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

Exact solutions of Kondo problems in higher-order fermions

Peng Song Department of Physics, Nanjing University, Nanjing 210093, China National Laboratory of Solid State Microstructures and Collaborative Innovation Center of Advanced Microstructures, Nanjing University, Nanjing 210093, China    Haodong Ma Department of Physics, Nanjing University, Nanjing 210093, China National Laboratory of Solid State Microstructures and Collaborative Innovation Center of Advanced Microstructures, Nanjing University, Nanjing 210093, China    Rui Wang [email protected] Department of Physics, Nanjing University, Nanjing 210093, China National Laboratory of Solid State Microstructures and Collaborative Innovation Center of Advanced Microstructures, Nanjing University, Nanjing 210093, China    Baigeng Wang [email protected] Department of Physics, Nanjing University, Nanjing 210093, China National Laboratory of Solid State Microstructures and Collaborative Innovation Center of Advanced Microstructures, Nanjing University, Nanjing 210093, China
Abstract

The conformal field theory (CFT) approach to Kondo problems, originally developed by Affleck and Ludwig (AL), has greatly advanced the fundamental knowledge of Kondo physics. The CFT approach to Kondo impurities is based on a necessary approximation, i.e., the linearization of the low-lying excitations in a narrow energy window about the Fermi surface. This treatment works well in normal metal baths, but encounters fundamental difficulties in systems with Fermi points and high-order dispersion relations. Prominent examples of such systems are the recently-proposed topological semimetals with emergent higher-order fermions. Here, we develop a new CFT technique that yields exact solutions to the Kondo problems in higher-order fermion systems. Our approach does not require any linearization of the low-lying excitations, and more importantly, it rigorously bosonizes the entire energy spectrum of the higher-order fermions. Therefore, it provides a more solid theoretical base for evaluating the thermodynamic quantities at finite temperatures. Our work significantly broadens the scope of CFT techniques and brings about unprecedented applications beyond the reach of conventional methods.

I Introduction

The Kondo effect Kondo (1964), which treats a quantum magnetic impurity in normal metals, invoked fundamental developments in condensed matter physics. Among these progresses, the conformal field theory (CFT) approach developed by Affleck and Ludwig (AL) unveiled the physical nature of the Kondo problem, and demonstrated the elegant connections between conformal symmetry and the Kondo fixed point Affleck and Ludwig (1991a, b, c); Affleck et al. (1992); Affleck and Ludwig (1993); Ludwig (1994); Affleck (1995); Delft (1995). In their original approach, AL mapped the multi-channel Kondo problem to an exactly-solvable CFT in the complex plane. In order to establish the mapping, they linearized the low-lying excitations in a narrow energy window ΛvF<ϵ<+ΛvF-\Lambda v_{F}<\epsilon<+\Lambda v_{F} about the Fermi surface, with vFv_{F} being the Fermi velocity and Λ\Lambda being an artificially introduced cut-off, and ignored the higher-energy excitations, as a necessary approximation (Fig. 1(a)). Although this treatment works well for normal metals with well-defined Fermi surfaces, it will encounter difficulties in more exotic thermals baths, including the pseudogapped systems Yu et al. (2020a) and those with Fermi points, such as topological semimetals.

On the other hand, past years have witnessed increasing interests in the Kondo problem in topological materials with emergent particles, including graphene, Weyl and Dirac semimetals. In particular, most recently, even more exotic Weyl and Dirac semimetals which exhibit effective fermions with higher-order dispersion relations have been proposed. These higher-order dispersion relations take a form that is linear in one direction, but quadratic or cubic in the orthogonal plane Fang et al. (2012); Yang and Nagaosa (2014); Yang et al. (2015); Huang et al. (2016); Gao et al. (2016); Bradlyn et al. (2016); Liu and Zunger (2017); Ezawa (2017); Yu et al. (2018, 2019); Isobe and Fu (2019); Song et al. (2020); Wu et al. (2020); Wang et al. (2020); Wu et al. (2021). Such materials are named quadratic and cubic Weyl/Dirac semimetals respectively Yang and Nagaosa (2014); Liu and Zunger (2017); Song et al. (2020), with their corresponding emergent fermions referred to as quadratic and cubic Weyl/Dirac fermions respectively Fang et al. (2012); Yang and Nagaosa (2014); Gao et al. (2016), and they are classified as Weyl/Dirac according to their 2-fold/4-fold band degeneracies at their band crossings Yang and Nagaosa (2014); Yang et al. (2015); Gao et al. (2016). For example, in three dimensions, the cubic Weyl/Dirac semimetals display the dispersion relation, in its simplest form,

ϵμ(p)=a1μp3+a2μpx,\epsilon_{\mu}(\vec{p})=a_{1\mu}p^{3}+a_{2\mu}p_{x}\,, (1)

where μ\mu is the band index, a1μa_{1\mu} and a2μa_{2\mu} are dispersion coefficients, p|p|=(px2+py2+pz2)12p\equiv|\vec{p}|=(p_{x}^{2}+p_{y}^{2}+p_{z}^{2})^{\frac{1}{2}}. A number of materials for the quadratic and cubic Weyl/Dirac semimetals have been recently proposed and extensively studied. Examples of quadratic Weyl/Dirac semimetals include HgCr2Se4, SiSr2, band-inverted α\alpha-Sn and PdSb2 Fang et al. (2012); Huang et al. (2016); Bradlyn et al. (2016), and examples of cubic Weyl/Dirac semimetals include LiOS{}_{\text{S}}O3 and quasi-one-dimensional molybdenum monochalcogenide compounds AI{}^{\text{I}}(MoXVI{}^{\text{VI}})3, where AI{}^{\text{I}} = Na, K, Rb, In or Tl, XVI{}^{\text{VI}} = S, Se or Te Liu and Zunger (2017); Yu et al. (2018); Song et al. (2020); Wu et al. (2020). Moreover, various novel quantum phenomena are predicted to take place in these semimetals, such as charge density wave, non-Fermi liquid, and topological superconductivity Yang and Nagaosa (2014); Ezawa (2017); Yu et al. (2018); Song et al. (2020); Wu et al. (2020); Wang et al. (2020); Potel et al. (1980); Armici et al. (1980); Huang et al. (1983); Tarascon et al. (1984); Hor et al. (1985a, b); Brusetti et al. (1988); Tessema et al. (1991); Brusetti et al. (1994); Petrović et al. (2010). It is therefore timely and of theoretical importance to study Kondo problems in these higher-order fermion (KHOF) systems, in particular, using analytically exact methods.

In ideal KHOF systems, the Fermi energy is located at the Dirac point (DP), with higher-order dispersions along certain directions in momentum space. In these cases, the aforementioned linearization approximation in AL’s original CFT approach becomes inapplicable. On one hand, it generates an artificial flat bands along certain directions, as shown by Fig. 1(b). Clearly, the flat band is not sufficient to capture the realistic low-energy excitations of the higher-order fermions. On the other hand, in contrast to the Kondo problems in normal metals with a single band, anisotropic multi-bands with touching nodes must be taken into account. These unique features of the KHOF systems pose severe challenges for the conventional CFT approach. Does there exist any alternative CFT scheme that overcome these difficulties?

In this work, we propose a full-energy mapping CFT (FEMCFT) method. This method solves all the above problems, and produces exact solutions to Kondo problems in a large class of exotic semimetals, in particular, the KHOF models. In contrast with the conventional CFT scheme that only establishes conformal invariance for the linearized low-lying excitations within some artificially-introduced energy cut-off Λ\Lambda near the Fermi surface, our approach is able to map the full energy spectrum of the KHOF model into a form that observes conformal invariance. Consequently, instead of artificially introducing a cut-off Λ\Lambda, and only mapping the low energy degrees of freedom of the bath bounded by Λ\Lambda to the complex plane, we can map the full energy spectrum into the complex plane without introducing any artificial cut-offs in the spectrum, as shown by Fig. 1(c) and 1(d). As a result, our approach is free from the flat band issue shown in Fig. 1(b). An additional advantage of our method is that, by rigorously taking into account the entire energy spectrum of the bath, we can make more accurate predictions about the low-energy fixed point and the thermodynamic quantities at finite temperatures, compared to conventional CFT techniques.

We emphasize that our FEMCFT approach is analytically exact for ideal isotropic KHOF systems. It also works well for highly realistic anisotropic KHOF materials, after systematically treating the anisotropic effects as corrections to the isotropic parts. Our work therefore rigorously solves the KHOF and related models for higher-order topological semimetals, and significantly advances the scope of CFT approaches to many-body resonances, bringing about new applications in previously inaccessible systems.

The remaining part of the manuscript is organized as follows. In Sec.II, we define the general Hamiltonian for the KHOF systems, which describes a multichannel Kondo impurity in higher-order Weyl/Dirac systems. In Sec.III, we present our FEMCFT approach in the isotropic limit, which establishes an exact mapping of an ideal isotropic KHOF system to a 2D CFT in the complex plane. The calculations of thermodynamic quantities are discussed in Sec.IV. In Sec.V, we tackle realistic anisotropic KHOF systems, and demonstrate our method with an explicit example, namely, the Kondo problem in the anisotropic cubic Weyl/Dirac fermion system in three dimensions, whose dispersion is given by (1), which has attracted great interests recently.

Refer to caption
Figure 1: Figure (a) denotes AL’s original approach applied to normal metals, in which an artificial cut-off ±ΛvF\pm\Lambda v_{F} about the Fermi surface has to be introduced. Only low-lying excitations satisfying ΛvF<ϵ<+ΛvF-\Lambda v_{F}<\epsilon<+\Lambda v_{F}, which are approximated to be linear, are mapped into the complex plane denoted by Figure (d), although the impurity also couples to the higher-energy regimes of the bath. Figure (b) shows how AL’s original approach becomes inapplicable to KHOF systems, which have their Fermi points located exactly at their Dirac points. AL’s cut-off becomes infinitesimally narrow and their linearization approximation produces a flat band; no degrees of freedom in the KHOF model can be mapped into the complex plane. Figure (c) denotes our FEMCFT approach applied to KHOF systems. No energy cut-offs and linearization approximations are required, and the full energy spectrum is mapped into the complex plane, establishing an exact CFT solution to KHOF models.

II General Hamiltonian for KHOF systems

For the sake of generality, we consider a KHOF system in nn spatial dimensions, whose Hamiltonian is given by

H=H0+HI,H=H_{0}+H_{I}\,, (2)

where H0H_{0} is the bath Hamiltonian, and HIH_{I} is the interaction Hamiltonian. H0H_{0} is given by

H0=dnpcα,i,μ(p)cα,i,μ(p)ϵμ(p),H_{0}=\int d^{n}\vec{p}\,c^{\dagger}_{\alpha,i,\mu}(\vec{p})c_{\alpha,i,\mu}(\vec{p})\epsilon_{\mu}(\vec{p})\,, (3)

where p=(p1,,pn)\vec{p}=(p_{1},...,p_{n}), cα,i,μ(p)c^{\dagger}_{\alpha,i,\mu}(\vec{p}) and cα,i,μ(p)c_{\alpha,i,\mu}(\vec{p}) are conduction electron creation and annihilation operators respectively, with α{,}\alpha\in\{\uparrow,\downarrow\} labelling the up/down spin index, i{1,2,,k}i\in\{1,2,...,k\} labelling the channel index, and μ{+,}\mu\in\{+,-\} labelling the band index. Summation over repeated indices is implied.

We consider a general higher-order dispersion ϵμ(p)\epsilon_{\mu}(\vec{p}) of the form

ϵμ(p)=a1μpn+a2μp1,\epsilon_{\mu}(\vec{p})=a_{1\mu}p^{n}+a_{2\mu}p_{1}\,, (4)

where p|p|p\equiv|\vec{p}|. The constants a1μa_{1\mu} and a2μa_{2\mu} satisfy a1+=a1a_{1+}=-a_{1-}, a2+=a2a_{2+}=-a_{2-}. The dispersion (4) is of particular interest, because for n=3n=3, it reduces to a cubic fermion system in three-dimensions (1). For n=2n=2, it produces a quadratic band crossing point, which also has attracted great interests in the past decade Sun et al. (2009); Uebelacker and Honerkamp (2011); Wang et al. (2015). Moreover, for n3n\geq 3, it describes even higher order emergent fermions which can be realized in artificial electric circuits with auxiliary dimensions Yu et al. (2020).

HIH_{I} describes the exchange interaction between an impurity spin and the bath electrons:

HI=λiS(2π)nμ,νdnpdnpcα,i,μ(p)σαβ2cβ,i,ν(p),H_{I}=\frac{\lambda_{i}\vec{S}}{(2\pi)^{n}}\cdot\sum_{\mu,\nu}\int d^{n}\vec{p}\,d^{n}\vec{p}^{\prime}c^{\dagger}_{\alpha,i,\mu}(\vec{p})\frac{\vec{\sigma}_{\alpha\beta}}{2}c_{\beta,i,\nu}(\vec{p}^{\prime})\,, (5)

where S\vec{S} is the spin of the magnetic impurity, λi\lambda_{i} is the Kondo coupling constant for the ii-th channel.

Eq. (2) - (5) constitute the most general KHOF model in nn-dimensions. Our focus is to look for exact CFT solutions to Eq. (2) - (5).

III FEMCFT approach in the isotropic limit

In this section, we outline our FEMCFT approach for an ideal isotropic KHOF system, i.e. the case where a2μ=0a_{2\mu}=0 in (4), so that

ϵμ(p)=a1μpn.\displaystyle\epsilon_{\mu}(p)=a_{1\mu}p^{n}\,. (6)

More details of the derivations in this section can be found in Appendix B. We shall treat the anisotropic case a2μ0a_{2\mu}\neq 0 in Sec.V.

We expand cα,i,μ(p)c_{\alpha,i,\mu}(\vec{p}) as a linear combination of the spherical harmonics, which form a complete set of orthonormal functions on the (n1)(n-1)- sphere Sn1S^{n-1}:

cα,i,μ(p)=1pn12\displaystyle c_{\alpha,i,\mu}(\vec{p})=\frac{1}{p^{\frac{n-1}{2}}}\cdot
l1,,ln1Yl1,,ln1(θ1,,θn1)cl1,,ln1,α,i,μ(p).\displaystyle\sum_{l_{1},...,l_{n-1}}Y_{l_{1},...,l_{n-1}}(\theta_{1},...,\theta_{n-1})c_{l_{1},...,l_{n-1},\alpha,i,\mu}(p)\,. (7)

Here Yl1,,ln1(θ1,,θn1)Y_{l_{1},...,l_{n-1}}(\theta_{1},...,\theta_{n-1}) are the spherical harmonics in nn dimensions, see Appendix A and Higuchi (1987) for detailed derivations of their important properties relevant for this work. θ1,,θn1\theta_{1},...,\theta_{n-1} are the n1n-1 angular coordinates of Sn1S^{n-1}, with 0θ1<2π0\leq\theta_{1}<2\pi, 0θjπ0\leq\theta_{j}\leq\pi for j=2,,n1j=2,...,n-1. The integers |l1|l2ln1|l_{1}|\leq l_{2}\leq...\leq l_{n-1} denote different partial waves, and are the analogues of the quantum numbers m,lm,l in the three-dimensional spherical harmonics Ylm(θ,ϕ)Y_{l}^{m}(\theta,\phi): in 3 dimensions, n=3n=3, l1m,l2ll_{1}\equiv m,l_{2}\equiv l, θ1ϕ\theta_{1}\equiv\phi is the azimuthal angle, and θ2θ\theta_{2}\equiv\theta is the polar angle. H0H_{0} and HIH_{I} can be written in terms of cl1,,ln1,α,i,μ(p)c_{l_{1},...,l_{n-1},\alpha,i,\mu}(p) as

HI=λiSΓ(n2)2nπn2μ,ν\displaystyle H_{I}=\frac{\lambda_{i}\vec{S}}{\Gamma(\frac{n}{2})2^{n}\pi^{\frac{n}{2}}}\cdot\sum_{\mu,\nu}\int dpdppn12pn12\displaystyle dpdp^{\prime}p^{\frac{n-1}{2}}p^{\prime\frac{n-1}{2}}
c0,,0,α,i,μ(p)σαβc0,,0,β,i,ν(p),\displaystyle c^{\dagger}_{0,...,0,\alpha,i,\mu}(p)\vec{\sigma}_{\alpha\beta}c_{0,...,0,\beta,i,\nu}(p^{\prime})\,, (8)
H0=𝑑pc0,,0,α,i,μ(p)c0,,0,α,i,μ(p)ϵμ(p).\displaystyle H_{0}=\int dp\,c^{\dagger}_{0,...,0,\alpha,i,\mu}(p)c_{0,...,0,\alpha,i,\mu}(p)\epsilon_{\mu}(p)\,. (9)

