 0, reflecting the phonon squeezed state.

Appendix C Estimation of charge doping

Refer to caption
Figure 10: Estimation of electron doping and band shift. (a) Evolution of K 3p3p corelevel peaks with potassium dosing. (b) Constant energy contour at Fermi level after potassium dosing, illustrating the quasi-1D nature of the conduction band (CB) and the valence band (VB). (c) The change of the momentum width along XΓXX-\mathit{\Gamma}-X direction. (d) Evolution of bands at Γ\mathit{\Gamma} point with dosing cycles.

The charge doping per unit cell in Ta2NiSe5 is quantitatively estimated through the changes of both its band structure and the potassium corelevel peak. As shown in Fig. 10(a), the relative amount of potassium deposited on the sample surface is estimated through the increment of the peak area of potassium 3p3p corelevel (Gaussian peak fitting). We then estimate the charge per unit cell using Luttinger theorem, by comparing the area of Fermi surface volume to the size of the Brillouin Zone. Ta2NiSe5 has a quasi-1D shape of Fermi surface [Fig. 10(b)], therefore only the increment along XΓXX-\mathit{\Gamma}-X direction is needed. The change of the momentum width along XΓXX-\mathit{\Gamma}-X direction is estimated to be \mathrm{\sim}0.01 Å-1 for each dosing cycle [Fig. 10(c)], which contributes to a Brillouin Zone portion of roughly 0.01 Å-1/1.79 Å-1=0.56%. Given the existence of two conduction bands and one valence band, as well as the spin degeneracy, the total amount of dosage for each cycle is estimated to be roughly 0.56%*6=0.034 electrons per unit cell. We explicitly used the heavy dosing spectra (cycle 14-19), where the conduction and valence bands (CB and VB) move linearly toward higher binding energy. In this region, the influence of interaction or fluctuation is considered to be minimum.

Refer to caption
Figure 11: Estimation of electronic compressibility for 12.5% (blue) and 25% (red) hole doping, simulated using ED for various EPC strengths.

Appendix D Negative Electronic Compressibility

We estimate the electronic compressibility κdn/dμ\kappa\propto{dn}/{d\mu} through the second derivative of total energy with respect to carrier concentration:

κ1d2Edn2E(n+Δn)+E(nΔn)2E(n)Δn2\displaystyle\kappa^{-1}\propto\frac{d^{2}E}{dn^{2}}\approx\frac{E\left(n+\Delta n\right)+E\left(n-\Delta n\right)-2E\left(n\right)}{\Delta n^{2}} (13)

Due to the finite system size in our simulation, we employ a coarse-grained Δn=0.125\Delta n=0.125. With such a large spacing between neighboring doping levels, the estimated value can be regarded as an average of κ1\kappa^{-1} within a doping interval. The evaluated κ\kappa turns below zero for strong EPC for 12.5% doping, while this effect is screened by carriers at larger dopings. Although one cannot compare the quantitative values due to the coarse-grained derivatives, this doping dependence trend matches the low-temperature experiments in Sec. IV. Here, the simulation comes from the pp-type doping, since the effective mass of the electron pocket is smaller and, accordingly, the NEC doping window is narrower. Our discrete doping regime averages over a 12.5% doping regime, which smears out the NEC on the electron-doped side.

Appendix E Hybridization Gap

We analyze the hybridization gap of the model Hamiltonian in Eq. (6), to distinguish from the bare band gap EgE_{g} caused by the overall band separation. Specifically, we define the hybridization gap ΔEhyb\Delta E_{\rm hyb} as twice the off-diagonal element of the mean-field Hamiltonian at kFk_{\rm F}, while the Fermi momentum kFk_{\rm F} is determined by minimizing |Eg||E_{g}|. If the bare bands overlap, the mean-field Hamiltonian has identical diagonal terms at kFk_{F}, leading to an actual band gap identical to the hybridization gap ΔEhyb\Delta E_{\rm hyb}; otherwise, ΔEhyb\Delta E_{\rm hyb} reflects the contribution of hybridization and the single-particle gap becomes Eg2+ΔEhyb2\sqrt{E_{g}^{2}+\Delta E_{\rm hyb}^{2}}.

