Next Article in Journal
Crystal Chemistry and Genetic Implications of Pink Tourmalines from Distinct Pegmatite Provinces
Previous Article in Journal
Impact of Inter-Modular Connections on Progressive Compressive Behavior of Prefabricated Column-Supported Volumetric Modular Steel Frames
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Delicate Competition Between Different Excitonic Orderings in Ta2NiSe5

1
Fakultät für Physik, Universität Duisburg-Essen, Loatharstrasse 1, 47057 Duisburg, Germany
2
Jožef Stefan Institute, Jamova 39, 1000 Ljubljana, Slovenia
3
Faculty of Mathematics and Physics, University of Ljubljana, Jadranska 19, 1000 Ljubljana, Slovenia
*
Author to whom correspondence should be addressed.
Crystals 2025, 15(5), 414; https://doi.org/10.3390/cryst15050414
Submission received: 1 April 2025 / Revised: 23 April 2025 / Accepted: 24 April 2025 / Published: 28 April 2025
(This article belongs to the Special Issue State of Art of Excitonic Insulators and Topological Materials)

Abstract

:
We investigate the energetics of the quasi-one-dimensional layered compound Ta 2 NiSe 5 in the excitonic phase within the six-band model using Hartree–Fock calculations. We calculate energies of states with different kinds of excitonic order and show that they differ by less than meV and depend sensitively on precise values of interchain hopping matrix elements.

1. Introduction

Ta 2 NiSe 5 (TNS) is a layered transition metal dichalcogenide compound that undergoes a semi-metal to semi-conductor phase transition followed by a structural transition from a high-temperature orthorhombic phase to a low-temperature monoclinic phase at T c = 325 K [1,2]. A gap of 0.15 eV opens across the Fermi level between the valence and conduction band due to this transition [3]. In the monoclinic phase, the Ta and Se atoms undergo a shear-like displacement, and the reflection symmetries are broken [4], but the magnitude of the displacement is moderate with the monoclinic angle being small (0.7 degrees only) [1]. The size of the gap exceeds what one finds in standard LDA calculations in monoclinic phase [5] (see, however, [4,6]). The large gap and peculiar two-hump shape of the valence band inspired the interpretations [7,8,9] that the dominant instability is of electronic origin and the low-temperature state was proposed to be thought as an excitonic insulator [10,11,12,13,14]. Additionally, short time scales and peculiar fluence dependences observed in pump–probe time-resolved studies [15,16,17,18,19] as well as the vanishing of thermopower at low temperatures [20] have been interpreted in terms of electronic degrees of freedom and argued to point to the excitonic ground state.
In its purest form, as relevant in heterostructure/bilayer systems [21,22,23,24,25,26], the excitonic phase transition results in the spontaneous breaking of the U(1) symmetry associated with the separate conservation of charge in valence/conduction orbital (or layer) subspaces coupled by Coulomb repulsion. Recent theoretical work [27] clarified the nature of symmetry breaking in the putative excitonic phase and pointed out that in the case of TNS, excitonic condensation leads to spontaneous hybridization terms that break mirror symmetries only.
A different line of research has emphasized the existence of structural instabilities [4,6] without explicit reference to excitonic ordering. Notably, the monoclinic structure possesses lower energy, and its deformation aligns with the freezing of an unstable B 2 g mode. Theoretical work in Ref. [28] also concludes that Ta 2 NiS 5 and Ta 2 NiSe 5 are not fundamentally dissimilar, suggesting that many properties might be shared. Consistent with this line of thought, some experimental pump–probe ARPES (Angle-Resolved Photoemission Spectroscopy) studies [29] argue against the presence of excitonic effects. This is challenged by a recent nonequilibrium Raman study [30] that found a nonequilibrium state without a gap but with monoclinic distortion.
In our previous work [31], we studied the collective modes in TNS using a realistic model that treat the electronic and lattice instability on an equal footing within the Hartree–Fock treatment. In the excitonic phase, we saw the signature of a massive phase mode due to the breaking of discrete lattice symmetries, and we studied the interplay with the electron–phonon interaction that led to the increased mass of the phase mode, which also stabilized excitonic ordering.
In this paper, we inspect the excitonic ordering in the absence of electron–phonon interaction more closely. We find that within the Hartree–Fock approximation, one can stabilize excitonic ordering distinct from the one that is compatible with structural distortions in the monoclinic phase. We evaluate the energies of those ordered states and find that their energies are close (the energy differences are less than meV). Depending on precise values of the overall interchain hopping amplitude (that are all compatible with the orthorhombic symmetry), two different kinds of excitonic ordering may be found in the ground state. For the choice of hopping matrix elements exploited in Refs. [27,31], the ordering compatible with monoclinic distortion is found.