Next we perform a change of variables on the fields cl1,,ln1,α,i,μ(p)c_{l_{1},...,l_{n-1},\alpha,i,\mu}(p) from pp to ϵ\epsilon, by defining

cl1,,ln1,α,i,μ(ϵμ)|dϵμ(p)dp|12cl1,,ln1,α,i,μ(p).c_{l_{1},...,l_{n-1},\alpha,i,\mu}(\epsilon_{\mu})\equiv\left|\frac{d\epsilon_{\mu}(p)}{dp}\right|^{-\frac{1}{2}}c_{l_{1},...,l_{n-1},\alpha,i,\mu}(p)\,. (10)

In this way, the cl1,,ln1,α,i,μ(ϵμ)c_{l_{1},...,l_{n-1},\alpha,i,\mu}(\epsilon_{\mu}) fields also satisfy the proper fermionic anti-commutation relations. With the definition

Φ0,,0,α,i(ϵ){c0,,0,α,i,+(ϵ)if ϵ0c0,,0,α,i,(ϵ)if ϵ<0,\displaystyle\Phi_{0,...,0,\alpha,i}(\epsilon)\equiv\begin{cases}c_{0,...,0,\alpha,i,+}(\epsilon)&\text{if }\epsilon\geq 0\\ c_{0,...,0,\alpha,i,-}(\epsilon)&\text{if }\epsilon<0\end{cases}\,, (11)

we combine the fields from the + and - bands into one single composite fermionic field Φ0,,0,α,i(ϵ)\Phi_{0,...,0,\alpha,i}(\epsilon). In terms of Φ0,,0,α,i(ϵ)\Phi_{0,...,0,\alpha,i}(\epsilon), H0H_{0} and HIH_{I} become

H0=𝑑ϵΦ0,,0,α,i(ϵ)Φ0,,0,α,i(ϵ)ϵ,\displaystyle H_{0}=\int_{-\infty}^{\infty}d\epsilon\,\Phi^{\dagger}_{0,...,0,\alpha,i}(\epsilon)\Phi_{0,...,0,\alpha,i}(\epsilon)\epsilon\,, (12)
HI\displaystyle H_{I} =λiSanΓ(n2)2nπn2\displaystyle=\frac{\lambda_{i}\vec{S}}{an\Gamma(\frac{n}{2})2^{n}\pi^{\frac{n}{2}}}
dϵdϵΦ0,,0,α,i(ϵ)σαβΦ0,,0,β,i(ϵ),\displaystyle\cdot\int_{-\infty}^{\infty}d\epsilon\int_{-\infty}^{\infty}d\epsilon^{\prime}\Phi^{\dagger}_{0,...,0,\alpha,i}(\epsilon)\vec{\sigma}_{\alpha\beta}\Phi_{0,...,0,\beta,i}(\epsilon^{\prime})\,, (13)

where a|a1μ|a\equiv|a_{1\mu}|. Notice that we have combined the ++ and - bands into a single band, and eliminated the band index μ{+,}\mu\in\{+,-\} from our model.

We then define the left and right moving fields, respectively, as

Ψαi(r)𝑑ϵeiϵrΦ0,,0,α,i(ϵ)\displaystyle\Psi_{\leftarrow\alpha i}(r)\equiv\int_{-\infty}^{\infty}d\epsilon\,e^{-i\epsilon r}\Phi_{0,...,0,\alpha,i}(\epsilon)
Ψαi(r)𝑑ϵe+iϵrΦ0,,0,α,i(ϵ).\displaystyle\Psi_{\rightarrow\alpha i}(r)\equiv\int_{-\infty}^{\infty}d\epsilon\,e^{+i\epsilon r}\Phi_{0,...,0,\alpha,i}(\epsilon)\,. (14)

Also, by introducing the imaginary time τit\tau\equiv it, we define the complex plane \mathbb{C}, to which we map our KHOF model, by

={zτ+ir}.\mathbb{C}=\{z\equiv\tau+ir\}\,. (15)

i.e. The horizontal axis of \mathbb{C} is the imaginary time τ\tau, and the vertical axis of \mathbb{C} is rr introduced in (III). We then view Ψ/αi(r)\Psi_{\leftarrow/\rightarrow\alpha i}(r) in the Heisenberg picture, so that they now have time-dependence, and live on \mathbb{C}. Ψαi(τ,r)\Psi_{\leftarrow\alpha i}(\tau,r) and Ψαi(τ,r)\Psi_{\rightarrow\alpha i}(\tau,r) are related to each other by

Ψαi(τ,r)=Ψαi(τ,r),\Psi_{\rightarrow\alpha i}(\tau,r)=\Psi_{\leftarrow\alpha i}(\tau,-r)\,, (16)

so Ψαi(τ,r)\Psi_{\rightarrow\alpha i}(\tau,r) can be eliminated in terms of Ψαi(τ,r)\Psi_{\leftarrow\alpha i}(\tau,r). H0H_{0} and HIH_{I} can be written in terms as Ψαi(τ,r)\Psi_{\leftarrow\alpha i}(\tau,r) as

H0=12π𝑑r(Ψαi(τ,r)irΨαi(τ,r)),\displaystyle H_{0}=\frac{1}{2\pi}\int^{\infty}_{-\infty}dr\left(\Psi^{\dagger}_{\leftarrow\alpha i}(\tau,r)i\frac{\partial}{\partial r}\Psi_{\leftarrow\alpha i}(\tau,r)\right)\,, (17)
HI=\displaystyle H_{I}= λiSanΓ(n2)2nπn2Ψαi(τ,0)σαβΨβi(τ,0).\displaystyle\frac{\lambda_{i}\vec{S}}{an\Gamma(\frac{n}{2})2^{n}\pi^{\frac{n}{2}}}\cdot\Psi^{\dagger}_{\leftarrow\alpha i}(\tau,0)\vec{\sigma}_{\alpha\beta}\Psi_{\leftarrow\beta i}(\tau,0)\,. (18)

The nn-dimensional KHOF model is thus mapped into the complex plane \mathbb{C} defined in (15), with H0H_{0} and HIH_{I} taking the forms of (17) and (18) respectively. These are in similar forms as the Hamiltonians mapped from single band normal metals. In particular, we see from (18) that the Kondo exchange coupling remains short-ranged in the new fermionic degrees of freedom; the impurity spin S\vec{S} only couples to the new fermionic field Ψαi(τ,r)\Psi_{\leftarrow\alpha i}(\tau,r) at r=0r=0. This means that HIH_{I} is only confined to the boundary r=0r=0, with H=H0H=H_{0} in the bulk r0r\neq 0, and the problem is suitable for further analysis by techniques in 2D boundary CFT.

We note that our FEMCFT approach is analytically exact; we have not introduced any artificial cut-off Λ\Lambda in the momentum or energy. Our integrals 0𝑑p\int_{0}^{\infty}dp and 𝑑ϵ\int_{-\infty}^{\infty}d\epsilon are over the entire spectrum. Also, we have transformed the 2-band KHOF system into an effective one-band system. This further facilitates the analysis of KHOF systems using CFT techniques.

IV Thermodynamic Quantities

After mapping KHOF systems into the form (17) + (18) on the complex plane via our FEMCFT approach, we can proceed to determine the thermodynamic quantities at T>0T>0. This is carried out via AL’s standard CFT techniques, which we briefly summarize here. More details on their approach can be found in Affleck and Ludwig (1991a, b, c); Affleck et al. (1992); Affleck and Ludwig (1993); Ludwig (1994); Affleck (1995); Delft (1995). The underlying geometry of the T>0T>0 physics is the infinite cylinder with circumference β=1/T\beta=1/T. The system is further mapped from the complex plane onto the infinite cylinder via a conformal map. The complete list of boundary operators for the system with a particular boundary condition can be then obtained by applying “double fusion” to the free system (i.e. the system with trivial boundary condition). From the list of boundary operators, we can determine the leading irrelevant operator with coupling constant λ~\tilde{\lambda}, whose Green’s functions allow us to compute the thermodynamic quantities of interest. The circumference of the cylinder β=1/T\beta=1/T enters the calculation of the thermodynamic quantities as a finite size of the system, giving rise to their TT-dependences. For example, the resistivity ρ\rho is found as ρ=ρU{1αλ~2T2+}\rho=\rho_{U}\left\{1-\alpha\tilde{\lambda}^{2}T^{2}+...\right\} in the Fermi liquid case and ρ=ρU1S(1)2{1+αλ~T22+k+}\rho=\rho_{U}\frac{1-S^{(1)}}{2}\left\{1+\alpha\tilde{\lambda}T^{\frac{2}{2+k}}+...\right\} in the non-Fermi liquid case. Here kk is the number of channels, S(1)=cos{π(2s+1)/(2+k)}cos{π/(2+k)}S^{(1)}=\frac{\cos\{\pi(2s+1)/(2+k)\}}{\cos\{\pi/(2+k)\}} with ss being the impurity spin, ρU\rho_{U} is the unitary limit resistivity, i.e. the greatest resistivity possibly achievable, and α\alpha is a dimensionless constant, with α=4π\alpha=4\sqrt{\pi} in the special case k=2k=2. These results are in excellent agreements with results obtained from numerical renormalization group (NRG) analysis.

We remark that, although our T>0T>0 thermodynamic quantities take the same form as AL’s, they are valid over greater ranges of temperatures compared to AL’s. This is because AL’s original approach to Kondo problems in normal metals requires the introduction of a narrow cutoff Λ\Lambda about the Fermi surface. Thus in their calculated thermodynamic quantities, AL can only consider temperatures TT satisfying TΛvFT\ll\Lambda v_{F}, and ignore terms of order T/ΛvFT/\Lambda v_{F}. In comparison, we did not introduce any artificial cut-offs in the spectrum. As a result, our calculated thermodynamic quantities are more accurate and are valid over greater ranges of temperatures, without the restriction of TΛvFT\ll\Lambda v_{F}. Moreover, conventional CFT approaches are not applicable at all in KHOF systems with coinciding Fermi energies and DPs. Our theory fills this research gap, and enables an exact CFT analysis for these novel phases with emergent higher-order fermions.

V Realistic cases with anisotropy

Realistic topological semimetals with higher order fermions are always anisotropic. In this section, we therefore present a general framework to treat this anisotropy, followed by an explicit application of our framework to a concrete example, namely the anisotropic cubic fermion model in three dimensions, realized by setting n=3n=3 and a2μ0a_{2\mu}\neq 0 in (4). More details of the derivations can be found in Appendix C.

For anisotropic KHOF systems, HIH_{I} remains the same as before, since the dispersion relation does not enter the expression of HIH_{I}. However, H0H_{0} can no longer be reduced to the simple form (9), due to the lack of spherical symmetry. In general, we can express H0H_{0} into a matrix form in the partial wave basis, i.e.,

H0=𝑑pCα,i,μ(p)Aμ(p)Cα,i,μ(p),H_{0}=\int dp\,C^{\dagger}_{\alpha,i,\mu}(p)A_{\mu}(p)C_{\alpha,i,\mu}(p),\ (19)

where Cα,i,μ(p)C_{\alpha,i,\mu}(p) is a column vector whose LL-th element is given by cL,α,i,μ(p)c_{L,\alpha,i,\mu}(p), where for brevity we have used the multi-index notation L(l1,,ln1)L\equiv(l_{1},...,l_{n-1}). Cα,i,μ(p)C^{\dagger}_{\alpha,i,\mu}(p) is the Hermitian conjugate of Cα,i,μ(p)C_{\alpha,i,\mu}(p) and Aμ(p)A_{\mu}(p) is a matrix whose (L,L)(L,L^{\prime})-th element, AL,L,μ(p)A_{L,L^{\prime},\mu}(p), describes the coupling between the LL and LL^{\prime}-th partial waves. We remark that for all Hermitian systems, the only partial waves that couple to the impurity are those with l1=0l_{1}=0. Thus the multi-index LL only includes (l1=0,l2,,ln1)(l_{1}=0,l_{2},...,l_{n-1}). In particular, in three dimensions, LL only includes (m=0,l)l(m=0,l)\rightarrow l, i.e. the multi-index LL reduces to the single index ll: L=lL=l.

Refer to caption
Figure 2: The couplings between the impurity and the different partial waves in the three-dimensional KHOF system with dispersion Eq.(22) is illustrated by a hierarchical one-dimensional chain. Only the m=0m=0 partial waves are relevant, which is a property universal for all Hermitian systems. The impurity only couples directly to the l=0l=0, m=0m=0 partial wave, which further couples to higher-order ll-components via the nearest “hoppings”. The hopping coefficient decays with ll. Clearly, the anisotropy is manifested by the higher-order corrections to the isotropic component. By integrating out the l0l\neq 0 components, the impurity is effectively coupled to the renormalized l=0l=0 m=0m=0 component, where the anisotropic effects have been absorbed as corrections to the isotropic part.

We separate H0H_{0} in (19) into two terms

H0=H00+Σ0,H_{0}=H_{00}+\Sigma_{0}\,, (20)

where the term H00H_{00} consists of only the L=(0,,0)L=(0,...,0) partial waves, thus describing the “isotropic part” of H0H_{0}. Σ0\Sigma_{0} consists of all the remaining partial waves with L(0,,0)L\neq(0,...,0), and captures the “anisotropic part” of H0H_{0}, as well as its couplings to H00H_{00}, as shown by left picture of Fig.2.

Due to anisotropy, partial waves with different quantum numbers are coupled to each other, forming a hierarchical structure depicted in the left picture of Fig. 2 (in general, different models can give rise to different coupling hierarchical structures. The left picture of Fig. 2 shows the case for the cubic fermion model in three dimensions (22), which suffices to illustrate our framework). As shown in Fig. 2, the impurity firstly couples to the isotropic sector, which further couples to the higher components. Owing to the hierarchical nature of the couplings, we can integrate out the anisotropic part by down-folding the Hamiltonian matrix in Eq.(19) to the isotropic subspace. It can be shown that in the static limit, the anisotropic effects of Σ0\Sigma_{0} can be casted into an additional renormalization term ϵΣμ(p)\epsilon_{\Sigma\mu}(p), which serves as a modification to the isotropic part of the dispersion relation ϵμ(p)\epsilon_{\mu}(p).

To further enable a CFT analysis, we approximate the value of ϵΣμ(p)\epsilon_{\Sigma\mu}(p) by evaluating it at the Fermi wave number, pFp_{F}. This fully captures the low-energy Kondo physics, since only the excitations near the Fermi point are important at low temperatures. It should be noted that, although this approximation is in the same spirit as AL’s linearization approximation near the Fermi surface, it is made only to the anisotropic part of the dispersion. Our method still admits an exact treatment of the isotropic part by mapping the entire spectrum into a two-dimensional CFT. Therefore, it is free from the flat band issue illustrated in Fig. 1(b), and overcomes the major difficulty present in conventional CFT techniques.

The above procedures generate an renormalized bath coupled to the impurity, as indicated by the right picture of Fig. 2. The renormalized bath enjoys the effective dispersion relation ϵ~μ(p)ϵμ(p)+ϵΣμ(pF)\tilde{\epsilon}_{\mu}(p)\equiv\epsilon_{\mu}(p)+\epsilon_{\Sigma\mu}(p_{F}), which contains the isotropic part of the dispersion, as well as the corrections from the anisotropy. Correspondingly, the bath is now effectively described by

H0=H00+Σ0𝑑pc0,,0,α,i,μ(p)c0,,0,α,i,μ(p)ϵ~μ(p).\displaystyle H_{0}=H_{00}+\Sigma_{0}\approx\int dp\,c^{\dagger}_{0,...,0,\alpha,i,\mu}(p)c_{0,...,0,\alpha,i,\mu}(p)\tilde{\epsilon}_{\mu}(p)\,. (21)

The above steps constitute a general framework that takes into account the anisotropy in the CFT analysis.

We now apply our approach to a concrete example, namely the cubic fermion system in three dimensions. This is realized by setting n=3n=3 and turning on a2μa_{2\mu} in (4), namely

ϵ(p)=a1μp3+a2μp1,\epsilon(\vec{p})=a_{1\mu}p^{3}+a_{2\mu}p_{1}\,, (22)

which is linear in the p1p_{1}-direction but cubic in the orthogonal plane, a typical low-energy dispersion of cubic fermion materials.

