Exact solutions of Kondo problems in higher-order fermions
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 about the Fermi surface, with being the Fermi velocity and 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,
(1) |
where is the band index, and are dispersion coefficients, . 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 Sn and PdSb2 Fang et al. (2012); Huang et al. (2016); Bradlyn et al. (2016), and examples of cubic Weyl/Dirac semimetals include LiOO3 and quasi-one-dimensional molybdenum monochalcogenide compounds A(MoX)3, where A = Na, K, Rb, In or Tl, X = 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 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 , and only mapping the low energy degrees of freedom of the bath bounded by 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.

II General Hamiltonian for KHOF systems
For the sake of generality, we consider a KHOF system in spatial dimensions, whose Hamiltonian is given by
(2) |
where is the bath Hamiltonian, and is the interaction Hamiltonian. is given by
(3) |
where , and are conduction electron creation and annihilation operators respectively, with labelling the up/down spin index, labelling the channel index, and labelling the band index. Summation over repeated indices is implied.
We consider a general higher-order dispersion of the form
(4) |
where . The constants and satisfy , . The dispersion (4) is of particular interest, because for , it reduces to a cubic fermion system in three-dimensions (1). For , 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 , it describes even higher order emergent fermions which can be realized in artificial electric circuits with auxiliary dimensions Yu et al. (2020).
describes the exchange interaction between an impurity spin and the bath electrons:
(5) |
where is the spin of the magnetic impurity, is the Kondo coupling constant for the -th channel.
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 in (4), so that
(6) |
More details of the derivations in this section can be found in Appendix B. We shall treat the anisotropic case in Sec.V.
We expand as a linear combination of the spherical harmonics, which form a complete set of orthonormal functions on the sphere :
(7) |
Here are the spherical harmonics in dimensions, see Appendix A and Higuchi (1987) for detailed derivations of their important properties relevant for this work. are the angular coordinates of , with , for . The integers denote different partial waves, and are the analogues of the quantum numbers in the three-dimensional spherical harmonics : in 3 dimensions, , , is the azimuthal angle, and is the polar angle. and can be written in terms of as
(8) |
(9) |
Next we perform a change of variables on the fields from to , by defining
(10) |
In this way, the fields also satisfy the proper fermionic anti-commutation relations. With the definition
(11) |
we combine the fields from the + and - bands into one single composite fermionic field . In terms of , and become
(12) |
(13) |
where . Notice that we have combined the and bands into a single band, and eliminated the band index from our model.
We then define the left and right moving fields, respectively, as
(14) |
Also, by introducing the imaginary time , we define the complex plane , to which we map our KHOF model, by
(15) |
i.e. The horizontal axis of is the imaginary time , and the vertical axis of is introduced in (III). We then view in the Heisenberg picture, so that they now have time-dependence, and live on . and are related to each other by
(16) |
so can be eliminated in terms of . and can be written in terms as as
(17) |
(18) |
The -dimensional KHOF model is thus mapped into the complex plane defined in (15), with and 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 only couples to the new fermionic field at . This means that is only confined to the boundary , with in the bulk , 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 in the momentum or energy. Our integrals and 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 . 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 physics is the infinite cylinder with circumference . 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 , whose Green’s functions allow us to compute the thermodynamic quantities of interest. The circumference of the cylinder enters the calculation of the thermodynamic quantities as a finite size of the system, giving rise to their -dependences. For example, the resistivity is found as in the Fermi liquid case and in the non-Fermi liquid case. Here is the number of channels, with being the impurity spin, is the unitary limit resistivity, i.e. the greatest resistivity possibly achievable, and is a dimensionless constant, with in the special case . These results are in excellent agreements with results obtained from numerical renormalization group (NRG) analysis.
We remark that, although our 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 about the Fermi surface. Thus in their calculated thermodynamic quantities, AL can only consider temperatures satisfying , and ignore terms of order . 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 . 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 and in (4). More details of the derivations can be found in Appendix C.
