Emergent Spinon Dispersion and Symmetry Breaking in Two-Channel Kondo Lattices
Abstract
Two-channel Kondo lattice serves as a model for a growing family of heavy-fermion compounds. We employ a dynamical large- technique and go beyond the independent bath approximation to study this model both numerically and analytically using renormalization group ideas. We show that the Kondo effect induces dynamic magnetic correlations that lead to an emergent spinon dispersion. Furthermore, we develop a quantitative framework that interpolates between infinite dimension where the channel-symmetry broken results of mean-field theory are confirmed, and one-dimension where the channel symmetry is restored and a critical fractionalized mode is found.
The screening of a magnetic impurity by the conduction electrons in a metal is governed by the Kondo effect. The multichannel version is when several channels compete for a single impurity, as a result of which the spin is frustrated and a new critical ground state formed with a fractional residual impurity entropy. In the two-channel case, this entropy corresponds to a Majorana fermion. If the channel symmetry is broken, the weaker channels decouple and the stronger-coupled channels win to screen the impurity at low temperature Andrei and Destri (1984); Affleck et al. (1992); Emery and Kivelson (1992); Affleck and Ludwig (1993).
While the case of a single impurity is well understood, much less is known about Kondo lattices where a lattice of spins is screened by conduction electrons Hewson (1993); Si et al. (2014); Coleman (2015), especially if multiple conduction channels are involved Cox and Jarrell (1996). The most established fact is the prediction of a large Fermi surface (FS) in the Kondo-dominated regime of the single-channel Kondo lattice Oshikawa (2000). In the multichannel case, the continuous channel symmetry naturally leads to new patterns of entanglement which are potentially responsible for the non-Fermi liquid physics Jarrell et al. (1996, 1997), symmetry breaking, and possibly fractionalized order parameter Komijani et al. (2018). This partly arises from the fact that the residual entropy seen in the impurity has to eventually disappear at zero temperature in the case of a lattice.
Beside fundamental interest, a pressing reason for studying this physics is that the multichannel Kondo lattice (MCKL), and in particular 2CKL, seems to be an appropriate model for several heavy-fermion compounds, e.g. the family of PrTr2Zn20 (Tr=Ir,Rh) Onimaru and Kusunose (2016); Patri and Kim (2020) as well as recent proposals that MCKLs may support nontrivial topology Hu et al. (2021); Kornjaca et al. (2021) and non-Abelian Kondo anyons Lopes et al. (2020); Komijani (2020).
The MCKL model is described by the Hamiltonian
(1) |
where is the Hamiltonian of the conduction electrons and Einstein summation over spin and channel indices is assumed. This model has SU() spin and SU() channel symmetries and we are interested in analyzing the effect of a channel symmetry breaking , where and ’s act as Pauli matrices in the channel space Hoshino et al. (2011). At first look, at least certain deformation of the MCKL can be thought of as a channel magnet. (A naïve strong coupling limit is not a spin-singlet, but the Nozières doublet. See supplementary materials SM for a deformation that changes this.) In the limit SM , the spin is quenched due to formation of Kondo singlet with either (for ) of the channels, leading to a doublet over which acts like Schauerte et al. (2005); SM . Interaction among adjacent doublets leads to a “channel magnet” . While channel Weiss-field favors a channel anti-ferromagnetic (channel AFM) super-exchange interaction, the mean-field theory predicts a variety of channel ferromagnetic (channel FM) and channel AFM solutions [Fig. 1(b)] depending on the conduction filling.

On the other hand, some differences to a channel magnet are expected since the winning channel has a larger FS Komijani et al. (2018); Wugalter et al. (2020) and the order parameter is strongly dissipated by coupling to fermionic degrees of freedom. Although a channel-symmetry broken ground state is predicted by both single-site dynamical mean-field theory (DMFT) Hoshino et al. (2011); Hoshino and Kuramoto (2015) and static mean-field theory Van Dyke et al. (2019); Zhang et al. (2018); Wugalter et al. (2020), it has not been observed in recent cluster DMFT studies Inui and Motome (2020). Furthermore, the effective theory of fluctuations in the large- limit Wugalter et al. (2020) predicts a disordered phase below the lower critical dimension but the nature of this quantum paramagnet is unclear. In 1D, Andrei and Orignac have used non-Abelian bosonization to show Andrei and Orignac (2000) that the ground state is gapless and fractionalized (dispersing Majoranas for ), a prediction that contradicts the analysis by Emery and Kivelson Emery and Kivelson (1993), and has not been confirmed by the density matrix renormalization group calculations Schauerte et al. (2005).