This model admits a simple coupling hierarchical structure as shown in Fig. 2. Here, the impurity is coupled only to the isotropic component, which is further connected to the higher-order partial waves via nearest-neighbour hoppings. Moreover, it can be shown that the hopping strength decays as ll increases. As a result, Aμ(p)A_{\mu}(p) in (19) can be casted into the simple tridiagonal form

Aμ(p)=(a1μp3a2μp3a2μp3a1μp32a2μp152a2μp15a1μp33a2μp353a2μp35a1μp34a2μp63).\displaystyle A_{\mu}(p)=\left(\begin{array}[]{@{}cccccc@{}}a_{1\mu}p^{3}&\frac{a_{2\mu}p}{\sqrt{3}}&&&&\\ \frac{a_{2\mu}p}{\sqrt{3}}&a_{1\mu}p^{3}&\frac{2a_{2\mu}p}{\sqrt{15}}&&&\\ &\frac{2a_{2\mu}p}{\sqrt{15}}&a_{1\mu}p^{3}&\frac{3a_{2\mu}p}{\sqrt{35}}&&\\ &&\frac{3a_{2\mu}p}{\sqrt{35}}&a_{1\mu}p^{3}&\frac{4a_{2\mu}p}{\sqrt{63}}&\\ &&&\ddots&\ddots&\ddots\end{array}\right)\,. (28)

We then integrate out the higher-order partial waves by down-folding the matrix Aμ(p)A_{\mu}(p) to the isotropic subspace. Interestingly, because each entry of Aμ(p)A_{\mu}(p) is proportional to pp, the renormalization ϵΣμ(p)\epsilon_{\Sigma_{\mu}}(p) is also found to be proportional to pp, and thus vanishes at the DP with pF=0p_{F}=0. This essentially indicates that the contributions from different high-order partial waves display a “destructive interference”, in the sense that they completely cancel with each other at the DP. This is an interesting feature of cubic Dirac fermions, which greatly simplifies the treatment of anisotropy.

The effective dispersion relation is eventually obtained as

ϵ~μ(p)=a1μp3,\tilde{\epsilon}_{\mu}(p)=a_{1\mu}p^{3}\,, (29)

which is clearly of the form (6) in three dimensions. As a result, all subsequent derivations following Eq. (6) hold, and our FEMCFT approach for isotropic KHOF systems applies. H0H_{0} and HIH_{I} will be mapped to the forms of (17) and (18) respectively. Correspondingly, the impurity ground state and the thermodynamic quantities can be readily obtained using the previously discussed methods.

VI Conclusion

In conclusion, we have developed a full-energy mapping conformal field theory (FEMCFT) approach to tackle Kondo problems in higher order fermion and related systems. Our FEMCFT approach is capable of solving these models exactly, which are inaccessible by using conventional CFT approaches. Moreover, the FEMCFT approach is able to make more accurate predictions about the thermodynamic quantities over greater ranges of temperatures. We applied our FEMCFT approach to a specific KHOF system, namely the cubic Weyl/Dirac fermion system in three dimensions. The high efficiency of our approach to realistic KHOF systems is clearly justified.

The FEMCFT approach significantly broadens the scope of existing CFT methods in the study of Kondo problems. We anticipate developments of novel CFT techniques for treating Kondo problems in more complicated systems, such as those displaying pseudogaps, as possible future directions of research. Such advancements would undoubtedly further fortify the connections between the mathematically elegant CFTs and the physically intriguing fixed points in strong-correlated problems, such as novel quantum criticalities and many-body resonances in strong-coupling limits.

Acknowledgements.
We are grateful to W. Su, Bin Tai, Feng Tang, Tigran Sedrakyan, and Y. X. Zhao for fruitful discussions. This work was supported by the Jiangsu Postdoctoral Research Grant (Grant No. 2020Z019), the Youth Program of National Natural Science Foundation of China (No. 11904225) and the National Key R&D Program of China (Grant No. 2017YFA0303200).

Appendix A Spherical Harmonics in Higher Dimensions

We review the properties of the nn-dimensional spherical harmonics used in the main section. Some of this material can also be found in Higuchi (1987).

The nn-dimensional spherical harmonics are the eigenfunctions of ΔSn1\Delta_{S^{n-1}}, the Laplace-Beltrami operator on the sphere Sn1S^{n-1}, which is the angular part of the nn-dimensional Laplacian operator. The Laplace-Beltrami operator can be defined recursively:

ΔSj=\displaystyle\Delta_{S^{j}}= (1sinj1θj)θj(sinj1θjθj)+ΔSj1sin2θj\displaystyle\left(\frac{1}{\sin^{j-1}\theta_{j}}\right)\frac{\partial}{\partial\theta_{j}}\left(\sin^{j-1}\theta_{j}\frac{\partial}{\partial\theta_{j}}\right)+\frac{\Delta_{S^{j-1}}}{\sin^{2}\theta_{j}} (30)

for 2jn12\leq j\leq n-1, and

ΔS1=2θ1 2,\Delta_{S^{1}}=\frac{\partial^{2}}{\partial\theta_{1}^{\,2}}\,, (31)

where θ1,,θn1\theta_{1},...,\theta_{n-1} are the n1n-1 angular coordinates in the nn-dimensional spherical coordinates, with 0θ1<2π0\leq\theta_{1}<2\pi, 0θ2,,θn1π0\leq\theta_{2},...,\theta_{n-1}\leq\pi. An nn-dimensional spherical harmonic Y(θ1,,θn1)Y(\theta_{1},...,\theta_{n-1}) of polynomial degree ln1l_{n-1}, where ln1l_{n-1} is a non-negative integer, has eigenvalue ln1(ln1+n2)-l_{n-1}(l_{n-1}+n-2):

ΔSn1Y(θ1,,θn1)=\displaystyle\Delta_{S^{n-1}}\,Y(\theta_{1},...,\theta_{n-1})= ln1(ln1+n2)\displaystyle-l_{n-1}(l_{n-1}+n-2)
Y(θ1,,θn1).\displaystyle Y(\theta_{1},...,\theta_{n-1})\,. (32)

This equation can be solved by separation of variables. By writing

Y(θ1,,θn1)=Y(θ1,,θn2)Θn1(θn1),Y(\theta_{1},...,\theta_{n-1})=Y^{\prime}(\theta_{1},...,\theta_{n-2})\Theta_{n-1}(\theta_{n-1})\,, (33)

where Y(θ1,,θn2)Y^{\prime}(\theta_{1},...,\theta_{n-2}) is a spherical harmonic of degree ln2l_{n-2} on Sn2S^{n-2}, and Θn1(θn1)\Theta_{n-1}(\theta_{n-1}) is a function that only depends on θn1\theta_{n-1}, to be further determined. Because Y(θ1,,θn2)Y^{\prime}(\theta_{1},...,\theta_{n-2}) is a factor of Y(θ1,,θn1)Y(\theta_{1},...,\theta_{n-1}), its polynomial degree must not be larger than that of Y(θ1,,θn1)Y(\theta_{1},...,\theta_{n-1}), thus ln2ln1l_{n-2}\leq l_{n-1}. We can now write (A) as

sin4nθn1Θn1θn1(sinn2θn1Θn1θn1)\displaystyle\frac{\sin^{4-n}\theta_{n-1}}{\Theta_{n-1}}\frac{\partial}{\partial\theta_{n-1}}\left(\sin^{n-2}\theta_{n-1}\frac{\partial\Theta_{n-1}}{\partial\theta_{n-1}}\right)
+\displaystyle+ ln1(ln1+n2)(sin2θn1)+1YΔSn2Y=0.\displaystyle l_{n-1}(l_{n-1}+n-2)(\sin^{2}\theta_{n-1})+\frac{1}{Y^{\prime}}\Delta_{S^{n-2}}Y^{\prime}=0\,. (34)

The first two terms only depend on θn1\theta_{n-1}, and the third term only depend on θ1,,θn2\theta_{1},...,\theta_{n-2}. Since Y(θ1,,θn2)Y^{\prime}(\theta_{1},...,\theta_{n-2}) is a spherical harmonic of degree ln2l_{n-2} on Sn2S^{n-2}, we have

ΔSn2Y(θ1,,θn2)=\displaystyle\Delta_{S^{n-2}}\,Y^{\prime}(\theta_{1},...,\theta_{n-2})= ln2(ln2+n3)\displaystyle-l_{n-2}(l_{n-2}+n-3)\cdot
Y(θ1,,θn2),\displaystyle Y^{\prime}(\theta_{1},...,\theta_{n-2})\,, (35)

so the third term in (A) must equal to ln2(ln2+n3)-l_{n-2}(l_{n-2}+n-3). As a result, the first two terms of (A) must equal to +ln2(ln2+n3)+l_{n-2}(l_{n-2}+n-3):

0=sin4nθn1θn1(sinn2θn1Θn1θn1)+Θn1\displaystyle 0=\sin^{4-n}\theta_{n-1}\frac{\partial}{\partial\theta_{n-1}}\left(\sin^{n-2}\theta_{n-1}\frac{\partial\Theta_{n-1}}{\partial\theta_{n-1}}\right)+\Theta_{n-1}\cdot
{ln1(ln1+n2)(sin2θn1)ln2(ln2+n3)}.\displaystyle\left\{l_{n-1}(l_{n-1}+n-2)(\sin^{2}\theta_{n-1})-l_{n-2}(l_{n-2}+n-3)\right\}\,. (36)

Now define

P¯cab(θ)\displaystyle{}_{b}\bar{P}_{c}^{a}(\theta)\equiv (2c+b1)(a+b+c2)!2(ca)!\displaystyle\sqrt{\frac{(2c+b-1)(a+b+c-2)!}{2(c-a)!}}
sin2b2(θ)Pc+b22(a+b22)(cosθ),\displaystyle\sin^{\frac{2-b}{2}}(\theta)P^{-(a+\frac{b-2}{2})}_{c+\frac{b-2}{2}}(\cos\theta)\,, (37)

where Plm(x)P^{m}_{l}(x) is the associated Legendre function of the first kind. In our case m,lm,l are generalized to take on integer or half-integer values. By noting that Plm(x)P^{m}_{l}(x) solves the Legendre equation:

(1x2)d2dx2Plm(x)2xddxPlm(x)\displaystyle(1-x^{2})\frac{d^{2}}{dx^{2}}P^{m}_{l}(x)-2x\frac{d}{dx}P^{m}_{l}(x)
+\displaystyle+ {l(l+1)m21x2}Plm(x)=0,\displaystyle\left\{l(l+1)-\frac{m^{2}}{1-x^{2}}\right\}P^{m}_{l}(x)=0\,, (38)

the solution to equation (A) is

Θn1=P¯ln1ln2n1(θn1).\Theta_{n-1}={}_{n-1}\bar{P}_{l_{n-1}}^{l_{n-2}}(\theta_{n-1})\,. (39)

Equation (A) is in the same form as (A) and can be solved recursively by repeating the procedure up to now, yielding Y(θ1,,θn2)=Y′′(θ1,,θn3)Θn2(θn2)Y^{\prime}(\theta_{1},...,\theta_{n-2})=Y^{\prime\prime}(\theta_{1},...,\theta_{n-3})\Theta_{n-2}(\theta_{n-2}) and Θn2(θn2)=P¯ln2ln3n2(θn2)\Theta_{n-2}(\theta_{n-2})={}_{n-2}\bar{P}_{l_{n-2}}^{l_{n-3}}(\theta_{n-2}). Eventually we obtain Θ2(θ2)=P¯l2l12(θ2)\Theta_{2}(\theta_{2})={}_{2}\bar{P}_{l_{2}}^{l_{1}}(\theta_{2}), and the equation

ΔS1Θ1(θ1)=l12Θ1(θ1)\displaystyle\Delta_{S^{1}}\,\Theta_{1}(\theta_{1})=-l_{1}^{2}\,\Theta_{1}(\theta_{1}) (40)

where ΔS1\Delta_{S^{1}} is given by (31). Its solution is

Θ1(θ1)=eil1θ1.\Theta_{1}(\theta_{1})=e^{il_{1}\theta_{1}}\,. (41)

Thus we arrive at

Yl1,,ln1(θ1,,θn1)=\displaystyle Y_{l_{1},...,l_{n-1}}(\theta_{1},...,\theta_{n-1})= Aj=1n1Θj(θj)\displaystyle A\prod^{n-1}_{j=1}\Theta_{j}(\theta_{j})
=\displaystyle= Aeil1θ1j=2n1P¯ljlj1j(θj),\displaystyle Ae^{il_{1}\theta_{1}}\prod^{n-1}_{j=2}{}_{j}\bar{P}_{l_{j}}^{l_{j-1}}(\theta_{j})\,, (42)

where |l1|l2ln1|l_{1}|\leq l_{2}\leq...\leq l_{n-1}. The integers l1,l2,ln1l_{1},l_{2},...l_{n-1} are analogues of the quantum numbers m,lm,l in the 3-dimensional spherical harmonics Ylm(θ,ϕ)Y_{l}^{m}(\theta,\phi). Indeed, |l1|l2|l_{1}|\leq l_{2} is due to the same reason as |m|l|m|\leq l in the 3D case. In particular, in 3D, n=3n=3, l1m,l2l,θ1ϕl_{1}\equiv m,l_{2}\equiv l,\theta_{1}\equiv\phi is the azimuthal angle, and θ2θ\theta_{2}\equiv\theta is the polar angle. Also, AA is a normalization constant to be determined. Before determining AA, let us show that the nn-dimensional spherical harmonics satisfy the orthonormality condition:

Yl1,,ln1(θ1,,θn1)Yl1,,ln1(θ1,,θn1)𝑑Ωn1\displaystyle\int Y^{*}_{l_{1},...,l_{n-1}}(\theta_{1},...,\theta_{n-1})Y_{l^{\prime}_{1},...,l^{\prime}_{n-1}}(\theta_{1},...,\theta_{n-1})d\Omega_{n-1}
=\displaystyle= δl1l1δl2l2δln1ln1.\displaystyle\delta_{l_{1}l^{\prime}_{1}}\delta_{l_{2}l^{\prime}_{2}}...\delta_{l_{n-1}l^{\prime}_{n-1}}\,. (43)

The normality condition in (A) can help us in determining the normalization constant AA. Consider the integral

Yl1,,ln1(θ1,,θn1)Yl1,,ln1(θ1,,θn1)𝑑Ωn1\displaystyle\int Y^{*}_{l_{1},...,l_{n-1}}(\theta_{1},...,\theta_{n-1})Y_{l^{\prime}_{1},...,l^{\prime}_{n-1}}(\theta_{1},...,\theta_{n-1})d\Omega_{n-1}
=\displaystyle= |A|2eil1θ1eil1θ1j=2n1P¯ljlj1j(θj)P¯ljlj1j(θj)dΩn1.\displaystyle|A|^{2}\int e^{-il_{1}\theta_{1}}e^{il^{\prime}_{1}\theta_{1}}\prod^{n-1}_{j=2}{}_{j}\bar{P}_{l_{j}}^{l_{j-1}}(\theta_{j})\,{}_{j}\bar{P}_{l^{\prime}_{j}}^{l^{\prime}_{j-1}}(\theta_{j})\,d\Omega_{n-1}\,. (44)

The θ1\theta_{1} integral is

02πei(l1l1)θ1𝑑θ1=2πδl1l1.\displaystyle\int^{2\pi}_{0}e^{-i(l_{1}-l^{\prime}_{1})\theta_{1}}d\theta_{1}=2\pi\delta_{l_{1}l^{\prime}_{1}}\,. (45)

The θj\theta_{j} integral, for 2jn12\leq j\leq n-1, is

0πP¯ljlj1j(θj)P¯ljlj1j(θj)sinj1θjdθj.\displaystyle\int^{\pi}_{0}{}_{j}\bar{P}_{l_{j}}^{l_{j-1}}(\theta_{j})\,{}_{j}\bar{P}_{l^{\prime}_{j}}^{l^{\prime}_{j-1}}(\theta_{j})\,\sin^{j-1}\theta_{j}\,d\theta_{j}\,. (46)

We first tackle the case in which we have at least one ljljl_{j}\neq l_{j}^{\prime}. Let jj be the smallest integer for which ljljl_{j}\neq l^{\prime}_{j} occurs. If j=1j=1, then by (45), (A) integrates to 0. If 2jn12\leq j\leq n-1, because jj is the smallest integer such that ljljl_{j}\neq l^{\prime}_{j} occurs, we have lj1=lj1l_{j-1}=l^{\prime}_{j-1}. We consider (46) in the special case lj1=lj1l_{j-1}=l^{\prime}_{j-1}, which is shown in (Higuchi, 1987) to equal to

0πP¯ljlj1j(θj)P¯ljlj1j(θj)sinj1θjdθj=δljlj.\displaystyle\int^{\pi}_{0}{}_{j}\bar{P}_{l_{j}}^{l_{j-1}}(\theta_{j})\,{}_{j}\bar{P}_{l^{\prime}_{j}}^{l_{j-1}}(\theta_{j})\,\sin^{j-1}\theta_{j}\,d\theta_{j}=\delta_{l_{j}l^{\prime}_{j}}\,. (47)

Thus (46) again integrates to 0, and orthogonality in (A) is proven. Next, by imposing normality in (A), we determine the normalization constant AA. Consider the special case where lj=ljl_{j}=l^{\prime}_{j} for all 1jn11\leq j\leq n-1. By (45) and (47), (A) integrates to |A|22π|A|^{2}2\pi. We want this to normalize to 1, so

A=12π.A=\frac{1}{\sqrt{2\pi}}\,. (48)

Thus, the nn-dimensional spherical harmonics take the form

Yl1,,ln1(θ1,,θn1)=12πeil1θ1j=2n1P¯ljlj1j(θj),\displaystyle Y_{l_{1},...,l_{n-1}}(\theta_{1},...,\theta_{n-1})=\frac{1}{\sqrt{2\pi}}e^{il_{1}\theta_{1}}\prod^{n-1}_{j=2}{}_{j}\bar{P}_{l_{j}}^{l_{j-1}}(\theta_{j})\,, (49)

and satisfy the orthonormality condition (A).

Next we prove another useful identity

l1,,ln1Yl1,,ln1(θ1,,θn1)Yl1,,ln1(θ1,,θn1)\displaystyle\sum_{\begin{subarray}{c}l_{1},...,\\ l_{n-1}\end{subarray}}Y_{l_{1},...,l_{n-1}}(\theta_{1},...,\theta_{n-1})Y^{*}_{l_{1},...,l_{n-1}}(\theta^{\prime}_{1},...,\theta^{\prime}_{n-1})
=1sinn2θn1sinn3θn2sinθ2δ(θ1θ1)\displaystyle=\frac{1}{\sin^{n-2}\theta_{n-1}\sin^{n-3}\theta_{n-2}...\sin\theta_{2}}\delta(\theta_{1}-\theta^{\prime}_{1})...
δ(θn1θn1).\displaystyle\cdot\delta(\theta_{n-1}-\theta^{\prime}_{n-1})\,. (50)

This is a result of the completeness property of the spherical harmonics. By completeness, we can express any function V(θ1,,θn1)V(\theta_{1},...,\theta_{n-1}) on Sn1S^{n-1} as a linear combination of Yl1,,ln1(θ1,,θn1)Y_{l_{1},...,l_{n-1}}(\theta_{1},...,\theta_{n-1}):

V(θ1,,θn1)=l1,,ln1Vl1,,ln1Yl1,,ln1(θ1,,θn1).\displaystyle V(\theta_{1},...,\theta_{n-1})=\sum_{\begin{subarray}{c}l_{1},...,\\ l_{n-1}\end{subarray}}V_{l_{1},...,l_{n-1}}Y_{l_{1},...,l_{n-1}}(\theta_{1},...,\theta_{n-1})\,. (51)

Using the orthonormality condition (A), we can determine the coefficients Vl1,,ln1V_{l_{1},...,l_{n-1}} by

Vl1,,ln1=\displaystyle V_{l_{1},...,l_{n-1}}=\int Yl1,,ln1(θ1,,θn1)V(θ1,,θn1)\displaystyle Y^{*}_{l_{1},...,l_{n-1}}(\theta_{1},...,\theta_{n-1})V(\theta_{1},...,\theta_{n-1})
dΩn1.\displaystyle d\Omega_{n-1}\,. (52)

Substitute (A) into (51), and also by noting that

dΩn1=sinn2θn1sinn3θn2sinθ2dθ1dθ2dθn1,\displaystyle d\Omega_{n-1}=\sin^{n-2}\theta_{n-1}\sin^{n-3}\theta_{n-2}...\sin\theta_{2}d\theta_{1}d\theta_{2}...d\theta_{n-1}\,, (53)

we get

V(θ1,,θn1)\displaystyle V(\theta_{1},...,\theta_{n-1})
=\displaystyle= l1,,ln1(Yl1,,ln1(θ1,,θn1)V(θ1,,θn1)𝑑Ωn1)\displaystyle\sum_{\begin{subarray}{c}l_{1},...,\\ l_{n-1}\end{subarray}}\left(\int Y^{*}_{l_{1},...,l_{n-1}}(\theta^{\prime}_{1},...,\theta^{\prime}_{n-1})V(\theta^{\prime}_{1},...,\theta^{\prime}_{n-1})d\Omega^{\prime}_{n-1}\right)
Yl1,,ln1(θ1,,θn1)\displaystyle Y_{l_{1},...,l_{n-1}}(\theta_{1},...,\theta_{n-1})
=\displaystyle= {l1,,ln1Yl1,,ln1(θ1,,θn1)Yl1,,ln1(θ1,,θn1)}\displaystyle\int\left\{\sum_{\begin{subarray}{c}l_{1},...,\\ l_{n-1}\end{subarray}}Y^{*}_{l_{1},...,l_{n-1}}(\theta^{\prime}_{1},...,\theta^{\prime}_{n-1})Y_{l_{1},...,l_{n-1}}(\theta_{1},...,\theta_{n-1})\right\}
V(θ1,,θn1)sinn2θn1sinn3θn2sinθ2\displaystyle V(\theta^{\prime}_{1},...,\theta^{\prime}_{n-1})\sin^{n-2}\theta^{\prime}_{n-1}\sin^{n-3}\theta^{\prime}_{n-2}...\sin\theta^{\prime}_{2}
dθ1dθn1.\displaystyle d\theta^{\prime}_{1}...d\theta^{\prime}_{n-1}\,. (54)

In order for this equation to hold, we must require the expression in the braces to equal to the expression on the right hand side of (A), thus proving (A).

Lastly, we shall prove

Yl1,,ln1(θ1,,θn1)𝑑Ωn1\displaystyle\int Y_{l_{1},...,l_{n-1}}(\theta_{1},...,\theta_{n-1})d\Omega_{n-1}
=\displaystyle= (2πn2Γ(n2))12δl10δl20δln10.\displaystyle\left(\frac{2\pi^{\frac{n}{2}}}{\Gamma(\frac{n}{2})}\right)^{\frac{1}{2}}\delta_{l_{1}0}\delta_{l_{2}0}...\delta_{l_{n-1}0}\,. (55)

From (49), we know that

Y0,,0(θ1,,θn1)=(Γ(n2)2πn2)12Y_{0,...,0}(\theta_{1},...,\theta_{n-1})=\left(\frac{\Gamma(\frac{n}{2})}{2\pi^{\frac{n}{2}}}\right)^{\frac{1}{2}} (56)

is a constant, where Γ\Gamma is the Gamma function. By the orthonormality condition (A), it is the only nn-dimensional spherical harmonic Yl1,,ln1(θ1,,θn1)Y_{l_{1},...,l_{n-1}}(\theta_{1},...,\theta_{n-1}) that is a constant. Thus

Y0,,0(θ1,,θn1)𝑑Ωn1\displaystyle\int Y_{0,...,0}(\theta_{1},...,\theta_{n-1})\,d\Omega_{n-1}
=\displaystyle= Y0,,0(θ1,,θn1)𝑑Ωn1=(Γ(n2)2πn2)122πn2Γ(n2)\displaystyle Y_{0,...,0}(\theta_{1},...,\theta_{n-1})\int d\Omega_{n-1}=\left(\frac{\Gamma(\frac{n}{2})}{2\pi^{\frac{n}{2}}}\right)^{\frac{1}{2}}\frac{2\pi^{\frac{n}{2}}}{\Gamma(\frac{n}{2})}
=\displaystyle= (2πn2Γ(n2))12.\displaystyle\left(\frac{2\pi^{\frac{n}{2}}}{\Gamma(\frac{n}{2})}\right)^{\frac{1}{2}}\,. (57)

Next, consider the integral

Y0,,0(θ1,,θn1)Yl1,,ln1(θ1,,θn1)𝑑Ωn1\displaystyle\int Y_{0,...,0}(\theta_{1},...,\theta_{n-1})Y_{l_{1},...,l_{n-1}}(\theta_{1},...,\theta_{n-1})d\Omega_{n-1} (58)

where Yl1,,ln1(θ1,,θn1)Y_{l_{1},...,l_{n-1}}(\theta_{1},...,\theta_{n-1}) has at least one lj0l_{j}\neq 0. By the orthogonality condition (A), this integral equals 0. Thus we have

0\displaystyle 0 =Y0,,0(θ1,,θn1)Yl1,,ln1(θ1,,θn1)𝑑Ωn1\displaystyle=\int Y_{0,...,0}(\theta_{1},...,\theta_{n-1})Y_{l_{1},...,l_{n-1}}(\theta_{1},...,\theta_{n-1})d\Omega_{n-1}
=Y0,,0(θ1,,θn1){Yl1,,ln1(θ1,,θn1)𝑑Ωn1}\displaystyle=Y_{0,...,0}(\theta_{1},...,\theta_{n-1})\left\{\int Y_{l_{1},...,l_{n-1}}(\theta_{1},...,\theta_{n-1})d\Omega_{n-1}\right\} (59)

Since Y0,,0(θ1,,θn1)Y_{0,...,0}(\theta_{1},...,\theta_{n-1}) is a non-zero constant, the integral in the braces must equal to 0, completing the proof of (A).

Appendix B Additional details on the derivations of the full-energy mapping CFT approach in the isotropic limit

In this section we provide some additional details on the derivations of the full-energy mapping CFT (FEMCFT) approach in the isotropic limit.

B.1 Derivations of equations (7) and (8) of the main text

First we note that cα,i,μ(p)c_{\alpha,i,\mu}(\vec{p}) in the Hamiltonians (2) and (4) of the main text are fermionic fields, so they satisfy the usual anticommutation relation

{cα,i,μ(p),cβ,j,ν(p)}=δαβδijδμνδ(n)(pp)\{c^{\dagger}_{\alpha,i,\mu}(\vec{p}),\,c_{\beta,j,\nu}(\vec{p}^{\prime})\}=\delta_{\alpha\beta}\delta_{ij}\delta_{\mu\nu}\delta^{(n)}(\vec{p}-\vec{p}^{\prime}) (60)

in nn dimensions. We then expand cα,i,μ(p)c_{\alpha,i,\mu}(\vec{p}) as a linear combination of the nn-dimensional spherical harmonics, which form a complete set of orthonormal functions on the (n1)(n-1)-sphere Sn1S^{n-1}. The expansion is given by equation (6) of the main text:

cα,i,μ(p)=1pn12\displaystyle c_{\alpha,i,\mu}(\vec{p})=\frac{1}{p^{\frac{n-1}{2}}}\cdot
l1,,ln1Yl1,,ln1(θ1,,θn1)cl1,,ln1,α,i,μ(p).\displaystyle\sum_{l_{1},...,l_{n-1}}Y_{l_{1},...,l_{n-1}}(\theta_{1},...,\theta_{n-1})c_{l_{1},...,l_{n-1},\alpha,i,\mu}(p)\,. (61)

where Yl1,,ln1(θ1,,θn1)Y_{l_{1},...,l_{n-1}}(\theta_{1},...,\theta_{n-1}) are the spherical harmonics in nn dimensions discussed in the previous section. Note that the coefficients in the expansion (B.1) are only functions of p|p|p\equiv|\vec{p}|, and do not depend on any of the angular coordinates θ1,,θn1\theta_{1},...,\theta_{n-1}. Also, a factor of 1/pn121/p^{\frac{n-1}{2}} has been factored out from each coefficient in the expansion, so that the remaining part of the coefficient, the fields cl1,,ln1,α,i,μ(p)c_{l_{1},...,l_{n-1},\alpha,i,\mu}(p), satisfy the proper fermionic anticommutation relation

{cl1,,ln1,α,i,μ(p),cl1,,ln1,β,j,ν(p)}\displaystyle\{c^{\dagger}_{l_{1},...,l_{n-1},\alpha,i,\mu}(p),\,c_{l^{\prime}_{1},...,l^{\prime}_{n-1},\beta,j,\nu}(p^{\prime})\}
=\displaystyle= δl1l1δln1ln1δαβδijδμνδ(pp).\displaystyle\delta_{l_{1}l^{\prime}_{1}}...\delta_{l_{n-1}l^{\prime}_{n-1}}\delta_{\alpha\beta}\delta_{ij}\delta_{\mu\nu}\delta(p-p^{\prime})\,. (62)

Also, one can show that cl1,,ln1,α,i,μ(p)c_{l_{1},...,l_{n-1},\alpha,i,\mu}(p) are related to cα,i,μ(p)c_{\alpha,i,\mu}(\vec{p}) by

cl1,,ln1,α,i,μ(p)=\displaystyle c_{l_{1},...,l_{n-1},\alpha,i,\mu}(p)=
pn12𝑑Ωn1Yl1,,ln1(θ1,,θn1)cα,i,μ(p),\displaystyle p^{\frac{n-1}{2}}\int d\Omega_{n-1}Y^{*}_{l_{1},...,l_{n-1}}(\theta_{1},...,\theta_{n-1})c_{\alpha,i,\mu}(\vec{p})\,, (63)

We shall write the Hamiltonian in terms of these new fermionic fields cl1,,ln1,α,i,μ(p)c_{l_{1},...,l_{n-1},\alpha,i,\mu}(p). By substituting (B.1) into HIH_{I} given by equation (4) of the main text, and by using dnp=pn1dpdΩn1d^{n}\vec{p}=p^{n-1}dpd\Omega_{n-1} and (A), HIH_{I} can be written into the form of equation (7) of the main text:

HI=λSΓ(n2)2nπn2μ,ν\displaystyle H_{I}=\frac{\lambda\vec{S}}{\Gamma(\frac{n}{2})2^{n}\pi^{\frac{n}{2}}}\cdot\sum_{\mu,\nu}\int dpdppn12pn12\displaystyle dpdp^{\prime}p^{\frac{n-1}{2}}p^{\prime\frac{n-1}{2}}
c0,,0,α,i,μ(p)σαβc0,,0,β,i,ν(p).\displaystyle c^{\dagger}_{0,...,0,\alpha,i,\mu}(p)\vec{\sigma}_{\alpha\beta}c_{0,...,0,\beta,i,\nu}(p^{\prime})\,. (64)

This shows that the only partial waves that couple to the impurity S\vec{S} are those with (l1,,ln1)=(0,,0)(l_{1},...,l_{n-1})=(0,...,0).

Similarly, substitute (B.1) into H0H_{0} given in equation (2) of the main text, and use the orthonormality condition (A), we get

H0=l1,,ln1𝑑pcl1,,ln1,α,i,μ(p)cl1,,ln1,α,i,μ(p)ϵμ(p).\displaystyle H_{0}=\sum_{\begin{subarray}{c}l_{1},...,\\ l_{n-1}\end{subarray}}\int dp\,c^{\dagger}_{l_{1},...,l_{n-1},\alpha,i,\mu}(p)c_{l_{1},...,l_{n-1},\alpha,i,\mu}(p)\epsilon_{\mu}(p)\,. (65)

Equation (B.1) shows that the only partial waves that couple to the impurity S\vec{S} in HIH_{I} are those with (l1,,ln1)=(0,,0)(l_{1},...,l_{n-1})=(0,...,0), and equation (65) shows that in the bath, there is only coupling among partial waves of the same quantum numbers. Thus no partial waves with {l1,,ln1}{0,,0}\{l_{1},...,l_{n-1}\}\neq\{0,...,0\} will couple to the impurity, either directly or indirectly. As a result, we only need to consider partial waves with {l1,,ln1}={0,,0}\{l_{1},...,l_{n-1}\}=\{0,...,0\} in H0H_{0}. Thus H0H_{0} reduces to equation (8) of the main text:

H0=𝑑pc0,,0,α,i,μ(p)c0,,0,α,i,μ(p)ϵμ(p).\displaystyle H_{0}=\int dp\,c^{\dagger}_{0,...,0,\alpha,i,\mu}(p)c_{0,...,0,\alpha,i,\mu}(p)\epsilon_{\mu}(p)\,. (66)

B.2 Derivations of the anticommutation relations for the c(ϵ)c(\epsilon) and the Φ\Phi fields

In this section, we show that the field cl1,,ln1,α,i,μ(ϵμ)c_{l_{1},...,l_{n-1},\alpha,i,\mu}(\epsilon_{\mu}) defined in equation (9) of the main text and the field Φ0,,0,α,i(ϵ)\Phi_{0,...,0,\alpha,i}(\epsilon) defined in equation (10) of the main text are fermionic fields satisfying the proper anticommutation relations. Due to the delta function identity

δ(ϵμϵμ)=δ(pp)|dϵμ(p)dp|p=p|,\delta(\epsilon_{\mu}-\epsilon^{\prime}_{\mu})=\frac{\delta(p-p^{\prime})}{\left|\frac{d\epsilon_{\mu}(p)}{dp}|_{p=p^{\prime}}\right|}\,, (67)

the field cl1,,ln1,α,i,μ(ϵμ)c_{l_{1},...,l_{n-1},\alpha,i,\mu}(\epsilon_{\mu}) obeys the anticommutation relation

{cl1,,ln1,α,i,μ(ϵμ),cl1,,ln1,β,j,ν(ϵν)}\displaystyle\{c^{\dagger}_{l_{1},...,l_{n-1},\alpha,i,\mu}(\epsilon_{\mu}),\,c_{l^{\prime}_{1},...,l^{\prime}_{n-1},\beta,j,\nu}(\epsilon^{\prime}_{\nu})\}
=\displaystyle= δl1l1δln1ln1δαβδijδμνδ(ϵμϵν).\displaystyle\delta_{l_{1}l^{\prime}_{1}}...\delta_{l_{n-1}l^{\prime}_{n-1}}\delta_{\alpha\beta}\delta_{ij}\delta_{\mu\nu}\delta(\epsilon_{\mu}-\epsilon^{\prime}_{\nu})\,. (68)

As a result, the composite field Φ0,,0,α,i(ϵ)\Phi_{0,...,0,\alpha,i}(\epsilon) satisfies the anticommutation relation

{Φ0,,0,α,i(ϵ),Φ0,,0,β,j(ϵ)}=δαβδijδ(ϵϵ)\displaystyle\{\Phi^{\dagger}_{0,...,0,\alpha,i}(\epsilon),\Phi_{0,...,0,\beta,j}(\epsilon^{\prime})\}=\delta_{\alpha\beta}\delta_{ij}\delta(\epsilon-\epsilon^{\prime}) (69)

as well.

