This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Pinch-point spectral singularity from the interference of topological loop states

Masafumi Udagawa1 [email protected]    Hiroki Nakai2 [email protected]    Chisa Hotta2 [email protected] 1Department of Physics, Gakushuin University, Mejiro, Tosshima-ku, Tokyo 171-8588, Japan 2Graduate School of Arts and Sciences, University of Tokyo, Meguro-ku, Tokyo 153-8902, Japan
Abstract

Pinch point is a spectral discontinuity found in the neutron diffraction image of spin ice. Similar spectral singularity is commonly observed in a broad range of systems that have a close connection with flat bands. We focus on the electron flat band and its two topologically distinct classes of wavefunction: the compact localized state (CLS), and the non-contractible loop state (NLS). We establish their simple mathematical relationship, showing that different Bloch NLSs can be derived as momentum derivatives of a Bloch CLS, depending on the approaching direction toward the singular point. This CLS-NLS correspondence helps visualize the pinch point as an interference pattern among NLSs through a “polarizer”, which encodes the information about the location of singular momentum and the experimental techniques like spin-polarized photoemission spectroscopy. It helps extract topological information knit to microscopic electronic and magnetic structures.

Introduction. Characterization of quantum states using the language of topology often encounters difficulty in confirming it experimentally via local observables. One of the few exceptions to date is the integer quantum Hall systems [1] whose Hall conductivity is quantized in units of e2/he^{2}/h. In this light, the global loop states of flat bands may offer another example because they should have strong relevance to the observable singularity of the dynamical structure factor called pinch points.

The pinch point originally drew attention as a singularity of the magnetic structure factor of spin ice compounds, Ho2Ti2O7 [2, 3]. The spin ice state consists of a macroscopic number of classical states that satisfy a local constraint known as the ice rule, wherein the spin configuration within a tetrahedron sums to zero [4, 5]. From a field-theory point of view, the ice rule is a divergence-free condition in analogy with the divergence-free magnetic field lines in electromagnetism, which produces the algebraic dipolar spin correlation [6] and a resultant pinch point singularity at 𝒌=0{\bf\it k}^{*}=0 [7, 8]. However, pinch point is not a unique concept of spin ice.

The common origin of pinch points underlying a variety of systems is a flat band. It is known that the magnetic correlation of spin ice is accurately described within a large-𝒩{\cal N} approximation, which attributes the magnetic structure factor of spin ice to the flat band eigenfunctions in momentum space  [7, 9, 8, 10]. This description offers two perspectives. Firstly, it highlights the ubiquity of pinch points in various systems ranging from frustrated magnets hosting classical spin liquids [11, 9, 12, 13, 14], tensor spin liquids [15], water ice [16, 17], ferroelectric models [18], to the materials with electronic flat bands  [19, 20, 21, 22, 23, 24]. Secondly, there is a rich mathematical structure underlying the flat band eigenstates  [25, 26, 27, 28, 29, 30, 31, 32, 33, 34, 27, 35, 36, 37, 38, 39] including global loop states [25, 26, 27, 28], quantum geometry [29, 30, 31, 32], and singularity at band-touching points [33]. Despite the knowledge accumulated on flat band and pinch point physics, their mutual relationship has not yet been clearly established.

Flat band phenomena.  Let us illustrate what is known about flat bands using the kagome lattice. On the flat-band energy level, there exist states of at least the number of unit cells, NcN_{c}. The absence of momentum dependence implies that these states can be written as localized states in real space, typically as the closed loops of finite length called a compact loop state (CLS).

Suppose that we have a single CLS state positioned at 𝑹0{\bf\it R}_{0}, denoted as |u𝑹0CLS|u^{\rm CLS}_{{\bf\it R}_{0}}\rangle, where one can generate its NcN_{c} copies at {𝑹l}\{{\bf\it R}_{l}\} by translation. By superposing them with a phase factor ei𝒌𝑹le^{i{\bf\it k}\cdot{\bf\it R}_{l}}, one naturally expects that NcN_{c} independent Bloch states constitute a flat band as

|u𝒌CLS1Ncl=0Nc1ei𝒌𝑹l|u𝑹lCLS.\displaystyle|u_{{\bf\it k}}^{\rm CLS}\rangle\equiv\frac{1}{\sqrt{N_{c}}}\sum_{l=0}^{N_{c}-1}e^{i{\bf\it k}\cdot{\bf\it R}_{l}}|u^{\rm CLS}_{{\bf\it R}_{l}}\rangle. (1)

However, in a popular example of a tight-binding model on a kagome lattice, they fail to span a complete basis because l=0Nc1|u𝑹lCLS=0\sum_{l=0}^{N_{c}-1}|u^{\rm CLS}_{{\bf\it R}_{l}}\rangle=0 (non-local constraint), resulting in only Nc1N_{c}-1 independent Bloch states in Eq. (1). The missing state is located at a momentum 𝒌=0{\bf\it k}^{*}=0.

Bergmann et al. [26] identified these missing states as non-contractible loop states (NLS): a global loop state winding around the system, topologically different from CLS. The kagome lattice has two NLSs extending in different directions, and one of them complements the missing flat band state, whereas the other belongs to the dispersive band. The presence of two NLS ensures that the dispersive band touches the flat band at 𝒌=0{\bf\it k}^{*}=0.

Indeed, the following facts on the “singular momentum” 𝒌{\bf\it k^{*}} have been recursively discussed [40]; (i) vanishing norm of the CLS u𝒌CLS|u𝒌CLS=0\langle u^{\rm CLS}_{{\bf\it k}^{*}}|u^{\rm CLS}_{{\bf\it k}^{*}}\rangle=0, (ii) existence of NLSs, (iii) the discontinuity of flat band wave functions, CLS (𝒌𝒌{\bf\it k}\neq{\bf\it k^{*}}) and NLS(𝒌=𝒌{\bf\it k}={\bf\it k^{*}}), (iv) band touching, and (v) pinch point. Rhim and Yang referred to (i) - (iii) as different facets of the same physics, while they further showed counter-example; (iv) exists but (ii) and (iii) do not [28, 32]. There is a list of how the flat band phenomena like (ii) and (iv) are interpreted within a large-𝒩{\cal N} scheme in the classical spin liquid state of magnets [40]. Although (i)-(v) are commonly observed in flat band systems, their mutual relationships are not necessarily clear.

This Letter aims at establishing a simple relationship between NLS and CLS and make transparent connections among the aforementioned flat band phenomena. Through a consistent characterization, we can see how the pinch point spectrum arises from NLS interference, thereby enabling the extraction of topological information from observable pinch points.

Preliminaries.  As the simplest platform, we consider a tight-binding Hamiltonian with nearest-neighbor hopping terms, =ti,j(|ij|+H.c.)\mathcal{H}=-t\sum_{\langle i,j\rangle}(|i\rangle\langle j|+{\rm H.c.}), on kagome and pyrochlore lattices (Fig. 1(a)) for spatial dimensions, d=2d=2 and 33, and number of sublattices, ns=3n_{s}=3 and 44, respectively. Let the lattice vector be given as 𝑹l=μ=1dlμ𝒂μ,{\bf\it R}_{\>l}=\sum_{\mu=1}^{d}l_{\mu}{\bf\it a}_{\mu},\; lμl_{\mu}\in{\mathbb{Z}} with a unit vector 𝒂μ{\bf\it a}_{\mu} in the μ\mu-th direction, along which we impose periodic boundary conditions. The reciprocal lattice vectors 𝒃η{\bf\it b}_{\eta} satisfy 𝒂μ𝒃η=2πδμη{\bf\it a}_{\mu}\cdot{\bf\it b}_{\eta}=2\pi\delta_{\mu\eta}. In kagome and pyrochlore lattices, sublattice vectors can be chosen as 𝒓0=0{\bf\it r}_{0}=0 and 𝒓μ=𝒂μ/2{\bf\it r}_{\mu}={\bf\it a}_{\mu}/2 for μ=1,,ns1\mu=1,\cdots,n_{s}-1.

Refer to caption
Figure 1: (a) Kagome and pyrochlore lattices with unit vectors 𝒂μ{\bf\it a}_{\mu} and sublattice indices. Hexagonal compact localized states (CLSs) are shown. (b) Non-contractible loop states (NLSs), |u2(1)NLS|u^{\rm NLS}_{2(1)}\rangle. (c) Bloch NLSs, |ψ𝒌,1(2)NLS|\psi^{\rm NLS}_{{\bf\it k^{*}},1(2)}\rangle, obtained by superposing the single-loop NLS |uL2(1)NLS|u^{\rm NLS}_{L_{2(1)}}\rangle in the direction 1(2)1(2).