For anisotropic KHOF systems, remains the same as before, since the dispersion relation does not enter the expression of . However, can no longer be reduced to the simple form (9), due to the lack of spherical symmetry. In general, we can express into a matrix form in the partial wave basis, i.e.,
(19) |
where is a column vector whose th element is given by , where for brevity we have used the multi-index notation . is the Hermitian conjugate of and is a matrix whose th element, , describes the coupling between the and th partial waves. We remark that for all Hermitian systems, the only partial waves that couple to the impurity are those with . Thus the multi-index only includes . In particular, in three dimensions, only includes , i.e. the multi-index reduces to the single index : .

We separate in (19) into two terms
(20) |
where the term consists of only the partial waves, thus describing the “isotropic part” of . consists of all the remaining partial waves with , and captures the “anisotropic part” of , as well as its couplings to , 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 can be casted into an additional renormalization term , which serves as a modification to the isotropic part of the dispersion relation .
To further enable a CFT analysis, we approximate the value of by evaluating it at the Fermi wave number, . 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 , which contains the isotropic part of the dispersion, as well as the corrections from the anisotropy. Correspondingly, the bath is now effectively described by
(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 and turning on in (4), namely
(22) |
which is linear in the 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 increases. As a result, in (19) can be casted into the simple tridiagonal form
(28) |
We then integrate out the higher-order partial waves by down-folding the matrix to the isotropic subspace. Interestingly, because each entry of is proportional to , the renormalization is also found to be proportional to , and thus vanishes at the DP with . 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
(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. and 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 -dimensional spherical harmonics used in the main section. Some of this material can also be found in Higuchi (1987).
The -dimensional spherical harmonics are the eigenfunctions of , the Laplace-Beltrami operator on the sphere , which is the angular part of the -dimensional Laplacian operator. The Laplace-Beltrami operator can be defined recursively:
(30) |
for , and
(31) |
where are the angular coordinates in the -dimensional spherical coordinates, with , . An -dimensional spherical harmonic of polynomial degree , where is a non-negative integer, has eigenvalue :
(32) |
This equation can be solved by separation of variables. By writing
(33) |
where is a spherical harmonic of degree on , and is a function that only depends on , to be further determined. Because is a factor of , its polynomial degree must not be larger than that of , thus . We can now write (A) as
(34) |
The first two terms only depend on , and the third term only depend on . Since is a spherical harmonic of degree on , we have
(35) |
so the third term in (A) must equal to . As a result, the first two terms of (A) must equal to :
(36) |
Now define
(37) |
where is the associated Legendre function of the first kind. In our case are generalized to take on integer or half-integer values. By noting that solves the Legendre equation:
(38) |
the solution to equation (A) is
(39) |
Equation (A) is in the same form as (A) and can be solved recursively by repeating the procedure up to now, yielding and . Eventually we obtain , and the equation
(40) |
where is given by (31). Its solution is
(41) |
Thus we arrive at
(42) |
where . The integers are analogues of the quantum numbers in the 3-dimensional spherical harmonics . Indeed, is due to the same reason as in the 3D case. In particular, in 3D, , is the azimuthal angle, and is the polar angle. Also, is a normalization constant to be determined. Before determining , let us show that the -dimensional spherical harmonics satisfy the orthonormality condition:
(43) |
The normality condition in (A) can help us in determining the normalization constant . Consider the integral
(44) |
The integral is
(45) |
The integral, for , is
(46) |
We first tackle the case in which we have at least one . Let be the smallest integer for which occurs. If , then by (45), (A) integrates to 0. If , because is the smallest integer such that occurs, we have . We consider (46) in the special case , which is shown in (Higuchi, 1987) to equal to
(47) |
Thus (46) again integrates to 0, and orthogonality in (A) is proven. Next, by imposing normality in (A), we determine the normalization constant . Consider the special case where for all . By (45) and (47), (A) integrates to . We want this to normalize to 1, so
(48) |
Thus, the -dimensional spherical harmonics take the form
(49) |
and satisfy the orthonormality condition (A).
