In condensed systems, Weyl quasiparticles have a massless nature and exhibit various relativistic chiral phenomena such as Klein tunneling, chiral anomaly, and Fermi arc surface states. However, in photonic systems, Weyl points (WPs) are protected by the D2d symmetry, often leading to multiple chiral WPs at the same energy level, which makes generating chirality-related effects challenging. To overcome this hinderance, a perturbation that breaks mirror symmetry in the metallic saddle structure was introduced. This perturbation effectively separates the energies of distinct chiral WPs, enabling the experimental measurement of the spectral intensity for each Weyl band and the assessment of chirality imbalance among the WPs. By maintaining time-reversal symmetry, the present study offers an approach for investigating the imbalance in the chirality of pseudo-fermionic fields in photonic materials.

Topological materials have garnered significant attention in the field of condensed matter systems in recent years. Particularly, topological insulators and topological Weyl semimetals have emerged as highly intriguing topics, which makes traditional definitions and interpretations of metals and insulators challenging. One crucial discovery is the dispersion relation observed in the three-dimensional energy band structure of certain materials. This dispersion relation arises from the linear crossing of two energy bands and satisfies the relativistic Weyl equation near the crossing point.1 Such a band degenerate point is called a Weyl point (WP), and its dispersion relation exhibits the quasiparticle excitation characteristics of chiral Weyl fermions.2,3 Additionally, Dirac semimetals4,5 and nodal chain semimetals6–8 exist according to the distribution characteristics of the cross-point. In the surface energy band structure of the Weyl material, the electron dispersion relation is expressed as a segment of open arcs connecting the projection positions of different chiral WPs, also known as Fermi arc surface states.9,10 Any point on the arc is topologically nontrivial and protected topologically, and they have stable and low-loss transport properties such as negative magnetoresistance, giant magnetoresistance effects, chiral anomalies, and very high carrier mobility, which are of great interest for future low-energy lossless devices or photonic devices.11–13 

Analogous to the topological energy band structures in electronic systems,14,15 the degeneracy features are extended to photonic16–19 and phononic systems,20,21 opening up research directions in topological photonics and topological acoustics. In 2018, the University of Birmingham introduced an electromagnetic metamaterial, namely, a metallic saddle-atom structure,22,23 comprising a pair of resonant rings with opposite opening directions. Within the energy bandgap, this structure features four WPs of the same frequency, referred to as ideal WPs. Each WP carries an integer “charge,” known as the Chern number, corresponding to its chirality. Similar to magnetic monopoles, WPs exist only in pairs with opposite charges. Isolated ideal WPs exhibit absolute topological stability, meaning that any perturbation will solely affect the position of the cross-point in momentum-energy space. WP degeneracy can be eliminated only by bringing different chiral WPs together at the same position, leading to their cancelation.

This Letter differs from previous research, as it focuses on examining the effect of mirror symmetry-breaking in ideal Weyl materials. In particular, we investigated the effect of introducing a small deformation to saddle-shaped metallic atoms within electromagnetic crystals. This deformation is applied to break the mirror symmetry in the xz and yz directions while maintaining D2d symmetry. The frequency of the WPs produces a chirality-dependent separation, leading to separation of the density of states (DOS) in the Weyl energy band. The DOS can be used in photonic systems to describe the degree of transport intensity in the material. The results show that D O S 0 = D O S + D O S 0 (± representing the different chirality of WPs), a phenomenon referred to as chirality imbalance in Weyl semimetals. The distribution of DOS curves, illustrating the chiral separation, is determined through theoretical calculations. To experimentally validate our theoretical findings, we applied fast Fourier transform (FFT) to the field diagram, revealing a strong agreement between the theory and the experiment. Additionally, we conducted experimental examinations of the Fermi arc surface states on the ⟨001⟩ crystal plane. Our findings provide evidence of chirality imbalance in photonic systems.24–27 This imbalance opens up an avenue for studying various effects of chiral physics, including Klein tunneling, chiral responses, and others. The results present an approach to investigate the properties of chiral materials in photonic Weyl materials, offering a perspective on the study of chiral materials in photonic Weyl systems.