Resolving these issues requires a technique that is applicable to arbitrary dimensions and goes beyond static mean field and DMFT by capturing both quantum and spatial fluctuations. Here, we show that the dynamical large- approach, recently applied successfully to study Kondo lattices Rech et al. (2006); Komijani and Coleman (2018, 2019); Wang et al. (2020); Shen et al. (2020); Wang and Yang (2021); Drouin-Touchette et al. (2021); Han et al. (2021), is precisely such a technique.
We assume the spins transform as a spin- representation of SU(). In the impurity case We use 0D to refer to the impurity problem , the spin is fully screened for whereas it is overscreened and underscreened for and , respectively Zinn-Justin and Andrei (1998). The focus of this Letter is on the Kondo-dominated regime of the double-screened case which is schematically shown in Fig. 1(a). We use Schwinger bosons to form a symmetric representation of spins with the size . We then rescale and treat the model (1) in the large- limit, by sending , but keeping and constant. The constraint is imposed on average via a uniform Lagrange multiplier .
In the present large- limit, the Ruderman-Kittel-Kasuya-Yosida (RKKY) interaction is [inset of Fig. 2(a)] and we need to include an explicit Heisenberg interaction between nearest neighbors to couple the impurities. Nevertheless, we will show that an infinitesimal is sufficient to produce significant magnetic correlations due to a novel variant of RKKY interaction. For simplicity we limit ourselves to ferromagnetic correlations .
For a site lattice, the Lagrangian becomes Parcollet and Georges (1997); Komijani and Coleman (2018)
Here, ’s are bosonic spinons and ’s are Grassmannian holons that mediate the local Kondo interaction. In momentum space, the electrons and bosons have dispersions and , respectively. is the (assumed to be homogeneous) nearest neighbor hopping of spinons due to large- decoupling of term Komijani and Coleman (2018). Here, we focus on a half-filled conduction band , but similar results are obtained at other commensurate fillings SM . In the large- limit the dynamics is dominated by the non-crossing Feynman diagrams, resulting in boson and holon self-energies []
(3) |
whereas is and thus the electrons propagator remains bare, with complex frequency. Equations (3) together with the Dyson equations and form a set of coupled integral equations that are solved iteratively and self-consistently, while is adjusted to satisfy the constraint. Thermodynamic variables are then computed from Green’s functions Rech et al. (2006); Komijani and Coleman (2018).
First, we study the case in which is absent, or . In this limit, the self-energies remain local and the problem reduces to the impurity problem Parcollet and Georges (1997). It has never been studied whether the large- overscreened impurities are susceptible to symmetry breaking Affleck et al. (1992). To do so, we assume that half of channels are coupled to the impurity with and the other half with . This corresponds to a uniform symmetry breaking deformation of the Lagrangian.

Figure 1(c) shows the entropy of the 2CK impurity model as a function of channel asymmetry, verifying that the impurity is indeed critical with respect to channel symmetry breaking. In symmetric 2CK, the ground state entropy at large- is fractional with a universal dependence on Parcollet and Georges (1997); SM .
Next, we focus on finite case for two settings of 1D and , which correspond to a Bethe lattice with coordination numbers and . In 1D, and depend on and , but in , self-energies have no spatial dependence and the Green’s functions of spinons/electrons obey .

Importantly, the criticality of overscreened impurity solution ensures that an infinitesimal spinon hopping seed can get an O(1) amplification [inset of Fig. 2(a)] and dispersions for spinons and holons are dynamically generated. Restricting ourselves to translationally invariant solutions with lattice periodicity , this effect can be succinctly represented by the zero-frequency spinon and holon effective dispersion and , shown in Figs. 2(a) and 2(b) for various temperatures. This emergent spinon dispersion is independent of the choice of the seed and agrees qualitatively with the finite results SM . The consumption of the residual entropy in the lattice by the emerging dispersion is visible in the inset of Fig. 2(b). We stress that in 1D, this apparent transition most likely becomes a crossover when N is finite Read (1985). In the case of D, the system is prone to spin or channel magnetization, as discussed later. Such symmetry breakings would consume the residual entropy SM .
Figures 3(a) and 3(b) shows the finite frequency spectral function of spinons and holons, respectively. Both are dominated by a sharp mode with emergent Lorentz invariance. The spinons are gapless and linearly dispersing and the holons form a FS. The temperature collapse of Fig. 3(c) confirms that the spectra are critical with the local spectra obeying a behavior. Figure 3(d) shows similar collapse for the case of infinite-coordination Bethe lattice (D). A marked difference between the two cases is that for 1D, which leads to minima at , whereas in D, manifested as a peak at .