2. Model and Methods

TNS is a layered compound where the different layers stack up in the ‘b’ direction and a single layer lies in the ‘ac’ plane. In each ‘ac’ plane layer one finds Ta-Ni-Ta chains running along a direction and that are weakly hybridized in the c-direction and these chains within a 2D layer make TNS a quasi-one-dimensional-layered compound. The unit cell contains 4 Ta and 2 Ni atoms that constitute elements of two Ta-Ni-Ta chains. We consider a realistic minimal model of a 2D layer of TNS following Ref. [27]. We consider low-energy bands spanned by one Wannier d-orbital per each Ta and Ni atom using the 6-atom unit cell (see Figure 1). In the unit cell, the four Ta atoms are labeled 1 to 4, and the two Ni atoms are labeled 5 and 6. The 6 atoms are distributed in two chains as (1, 2, 5) and (3, 4, 6) as marked by green lines in Figure 1. In the high-temperature orthorhombic structure (Cmcm space group), the symmetry group contains inversion I and four reflection symmetries for planes parallel and perpendicular to the Ta-Ni-Ta chain ( σ , A / B ) as marked in Figure 1. The reflection symmetries are broken as we enter the monoclinic phase (C2/c space group) due to the shear displacement of the Ta atoms, but the inversion symmetry and the product of the reflection symmetries are preserved [6,27].
The six bands of TNS are given by Hamiltonian
H ^ kin = R δ Ψ ^ R + δ σ T ( δ ) Ψ ^ R σ ,
where the sum is taken over all unit cells R and neighboring cells δ as parametrized by the matrix of tight-binding hopping elements T. We have introduced the spinor Ψ ^ R σ = { c ^ 1 σ R , , c ^ 6 σ R } and c ^ i σ R is the annihilation operator in unit cell R , orbital i and spin σ . In Figure 1, the crystalline structure and unit cell are depicted.
The matrix elements of the kinetic energy T ( δ ) obtained using density-functional theory and Wannierization technique [27,32] are given in Table 1. We note that nearest neighbour hopping is considered in the vertical and horizontal directions. Hopping within one Ta-Ni-Ta chain is denoted as “intra-chain” while hopping between chains is denoted as ‘inter-chain’.
The electronic interaction energy is given by
H ^ int = U R i n ^ i R n ^ i R + V δ i { 1 , 2 } R σ σ n ^ i σ R + δ i n ^ 5 σ R + V δ i { 3 , 4 } R σ σ n ^ i σ R + δ i n ^ 6 σ R ,
where the first term represents the on-site Hubbard interaction for all atoms. The second and third terms describe the repulsion between nearest neighbours, where δ i sums over all nearest neighbours of both Ni atoms. In momentum space, the interaction reads
H ^ int = k , q V q i j n ^ i k n ^ j k + q ,
where we have introduced the full density operator at momentum k for orbital i as n ^ i k = q σ c ^ i σ k + q c ^ i σ q . The interaction vertex matrix is given by
V k = U / 2 0 0 0 V 1 ( k ) 0 0 U / 2 0 0 V 1 ( k ) 0 0 0 U / 2 0 0 V 2 ( k ) 0 0 0 U / 2 0 V 2 ( k ) V 1 * ( k ) V 1 * ( k ) 0 0 U / 2 0 0 0 V 2 * ( k ) V 2 * ( k ) 0 U / 2 ,
with V 1 = V ( 1 + exp i k x ) and V 2 = V ( 1 + exp i k x ) .
We solve the problem using the Hartree Fock method as used in Refs. [27,31]. The Hartree contribution to the self-energy is given by
Σ i i H ( t ) = V k = 0 i i n i ( t ) intraorbital 2 V k = 0 i j n j ( t ) interorbital ,
where we have assumed the translational invariance n i , R = n i and employed the Einstein notation. The Fock term is given as
Σ i j , k F ( t ) = q V q i j ϕ ^ i j , k q ( t ) .
Excitonic order is given by nonvanishing expectation value of ϕ ^ R = { ϕ ^ 15 R , ϕ ^ 25 R , ϕ ^ 36 R , ϕ ^ 46 R } . Here ϕ ^ i j R = σ c ^ i σ R c ^ j σ R + c ^ i σ R ± a c ^ j σ R , and a is a shift in the x direction. We choose + ( ) for { i j } = 15 , 25 ( 36 , 46 ) , respectively. In the ground state, the components of the order parameter ϕ R = ϕ ^ R are real. Their signs (total 2 4 = 16 choices of which subset (1111), (1-1-11), (1-11-1), and (1-111) corresponds to physically different cases) determine distinct excitonic orderings. Among these, the ordering that is described by ϕ R = ϕ 0 { 1 , 1 , 1 , 1 } breaks σ , A / B symmetries consistent with the monoclinic distortion [6,27]. We investigate if (and when) this ordering has the lowest energy.