CLS-NLS correspondence.  Let us continue with the kagome lattice. We may choose a primitive CLS around a hexagon as (Fig. 1(a)),

|u𝑹lCLS\displaystyle|u^{\rm CLS}_{{\bf\it R}_{l}}\rangle =(|𝑹l,1|𝑹l,2)+(|𝑹l+𝒂1,2|𝑹l+𝒂1,0)\displaystyle=(|{\bf\it R}_{l},1\rangle-|{\bf\it R}_{l},2\rangle)+(|{\bf\it R}_{l}+{\bf\it a}_{1},2\rangle-|{\bf\it R}_{l}+{\bf\it a}_{1},0\rangle)
+(|𝑹l+𝒂2,0|𝑹l+𝒂2,1),\displaystyle+(|{\bf\it R}_{l}+{\bf\it a}_{2},0\rangle-|{\bf\it R}_{l}+{\bf\it a}_{2},1\rangle), (2)

where |𝑹l,μ|{\bf\it R}_{l},\mu\rangle denotes the electron state on the μ\mu-th sublattice of the ll-th unit cell. Similarly, we can write down a single global NLS LμL_{\mu} as

|uLμNLS=lLμ(|𝑹l,0|𝑹l,μ),|u^{\rm NLS}_{L_{\mu}}\rangle=\sum_{l\in{\rm L}_{\mu}}(|{\bf\it R}_{l},0\rangle-|{\bf\it R}_{l},\mu\rangle), (3)

winding around a global loop L1(2)L_{1(2)} in the direction 1(2)(Fig. 1(b)). Both CLS and NLS wavefunctions explicitly satisfy the divergence-free condition [41], i.e. the coefficients of |𝑹l,μ|{\bf\it R}_{l},\mu\rangle sum to zero in each unit cell.

Once we choose an elementary CLS, its translation and superposition will provide a set of Bloch CLS,

|u𝒌CLS=1Ncl=0Nc1ei𝒌𝑹l(μ=0nsvμ,𝒌|𝑹l,μ),\displaystyle|u^{\rm CLS}_{{\bf\it k}}\rangle=\frac{1}{\sqrt{N_{c}}}\sum_{l=0}^{N_{c}-1}e^{i{\bf\it k}\cdot{\bf\it R}_{l}}\Big{(}\sum_{\mu=0}^{n_{s}}v_{\mu,\bm{k}}|{\bf\it R}_{l},\mu\rangle\Big{)}, (4)
{v0,𝒌=eik2eik1,v1,𝒌=1eik2,v2,𝒌=eik11.\displaystyle\left\{\begin{array}[]{l}v_{0,{\bf\it k}}=e^{-ik_{2}}-e^{-ik_{1}},\\ v_{1,{\bf\it k}}=1-e^{-ik_{2}},\\ v_{2,{\bf\it k}}=e^{-ik_{1}}-1.\end{array}\right. (8)

In determining the vector, 𝒗𝒌=(v0,𝒌,v1,𝒌,v2,𝒌){\bf\it v}_{{\bf\it k}}=(v_{0,{\bf\it k}},v_{1,{\bf\it k}},v_{2,{\bf\it k}}), the overlap of neighboring |u𝑹lCLS|u^{\rm CLS}_{{\bf\it R}_{l}}\rangle’s, each spreading over a few unit cells, is essential. For example, |𝑹l+𝒂1,2|{\bf\it R}_{l}+{\bf\it a}_{1},2\rangle is shared by two CLSs, |u𝑹lCLS|u^{\rm CLS}_{{\bf\it R}_{l}}\rangle and |u𝑹l+𝒂1CLS|u^{\rm CLS}_{{\bf\it R}_{l}+{\bf\it a}_{1}}\rangle, and in the Bloch state |u𝒌CLS|u^{\rm CLS}_{{\bf\it k}}\rangle, the former gives ei𝒌(𝑹l+𝒂1)eik1|𝑹l+𝒂1,2e^{i{\bf\it k}\cdot({\bf\it R}_{l}+{\bf\it a}_{1})}e^{-ik_{1}}|{\bf\it R}_{l}+{\bf\it a}_{1},2\rangle, while the latter, ei𝒌(𝑹l+𝒂1)|𝑹l+𝒂1,2-e^{i{\bf\it k}\cdot({\bf\it R}_{l}+{\bf\it a}_{1})}|{\bf\it R}_{l}+{\bf\it a}_{1},2\rangle. Their sum yields v2,𝒌v_{2,{\bf\it k}} in Eq. (8), where the factor eik1e^{-ik_{1}} appears because CLSs are shared in the direction 1. 𝒗𝒌{\bf\it v}_{{\bf\it k}} is also concerned with the norm of the Bloch CLS: (u𝒌CLS|u𝒌CLS)12(\langle u^{\rm CLS}_{{\bf\it k}}|u^{\rm CLS}_{{\bf\it k}}\rangle)^{\frac{1}{2}} =|𝒗𝒌|=2[sin2k2k12+sin2k22+sin2k12]1/2=|{\bf\it v}_{{\bf\it k}}|=2[\sin^{2}\frac{k_{2}-k_{1}}{2}+\sin^{2}\frac{k_{2}}{2}+\sin^{2}\frac{k_{1}}{2}]^{1/2}. Then, |𝒗𝒌|=0|{\bf\it v}_{{\bf\it k}}|=0 at 𝒌=0{\bf\it k}=0, where |u𝒌CLS|u^{\rm CLS}_{{\bf\it k}}\rangle cannot be defined, i.e. 𝒌=𝒌=0{\bf\it k}={\bf\it k^{*}}=0 gives a singular momentum.

We are now ready to introduce the CLS-NLS correspondence. The first observation on Eq. (4) is that if we replace the coefficient vμ,𝒌v_{\mu,\bm{k}} with one of its momentum derivatives, vμ,𝒌kη\frac{\partial v_{\mu,{\bf\it k}}}{\partial k_{\eta}}, and take 𝒌𝒌{\bf\it k}\to{\bf\it k^{*}}, we find for example,

|ψ𝒌,1NLS\displaystyle|\psi^{\rm NLS}_{{\bf\it k}^{*},1}\rangle 1Ncl=0Nc1ei𝒌𝑹lμ=0ns1vμ,𝒌k1|𝑹l,μ\displaystyle\equiv\frac{1}{\sqrt{N_{c}}}\sum_{l=0}^{N_{c}-1}e^{i{\bf\it k}^{*}\cdot{\bf\it R}_{l}}\sum_{\mu=0}^{n_{s}-1}\frac{\partial v_{\mu,{\bf\it k}^{*}}}{\partial k_{1}}|{\bf\it R}_{l},\mu\rangle
=iNcl=0Nc1(|𝑹l,0|𝑹l,2),\displaystyle=\frac{i}{\sqrt{N_{c}}}\sum_{l=0}^{N_{c}-1}(|{\bf\it R}_{l},0\rangle-|{\bf\it R}_{l},2\rangle), (9)

which is nothing but a Bloch NLS, i.e. the equal-weight superposition of the one-loop NLS |uL2NLS|u^{\rm NLS}_{L_{2}}\rangle (Eq. (3)) translated one by one in the direction 11. The factor vμ,𝒌k1\frac{\partial v_{\mu,{\bf\it k}^{*}}}{\partial k_{1}} is proportional to the weight of NLS state at sublattice μ\mu. Similarly, the k2k_{2} derivative results in another Bloch NLS |ψ𝒌,2NLS|\psi^{\rm NLS}_{{\bf\it k}^{*},2}\rangle, which superposes another one-loop NLS, |uL1NLS|u^{\rm NLS}_{L_{1}}\rangle, translated in the direction 22. To see why the kk-derivative can extract NLS from CLS, let us look at their relationships in Fig. 1(b). Along the one-loop NLS, |uL2NLS|u^{\rm NLS}_{L_{2}}\rangle, each site belonging to either μ=0\mu=0 or 22 hinges two adjacent CLSs as marked with ovals. Acting on the phase factor eik1e^{-ik_{1}} assigned to these hinge sites, the kk-derivative extracts the one-loop NLS |uL2NLS|u^{\rm NLS}_{L_{2}}\rangle.