Next we prove another useful identity
(50) |
This is a result of the completeness property of the spherical harmonics. By completeness, we can express any function on as a linear combination of :
(51) |
Using the orthonormality condition (A), we can determine the coefficients by
(52) |
Substitute (A) into (51), and also by noting that
(53) |
we get
(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
(55) |
From (49), we know that
(56) |
is a constant, where is the Gamma function. By the orthonormality condition (A), it is the only -dimensional spherical harmonic that is a constant. Thus
(57) |
Next, consider the integral
(58) |
where has at least one . By the orthogonality condition (A), this integral equals 0. Thus we have
(59) |
Since 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 in the Hamiltonians (2) and (4) of the main text are fermionic fields, so they satisfy the usual anticommutation relation
(60) |
in dimensions. We then expand as a linear combination of the dimensional spherical harmonics, which form a complete set of orthonormal functions on the -sphere . The expansion is given by equation (6) of the main text:
(61) |
where are the spherical harmonics in dimensions discussed in the previous section. Note that the coefficients in the expansion (B.1) are only functions of , and do not depend on any of the angular coordinates . Also, a factor of has been factored out from each coefficient in the expansion, so that the remaining part of the coefficient, the fields , satisfy the proper fermionic anticommutation relation
(62) |
Also, one can show that are related to by
(63) |
We shall write the Hamiltonian in terms of these new fermionic fields . By substituting (B.1) into given by equation (4) of the main text, and by using and (A), can be written into the form of equation (7) of the main text:
(64) |
This shows that the only partial waves that couple to the impurity are those with .
Similarly, substitute (B.1) into given in equation (2) of the main text, and use the orthonormality condition (A), we get
(65) |
Equation (B.1) shows that the only partial waves that couple to the impurity in are those with , 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 will couple to the impurity, either directly or indirectly. As a result, we only need to consider partial waves with in . Thus reduces to equation (8) of the main text:
(66) |
B.2 Derivations of the anticommutation relations for the and the fields
In this section, we show that the field defined in equation (9) of the main text and the field defined in equation (10) of the main text are fermionic fields satisfying the proper anticommutation relations. Due to the delta function identity
(67) |
the field obeys the anticommutation relation
(68) |
As a result, the composite field satisfies the anticommutation relation
(69) |
as well.
B.3 Writing the Hamiltonian in terms of the composite field
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 . 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 is positive), we get
(70) |
where we have relabelled the dummy variables and in the first term and in the second term respectively. Then by using the definition of given by equation (10) of the main text, becomes
(71) |
which is equation (11) of the main text.
As for , 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
(72) |
where
(73) |
Using equation (5) of the main text, and cancel off each other, and similarly and cancel off each other. We get
(74) |
where . Writing out the sum over explicitly and relabelling the dummy variables and , we get
(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 be the imaginary time , and define the complex plane to which we map our KHOF model as equation (14) of the main text,
(76) |
View the fields defined in equation (13) of the main text in the Heisenberg picture, so that they now have time-dependence, and thus live on . Using equation (13) of the main text, we can see that 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, given by equation (11) of the main text can be written as
(77) |
We note that in (B.4). One can also show that satisfy the anticommutation relations
(78) |
Now equation (17) of the main text tells us that is only confined to the horizontal axis . At , and we have the Heisenberg equations of motion
(79) |
for each . By substituting in the form of (B.4) into the Heisenberg equation of motion (79), and applying the anticommutation relation (78) and the identity for three arbitrary operators and , the Heisenberg equation of motion reduces to
(80) |
These are exactly the Cauchy-Riemann equations for holomorphic functions and antiholomorphic functions respectively, which imply that is a holomorphic function, and is an antiholomorphic function, on the upper half plane (since in (B.4), is non-negative, live on the upper half plane). Equivalently, is a holomorphic function on the lower complex plane . Also, by the definition of given by equation (13) of the main text, , i.e. the two holomorphic functions on the upper complex plane and on the lower complex plane agree on the horizontal -axis . Thus they are analytic continuations of each other to the entire complex plane. This fact allow us to eliminate in terms of because the former is simply the analytic continuation of the latter into the lower complex plane:
(81) |
Thus (B.4) becomes
(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 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 .
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 , remains unchanged, since the dispersion relation does not enter the expression of . However, now in , we not only have to expand as a linear combination of the spherical harmonics according to (B.1), but also need to do so for as well:
(83) |
Substitute (B.1), (83) into equation (2) of the main text, we get
(84) |
where for brevity we used the multi-index notation . The coupling strengths are given by the hopping integral
(85) |
We note that depends on the dispersion relation of the system, because different observes different expansions in terms of spherical harmonics (83), resulting in different sets of appearing in (C). As a concrete example, we calculate for the dispersion relation given by equation (21) of the main text in 3 dimensions,
(86) |
where are real constants satisfying
(87) |
so that
(88) |
WLOG align in the -direction, so . Thus (86) reads
(89) |
whose expansion (83) in terms of spherical harmonics in 3 dimensions is
(90) |
This shows that for 3-dimensional systems with dispersion relation (86), the only non-zero spherical harmonics in the expansion (83) are and , which enter (C) as . Thus the only non-zero are and . They are given by
(91) |
where we have used and the orthogonality condition (A), and
(92) |
where we have used
(93) |
in the -integral, the orthogonality relation
(94) |
and the recurrence formula
(95) |
for the associated Legendre functions of the first kind, , in the -integral.
The fact that the only non-zero coupling strengths 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 . We have showed in (III) that the impurity only couples to the partial wave. Anisotropy does not change this fact, since anisotropy does not alter , as previously mentioned. Now (91) and (C) both have a factor of , implying that there is no coupling between partial waves with . Thus the partial wave that couples to the impurity will only couple to other partial waves with . We can hence ignore all partial waves with from . 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, is real, so it can be expanded in terms of real spherical harmonics. Thus all in the hopping integral (C) are real, and are independent of : indeed, from (49), we see that real spherical harmonics have no dependence. Thus the -integral in (C) only involve and , and equals to , which forbids coupling between partial waves with different , which is the analogue of the quantum number in three dimensions. (ii) There is only nearest-neighbor hopping. This can be seen from the factor in the braces of (C): and implies hopping can only occur between partial waves whose and differ by 1, resulting in only nearest-neighbor hopping. (iii) The nearest-neighbor hopping strength in (ii) decays with increasing . Due to (i), we are only interested in the partial waves. Substitute into (C), we obtain
(96) |
In other words, the coupling between the adjacent -th and -th partial waves is given by
(97) |
By computing its derivative with respect to , we see that the coupling strength monotonically decreases as increases, for all .
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 increases.
It is instructive to express the general in (84) into a matrix form in the partial wave basis, as shown in equation (18) of the main text:
(98) |
where is a matrix whose th element, , is given by
(99) |
is a column vector whose th element is given by , and is its Hermitian conjugate. Recall that property (i), which holds for all Hermitian systems, allows us to only include partial waves with in . Thus here, the multi-indices and only includes . In particular, in 3 dimensions, we have the property that the multi-indices and only includes , i.e. in 3 dimensions the multi-indices and reduce to the single indices and : , .
In our specific example (86), is the tridiagonal matrix:
(105) |
where the partitioning denoted by the horizontal and vertical lines will be explained shortly. We have due to (87). In general, is a matrix, but by property (iii), since the hopping strength decays as the row/column number increases, in practice we can truncate it to an appropriate size for further calculations.
We next partition as shown in (105): the first block consists of only the (1,1)-th element, the second block consists of the remaining square matrix, which we denote by , and the third and fourth parts consist of the column vector, denoted by , and its transpose, the row vector respectively. This corresponds to separating the matrix form of in (19) into four terms:
(106) |
We shall now formally define the objects that appear in (C). First of all, define the isotropic kernel of the dispersion relation as
(107) |
can be seen as the “isotropic part” of the dispersion relation. Next, let be the multi-index obtained from by excluding . is the column vector whose th element is given by , i.e. is obtained from by excluding its first entry . is the Hermitian conjugate of , i.e. is the row vector obtained from by excluding its first entry . is the matrix whose th element is defined by
(108) |
i.e. is the matrix obtained from by removing its first row and first column, which corresponds to and respectively. In our example, is the bottom right block matrix in (105). is the column vector whose th entry is given by
(109) |
where in the third equality we have used the fact that is a real constant given by (56), in the fourth equality we have expanded the dispersion relation in terms of the spherical harmonics (83), and in the last equality we have used the orthogonality relation (A). Recall that does not include . In our 3-dimensional example, the only non-zero is , according to (90). Thus for our example
(110) |
as can be seen from the first column of the matrix (105), excluding its first entry. is the transpose pf . Also in the derivation of the first term of (C) we have used
where in the last equality, we have used the definition of the kernel in (107).
Let us define the first term of in (C) to be , and define the sum of the remaining three terms of in (C) to be :
(112) |
(113) |
(114) |
We make these definitions because is the “isotropic part” of , whereas is the “anisotropic part” of . Indeed, only includes the partial wave, and contains the remaining partial waves. For systems with isotropic dispersion relations, only the partial wave couples to the impurity, and we can ignore all partial waves, so and . All contributions to the Hamiltonian due to anisotropy of the system are captured in .
We shall now write the partition function for our system. Let be the Grassmann variable obtained from the operator acting on a coherent state. We have
where is the part of the action that only involves the partial wave, and is the part of the action that involves the remaining partial waves. As previously discussed, in our Hamiltonian , and only includes the partial wave, so and appears in , whereas includes the remaining partial waves, so appears in . We thus have
and
where is the column vector whose th element is given by , is the row vector whose th element is given by ,
(118) |
and
(119) |
We then integrate out the Grassmann field , in (C), after which the second factor of (C) becomes
(120) |
where we have taken the static limit, , 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 is gapped. Notice that the factor in Eq. (C) now only includes the partial wave. Thus the partition function (C) becomes
(121) |
where
(122) |
This means that contributes an additional term of
(123) |
to the braced integrand of given in (C). In terms of the Hamiltonian, this term is
(124) |
Thus we can write as
(125) |
where in the second equality we have used the definition of given in (113), in the third equality we have made the definition
(126) |
and in the last equality we have made the definition
(127) |
We call the effective dispersion relation of the system.
We note that the kernel satisfies the relation
(128) |
because
(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 . Similarly, satisfies the relation
(130) |
To see this, notice that in the definition of given in (126), each of the factors and flips a sign when flips a sign: indeed, the entries of and are just the entries of the matrix defined in (C), whose integrand contains a factor of , which flips a sign as flips a sign. By matrix inverse properties we also have
(131) |
completing the proof. Thus the effective dispersion relation also satisfies the relation
(132) |
Let us compute the effective dispersion relation for our example. The kernel is given by
(133) |
As for defined in (126), since in our example only has its first entry being non-zero, as seen from (110), becomes
(134) |
where denotes the (1,1)-th element of . Recall that is the lower right block matrix of in (105):
(139) |
By the discussion below (105), we may truncate , and thus , to a desired size for actual calculations. Let denote the truncated matrix of , i.e. is obtained from by taking only its first rows and first columns. Because is a tridiagonal matrix, one can show that
(140) |
where is the determinant of , which can be computed from the recurrence relation
(141) |
with , and is a polynomial in that satisfies the same recurrence relation
(142) |
but with different initial terms . Thus in our example,
(143) |
where .
We view the term , which captures all anisotropic effects, as a correction to the isotropic term of the dispersion relation, the kernel . We approximate by evaluating its value at , before adding it to the kernel . satisfies the property that , i.e. is a non-negative real root of ( needs to be non-negative because it is a special value of ). In general, may have no non-negative real root, in which case the system is gapped and we do not consider such gapped systems. Also, may have more than one non-negative real root. In such cases we pick the non-negative real root of that gives rise to the greatest density of states (DOS) to be the at which we evaluate . In other words, we pick the non-negative real root that results in the least . If there exists more than one non-negative real root of that give rise to the same greatest DOS, we evaluate at each of them - this will result in channel-multiplying.
In our example, given by (143) has more than one non-negative real root, with the smallest of them giving rise to the least , thus the greatest DOS. Hence there is only one at which we evaluate , namely this smallest non-negative real root of . For all even values of , this smallest non-negative real root of is 0. For odd values of , as , the smallest non-negative real root also approaches 0. Thus for our example, , , and our effective dispersion relation becomes
(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
- Kondo (1964) J. Kondo, Progress of Theoretical Physics 32, 37 (1964).
- Affleck and Ludwig (1991a) I. Affleck and A. W. W. Ludwig, Nuclear Physics B 352, 849 (1991a).
- Affleck and Ludwig (1991b) I. Affleck and A. W. W. Ludwig, Nuclear Physics B 360, 641 (1991b).
- Affleck and Ludwig (1991c) I. Affleck and A. W. W. Ludwig, Phys. Rev. Lett. 67, 3160 (1991c).
- Affleck et al. (1992) I. Affleck, A. W. W. Ludwig, H.-B. Pang, and D. L. Cox, Physical Review B 45, 7918 (1992).
- Affleck and Ludwig (1993) I. Affleck and A. W. W. Ludwig, Phys. Rev. B 48, 7297 (1993).
- Ludwig (1994) A. W. W. Ludwig, International Journal of Modern Physics B 8, 347 (1994).
- Affleck (1995) I. Affleck, arXiv e-prints , cond-mat/9512099 (1995), arXiv:cond-mat/9512099 [cond-mat] .
- Delft (1995) J. V. Delft, 2-CHANNEL Kondo Scaling in Metal Nanoconstrictions: a Conformal Field Theory Calculation of Scaling Function., Ph.D. thesis, CORNELL UNIVERSITY. (1995).
- Yu et al. (2020a) Z. Yu, F. Zamani, P. Ribeiro, and S. Kirchner, Phys. Rev. B 102, 115124 (2020a).
- Fang et al. (2012) C. Fang, M. J. Gilbert, X. Dai, and B. A. Bernevig, Phys. Rev. Lett. 108, 266802 (2012).
- Yang and Nagaosa (2014) B.-J. Yang and N. Nagaosa, Nature Communications 5, 4898 (2014).
- Yang et al. (2015) B.-J. Yang, T. Morimoto, and A. Furusaki, Phys. Rev. B 92, 165120 (2015).
- Huang et al. (2016) S.-M. Huang, S.-Y. Xu, I. Belopolski, C.-C. Lee, G. Chang, T.-R. Chang, B. Wang, N. Alidoust, G. Bian, M. Neupane, D. Sanchez, H. Zheng, H.-T. Jeng, A. Bansil, T. Neupert, H. Lin, and M. Z. Hasan, Proceedings of the National Academy of Sciences of the United States of America 113, 1180—1185 (2016).
- Gao et al. (2016) Z. Gao, M. Hua, H. Zhang, and X. Zhang, Phys. Rev. B 93, 205109 (2016).
- Bradlyn et al. (2016) B. Bradlyn, J. Cano, Z. Wang, M. G. Vergniory, C. Felser, R. J. Cava, and B. A. Bernevig, Science 353, 10.1126/science.aaf5037 (2016).
- Liu and Zunger (2017) Q. Liu and A. Zunger, Phys. Rev. X 7, 021019 (2017).
- Ezawa (2017) M. Ezawa, Phys. Rev. B 96, 161202 (2017).
- Yu et al. (2018) W. C. Yu, X. Zhou, F.-C. Chuang, S. A. Yang, H. Lin, and A. Bansil, Phys. Rev. Materials 2, 051201 (2018).