3. Results

We set the intra-band interaction U = 2.5 eV and the inter-band interaction V = 0.785 eV, which leads to the excitonic phase with a gap that agrees with the experimental band gap. In practice, our calculations were performed at β = 1 / T = 100 , but we checked that to the reported number of digits, the energies do not change upon further cooling down, so the results are characteristic of the ground state. When we do not write out the unit of energy explicitly, we have eV in mind.
We calculate the energies of distinct symmetry broken states for parameters in Table 1, and we find three distinct energies corresponding to phases (1-1-11), (11-1-1), and (1-111) with non-degenerate energies as presented in Table 2. The symmetry broken phases of the different excitonic orderings in Table 2 are pictorially represented in Figure 2. Phases (1111) and (1-1-11) are degenerate. If one investigates a single Ta-Ni-Ta chain, orderings 1-1 and 11 are degenerate, and only the relative signs of interchain hybridization affect the energetics. While all components of the order parameter have the same magnitude in the case of orderings (1-1-11), (11-1-1) with value ϕ 0 = 0.1476 , two distinct magnitudes are found for the ordering (1-111) with components | ϕ 15 | = | ϕ 46 | = 0.1514 , and | ϕ 25 | = | ϕ 36 | = 0.1387 .
For our chosen model in Table 1, we indeed see that the monoclinic phase (1, −1, −1, 1) is the ground state. However, the energy of other symmetry-broken phases is larger by less than a meV. We further show in Figure 3a–c the band structure corresponding to the different phases. All symmetry-broken phases show the excitonic gap with nominal changes in the bandgap (around 0.6 per cent) at the ‘ Γ ’ point as presented in Table 2. The orbital weights of Ta (in blue) and Ni (in red) show that the maximum hybridization occurs along the ‘Z- Γ -X’ path and affects the higher conduction bands of Ta. Moreover, the energy gap between the maximally hybridized bands ( E ex ) is more than twice the bare band gap E gap as shown in Table 2. The behavior of excitonic hybridization is similar for the different phases.
We investigated why energy differs for various phases. If we decouple the two chains, i.e., switch off the interchain matrix elements, all energies would be degenerate, and therefore the inter-chain hoppings determine the ground state configuration. The degeneracy of energy for a decoupled chain can be explained analytically by a simple three-chain model as in Ref. [33].
The monoclinic phase (1-1-11) is the ground state for the choice of model in Table 1. We comment here that while the different hopping parameters in the orthorhombic phase should respect the crystal symmetry, which restrains the choice of their relative phases, it is still possible to have different choices of overall signs of the hopping amplitude (+/−) in Table 1 without disturbing the crystal symmetry. The precise magnitude of the hoppings (including the overall sign) relies on the details of the first-principle calculation and Wannier construction, e.g., choice and the orientation of the Wannier orbitals.
It turns out that the ground state ordering varies with the precise choice of the overall interchain hopping amplitudes. To demonstrate this, we first varied the signs of the interchain hoppings and presented the energies for the different orderings and hopping sign configurations in Table 3. To avoid confusion, we again stress the nomenclature: we refer to the different symmetry broken phases as different orderings (or different “kinds” of ordering), and a certain combination of the overall hopping sign as a configuration. Hence, for each configuration, we explore several different orderings. We find two different ground state orderings, namely (1-1-11) and (11-1-1), that vary as the signs of the hoppings are varied.
One may also present the results graphically. In Figure 4, we plot the energies for different orderings and indicate the configuration by different symbols. We see that configurations II and V behave oppositely to the others. In II and V, (11-1-1) has a lower energy than the ordering (1-1-11), whereas for the other configurations, the opposite statement holds.
Since Ta-Ta ( t 23 ) and Ni-Ni ( t 65 ) interchain hopping strength determines the ground state of the system, we tune these parameters around those corresponding to configuration I and plot the resulting dependence of δ E = E ( 1 1 11 ) E ( 11 1 1 ) in Figure 5. We vary t 23 by tuning a parameter ξ as t 23 = 0.02 eV ξ and show the data in red (keeping all the other parameters as in configuration I). The same exercise is repeated by varying instead t 56 = 0.03 ξ and the data is shown in black. We see a roughly linear dependence of δ E with the hopping strength and the crossing of the zero line clearly indicates that for a certain hopping the ground state energy flips from (1-1-11) to (11-1-1) in both the cases. Notice that the ground state flips at a nonvanishing value of the respective hopping, i.e., not only the signs of the hoppings but their precise magnitudes set the ground state.
In Figure 6, we show the dependence of the magnitude of the order parameter ϕ 0 with ξ for the ordering (1-1-11) (in solid line) and (11-1-1) (in dashed line) when t 23 is varied (left in red) and t 56 is varied (right in black). We see that for a constant Ni-Ni interchain hopping t 56 = 0.03 , the value of the order parameter increases for (1-1-11) (and decreases for (11-1-1)) with increasing Ta-Ta interchain hopping, while for a constant Ta-Ta interchain hopping t 23 = 0.02 , the behavior is rather opposite.
Our results point out that different orderings are energetically close. The smallness of energy differences stems from the smallness of interchain hybridization matrix elements, and this follows physically from the fact that associated Wannier orbitals are far in the real space. From that point of view, it is clear that our finding of small energy differences between distinct orderings may be robust to the changes in other parameters of calculations, provided one remains in the excitonic phase and the interchain matrix elements remain small. We explicitly checked that changing intra-chain hopping for 10meV leads to an overall energy change of 10 meV, but the energy balance between the phases that is determined by the interchain parameters remained unchanged.

