Optical signatures of multifold fermions in the chiral topological semimetal CoSi

Edited by John DiTusa, Louisiana State University, Baton Rouge, LA, and accepted by Editorial Board Member Zachary Fisk September 21, 2020 (received for review May 28, 2020)
October 19, 2020
117 (44) 27104-27110

Significance

We present a comprehensive combined experimental and theoretical study of optical conductivity in the chiral topological semimetal CoSi based on the development of high-quality crystals. We reveal the presence and the energy range of various exotic multifold quasiparticles in the optical responses and provide experimental evidence for the realization of fourfold spin-3/2 fermions, which were not directly observed previously. Our work is critical to interpreting future optical and transport responses of multifold fermion materials. We believe the methods used in this work will not only stimulate future research in this class of materials but will also provide a strategy for addressing optical signatures of chiral topological fermions in solids.

Abstract

We report the optical conductivity in high-quality crystals of the chiral topological semimetal CoSi, which hosts exotic quasiparticles known as multifold fermions. We find that the optical response is separated into several distinct regions as a function of frequency, each dominated by different types of quasiparticles. The low-frequency intraband response is captured by a narrow Drude peak from a high-mobility electron pocket of double Weyl quasiparticles, and the temperature dependence of the spectral weight is consistent with its Fermi velocity. By subtracting the low-frequency sharp Drude and phonon peaks at low temperatures, we reveal two intermediate quasilinear interband contributions separated by a kink at 0.2 eV. Using Wannier tight-binding models based on first-principle calculations, we link the optical conductivity above and below 0.2 eV to interband transitions near the double Weyl fermion and a threefold fermion, respectively. We analyze and determine the chemical potential relative to the energy of the threefold fermion, revealing the importance of transitions between a linearly dispersing band and a flat band. More strikingly, below 0.1 eV our data are best explained if spin-orbit coupling is included, suggesting that at these energies, the optical response is governed by transitions between a previously unobserved fourfold spin-3/2 node and a Weyl node. Our comprehensive combined experimental and theoretical study provides a way to resolve different types of multifold fermions in CoSi at different energy. More broadly, our results provide the necessary basis to interpret the burgeoning set of optical and transport experiments in chiral topological semimetals.
Topological semimetals are metals defined by topologically protected degeneracies. In the solid state, their simplest realization features two bands that cross at a single node (13), known as a Weyl node. Weyl nodes are degeneracies not protected by the crystalline symmetry, and the excitations around them behave as spin-1/2 quasiparticles, with a linear relationship between energy and momentum. They are found in a family of noncentrosymmetric transition monopnictides, including TaAs, TaP, NbAs, and NbP (411), as well as in the magnetic compounds Co2MnGa (12) and Co3Sn2S2 (13, 14).
In general, crystal symmetries can protect band crossings with higher degeneracies at high-symmetry points, known as multifold crossings, around which the bands disperse linearly (15, 16). In particular, chiral crystals with nonsymmorphic symmetries that lack inversion and mirror symmetry were predicted to realize a variety of multifold crossings: three-, four-, or sixfold crossings. The dispersion close to each multifold degeneracy is described by a higher-spin quasiparticle, such as a spin-1 fermion in the case of a threefold crossing. Analogous to the case of Weyl nodes, multifold nodes are monopoles of Berry curvature with integer charge. However, the topological charge (i.e., Chern number) of multifold crossings is higher than that of Weyl nodes, which have topological charge ±1. Remarkably, multifold fermions have been shown to exist in the chiral semimetals CoSi, RhSi, PtAl, and PdGa, all of which crystallize in the chiral space group 198 (1723). CoSi and RhSi host a threefold spin-1 fermion at the zone center (the Γ point), which is expected to weakly split by the spin-orbit interaction into a fourfold spin-3/2 and a twofold spin-1/2 Weyl fermion. At the zone boundary (the R point), they host a double spin-1/2 Weyl fermion, which is expected to split into a sixfold spin-1 fermion and a twofold degenerate Kramers pair (17, 18). However, the splitting of the spin 3/2 multifold and the Weyl node was not resolved in the previous photoemission experiments due to the small energy scales that are involved (1921). The threefold nodes in these materials are fundamentally different from the triply degenerate point in achiral materials such as in MoP (24), which have zero Chern number and occur away from time-reversal invariant momenta (2428).
The multifold fermions in chiral topological semimetals are responsible for unusual and interesting optical responses such as gyrotropy (29, 30) and the quantized circular photogalvanic effect (17, 3134). The optical conductivity is a particularly useful tool to probe the multifold node at the zone center since it is expected to dominate the response at low frequencies (35). An accurate measurement of the optical conductivity is also essential to precisely extract nonlinear optical responses such as second harmonic generation (36, 37) and photogalvanic effect precisely (3844). Clarifying the nonlinear responses in these noncentrosymmetric topological semimetals acts to certify the presence and energy range where topological nodes are active (43, 45, 46). Determining the carrier lifetime and energy range at which topological crossings are activated is the key to observing the quantized circularly photogalvanic effect and to using them in the next generation of efficient topological optoelectronics (31).
Linearly dispersing bands result in a characteristic low-frequency optical conductivity given by σ1(ω)ωd2, where d is the spatial dimension (47). Although the expected frequency-independent conductivity was observed in graphite (48) and graphene (49), the observation of a linear ω-dependence characteristic of three-dimensional linear bands (50) is often challenging. The latter has been the subject of various experimental studies: for example, in the Dirac semimetals Cd3As2 (51, 52) and ZrTe5 (5355) and the Weyl semimetal TaAs (56, 57). However, in Cd3As2 the band structure at low energy deviates from linearity, resulting in a sublinear frequency dependence (51). In ZrTe5, the band structure is quite sensitive to the growth method, and the bands can disperse quadratically along one direction, leading to a different frequency dependence of the optical conductivity at low temperature (5355). In the case of TaAs, coexistence of trivial electron and hole pockets and the small energy difference between two kinds of Weyl nodes complicates the data analysis (56, 57). In RhSi, despite promising initial work (43, 58), good agreement between theory and experiment on the linear scaling of the conductivity, which would signal multifold fermions, has remained elusive. Furthermore, in many materials studied previously, the Lifshitz energy (i.e., the energy below which topological quasiparticles are excited) is so low that it could not be resolved (5157), a circumstance that makes these systems not ideal for studying linear optical conductivity, in particular the effect of spin-orbit coupling or correlation on the optical conductivity response.
