Proceedings of the National Academy of Sciences of the United States of America
Home Evidence of ideal excitonic insulator in bulk MoS2 under pressure
Evidence of ideal excitonic insulator in bulk MoS<sub>2</sub> under pressure
Evidence of ideal excitonic insulator in bulk MoS2 under pressure

Edited by Lu Jeu Sham, University of California, San Diego, La Jolla, CA, and approved February 4, 2021 (received for review May 20, 2020)

Author contributions: D.V., E.M., and M.R. designed research; S.S.A., D.V., and M.R. performed research; S.S.A., D.V., and M.R. analyzed data; and S.S.A. and M.R. wrote the paper.

1S.S.A. and D.V. contributed equally to this work.

Article Type: Research Article Article History
Abstract

We claim that MoS2

under pressure becomes the long-sought “excitonic insulator” (EI). This is a permanent condensate of excitons, electron–hole pairs bound by Coulomb interaction, which form in the absence of optical excitation. A surge of experimental claims has recently addressed layered materials, because of reduced Coulomb screening. However, the transition to the putative EI is ubiquitously accompanied by the softening of a phonon inducing a structural change; therefore it remains unclear whether the observed phase is excitonic or instead stabilized by electron–phonon interaction. Our calculations show that MoS2
for a range of applied pressure is unstable against the generation of excitons but stable against lattice distortion: We predict that the EI is an antiferroelectric, electronic density wave.

Spontaneous condensation of excitons is a long-sought phenomenon analogous to the condensation of Cooper pairs in a superconductor. It is expected to occur in a semiconductor at thermodynamic equilibrium if the binding energy of the excitons—electron (e

) and hole (h
) pairs interacting by Coulomb force—overcomes the band gap, giving rise to a new phase: the “excitonic insulator” (EI). Transition metal dichalcogenides are excellent candidates for the EI realization because of reduced Coulomb screening, and indeed a structural phase transition was observed in few-layer systems. However, previous work could not disentangle to which extent the origin of the transition was in the formation of bound excitons or in the softening of a phonon. Here we focus on bulk MoS2
and demonstrate theoretically that at high pressure it is prone to the condensation of genuine excitons of finite momentum, whereas the phonon dispersion remains regular. Starting from first-principles many-body perturbation theory, we also predict that the self-consistent electronic charge density of the EI sustains an out-of-plane permanent electric dipole moment with an antiferroelectric texture in the layer plane: At the onset of the EI phase, those optical phonons that share the exciton momentum provide a unique Raman fingerprint for the EI formation. Finally, we identify such fingerprint in a Raman feature that was previously observed experimentally, thus providing direct spectroscopic confirmation of an ideal excitonic insulator phase in bulk MoS2
above 30 GPa.

Keywords
Ataei,Varsano,Molinari,and Rontani: Evidence of ideal excitonic insulator in bulk MoS2 under pressure

The long-sought excitonic insulator (EI) is a permanent Bose–Einstein condensate of excitons in the absence of optical excitation, hosted in a narrow-gap semiconductor or a semimetal (1234). As the exciton condensate shares similarities with the superconductor ground state (5), it may exhibit macroscopic quantum coherence and exotic low-energy excitations (6789101112). These intriguing features are linked to the arbitrariness of the phase of the condensate wave function, φ (defined in Eq. 2 below): Whereas in the superconductor this phase degeneracy is protected by the conservation of electronic charge, in the EI it is contingent on the preservation of excitons (7, 11) and hence lifted by those terms in the Hamiltonian that annihilate or create e-h pairs. This is the case of e-phonon (13) and spin–orbit (14) interactions, which pin φ while hybridizing conduction and valence bands (remarkably, the combination of spin–orbit coupling and other factors may lead to a topological insulator whose character is inherited by the excitonic state) (14). Other mechanisms that act as sources/sinks for excitons include interband hybridization and Coulomb interaction terms allowed by symmetry (11), disorder (15), and environmental fluctuations of the electrostatic potential (8). So far, the most accomplished EIs were realized in bilayer heterostructures in the presence of a magnetic field, requiring both low temperature and complex engineering to maximize the impact of e-h correlations as well as the degeneracy of φ (10, 16). A related concept aims to achieve the temporary condensation of indirect excitons, made of spatially separated e and h, through the optical pumping of artificial bilayers designed to maximize the exciton lifetime (171819).

Recently, layered materials (20212223) renewed the promise of the EI because of the enhanced Coulomb interactions, and hence exciton binding, due to their reduced dimensionality. In particular, the indirect character of excitons—in reciprocal (20, 21) and real (22, 23) space for TiSe2 and Ta2NiSe5, respectively—led to inherently weaker screening and hence stronger e-h attraction, thus potentially stabilizing the EI phase. In those systems, the putative transition to the EI was accompanied by a lattice instability (2425262728) when lowering the temperature—a singularity in the phonon density of states at vanishing energy—that in turn created e-h pairs through e-phonon interaction. In contrast, the transition to the ideal EI is purely electronic, with only small adjustments of the lattice (4, 29).

Here, we follow an early suggestion by Hromadová et al. (30) and focus on bulk MoS2 under hydrostatic pressure (30313233). We use many-body perturbation theory from first principles (34, 35) to demonstrate that MoS2 is unstable against exciton condensation but stable against lattice distortion. Building a self-consistent effective-mass model on top of ab initio calculations, we show that the true ground state is an ideal, antiferroelectric EI with a distinctive Raman fingerprint that has already been observed (36).

In bulk MoS2, the pressure (P) closes the indirect gap, G, between the top of the filled valence band, located at the center of the Brillouin zone (Γ point), and the bottom of the six-degenerate valleys of the empty conduction band—placed at Λ points (approximately midway between Γ and K; see Fig. 1C for P= 34 GPa). The energy landscape along one of the ΓΛ cuts (sketched in Fig. 1A) favors the Coulomb binding of an e, located at Λ, with an h, placed at Γ, creating an exciton of finite momentum |q|=ΓΛ and binding energy Eb. Whereas ordinarily Eb<G, it may occur that Eb>G above a critical pressure, a condition that makes the semiconductor unstable against the condensation of excitons. This is actually the case, as we show below from first principles.

Indirect-gap MoS2 as a candidate excitonic insulator. (A) Sketch of the excitonic insulator instability, adapted from Walter Kohn’s original proposal (4). An exciton binds an electron at the conduction band bottom, located at Λ in k space, with a hole at the valence band top at Γ. If the exciton binding energy, Eb, is larger than the indirect gap, G, then the system is unstable against the spontaneous generation of excitons. The reconstructed many-body ground state—a condensate of excitons at thermodynamic equilibrium—is the excitonic insulator. (B) Model of the 2Hc crystal structure from different views. The violet (yellow) color labels Mo (S) atoms. The dashed frame appearing in side and top views is the unitary cell of the layered structure, with a and c being the in- and out-of-plane lattice constants, respectively. (C) Lowest conduction and topmost valence energy band as a function of wave vector in the kz=0 plane, as obtained from first-principles many-body perturbation theory (GW) at a pressure of 34 GPa.
Fig. 1.

Indirect-gap MoS2 as a candidate excitonic insulator. (A) Sketch of the excitonic insulator instability, adapted from Walter Kohn’s original proposal (4). An exciton binds an electron at the conduction band bottom, located at Λ in k space, with a hole at the valence band top at Γ. If the exciton binding energy, Eb, is larger than the indirect gap, G, then the system is unstable against the spontaneous generation of excitons. The reconstructed many-body ground state—a condensate of excitons at thermodynamic equilibrium—is the excitonic insulator. (B) Model of the 2Hc crystal structure from different views. The violet (yellow) color labels Mo (S) atoms. The dashed frame appearing in side and top views is the unitary cell of the layered structure, with a and c being the in- and out-of-plane lattice constants, respectively. (C) Lowest conduction and topmost valence energy band as a function of wave vector in the kz=0 plane, as obtained from first-principles many-body perturbation theory (GW) at a pressure of 34 GPa.

So far, ultrahigh pressure has been used as a handle to make MoS2 superconducting (33) (at P 90 GPa), although the pairing mechanism remains unclear (373839). The putative EI must be searched at lower pressure (P 25 GPa), close to the semiconductor–semimetal transition that was observed by several groups (31, 32, 404142). Near this boundary, theory (30, 43)—including our own calculations (SI Appendix, Fig. S1)—predicts an isostructural transition from the 2Hc (Fig. 1B) to the 2Ha (SI Appendix, Fig. S2) phase, which does not affect the crystal space group D6h4, as the two structures transform into each other through the sliding of the layers in the unit cell (the layer unit is made of one Mo and two S atoms, represented respectively by violet and yellow balls in the sketch of Fig. 1B). Raman and X-ray spectroscopic observations (31, 33, 36, 41, 44) suggest that 2Hc and 2Ha phases coexist in diamond-anvil cells, in a range that varies between 25 and 50 GPa in powders but has narrower extension (4 GPa) in single crystals. Importantly, we find that both 2Hc and 2Ha polytypes experience a similar excitonic instability—unrelated to the structural transition, as the electronic bands of the two phases are basically identical close to the Fermi energy. Below, we discuss the 2Hc stacking and leave the analysis of 2Ha to SI Appendix, Figs. S3 and S4.

