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

Metamagnetic transition in the two ff orbitals Kondo lattice model

Christopher Thomas Sébastien Burdin Claudine Lacroix Instituto de Física, UFRGS, 91501-970 Porto Alegre, Brazil Université de Bordeaux, CNRS, LOMA, UMR 5798, F-33400 Talence, France Institut Néel, CNRS and Université Grenoble-Alpes, Boite Postale 166, 38042 Grenoble Cedex 09, France
Abstract

In this work, we study the effects of a transverse magnetic field in a Kondo lattice model with two ff orbitals interacting with the conduction electrons. The ff electrons that are present on the same site interact through Hund’s coupling, while on neighboring sites they interact through intersite exchange. We consider here that part of ff electrons are localized (orbital 1) while another part (orbital 2) are delocalized, as it is frequent in uranium systems. Then, only electrons in the localized orbital 1 interact through exchange interaction with the neighboring ones, while electrons in orbital 2 are coupled with conduction electrons through a Kondo interaction. We obtain a solution where ferromagnetism and Kondo effect coexist for small values of an applied transverse magnetic field for T0T\rightarrow 0. Increasing the transverse field, two situations can be obtained when Kondo coupling vanishes: first, a metamagnetic transition occurs just before or at the same time of the fully polarized state, and second, a metamagnetic transition occurs when the spins are already pointing out along the magnetic field.

1 Introduction

The duality, local versus non-local character, of strongly correlated ff-orbital electrons is a crucial microscopic quantum phenomenon. It generates unconventional macroscopic properties in rare earth and actinide heavy fermion compounds [1, 2, 3]. Among numerous fascinating examples, a family of uranium based compounds recently revealed the possibility of a new paradigm, where superconductivity and ferromagnetic order can coexist [4, 5]. In these materials, application of a magnetic field perpendicular to the easy axis can generate a metamagnetic transition and a surprising enhancement of superconductivity inside a ferromagnetic phase.

In order to understand the microscopic origin of this unanticipated behavior, we may first be inspired from the physics of heavy fermion cerium compounds, that can be described by a Kondo lattice model [6]. In this case, the 4f14f^{1} Kondo ions can form a magnetically ordered lattice for small values of pressure where Ruderman-Kittel-Kasuya-Yosida (RKKY) interaction is predominant and go for a coherent Fermi Liquid state at higher pressure where local Kondo screening dominates. This pressure induced quantum transition can be well explained invoking Doniach’s phase diagram [7]. The itinerant contribution of a priori local ff electronic orbitals in the formation of the coherent Fermi liquid Kondo state is a signature of duality. It is also characterized by a large effective mass which can be revealed, for example, by big values of the specific heat Sommerfeld coefficient. On the other hand, ff electrons may be fully localized in the magnetically ordered state. Usually, both regimes are separated by a quantum critical point  [8, 9], close to which a rich diversity of physical properties emerges, as for example, superconductivity in cerium based compounds [10].

Revealing experimentally the passage from fully localized to itinerant and dual behavior is crucial and challenging. In CeRhIn5, which is antiferromagnetic (AF) at ambient pressure, it was detected with de Haas-van Alphen experiment that these changes occur exactly at 2.4 GPa [11, 12]. It was observed that the large band signal indicates an increase of the effective mass from 5m0 to 60m0 which abruptly “disappears” as for a first order transition. In CeRu2Si2, a clear Fermi surface reconstruction is observed along the metamagnetic transition [13, 14], where a short Fermi surface takes places from the large Fermi surface, indicating a itinerant to localized behavior. In YbRh2Si2 [15], the authors studied the modification of the Fermi surface using high-resolution Compton scattering and showed a strong variation of the Fermi surface topology between low and high temperature regimes. They observed a clear enlargement of Fermi surface as temperature decreases, which is a signature of a coherent Kondo lattice ground state.

On the other side, the actinides have a partially filled 5ff shell and their behavior is different from the lanthanides behavior. At the beginning, the 5ff spatial wave function has a larger extent than the 4ff. Furthermore, while the valence is 4f1f^{1} in cerium based Kondo lattice systems, the valence of uranium based compounds fluctuates between 5f2f^{2} and 5f3f^{3}. As a consequence, a microscopic phenomenological description of duality in these materials requires an adaptation of the Kondo lattice model, where magnetic ions are described by composite local f2f^{2} multiplet states coupled to conduction electrons that account for the f2f^{2}-f3f^{3} valence fluctuations [16, 17]. Also, on general theoretical grounds, it is known that the possible presence of different channels for the conduction electrons might cause underscreeened or overscreened Kondo effect, depending on the comparison between the number of channels and the effective spin of the localized electronic multiplets [18]. In the uranium compounds, the 5f5f electrons are considered to be less localized than the 4f4f electrons in rare earth (lanthanides) compounds, allowing the possibility of underscreened Kondo effect, where the 5f5f electrons are only partially screened [16, 17]. Some models consider the 5f25f^{2} pair of orbitals as localized, while in other models one of the 5f5f electrons is itinerant, which is necessary to describe the superconductivity in U-based compounds [19, 20], where ff electrons are considered to be the key for both superconductivity and ferromagnetism, when they exist.

More precisly, in the group of the ferromagnetic superconductors, UGe2 [21], URhGe [22], and UCoGe [23], superconductivity appears at low temperatures in the region where they are ferromagnetic. While for UGe2 the superconducting state appears for high pressure, for URhGe and UCoGe the superconducting state already appears at ambient pressure and it persists for higher values of pressure. From the point of view of the magnetic response, the magnetic susceptibility χa/b/c\chi_{a/b/c} is anisotropic for UGe2, with aa-axis being the easy magnetic axis, and for UCoGe, where cc-axis is the easy magnetic axis. For URhGe, the evidence of anisotropy appears only below 50 K, where χbχc\chi_{b}\neq\chi_{c} with cc being the easy axis. UGe2 and UCoGe are considered to be Ising ferromagnets. For URhGe, the combination between ferromagnetism along the c- and b-axes is the core of its extremely high field-reentrant superconductivity for magnetic fields applied along bb.

The low value of magnetic moment per atom in UGe2, URhGe, and UCoGe compounds (1.5, 0.4, and 0.06 μB/U\mu_{B}/U, respectively) indicates duality between localized and itinerant character of the 5f5f electrons in UGe2 [24], while UCoGe has an itinerant behavior. This situation is also present in the uranium monochalcogenides US, USe and UTe [25], although their Curie temperatures are much greater and they do not become superconductors at smaller temperature.

A new fascinating compound, UTe2, was recently added into the family of uranium superconductors [26]. However, differently from the previous one, UTe2 does not order ferromagnetically. Nevertheless, the compound presents superconductivity coexisting with ferromagnetic fluctuations [27], and a rich phase diagram with different superconducting phases [28]. What is also interesting is that the compound presents a metamagnetic transition for higher magnetic field, leading the system from paramagnetic to polarized paramagnetic [29, 30]. Earlier first principle calculations with local density approximation have shown that UTe2 was a Kondo semiconductor with small gap (130\sim 130K) and flat bands around the Fermi Surface coming mainly from the 5ff electrons [31]. Other authors found a metallic ground state with moderate values of Coulomb repulsion (1.0\sim 1.0 eV) and an insulator-metal transition by increasing the Coulomb interaction [32]. More recently [33], density functional theory with exact diagonalization calculations indicate that ff-valence in UTe2 is close to 5f3f^{3} (Coulomb interaction is about 3 eV).

The interplay between Kondo screening and ferromagnetism has been studied by different authors for spin-1/2 or spin-1 Kondo lattice models. Mean-field approaches result in phase diagrams where ferromagnetism is predominant for small interaction (|J|/D|J|/D) and for the number of conduction electrons per site from zero to around 0.5, while Kondo phase appears for stronger interaction and in a larger range of conduction electrons density [34]. Other approaches report similar results [35], without coexisting phases. The dynamical mean field theory calculations and further mean-field decouplings found coexisting phases [36, 37, 38, 39], when ff-ff ferromagnetic exchange interaction is added to the Kondo lattice model [40, 41]. For URhGe, the density-matrix renormalization group calculations [42, 43] also addressed this competition, but no coexistence is reported. In a recent paper [44] it was shown, using a fermionic mean field approximation, that under applied magnetic field a metamagnetic transition may occur in the Kondo lattice; in the present paper we study the effect of magnetic field on 2-ff-orbitals Kondo lattice, which we think is more appropriate to U compounds which contain both localized and itinerant ff-electrons.

One guideline for the theoretical work presented hereafter is URhGe and more precisely its magnetization, which displays a rapid variation when a transverse magnetic field is applied [45]. In a more complete description, at low temperature, magnetic field perpendicular to the easy axis first destroys the superconducting phase around 22 T, and at higher field a reentrant superconducting phase appears between 8\sim 8 T and 13\sim 13 T . Above the superconducting transition temperature, in the ferromagnetic phase the spin-reorientation occurs at HR12H_{R}\approx 12 T. The maximum value of TSCT_{\mathrm{SC}} in the reentrant phase coincides approximatively with HRH_{R}. We study the effect of a transverse magnetic field using a modified Kondo lattice model adapted for this uranium system. Although a superconductiting phase is observed experimentally around the magnetization change, we do not explore this fact, and let it for a future work. This is justified since TSCT_{SC} is much smaller than the Curie temperature TCT_{C} (respectively 0.25K0.25K and 9.5K9.5K).

The paper is organized as follows: In section 2 the model used to discuss the Kondo and ferromagnetic interactions in a transverse field is presented. The results and discussions are presented in sections 3 and 4.

