Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Possible realization of optical Dirac points in woodpile photonic crystals

Open Access Open Access

Abstract

The simulation of fermionic relativistic physics, e.g., Dirac and Weyl physics, has led to the discovery of many unprecedented phenomena in photonics, of which the optical-frequency realization is, however, still challenging. Here, surprisingly, we discover that the woodpile photonic crystals commonly used for optical frequency applications host exotic fermion-like relativistic degeneracies: a Dirac nodal line and a fourfold quadratic point, as protected by the nonsymmorphic crystalline symmetry. Deforming the woodpile photonic crystal leads to the emergence of type-II Dirac points from the fourfold quadratic point. Such type-II Dirac points can be detected by its anomalous refraction property which is manifested as a giant birefringence in a slab setup. Our findings provide a promising route towards 3D optical Dirac physics in all-dielectric photonic crystals.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Fermionic relativistic waves described by the Dirac and Weyl equations [15] (and beyond [6]) have many nontrivial properties as discovered in condensed materials [7]. Recently, photonic crystals (PCs) became a versatile platform to simulate such relativistic waves [823]. Photonic relativistic waves have been harnessed for a number of fundamental phenomena and applications such as Zitterbewegung [24], pseudodiffusive transport [25], zero-index metamaterials [26], synthetic magnetic fields for photons [27], and anomalous refraction [17]. In particular, fermion-like relativistic (i.e., even-fold band degeneracy) waves are more appealing than the boson-like (i.e., odd-fold band degeneracy) counterparts, since they are closely related to photonic topological insulators [2840]. At optical frequencies, the magneto-optical and bianisotropic effects of natural materials, which are often used to create photonic topological insulators, become negligible. One then has to turn to all-dielectric PCs as low-dissipation optical crystalline materials. Recent reports on the experimental observations of photonic Weyl points at the infrared and even optical frequency regimes [1821] indicate that three-dimensional (3D) dielectric PCs still hold the promise toward Weyl physics. However, realizing 3D optical Dirac points (DPs) as a potential pathway towards optical 3D topological insulators is more challenging, since in PCs the spin degeneracy of photons is broken. Space symmetry must be leveraged to simulate both the fermion-like Kramers degeneracy and the parity inversion [17]. With the limited types of available optical-frequency 3D PCs in the current technology, it is unknown which can lead to 3D photonic DPs.

In this article, we illustrate a practical route towards 3D optical DPs: using woodpile-like PCs—a prototype 3D optical-frequency PCs that have been fabricated with high quality [4146] [Fig. 1(a)]. Surprisingly, we find that the woodpile PCs host two types of exotic band degeneracies in the lowest photonic bands: a Dirac nodal line and a fourfold quadratic point (FQP), as protected by the nonsymmorphic crystalline symmetry. Starting from the motherboard of the woodpile PCs, type-II DPs can be created by deforming the unit-cell geometry. Interestingly, we find that the type-II DPs exhibit anomalous birefringence. Such birefringence is maximized when the incident light excite exactly the type-II DPs. These findings provide a promising path towards optical-frequency Dirac physics and the potential realization of 3D optical topological insulators in all-dielectric PCs that are compatible with optoelectronic integration and nano-photonic applications [44,47].

 figure: Fig. 1.

Fig. 1. (a) Optical-frequency woodpile PCs: layer-by-layer stacking of dielectric (colored) logs. (b) Lattice vectors of the undeformed woodpile PC, ${\vec a}_j$, are shown together with those of the deformed woodpile PC, ${\vec b}_j$ ($j=1,2,3$). (c) The top-view (upper) and the side-view (lower) of the unit-cell of the deformed (green) woodpile PC. (d) The relationship between the Brillouin zones of the undeformed (gray) and the deformed woodpile PCs. (e) The low-lying photonic bands of the undeformed woodpile PC with $l_x=l_y=b_x/2$, $w_x=w_y=0.4b_x$, $h=0.2b_x$, and $\varepsilon =5.06$ (TiO$_2$). The connection between photonic bands in the tetragonal and face-centered cubic Brillouin zone is given in Appendix A. Inset: the Dirac dispersion at the $A$ point (labeled by the star). The $M$-$A$ line (red) is a Dirac nodal line on which each point has Dirac-like dispersions.

Download Full Size | PDF

2. Woodpile space symmetry

Structures of the undeformed and deformed woodpile PCs, which consist of four layers of rectangular rod stacked on top of each other with a relative $90^{\circ }$ in plane rotation, are illustrated in Figs. 1(a)–1(c) together with their unit-cells and lattice vectors. Note that the undeformed woodpile PCs exhibit tetragonal symmetry of space group $P4_2/mcm$. The two corresponding Brillouin zones are illustrated in Fig. 1(d). The deformed woodpile PC has a unit-cell twice as large as the unit-cell of the undeformed woodpile PC. These two PCs can be described in an unified fashion using the unit-cell of the deformed woodpile PC. In this study, we set the lattice constant as $|{\vec b}_1|=|{\vec b}_2|=b_x=0.8~\mu$m and $|{\vec b}_3|=b_z=0.8b_x$. The deformation of the woodpile PC can be feasibly realized by tuning the distances between the dielectric logs, $l_x$ and $l_y$, or by tuning the width of the logs, $w_x$ and $w_y$. The undeformed woodpile represents the limit with $w_x=w_y$ and $l_x=l_y=b_x/2$. We consider mostly the woodpile PCs made of TiO$_2$ which can be directly generalized to other dielectric materials such as silicon and GaAs.

The undeformed woodpile PC is a spiral stacking of the dielectric logs, with a four fold screw symmetry $S_{\frac {\pi }{2}}:=(x, y, z) \to (y, -x, z + \frac {b_z}{4})$. We list here only the most relevant symmetries: $M_x := (x,y,z) \to (-x-\frac {b_x}{2}, y, z)$ and $M_y := (x,y,z) \to (x, -y-\frac {b_x}{2}, z)$, the glide symmetry $G_z := (x, y, z) \to (x + \frac {b_x}{2}, y, -z-\frac {b_z}{4})$, the fourfold screw symmetries $S_{\frac {\pi }{2}}$, and the point group symmetry $S_4 := (x,y,z) \to (y, - x, -z - \frac {b_z}{2})$. The deformed woodpile PCs may break most of the above symmetries, leaving some of the mirror or glide symmetries unchanged. Note that in this Letter, we use the capital letters for the symmetry operators while the small letters for their eigenvalues.

3. Symmetry-enriched degeneracy

Quite different from the simulation of Weyl points, the realization of the synthetic Kramers degeneracy ${\cal T}_p^{2}=-1$ for photons is crucial for the simulation of the Dirac and quadratic points as well as the Dirac nodal lines. Here, the synthetic Kramers degeneracy is realized via the nonsymmorphic crystalline symmetries. For instance, the double degeneracy on the $k_z=\frac {\pi }{b_z}$ plane [the $A$-$R$-$Z$ line in Fig. 1(e)] can be understood by constructing an anti-unitary operator $\Theta _{\pi }\equiv S_{\frac {\pi }{2}}^{2}*{\cal T}$ which is invariant at the $k_z=\frac {\pi }{b_z}$ plane and yields $\left (\Theta _{\pi }\right )^{2}\Psi _{n,{\vec k}}=e^{ik_zb_z}\Psi _{n,{\vec k}}$ for all photonic Bloch states $\Psi _{n,{\vec k}}\equiv ({\vec E}_{n,{\vec k}}, {\vec H}_{n,{\vec k}})^{T}$ (including both the electric field ${\vec E}$ and the magnetic field ${\vec H}$ for the $n$-th band with a wavevector ${\vec k}$). Hence,

$$\left.\Theta_\pi^{2}\right|_{k_z=\frac{\pi}{b_z}} ={-}1,$$
leads to the synthetic Kramers degeneracy for all photonic bands on the $k_z=\frac {\pi }{b_z}$ plane. Similarly, the double degeneracies on the $k_x=\frac {\pi }{b_x}$ [the $X$-$M$-$A$-$R$ line in Fig. 1(e)] and $k_y=\frac {\pi }{b_x}$ planes are induced by the other screw symmetries of the woodpile PCs (see Appendix B for details).

4. Photonic Dirac nodal line

The $M$-$A$ line is a nodal line composed of an “infinite” number of two-dimensional (2D) DPs [see Fig. 1(e) and Fig. 2]. The two fundamental elements in the Dirac equation, the spin and orbital degrees-of-freedom, are associated with the four degenerate states on the line. The field profiles of these four modes [Fig. 2(a)] indicate that they are the electric and magnetic dipole modes which can be denoted by the mirror symmetry eigenvalues as $|{m_x,m_y}\rangle$ ($m_x,m_y=\pm 1$). The fourfold degeneracy is dictated by $\Theta _z\equiv G_z*{\cal T}$ and $S_{\pi }\equiv S_{\frac {\pi }{2}}^{2}$ which are invariant operators on the $M$-$A$ line, as manifested by the following symmetric transformations (see Appendix C for more details),

$$\Theta_z|{m_x,m_y}\rangle = |{-m_x,m_y}\rangle,$$
$$S_{\pi}|{m_x,m_y}\rangle = |{-m_x,-m_y}\rangle,$$
$$\Theta_zS_{\pi}|{m_x,m_y}\rangle = |{m_x,-m_y}\rangle .$$