The cubic chiral crystal CoSi is a promising material to reveal the signature of multifold fermions in an optical conductivity experiment. It has a low carrier density and a large Lifshiz energy of 0.6 eV (18, 19, 21, 59), and the multifold nodes located at the zone center and zone boundary are significantly split by an energy difference 0.2 eV, with no other bands expected at the Fermi level (1921). Below 0.2 eV, interband excitations near the nodes at the zone boundary are expected to be Pauli blocked, leaving only the linearly dispersing multifold fermions at the zone center (35, 60). These properties motivate the choice of CoSi as an ideal candidate to display a clean linear relation between the conductivity and frequency and also, to reveal why and how deviations could occur. Optical conductivity on CoSi was first measured more than two decades ago, but an understanding of neither the Drude response nor the interband excitations was provided (61, 62), most likely because it is a weakly correlated semimetal, as evidenced by a normal metallic Sommerfeld constant (63). Most importantly, the topological properties of CoSi and their implications for the optical response were not addressed since the topology of CoSi was not known until very recently (18, 19, 21, 59).
In the present work, we used Fourier-transform infrared (FTIR) spectroscopy to measure the conductivity over a broad range of 40 to 50,000 cm1, with temperature dependence from 10 to 300 K. Based on improvements of the sample quality, we observed a Drude peak width of 2 meV and low onset energy of interband excitation of 20 meV, which are both around one order of magnitude lower than previous works (61, 62). Due to such improvement, we establish that the optical conductivity in CoSi is determined by the existence of multifold fermions at low frequencies. The temperature-dependent spectral weight of the Drude response is consistent with the Fermi velocity of the electron pocket consisting of the double Weyl quasiparticles. By subtracting a single sharp Drude peak and four narrow phonon peaks from the real part of the conductivity at 10 K, we reveal two approximately linear conductivity regimes separated by a kink at 0.2 eV, culminating in a sharp peak at 0.56 eV. Using density functional theory (DFT) and tight-binding calculations, we link the two linear regimes to multifold excitations close to the Γ (ω<0.2 eV) and R (ω>0.2 eV) points and the peak to a saddle point in the band structure at the M point. Crucially, below 0.2 eV our calculations and measurements reveal a slight deviation from a perfect linear conductivity, consistent with a chemical potential that crosses a nearly flat band of a threefold node at the zone center. This observation, combined with the band splitting due to spin-orbit coupling, suggests that the optical transitions below 0.2 eV involve the spin-3/2 and a spin-1/2 Weyl nodes at the zone center. Our work highlights that the location of the chemical potential with respect to the multifold fermion is critical to interpreting future optical and transport responses of multifold fermion materials, due to the presence of a threefold fermion with an approximately flat band at the zone center.
High-quality single crystals of CoSi have been synthesized with a chemical vapor transport (CVT) method (59) and a flux method (64). Since it was found that the Drude response is significantly sharper for unpolished samples (56, 57), all of the reflectivity measurements on CoSi were performed on as-grown flat shiny facets. The data in Figs. 1, 2, and 4 are all measured on the (001) facet of CoSi grown by CVT. Part of the data measured on the flux-grown sample is shown in Fig. 5. (A complete dataset on the flux-grown sample is shown in SI Appendix, section A.) The facet direction is confirmed by X-ray Laue diffraction and second-harmonic generation measurement (36).
Fig. 1.
(A) Temperature-dependent reflectivity spectra of a (001)-oriented CoSi crystal grown by CVT. Inset shows the reflectivity from 40 to 50,000 cm1 at room temperature. (B) Temperature dependence of the real part of the dielectric function ε1(ω). Inset shows an enlarged view to emphasize the zero crossings of ε1(ω), which correspond to screened plasma frequencies at different temperatures. (C) Temperature-dependent loss function. Inset shows the screened plasma frequency of free carriers obtained from the zero crossings of ε1(ω) and the peak in the loss function as a function of temperature.
Fig. 2.
(A) Temperature-dependent optical conductivity spectra σ1(ω) of a CVT-grown CoSi (001) crystal. (B) Measured optical conductivity at 10 K along with the Wannier tight-binding (TB) calculation with and without spin-orbit coupling (SOC) at 10 K.
Fig. 1A shows the measured temperature-dependent reflectivity spectra R(ω) of CoSi over a wide frequency range. In the low-frequency region (<1,000cm1), R(ω) shows a typical metallic response with a rather sharp plasma edge, below which it rapidly approaches unity. The plasma edge exhibits a rather strong temperature dependence. As the temperature decreases, it shifts continuously toward lower frequency and becomes steeper, indicating reductions in both the carrier density and scattering rate. The low value of the plasma edge (700cm1) generally suggests a very small carrier density, consistent with the tiny Fermi surface in this material (19, 21, 64). In addition, we identify four sharper phonon peaks that are located at about 225, 305, 345, and 420 cm1 and a stronger temperature dependence of the plasma edge than previous works (61, 62).
Fig. 1B shows the real part of the dielectric function ε1(ω). At low frequencies, ε1(ω) is negative (a defining property of a metal) and can be described by the Drude model ε(ω)=εωp2/(ω2+iωγ), where ε is the high-frequency dielectric constant, ωp is the Drude plasma frequency, and γ is the electronic scattering rate. Fig. 1 B, Inset shows an enlarged view to emphasize the zero crossing of ε1(ω) at different temperatures. The zero crossing of ε1(ω) corresponds to the screened plasma frequency ωpscr of free carriers, which is related to the Drude plasma frequency through ωpscr=ωp/ε. Fig. 1C shows the temperature-dependent loss function with the peak around 700 cm1 being the screened plasma frequency. We observe that the temperature dependence of ωpscr in the dielectric function and the loss function agree well with the minimum of the plasma edge of R(ω)700cm1. As shown in Fig. 1 C, Inset, the screened plasma frequency decreases from about 740 cm1 at 300 K to 625 cm1 at 10 K.
Fig. 2A displays the temperature dependence of the real part of the optical conductivity σ1(ω) in the infrared range. In the low-frequency region, the free carrier contribution to σ1(ω) is seen as a Drude-like peak centered at zero frequency. Upon cooling, the Drude peak becomes narrower and loses spectral weight, implying that both the quasiparticle scattering rate and carrier density drop with decreasing temperature. This is consistent with the reflectivity analysis shown in Fig. 1. Along with the tail of the Drude peak, the interband optical conductivity increases approximately linearly with ω (up to about 2,000 cm1 at 300 K). More interestingly, at low temperatures an upturn kink emerges around 1,600 cm1 (0.2 eV) in the spectrum of σ1(ω), resulting in two regions of quasilinear behavior at higher and lower frequencies. The upturn of σ1(ω) indicates that new interband excitations become allowed above 0.2 eV. Fig. 2B shows the σ1(ω) spectrum at 10 K on a larger energy scale up to 1.5 eV. It reveals a sharp peak at 0.56 eV that most likely arises from vertical transitions at the corresponding energy, between bands that disperse nearly parallel in a large region of momentum space.