The second observation is that such kk-derivative naturally appears in the process of taking the limit, 𝒌𝒌{\bf\it k}\to{\bf\it k}^{*}. To see this, we introduce the normalized Bloch CLS,

|ψ𝒌CLS=1Ncl=0Nc1ei𝒌𝑹lμ=0nsvμ,𝒌|𝒗𝒌||𝑹l,μ.|\psi^{\rm CLS}_{{\bf\it k}}\rangle=\frac{1}{\sqrt{N_{c}}}\sum_{l=0}^{N_{c}-1}e^{i{\bf\it k}\cdot{\bf\it R}_{l}}\sum_{\mu=0}^{n_{s}}\frac{v_{\mu,\bm{k}}}{|{\bf\it v}_{{\bf\it k}}|}|{\bf\it R}_{l},\mu\rangle. (10)

In approaching the singular point as 𝒌=𝒌+δkη𝒃η{\bf\it k}={\bf\it k}^{*}+\delta k_{\eta}{\bf\it b}_{\eta} and δkη0\delta k_{\eta}\to 0, the denominator vanishes linearly as |𝒗𝒌|Cδkη|{\bf\it v}_{{\bf\it k}}|\to C\delta k_{\eta}. Given that 𝒗𝒌=0{\bf\it v}_{{\bf\it k}^{*}}=0, the coefficient of Eq. (10) approaches

vμ,𝒌|𝒗𝒌|vμ,𝒌+δkη𝒃ηvμ,𝒌Cδkηvμ,𝒌kη.\displaystyle\frac{v_{\mu,\bm{k}}}{|{\bf\it v}_{{\bf\it k}}|}\to\frac{v_{\mu,{\bf\it k}^{*}+\delta k_{\eta}{\bf\it b}_{\eta}}-v_{\mu,{\bf\it k}^{*}}}{C\delta k_{\eta}}\propto\frac{\partial v_{\mu,{\bf\it k}^{*}}}{\partial k_{\eta}}. (11)

Combing Eqs. (9) and (11), it is naturally shown that the Bloch NLS |ψ𝒌,ηNLS|\psi^{\rm NLS}_{{\bf\it k}^{*},\eta}\rangle is obtained as the 𝒌𝒌{\bf\it k}\to{\bf\it k}^{*} limit of normalized Bloch CLS. Because Eq. (11) means that different NLS state appears depending on the direction η\eta to approach 𝒌{\bf\it k}^{*}, there is a discontinuity of flat band wavefunction at 𝒌=𝒌{\bf\it k}={\bf\it k}^{*}.

One can approach the singular point from an arbitrary direction as 𝒌=𝒌+(δk1𝒃1+δk2𝒃2)/(2π){\bf\it k}={\bf\it k}^{*}+(\delta k_{1}{\bf\it b}_{1}+\delta k_{2}{\bf\it b}_{2})/(2\pi), and the Bloch CLS converges to a linearly combined two Bloch NLSs as

|ψ𝒌CLSC1δk1|ψ𝒌,1NLS+C2δk2|ψ𝒌,2NLS.|\psi^{\rm CLS}_{{\bf\it k}}\rangle\to C_{1}\delta k_{1}|\psi^{\rm NLS}_{{\bf\it k}^{*},1}\rangle+C_{2}\delta k_{2}|\psi^{\rm NLS}_{{\bf\it k}^{*},2}\rangle. (12)

The simplest CLS-NLS correspondence, |ψ𝒌CLS|ψ𝒌,1(2)NLS|\psi^{\rm CLS}_{{\bf\it k}}\rangle\rightarrow|\psi^{\rm NLS}_{{\bf\it k}^{*},1(2)}\rangle, is attained when approaching 𝒌{\bf\it k}^{*} along the reciprocal vector, 𝒃1(2){\bf\it b}_{1(2)}, where we have k1𝒗𝒌=(v0,𝒌k1,v1,𝒌k1,v2,𝒌k1)(+1,0,1)\frac{\partial}{\partial k_{1}}{\bf\it v}_{{\bf\it k}^{*}}=(\frac{\partial v_{0,{\bf\it k}^{*}}}{\partial k_{1}},\frac{\partial v_{1,{\bf\it k}^{*}}}{\partial k_{1}},\frac{\partial v_{2,{\bf\it k}^{*}}}{\partial k_{1}})\propto(+1,0,-1) for |ψ𝒌,1NLS|\psi^{\rm NLS}_{{\bf\it k}^{*},1}\rangle, and k2𝒗𝒌(+1,1,0)\frac{\partial}{\partial k_{2}}{\bf\it v}_{{\bf\it k}^{*}}\propto(+1,-1,0) for |ψ𝒌,2NLS|\psi^{\rm NLS}_{{\bf\it k}^{*},2}\rangle. They are the weight of μ=(0,1,2)\mu=(0,1,2) sublattices constituting the associated one-loop NLS states, |uL2NLS|u^{\rm NLS}_{L_{2}}\rangle and |uL1NLS|u^{\rm NLS}_{L_{1}}\rangle, respectively (see Fig. 1(b)).

There might be a small complication in matching |ψ𝒌,1(2)NLS|\psi^{\rm NLS}_{{\bf\it k}^{*},1(2)}\rangle with |uL2(1)NLS|u^{\rm NLS}_{L_{2(1)}}\rangle due to the switch in indices, η=1\eta=1 and 22. To explain this further, when approaching along 𝒃1{\bf\it b}_{1}, we find the NLS Bloch state |ψ𝒌,1NLS|\psi^{\rm NLS}_{{\bf\it k}^{,}1}\rangle consisting of loops, |uL2NLS|u^{\rm NLS}_{L_{2}}\rangle, running in direction 2, orthogonal to 𝒃1{\bf\it b}_{1} (𝒃1𝒂2=0{\bf\it b}_{1}\cdot{\bf\it a}_{2}=0). In other words, the loop |uL2NLS|u^{\rm NLS}_{L_{2}}\rangle regarded as a wavefront “propagates” in the direction 1 to form |ψ𝒌,1NLS|\psi^{\rm NLS}_{{\bf\it k}^{*},1}\rangle. Therefore, η\eta indexes the direction of the wavefront and not the winding direction of the loop state.

Degenerate flat bands at d=3d=3.  We can generalize the above arguments to three and higher dimensions. One possible complication is the degeneracy of flat bands. Suppose the system has nfn_{f} flat bands, then we can choose a set of nfn_{f} linearly independent normalized Bloch CLS {|ψ𝒌CLS(j)}\{|\psi^{{\rm CLS}(j)}_{{\bf\it k}}\rangle\}, with j=1,2,nfj=1,2,\cdots n_{f}. If all these states simultaneously vanish at 𝒌{\bf\it k}^{*}, this momentum is singular. In approaching 𝒌{\bf\it k}^{*} along 𝒃η{\bf\it b}_{\eta}, each normalized Bloch CLS converges to a corresponding Bloch NLS, whose form |ψ𝒌,ηNLS(j)|\psi^{{\rm NLS}(j)}_{{\bf\it k}^{*},\eta}\rangle, depends on the choice of {|ψ𝒌CLS(j)}\{|\psi^{{\rm CLS}(j)}_{{\bf\it k}}\rangle\}. However, at least |ψ𝒌,ηNLS(j)|\psi^{{\rm NLS}(j)}_{{\bf\it k}^{*},\eta}\rangle must live in the space orthogonal to 𝒃η{\bf\it b}_{\eta}, so that η\eta indexes the direction of the wavefront again, i.e. |ψ𝒌,ηNLS(j)|\psi^{{\rm NLS}(j)}_{{\bf\it k}^{*},\eta}\rangle is the linear combination of one-loop NLS’s in the subspace spanned by the lattice vectors 𝒂μ{\bf\it a}_{\mu}’s, which satisfy 𝒃η𝒂μ=0{\bf\it b}_{\eta}\cdot{\bf\it a}_{\mu}=0.

As an example, we consider the pyrochlore lattice at d=3d=3, which hosts two-fold degenerate flat bands. As a counterpart of Eq. (8), we can choose two CLSs |u𝒌CLS(1)|u^{\rm CLS(1)}_{{\bf\it k}}\rangle and |u𝒌CLS(2)|u^{\rm CLS(2)}_{{\bf\it k}}\rangle, as their basis set, given as