B.3 Writing the Hamiltonian in terms of the composite field Φ\Phi

In this section, we show the detailed derivations of equations (11) and (12) of the main text, i.e. how to write the Hamiltonian in terms of the composite fermion field Φ0,,0,α,i(ϵ)\Phi_{0,...,0,\alpha,i}(\epsilon). We tackle equation (11) of the main text first. By substituting equation (9) of the main text into equation (8) of the main text (i.e. equation (66) in this appendix), and using equation (5) of the main text (WLOG assume a1+a_{1+} is positive), we get

H0=\displaystyle H_{0}= 0𝑑ϵc0,,0,α,i,+(ϵ)c0,,0,α,i,+(ϵ)ϵ\displaystyle\int_{0}^{\infty}d\epsilon\,c^{\dagger}_{0,...,0,\alpha,i,+}(\epsilon)c_{0,...,0,\alpha,i,+}(\epsilon)\epsilon
+\displaystyle+ 0𝑑ϵc0,,0,α,i,(ϵ)c0,,0,α,i,(ϵ)ϵ,\displaystyle\int_{-\infty}^{0}d\epsilon\,c^{\dagger}_{0,...,0,\alpha,i,-}(\epsilon)c_{0,...,0,\alpha,i,-}(\epsilon)\epsilon\,, (70)

where we have relabelled the dummy variables ϵ+ϵ\epsilon_{+}\rightarrow\epsilon and ϵϵ\epsilon_{-}\rightarrow\epsilon in the first term and in the second term respectively. Then by using the definition of Φ0,,0,α,i(ϵ)\Phi_{0,...,0,\alpha,i}(\epsilon) given by equation (10) of the main text, H0H_{0} becomes

H0=𝑑ϵΦ0,,0,α,i(ϵ)Φ0,,0,α,i(ϵ)ϵ,\displaystyle H_{0}=\int_{-\infty}^{\infty}d\epsilon\,\Phi^{\dagger}_{0,...,0,\alpha,i}(\epsilon)\Phi_{0,...,0,\alpha,i}(\epsilon)\epsilon\,, (71)

which is equation (11) of the main text.

As for HIH_{I}, by substituting equation (9) of the main text into equation (7) of the main text (i.e. equation (B.1) in this appendix), we get

HI=λSΓ(n2)2nπn2μ,ν𝑑ϵμ𝑑ϵνpn12pn12\displaystyle H_{I}=\frac{\lambda\vec{S}}{\Gamma(\frac{n}{2})2^{n}\pi^{\frac{n}{2}}}\cdot\sum_{\mu,\nu}\int d\epsilon_{\mu}\int d\epsilon^{\prime}_{\nu}p^{\frac{n-1}{2}}p^{\prime\frac{n-1}{2}}
|dϵμ(p)dp|12|dϵν(p)dp|12c0,,0,α,i,μ(ϵμ)σαβc0,,0,β,i,ν(ϵν),\displaystyle\left|\frac{d\epsilon_{\mu}(p)}{dp}\right|^{-\frac{1}{2}}\left|\frac{d\epsilon^{\prime}_{\nu}(p^{\prime})}{dp^{\prime}}\right|^{-\frac{1}{2}}c^{\dagger}_{0,...,0,\alpha,i,\mu}(\epsilon_{\mu})\vec{\sigma}_{\alpha\beta}c_{0,...,0,\beta,i,\nu}(\epsilon^{\prime}_{\nu})\,, (72)

where

𝑑ϵμ\displaystyle\int d\epsilon_{\mu} {0𝑑ϵ+if μ=+0𝑑ϵif μ=,\displaystyle\equiv\begin{cases}\int_{0}^{\infty}d\epsilon_{+}&\text{if }\mu=+\\ \int_{-\infty}^{0}d\epsilon_{-}&\text{if }\mu=-\end{cases}\,,
𝑑ϵν\displaystyle\int d\epsilon^{\prime}_{\nu} {0𝑑ϵ+if ν=+0𝑑ϵif ν=.\displaystyle\equiv\begin{cases}\int_{0}^{\infty}d\epsilon^{\prime}_{+}&\text{if }\nu=+\\ \int_{-\infty}^{0}d\epsilon^{\prime}_{-}&\text{if }\nu=-\end{cases}\,. (73)

Using equation (5) of the main text, |dϵμ(p)dp|12\left|\frac{d\epsilon_{\mu}(p)}{dp}\right|^{-\frac{1}{2}} and pn12p^{\frac{n-1}{2}} cancel off each other, and similarly |dϵν(p)dp|12\left|\frac{d\epsilon^{\prime}_{\nu}(p^{\prime})}{dp^{\prime}}\right|^{-\frac{1}{2}} and pn12p^{\prime\frac{n-1}{2}} cancel off each other. We get

HI=\displaystyle H_{I}= λSanΓ(n2)2nπn2\displaystyle\frac{\lambda\vec{S}}{an\Gamma(\frac{n}{2})2^{n}\pi^{\frac{n}{2}}}
μ,νdϵμdϵνc0,,0,α,i,μ(ϵμ)σαβc0,,0,β,i,ν(ϵν),\displaystyle\cdot\sum_{\mu,\nu}\int d\epsilon_{\mu}\int d\epsilon^{\prime}_{\nu}c^{\dagger}_{0,...,0,\alpha,i,\mu}(\epsilon_{\mu})\vec{\sigma}_{\alpha\beta}c_{0,...,0,\beta,i,\nu}(\epsilon^{\prime}_{\nu})\,, (74)

where a|a1μ|=|a1ν|a\equiv|a_{1\mu}|=|a_{1\nu}|. Writing out the sum over μ,ν\mu,\nu explicitly and relabelling the dummy variables ϵ+ϵ\epsilon_{+}\rightarrow\epsilon and ϵϵ\epsilon_{-}\rightarrow\epsilon, we get

HI=λSanΓ(n2)2nπn2\displaystyle H_{I}=\frac{\lambda\vec{S}}{an\Gamma(\frac{n}{2})2^{n}\pi^{\frac{n}{2}}}
{0dϵ0dϵc0,,0,α,i,+(ϵ)σαβc0,,0,β,i,+(ϵ)\displaystyle\cdot\left\{\int_{0}^{\infty}d\epsilon\int_{0}^{\infty}d\epsilon^{\prime}c^{\dagger}_{0,...,0,\alpha,i,+}(\epsilon)\vec{\sigma}_{\alpha\beta}c_{0,...,0,\beta,i,+}(\epsilon^{\prime})\right.
+0𝑑ϵ0𝑑ϵc0,,0,α,i,+(ϵ)σαβc0,,0,β,i,(ϵ)\displaystyle+\int_{0}^{\infty}d\epsilon\int_{-\infty}^{0}d\epsilon^{\prime}c^{\dagger}_{0,...,0,\alpha,i,+}(\epsilon)\vec{\sigma}_{\alpha\beta}c_{0,...,0,\beta,i,-}(\epsilon^{\prime})
+0𝑑ϵ0𝑑ϵc0,,0,α,i,(ϵ)σαβc0,,0,β,i,+(ϵ)\displaystyle+\int_{-\infty}^{0}d\epsilon\int_{0}^{\infty}d\epsilon^{\prime}c^{\dagger}_{0,...,0,\alpha,i,-}(\epsilon)\vec{\sigma}_{\alpha\beta}c_{0,...,0,\beta,i,+}(\epsilon^{\prime})
+0dϵ0dϵc0,,0,α,i,(ϵ)σαβc0,,0,β,i,(ϵ)}\displaystyle+\left.\int_{-\infty}^{0}d\epsilon\int_{-\infty}^{0}d\epsilon^{\prime}c^{\dagger}_{0,...,0,\alpha,i,-}(\epsilon)\vec{\sigma}_{\alpha\beta}c_{0,...,0,\beta,i,-}(\epsilon^{\prime})\right\}
=λSanΓ(n2)2nπn2\displaystyle=\frac{\lambda\vec{S}}{an\Gamma(\frac{n}{2})2^{n}\pi^{\frac{n}{2}}}
dϵdϵΦ0,,0,α,i(ϵ)σαβΦ0,,0,β,i(ϵ).\displaystyle\cdot\int_{-\infty}^{\infty}d\epsilon\int_{-\infty}^{\infty}d\epsilon^{\prime}\Phi^{\dagger}_{0,...,0,\alpha,i}(\epsilon)\vec{\sigma}_{\alpha\beta}\Phi_{0,...,0,\beta,i}(\epsilon^{\prime})\,. (75)

which is equation (12) of the main text.

B.4 Derivation of equation (16) and (17) of the main text

We first derive equation (17) of the main text. Let τ\tau be the imaginary time τit\tau\equiv it, and define the complex plane \mathbb{C} to which we map our KHOF model as equation (14) of the main text,

={zτ+ir}.\mathbb{C}=\{z\equiv\tau+ir\}\,. (76)

View the Ψ/αi(r)\Psi_{\leftarrow/\rightarrow\alpha i}(r) fields defined in equation (13) of the main text in the Heisenberg picture, so that they now have time-dependence, and thus live on \mathbb{C}. Using equation (13) of the main text, we can see that HIH_{I} given by equation (12) of the main text can be written in the form of equation (17) of the main text.

We now derive equation (16) of the main text. Using equation (13) of the main text, H0H_{0} given by equation (11) of the main text can be written as

H0=12π0dr\displaystyle H_{0}=\frac{1}{2\pi}\int^{\infty}_{0}dr\,\cdot
(Ψαi(r)irΨαi(r)Ψαi(r)irΨαi(r)).\displaystyle\left(\Psi^{\dagger}_{\leftarrow\alpha i}(r)i\frac{\partial}{\partial r}\Psi_{\leftarrow\alpha i}(r)-\Psi^{\dagger}_{\rightarrow\alpha i}(r)i\frac{\partial}{\partial r}\Psi_{\rightarrow\alpha i}(r)\right)\,. (77)

We note that 0r0\leq r\leq\infty in (B.4). One can also show that Ψ/αi(r)\Psi_{\leftarrow/\rightarrow\alpha i}(r) satisfy the anticommutation relations

{ΨXαi(r),ΨXβj(r)}=2πδXXδαβδijδ(rr).\displaystyle\{\Psi^{\dagger}_{X\alpha i}(r),\Psi_{X^{\prime}\beta j}(r^{\prime})\}=2\pi\delta_{XX^{\prime}}\delta_{\alpha\beta}\delta_{ij}\delta(r-r^{\prime})\,. (78)

Now equation (17) of the main text tells us that HIH_{I} is only confined to the horizontal axis r=0r=0. At r0r\neq 0, H=H0H=H_{0} and we have the Heisenberg equations of motion

itΨXαi(τ,r)=[ΨXαi(τ,r),H0]\displaystyle i\frac{\partial}{\partial t}\Psi_{X\alpha i}(\tau,r)=[\Psi_{X\alpha i}(\tau,r),H_{0}] (79)

for each X{,}X\in\{\leftarrow,\rightarrow\}. By substituting H0H_{0} in the form of (B.4) into the Heisenberg equation of motion (79), and applying the anticommutation relation (78) and the identity [A,BC]={A,B}CB{A,C}[A,BC]=\{A,B\}C-B\{A,C\} for three arbitrary operators A,BA,B and CC, the Heisenberg equation of motion reduces to

z¯Ψαi(τ,r)=0,zΨαi(τ,r)=0.\displaystyle\partial_{\bar{z}}\Psi_{\leftarrow\alpha i}(\tau,r)=0\,,\qquad\,\partial_{z}\Psi_{\rightarrow\alpha i}(\tau,r)=0\,. (80)

These are exactly the Cauchy-Riemann equations for holomorphic functions and antiholomorphic functions respectively, which imply that Ψαi(t,r)\Psi_{\leftarrow\alpha i}(t,r) is a holomorphic function, and Ψαi(t,r)\Psi_{\rightarrow\alpha i}(t,r) is an antiholomorphic function, on the upper half plane +={zτ+ir|r0}\mathbb{C}_{+}=\{z\equiv\tau+ir|r\geq 0\} (since in (B.4), rr is non-negative, Ψ/αi(τ,r)\Psi_{\leftarrow/\rightarrow\alpha i}(\tau,r) live on the upper half plane). Equivalently, Ψαi(τ,r)\Psi_{\rightarrow\alpha i}(\tau,r) is a holomorphic function on the lower complex plane ={z¯τir|r0}\mathbb{C}_{-}=\{\bar{z}\equiv\tau-ir|r\geq 0\}. Also, by the definition of Ψ/αi(r)\Psi_{\leftarrow/\rightarrow\alpha i}(r) given by equation (13) of the main text, Ψαi(τ,0)=Ψαi(τ,0)\Psi_{\leftarrow\alpha i}(\tau,0)=\Psi_{\rightarrow\alpha i}(\tau,0), i.e. the two holomorphic functions Ψαi(τ,r)\Psi_{\leftarrow\alpha i}(\tau,r) on the upper complex plane and Ψαi(τ,r)\Psi_{\rightarrow\alpha i}(\tau,r) on the lower complex plane agree on the horizontal τ\tau-axis r=0r=0. Thus they are analytic continuations of each other to the entire complex plane. This fact allow us to eliminate Ψαi(τ,r)\Psi_{\rightarrow\alpha i}(\tau,r) in terms of Ψαi(τ,r)\Psi_{\leftarrow\alpha i}(\tau,r) because the former is simply the analytic continuation of the latter into the lower complex plane:

Ψαi(τ,r)=Ψαi(τ,r).\Psi_{\rightarrow\alpha i}(\tau,r)=\Psi_{\leftarrow\alpha i}(\tau,-r)\,. (81)

Thus (B.4) becomes

H0=12π𝑑r(Ψαi(τ,r)irΨαi(τ,r)),\displaystyle H_{0}=\frac{1}{2\pi}\int^{\infty}_{-\infty}dr\left(\Psi^{\dagger}_{\leftarrow\alpha i}(\tau,r)i\frac{\partial}{\partial r}\Psi_{\leftarrow\alpha i}(\tau,r)\right)\,, (82)

which is equation (16) of the main text.

Appendix C Additional details on the applications of the FEMCFT method in anisotropic materials

In this section, we provide additional details on the applications of our FEMCFT method to anisotropic KHOF models. When we describe our general method, we shall keep the dimension nn of our KHOF model to be arbitrary. From time to time, we illustrate our procedure with the cubic fermion system in three dimensions, whose dispersion is given by equation (21) of the main text. At such points we shall let n=3n=3.