Protected by time-reversal symmetry and D2d symmetry [Fig. 1(a)], this metallic saddle-shaped structure possesses two pairs of distinct chiral Weyl points. The WPs are distributed along the four diagonal directions of the square lattice, possessing the same frequency [as shown by the purple plane in Fig. 1(a)]. These four WPs exhibit an identical DOS distribution, and their degeneracy originates from the energy band crossing between the transverse and longitudinal modes in the Weyl metamaterial. Their stability is safeguarded by the second-order rotational symmetry along the diagonal directions within the D2d operation.

FIG. 1.

Chirality imbalance resulting from a broken mirror symmetry. (a) Schematic representation of a saddle-shaped metallic coil with a mirror symmetry. The projection of the sample in the Z direction reveals a square shape. The four Weyl points have equal frequencies, and the density of states is the same for WPs with different chirality. (b) Sample with a broken mirror symmetry, which undergoes compression along one of the diagonals, resulting in a diamond structure upon projection in the z direction. The WPs energetically separate, and the density of states exhibits a distinct segregation based on chirality.

FIG. 1.

Chirality imbalance resulting from a broken mirror symmetry. (a) Schematic representation of a saddle-shaped metallic coil with a mirror symmetry. The projection of the sample in the Z direction reveals a square shape. The four Weyl points have equal frequencies, and the density of states is the same for WPs with different chirality. (b) Sample with a broken mirror symmetry, which undergoes compression along one of the diagonals, resulting in a diamond structure upon projection in the z direction. The WPs energetically separate, and the density of states exhibits a distinct segregation based on chirality.

Close modal

The eigenvalues of the second-order rotational operators for the electromagnetic longitudinal and transverse modes are +1 and −1, respectively. The observed frequency agreement among WPs of the same chirality is a direct consequence of the C2 symmetry and time-reversal symmetry along the z-direction. Furthermore, the identical frequency characteristics exhibited by different chiral WPs are preserved by the mirror symmetry about the xz and yz planes within the D2d operation. To disrupt the identical frequency properties of the WPs, a structural deformation was introduced, as depicted in Fig. 1(b). This involves modifying the distances, l1 and l2, between the columns aligned along the diagonal and antidiagonal directions, respectively. Particularly, the length l2 of the square structure was reduced in the antidiagonal direction (from Γ M ). When viewed from the top, the structure undergoes degeneration thereby transforming the structure into a diamond-shaped configuration. This structural transformation breaks the mirror symmetries Mx and My within the structure, leading to WPs with a chiral frequency separation [represented by the blue and red planes below Fig. 1(b)]. Consequently, a separation exists in the DOS. The extent of the structural deformation can be quantified from the θ value, which characterizes the degree of breaking. The magnitude of θ directly influences the difference in the frequency between distinct chiral WPs and affects the bandgap width.28 

We also investigated the energy band structure of the superconformal material after deformation. We designed samples with the following parameters in CST STUDIO SUITE: ax = ay=1.7, hc = 1.524, and h0 = 0.508 mm, width of the metal connecting plate w=0.325, rout = 0.3, and rin = 0.2 mm, length of the two diagonals l1 = 0.9, l2 = 0.84, and θ = 86°. To obtain the experimental results and simplify the calculation, the metal structure inside the unit cell was set as a loss-free perfect electric conductor (PEC) and the rest is set as printed circuit board (PCB) with a dielectric constant of ε = 3.3. The energy band structure of the highly symmetric path in the first Brillouin zone can be obtained by using the eigenmode solver of CST [Fig. 2(a)], where kx-ky-kz is the Bloch wave vector in three directions (the data were normalized to π/axπ/ayπ/az), and two different chiral WPs appear in the diagonal and antidiagonal directions in red dot and blue dot. The positions of the two external points are f1 = 21.4 GHz, −kx = ky = 0.467 and f2 = 21.8 GHz, kx = ky = 0.467. As for the frequency of the WPs, an evident blue shift was observed along the diagonal of the longer side, and a red shift along the diagonal of the shorter side. This phenomenon arises because the symmetry operation solely disrupts the mirror symmetry in the x-y direction, while preserving the rotational symmetries of C2 (along the z-direction) and 2 C 2 C (along the Γ and Γ′ directions) within the D2d operation in the unit cell. These results confirmed the presence of the WPs.

FIG. 2.

Band profile and iso-frequency in a saddle-shaped metallic structure. (a) Simulated band structure in CST along the high-symmetry line. The right side of the pictures are partial enlarged drawing of the Weyl cone. (b) The state density curve of Weyl cones derived from Eq. (1). The red and blue lines represent the DOS of WP1 (+) and WP2 (−), respectively. (c) Band structure in the first Brillouin zones. (c) and (d) Iso-frequency map at frequencies of 21 GHz (less than that of WP2) and 22 GHz (higher than that of WP1), wherein the circle represents the size of the Weyl cone.

FIG. 2.

Band profile and iso-frequency in a saddle-shaped metallic structure. (a) Simulated band structure in CST along the high-symmetry line. The right side of the pictures are partial enlarged drawing of the Weyl cone. (b) The state density curve of Weyl cones derived from Eq. (1). The red and blue lines represent the DOS of WP1 (+) and WP2 (−), respectively. (c) Band structure in the first Brillouin zones. (c) and (d) Iso-frequency map at frequencies of 21 GHz (less than that of WP2) and 22 GHz (higher than that of WP1), wherein the circle represents the size of the Weyl cone.

Close modal
From the energy band relationship, the distribution of the DOS of the Weyl cone is quadratically29 related to the frequency (supplementary material 1), which is expressed as Eq. (1). Notably, we fully considered the effect of air noise on DOS and, therefore, modified the equation for the DOS curve to
DOS = D O S Weyl + D O S noise .
(1)

When measuring the electric field intensity distribution between the air field and the sample, an air noise intensity approximately one-tenth of the experimental intensity was observed. Using the correction equation, two DOS curves associated with chirality (+ red, − blue) were derived [as shown on Fig. 2(a)]. The vertical axis represents the intensity on a normalized logarithmic scale, while the horizontal axis represents the frequency, corresponding to the bandgap region. As the linear distribution of WPs is confined to narrow bands, the present study focused on a frequency range of 3 GHz. Within this range, the intensity reversal within the bandgap suggests the presence of chirality imbalance. At a frequency of 21 GHz (below the low-frequency WPs), the band 2 energy band exhibits an elliptical curve, with WP1 (WP3) being larger than WP2 (WP4). As the frequency gradually increases, the size of the Weyl cone reverses. At a frequency of 22 GHz (above the high-frequency WPs), the band 3 energy band displays an iso-frequency elliptical curve, where WP2 (WP4) surpasses the iso-frequency curve of WP1 (WP3).

To investigate the chiral segregation of DOS in the WPs, measurements of the intensity were performed from iso-frequency maps of distinct chiral WPs, and momentum-resolved spectral intensity was determined at these points. During the experiment, a layered structural sample measuring 50 × 100 × 10 was created [Fig. 3(a)]. A dipole antenna was positioned at the center of the sample's lower surface to stimulate the electric field. Additionally, a stripped coaxial cable, devoid of its metal casing, was affixed to a stepper motor that scanned the upper surface of the sample as it moved (supplementary material 2). A 30 × 30 lattice width in the central region was scanned to exclude any influence from the Fermi arc on the intensity extraction. After applying FFT, an iso-frequency map of the Weyl energy band was prepared [Fig. 3(b)]. Additional iso-frequency surfaces can be found in supplementary material 3. To enhance measurement accuracy, the intensity in the designated region was determined by considering a length of 0.17×π/ai (where i represents x or y) both above and below each WP, and the DOS of the WPs was considered proportional to S P FFT W P ζ d s, where S denotes the area of a square, P represents the spectral intensity of each point in the area, and ζ=1–2–3–4 represents the four quadrants in the coordinate system (supplementary material 3). Figure 3(c) shows the DOS curves for different frequencies, unveiling a conspicuous segregation between positive and negative chiral WPs. Noteworthy, the simulation showed an intensity shift when traversing low-frequency WPs and high-frequency WPs. Conversely, in the experiment, the low-frequency signal is accompanied by heightened noise levels, which makes the distinction of energy band information from the noise challenging and thus resulting in the absence of evident intensity changes. At a frequency of 21.3 GHz on the iso-frequency diagram, all four WPs exhibit evident and nearly uniform intensities. This uniformity is further reflected in the DOS spectral line, wherein the intensity profile remains largely constant [(PWP2+PWP4) − (PWP1+PWP3) ≈ 0] [Fig. 3(c)]. However, upon crossing the high Weyl frequency (21.8 GHz), discernible sets of intensity curves emerge on the iso-frequency diagram [(PWP2+PWP4) − (PWP1+PWP3) > 0]. Figure 3(d) illustrates the direct manifestation of DOS segregation on the iso-frequency diagram, thereby substantiating the chirality imbalance and aligning with the earlier observations shown in Figs. 2(c) and 2(d).

FIG. 3.

Experiment measurement of the top surface. (a) Top surface of the sample, which was fabricated with a printed circuit board. (b) Iso-frequency maps at frequencies of 21.3 and 22 GHz. Inset: the white circle is the line cone. (c) Experimentally measured density curves of states derived from the iso-frequency map. The regular triangle (black and blue) and inverted triangle (red and green) represent chiral+ and chiral−, respectively. (d) Four similar intensities appear on the state density curves at each point in 21.3 GHz (ratio ≈ 1). However, the split of intensities is shown in 22 GHz (ratio ≈ 5).

FIG. 3.

Experiment measurement of the top surface. (a) Top surface of the sample, which was fabricated with a printed circuit board. (b) Iso-frequency maps at frequencies of 21.3 and 22 GHz. Inset: the white circle is the line cone. (c) Experimentally measured density curves of states derived from the iso-frequency map. The regular triangle (black and blue) and inverted triangle (red and green) represent chiral+ and chiral−, respectively. (d) Four similar intensities appear on the state density curves at each point in 21.3 GHz (ratio ≈ 1). However, the split of intensities is shown in 22 GHz (ratio ≈ 5).

Close modal

The Fermi arc, a crucial characteristic of Weyl semimetals, is preserved by the bulk state of the semimetal. The shape of the Fermi arc state can be predicted by truncating the infinite periodicity in the z-direction. The experimental setup involved a ten-layer cyclic structure in CST along the z-direction, where PEC boundaries were applied to the zmax and zmin surfaces to simulate open boundaries in real space, located a distance of 4.5 mm (equivalent to one cycle length). The xy plane followed periodic boundary conditions, while ky was fixed at 0.467 and the value of kx= [ 1 , 1 ]. With this setup, we derived the projected energy bands passing through different chiral WPs, comprising two components: a bulk band originating from the Weyl cone and a Fermi surface state resulting from the finite period. To simplify the calculation, we utilized a single-cell structure in the simulation, selecting 11 discrete points along the kz direction and projecting them onto the ky = 0.467 direction. By applying this procedure, the light gray region (representing the bulk state) was determined in the middle of the figure. Notably, the structures of the upper and lower connected surfaces were not identical. Therefore, in this experiment, the zmax and zmin surfaces were measured, and a lattice size of 98 × 48 in the sample area was scanned. The Brillouin zone was set to ky = −0.467, and the value of kx =  [ 1 , 1 ]. The results of (zmax) and (zmin), displayed on the right side of Figs. 4(a) and 4(b), respectively, suggest the separation of the two distinct chiral WPs at frequencies f1 = 21.8 and f2 = 21.4 GHz, consistent with the simulation outcomes (supplementary material 4). On the two different surfaces, surface Fermi arcs extended in diverse directions, connecting the two WPs. Together, these Fermi arcs constituted the spiral surface state observed in Weyl semimetals.

FIG. 4.

Experimental observation of the topological surface states. (a)–(c) Left: tangent position of the projected band. The plane of Cut-I is the projected area, which passes through two different chirality Weyl points (WPs). On the contrary, the plane of Cut-II passes through the same chirality WPs. (a)–(c) Middle: simulated band structures in CST, where the Fermi arcs are drawn in red line (top surface topological surface states), blue line (bottom surface topological surface states), and orange line (top surface and bottom surface coexistence). The gray area represents the bulk band. (a)–(c) Right: Measured band structures along Cut-I and Cut-II, which were obtained by the superposition of the iso-frequency diagram along the frequency. The black area of the right pictures in (b) and (c) represents the light line.

FIG. 4.

Experimental observation of the topological surface states. (a)–(c) Left: tangent position of the projected band. The plane of Cut-I is the projected area, which passes through two different chirality Weyl points (WPs). On the contrary, the plane of Cut-II passes through the same chirality WPs. (a)–(c) Middle: simulated band structures in CST, where the Fermi arcs are drawn in red line (top surface topological surface states), blue line (bottom surface topological surface states), and orange line (top surface and bottom surface coexistence). The gray area represents the bulk band. (a)–(c) Right: Measured band structures along Cut-I and Cut-II, which were obtained by the superposition of the iso-frequency diagram along the frequency. The black area of the right pictures in (b) and (c) represents the light line.

Close modal

Figure 4(c) shows the results of the projected cross section along the diagonal direction. The simulated energy band results reveal the same frequency, f1 = 21.8 GHz, and exhibit a symmetrical shape of the WPs. The orange curve represents the calculated surface Fermi arc. It is worth noting that the Fermi arc exhibits a twofold degeneracy, and the eigen electric field is localized on both the upper and lower surfaces, mainly because the two surfaces of ⟨001⟩ are equivalent when kx = ky.

In this study, breaking of the mirror symmetry reveals the splitting of frequencies in WPs with different chirality from the central position f0. Particularly, the “+” chiral WPs exhibit a blue-shift to f1, while the “−” chiral WPs display a red shift to f2, depending on their chirality. The distribution of the DOS was investigated both theoretically and experimentally. Notably, we observed that the DOS of “−” chiral WPs surpasses that of “+” chiral WPs when the frequency is below that of the “−” chiral WPs. Additionally, a flip in the DOS occurs across the “+” chiral WPs. The DOS plays a crucial role in describing the transport properties of electron systems. By establishing a connection between WPs and chirality transport, the findings of this study provide insights into the phenomenon of imbalanced chirality in real space within photonic Weyl semimetals. This research opens up avenues for further investigations into the chiral physics of pseudo-fermionic fields in photonic materials.

See the supplementary material that includes the theoretical formula for calculating the density of states (DOS) and the experimental measurement and post-processing methods.

This work was supported by the Natural Science Foundation of China (NSFC) (Grant No. 12074279), the Major Program of Natural Science Research of Jiangsu Higher Education Institutions (Grant No.18KJA140003), and the Priority Academic Program Development (PAPD) of Jiangsu Higher Education Institutions.

The authors have no conflicts to disclose.

Xiaoxi Zhou: Conceptualization (lead); Data curation (lead); Formal analysis (lead); Funding acquisition (lead); Investigation (lead); Methodology (lead); Supervision (equal); Writing – original draft (lead); Writing – review & editing (lead). Shanshan Li: Conceptualization (supporting); Data curation (supporting); Formal analysis (supporting); Investigation (supporting); Methodology (supporting). Chuandeng Hu: Conceptualization (supporting); Formal analysis (supporting); Investigation (supporting); Methodology (supporting); Writing – review & editing (supporting). Gang Wang: Formal analysis (supporting); Funding acquisition (supporting); Methodology (supporting); Supervision (supporting); Writing – review & editing (supporting). Bo Hou: Conceptualization (equal); Formal analysis (lead); Funding acquisition (lead); Investigation (lead); Methodology (lead); Supervision (lead); Writing – original draft (lead); Writing – review & editing (lead).

The data that support the findings of this study are available from the corresponding authors upon reasonable request.