Results

The indirect gap of 2HcMoS2 is sensitive to pressure, as its value drops from 1.31 eV at P=0 (Fig. 2A) to only 9 meV at P=34 GPa (Fig. 2C), close to the semimetal limit. With respect to the accurate band structure calculated within the GW approximation (circles in Fig. 2 AC) (Materials and Methods), density functional theory (triangles) underestimates the gap of about 0.4 eV at P=0. However, as pressure reduces the out-of-plane lattice parameter c (SI Appendix, Fig. S1), forcing sulfur orbitals belonging to adjacent layers to overlap (45), virtual e-h pairs start tunneling among layers, screening effectively Coulomb interaction at long wavelength. This reduces the GW energy correction to the density functional theory (DFT) bandgap, as evident in Fig. 2C. Consistently, the conduction band increases its dispersion along the kz direction (Fig. 2F), as well as the other axes of the effective mass tensor (Fig. 2 D and E; dots and lines are GW data and effective-mass fits, respectively). Overall, the semiconductor becomes progressively more isotropic as it turns into a semimetal, losing its two-dimensional character.

Closing the gap by applying pressure. (A–C) Band structure along the Γ−Λ−K cut of the Brillouin zone at pressure P= 0 (A), 25 (B), and 34 GPa (C). Band energies obtained from first principles including the quasiparticle GW corrections beyond DFT (circles) are compared to bare DFT data (triangles, PBE functional). Lines are guides to the eye. (D–F) Band dispersion of conduction and valence bands close to Λ and Γ points, respectively, for P= 0 GPa (black color), 25 GPa (red), and 34 GPa (blue). In E and F the conduction band has been rigidly translated by the wave vector −ΓΛ→. GW predictions (dots) are shown together with effective-mass fits (curves). The directions of the cuts (shown as red arrows in the Brillouin zone) are the principal axes of the effective-mass tensor, two being in the kz=0 plane (D and E) and one being parallel to the kz axis (F).
Fig. 2.

Closing the gap by applying pressure. (A–C) Band structure along the ΓΛK cut of the Brillouin zone at pressure P= 0 (A), 25 (B), and 34 GPa (C). Band energies obtained from first principles including the quasiparticle GW corrections beyond DFT (circles) are compared to bare DFT data (triangles, PBE functional). Lines are guides to the eye. (DF) Band dispersion of conduction and valence bands close to Λ and Γ points, respectively, for P= 0 GPa (black color), 25 GPa (red), and 34 GPa (blue). In E and F the conduction band has been rigidly translated by the wave vector ΓΛ. GW predictions (dots) are shown together with effective-mass fits (curves). The directions of the cuts (shown as red arrows in the Brillouin zone) are the principal axes of the effective-mass tensor, two being in the kz=0 plane (D and E) and one being parallel to the kz axis (F).

Exciton Binding and Instability.

The exciton candidate for the instability has a finite center-of-mass momentum q; i.e., it travels in space. We compute its excitation energy—the difference between the GW bandgap and the binding energy—by solving the Bethe–Salpeter equation from first principles (Materials and Methods). The dispersion exhibits a dip for q=Λ, whose energy is first positive at P=0 (1.26 eV, black dots in Fig. 3A) but then quickly lowers with P, eventually changing sign close to the semimetal threshold (27 meV at P=34 GPa, blue diamonds). This negative value signals that excitons spontaneously form, which leads to a reconstructed many-body phase of lower energy.

Excitonic instability. (A) Excitation energy of the lowest exciton vs. center-of-mass momentum q along the ΓK direction. Data are obtained from first principles for P= 0 (dots), 25 (squares), and 34 GPa (diamonds). Note that the K point position expressed in units of Bohr−1 shifts with P. Solid lines are guides to the eye. At P= 34 GPa the excitation energy is negative for q=ΓΛ, which points to the instability against exciton condensation (the dashed line highlights the energy zero). (B) Binding energy of the exciton having momentum q=ΓΛ (black circles, left vertical axis) and macroscopic static dielectric constant (red circles, right axis) vs. P. The latter is obtained through the inverse dielectric matrix, as 1/ [ϵ−1(q=0)]G=G′=0 (G is the reciprocal lattice vector). (C and D) Wave function square modulus of the lowest exciton with q=ΓΛ at P=0. The plot shows the conditional probability to locate the bound electron (green contour map), provided the hole position is fixed (black dot), either in (C) or out (D) of plane. The violet (yellow) color in the stick-and-ball skeleton points to Mo (S) atoms.
Fig. 3.

Excitonic instability. (A) Excitation energy of the lowest exciton vs. center-of-mass momentum q along the ΓK direction. Data are obtained from first principles for P= 0 (dots), 25 (squares), and 34 GPa (diamonds). Note that the K point position expressed in units of Bohr1 shifts with P. Solid lines are guides to the eye. At P= 34 GPa the excitation energy is negative for q=ΓΛ, which points to the instability against exciton condensation (the dashed line highlights the energy zero). (B) Binding energy of the exciton having momentum q=ΓΛ (black circles, left vertical axis) and macroscopic static dielectric constant (red circles, right axis) vs. P. The latter is obtained through the inverse dielectric matrix, as 1/[ϵ1(q=0)]G=G=0 (G is the reciprocal lattice vector). (C and D) Wave function square modulus of the lowest exciton with q=ΓΛ at P=0. The plot shows the conditional probability to locate the bound electron (green contour map), provided the hole position is fixed (black dot), either in (C) or out (D) of plane. The violet (yellow) color in the stick-and-ball skeleton points to Mo (S) atoms.

The softening of the exciton shown in Fig. 3A validates from first principles the seminal prediction by des Cloizeaux (2): The binding energy remains finite even if the gap vanishes, as explicitly shown in Fig. 3B (black dots). The reason is that conduction and valence band profiles are almost unaffected by P (Fig. 2), as the band edges are displaced in k space, which prevents the macroscopic dielectric constant from diverging (red dots in Fig. 3B). Were the closing gap direct, metal-like screening would dissociate the exciton.

The square modulus of the exciton wave function is illustrated in Fig. 3 C and D, as the conditional probability density to locate the bound electron (green contour map), provided the hole is fixed (black dot). Note that the center-of-mass motion does not appear in this frame. The probability extends tens of angstroms—the feature of Wannier excitons familiar from bulk semiconductors—both in and out of plane, as apparent in Fig. 3 C and D, respectively (the Bohr radius is 50 Å at 34 GPa, as shown in SI Appendix, Fig. S5). The exciton becomes lighter and more isotropic with pressure, i.e., more delocalized in real space (here shown at P=0).

Two-Band Model.

The major source of numerical error is the finite sampling of the Brillouin zone (14), since the exciton is significantly localized in k space while the computational load prevents us from refining the mesh (Materials and Methods). However, the specific features of the exciton provide us with a workaround, since 1) the wave function is spanned essentially by those e and h states that are close to the edges of the lowest conduction and highest valence band, respectively (Fig. 1A) and 2) the spin degree of freedom is irrelevant, the exciton energy being fourfold degenerate within numerical accuracy (spin–orbit coupling is fully included in the calculation). Therefore, we may afford ultradense k-space sampling by replacing the first-principles Bethe–Salpeter equation with its spinless two-band counterpart within the effective mass approximation (35), the mass tensor being extracted from Fig. 2 DF and the dielectric constant from Fig. 3B (Materials and Methods and SI Appendix, Fig. S5). The resulting excitation energy, at the semimetal threshold, is 8 meV.

The Excitonic Insulator Phase.

Close to the semiconductor–semimetal boundary, the ground state undergoes a reconstruction from the “normal” phase, Φ0, which is either insulating or semimetallic, to the excitonic insulator, ΨEI. In the following, we highlight the essential features of ΨEI within the simpler two-band model (as a mnemonic, we adopt the apex “0” to identify quantities of interest defined within this model). Then, we take into account the EI multivalley nature by adapting the theory first proposed for the candidate material TiSe2 (46).