The left panel of Fig. 12 shows the hybridization gap ΔEhyb\Delta E_{\rm hyb} determined by the mean-field theory. A strong Coulomb interaction cannot stabilize a sizeable hybridization gap comparable to experimental results (0.3\sim 0.3 eV), indicating that the critical interaction VcV_{c} (defined in Fig. 9 of the main text) is infinite. In contrast, electron-phonon coupling can efficiently open a sizeable ΔEhyb\Delta E_{\rm hyb}. This distinction originates from the Hartree term of the Coulomb interaction [same as Eqs. (2) and (3)], which shifts the two bands away from each other and competes with the exciton formation. To reveal its impact, we further conduct an artificial mean-field simulation by removing the Hartree terms. This treatment removes the impact of band shifting and maximizes the capability of the Coulomb interaction VV to open the hybridization gap. Thus, with no or minimal contribution of the electron-phonon coupling gg, the Coulomb interaction can open a gap comparable to experiment, but it requires an extremely strong interaction V1V\sim 1 eV. We set tTaNi=0t_{\rm TaNi}=0 for all mean-field simulations except the results in Fig. 9(f) of the main text.

Refer to caption
Figure 12: Left panel: The hybridization gap ΔEhyb\Delta E_{\rm hyb} induced by the combined electron-electron interaction VV and electron-phonon coupling gg. Right panel: Same as the left panel but artificially removing the Hartree term in the mean-field solution.