4. Conclusions

We explored how energies of symmetry inequivalent excitonic insulator ordering vary with interchain hopping magnitude within Hartree–Fock calculations. We find that symmetry inequivalent orderings have energies that differ by only a small value of order less than meV and that the precise ordering pattern depends sensitively on the value of the interchain hopping amplitudes. The presence of such a sensitive balance in energy between different orders hints that other degrees of freedom should have an important role in stabilizing the ground state observed in the actual material. This agrees with previous work, suggesting that the intertwined orders with both electronic and lattice should be considered to understand the nature of the ground state [31]. Finally, the situation opens the possibility of using external stimuli, e.g., strain engineering or laser excitations, to manipulate the state of matter.
Our method could be further used to study the energy balance between different excitonic orderings in another promising excitonic insulator candidate Ta 2 Pd 3 Te 5 , which is also a quasi-one-dimensional layered compound with Ta-Pd-Ta chains in the ’bc’ plane [34,35].

Author Contributions

Conceptualization, J.M. and B.C.; methodology, D.G.; software, D.G.; validation, B.C.; formal analysis, B.C.; investigation, B.C.; resources, B.C. and J.M.; data curation, B.C.; writing—original draft preparation, B.C.; writing—review and editing, J.M. and D.G.; visualization, B.C.; supervision, J.M. and D.G.; project administration, J.M. and D.G.; funding acquisition, J.M. and D.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Slovenian Research Agency (ARIS) grant number P1-0044, J1-2455, J1-2458 and MN-0016-106.

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by authors on request.

Acknowledgments

We thank G. Mazza for helpful discussions.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