Here, the “orbital” degree-of-freedom are associated with the parity, ${\cal P}=m_xm_y$. For example, the two electric dipole modes $|{m_x=-1,m_y=1}\rangle$ and $|{m_x=1,m_y=-1}\rangle$ constitute the odd-parity “antiparticle” sector of the Dirac equation, whereas the magnetic dipole modes $|{1,1}\rangle$ and $|{-1,-1}\rangle$ comprise the even-parity “particle” sector. The “spin” states for the particle sectors (‘$p$’) and antiparticle (‘$a$’) are constructed, respectively, as

$$\begin{aligned}& |{p, \uparrow}\rangle = \frac{1}{\sqrt{2}}(|{1,1}\rangle + i |{-1,-1}\rangle) \equiv \frac{|{x^{2}-y^{2}}\rangle+i|{2xy}\rangle}{\sqrt{2}},\end{aligned}$$
$$\begin{aligned}& |{p, \downarrow}\rangle = \frac{1}{\sqrt{2}} (|{1,1}\rangle - i |{-1,-1}\rangle) \equiv \frac{|{x^{2}-y^{2}}\rangle-i|{2xy}\rangle}{\sqrt{2}}, \end{aligned}$$
$$\begin{aligned}& |{a, \uparrow}\rangle = \frac{1}{\sqrt{2}}(|{-1,1}\rangle + i |{1,-1}\rangle) \equiv \frac{|{x}\rangle+i|{y}\rangle}{\sqrt{2}},\end{aligned}$$
$$\begin{aligned}& |{a, \downarrow}\rangle = \frac{1}{\sqrt{2}} (|{-1,1}\rangle - i |{1,-1}\rangle) \equiv \frac{|{x}\rangle-i|{y}\rangle}{\sqrt{2}}. \end{aligned}$$

In the above equations, we have denoted the mirror eigenstates as $| {-1,1}\rangle\equiv | {x}\rangle$, $| {1,-1}\rangle\equiv | {y}\rangle$, $| {1,1}\rangle\equiv | {x^{2}-y^{2}}\rangle$ and $| {-1,-1}\rangle\equiv | {2xy}\rangle$ to elucidate the spatial symmetry of the eigenstates. We find that the two spin states carry finite and opposite total angular momenta (including both spin and orbital angular momentum) of photon, which is a natural generalization of the concept of emulating fermion-like spin with photonic spin [33] or orbital angular momentum [36,38] in previous studies.

 figure: Fig. 2.

Fig. 2. (a) Field profiles of the four degenerate modes on the $M$-$A$ line. (b) and (c): Dispersion of the Dirac nodal line in the (b) $k_x$-$k_z$ and (c) $k_x$-$k_y$ planes. In (b) an isofrequency plane (the blue-gray sheet) is plotted in order to show the isofrequency contours (the red curves). The Dirac nodal line is labeled by the black curve. Parameters are the same as in Fig. 1.

Download Full Size | PDF

Photonic simulation of a fermionic Hamiltonian $\hat {{\cal H}}_F$ is the following mapping from the photonic Hamiltonian $\hat {{\cal H}}_{EM}$ to the fermionic one (resembling the mapping between the Dirac equation and the Klein-Gordon equation [48]),

$$\hat{{\cal H}}_{EM} := (\hat{{\cal H}}_F)^{2} .$$

From the Maxwell equation, $c^{2}\boldsymbol {\nabla }\times \frac {1}{\varepsilon }\boldsymbol {\nabla }\times {\vec H}=\omega ^{2}{\vec H}$ ($c$ is the speed of light in vacuum, $\varepsilon$ the relative permittivity, and ${\vec H}$ is the magnetic field), the photonic Hamiltonian can be defined as the Hermitian operator $\hat {{\cal H}}_{EM}:=c^{2}\boldsymbol {\nabla }\times \frac {1}{\varepsilon }\boldsymbol {\nabla }\times$ [45].

In the basis of the four states, $| {\rho,\uparrow }\rangle$, $| {\rho,\downarrow }\rangle$ ($\rho =a,p$), the above mapping (together with the ${\vec k}\cdot {\vec P}$ theory in Appendix D) yields the following fermion-like Hamiltonian,

$$\hat{{\cal H}}_F^{DL} = \omega_0 + v \left( \begin{array}{cc} 0 & \hat{{\cal A}} \\ \hat{{\cal A}}^{{\dagger}} & 0 \end{array}\right),$$
where $\hat {{\cal A}} \equiv \gamma _x q_x \sigma _x + \gamma _y q_y \sigma _y$, $q_j=k_j-\pi /b_x$ ($j=x,y$), and $\sigma _j (j=x,y,z)$ refers to Pauli matrix. The coefficients $\omega _0$, $v$, $\gamma _x$ and $\gamma _y$ are $k_z$-dependent, making the Dirac nodal line. At the $M$ and $A$ points, the $S_4$ symmetry imposes additional constraints, $\gamma _x=\gamma _y$, leading to an isotropic 2D DP as shown in Fig. 1(e). Away from these two points, the $S_4$ symmetry is ineffective, yielding unconventional 2D DPs as shown in Fig. 2(c).

Since the particle and antiparticle sectors correspond to the magnetic and electric dipole modes, respectively, the magneto-electric coupling in the Maxwell equations naturally guarantee the linear in ${\vec q}$ “spin-orbit couplings”. The coupling coefficients can be written as (Here, $j=x,y$, and UC stands for the unit cell)

$$v\tilde{\gamma}_j = \frac{c }{\sqrt{N_EN_H}}\int_{UC} d{\vec r}~ ({\vec E}_{\uparrow, p}^{{\ast}}\times {\vec H}_{\downarrow, a} + {\vec E}_{\downarrow, a} \times {\vec H}_{\uparrow, p}^{{\ast}}) \cdot {\vec n}_j,$$
where $\tilde {\gamma }_x=\gamma _x$ and $\tilde {\gamma }_y=-i\gamma _y$; $N_E\equiv \int _{UC}d{\vec r} \varepsilon ({\vec r})|{\vec E}|^{2}$, $N_H\equiv \int _{UC}d{\vec r} |{\vec H}|^{2}$; ${\vec n}_{j}$ is the unit vector along the $j=x,y$ direction; the integration is within a unit-cell. Interestingly, the above expression is similar to the Poynting vector between the particle and antiparticle sectors. A general form of the Dirac velocity tensor is presented in Appendix D where the “selection rules” due to the mirror and glide symmetries are discussed.

5. Photonic fourfold quadratic point

The $Z$ point is a photonic FQP which is induced by the fourfold screw symmetry $S_{\frac {\pi }{2}}$ as

$$(\Theta_{\frac{\pi}{2}})^{4} =\left. e^{ik_zb_z}\right|_{k_z=\frac{\pi}{b_z}} ={-} 1 ,$$
where $\Theta _{\frac {\pi }{2}}\equiv S_{\frac {\pi }{2}}*{\cal T}$ transforms $(k_x,k_y,k_z)$ to $(-k_y, k_x, -k_z)$ and is an invariant operator at the $Z$ point. The above indicates that the quadruplet consist of $| {\Psi }\rangle$, $\Theta _{\frac {\pi }{2}}| {\Psi }\rangle$, $(\Theta _{\frac {\pi }{2}})^{2}| {\Psi }\rangle$, and $(\Theta _{\frac {\pi }{2}})^{3}| {\Psi }\rangle$. If $| {\Psi }\rangle$ is labeled as $| {m_x,m_y,g_z}\rangle$ ($g_z=\pm 1$ is the eigenvalue of $G_z$), then
$$\Theta_{\frac{\pi}{2}}|{\Psi}\rangle=|{m_y,m_x,g_z}\rangle,$$
$$(\Theta_{\frac{\pi}{2}})^{2}|{\Psi}\rangle=|{m_x,m_y,-g_z}\rangle,$$
$$(\Theta_{\frac{\pi}{2}})^{3}|{\Psi}\rangle=|{m_y,m_x,-g_z}\rangle.$$

The field patterns of the eigenstates are shown in Fig. 3(a), indicating $m_x=-m_y$. The fourfold degeneracy at the $Z$ point is protected by the screw symmetry $S_{\frac {\pi }{2}}$ and hence independent of the specific parameters of the woodpile PC (see Appendix E for details).

 figure: Fig. 3.

Fig. 3. Quadratic degeneracy points: (a) Photonic bands in the $k_y=0$ plane for the same parameters as in Fig. 1. The orange (blue) band has $m_y=+1$ (−1). The $Z$ point (indicated by the arrow) is a FQP with four eigenstates of different mirror ($m_{x/y}$) and glide ($g_z$) symmetries (illustrated in the inset). (b) and (c): Dispersion of the FQP in the (b) $k_x$-$k_y$ and (c) $k_y$-$k_z$ planes.

Download Full Size | PDF

The photonic system simulates the following fermion-like Hamiltonian of a 3D FQP,

$$\begin{aligned} & \hat{{\cal H}}_F^{Z} = \omega_Z + v_z q_z\hat{\tau}_y\\ &\quad \quad \quad + f_0 [ (q_x^{2}-q_y^{2}) \hat{\sigma}_z + 2 f_1 q_x q_y \hat{\sigma}_x + f_2 q_\parallel^{2} ] , \end{aligned}$$
where $\omega _Z$ is the frequency at the $Z$ point, $v_z$ is the group velocity along the $z$ direction, and $q_z=k_z-\frac {\pi }{b_z}$. $\tau _z=\pm 1$ labels the $g_z=\pm 1$ states, while $\sigma _z=\pm 1$ labels the $m_y=\pm 1$ states. Note that both $\hat {\sigma _j}$ and $\hat {\tau _j} (j=x,y,z)$ refer to Pauli matrix. The coefficients $f_i$ ($i=0,1,2$) depend on the geometry and materials of the PC (see Appendix F for details).