References

  • Fradkin et al. [2015] E. Fradkin, S. A. Kivelson, and J. M. Tranquada, Colloquium: Theory of Intertwined Orders in High Temperature Superconductors, Rev. Mod. Phys 87, 457 (2015).
  • Fernandes et al. [2014] R. Fernandes, A. Chubukov, and J. Schmalian, What Drives Nematic Order in Iron-Based Superconductors?, Nat. Phys. 10, 97 (2014).
  • Schmalian [2013] J. Schmalian, Settling the Chicken and Egg Question in Iron Based Superconductors, J. Condens. Matter Phys.  (2013).
  • Mott [1961] N. F. Mott, The Transition to the Metallic State, Philos. Mag. 6, 287 (1961).
  • Knox [1983] R. Knox, Introduction to Exciton Physics, Collective Excitations in Solids , 183 (1983).
  • Kogar et al. [2017] A. Kogar, M. S. Rak, S. Vig, A. A. Husain, F. Flicker, Y. I. Joe, L. Venema, G. J. MacDougall, T. C. Chiang, E. Fradkin, et al.Signatures of Exciton Condensation in a Transition Metal Dichalcogenide, Science 358, 1314 (2017).
  • Halperin and Rice [1968] B. Halperin and T. Rice, Possible Anomalies at a Semimetal-semiconductor Transition, Rev. Mod. Phys. 40, 755 (1968).
  • Seitz et al. [1968] F. Seitz, D. Turnbull, and H. Ehrenreich, Solid State Physics, Vol. 21 (Academic Press, 1968) p. 115.
  • Alexandradinata et al. [2020] A. Alexandradinata, N. Armitage, A. Baydin, W. Bi, Y. Cao, H. J. Changlani, E. Chertkov, E. H. Neto, L. Delacretaz, I. E. Baggari, et al.The Future of the Correlated Electron Problem, arXiv:2010.00584  (2020).
  • Di Salvo et al. [1986] F. Di Salvo, C. Chen, R. Fleming, J. Waszczak, R. Dunn, S. Sunshine, and J. A. Ibers, Physical and Structural Properties of the New Layered Compounds Ta2NiS5 and Ta2NiSe5, J. Less-common Met. 116, 51 (1986).
  • Wakisaka et al. [2009] Y. Wakisaka, T. Sudayama, K. Takubo, T. Mizokawa, M. Arita, H. Namatame, M. Taniguchi, N. Katayama, M. Nohara, and H. Takagi, Excitonic Insulator State in Ta2NiSe5 Probed by Photoemission Spectroscopy, Phys. Rev. Lett. 103, 026402 (2009).
  • Lu et al. [2017] Y. Lu, H. Kono, T. Larkin, A. Rost, T. Takayama, A. Boris, B. Keimer, and H. Takagi, Zero-gap Semiconductor to Excitonic Insulator Transition in Ta2NiSe5, Nat. Commun. 8, 1 (2017).
  • Yan et al. [2019] J. Yan, R. Xiao, X. Luo, H. Lv, R. Zhang, Y. Sun, P. Tong, W. Lu, W. Song, X. Zhu, et al.Strong Electron-phonon Coupling in the Excitonic Insulator Ta2NiSe5, Inorg. Chem. 58, 9036 (2019).
  • Volkov et al. [2021] P. A. Volkov, M. Ye, H. Lohani, I. Feldman, A. Kanigel, and G. Blumberg, Critical Charge Fluctuations and Emergent Coherence in a Strongly Correlated Excitonic Insulator, npj Quantum Mater. 6, 1 (2021).
  • Kim et al. [2021] K. Kim, H. Kim, J. Kim, C. Kwon, J. S. Kim, and B. Kim, Direct Observation of Excitonic Instability in Ta2NiSe5, Nat. Commun. 12, 1 (2021).
  • Mazza et al. [2020] G. Mazza, M. Rösner, L. Windgätter, S. Latini, H. Hübener, A. J. Millis, A. Rubio, and A. Georges, Nature of Symmetry Breaking at the Excitonic Insulator Transition: Ta2NiSe5, Phys. Rev. Lett. 124, 197601 (2020).
  • Medarde [1997] M. L. Medarde, Structural, Magnetic and Electronic Properties of rnio3 Perovskites (R= rare earth), J. Condens. Matter Phys. 9, 1679 (1997).
  • Chu et al. [1973] C. Chu, J. Harper, T. Geballe, and R. Greene, Pressure Dependence of the Metal-insulator Transition in Tetrathiofulvalinium Tetracyanoquinodimethane (TTF-TCNQ), Phys. Rev. Lett. 31, 1491 (1973).
  • Wakisaka et al. [2012] Y. Wakisaka, T. Sudayama, K. Takubo, T. Mizokawa, N. Saini, M. Arita, H. Namatame, M. Taniguchi, N. Katayama, M. Nohara, et al.Photoemission Spectroscopy of Ta2NiSe5, J. Supercond. Nov. Magn. 25, 1231 (2012).