To determine the origin of the observed peak, we performed first-principles DFT band structure calculations (Materials and Methods). We find the relaxed CoSi lattice constant to be a=4.485 Å, which matches well with previous theoretical and experimental work (21, 60, 63, 65). The calculated electronic band structure is shown in Fig. 3A and is also in agreement with previous reports (18, 60, 65). Fig. 3 A, Lower shows the electronic bands calculated without spin-orbit coupling. As previously reported (18), in the absence of spin-orbit coupling, CoSi hosts a threefold spin-1 fermion at Γ and a double Weyl fermion at the zone boundary R point, which may be viewed as two degenerate spin-1/2 Weyl fermions with equal Chern number. The effect of spin-orbit coupling is to split the threefold node at Γ into a fourfold spin-3/2 fermion and a twofold Weyl fermion, as expected from the addition rules for angular momenta (17, 18). At the R point, the double Weyl node splits into a sixfold multifold fermion, which can be viewed as a double spin-1 multifold fermion, and a twofold degenerate Kramers pair. The energetic splitting of the multifold nodes is a measure of the strength of spin-orbit coupling when compared with the bandwidth; at Γ, we find a spin-orbit splitting of 18.1 meV. Note that the band structure exhibits a saddle point at the M point.
Fig. 3.
(A) Band structure of CoSi with spin-orbit coupling. Lower shows the band structure without spin-orbit coupling in the same shaded energy window of Upper. (B) JDOS as a function of energy. (C) Momentum-resolved matrix element that contributes to the 0.62-eV interband transition along the ΓRM direction.
Furthermore, we calculated the maximally localized Wannier functions that represent the DFT valence Bloch states using Wannier90 (66), and we constructed a tight-binding model in this Wannier function basis by fitting to the DFT band structure. We use this model to calculate the interband contribution to the optical conductivity, defined as
σab(ω)=πe2mnd3k(2π)3fnmEnmvnmavmnbδ(Emnω),
[1]
where fnm(k)=fn(k)fm(k) is the occupation difference between Bloch states with band indices n and m at momentum k, Enm(k)=En(k)Em(k) is the energy difference between these two bands, and vnma(k) is the velocity matrix element along the a direction. With these Wannier orbitals, this model can describe the bands up to 1.5 eV above the Fermi level accurately. As a consequence of cubic symmetry, the conductivity tensor σab(ω) has only one independent component.
The optical conductivity, calculated for the band structure shown in Fig. 3A, is shown in Fig. 2B, in addition to the experimental result at 10 K. For comparison, we present the optical conductivity calculation with and without spin-orbit coupling with a Gaussian broadening of 5 meV, which is close to the Drude peak width at low temperature. (SI Appendix, sections B and C have more details.) The difference between the two calculations is relatively small, as the spin-orbit coupling strength is weak (20 meV) in this material. This suggests that a model without spin-orbit coupling is sufficient to describe the coarse features, broader than this energy scale, while spin-orbit coupling is needed to describe finer structure of the response. As mentioned above, the calculation is restricted to the interband contribution to the conductivity, whereas the experimental measurement also shows the (intraband) Drude response, as well as the phonon contribution in the low-energy regime.
Let us first focus on the measured peak around 0.6 eV, which is only due to interband transitions. The calculated optical conductivity shows a peak at ω0.62 eV, the position of which matches well with the experimentally observed peak. Contributions to the conductivity peak come from all transitions at the peak energy range, which are not Pauli blocked and have nonzero velocity matrix elements.
The joint density of states (JDOS), shown in Fig. 3B for the case of vanishing spin-orbit coupling, is a measure of the number of transitions at a given energy and is thus a significant indicator for the origins of the observed peak. As is clear from Fig. 3B, the JDOS exhibits a shoulder-like feature at 0.62 eV, characteristic of a saddle point, which suggests that the interband contributions primarily originate from the M point. The JDOS is a momentum-integrated quantity; therefore, to unambiguously determine which interband transitions give rise to the peak, we plot the (momentum-resolved) integrand of Eq. 1 in Fig. 3C and the momentum-resolved contribution in SI Appendix, section D. In particular, in Fig. 3C we show the quantity |V(k)|2=nmfnm(k)Emn|vnma(k)|2δ(Emnω) for fixed energy ω=0.62 eV along ΓR and RM, which is a measure of the matrix elements of all allowed transitions at the peak energy. Fig. 3C clearly confirms that the interband transitions giving rise to the peak originate mainly from the vicinity of the M point, which was assigned to a different origin in previous works (43, 58, 60). Note that the calculated peak in Fig. 2B becomes lower and wider with increasing broadening (SI Appendix, section C).
A second feature of the calculated conductivity shown in Fig. 2B is a dip at around 0.25 eV. This is due to the curvature of the middle band in the threefold node at the Γ point. As shown in Fig. 3A, when the energy of the incoming photon is very small, the allowed transitions should occur between the lower band and the middle band at momenta right near Γ. Away from Γ, the middle band curves downward in energy and becomes occupied, thus blocking transitions from the lower band to the middle band and providing the downward dip in the spectrum. As the energy of the incoming photon increases further, the transitions around the R point become activated, providing the recovery from the dip and the continuation of the spectrum upward. However, this dip is not observed in experiments, probably due to a short lifetime of the hot carriers around 0.25 eV as decreasing the quasiparticle lifetime by broadening the Dirac delta function smears out this dip feature (60) (SI Appendix, section C).
Finally, we focus on the optical conductivity below 0.2 eV. In order to compare the calculation results with experiment below 0.2 eV accurately, we need to extract the interband contribution from the experimental data by subtracting the sharp Drude and phonon responses at low energy. In order to quantify the temperature dependence of the Drude response, in Fig. 4A we fit the low-energy part of σ1(ω) using the Drude model. Fig. 4 A, Inset shows the dc resistivity ρ1/σ1(ω=0), derived from the fitted zero-frequency value (open red circles), which agrees well with the transport data (solid black curve), indicating that the subtraction of the Drude response is reliable. As shown in SI Appendix, section A, the linear Hall resistivity is dominated by a high-mobility electron channel, which indicates that the narrow Drude response comes from the electron pocket with double Weyl quasiparticles at R. The corresponding free carrier weight, SDrude=π2Z0ωp2, where Z0 is the vacuum impedance, and the scattering rate, γ, are displayed in Fig. 4 B and C, respectively. The free carrier weight shows a T2 temperature dependence as expected for massless Dirac electrons (67). The coefficient before the T2 temperature dependence could be used to extract the Fermi velocity of the Dirac fermions (double Weyl fermions in this case) by the relation ωp(T)2=ωp02+2π9vFRe23(kBT)2 (67, 68). We estimated that vFR for the double Weyl fermions is around 1.4 ×105 m/s, which agrees with the DFT calculation in Fig. 3 A, Lower within 4%. The scattering rate decreases as the temperature is lowered, following a T2 temperature dependence at low temperatures, as shown by the red solid curve through the data points in Fig. 4C. Moreover, the value of γ becomes extremely small at low temperatures, which is at least one order of magnitude narrower than the value reported in RhSi (43, 46, 58) and in previous studies of CoSi (61, 62).
Fig. 4.
(A) Optical conductivity of CoSi at 10 K (blue); thin solid lines represent fit contributions of the Drude term (green) and phonon modes (gray), as well as the remaining interband contribution (subtracting Drude and phonons) to σ1(ω) (red). The black dashed line indicates the linear conductivity below 0.2 eV. (Inset) Comparison of the dc resistivity, ρab (solid black line), with the zero-frequency values of the Drude fits to the conductivity data, 1/σ1(ω=0) (red circles). Temperature dependences of (B) the Drude weight and (C) the electronic scattering rate.
The most striking feature of the optical conductivity is its approximately linear behavior up to 0.2 eV, obtained after subtracting the single sharp Drude peak and four narrow phonons, as indicated by the black dashed line in Fig. 4A. After Drude subtraction, the optical conductivity derives from vertical interband transitions, and in the case of CoSi we expect that, at low frequencies, these transitions occur between bands associated with the multifold fermions. Below 0.2 eV, vertical transitions at the R point are Pauli blocked, as is clear from Fig. 3A, except for the tiny 11-meV peak shown in Fig. 2B, which is associated with the interband transitions between spin-orbit split bands along the RM line (SI Appendix, section E has more discussion). The contribution from these transitions is much smaller, however, than the vertical transitions from the multifold fermions at the Γ point (35). Therefore, the linear conductivity below 0.2 eV is mainly attributed to these interband transitions near the Γ point.
For multifold fermions with linear band dispersion, the optical conductivity at low frequencies (where the linear approximation holds) was shown to have a linear frequency dependence (35), which can be understood qualitatively from dimensional analysis, σ1(ω)ωd2. In the isotropic limit of the dispersion (i.e., when full rotation symmetry is present), the slope of the conductivity of a single multifold fermion only depends on the Fermi velocity vFΓ and is inversely proportional to it (35).
To quantitatively understand the linear slope of the observed conductivity and to determine whether it originates from the low-energy multifold fermions, we now analyze the low-frequency regime of the optical conductivity in detail. As a first step, we consider a kp model for the threefold spin-1 fermion node shown in Fig. 3 A, Lower in the absence of spin-orbit coupling. Interband transitions are allowed from the partially occupied lower linear band to the central flat band but are forbidden between the two linearly dispersing bands due to angular momentum selection rules (35). For this model, the analytic formula for the optical conductivity at T0 K is e2ω3πvFΓθ(ωE0) (35), where E0 is the energy of the threefold node measured from the Fermi energy. Using a Fermi velocity vFΓ = 1.9 ×105 m/s from a fit to the linear band in Fig. 3A (details of the fit are included in SI Appendix, section F), we plot the analytical conductivity in SI Appendix, Fig. S9 (orange curve) in SI Appendix, section F, together with the experimental result. The linear model falls lower than the experimental curves for both samples, failing to capture the shoulder-like features around 50 meV in the data, mainly due to the absence of quadratic corrections to the flat band.
To better understand the origin of the discrepancies, we go beyond the linear model and compare the optical conductivity data with that obtained by the Wannier tight-binding model based on DFT introduced above. The lower-energy regime of optical conductivity in Fig. 2B of the Wannier tight-binding models is shown in Fig. 5. Overall, this analysis establishes that the chemical potential lies below the threefold node at Γ and highlights that deviations from a linear band structure near multifold fermions have an important qualitative impact on the optical conductivity. A simpler four-band tight-binding model for CoSi (17, 32, 35) that captures the symmetries of space group 198 with three parameters that we fit to the first-principles band structure in Fig. 3A reaches the same conclusions (SI Appendix, section G).
Fig. 5.
Interband optical conductivity in CoSi from measurement in both CVT- and flux-grown samples (red and black curves, respectively) along with Wannier tight-binding (TB) calculations without and with spin-orbit coupling (SOC).
Next, we assess the role of spin-orbit coupling at low frequencies. The results of DFT-based Wannier tight-binding model (with spin-orbit coupling) are shown as the green curve in Fig. 5. We subtracted the small peak feature at 11 meV shown in Fig. 2B since it originates from the interband excitations between spin-orbit split bands along the RM line (SI Appendix, section E). This calculation displays close agreement with the measured conductivity in the flux-grown sample below 0.1 eV. This agreement supports that the low-frequency shoulder below 0.1 eV observed in experiments arises from transitions between the spin-orbit split linear and flat threefold fermion bands, which are composed of a spin-3/2 fourfold node and a spin-1/2 Weyl node.
Finally, we discuss the distinctions between the two crystals studied. Our calculated conductivity agrees well with the interband conductivity data from the flux-grown CoSi sample, while the conductivity data from the CVT-grown sample are higher at low frequencies (by 80 Ω1cm1 below 0.1 eV). It turns out that this difference is not due to the observation that the carrier density in the CVT-grown sample is 10 % higher than the flux-grown sample by comparing the Drude spectral weight. As shown in SI Appendix, sections G and H, assuming a higher chemical potential does not reach as good agreement as the flux-grown sample. This is an indication that either the additional contribution in the CVT-grown sample is not coming from interband transition (46, 58) or it is related to surface Fermi arcs as (001) (CVT) sample has surface states but (111) (flux) sample does not (69). Nevertheless, the average slopes of both samples and the calculations are comparable after shifting down the experimental result on the CVT sample by 80 Ω1cm1 (Fig. 5), suggesting that there is agreement concerning the Fermi velocity (SI Appendix, section H has more discussions).
In summary, we revealed the topological origin of the optical conductivity in CoSi based on the development of high-quality crystals. Our analysis shows that locating the chemical potential accurately with respect to the central flat band is crucial to understanding the optical response. We have shown that both intraband and interband optical conductivities in different frequency regimes are dominated by the existence of different topological multifold fermions, as well as a band structure saddle point. Most notably, we provide evidence of the existence of the fourfold spin-3/2 quasiparticles. Our results not only shed light on the interpretation of circular photogalvanic measurement in this family of materials (17, 31, 43, 45, 46) but also pave the way to study optical signatures in other chiral topological semimetals in the future (16, 34).

Materials and Methods

Millimeter-sized single crystals of CoSi have been synthesized with both a CVT method (59) and a flux method (64). Temperature-dependent infrared reflectance measurements were performed on as-grown flat shiny facets using a Bruker VERTEX 70v FTIR spectrometer with an in situ gold overfilling technique (70) (SI Appendix, section A has details). We performed first-principles DFT band structure calculations with the Quantum Espresso code (7173). We used norm-conserving pseudopotentials generated with the OPIUM code (74), the Perdew–Burke–Ernzerhof generalized gradient approximation functional (75), and spin-orbit coupling to relax the structure and to calculate the band structures.

Data Availability

All study data are included in the article and SI Appendix.

Acknowledgments

We thank F. de Juan, C. L. Kane, and Y. Zhang for helpful discussion. Z.N. and L.W. are supported by Army Research Office Grant W911NF1910342. B.X. and C.B. are supported by Schweizerische Nationalfonds Grant 200020-172611. Z.F., J.W.F.V., and E.J.M. are supported by the US National Science Foundation under grant DMR-1720530 (Materials Research Science and Engineering Center program). M.-Á.S.-M. acknowledges support from the European Union’s Horizon 2020 Research and Innovation Programme under Marie Sklodowska-Curie Grant 754303 and the GreQuE Cofund Programme. K.M. and C.F. acknowledge financial support from European Research Council Advanced Grant 742068 TOP-MAT and Deutsche Forschungsgemeinschaft Projects 258499086 and FE 63330-1. K.W. and J.P. are supported by the Gordon and Betty Moore Foundation’s EPiQS Initiative Grant GBMF9071 and the Maryland Quantum Materials Center. A.G.G. is supported by ANR Grant ANR-18-CE30-0001-01 (TOPODRIVE) and the European Union’s Horizon 2020 Research and Innovation Programme Grant 829044 (SCHINES). A.M.R. was supported by Department of Energy, Office of Basic Energy Sciences Grant DE-FG02-07ER46431. Computational support was provided by the National Energy Research Scientific Computing Center.

Supporting Information

Appendix (PDF)

References

1
S. Murakami, Phase transition between the quantum spin Hall and insulator phases in 3D: Emergence of a topological gapless phase. New J. Phys. 9, 356 (2007).
2
X. Wan, A. M. Turner, A. Vishwanath, S. Y. Savrasov, Topological semimetal and Fermi-arc surface states in the electronic structure of pyrochlore iridates. Phys. Rev. B 83, 205101 (2011).
3
A. Burkov, L. Balents, Weyl semimetal in a topological insulator multilayer. Phys. Rev. Lett. 107, 127205 (2011).
4
H. Weng, C. Fang, Z. Fang, B. A. Bernevig, X. Dai, Weyl semimetal phase in noncentrosymmetric transition-metal monophosphides. Phys. Rev. X 5, 011029 (2015).
5
S. M. Huang et al., A Weyl fermion semimetal with surface Fermi arcs in the transition metal monopnictide TaAs class. Nat. Commun. 6, 7373 (2015).
6
S. Y. Xu et al., Discovery of a Weyl fermion semimetal and topological Fermi arcs. Science 349, 613–617 (2015).
7
B. Lv et al., Experimental discovery of Weyl semimetal TaAs. Phys. Rev. X 5, 031013 (2015).
8
B. Lv et al., Observation of Weyl nodes in TaAs. Nat. Phys. 11, 724–727 (2015).
9
S. Y. Xu et al., Discovery of a Weyl fermion state with Fermi arcs in niobium arsenide. Nat. Phys. 11, 748–754 (2015).
10
L. Yang et al., Weyl semimetal phase in the non-centrosymmetric compound TaAs. Nat. Phys. 11, 728–732 (2015).
11
N. Xu et al., Observation of Weyl nodes and Fermi arcs in tantalum phosphide. Nat. Commun. 7, 11006 (2016).
12
I. Belopolski et al., Discovery of topological Weyl fermion lines and drumhead surface states in a room temperature magnet. Science 365, 1278–1281 (2019).
13
D. Liu et al., Magnetic Weyl semimetal phase in a Kagomé crystal. Science 365, 1282–1285 (2019).
14
N. Morali et al., Fermi-arc diversity on surface terminations of the magnetic Weyl semimetal Co3Sn2S2. arXiv:1903.00509 (1 March 2019).
15
B. J. Wieder, Y. Kim, A. Rappe, C. Kane, Double Dirac semimetals in three dimensions. Phys. Rev. Lett. 116, 186402 (2016).
16
B. Bradlyn et al., Beyond Dirac and Weyl fermions: Unconventional quasiparticles in conventional crystals. Science 353, aaf5037 (2016).
17
G. Chang et al., Unconventional chiral fermions and large topological Fermi arcs in RhSi. Phys. Rev. Lett. 119, 206401 (2017).
18
P. Tang, Q. Zhou, S. C. Zhang, Multiple types of topological fermions in transition metal silicides. Phys. Rev. Lett. 119, 206402 (2017).
19
Z. Rao et al., Observation of unconventional chiral fermions with long Fermi arcs in CoSi. Nature 567, 496–499 (2019).
20
D. S. Sanchez et al., Topological chiral crystals with helicoid-arc quantum states. Nature 567, 500–505 (2019).
21
D. Takane et al., Observation of chiral fermions with a large topological charge and associated Fermi-arc surface states in CoSi. Phys. Rev. Lett. 122, 076402 (2019).
22
N. B. Schröter et al., Chiral topological semimetal with multifold band crossings and long Fermi arcs. Nat. Phys., 15, 759–765 (2019).
23
N. Schröter et al., Observation and control of maximal Chern numbers in a chiral topological semimetal. arXiv:1907.08723 (19 July 2019).
24
B. Lv et al., Observation of three-component fermions in the topological semimetal molybdenum phosphide. Nature 546, 627–631 (2017).
25
H. Weng, C. Fang, Z. Fang, X. Dai, Topological semimetals with triply degenerate nodal points in θ-phase tantalum nitride. Phys. Rev. B 93, 241202 (2016).
26
H. Weng, C. Fang, Z. Fang, X. Dai, Coexistence of Weyl fermion and massless triply degenerate nodal points. Phys. Rev. B 94, 165201 (2016).
27
Z. Zhu, G. W. Winkler, Q. Wu, J. Li, A. A. Soluyanov, Triple point topological metals. Phys. Rev. X 6, 031003 (2016).
28
G. Chang et al., Nexus fermions in topological symmorphic crystalline metals. Sci. Rep. 7, 1688 (2017).
29
J. Ma, D. Pesin, Chiral magnetic effect and natural optical activity in metals with or without Weyl points. Phys. Rev. B 92, 235205 (2015).
30
S. Zhong, J. E. Moore, I. Souza, Gyrotropic magnetic effect and the magnetic moment on the Fermi surface. Phys. Rev. Lett. 116, 077201 (2016).
31
F. de Juan, A. G. Grushin, T. Morimoto, J. E. Moore, Quantized circular photogalvanic effect in Weyl semimetals. Nat. Commun. 8, 15995 (2017).
32
F. Flicker et al., Chiral optical response of multifold fermions. Phys. Rev. B 98, 155145 (2018).
33
F. de Juan et al., Difference frequency generation in topological semimetals. Phys. Rev. Res. 2, 012017 (2020).
34
G. Chang et al., Topological quantum properties of chiral crystals. Nat. Mater. 17, 978–985 (2018).
35
M. Á. Sánchez-Martínez, F. de Juan, A. G. Grushin, Linear optical conductivity of chiral multifold fermions. Phys. Rev. B 99, 155145 (2019).
36
L. Wu et al., Giant anisotropic nonlinear optical response in transition metal monopnictide Weyl semimetals. Nat. Phys. 13, 350–355 (2017).
37
S. Patankar et al., Resonance-enhanced optical nonlinearity in the Weyl semimetal TaAs. Phys. Rev. B 98, 165113 (2018).
38