2 The model

We have in mind the objective of modeling uranium compounds with a valence fluctuating between 5f2f^{2} and 5f3f^{3}. We will thus consider a Kondo lattice Hamiltonian where the Kondo ions are described by 5f2f^{2} local multiplets. These two 5ff orbitals (labeled as orbitals 11 and 22) are centered on each lattice site ii and we define the spin of each of these electrons as 𝐒i1\mathbf{S}_{i1} and 𝐒i2\mathbf{S}_{i2}. The intrasite interaction should in principle include both the local Coulomb repulsion and the Hund’s coupling. However, by enforcing the localization and describing the corresponding degrees of freedom with effective spin operators, the Coulomb repulsion is implicitly assumed to be strong.

Furthermore, we also want to describe the duality of the localized ff-electrons. We consider here that part of ff-electrons is fully localized (orbital 11) while another part (orbital 22) can be partially delocalized, as it is frequent in U systems. Then, only electrons in the fully localized orbital 11 are supposed to interact through ferromagnetic intersite exchange interaction, while electrons in orbital 22 are coupled with conduction electrons through a Kondo interaction. The itinerant electrons emerge from the f2f^{2}-f3f^{3} valence fluctuation effects as well as from other electronic bands. In figure 1 we can see some possible solutions that can be described by the proposed model. In the ferromagnetic phase spins with all components (cc, and both ff orbitals) exhibit a non-zero magnetization. In the fully polarized phase, all spins are aligned with the transverse magnetic field. In the Kondo phase, the spin of the ff-electrons in orbital 2 couple with the spin of the conduction electrons through Kondo interaction and the spins of orbital 1 remain paramagnetic and finally, the mixed state where Kondo and ferromagnetism are both present.

The generalized Kondo lattice Hamiltonian of the system is written as:

H\displaystyle H =Hintersite+Hlocal+HKondo+Hc,\displaystyle=H_{\mathrm{intersite}}+H_{\mathrm{local}}+H_{\mathrm{Kondo}}+H_{c}~{}, (1)

where the first term is intersite exchange that we consider here as an Ising interaction between localized spins in orbital 1:

Hintersite\displaystyle H_{\mathrm{intersite}} =JijSi1zSj1z.\displaystyle=-J\sum_{\langle{ij}\rangle}S_{i1}^{z}S_{j1}^{z}\,~{}. (2)

This Ising exchange interaction describes phenomenologically the large magnetic uniaxial anisotropy observed experimentally in the uranium based ferromagnetic superconductors. The second term in the Hamiltonian includes the on-site interactions: Hund’s coupling (JH0J_{H}\geq 0) between electrons in orbitals 11 and 22, and applied magnetic field, which can be in any direction:

Hlocal\displaystyle H_{\mathrm{local}} =JHi𝐒i1𝐒i2𝐡i(𝐒i1+𝐒i2).\displaystyle=-J_{\mathrm{H}}\sum_{i}\mathbf{S}_{i1}\cdot\mathbf{S}_{i2}-\mathbf{h}\cdot\sum_{i}\big{(}\mathbf{S}_{i1}+\mathbf{S}_{i2}\big{)}\,~{}. (3)

The third term is the local Kondo coupling (JK0J_{K}\geq 0) between ff-electrons in orbital 2 and cc-electrons:

HKondo\displaystyle H_{\mathrm{K}ondo} =JKi𝐒i2𝐬i,\displaystyle=J_{K}\sum_{i}\mathbf{S}_{i2}\cdot\mathbf{s}_{i}\,~{}, (4)

where 𝐬i\mathbf{s}_{i} denotes the local spin density of the cc-electrons. The last term describes the conduction band:

Hc\displaystyle H_{c} =𝐤σ(ϵ𝐤μ)c𝐤σc𝐤σ+Nμnc,\displaystyle=\sum_{{\bf k}\sigma}(\epsilon_{{\bf k}}-\mu)c_{{\bf k}\sigma}^{{\dagger}}c_{{\bf k}\sigma}+N\mu n_{c}\,~{}, (5)

where σ=,\sigma=\uparrow,\downarrow is the spin component and 𝐤{\bf k} is the electron momentum. ϵ𝐤\epsilon_{{\bf k}} is the dispersion relation, μ\mu is the chemical potential associated with the averaged total number of conduction electrons per site ncn_{c}, and NN denotes the number of lattice sites.

Refer to caption
Figure 1: Some possible ordered phases of the system. The moments of the 3 types of electrons ( f1f_{1}, f2f_{2} and cc electrons) are indicated separately. From left to right, from top to bottom. Ferromagnetic phase: the averaged magnetizations for ff- and c-c electrons is non zero. Fully Polarized phase : particular ferromagnetic phase where the magnetizations are saturated and aligned along the external magnetic field; the ff-electron magnetisation per site is S=1S=1. The cc electrons couple antiparallel to the ff ones since local Kondo exchange is negative. Pure Kondo state: ff-electrons in orbital 2 and cc-electrons form Kondo singlets, and the ff-electrons in orbital 1 are paramagnetic. Mixed state: coexistence of ferromagnetic order (non-zero magnetization) and Kondo screening of electrons in orbital 2.

In the following, we will consider that there are 2 ff electrons on each site, one in each orbital. Thus in the absence of Kondo interaction, these two electrons are coupled by Hund’s interaction in a S=1S=1 state and they interact with neighboring sites through Ising exchange. The problem is solved in two steps: first, we solve the ff-electrons problem, with a magnetic field applied in the xzx-z plane; second, the Kondo coupling is added for orbital 2.

2.1 Solution for the magnetic part

We first treat the effect of the Ising and Hund’s interactions in orbitals 1 and 2 in the presence of magnetic field.

HfHintersite+Hlocal,\displaystyle H_{f}\equiv H_{\mathrm{intersite}}+H_{\mathrm{local}}\,, (6)

where the different terms are defined above.

In a mean field decoupling , we introduce the effective field acting on electrons in orbital 1 and 2 respectively, hw1h_{w1} and hw2h_{w2}. The purely ff-orbital part of the Hamiltonian is thus approximated as:

HfHfMFE0i(𝐡w1𝐒i1+𝐡w2𝐒i2),\displaystyle H_{f}\approx H^{\mathrm{MF}}_{f}\equiv E_{0}-\sum_{i}\left(\mathbf{h}_{w1}\cdot\mathbf{S}_{i1}+\mathbf{h}_{w2}\cdot\mathbf{S}_{i2}\right)\,, (7)

where E0=NJ2(m1z)2+NJH(m1xm2x+m1zm2z)E_{0}=N\frac{J^{\prime}}{2}(m_{1}^{z})^{2}+NJ_{H}(m_{1}^{x}m_{2}^{x}+m_{1}^{z}m_{2}^{z}) and m1(2)z=1NiSi1(2)zm_{1(2)}^{z}=\frac{1}{N}\sum_{i}\langle{S_{i1(2)}^{z}}\rangle and m1(2)x=1NiSi1(2)xm_{1(2)}^{x}=\frac{1}{N}\sum_{i}\langle{S_{i1(2)}^{x}}\rangle are the components of magnetization in the zz and xx directions. Note the redefinition J=2zJJ^{\prime}=2zJ, where zz is the number of nearest neighbors.

The effective fields are defined as

𝐡w1=(hx+JHm2x0hz+Jm1z+JHm2z),\displaystyle\mathbf{h}_{w1}=\left({\begin{array}[]{c}h_{x}+J_{H}m_{2}^{x}\\ 0\\ h_{z}+J^{\prime}m_{1}^{z}+J_{H}m_{2}^{z}\end{array}}\right)\,, (11)
𝐡w2=(hx+JHm1x0hz+JHm1z).\displaystyle\mathbf{h}_{w2}=\left({\begin{array}[]{c}h_{x}+J_{H}m_{1}^{x}\\ 0\\ h_{z}+J_{H}m_{1}^{z}\end{array}}\right)\,. (15)

We can also identify the angles of magnetization for both orbitals 1 and 2 with respect to x-direction:

tanθ1\displaystyle\tan{\theta_{1}} =m1zm1x=hz+Jm1z+JHm2zhx+JHm2x,\displaystyle=\frac{m_{1}^{z}}{m_{1}^{x}}=\frac{h_{z}+J^{\prime}m_{1}^{z}+J_{H}m_{2}^{z}}{h_{x}+J_{H}m_{2}^{x}}\,, (16)
tanθ2\displaystyle\tan{\theta_{2}} =m2zm2x=hz+JHm1zhx+JHm1x.\displaystyle=\frac{m_{2}^{z}}{m_{2}^{x}}=\frac{h_{z}+J_{H}m_{1}^{z}}{h_{x}+J_{H}m_{1}^{x}}\,. (17)

These definitions help us to have a more compact form for the set of self-consistent mean-field equations, that can be written as

m1z\displaystyle m_{1}^{z} =sinθ12tanhβ𝐡w12,\displaystyle=\frac{\sin{\theta_{1}}}{2}\tanh{\frac{\beta\|\mathbf{h}_{w1}\|}{2}}\,, (18)
m1x\displaystyle m_{1}^{x} =cosθ12tanhβ𝐡w12,\displaystyle=\frac{\cos{\theta_{1}}}{2}\tanh{\frac{\beta\|\mathbf{h}_{w1}\|}{2}}\,, (19)
m2z\displaystyle m_{2}^{z} =sinθ22tanhβ𝐡w22,\displaystyle=\frac{\sin{\theta_{2}}}{2}\tanh{\frac{\beta\|\mathbf{h}_{w2}\|}{2}}\,, (20)
m2x\displaystyle m_{2}^{x} =cosθ22tanhβ𝐡w22.\displaystyle=\frac{\cos{\theta_{2}}}{2}\tanh{\frac{\beta\|\mathbf{h}_{w2}\|}{2}}\,. (21)

The problem can be solved considering the external magnetic field in the xx - zz plane. However, as we will consider the Kondo coupling between the ff electrons in orbital 2 with the conduction electrons, the magnetization of orbital 2, m2fm_{2}^{f}, has to be calculated self-consistently in the presence of Kondo coupling. This is presented below.