  • Seki et al. [2014] K. Seki, Y. Wakisaka, T. Kaneko, T. Toriyama, T. Konishi, T. Sudayama, N. Saini, M. Arita, H. Namatame, M. Taniguchi, et al.Excitonic Bose-Einstein Condensation in Ta2NiSe5 above Room Temperature, Phys. Rev. B 90, 155116 (2014).
  • Mor et al. [2017] S. Mor, M. Herzog, D. Golež, P. Werner, M. Eckstein, N. Katayama, M. Nohara, H. Takagi, T. Mizokawa, C. Monney, et al.Ultrafast Electronic Band Gap Control in an Excitonic Insulator, Phys. Rev. Lett. 119, 086401 (2017).
  • Okazaki et al. [2018] K. Okazaki, Y. Ogawa, T. Suzuki, T. Yamamoto, T. Someya, S. Michimae, M. Watanabe, Y. Lu, M. Nohara, H. Takagi, et al.Photo-induced Semimetallic States Realised in Electron-hole Coupled Insulators, Nat. Commun. 9, 4322 (2018).
  • Tang et al. [2020] T. Tang, H. Wang, S. Duan, Y. Yang, C. Huang, Y. Guo, D. Qian, and W. Zhang, Non-Coulomb Strong Electron-hole Binding in Ta2NiSe5 Revealed by Time-and Angle-resolved Photoemission Spectroscopy, Phys. Rev. B 101, 235148 (2020).
  • Baldini et al. [2020] E. Baldini, A. Zong, D. Choi, C. Lee, M. H. Michael, L. Windgaetter, I. I. Mazin, S. Latini, D. Azoury, B. Lv, et al.The Spontaneous Symmetry Breaking in Ta2NiSe5 is Structural in Nature, arXiv:2007.02909  (2020).
  • Watson et al. [2020] M. D. Watson, I. Marković, E. A. Morales, P. Le Fèvre, M. Merz, A. A. Haghighirad, and P. D. King, Band Hybridization at the Semimetal-semiconductor Transition of Ta2NiSe5 Enabled by Mirror-symmetry Breaking, Phys. Rev. Res. 2, 013236 (2020).
  • Fukutani et al. [2021] K. Fukutani, R. Stania, C. Il Kwon, J. S. Kim, K. J. Kong, J. Kim, and H. W. Yeom, Detecting Photoelectrons from Spontaneously Formed Excitons, Nat. Phys. 17, 1024 (2021).
  • Lee et al. [2019] J. Lee, C.-J. Kang, M. J. Eom, J. S. Kim, B. I. Min, and H. W. Yeom, Strong Interband Interaction in the Excitonic Insulator Phase of Ta2NiSe5, Phys. Rev. B 99, 075408 (2019).
  • Larkin et al. [2017] T. Larkin, A. Yaresko, D. Pröpper, K. Kikoin, Y. Lu, T. Takayama, Y.-L. Mathis, A. Rost, H. Takagi, B. Keimer, et al.Giant Exciton Fano Resonance in Quasi-one-dimensional Ta2NiSe5, Phys. Rev. B 95, 195144 (2017).
  • Nakano et al. [2018] A. Nakano, T. Hasegawa, S. Tamura, N. Katayama, S. Tsutsui, and H. Sawa, Antiferroelectric Distortion with Anomalous Phonon Softening in the Excitonic Insulator Ta2NiSe5, Phys. Rev. B 98, 045139 (2018).
  • Borisenko et al. [2008] S. Borisenko, A. Kordyuk, A. Yaresko, V. Zabolotnyy, D. Inosov, R. Schuster, B. Büchner, R. Weber, R. Follath, L. Patthey, et al.Pseudogap and Charge Density Waves in Two Dimensions, Phys. Rev. Lett. 100, 196402 (2008).
  • Chen et al. [2015] P. Chen, Y.-H. Chan, X.-Y. Fang, Y. Zhang, M.-Y. Chou, S.-K. Mo, Z. Hussain, A.-V. Fedorov, and T.-C. Chiang, Charge Density Wave Transition in Single-layer Titanium Diselenide, Nat. Commun. 6, 8943 (2015).
  • He et al. [2021] Y. He, S.-D. Chen, Z.-X. Li, D. Zhao, D. Song, Y. Yoshida, H. Eisaki, T. Wu, X.-H. Chen, D.-H. Lu, et al.Superconducting Fluctuations in Overdoped Bi2Sr2CaCu2O8+δ, Phys. Rev. X 11, 031068 (2021).
  • Alloul et al. [1989] H. Alloul, T. Ohno, and P. Mendels, NMR evidence for a fermi-liquid behavior in YBa2Cu3O6+x, Phys. Rev. Lett. 63, 1700 (1989).
  • Timusk and Statt [1999] T. Timusk and B. Statt, The pseudogap in high-temperature superconductors: an experimental survey, Rep. Prog. Phys. 62, 61 (1999).
  • Sobota et al. [2021] J. A. Sobota, Y. He, and Z.-X. Shen, Angle-resolved Photoemission Studies of Quantum Materials, Rev. Mod. Phys. 93, 025006 (2021).
  • Singh et al. [2022] A. Singh, H. Huang, J. Xie, J. Okamoto, C. Chen, T. Watanabe, A. Fujimori, M. Imada, and D. Huang, Unconventional exciton evolution from the pseudogap to superconducting phases in cuprates, Nat. Commun. 13, 7906 (2022).