We have considered the FEMCFT approach in the isotropic limit, in which the dispersion relation of our KHOF system is isotropic. We now consider the changes that a KHOF system with an anisotropic dispersion relation brings about. In terms of the Hamiltonian H=H0+HIH=H_{0}+H_{I}, HIH_{I} remains unchanged, since the dispersion relation does not enter the expression of HIH_{I}. However, now in H0H_{0}, we not only have to expand cα,i,μ(p)c_{\alpha,i,\mu}(\vec{p}) as a linear combination of the spherical harmonics according to (B.1), but also need to do so for ϵμ(p)\epsilon_{\mu}(\vec{p}) as well:

ϵμ(p)=l1,,ln1Yl1,,ln1(θ1,,θn1)ϵl1,,ln1,μ(p).\epsilon_{\mu}(\vec{p})=\sum_{l_{1},...,l_{n-1}}Y_{l_{1},...,l_{n-1}}(\theta_{1},...,\theta_{n-1})\epsilon_{l_{1},...,l_{n-1},\mu}(p)\,. (83)

Substitute (B.1), (83) into equation (2) of the main text, we get

H0=L,L,L′′VL,L,L′′𝑑pcL,α,i,μ(p)cL,α,i,μ(p)ϵL′′,μ(p),\displaystyle H_{0}=\sum_{L,L^{\prime},L^{\prime\prime}}V_{L,L^{\prime},L^{\prime\prime}}\int dp\,c^{\dagger}_{L,\alpha,i,\mu}(p)c_{L^{\prime},\alpha,i,\mu}(p)\epsilon_{L^{\prime\prime},\mu}(p)\,, (84)

where for brevity we used the multi-index notation L(l1,,ln1),L(l1,,ln1),L′′(l1′′,,ln1′′)L\equiv(l_{1},...,l_{n-1}),L^{\prime}\equiv(l^{\prime}_{1},...,l^{\prime}_{n-1}),L^{\prime\prime}\equiv(l^{\prime\prime}_{1},...,l^{\prime\prime}_{n-1}). The coupling strengths VL,L,L′′V_{L,L^{\prime},L^{\prime\prime}} are given by the hopping integral

VL,L,L′′=\displaystyle V_{L,L^{\prime},L^{\prime\prime}}=
𝑑Ωn1YL(θ1,,θn1)YL(θ1,,θn1)YL′′(θ1,,θn1).\displaystyle\int d\Omega_{n-1}Y^{*}_{L}(\theta_{1},...,\theta_{n-1})Y_{L^{\prime}}(\theta_{1},...,\theta_{n-1})Y_{L^{\prime\prime}}(\theta_{1},...,\theta_{n-1})\,. (85)

We note that VL,L,L′′V_{L,L^{\prime},L^{\prime\prime}} depends on the dispersion relation ϵμ(p)\epsilon_{\mu}(\vec{p}) of the system, because different ϵμ(p)\epsilon_{\mu}(\vec{p}) observes different expansions in terms of spherical harmonics (83), resulting in different sets of YL′′(θ1,,θn1)Y_{L^{\prime\prime}}(\theta_{1},...,\theta_{n-1}) appearing in (C). As a concrete example, we calculate VL,L,L′′V_{L,L^{\prime},L^{\prime\prime}} for the dispersion relation given by equation (21) of the main text in 3 dimensions,

ϵμ(p)=a1μp3+a2μp1,\epsilon_{\mu}(\vec{p})=a_{1\mu}p^{3}+a_{2\mu}p_{1}\,, (86)

where a1μ,a2μa_{1\mu},a_{2\mu} are real constants satisfying

a1+=a1anda2+=a2,\displaystyle a_{1+}=-a_{1-}\qquad\text{and}\qquad a_{2+}=-a_{2-}\,, (87)

so that

ϵ+(p)=ϵ(p).\epsilon_{+}(\vec{p})=-\epsilon_{-}(\vec{p})\,. (88)

WLOG align p1p_{1} in the zz-direction, so p1=pz=pcosθp_{1}=p_{z}=p\cos\theta. Thus (86) reads

ϵμ(p)=a1μp3+a2μpcosθ,\epsilon_{\mu}(\vec{p})=a_{1\mu}p^{3}+a_{2\mu}p\cos\theta\,, (89)

whose expansion (83) in terms of spherical harmonics in 3 dimensions is

ϵμ(p)=a1μp34πY00(θ,ϕ)+a2μp4π/3Y10(θ,ϕ).\epsilon_{\mu}(\vec{p})=a_{1\mu}p^{3}\sqrt{4\pi}Y^{0}_{0}(\theta,\phi)+a_{2\mu}p\sqrt{4\pi/3}Y^{0}_{1}(\theta,\phi)\,. (90)

This shows that for 3-dimensional systems with dispersion relation (86), the only non-zero spherical harmonics in the expansion (83) are Y00Y^{0}_{0} and Y10Y^{0}_{1}, which enter (C) as YL′′Y_{L^{\prime\prime}}. Thus the only non-zero VL,L,L′′V_{L,L^{\prime},L^{\prime\prime}} are VL,L,(00)V_{L,L^{\prime},(00)} and VL,L,(01)V_{L,L^{\prime},(01)}. They are given by

VL,L,(00)=𝑑Ω2Ylm(θ,ϕ)Ylm(θ,ϕ)Y00(θ,ϕ)=δllδmm4π,V_{L,L^{\prime},(00)}=\int d\Omega_{2}Y^{m*}_{l}(\theta,\phi)Y^{m^{\prime}}_{l^{\prime}}(\theta,\phi)Y^{0}_{0}(\theta,\phi)=\frac{\delta_{ll^{\prime}}\delta_{mm^{\prime}}}{\sqrt{4\pi}}\,, (91)

where we have used Y00(θ,ϕ)=1/4πY^{0}_{0}(\theta,\phi)=1/\sqrt{4\pi} and the orthogonality condition (A), and

VL,L,(01)=𝑑Ω2Ylm(θ,ϕ)Ylm(θ,ϕ)Y10(θ,ϕ)\displaystyle V_{L,L^{\prime},(01)}=\int d\Omega_{2}Y^{m*}_{l}(\theta,\phi)Y^{m^{\prime}}_{l^{\prime}}(\theta,\phi)Y^{0}_{1}(\theta,\phi)
=\displaystyle= 3(l|m|)!(l|m|)!4π(2l+1)(2l+1)(l+|m|)!(l+|m|)!(l+m)!(lm)!\displaystyle\sqrt{\frac{3(l-|m|)!(l^{\prime}-|m|)!}{4\pi(2l+1)(2l^{\prime}+1)(l+|m|)!(l^{\prime}+|m|)!}}\frac{(l^{\prime}+m)!}{(l^{\prime}-m)!}
{(lm)δl+1,l+(l+m)δl1,l}δmm,\displaystyle\cdot\{(l^{\prime}-m)\delta_{l+1,l^{\prime}}+(l+m)\delta_{l-1,l^{\prime}}\}\delta_{mm^{\prime}}\,, (92)

where we have used

02π𝑑ϕei(mm)ϕ=2πδmm\int_{0}^{2\pi}d\phi\,e^{-i(m-m^{\prime})\phi}=2\pi\delta_{mm^{\prime}} (93)

in the ϕ\phi-integral, the orthogonality relation

0πPlm(cosθ)Plm(cosθ)sinθdθ=2(l+m)!(2l+1)(lm)!δll\displaystyle\int_{0}^{\pi}P^{m}_{l}(\cos\theta)P^{m}_{l^{\prime}}(\cos\theta)\sin\theta d\theta=\frac{2(l+m)!}{(2l+1)(l-m)!}\delta_{ll^{\prime}} (94)

and the recurrence formula

(2l+1)cosθPlm(cosθ)\displaystyle(2l+1)\cos\theta P^{m}_{l}(\cos\theta)
=\displaystyle= (l+m)Pl1m(cosθ)+(lm+1)Pl+1m(cosθ)\displaystyle(l+m)P^{m}_{l-1}(\cos\theta)+(l-m+1)P^{m}_{l+1}(\cos\theta) (95)

for the associated Legendre functions of the first kind, Plm(cosθ)P^{m}_{l}(\cos\theta), in the θ\theta-integral.

The fact that the only non-zero coupling strengths VL,L,L′′V_{L,L^{\prime},L^{\prime\prime}} are (91) and (C) gives rise to three interesting characteristics in our model. (i) The only partial waves that couple to the impurity are those with m=0m=0. We have showed in (III) that the impurity only couples to the (m,l)=(0,0)(m,l)=(0,0) partial wave. Anisotropy does not change this fact, since anisotropy does not alter HIH_{I}, as previously mentioned. Now (91) and (C) both have a factor of δmm\delta_{mm^{\prime}}, implying that there is no coupling between partial waves with mmm\neq m^{\prime}. Thus the (m,l)=(0,0)(m,l)=(0,0) partial wave that couples to the impurity will only couple to other partial waves with m=0m=0. We can hence ignore all partial waves with m0m\neq 0 from H0H_{0}. We note in particular that this a characteristic universal to all Hermitian systems in any dimension, not only unique to our example. This is because for Hermitian systems, ϵ(p)\epsilon(\vec{p}) is real, so it can be expanded in terms of real spherical harmonics. Thus all YL′′Y_{L^{\prime\prime}} in the hopping integral (C) are real, and are independent of θ1\theta_{1}: indeed, from (49), we see that real spherical harmonics have no θ1\theta_{1} dependence. Thus the θ1\theta_{1}-integral in (C) only involve YLY^{*}_{L} and YLY_{L^{\prime}}, and equals to 02π𝑑θ1ei(l1l1)θ1=2πδl1l1\int_{0}^{2\pi}d\theta_{1}e^{-i(l_{1}-l^{\prime}_{1})\theta_{1}}=2\pi\delta_{l_{1}l^{\prime}_{1}}, which forbids coupling between partial waves with different l1l_{1}, which is the analogue of the quantum number mm in three dimensions. (ii) There is only nearest-neighbor hopping. This can be seen from the factor in the braces of (C): δl+1,l\delta_{l+1,l^{\prime}} and δl1,l\delta_{l-1,l^{\prime}} implies hopping can only occur between partial waves whose ll and ll^{\prime} differ by 1, resulting in only nearest-neighbor hopping. (iii) The nearest-neighbor hopping strength in (ii) decays with increasing ll. Due to (i), we are only interested in the m=0m=0 partial waves. Substitute m=0m=0 into (C), we obtain

V(0l),(0l),(01)=34π(2l+1)(2l+1)(lδl+1,l+lδl1,l).V_{(0l),(0l^{\prime}),(01)}=\sqrt{\frac{3}{4\pi(2l+1)(2l^{\prime}+1)}}(l^{\prime}\delta_{l+1,l^{\prime}}+l\delta_{l-1,l^{\prime}})\,. (96)

In other words, the coupling between the adjacent (m=0,l)(m=0,l)-th and (m=0,l+1)(m=0,l+1)-th partial waves is given by

V(0l),(0,l+1),(01)=34π(2l+1)(2l+3)(l+1).V_{(0l),(0,l+1),(01)}=\sqrt{\frac{3}{4\pi(2l+1)(2l+3)}}(l+1)\,. (97)

By computing its derivative with respect to ll, we see that the coupling strength V(0l),(0,l+1),(01)V_{(0l),(0,l+1),(01)} monotonically decreases as ll increases, for all l0l\geq 0.

The three properties (i), (ii), (iii) discussed above are depicted pictorially in the left picture of Figure 2 in the main text. In particular, property (iii) allows us to truncate the chain of partial waves in the figure, since the coupling strength decays as ll increases.

It is instructive to express the general H0H_{0} in (84) into a matrix form in the partial wave basis, as shown in equation (18) of the main text:

H0=𝑑pCα,i,μ(p)Aμ(p)Cα,i,μ(p),H_{0}=\int dp\,C^{\dagger}_{\alpha,i,\mu}(p)A_{\mu}(p)C_{\alpha,i,\mu}(p)\,, (98)

where Aμ(p)A_{\mu}(p) is a matrix whose (L,L)(L,L^{\prime})-th element, AL,L,μ(p)A_{L,L^{\prime},\mu}(p), is given by

AL,L,μ(p)=L′′VL,L,L′′ϵL′′μ(p)\displaystyle A_{L,L^{\prime},\mu}(p)=\sum_{L^{\prime\prime}}V_{L,L^{\prime},L^{\prime\prime}}\epsilon_{L^{\prime\prime}\mu}(p)
=\displaystyle= 𝑑Ωn1YL(θ1,,θn1)YL(θ1,,θn1)ϵμ(p),\displaystyle\int d\Omega_{n-1}Y^{*}_{L}(\theta_{1},...,\theta_{n-1})Y_{L^{\prime}}(\theta_{1},...,\theta_{n-1})\epsilon_{\mu}(\vec{p})\,, (99)

Cα,i,μ(p)C_{\alpha,i,\mu}(p) is a column vector whose LL-th element is given by cL,α,i,μ(p)c_{L,\alpha,i,\mu}(p), and Cα,i,μ(p)C^{\dagger}_{\alpha,i,\mu}(p) is its Hermitian conjugate. Recall that property (i), which holds for all Hermitian systems, allows us to only include partial waves with l1=0l_{1}=0 in H0H_{0}. Thus here, the multi-indices LL and LL^{\prime} only includes (l1=0,l2,,ln1)(l_{1}=0,l_{2},...,l_{n-1}). In particular, in 3 dimensions, we have the property that the multi-indices LL and LL^{\prime} only includes (m=0,l)l(m=0,l)\rightarrow l, i.e. in 3 dimensions the multi-indices LL and LL^{\prime} reduce to the single indices ll and ll^{\prime}: L=lL=l, L=lL^{\prime}=l^{\prime}.

In our specific example (86), Aμ(p)A_{\mu}(p) is the tridiagonal matrix:

Aμ(p)=(a1μp3a2μp3a2μp3a1μp32a2μp152a2μp15a1μp33a2μp353a2μp35a1μp34a2μp63),\displaystyle A_{\mu}(p)=\left(\begin{array}[]{@{}c|ccccc@{}}a_{1\mu}p^{3}&\frac{a_{2\mu}p}{\sqrt{3}}&&&&\\ \hline\cr\frac{a_{2\mu}p}{\sqrt{3}}&a_{1\mu}p^{3}&\frac{2a_{2\mu}p}{\sqrt{15}}&&&\\ &\frac{2a_{2\mu}p}{\sqrt{15}}&a_{1\mu}p^{3}&\frac{3a_{2\mu}p}{\sqrt{35}}&&\\ &&\frac{3a_{2\mu}p}{\sqrt{35}}&a_{1\mu}p^{3}&\frac{4a_{2\mu}p}{\sqrt{63}}&\\ &&&\ddots&\ddots&\ddots\end{array}\right)\,, (105)

where the partitioning denoted by the horizontal and vertical lines will be explained shortly. We have A+(p)=A(p)A_{+}(p)=-A_{-}(p) due to (87). In general, Aμ(p)A_{\mu}(p) is a ×\infty\times\infty matrix, but by property (iii), since the hopping strength decays as the row/column number L=lL=l increases, in practice we can truncate it to an appropriate size for further calculations.

We next partition Aμ(p)A_{\mu}(p) as shown in (105): the first block consists of only the (1,1)-th element, the second block consists of the remaining (1)×(1)(\infty-1)\times(\infty-1) square matrix, which we denote by Mμ(p)M_{\mu}(p), and the third and fourth parts consist of the (1)×1(\infty-1)\times 1 column vector, denoted by Nμ(p)N_{\mu}(p), and its transpose, the 1×(1)1\times(\infty-1) row vector NμT(p)N^{T}_{\mu}(p) respectively. This corresponds to separating the matrix form of H0H_{0} in (19) into four terms:

H0=𝑑p\displaystyle H_{0}=\int dp (Γ(n/2)2πn2c0,,0,α,i,μ(p)c0,,0,α,i,μ(p)ϵμ(p)\displaystyle\left(\frac{\Gamma(n/2)}{2\pi^{\frac{n}{2}}}c^{\dagger}_{0,...,0,\alpha,i,\mu}(p)c_{0,...,0,\alpha,i,\mu}(p)\epsilon_{\mu}(p)\right.
+𝒞α,i,μ(p)Mμ(p)𝒞α,i,μ(p)\displaystyle+\mathcal{C}^{\dagger}_{\alpha,i,\mu}(p)M_{\mu}(p)\mathcal{C}_{\alpha,i,\mu}(p)
+𝒞α,i,μ(p)Nμ(p)c0,,0,α,i,μ(p)\displaystyle+\mathcal{C}^{\dagger}_{\alpha,i,\mu}(p)N_{\mu}(p)c_{0,...,0,\alpha,i,\mu}(p)
+c0,,0,α,i,μ(p)NμT(p)𝒞α,i,μ(p)).\displaystyle+\left.c^{\dagger}_{0,...,0,\alpha,i,\mu}(p)N^{T}_{\mu}(p)\mathcal{C}_{\alpha,i,\mu}(p)\right)\,. (106)

We shall now formally define the objects that appear in (C). First of all, define the isotropic kernel ϵμ(p)\epsilon_{\mu}(p) of the dispersion relation ϵμ(p)\epsilon_{\mu}(\vec{p}) as

ϵμ(p)𝑑Ωn1ϵμ(p).\displaystyle\epsilon_{\mu}(p)\equiv\int d\Omega_{n-1}\epsilon_{\mu}(\vec{p})\,. (107)

ϵμ(p)\epsilon_{\mu}(p) can be seen as the “isotropic part” of the dispersion relation. Next, let (ł1,,ln1)(0,,0)\mathcal{L}\equiv(\l_{1},...,l_{n-1})\neq(0,...,0) be the multi-index obtained from LL by excluding L=(0,,0)L=(0,...,0). 𝒞α,i,μ(p)\mathcal{C}_{\alpha,i,\mu}(p) is the column vector whose \mathcal{L}-th element is given by c,α,i,μ(p)c_{{\mathcal{L}},\alpha,i,\mu}(p), i.e. 𝒞α,i,μ(p)\mathcal{C}_{\alpha,i,\mu}(p) is obtained from Cα,i,μ(p)C_{\alpha,i,\mu}(p) by excluding its first entry c0,,0,α,i,μ(p)c_{0,...,0,\alpha,i,\mu}(p). 𝒞α,i,μ(p)\mathcal{C}^{\dagger}_{\alpha,i,\mu}(p) is the Hermitian conjugate of 𝒞α,i,μ(p)\mathcal{C}_{\alpha,i,\mu}(p), i.e. 𝒞α,i,μ(p)\mathcal{C}^{\dagger}_{\alpha,i,\mu}(p) is the row vector obtained from Cα,i,μ(p)C^{\dagger}_{\alpha,i,\mu}(p) by excluding its first entry c0,,0,α,i,μ(p)c^{\dagger}_{0,...,0,\alpha,i,\mu}(p). Mμ(p)M_{\mu}(p) is the matrix whose (,)(\mathcal{L},\mathcal{L^{\prime}})-th element M,,μ(p)M_{\mathcal{L},\mathcal{L^{\prime}},\mu}(p) is defined by

M,,μ(p)A,,μ(p),M_{\mathcal{L},\mathcal{L^{\prime}},\mu}(p)\equiv A_{\mathcal{L},\mathcal{L^{\prime}},\mu}(p)\,, (108)

i.e. Mμ(p)M_{\mu}(p) is the matrix obtained from Aμ(p)A_{\mu}(p) by removing its first row and first column, which corresponds to L=(0,,0)L=(0,...,0) and L=(0,,0)L^{\prime}=(0,...,0) respectively. In our example, Mμ(p)M_{\mu}(p) is the bottom right block matrix in (105). Nμ(p)N_{\mu}(p) is the column vector whose \mathcal{L}-th entry N,μ(p)N_{\mathcal{L},\mu}(p) is given by

N,μ(p)A,(0,,0),μ(p)\displaystyle N_{\mathcal{L},\mu}(p)\equiv A_{\mathcal{L},(0,...,0),\mu}(p)
=\displaystyle= 𝑑Ωn1Y(θ1,,θn1)Y0,,0(θ1,,θn1)ϵμ(p)\displaystyle\int d\Omega_{n-1}Y^{*}_{\mathcal{L}}(\theta_{1},...,\theta_{n-1})Y_{0,...,0}(\theta_{1},...,\theta_{n-1})\epsilon_{\mu}(\vec{p})
=\displaystyle= (Γ(n/2)2πn2)12𝑑Ωn1Y(θ1,,θn1)ϵμ(p)\displaystyle\left(\frac{\Gamma(n/2)}{2\pi^{\frac{n}{2}}}\right)^{\frac{1}{2}}\int d\Omega_{n-1}Y^{*}_{\mathcal{L}}(\theta_{1},...,\theta_{n-1})\epsilon_{\mu}(\vec{p})
=\displaystyle= (Γ(n/2)2πn2)12𝑑Ωn1Y(θ1,,θn1)\displaystyle\left(\frac{\Gamma(n/2)}{2\pi^{\frac{n}{2}}}\right)^{\frac{1}{2}}\int d\Omega_{n-1}Y^{*}_{\mathcal{L}}(\theta_{1},...,\theta_{n-1})
LYL(θ1,,θn1)ϵL,μ(p)\displaystyle\sum_{L^{\prime}}Y_{L^{\prime}}(\theta_{1},...,\theta_{n-1})\epsilon_{L^{\prime},\mu}(p)
=\displaystyle= (Γ(n/2)2πn2)12ϵ,μ(p),\displaystyle\left(\frac{\Gamma(n/2)}{2\pi^{\frac{n}{2}}}\right)^{\frac{1}{2}}\epsilon_{\mathcal{L},\mu}(p)\,, (109)

where in the third equality we have used the fact that Y0,0(θ1,,θn1)Y_{0,...0}(\theta_{1},...,\theta_{n-1}) is a real constant given by (56), in the fourth equality we have expanded the dispersion relation ϵμ(p)\epsilon_{\mu}(\vec{p}) in terms of the spherical harmonics (83), and in the last equality we have used the orthogonality relation (A). Recall that ϵ,μ(p)\epsilon_{\mathcal{L},\mu}(p) does not include ϵ0,,0,μ(p)\epsilon_{0,...,0,\mu}(p). In our 3-dimensional example, the only non-zero ϵ,μ(p)\epsilon_{\mathcal{L},\mu}(p) is ϵ0,1,μ(p)=a2μp4π/3\epsilon_{0,1,\mu}(p)=a_{2\mu}p\sqrt{4\pi/3}, according to (90). Thus for our example

Nμ(p)=(Γ(32)2π32)12(a2μp4π30)=(a2μp30),\displaystyle N_{\mu}(p)=\left(\frac{\Gamma(\frac{3}{2})}{2\pi^{\frac{3}{2}}}\right)^{\frac{1}{2}}\begin{pmatrix}\frac{a_{2\mu}p\sqrt{4\pi}}{\sqrt{3}}\\ \vdots\\ \text{\huge 0}\\ \vdots\\ \end{pmatrix}=\begin{pmatrix}\frac{a_{2\mu}p}{\sqrt{3}}\\ \vdots\\ \text{\huge 0}\\ \vdots\\ \end{pmatrix}\,, (110)

as can be seen from the first column of the matrix (105), excluding its first entry. NμT(p)N^{T}_{\mu}(p) is the transpose pf Nμ(p)N_{\mu}(p). Also in the derivation of the first term of (C) we have used

𝑑pc0,,0,α,i,μ(p)A(0,,0),(0,,0),μ(p)c0,,0,α,i,μ(p)\displaystyle\int dp\,c^{\dagger}_{0,...,0,\alpha,i,\mu}(p)A_{(0,...,0),(0,...,0),\mu}(p)c_{0,...,0,\alpha,i,\mu}(p)
=\displaystyle= 𝑑pc0,,0,α,i,μ(p)c0,,0,α,i,μ(p)\displaystyle\int dp\,c^{\dagger}_{0,...,0,\alpha,i,\mu}(p)c_{0,...,0,\alpha,i,\mu}(p)
dΩn1Y0,,0(θ1,,θn1)Y0,0(θ1,,θn1)ϵμ(p)\displaystyle\cdot\int d\Omega_{n-1}Y^{*}_{0,...,0}(\theta_{1},...,\theta_{n-1})Y_{0,...0}(\theta_{1},...,\theta_{n-1})\epsilon_{\mu}(\vec{p})
=\displaystyle= Γ(n/2)2πn2𝑑pc0,,0,α,i,μ(p)c0,,0,α,i,μ(p)𝑑Ωn1ϵμ(p)\displaystyle\frac{\Gamma(n/2)}{2\pi^{\frac{n}{2}}}\int dp\,c^{\dagger}_{0,...,0,\alpha,i,\mu}(p)c_{0,...,0,\alpha,i,\mu}(p)\int d\Omega_{n-1}\epsilon_{\mu}(\vec{p})
=\displaystyle= Γ(n/2)2πn2𝑑pc0,,0,α,i,μ(p)c0,,0,α,i,μ(p)ϵμ(p),\displaystyle\frac{\Gamma(n/2)}{2\pi^{\frac{n}{2}}}\int dp\,c^{\dagger}_{0,...,0,\alpha,i,\mu}(p)c_{0,...,0,\alpha,i,\mu}(p)\epsilon_{\mu}(p)\,,

where in the last equality, we have used the definition of the kernel ϵμ(p)\epsilon_{\mu}(p) in (107).

Let us define the first term of H0H_{0} in (C) to be H00H_{00}, and define the sum of the remaining three terms of H0H_{0} in (C) to be Σ0\Sigma_{0}:

H0=H00+Σ0,H_{0}=H_{00}+\Sigma_{0}\,, (112)
H00Γ(n/2)2πn2𝑑pc0,,0,α,i,μ(p)c0,,0,α,i,μ(p)ϵμ(p),H_{00}\equiv\frac{\Gamma(n/2)}{2\pi^{\frac{n}{2}}}\int dp\,c^{\dagger}_{0,...,0,\alpha,i,\mu}(p)c_{0,...,0,\alpha,i,\mu}(p)\epsilon_{\mu}(p)\,, (113)
Σ0𝑑p\displaystyle\Sigma_{0}\equiv\int dp (𝒞α,i,μ(p)Mμ(p)𝒞α,i,μ(p)\displaystyle\left(\mathcal{C}^{\dagger}_{\alpha,i,\mu}(p)M_{\mu}(p)\mathcal{C}_{\alpha,i,\mu}(p)\right.
+𝒞α,i,μ(p)Nμ(p)c0,,0,α,i,μ(p)\displaystyle+\mathcal{C}^{\dagger}_{\alpha,i,\mu}(p)N_{\mu}(p)c_{0,...,0,\alpha,i,\mu}(p)
+c0,,0,α,i,μ(p)NμT(p)𝒞α,i,μ(p)).\displaystyle+\left.c^{\dagger}_{0,...,0,\alpha,i,\mu}(p)N^{T}_{\mu}(p)\mathcal{C}_{\alpha,i,\mu}(p)\right)\,. (114)

We make these definitions because H00H_{00} is the “isotropic part” of H0H_{0}, whereas Σ0\Sigma_{0} is the “anisotropic part” of H0H_{0}. Indeed, H00H_{00} only includes the L=(0,,0)L=(0,...,0) partial wave, and Σ0\Sigma_{0} contains the remaining (0,,0)\mathcal{L}\neq(0,...,0) partial waves. For systems with isotropic dispersion relations, only the L=(0,,0)L=(0,...,0) partial wave couples to the impurity, and we can ignore all 0\mathcal{L}\neq 0 partial waves, so Σ0=0\Sigma_{0}=0 and H0=H00H_{0}=H_{00}. All contributions to the Hamiltonian due to anisotropy of the system are captured in Σ0\Sigma_{0}.

We shall now write the partition function 𝒵\mathcal{Z} for our system. Let ψL,α,i,μ(p)\psi_{L,\alpha,i,\mu}(p) be the Grassmann variable obtained from the operator cL,α,i,μ(p)c_{L,\alpha,i,\mu}(p) acting on a coherent state. We have

𝒵=D(ψ¯,ψ)eS[ψ¯,ψ]\displaystyle\mathcal{Z}=\int D(\bar{\psi},\psi)e^{-S[\bar{\psi},\psi]}
=\displaystyle= D(ψ0¯,ψ0)eS[ψ0¯,ψ0]D(ψ¯,ψ)eS[ψ¯,ψ],\displaystyle\int D(\bar{\psi_{0}},\psi_{0})e^{-S[\bar{\psi_{0}},\psi_{0}]}\int D(\bar{\psi_{\mathcal{L}}},\psi_{\mathcal{L}})e^{-S[\bar{\psi_{\mathcal{L}}},\psi_{\mathcal{L}}]}\,,

where S[ψ0¯,ψ0]S[\bar{\psi_{0}},\psi_{0}] is the part of the action that only involves the L=(0,,0)L=(0,...,0) partial wave, and S[ψ¯,ψ]S[\bar{\psi_{\mathcal{L}}},\psi_{\mathcal{L}}] is the part of the action that involves the remaining (0,,0)\mathcal{L}\neq(0,...,0) partial waves. As previously discussed, in our Hamiltonian H=H00+Σ0+HIH=H_{00}+\Sigma_{0}+H_{I}, H00H_{00} and HIH_{I} only includes the L=(0,,0)L=(0,...,0) partial wave, so H00H_{00} and HIH_{I} appears in S[ψ0¯,ψ0]S[\bar{\psi_{0}},\psi_{0}], whereas Σ0\Sigma_{0} includes the remaining (0,,0)\mathcal{L}\neq(0,...,0) partial waves, so Σ0\Sigma_{0} appears in S[ψ¯,ψ]S[\bar{\psi_{\mathcal{L}}},\psi_{\mathcal{L}}]. We thus have

S[ψ0¯,ψ0]=\displaystyle S[\bar{\psi_{0}},\psi_{0}]= α,β,i,μ𝑑τ\displaystyle\sum_{\alpha,\beta,i,\mu}\int d\tau
{dpψ¯0,,0,α,i,μ(p)τψ0,,0,α,i,μ(p)\displaystyle\left\{\int dp\,\bar{\psi}_{0,...,0,\alpha,i,\mu}(p)\partial_{\tau}\psi_{0,...,0,\alpha,i,\mu}(p)\right.
+𝑑pH00(ψ¯0,,0,α,i,μ(p),ψ0,,0,α,i,μ(p))\displaystyle+\int dpH_{00}(\bar{\psi}_{0,...,0,\alpha,i,\mu}(p),\psi_{0,...,0,\alpha,i,\mu}(p))
+dpdpHI(ψ¯0,,0,α,i,μ(p),ψ0,,0,β,i,μ(p))}\displaystyle+\left.\int dp\,dp^{\prime}H_{I}(\bar{\psi}_{0,...,0,\alpha,i,\mu}(p),\psi_{0,...,0,\beta,i,\mu}(p^{\prime}))\right\}

and

S[ψ¯,ψ]=α,i,μ𝑑τ𝑑p\displaystyle S[\bar{\psi_{\mathcal{L}}},\psi_{\mathcal{L}}]=\sum_{\alpha,i,\mu}\int d\tau dp\, {¯α,i,μ(p)(τ+Mμ(p))α,i,μ(p)\displaystyle\{\bar{\mathcal{F}}_{\alpha,i,\mu}(p)(\partial_{\tau}+M_{\mu}(p))\mathcal{F}_{\alpha,i,\mu}(p)
+¯α,i,μ(p)𝒩α,i,μ(p)\displaystyle+\bar{\mathcal{F}}_{\alpha,i,\mu}(p)\mathcal{N}_{\alpha,i,\mu}(p)
+𝒩¯α,i,μT(p)α,i,μ(p)},\displaystyle+\bar{\mathcal{N}}^{T}_{\alpha,i,\mu}(p)\mathcal{F}_{\alpha,i,\mu}(p)\}\,,

where α,i,μ(p)\mathcal{F}_{\alpha,i,\mu}(p) is the column vector whose \mathcal{L}-th element is given by ψ,α,i,μ(p)\psi_{{\mathcal{L}},\alpha,i,\mu}(p), ¯α,i,μ(p)\bar{\mathcal{F}}_{\alpha,i,\mu}(p) is the row vector whose \mathcal{L}-th element is given by ψ¯,α,i,μ(p)\bar{\psi}_{{\mathcal{L}},\alpha,i,\mu}(p),

𝒩α,i,μ(p)Nμ(p)ψ0,,0,α,i,μ(p),\mathcal{N}_{\alpha,i,\mu}(p)\equiv N_{\mu}(p)\psi_{0,...,0,\alpha,i,\mu}(p)\,, (118)

and

𝒩¯α,i,μT(p)ψ¯0,,0,α,i,μ(p)NμT(p).\bar{\mathcal{N}}^{T}_{\alpha,i,\mu}(p)\equiv\bar{\psi}_{0,...,0,\alpha,i,\mu}(p)N^{T}_{\mu}(p)\,. (119)

We then integrate out the Grassmann field ¯α,i,μ(p)\bar{\mathcal{F}}_{\alpha,i,\mu}(p), α,i,μ(p)\mathcal{F}_{\alpha,i,\mu}(p) in (C), after which the second factor of (C) becomes

D(ψ¯,ψ)eS[ψ¯,ψ]\displaystyle\int D(\bar{\psi_{\mathcal{L}}},\psi_{\mathcal{L}})e^{-S[\bar{\psi_{\mathcal{L}}},\psi_{\mathcal{L}}]}
=\displaystyle= eα,i,μ𝑑τ𝑑p{𝒩¯α,i,μT(p)Mμ1(p)𝒩α,i,μ(p)}\displaystyle e^{\sum_{\alpha,i,\mu}\int d\tau dp\{\bar{\mathcal{N}}^{T}_{\alpha,i,\mu}(p)M^{-1}_{\mu}(p)\mathcal{N}_{\alpha,i,\mu}(p)\}}
\displaystyle\equiv eα,i,μ𝑑τ𝑑p{ψ¯0,,0,α,i,μ(p)NμT(p)Mμ1(p)Nμ(p)ψ0,,0,α,i,μ(p)}.\displaystyle e^{\sum_{\alpha,i,\mu}\int d\tau dp\{\bar{\psi}_{0,...,0,\alpha,i,\mu}(p)N^{T}_{\mu}(p)M^{-1}_{\mu}(p)N_{\mu}(p)\psi_{0,...,0,\alpha,i,\mu}(p)\}}\,. (120)

where we have taken the static limit, τiw0\partial_{\tau}\rightarrow-iw\rightarrow 0, in (C). This is a good approximation that well captures the renormalization of the isotropic part in low-energy window, as long as the system described by Mμ(p)M_{\mu}(p) is gapped. Notice that the factor in Eq. (C) now only includes the L=(0,,0)L=(0,...,0) partial wave. Thus the partition function (C) becomes

𝒵=D(ψ0¯,ψ0)e(S[ψ0¯,ψ0]+SΣ0),\displaystyle\mathcal{Z}=\int D(\bar{\psi_{0}},\psi_{0})e^{-(S[\bar{\psi_{0}},\psi_{0}]+S_{\Sigma_{0}})}\,, (121)

where

SΣ0α,i,μ𝑑τ𝑑p\displaystyle S_{\Sigma_{0}}\equiv-\sum_{\alpha,i,\mu}\int d\tau dp {ψ¯0,,0,α,i,μ(p)NμT(p)Mμ1(p)Nμ(p)\displaystyle\left\{\bar{\psi}_{0,...,0,\alpha,i,\mu}(p)N^{T}_{\mu}(p)M^{-1}_{\mu}(p)N_{\mu}(p)\right.
ψ0,,0,α,i,μ(p)}.\displaystyle\left.\psi_{0,...,0,\alpha,i,\mu}(p)\right\}\,. (122)

This means that SΣ0S_{\Sigma_{0}} contributes an additional term of

𝑑pψ¯0,,0,α,i,μ(p)NμT(p)Mμ1(p)Nμ(p)ψ0,,0,α,i,μ(p)\displaystyle-\int dp\,\bar{\psi}_{0,...,0,\alpha,i,\mu}(p)N^{T}_{\mu}(p)M^{-1}_{\mu}(p)N_{\mu}(p)\psi_{0,...,0,\alpha,i,\mu}(p) (123)

to the braced integrand of S[ψ0¯,ψ0]S[\bar{\psi_{0}},\psi_{0}] given in (C). In terms of the Hamiltonian, this term is

𝑑pc0,,0,α,i,μ(p)NμT(p)Mμ1(p)Nμ(p)c0,,0,α,i,μ(p).\displaystyle-\int dp\,c^{\dagger}_{0,...,0,\alpha,i,\mu}(p)N^{T}_{\mu}(p)M^{-1}_{\mu}(p)N_{\mu}(p)c_{0,...,0,\alpha,i,\mu}(p)\,. (124)

Thus we can write H0H_{0} as

H0=H00+Σ0\displaystyle H_{0}=H_{00}+\Sigma_{0}
=\displaystyle= Γ(n/2)2πn2𝑑pc0,,0,α,i,μ(p)c0,,0,α,i,μ(p)ϵμ(p)\displaystyle\frac{\Gamma(n/2)}{2\pi^{\frac{n}{2}}}\int dp\,c^{\dagger}_{0,...,0,\alpha,i,\mu}(p)c_{0,...,0,\alpha,i,\mu}(p)\epsilon_{\mu}(p)
𝑑pc0,,0,α,i,μ(p)NμT(p)Mμ1(p)Nμ(p)c0,,0,α,i,μ(p)\displaystyle-\int dp\,c^{\dagger}_{0,...,0,\alpha,i,\mu}(p)N^{T}_{\mu}(p)M^{-1}_{\mu}(p)N_{\mu}(p)c_{0,...,0,\alpha,i,\mu}(p)
\displaystyle\equiv Γ(n/2)2πn2𝑑pc0,,0,α,i,μ(p)c0,,0,α,i,μ(p){ϵμ(p)+ϵΣμ(p)}\displaystyle\frac{\Gamma(n/2)}{2\pi^{\frac{n}{2}}}\int dp\,c^{\dagger}_{0,...,0,\alpha,i,\mu}(p)c_{0,...,0,\alpha,i,\mu}(p)\{\epsilon_{\mu}(p)+\epsilon_{\Sigma\mu}(p)\}
\displaystyle\equiv 𝑑pc0,,0,α,i,μ(p)c0,,0,α,i,μ(p)ϵ~μ(p),\displaystyle\int dp\,c^{\dagger}_{0,...,0,\alpha,i,\mu}(p)c_{0,...,0,\alpha,i,\mu}(p)\tilde{\epsilon}_{\mu}(p)\,, (125)

where in the second equality we have used the definition of H00H_{00} given in (113), in the third equality we have made the definition

ϵΣμ(p)2πn2Γ(n/2)NμT(p)Mμ1(p)Nμ(p),\epsilon_{\Sigma\mu}(p)\equiv-\frac{2\pi^{\frac{n}{2}}}{\Gamma(n/2)}N^{T}_{\mu}(p)M^{-1}_{\mu}(p)N_{\mu}(p)\,, (126)

and in the last equality we have made the definition

ϵ~μ(p)Γ(n/2)2πn2(ϵμ(p)+ϵΣμ(p)).\tilde{\epsilon}_{\mu}(p)\equiv\frac{\Gamma(n/2)}{2\pi^{\frac{n}{2}}}\left(\epsilon_{\mu}(p)+\epsilon_{\Sigma\mu}(p)\right)\,. (127)

We call ϵ~μ(p)\tilde{\epsilon}_{\mu}(p) the effective dispersion relation of the system.

We note that the kernel ϵ\epsilon satisfies the relation

ϵ+(p)=ϵ(p),\epsilon_{+}(p)=-\epsilon_{-}(p)\,, (128)

because

ϵ+(p)𝑑Ωn1ϵ+(p)=𝑑Ωn1ϵ(p)ϵ(p),\displaystyle\epsilon_{+}(p)\equiv\int d\Omega_{n-1}\epsilon_{+}(\vec{p})=-\int d\Omega_{n-1}\epsilon_{-}(\vec{p})\equiv-\epsilon_{-}(p)\,, (129)

where in the first and last equalities we have used the definition of the kernel given in (107), and in the second equality we have used the relation ϵ+(p)=ϵ(p)\epsilon_{+}(\vec{p})=-\epsilon_{-}(\vec{p}). Similarly, ϵΣμ(p)\epsilon_{\Sigma\mu}(p) satisfies the relation

ϵΣ+(p)=ϵΣ(p).\epsilon_{\Sigma+}(p)=-\epsilon_{\Sigma-}(p)\,. (130)

To see this, notice that in the definition of ϵΣμ(p)\epsilon_{\Sigma\mu}(p) given in (126), each of the factors NμT(p),Mμ1(p),N^{T}_{\mu}(p),M^{-1}_{\mu}(p), and Nμ(p)N_{\mu}(p) flips a sign when μ\mu flips a sign: indeed, the entries of NμT(p),Mμ(p)N^{T}_{\mu}(p),M_{\mu}(p) and Nμ(p)N_{\mu}(p) are just the entries of the matrix AA defined in (C), whose integrand contains a factor of ϵμ(p)\epsilon_{\mu}(\vec{p}), which flips a sign as μ\mu flips a sign. By matrix inverse properties we also have

M+1(p)=(M(p))1=M1(p),\displaystyle M^{-1}_{+}(p)=(-M_{-}(p))^{-1}=-M^{-1}_{-}(p)\,, (131)

completing the proof. Thus the effective dispersion relation ϵ~μ(p)\tilde{\epsilon}_{\mu}(p) also satisfies the relation

ϵ~+(p)=ϵ~(p),\tilde{\epsilon}_{+}(p)=-\tilde{\epsilon}_{-}(p)\,, (132)

due to (127), (128) and (130).

Let us compute the effective dispersion relation ϵ~μ(p)\tilde{\epsilon}_{\mu}(p) for our example. The kernel is given by

ϵμ(p)\displaystyle\epsilon_{\mu}(p) =𝑑Ω2ϵμ(p)\displaystyle=\int d\Omega_{2}\,\epsilon_{\mu}(\vec{p})
=𝑑Ω2{a1μ(px2+py2+pz2)32+a2μpz}\displaystyle=\int d\Omega_{2}\,\left\{a_{1\mu}(p_{x}^{2}+p_{y}^{2}+p_{z}^{2})^{\frac{3}{2}}+a_{2\mu}p_{z}\right\}
=02π𝑑ϕ0πsin(θ)𝑑θ{a1μp3+a2μpcos(θ)}\displaystyle=\int_{0}^{2\pi}d\phi\int_{0}^{\pi}\sin(\theta)d\theta\,\left\{a_{1\mu}p^{3}+a_{2\mu}p\cos(\theta)\right\}
=4πa1μp3.\displaystyle=4\pi a_{1\mu}p^{3}\,. (133)

As for ϵΣμ(p)\epsilon_{\Sigma\mu}(p) defined in (126), since Nμ(p)N_{\mu}(p) in our example only has its first entry being non-zero, as seen from (110), ϵΣμ(p)\epsilon_{\Sigma\mu}(p) becomes

ϵΣμ(p)=4π3a2μ2p2(Mμ1(p))11,\displaystyle\epsilon_{\Sigma\mu}(p)=-\frac{4\pi}{3}a_{2\mu}^{2}p^{2}\left(M^{-1}_{\mu}(p)\right)_{11}\,, (134)

where (Mμ1(p))11\left(M^{-1}_{\mu}(p)\right)_{11} denotes the (1,1)-th element of Mμ1(p)M^{-1}_{\mu}(p). Recall that Mμ(p)M_{\mu}(p) is the lower right block matrix of Aμ(p)A_{\mu}(p) in (105):

Mμ(p)=(a1μp32a2μp152a2μp15a1μp33a2μp353a2μp35a1μp34a2μp63).\displaystyle M_{\mu}(p)=\left(\begin{array}[]{@{}ccccc@{}}a_{1\mu}p^{3}&\frac{2a_{2\mu}p}{\sqrt{15}}&&&\\ \frac{2a_{2\mu}p}{\sqrt{15}}&a_{1\mu}p^{3}&\frac{3a_{2\mu}p}{\sqrt{35}}&&\\ &\frac{3a_{2\mu}p}{\sqrt{35}}&a_{1\mu}p^{3}&\frac{4a_{2\mu}p}{\sqrt{63}}&\\ &&\ddots&\ddots&\ddots\end{array}\right)\,. (139)

By the discussion below (105), we may truncate Aμ(p)A_{\mu}(p), and thus Mμ(p)M_{\mu}(p), to a desired size for actual calculations. Let Mμr(p)M^{r}_{\mu}(p) denote the r×rr\times r truncated matrix of Mμ(p)M_{\mu}(p), i.e. Mμr(p)M^{r}_{\mu}(p) is obtained from Mμ(p)M_{\mu}(p) by taking only its first rr rows and first rr columns. Because Mμr(p)M^{r}_{\mu}(p) is a tridiagonal matrix, one can show that

(Mμr(p))111=Gμr(p)Dμr(p),\left(M^{r}_{\mu}(p)\right)^{-1}_{11}=\frac{G^{r}_{\mu}(p)}{D^{r}_{\mu}(p)}\,, (140)

where Dμr(p)D^{r}_{\mu}(p) is the determinant of Mμr(p)M^{r}_{\mu}(p), which can be computed from the recurrence relation

Dμr(p)=a1μp3Dμr1(p)(a2μpr)2(2r1)(2r+1)Dμr2(p),\displaystyle D^{r}_{\mu}(p)=a_{1\mu}p^{3}D^{r-1}_{\mu}(p)-\frac{(a_{2\mu}pr)^{2}}{(2r-1)(2r+1)}D^{r-2}_{\mu}(p)\,, (141)

with Dμ0(p)=1,Dμ1(p)=a1μp3D^{0}_{\mu}(p)=1,D^{1}_{\mu}(p)=a_{1\mu}p^{3}, and Gμr(p)G^{r}_{\mu}(p) is a polynomial in pp that satisfies the same recurrence relation

Gμr(p)=a1μp3Gμr1(p)(a2μpr)2(2r1)(2r+1)Gμr2(p),\displaystyle G^{r}_{\mu}(p)=a_{1\mu}p^{3}G^{r-1}_{\mu}(p)-\frac{(a_{2\mu}pr)^{2}}{(2r-1)(2r+1)}G^{r-2}_{\mu}(p)\,, (142)

but with different initial terms Gμ0(p)=0,Gμ1(p)=1G^{0}_{\mu}(p)=0,G^{1}_{\mu}(p)=1. Thus in our example,

ϵ~μ(p)=a1μp3(a2μp)23Gμr(p)Dμr(p),\tilde{\epsilon}_{\mu}(p)=a_{1\mu}p^{3}-\frac{(a_{2\mu}p)^{2}}{3}\frac{G^{r}_{\mu}(p)}{D^{r}_{\mu}(p)}\,, (143)

where rr\rightarrow\infty.

We view the term ϵΣμ(p)\epsilon_{\Sigma\mu}(p), which captures all anisotropic effects, as a correction to the isotropic term of the dispersion relation, the kernel ϵμ(p)\epsilon_{\mu}(p). We approximate ϵΣμ(p)\epsilon_{\Sigma\mu}(p) by evaluating its value at pFp_{F}, before adding it to the kernel ϵμ(p)\epsilon_{\mu}(p). pFp_{F} satisfies the property that ϵ~μ(pF)=0\tilde{\epsilon}_{\mu}(p_{F})=0, i.e. pFp_{F} is a non-negative real root of ϵ~μ(p)\tilde{\epsilon}_{\mu}(p) (pFp_{F} needs to be non-negative because it is a special value of p|p|p\equiv|\vec{p}|). In general, ϵ~μ(p)\tilde{\epsilon}_{\mu}(p) may have no non-negative real root, in which case the system is gapped and we do not consider such gapped systems. Also, ϵ~μ(p)\tilde{\epsilon}_{\mu}(p) may have more than one non-negative real root. In such cases we pick the non-negative real root of ϵ~μ(p)\tilde{\epsilon}_{\mu}(p) that gives rise to the greatest density of states (DOS) to be the pFp_{F} at which we evaluate ϵΣμ(pF)\epsilon_{\Sigma\mu}(p_{F}). In other words, we pick the non-negative real root that results in the least |dϵ~μ(pF)/dp||d\tilde{\epsilon}_{\mu}(p_{F})/dp|. If there exists more than one non-negative real root of ϵ~μ(p)\tilde{\epsilon}_{\mu}(p) that give rise to the same greatest DOS, we evaluate ϵΣμ(p)\epsilon_{\Sigma\mu}(p) at each of them - this will result in channel-multiplying.

In our example, ϵ~μ(p)\tilde{\epsilon}_{\mu}(p) given by (143) has more than one non-negative real root, with the smallest of them giving rise to the least |dϵ~μ(p)/dp||d\tilde{\epsilon}_{\mu}(p)/dp|, thus the greatest DOS. Hence there is only one pFp_{F} at which we evaluate ϵΣμ(pF)\epsilon_{\Sigma\mu}(p_{F}), namely this smallest non-negative real root of ϵ~μ(p)\tilde{\epsilon}_{\mu}(p). For all even values of rr, this smallest non-negative real root of ϵ~μ(p)\tilde{\epsilon}_{\mu}(p) is 0. For odd values of rr, as rr\rightarrow\infty, the smallest non-negative real root also approaches 0. Thus for our example, pF=0p_{F}=0, ϵΣμ(0)=0\epsilon_{\Sigma\mu}(0)=0, and our effective dispersion relation becomes

ϵ~μ(p)=a1μp3.\tilde{\epsilon}_{\mu}(p)=a_{1\mu}p^{3}\,. (144)

Note that the effective dispersion relation (144) is in the form of equation (5) of the main text in three dimensions. As a result, all subsequent derivations following equation (5) of the main text hold, and our FEMCFT approach applies.

References