Within the two-band model (3), ΨEI0 is formally analogous to the superconductor wave function (5),

ΨEI0=k[uk0+vk0eiφb^k+âk]Φ0  , [1]
provided the Cooper pairs of the metal are replaced with the e-h pair excitations of the normal state, b^k+âkΦ0. Here b^k+ creates an electron with momentum k +ΓΛ and energy εb(k) in the conduction band, âk annihilates an electron with momentum k and energy εa(k) in the valence band, uk0 and vk0 are positive coherence factors [(uk0)2+(vk0)2=1], and φ is the phase of the condensate wave function, ζk0=uk0vk0eiφ=Δk0/2Ek, with Δk0 being the excitonic gap function and Ek={[εb(k)εa(k)]2/4+|Δk0|2}1/2. The value of φ is—ideally—arbitrary and solely fixed by the spontaneous breaking of the conservation law for e-h pairs, as
ΨEI0b^k+âkΨEI0=ζk0eiφ. [2]
The EI band structure is obtained by solving the pseudo-Bethe–Salpeter equation for ζk0 self-consistently,
2Ekζk0kW(kk)ζk0=0, [3]
where W(q) is the screened Coulomb interaction and the minimum value of 2Ek is the bandgap (Materials and Methods). Reassuringly, Eq. 3 turns into the Bethe–Salpeter equation for the zero-energy exciton at the onset of the EI phase (Δk00+). As a consequence of the condensation energy gain, the EI conduction and valence bands (circles in Fig. 4A) are flattened and distorted with respect to those of the pristine semiconductor (dashed curves), the gap widening by 15 meV at P=34 GPa.

Antiferroelectric excitonic insulator. (A) Band structure of the excitonic (solid curves) and pristine (dashed curves) insulator along one of the six equivalent Γ−Λ directions in the kz=0 plane of the Brillouin zone at P= 34 GPa. The original conduction bands are folded from Λ valleys to Γ and renormalized together with the valence band. The new band structure at Γ is replicated at Λ, since both Γ and Λ points belong to the EI reciprocal lattice. Apart from spin degeneracy, EI renormalized bands exhibit an additional orbital degeneracy reminiscent of the pristine multivalley structure: Bands (solid curves) from top to bottom are respectively one-, three-, two-, and onefold degenerate, respectively. The twofold degenerate band, which overlaps with the pristine conduction band (dashed curve), is actually split due to the tiny anisotropy of Δk in the kx, ky plane (splitting hardly visible in the plot). Circles point to EI conduction and valence bands obtained within the two-band model. (B) Antiferroelectric structure. A permanent out-of-plane electric dipole, Pz(x,y), spontaneously develops and exhibits an in-plane modulation that breaks inversion symmetry. This dipole, which averages to zero over the unitary cell of the superstructure (solid frame), is perpendicular to the plane and depicted as a red arrow of varying sign and modulus. The superstructure cell contains 72 atoms against 6 of the original cell (dashed frame) (here ΓΛ≈2π/(3a)). Violet and blue dots are Mo atoms, respectively, in the top and bottom layer (S atoms are not shown). (C) Overlap charge density of the periodic part of pristine conduction and valence band Bloch states, respectively, at Λ and Γ, shown in the 2Hc cell. The red (blue) color points to a surplus (deficit) of charge. The depicted charge displacement, which is associated with the polarization of condensed excitons, is the origin of the permanent dipole Pz shown in B. (D) Maximum local value of Pz vs. P.
Fig. 4.

Antiferroelectric excitonic insulator. (A) Band structure of the excitonic (solid curves) and pristine (dashed curves) insulator along one of the six equivalent ΓΛ directions in the kz=0 plane of the Brillouin zone at P= 34 GPa. The original conduction bands are folded from Λ valleys to Γ and renormalized together with the valence band. The new band structure at Γ is replicated at Λ, since both Γ and Λ points belong to the EI reciprocal lattice. Apart from spin degeneracy, EI renormalized bands exhibit an additional orbital degeneracy reminiscent of the pristine multivalley structure: Bands (solid curves) from top to bottom are respectively one-, three-, two-, and onefold degenerate, respectively. The twofold degenerate band, which overlaps with the pristine conduction band (dashed curve), is actually split due to the tiny anisotropy of Δk in the kx, ky plane (splitting hardly visible in the plot). Circles point to EI conduction and valence bands obtained within the two-band model. (B) Antiferroelectric structure. A permanent out-of-plane electric dipole, Pz(x,y), spontaneously develops and exhibits an in-plane modulation that breaks inversion symmetry. This dipole, which averages to zero over the unitary cell of the superstructure (solid frame), is perpendicular to the plane and depicted as a red arrow of varying sign and modulus. The superstructure cell contains 72 atoms against 6 of the original cell (dashed frame) (here ΓΛ2π/(3a)). Violet and blue dots are Mo atoms, respectively, in the top and bottom layer (S atoms are not shown). (C) Overlap charge density of the periodic part of pristine conduction and valence band Bloch states, respectively, at Λ and Γ, shown in the 2Hc cell. The red (blue) color points to a surplus (deficit) of charge. The depicted charge displacement, which is associated with the polarization of condensed excitons, is the origin of the permanent dipole Pz shown in B. (D) Maximum local value of Pz vs. P.

Multivalley Effects.

As e-h pairs may be formed by exciting an electron from the valence band to any one of the six conduction band valleys, Λi, the condensate wave function is multicomponent (46), ΨEIb^ik+âkΨEI=ζik, with b^ik+ creating an electron with momentum k +ΓΛi and energy εib(k) in the ith valley (i=1,,6). In principle, one must solve up to six coupled equations for ζik to account for the distortion of the condensate in k space, due to intervalley coupling. Nevertheless, we note that Δk0 has hardly any angular dependence in the kx, ky plane (the maximum amplitude of the azimuthal modulation is smaller than 0.07 meV) (SI Appendix, Fig. S6), whereas εib(k) depends on the angle between ΓΛi and (kx,ky) due to mass anisotropy. As Coulomb interaction protects the cylindrical symmetry of ζik, and since the bare-band anisotropy has negligible effect at valley bottom k0 (where the value of ζik is largest), we neglect the azimuthal dependence of ζik and obtain (Materials and Methods)

ζik=ΨEIb^ik+âkΨEI=16uk0vk0eiφi,i=1,,6. [4]
Here only the magnitude of ζik is fixed (from the self-consistent solution of Eq. 3), whereas the six phases φi remain undetermined. This is sufficient to compute the band structure of the EI (Fig. 4A), as the ground-state energy is independent from φi (Materials and Methods).

There are now one valence and six conduction bands (solid thin lines in Fig. 4A), in place of the two bands (circles) of the superconductor-like model. Some of the conduction bands are degenerate, the degeneracy being respectively one, three, two, and one, from the topmost conduction to the valence band. Importantly, the band structure at Γ is replicated at Λ, as the electronic charge exhibits a supermodulation in real space that we discuss below, the corresponding unit cell (solid frame in Fig. 4B) being larger than the cell of the crystal lattice (dashed frame). As a consequence, bands are folded into the smaller Brillouin zone (SI Appendix, Fig. S6), changing the gap from indirect to direct. Only the valence and topmost conduction bands repel each other, in agreement with the two-band model (circles), whereas the remaining bands, which are unaffected by the presence of the exciton condensate, replicate at Γ the bare valleys and hence reduce the direct gap. Since the location of the valence band top is slightly displaced from Γ along the kz axis (SI Appendix, Fig. S7), by 0.02Bohr1, the actual EI gap is indirect and around 5 meV, smaller than the direct gap at Γ. Note that in Fig. 4A the twofold degenerate band, which almost overlaps with the bare conduction band (dashed curve), splits due to the tiny anisotropy of Δk in the kx, ky plane (the splitting is hardly visible in the plot).

Antiferroelectric Excitonic Insulator.

The EI ground state is invariant under time reversal, and hence the phases of the condensate components that live in two antipodal valleys must have opposite sign (modulus a multiple integer of 2π); i.e., φ1=φ4, φ3=φ6, and φ5=φ2 (SI Appendix, Fig. S6 and Materials and Methods). This constraint leads to the formation of a purely electronic, self-sustained charge density wave, Δϱ(r), which breaks the inversion symmetry of the pristine crystal (the proof is given in Materials and Methods). The total wave Δϱ is the coherent superposition of three contributions, Δϱ=Δϱ1,4+Δϱ3,6+Δϱ5,2, each one originating from a couple of antipodal valleys. For example,

Δϱ1,4(r)=86kuk0vk0×ReψΓ(r)ψΛ1*(r)exp[i(ΓΛ1rφ1)] [5]
exhibits the new periodicity 2π/|ΓΛ1| given by the momentum of those excitons that condense in valleys 1 and 4, and similarly Δϱ3,6 and Δϱ5,2 display an analogous modulation along directions ΓΛ3 and ΓΛ5 with phase shifts φ3 and φ5, respectively. Here ψΓ and ψΛ1 are the periodic envelopes of Bloch states respectively at Γ and Λ1, ψΛ4=ψΛ1*, and the spin has been factored out, since the lattice space group contains a center of inversion and a unique z axis (47). It is clear that the total amount of charge displaced from the pristine background, as well as the amplitude of the charge modulation, is driven by the condensate through kuk0vk0.