{v0,𝒌(1)=eik3eik1,v1,𝒌(1)=1eik3,v2,𝒌(1)=0,v3,𝒌(1)=eik11,{v0,𝒌(2)=eik3eik2,v1,𝒌(2)=0,v2,𝒌(2)=1eik3,v3,𝒌(2)=eik21,\displaystyle\left\{\begin{array}[]{l}v^{(1)}_{0,{\bf\it k}}=e^{-ik_{3}}-e^{-ik_{1}},\\ v^{(1)}_{1,{\bf\it k}}=1-e^{-ik_{3}},\\ v^{(1)}_{2,{\bf\it k}}=0,\\ v^{(1)}_{3,{\bf\it k}}=e^{-ik_{1}}-1,\end{array}\right.\left\{\begin{array}[]{l}v^{(2)}_{0,{\bf\it k}}=e^{-ik_{3}}-e^{-ik_{2}},\\ v^{(2)}_{1,{\bf\it k}}=0,\\ v^{(2)}_{2,{\bf\it k}}=1-e^{-ik_{3}},\\ v^{(2)}_{3,{\bf\it k}}=e^{-ik_{2}}-1,\end{array}\right. (21)

each constructed from the primitive CLSs localized around hexagons on the sliced kagome plane, 𝒂1{\bf\it a}_{1}-𝒂3{\bf\it a}_{3} and 𝒂2{\bf\it a}_{2}-𝒂3{\bf\it a}_{3}, respectively (Fig 1 (a)). Both vanish at 𝒌=𝒌=0{\bf\it k}={\bf\it k}^{*}=0, that marks the singularity. For instance, in approaching 𝒌{\bf\it k}^{*} along 𝒃3{\bf\it b}_{3}, the linear combination of normalized Bloch CLSs, α|ψ𝒌CLS(1)+β|ψ𝒌CLS(2)\alpha|\psi^{\rm CLS(1)}_{{\bf\it k}}\rangle+\beta|\psi^{\rm CLS(2)}_{{\bf\it k}}\rangle, converges to the superposition of |uL1NLS|u^{\rm NLS}_{L_{1}}\rangle and |uL2NLS|u^{\rm NLS}_{L_{2}}\rangle winding along directions 1 and 2, and hence living on the [111] kagome plane orthogonal to 𝒃3{\bf\it b}_{3} 111See Sec. A in Supplemental Material. To see this, it is enough to notice k3(α𝒗𝒌(1)+β𝒗𝒌(2))(α+β,α,β,0)\frac{\partial}{\partial k_{3}}(\alpha{\bf\it v}^{(1)}_{{\bf\it k}^{*}}+\beta{\bf\it v}^{(2)}_{{\bf\it k}^{*}})\propto(\alpha+\beta,-\alpha,-\beta,0), from Eq. (21), namely the two NLSs are placed on sublattices, μ=0,1,2\mu=0,1,2.

Refer to caption
Figure 2: Structure factors of flat bands: (a) kagome and (b,c,d) pyrochlore lattice. A dashed hexagon in (b) and yellow solid lines in (c) denote the line 𝒮(𝒌)=8/3{\cal S}(\bm{k})=8/3. In panel (c) the dark region 𝒮(𝒌)0.01{\cal S}(\bm{k})\leq 0.01 is shown near the Γ\Gamma-point which extends toward 12 different singular points as dark cones, whose slices are shown in the inset. Panel (d) shows the case where the removed electron spin varies its angle θe\theta_{e} about the zz-axis for the spin-orbit coupled flat bands. Inset shows spin-rotation axes 𝒎μ{\bf\it m}_{\mu} pointing toward the center of the tetrahedron.

From NLS to Pinch point. In diffraction measurements, we are concerned with the structure factor, which involves assessing the matrix element containing spatial details regarding sublattice coordinates. For example, in photoemission spectroscopy [43, 44, 45, 46], we are interested in the operator,

a𝒌=1Ncl=0Nc1μ=0ns1ei𝒌(𝑹l+𝒓μ)cl,μ,a^{{\dagger}}_{{\bf\it k}}=\frac{1}{\sqrt{N_{c}}}\sum_{l=0}^{N_{c}-1}\sum_{\mu=0}^{n_{s}-1}e^{i{\bf\it k}\cdot({\bf\it R}_{l}+{\bf\it r}_{\mu})}c^{{\dagger}}_{l,\mu}, (22)

where cl,μc^{{\dagger}}_{l,\mu} creates an electron in the state |𝑹l,μ|{\bf\it R}_{l},\mu\rangle. We now start from the NeN_{e} electron ground state, |ΨNe|\Psi^{N_{e}}\rangle, and examine the probability that one electron is added to the empty flat band state |ψ𝒌CLS|\psi^{\rm CLS}_{{\bf\it k}}\rangle. The charge structure factor measured through the photoemission spectroscopy experiment is evaluated by the matrix element, 𝒮(𝒌)=|ΨNe|c𝒌CLSa𝒌|ΨNe|2\mathcal{S}({\bf\it k})=|\langle\Psi^{N_{e}}|c^{\rm CLS}_{{\bf\it k}}a^{\dagger}_{{\bf\it k}}|\Psi^{N_{e}}\rangle|^{2}, which can be rewritten using the normalized Bloch function, Eq. (10) as

𝒮(𝒌)\displaystyle\mathcal{S}({\bf\it k}) =|μei𝒌𝒓μvμ,𝒌|𝒗𝒌||2𝒌𝒌+𝑮η|𝑷𝒗𝒌kη|2,\displaystyle=\bigg{|}\sum_{\mu}e^{-i{{\bf\it k}\cdot{{\bf\it r}}_{\mu}}}\>\frac{v_{\mu,\bm{k}}}{|{\bf\it v}_{{\bf\it k}}|}\bigg{|}^{2}\xrightarrow[\begin{subarray}{c}{\bf\it k}\to{\bf\it k}^{*}+{\bf\it G}\\ \eta\end{subarray}]{}\bigg{|}\,{{\bf\it P}}\cdot\frac{\partial{\bf\it v}_{{\bf\it k}^{*}}}{\partial k_{\eta}}\,\bigg{|}^{2}, (23)
𝑷\displaystyle{\bf\it P} =(ei(𝒌+𝑮)𝒓μ).\displaystyle=(e^{-i({\bf\it k}^{*}+{\bf\it G})\cdot{{\bf\it r}}_{\mu}}). (24)

The vanishing of the norm, |𝒗𝒌|=0|{\bf\it v}_{{\bf\it k}^{*}}|=0, gives the discontinuity of 𝒮(𝒌)\mathcal{S}({\bf\it k}) at 𝒌=𝒌{\bf\it k}={\bf\it k}^{*}, implying nothing but the pinch point singularity. This singularity potentially appears at a set of momenta, 𝒌+𝑮{\bf\it k}^{*}+{\bf\it G}, shifted by arbitrary reciprocal vector 𝑮=μgμ𝒃μ{\bf\it G}=\sum_{\mu}g_{\mu}{\bf\it b}_{\mu}, with integers gμg_{\mu}. Whereas, the profile of 𝒮(𝒌𝒌+𝑮)\mathcal{S}({\bf\it k}\sim{\bf\it k}^{*}+{\bf\it G}) may differ depending on 𝑮{\bf\it G}, because the phase factor ei𝒌𝒓μe^{-i{{\bf\it k}}\cdot{{\bf\it r}}_{\mu}} has different periodicities about 𝑮{\bf\it G} 222The dependence of spectrum on phase factor is systematically studied by Conlon and Chalker  [53] in different context..

Indeed, the profile is determined by the two factors that appear on the r.h.s. of Eq. (23). One factor is the vector 𝑷(ei(𝒌+𝑮)𝒓μ){\bf\it P}\equiv(e^{-i({{\bf\it k}^{*}}+{\bf\it G})\cdot{{\bf\it r}}_{\mu}}) and the other is the weight of the NLSs, (𝒗𝒌/kη)(\partial{\bf\it v}_{{\bf\it k}^{*}}\big{/}\partial k_{\eta}). The former depends on the choice of 𝒌+𝑮{\bf\it k}^{*}+{\bf\it G} but not on the approaching direction η\eta to 𝒌{\bf\it k}^{*}, whereas for the latter the dependence is vice versa. The inner product of these two factors governs the profile of 𝒮(𝒌)\mathcal{S}({\bf\it k}), indicating that the interference of dd independent NLSs occurs according to the selection rule encoded in 𝑷{\bf\it P}. We call the vector 𝑷{\bf\it P} “a polarizer” as it produces different images by varying the choice of NLSs when 𝒌{\bf\it k}^{*} is translated by 𝑮{\bf\it G}.

Examples. Revisiting the kagome lattice, we focus on the singular point 𝒌=𝒃2{\bf\it k}={\bf\it b}_{2}, where the polarizer takes the form: 𝑷=(ei𝒃2𝒓0,ei𝒃2𝒓1,ei𝒃2𝒓2)=(1,1,1){{\bf\it P}}=(e^{-i{{\bf\it b}_{2}}\cdot{{\bf\it r}}_{0}},e^{-i{{\bf\it b}_{2}}\cdot{{\bf\it r}}_{1}},e^{-i{{\bf\it b}_{2}}\cdot{{\bf\it r}}_{2}})=(1,1,-1). If we approach 𝒌{\bf\it k}^{*} along 𝒃2{\bf\it b}_{2} (vertical line in Fig. 2(a)), the NLSs are chosen as 𝒗𝒌k2(1,1,0)\frac{\partial{\bf\it v}_{{\bf\it k}^{*}}}{\partial k_{2}}\propto(1,-1,0), and 𝒮(𝒌)\mathcal{S}({\bf\it k}) vanishes as |11+1(1)+(1)0|2=0\propto|1\cdot 1+1\cdot(-1)+(-1)\cdot 0|^{2}=0. Along 𝒃1{\bf\it b}_{1}, it yields 𝒮(𝒌)22=4\mathcal{S}({\bf\it k})\sim 2^{2}=4. When approaching along a general direction, 𝒌(α𝒃1+β𝒃2){\bf\it k}\parallel(\alpha{\bf\it b}_{1}+\beta{\bf\it b}_{2}), only the 𝒃2{\bf\it b}_{2} component contributes according to Eq. (12), and 𝒮(𝒌)\mathcal{S}({\bf\it k}) takes a finite value. With these simple considerations we can reproduce a bowtie structure.