2.2 Solution in the presence of Kondo coupling

We now consider the full Hamiltonian defined by eq. (1), where the purely ff-orbital part is approximated by the mean-field expression eq. (7):

HHMF1E0+HfMF+HKondo+Hc,\displaystyle H\approx H^{\mathrm{MF1}}\equiv E_{0}+H^{\mathrm{MF}}_{f}+H_{\mathrm{Kondo}}+H_{c}~{}, (22)

First it is necessary to make a basis change of the spin quantization axis: in the presence of the external magnetic field, the ff magnetic moments tend to point towards the direction of magnetic field, but they are not aligned with field since the Ising exchange acts as a local anisotropy (see above eqs. (18)-(21)). For this reason, we will change the quantization axis for the spins. We use the symbol tilde to represent this new quantization axis: z~\tilde{z} and σ~\tilde{\sigma} correspond to the zz axis rotated by angle θ2{\theta_{2}} defined in the previous section. Invoking the spin rotational invariance of HKondoH_{\mathrm{Kondo}} and HcH_{c}, the mean-field Hamiltonian is thus rewritten as

HMF1=\displaystyle H^{\mathrm{MF1}}= E0+Nμnc𝐡w1i𝐒i1hw2iSi2z~\displaystyle E_{0}+N\mu n_{c}-\mathbf{h}_{w1}\cdot\sum_{i}\mathbf{S}_{i1}-h_{w2}\sum_{i}S_{i2}^{\tilde{z}}
+JKi𝐒~i2𝐬~i+𝐤σ~(ϵ𝐤μ)c𝐤σ~c𝐤σ~.\displaystyle+J_{K}\sum_{i}\tilde{\mathbf{S}}_{i2}\cdot\tilde{\mathbf{s}}_{i}+\sum_{{\bf k}\tilde{\sigma}}(\epsilon_{{\bf k}}-\mu)c_{{\bf k}\tilde{\sigma}}^{{\dagger}}c_{{\bf k}\tilde{\sigma}}~{}. (23)

The Kondo effect is treated within usual mean field approximation using fermionic representation of the spin operators :

Si2z~=\displaystyle S_{i2}^{\tilde{z}}= 12(fi2~fi2~fi2~fi2~)\displaystyle\frac{1}{2}\left(f_{i2\tilde{\uparrow}}^{\dagger}f_{i2\tilde{\uparrow}}-f_{i2\tilde{\downarrow}}^{\dagger}f_{i2\tilde{\downarrow}}\right) (24)
Si2y~=\displaystyle S_{i2}^{\tilde{y}}= i2(fi2~fi2~fi2~fi2~)\displaystyle\frac{i}{2}\left(f_{i2\tilde{\downarrow}}^{\dagger}f_{i2\tilde{\uparrow}}-f_{i2\tilde{\uparrow}}^{\dagger}f_{i2\tilde{\downarrow}}\right) (25)
Si2x~=\displaystyle S_{i2}^{\tilde{x}}= 12(fi2~fi2~+fi2~fi2~),\displaystyle\frac{1}{2}\left(f_{i2\tilde{\downarrow}}^{\dagger}f_{i2\tilde{\uparrow}}+f_{i2\tilde{\uparrow}}^{\dagger}f_{i2\tilde{\downarrow}}\right)~{}, (26)

where the Abrikosov fermion annihilation (creation) operators fi2σ~()f_{i2\tilde{\sigma}}^{(\dagger)} satisfy the local constraints σ~fi2σ~fi2σ~=1\sum_{\tilde{\sigma}}f_{i2\tilde{\sigma}}^{\dagger}f_{i2\tilde{\sigma}}=1. Within the mean-field approximation, this constraint is satisfied on average, iσ~fi2σ~fi2σ~=N\sum_{i\tilde{\sigma}}\langle{f_{i2\tilde{\sigma}}^{\dagger}f_{i2\tilde{\sigma}}}\rangle=N, by introducing an effective energy level ϵ2f\epsilon^{f}_{2}. Finally the mean-field approximations for the Hamiltonian give

HHMF2\displaystyle H\approx H^{\mathrm{MF2}} =E0MF2𝐡w1i𝐒i1+iσ~ϵ2σ~ffi2σ~fi2σ~\displaystyle=E_{0}^{\mathrm{MF2}}-\mathbf{h}_{w1}\cdot\sum_{i}\mathbf{S}_{i1}+\sum_{i\tilde{\sigma}}\epsilon^{f}_{2\tilde{\sigma}}f_{i2\tilde{\sigma}}^{{\dagger}}f_{i2\tilde{\sigma}}
+iσ~Λ2σ¯~(ciσ~fi2σ~+fi2σ~ciσ~)+𝐤σ~ϵ𝐤cc𝐤σ~c𝐤σ~,\displaystyle+\sum_{i\tilde{\sigma}}\Lambda_{2}^{\tilde{\bar{\sigma}}}\Big{(}c_{i\tilde{\sigma}}^{{\dagger}}f_{i2\tilde{\sigma}}+f_{i2\tilde{\sigma}}^{{\dagger}}c_{i\tilde{\sigma}}\Big{)}+\sum_{{\bf k}\tilde{\sigma}}\epsilon_{{\bf k}}^{c}c_{{\bf k}\tilde{\sigma}}^{{\dagger}}c_{{\bf k}\tilde{\sigma}}\,, (27)

where λi2σ~=fi2σ~ciσ~\lambda_{i2}^{\tilde{\sigma}}=\langle{f_{i2\tilde{\sigma}}^{{\dagger}}c_{i\tilde{\sigma}}}\rangle and

E0MF2\displaystyle E_{0}^{\mathrm{MF2}} =E0+NμncNϵ2f+JKNλ2~λ2~,\displaystyle=E_{0}+N\mu n_{c}-N\epsilon^{f}_{2}+J_{K}N\lambda_{2}^{\tilde{\uparrow}}\lambda_{2}^{\tilde{\downarrow}}\,, (28)
ϵ2σ~f\displaystyle\epsilon^{f}_{2\tilde{\sigma}} =ϵ2fσ~hw2,\displaystyle=\epsilon^{f}_{2}-\tilde{\sigma}h_{w2}\,, (29)
Λ2σ~\displaystyle\Lambda_{2}^{\tilde{\sigma}} =JK2λ2σ~,\displaystyle=-\frac{J_{K}}{2}\lambda_{2}^{\tilde{\sigma}}\,, (30)
ϵ𝐤c\displaystyle\epsilon_{{\bf k}}^{c} =ϵ𝐤μ.\displaystyle=\epsilon_{{\bf k}}-\mu\,. (31)

σ~=±1/2\tilde{\sigma}=\pm 1/2 and hw2=𝐡w2h_{w2}=\|\mathbf{h}_{w2}\|. With the first mean-field approximation that we used, the momenta 𝐒i1\mathbf{S}_{i1} describing the local electronic orbital 1 are effectively decoupled from both the orbital 2 and the conduction electrons. However, the Hund interaction between orbitals 1 and 2 is still implicitly present through the effective fields 𝐡w1\mathbf{h}_{w1} and 𝐡w2\mathbf{h}_{w2} given by solving eqs. (11) and (15) and the self-consistent relations (18) and (21). The second mean-field approximation replaces the Kondo interaction by an effective hybridization between orbital 2 and conduction electrons. It describes qualitatively and quantitatively the Kondo-singlet correlations that can occur at low temperature. The mean-field Hamiltonian can be diagonalized using the space momentum representation 𝐤{\bf k}:

HMF2\displaystyle H^{\mathrm{MF2}} =E0MF2𝐡w1i𝐒i1+𝐤σ~(c𝐤σ~,f𝐤2σ~)H𝐤σ~cf2(c𝐤σ~f𝐤2σ~),\displaystyle=E_{0}^{\mathrm{MF2}}-\mathbf{h}_{w1}\cdot\sum_{i}\mathbf{S}_{i1}+\sum_{{\bf k}\tilde{\sigma}}\left(c_{{\bf k}\tilde{\sigma}}^{{\dagger}},f_{{\bf k}2\tilde{\sigma}}^{{\dagger}}\right)H_{{\bf k}\tilde{\sigma}}^{cf_{2}}\left(\begin{array}[]{c}c_{{\bf k}\tilde{\sigma}}\\ f_{{\bf k}2\tilde{\sigma}}\end{array}\right)~{}, (35)

with

H𝐤σ~cf2(ϵ𝐤cΛ2σ~Λ2σ~ϵ2σ~f).\displaystyle H_{{\bf k}\tilde{\sigma}}^{cf_{2}}\equiv\left(\begin{array}[]{cc}\epsilon_{{\bf k}}^{c}&\Lambda_{2}^{\tilde{\sigma}}\\ {}{}\hfil&{}{}\hfil\\ \Lambda_{2}^{\tilde{\sigma}}&\epsilon^{f}_{2\tilde{\sigma}}\end{array}\right)~{}. (39)

From HMF2H^{\mathrm{MF2}} and using the above expression of the block H𝐤σ~cf2H_{{\bf k}\tilde{\sigma}}^{cf_{2}}, the one body Green’s functions for orbital 2 and conduction electrons can be written as

g𝐤σ~cc(iω)\displaystyle g_{{\bf k}\tilde{\sigma}}^{cc}(i\omega) =iωϵ2σ~f(iωϵ𝐤c)(iωϵ2σ~f)(Λ2σ¯~)2,\displaystyle=\frac{i\omega-\epsilon^{f}_{2\tilde{\sigma}}}{(i\omega-\epsilon_{{\bf k}}^{c})(i\omega-\epsilon^{f}_{2\tilde{\sigma}})-(\Lambda_{2}^{\tilde{\bar{\sigma}}})^{2}}\,, (40)
g𝐤σ~f2f2(iω)\displaystyle g_{{\bf k}\tilde{\sigma}}^{f_{2}f_{2}}(i\omega) =iωϵ𝐤c(iωϵ𝐤c)(iωϵ2σ~f)(Λ2σ¯~)2,\displaystyle=\frac{i\omega-\epsilon_{{\bf k}}^{c}}{(i\omega-\epsilon_{{\bf k}}^{c})(i\omega-\epsilon^{f}_{2\tilde{\sigma}})-(\Lambda_{2}^{\tilde{\bar{\sigma}}})^{2}}\,, (41)
g𝐤σ~cf2(iω)\displaystyle g_{{\bf k}\tilde{\sigma}}^{cf_{2}}(i\omega) =g𝐤σ~f2c(iω)=Λ2σ¯~(iωϵ𝐤c)(iωnϵ2σ~f)(Λ2σ¯~)2.\displaystyle=g_{{\bf k}\tilde{\sigma}}^{f_{2}c}(i\omega)=\frac{-\Lambda_{2}^{\tilde{\bar{\sigma}}}}{(i\omega-\epsilon_{{\bf k}}^{c})(i\omega_{n}-\epsilon^{f}_{2\tilde{\sigma}})-(\Lambda_{2}^{\tilde{\bar{\sigma}}})^{2}}\,. (42)

From the Green functions we can calculate the self-consistent parameters:

nσ~c\displaystyle n_{\tilde{\sigma}}^{c} =𝐤𝑑ωf(ω)ρ𝐤σ~c(ω),\displaystyle=\sum_{{\bf k}}\int_{-\infty}^{\infty}d\omega f(\omega)\rho_{{\bf k}\tilde{\sigma}}^{c}(\omega)\,, (43)
nσ~f2\displaystyle n_{\tilde{\sigma}}^{f_{2}} =𝐤𝑑ωf(ω)ρ𝐤σ~f2(ω),\displaystyle=\sum_{{\bf k}}\int_{-\infty}^{\infty}d\omega f(\omega)\rho_{{\bf k}\tilde{\sigma}}^{f_{2}}(\omega)\,, (44)
λ2σ~\displaystyle\lambda_{2}^{\tilde{\sigma}} =𝐤𝑑ωf(ω)ρ𝐤σ~cf2(ω).\displaystyle=\sum_{{\bf k}}\int_{-\infty}^{\infty}d\omega f(\omega)\rho_{{\bf k}\tilde{\sigma}}^{cf_{2}}(\omega)\,. (45)

where the the spectral density functions are

ρ𝐤σ~c(ω)\displaystyle\rho_{{\bf k}\tilde{\sigma}}^{c}(\omega) =1π[g𝐤σ~cc(iω)],\displaystyle=-\frac{1}{\pi}\Im{\big{[}g_{{\bf k}\tilde{\sigma}}^{cc}(i\omega)\big{]}}\,, (46)
ρ𝐤σ~f2(ω)\displaystyle\rho_{{\bf k}\tilde{\sigma}}^{f_{2}}(\omega) =1π[g𝐤σ~f2f2(iω)],\displaystyle=-\frac{1}{\pi}\Im{\big{[}g_{{\bf k}\tilde{\sigma}}^{f_{2}f_{2}}(i\omega)\big{]}}\,, (47)
ρ𝐤σ~cf2(ω)\displaystyle\rho_{{\bf k}\tilde{\sigma}}^{cf_{2}}(\omega) =1π[g𝐤σ~cf2(iω)].\displaystyle=-\frac{1}{\pi}\Im{\big{[}g_{{\bf k}\tilde{\sigma}}^{cf_{2}}(i\omega)\big{]}}\,. (48)

The parameters defined in equations (43)(44)(45) are solved considering a square band (i.e., a constant density of states) for the conduction electrons. The only 𝐤{\bf k} dependence comes from the dispersion relation ϵ𝐤\epsilon_{{\bf k}}, in this way, the sum over 𝐤{\bf k} is changed by an integral over energy as 𝐤DDρ0(ϵ)𝑑ϵ\sum_{{\bf k}}\rightarrow\int_{-D}^{D}\rho_{0}(\epsilon)d\epsilon. The non-interacting conduction electrons density of states is taken as ρ0=1/2D\rho_{0}=1/2D in the interval [D:D][-D:D]. DD is the half of the value of the bandwidth.

Using the above expressions for the Green’s functions and invoking the relation eq. (31), the mean-field self-consistent equations can be rewritten as

nσ~c\displaystyle n_{\tilde{\sigma}}^{c} =ρ0DD𝑑ϵ(E+σ~(ϵ)ϵ2σ~f)f(E+σ~(ϵ))(Eσ~(ϵ)ϵ2σ~f)f(Eσ~(ϵ))ΔE(ϵ),\displaystyle=\rho_{0}\int_{-D}^{D}d\epsilon\frac{\left(E_{+}^{\tilde{\sigma}}(\epsilon)-\epsilon^{f}_{2\tilde{\sigma}}\right)f(E_{+}^{\tilde{\sigma}}(\epsilon))-\left(E_{-}^{\tilde{\sigma}}(\epsilon)-\epsilon^{f}_{2\tilde{\sigma}}\right)f(E_{-}^{\tilde{\sigma}}(\epsilon))}{\Delta E(\epsilon)}\,, (49)
nσ~f2\displaystyle n_{\tilde{\sigma}}^{f_{2}} =ρ0DD𝑑ϵ(E+σ~(ϵ)ϵ𝐤c)f(E+σ~(ϵ))(Eσ~(ϵ)ϵ𝐤c)f(Eσ~(ϵ))ΔE(ϵ),\displaystyle=\rho_{0}\int_{-D}^{D}d\epsilon\frac{\left(E_{+}^{\tilde{\sigma}}(\epsilon)-\epsilon_{{\bf k}}^{c}\right)f(E_{+}^{\tilde{\sigma}}(\epsilon))-\left(E_{-}^{\tilde{\sigma}}(\epsilon)-\epsilon_{{\bf k}}^{c}\right)f(E_{-}^{\tilde{\sigma}}(\epsilon))}{\Delta E(\epsilon)}\,, (50)
Λ2σ~\displaystyle\Lambda_{2}^{\tilde{\sigma}} =ρ0JK2Λ2σ~DD𝑑ϵf(E+σ~(ϵ))f(Eσ~(ϵ))ΔE(ϵ).\displaystyle=\frac{\rho_{0}J_{K}}{2}\Lambda_{2}^{\tilde{\sigma}}\int_{-D}^{D}d\epsilon\frac{f(E_{+}^{\tilde{\sigma}}(\epsilon))-f(E_{-}^{\tilde{\sigma}}(\epsilon))}{\Delta E(\epsilon)}\,. (51)

where f(ω)=(1+eβω)1f(\omega)=(1+e^{\beta\omega})^{-1} denotes the Fermi-Dirac function, and

E±σ~(ϵ)\displaystyle E_{\pm}^{\tilde{\sigma}}(\epsilon) ϵ+ϵ2σ~f±(ϵϵ2σ~f)2+4(Λ2σ¯~)22,\displaystyle\equiv\frac{\epsilon+\epsilon^{f}_{2\tilde{\sigma}}\pm\sqrt{(\epsilon-\epsilon^{f}_{2\tilde{\sigma}})^{2}+4(\Lambda_{2}^{\tilde{\bar{\sigma}}})^{2}}}{2}\,, (52)
ΔE(ϵ)\displaystyle\Delta E(\epsilon) E2σ~(ϵ)E3σ~(ϵ).\displaystyle\equiv E_{2}^{\tilde{\sigma}}(\epsilon)-E_{3}^{\tilde{\sigma}}(\epsilon)~{}. (53)

We note that the mean-field equation (51) has a trivial solution Λ2σ~=0\Lambda_{2}^{\tilde{\sigma}}=0 which is realized in the non-Kondo phases where the local ff and the conduction electrons are decoupled. When a solution Λ2σ~0\Lambda_{2}^{\tilde{\sigma}}\neq 0 exists and is energetically stable, a Kondo phase is realized. On top of this usual mean-field description of Kondo effect, we also consider here the possibility of magnetic ordering, which may coexist or not with the Kondo effect. From equation (50) we can obtain the magnetization of orbital 2 in the rotated direction (noted with tilda), M2fz~=12(n~f2n~f2)M_{2}^{f\tilde{z}}=\frac{1}{2}(n_{\tilde{\uparrow}}^{f_{2}}-n_{\tilde{\downarrow}}^{f_{2}}). However, this magnetization is in the direction of the effective field hw2h_{w2} and not in the original zz or xx directions. The magnetization in the initial cartesian coordinates is thus given by

m2z\displaystyle m_{2}^{z} =M2fz~sinθ2,\displaystyle=M_{2}^{f\tilde{z}}\sin{\theta_{2}}\,, (54)
m2x\displaystyle m_{2}^{x} =M2fz~cosθ2,\displaystyle=M_{2}^{f\tilde{z}}\cos{\theta_{2}}\,, (55)

where θ2\theta_{2} was defined previously in equation (17).