1.
H.
Weyl
, “
Gravitation and the electron
,”
Proc. Natl. Acad. Sci. U. S. A.
15
,
323
334
(
1929
).
2.
X. G.
Wan
,
A. M.
Turner
,
A.
Vishwanath
, and
S. Y.
Savrasov
, “
Topological semimetal and Fermi-arc surface states in the electronic structure of pyrochlore iridates
,”
Phys. Rev. B
83
(
20
),
205101
(
2011
).
3.
N. P.
Armitage
,
E. J.
Mele
, and
A.
Vishwanath
, “
Weyl and Dirac semimetals in three-dimensional solids
,”
Rev. Mod. Phys.
90
(
1
),
015001
(
2018
).
4.
Z. K.
Liu
,
B.
Zhou
,
Y.
Zhang
,
Z. J.
Wang
,
H. M.
Weng
,
D.
Prabhakaran
,
S. K.
Mo
,
Z. X.
Shen
,
Z.
Fang
,
X.
Dai
,
Z.
Hussain
, and
Y. L.
Chen
, “
Discovery of a three-dimensional topological Dirac semimetal, Na3Bi
,”
Science
343
(
6173
),
864
867
(
2014
).
5.
S. Y.
Xu
,
C.
Liu
,
S. K.
Kushwaha
,
R.
Sankar
,
J. W.
Krizan
,
I.
Belopolski
,
M.
Neupane
,
G.
Bian
,
N.
Alidoust
,
T. R.
Chang
,
H. T.
Jeng
,
C. Y.
Huang
,
W. F.
Tsai
,
H.
Lin
,
P. P.
Shibayev
,
F. C.
Chou
,
R. J.
Cava
, and
M. Z.
Hasan
, “
Observation of Fermi arc surface states in a topological metal
,”
Science
347
(
6219
),
294
298
(
2015
).
6.
Q. H.
Yan
,
R. J.
Liu
,
Z. B.
Yan
,
B. Y.
Liu
,
H. S.
Chen
,
Z.
Wang
, and
L.
Lu
, “
Experimental discovery of nodal chains
,”
Nat. Phys.
14
(
5
),
461
464
(
2018
).
7.
R.
Yu
,
H. M.
Weng
,
Z.
Fang
,
X.
Dai
, and
X.
Hu
, “
Topological node-line semimetal and Dirac semimetal state in antiperovskite Cu3PdN
,”
Phys. Rev. Lett.
115
(
3
),
036807
(
2015
).
8.
R.
Yu
,
Q. S.
Wu
,
Z.
Fang
, and
H. M.
Weng
, “
From nodal chain semimetal to Weyl semimetal in HfC
,”
Phys. Rev. Lett.
119
(
3
),
036401
(
2017
).
9.
B. A.
Yang
,
Q. H.
Guo
,
B.
Tremain
,
L. E.
Barr
,
W. L.
Gao
,
H. C.
Liu
,
B.
Beri
,
Y. J.
Xiang
,
D. Y.
Fan
,
A. P.
Hibbins
, and
S.
Zhang
, “
Direct observation of topological surface-state arcs in photonic metamaterials
,”
Nat. Commun.
8
,
97
(
2017
).
10.
B. Q.
Lv
,
H. M.
Weng
,
B. B.
Fu
,
X. P.
Wang
,
H.
Miao
,
J.
Ma
,
P.
Richard
,
X. C.
Huang
,
L. X.
Zhao
,
G. F.
Chen
,
Z.
Fang
,
X.
Dai
,
T.
Qian
, and
H.
Ding
, “
Experimental discovery of Weyl semimetal TaAs
,”
Phys. Rev. X.
5
(
3
),
031013
(
2015
).
11.
L. C.
Yang
,
G. R.
Li
,
X. M.
Gao
, and
L.
Lu
, “
Topological-cavity surface-emitting laser
,”
Nat. Photonics
16
(
4
),
279
(
2022
).
12.
X. L.
Liu
,
L. J.
Zhao
,
D.
Zhang
, and
S. H.
Gao
, “
Topological cavity laser with valley edge states
,”
Opt. Express
30
(
4
),
4965
4977
(
2022
).
13.
M.
Hafezi
,
E. A.
Demler
,
M. D.
Lukin
, and
J. M.
Taylor
, “
Robust optical delay lines with topological protection
,”
Nat. Phys.
7
(
11
),
907
912
(
2011
).
14.
C.
Zhang
,
Y.
Zhang
,
H. Z.
Lu
, et al “
Cycling Fermi arc electrons with Weyl orbits
,”
Nat. Rev. Phys.
3
,
660
670
(
2021
).
15.
S. Y.
Xu
,
I.
Belopolski
,
N.
Alidoust
,
M.
Neupane
,
G.
Bian
,
C. L.
Zhang
,
R.
Sankar
,
G. Q.
Chang
,
Z. J.
Yuan
,
C. C.
Lee
,
S. M.
Huang
,
H.
Zheng
,
J.
Ma
,
D. S.
Sanchez
,
B. K.
Wang
,
A.
Bansil
,
F. C.
Chou
,
P. P.
Shibayev
,
H.
Lin
,
S.
Jia