6. Photonic Dirac points

There are two means to split the FQP to yield a pair of DPs: either tuning $w_y/w_x$ or $l_y/l_x$ away from unity. For these tuning, $S_{\frac {\pi }{2}}$ is broken while $(\Theta _{\frac {\pi }{2}})^{2}=S_{\pi }$ is preserved. From Eq. (7c), the degeneracy between states of different $m_{x/y}$ is then lifted. This can be described by a constant perturbation $\Delta _zf_0 \sigma _z$ that splits the FQP to a pair of DPs emerging at two wavevectors ${\vec K}_{\pm } = (0, K_y^{\pm }, \frac {\pi }{b_z})$ with $K_y^{\pm } = \pm \sqrt {|\Delta _z|}$. The DPs are described by the following Hamiltonian,

$$\hat{{\cal H}}_F^{DP\pm} = \omega_D \pm v_0 \delta k_y + v_z \delta k_z\hat{\tau}_y \pm v_x \delta k_x \hat{\sigma}_x \pm v_y \delta k_y \sigma_z ,$$
where ${\vec \delta k}={\vec k}-{\vec K}_{\pm }$. Here, the coefficients are given by $\omega _D=\omega _Z+f_2f_0|\Delta _z|$, $v_0=2f_2f_0 K_y^{+}$, $v_x=2f_1f_0 K_y^{+}$, and $v_y=-2f_0K_y^{+}$. With the parameters adopted, we find that $|v_0|>|v_y|$, thus the Dirac cones are of the type-II nature [see Fig. 4(a)].

 figure: Fig. 4.

Fig. 4. (a) Photonic dispersion near a type-II DP for $l_x=l_y=0.5$, $w_x=0.2$, $w_y=0.3$, $h=0.25$, and $\varepsilon =11$. (b) Illustration of the refraction angle measurement. $\theta _i$ and $\phi _i$ are angles of incidence, while $\delta \phi _o$ is the angular difference between the two refraction beams. (c) $\delta \phi _o$ vs. $\theta _i$ and $\omega$ at $\phi _i=89.5^{\circ }$ as a signature of the type-II DP. (d) The group velocity along $x$ direction $v_x$ for the beams correspond to the lower branch as a function of $\phi _i$ for $\theta _i=47.5$ (aligned with the DP) and $\theta _i=45$ (misaligned with the DP).

Download Full Size | PDF

7. Anomalous refraction

The band degeneracies have been detected via transmission measurements [9,11,1821]. Here, we show that they can also be detected via unconventional refraction, specifically, the birefringence. Birefringence emerges here because the type-II DPs support two branches of propagating refraction beams with different group velocities. The setup for measuring the birefringence is illustrated in Fig. 4(b). A Gaussian beam is shed on the PC slab and is detected at the bottom of the slab. The drifted beam centers at the detection plane, ${\vec r}_o=r_{o}(\cos \phi _o,\sin \phi _o)$, gives the azimuth refraction angle $\phi _o$. In the theory of refraction, the wavevector in the PC is determined by frequency and parallel wavevector matching. We find that the refraction angle is determined by the group velocities as $\phi _o=\arctan (v_{g,y}/v_{g,x})$ and $\theta _o=\arctan (\sqrt {v_{g,x}^{2}+v_{g,y}^{2}}/v_{g,z})$ (see Appendix G for details).

Considering refraction on the (001) surface, type-II dispersion yields birefringence with two refraction beams [17], characterized by two displacement vectors ${\vec r}_{o,\pm }$. We show that the DPs can be detected by studying the difference in the azimuth refraction angle $\delta \phi _o\equiv |\phi _{o,+}-\phi _{o,-}|$. Figure 4(c) shows that the DP at ${\vec K}_+$ can be identified as the point with maximum $\delta \phi _o$ when frequency $\omega$ and angle of incidence $\theta _i$ are swept in large ranges at a fixed azimuth incidence angle $\phi _i$. This phenomenon is a signature of the type-II DPs: as the excited wavevector approaches the DP along the $k_x$ direction, the difference in $v_{g,x}$ for the upper and lower branches changes abruptly. In contrast, away from the DP, such change is gradual. Similar scenario happens when $\phi _i$ is swept at fixed $\theta _i$ and $\omega$, as shown in Fig. 4(d). Therefore, the difference in the azimuth refraction angle $\delta \phi _o$ is a signature of the DPs.

8. Conclusion

We have shown that the crystalline symmetry of woodpile PCs can enable the emergence of the 3D optical quadratic point and DPs. The DPs exhibit anomalous refraction and birefringence that can enable experimental detection of them. This study offers a guide for the realization of optical Dirac nodal line, FQP and DPs in 3D all-dielectric PCs. It also provides the inspiration and stimulation towards future realization of optical 3D topological insulators in all-dielectric PCs.

Appendix A: Connection between photonic bands in the tetragonal and face-centered cubic Brillouin zones

Here, we illustrate the equivalence of photonic band in the scheme of tetragonal (TET) and face-centered-cubic (FCC) unit-cells under the case with $l_x=l_y=0.5$. For simplicity, we assume that the all lattice constants along the three directions are 1, i.e., $b_x=b_z=1$. As mentioned in the main text, the FCC unit-cell is the primitive unit-cell with only two logs, while the TET unit-cell has four logs. For the TET scheme, the lattice vectors are defined as follows:

$${\vec b}_1=(1,\ 0,\ 0), {\vec b}_2=(0,\ 1,\ 0), {\vec b}_3=(0,\ 0,\ 1).$$

While for the FCC scheme, we choose the lattice vexctors as follows:

$${\vec a}_1=\frac{1}{2}(-{\vec b}_1+{\vec b}_2+{\vec b}_3), {\vec a}_2=\frac{1}{2}({\vec b}_1+{\vec b}_2+{\vec b}_3), {\vec a}_3={\vec b}_2.$$

Thus $|{\vec a}_1|=|{\vec a}_2|=\frac {\sqrt {3}}{2}$, $|{\vec {a}_3}|=1$, and the volume of the FCC unit-cell is $\Omega =|\vec {a}_3 \cdot (\vec {a}_1 \times \vec {a}_2)|=\frac {1}{2}$. It is known to all that the first Brillouin zone reduced by half when the unit-cell is doubled. In order to match the high symmetry points between the TET and FCC Brillouin zone, we need to rotate the unit-cell with 45 degree around $z-axis$, i.e., multiple the lattice vectors with rotation matrix $R$,

$$\left( \begin{matrix} \frac{1}{\sqrt{2}} & \frac{1}{\sqrt{2}} & 0 \\ \frac{1}{\sqrt{2}} & -\frac{1}{\sqrt{2}} & 0 \\ 0 & 0 & 1 \\ \end{matrix} \right).$$

Therefore, the new lattice vectors are:

$${\vec{a}_1}^{'}=(0,\ -\frac{\sqrt{2}}{2}, \frac{1}{2}), {\vec{a}_2}^{'}=(\frac{\sqrt{2}}{2},\ 0,\ \frac{1}{2}), {\vec{a}_3}^{'}=(\frac{\sqrt{2}}{2},\ -\frac{\sqrt{2}}{2},\ 0).$$

Finally, the new reciprocal lattice basis of FCC unit-cell can be derived as:

$$\begin{aligned}&\vec{r}_{b1} =\frac{2\pi}{\Omega}({\vec{a}_2}^{'} \times {\vec{a}_3}^{'}) =2\pi (\frac{\sqrt{2}}{2}\ \frac{\sqrt{2}}{2}\ -1), \end{aligned}$$
$$\begin{aligned}&\vec{r}_{b2} =\frac{2\pi}{\Omega}({\vec{a}_3}^{'} \times {\vec{a}_1}^{'}) =2\pi (-\frac{\sqrt{2}}{2}\ -\frac{\sqrt{2}}{2}\ -1), \end{aligned}$$
$$\begin{aligned}&\vec{r}_{b3} =\frac{2\pi}{\Omega}({\vec{a}_1}^{'} \times {\vec{a}_2}^{'}) =2\pi (-\frac{\sqrt{2}}{2}\ \frac{\sqrt{2}}{2}\ 1), \end{aligned}$$
and the correspondence of the high symmetry points between FCC and TET Brillouin zone are list as follows: point $Z_0$, $X_0$, $W_0$ in the Brillouin zone under the FCC scheme refer to the point $\Gamma$, $M$, $A$ in the Brillouin zone under the TET scheme, respectively.

Figure 5 shows the photonic bands below the fundamental gaps of woodpile in scheme of FCC and TET unit-cells, respectively. It should be noted there are four bands below the fundamental gap in the TET scheme, while there are only two bands in FCC scheme. In spite of the difference of the amount of the bands in two schemes, we remarked that the equivalence of these two schemes can be verified by band folding analysis. The fourfold degeneracy at the $Z$ point in the scheme of TET unit-cell originates from the band folding.

 figure: Fig. 5.

Fig. 5. Correspondence of photonic bands of woodpile between two schemes: TET (blue-star) and FCC (black line) unit-cell, respectively. The parameter setting are list as follows: $l_x=l_y=0.5,w_x=w_y=0.25,\epsilon =13$.

Download Full Size | PDF

Appendix B: Other Kramers degeneracies

The Kramers degeneracy on the $k_i=\pi (i=x,y)$ plane can be understood via the construction of the anti-unitary operators $\Theta _x\equiv S_x*{\cal T}: (x,y,z, t)\rightarrow \left (x+\frac {1}{2},-y+\frac {1}{2},-z-\frac {1}{4}, -t\right )$ and $\Theta _y\equiv S_y*{\cal T}: (x,y,z,t) \rightarrow \left (-x+\frac {1}{2},y+\frac {1}{2},-z+\frac {1}{4}, -t\right )$, respectively, where $S_{x} : (x,y,z)\rightarrow \left (x+\frac {1}{2},-y+\frac {1}{2},-z-\frac {1}{4}\right )$ and $S_y : (x,y,z) \rightarrow \left (-x+\frac {1}{2},y+\frac {1}{2},-z+\frac {1}{4}\right )$ are the two-fold screw symmetries. Let us consider the anti-unitary operator $\Theta _x$ as an example. $\Theta _x$ transforms ${\vec k}$ to $(-k_x, k_y, k_z)$ is a symmetry operator on the $k_x=\pi$ plane. The square of this operator is $\left (\Theta _{x}\right )^{2}: (x,y,z,t) \rightarrow (x+1,y,z,t)$, which yields $\left (\Theta _{x}\right )^{2} =e^{ik_x}$ for all photonic Bloch states(including both the electric field ${\vec E}$ and the magnetic field ${\vec H}$). Hence, on the $k_x=\pi$ plane we have $\Theta _x^{2}=-1$ which explains the double degeneracy on the $k_x=\pi$ plane. Specifically, there exist the following commutation relations between anti-unitary operator $\Theta _{x}$ and operator $M_x$ for the $k_x=\pi$ plane,

$$M_x\Theta_x ={-} \Theta_x M_x .$$

Thus for an eigenstate of $M_x$ with eigenvalue $m_x$, labeled as $| {m_x}\rangle$, we have

$$M_x\Theta_x|{m_x}\rangle ={-} \Theta_x M_x|{m_x}\rangle ={-} m_x \Theta_x |{m_x}\rangle .$$

Hence $\Theta _x| {m_x}\rangle$ is also an eigenstate of $M_x$ with opposite eigenvalue. These are the two degenerate Bloch states on the $k_x=\pi$ plane.