Q. Ma et al., Direct optical detection of Weyl fermion chirality in a topological semimetal. Nat. Phys. 13, 842–847 (2017).
39
G. B. Osterhoudt et al., Colossal mid-infrared bulk photovoltaic effect in a type-I Weyl semimetal. Nat. Mater. 18, 471 (2019).
40
Z. Ji et al., Spatially dispersive circular photogalvanic effect in a Weyl semimetal. Nat. Mater. 1, 955–962 (2019).
41
J. Ma et al., Nonlinear photoresponse of type-II Weyl semimetals. Nat. Mater. 18, 476–481 (2019).
42
N. Sirica et al., Tracking ultrafast photocurrents in the Weyl semimetal TaAs using THz emission spectroscopy. Phys. Rev. Lett. 122, 197401 (2019).
43
D. Rees et al., Helicity-dependent photocurrents in the chiral Weyl semimetal RhSi. Sci. Adv. 6, eaba0509 (2020).
44
Y. Gao et al., Chiral terahertz wave emission from the Weyl semimetal TaAs. Nat. Commun. 11, 720 (2020).
45
Z. Ni et al., Giant topological longitudinal circular photo-galvanic effect in the chiral multifold semimetal CoSi. arXiv:2006.09612 (17 June 2020).
46
Z. Ni et al., Linear and nonlinear optical responses in the chiral multifold semimetal RhSi. arXiv:2005.13473 (27 May 2020).
47
Á. Bácsi, A. Virosztek, Low-frequency optical conductivity in graphene and in other scale-invariant two-band systems. Phys. Rev. B 87, 125425 (2013).
48
A. Kuzmenko, E. Van Heumen, F. Carbone, D. Van Der Marel, Universal optical conductance of graphite. Phys. Rev. Lett. 100, 117401 (2008).
49
R. R. Nair et al., Fine structure constant defines visual transparency of graphene. Science 320, 1308 (2008).
50
P. Hosur, S. Parameswaran, A. Vishwanath, Charge transport in Weyl semimetals. Phys. Rev. Lett. 108, 046602 (2012).
51
A. Akrap et al., Magneto-optical signature of massless Kane electrons in Cd3As2. Phys. Rev. Lett. 117, 136401 (2016).
52
D. Neubauer et al., Interband optical conductivity of the [001]-oriented Dirac semimetal Cd3As2. Phys. Rev. B 93, 121202 (2016).
53
R. Chen et al., Optical spectroscopy study of the three-dimensional Dirac semimetal ZrTe5. Phys. Rev. B 92, 075107 (2015).
54
B. Xu et al., Temperature-driven topological phase transition and intermediate Dirac semimetal phase in ZrTe5. Phys. Rev. Lett. 121, 187401 (2018).
55
E. Martino et al., Two-dimensional conical dispersion in ZrTe5 evidenced by optical spectroscopy. Phys. Rev. Lett. 122, 217402 (2019).
56
B. Xu et al., Optical spectroscopy of the Weyl semimetal TaAs. Phys. Rev. B 93, 121110 (2016).
57
S.-i. Kimura et al., Optical signature of Weyl electronic structures in tantalum pnictides TaPn (Pn= P, As). Phys. Rev. B 96, 075119 (2017).
58
L. Maulana et al., Optical conductivity of multifold fermions: The case of RhSi. Phys. Rev. Res. 2, 023018 (2020).
59
D. S. Sanchez et al., Topological chiral crystals with helicoid-arc quantum states. Nature 567, 500–505 (2019).
60
T. Habe, Dynamical conductivity in the multiply degenerate point-nodal semimetal CoSi. Phys. Rev. B 100, 245131 (2019).
61
D. van der Marel, A. Damascelli, K. Schulte, A. Menovsky, Spin, charge, and bonding in transition metal mono-silicides. Phys. B Condens. Matter 244, 138–147 (1998).
62
F. P. Mena et al., Suppressed reflectivity due to spin-controlled localization in a magnetic semiconductor. Phys. Rev. B 73, 085205 (2006).
63
A. E. Petrova et al., Elastic, thermodynamic, and electronic properties of MnSi, FeSi, and CoSi. Phys. Rev. B 82, 155124 (2010).
64
X. Xu et al., Crystal growth and quantum oscillations in the topological chiral semimetal CoSi. Phys. Rev. B 100, 045104 (2019).
65
D. Pshenay-Severin, Y. V. Ivanov, A. Burkov, A. Burkov, Band structure and unconventional electronic topology of CoSi. J. Phys. Condens. Matter 30, 135501 (2018).
66
A. A. Mostofi et al., An updated version of wannier90: A tool for obtaining maximally-localised Wannier functions. Comput. Phys. Commun. 185, 2309–2310 (2014).
67
P. E. C. Ashby, J. P. Carbotte, Chiral anomaly and optical absorption in Weyl semimetals. Phys. Rev. B 89, 245121 (2014).
68
K. Wang et al., Unconventional free charge in the correlated semimetal Nd2Ir2O7. Nat. Phys., (2020).
69
G. Chang et al., Unconventional photocurrents from surface Fermi arcs in topological chiral semimetals. Phys. Rev. Lett. 124, 166404 (2020).
70
C. C. Homes, M. Reedyk, D. A. Cradles, T. Timusk, Technique for measuring the reflectance of irregular, submillimeter-sized samples. Appl. Opt. 32, 2976–2983 (1993).
71
NJ. Ramer, A. M. Rappe, Designed nonlocal pseudopotentials for enhanced transferability. Phys. Rev. B 59, 12471–12478 (1999).
72
A. M. Rappe, K. M. Rabe, E. Kaxiras, J. D. Joannopoulos, Optimized pseudopotentials. Phys. Rev. B 41, 1227–1230 (1990).
73
P. Giannozzi et al., QUANTUM ESPRESSO: A modular and open-source software project for quantum simulations of materials. J. Phys. Condens. Matter 21, 395502 (2009).
74
Opium - pseudopotential generation project, Version 4.1. http://opium.sourceforge.net. Accessed 5 December 2019.
75
J. P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865–3868 (1996).