, and
M. Z.
Hasan
, “
Discovery of a Weyl fermion semimetal and topological Fermi arcs
,”
Science
349
(
6248
),
613
617
(
2015
).
16.
J. B.
Pendry
,
D.
Schurig
, and
D. R.
Smith
, “
Controlling electromagnetic fields
,”
Science
312
(
5781
),
1780
1782
(
2006
).
17.
G. G.
Liu
,
P. H.
Zhou
,
Y. H.
Yang
,
H. R.
Xue
,
X.
Ren
,
X.
Lin
,
H. X.
Sun
,
L.
Bi
,
Y. D.
Chong
, and
B. L.
Zhang
, “
Observation of an unpaired photonic Dirac point
,”
Nat. Commun.
11
(
1
),
1873
(
2020
).
18.
L.
Lu
,
Z. Y.
Wang
,
D. X.
Ye
,
L. X.
Ran
,
L.
Fu
,
J. D.
Joannopoulos
, and
M.
Soljacic
, “
Experimental observation of Weyl points
,”
Science
349
(
6248
),
622
624
(
2015
).
19.
D. Y.
Wang
,
B.
Yang
,
W. L.
Gao
,
H. W.
Jia
,
Q. L.
Yang
,
X. Y.
Chen
,
M. G.
Wei
,
C. X.
Liu
,
M.
Navarro-Cia
,
J. G.
Han
,
W. L.
Zhang
, and
S.
Zhang
, “
Photonic Weyl points due to broken time-reversal symmetry in magnetized semiconductor
,”
Nat. Phys.
15
(
11
),
1150
(
2019
).
20.
S. Y.
Yu
,
X. C.
Sun
,
X.
Ni
,
Q.
Wang
,
X. J.
Yan
,
C.
He
,
X. P.
Liu
,
L.
Feng
,
M. H.
Lu
, and
Y. F.
Chen
, “
Surface phononic graphene
,”
Nat. Mater.
15
(
12
),
1243
1247
(
2016
).
21.
X. Q.
Huang
,
W. Y.
Deng
,
F.
Li
,
J. Y.
Lu
, and
Z. Y.
Liu
, “
Ideal type-II Weyl phase and topological transition in phononic crystals
,”
Phys. Rev. Lett.
124
(
20
),
206802
(
2020
).
22.
B.
Yang
,
Q. H.
Guo
,
B.
Tremain
,
R. J.
Liu
,
L. E.
Barr
,
Q. H.
Yan
,
W. L.
Gao
,
H. C.
Liu
,
Y. J.
Xiang
,
J.
Chen
,
C.
Fang
,
A.
Hibbins
,
L.
Lu
, and
S.
Zhang
, “
Ideal Weyl points and helicoid surface states in artificial photonic crystal structures
,”
Science
359
(
6379
),
1013
1016
(
2018
).
23.
H. W.
Jia
,
R. X.
Zhang
,
W. L.
Gao
,
Q. H.
Guo
,
B.
Yang
,
J.
Hu
,
Y. G.
Bi
,
Y. J.
Xiang
,
C. X.
Liu
, and
S.
Zhang
, “
Observation of chiral zero mode in inhomogeneous three-dimensional Weyl metamaterials
,”
Science
363
(
6423
),
148
151
(
2019
).
24.
Q.
Li
,
D. E.
Kharzeev
,
C.
Zhang
,
Y.
Huang
,
I.
Pletikosic
,
A. V.
Fedorov
,
R. D.
Zhong
,
J. A.
Schneeloch
,
G. D.
Gu
, and
T.
Valla
, “
Chiral magnetic effect in ZrTe5
,”
Nat. Phys.
12
(
6
),
550
(
2016
).
25.
D. E.
Kharzeev
, “
The chiral magnetic effect and anomaly-induced transport
,”
Prog. Part. Nucl. Phys.
75
,
133
151
(
2014
).
26.
D. T.
Son
and
B. Z.
Spivak
, “
Chiral anomaly and classical negative magnetoresistance of Weyl metals
,”
Phys. Rev. B
88
(
10
),
104412
(
2013
).
27.
A. A.
Burkov
, “
Chiral anomaly and diffusive magnetotransport in Weyl metals
,”
Phys. Rev. Lett.
113
(
24
),
247203
(
2014
).
28.
X. X.
Zhou
,
C. D.
Hu
,
W. X.
Lu
,
Y.
Lai
, and
B.
Hou
, “
Numerical design of frequency-split Weyl points in Weyl metamaterial
,”
Acta Phys. Sin.
69
(
15
),
154204
(
2020
).
29.
Y.
Yang
,
W. L.
Gao
,
L. B.
Xia
,
H.
Cheng
,
H. W.
Jia
,
Y. J.
Xiang
, and
S.
Zhang
, “
Spontaneous emission and resonant scattering in transition from type I to type II photonic Weyl systems
,”
Phys. Rev. Lett.
123
(
3
),
033901
(
2019
).

Supplementary Material