Abbreviation

The following abbreviation is used in this manuscript:
TNS Ta 2 NiSe 5

References

  1. Sunshine, S.A.; Ibers, J.A. Structure and physical properties of the new layered ternary chalcogenides tantalum nickel sulfide (Ta2NiS5) and tantalum nickel selenide (Ta2NiSe5). Inorg. Chem. 1985, 24, 3611–3614. [Google Scholar] [CrossRef]
  2. Di Salvo, F.; Chen, C.; Fleming, R.; Waszczak, J.; Dunn, R.; Sunshine, S.; Ibers, J.A. Physical and structural properties of the new layered compounds Ta2NiS5 and Ta2NiSe5. J. Less Common Met. 1986, 116, 51–61. [Google Scholar] [CrossRef]
  3. Lu, Y.; Kono, H.; Larkin, T.; Rost, A.; Takayama, T.; Boris, A.; Keimer, B.; Takagi, H. Zero-gap semiconductor to excitonic insulator transition in Ta2NiSe5. Nat. Commun. 2017, 8, 14408. [Google Scholar] [CrossRef] [PubMed]
  4. Subedi, A. Orthorhombic-to-monoclinic transition in Ta2NiSe5 due to a zone-center optical phonon instability. Phys. Rev. Mater. 2020, 4, 083601. [Google Scholar] [CrossRef]
  5. Lee, J.; Kang, C.J.; Eom, M.J.; Kim, J.S.; Min, B.I.; Yeom, H.W. Strong interband interaction in the excitonic insulator phase of Ta2NiSe5. Phys. Rev. B 2019, 99, 075408. [Google Scholar] [CrossRef]
  6. Watson, M.D.; Marković, I.; Morales, E.A.; Le Fèvre, P.; Merz, M.; Haghighirad, A.A.; King, P.D. Band hybridization at the semimetal-semiconductor transition of Ta2NiSe5 enabled by mirror-symmetry breaking. Phys. Rev. Res. 2020, 2, 013236. [Google Scholar] [CrossRef]
  7. Wakisaka, Y.; Sudayama, T.; Takubo, K.; Mizokawa, T.; Arita, M.; Namatame, H.; Taniguchi, M.; Katayama, N.; Nohara, M.; Takagi, H. Excitonic insulator state in Ta2NiSe5 probed by photoemission spectroscopy. Phys. Rev. Lett. 2009, 103, 026402. [Google Scholar] [CrossRef]
  8. Wakisaka, Y.; Sudayama, T.; Takubo, K.; Mizokawa, T.; Saini, N.; Arita, M.; Namatame, H.; Taniguchi, M.; Katayama, N.; Nohara, M.; et al. Photoemission spectroscopy of Ta2NiSe5. J. Supercond. Nov. Magn. 2012, 25, 1231–1234. [Google Scholar] [CrossRef]
  9. Seki, K.; Wakisaka, Y.; Kaneko, T.; Toriyama, T.; Konishi, T.; Sudayama, T.; Saini, N.; Arita, M.; Namatame, H.; Taniguchi, M.; et al. Excitonic Bose–Einstein condensation in Ta2NiSe5 above room temperature. Phys. Rev. B 2014, 90, 155116. [Google Scholar] [CrossRef]
  10. Des Cloizeaux, J. Exciton instability and crystallographic anomalies in semiconductors. J. Phys. Chem. Solids 1965, 26, 259–266. [Google Scholar] [CrossRef]
  11. Jérome, D.; Rice, T.; Kohn, W. Excitonic insulator. Phys. Rev. 1967, 158, 462–475. [Google Scholar] [CrossRef]
  12. Kuneš, J. Excitonic condensation in systems of strongly correlated electrons. J. Phys. Condens. Matter 2015, 27, 333201. [Google Scholar] [CrossRef]
  13. Kaneko, T.; Toriyama, T.; Konishi, T.; Ohta, Y. Electronic structure of Ta2NiSe5 as a candidate for excitonic insulators. J. Phys. Conf. Ser. 2012, 400, 032035. [Google Scholar] [CrossRef]
  14. Kaneko, T.; Toriyama, T.; Konishi, T.; Ohta, Y. Orthorhombic-to-monoclinic phase transition of Ta2NiSe5 induced by the Bose–Einstein condensation of excitons. Phys. Rev. B 2013, 87, 035121. [Google Scholar] [CrossRef]
  15. Mor, S.; Herzog, M.; Golež, D.; Werner, P.; Eckstein, M.; Katayama, N.; Nohara, M.; Takagi, H.; Mizokawa, T.; Monney, C.; et al. Ultrafast Electronic Band Gap Control in an Excitonic Insulator. Phys. Rev. Lett. 2017, 119, 086401. [Google Scholar] [CrossRef]
  16. Mor, S.; Herzog, M.; Noack, J.; Katayama, N.; Nohara, M.; Takagi, H.; Trunschke, A.; Mizokawa, T.; Monney, C.; Stähler, J. Inhibition of the photoinduced structural phase transition in the excitonic insulator Ta2NiSe5. Phys. Rev. B 2018, 97, 115154. [Google Scholar] [CrossRef]
  17. Okazaki, K.; Ogawa, Y.; Suzuki, T.; Yamamoto, T.; Someya, T.; Michimae, S.; Watanabe, M.; Lu, Y.; Nohara, M.; Takagi, H.; et al. Photo-induced semimetallic states realised in electron–hole coupled insulators. Nat. Commun. 2018, 9, 4322. [Google Scholar] [CrossRef]
  18. Bretscher, H.M.; Andrich, P.; Telang, P.; Singh, A.; Harnagea, L.; Sood, A.K.; Rao, A. Ultrafast melting and recovery of collective order in the excitonic insulator Ta2NiSe5. Nat. Commun. 2021, 12, 1699. [Google Scholar] [CrossRef]