Finally, we can solve self-consistently our set of six equations (49), (50), (18), (54), and (51), and determine the parameters μ\mu, ϵ2f\epsilon^{f}_{2}, m1zm_{1}^{z}, m2zm_{2}^{z}, λ~2\tilde{\lambda}_{2}^{\uparrow}, and λ~2\tilde{\lambda}_{2}^{\downarrow}, respectively. We have checked the self-consistent solutions looking for the parameters that minimize the mean-field Hamiltonian presented in equation (35) for the case T0T\rightarrow 0. The numerical results are presented in the next section.

3 Results

We explore the phase diagram for various sets of the model’s parameters. First, we fix the number of particles in orbital 2, n2f1\langle{n_{2f}}\rangle\equiv 1, and the number of conduction electrons ncnc=0.8\langle{n_{c}}\rangle\equiv n_{c}=0.8, with the help of the auxiliary Lagrangian’s multipliers, ϵ2f\epsilon^{f}_{2} and μ\mu. In the last part, we will explore the effect of variation of ncn_{c}. The magnetic field is taken in the xx direction, fixing hz=0h_{z}=0. The energies are scaled with respect to the half bandwidth D=1D=1, using kB=1k_{\mathrm{B}}=1 and T=0.0001T=0.0001 for all figures.

3.1 Effect of magnetic field and JKJ_{K} for J=JH=0J^{\prime}=J_{H}=0.

First we study here the effect of a magnetic field and Kondo coupling in the absence of both Hund and intersite interactions, JH=0J_{H}=0 and J=0J^{\prime}=0. In this case the local spins 𝐒i1\mathbf{S}_{i1} are fully decoupled from the orbital 2 and the conduction electrons. The model corresponds to an usual Kondo lattice for orbital 2, and we use the Abrikosov fermions fi2f_{i2} to describe the local spins 𝐒i2\mathbf{S}_{i2}. Since rotational symmetry is preserved at zero field when J=0J^{\prime}=0, we arbitrarily choose the xx-axis along the direction of the applied magnetic field.

Refer to caption
Figure 2: (𝐚\mathbf{a}) Variation of the magnetization and Kondo mean field parameter as a function of hxh_{x}. λλ\lambda_{\uparrow}\neq\lambda_{\downarrow} while m2xm_{2}^{x} is finite and not saturated. (𝐛\mathbf{b}) The magnetization m2xm_{2}^{x} as a function of hxh_{x}. Higher values of JKJ_{K} allows longer plateau behaviour. (𝐜\mathbf{c}) m2xm_{2}^{x} as a function of JKJ_{K}. At the same point where λ\lambda_{\uparrow} and λ\lambda_{\downarrow} have a discontinuity, m2xm_{2}^{x} has a step from its saturated value to m2x=(1nc)/2m_{2}^{x}=(1-n_{c})/2 indicating the mixed phase. It goes continually to zero for large values of JKJ_{K} . m1zm_{1}^{z} and m2zm_{2}^{z} are zero for all figures since J=JH=0J^{\prime}=J_{H}=0.

The figure 2(𝐚)(\mathbf{a}) shows the variation of the effective Kondo hybridizations λσ\lambda_{\sigma} and the magnetizations for both orbitals as a function of the magnetic field hxh_{x}, for JK=1.0J_{K}=1.0 and J=JH=0J=J_{H}=0. Here, orbital 1 is fully decoupled from the other electrons, therefore in the ground state its magnetization saturates to its maximal value as soon as hx0h_{x}\neq 0. The magnetization of orbital 2 is more interesting and we can identify three regimes depending on intensity of the applied magnetic field: at relatively small hxh_{x}, we find that m2xm_{2}^{x} increases linearly with hxh_{x}, revealing a Fermi-liquid regime with a finite susceptibility that correspond to the usual Kondo coherent regime. On the other side, the magnetization m2xm_{2}^{x} saturates to its maximal value when the applied field is higher than a critical value hK{h_{K}^{\star}}. hK{h_{K}^{\star}} is defined as the critical field necessary to destroy the Kondo hybridization and it coincides with the saturation of m2xm_{2}^{x} only in some cases (see section 3.3). The intermediate regime of magnetic fields, above the linear response and below the critical value hK{h_{K}^{\star}}, we find a magnetization plateau at the value m2x(1nc)/2m_{2}^{x}\approx(1-n_{c})/2 independently from hxh_{x} and JKJ_{K} as can be seen in figures 2(b) and 2(c).

The three magnetization regimes described above can be interpreted as follows: for relatively small hxh_{x}, the local Kondo spins 𝐒i2\mathbf{S}_{i2} and the conduction electrons form a coherent Kondo Fermi liquid state characterized by a constant susceptibility and a linear magnetization. The occurrence of an intermediate plateau regime at higher values of the field was predicted previously for a Kondo lattice (see Figure 5 in [46]). It can be explained using a strong Kondo coupling picture: independently from the magnetic field, in this Kondo phase a fraction ncn_{c} of the Kondo spins are screened and form local Kondo singlets with the conduction electrons. The remaining fraction 1nc1-n_{c} of unscreened Kondo spins can thus be magnetized without breaking the macroscopic coherent Kondo state. This intermediate Kondo regime with a magnetization plateau is energetically favorable as long as the Kondo singlet energy (typically the Kondo temperature TKT_{K}) is higher than the Zeeman energy (proportional to hxh_{x}).

This plateau corresponds to a Kondo phase with λ\lambda_{\uparrow} and λ\lambda_{\downarrow} slightly different from each other but both non-zero (see figure 2(𝐚{\mathbf{a}})). In the Kondo phase, λ\lambda_{\uparrow} and λ\lambda_{\downarrow} take relatively close values that differ as soon as hxh_{x} is non-zero and both vanish above the same critical field hK{h_{K}^{\star}} which marks the breakdown of Kondo effect.

The variation of hK{h_{K}^{\star}} is depicted in figure 3(a) and in its inset. After an exponential behavior (exp(a/JK)\sim\exp(-a/J_{K})) at small Kondo coupling, it goes linearly with JKJ_{K} for strong Kondo interaction. From the inset, it is possible to verify that the Kondo critical field scales as TKT_{K} for hx=0h_{x}=0 [46]. The figures 3(b) and 3(c) present the variation of the fcf-c effective hybridization λ\lambda_{\uparrow} as a function of JKJ_{K} for fixed values of hxh_{x}, and as a function of hxh_{x} for fixed values of JKJ_{K}, respectively. We find that λ\lambda_{\uparrow} vanishes at small JKJ_{K} as soon as hxh_{x} is non-zero, it abruptly jumps to a finite value around a critical value of JKJ_{K} and then increases continuously with JKJ_{K}. When fixing the Kondo coupling, we find that the effective hybridization is almost constant as hxh_{x} increases and vanishes abruptly at hxhKh_{x}\equiv h_{K}^{\star}. This is consistent with the abrupt jump observed when increasing JKJ_{K}, reflecting the fact that hKh_{K}^{\star} depends on JKJ_{K} in a monotonous way.

Refer to caption
Figure 3: (a) The critical magnetic field, hKh_{K}^{\star}, as a function of JKJ_{K}. The critical field follows an exponential behavior (exp(a/JK)\sim\exp(-a/J_{K})) for small values of JKJ_{K} (inset). For higher values, the JKJ_{K} dependence is linear, as it can also be seen in the inset. (b) λ\lambda_{\uparrow} as a function of JKJ_{K} for different values of hxh_{x}. The effective hybridization is zero for small values of JKJ_{K}. For hx0h_{x}\neq 0, there is a critical value of JKJ_{K} where λ\lambda_{\uparrow} increases abruptly. (c) λ\lambda_{\uparrow} as a function of hxh_{x} for different values of JKJ_{K}. As JKJ_{K} increases, the absolute value of λ\lambda_{\uparrow} increases and the field hxh_{x} above which Kondo pairing is destroyed increases also. J=JH=0J^{\prime}=J_{H}=0 for all figures.

3.2 Effect of the intersite magnetic interaction JJ^{\prime} for JH=0J_{H}=0

Here, the effect of a magnetic field along the xx direction is analyzed by considering also the intersite Ising-like interaction JJ^{\prime} for the orbital 1. The Hund’s coupling is not considered here, JH=0J_{H}=0. Therefore the orbital 1 is fully decoupled from the other electrons which, on their side, form a Kondo lattice system. The model for the orbital 2 and the conduction electrons is similar to the one discussed in section 3.1 and we thus choose here to fix JK=1.0J_{K}=1.0 which correspond to hK0.036h_{K}^{\star}\approx 0.036. The evolution of the magnetizations of the two orbitals as well as the fcf-c effective hybridization as a function of the magnetic field hxh_{x} are depicted in figure 4. We consider that the intersite magnetic exchange between electrons in orbital 1 is of the same order of magnitude as the critical field hKh_{K}^{\star}. The resultas for J=0.05J^{\prime}=0.05 in  4(a), and J=0.1J^{\prime}=0.1 in  4(b) are presented in figure 4.

Refer to caption
Figure 4: Variation of magnetization and Kondo coupling as a function of hxh_{x} for different values of JJ^{\prime} with JK=1.0J_{K}=1.0 and JH=0J_{H}=0. (a) J=0.05J^{\prime}=0.05. (b) J=0.1J^{\prime}=0.1. The inclusion of Ising-like interaction gives values of 𝐦𝟏\mathbf{m_{1}} different from zero for hx=0h_{x}=0, but it does not affect what happens in orbital 2. Increasing JJ^{\prime} increases the critical value where m1zm_{1z} goes to zero, and consequently, the region where m1xm_{1x} is different of its saturated value.

The Kondo lattice formed by electrons in the orbital 2 is discussed in detail in section 3.1 with the three regimes, linear, plateau, and saturation. We now focus on the analysis of orbital 1, which is saturated to its maximal value 𝐦𝟏=1/2\|\mathbf{m_{1}}\|=1/2, but not necessarily in the same direction as m2m_{2}. The Ising-like interaction JJ^{\prime} favors a magnetization along the zz direction, while the transverse magnetic field favors alignment along xx. By exploiting the fact that orbital 1 is decoupled from other electrons we can solve exactly the mean-field equation for 𝐦𝟏\mathbf{m_{1}} at zero temperature. Indeed, for JH=0J_{H}=0 and with a field oriented along xx axis, Eq. 16 has two possible solutions: either m1z=0m_{1}^{z}=0 and m1x=1/2m_{1}^{x}=1/2, which is realized at sufficiently large field. Or m1x=hx/Jm_{1}^{x}=h_{x}/J^{\prime} and m1z=14(hxJ)2m_{1}^{z}=\sqrt{\frac{1}{4}-\left(\frac{h_{x}}{J^{\prime}}\right)^{2}} at fields hxh_{x} lower than a critical value hM=J/2h_{M}^{\star}=J^{\prime}/2. This linear increase of m1xm_{1}^{x} with hxh_{x} is depicted in figure 4(a), corresponding to a gradual rotation of the magnetization until its complete alignment along xx axis at hxhMh_{x}\geq h_{M}^{\star}. The model parameters used for the plots in figure 4(a) correspond to hM=0.025h_{M}^{\star}=0.025, which is slightly lower than the other critical field hK0.036h_{K}^{\star}\approx 0.036 characterizing the vanishing of the Kondo fcf-c hybridization and the full magnetization of orbital 2. However, by increasing the value of the Ising interaction JJ^{\prime}, the alignment of 𝐦𝟏\mathbf{m_{1}} along xx axis can also be obtained for higher critical field, inside the non-Kondo regime (hM>hKh_{M}^{\star}>h_{K}^{\star}), as can be seen in figure 4(b) .

3.3 Effect of Hund’s coupling

We now extend our study by considering the effect of the transverse magnetic field in the presence of local Hund’s coupling and intersite Ising interaction. In this case, orbital 1 and 2 interact by the Hund’s coupling and orbital 2 is coupled with the conduction electrons by Kondo interaction. In the presence of magnetic field, the magnetization of orbital 2 will no longer be in the field direction because it is coupled to the magnetization of orbital 1. Moreover, the Weiss mean-field resulting from Hund’s coupling will add to the applied field and it is expected to weaken the effective Kondo hybridization , destroying it for a critical field smaller than hK{h_{K}^{\star}} defined above.

Refer to caption
Figure 5: Variation of magnetization and Kondo coupling as a function of hxh_{x} for different values of JKJ_{K} and JHJ_{H} for fixed J=0.1J^{\prime}=0.1. (𝐚)(\mathbf{a}) JK=1.0J_{K}=1.0 and JH=0.02J_{H}=0.02 ; hK<hMh_{K}^{\star}<h_{M}^{\star} in this case. (𝐛\mathbf{b}) JK=1.3J_{K}=1.3 and JH=0.07J_{H}=0.07; hK>hMh_{K}^{\star}>h_{M}^{\star}. (𝐜\mathbf{c}) JK=1.0J_{K}=1.0 and JH=0.1J_{H}=0.1. When JHJ_{H} is tuned, all parameters are linked. The Hund’s coupling allows ta non-zero component of magnetization in zz direction for the orbital 2. This effect increases the global magnetization and can destroy the Kondo coupling.

The evolutions of the various mean-field parameters as functions of the magnetic field hxh_{x} are depicted in figure 5. We consider two set of parameters corresponding to either hM>hKh_{M}^{\star}>h_{K}^{\star} (Figure 5(𝐚)(\mathbf{a})) or hK>hMh_{K}^{\star}>h_{M}^{\star} (Figure 5(𝐛)(\mathbf{b})) described previously for JHJ_{H}.

The behaviors depicted in figure 5(a) for small JH=0.02J_{H}=0.02 is similar with the one depicted in figure 4(b) for JH=0J_{H}=0. For example, both plateau and saturated regimes are observed for the magnetization of orbital 2, but now for the modulus of m2m_{2} (shown in Figure 6 (𝐚)\mathbf{(a)}). Here, the small but non-zero zz-component comes from a combined effect of Hund’s coupling with orbital 1 together with the Ising intersite interaction ; the variation with field is no longer linear behavior as it was for JH=0.0J_{H}=0.0. Also, some slight discontinuities are observed for the magnetizations m1xm_{1}^{x} and m1zm_{1}^{z} at the critical field hKh_{K}^{\star}, which marks the vanishing of the Kondo fcf-c hybridization.

The figure 5 (𝐛)\mathbf{(b)} obtained for JK=1.3J_{K}=1.3 and JH=0.07J_{H}=0.07 shows an increase of hKh_{K}^{\star} since the Kondo coupling is increased. The discontinuities of m1xm_{1}^{x} and m1zm_{1}^{z} are not observed, and the zz-component of 𝐦2\mathbf{m}_{2} is larger. Also, the stronger value of JKJ_{K} allows stabilization of the mixed state with a magnetic moments of orbital 1 completely aligned to the magnetic field hxh_{x}. The figure 5 (𝐜\mathbf{c}) is obtained for JK=1.0J_{K}=1.0 and JH=0.1J_{H}=0.1. The resulting Weiss field induced by the fully polarized orbital 1 on the orbital 2 has stronger intensity than the previous cases reported before.The presence of Ising interaction together with Hund’s coupling favors magnetizations aligned along zz direction for small hxh_{x}, while the alignment is along xx direction above a critical field hMh_{M}^{\star}. The rotations of the magnetizations induced by the field are continuous and gradual.

The variation of the magnetization at orbital 2 as a function of the transverse magnetic field and the number of conduction electron is depicted in figure 6. For the three values of ncn_{c} shown in figure 6(a), we can see the presence of the plateau already discussed at (1nc)/2\sim(1-n_{c})/2. In figure 6(b), the linear variation of m2m_{2} is observed for some range of ncn_{c}, where the plateau is present.

Refer to caption
Figure 6: (𝐚)\mathbf{(a)} Variation of m2m_{2} as function of hxh_{x} for different values of conduction electrons and, (𝐛)\mathbf{(b)} Variation of m2m_{2} as function of the conduction electron number and different values of hxh_{x} for JK=1.0J_{K}=1.0, J=0.1J^{\prime}=0.1, and JH=0.02J_{H}=0.02.

The figure 7 shows the phase diagrams as a function of occupation number, ncn_{c}, and the transverse magnetic field, hxh_{x}, for fixed J=0.1J^{\prime}=0.1 and four combinations of JKJ_{K} and JHJ_{H}. Following the description made in the subsection 3.1, we can relate the different phases with the schematic pictures presented in figure 1. In the white part of Figure 7, the F1 phase is defined as a ferromagetic phase where the zz component of the magnetization is non-zero. In the yellow region, the F2 phase corresponds to a magnetization completely aligned with the applied field. It can be saturated for both orbitals, leading to the fully polarized state. In the red region, FK1 indicates a mixed state: ferromagnetic (non fully polarized) order and Kondo hybridization coexist. Another mixted state is obtained, FK2, which appears in the orange region: here, the magnetization oriented along xx direction coexists with Kondo hybridization. The two diagrams represented by figures 7(𝐚)\mathbf{(a)} and (𝐜)\mathbf{(c)}, which mimic two different values of JHJ_{H}, represent qualitatively similar situations: in both cases, FK1 can be present for sufficiently large value of electronic filling ncn_{c} and small field; this phase is limited by the critical field hKh_{K}^{\star} above which the field always abruptly destroys the Kondo hybridization. Also, here, an increase of JHJ_{H} results in a decrease of hKh_{K}^{\star}, which vanishes at small ncn_{c} where non Kondo state can be formed. In all cases, a fully polarized non-Kondo state F2 is realized for strong applied manetic field. We now focus onto the diagrams obtained for larger values of Kondo interaction. In 7(𝐛)\mathbf{(b)}, an intermediate situation is observed with JKJ_{K} slightly stronger than the one in 7(𝐚)\mathbf{(a)}. In this case, a direct transition from FK1 to F2 may be realized by applying a magnetic field if ncn_{c} is sufficiently large. Such a FK1-F2 transition corresponds to hK=hMh_{K}^{\star}=h_{M}^{\star} and, as a consequence, all parameters are expected to vary abruptly, especially magnetization (value and orientation), electronic density of states, Fermi-surface. Figure 7(𝐝)\mathbf{(d)} depicts a situation obtained where both Kondo and Hund interactions are relatively stronger than in other figures. In this case, a cascade of transitions FK1-FK2-F2 can be induced by the magnetic field: as a first step, from FK1 to FK2, the magnetic field continuously turns the magnetization to xx direction, preserving continuously the non-zero Kondo hybridization just above hMh_{M}^{\star}. Then, in the intermediate Kondo phase FK2, the xx component of 𝐦𝟐{\bf m_{2}} reaches the plateau behaviour. The breakdown of Kondo effect can then be realized (FK2-F2 transition) at the critical field hK>hMh_{K}^{\star}>h_{M}^{\star}.