Importantly, the arbitrariness of the phases φ1, φ3, and φ5 points to a huge, continuous degeneracy of the ground state. Since the effect of any given two arbitrary phases is merely to rigidly shift the charge pattern Δϱ with respect to the frame origin (Materials and Methods), in the following we take φ1=φ3=φ5=0. The resulting density wave is slightly distorted in the generic case, in which all three phases take arbitrary values (discussion below).

Fig. 4C shows the overlap charge density of the envelopes obtained from first principles, σψΓσ*(r)ψΛ1σ(r)+c.c., which is proportional to Δϱ1,4 in the unit cell at the origin (we have added the subscript σ to ψ since the numerical envelopes are generically spinors in the presence of spin–orbit coupling). The density wave shows an asymmetric pattern—transferring charge mainly between the two Mo atoms, which breaks the inversion symmetry with respect to the origin of the cell (the red [blue] contour map points to a surplus [deficit] of charge). This charge transfer sets a local electric dipole with an in-plane texture, P1,4(x,y), as Δϱ1,4 is modulated by exp[i(ΓΛ1r)]. This dipole may be regarded as the polarization of the excitons coherently built in the condensate (8). Since the contributions to the dipole due to the remaining valleys, P3,6 and P5,2, are obtained by rotating P1,4 by respectively 2π/3 and 2π/3 along the z axis, the total dipole P=P1,4+P3,6+P5,2 is parallel to the z axis. We evaluate this parallel component, Pz, through direct integration over the unit cell (Fig. 4D and Materials and Methods).

The overall charge pattern, Pz(x,y), exhibits an antiferroelectric texture that breaks inversion symmetry. This is shown in Fig. 4B, where local dipoles, which point out of the plane, are depicted as red arrows having length proportional to |Pz|. The electric dipole averages to zero over the unitary cell of the superstructure (solid frame), which contains 72 atoms (with ΓΛ2π/(3a)) against 6 of the original cell (dashed frame). The reconstructed Brillouin zone, which is again hexagonal in the plane but rotated by π/6 (SI Appendix, Fig. S6D), is spanned by any two independent vectors chosen among the ΓΛis. In the generic, degenerate case that φ1, φ3, and φ5 take arbitrary values, we expect a reduction of the maximum local value of |Pz| up to 2/3, together with a variable tilt of the dipole in the plane.

Semiconductor–Semimetal Cross-Over.

The formation of a Fermi surface, made of six e pockets in the Λ valleys and one h pocket at Γ, signals the transition from the semiconductor (Fig. 5B) to the semimetal (Fig. 5D) occurring in the absence of excitonic effects. Fig. 5 B and D shows one of the conduction valleys, displaced by ΓΛ in k space, and the valence band, the filled states being shadowed by gray color. As the free carriers populating the Fermi pockets effectively screen the e-h attraction, we replace the long-range Coulomb force W in Eq. 3 with the vertex interaction proposed by Kozlov and Maksimov (48) to establish self-consistently the range of the force; besides, we extrapolate P-dependent masses from first principles (Materials and Methods).

Excitonic insulator phase diagram. (A) Phase diagram in the P–T space. Lines are guides to the eye. The shadowed area highlighted in cyan (green) color is the excitonic gapped (gapless) phase. The vertical dashed line points to the semiconductor–semimetal boundary in the absence of excitonic effects. (B–E) Bare energy bands and wave function of the exciton driving the instability in the e-h center-of-mass frame, evaluated in reciprocal space along the ΓΛ direction. The e-h pair of wave vector q is made of a hole with momentum −q and an electron with momentum q+ΓΛ (in B and D the bare conduction band has been displaced by the vector −ΓΛ→ and the shadowed region highlights occupied states). Going from P=34 GPa (B and C) to P=34.12 GPa (D and E), a Fermi surface forms as conduction and valence band overlap in energy. Consequently, plasmonic features appear in the exciton wave function, the spectral weight accumulating close to the Fermi surface (E). In D the Fermi energy is negative as a consequence of the sixfold valley degeneracy at Λ.
Fig. 5.

Excitonic insulator phase diagram. (A) Phase diagram in the PT space. Lines are guides to the eye. The shadowed area highlighted in cyan (green) color is the excitonic gapped (gapless) phase. The vertical dashed line points to the semiconductor–semimetal boundary in the absence of excitonic effects. (BE) Bare energy bands and wave function of the exciton driving the instability in the e-h center-of-mass frame, evaluated in reciprocal space along the ΓΛ direction. The e-h pair of wave vector q is made of a hole with momentum q and an electron with momentum q+ΓΛ (in B and D the bare conduction band has been displaced by the vector ΓΛ and the shadowed region highlights occupied states). Going from P=34 GPa (B and C) to P=34.12 GPa (D and E), a Fermi surface forms as conduction and valence band overlap in energy. Consequently, plasmonic features appear in the exciton wave function, the spectral weight accumulating close to the Fermi surface (E). In D the Fermi energy is negative as a consequence of the sixfold valley degeneracy at Λ.

The resulting EI phase extends over a narrow interval of 0.35 GPa, reaching a maximum critical temperature of T60 K at P34.05 GPa, which is the semiconductor–semimetal boundary in the normal state (vertical dashed line in Fig. 5A). Importantly, the downward shift of the valence band shown in Fig. 4A opens/widens the gap over a pressure range that extends to values that would lead to a semimetal for ζik=0. In the PT diagram of Fig. 5A, the gapped excitonic phase, highlighted as a shadowed cyan area, is the overwhelming part of the larger region that sustains a finite condensate of excitons, ζik>0. The remaining excitonic region—the thin green slice located between P 34.19 and 43.22 GPa—is gapless (SI Appendix, Fig. S7) and ends on the semimetal frontier where ζik=0. Here the critical pressure is equivalent to an amount of free carriers (the density per species is 1.1107Bohr3) comparable to the maximum number of excitons in the condensate (2.2107Bohr3). This overall behavior is in stark contrast with that of the EI candidate TiSe2, which has a multivalley structure similar to that of MoS2 but remains a semimetal due to the unintentional doping of Ti atoms (46).

The exciton responsible for the instability of the conventional semiconductor exhibits a mixed transverse–longitudinal polarization (49), due to the small C2 symmetry of the ΓΛ line (this is also the case of the displacement vectors of the vibrational mode of Fig. 6D). As one moves from the semiconductor to the semimetal, the exciton smoothly turns into a plasmon (4), as illustrated by the wave function in the e-h center-of-mass frame (Materials and Methods). Whereas in the semiconductor (Fig. 5C) the amplitude is Lorentzian-like in k space, similar to that of a familiar Wannier exciton in the bulk, in the semimetal it acquires plasmonic features, as the wave function accumulates close to the Fermi surface (Fig. 5E). Outside the EI phase, this exciton–plasmon dissolves into the continuum of e-h excitations. Note that there may be other long-lived interband plasmons, since small gaps open in the e-h energy continuum due to the degeneracy of Λ valleys. Were there only one valley, then the Fermi energy would be at the crossing of a and b bands (ignoring the mass anisotropy; cf. Fig. 5D) and the e-h excitation spectrum would be gapless.

Phonon dispersion and Raman fingerprint. (A and B) Dispersion of the lowest-energy phonon modes for P=0 (A) and 34 GPa (B), respectively, computed from first principles. All modes harden with P. The red dot points to the lowest optical mode that is folded from Λ into Γ through the excitonic insulator phase transition. (C) Raman spectrum of the normal phase from first principles, for pressures P= 0, 15, 20, 25, and 35 GPa, respectively, from bottom to top. The peaks are broadened using Gaussians with a standard deviation of 2 cm−1. The bright peak at lower (higher) frequency has E2g1 (A1g) symmetry. The plot compares with figure 4b of ref. 36. (D) Displacement vectors for the mode labeled as a red dot in B, as viewed in the excitonic insulator reconstructed cell along the ΓΛ direction (parallel to the y axis in the adopted frame) (47). The superlattice constant is 3a. The violet (yellow) color labels Mo (S) atoms. This mode is Raman active and degenerate with the one folded from Λ′ to Γ.
Fig. 6.

Phonon dispersion and Raman fingerprint. (A and B) Dispersion of the lowest-energy phonon modes for P=0 (A) and 34 GPa (B), respectively, computed from first principles. All modes harden with P. The red dot points to the lowest optical mode that is folded from Λ into Γ through the excitonic insulator phase transition. (C) Raman spectrum of the normal phase from first principles, for pressures P= 0, 15, 20, 25, and 35 GPa, respectively, from bottom to top. The peaks are broadened using Gaussians with a standard deviation of 2 cm1. The bright peak at lower (higher) frequency has E2g1 (A1g) symmetry. The plot compares with figure 4b of ref. 36. (D) Displacement vectors for the mode labeled as a red dot in B, as viewed in the excitonic insulator reconstructed cell along the ΓΛ direction (parallel to the y axis in the adopted frame) (47). The superlattice constant is 3a. The violet (yellow) color labels Mo (S) atoms. This mode is Raman active and degenerate with the one folded from Λ to Γ.

Raman Fingerprint.

Were ion displacements responsible for the building of electric dipoles in place of excitons, the frequency of the phonon of momentum q =ΓΛ and consistent symmetry would soften (or at least exhibit a dip) at the onset of the new phase (50). The phonon dispersion obtained from first principles, respectively at P=0 (Fig. 6A) and 34 GPa (Fig. 6B), shows the opposite behavior, with all low-energy modes hardening with P (see Materials and Methods and SI Appendix, Fig. S8 for the 2Ha phase). Therefore, the antiferroelectricity has a purely electronic origin. This prediction is consistent with recent diffraction measurements, which ruled out any periodic lattice distortion above 40 K (44).

The evolution of Raman spectrum with pressure, as obtained from first principles in Fig. 6C (structure 2Hc) and SI Appendix, Fig. S9 (2Ha), compares with observed data with the exception of the E peak at 174 cm1 (figure 4b in ref. 36), which appears below 150 K and above 30 GPa but is missed by the theory for the normal phase. Cao et al. (36) proposed this mode is a transverse acoustic phonon of finite momentum, which becomes bright at the onset of a charge density wave, due to the reconstruction of the Brillouin zone. Whereas the first-principles spectrum for the excitonic phase is presently out of reach, below we confirm the essence of Cao’s explanation by identifying E as the lowest optical phonon at Λ. This is the fingerprint of the antiferroelectric charge density wave associated with exciton condensation.

The symmetry group of the antiferroelectric ground state depicted in Fig. 4B includes only the identity operation. Therefore, all 216 vibrational modes are in principle infrared and/or Raman active. However, since the EI critical temperature is relatively low and the E peak is extremely bright, we expect that the new mode is an optical phonon of momentum Λ, which is Raman active through the folding into the zone center and strongly couples with P. Since P(x,y) originates everywhere in the cell from the interlayer vertical displacement of the charge between two neighbor Mo atoms, it will mainly couple with those optical oscillations of Mo atoms that occur along the z axis. In fact, these vibrations linearly change the Mo-Mo distance and hence the local dipole strength, whereas the amount of displaced charge, which is ruled by the long-range part of Coulomb interaction, changes weakly with the oscillation. From direct inspection of phonon eigenvectors, there is one candidate only below 400 cm1, i.e., the lowest optical mode of frequency 164 cm1 located at Λ, which is highlighted by a red dot in Fig. 6B. As shown by the displacement vectors in the EI reconstructed cell displayed in Fig. 6D, the Mo atoms oscillate out of phase along the z direction with an in-plane modulation of period 3a along the ΓΛ direction (parallel to the vertical axis of Fig. 4B), hence matching the periodicity of Pz(x,y) in the plane. This superlattice vibration is twice degenerate, due to the additional folding of the phonon with independent wave vector ΓΛ. Note that the observed intensity of the E mode is constant up to 60 K, which compares with the EI critical temperature. In summary, the E mode points to the excitonic insulator in the PT space.

Discussion

Both 2Hc and 2Ha phases coexist (313233, 41) in the region of visibility of the E mode, which extends between 30 and 50 GPa at 5 K (36). The lower bound agrees with our prediction, since in the 2Ha structure the EI sets in at P28 GPa (SI Appendix, Fig. S4) with a mode frequency of 166 cm1 (SI Appendix, Fig. S8). The upper bound of 50 GPa is larger than our expectation of 34 GPa for the 2Hc phase. However, recent diffraction measurements on single crystals (44), although available only at temperatures higher than 40 K, suggest that the critical upper pressure could be actually much lower, being artificially enhanced in powders due to the deviatoric stress field applied to randomly oriented crystallites.

In addition, other Raman features unexplained so far (36) point to the EI scenario: 1) the observation of modes supposedly forbidden or silent and 2) the anomalous frequency variation of the out-of-plane A1g mode accompanying the onset of the E mode. Since the understanding of the available electrical transport measurements (31, 36) is complicated by the mixture of phases in the high-pressure cell, we do not speculate on the origin of the resistivity peak that was tentatively assigned (31) to the EI.

The huge degeneracy of the EI ground state, associated with condensate phases φ1, φ3, and φ5, points to the emergence of acoustic-like electronic excitations—collective phase modes that, if gapless, would manifest exciton superfluidity (4). Within the two-band model of an isotropic semimetal, Kozlov and Maksimov (51) predicted that the “excitonic sound” velocity, cexciton=kF(3mamb)1/2, is proportional to the Fermi wave vector in the normal phase, kF (ma and mb are valence and conduction band masses). By taking average values at the EI/semimetal boundary, we estimate cexciton2104 m/s, which is much higher than the sound velocity of the stiffest acoustic phonon branch (Fig. 6B), cphonon8103 m/s. Therefore, the phase mode of the exciton condensate should be experimentally accessible, even if it might be gapped due to the mechanisms mentioned in the Introduction and not included in our analysis of the EI phase.

Conclusion

In summary, we have demonstrated that a real excitonic insulator phase sets in between the semiconducting and semimetallic phases of MoS2, building on calculations from first principles and available spectroscopic data. These findings call for further investigation of some fascinating possibilities. A first question is the manifestation of the macroscopic quantum coherence of the exciton condensate, which might occur through the observation of low-lying collective modes associated with the oscillation of the condensate phase φ(r,t). Another issue is whether the superconductivity observed above 90 GPa is related to the excitonic phase, as the overscreening action of surviving exciton–plasmons might act as unconventional glue for Cooper pairs. We hope our study may stimulate further work along these paths.

Materials and Methods

Computational Details of Ground-State Calculation from First Principles.

The lattice parameters and the ground-state electronic structure for the three values of pressure were obtained within DFT, with a plane wave basis set as implemented in the Quantum ESPRESSO package (52, 53), using the generalized gradient approximation Perdew–Burke–Ernzerhof (PBE) parameterization (54). A kinetic energy cutoff of 100 Ry was adopted for the wave functions, and fully relativistic norm-conserving pseudopotentials (55) were used to take into account spin–orbit interaction. van der Waals interactions, included by using the Grimme approximation method, were found to be relevant only at zero pressure, as already shown in ref. 30.

Phonons.

Phonon dispersions were calculated by using a density functional perturbation theory approach (56). We used a 10 × 10 × 3 Monkhorst–Pack grid for the integration in the Brillouin zone; the dynamical matrix at a given point of the Brillouin zone was obtained from a Fourier interpolation of the dynamical matrices computed on a 5 × 5 × 1 q-point mesh.

Quasiparticles and Excitons.

Many-body calculations (34, 57, 58) were performed by using the Yambo code (59, 60). Quasiparticle corrections to the Kohn–Sham energies were evaluated using the G0W0 approximation for the self-energy, the dynamical dielectric screening being accounted for within the plasmon-pole approximation (61). To speed up the convergence of quasiparticle energies with respect to the number of empty bands in the sum over states occurring in the calculation of the polarizability and the self-energy, we have adopted the scheme proposed in ref. 62. Fifty empty bands were used to build the polarizability and to integrate the self-energy (SI Appendix, Fig. S10); the Brillouin zone was sampled by using a 27 × 27 × 3 k-point grid. Quasiparticle energies were converged by using 68- and 15-Ry kinetic energy cutoffs for the exchange and correlation parts of the self-energy (SI Appendix, Fig. S11), respectively. Excitation energies and dispersion of the lowest exciton with finite wavevector q were calculated by solving the Bethe–Salpeter equation (BSE) using the Yambo code where the finite-q BSE was implemented as described in refs. 63 and 64. The static screening in the direct term was calculated within the random phase approximation with the inclusion of local field effects; the Tamm–Dancoff approximation for the Bethe–Salpeter Hamiltonian was employed, after having verified that the correction introduced by coupling the resonant and the antiresonant part was negligible for q = 0. Converged excitation energies were obtained considering respectively three valence and five conduction bands in the Bethe–Salpeter matrix, the irreducible Brillouin zone being sampled with a 27 × 27 × 3 k-point grid (SI Appendix, Fig. S12).

Computational Details of the Two-Band Model.

The effective-mass framework builds on the knowledge of conduction

εb(k)=G2+22(k+ΓΛ)2mb+k2mb+kz2mbz [6]
and valence
εa(k)=G222k2ma+k2ma+kz2maz [7]
energy bands. Here G>0 (G<0) is the indirect bandgap (band overlap) for pressures below (above) the semiconductor–semimetal threshold—in the absence of excitonic effects—and the momentum components, k, k, kz, are projected along the principal axes of the effective mass tensor (65), the corresponding masses being mi, mi, miz, with i=a,b. These axes are respectively parallel (k) and perpendicular (in-plane (k) and out-of-plane (kz)) to the ΓΛ direction, the axis origin being placed at the band edge. We emphasize that all parameters of the two-band model, for a given pressure, are fixed and obtained from first principles. In particular, the bandgap and the effective masses are extracted from GW bands, as illustrated in Fig. 2 DF, and hence include the mean-field renormalization due to e-e interactions. The (modulus) of the screened e-h Coulomb attraction in momentum space,
W(q)=1κr4πe2Ω1q2, [8]
depends on the static dielectric constant, κr, which is obtained as the inverse of the first-principles dielectric tensor, 1/[ϵ1(q=0)]G=G=0, in the long-wavelength, macroscopic limit, as illustrated in Fig. 3B (here Ω is the crystal volume and G the reciprocal lattice vector).

In the semimetal, the P-dependent values of G, mi, mi, miz, and κr are derived as linear extrapolations of first-principles data at P= 25 and 34 GPa, respectively. Since free e and h carriers effectively screen the interaction by adding a metal-like, intraband contribution to the polarizability, we modify the dressed Coulomb potential as

W(q)=1κr+4πe2D(εF)/q24πe2Ω1q2. [9]
Here the Thomas–Fermi term, proportional to the density of states, D(ε), evaluated at the Fermi energy, εF, removes the long-wavelength divergence of W. We obtain numerically D through the summation of Gaussian functions over a fine grid in k space, as well as εF by imposing overall charge neutrality (we take into account the sixfold degeneracy of the conduction band).

Two-Band Bethe–Salpeter Equation.

In the semiconductor, the exciton wave function is

exciton=kϕkb^k+âkΦ0, [10]
where ϕk is the probability amplitude of a bound e-h pair in momentum space. The Bethe–Salpeter equation of motion for ϕk is
εb(k)εa(k)ϕkkW(kk)ϕk=εexcϕk, [11]
where εexc is the excitation energy of the exciton, whose negative value signals the instability. We solve this equation by numerical discretization in k space and assess convergence by refining the mesh as well as varying the momentum cutoff. Note that the singularity of Coulomb potential for q0 is harmless, as we integrate W over a small parallelepiped, in a semianalytical, accurate manner. We have benchmarked the convergence of our calculations against known analytical or high-precision results, as shown for bulk Wannier excitons in SI Appendix, Fig. S13 and for anisotropic excitons with a well-defined azimuthal quantum number (66) in SI Appendix, Fig. S14.

In the semimetal ground state, a small area of k space around the origin is populated by electrons in band b and holes in band a. In addition, due to band anisotropy (67), nearby regions exist populated by either electrons or holes only, which prevents from exciting e-h pairs due to Pauli exclusion principle. Therefore, the Bethe–Salpeter equation of motion must be modified as (4)

εb(k)εa(k)ϕkkW(kk)na(k)nb(k)ϕk=εexcϕk, [12]
where ni(k) is the occupancy factor of the ith band in the normal ground state, which takes either 0 or 1 as a value. The “counting” prefactor of W, nanb, removes scattering channels forbidden by Pauli blocking and is responsible of the plasmon-like features shown Fig. 5E. Note that in the semiconductor, na(k)=1 and nb(k)=0, and hence one regains the standard form of Eq. 11.

Self-Consistent Theory of the Excitonic Insulator within the Two-Band Model.

The EI bands (circles in Fig. 4A) are Ebk=εb(k)+εa(k)/2+Ek and Eak=εb(k)+εa(k)/2Ek, with Ek being fixed by the solution of the gap Eq. 3 for Δk0 (through ζk0). Eq. 3 is solved self-consistently by numerical recursion, exploiting the exciton wave function ϕk as a seed (35). If the semimetal is the normal ground state, the gap equation maintains the form 3 of the main text, provided that 1) the summation over k is limited to those points whose occupancies are such that na(k)nb(k)0 to comply with Fermi statistics (67) and 2) the dressed Coulomb interaction W is renormalized by a vertex correction associated with the EI ground state (48), as the opening of the many-body gap significantly enhances the e-h attraction—by suppressing screening—with respect to the gapless normal phase. Therefore, following Kozlov and Maksimov (48), for small momentum transfer q the dressed interaction W appearing in Eq. 3 takes the self-consistent form

W(q)=11+α/(ΔkF0)21κr4πe2Ω1q2, [13]
where the gap function at the Fermi surface, ΔkF0, which is determined recursively, removes the long-wavelength divergence as one approaches the EI–semimetal boundary. Here ΔkF0 is an average value defined as ΔkF0=ΔkxF,0,00Δ0,kyF,00Δ0,0,kzF01/3, with kxF given implicitly by εF=εb(kxF,0,0), and similarly for kyF and kzF. The constant α, for given band overlap G<0, is α=[G0(εFG/2)3/2]1/2, where G0=9.38 meV is the maximum magnitude of the band overlap at which e-h pairing takes place. We neglect the modification of Eq. 13 for large momentum transfer, as it turns out to be irrelevant numerically. Whereas the vertex form 13 was originally proposed (48) for the case of spherically symmetric e and h pockets, we note that, at the semiconductor–semimetal threshold, the exciton responsible for the instability is essentially isotropic (SI Appendix, Fig. S5). At finite temperature, T, the gap equation takes the form
2Ekζk0kW(kk)ζk0fF(EakεF)fF(EbkεF)=0, [14]
where fF(x)=1/[1+exp(βx)] is the Fermi distribution function, with β=1/kBT and kB being the Boltzmann constant, and we neglect the small renormalization of the chemical potential due to the presence of the exciton condensate.

Multivalley Band Structure.

The calculation of the EI band structure relies on the theory by Monney et al. (46) to include valley degeneracy. This approach, based on Green functions, generalizes to multiple bands the original theory by Jérome et al. (3). For every k point, the EI band energies (solid lines in Fig. 4A and SI Appendix, Fig. S7) are found as the seven roots of the equation

zεa(k)i=16Δi(k)2zεib(k)=0 [15]
(cf. equation 8 of ref. 46), after the magnitudes of the excitonic gap components, Δi(k), are obtained as follows. The gap function is defined as
Δi(p)=kW(k)ζik+p, [16]
with ζik, apart from a phase factor, being the equal-time interband excitonic coherence Fi(k,t,t) defined in equation 4 of ref. 46,
ζik=iFi(k,t+δ,t)=12πi   dωFi(k,ω)eiωδ, [17]
and δ0+ being a positive infinitesimal quantity. The integral 17 is evaluated through contour integration, the Fourier transform Fi(k,ω) being derived from the equations of motion of Green functions (46) as
Fi(k,ω)=Δi(k)ωεa(k)jiΔj(k)2ωεjb(k)1×ωεib(k)Δi(k)2ωεa(k)jiΔj(k)2ωεjb(k)11. [18]
Whereas this expression would generically lead to an intractable system of six coupled equations for the Δis, we exploit the high symmetry of the problem to simplify the form of Fi(k,ω) and recover a single gap equation. As discussed in the main text and SI Appendix, Fig. S6, the symmetrizing effect of e-h attraction makes Δk0 almost independent from the azimuthal angle φk, with k (k,φk,kz) being expressed in cylindrical coordinates (k is the in-plane radial distance and kz the component along the z axis). Therefore, it is natural to assume that Δi has cylindrical symmetry, Δi(k)=Δ(k,kz)eiφi. Since we are mainly interested in the region k0, we also neglect the azimuthal dependence of εib(k) in the denominator of Fi, obtaining
Fi(ω)=Δiωεa5Δ2ωεibωεibΔ2ωεa5Δ2ωεib, [19]
where we omitted the dependence of terms on k in the notation. Eq. 19 is now easily integrated, giving a single self-consistent gap equation. This has the same form as Eq. 3 of the two-band model, provided that Δk0 is replaced with 6Δi(k).

Ground-State Wave Function.

The contour integration of equal-time Green functions provides us with all interband coherences and band populations, i.e., ΨEIb^jk+b^ikΨEI=Δi*Δj/2E(E+εb/2εa/2), ΨEIb^ik+b^ikΨEI=(v0)2/6, ΨEIâk+âkΨEI=(u0)2, where we omitted the dependence of right-hand-side terms on k to ease the notation, neglected the in-plane anisotropy of the valence band, εib=εb, and put E={[εbεa]2/4+|Δ0|2}1/2. This allows us to write explicitly the ground-state wave function,

ΨEI=kγ^k+vacuum  , [20]
in terms of Bogoliubov–Valatin-like creation operators, γ^+, which are defined as
γ^k+=uk0âk++vk06i=16eiφib^ik+. [21]
As discussed in the main text, time reversal symmetry limits the number of independent condensate phases to three: φ1, φ3, and φ5 (recall that uk0=uk0 and vk0=vk0 are real positive quantities) (SI Appendix, Fig. S6C).