In a similar way, the anti-unitary operator $\Theta _y$ is a symmetry operator on the $k_y=\pi$ plane. The square of this operator $\left (\Theta _{y}\right )^{2}: (x,y,z,t) \rightarrow (x,y+1,z,t)$ yields $\Theta _y^{2} = e^{i k_y}$ for all Bloch states. Thus on the $k_y=\pi$ plane we have $\Theta _y^{2} =-1$, which yields the Kramers degeneracy.

Moreover, we find that $\Theta _{gx}\equiv G_x*{\cal T} : (x, y, z, t) \to (-x, y + \frac {1}{2}, z + \frac {1}{2}, -t)$ and $\Theta _{gy}\equiv G_y*{\cal T} : (x, y, z, t) \to (x + \frac {1}{2}, -y, z + \frac {1}{2}, -t)$ can also yield Kramers degeneracy, where $G_x : (x, y, z) \to (-x, y + \frac {1}{2}, z + \frac {1}{2})$ and $G_y : (x, y, z) \to (x + \frac {1}{2}, - y , z + \frac {1}{2})$ are two glide symmetries. Since $\Theta _{gx}$ transforms ${\vec k}$ into $(k_x,-k_y,-k_z)$, it is a symmetry operator only for the four lines: $k_y, k_z=0, \pi$. The square of this operator $\left (\Theta _{y}\right )^{2}: (x,y,z,t) \rightarrow (x,y+1,z+1,t)$ yields $\Theta _{gx}^{2}=e^{i(k_y+k_z)}$ for all Bloch states. Therefore it leads to double degeneracy for the two lines: $(k_x, 0, \pi )$ and $(k_x, \pi, 0)$. Similarly, $\Theta _{gy}$ is a symmetry operator for the four lines: $k_x, k_z=0,\pi$. The square of this operator $\left (\Theta _{y}\right )^{2}: (x,y,z,t) \rightarrow (x,y+1,z+1,t)$ yields $\Theta _{gy}^{2}=e^{i(k_x+k_z)}$ which reuslt in Kramers degeneracy on the two lines: $(0, k_y, \pi )$ and $(\pi, k_y, 0)$. Notice that these lines crossing at the $Z$ point $(0,0,\pi )$, where both $\Theta _{gx}$ and $\Theta _{gy}$ are effective.

Appendix C: Proof of the fourfold degeneracy on the $M$-$A$ line

The fourfold degeneracy on the $M$-$A$ line can be understood as follows: Any Bloch state on this line can be labeled with the eigenvalues $m_x$ and $m_y$ of the two mirror operators $M_x$ and $M_y$, respectively. We shall prove that $|{m_x,m_y}\rangle$, $\Theta _z|{m_x,m_y}\rangle$ ($\Theta _z\equiv G_z*{\cal T}$), $S_\pi |{m_x,m_y}\rangle$, and $\Theta _zS_\pi |{m_x,m_y}\rangle$ are distinct from each other. We find that such degeneracy is essentially related to the following commutation relationships on the $M$-$A$ line,

$$\begin{aligned}& [M_x, \Theta_z]_{+} = 0, \quad [M_y, \Theta_z]_{-}=0,\\ & [M_x, S_\pi]_{+} = 0 ,\quad [M_y, S_\pi]_{+} = 0 , \end{aligned}$$
where $[A, B]_{\pm } = AB \pm BA$. Thus for an eigenstate labeled with $m_x$ and $m_y$ of mirror operator $M_x$ and $M_y$, we have
$$M_x \Theta_z |{m_x,m_y}\rangle={-}\Theta_z M_x|{m_x,m_y}\rangle={-}m_x \Theta_z |{m_x,m_y}\rangle,$$
$$M_y \Theta_z |{m_x,m_y}\rangle=\Theta_z M_y|{m_x,m_y}\rangle=m_y \Theta_z |{m_x,m_y}\rangle,$$
$$M_x S_{\pi} |{m_x,m_y}\rangle={-}S_{\pi} M_x|{m_x,m_y}\rangle={-}m_x S_{\pi} |{m_x,m_y}\rangle,$$
$$M_y S_{\pi} |{m_x,m_y}\rangle={-}S_{\pi} M_y|{m_x,m_y}\rangle={-}m_y S_{\pi} |{m_x,m_y}\rangle.$$

From these relations, we find that: (i) $\Theta _z |{m_x,m_y}\rangle$ is also an eigenstates of $M_x$ with opposite eigenvalue, i.e., $\Theta _z|{m_x,m_y}\rangle = | {-m_x,m_y}\rangle$; (ii) $S_{\pi } |{m_x,m_y}\rangle$ is also an eigenstates of $M_x (M_y)$ with opposite eigenvalue, i.e., $S_\pi |{m_x,m_y}\rangle=| {-m_x,-m_y}\rangle$. (ii) $\Theta _z S_\pi$ is also an eigenstates of $M_y$ with opposite eigenvalue, i.e., $\Theta _z S_\pi |{m_x,m_y}\rangle = | {m_x,-m_y}\rangle$. It is evident that these four states, i.e., $|{m_x,m_y}\rangle$, $\Theta _z|{m_x,m_y}\rangle$, $S_\pi |{m_x,m_y}\rangle$, and $\Theta _zS_\pi |{m_x,m_y}\rangle$ are distinct from each other and they are related by symmetry operators. Therefore, they are degenerate states.

Appendix D: ${\vec k}\cdot {\vec P}$ Hamiltonian of the Dirac nodal line

The photonic Hamiltonian is obtained by applying the ${\vec k}\cdot {\vec P}$ method to the Maxwell equation,

$$\boldsymbol{\nabla}\times\frac{1}{\varepsilon}\boldsymbol{\nabla}\times {\vec H}_{n^{\prime},{\vec k}} = \frac{\omega_{n^{\prime},{\vec k}}^{2}}{c^{2}} {\vec H}_{n^{\prime},{\vec k}} .$$

The photonic Hamiltonian $\hat {{\cal H}}_{EM}\equiv c^{2} \boldsymbol {\nabla }\times \frac {1}{\varepsilon }\boldsymbol {\nabla }\times$ around a high symmetry ${\vec K}$ is obtained as follows. We first expand the Bloch states at a wavevector ${\vec k}$ in the basis formed by the four Bloch states at the high symmetry point, i.e., (${\vec q}\equiv {\vec k}-{\vec K}$)

$${\vec H}_{n^{\prime}, {\vec q}} = \sum_{n} e^{i{\vec q}\cdot{\vec r}} C_n {\vec H}_{n^{\prime}} ,$$
where ${\vec H}_{n^{\prime }}$ is the magnetic field at the ${\vec q}=0$ point (i.e., the high symmetry point) and $C_n$ are the coefficients to be solved by diagonalizing the Hamiltonian (and normalized as $\sum _n|C_n|^{2}=1)$. The magnetic fields are normalized such that (UC stands for unit-cell)
$$\int_{UC} d{\vec r} {\vec H}_{n}^{{\ast}} \cdot {\vec H}_{n^{\prime}} = \delta_{n,n^{\prime}} .$$

The Hamiltiona is written in the basis of the Bloch states ${\vec H}_{n^{\prime }}$, and we find that

$$\hat{{\cal H}}_{EM} = \omega_0^{2}\delta_{n,n^{\prime}} +\sum_{\alpha} q_\alpha P_{n,n^{\prime}}^{\alpha} + \sum_{\alpha,\beta} W_{n\alpha,n^{\prime}\beta}q_\alpha q_\beta .$$

Direct calculation yields,

$$\begin{aligned}& P_{n,n^{\prime}}^{\alpha} = i \sum_\nu \int_{UC}\frac{d{\vec r}}{\varepsilon({\vec r})} [H^{{\ast}}_{n,\nu} \partial_\nu H_{n^{\prime},\alpha} - H_{n^{\prime},\nu} \partial_\nu H_{n,\alpha}^{{\ast}}\\ & \quad\quad \quad\quad -H^{{\ast}}_{n,\nu} \partial_\alpha H_{n^{\prime},\nu} + H_{n^{\prime},\nu} \partial_\alpha H^{{\ast}}_{n,\nu} ] . \end{aligned}$$

We shall use the Maxwell equation to simplify the above results

$$\partial_\nu H_{n,\alpha} = \sum_\mu - i \frac{\omega_n}{c} E_{n,\mu}\epsilon_{\nu\alpha\mu} \varepsilon({\vec r}),$$
where $\omega _n=\omega _0$ at the degenerate point, $\epsilon _{\nu \alpha \mu }$ is the Levi-Civita tensor, ${\vec E}_{n}$ is the electric field of the Bloch states at ${\vec q}=0$ satisfying the normalization condition of
$$\int_{UC} d{\vec r} \varepsilon({\vec r}) {\vec E}_{n}^{{\ast}} \cdot {\vec E}_{n^{\prime}} = \delta_{n,n^{\prime}} .$$