Refer to caption
Figure 7: Phase diagram as a function of the conduction electron occupation, ncn_{c}, and the transverse magnetic field, hxh_{x}, for J=0.1J^{\prime}=0.1, T=0.0001T=0.0001, and different values of JKJ_{K} and JHJ_{H}. Depending on the interaction parameters different phases are stabilized: ferromagnetism and Kondo effect may compete or coexist. Strong values of the applied field always induce a fully polarized state, F2, with magnetization aligned along the field. At zero field, a ferromagnetic phase is always found due to the Ising interaction JJ^{\prime}, coexisting with Kondo screening in the FK1 state at sufficiently large ncn_{c}. (a) and (c) are obtained for relatively small JK=1.0J_{K}=1.0. Here, hK<hMh_{K}^{\star}<h_{M}^{\star} and the field-induced breakdown of Kondo effect, FK1-F1, is realized in a ferromagnetic phase where the magnetization is not fully aligned along the field. (b) is obtained for a slightly larger value of Kondo interaction. Here, a direct transition FK1-F2 can be realized, reflecting hK=hMh_{K}^{\star}=h_{M}^{\star}, and corresponding to a simultaneous breakdown of Kondo effect and allignement of the magnetizations along the field direction. (d) is obtained for relatively stronger values of JKJ_{K} and JHJ_{H}. Here, a cascade of transitions may be realized, FK1-FK2-F2. In the intermediate mixed Kondo phase FK2, the magnetization is aligned along the applied field.

4 Discussion and conclusion

The description presented in this paper is intended to shed light on the discussion of the metamagnetic transition in the URhGe compound. With this proposal in mind, we studied the effect of the transverse magnetic field in a model with two ff-electron orbitals that interact with onsite and intersite exchange, and are in contact with a background of conduction electrons. Here, the interplay between Kondo effect and ferromagnetic order has a strong influence on the properties of the system. We focused on the tunability of the ground state properties by applying a transverse external magnetic field. The two main possible effects that might be induced by applying a transverse magnetic field in this system are: breaking of the Kondo effect, and rotating the magnetization axis along the direction of the field. We obtained and characterized different scenarios for the predicted field-induced transitions. When the critical field hKh_{K}^{\star} (field necessary to destroy Kondo effect) is smaller than hMh_{M}^{\star} (field necessary to rotate the spins along xx axis) (see, e.g., Figure 7(a)), the Kondo effect is destroyed before the field completes the rotation of the magnetization along xx. In this case, we predict an abrupt transition from phase FK1 to phase F1 (at hKh_{K}^{\star}), followed by a continuous rotation of the magnetization as a function of the transverse field (ended at hMh_{M}^{\star}) up to the fully polarized F2 phase. On the other hand, when hKh_{K}^{\star} is bigger than hMh_{M}^{\star} (see, e.g., Figure 7(d)), the complete rotation of the spins occurs inside the Kondo phase (m2zm_{2z} is zero but m2xm_{2x} is not saturated) from phase FK1 to phase FK2 and the metamagnetic transition indicating the breakdown of Kondo effect is obtained only when the magnetization is already along the xx direction, from phase FK2 to F2. For an intermediate set of parameters (see, e.g., Figure 7(b)), the critical fields can coincide, hK=hMh_{K}^{\star}=h_{M}^{\star}, and the complete rotation of the magnetization is expected to occur at the same field as the Kondo breakdown metamagnetic transition, from phase FK1 to phase F2.

These different situations, resulting from different values of the dimensionless ratio hK/hMh_{K}^{\star}/h_{M}^{\star}, correspond to different scenarii, which are associated with different experimental signatures: focusing on the magnetic signatures, Kondo breakdown at hKh_{K}^{\star} is expected to be revealed by a metamagnetic transition, while hMh_{M}^{\star} marks the full rotation of the magnetization along the direction of the applied field. Analyzing magnetization curves as a function of transverse field should thus provide a clear way to discriminate between the different scenarii. Other signatures of Kondo effect, like the expected increase of the effective mass, also need to be analyzed coherently with the magnetization.

Regarding the discussion of the experimental results in URhGe, we identify our theoretical parameter hMh_{M}^{\star} with HRH_{R} (spin-reorientation field). This is consistent with the phenomenological analysis given in [47], using Landau free energy expansion. Experimentally, an increase of the effective mass around HRH_{R} is revealed in different contexts: it is reported by specific heat [48] that γ\gamma has a peak, and the constant AA of the T2T^{2} relation on the resistivity, has similar behavior [49]. Assuming the Kadowaki-Woods relation Am\sqrt{A}\sim m^{*} and considering that the effective mass has two main contributions, a band mass mBm_{B} and a magnetic contribution mm^{**}, m=mB+mm^{*}=m_{B}+m^{**}, it was proposed  [48] that, when the effective mass increases, it is the magnetic part due to the magnetic fluctuations that is responsible of this increase. The band mass contribution is considered as field independent and does not change across the transition [50]. Also, the Fermi surface presents a variation of around 7% detected by quantum oscillations (Shubnikov-Das Haas experiment) at HRH_{R} and it is reported that the effective mass decreases when crossing HRH_{R} and stays constant for higher fields [51]. Measurements of Hall effect also indicates a Fermi surface reconstruction [52], together with the results of thermoeletric power that changes sign around HRH_{R} [49]. ARPES shows itinerant behavior for the 5f5f electrons in the ferromagnetic state [53], but there are no result above to the metamagnetic transition.

On the basis of these experimental results, we propose two possible descriptions : first, a scenario with hM<hKh_{M}^{\star}<h_{K}^{\star} as depicted in Figure 7(d), can be realized if experimentaly Kondo effect is established above and below HRH_{R}. However, in this configuration, we would not observe any abrupt variation in the magnetization at that field, and the metamagnetic transition might occur only for higher values of the transverse field when hx=hKh_{x}=h_{K}^{\star}. The second scenario corresponds to hK=hMh_{K}^{\star}=h_{M}^{\star} (see for example 7(b)), where the rotation is accompanied with a sudden variation of the magnetization along the applied transversal field. This could be the case if the effective mass has an abrupt change across the metamagnetic transition.

The Das Haas-van Alphen experiment on URhGe could help in understanding the nature of the metamagnetic transition discussed here, analyzing how the effective band mass changes across HRH_{R}. Finally, we did not take into account any possible effect on the metamagnetic transition due to the reentrant superconducting phase, but we strongly believe that the interaction of the itinerant and localized electrons is important for the description of the magnetic and superconducting effects in the uranium compounds.

ACKNOWLEDGMENTS

C.T. acknowledges partial support provided by the Brazilian-France Agreement CAPES-COFECUB (No. 88881.192345/2018-01). We would like to thank the COTEC (CBPF) for making available their facilities for the numerical calculations on the Cluster HPC.