Information & Authors

Information

Published in

Go to Proceedings of the National Academy of Sciences
Proceedings of the National Academy of Sciences
Vol. 117 | No. 44
November 3, 2020
PubMed: 33077590

Classifications

Data Availability

All study data are included in the article and SI Appendix.

Submission history

Published online: October 19, 2020
Published in issue: November 3, 2020

Keywords

  1. topological semimetal
  2. optical conductivity
  3. multifold fermions

Acknowledgments

We thank F. de Juan, C. L. Kane, and Y. Zhang for helpful discussion. Z.N. and L.W. are supported by Army Research Office Grant W911NF1910342. B.X. and C.B. are supported by Schweizerische Nationalfonds Grant 200020-172611. Z.F., J.W.F.V., and E.J.M. are supported by the US National Science Foundation under grant DMR-1720530 (Materials Research Science and Engineering Center program). M.-Á.S.-M. acknowledges support from the European Union’s Horizon 2020 Research and Innovation Programme under Marie Sklodowska-Curie Grant 754303 and the GreQuE Cofund Programme. K.M. and C.F. acknowledge financial support from European Research Council Advanced Grant 742068 TOP-MAT and Deutsche Forschungsgemeinschaft Projects 258499086 and FE 63330-1. K.W. and J.P. are supported by the Gordon and Betty Moore Foundation’s EPiQS Initiative Grant GBMF9071 and the Maryland Quantum Materials Center. A.G.G. is supported by ANR Grant ANR-18-CE30-0001-01 (TOPODRIVE) and the European Union’s Horizon 2020 Research and Innovation Programme Grant 829044 (SCHINES). A.M.R. was supported by Department of Energy, Office of Basic Energy Sciences Grant DE-FG02-07ER46431. Computational support was provided by the National Energy Research Scientific Computing Center.

Notes

This article is a PNAS Direct Submission. J.D. is a guest editor invited by the Editorial Board.

Authors

Affiliations

Fribourg Center for Nanomaterials, Department of Physics, University of Fribourg, CH-1700 Fribourg, Switzerland;
Department of Chemistry, University of Pennsylvania, Philadelphia, PA 19104-6323;
Miguel-Ángel Sánchez-Martínez1 https://orcid.org/0000-0003-4907-7466
Institut Néel, CNRS and Université Grenoble Alpes, 38042 Grenoble, France;
Jorn W. F. Venderbos
Department of Physics and Astronomy, University of Pennsylvania, Philadelphia, PA 19104;
Department of Physics, Drexel University, Philadelphia, PA 19104;
Department of Materials Science and Engineering, Drexel University, Philadelphia, PA 19104;
Zhuoliang Ni
Department of Physics and Astronomy, University of Pennsylvania, Philadelphia, PA 19104;
Tian Qiu
Department of Chemistry, University of Pennsylvania, Philadelphia, PA 19104-6323;
Max Planck Institut fur Chemische Physik fester Stoffe, 01187 Dresden, Germany;
Kefeng Wang
Maryland Quantum Materials Center, Department of Physics, University of Maryland, College Park, MD 20742;
Johnpierre Paglione
Maryland Quantum Materials Center, Department of Physics, University of Maryland, College Park, MD 20742;
Canadian Institute for Advanced Research, Toronto, Ontario M5G 1Z8, Canada
Fribourg Center for Nanomaterials, Department of Physics, University of Fribourg, CH-1700 Fribourg, Switzerland;
Max Planck Institut fur Chemische Physik fester Stoffe, 01187 Dresden, Germany;
Canadian Institute for Advanced Research, Toronto, Ontario M5G 1Z8, Canada
Eugene J. Mele
Department of Physics and Astronomy, University of Pennsylvania, Philadelphia, PA 19104;
Institut Néel, CNRS and Université Grenoble Alpes, 38042 Grenoble, France;
Department of Chemistry, University of Pennsylvania, Philadelphia, PA 19104-6323;
Department of Physics and Astronomy, University of Pennsylvania, Philadelphia, PA 19104;

Notes

2
To whom correspondence may be addressed. Email: [email protected].
Author contributions: L.W. designed research; B.X. and C.B. performed experiments; B.X. and L.W. analyzed data; Z.F., M.-Á.S.-M., J.W.F.V., T.Q., E.J.M., A.G.G., and A.M.R. performed calculation; Z.N. characterized sample orientation; K.M., K.W., J.P., and C.F. grew samples; and B.X., Z.F., J.W.F.V., A.G.G., and L.W. wrote the paper.
1
B.X., Z.F., and M.-Á.S.-M. contributed equally to this work.

Competing Interests

The authors declare no competing interest.

Metrics & Citations

Metrics

Note: The article usage is presented with a three- to four-day delay and will update daily once available. Due to ths delay, usage data will not appear immediately following publication. Citation information is sourced from Crossref Cited-by service.


Citation statements




Altmetrics

Citations

Export the article citation data by selecting a format from the list below and clicking Export.

Cited by

    Loading...

    View Options

    View options

    PDF format

    Download this article as a PDF file

    DOWNLOAD PDF

    Login options

    Check if you have access through your login credentials or your institution to get full access on this article.

    Personal login Institutional Login

    Recommend to a librarian

    Recommend PNAS to a Librarian

    Purchase options

    Purchase this article to access the full text.

    Single Article Purchase

    Optical signatures of multifold fermions in the chiral topological semimetal CoSi
    Proceedings of the National Academy of Sciences
    • Vol. 117
    • No. 44
    • pp. 27063-27746

    Media

    Figures

    Tables

    Other

    Share

    Share

    Share article link

    Share on social media