- Yu et al. (2019) Z.-M. Yu, W. Wu, X.-L. Sheng, Y. X. Zhao, and S. A. Yang, Phys. Rev. B 99, 121106 (2019).
- Isobe and Fu (2019) H. Isobe and L. Fu, Phys. Rev. Research 1, 033206 (2019).
- Song et al. (2020) Z. Song, B. Li, C. Xu, S. Wu, B. Qian, T. Chen, P. K. Biswas, X. Xu, and J. Sun, Journal of Physics: Condensed Matter 32, 215402 (2020).
- Wu et al. (2020) W. Wu, Z.-M. Yu, X. Zhou, Y. X. Zhao, and S. A. Yang, Phys. Rev. B 101, 205134 (2020).
- Wang et al. (2020) J.-R. Wang, W. Li, and C.-J. Zhang, Phys. Rev. B 102, 085132 (2020).
- Wu et al. (2021) W. Wu, Y. Liu, Z.-M. Yu, Y. X. Zhao, W. Gao, and S. A. Yang, Higher-order nodal points in two dimensions (2021), arXiv:2105.08424 [cond-mat.mes-hall] .
- Potel et al. (1980) M. Potel, R. Chevrel, M. Sergent, J. Armici, M. Decroux, and . Fischer, Journal of Solid State Chemistry 35, 286 (1980).
- Armici et al. (1980) J. Armici, M. Decroux, . Fischer, M. Potel, R. Chevrel, and M. Sergent, Solid State Communications 33, 607 (1980).
- Huang et al. (1983) S. Huang, J. Mayerle, R. Greene, M. Wu, and C. Chu, Solid State Communications 45, 749 (1983).
- Tarascon et al. (1984) J. Tarascon, F. DiSalvo, and J. Waszczak, Solid State Communications 52, 227 (1984).
- Hor et al. (1985a) P. Hor, R. Meng, C. Chu, J. Tarascon, and M. Wu, Physica B+C 135, 245 (1985a).
- Hor et al. (1985b) P. Hor, W. Fan, L. Chou, R. Meng, C. Chu, J. Tarascon, and M. Wu, Solid State Communications 55, 231 (1985b).
- Brusetti et al. (1988) R. Brusetti, P. Monceau, M. Potel, P. Gougeon, and M. Sergent, Solid State Communications 66, 181 (1988).
- Tessema et al. (1991) G. X. Tessema, Y. T. Tseng, M. J. Skove, E. P. Stillwell, R. Brusetti, P. Monceau, M. Potel, and P. Gougeon, Phys. Rev. B 43, 3434 (1991).
- Brusetti et al. (1994) R. Brusetti, A. Briggs, O. Laborde, M. Potel, and P. Gougeon, Phys. Rev. B 49, 8931 (1994).
- Petrović et al. (2010) A. P. Petrović, R. Lortz, G. Santi, M. Decroux, H. Monnard, O. Fischer, L. Boeri, O. K. Andersen, J. Kortus, D. Salloum, P. Gougeon, and M. Potel, Phys. Rev. B 82, 235128 (2010).
- Sun et al. (2009) K. Sun, H. Yao, E. Fradkin, and S. A. Kivelson, Phys. Rev. Lett. 103, 046811 (2009).
- Uebelacker and Honerkamp (2011) S. Uebelacker and C. Honerkamp, Phys. Rev. B 84, 205122 (2011).
- Wang et al. (2015) R. Wang, B. Wang, L. Sheng, D. Y. Xing, and J. Wang, Phys. Rev. B 92, 195151 (2015).
- Yu et al. (2020) R. Yu, Y. X. Zhao, and A. P. Schnyder, National Science Review 7, 1288 (2020), https://academic.oup.com/nsr/article-pdf/7/8/1288/38882379/nwaa065.pdf .
- Higuchi (1987) A. Higuchi, Journal of Mathematical Physics 28, 1553 (1987).