What is the effect of channel symmetry breaking on the volume of FS? According to Luttinger’s theorem, the FS volume is related to electron phase shift for a dimensional lattice. From case of the Ward identity Coleman et al. (2005), the electron phase shift is related to that of holons , which itself is defined as
(4) |
The locus of points at which changes sign defines a holon FS which generalizes to any dimension. In 1D, holons are occupied for . So, we find that and the total change in electron FS is , corresponding to a large FS in the critical phase. We use Eq. (4) to study the effect of a uniform symmetry breaking field . Figure 4(a) shows how FSs of slightly favored and disfavored channels evolve as a function of in the two cases. In 1D, the FS asymmetry disappears, restoring a channel symmetric criticality at low , consistent with the Mermin-Wagner theorem. On the other hand, in D the asymmetry grows and one channel totally decouples from the spins, with gapped spinons and also gapped holons for both channels. The exponents are related to ; varying in Eq. (4) we find
(5) |
Assuming , the holon FS is unstable against symmetry breaking when diverges. This regime coincides with when the symmetry breaking term is relevant, in the renormalization group (RG) sense. On the other hand instability of the entire holon FS requires the divergence of , i.e. which is a more stringent condition and agrees with Fig. 4(a), confirming as the marginal dimension.
Figure 4(a) shows that the symmetry breaking is relevant in D, but is irrelevant in 1D. To establish this from the microscopic model, one has to access the infrared (IR) fixed point. From the numerics we see that the system flows to a critical IR fixed point, in which spinons and holons are critical in addition to electrons. For an impurity and are reasonable at . The exponents are known Parcollet and Georges (1997); SM :
(6) |
and coincide with those of the D in the small regime we are interested here SM . In the presence of a dimensionless , the RG analysis predicts a dynamical scale [cf. Fig. 1(c)].
The 1D case is more subtle; as , we see from Fig. 2 that and at the IR fixed point This is not the case for the spinon’s -mode, ≠0() (±π/a). This means that the Kondo coupling flows to strong coupling at , to weak coupling at , and gets critical at , while the spinons are gapless at . At these momenta, the Dyson equation has the scale-invariant form .
We can obtain a low-energy description by expanding fields near zero energy, e.g. for electrons and holons. In 1+1 dimensions, the conformal invariance of the fixed point dictates the following form for the Green’s functions :
(7) |
where and . The is obtained from by . These Green’s functions can be conformally mapped to finite- via replacement. Furthermore, in terms of , they have the Fourier transforms:
(8) | |||||
where . From matching the powers of frequency in Eqs. (3), (7) and (8), we conclude that in order to satisfy the self-consistency. Moreover, from the matching of the amplitudes of the Green’s functions we find SM
(9) |
Note that , ensuring that channel symmetry breaking perturbations are irrelevant in 1D. These are in excellent agreement with the exponents extracted from scaling [Fig. 4(b)] and we have established a semianalytical framework to interpolate between 1D and .
The emergent dispersion in Fig. 2, the scaling dimensions in Eq. (9), and their relation to symmetry breaking in Fig. 4 are the central results of this Letter. In the following we discuss some of the implications of these results for physical observables that are independent of our fractionalized description, leaving the details to SM .

The fractionalization or contraction are related to order parameter fractionalization Komijani et al. (2018); Tsvelik and Coleman (2021). In the long time/distance limit, correlation functions of and that of are given by and , respectively and thus, have exponents that add up to zero. On the other hand, correlators of gauge-invariant operators and are exactly equal since both can be constructed by taking derivatives of free energy with respect to , either before or after Hubbard-Stratonovitch transformation. A diagrammatic proof of this equivalence is provided in SM . Scaling analysis gives and up to a constant shift coming from the regular part of free energy.
Another nontrivial feature of 2CK impurity fixed point is its magnetic instability Affleck et al. (1992) whose large- incarnation is for the impurity (or ) in Eq. (6). From Eq. (9), we see that this also holds for 1D 2CKL for . This is reflected in the divergence of the uniform static magnetic susceptibilities as a function of . Using scaling analysis and up to a constant shift, in good agreement with numerics SM . Note that this critical spin behavior is different from the gapped spin sector observed in Emery and Kivelson (1993); Schauerte et al. (2005), but is qualitatively consistent with Andrei and Orignac (2000).
Lastly, the fact that the fixed point discussed above is IR stable follows from the fact that the interaction is exactly marginal due to and that vertex corrections remain . The 1+1D correlators (7) can be obtained from three sets of decoupled Luttinger liquids for each of the fields with fine-tuned Luttinger parameters that give the correct exponents. Such a spinon-holon theory will have a Virasoro central charge . On the other hand the coset theory of Andrei and Orignac (2000); Azaria et al. (1998); SM predicts . We have used heat-capacity and the excitation velocities to compute the central charge according to as a function of and found SM . Note that there is no contradiction with the -theorem since the UV theory is not Lorentz invariant due to ferromagnetism. The discrepancy with is likely rooted in inability of Schwinger bosons to capture gapless spin liquids Arovas and Auerbach (1988).