  19. Saha, T.; Golež, D.; De Ninno, G.; Mravlje, J.; Murakami, Y.; Ressel, B.; Stupar, M.; Ribič, P.R. Photoinduced phase transition and associated timescales in the excitonic insulator Ta2NiSe5. Phys. Rev. B 2021, 103, 144304. [Google Scholar] [CrossRef]
  20. Nakano, A.; Nagai, T.; Katayama, N.; Sawa, H.; Taniguchi, H.; Terasaki, I. Exciton Transport in the Electron–Hole System Ta2NiSe5. J. Phys. Soc. Jpn. 2019, 88, 113706. [Google Scholar] [CrossRef]
  21. Eisenstein, J. Exciton Condensation in Bilayer Quantum Hall Systems. Annu. Rev. Condens. Matter Phys. 2014, 5, 159–181. [Google Scholar] [CrossRef]
  22. Li, J.I.A.; Taniguchi, T.; Watanabe, K.; Hone, J.; Dean, C.R. Excitonic superfluid phase in double bilayer graphene. Nat. Phys. 2017, 13, 751–755. [Google Scholar] [CrossRef]
  23. Liu, X.; Watanabe, K.; Taniguchi, T.; Halperin, B.I.; Kim, P. Quantum Hall drag of exciton condensate in graphene. Nat. Phys. 2017, 13, 746–750. [Google Scholar] [CrossRef]
  24. Wang, Z.; Rhodes, D.A.; Watanabe, K.; Taniguchi, T.; Hone, J.C.; Shan, J.; Mak, K.F. Evidence of high-temperature exciton condensation in two-dimensional atomic double layers. Nature 2019, 574, 76–80. [Google Scholar] [CrossRef]
  25. Gupta, S.; Kutana, A.; Yakobson, B.I. Heterobilayers of 2D materials as a platform for excitonic superfluidity. Nat. Commun. 2020, 11, 2989. [Google Scholar] [CrossRef] [PubMed]
  26. Nilsson, F.; Kuisma, M.; Pakdel, S.; Thygesen, K.S. Excitonic Insulators and Superfluidity in Two-Dimensional Bilayers without External Fields. J. Phys. Chem. Lett. 2023, 14, 2277–2283. [Google Scholar] [CrossRef]
  27. Mazza, G.; Rösner, M.; Windgätter, L.; Latini, S.; Hübener, H.; Millis, A.J.; Rubio, A.; Georges, A. Nature of symmetry breaking at the excitonic insulator transition: Ta 2 NiSe 5. Phys. Rev. Lett. 2020, 124, 197601. [Google Scholar] [CrossRef] [PubMed]
  28. Windgätter, L.; Rösner, M.; Mazza, G.; Hübener, H.; Georges, A.; Millis, A.J.; Latini, S.; Rubio, A. Common microscopic origin of the phase transitions in Ta2NiS5 and the excitonic insulator candidate Ta2NiSe5. npj Comput. Mater. 2021, 7, 210. [Google Scholar] [CrossRef]
  29. Baldini, E.; Zong, A.; Choi, D.; Lee, C.; Michael, M.H.; Windgaetter, L.; Mazin, I.I.; Latini, S.; Azoury, D.; Lv, B.; et al. The spontaneous symmetry breaking in Ta2NiSe5 is structural in nature. Proc. Natl. Acad. Sci. 2023, 120, e2221688120. [Google Scholar] [CrossRef]
  30. Katsumi, K.; Alekhin, A.; Souliou, S.M.; Merz, M.; Haghighirad, A.A.; Le Tacon, M.; Houver, S.; Cazayous, M.; Sacuto, A.; Gallais, Y. Disentangling Lattice and Electronic Instabilities in the Excitonic Insulator Candidate Ta2NiSe5 by Nonequilibrium Spectroscopy. Phys. Rev. Lett. 2023, 130, 106904. [Google Scholar] [CrossRef]
  31. Chatterjee, B.; Mravlje, J.; Golež, D. Collective modes and Raman response in Ta2NiSe5. Phys. Rev. B 2025, 111, L121106. [Google Scholar] [CrossRef]
  32. Mostofi, A.A.; Yates, J.R.; Lee, Y.S.; Souza, I.; Vanderbilt, D.; Marzari, N. wannier90: A tool for obtaining maximally localised Wannier functions. Comput. Phys. Commun. 2008, 178, 685–699. [Google Scholar] [CrossRef]
  33. Sun, Z.; Kaneko, T.; Golež, D.; Millis, A.J. Second-order Josephson effect in excitonic insulators. Phys. Rev. Lett. 2021, 127, 127702. [Google Scholar] [CrossRef]
  34. Huang, J.; Jiang, B.; Yao, J.; Yan, D.; Lei, X.; Gao, J.; Guo, Z.; Jin, F.; Li, Y.; Yuan, Z.; et al. Evidence for an excitonic insulator state in Ta2Pd3Te5. Phys. Rev. X 2024, 14, 011046. [Google Scholar] [CrossRef]
  35. Zhang, P.; Dong, Y.; Yan, D.; Jiang, B.; Yang, T.; Li, J.; Guo, Z.; Huang, Y.; Haobo; Li, Q.; et al. Spontaneous gap opening and potential excitonic states in an ideal Dirac semimetal Ta2Pd3Te5. Phys. Rev. X 2024, 14, 011047. [Google Scholar] [CrossRef]
Figure 1. Monolayer (in the ac plane) of Ta 2 NiSe 5 (orthorhombic phase) with Ta (blue), Ni (red) and Se (yellow) atoms (left) and the 6-band minimal model (right) with the unit cell indicated with a box. The two Ta-Ni-Ta chains (1, 2, 5), and (3, 4, 6) in the unit cell are marked with green lines. The two reflection lines and the inversion center (I) are marked.