  • de la Torre et al. [2021] A. de la Torre, D. M. Kennes, M. Claassen, S. Gerber, J. W. McIver, and M. A. Sentef, Colloquium: Nonthermal Pathways to Ultrafast Control in Quantum Materials, Rev. Mod. Phys. 93, 041002 (2021).
  • Fukutani et al. [2019] K. Fukutani, R. Stania, J. Jung, E. F. Schwier, K. Shimada, C. I. Kwon, J. S. Kim, and H. W. Yeom, Electrical Tuning of the Excitonic Insulator Ground State of Ta2NiSe5, Phys. Rev. Lett. 123, 206401 (2019).
  • Chen et al. [2020] L. Chen, T. Han, C. Cai, Z. Wang, Y. Wang, Z. Xin, and Y. Zhang, Doping-controlled Transition from Excitonic Insulator to Semimetal in Ta2NiSe5, Phys. Rev. B 102, 161116 (2020).
  • Wen et al. [2020] W. Wen, G. Zhao, C. Hong, Z. Song, and R.-H. He, 3D Negative Electronic Compressibility as a New Emergent Phenomenon, J. Supercond. Nov. Magn. 33, 229 (2020).
  • Jérome et al. [1967] D. Jérome, T. Rice, and W. Kohn, Excitonic Insulator, Phys. Rev. 158, 462 (1967).
  • Furness et al. [2020] J. W. Furness, A. D. Kaplan, J. Ning, J. P. Perdew, and J. Sun, Accurate and Numerically Efficient r2SCAN Meta-generalized Gradient Approximation, J. Phys. Chem. 11, 8208 (2020).
  • Pokharel et al. [2022] K. Pokharel, C. Lane, J. W. Furness, R. Zhang, J. Ning, B. Barbiellini, R. S. Markiewicz, Y. Zhang, A. Bansil, and J. Sun, Sensitivity of the Electronic and Magnetic Structures of Cuprate Superconductors to Density Functional Approximations, npj Comput. Mater. 8, 31 (2022).
  • Giustino [2017] F. Giustino, Electron-phonon Interactions from First Principles, Rev. Mod. Phys. 89, 015003 (2017).
  • Kaneko et al. [2013] T. Kaneko, T. Toriyama, T. Konishi, and Y. Ohta, Orthorhombic-to-monoclinic Phase Transition of Ta2NiSe5 Induced by the Bose-Einstein Condensation of Excitons, Phys. Rev. B 87, 035121 (2013).
  • Grilli and Castellani [1994] M. Grilli and C. Castellani, Electron-phonon Interactions in the Presence of Strong Correlations, Phys. Rev. B 50, 16880 (1994).
  • Castellani et al. [1995] C. Castellani, C. Di Castro, and M. Grilli, Singular Quasiparticle Scattering in the Proximity of Charge Instabilities, Phys. Rev. Lett. 75, 4650 (1995).
  • [48] Note that the lattice order parameter XMFTX_{\rm MFT} is evaluated through minimizing the mean-field energy rather than using the fluctuation-informed displacement XEDX_{\rm ED} in Eq. (10).
  • [49] The hybridization gap is not directly the order parameter. A more accurate definition is presented in Appendix E.
  • Pouget et al. [1985] J. Pouget, C. Noguera, A. Moudden, and R. Moret, Structural Study of the Charge-density-wave Phase Transition of the Blue Bronze: K0.3MoO3, J. Phys. 46, 1731 (1985).
  • Schäfer et al. [2001] J. Schäfer, E. Rotenberg, S. D. Kevan, P. Blaha, R. Claessen, and R. Thorne, High-Temperature Symmetry Breaking in the Electronic Band Structure of the Quasi-One-Dimensional Solid NbSe3, Phys. Rev. Lett. 87, 196403 (2001).
  • Chatterjee et al. [2015] U. Chatterjee, J. Zhao, M. Iavarone, R. Di Capua, J. Castellan, G. Karapetrov, C. Malliakas, M. G. Kanatzidis, H. Claus, J. Ruff, et al.Emergence of Coherence in the Charge-density-wave State of 2H-NbSe2, Nat. Commun. 6, 6313 (2015).
  • Yokoya et al. [2005] T. Yokoya, T. Kiss, A. Chainani, S. Shin, and K. Yamaya, Role of Charge-density-wave Fluctuations on the Spectral Function in a Metallic Charge-density-wave System, Phys. Rev. B 71, 140504 (2005).
  • Zhu et al. [2015] X. Zhu, Y. Cao, J. Zhang, E. Plummer, and J. Guo, Classification of charge density waves based on their nature, Proc. Natl. Acad. Sci. U.S.A. 112, 2367 (2015).
  • Birgeneau et al. [1974] R. Birgeneau, J. Kjems, G. Shirane, and L. Van Uitert, Cooperative Jahn-Teller Phase Transition in PrAlO3, Phys. Rev. B 10, 2512 (1974).
  • Liu et al. [2021] Q. Liu, D. Wu, Z. Li, L. Shi, Z. Wang, S. Zhang, T. Lin, T. Hu, H. Tian, J. Li, et al.Photoinduced Multistage Phase Transitions in Ta2NiSe5, Nat. Commun. 12, 2050 (2021).
  • Giannozzi et al. [2009] P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, G. L. Chiarotti, M. Cococcioni, I. Dabo, et al.QUANTUM ESPRESSO: A Modular and Open-source Software Project for Quantum Simulations of Materials, J. Condens. Matter Phys. 21, 395502 (2009).
  • Grimme [2006] S. Grimme, Semiempirical GGA-type Density Functional Constructed with a Long-range Dispersion Correction, J. Comput. Chem. 27, 1787 (2006).
  • Togo and Tanaka [2015] A. Togo and I. Tanaka, First Principles Phonon Calculations in Materials Science, Scr. Mater. 108, 1 (2015).
  • Poilblanc [1991] D. Poilblanc, Twisted Boundary Conditions in Cluster Calculations of the Optical Conductivity in Two-dimensional Lattice Models, Phys. Rev. B 44, 9562 (1991).
Refer to caption
Fig. S1: Photoemission data in 3D Brillouin Zone (BZ). (a) Illustration of BZ of Ta2NiSe5at high-temperature orthorhombic phase. (b) Out of plane photoemission map along the ky direction by changing the photon energy of the incident beam. (c) In-plane photoemission map at Γ\mathit{\Gamma} plane (49eV) with linear horizontal (LH) and linear vertical (LV) polarizations of the incident beam. The quasi-1D electronic structure of the system is clearly demonstrated. As the shearing of the Ta atoms in the low-temperature monoclinic phase is very small (less than 1%), the BZ can be treated as the same for the whole temperature range.
Refer to caption
Fig. S2: Determination of the pseudogap state at high-temperature phase of Ta2NiSe5. (a) High symmetry cut XΓXX-\mathit{\Gamma}-X at different Brillouin Zones (b) High symmetry cut XΓXX-\mathit{\Gamma}-X at different photon energies of the incident beam (different ky values). The spectrum weight depletion around EFE_{F} is constantly observed.
Refer to caption
Fig. S3: Complete potassium dosing data for both LH and LV polarizations. The dosing experiment was performed with a photon energy of the incident beam at 84eV (also at the Γ\mathit{\Gamma} plane in the 3D Brillouin Zone as illustrated in Fig. S1), due to the availability of both the polarizations of the light at the beamline. Besides the 3-stage evolution illustrated in LH channel (Sec. IV of main text), the LV channel illustrates the evolution of the valence band with the signal mainly from Ni 3d orbitals. The tip-like valence band top was gradually recovered, consistent with the DFT calculation for the semi-metallic normal state of the system (Sec. III of main text). A second set of bands can be seen in the background (most distinguishable in 16-19), suggesting the phase separation nature of the system during the transition from the monoclinic phase to the orthorhombic phase. Such a behaviour further indicates the structural instability of the system, as a natural result of the presence of fluctuating lattice.
Refer to caption
Fig. S4: Role of electron-phonon coupling and the direct Coulomb interaction in the system. (a) At the orthorhombic phase without any phonon fluctuations, the increasing Coulomb interaction (increasing V) only leads to a relative band shift without mediating exciton formation. (b) With finite e-ph coupling (g=70meVg=70meV) and phonon fluctuations, the Coulomb interaction enhances gap opening and spectral depletion.