Using Eq. (24), we find that

$$P_{n,n^{\prime}}^{\alpha} = 2\omega_0 c\int_{UC}d{\vec r} [{\vec E}_{n^{\prime}}\times {\vec H}_{n}^{{\ast}} + {\vec E}_{n}^{{\ast}} \times {\vec H}_{n^{\prime}}]\cdot {\vec n}_\alpha$$
where $\alpha =(x,y,z)$ and ${\vec n}_{\alpha }$ is the unit vector along the $\alpha$ direction. And
$$W_{n\alpha,n^{\prime}\beta} = c^{2} \int_{UC} \frac{d{\vec r}}{\varepsilon({\vec r})} [\delta_{\alpha\beta} ({\vec H}_n^{{\ast}}\cdot{\vec H}_{n^{\prime}}) - H_{n\alpha}^{{\ast}} H_{n^{\prime}\beta}] .$$

The photonic Hamiltonian $\hat {{\cal H}}_{EM}$ is connected with the simulated fermion-like as follows, ($\hbar \equiv 1$)

$$\hat{{\cal H}}_{EM} := (\hat{{\cal H}}_F)^{2} .$$

Therefore, near the degenerate point, we have

$$\hat{{\cal H}}_F = \omega_0 + \sum_\alpha \hat{v}^{\alpha} q_\alpha + \sum_{\alpha,\beta} \hat{w}_{\alpha,\beta}q_\alpha q_\beta + \cdots,$$
where $\cdots$ represents higher order terms, and
$$\begin{aligned}& \hat{v}^{\alpha} = \frac{1}{2\omega_0} P_{n,n^{\prime}}^{\alpha} ,\\ & \hat{w}_{\alpha,\beta} = \frac{1}{2\omega_0} \hat{W}_{\alpha,\beta} . \end{aligned}$$

One can easily verify that $\hat {v}^{\alpha }$ and $\hat {w}_{\alpha,\beta }$ are Hermitian operators.

We now examine the constraints on the matrix element of $\hat {v}$ imposed by the mirror and/or glide symmetry. Explicitly, we have

$$v_{n,n^{\prime}}^{\alpha} = c \sum_{\mu,\nu} \epsilon_{\alpha \mu\nu} \int_{UC}d{\vec r} [ E_{n^{\prime},\mu} H_{n,\nu}^{{\ast}} + E_{n,\mu}^{{\ast}} H_{n^{\prime},\nu}] .$$

If, say, there is a mirror or glide symmetry, labeled as $\hat {F}_{x}$, the mirror operation acts on the electric and magnetic fields as follows:

$$\begin{aligned}& \hat{F}_x H_x({\vec r}) = H_x(\hat{F}_x{\vec r}) ,\\ & \hat{F}_x H_y({\vec r}) ={-} H_y(\hat{F}_x{\vec r}) ,\\ & \hat{F}_x H_z({\vec r}) ={-} H_z(\hat{F}_x{\vec r}) ,\\ & \hat{F}_x E_x({\vec r}) ={-} E_x(\hat{F}_x{\vec r}) ,\\ & \hat{F}_x E_y({\vec r}) = E_y(\hat{F}_x{\vec r}) ,\\ & \hat{F}_x E_z({\vec r}) = E_z(\hat{F}_x{\vec r}) . \end{aligned}$$

Thus, the velocity matrix element $v_{n,n^{\prime }}^{\alpha }$ is nonzero only when the $n$ and $n^{\prime }$ states carry opposite mirror eigenvalue $m_\alpha$ (or glide eigenvalue $g_\alpha$).

The invariance of the Hamiltonian under a symmetry operation ${\cal S}$ implies that

$${\cal S} {\cal H}({\vec k}) {\cal S}^{{-}1} = {\cal H}({\cal S}{\vec k}) .$$

For example, the time-reversal symmetry symmetry ${\cal T}=-{\cal K}$ yields that the ${\vec k}\cdot {\vec P}$ around a time-reversal invariant momentum has

$$(v_{n,n^{\prime}}^{\alpha} )^{{\ast}} ={-} v_{n,n^{\prime}}^{\alpha}.$$

Hence the matrix element of the $q$ linear terms are purely imaginary. One can prove that the $q$ quadratic terms are purely real.

We now derive the ${\vec k}\cdot {\vec P}$ Hamiltonian for the Dirac nodal line. Around a point on the $M$-$A$ line, the four degenerate states are chosen as the $|{m_x,m_y}\rangle$ for $m_x,m_y=\pm 1$. In the basis of $(| {1,1}\rangle, |{-1,-1}\rangle, |{-1,1}\rangle, |{1,-1}\rangle)^{T}$ the $q$ linear Hamiltonian is written as

$$\hat{{\cal H}}_F^{DL} =\omega_0 + \left( \begin{array}{cccc} 0 & 0 & a_1 q_x & a_2 q_y \\ 0 & 0 & b_1 q_y & b_2 q_x \\ a_1^{{\ast}} q_x & b_1^{{\ast}} q_y & 0 & 0 \\ a_2^{{\ast}} q_y & b_2^{{\ast}} q_x & 0 & 0 \end{array}\right) ,$$
where $a_i$ and $b_i$ are the ${\vec k}\cdot {\vec P}$ coefficients which can be calculated from the photonic Bloch functions ${\vec H}_{n}$ at the degeneracy point $(\pi,\pi,k_z)$ (hence these coefficients are $k_z$ dependent). Those coefficients are restricted by the symmetry via Eq. (33). Two symmetry operations are relevant: $\Theta _z$ and $S_\pi$. In the chosen basis, these operations are manifested as the following matrices:
$$\Theta_z = \tau_y {\cal K}, \quad S_\pi = \sigma_y e^{ik_z z/2} .$$

Their effects on the wavevectors are

$$\Theta_z {\vec k} = ({-}k_x, -k_y, k_z), \quad S_\pi {\vec k} = ( - k_x, - k_y, k_z) .$$

We thus obtain from Eq. (33) that

$$b_1 = a_2, \quad b_2 ={-} a_1 .$$

It is convenient to define $v=\sqrt {|b_1|^{2}+|b_2|^{2}}=\sqrt {|a_1|^{2}+|a_2|^{2}}$, $\gamma _x=a_1/v$, and $\gamma _y=a_2/v$. In the new basis of $(| {\uparrow,p}\rangle, | {\downarrow,p}\rangle, | {\uparrow,a}\rangle, | {\downarrow,a}\rangle)^{T}$, the ${\vec k}\cdot {\vec P}$ Hamiltonian can be further simplified as

$$\begin{aligned}& \hat{{\cal H}}_F^{DL} = \omega_0 + v \left( \begin{array}{cc} 0 & \hat{{\cal A}} \\ \hat{{\cal A}}^{{\dagger}} & 0 \end{array}\right) + {\cal O}(q^{2}) ,\\ & \hat{{\cal A}} \equiv \gamma_x q_x \sigma_x + \gamma_y q_y \sigma_y . \end{aligned}$$

The above Hamiltonian applies for the whole $M$-$A$ line, where the coefficients $\omega _0$, $v$, $\gamma _x$ and $\gamma _y$ are $k_z$-dependent. At the M or A point, the $S_4$ symmetry also holds, thus there will be additional constraints on the coefficients. Let’s consider such constraints in the original basis for the Hamiltonian Eq. (35). $S_4$ is manifested as $-\sigma _0$ in the even-parity doublet, whereas $S_4=-i\sigma _y$ in the odd-parity doublet. Imposing (33) we find that for the M and A points, $\gamma _x=\gamma _y$. In addition, for these points, the time-reversal symmetry dictates that $\gamma _x=\gamma _y$ are purely imaginary coefficients. Therefore, at these points the Dirac point is doubly degenerate in the $k_x$-$k_y$ plane, while away from these points the dispersion in the $k_x$-$k_y$ plane is generally non-degenerate. The double degeneracy for generic points on the MA line (except the M and A points) are restored only for $q_x=0$ or $q_y=0$ (i.e., at the $k_x=\pi$ or $k_y=\pi$ plane), where the nonsymmorphic screw symmetry ensures double degeneracy.

For the quadratic degeneracy at the $Z$ point, the eigenstates can be labeled as $| {m_x,m_y,g_z}\rangle$ for $m_x=-m_y=\pm 1, g_z=\pm 1$. Using the basis of $(| {-1,1,1}\rangle, | {1,-1,1}\rangle, | {-1,1,-1}\rangle, | {1,-1,-1}\rangle)^{T}$, the $q$ linear term comes simply as $v_z q_z\tau _y$, because it is finite only between states with opposite $g_z$. Considering the $S_{\frac {\pi }{2}}$ symmetry, the higher order terms are

$$f_0 f_2 q_{{\parallel}}^{2} + f_0 \left( \begin{array}{cccc} q_x^{2} - q_y^{2} & 2 f_1 q_x q_y & 0 & 0 \\ 2f_1 q_x q_y & q_y^{2}-q_x^{2} & 0 & 0 \\ 0 & 0 & q_x^{2} - q_y^{2} & 2f_1 q_x q_y \\ 0 & 0 & 2f_1 q_x q_y & q_y^{2} - q_x^{2} \end{array}\right) .$$

The above Hamiltonian recovers Eq. (8) in the main text when written in the basis of spin states that carrying angular momentum. When the $S_{\frac {\pi }{2}}$ symmetry is broken the degeneracy between states with different $m_x$ at $q_z=0$ is split by a constant term $\Delta _zf_0\sigma _z$ with $\Delta _z$ characterizing the strength of the perturbation.

Appendix E: Proof of the fourfold quadratic degeneracy on the $Z$ point

The $S_{\frac {\pi }{2}}$ operator transforms the eigenstate of $M_x$ with the eigenvalue $m_x$ at ${\vec k}$ to a Bloch state at ${\vec k}^{\prime }$ as the eigenstate of $M_y$ with the same eigenvalue, i.e.,

$$\Theta_{\frac{\pi}{2}}M_x\Theta_{-\frac{\pi}{2}}=M_y .$$

Thus $\Theta _{\frac {\pi }{2}}$ is a symmetry operator only for the $Z$ and $A$ points. Analogous to the Kramers degeneracy, the above yields fourfold degeneracy at the $Z$ point. The four degenerate states are $| {\Psi }\rangle$, $\Theta _{\frac {\pi }{2}}| {\Psi }\rangle$, $(\Theta _{\frac {\pi }{2}})^{2}| {\Psi }\rangle$, and $(\Theta _{\frac {\pi }{2}})^{3}| {\Psi }\rangle$.

The four degenerate states can be labeled by the eigenvalue of $M_x$, $M_y$ and $\tilde {G}_z=e^{i\frac {\pi }{4}}G_z$. We find that

$$[\Theta_{\frac{\pi}{2}}, \tilde{G}_z]_{-} = 0, \quad [(\Theta_{\frac{\pi}{2}})^{2}, \tilde{G}_z]_{+} =0 ,$$
$$[(\Theta_{\frac{\pi}{2}})^{2}, M_x]_{-} = 0, \quad [(\Theta_{\frac{\pi}{2}})^{2}, M_y]_{-} = 0 .$$

Thus $| {\Psi }\rangle$ and $\Theta _{\frac {\pi }{2}}| {\Psi }\rangle$ carry distinct mirror eigenvalues but the same $\tilde {G}_z$ eigenvalue, whereas $| {\Psi }\rangle$ and $(\Theta _{\frac {\pi }{2}})^{2}| {\Psi }\rangle$ carry the same mirror eigenvalues but different $\tilde {G}_z$ eigenvalues. For simplify, we label the Bloch states, which is the eigenstates of $M_x$, $M_y$ and $G_z$ with eigenvalue $m_x$, $m_y$, and $g_z$, respectively, as $| {m_x,m_y,g_z}\rangle$. According to the above commute (anticommute) relationships, we find that $\Theta _{\frac {\pi }{2}}| {\Psi }\rangle=| {m_y,m_x,g_z}\rangle$, $(\Theta _{\frac {\pi }{2}})^{2}| {\Psi }\rangle=| {m_x,m_y,-g_z}\rangle$, and $(\Theta _{\frac {\pi }{2}})^{3}| {\Psi }\rangle=| {m_y,m_x,-g_z}\rangle$. It is obviously that these four states, i.e., $| {m_x,m_y,g_z}\rangle$, $| {m_y,m_x,g_z}\rangle$, $| {m_x,m_y,-g_z}\rangle$, and $| {m_y,m_x,-g_z}\rangle$ are distinct from each other and they are related by symmetry operations. Therefore, we prove the fourfold degeneracy at $Z$ point.

Appendix F: Varying geometries and materials

Here we show the emergent Dirac physics are robust to the material and geometry of the photonic crystals. We calculate the photonic bands along the high symmetry lines with three different geometric/material parameter settings. The results of the photonic bands are shown in Fig. 6, where the Dirac nodal line and the quadratic point holds for all parameters. This is because such double degeneracy are guranteed the lattice symmetry. This also indicates the stableness of the Dirac points and implies that such a symmetry-guided method can also be effective in other classicl/bosonic systems.

 figure: Fig. 6.

Fig. 6. Photonic bands along the high symmetry lines for various geometry/material parameters. (a) $l_x=l_y=0.5,w_x=w_y=0.2,h=0.25,\epsilon =8$, (b) $l_x=l_y=0.5,w_x=w_y=0.2,h=0.15,\epsilon =11$, (c) $l_x=l_y=0.5,w_x=w_y=0.1,h=0.25,\epsilon =11$.

Download Full Size | PDF

Appendix G: Optical properties of type-II Dirac points

The type-II DPs offer special band structures that may enable in the manipulation of light in ways that cannot be achieved in uniform dielectric materials. The refraction properties can be determined by matching the frequency and the wavevector parallel to the interface. Here, we consider refraction on the $(001)$ surface where light is injected from air (above the PC). The wavevector in the air is determined by the frequency and angle of incidence ($\theta _i$ and $\phi _i$) via

$$k_x = \frac{\omega}{c} \sin\theta_i\cos\phi_i,$$
$$k_y =\frac{\omega}{c} \sin\theta_i\sin\phi_i,$$
$$k_z = \frac{\omega}{c} \cos\theta_i .$$

The wavevector $k_x$ and $k_y$ remains the same in the PC, the quantity to be found is the wavevector along $z$ direction in the PC. We shall denote the wavevector in the PC as ${\vec q}$. So far we have

$$q_x = k_x, \quad q_y = k_y .$$

For the Dirac points, $q_z$ is obtained by solving the equation,

$$\delta \omega = v_z q_z\tau_z + f_0 \Bigg[\beta q_\parallel^{2} \pm \sqrt{(\Delta_z +q_x^{2} -q_y^{2})^{2} + 4 q_x^{2} q_y^{2} }\bigg],$$
for the $\pm$ branches with $\tau _z=-1$ as required by that the group velocity along the $z$ direction should be negative. The above equation can be solved straightforwardly.

The group velocity along the $x$ and $y$ directions are then,

$$v_{g,x} = \frac{\partial\omega}{\partial q_x} = f_0 \bigg(2\beta q_x \pm \frac{2(\Delta_z + q_x^{2} -q_y^{2}) q_x + 4 q_y^{2} q_x}{\sqrt{(\Delta_z +q_x^{2} -q_y^{2})^{2} + 4 q_x^{2} q_y^{2} } }\bigg) ,$$
$$v_{g,y} = \frac{\partial\omega}{\partial q_y} = f_0 \bigg(2\beta q_y \pm \frac{-2(\Delta_z + q_x^{2} -q_y^{2}) q_y + 4 q_x^{2} q_y}{\sqrt{(\Delta_z +q_x^{2} -q_y^{2})^{2} + 4 q_x^{2} q_y^{2} } }\bigg) .$$

The refraction angle $\phi _o$ is then determined as

$$\phi_o = {\rm Arg}[v_{g,x} + i v_{g,y}],$$
for the $\pm$ branches. The crucial physics is that around the type-II Dirac point the $v_{g,x}$ becomes significant for the two branches and they are of opposite sign, while the $v_{g,y}$ do not change significantly. This yields a large variation of the refraction angle across the Dirac point, such variation goes to opposite direction for the two branches.

Funding

National Natural Science Foundation of China (11904060, 12074281, 12125504); Natural Sciences and Engineering Research Council of Canada (06089-2016).

Acknowledgments

HXW and JHJ acknowledge supports from the Jiangsu provincial distinguished professor funding. JHJ also thanks Zhi Hong Hang, Jie Luo, Zhengyou Liu, and Huanyang Chen for many insightful discussions, as well as the University of Toronto and the Weizmann Institute of Science for hospitality. HYK acknowledges support from the Canada Research Chairs program. GYG acknowledge support from the Ministry of Science and Technology as well as National Center for Theoretical Sciences, Taiwan.

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. G. E. Volovik, The Universe in a Helium Droplet, (Oxford University, 2003).

2. X. 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). [CrossRef]  

3. Z. Wang, Y. Sun, X.-Q. Chen, C. Franchini, G. Xu, H. Weng, X. Dai, and Z. Fang, “Dirac semimetal and topological phase transitions in A3Bi (A = Na, K, Rb),” Phys. Rev. B 85(19), 195320 (2012). [CrossRef]  

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). [CrossRef]  

5. A. A. Soluyanov, D. Gresch, Z. Wang, Q. Wu, M. Troyer, X. Dai, and B. A. Bernevig, “Type-II Weyl semimetals,” Nature (London) 527(7579), 495–498 (2015). [CrossRef]  

6. B. Bradlyn, J. Cano, Z. Wang, M. G. Vergniory, C. Felser, R. J. Cava, and B. A. Bernevig, “Beyond Dirac and Weyl fermions: unconventional quasiparticles in conventional crystals,” Science 353(6299), aaf5037 (2016). [CrossRef]  

7. N. P. Armitage, E. J. Mele, and A. Vishwanath, “Weyl and Dirac semimetals in three-dimensional solids,” Rev. Mod. Phys. 90(1), 015001 (2018). [CrossRef]  

8. L. Lu, L. Fu, J. D. Joannopoulos, and M. Soljačić, “Weyl points and line nodes in gyroid photonic crystals,” Nat. Photonics 7(4), 294–299 (2013). [CrossRef]  

9. L. Lu, Z. Wang, D. Ye, L. Ran, L. Fu, J. D. Joannopoulos, and M. Soljačić, “Experimental observation of Weyl points,” Science 349(6248), 622–624 (2015). [CrossRef]  

10. W. Gao, B. Yang, M. Lawrence, F. Fang, B. Béri, and S. Zhang, “Photonic Weyl degeneracies in magnetized plasma,” Nat. Commun. 7(1), 12435 (2016). [CrossRef]  

11. D. Wang, B. Yang, W. Gao, H. Jia, Q. Yang, X. Chen, M. Wei, C. Liu, M. Navarro-Cía, J. Han, W. Zhang, and S. Zhang, “Photonic Weyl points due to broken time-reversal symmetry in magnetized semiconductor,” Nat. Phys. 15(11), 1150–1155 (2019). [CrossRef]  

12. W.-J. Chen, M. Xiao, and C. T. Chan, “Photonic crystals possessing multiple Weyl points and the experimental observation of robust surface states,” Nat. Commun. 7(1), 13038 (2016). [CrossRef]  

13. H.-X. Wang, L. Xu, H. Y. Chen, and J.-H. Jiang, “Three-dimensional photonic Dirac points stabilized by point group symmetry,” Phys. Rev. B 93(23), 235155 (2016). [CrossRef]  

14. M. Xiao, Q. Lin, and S. Fan, “Hyperbolic Weyl point in reciprocal chiral metamaterials,” Phys. Rev. Lett. 117(5), 057401 (2016). [CrossRef]  

15. Q. Lin, M. Xiao, L. Yuan, and S. Fan, “Photonic Weyl point in a two-dimensional resonator lattice with a synthetic frequency dimension,” Nat. Commun. 7(1), 13731 (2016). [CrossRef]  

16. S. Peng, R. Zhang, V. H. Chen, E. T. Khabiboulline, P. Braun, and H. A. Atwater, “Three-Dimensional Single Gyroid Photonic Crystals with a Mid-Infrared Bandgap,” ACS Photonics 3(6), 1131–1137 (2016). [CrossRef]  

17. H.-X. Wang, Y. Chen, Z. H. Hang, H.-Y. Kee, and J.-H. Jiang, “Type-II Dirac photons,” npj Quantum Mater. 2(1), 54 (2017). [CrossRef]  

18. J. Noh, S. Huang, D. Leykam, Y. D. Chong, K. P. Chen, and M. C. Rechtsman, “Experimental observation of optical Weyl points and Fermi arc-like surface states,” Nat. Phys. 13(6), 611–617 (2017). [CrossRef]  

19. E. Goi, Z. Yue, B. P. Cumming, and M. Gu, “Observation of type-I Photonic Weyl points in optical frequencies,” Laser Photonics Rev. 12(2), 1700271 (2018). [CrossRef]  

20. S. Vaidva, J. Noh, A. Cerian, C. Jörg, G. v. Freymann, and M. C. Rechtsman, “Observation of a charge-2 photonic Weyl point in the infrared,” Phys. Rev. Lett. 125(25), 253902 (2020). [CrossRef]  

21. C. Jörg, S. Vaidva, J. Noh, A. Cerian, S. Augustine, G. v. Freymann, and M. C. Rechtsman, “Observation of quadratic (charge-2) Weyl point splitting in near-infrared photonic crystals,” Laser Photonics Rev. 16(1), 2100452 (2022). [CrossRef]  

22. Q. Guo, B. Yang, L. Xia, W. Gao, H. Liu, J. Chen, Y. Xiang, and S. Zhang, “Three dimensional photonic Dirac points in metamaterials,” Phys. Rev. Lett. 119(21), 213901 (2017). [CrossRef]  

23. M.-L. Chang, M. Xiao, W.-J. Chen, and C. T. Chan, “Multi Weyl points and the sign change of their topological charges in woodpile photonic crystals,” Phys. Rev. B 95(12), 125136 (2017). [CrossRef]  

24. X. Zhang, “Observing Zitterbewegung for Photons near the Dirac Point of a Two-Dimensional Photonic Crystal,” Phys. Rev. Lett. 100(11), 113903 (2008). [CrossRef]  

25. R. A. Sepkhanov, Y. B. Bazaliy, and C. W. J. Beenakker, “Extremal transmission at the Dirac point of a photonic band structure,” Phys. Rev. A 75(6), 063813 (2007). [CrossRef]  

26. X. Q. Huang, Y. Lai, Z. H. Hang, H. H. Zheng, and C. T. Chan, “Dirac cones induced by accidental degeneracy in photonic crystals and zero-refractive-index materials,” Nat. Mater. 10(8), 582–586 (2011). [CrossRef]  

27. M. C. Rechtsman, J. M. Zeuner, A. Tünnermann, S. Nolte, M. Segev, and A. Szameit, “Strain-induced pseudomagnetic field and photonic Landau levels in dielectric structures,” Nat. Photonics 7(2), 153–158 (2013). [CrossRef]  

28. F. D. M. Haldane and S. Raghu, “Possible Realization of Directional Optical Waveguides in Photonic Crystals with Broken Time-Reversal Symmetry,” Phys. Rev. Lett. 100(1), 013904 (2008). [CrossRef]  

29. Z. Wang, Y. D. Chong, J. D. Joannopoulos, and M. Soljačić, “Reflection-Free One-Way Edge Modes in a Gyromagnetic Photonic Crystal,” Phys. Rev. Lett. 100(1), 013905 (2008). [CrossRef]  

30. Z. Wang, Y. Chong, J. D. Joannopoulos, and M. Soljačić, “Observation of unidirectional backscattering immune topological electromagnetic states,” Nature 461(7265), 772–775 (2009). [CrossRef]  

31. M. Hafezi, S. Mittal, J. Fan, A. Migdall, and J. M. Taylor, “Imaging topological edge states in silicon photonics,” Nat. Photonics 7(12), 1001–1005 (2013). [CrossRef]  

32. M. C. Rechtsman, J. M. Zeuner, Y. Plotnik, Y. Lumer, D. Podolsky, F. Dreisow, S. Nolte, M. Segev, and A. Szameit, “Photonic Floquet topological insulators,” Nature 496(7444), 196–200 (2013). [CrossRef]  

33. A. B. Khanikaev, S. H. Mousavi, W.-K. Tse, M. Kargarian, A. H. MacDonald, and G. Shvets, “Photonic topological insulators,” Nat. Mater. 12(3), 233–239 (2013). [CrossRef]  

34. W.-J. Chen, S.-J. Jiang, X.-D. Chen, J.-W. Dong, and C. T. Chan, “Experimental realization of photonic topological insulator in a uniaxial metacrystal waveguide,” Nat. Commun. 5(1), 5782 (2014). [CrossRef]  

35. T. Ma, A. B. Khanikaev, S. H. Mousavi, and G. Shvets, “Guiding electromagnetic waves around sharp corners: topologically protected photonic transport in metawaveguides,” Phys. Rev. Lett. 114(12), 127401 (2015). [CrossRef]  

36. L.-H. Wu and X. Hu, “Scheme for Achieving a Topological Photonic Crystal by Using Dielectric Material,” Phys. Rev. Lett. 114(22), 223901 (2015). [CrossRef]  

37. C. He, X.-C. Sun, X.-P. Liu, M.-H. Lu, Y. Chen, L. Feng, and Y.-F. Chen, “Photonic topological insulator with broken time-reversal symmetry,” Proc. Natl. Acad. Sci. U. S. A. 113(18), 4924–4928 (2016). [CrossRef]  

38. L. Xu, H.-X. Wang, Y. D. Xu, H. Y. Chen, and J.-H. Jiang, “Accidental degeneracy in photonic bands and topological phase transitions in two-dimensional core-shell dielectric photonic crystals,” Opt. Express 24(16), 18059 (2016). [CrossRef]  

39. L. Lu, C. Fang, L. Fu, S. G. Johnson, J. D. Joannopoulos, and M. Soljačić, “Symmetry-protected topological photonic crystal in three dimensions,” Nat. Phys. 12(4), 337–340 (2016). [CrossRef]  

40. A. Slobozhanyuk, S. H. Mousavi, X. Ni, D. Smirnova, Y. S. Kivshar, and A. B. Khanikaev, “Three-dimensional all-dielectric photonic topological insulator,” Nat. Photonics 11(2), 130–136 (2017). [CrossRef]  

41. S. Y. Lin, J. G. Fleming, D. L. Hetherington, B. K. Smith, R. Biswas, K. M. Ho, M. M. Sigalas, W. Zubrzycki, S. R. Kurtz, and J. Bur, “A three-dimensional photonic crystal operating at infrared wavelengths,” Nature (London) 394(6690), 251–253 (1998). [CrossRef]  

42. S. Noda, K. Tomoda, N. Yamamoto, and A. Chutinan, “Full three-dimensional photonic bandgap crystals at near-Infrared wavelengths,” Science 289(5479), 604–606 (2000). [CrossRef]  

43. M. Deubel, G. von Freymann, M. Wegener, S. Pereira, K. Busch, and C. M. Soukoulis, “Direct laser writing of three-dimensional photonic-crystal templates for telecommunications,” Nat. Mater. 3(7), 444–447 (2004). [CrossRef]  

44. S. Noda, M. Fujita, and T. Asano, “Spontaneous-emission control by photonic crystals and nanocavities,” Nat. Photonics 1(8), 449–458 (2007). [CrossRef]  

45. J. D. Joannopoulos, S. G. Johnson, J. N. Winn, and R. D. Meade, Photonic crystals: molding the flow of light, (Princeton University, 2011).