Figure 1. Monolayer (in the ac plane) of Ta 2 NiSe 5 (orthorhombic phase) with Ta (blue), Ni (red) and Se (yellow) atoms (left) and the 6-band minimal model (right) with the unit cell indicated with a box. The two Ta-Ni-Ta chains (1, 2, 5), and (3, 4, 6) in the unit cell are marked with green lines. The two reflection lines and the inversion center (I) are marked.
Crystals 15 00414 g001
Figure 2. Schematic representation of the different excitonic ordering for TNS monolayer (in the ac plane), (a) (1-1-11), (b) (1-111), and (c) (11-1-1), with non-degenerate energies as shown in Table 2. The thickness of the lines connecting the atoms indicates the absolute value of the hybridization. The two reflection lines and the inversion center (I) are marked.
Figure 2. Schematic representation of the different excitonic ordering for TNS monolayer (in the ac plane), (a) (1-1-11), (b) (1-111), and (c) (11-1-1), with non-degenerate energies as shown in Table 2. The thickness of the lines connecting the atoms indicates the absolute value of the hybridization. The two reflection lines and the inversion center (I) are marked.
Crystals 15 00414 g002
Figure 3. Band structure for the different symmetry broken phases of the order parameter ϕ R (a) (1-1-11), (b) (11-1-1), and (c) (1-111), with non-degenerate energies in the hoppings consistent with Table 1. The orbital weights are indicated for Ni (in red) and Ta (in blue). (d) The 3-chain model with all the interchain hoppings switched off. The gap size E gap and the gap between maximally hybridized bands E ex are marked by (a). (e) The k-path is illustrated in the 3-dimensional Brillouin zone.
Figure 3. Band structure for the different symmetry broken phases of the order parameter ϕ R (a) (1-1-11), (b) (11-1-1), and (c) (1-111), with non-degenerate energies in the hoppings consistent with Table 1. The orbital weights are indicated for Ni (in red) and Ta (in blue). (d) The 3-chain model with all the interchain hoppings switched off. The gap size E gap and the gap between maximally hybridized bands E ex are marked by (a). (e) The k-path is illustrated in the 3-dimensional Brillouin zone.
Crystals 15 00414 g003
Figure 4. Non-degenerate energies for the different orderings and configurations presented in Table 3.
Figure 4. Non-degenerate energies for the different orderings and configurations presented in Table 3.
Crystals 15 00414 g004
Figure 5. Change in energy δ E = E ( 1 - 1 - 11 ) E ( 11 - 1 - 1 ) varying the hopping strength of Ta-Ta interchain hopping only (red) and Ni-Ni interchain hopping only (black) by a dimensionless parameter ξ . Configuration I corresponds to the left-most black and right-most red symbol, where t 23 = 0.02 eV and t 56 = 0.03 eV.
Figure 5. Change in energy δ E = E ( 1 - 1 - 11 ) E ( 11 - 1 - 1 ) varying the hopping strength of Ta-Ta interchain hopping only (red) and Ni-Ni interchain hopping only (black) by a dimensionless parameter ξ . Configuration I corresponds to the left-most black and right-most red symbol, where t 23 = 0.02 eV and t 56 = 0.03 eV.
Crystals 15 00414 g005
Figure 6. Change in magnitude of order parameter ϕ 0 for the phases (1-1-11) (in solid line), (11-1-1) (in dashed line) varying the hopping strength of Ta-Ta interchain hopping only (left in red), and Ni-Ni interchain hopping only (right in black). On the x-axis, we show the renormalized hopping ξ = t i j / t i j I , where t i j is either t 56 or t 23 , and t i j I denotes the respective hopping values in configuration I.
Figure 6. Change in magnitude of order parameter ϕ 0 for the phases (1-1-11) (in solid line), (11-1-1) (in dashed line) varying the hopping strength of Ta-Ta interchain hopping only (left in red), and Ni-Ni interchain hopping only (right in black). On the x-axis, we show the renormalized hopping ξ = t i j / t i j I , where t i j is either t 56 or t 23 , and t i j I denotes the respective hopping values in configuration I.
Crystals 15 00414 g006
Table 1. Elements of the hopping matrix T ( δ ) taken from Ref. [27].
Table 1. Elements of the hopping matrix T ( δ ) taken from Ref. [27].