Even in the pyrochlore model with degenerate flat bands, we can identify the dark line where 𝒮(𝒌){\cal S}({\bf\it k}) vanishes; we may target 𝒌=𝒃3{\bf\it k}={\bf\it b}_{3} whose polarizer is 𝑷=(ei𝒃3𝒓0,ei𝒃3𝒓1,ei𝒃3𝒓2,ei𝒃3𝒓3)=(1,1,1,1){\bf\it P}=(e^{-i{{\bf\it b}_{3}}\cdot{{\bf\it r}}_{0}},e^{-i{{\bf\it b}_{3}}\cdot{{\bf\it r}}_{1}},e^{-i{{\bf\it b}_{3}}\cdot{{\bf\it r}}_{2}},e^{-i{{\bf\it b}_{3}}\cdot{{\bf\it r}}_{3}})=(1,1,1,-1). Approaching this point parallel to 𝒃3{\bf\it b}_{3}, NLSs have weights, k3(α𝒗𝒌(1)+β𝒗𝒌(2))(α+β,α,β,0)\frac{\partial}{\partial k_{3}}(\alpha{\bf\it v}^{(1)}_{{\bf\it k}^{*}}+\beta{\bf\it v}^{(2)}_{{\bf\it k}^{*}})\propto(\alpha+\beta,-\alpha,-\beta,0), which is orthogonal to the polarizer and 𝒮(𝒌)=𝑷𝒗𝒌k3=0{\cal S}({\bf\it k})={{\bf\it P}}\cdot\frac{\partial{\bf\it v}_{{\bf\it k}^{*}}}{\partial k_{3}}=0. This result is confirmed by the exact analytical form of 𝒮(𝒌){\cal S}({\bf\it k}) given in Supplemental Material 333See Sec. B and Eq. (S16) in Supplemental Material. Figure 2(c) visualizes the low-intensity region, 𝒮(𝒌)0.01{\cal S}({\bf\it k})\leq 0.01. This region develops around dark lines branching from the Γ\Gamma-point toward 12 different singular points, where the two dark cones touch.

Spin-dependent polarizer. One caveat is that the function of the polarizer depends on models and experimental probes. As an interesting case study, we consider the spin-polarized photoemission spectroscopy applied to the spin-orbit coupled flat band state on the pyrochlore lattice [49, 50]. Suppose we have a half-filled flat band in the ground state, with sublattice spins polarized as |𝒔μ(𝒎μ𝝈)||{\bf\it s}_{\mu}\rangle\equiv({\bf\it m}_{\mu}\cdot\bm{\sigma})|\uparrow\rangle, where 𝒎μ{\bf\it m}_{\mu} are the axes shown in Fig. 2(d). In the inverse photoemission spectroscopy to remove an electron from this ground state, the polarizer takes the form, P=ei(𝒌+𝑮)𝒓μ𝒔e|𝒔μP=e^{-i({{\bf\it k}^{*}}+{\bf\it G})\cdot{{\bf\it r}}_{\mu}}\langle{\bf\it s}_{e}|{\bf\it s}_{\mu}\rangle, where |𝒔e|{\bf\it s}_{e}\rangle stands for the spin state of the removed electron. Due to the additional spin-dependent matrix element 𝒔e|𝒔μ\langle{\bf\it s}_{e}|{\bf\it s}_{\mu}\rangle, the dark line rotates with the spin orientation of the removed electron, θe\theta_{\rm e}, as shown in Fig. 2(d) 444See Sec. C in Supplemental Material. By carefully preparing a polarizer with dressed matrix elements, the information drawn from the spectrum can be used to clarify the microscopic spin structure of the state at focus.

Refer to caption
Figure 3: CLS-NLS correspondence and the relationships with the nature of flat bands and pinch points, (i)-(v).

Summary and Discussion. We have established the CLS-NLS correspondence, identifying the Bloch NLS at the singular 𝒌{\bf\it k}^{*} point as the momentum derivative of the Bloch CLS. The CLS-NLS correspondence makes a transparent connection among the flat band phenomena (i)-(v) as shown in Fig. 3. Here, (i) the vanishing of the norm of Bloch CLS enables the operation of momentum derivative in approaching the singular momentum, from which we reach the form of NLS in (ii) 555The combination of (i) and (ii) account for (iv) the presence of band touching [26, 28]. . NLS thus obtained is selected from among dd dependent choices, depending on the direction in approaching 𝒌{\bf\it k}^{*} , which naturally explains (iii) the discontinuity of the flat band wavefunction. Finally, the NLS tomography around the singular momentum provides a simple way to describe (v) the structure of pinch point, as an interference pattern of the NLS.

To note, any flat bands including non-singular ones host CLSs, given a general construction [28], and supported by the arbitrarity of the choice of CLSs, “NLS” is derived from our formula. However, when CLSs span a complete flat band basis, such “NLS” is contractible and (ii) and (iii) break down. Band touching may occur in that case as well.

The CLS-NLS correspondence guarantees that we can always have NLS for singular flat bands, and can use it for identifying pinch points, or to extract from pinch points the electronic or magnetic structures of more complex systems with multi-degrees of freedom.

Although the system with perfect flat bands and ideal pinch point may seem to be a specific example, in reality there are a wide variety of materials with nearly flat bands in its vicinity. These classes of bands are allowed to host finite Chern numbers or may give another spectral feature known as half moons or “shadow pinch point” [53, 54, 55, 56, 57]. The variety of spectral patterns found there can be diagnosed using in a language of pinch points, and thus our CLS-NLS correspondence will provide a guiding principle to understand a broad range of systems close to flat band.

Acknowledgements.
H.N. was supported by a Grant-in-Aid for JSPS Research Fellow (Grant No. JP23KJ0783). This work is supported by KAKENHI Grant No. JP20H05655, JP21H05191, JP21K03440 and JP22H01147 from JSPS of Japan.

References

Supplemental Material

.1 Convention of pyrochlore lattice and the choice of reciprocal lattice vectors

We take the lattice vectors of the pyrochlore lattice as 𝒂1=(1,1,0)/2,𝒂2=(0,1,1)/2,𝒂3=(1,0,1)/2{\bf\it a}_{1}=(1,-1,0)/2,{\bf\it a}_{2}=(0,-1,1)/2,{\bf\it a}_{3}=(1,0,1)/2, which gives 𝒃1=(1,1,1)2π,𝒃2=(1,1,1)2π,𝒃3=(1,1,1)2π{\bf\it b}_{1}=(1,-1,-1)2\pi,{\bf\it b}_{2}=(-1,-1,1)2\pi,{\bf\it b}_{3}=(1,1,1)2\pi shown in Fig. S1(a). In the pyrochlore lattice, we have four independent kagome planes which are denoted as [1 1 1],[1 11],[1 1 1][-1\ 1\,1],[1\,1-\!1],[1\,1\,1], and [11 1][1-\!1\,1] by the vectors normal to these planes. If we construct a hexagonal CLS on the four species of hexagons on these planes, the four that share half of the edges and form a polygon are not independent of each other. Namely, the summation over them cancels out (see Fig. S1(b)). In this sense, only three of the four CLSs are linearly independent.

Refer to caption
Figure S1: (a) Lattice conventions: the directions of lattice vectors and reciprocal lattice vectors. (b) Four hexagons forming a polygon. The hexagons share their vertices and by summing up the four CLS states on these hexagons, they cancel out. According to the vectors normal to the hexagons we label the surfaces by [11 1],[1 11],[1 1 1][-11\,1],[1\,1-1],[1\,1\,1], and [11 1][1-1\,1]. (c) The [1 1 1] plane spanned by 𝒂1{\bf\it a}_{1} and 𝒂2{\bf\it a}_{2} supports the NLSs living on the sublattices μ=0,1,2\mu=0,1,2.

By using them we can construct a Bloch CLS confined within one sheet of kagome plane in each direction, as we do in the main text Eq. (4), which are shown schematically in Fig. S1(c) as colored hexagons. We can also visualize one-loop NLS on each plane. When we choose the [111][111] kagome plane, the choices of three sublattices are μ=0,1,2\mu=0,1,2, and we find two NLSs,

|u1NLS=n(|n𝒂1,0|n𝒂1,1),\displaystyle|u_{1}^{\rm NLS}\rangle=\sum_{n}(|n{\bf\it a}_{1},0\rangle-|n{\bf\it a}_{1},1\rangle),
|u2NLS=n(|n𝒂2,0+|n𝒂2,2),\displaystyle|u_{2}^{\rm NLS}\rangle=\sum_{n}(-|n{\bf\it a}_{2},0\rangle+|n{\bf\it a}_{2},2\rangle), (S1)

which constitute the Bloch NLS, when approaching the singular 𝒌{\bf\it k}^{*} point along 𝒃3{\bf\it b}_{3}.

Similarly, for the [111][11-1] plane, we find one-loop NLSs,

|u1NLS=n(|n𝒂1,0|n𝒂1,1),\displaystyle|u_{1}^{\rm NLS}\rangle=\sum_{n}(|n{\bf\it a}_{1},0\rangle-|n{\bf\it a}_{1},1\rangle),
|u3NLS=n(|n𝒂3,0+|n𝒂3,3),\displaystyle|u_{3}^{\rm NLS}\rangle=\sum_{n}(-|n{\bf\it a}_{3},0\rangle+|n{\bf\it a}_{3},3\rangle), (S2)

and for the [111][-111] plane,

|u2NLS=n(|n𝒂2,0|n𝒂2,2),\displaystyle|u_{2}^{\rm NLS}\rangle=\sum_{n}(|n{\bf\it a}_{2},0\rangle-|n{\bf\it a}_{2},2\rangle),
|u3NLS=n(|n𝒂3,0+|n𝒂3,3).\displaystyle|u_{3}^{\rm NLS}\rangle=\sum_{n}(-|n{\bf\it a}_{3},0\rangle+|n{\bf\it a}_{3},3\rangle). (S3)

.2 Exact form of the spectral function

.2.1 general

If two or more flat bands exist, the formula of the structure factor must be generalized. The expression for the nfn_{f}-fold degenerate flat bands takes the following form,

S(𝒌)=μ,μ=0,ns1Sμ,μ(𝒌),\displaystyle S({\bf\it k})=\sum_{\mu,\mu^{\prime}=0,\cdots n_{s}-1}S_{\mu,\mu^{\prime}}({\bf\it k}),
Sμ,μ(𝒌)=PμPμm,m=1,nfvμ,𝒌(m)(U𝒌1)mmvμ,𝒌(m).\displaystyle S_{\mu,\mu^{\prime}}({\bf\it k})=P_{\mu}P^{*}_{\mu^{\prime}}\sum_{m,m^{\prime}=1,\cdots n_{f}}v^{(m)}_{\mu,{\bf\it k}}(U^{-1}_{{\bf\it k}})_{mm^{\prime}}v^{(m^{\prime})*}_{\mu^{\prime},{\bf\it k}}. (S4)

Here, U(𝒌)U({\bf\it k}) is the Gram matrix composed of the degenerate Bloch CLS bases, which are not necessarily orthogonal or normalized;

U𝒌=[(𝒗𝒌(1)𝒗𝒌(1))(𝒗𝒌(1)𝒗𝒌(2))(𝒗𝒌(1)𝒗𝒌(nf))(𝒗𝒌(2)𝒗𝒌(1))(𝒗𝒌(2)𝒗𝒌(2))(𝒗𝒌(2)𝒗𝒌(nf))(𝒗𝒌(nf)𝒗𝒌(1))(𝒗𝒌(nf)𝒗𝒌(2))(𝒗𝒌(nf)𝒗𝒌(nf))].U_{{\bf\it k}}\!=\!\begin{bmatrix}({{\bf\it v}}^{(1)*}_{{\bf\it k}}\cdot{{\bf\it v}}^{(1)}_{{\bf\it k}})&({{\bf\it v}}^{(1)*}_{{\bf\it k}}\cdot{{\bf\it v}}^{(2)}_{{\bf\it k}})&\cdots&({{\bf\it v}}^{(1)*}_{{\bf\it k}}\cdot{{\bf\it v}}^{(n_{f})}_{{\bf\it k}})\\ ({{\bf\it v}}^{(2)*}_{{\bf\it k}}\cdot{{\bf\it v}}^{(1)}_{{\bf\it k}})&({{\bf\it v}}^{(2)*}_{{\bf\it k}}\cdot{{\bf\it v}}^{(2)}_{{\bf\it k}})&\cdots&({{\bf\it v}}^{(2)*}_{{\bf\it k}}\cdot{{\bf\it v}}^{(n_{f})}_{{\bf\it k}})\\ \vdots&\vdots&\ddots&\vdots\\ ({{\bf\it v}}^{(n_{f})*}_{{\bf\it k}}\cdot{{\bf\it v}}^{(1)}_{{\bf\it k}})&({{\bf\it v}}^{(n_{f})*}_{{\bf\it k}}\cdot{{\bf\it v}}^{(2)}_{{\bf\it k}})&\cdots&({{\bf\it v}}^{(n_{f})*}_{{\bf\it k}}\cdot{{\bf\it v}}^{(n_{f})}_{{\bf\it k}})\end{bmatrix}. (S5)

PμP_{\mu} is the sublattice μ\mu component of the polarizer, and takes the form Pμ=ei𝒌𝒓μP_{\mu}=e^{-i{\bf\it k}\cdot{\bf\it r}_{\mu}} in the simplest case of photoemission spectroscopy, considered in the main text. If the flat band is not degenerate, the expression Eq. (S4), combined with Eq. (S5), is reduced to Eq. (12) in the main text.

.2.2 kagome lattice

Only a single flat band exists for the kagome lattice model. As written as Eq. (5) in the main text, the Bloch CLS takes the form,

𝒗𝒌=[v0,𝒌v1,𝒌v2,𝒌]=[eik2eik11eik2eik11].\displaystyle{\bf\it v}_{{\bf\it k}}=\begin{bmatrix}v_{0,{\bf\it k}}\\ v_{1,{\bf\it k}}\\ v_{2,{\bf\it k}}\end{bmatrix}=\begin{bmatrix}e^{-ik_{2}}-e^{-ik_{1}}\\ 1-e^{-ik_{2}}\\ e^{-ik_{1}}-1\end{bmatrix}. (S6)

Putting this expression into Eq. (S4), we obtain

𝒮(𝒌)=|sink1k22+sink22sink12|2sin2k1k22+sin2k22+sin2k12.\mathcal{S}({\bf\it k})=\frac{\Big{|}\sin\frac{k_{1}-k_{2}}{2}\!+\!\sin\frac{k_{2}}{2}\!-\!\sin\frac{k_{1}}{2}\Big{|}^{2}}{\sin^{2}\frac{k_{1}-k_{2}}{2}\!+\!\sin^{2}\frac{k_{2}}{2}\!+\!\sin^{2}\frac{k_{1}}{2}}. (S7)

By choosing the lattice coordinates, 𝒂1=(1,0){\bf\it a}_{1}=(1,0) and 𝒂2=(12,32){\bf\it a}_{2}=(\frac{1}{2},\frac{\sqrt{3}}{2}), which gives k1=kxk_{1}=k_{x} and k2=kx+3ky2k_{2}=\frac{k_{x}+\sqrt{3}k_{y}}{2}, we obtain S(𝒌)S({\bf\it k}) as depicted in Fig. 2(a) in the main text.

.2.3 pyrochlore lattice

A pyrochlore lattice has two-fold degenerate flat bands (except for the spin degeneracy). The two basis vectors of Bloch CLS can be chosen as

𝒗𝒌(1)=[v0,𝒌(1)v1,𝒌(1)v2,𝒌(1)v3,𝒌(1)]=[eik3eik11eik30eik11],\displaystyle{\bf\it v}^{(1)}_{{\bf\it k}}=\begin{bmatrix}v^{(1)}_{0,{\bf\it k}}\\ v^{(1)}_{1,{\bf\it k}}\\ v^{(1)}_{2,{\bf\it k}}\\ v^{(1)}_{3,{\bf\it k}}\end{bmatrix}=\begin{bmatrix}e^{-ik_{3}}-e^{-ik_{1}}\\ 1-e^{-ik_{3}}\\ 0\\ e^{-ik_{1}}-1\end{bmatrix}, (S8)
𝒗𝒌(2)=[v0,𝒌(2)v1,𝒌(2)v2,𝒌(2)v3,𝒌(2)]=[eik3eik201eik3eik21].\displaystyle{\bf\it v}^{(2)}_{{\bf\it k}}=\begin{bmatrix}v^{(2)}_{0,{\bf\it k}}\\ v^{(2)}_{1,{\bf\it k}}\\ v^{(2)}_{2,{\bf\it k}}\\ v^{(2)}_{3,{\bf\it k}}\end{bmatrix}=\begin{bmatrix}e^{-ik_{3}}-e^{-ik_{2}}\\ 0\\ 1-e^{-ik_{3}}\\ e^{-ik_{2}}-1\end{bmatrix}. (S9)

𝒗𝒌(1){\bf\it v}^{(1)}_{{\bf\it k}} and 𝒗𝒌(2){\bf\it v}^{(2)}_{{\bf\it k}} are respectively constructed from the primitive CLSs localized on kagome planes, 𝒂1𝒂3{\bf\it a}_{1}-{\bf\it a}_{3} and 𝒂2𝒂3{\bf\it a}_{2}-{\bf\it a}_{3}, as defined in Sec. .1.

By inputting them into Eqs. (S4) and (S5), we find the diagonal components of Sμ,μ(𝒌)S_{\mu,\mu^{\prime}}({\bf\it k}) satisfy

μ=0,1,2,3Sμ,μ(𝒌)=2.\displaystyle\sum_{\mu=0,1,2,3}S_{\mu,\mu}({\bf\it k})=2. (S10)

For the off-diagonal components, μμ\mu\not=\mu^{\prime}, we obtain

Sμ,μ(𝒌)=Ω𝒌1[μ′′=0,1,2,3sinkμkμ′′2sinkμ′′kμ2]1,\displaystyle S_{\mu,\mu^{\prime}}({\bf\it k})=\Omega_{{\bf\it k}}^{-1}\bigg{[}\sum_{\mu^{\prime\prime}=0,1,2,3}\sin\frac{k_{\mu}-k_{\mu^{\prime\prime}}}{2}\sin\frac{k_{\mu^{\prime\prime}}-k_{\mu^{\prime}}}{2}\bigg{]}^{-1},

where k0=0k_{0}=0, and

Ω𝒌\displaystyle\Omega_{{\bf\it k}} =\displaystyle= sin2k12+sin2k22+sin2k32+sin2k1k22+sin2k1k32+sin2k2k32,\displaystyle\sin^{2}\frac{k_{1}}{2}+\sin^{2}\frac{k_{2}}{2}+\sin^{2}\frac{k_{3}}{2}+\sin^{2}\frac{k_{1}-k_{2}}{2}+\sin^{2}\frac{k_{1}-k_{3}}{2}+\sin^{2}\frac{k_{2}-k_{3}}{2}, (S12)

and each component is given as

S01(𝒌)=Ω𝒌1[sink22sink1k22+sink32sink1k32],\displaystyle S_{01}({\bf\it k})=\Omega_{{\bf\it k}}^{-1}\bigl{[}\sin\frac{k_{2}}{2}\sin\frac{k_{1}-k_{2}}{2}+\sin\frac{k_{3}}{2}\sin\frac{k_{1}-k_{3}}{2}\bigr{]},
S02(𝒌)=Ω𝒌1[sink12sink1k22+sink32sink2k32],\displaystyle S_{02}({\bf\it k})=\Omega_{{\bf\it k}}^{-1}\bigl{[}-\sin\frac{k_{1}}{2}\sin\frac{k_{1}-k_{2}}{2}+\sin\frac{k_{3}}{2}\sin\frac{k_{2}-k_{3}}{2}\bigr{]},
S03(𝒌)=Ω𝒌1[sink12sink1k32sink22sink2k32],\displaystyle S_{03}({\bf\it k})=\Omega_{{\bf\it k}}^{-1}\bigl{[}-\sin\frac{k_{1}}{2}\sin\frac{k_{1}-k_{3}}{2}-\sin\frac{k_{2}}{2}\sin\frac{k_{2}-k_{3}}{2}\bigr{]},
S12(𝒌)=Ω𝒌1[sink12sink22sink1k32sink2k32],\displaystyle S_{12}({\bf\it k})=\Omega_{{\bf\it k}}^{-1}\bigl{[}-\sin\frac{k_{1}}{2}\sin\frac{k_{2}}{2}-\sin\frac{k_{1}-k_{3}}{2}\sin\frac{k_{2}-k_{3}}{2}\bigr{]},
S13(𝒌)=Ω𝒌1[sink12sink32+sink1k22sink2k32],\displaystyle S_{13}({\bf\it k})=\Omega_{{\bf\it k}}^{-1}\bigl{[}-\sin\frac{k_{1}}{2}\sin\frac{k_{3}}{2}+\sin\frac{k_{1}-k_{2}}{2}\sin\frac{k_{2}-k_{3}}{2}\bigr{]},
S23(𝒌)=Ω𝒌1[sink22sink32sink1k22sink1k32].\displaystyle S_{23}({\bf\it k})=\Omega_{{\bf\it k}}^{-1}\bigl{[}-\sin\frac{k_{2}}{2}\sin\frac{k_{3}}{2}-\sin\frac{k_{1}-k_{2}}{2}\sin\frac{k_{1}-k_{3}}{2}\bigr{]}.
(S13)

The other components are given as Sμ,μ(𝒌)=Sμ,μ(𝒌)S_{\mu,\mu^{\prime}}({\bf\it k})=S_{\mu^{\prime},\mu}({\bf\it k}). We finally obtain the exact form of the structure factor as

S(𝒌)\displaystyle S({\bf\it k}) =\displaystyle= 2(1+S01(𝒌)+S02(𝒌)+S03(𝒌)+S12(𝒌)+S13(𝒌)+S23(𝒌)).\displaystyle 2(1+S_{01}({\bf\it k})+S_{02}({\bf\it k})+S_{03}({\bf\it k})+S_{12}({\bf\it k})+S_{13}({\bf\it k})+S_{23}({\bf\it k})). (S14)

For a reciprocal vector 𝒌=(k1𝒃1+k2𝒃2+k3𝒃3)/2π{\bf\it k}=(k_{1}{\bf\it b}_{1}+k_{2}{\bf\it b}_{2}+k_{3}{\bf\it b}_{3})/2\pi, with k1=(kxky)/2k_{1}=(k_{x}-k_{y})/2, k2=(ky+kz)/2k_{2}=(-k_{y}+k_{z})/2, and k3=(kx+kz)/2k_{3}=(k_{x}+k_{z})/2. Putting them into Eqs. (S12), (S13) and (S14), we obtain the structure factor in a symmetric form,

Ω𝒌=sin2kx+ky4+sin2ky+kz4+sin2kz+kx4+sin2kxky4+sin2kykz4+sin2kzkx4,\displaystyle\Omega_{{\bf\it k}}=\sin^{2}{\frac{k_{x}+k_{y}}{4}}+\sin^{2}{\frac{k_{y}+k_{z}}{4}}+\sin^{2}{\frac{k_{z}+k_{x}}{4}}+\sin^{2}{\frac{k_{x}-k_{y}}{4}}+\sin^{2}{\frac{k_{y}-k_{z}}{4}}+\sin^{2}{\frac{k_{z}-k_{x}}{4}}, (S15)
S(𝒌)=2Ω𝒌[1\displaystyle S({\bf\it k})=\frac{2}{\Omega_{{\bf\it k}}}\Bigl{[}1- (sin(kxky4)sin(kxkz4)+sin(kykx4)sin(kykz4)+sin(kzkx4)sin(kzky4)\displaystyle\Bigl{(}\sin{\frac{k_{x}-k_{y}}{4}}\sin{\frac{k_{x}-k_{z}}{4}}+\sin{\frac{k_{y}-k_{x}}{4}}\sin{\frac{k_{y}-k_{z}}{4}}+\sin{\frac{k_{z}-k_{x}}{4}}\sin{\frac{k_{z}-k_{y}}{4}}
+sin(kx+ky4)sin(kx+kz4)+sin(ky+kx4)sin(ky+kz4)+sin(kz+kx4)sin(kz+ky4))\displaystyle+\sin{\frac{k_{x}+k_{y}}{4}}\sin{\frac{k_{x}+k_{z}}{4}}+\sin{\frac{k_{y}+k_{x}}{4}}\sin{\frac{k_{y}+k_{z}}{4}}+\sin{\frac{k_{z}+k_{x}}{4}}\sin{\frac{k_{z}+k_{y}}{4}}\Bigr{)}
+sin(kx+ky4)(sin(kzkx4)+sin(kzky4))+sin(ky+kz4)(sin(kxky4)+sin(kxkz4))\displaystyle+\sin{\frac{k_{x}+k_{y}}{4}}\Bigl{(}\sin{\frac{k_{z}-k_{x}}{4}}+\sin{\frac{k_{z}-k_{y}}{4}}\Bigr{)}+\sin{\frac{k_{y}+k_{z}}{4}}\Bigl{(}\sin{\frac{k_{x}-k_{y}}{4}}+\sin{\frac{k_{x}-k_{z}}{4}}\Bigr{)}
+sin(kz+kx4)(sin(kykx4)+sin(kykz4))].\displaystyle+\sin{\frac{k_{z}+k_{x}}{4}}\Bigl{(}\sin{\frac{k_{y}-k_{x}}{4}}+\sin{\frac{k_{y}-k_{z}}{4}}\Bigr{)}\Bigr{]}. (S16)

From these expressions, we can obtain the plots in Fig. 2(b) and (c) in the main text.

.3 Spin-orbit coupled flat band states of the pyrochlore lattice

We introduce the spin-orbit coupled flat band model on a pyrochlore lattice [49, 50], which was originally introduced to address the exotic low-temperature insulating phase of CsW2O6 [58]. The tight-binding part of the model contains two parameters, spin-orbit coupling (SOC) λ\lambda and a standard transfer integral tt, and the Hamiltonian is written as

kin=teffi,j𝒄iUij(θ)𝒄j,\displaystyle\mathcal{H}_{\rm kin}=-t_{\rm eff}\sum_{\langle i,j\rangle}{\bf\it c}^{\dagger}_{i}U_{ij}(\theta){\bf\it c}_{j}, (S17)

with 𝒄i=(ci,ci)T{\bf\it c}_{i}=(c_{i\uparrow},c_{i\downarrow})^{\rm T} and teff=t2+λ2t_{\rm eff}=\sqrt{t^{2}+\lambda^{2}}. An SU(2) matrix Uij(θ)=eiθ𝝂^ij𝝈/2U_{ij}(\theta)=e^{-i\theta\hat{{\bf\it\nu}}_{ij}\cdot{\bf\it\sigma}/2} rotates the electron spin by angle θ=2arctan(λ/t)\theta=2\arctan(\lambda/t) about the 𝝂^ij\hat{{\bf\it\nu}}_{ij}-axis when it hops from site jj to ii, and this angle varies with λ/t\lambda/t. Here, we take teff=t2+λ2=1t_{\rm eff}=\sqrt{t^{2}+\lambda^{2}}=1 as an energy unit. When θ=2arctan(22)\theta=-2\arctan(2\sqrt{2}) we have SOC flat bands.

In the SOC flat band state, the relative angle of spin orientation of the four sublattices is quenched while totally the SU(2) symmetry is kept. These spin configurations are given by setting a global spinor state 𝝌{\bf\it\chi} and applying a set of operator Γμ\Gamma_{\mu}, which is the π\pi rotation of the spins about the axis 𝒎μ\bm{m}_{\mu} that points from the sublattice to the center of the tetrahedron, given as

𝒎0=(1,1,1)/2,𝒎1=(1,1,1)/2,\displaystyle\bm{m}_{0}=(1,-1,1)/2,\quad\bm{m}_{1}=(-1,1,1)/2,
𝒎2=(1,1,1)/2,𝒎3=(1,1,1)/2.\displaystyle\bm{m}_{2}=(1,1,-1)/2,\quad\bm{m}_{3}=(-1,-1,-1)/2. (S18)

When we set 𝝌=(1,0)T{\bf\it\chi}=(1,0)^{T} pointing in the zz-direction of the pyrochlore xyzxyz-axis, the orientation of spins on four sublattices yields,

|𝒔μ\displaystyle|{\bf\it s}_{\mu}\rangle =(cos(αμ/2),eiϕμ/2sin(αμ/2))T,\displaystyle=(\cos(\alpha_{\mu}/2),e^{i\phi_{\mu}/2}\sin(\alpha_{\mu}/2))^{T},
(αμ,ϕμ)={(2π/3,π/4)μ=0(2π/3,3π/4)μ=1(π/3,π/4)μ=2(π/3,5π/4)μ=3,\displaystyle(\alpha_{\mu},\phi_{\mu})=\left\{\begin{array}[]{ll}(2\pi/3,-\pi/4)&\mu=0\\ (2\pi/3,3\pi/4)&\mu=1\\ (\pi/3,\pi/4)&\mu=2\\ (\pi/3,5\pi/4)&\mu=3\\ \end{array}\right., (S23)

where αμ\alpha_{\mu} and ϕμ\phi_{\mu} are the zenithal and azimuthal angles of spins.

If finite Hubbard interaction is considered, the NeN_{e}-electron ground state |ΨNe|\Psi^{N_{e}}\rangle is still exactly obtained at quarter-filling for θ=2arctan(22)\theta=-2\arctan(2\sqrt{2}), where the lowest flat band is half occupied. The ground state can be written as

|ΨNe=mFB𝒌c𝒌,m,𝒔μ|0,\displaystyle|\Psi^{N_{e}}\rangle=\prod_{m\in{\rm FB}}\prod_{{\bf\it k}}c^{{\dagger}}_{{\bf\it k},m,{\bf\it s}_{\mu}}|0\rangle, (S24)

where

c𝒌,m,𝒔μ=1Ncl=0Nc1\displaystyle c^{{\dagger}}_{{\bf\it k},m,{\bf\it s}_{\mu}}=\frac{1}{\sqrt{N_{c}}}\sum_{l=0}^{N_{c}-1} ei𝒌𝑹lμ=0ns1v~μ,𝒌(m)|𝒗~𝒌(m)|σ=,cl,μ,σσ|𝒔μ.\displaystyle e^{i{\bf\it k}\cdot{\bf\it R}_{l}}\sum_{\mu=0}^{n_{s}-1}\frac{\tilde{v}^{(m)}_{\mu,\bm{k}}}{|{\bf\it\tilde{v}}^{(m)}_{{\bf\it k}}|}\sum_{\sigma=\uparrow,\downarrow}c^{{\dagger}}_{l,\mu,\sigma}\langle\sigma|{\bf\it s}_{\mu}\rangle. (S25)

c𝒌,m,𝒔μc^{{\dagger}}_{{\bf\it k},m,{\bf\it s}_{\mu}} creates an electron at the mm-th flat band with the spin state |𝒔μ|{\bf\it s}_{\mu}\rangle. Here, the coefficients of flat band Bloch function, 𝒗~𝒌(m){\bf\it\tilde{v}}^{(m)}_{{\bf\it k}} are chosen to be orthogonal to each other.

In the inverse spin-polarized photoemission spectroscopy, we take out one electron at the spin state |𝒔e|{\bf\it s}_{e}\rangle. The corresponding matrix element is evaluated as

𝒮(𝒌)=mFB|ΨNe|a𝒌c𝒌,m,𝒔μ|ΨNe|2,\displaystyle\mathcal{S}({\bf\it k})=\sum_{m\in{\rm FB}}|\langle\Psi^{N_{e}}|a^{\dagger}_{{\bf\it k}}c_{{\bf\it k},m,{\bf\it s}_{\mu}}|\Psi^{N_{e}}\rangle|^{2}, (S26)

where the creation operator of plane wave state is written as

a𝒌=1Ncl=0Nc1μ=0ns1ei𝒌(𝑹l+𝒓μ)σ=,cl,μ,σσ|𝒔e.a^{{\dagger}}_{{\bf\it k}}=\frac{1}{\sqrt{N_{c}}}\sum_{l=0}^{N_{c}-1}\sum_{\mu=0}^{n_{s}-1}e^{i{\bf\it k}\cdot({\bf\it R}_{l}+{\bf\it r}_{\mu})}\sum_{\sigma=\uparrow,\downarrow}c^{{\dagger}}_{l,\mu,\sigma}\langle\sigma|{\bf\it s}_{e}\rangle. (S27)

From Eqs. (S25) and (S27), we find

ΨNe|c𝒌,m,𝒔μa𝒌|ΨNe=ei𝒌𝒓μ𝒔e|𝒔μmFBv~μ,𝒌(m)|𝒗~𝒌(m)|.\displaystyle\langle\Psi^{N_{e}}|c^{{\dagger}}_{{\bf\it k},m,{\bf\it s}_{\mu}}a_{{\bf\it k}}|\Psi^{N_{e}}\rangle=e^{-i{\bf\it k}\cdot{\bf\it r}_{\mu}}\langle{\bf\it s}_{e}|{\bf\it s}_{\mu}\rangle\sum_{m\in{\rm FB}}\frac{\tilde{v}^{(m)}_{\mu,\bm{k}}}{|{\bf\it\tilde{v}}^{(m)}_{{\bf\it k}}|}. (S28)

The prefactor Pμ=ei𝒌𝒓μ𝒔e|𝒔μP_{\mu}=e^{-i{\bf\it k}\cdot{\bf\it r}_{\mu}}\langle{\bf\it s}_{e}|{\bf\it s}_{\mu}\rangle accounts for the polarizer in this case. We define the zenithal and azimuthal angles of incident spin σ\sigma as (θe,ϕe)(\theta_{\rm e},\phi_{\rm e}) and set ϕe=0\phi_{\rm e}=0, and obtain the spectrum shown in Fig. 2(d) in the main text for θe=0,π/2\theta_{\rm e}=0,\pi/2, and π\pi.