46. L. A. Ibbotson, A. Demetriadou, S. Croxall, O. Hess, and J. J. Baumberg, “Optical nano-woodpiles: large-area metallic photonic crystals and metamaterials,” Sci. Rep. 5(1), 8313 (2015). [CrossRef]  

47. M. I. Shalaev, W. Walasik, A. Tsukernik, Y. Xu, and N. M. Litchinitser, “Robust topologically protected transport in photonic crystals at telecommunication wavelengths,” Nat. Nanotechnol. 14(1), 31–34 (2019). [CrossRef]  

48. P. A. M. Dirac, “The quantum theory of the electron,” Proc. Roy. Soc. A 117(778), 610–624 (1928). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (6)

Fig. 1.
Fig. 1. (a) Optical-frequency woodpile PCs: layer-by-layer stacking of dielectric (colored) logs. (b) Lattice vectors of the undeformed woodpile PC, ${\vec a}_j$, are shown together with those of the deformed woodpile PC, ${\vec b}_j$ ($j=1,2,3$). (c) The top-view (upper) and the side-view (lower) of the unit-cell of the deformed (green) woodpile PC. (d) The relationship between the Brillouin zones of the undeformed (gray) and the deformed woodpile PCs. (e) The low-lying photonic bands of the undeformed woodpile PC with $l_x=l_y=b_x/2$, $w_x=w_y=0.4b_x$, $h=0.2b_x$, and $\varepsilon =5.06$ (TiO$_2$). The connection between photonic bands in the tetragonal and face-centered cubic Brillouin zone is given in Appendix A. Inset: the Dirac dispersion at the $A$ point (labeled by the star). The $M$-$A$ line (red) is a Dirac nodal line on which each point has Dirac-like dispersions.
Fig. 2.
Fig. 2. (a) Field profiles of the four degenerate modes on the $M$-$A$ line. (b) and (c): Dispersion of the Dirac nodal line in the (b) $k_x$-$k_z$ and (c) $k_x$-$k_y$ planes. In (b) an isofrequency plane (the blue-gray sheet) is plotted in order to show the isofrequency contours (the red curves). The Dirac nodal line is labeled by the black curve. Parameters are the same as in Fig. 1.
Fig. 3.
Fig. 3. Quadratic degeneracy points: (a) Photonic bands in the $k_y=0$ plane for the same parameters as in Fig. 1. The orange (blue) band has $m_y=+1$ (−1). The $Z$ point (indicated by the arrow) is a FQP with four eigenstates of different mirror ($m_{x/y}$) and glide ($g_z$) symmetries (illustrated in the inset). (b) and (c): Dispersion of the FQP in the (b) $k_x$-$k_y$ and (c) $k_y$-$k_z$ planes.
Fig. 4.
Fig. 4. (a) Photonic dispersion near a type-II DP for $l_x=l_y=0.5$, $w_x=0.2$, $w_y=0.3$, $h=0.25$, and $\varepsilon =11$. (b) Illustration of the refraction angle measurement. $\theta _i$ and $\phi _i$ are angles of incidence, while $\delta \phi _o$ is the angular difference between the two refraction beams. (c) $\delta \phi _o$ vs. $\theta _i$ and $\omega$ at $\phi _i=89.5^{\circ }$ as a signature of the type-II DP. (d) The group velocity along $x$ direction $v_x$ for the beams correspond to the lower branch as a function of $\phi _i$ for $\theta _i=47.5$ (aligned with the DP) and $\theta _i=45$ (misaligned with the DP).
Fig. 5.
Fig. 5. Correspondence of photonic bands of woodpile between two schemes: TET (blue-star) and FCC (black line) unit-cell, respectively. The parameter setting are list as follows: $l_x=l_y=0.5,w_x=w_y=0.25,\epsilon =13$.
Fig. 6.
Fig. 6. Photonic bands along the high symmetry lines for various geometry/material parameters. (a) $l_x=l_y=0.5,w_x=w_y=0.2,h=0.25,\epsilon =8$, (b) $l_x=l_y=0.5,w_x=w_y=0.2,h=0.15,\epsilon =11$, (c) $l_x=l_y=0.5,w_x=w_y=0.1,h=0.25,\epsilon =11$.

Equations (64)

Equations on this page are rendered with MathJax. Learn more.

Θ π 2 | k z = π b z = 1 ,
Θ z | m x , m y = | m x , m y ,
S π | m x , m y = | m x , m y ,
Θ z S π | m x , m y = | m x , m y .
| p , = 1 2 ( | 1 , 1 + i | 1 , 1 ) | x 2 y 2 + i | 2 x y 2 ,
| p , = 1 2 ( | 1 , 1 i | 1 , 1 ) | x 2 y 2 i | 2 x y 2 ,
| a , = 1 2 ( | 1 , 1 + i | 1 , 1 ) | x + i | y 2 ,
| a , = 1 2 ( | 1 , 1 i | 1 , 1 ) | x i | y 2 .
H ^ E M := ( H ^ F ) 2 .
H ^ F D L = ω 0 + v ( 0 A ^ A ^ 0 ) ,
v γ ~ j = c N E N H U C d r   ( E , p × H , a + E , a × H , p ) n j ,
( Θ π 2 ) 4 = e i k z b z | k z = π b z = 1 ,
Θ π 2 | Ψ = | m y , m x , g z ,
( Θ π 2 ) 2 | Ψ = | m x , m y , g z ,
( Θ π 2 ) 3 | Ψ = | m y , m x , g z .
H ^ F Z = ω Z + v z q z τ ^ y + f 0 [ ( q x 2 q y 2 ) σ ^ z + 2 f 1 q x q y σ ^ x + f 2 q 2 ] ,
H ^ F D P ± = ω D ± v 0 δ k y + v z δ k z τ ^ y ± v x δ k x σ ^ x ± v y δ k y σ z ,
b 1 = ( 1 ,   0 ,   0 ) , b 2 = ( 0 ,   1 ,   0 ) , b 3 = ( 0 ,   0 ,   1 ) .
a 1 = 1 2 ( b 1 + b 2 + b 3 ) , a 2 = 1 2 ( b 1 + b 2 + b 3 ) , a 3 = b 2 .
( 1 2 1 2 0 1 2 1 2 0 0 0 1 ) .
a 1 = ( 0 ,   2 2 , 1 2 ) , a 2 = ( 2 2 ,   0 ,   1 2 ) , a 3 = ( 2 2 ,   2 2 ,   0 ) .
r b 1 = 2 π Ω ( a 2 × a 3 ) = 2 π ( 2 2   2 2   1 ) ,
r b 2 = 2 π Ω ( a 3 × a 1 ) = 2 π ( 2 2   2 2   1 ) ,
r b 3 = 2 π Ω ( a 1 × a 2 ) = 2 π ( 2 2   2 2   1 ) ,
M x Θ x = Θ x M x .
M x Θ x | m x = Θ x M x | m x = m x Θ x | m x .
[ M x , Θ z ] + = 0 , [ M y , Θ z ] = 0 , [ M x , S π ] + = 0 , [ M y , S π ] + = 0 ,
M x Θ z | m x , m y = Θ z M x | m x , m y = m x Θ z | m x , m y ,
M y Θ z | m x , m y = Θ z M y | m x , m y = m y Θ z | m x , m y ,
M x S π | m x , m y = S π M x | m x , m y = m x S π | m x , m y ,
M y S π | m x , m y = S π M y | m x , m y = m y S π | m x , m y .
× 1 ε × H n , k = ω n , k 2 c 2 H n , k .
H n , q = n e i q r C n H n ,
U C d r H n H n = δ n , n .
H ^ E M = ω 0 2 δ n , n + α q α P n , n α + α , β W n α , n β q α q β .
P n , n α = i ν U C d r ε ( r ) [ H n , ν ν H n , α H n , ν ν H n , α H n , ν α H n , ν + H n , ν α H n , ν ] .
ν H n , α = μ i ω n c E n , μ ϵ ν α μ ε ( r ) ,
U C d r ε ( r ) E n E n = δ n , n .
P n , n α = 2 ω 0 c U C d r [ E n × H n + E n × H n ] n α
W n α , n β = c 2 U C d r ε ( r ) [ δ α β ( H n H n ) H n α H n β ] .
H ^ E M := ( H ^ F ) 2 .
H ^ F = ω 0 + α v ^ α q α + α , β w ^ α , β q α q β + ,
v ^ α = 1 2 ω 0 P n , n α , w ^ α , β = 1 2 ω 0 W ^ α , β .
v n , n α = c μ , ν ϵ α μ ν U C d r [ E n , μ H n , ν + E n , μ H n , ν ] .
F ^ x H x ( r ) = H x ( F ^ x r ) , F ^ x H y ( r ) = H y ( F ^ x r ) , F ^ x H z ( r ) = H z ( F ^ x r ) , F ^ x E x ( r ) = E x ( F ^ x r ) , F ^ x E y ( r ) = E y ( F ^ x r ) , F ^ x E z ( r ) = E z ( F ^ x r ) .
S H ( k ) S 1 = H ( S k ) .
( v n , n α ) = v n , n α .
H ^ F D L = ω 0 + ( 0 0 a 1 q x a 2 q y 0 0 b 1 q y b 2 q x a 1 q x b 1 q y 0 0 a 2 q y b 2 q x 0 0 ) ,
Θ z = τ y K , S π = σ y e i k z z / 2 .
Θ z k = ( k x , k y , k z ) , S π k = ( k x , k y , k z ) .
b 1 = a 2 , b 2 = a 1 .
H ^ F D L = ω 0 + v ( 0 A ^ A ^ 0 ) + O ( q 2 ) , A ^ γ x q x σ x + γ y q y σ y .
f 0 f 2 q 2 + f 0 ( q x 2 q y 2 2 f 1 q x q y 0 0 2 f 1 q x q y q y 2 q x 2 0 0 0 0 q x 2 q y 2 2 f 1 q x q y 0 0 2 f 1 q x q y q y 2 q x 2 ) .
Θ π 2 M x Θ π 2 = M y .
[ Θ π 2 , G ~ z ] = 0 , [ ( Θ π 2 ) 2 , G ~ z ] + = 0 ,
[ ( Θ π 2 ) 2 , M x ] = 0 , [ ( Θ π 2 ) 2 , M y ] = 0 .
k x = ω c sin θ i cos ϕ i ,
k y = ω c sin θ i sin ϕ i ,
k z = ω c cos θ i .
q x = k x , q y = k y .
δ ω = v z q z τ z + f 0 [ β q 2 ± ( Δ z + q x 2 q y 2 ) 2 + 4 q x 2 q y 2 ] ,
v g , x = ω q x = f 0 ( 2 β q x ± 2 ( Δ z + q x 2 q y 2 ) q x + 4 q y 2 q x ( Δ z + q x 2 q y 2 ) 2 + 4 q x 2 q y 2 ) ,
v g , y = ω q y = f 0 ( 2 β q y ± 2 ( Δ z + q x 2 q y 2 ) q y + 4 q x 2 q y ( Δ z + q x 2 q y 2 ) 2 + 4 q x 2 q y 2 ) .
ϕ o = A r g [ v g , x + i v g , y ] ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.