CasesHopping Matrix Elements t( δ )
Intra-chain Ta-Ta T i i ( a x , 0) = T i i ( a x , 0) = −0.72 eV
T i i (0, 0) = 1.35 eV, i = 1,...,4
Intra-chain Ni-Ni T i i ( a x , 0) = T i i ( a x , 0) = 0.30 eV,
T i i (0, 0) = −0.36 eV, i = 5, 6
Intra-chain Ta-Ni T 15 ( a x , 0) = T 25 ( a x , 0) = T 25 (0) = T 15 (0)
= T 36 (0) = T 46 (0) = T 36 ( a x , 0), T 46 ( a x , 0) = −0.035 eV
Inter-chain Ta-Ni T 35 ( a x , 0) = T 35 ( a x , 0), T 45 ( a x , a y ) =
T 45 ( a x , a y ) = −0.04 eV,
T 26 ( a x , 0) = T 26 ( a x , 0) =
T 16 ( a x , a y ) = T 16 ( a x , a y ) = T 45 ( a x , a y )
Inter-chain Ta-Ta T 23 (0) = T 23 ( a x , 0) = T 41 ( a x , a y ) =
T 41 (0, a y ) = 0.02 eV
Inter-chain Ni-Ni T 65 (0) = T 65 ( a x , 0) = T 65 ( a x , a y ) =
T 65 (0, a y ) = −0.03 eV
We note that in the above notation T i j (0, 0) indicates hopping within the unit cell, T i j ( a x , 0) or T i j ( a x , 0) indicates horizontal hopping (along ‘a’ direction) to the right or left in the nearest neighbour unit cell. T i j ( a y , 0) or T i j ( a y , 0) indicates vertical hopping (along ‘c’ direction) to up or down in the nearest neighbour unit cell. A detailed pictorial representation is available in Ref. [27].
Table 2. The non-degenerate total energies ( E tot ), bare band gap E gap , energy difference between the maximally hybridized band E ex for different orderings and hoppings given in Table 1 E G denote the ground state energy. All energies are in eV.
Table 2. The non-degenerate total energies ( E tot ), bare band gap E gap , energy difference between the maximally hybridized band E ex for different orderings and hoppings given in Table 1 E G denote the ground state energy. All energies are in eV.
Ordering E tot E gap E ex Comment
(1-1-11)1.7652890.1380.308 E G
(11-1-1)1.7661090.1330.328 E G + 8 × 10 4
(1-111)1.7656790.1390.340 E G + 3.9 × 10 4
Table 3. Dependence of ground state ordering on interchain hopping signs. The hoppings are expressed in eV. Configuration I corresponds to the case presented in Table 1 and is taken as the reference configuration. In the last column, we mention which of the hopping signs differ from those of configuration I. Column GS indicates the ordering in the ground state. Δ E shows the energy difference (in 10 4 eV) between the ground state ordering and the lowest excited ordered state of the same configuration.
Table 3. Dependence of ground state ordering on interchain hopping signs. The hoppings are expressed in eV. Configuration I corresponds to the case presented in Table 1 and is taken as the reference configuration. In the last column, we mention which of the hopping signs differ from those of configuration I. Column GS indicates the ordering in the ground state. Δ E shows the energy difference (in 10 4 eV) between the ground state ordering and the lowest excited ordered state of the same configuration.
Config * T 15 ( 0 ) T 35 ( a x , 0 ) T 23 ( 0 ) T 65 ( 0 ) GS Δ EComment
I 0.035 0.04 0.02 0.03 (1-1-11)3.9-
II 0.035 0.04 0.02 0.03 (11-1-1)7.2Ta-Ta interchain
III 0.035 0.04 0.02 0.03 (1-1-11)3.9Ta-Ni intrachain
IV 0.035 0.04 0.02 0.03 (1-1-11)3.9Ta-Ni interchain
V 0.035 0.04 0.02 0.03 (11-1-1)7.4Ni-Ni interchain
VI 0.035 0.04 0.02 0.03 (1-1-11)4.3Ta-Ta, Ni-Ni interchain
* Config refers to configuration.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chatterjee, B.; Golež, D.; Mravlje, J. Delicate Competition Between Different Excitonic Orderings in Ta2NiSe5. Crystals 2025, 15, 414. https://doi.org/10.3390/cryst15050414

AMA Style

Chatterjee B, Golež D, Mravlje J. Delicate Competition Between Different Excitonic Orderings in Ta2NiSe5. Crystals. 2025; 15(5):414. https://doi.org/10.3390/cryst15050414

Chicago/Turabian Style

Chatterjee, Banhi, Denis Golež, and Jernej Mravlje. 2025. "Delicate Competition Between Different Excitonic Orderings in Ta2NiSe5" Crystals 15, no. 5: 414. https://doi.org/10.3390/cryst15050414

APA Style

Chatterjee, B., Golež, D., & Mravlje, J. (2025). Delicate Competition Between Different Excitonic Orderings in Ta2NiSe5. Crystals, 15(5), 414. https://doi.org/10.3390/cryst15050414

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop