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

Tunnelling theory of Weyl semimetals in proximity to a metallic band

L. Goutte [email protected]    T. Pereg-Barnea Department of Physics and Centre for the Physics of Materials
McGill University, Montréal, Québec, H3A 2T8, Canada
Abstract

We study the effects of tunnelling on the band structure and Fermi arc of a time-reversal broken Weyl semimetal (WSM). When coupled to a non-magnetic parabolic band, the WSM’s chiral arc state lowers in energy and forms, together with a previously extended state, a noticeable spin-dependent asymmetry in the interface spectrum in the vicinity of the Weyl nodes. We study these effects with a lattice model which we solve numerically on a finite sample and analytically through using an ansatz on an infinite sample. Our ansatz agrees very well with the numerical simulation as it accurately describes the behaviour of the chiral state, from its energy asymmetry to the spin canting at the interface. We find that the tunnelling effectively increases the Fermi arc length, allowing for the presence of interface states beyond the bare Weyl nodes. These additional states may carry current along the interface and their contribution can be detected in the conductance. Across the interface, the spin-independent conductance reproduces the results of an electron tunnelling experiment to reveal the WSM’s density of states. Besides conductivity, the effect of tunneling between the WSM and the metallic band can be seen in quantum oscillations experiments which we briefly comment about.

Weyl semimetals, surface tunnelling
preprint: APS/123-QED

I Introduction

Weyl semimetals (WSMs) are materials whose low-energy excitations are Weyl fermions [1, 2, 3]. While these particles have their roots in high-energy physics as solutions to the massless three-dimensional Dirac equation in a chiral basis, WSMs present an elegant way of accessing their properties in the condensed matter regime. A growing interest in these materials culminated with their physical realization in TaAs [4] and TaNb[5], with additional predictions of type-II WSMs in WTe2\text{WTe}_{2} [6] and MoTe2\text{MoTe}_{2} [7]. On the theoretical side, the WSM’s classification as a gapless topological phase makes it an appealing object of study with deep connections to topological Chern insulators [8] and novel properties in the presence of superconductivity [9, 10] and external magnetic fields [11], to name but a few.

The Weyl Hamiltonian describes a linear crossing of two non-degenerate bands. For a pair of such bands to touch, one must in general tune three independent parameters, one for each Pauli matrix. In three spatial dimensions with three independent momenta Weyl points are therefore robust against weak perturbations. Near these points/nodes the bulk energy disperses linearly and the physics are governed by the Weyl Hamiltonian:

H=𝐯0𝐤±v𝐤𝝈,H=\hbar\mathbf{v}_{0}\cdot\mathbf{k}\pm\hbar v\mathbf{k}\cdot\bm{\sigma}, (1)

where ±\pm denotes the node’s chirality, vv is the effective Fermi velocity, 𝐤\mathbf{k} is the momentum and 𝝈\bm{\sigma} is the vector of Pauli matrices acting in spin space. The first term, proportional to the unit matrix, breaks Lorentz invariance and tilts the dispersion. For type-I WSMs, it can be ignored, leaving only the second term. The latter has a linear dispersion that, while strongly reminiscent of two-dimensional graphene, will not open a gap in the presence of small perturbations. Each Weyl node is also a monopole of Berry curvature, leading to a chiral anomaly which manifests itself in many exotic properties such as the Quantum anomalous Hall effect, negative magnetoresistance [12], the chiral magnetic effect [13], and high carrier mobility [14].

In lattice systems, a Weyl semimetal hosts pairs of Weyl nodes along a given nodal direction [15, 16, 1]. This is required by either time-reversal (𝒯\mathcal{T}) or inversion (\mathcal{I}) symmetry and the fact that the total Berry flux in the first Brillouin zone (BZ) must vanish. One can slice the system along the nodal direction and assign a Chern number to each two dimensional slice of momentum space: if a plane is pierced by Berry flux it will be topologically non-trivial, and vice-versa. Therefore, the bulk-boundary correspondence implies the presence of topologically protected surface states in between the Weyl nodes only. At the Fermi level, then, an open system will host a Fermi arc – a projection of zero-energy chiral surface states connecting pairs of opposite chirality Weyl nodes and dispersing linearly away from the Fermi level. In this sense, gapless topological phases are intermediaries between genuine trivial and topological phases of matter and can even be realized by a repeated stacking of the two [16].

While a growing number of their properties are known, such as the effect of impurities and defects [17, 18, 19, 20], the manoeuvrability and theoretical richness of these materials further motivates the analytical study of tunnelling in WSMs. In what follows, we investigate the effect of tunnelling by constructing a tight-binding model of a time reversal symmetry (𝒯\mathcal{T}) broken WSM coupled to a non-magnetic band. In describing a single Fermi arc, the 𝒯\mathcal{T}-broken WSM displays all aforementioned properties while providing a minimal model to serve as a building block for setups with more pairs of nodes. Likewise, our choice of a simple tunnelling potential and featureless band are intentional: we seek to draw out the bare properties of a WSM in contact with a non-topological material.

The remaining sections are structured as follows. In Sec. II, we present the WSM and non-magnetic band models along with the specific form of surface tunnelling. The numerical results of a finite lattice model are then presented in Sec. III. In Sec. IV we derive an infinite lattice theory with an interface to model the spectra, spin canting and interface arcs in a lattice framework, while Sec. V presents a simpler continuum model. We finish in Sec. VI by investigating the novel transport properties of the coupled system both along and across the interface in the Landauer-Buttiker and electron tunnelling formalism, respectively. Directions for further study are briefly touched upon in the conclusion, Sec. VII, and relevant technical details are included in the appendices.

II Model

II.1 Weyl semimetal

We consider a minimal Hamiltonian which captures the Fermi arc feature. This can be achieved either by breaking 𝒯\mathcal{T} while preserving \mathcal{I} or vice-versa. In order to work with smaller matrices, we choose the former. Explicitly, then, our Hamiltonian must satisfy H(𝐤)=σzH(𝐤)σzH\left(\mathbf{k}\right)=\sigma_{z}H\left(-\mathbf{k}\right)\sigma_{z} and H(𝐤)σyH(𝐤)σyH\left(\mathbf{k}\right)\neq\sigma_{y}H^{*}\left(-\mathbf{k}\right)\sigma_{y}. A simple tight-binding Hamiltonian which abides by these symmetries is (=lattice constant=1\hbar=\text{lattice constant}=1) [1]

Hw=𝐤𝐜𝐤wbulk(𝐤)𝐜𝐤,\displaystyle{H}_{w}=\sum_{\mathbf{k}}\mathbf{c}^{\dagger}_{\mathbf{k}}\mathcal{H}^{\mathrm{bulk}}_{w}\left(\mathbf{k}\right)\mathbf{c}_{\mathbf{k}}, (2a)
wbulk(𝐤)=txsin(kx)σx+tysin(ky)σy+tzm(𝐤)σz,\displaystyle\mathcal{H}^{\mathrm{bulk}}_{w}\left(\mathbf{k}\right)=t_{x}\sin{k_{x}}\sigma_{x}+t_{y}\sin{k_{y}}\sigma_{y}+t_{z}m\left(\mathbf{k}\right)\sigma_{z}, (2b)
m(𝐤)=(2+γcos(kx)cos(ky)cos(kz)).\displaystyle m\left(\mathbf{k}\right)=\left(2+\gamma-\cos{k_{x}}-\cos{k_{y}}-\cos{k_{z}}\right). (2c)

Here, 𝐜𝐤=(c𝐤,,c𝐤,)\mathbf{c}_{\mathbf{k}}=\left(c_{\mathbf{k},\uparrow},c_{\mathbf{k},\downarrow}\right)^{\top} is an annihilation operator in momentum space, tst_{s} (s=x,y,zs=x,y,z) is the strength of hopping in the ss-direction and σ{\sigma} are the Pauli spin matrices. We further set tx=ty=tz=t>0t_{x}=t_{y}=t_{z}=t>0 for simplicity. The Hamiltonian (2) admits the bulk energies

E±=±t[sin2kx+sin2ky+m2(𝐤)]12.E_{\pm}=\pm t\left[\sin^{2}{k_{x}}+\sin^{2}{k_{y}}+m^{2}\left(\mathbf{k}\right)\right]^{\frac{1}{2}}. (3)

These vanish at 𝐤w±=(0,0,±arccos(γ))(0,0,±kw)\mathbf{k}^{\pm}_{w}=\left(0,0,\pm\arccos{\gamma}\right)\equiv\left(0,0,\pm k_{w}\right) – the aforementioned Weyl nodes. We emphasize the importance of the cos(kx,y)\cos{k_{x,y}} terms, without which there would be more than two nodes in the BZ for a given γ\gamma.

These gapless bulk momenta 𝐤w±\mathbf{k}^{\pm}_{w} suggest that HwH_{w} exhibits different phases that depend solely on the arc length parameter γ\gamma. For γ>1\gamma>1, m(𝐤)>0m\left(\mathbf{k}\right)>0 for all 𝐤\mathbf{k} and the system is trivially gapped. As γ\gamma decreases to 11, a pair of Weyl nodes appear at the origin and move outward along kzk_{z} as γ\gamma decreases further. This defines a gapless topological phase whereby a nonzero Berry flux flows within the momentum range |kz|<kw|k_{z}|<k_{w} from the node of negative chirality to the one of positive chirality. Consequently, the Chern number – defined for a fixed kzk_{z} – is nonzero between the nodes, and zero beyond them. When γ1\gamma\leq-1, the Weyl nodes reach the BZ boundaries and disappear, leaving the bulk dispersion with an inverted band gap. Between 5<γ<1-5<\gamma<-1, the same process occurs for Weyl nodes with (kx,ky)=(0,π)(k_{x},k_{y})=(0,\pi), (π,0)(\pi,0), and (π,π)(\pi,\pi), until γ<5\gamma<-5 where the system is again gapped and trivial for all 𝐤\mathbf{k}. In all numerical results that follow, we take γ=0\gamma=0 (kw=π/2k_{w}=\pi/2), well within the gapless topological regime and with a Fermi arc length karc=πk_{\mathrm{arc}}=\pi. The bare WSM’s surface spectrum, Fermi arc, and topological phases are shown in Fig. 1.

Refer to caption
Figure 1: The minimal Weyl semimetal model. (a) Spectral function at y=Ly1y=L_{y}-1 of a WSM open in yy plotted in the EE-kxk_{x} plane for fixed kz=0k_{z}=0 and (b) kz=π/2k_{z}=\pi/2. (c) WSM spectral function plotted in the kxk_{x}-kzk_{z} plane for fixed E=0E=0 showing the Fermi arc. (d) Phase diagram of Eq. (2) with lower band’s Chern numbers. The phase boundaries γ=cos(kx)\gamma=\cos{k_{x}}, γ=cos(kx)2\gamma=\cos{k_{x}}-2 and γ=cos(kz)4\gamma=\cos{k_{z}}-4 are plotted in blue. We will work at γ=0\gamma=0 (dashed orange line).

II.2 Tunnelling

Refer to caption
Figure 2: (a) Schematic of the WSM-metal system. Only the rightmost surface, or interface (with the Fermi arc shown as a white line and the nodes as white crosses) is linked to the metal via tunnelling Δ\Delta. (b) Physical representation of the system as a chain (𝐑w\mathbf{R}_{w}, 𝐡w\mathbf{h}_{w}, etc. defined in App. A). (c) By integrating out the metal degrees of freedom, the chain is simplified into a single semi-infinite chain with a single edge site of energy hΔ=TGmTh_{\Delta}=T^{\dagger}G_{m}T (shaded with grey line).

To draw out the tunnelling properties of the Weyl semimetal, we couple it to a simple parabolic band via non-magnetic surface tunnelling. The band’s Hamiltonian is spin-independent and reads

Hm=𝐤𝐝𝐤mbulk(𝐤)𝐝𝐤,\displaystyle H_{m}=\sum_{\mathbf{k}}\mathbf{d}^{\dagger}_{\mathbf{k}}\mathcal{H}^{\mathrm{bulk}}_{m}\left(\mathbf{k}\right)\mathbf{d}_{\mathbf{k}}, (4a)
mbulk(𝐤)=2tm(cos(kx)+cos(ky)+cos(kz))μ,\displaystyle\mathcal{H}^{\mathrm{bulk}}_{m}\left(\mathbf{k}\right)=-2t_{m}\left(\cos{k_{x}}+\cos{k_{y}}+\cos{k_{z}}\right)-\mu, (4b)

where tmt_{m} is the hopping amplitude, μ\mu the chemical potential and 𝐝𝐤=(d𝐤,,d𝐤,)\mathbf{d}_{\mathbf{k}}=\left(d_{\mathbf{k},\uparrow},d_{\mathbf{k},\downarrow}\right)^{\top} is an annihilation operator in momentum space. For brevity, we equivalently refer to this non-magnetic parabolic band as “metal”, though one may of course tune μ\mu to achieve a semi-conductor or an insulator, as discussed in App. B, where we also consider two parabolic bands.

We now introduce a tunnelling Hamiltonian which couples the surface of the WSM to the surface of the metal. We proceed with open boundary conditions in the yy-direction and keep well-defined momenta perpendicular to the surface, 𝐤=(kx,kz)\mathbf{k}_{\perp}=\left(k_{x},k_{z}\right). The WSM (metal) side runs from y=Ly+1y=-L_{y}+1 to 0 (y=1y=1 to LyL_{y}), defining an interface between the WSM’s y=0y=0 and metal’s y=1y=1 sites. The Hamiltonian for the full (finite-sized) system is therefore

H=𝐤y,y=Ly+1Ly𝐟𝐤,y(𝐤)y,y𝐟𝐤,y,\displaystyle{H}=\sum_{\mathbf{k_{\perp}}}\sum_{y,y^{\prime}=-L_{y}+1}^{L_{y}}\mathbf{f}_{\mathbf{k}_{\perp},y}^{\dagger}\mathcal{H}\left(\mathbf{k}_{\perp}\right)_{y,y^{\prime}}\mathbf{f}_{\mathbf{k}_{\perp},y^{\prime}}, (5a)
(𝐤)=(wopen(𝐤)TTmopen(𝐤)),\displaystyle\mathcal{H}\left(\mathbf{k}_{\perp}\right)=\begin{pmatrix}\mathcal{H}^{\mathrm{open}}_{w}\left(\mathbf{k}_{\perp}\right)&T^{\dagger}\\ T&\mathcal{H}^{\mathrm{open}}_{m}\left(\mathbf{k}_{\perp}\right)\end{pmatrix}, (5b)

where

𝐟𝐤,y={𝐜𝐤,yLy+1y0𝐝𝐤,y1yLy\mathbf{f}_{\mathbf{k}_{\perp},y}=\begin{cases}\mathbf{c}_{\mathbf{k}_{\perp},y}&-L_{y}+1\leq y\leq 0\\ \mathbf{d}_{\mathbf{k}_{\perp},y}&1\leq y\leq L_{y}\end{cases} (6)

and open\mathcal{H}^{\mathrm{open}} is the partial-in-yy Fourier transform of bulk\mathcal{H}^{\mathrm{bulk}}. The full form of Eq. (5) is shown in App. A. The surface tunnelling term is also non-magnetic and takes the form (T)y,y=Δδ0,Ly1\left(T\right)_{y,y^{\prime}}=\Delta\delta_{0,L_{y}-1}, or

T=(0Δ00),T=\begin{pmatrix}0&\dots&\Delta\\ \vdots&\ddots&\vdots\\ 0&\dots&0\end{pmatrix}, (7)

where, for simplicity, we have assumed that Δ\Delta is a real constant that modulates the tunnelling strength between interface sites y=0y=0 and y=1y=1. Physically, the tunnelling strength can be modified either by varying the metal bandwidth tmt_{m} or changing the interface thickness, as suggested by Fig. 2. There are therefore two competing energy scales at the WSM’s interface: the interlayer hopping tt pulling the electron towards the bulk and the tunnelling strength Δ\Delta pulling the electron towards the metal.

Before moving on to the finite lattice simulations, we note that the metal dynamics can be exactly integrated out to make way for a modified WSM propagator [21, 22]. More precisely, the effective Green’s function becomes

Geff(iωn)=[Gw1(iωn)TGm(iωn)T]1G_{\mathrm{eff}}\left(i\omega_{n}\right)=\left[G_{w}^{-1}\left(i\omega_{n}\right)-T^{\dagger}G_{m}\left(i\omega_{n}\right)T\right]^{-1} (8)

where ωn\omega_{n} is the Matsubara frequency and Gw,m=(iωnw,mopen)1G_{w,m}=\left(i\omega_{n}-\mathcal{H}^{\mathrm{open}}_{w,m}\right)^{-1} are the bare Green’s functions. Substituting in Eqs. (4) and (7) and yields, after some algebra,

TGm(iωn)T=Δ2(iωnhm)24tm2δy,Ly1δy,y,T^{\dagger}G_{m}\left(i\omega_{n}\right)T=-\frac{\Delta^{2}}{\sqrt{\left(i\omega_{n}-h_{m}\right)^{2}-4t_{m}^{2}}}\delta_{y,L_{y}-1}\delta_{y,y^{\prime}}, (9)

where hm=2tm(cos(kx)+cos(kz))μh_{m}=-2t_{m}\left(\cos{k_{x}}+\cos{k_{z}}\right)-\mu. Thus, surface tunnelling simply shifts the same-site hopping of the last site (Fig. 2c).

III Finite lattice model

Refer to caption
Figure 3: Interface density of states for the coupled WSM-metal system for both spins. The numerical results (simulated on a Ly=30L_{y}=30 size chain sampled at 100100 momentum points) are shown in warm colours whereas the chiral state’s infinite lattice model (Sec. IV) is plotted in blue. In all plots where the infinite lattice theory obstructs the numerical results (e.g. the top left), the agreement is near exact. The columns correspond to (a) the spectrum along kxk_{x} at kz=0k_{z}=0, (b) the spectrum along kxk_{x} at the Weyl point kz=+π/2k_{z}=+\pi/2, and the emergent interface arcs at (c) E=0E=0 and (d) E=0.5E=0.5. The rows are set in increasing order of Δ=0\Delta=0, 11, 2.32.3 going down. The bulk energy edges Ebulk=±t[sin2kx+(1+γcos(kx)cos(kz))]12E_{\mathrm{bulk}}=\pm t\left[\sin^{2}{k_{x}}+\left(1+\gamma-\cos{k_{x}}-\cos{k_{z}}\right)\right]^{\frac{1}{2}} are denoted by dashed white lines, as is the Fermi surface in the E=0.5E=0.5 interface plots. The bare Weyl nodes 𝐤,w±=(0,±π/2)\mathbf{k}^{\pm}_{\perp,w}=(0,\pm\pi/2) are white crosses. The fixed parameters used for these and all other plots are t=1t=1, γ=0\gamma=0, tm=0.5t_{m}=0.5, μ=4\mu=-4, unless otherwise specified.

We now turn to the numerical results of Eq. (5) on a finite lattice. Keeping the system open in yy with the quantum numbers kxk_{x} and kzk_{z}, the spectral function is obtained by evaluating A(E,𝐤)=π1Im[Tr(G)]A\left(E,\mathbf{k}_{\perp}\right)=-{\pi}^{-1}\mathrm{Im}\left[\mathrm{Tr}\left(G\right)\right] with the Green’s function

G(E,𝐤)=[E+i0+(𝐤)]1.G\left(E,\mathbf{k}_{\perp}\right)=\left[{E+i0^{+}-\mathcal{H}\left(\mathbf{k}_{\perp}\right)}\right]^{-1}. (10)

The WSM’s interface density of states (IDOS), displayed in Fig. 3, is found by tracing over the y=0y=0 site only.

At kz=0k_{z}=0 (Fig. 3a), we are exactly in between the Weyl nodes. Without tunnelling, only the so-called chiral state is present and localized to the interface, residing on the Fermi arc and dispersing as E=tsin(kx)E=-t\sin{k_{x}} with a spin σx=1\sigma_{x}=-1. With tunnelling, there are two noticeable effects. First, the chiral state lowers its energy as it is now able to hop to the metal side, spreading its wavefunction. Indeed, this lowering of energy captured by Eq. (9) is a prevailing effect throughout this work. By that same token, a previously extended state enters the bulk gap from the upper bulk band and localizes to the interface. Contrary to the chiral state, this so-called emergent interface state does not have a uniform spin polarization.

At the Weyl nodes (Fig. 3b), the Fermi arc terminates and there are no interface states for Δ=0\Delta=0. As tunnelling is increased, however, the chiral state can be seen along the Weyl node’s upper cone. When Δ\Delta increases beyond the interlayer hopping tt the chiral state detaches from the Weyl cone and forms, together with the previously discussed emergent interface state, a noticeable asymmetry in the interface density of states with respect to kxk_{x} reflection (Fig. 3b.iii). This striking asymmetry is of particular interest. Physically, it suggests that tunnelling modifies the group velocity along the interface to produce additional left- and right-flowing current in an energy range between the chiral and emergent interface states’ intersections with the bulk dispersion. Naively, this is surprising because one may not expect the breaking of translation symmetry in yy to induce an asymmetry in the xx-direction. However, one must remember that the physics on a single surface are not in fact symmetric in kxk_{x} to begin with, as evidenced by the linearly dispersing chiral state at the interface. Therefore, although the spectral function is symmetric in kxk_{x} when traced over all sites, the localized tunnelling term in yy will explicitly break this symmetry.

By plotting A(E,𝐤)A\left(E,\mathbf{k}_{\perp}\right) in the surface BZ, we see that the zero-energy interface Fermi arc (Fig. 3c and d) will curve in the presence of tunnelling [23]. While still terminating at the Weyl nodes 𝐤,w±=(0,±kw)\mathbf{k}^{\pm}_{\perp,w}=(0,\pm k_{w}), it does go beyond kz=±kwk_{z}=\pm k_{w} at zero energy, signifying the existence of interface states in a region of parameters outside the bare Fermi arc. This is illustrated by the previously discussed chiral state’s presence at kz=π/2k_{z}=\pi/2 and will have important transport consequences come Sec. VI.1.

These results are robust to changes in the metal’s form. In fact, we find that equivalent behaviour may be obtained simply by coupling the WSM to a constant energy reservoir tm=0t_{m}=0, μ=M\mu=-M. A more realistic setup in which the WSM is coupled to a two-band bulk insulator will yield two copies of the dispersions found in Fig. 3, one for positive and one for negative energy (see App. B).

As seen in the numerics above, a new closed orbit of low energy states appears on the interface. This closed orbit should be apparent in quantum oscillations experiments as it leads to oscillations with frequency which matches the enclosed momentum space area. These oscillations should be contrasted to the arc/node oscillations suggested by Ref. [24] and studied in Ref. [21]. The latter oscillations result from closed orbits which include both the surface (interface) and bulk states, meaning their frequency depends on the slab depth. By contrast, the new orbit seen in Fig. 3d.iii contains only interface states and its frequency is depth independent.

IV Infinite lattice theory with an interface

The physics at the interface seen in the lattice model above can be described in an infinite model and treated analytically with the help of an ansatz. We take LyL_{y}\rightarrow\infty and impose ψ0\psi\rightarrow 0 at y±y\to\pm\infty on both sides of the interface. Therefore, this theory effectively consists of two semi-infinite slabs connected by surface tunnelling Δ\Delta.

Seeking states 𝝋\bm{\varphi} exponentially localized to the interface, we make the ansatz

𝝋={𝝋w(y)=eikxx+ikzzyϕwy=,,1,0𝝋m(y)=eikxx+ikzzmy+1ϕmy=1,2,,\bm{\varphi}=\begin{cases}\bm{\varphi}_{w}(y)=e^{ik_{x}x+ik_{z}z}\ell^{y}\bm{\phi}_{w}&y=-\infty,\dots,-1,0\\ \bm{\varphi}_{m}(y)=e^{ik_{x}x+ik_{z}z}\ell_{m}^{-y+1}\bm{\phi}_{m}&y=1,2,\dots,\infty\end{cases} (11)

where ϕw,m\bm{\phi}_{w,m} are spinors carrying the overall normalization. Note that this ansatz assumes a constant spin direction and is therefore suitable for the chiral state found above but is not completely general. To simplify our problem slightly, we rotate our states by π/2\pi/2 about yy axis in spin space. Defining g1tsin(kx)g_{1}\equiv t\sin{k_{x}} and g3t(2+γcos(kx)cos(kz))g_{3}\equiv t\left(2+\gamma-\cos{k_{x}}-\cos{k_{z}}\right) leads to the change

𝐡w\displaystyle\mathbf{h}_{w} =g1σzg3σx,\displaystyle=g_{1}\sigma_{z}-g_{3}\sigma_{x}, (12a)
𝐑w\displaystyle\mathbf{R}_{w} =t(σx+iσy)/2,\displaystyle=t(\sigma_{x}+i\sigma_{y})/2, (12b)

for the same-site and nearest-neighbour hopping matrices, respectively. The metal and tunnelling components are unchanged.

IV.1 Δ=0\Delta=0

As a first test of validity, we take the Δ=0\Delta=0 case and recover the chiral state and Fermi arc of the finite lattice model. We do not impose any boundary conditions at the yy-termination but instead just look for exponentially localized states which solve the bulk difference equations. In a lattice formalism, wopen𝝋w=E𝝋w\mathcal{H}^{\mathrm{open}}_{w}\bm{\varphi}_{w}=E\bm{\varphi}_{w} produces a set of coupled difference equations relating 𝝋w(y)\bm{\varphi}_{w}(y) to its nearest neighbours 𝝋w(y±1)\bm{\varphi}_{w}(y\pm 1) [25]:

E𝝋w(y)=𝐡w𝝋w(y)+𝐑w𝝋w(y+1)+𝐑w𝝋w(y1).E\bm{\varphi}_{w}(y)=\mathbf{h}_{w}\bm{\varphi}_{w}(y)+\mathbf{R}^{\dagger}_{w}\bm{\varphi}_{w}(y+1)+\mathbf{R}_{w}\bm{\varphi}_{w}(y-1). (13)

Plugging in Eq. (11), we obtain the matrix equation

0=(Eg1σz+g3σxt1σ+tσ)ϕw\displaystyle 0=\left(E-g_{1}\sigma_{z}+g_{3}\sigma_{x}-t\ell^{-1}\sigma_{+}-t\ell\sigma_{-}\right)\bm{\phi}_{w} (14)

where σ±=(σx±iσy)/2\sigma_{\pm}=\left(\sigma_{x}\pm i\sigma_{y}\right)/2. Setting the determinant of Eq. (14) to zero yields the ratio of spins and energy, respectively:

ϕwϕw=E+g1tg3=t1g3Eg1,\displaystyle\frac{\phi^{\uparrow}_{w}}{\phi^{\downarrow}_{w}}=\frac{E+g_{1}}{t\ell-g_{3}}=\frac{t\ell^{-1}-g_{3}}{E-g_{1}}, (15a)
E=±[g12+g32+t2g3t(+1)]12.\displaystyle E=\pm\left[g_{1}^{2}+g_{3}^{2}+t^{2}-g_{3}t\left(\ell+\ell^{-1}\right)\right]^{\frac{1}{2}}. (15b)

Guided by the previous section we make the assumption ϕw/ϕw=0\phi^{\uparrow}_{w}/\phi^{\downarrow}_{w}=0 (spin in the x-x direction) and find a solution with =t/g3\ell=t/g_{3} and E=g1E=-g_{1}. To satisfy the boundary condition at -\infty, we impose Re()>1\mathrm{Re}\left(\ell\right)>1, or g3<tg_{3}<t. The familiar Fermi arc condition γ<cos(kz)\gamma<\cos{k_{z}} then follows naturally111For a bulk state of energy EE, Re()>1\mathrm{Re}\left(\ell\right)>1 implies Ebulk<E<Ebulk-E_{\mathrm{bulk}}<E<E_{\mathrm{bulk}} where Ebulk(𝐤)=[g12+(g3t)2]12E_{\mathrm{bulk}}(\mathbf{k}_{\perp})=\left[g_{1}^{2}+(g_{3}-t)^{2}\right]^{\frac{1}{2}} is the bulk edge.. We have therefore recovered the aforementioned chiral state: a uni-directional interface state on the Fermi arc.

IV.2 Δ>0\Delta>0

We now allow for tunnelling at the interface between the y=0y=0 and y=1y=1 sites. There are then four difference equations, one for each type of site: the Weyl bulk, Weyl interface, metal interface, and metal bulk. Substituting in the supposed forms of 𝝋w,m(y)\bm{\varphi}_{w,m}(y), the difference equations are, respectively:

0=(Eg1σz+g3σxtσt1σ+)ϕw,\displaystyle 0=\left(E-g_{1}\sigma_{z}+g_{3}\sigma_{x}-t\ell\sigma_{-}-t\ell^{-1}\sigma_{+}\right)\bm{\phi}_{w}, (16a)
0=(Eg1σz+g3σxt1σ+)ϕwΔϕm,\displaystyle 0=\left(E-g_{1}\sigma_{z}+g_{3}\sigma_{x}-t\ell^{-1}\sigma_{+}\right)\bm{\phi}_{w}-\Delta\bm{\phi}_{m}, (16b)
0=(Ehm+tmm1)ϕmΔϕw,\displaystyle 0=\left(E-h_{m}+t_{m}\ell_{m}^{-1}\right)\bm{\phi}_{m}-\Delta\bm{\phi}_{w}, (16c)
0=(Ehm+tmm1+tmm)ϕm.\displaystyle 0=\left(E-h_{m}+t_{m}\ell_{m}^{-1}+t_{m}\ell_{m}\right)\bm{\phi}_{m}. (16d)

Eq. (16c) which has no matrix structure and is hence the same for both components of the spinor requires the spinor direction to be the same on both sides of the interface. Moreover, it determines the magnitude ratio:

ϕm=ΔEhm+tmm1ϕw.\bm{\phi}_{m}=\frac{\Delta}{E-h_{m}+t_{m}\ell_{m}^{-1}}\bm{\phi}_{w}. (17)

Together with Eq. (16b), a final relation is obtained:

(EΔ2Ehm+tmm1g1σz+g3σxt1σ+)ϕ𝒘=0.\left(E-\frac{\Delta^{2}}{E-h_{m}+t_{m}\ell_{m}^{-1}}-g_{1}\sigma_{z}+g_{3}\sigma_{x}-t\ell^{-1}\sigma_{+}\right)\bm{\phi_{w}}=0. (18)

Eq. (18) is similar in form and purpose to the effective surface Green’s function (8) except it is purely in spin space since the ansatze and vanishing boundary conditions took care of position dependencies. It can be interpreted as an eigenvalue problem for the matrix g1σzg3σx+1σ+g_{1}\sigma_{z}-g_{3}\sigma_{x}+\ell^{-1}\sigma_{+} whose eigenvalues are

EΔEΔ2Ehm+tmm1.E_{\Delta}\equiv E-\frac{\Delta^{2}}{E-h_{m}+t_{m}\ell_{m}^{-1}}. (19)

With this, its energy bands are twofold and defined by the implicit equation

EηΔ2Eηhm+tmm1=η(g12+g32tg31)12E_{\eta}-\frac{\Delta^{2}}{E_{\eta}-h_{m}+{t_{m}}\ell_{m}^{-1}}=\eta\left({g_{1}^{2}+g_{3}^{2}-tg_{3}\ell^{-1}}\right)^{\frac{1}{2}} (20)

where η=±\eta=\pm is the band index and \ell (m\ell_{m}) is itself a function of energy through Eq. (16a) [Eq. (16d)]:

±\displaystyle\ell_{\pm} =Q±Q21,\displaystyle=Q\pm\sqrt{Q^{2}-1}, (21a)
m,±\displaystyle\ell_{m,\pm} =P±P21.\displaystyle=P\pm\sqrt{P^{2}-1}. (21b)

Here, Q=g12+g32+t2E22g3tQ=\frac{g_{1}^{2}+g_{3}^{2}+t^{2}-E^{2}}{2g_{3}t} and P=hmE2tmP=\frac{h_{m}-E}{2t_{m}}. While it may seem at first glance that the energies are symmetric in kxk_{x} due to the even parity of both \ell and m\ell_{m} with respect to kxk_{x}, one must be careful in choosing the appropriate branch η\eta, such that the state indeed decays away from the interface. In general, the branch may vary as a function of 𝐤\mathbf{k}_{\perp}. The infinite lattice theory Eq. (20) is compared to the finite model in Fig. 3.

Another remarkable consequence of tunnelling is spin canting. At first glance, one should not expect that non-magnetic tunnelling to a non-magnetic metal should cause the polarized spins at the interface to cant. Indeed, this plain intuition is seemingly supported by Eq. (17) and agrees with the finite lattice model for Δt\Delta\lesssim t. To dress a more complete picture, however, we must consider the ratio of spins of the interface state which, in light of Eq. (18), is

ϕwϕw=EΔ+g1g3.\frac{\phi^{\uparrow}_{w}}{\phi^{\downarrow}_{w}}=\frac{E_{\Delta}+g_{1}}{-g_{3}}. (22)

If Δ>0\Delta>0 and E<hm+tmm1E<h_{m}+t_{m}\ell_{m}^{-1}, we expect the interface spin to cant away from ϕw/ϕw=0{\phi^{\uparrow}_{w}}/{\phi^{\downarrow}_{w}}=0 (σx=1\sigma_{x}=-1) towards ϕw/ϕw=1{\phi^{\uparrow}_{w}}/{\phi^{\downarrow}_{w}}=-1 (σz=+1\sigma_{z}=+1). Solving for ϕw/ϕw{\phi^{\uparrow}_{w}}/{\phi^{\downarrow}_{w}} together with Eq. (20) yields the spins of the chiral state as they vary with tunnelling, shown in Fig. 4. We therefore conclude that non-magnetic surface tunnelling to a non-magnetic metal can in fact induce a change in the spins of the WSM’s chiral states. It is also of note that without HwH_{w}’s cos(ky)\cos{k_{y}} term, spin canting is absent and the interface state will remain in a σx=1\sigma_{x}=-1 eigenstate independent of tunnelling.

Refer to caption
Figure 4: The chiral state’s spin at the interface, fixed at the Weyl node kz=π/2{k}_{z}=\pi/2 and kx=0.7k_{x}=-0.7 on a lattice of size Ly=30L_{y}=30. Varying Δ\Delta cants the spin from σx=1\sigma_{x}=-1 towards σz=+1\sigma_{z}=+1 (black arrows), matching the prediction of Eq. (22). The solid (dashed) lines correspond to numerical (infinite lattice theory) results. In the absence of the cos(ky)\cos{k_{y}} term, the spins are unchanged with tunnelling (faded red and blue points). Note that σy\expectationvalue{\sigma_{y}} is always zero (see App. C).

V continuum interface theory

A simplified model that can capture the effect of tunneling is a linearized continuum model which is valid at long distances. We note however that in order to satisfy the boundary conditions at the interface, it is important to keep the second derivative in the yy-direction, as can be seen below.

V.1 Δ=0\Delta=0

We first consider the Δ=0\Delta=0 case, a semi-infinite WSM slab in the continuum limit. Keeping 𝒪(ky2)\mathcal{O}\left(k_{y}^{2}\right) terms in the WSM Hamiltonian, letting ky=iyk_{y}=-i\partial_{y}, and multiplying by iσyi\sigma_{y} throughout yields the differential equation (t=1t=1):

y𝝍+σx2y2𝝍=iσy(Etsin(kx)σx)𝝍+hzσx𝝍{\partial_{y}\bm{\psi}}+\frac{\sigma_{x}}{2}{\partial^{2}_{y}\bm{\psi}}=i\sigma_{y}(E-t\sin{k_{x}}\sigma_{x})\bm{\psi}+h_{z}\sigma_{x}\bm{\psi} (23)

where hzg3th_{z}\equiv g_{3}-t. To hone in on the objective interface states, we take the interface to be at y=0y=0 and make the ansatz 𝝍eκyϕ{\bm{\psi}}\propto e^{\kappa y}\bm{\phi}, where ϕ\bm{\phi} is an unspecified spinor and Re(κ)>0\mathrm{Re}\left(\kappa\right)>0 such that 𝝍0\bm{\psi}\to 0 as yy\to\infty. The differential equation (23) admits four solutions for κ\kappa, of which two have a putatively positive real part:

κ±2=2(1+hz)±2[1+2hz+E2sin2kx]12.\kappa_{\pm}^{2}=2\left(1+h_{z}\right)\pm 2\left[{1+2h_{z}+E^{2}-\sin^{2}{k_{x}}}\right]^{\frac{1}{2}}. (24)

For a state of energy EE, Re(κ)>0\mathrm{Re}\left(\kappa\right)>0 translates to E2<sin2kx+hz2E^{2}<\sin^{2}{k_{x}}+h_{z}^{2}, in agreement with the infinite lattice theory. Eq. (24) sheds light on the fact that for a given eigenvector with energy EE satisfying Eq. (23), there is a distinct eigenvector with equal and opposite energy E-E which is also a solution. Therefore, there are two κ\kappa values per energy.

To determine which solution is correct, we impose the boundary condition 𝝍(0)=0\bm{\psi}(0)=0. In general, one has the superposition 𝝍eκ+yϕκ++αeκyϕκ{\bm{\psi}}\propto e^{\kappa_{+}y}\bm{\phi}_{\kappa_{+}}+\alpha e^{\kappa_{-}y}\bm{\phi}_{\kappa_{-}}. Therefore, α=1\alpha=-1 and ϕκ+=ϕκ\bm{\phi}_{\kappa_{+}}=\bm{\phi}_{\kappa_{-}}. Equating the ratio of spins, the latter condition can be surmised as

E+hzκ+2/2sin(kx)+κ+=E+hzκ2/2sin(kx)+κ.\frac{E+h_{z}-\kappa_{+}^{2}/2}{\sin{k_{x}}+\kappa_{+}}=\frac{E+h_{z}-\kappa_{-}^{2}/2}{\sin{k_{x}}+\kappa_{-}}. (25)

After some algebra, we recover the aforementioned chiral state of energy E=tsin(kx)E=-t\sin{k_{x}}, leading to the decay parameters κ±=1±1+2hz\kappa_{\pm}=1\pm\sqrt{1+2h_{z}} and spin in the negative xx-direction:

𝝍chiraleikx+ikzz(eκ+yeκy)(11).\bm{\psi}_{\mathrm{chiral}}\propto e^{ik_{x}+ik_{z}z}\left(e^{\kappa_{+}y}-e^{\kappa_{-}y}\right)\begin{pmatrix}1\\ -1\end{pmatrix}. (26)

The condition of Re(κ)>0\mathrm{Re}\left(\kappa\right)>0 leads to γ<cos(kz)\gamma<\cos{k_{z}}, which is the familiar arc condition. At the surface BZ origin 𝐤,0=(0,0)\mathbf{k}_{\perp,0}=\left(0,0\right), the chiral state’s decay length is on the order of a lattice length, pointing to a strongly localized state which may therefore well be described by a continuum interface theory. At the surface Weyl points 𝐤,w±=(0,±kw)\mathbf{k}^{\pm}_{\perp,w}=(0,\pm k_{w}), however, κ=0\kappa_{-}=0 and the chiral state’s decay length diverges, as expected from the absence of such surface states at the Weyl node.

V.2 Δ>0\Delta>0

In order to get a simple analytical result, we imagine coupling the WSM to a quantum dot of energy MM. Here, we model the metal as a flat band since it is well above the WSM and only states with the same energy are relevant. The continuum Hamiltonian reads

Hcont=(sin(kx)σx+hzσziσyy12σzy2ΔΔM).\displaystyle H_{\mathrm{cont}}=\begin{pmatrix}\sin{k_{x}}\sigma_{x}+h_{z}\sigma_{z}-i\sigma_{y}\partial_{y}-\frac{1}{2}\sigma_{z}\partial^{2}_{y}&\Delta\\ \Delta&M\end{pmatrix}. (27)

Once again, we focus on solutions bound to the interface 𝝍weκyϕw\bm{\psi}_{w}\propto e^{\kappa y}\bm{\phi}_{w} (𝝍meκmyϕm\bm{\psi}_{m}\propto e^{-\kappa_{m}y}\bm{\phi}_{m}), leading to four differential equations. The first two restrict the metal spins to be identical to the Weyl spins up to a scalar factor:

ϕm=ΔEMϕw.{\bm{\phi}_{m}}=\frac{\Delta}{E-M}{\bm{\phi}_{w}}. (28)

The remaining two equations reduce to a 2×22\times 2 matrix equation expressed in the basis of Weyl spins ϕw\bm{\phi}_{w}:

(EΔ2EMsin(kx)σxhzσz+κ22σz+iκσy)ϕw=0,\displaystyle\left(E-\frac{\Delta^{2}}{E-M}-\sin{k_{x}}\sigma_{x}-h_{z}\sigma_{z}+\frac{\kappa^{2}}{2}\sigma_{z}+i\kappa\sigma_{y}\right){\bm{\phi}_{w}}=0, (29)

which is the continuum form of Eq. (18). When Δ=0\Delta=0 and EME\neq M, it is not difficult to see that the bare WSM surface chiral state is recovered. For Δ>0\Delta>0, the physics are identical to the Δ=0\Delta=0 case with the substitution EEΔ2/(EM)=EΔE\to E-{\Delta^{2}}/(E-M)=E_{\Delta} 222In fact, the effective surface propagator Eq. (9) exactly reduces to Δ2/(EM)-\Delta^{2}/\left(E-M\right) when tm=0t_{m}=0 and μ=M\mu=-M.. For instance, the decay parameters are now

κ±2a2t=2(t+hz)±2(t2+2thzt2sin2kx+EΔ2)12,\kappa_{\pm}^{2}a^{2}t=2\left(t+h_{z}\right)\pm 2\left({t^{2}+2th_{z}-t^{2}\sin^{2}{k_{x}}+E_{\Delta}^{2}}\right)^{\frac{1}{2}}, (30)

where we have re-inserted the energy scale tt and the lattice constant aa.

The continuum interface theory therefore hints at a straightforward interpretation of the energy shift upon tunnelling. Indeed, seeing as the only effect of Δ\Delta was to shift the energies, the chiral band’s energy in the continuum theory is defined by EΔ=tsin(kx)E_{\Delta}=-t\sin{k_{x}}, or

E=Mtsin(kx)212[(M+tsin(kx))2+4Δ2]12.E=\frac{M-t\sin{k_{x}}}{2}-\frac{1}{2}\left[{\left(M+t\sin{k_{x}}\right)^{2}+4\Delta^{2}}\right]^{\frac{1}{2}}. (31)

In regimes where the decay lengths κ±1a\kappa^{-1}_{\pm}\sim a, Eq. (31) is in agreement with finite lattice simulations, as shown in Fig. 5. As for the chiral state’s spin, it remains unchanged due to Eq. (25) still being satisfied and equal to 1-1 when EEΔ=tsin(kx)E\to E_{\Delta}=-t\sin{k_{x}}333In the infinite lattice theory, the replacement EEΔ=tsin(kx)=g1E\to E_{\Delta}=-t\sin{k_{x}}=-g_{1} in Eq.(22) also leads to a spin σx=1\sigma_{x}=-1 (rinterface=0r_{\mathrm{interface}}=0).. Therefore, the validity of Eq. (31) will depend wholly on whether or not the state is in a σx=1\sigma_{x}=-1 eigenstate, and any deviations in the bandstructure must reflect a changing spin in the lattice model. Since the spins do in fact cant for Δt\Delta\gtrsim t, this is the root of the continuum theory’s inaccuracy in this regime.

Refer to caption
Figure 5: Interface density of states for the WSM-metal system (Ly=30L_{y}=30) at kz=0k_{z}=0 for a tunnelling strength of (a) Δ=0\Delta=0 and (b) Δ=1.5\Delta=1.5. The blue line is the analytic chiral state dispersion EchiralE_{\mathrm{chiral}} whereas the dashed white lines represent the bulk energy gap Ebulk=±1E_{\mathrm{bulk}}=\pm 1. The metal energy is M=4M=4.

Another aspect captured by the continuum theory is the localization of bulk states at the interface to produce the emergent interface state, a typical feature of systems with boundary topologies [25]. Simply put, the lowering of energy with tunnelling will give the bulk state’s decay parameter a positive real part.

VI Transport

VI.1 Along the interface

We will now turn to the transport consequences of the previously described theory and numerics. We begin by analyzing the current along the interface, travelling in the xx-direction. We fix kzk_{z} and analyze transport in 2D, summing over all momenta at the end.

At Δ=0\Delta=0, the conductance at the Weyl node should vanish due to the gap closure and subsequent absence of uni-dimensional current-carrying states. For Δ>0\Delta>0, however, the presence of interface states near the Weyl node and the resulting spectral asymmetry in kxk_{x} [Fig. 3 (row 3, col. b)] suggests a jump in group velocity kxE\partial_{k_{x}}E across the Weyl point, leading to a nonzero conductance.

We verify our reasoning numerically via the Landauer-Büttiker formalism, where conductance along the interface 𝒢\mathcal{G}_{\parallel} is defined as [26]

𝒢(E)=e2hTr(GRΓlGAΓr).\mathcal{G}_{\parallel}(E)=\frac{e^{2}}{h}\mathrm{Tr}(G^{R}\Gamma_{l}G^{A}\Gamma_{r}). (32)

Here, GRG^{R} is the usual retarded Green’s function

GR=(EHΣR)1G^{R}=\left({E-H-\Sigma^{R}}\right)^{-1} (33)

with the lead self-energy ΣR=ΣlR+ΣrR\Sigma^{R}=\Sigma^{R}_{l}+\Sigma^{R}_{r} giving the quasiparticles a finite lifetime. The Γl\Gamma_{l} and Γr\Gamma_{r} operators describe the loss of electrons into the left and right leads, respectively:

Γl(r)=i(Σl(r)RΣl(r)A)=2Im(Σl(r)R).\Gamma_{l(r)}=i\left(\Sigma^{R}_{l(r)}-\Sigma^{A}_{l(r)}\right)=-2\,\mathrm{Im}(\Sigma^{R}_{l(r)}). (34)

In the simplest case, we place two leads, one on each of the xx-boundaries, which span the entire sample in the yy-direction. Since the leads are (the interface is) in the plane perpendicular to xx (yy), our construction forces the sample to be open in both the xx-direction and the yy-direction while still remaining periodic in zz. For any kzk_{z}, ΣlR\Sigma^{R}_{l} takes the form

(ΣlR)x,x;y,y=i2τδx0δxxδyy,\left(\Sigma^{R}_{l}\right)_{x,x^{\prime};y,y^{\prime}}=-\frac{i}{2\tau}\delta_{x0}\delta_{xx^{\prime}}\delta_{yy^{\prime}}, (35)

where τ\tau is the quasiparticle’s lifetime. For its part, ΣrR\Sigma^{R}_{r} admits a similar form with δx0\delta_{x0} replaced by δx,Lx1\delta_{x,L_{x}-1}. The tunnelling matrix TT, while unchanged in the yy-direction, now adopts a new diagonal sub-component in the xx-direction:

Tx,x;y,y=Δδxxδy0δyLy.T_{x,x^{\prime};y,y^{\prime}}={\Delta}\delta_{xx^{\prime}}\delta_{y0}\delta_{y^{\prime}L_{y}}. (36)

For Δ=0\Delta=0 (Fig. 6, panels a and b, top row), the e2/he^{2}/h quantized conductance for |kz|<kw|k_{z}|<k_{w} can be understood in the context of the quantum anomalous Hall effect, treating each constant kzk_{z} plane as a 2D quantum spin Hall insulator with one-dimensional edge states carrying 𝒢=e2/h\mathcal{G}_{\parallel}=e^{2}/h [27].

At kz=0k_{z}=0 (Fig. 6a), the surface tunnelling localizes a bulk state to within the gap, allowing for both left- and right-moving carriers to produce a “bump” in the conductance. One can reason by examining the juxtaposed spectrum. Above and below the bump energies (denoted by pink and green lines), there is only one left-moving state, whereas within it there are two left- and one right-mover. Without scattering between left and right movers, these states should contribute 2e2/h2e^{2}/h to the conductance in one direction and e2/he^{2}/h in the other direction. On the other hand, scattering may reduce the conductance since a left and right mover can hybridize. In our case, the scattering is provided by the leads and therefore the resulting conductance is between 1 and 2 quanta of conductance.

The effect of tunnelling is perhaps most pronounced at the Weyl node (Fig. 6b). As discussed, the bulk gap closes and the subsequent absence of interface states at Δ=0\Delta=0 leads to zero conductance at zero energy. However, with tunnelling there are now two interface states in the spectrum: the left-moving chiral state and the right-moving emergent interface state. The former will terminate at an energy EtermE_{\mathrm{term}} (green line), the intersection of Ebulk=(t2sin2kx+hz2)12E_{\mathrm{bulk}}=\left(t^{2}\sin^{2}{k_{x}}+h_{z}^{2}\right)^{\frac{1}{2}} with Eq. (20). Below EtermE_{\mathrm{term}}, there are no uni-directional carriers and the conductance is unchanged. For Eterm<E<0E_{\mathrm{term}}<E<0, only the chiral state is present, and there is a conductance e2/he^{2}/h. Note the deviation from e2/he^{2}/h due to the small amount of bulk states present near zero energy. Above this range, both the chiral and the emergent interface state are present and move in opposite directions – their sum is null (modulo scattering), and only bulk states contribute.

Refer to caption
Figure 6: Conductance 𝒢\mathcal{G}_{\parallel} of the WSM-metal system along the interface at (a) kz=0k_{z}=0 and (b) kz=π/2k_{z}=\pi/2 for Δ=0\Delta=0 (i, ii) and Δ=2.3\Delta=2.3 (iii, iv). To guide the physical intuition, the spectra are shown in the left panels (i, iii) and states are colored and shaded according to their yy-position, with the relevant interface states in dark magenta and bulk states in faint colours. Energies relevant to the discussion in Sec. VI.1 are denoted by full horizontal lines.

When experimentally measuring transport between leads, the measured quantity is a sum over all kzk_{z} momenta. We therefore define the total conductance along the interface,

𝒢(E)=1Lzkz𝒢(E,kz).\mathcal{G}_{\parallel}(E)=\frac{1}{L_{z}}\sum_{k_{z}}\mathcal{G}_{\parallel}(E,k_{z}). (37)

Summing over quantized conductance contributions on the zz-projected Fermi arc karczk_{\mathrm{arc}}^{z}, Eq. (37)’s minimum is fixed (Fig. 7):

min𝒢=e2hkarcz2π.\min{\mathcal{G}_{\parallel}}=\frac{e^{2}}{h}\frac{k_{\mathrm{arc}}^{z}}{2\pi}. (38)

To probe this signature, we vary the arc length along the kzk_{z}-direction, as shown by Fig.7b. In the minimal model, this can be done by applying a strong Zeeman-like magnetic field bz𝐞zb_{z}\mathbf{e}_{z} coupling to spin degrees of freedom, bringing the arc length to karcz2arccos((γ+bz))k_{\mathrm{arc}}^{z}\to 2\arccos{(\gamma+b_{z})} provided bzb_{z} is small enough not to change the overall topological phase and that its orbital effects may be neglected.

Refer to caption
Figure 7: (a) Conductance summed over all kzk_{z} for Δ=0\Delta=0 (i) and Δ=2.3\Delta=2.3 (ii). (b) Total conductance minimum (38) as a function of bare Fermi arc length. For Δ=0\Delta=0 (black crosses), karcz=2arccos(γ)k^{z}_{\mathrm{arc}}=2\arccos{\gamma} and the minimum conductance scales with the Fermi arc length (modulo scattering). For Δ=2.3\Delta=2.3 (red triangles), karcz>2arccos(γ)k_{\mathrm{arc}}^{z}>2\arccos{\gamma} and the conductance minimum is therefore increased relative to Δ=0\Delta=0.

VI.2 Across the interface

We complete our study with a simple analytical model for electron tunnelling across the interface. We set out to derive an expression for the conductance across the interface 𝒢,σ=dIσ/dV\mathcal{G}_{\perp,\sigma}=dI_{\sigma}/dV of a particle polarized with spin σ\sigma. A detailed derivation is included in App. D.

We begin by expressing the current IσI_{\sigma} of a particle with spin σ\sigma in terms of the retarded correlation function URσσU_{R}^{\sigma\sigma^{\prime}} [28, 29, 30]:

Iσ=2eImσURσσ(eV).I_{\sigma}=-{2e}\,\mathrm{Im}\sum_{\sigma^{\prime}}U_{R}^{\sigma\sigma^{\prime}}(-eV). (39)

URσσ(eV){U}^{\sigma\sigma^{\prime}}_{R}\left(-eV\right) is found by computing the Matsubara correlation function 𝒰σσ(iωn){\mathcal{U}}^{\sigma\sigma^{\prime}}(i\omega_{n}) and analytically continuing iωneV+i0+i\omega_{n}\to-eV+i0^{+}. At finite temperature β1\beta^{-1}, we have

𝒰σσ(iωn)=1β𝐤𝐪|T𝐤𝐪|2ipgwσσ(𝐤,ipiωn)gmσσ(𝐪,ip).\mathcal{U}^{\sigma\sigma^{\prime}}(i\omega_{n})=\frac{1}{\beta}\sum_{\mathbf{k}\mathbf{q}}|T_{\mathbf{k}\mathbf{q}}|^{2}\sum_{ip}g_{w}^{\sigma^{\prime}\sigma}(\mathbf{k},ip-i\omega_{n})g_{m}^{\sigma\sigma^{\prime}}(\mathbf{q},ip). (40)

where 𝐤\mathbf{k} (𝐪\mathbf{q}) is the momentum in the WSM (metal), T𝐤𝐪T_{\mathbf{k}\mathbf{q}} is the tunnelling matrix element, ωn\omega_{n} (pp) is a bosonic (fermionic) Matsubara frequency, and 𝒈w\bm{g}_{w} (𝒈m\bm{g}_{m}) is the Matsubara Green’s function for the bare WSM (metal). Since states bound to the interface will not contribute to tunnelling across of it, we may consider only bulk states. The bulk Green’s functions 𝒈m,w\bm{g}_{m,w} are therefore

𝒈m(𝐪,ip)\displaystyle\bm{g}_{m}(\mathbf{q},ip) =1ipξm,\displaystyle=\frac{1}{ip-\xi_{m}}, (41a)
𝒈w(𝐤,ip)\displaystyle\bm{g}_{w}(\mathbf{k},ip) =ip+wbulk(ipξw)(ip+ξw),\displaystyle=\frac{ip+\mathcal{H}^{\mathrm{bulk}}_{w}}{(ip-\xi_{w})(ip+\xi_{w})}, (41b)

with the WSM (metal) dispersion ξw\xi_{w} (ξm\xi_{m}). Setting |T𝐤𝐪|2=Δ2δ(𝐤𝐪)|T_{\mathbf{k}\mathbf{q}}|^{2}=\Delta^{2}\delta(\mathbf{k}_{\perp}-\mathbf{q}_{\perp}), we perform the Matsubara frequency summation ip(ipξ)1=βnF(ξ)\sum_{ip}\left({ip-\xi}\right)^{-1}=\beta n_{F}\left(\xi\right) [29], where nFn_{F} is the fermionic distribution, by splitting the denominator into partial fractions. Using Im(eV+i0+ξ)1=δ(eVξ)\mathrm{Im}\,(-eV+i0^{+}-\xi)^{-1}=-\delta(-eV-\xi), Eq. (39) becomes

I=2eΔ2\displaystyle I={2e\Delta^{2}} 𝐤,ky,qy{u𝐤2[nF(ξm)nF(ξw)]δ(eVξ)\displaystyle\sum_{\mathbf{k}_{\perp},k_{y},q_{y}}\Big{\{}{u}_{\mathbf{k}}^{2}\left[n_{F}(\xi_{m})-n_{F}(\xi_{w})\right]\delta(-eV-\xi_{-})
+v𝐤2[nF(ξm)nF(ξw)]δ(eVξ+)},\displaystyle+{v}_{\mathbf{k}}^{2}\left[n_{F}(\xi_{m})-n_{F}(-\xi_{w})\right]\delta(-eV-\xi_{+})\Big{\}}, (42)

where ξ±=ξm±ξw\xi_{\pm}=\xi_{m}\pm\xi_{w} and

u𝐤2\displaystyle{u}^{2}_{\mathbf{k}} =12(1+tsin(kx)/ξw),\displaystyle=\frac{1}{2}\left(1+t\sin{k_{x}}/{\xi_{w}}\right), (43a)
v𝐤2\displaystyle{v}^{2}_{\mathbf{k}} =12(1tsin(kx)/ξw).\displaystyle=\frac{1}{2}\left(1-t\sin{k_{x}}/{\xi_{w}}\right). (43b)

Note that we have chosen the quantization axis in the xx-direction for simplicity. More generally, the second term in Eqs. (43) is an odd function of kxk_{x}, kzk_{z}, and ξw\xi_{w} and will vanish when integrated over, leaving the current spin-independent.

To proceed, we imagine placing the metal band’s Fermi level μm\mu_{m} in the WSM’s upper band and largely above to parabolic band minimum. At low energies,

ξw=v(𝐤2+ky2)12\xi_{w}=v\left(\mathbf{k}_{\perp}^{2}+k_{y}^{2}\right)^{\frac{1}{2}} (44a)
and
ξm=μm+1m(2mμ~𝐤2)12qy,\xi_{m}={\mu}_{m}+\frac{1}{m}\left(2m\tilde{\mu}-\mathbf{k}_{\perp}^{2}\right)^{\frac{1}{2}}q_{y}, (44b)

where m=1/2tmm=1/2t_{m} and μ~=μ+μm+6tm\tilde{\mu}=\mu+\mu_{m}+6t_{m} (the lattice constant is still a=1a=1). The latter expression is found by expanding near ξm\xi_{m}’s intercept with μm\mu_{m} along qyq_{y}, the metal’s yy-momentum. We further consider a small positive applied voltage such that particles tunnel from the upper WSM band to the metal. Thus, only the first term of Eq. (VI.2) contributes. Replacing the sums by integrals, changing variables from kyk_{y} to ξw\xi_{w} and qyq_{y} to ξm\xi_{m}, and re-inserting \hbar, the current is now

I\displaystyle I =ehmΔ22πv20eV𝑑ξwd2𝐤(2π)2ξwu𝐤2\displaystyle=\frac{e}{h}\frac{m\Delta^{2}}{2\pi v^{2}}\int_{0}^{eV}d\xi_{w}\int\frac{d^{2}\mathbf{k}_{\perp}}{(2\pi)^{2}}\xi_{w}u^{2}_{\mathbf{k}}
×θ(ξwv|𝐤|)θ(2mμ~𝐤2)ξw2/v2𝐤22mμ~𝐤2.\displaystyle\times\frac{\theta(\xi_{w}-v|\mathbf{k}_{\perp}|)\theta(2m\tilde{\mu}-\mathbf{k}_{\perp}^{2})}{\sqrt{\xi_{w}^{2}/v^{2}-\mathbf{k}_{\perp}^{2}}\sqrt{2m\tilde{\mu}-\mathbf{k}_{\perp}^{2}}}. (45)

Note that we have applied the low-temperature limit nF(ξw)=θ(ξw)n_{F}(\xi_{w})=-\theta(\xi_{w}). The integral over d2𝐤d^{2}\mathbf{k}_{\perp} can be done analytically, yielding the conductance across the interface:

𝒢(eV)=e2hmΔ2(2π)2v2eVlog|ε+eVεeV|.\mathcal{G}_{\perp}\left(eV\right)=\frac{e^{2}}{h}\frac{m\Delta^{2}}{\left(2\pi\right)^{2}v^{2}}eV\log\left|\frac{\varepsilon+eV}{\varepsilon-eV}\right|. (46)

For eV2mv2μ~εeV\ll\sqrt{2mv^{2}\tilde{\mu}}\equiv\varepsilon, the leading order term is quadratic in VV:

𝒢(eV)e2h2mΔ2(2π)2v2ε(eV)2.\mathcal{G}_{\perp}\left(eV\right)\approx\frac{e^{2}}{h}\frac{2m\Delta^{2}}{\left(2\pi\right)^{2}v^{2}\varepsilon}\left(eV\right)^{2}. (47)

Eq. (47) maintains that tunnelling measurements with featureless metals reveal the density of states at the tunnelling energy, since the three-dimensional WSM’s linear dispersion corresponds to a density of states proportional to E2E^{2}.

VII Conclusion and discussion

Using both lattice and continuum frameworks, we have described the behaviour of a 𝒯\mathcal{T}-broken WSM’s interface in proximity to a non-magnetic band. When coupled to this band via non-magnetic surface tunnelling, the WSM’s chiral state lowers in energy and forms, together with a previously delocalized bulk state, a noticeable spin-dependent asymmetry in the interface spectrum across the Weyl nodes. To model this phenomenon, we derived a infinite lattice theory of the interface and compared it to finite lattice model numerical results. We found that the infinite lattice theory accurately described the behaviour of the chiral state in the entire Brillouin zone (BZ), from its energy asymmetry to its spin canting at the interface. The localization of bulk states and the curving of the Fermi arc was also captured by the infinite lattice theory. To build intuition, we also derived a simpler continuum theory of interface states which captured the physics near 𝐤,0\mathbf{k}_{\perp,0}. Using the Landauer-Büttiker formalism, we calculated the transport of Weyl electrons travelling along the interface. Due to the asymmetry and increased Fermi arc length which allows for the presence of interface states beyond 𝐤,w±\mathbf{k}^{\pm}_{\perp,w}, we found a quantized increase in conductance per kzk_{z} at the Weyl nodes due to tunnelling. We proposed a possible probe of this increase by relating the minimum in total conductance to the Fermi arc length. Finally, across the interface, the conductance reproduces a simple electron tunnelling experiment, revealing the WSM’s density of states.

The results obtained herein may also be understood in the context of a pseudo-magnetic theory whereby the Weyl node separation plays the role of a magnetic gauge field [31, 32]. Alternatively, one may also view tunnelling as a finite potential well. As tunnelling broadens to link more sites on either side of the interface and broadens the well, the number of bound states increases. Indeed, Cheng et al. consider a Dirac cone (intuitively thought of as two Weyl cones of opposite chirality) under a confining potential well and find qualitatively similar interface spectra shown herein, albeit with symmetric kxk_{x}-spectra [33].

Though this toy model described the minimal case of two Weyl nodes in a magnetic WSM, these nodes always come in pairs connected by Fermi arcs. It is therefore reasonable to expect that the results obtained herein will still manifest themselves in more complicated systems with, e.g., broken inversion symmetry and a greater number of Fermi arcs. Finally, the asymmetry is resolved if one also accounts for the Hamiltonian’s 𝒯\mathcal{T}-reversed partner σyHw(𝐤)σy\sigma_{y}H_{w}^{*}\left(-\mathbf{k}\right)\sigma_{y}, instead breaking inversion symmetry.

Acknowledgements.
We would like to thank C.-T. Chen, A. Grushin, and B. Levitan for helpful discussions. LG acknowledges the hospitality of the Houches School of Physics and financial support from the NSERC CGS-M scholarship. TPB acknowledges funding from NSERC and FRQNT.

Appendix A The full Hamiltonian

Recall the Hamiltonian for the full (finite-sized) system:

H=𝐤y,y=Ly+1Ly𝐟𝐤,y(𝐤)y,y𝐟𝐤,y,\displaystyle{H}=\sum_{\mathbf{k_{\perp}}}\sum_{y,y^{\prime}=-L_{y}+1}^{L_{y}}\mathbf{f}_{\mathbf{k}_{\perp},y}^{\dagger}\mathcal{H}\left(\mathbf{k}_{\perp}\right)_{y,y^{\prime}}\mathbf{f}_{\mathbf{k}_{\perp},y^{\prime}}, (48a)
(𝐤)=(wopen(𝐤)TTmopen(𝐤)),\displaystyle\mathcal{H}\left(\mathbf{k}_{\perp}\right)=\begin{pmatrix}\mathcal{H}^{\mathrm{open}}_{w}\left(\mathbf{k}_{\perp}\right)&T^{\dagger}\\ T&\mathcal{H}^{\mathrm{open}}_{m}\left(\mathbf{k}_{\perp}\right)\end{pmatrix}, (48b)

where

𝐟𝐤,y={𝐜𝐤,yLy+1y0𝐝𝐤,y1yLy\mathbf{f}_{\mathbf{k}_{\perp},y}=\begin{cases}\mathbf{c}_{\mathbf{k}_{\perp},y}&-L_{y}+1\leq y\leq 0\\ \mathbf{d}_{\mathbf{k}_{\perp},y}&1\leq y\leq L_{y}\end{cases} (49)

with 𝐜y=(c𝐤,y,,c𝐤,y,)\mathbf{c}_{y}=\left(c_{\mathbf{k}_{\perp},y,\uparrow},c_{\mathbf{k}_{\perp},y,\downarrow}\right)^{\top}. open\mathcal{H}^{\mathrm{open}} is the partial-in-yy Fourier transform of bulk\mathcal{H}^{\mathrm{bulk}}. There is translational invariance in both xx and zz, so each block is in general a function of 𝐤\mathbf{k}_{\perp}.

The bare Weyl Hamiltonian wopen\mathcal{H}^{\mathrm{open}}_{w} is

wopen=𝐂0𝐡w+𝐂1𝐑w+𝐂1𝐑w\mathcal{H}^{\mathrm{open}}_{w}=\mathbf{C}_{0}\otimes\mathbf{h}_{w}+\mathbf{C}_{1}\otimes\mathbf{R}_{w}+\mathbf{C}^{\dagger}_{1}\otimes\mathbf{R}^{\dagger}_{w} (50)

where 𝐂0\mathbf{C}_{0} is the LyL_{y}-sized identity and (𝐂1)y,y=δy+1,y\left(\mathbf{C}_{1}\right)_{y,y^{\prime}}=\delta_{y+1,y^{\prime}} is the displacement operator on the lattice. Here,

𝐡w=txsin(kx)σx+tz(2+γcos(kx)cos(kz))σz,\displaystyle\mathbf{h}_{w}=t_{x}\sin{k_{x}}\sigma_{x}+t_{z}\left(2+\gamma-\cos{k_{x}}-\cos{k_{z}}\right)\sigma_{z}, (51a)
𝐑w=ity2σytz2σz\displaystyle\mathbf{R}_{w}=\frac{it_{y}}{2}\sigma_{y}-\frac{t_{z}}{2}\sigma_{z} (51b)
are the spin-matrices corresponding to same-site and nearest-neighbour hopping, respectively.

The bare metal Hamiltonian mopen\mathcal{H}^{\mathrm{open}}_{m} is written in similar form (σ0\sigma_{0} is the identity matrix in spin):

mopen=hm𝐂0σ0tm𝐂1σ0tm𝐂1σ0\mathcal{H}^{\mathrm{open}}_{m}=h_{m}\mathbf{C}_{0}\otimes\sigma_{0}-t_{m}\mathbf{C}_{1}\otimes\sigma_{0}-t_{m}\mathbf{C}^{\dagger}_{1}\otimes\sigma_{0} (52)

with

hm=2tm(cos(kx)+cos(kz))μ.h_{m}=-2t_{m}\left(\cos{k_{x}}+\cos{k_{z}}\right)-\mu. (53)

Finally, the tunnelling term admits the simple form (T)y,y=Δδ0,Ly1\left(T\right)_{y,y^{\prime}}=\Delta\delta_{0,L_{y}-1}, or

T=(0Δ00),T=\begin{pmatrix}0&\dots&\Delta\\ \vdots&\ddots&\vdots\\ 0&\dots&0\end{pmatrix}, (54)

once again diagonal in spin.

Appendix B The interface spectrum for different band configurations

The asymmetry at the Weyl node illustrated in Fig. 3b is also apparent for different choices of μ\mu, tmt_{m}, and metal bandstructure. To convince ourselves of our specific model’s ubiquitous features, we display a few more metal configurations in Fig. 8a. We expect that in the more realistic setup of a WSM coupled to a two-band bulk insulator, both copies will be present: one for positive and one for negative energies. This is confirmed in what follows.

Refer to caption
Figure 8: The energies at the Weyl node for various parabolic band configurations at Δ=2.3\Delta=2.3. (a) The interface spectral function for tm=0.5t_{m}=0.5 at (from left to right, top to bottom): μ=4\mu=-4, μ=2\mu=-2, μ=0\mu=0, and μ=+2\mu=+2 as defined by Eq. (4). The WSM (metal) bulk edge is shown in dashed white (blue) lines. (b) The bandstructure for a WSM in contact with the two-band insulator [Eq. (55)]. States localized to the interface are shown in red. The insulator band gap is.

One may also imagine coupling the WSM to a two-band bulk insulator, i.e. two copies of the single bulk metal band separated by a gap. Keeping each individual band non-magnetic, the Hamiltonian is now

H=(wopenTTT+open0T0open)H=\begin{pmatrix}\mathcal{H}^{\mathrm{open}}_{w}&T^{\dagger}&T^{\dagger}\\ T&\mathcal{H}^{\mathrm{open}}_{+}&0\\ T&0&\mathcal{H}^{\mathrm{open}}_{-}\end{pmatrix} (55)

where ±open\mathcal{H}^{\mathrm{open}}_{\pm} represent the metal Hamiltonian mopen\mathcal{H}^{\mathrm{open}}_{m} with parameters tm,±=±tmt_{m,\pm}=\pm t_{m}, μ±=±μ\mu_{\pm}=\pm\mu and TT is the same as before. The resulting spectrum is shown in Fig. 8b. Unsurprisingly, we recover two copies of the previously observed, single-band asymmetry: one for positive energies and one for negative energies. The asymmetry is therefore resolved if one inverts both the momentum and energy.

Though the case of a band with zero bandwidth tm=0t_{m}=0 is not physically realistic, it still reproduces the same qualitative asymmetry. For mathematical simplification, therefore, we may set tm=0t_{m}=0 as is done in the continuum theory and spin canting discussion. We ultimately choose to work with μ=4\mu=-4 and tm=0.5t_{m}=0.5 and a single metal band due to the clear asymmetry across the Weyl node and separation between bulk WSM and metal dispersions.

Appendix C σy=0\expectationvalue{\sigma_{y}}=0 in an open WSM

In the bulk, it is clear that σy\expectationvalue{\sigma_{y}} may be any value. In particular, the spin-orbit coupling in a Weyl Hamiltonian of the form 𝐤𝝈\mathbf{k}\cdot\bm{\sigma} will tie the yy-momentum kyk_{y} to the spin in that same direction. Upon opening our system in the yy-direction, however, σy=0\expectationvalue{\sigma_{y}}=0 identically throughout. Similarly, opening the system in xx renders σx=0\expectationvalue{\sigma_{x}}=0.

One can see why this is the case by examining the finite-sized WSM Hamiltonian Eq. (50). Written in matrix form, the blocks are

𝐡w\displaystyle\mathbf{h}_{w} =(g3g1g1g3),\displaystyle=\begin{pmatrix}g_{3}&g_{1}\\ g_{1}&-g_{3}\end{pmatrix}, (56a)
𝐑w\displaystyle\mathbf{R}^{\dagger}_{w} =12(tztytytz)\displaystyle=\frac{1}{2}\begin{pmatrix}-t_{z}&t_{y}\\ -t_{y}&t_{z}\end{pmatrix} (56b)

in spin space, where tx,y,zt_{x,y,z} are real. Adding these blocks into the finite-sized matrix HwH_{w} leads to a real and Hermitian (or, symmetric) matrix, i.e. Hw=HwH_{w}^{\top}=H_{w}. We set out to prove that one can always find real eigenstates to a real symmetric matrix, thereby rendering σy=0\expectationvalue{\sigma_{y}}=0 identically as σy\sigma_{y} is purely imaginary.

To prove this, we start by noting that the eigenvalues of a symmetric (or, more generally, Hermitian) matrix are real. Take Hw|ψ=E|ψH_{w}\ket{\psi}=E\ket{\psi}. Adding it to its complex conjugate yields

Hw(|ψ+|ψ)=E(|ψ+|ψ).H_{w}\left(\ket{\psi}+\ket{\psi}^{*}\right)=E\left(\ket{\psi}+\ket{\psi}^{*}\right). (57)

Now, if |ψ=|ψ\ket{\psi}=-\ket{\psi}^{*} then |ψ\ket{\psi} is purely imaginary and we can therefore define |ψ=i|φ\ket{\psi}=i\ket{\varphi} with |φ\ket{\varphi} purely real, satisfying Hw|φ=E|φH_{w}\ket{\varphi}=E\ket{\varphi}. Otherwise, if |ψ+|ψ0\ket{\psi}+\ket{\psi}^{*}\neq 0 then it is necessarily real. Therefore, one may always find a complete set of real eigenvectors to a real symmetric matrix. Since the eigenvectors are purely real and the matrix σy\sigma_{y} contains only imaginary entries,

ψ|σy|ψ=0.\bra{\psi}\sigma_{y}\ket{\psi}=0. (58)

Of course, one can always perform a unitary rotation in spin space such that σxσy\sigma_{x}\rightarrow\sigma_{y} and σyσx\sigma_{y}\rightarrow-\sigma_{x}. In this case, an open system in xx exactly mirrors one open in yy before the rotation, and σx=0\expectationvalue{\sigma_{x}}=0 likewise follows.

Appendix D Derivation of conductance across the interface

The following procedure is similar to that performed by Cohen et al. [28, 29, 30] with one important caveat: the Hamiltonian’s off-diagonal elements may correlate different spins due to spin-orbit coupling.

To begin, we posit that the current from left to right will be in direct proportion to the number of electrons leaving the Weyl side and subsequently entering the metal side. To serve our general discussion, we write the tunnelling term as

HT=𝐤𝐪σT𝐤𝐪c𝐤σd𝐪σ+h.c.H_{T}=\sum_{\mathbf{k}\mathbf{q}\sigma}T_{\mathbf{k}\mathbf{q}}c^{\dagger}_{\mathbf{k}\sigma}d_{\mathbf{q}\sigma}+\mathrm{h.c.} (59)

where 𝐤\mathbf{k} (𝐪\mathbf{q}) is the momentum on the WSM (metal) side. We further assume operators on opposite sides are independent, {ci,dj}=0\{c_{i},d_{j}^{\dagger}\}=0. Next, define

Nσ=𝐤c𝐤σc𝐤σN_{\sigma}=\sum_{\mathbf{k}}c_{\mathbf{k}\sigma}^{\dagger}c_{\mathbf{k}\sigma} (60)

as the number of electrons with spin σ\sigma on the Weyl side. Since HwH_{w} is number conserving, only HTH_{T} fails to commute with NσN_{\sigma}:

iN˙σ=[Nσ,HT]=𝐤𝐪(T𝐤𝐪c𝐤σd𝐪σh.c.).i\hbar\dot{N}_{\sigma}=[N_{\sigma},H_{T}]=\sum_{\mathbf{k}\mathbf{q}}\left(T_{\mathbf{k}\mathbf{q}}c^{\dagger}_{\mathbf{k}\sigma}d_{\mathbf{q}\sigma}-\mathrm{h.c.}\right). (61)

Though we may impose our diffusive potential T𝐤𝐪=Δδ(𝐤𝐪)T_{\mathbf{k}\mathbf{q}}=\Delta\delta(\mathbf{k}_{\perp}-\mathbf{q}_{\perp}) to pick out 𝐤=𝐪\mathbf{k}_{\perp}=\mathbf{q}_{\perp}, we proceed with a general tunnelling term and substitute it at the end.

In linear response, the total current across the interface is

Iσ(t)=eN˙σ(t)=iet𝑑t[N˙σ(t),HT(t)]0I_{\sigma}(t)=-e\expectationvalue{\dot{N}_{\sigma}(t)}=ie\int_{-\infty}^{t}dt^{\prime}\expectationvalue{[\dot{N}_{\sigma}(t),H_{T}(t^{\prime})]}_{0} (62)

where 0\expectationvalue{\cdot}_{0} represents an average over uncoupled states. To reduce the clutter of notation, define

Cσ(t)=𝐤𝐪T𝐤𝐪c𝐤σ(t)d𝐪σ(t).C_{\sigma}(t)=\sum_{\mathbf{k}\mathbf{q}}T_{\mathbf{k}\mathbf{q}}c^{\dagger}_{\mathbf{k}\sigma}(t)d_{\mathbf{q}\sigma}(t). (63)

Making the bias μmμw=eV\mu_{m}-\mu_{w}=eV explicit, it can be written as a collection of four terms:

i[N˙σ(t),HT(t)]=\displaystyle i[\dot{N}_{\sigma}(t),H_{T}(t^{\prime})]= σ{eieV(tt)[Cσ(t),Cσ(t)]\displaystyle\sum_{\sigma^{\prime}}\Big{\{}e^{ieV(t-t^{\prime})}[C_{\sigma}(t),C^{\dagger}_{\sigma^{\prime}}(t^{\prime})]
eieV(tt)[Cσ(t),Cσ(t)]\displaystyle-e^{-ieV(t-t^{\prime})}[C^{\dagger}_{\sigma}(t),C_{\sigma^{\prime}}(t^{\prime})]
+eieV(t+t)[Cσ(t),Cσ(t)]\displaystyle+e^{ieV(t+t^{\prime})}[C_{\sigma}(t),C_{\sigma^{\prime}}(t^{\prime})]
eieV(t+t)[Cσ(t),Cσ(t)]},\displaystyle-e^{-ieV(t+t^{\prime})}[C^{\dagger}_{\sigma}(t),C^{\dagger}_{\sigma^{\prime}}(t^{\prime})]\Big{\}}, (64)

Only the first two terms are nonzero. Indeed, they represent quantum dot tunnelling, while the last two correspond to a Josephson type current, which is identically zero for the uncoupled states in the system we are considering. The first term has exactly the right form to be a retarded correlation function,

URσσ(ω)=𝑑teiωtθ(t)[Cσ(t),Cσ]0\displaystyle U_{R}^{\sigma\sigma^{\prime}}(\omega)=\int_{-\infty}^{\infty}dte^{-i\omega t}\theta(t)\expectationvalue{[C_{\sigma}(t),C^{\dagger}_{\sigma^{\prime}}]}_{0} (65)

where we have used the fact that 𝒰Rσσ(t)\mathcal{U}_{R}^{\sigma\sigma^{\prime}}(t) depends only on the time difference ttt-t^{\prime}. These are precisely the kinds of terms which appear in the current calculation. We may recast IσI_{\sigma} in terms of retarded correlation functions (re-inserting the factor of \hbar):

Iσ=2eσImURσσ(eV).I_{\sigma}=-\frac{2e}{\hbar}\sum_{\sigma^{\prime}}\mathrm{Im}U_{R}^{\sigma\sigma^{\prime}}(-eV). (66)

The Matsubara formalism provides an elegant way of finding URσσ(eV)U_{R}^{\sigma\sigma^{\prime}}(-eV) by first computing 𝒰σσ(iωn)\mathcal{U}^{\sigma\sigma^{\prime}}(i\omega_{n}) and subsequently letting iωneV+i0+i\omega_{n}\to-eV+i0^{+}. In this framework,

𝒰σσ(iωn)\displaystyle\mathcal{U}^{\sigma\sigma^{\prime}}(i\omega_{n}) =0β𝑑τeiωnτTτCσ(τ)Cσ0\displaystyle=-\int_{0}^{\beta}d\tau e^{i\omega_{n}\tau}\expectationvalue{T_{\tau}C_{\sigma}(\tau)C^{\dagger}_{\sigma^{\prime}}}_{0}
=𝐤𝐪T𝐤𝐪0β𝑑τeiωnτgwσσ(𝐤,τ)gmσσ(𝐪,τ),\displaystyle=\sum_{\mathbf{k}\mathbf{q}}T_{\mathbf{k}\mathbf{q}}\int_{0}^{\beta}d\tau e^{i\omega_{n}\tau}{g}_{w}^{\sigma^{\prime}\sigma}(\mathbf{k},-\tau){g}_{m}^{\sigma\sigma^{\prime}}(\mathbf{q},\tau), (67)

where τ=it\tau=it and ωn=2πn/β\omega_{n}=2\pi n/\beta is a bosonic Matsubara frequency because we are working with pairs of fermionic operators. TτT_{\tau} is simply an instruction to order the operators in increasing τ\tau starting from the right which, when combined with Wick’s theorem, gives rise to the uncoupled Green’s functions 𝒈w(𝐤,τ)\bm{g}_{w}(\mathbf{k},\tau) and 𝒈w(𝐪,τ)\bm{g}_{w}(\mathbf{q},\tau). Fourier transforming with g(τ)=β1ipeipτg(ip){g}(\tau)=\beta^{-1}\sum_{ip}e^{-ip\tau}g(ip) one has

𝒰σσ(iωn)=1β𝐤𝐪|T𝐤𝐪|2ipgwσσ(𝐤,ipiωn)gmσσ(𝐪,ip).\mathcal{U}^{\sigma\sigma^{\prime}}(i\omega_{n})=\frac{1}{\beta}\sum_{\mathbf{k}\mathbf{q}}|T_{\mathbf{k}\mathbf{q}}|^{2}\sum_{ip}g_{w}^{\sigma^{\prime}\sigma}(\mathbf{k},ip-i\omega_{n})g_{m}^{\sigma\sigma^{\prime}}(\mathbf{q},ip). (68)

To arrive at an analytic result, we take 𝒈\bm{g} to be the bulk Green’s function since states bound to the interface will not contribute to tunnelling across it, so we may consider only bulk states. Solving the equations of motion, we find

𝒈m(𝐪,ip)\displaystyle\bm{g}_{m}(\mathbf{q},ip) =1ipξm,\displaystyle=\frac{1}{ip-\xi_{m}}, (69a)
𝒈w(𝐤,ip)\displaystyle\bm{g}_{w}(\mathbf{k},ip) =ip+wbulk(ipξw)(ip+ξw).\displaystyle=\frac{ip+\mathcal{H}^{\mathrm{bulk}}_{w}}{(ip-\xi_{w})(ip+\xi_{w})}. (69b)

Note that the bare WSM (metal) dispersion ξw\xi_{w} (ξm\xi_{m}) depends on the momenta 𝐤\mathbf{k} (𝐪\mathbf{q}). With

|T𝐤𝐪|2=Δ2δ(𝐤𝐪),|T_{\mathbf{k}\mathbf{q}}|^{2}=\Delta^{2}\delta(\mathbf{k}_{\perp}-\mathbf{q}_{\perp}), (70)

the correlation function 𝒰σσ(iωn)\mathcal{U}^{\sigma\sigma^{{}^{\prime}}}(i\omega_{n}) can be written as a 2×22\times 2 matrix:

𝓤(iωn)=Δ2β𝐤,ky,qyipipiωn+wbulk(ipξm)(ipiωnξw)(ipiωn+ξw),\displaystyle\bm{\mathcal{U}}(i\omega_{n})=\frac{\Delta^{2}}{\beta}\sum_{\mathbf{k}_{\perp},k_{y},q_{y}}\sum_{ip}\frac{ip-i\omega_{n}+\mathcal{H}^{\mathrm{bulk}}_{w}}{(ip-\xi_{m})(ip-i\omega_{n}-\xi_{w})(ip-i\omega_{n}+\xi_{w})}, (71)

taking wopen\mathcal{H}^{\mathrm{open}}_{w} as Eq. (2). To perform the Matsubara frequency summation, we split the denominator via partial fractions and use the relation

1βip1ipξ=nF(ξ)\frac{1}{\beta}\sum_{ip}\frac{1}{ip-\xi}=n_{F}\left(\xi\right) (72)

for a fermionic frequency ip=iπ(2n+1)/βip=i\pi\left(2n+1\right)/\beta. Note that shifting ipip by a bosonic iωni\omega_{n} frequency will not change the summation. Now, upon analytic continuation, Im(ωξ+i0+)1=δ(ωξ)\mathrm{Im}\,(\omega-\xi+i0^{+})^{-1}=-\delta(\omega-\xi) relates each term to a statement of conservation of energy. We may therefore rewrite Eq. (71) as a product of occupation numbers and Dirac deltas, each weighted by spin-dependent factors. For an electron with spin σ\sigma, the current is therefore

Iσ=2eΔ2𝐤qy{u𝐤,σ2[nF(ξm)nF(ξw)]δ(eVξ)+v𝐤,σ2[nF(ξm)nF(ξw)]δ(eVξ+)},\displaystyle I_{\sigma}=\frac{2e\Delta^{2}}{\hbar}\sum_{\mathbf{k}q_{y}}\Big{\{}{u}_{\mathbf{k},\sigma}^{2}\left[n_{F}(\xi_{m})-n_{F}(\xi_{w})\right]\delta(-eV-\xi_{-})+{v}_{\mathbf{k},\sigma}^{2}\left[n_{F}(\xi_{m})-n_{F}(-\xi_{w})\right]\delta(-eV-\xi_{+})\Big{\}}, (73)

which closely resembles the superconducting case up to the modified spin-dependent coherence factors

u𝐤,σ2\displaystyle{u}^{2}_{\mathbf{k},\sigma} =12(1+σ|wbulk|σ/ξw),\displaystyle=\frac{1}{2}\left(1+{\bra{\sigma}\mathcal{H}^{\mathrm{bulk}}_{w}\ket{\sigma}}/{\xi_{w}}\right), (74a)
v𝐤,σ2\displaystyle{v}^{2}_{\mathbf{k},\sigma} =12(1σ|wbulk|σ/ξw).\displaystyle=\frac{1}{2}\left(1-{\bra{\sigma}\mathcal{H}^{\mathrm{bulk}}_{w}\ket{\sigma}}/{\xi_{w}}\right). (74b)

Here, kyk_{y} (qyq_{y}) is momentum along yy in the WSM (metal), ξw(kx,ky,kz)\xi_{w}(k_{x},k_{y},k_{z}) [ξm(kx,qy,kz)\xi_{m}(k_{x},q_{y},k_{z})] is the bare WSM (metal) dispersion, ξ±=ξm±ξw\xi_{\pm}=\xi_{m}\pm\xi_{w}, and nFn_{F} is the the usual fermionic occupation number. Eq. (73) has two terms: the first (second) corresponds to tunnelling from the upper (lower) WSM band to the metal band. Each term has three parts: the Dirac delta imposes energy conservation, nF(ξm)nF(ξw)n_{F}(\xi_{m})-n_{F}(\xi_{w}) counts the participating states available to tunnel, and u𝐤,σ2u^{2}_{\mathbf{k},\sigma}/v𝐤,σ2v^{2}_{\mathbf{k},\sigma} weigh the band according to the spin. All of this is proportional to Δ2\Delta^{2}, the amplitude of a tunnelling interaction in light of the exact action Eq. (8).

References

  • Armitage et al. [2018] N. P. Armitage, E. J. Mele, and A. Vishwanath, Weyl and dirac semimetals in three-dimensional solids, Rev. Mod. Phys. 90, 015001 (2018).
  • Shen [2012] S.-Q. Shen, Topological Insulators: Dirac Equation In Condensed Matters (Springer, 2012).
  • Bernevig [2015] B. Bernevig, It’s been a weyl coming., Nature 11, 698 (2015).
  • Xu et al. [2015a] S.-Y. Xu, I. Belopolski, N. Alidoust, M. Neupane, G. Bian, C. Zhang, R. Sankar, G. Chang, Z. Yuan, C.-C. Lee, S.-M. Huang, H. Zheng, J. Ma, D. S. Sanchez, B. Wang, A. Bansil, F. Chou, P. P. Shibayev, H. Lin, S. Jia, and M. Z. Hasan, Discovery of a weyl fermion semimetal and topological fermi arcs, Science 349, 613 (2015a)https://www.science.org/doi/pdf/10.1126/science.aaa9297 .
  • Xu et al. [2015b] S.-Y. Xu, N. Alidoust, I. Belopolski, Z. Yuan, G. Bian, T.-R. Chang, H. Zheng, V. N. Strocov, D. S. Sanchez, G. Chang, C. Zhang, D. Mou, Y. Wu, L. Huang, C.-C. Lee, S.-M. Huang, B. Wang, A. Bansil, H.-T. Jeng, T. Neupert, A. Kaminski, H. Lin, S. Jia, and M. Z. Hasan, Discovery of a weyl fermion state with fermi arcs in niobium arsenide, Nature 11, 748 (2015b).
  • Soluyanov et al. [2015] A. A. Soluyanov, D. Gresch, Z. Wang, Q. Wu, M. Troyer, X. Dai, and B. A. Bernevig, Type-ii weyl semimetals, Nature 527, 495 (2015).
  • Sun et al. [2015] Y. Sun, S.-C. Wu, M. N. Ali, C. Felser, and B. Yan, Prediction of weyl semimetal in orthorhombic mote2{\mathrm{mote}}_{2}Phys. Rev. B 92, 161107 (2015).
  • Yoshimura et al. [2016] Y. Yoshimura, W. Onishi, K. Kobayashi, T. Ohtsuki, and K.-I. Imura, Comparative study of weyl semimetal and topological/chern insulators: Thin-film point of view, Phys. Rev. B 94, 235414 (2016).
  • Meng and Balents [2012] T. Meng and L. Balents, Weyl superconductors, Phys. Rev. B 86, 054504 (2012).
  • Chen and Franz [2016] A. Chen and M. Franz, Superconducting proximity effect and majorana flat bands at the surface of a weyl semimetal, Phys. Rev. B 93, 201105 (2016).
  • Burkov [2022] A. A. Burkov, Topological properties of dirac and weyl semimetals (2022).
  • Liu et al. [2018] E. Liu, Y. Sun, N. Kumar, L. Muechler, A. Sun, L. Jiao, S.-Y. Yang, D. Liu, A. Liang, Q. Xu, J. Kroder, V. Süß, H. Borrmann, C. Shekhar, Z. Wang, C. Xi, W. Wang, W. Schnelle, S. Wirth, Y. Chen, S. T. B. Goennenwein, and C. Felser, Giant anomalous hall effect in a ferromagnetic kagome-lattice semimetal, Nature Physics 14, 1125 (2018).
  • O’Brien et al. [2017] T. E. O’Brien, C. W. J. Beenakker, and i. d. I. Adagideli, Superconductivity provides access to the chiral magnetic effect of an unpaired weyl cone, Phys. Rev. Lett. 118, 207701 (2017).
  • Shekhar et al. [2015] C. Shekhar, A. K. Nayak, Y. Sun, M. Schmidt, M. Nicklas, I. Leermakers, U. Zeitler, Y. Skourski, J. Wosnitza, Z. Liu, Y. Chen, W. Schnelle, H. Borrmann, Y. Grin, C. Felser, and B. Yan, Extremely large magnetoresistance and ultrahigh mobility in the topological weyl semimetal candidate nbp, Nature Physics 11, 645 (2015).
  • Burkov et al. [2011] A. A. Burkov, M. D. Hook, and L. Balents, Topological nodal semimetals, Phys. Rev. B 84, 235126 (2011).
  • Burkov and Balents [2011] A. A. Burkov and L. Balents, Weyl semimetal in a topological insulator multilayer, Phys. Rev. Lett. 107, 127205 (2011).
  • Rüßmann et al. [2018] P. Rüßmann, A. P. Weber, F. Glott, N. Xu, M. Fanciulli, S. Muff, A. Magrez, P. Bugnon, H. Berger, M. Bode, J. H. Dil, S. Blügel, P. Mavropoulos, and P. Sessi, Universal scattering response across the type-ii weyl semimetal phase diagram, Phys. Rev. B 97, 075106 (2018).
  • Wu et al. [2017] Y. Wu, H. Liu, H. Jiang, and X. C. Xie, Global phase diagram of disordered type-ii weyl semimetals, Phys. Rev. B 96, 024201 (2017).
  • Silva et al. [2022] W. C. Silva, W. N. Mizobata, J. E. Sanches, L. S. Ricco, I. A. Shelykh, M. de Souza, M. S. Figueira, E. Vernek, and A. C. Seridonio, Topological charge fano effect in multi-weyl semimetals, Physical Review B 10510.1103/physrevb.105.235135 (2022).
  • Kundu and Jalil [2022] A. Kundu and M. B. A. Jalil, Ruderman-kittel-kasuya-yosida (rkky) interaction in weyl semimetals with tilted energy dispersion (2022).
  • Borchmann and Pereg-Barnea [2017] J. Borchmann and T. Pereg-Barnea, Quantum oscillations in weyl semimetals: A surface theory approach, Phys. Rev. B 96, 125153 (2017).
  • Marchand and Franz [2012] D. J. J. Marchand and M. Franz, Lattice model for the surface states of a topological insulator with applications to magnetic and exciton instabilities, Phys. Rev. B 86, 155146 (2012).
  • Gorbar et al. [2015] E. V. Gorbar, V. A. Miransky, I. A. Shovkovy, and P. O. Sukhachov, Surface fermi arcs in 𝕫2{\mathbb{z}}_{2} weyl semimetals A3Bi{A}_{3}\mathrm{Bi} (a=Naa=\mathrm{Na}, k, rb), Phys. Rev. B 91, 235138 (2015).
  • Potter et al. [2014] A. Potter, I. Kimchi, and A. Vishwanath, Quantum oscillations from surface fermi arcs in weyl and dirac semimetals., Nature Communications 510.1038/ncomms6161 (2014).
  • Pereg-Barnea and Lin [2005] T. Pereg-Barnea and H.-H. Lin, Andreev edge state on semi-infinite triangular lattice: Detecting the pairing symmetry in na0.35coo2·yh2o, Europhysics Letters 69, 791 (2005).
  • Datta [1995] S. Datta, Electronic Transport in Mesoscopic Systems, Cambridge Studies in Semiconductor Physics and Microelectronic Engineering (Cambridge University Press, 1995).
  • Bernevig et al. [2006] B. A. Bernevig, T. L. Hughes, and S.-C. Zhang, Quantum spin hall effect and topological phase transition in hgte quantum wells, Science 314, 1757 (2006)https://www.science.org/doi/pdf/10.1126/science.1133734 .
  • Cohen et al. [1962] M. H. Cohen, L. M. Falicov, and J. C. Phillips, Superconductive tunneling, Phys. Rev. Lett. 8, 316 (1962).
  • Mahan [2000] G. D. Mahan, Many-Particle Physics (Physics of Solids and Liquids), 3rd ed. (Springer, 2000).
  • Ryndyk [2016] D. Ryndyk, Theory of Quantum Transport at Nanoscale, Springer Series in Solid-State Sciences (Springer Cham, 2016).
  • Grushin et al. [2016] A. G. Grushin, J. W. F. Venderbos, A. Vishwanath, and R. Ilan, Inhomogeneous weyl and dirac semimetals: Transport in axial magnetic fields and fermi arc surface states from pseudo-landau levels, Phys. Rev. X 6, 041046 (2016).
  • Grushin [2012] A. G. Grushin, Consequences of a condensed matter realization of lorentz-violating qed in weyl semi-metals, Phys. Rev. D 86, 045001 (2012).
  • Cheng et al. [2019] A. Cheng, T. Taniguchi, K. Watanabe, P. Kim, and J.-D. Pillet, Guiding dirac fermions in graphene with a carbon nanotube, Phys. Rev. Lett. 123, 216804 (2019).