Inversion Symmetry Breaking.

The ground-state wave function allows us to understand the symmetry breaking associated with exciton condensation. The inversion operator, I^, acts differently on bi and a Bloch states, since the envelope function at Γ is odd: I^âk+=âk+, I^b^1k+=b^4k+, etc. Therefore, the inverted ground state, I^ΨEI, is not proportional to the original one:

ΨEII^ΨEI=k{(uk0)2+(vk0)23cos(2φ1)+cos(2φ3)+cos(2φ5)}. [22]
The magnitude of the expression enclosed in curly brackets is less than one (unless φ1=φ3=φ5=±π/2, i.e., I^ΨEI=ΨEI); hence, in the thermodynamic limit, the overlap between I^ΨEI and ΨEI tends to zero as the two states become orthogonal. Since the ground state has a lower symmetry than the Hamiltonian, inversion symmetry is broken.

Charge Density Wave.

The form 5 of the purely electronic charge density wave, Δϱ=Δϱ1,4+Δϱ3,6+Δϱ5,2, is derived in a straightforward manner by averaging the density operator, ϱ^(r)=2ψ^(r)ψ^(r), over ΨEI, with the Fermi field operator, ψ^(r), being defined as

ψ^(r)=keikrψΓ(r)âk+i=16exp(iΓΛir)ψΛi(r)b^ik. [23]
Cross-terms proportional to ψ*ΛiψΛj average out to zero, once summed together, as the various ψΛis are obtained one from the other by either rotation by ±2π/3 along the z axis or complex conjugation. Apart from the envelope functions, which have the lattice periodicity, Δϱ depends on r through a sum over three exponentials, whose imaginary arguments are respectively (times the prefactor i) ΓΛ1rφ1, ΓΛ3rφ3, and ΓΛ5rφ5, as illustrated in the main text.

We show below that, for any given two condensate phases, say φ1 and φ3, there exist a lattice vector Rshift and a phase φ5=φ1φ3 such that a rigid translation of the density wave by Rshift provides the density wave corresponding to φ1=φ3=φ5=0; i.e., [Δϱ(rRshift)]φ1,φ3,φ5=[Δϱ(r)]0,0,0.

Let us construct explicitly Rshift as Rshift=RR, where R=nt2 and R=n(2t1+t2) are respectively parallel and perpendicular to ΓΛ1 (SI Appendix, Fig. S6C); n and n are integers to be determined; and t1, t2 are the primitive vectors that generate the hexagonal lattice in Mattheiss’s (47) coordinate frame. Since ΓΛ1 is generically not commensurable with the reciprocal lattice vectors, there exists an integer n such that ΓΛ1R=φ1 with arbitrary accuracy (4), modulus an integer multiple of 2π. Similarly, we may fix n such that ΓΛ3R=ΓΛ5R=φ3ΓΛ3R. Finally, we take φ5=φ3+2ΓΛ3R=φ3φ1. One may verify, by direct substitution into the expression Δϱ=Δϱ1,4+Δϱ3,6+Δϱ5,2, that [Δϱ(rRshift)]φ1,φ3,φ1φ3=[Δϱ(r)]0,0,0, Q.E.D.

This theorem implies that the set of charge density waves [Δϱ(r)]0,0,φ5 labeled by the continuous parameter φ5 spans all possible modulations of the electronic charge density of the EI, each realization having in turn a huge translational degeneracy, which is parameterized by the two continuous variables φ1 and φ3.

Antiferroelectric Order.

The electronic charge density wave of the EI ground state (Eq. 5) induces an out-of-plane electric dipole, Pz(Ri), in the ith cell of the pristine 2H phase located at Ri, with i=1,,N (N is the total number of cells). This is illustrated in Fig. 4B, where the dipoles Pz(Ri) are depicted as red arrows. The local dipole Pz(Ri) is given by the coherent superposition of three density waves, whose characteristic wave vectors are qi=ΓΛi, with i=1,3,5,

Pz(Ri)=Pz0Ωk46uk0vk0j=1,3,5cos(qjRi). [24]
The maximum value, Pz(0), is shown in Fig. 4D. Here Pz(Ri) is evaluated within the envelope function approximation, the factor Pz0 being derived from first principles through the overlap charge density of the periodic part of conduction and valence Bloch states at Γ and Λ, respectively, which is shown in Fig. 4C. The latter is numerically integrated over the pristine unit cell volume, Ωcell:
Pz0=σe  Ωcell   drzψ*Γσ(r)ψΛσ(r)+c.c., [25]
the frame origin being placed at the inversion center—the midpoint between the two Mo atoms of the 2H cell. As the charge displacement that gives rise to the dipole is essentially localized on Mo atoms (Fig. 4C), we expect Pz0 to be well defined. We obtain Pz0/e= 15.1 Bohr at P=34 GPa.

Acknowledgements

D.V. acknowledges the joint work with Davide Sangalli to implement the finite-momentum Bethe–Salpeter calculation into the Yambo code. This work was supported in part by the “MAterials design at the eXascale” (MaX) European Center of Excellence (www.max-centre.eu) funded by the European Union H2020-INFRAEDI-2018-1 program, Grant 824143. It was also supported by the Italian national program PRIN2017 2017BZPKSZ “Excitonic insulator in two-dimensional long-range interacting systems.” We acknowledge access to the Marconi supercomputing system based at Consorzio Interuniversitario per la Gestione del Centro di Calcolo Elettronico dell’Italia Nord-Orientale (CINECA), Italy, through Partnership for Advanced Computing in Europe and the Italian SuperComputing Resource Allocation program.

The authors declare no competing interest.
This article is a PNAS Direct Submission.
This article contains supporting information online at https://www.pnas.org/lookup/suppl/doi:10.1073/pnas.2010110118/-/DCSupplemental.

Data Availability

Many-body perturbation theory calculations were performed by means of the codes Yambo (www.yambo-code.org/) and Quantum ESPRESSO (www.quantum-espresso.org), which are both open source software. Results for the two-band model were obtained through a custom Fortran code that is available at Zenodo, doi.org/10.5281/zenodo.4455373. The data that support the findings of this study (crystal structures) are available at Zenodo, doi.org/10.5281/zenodo.4455373 (68).

References

L. V. Keldysh, Y. V. Kopaev, Possible instability of the semimetallic state against Coulomb interaction. Sov. Phys. Sol. State 6, 2219 (1965).

J. des Cloizeaux, Excitonic instability and crystallographic anomalies in semiconductors. J. Phys. Chem. Solid. 26, 259 (1965).

D. Jèrome, T. M. Rice, W. Kohn, Excitonic insulator. Phys. Rev. 158, 462 (1967).

W. Kohn, “Metals and insulators” in Many-Body Physics, C. de Witt, R. Balian, Eds. (Gordon & Breach, New York, NY, 1967), pp. 351411.

J. Bardeen, L. N. Cooper, J. R. Schrieffer, Theory of superconductivity. Phys. Rev. 108, 11751204 (1957).

B. I. Halperin, T. M. Rice, The excitonic state at the semiconductor-semimetal transition. Solid State Phys. 21, 115 (1968).

R. R. GuseÄŋnov, L. V. Keldysh, Nature of the phase transition under the condition of an “excitonic” instability in the electronic spectrum of a crystal. Sov. Phys. JETP 36, 1193 (1973).

T. Portengen, T. Östreich, L. J. Sham, Theory of electronic ferroelectricity. Phys. Rev. B 54, 17452 (1996).

L. Pitaevskii, S. Stringari, Bose-Einstein Condensation (Oxford University Press, Oxford, UK, 2003).

10 

J. P. Eisenstein, A. H. MacDonald, Bose-Einstein condensation of excitons in bilayer electron systems. Nature 432, 691694 (2004).

11 

P. B. Littlewood, “Exciton coherence” in Problems of Condensed Matter Physics, A. L. Ivanov, S. G. Tikhodeev, Eds. (International Series of Monographs on Physics, Oxford University Press, Oxford, UK, 2008), vol. 139, pp. 163181.

12 

M. Rontani, L. J. Sham, “Coherent exciton transport in semiconductors” in Novel Superfluids Vol. 2, K. H. Bennemann, J. B. Ketterson, Eds. (International Series of Monographs on Physics, Oxford University Press, Oxford, UK, 2014), vol. 157, pp. 423474.

13 

V. A. Volkov, Y. V. Kopaev, Theory of phase transitions in semiconductors of the A4B6 group. Sov. Phys. JETP 37, 11031108 (1974).

14 

D. Varsano, M. Palummo, E. Molinari, M. Rontani, A monolayer transition-metal dichalcogenide as a topological excitonic insulator. Nat. Nanotechnol. 15, 367372 (2020).