In summary, we have shown that the dynamical large- approach can capture symmetry breaking in multichannel Kondo impurities and lattices in the presence of both emergent and induced ferromagnetic correlations within an RG framework with explicit examples on 0D, 1D, and D. The scaling analysis enables an analytical solution to the critical exponents and susceptibilities which are in good quantitative agreement with numerics, and is applicable to higher dimensional CFTs. A determination of the upper and lower critical dimensions and the effect of antiferromagnetic correlations are left to a future work Ge and Komijani .
Acknowledgment—The authors acknowledge fruitful discussions with P. Coleman and N. Andrei. This work was performed in part at Aspen Center for Physics, which is supported by NSF Grant No. PHY-1607611. Computations for this research were performed on the Advanced Research Computing Cluster at the University of Cincinnati, and the Penn State University’s Institute for Computational and Data Sciences’ Roar supercomputer.
References
- Andrei and Destri (1984) N. Andrei and C. Destri, Phys. Rev. Lett. 52, 364 (1984).
- Affleck et al. (1992) I. Affleck, A. W. W. Ludwig, H.-B. Pang, and D. L. Cox, Phys. Rev. B 45, 7918 (1992).
- Emery and Kivelson (1992) V. J. Emery and S. Kivelson, Phys. Rev. B 46, 10812 (1992).
- Affleck and Ludwig (1993) I. Affleck and A. W. W. Ludwig, Phys. Rev. B 48, 7297 (1993).
- Hewson (1993) A. C. Hewson, The Kondo Problem to Heavy Fermions, Cambridge Studies in Magnetism (Cambridge University Press, 1993).
- Si et al. (2014) Q. Si, J. H. Pixley, E. Nica, S. J. Yamamoto, P. Goswami, R. Yu, and S. Kirchner, J. Phys. Soc. Jpn. 83, 061005 (2014).
- Coleman (2015) P. Coleman, Introduction to Many-Body Physics (Cambridge University Press, 2015).
- Cox and Jarrell (1996) D. L. Cox and M. Jarrell, J. Phys. Condens. Matter 8, 9825 (1996).
- Oshikawa (2000) M. Oshikawa, Phys. Rev. Lett. 84, 3370 (2000).
- Jarrell et al. (1996) M. Jarrell, H. Pang, D. L. Cox, and K. H. Luk, Phys. Rev. Lett. 77, 1612 (1996).
- Jarrell et al. (1997) M. Jarrell, H. Pang, and D. L. Cox, Phys. Rev. Lett. 78, 1996 (1997).
- Komijani et al. (2018) Y. Komijani, A. Toth, P. Chandra, and P. Coleman, arXiv:1811.11115 (2018).
- Onimaru and Kusunose (2016) T. Onimaru and H. Kusunose, J. Phys. Soc. Jpn. 85, 082002 (2016).
- Patri and Kim (2020) A. S. Patri and Y. B. Kim, Phys. Rev. X 10, 041021 (2020).
- Hu et al. (2021) H. Hu, L. Chen, C. Setty, S. E. Grefe, A. Prokofiev, S. Kirchner, S. Paschen, J. Cano, and Q. Si, arXiv:2110.06182 (2021).
- Kornjaca et al. (2021) M. Kornjaca, V. L. Quito, and F. Rebecca, arXiv:2104.11173 (2021).
- Lopes et al. (2020) P. L. S. Lopes, I. Affleck, and E. Sela, Phys. Rev. B 101, 085141 (2020).
- Komijani (2020) Y. Komijani, Phys. Rev. B 101, 235131 (2020).
- Hoshino et al. (2011) S. Hoshino, J. Otsuki, and Y. Kuramoto, Phys. Rev. Lett. 107, 247202 (2011).
- (20) See supplementary materials .
- Schauerte et al. (2005) T. Schauerte, D. L. Cox, R. M. Noack, P. G. J. van Dongen, and C. D. Batista, Phys. Rev. Lett. 94, 147201 (2005).
- Wugalter et al. (2020) A. Wugalter, Y. Komijani, and P. Coleman, Phys. Rev. B 101, 075133 (2020).
- Hoshino and Kuramoto (2015) S. Hoshino and Y. Kuramoto, J. Phys. Conf. Ser. 592, 012098 (2015).
- Van Dyke et al. (2019) J. S. Van Dyke, G. Zhang, and R. Flint, Phys. Rev. B 100, 205122 (2019).
- Zhang et al. (2018) G. Zhang, J. S. Van Dyke, and R. Flint, Phys. Rev. B 98, 235143 (2018).
- Inui and Motome (2020) K. Inui and Y. Motome, Phys. Rev. B 102, 155126 (2020).
- Andrei and Orignac (2000) N. Andrei and E. Orignac, Phys. Rev. B 62, R3596 (2000).
- Emery and Kivelson (1993) V. J. Emery and S. A. Kivelson, Phys. Rev. Lett. 71, 3701 (1993).
- Rech et al. (2006) J. Rech, P. Coleman, G. Zarand, and O. Parcollet, Phys. Rev. Lett. 96, 016601 (2006).
- Komijani and Coleman (2018) Y. Komijani and P. Coleman, Phys. Rev. Lett. 120, 157206 (2018).
- Komijani and Coleman (2019) Y. Komijani and P. Coleman, Phys. Rev. Lett. 122, 217001 (2019).
- Wang et al. (2020) J. Wang, Y.-Y. Chang, C.-Y. Mou, S. Kirchner, and C.-H. Chung, Phys. Rev. B 102, 115133 (2020).
- Shen et al. (2020) B. Shen, Y. Zhang, Y. Komijani, M. Nicklas, R. Borth, A. Wang, Y. Chen, Z. Nie, R. Li, X. Lu, et al., Nature 579, 51 (2020).
- Wang and Yang (2021) J. Wang and Y.-f. Yang, Phys. Rev. B 104, 165120 (2021).
- Drouin-Touchette et al. (2021) V. Drouin-Touchette, E. J. König, Y. Komijani, and P. Coleman, Phys. Rev. B 103, 205147 (2021).
- Han et al. (2021) R. Han, D. Hu, J. Wang, and Y.-f. Yang, Phys. Rev. B 104, 245132 (2021).
- (37) We use 0D to refer to the impurity problem, although a conduction band exists .
- Zinn-Justin and Andrei (1998) P. Zinn-Justin and N. Andrei, Nucl. Phys. B 528, 648 (1998).
- Parcollet and Georges (1997) O. Parcollet and A. Georges, Phys. Rev. Lett. 79, 4665 (1997).
- Read (1985) N. Read, J. Phys. C 18, 2651 (1985).
- Coleman et al. (2005) P. Coleman, I. Paul, and J. Rech, Phys. Rev. B 72, 094430 (2005).
- This is not the case for the spinon’s -mode, ≠0() (±π/a) This is not the case for the spinon’s -mode, , but this mode has negligible spectral weight .
- Tsvelik and Coleman (2021) A. M. Tsvelik and P. Coleman, arXiv:2112.07781 (2021).
- Azaria et al. (1998) P. Azaria, P. Lecheminant, and A. A. Nersesyan, Phys. Rev. B 58, R8881 (1998).
- Arovas and Auerbach (1988) D. P. Arovas and A. Auerbach, Phys. Rev. B 38, 316 (1988).
- (46) Y. Ge and Y. Komijani, in preparation .
Appendix A Supplementary materials
A.1 1. Strong coupling and channel magnet
We can use SUsp(2)SUch(2)U(1) symmetries to label various local states of the two-channel Kondo lattice. First, we consider a single-site two-channel Kondo model. The Hamiltonian
(10) |
can be written as
(11) |
in terms of the charge and Casimirs of spin and channel
(12) |
which are and . The energies of the 32 resulting states are listed in Table 1. The ground state is the Nozières doublet corresponding to the overscreened state. This Hamiltonian can be deformed to
(13) |
while preserving the symmetries of the Hamiltonian. is sufficient to change the ground state to the four spin-singlet channel-doublet states. These are states in which the impurity spin forms a spin-singlet with one of the channels. The remaining channel can be either empty or fully occupied, giving rise to the quartet. This quartet can be split by having a non-zero . For example will select a channel doublet with the other channel empty, and can be represented as
The operators act like Pauli matrices in the space of the doublet ,
(14) |
Having singled-out a channel doublet, we consider a two-channel Kondo lattice where these local ground states are coupled via the electron-hopping term. We can write
where and we have defined the Noziéres overscreened excited states ()
(15) | |||||
(16) |
and channel triplet excited states . For the sake of this section, we can project out the latter assuming that their energy is pushed further up. Therefore, the lowest excited states are the Noziéres states and the empty states . Treating the hopping via 2nd order perturbation theory,
(17) |
where and are projectors to ground state and excited states, respectively and . Assuming that the sites and are initially at and eventually at we find after a straightforward calculation
where . It can be easily shown that where is the exchange matrix and we have the relation
(18) |
This completes the derivation.
# | E | |||
1 | 0 | 0 | 3 | 0 |
1/2 | 1/2 | 1 | 8 | 0 |
0 | 0 | 2 | 2 | 0 |
0 | 1 | 0 | 3 | 0 |
# | E | |||
3/2 | 0 | 0 | 4 | 1 |
1 | 1/2 | 1 | 12 | 1/2 |
1/2 | 0 | 2 | 4 | 0 |
1/2 | 1 | 0 | 6 | 0 |
0 | 1/2 | 1 | 4 | -3/2 |
1/2 | 0 | 0 | 2 | -2 |
A.2 2. Details of numerical computation
Here we describe our algorithm to compute the low temperature Green’s functions. All other physical quantities can be computed from them thanks to the large- limit. The Green’s functions are computed from the self-consistency equations,
(19) | |||||
(20) |
We found that they are best represented in momentum and frequency domain. The solutions are solved on linear or logarithmic frequency grids and a linear momentum grid. Using that , we find
Thus, only a single sum over momentum is needed. Then we used Hilbert transform to recover the full self-energies, up to a constant in as discussed later this section. Next, the retarded Green’s functions are computed from the self-energies with set to in
(21) | |||||
(22) |
A search for is lastly used to satisfy the constraint
(23) |
The procedures above constitute the essential step in updating the self-consistency equations. Our main solver program is organized as follows:
-
0.
Initialize the self-energies to zero at a high temperature . Also initialize , say, to .
-
1.
At present temperature , initialize to a large value, say . Then,
-
(a)
update the self-consistency equations for , , and ;
-
(b)
search for a that gives a satisfying the constraint; then
-
(c)
reduce and repeat (a–b) until is small compared to , say ; then
-
(d)
repeat (a–b) until convergence.
-
(a)
-
2.
Reduce and rerun Step 1. Repeat until the desired temperature is reached.
Decreasing and slowly helps with the convergence. At low temperatures, frequency and momentum resolutions limits the convergence. The frequency grid need to be fine enough to resolve the sharp features due to small and . For us typically . The momentum resolution, or finite size effect, limits the lowest attainable to the order of the Fermi velocity of conduction electrons , being the linear dimension.
Strictly speaking, does not obey a Kramers-Kronig relation, whereas does. Consequently, obtained above is missing a real temperature-dependent constant . This can be conveniently absorbed into and need not be computed.
To improve efficiency, 1D calculations can utilize inversion symmetry. Thus, only half of the momentum grid is needed. At low temperatures, the computation can be further sped up with a frequency grid that is uniform at low frequencies and logarithmic at higher ones.

A.3 3. Other screenings and fillings in 1D
Here, we briefly present two other screening or filling cases in addition to the one presented in the Letter.
The first is the fully-screened case with half-filled conduction electrons, shown in Fig. 5. This has to be contrasted to the Fermi liquid regime of the ferromagnetically coupled Kondo lattice treated in the independent bath approximation in Ref. Komijani and Coleman, 2018. The spectrum of spinons and holons [Fig. 5(a,b)] show that ground state is gapped. Figure 5(c) shows that the ground state is paramagnetic, all spinons are confined to negative frequencies (as expected from spinon gap) and the electrons have a large FS, as . The plateau in is due to van Hove singularity of the conduction band density of states. Figure 5(d) shows the conduction electron self-energy ,
At zero temperature we can simplify this expression to:
(25) |
For the fully-screened case, spinons and holons develop a gap in their spectrum at zero temperature. Therefore, for this expression is equal to zero and is also gapped. This agrees with the numerical results shown in Fig. 5(d).
The second case is a two-channel Kondo case with a band of -filled conduction electrons in Fig. 6. For an infinitesimal the low energy spectrum shows unit cell-doubling and consequently two pairs of low-energy spinon and holon modes. A stronger breaks the symmetry between these two pairs.

A.4 4. Emergent dispersion in 1D
According to the self-consistency equations in Eqs. (19)–(22), if the input and are both local and , then the updated Green’s functions will remain local. Thus naively, without any Kondo lattice problem always reduces to a 0D impurity one. As shown in the Letter, 1D 2CKL is quite different from the impurity case when . Decreasing from large values will reduce the bandwidths of low energy spinon and holon modes, but below a certain , these bandwidths will cease to decease. This is the emergent dispersion dynamically generated due to an amplified RKKY interaction, as illustrated in Fig. 2.
The emergent dispersion lowers the free energy of the system compared to the local solution, as seen in the entropy inset of Fig. 2(b). At high temperature, denoted by , the 0D and infinitesimal- 1D′ state have the same entropy and free energy . At low temperature, denoted by , . Since , the solution of 1D′ gives a lower free energy. Therefore it is the genuine solution of the system.
This phenomenon is best demonstrated in a seeding numerical experiment for the zero static hopping case, i.e., . Before the self-consistency loop starts at each , one can add a tiny -dependent to the self-energy of spinons (or holons) used to construct the Green’s function, which is at Step 1 of the solver. The form of this seed ensures that it decreases with . The dependence is nonessential. At high temperature, is rapidly suppressed by the self-consistency iterations and the Green’s functions remain local, as seen in Fig. 7(a,b). Below a critical temperature, a part of the dispersion is exponentially amplified until it saturates. Since the seed magnitude diminishes with , this critical temperature must be finite and independent of .

Another way to observe this effect is to use an infinitesimal . The emergent dispersion will then dominate the shape of low energy modes. This is also manifested in and , shown in Fig. 2. Similar to the case with self-energy seeding, the temperature for emergent dispersion onset is finite as . Compared to the case with a finite in Fig. 7(c), they behave qualitatively the same. Therefore, at both finite and infinitesimal the system flow to the same fixed point.
Without , the Lagrangian of Eq. (Emergent Spinon Dispersion and Symmetry Breaking in Two-Channel Kondo Lattices) is invariant under simultaneous Galilean boosts of spinons and holons, that is . Consequently, the emergent dispersing mode here can freely translate in . In the numerics, the exact form of will determine the center of this dispersion. Different forms of , including random ones and longer range hopping terms, e.g. , all yield the same low energy spectrum up to momentum translations. Therefore, the emergent dispersion is robust. Without loss of generality, we set the center of emergent spinon dispersions to .
Another manifestation of the emergent dispersion is the behavior of effective as the system cools down. In 1D, variational principle applied to gives Parcollet and Georges (1997); Komijani and Coleman (2018)
(26) | |||||
(27) |
For a fixed , Fig. 7(c) shows that gradually decreases with until it settles at a finite zero-temperature value. For , we see that when the dispersion emerges, the effective has a steep drop to a value close to zero. This shows that no Heisenberg coupling is needed in 1D for the spinons to become mobile.
A.5 5. Susceptibilities
A.5.1 Channel Susceptibility

In the presence of channel anisotropy,
(28) |
the holon term in the Lagrangian is modified to
Here with are generators of SU() normalized according to . Expanding in small , the Lagrangian is changed by
(29) | |||||
up to . The channel susceptibility can be derived from taking derivatives of w.r.t. . We define
(30) |
where . Here the second term comes from the free path integral of the Hubbard-Stratonovitch field , which must be subtracted from the free energy. The first term gives
(31) | |||||
where denoting the linked clusters. In the large- limit, . Noting that , R.H.S. becomes
(32) |
The second term in Eq. (30) gives the same expression as Eq. (32) with an opposite sign and all . Therefore, we find for the channel susceptibility
(33) |
This expresses using correlators.

On the other hand, from the Feynman diagram in Fig. 8 we have
(34) |
where is defined as
(35) |
In momentum and frequency space
(36) |
Thus, . Therefore, as Fig. 8 suggests, the two approaches agree.
In Fig. 9(a,b), we show the static channel susceptibilities of the 0D and 1D 2CKL. As discussed in the Letter, in all dimensions the local susceptibility is , and in 1D the uniform susceptibility is . For 0D or , the changes from diverging to vanishing at low temperature as increases, according to Eq. (6). For 1D, the static channel susceptibilities are always vanishing according to Eq. (9). Since the Green’s functions we computed have both a scaling part and a regular part, only diverging scaling laws may be reliably extracted. The regular part will typically overwhelm the vanishing components, except at very low (zero) frequencies as is the case for [Fig. 4(a)].
A.5.2 Magnetic susceptibility
We show the magnetic susceptibilities in 1D for in Fig. 9(c). Here the uniform static magnetic susceptibility is diverging, revealing a magnetic instability as discussed in the Letter. The local static susceptibility is vanishing in this case, but will diverge for . The inset shows vs. . At low , becomes sharply peaked at .
A.6 6. Dispersion and ground state entropy in D
In the Bethe lattice, we find no emergent dispersion generated by the self-consistency solver. This comes from taking the infinite dimension limit first, which is a singular limit. Therefore, the dispersion is not a valid mechanism in this case to eliminate the extensive ground state entropy. However, the system is prone to other forms of symmetry breaking, such as spin or channel magnetization, and the entropy will then decrease to zero. An example is shown in Fig. 10.

A.7 7. Scaling analysis for impurity problem
For the sake of completeness, we review the scaling analysis for the multichannel Kondo problem using Schwinger bosons Parcollet and Georges (1997), which also applies to the infinite-coordination lattice problem we have studied in this Letter. For the spinon and holon we use the ansatz
(37) |
Using the self-energy formulae
(38) |
and that , we have
(39) |
Fourier transform of the Green’s function are
(40) | |||||
with similar expressions for and . can be computed from in the limit. The zero and the pole of cancel each other in this limit. Next we plug these into the Dyson equations:
In the scaling limit, the numerical solutions show that . For this equation to hold, the powers of frequency have to cancel from the left side, i.e. , and the amplitudes must match, i.e.
(41) |
Using , the ratio finally gives
(42) |
For the case of D Bethe lattice, we can write
(43) |
for where but . The solution is
(44) |
We are interested in the limit of this expression. We find not surprisingly that
(45) |
showing that in this limit the lattice is dominated by the local impurity solution and so, exponents are the same.
A.8 8. Ground state entropy of the impurity model
The ground state entropy of discussed in the introduction is specific to the spin-1/2 SU(2) two-channel Kondo model, i.e., . For generic , the result from Bethe ansatz and conformal field theory is Parcollet and Georges (1997)
(46) |
In the large- limit (keeping and constant), this gives
(47) |
where This fractional entropy is universally characterized by . It is fully consistent with the result of the large- theory as shown in Refs. Parcollet and Georges, 1997; Komijani and Coleman, 2018. Specifically, it agrees with Fig. 1(c).
A.9 9. Details of scaling analysis in 1D case
This section follows closely the impurity solution. We use the symbol for to refer to the low-momentum content of the spinons. We also expand fermionic field around the Fermi energy:
(48) | |||||
(49) |
The interaction term becomes
(50) |
This leads to the self-energies
(51) | |||||
In terms of these, the lattice Green’s functions are
(52) | |||||
(53) | |||||
(54) |
The ansatzes put forward in the Letter are
The Fourier transform of the imaginary-time function is straightforward:
(55) |
Defining and ,
(56) |
and we have
This integral is the version of a more general integral that we will encounter again. So, let us define
(57) | |||||
where we have used a shift to write it in terms of the function
(58) |
which by itself is written in terms of
Therefore, zeta-function is nothing but the Mellin transform of the Bessel function:
(59) |
valid for . Since we can express in terms of . Using the integral we readily find
(60) |
where from Eq. (59). Analytical continuation leads to
For the holons we have
This is the case of integral. Therefore,
(63) |
and similarly for left-movers
(64) |
We use these to find the Fourier transform of lattice Green’s functions. Defining shifted variables
and using , the lattice Green’s functions are
(65) | |||||
(66) | |||||
Assuming that the bare Green’s functions computed above are the exact ones, the boson self-energy is
(68) |
and the holon self-energy is
(69) |
To have conformal invariance, we have assumed all the velocities are the same , and holons and conduction electrons have the same Fermi wavevectors. The Fourier transform can be computed as before:
(70) | |||||
(71) |
For the self-consistency, we have to note that and we must choose . To first approximation, we neglect the contributions. The self-consistency, then leads to [Note that in the second line, the signs do not match for , therefore .]
where we defined . Eliminating between these equations, and using we find
(72) |
To solve this equation, we use the relation , the function can be written as
This equation, together with and can be used to prove that
(73) |
for . Using these we find the solution
(74) |
which leads to
(75) |

The conformal solutions to the large- 1D two-channel Kondo lattice we just derived are plotted in Fig. 11. Note that the singularities at in the effective holon dispersion is physically allowed since the holon is incoherent there. This “jump” diminishes as .
A.10 10. Large- limit of the Andrei-Orignac coset theory
According to Ref. Andrei and Orignac, 2000, using non-Abelian bosonization in the limit of large Heisenberg coupling between spins and away from any charge commensurate filling, the Hamiltonian can be written in the form of where are SU1(2) right/left-mover currents describing local moments and are SUK(2) right/left-mover currents describing the spin sector of fermions. Ref. Andrei and Orignac, 2000 shows that this system flows to the so-called chirally-stabilized fixed point, described by the coset theory,
For a dispersing Majorana model is predicted. Assuming that and are currents of SU() WZW models, we propose a generalization of this theory
In the limit of large- we find
(78) |
However, in addition to this non-magnetic part, there is also a residual magnetic contribution, which is given by
The extensive part of this, together with the decoupled charge () and channel contribution
(79) |
gives the of conduction electrons. The remaining O(N) part of the magnetic modes adds to the non-magnetic coset part to give the IR central charge
(80) |