15 

B. Remez, N. R. Cooper, Effects of disorder on the transport of collective modes in an excitonic condensate. Phys. Rev. B 101, 235129 (2020).

16 

A. Nandi, A. D. K. Finck, J. P. Eisenstein, L. N. Pfeiffer, K. W. West, Exciton condensation and perfect Coulomb drag. Nature 488, 481 (2012).

17 

L. V. Butov, C. W. Lai, A. L. Ivanov, A. C. Gossard, D. S. Chemla, Towards Bose–Einstein condensation of excitons in potential traps. Nature 417, 4752 (2002).

18 

A. A. High, Spontaneous coherence in a cold exciton gas. Nature 483, 584588 (2012).

19 

R. Anankine, Quantized vortices and four-component superfluidity of semiconductor excitons. Phys. Rev. Lett. 118, 127402 (2017).

20 

T. Rohwer, Collapse of long-range charge order tracked by time-resolved photoemission at high momenta. Nature 471, 490494 (2011).

21 

A. Kogar, Signatures of exciton condensation in a transition metal dichalcogenide. Science 358, 13141317 (2017).

22 

Y. F. Lu, Zero-gap semiconductor to excitonic insulator transition in Ta2NiSe5. Nat. Commun. 8, 14408 (2017).

23 

D. Werdehausen, Coherent order parameter oscillations in the ground state of the excitonic insulator Ta2NiSe5. Sci. Adv. 4, eaap8652 (2018).

24 

F. J. D. Salvo, D. E. Moncton, J. V. Waszczak, Electronic properties and superlattice formation in the semimetal TiSe2. Phys. Rev. B 14, 4321 (1976).

25 

H. Hedayat, Excitonic and lattice contributions to the charge density wave in 1T-TiSe2 revealed by a phonon bottleneck. Phys. Rev. Res. 1, 023029 (2019).

26 

J. S. Zhou, Anharmonicity and doping melt the charge density wave in single-layer TiSe2. Nano Lett. 20, 48094815 (2020).

27 

A. Nakano, Antiferroelectric distortion with anomalous phonon softening in the excitonic insulator Ta2NiSe5. Phys. Rev. B 98, 045139 (2018).

28 

J. Yan, Strong electron-phonon coupling in the excitonic insulator Ta2NiSe5. Inorg. Chem. 58, 90369042 (2019).

29 

W. Kohn, D. Sherrington, Two kinds of bosons and Bose condensates. Rev. Mod. Phys. 42, 111 (1970).

30 

L. Hromadová, R. Martonák, E. Tosatti, Structure change, layer sliding, and metallization in high-pressure MoS2. Phys. Rev. B 87, 144105 (2013).

31 

Z. H. Chi, Pressure-induced metallization of molybdenum disulfide. Phys. Rev. Lett. 113, 036802 (2014).

32 

A. P. Nayak, Pressure-induced semiconducting to metallic transition in multilayered molybdenum disulphide. Nat. Commun. 5, 3731 (2014).

33 

Z. Chi, Superconductivity in pristine 2Ha-MoS2 at ultrahigh pressure. Phys. Rev. Lett. 120, 037002 (2018).

34 

G. Onida, L. Reining, A. Rubio, Electronic excitations: Density-functional versus many-body Green’s-function approaches. Rev. Mod. Phys. 74, 601659 (2002).

35 

D. Varsano, Carbon nanotubes as excitonic insulators. Nat. Commun. 8, 1461 (2017).

36 

Z. Y. Cao, J. W. Hu, A. F. Goncharov, X. J. Chen, Nontrivial metallic state of MoS2. Phys. Rev. B 97, 214519 (2018).

37 

Y. Ge, A. Y. Liu, Phonon-mediated superconductivity in electron-doped single-layer MoS2: A first-principles prediction. Phys. Rev. B 87, 241408 (2013).

38 

R. Roldán, E. Cappelluti, F. Guinea, Interactions and superconductivity in heavily doped MoS2. Phys. Rev. B 88, 054515 (2013).

39 

M. Rösner, S. Haas, T. O. Wehling, Phase diagram of electron-doped dichalcogenides. Phys. Rev. B 90, 245105 (2014).

40 

R. Aksoy, X-ray diffraction study of molybdenum disulfide to 38.8 GPa. J. Phys. Chem. Solid. 67, 19141917 (2006).

41 

N. Bandaru, Effect of pressure and temperature on structural stability of MoS2. J. Phys. Chem. C 118, 32303235 (2014).

42 

Y. Zhuang, Pressure-induced permanent metallization with reversible structural transition in molybdenum disulfide. Appl. Phys. Lett. 110, 122103 (2017).

43 

M. Brotons-Gisbert, Optical and electronic properties of 2H-MoS2 under pressure: Revealing the spin-polarized nature of bulk electronic bands. Phys. Rev. Mater. 2, 054602 (2018).

44 

A. F. Goncharov, Structure and stability of 2Ha-MoS2 at high pressure and low temperatures. Phys. Rev. B 102, 064105 (2020).

45 

H. Guo, T. Yang, P. Tao, Y. Wang, Z. Zhang, High pressure effect on structure, electronic structure, and thermoelectric properties of MoS2. J. Appl. Phys. 113, 013709 (2013).

46 

C. Monney, Spontaneous exciton condensation in 1T-TiSe2: BCS-like approach. Phys. Rev. B 79, 045116 (2009).

47 

L. F. Mattheiss, Band structures of transition-metal-dichalcogenide layer compounds. Phys. Rev. B 8, 37193740 (1973).

48 

A. N. Kozlov, L. A. Maksimov, The metal-dielectric divalent crystal phase transition. Sov. Phys. JETP 21, 790795 (1965).

49 

R. S. Knox, Theory of Excitons (Solid State Physics, Academic Press, New York, NY, 1963), vol. suppl. 5.

50 

G. Grüner, Density Waves in Solids (CRC Press, Boca Raton, FL, 2018).

51 

A. N. Kozlov, L. A. Maksimov, Collective excitations in semimetals. Sov. Phys. JETP 22, 889893 (1966).

52 

P. Giannozzi, Quantum ESPRESSO: A modular and open-source software project for quantum simulations of materials. J. Phys. Condens. Matter 21, 395502 (2009).

53 

P. Giannozzi, Advanced capabilities for materials modelling with quantum espresso. J. Phys. Condens. Matter 29, 465901 (2017).

54 

J. P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865 (1996).

55 

D. Hamann, Optimized norm-conserving Vanderbilt pseudopotentials. Phys. Rev. B 88, 085117 (2013).

56 

S. Baroni, S. de Gironcoli, A. Dal Corso, P. Giannozzi, Phonons and related crystal properties from density-functional perturbation theory. Rev. Mod. Phys. 73, 515562 (2001).

57 

M. S. Hybertsen, S. G. Louie, Electron correlation in semiconductors and insulators: Band gaps and quasiparticle energies. Phys. Rev. B 34, 5390 (1986).

58 

G. Strinati, Application of the Green’s functions method to the study of the optical properties of semiconductors. Riv. Nuovo Cimento 11, 186 (1988).

59 

A. Marini, C. Hogan, M. Grüning, D. Varsano, Yambo: An ab initio tool for excited state calculations. Comput. Phys. Commun. 180, 13921403 (2009).

60 

D. Sangalli, Many-body perturbation theory calculations using the Yambo code. J. Phys. Condens. Matter 31, 325902 (2019).

61 

R. W. Godby, R. J. Needs, Metal-insulator transition in Kohn-Sham theory and quasiparticle theory. Phys. Rev. Lett. 62, 11691172 (1989).

62 

F. Bruneval, X. Gonze, Accurate GW self-energies in a plane-wave basis using only a few empty states: Towards large systems. Phys. Rev. B 78, 085125 (2008).

63 

M. Gatti, F. Sottile, Exciton dispersion from first principles. Phys. Rev. B 88, 155113 (2013).

64 

J. A. Soininen, E. L. Shirley, Effects of electron-hole interaction on the dynamic structure factor: Application to nonresonant inelastic x-ray scattering. Phys. Rev. B 61, 1642316429 (2000).

65 

G. L. Bir, G. E. Pikus, Symmetry and Strain-Induced Effects in Semiconductors (Wiley, New York, NY, 1974).

66 

T. G. Pedersen, S. Latini, K. S. Thygesen, H. Mera, B. K. Nikolić, Exciton ionization in multilayer transition-metal dichalcogenides. New J. Phys. 18, 073043 (2016).

67 

J. Zittartz, Anisotropy effects in the excitonic insulator. Phys. Rev. 162, 752758 (1967).

68 

S. Samaneh Ataei, D. Varsano, E. Molinari, M. Rontani, Source custom Fortran code and supporting data for “Evidence of ideal excitonic insulator in bulk MoS2 under pressure.” Zenodo. https://zenodo.org/record/4455373. Deposited 4 November 2020.