References

  • [1] G. Zwicknagl, A. N. Yaresko, P. Fulde, Microscopic description of origin of heavy quasiparticles in UPt 3, Physical Review B 65 (8) (2002) 081103. doi:10.1103/PhysRevB.65.081103.
  • [2] G. Zwicknagl, A. Yaresko, P. Fulde, Fermi surface and heavy masses for UPd 2 Al 3, Physical Review B 68 (5) (2003) 052508. doi:10.1103/PhysRevB.68.052508.
  • [3] P. Fulde, P. Thalmeier, G. Zwicknagl, Strongly correlated electrons, arXiv:cond-mat/0607165 (Jul. 2006). arXiv:cond-mat/0607165, doi:10.1016/S0081-1947(06)80003-2.
  • [4] A. D. Huxley, Ferromagnetic superconductors, Physica C: Superconductivity and its Applications 514 (2015) 368–377. doi:10.1016/j.physc.2015.02.026.
  • [5] D. Aoki, K. Ishida, J. Flouquet, Review of U-based Ferromagnetic Superconductors: Comparison between UGe 2 , URhGe, and UCoGe, Journal of the Physical Society of Japan 88 (2) (2019) 022001. doi:10.7566/JPSJ.88.022001.
  • [6] A. C. Hewson, The Kondo Problem to Heavy Fermions, 1st Edition, Cambridge University Press, 1993. doi:10.1017/CBO9780511470752.
  • [7] S. Doniach, Phase Diagram for the Kondo Lattice, in: R. D. Parks (Ed.), Valence Instabilities and Related Narrow-Band Phenomena, Springer US, Boston, MA, 1977, pp. 169–176. doi:10.1007/978-1-4615-8816-0_15.
  • [8] Q. Si, F. Steglich, Heavy Fermions and Quantum Phase Transitions, Science 329 (5996) (2010) 1161–1166. doi:10.1126/science.1191195.
  • [9] S. Kirchner, S. Paschen, Q. Chen, S. Wirth, D. Feng, J. D. Thompson, Q. Si, Colloquium : Heavy-electron quantum criticality and single-particle spectroscopy, Reviews of Modern Physics 92 (1) (2020) 011002. doi:10.1103/RevModPhys.92.011002.
  • [10] V. S. Zapf, E. J. Freeman, E. D. Bauer, J. Petricka, C. Sirvent, N. A. Frederick, R. P. Dickey, M. B. Maple, Coexistence of superconductivity and antiferromagnetism in CeRh 1 - x Co x In 5, Physical Review B 65 (1) (2001) 014506. doi:10.1103/PhysRevB.65.014506.
  • [11] R. Settai, H. Shishido, S. Ikeda, Y. Murakawa, M. Nakashima, D. Aoki, Y. Haga, H. Harima, Y. Onuki, Quasi-two-dimensional Fermi surfaces and the de Haas-van Alphen oscillation in both the normal and superconducting mixed states of CeCoIn 5, Journal of Physics: Condensed Matter 13 (27) (2001) L627–L634. doi:10.1088/0953-8984/13/27/103.
  • [12] H. Shishido, R. Settai, D. Aoki, S. Ikeda, H. Nakawaki, N. Nakamura, T. Iizuka, Y. Inada, K. Sugiyama, T. Takeuchi, K. Kindo, T. C. Kobayashi, Y. Haga, H. Harima, Y. Aoki, T. Namiki, H. Sato, Y. Ōnuki, Fermi Surface, Magnetic and Superconducting Properties of LaRhIn 5 and CeTIn 5 (T: Co, Rh and Ir), Journal of the Physical Society of Japan 71 (1) (2002) 162–173. doi:10.1143/JPSJ.71.162.
  • [13] H. Aoki, S. Uji, A. K. Albessard, Y. Ōnuki, Transition of f electron nature from itinerant to localized: Metamagnetic transition in CeRu 2 Si 2 studied via the de Haas–van Alphen effect, Physical Review Letters 71 (13) (1993) 2110–2113. doi:10.1103/PhysRevLett.71.2110.
  • [14] R. Daou, C. Bergemann, S. R. Julian, Continuous Evolution of the Fermi Surface of CeRu 2 Si 2 across the Metamagnetic Transition, Physical Review Letters 96 (2) (2006) 026401. doi:10.1103/PhysRevLett.96.026401.
  • [15] M. Güttler, K. Kummer, K. Kliemt, C. Krellner, S. Seiro, C. Geibel, C. Laubschat, Y. Kubo, Y. Sakurai, D. V. Vyalikh, A. Koizumi, Visualizing the Kondo lattice crossover in YbRh 2 Si 2 with Compton scattering, Physical Review B 103 (11) (2021) 115126. doi:10.1103/PhysRevB.103.115126.
  • [16] N. B. Perkins, M. D. Núñez-Regueiro, B. Coqblin, J. R. Iglesias, The underscreened Kondo lattice model applied to heavy fermion uranium compounds, Physical Review B 76 (2007) 125101. doi:dx.doi.org/10.1103/PhysRevB.76.125101.
  • [17] C. Thomas, A. S. da Rosa Simões, C. Lacroix, J. R. Iglesias, B. Coqblin, Effect of anisotropy in the S=1 underscreened Kondo lattice, Journal of Magnetism and Magnetic Materials 372 (2014) 247–252. doi:10.1016/j.jmmm.2014.07.028.
  • [18] Ph. Nozières, A. Blandin, Kondo effect in real metals, Journal de Physique 41 (3) (1980) 193–211. doi:10.1051/jphys:01980004103019300.
  • [19] M. Fidrysiak, D. Goc-Jagło, E. Kądzielawa-Major, P. Kubiczek, J. Spałek, Coexistent spin-triplet superconducting and ferromagnetic phases induced by Hund’s rule coupling and electronic correlations: Effect of the applied magnetic field, Physical Review B 99 (20) (2019) 205106. doi:10.1103/PhysRevB.99.205106.
  • [20] A. B. Shick, W. E. Pickett, Spin-orbit coupling induced degeneracy in the anisotropic unconventional superconductor UTe 2, Physical Review B 100 (13) (2019) 134502. doi:10.1103/PhysRevB.100.134502.
  • [21] S. S. Saxena, P. Agarwal, K. Ahilan, F. M. Grosche, R. K. Haselwimmer, M. J. Steiner, E. Pugh, I. R. Walker, S. R. Julian, P. Monthoux, G. G. Lonzarich, A. Huxley, I. Sheikin, D. Braithwaite, J. Flouquet, Superconductivity on the border of itinerant-electron ferromagnetism in UGe2, Nature 406 (6796) (2000) 587–92. arXiv:10949292, doi:10.1038/35020500.
  • [22] D. Aoki, A. Huxley, E. Ressouche, D. Braithwaite, J. Flouquet, J.-P. Brison, E. Lhotel, C. Paulsen, Coexistence of superconductivity and ferromagnetism in URhGe, Nature 413 (6856) (2001) 613–616. doi:10.1038/35098048.
  • [23] N. T. Huy, A. Gasparini, D. E. de Nijs, Y. Huang, J. C. P. Klaasse, T. Gortenmulder, A. de Visser, A. Hamann, T. Görlach, H. v. Löhneysen, Superconductivity on the Border of Weak Itinerant Ferromagnetism in UCoGe, Physical Review Letters 99 (6) (2007) 067006. doi:10.1103/PhysRevLett.99.067006.
  • [24] R. Troć, Z. Gajek, A. Pikul, Dualism of the 5 f electrons of the ferromagnetic superconductor UGe 2 as seen in magnetic, transport, and specific-heat data, Physical Review B 86 (22) (2012) 224403. doi:10.1103/PhysRevB.86.224403.
  • [25] J. Schoenes, B. Frick, O. Vogt, Transport properties of uranium monochalcogenide and monopnictide single crystals, Physical Review B 30 (11) (1984) 6578. doi:dx.doi.org/10.1103/PhysRevB.30.6578.
  • [26] S. Ran, C. Eckberg, Q.-P. Ding, Y. Furukawa, T. Metz, S. R. Saha, I.-L. Liu, M. Zic, H. Kim, J. Paglione, N. P. Butch, Nearly ferromagnetic spin-triplet superconductivity, Science 365 (6454) (2019) 684–687. doi:10.1126/science.aav8645.
  • [27] S. Sundar, S. Gheidi, K. Akintola, A. M. Côté, S. R. Dunsiger, S. Ran, N. P. Butch, S. R. Saha, J. Paglione, J. E. Sonier, Coexistence of ferromagnetic fluctuations and superconductivity in the actinide superconductor UTe 2, Physical Review B 100 (14) (2019) 140502. doi:10.1103/PhysRevB.100.140502.
  • [28] D. Aoki, F. Honda, G. Knebel, D. Braithwaite, A. Nakamura, D. Li, Y. Homma, Y. Shimizu, Y. J. Sato, J.-P. Brison, J. Flouquet, Multiple Superconducting Phases and Unusual Enhancement of the Upper Critical Field in UTe 2, Journal of the Physical Society of Japan 89 (5) (2020) 053705. doi:10.7566/JPSJ.89.053705.
  • [29] A. Miyake, Y. Shimizu, Y. J. Sato, D. Li, A. Nakamura, Y. Homma, F. Honda, J. Flouquet, M. Tokunaga, D. Aoki, Metamagnetic transition in heavy fermion superconductor UTe2, Journal of the Physical Society of Japan 88 (6) (2019) 063706. doi:10.7566/jpsj.88.063706.
  • [30] Q. Niu, G. Knebel, D. Braithwaite, D. Aoki, G. Lapertot, M. Vališka, G. Seyfarth, W. Knafo, T. Helm, J.-P. Brison, J. Flouquet, A. Pourret, Evidence of Fermi surface reconstruction at the metamagnetic transition of the strongly correlated superconductor UTe 2, Physical Review Research 2 (3) (2020) 033179. doi:10.1103/PhysRevResearch.2.033179.
  • [31] D. Aoki, A. Nakamura, F. Honda, D. Li, Y. Homma, Y. Shimizu, Y. J. Sato, G. Knebel, J.-P. Brison, A. Pourret, D. Braithwaite, G. Lapertot, Q. Niu, M. Vališka, H. Harima, J. Flouquet, Unconventional superconductivity in heavy fermion UTe2, Journal of the Physical Society of Japan 88 (4) (2019) 043702. doi:10.7566/jpsj.88.043702.
  • [32] J. Ishizuka, S. Sumita, A. Daido, Y. Yanase, Insulator-Metal Transition and Topological Superconductivity in UTe 2 from a First-Principles Calculation, Physical Review Letters 123 (21) (2019) 217001. doi:10.1103/PhysRevLett.123.217001.
  • [33] A. B. Shick, S.-i. Fujimori, W. E. Pickett, UTe 2 : A nearly insulating half-filled j = 5 2 5 f 3 heavy-fermion metal, Physical Review B 103 (12) (2021) 125136. doi:10.1103/PhysRevB.103.125136.
  • [34] C. Lacroix, M. Cyrot, Phase diagram of the Kondo lattice, Physical Review B 20 (5) (1979) 1969–1976. doi:10.1103/PhysRevB.20.1969.
  • [35] P. Fazekas, E. Müller-Hartmann, Magnetic and non-magnetic ground states of the Kondo lattice, Zeitschrift für Physik B Condensed Matter 85 (2) (1991) 285–300. doi:10.1007/BF01313231.
  • [36] G.-B. Li, G.-M. Zhang, L. Yu, Kondo screening coexisting with ferromagnetic order as a possible ground state for Kondo lattice systems, Physical Review B 81 (9) (2010) 094420. doi:10.1103/PhysRevB.81.094420.
  • [37] R. Peters, N. Kawakami, T. Pruschke, Spin-Selective Kondo Insulator: Cooperation of Ferromagnetism and the Kondo Effect, Physical Review Letters 108 (8) (2012) 086402. doi:10.1103/PhysRevLett.108.086402.
  • [38] Y. Liu, G.-M. Zhang, L. Yu, Weak ferromagnetism with the Kondo screening effect in the Kondo lattice systems, Physical Review B 87 (13) (2013) 134409. doi:10.1103/PhysRevB.87.134409.
  • [39] D. Golež, R. Žitko, Lifshitz phase transitions in the ferromagnetic regime of the Kondo lattice model, Physical Review B 88 (5) (2013) 054431. doi:10.1103/PhysRevB.88.054431.
  • [40] C. Thomas, A. S. da Rosa Simões, J. R. Iglesias, C. Lacroix, N. B. Perkins, B. Coqblin, Application of the S = 1 underscreened Anderson lattice model to Kondo uranium and neptunium compounds, Physical Review B 83 (1) (2011) 014415. doi:10.1103/PhysRevB.83.014415.
  • [41] B. H. Bernhard, C. Lacroix, Coexistence of magnetic order and Kondo effect in the Kondo-Heisenberg model, Physical Review B 92 (9) (2015) 094401. doi:10.1103/PhysRevB.92.094401.
  • [42] K. Suzuki, K. Hattori, Ground-state phase diagram of the $S = 1$ one-dimensional Kondo lattice model with a uniaxial anisotropy under transverse fields, Journal of the Physical Society of Japan 88 (2) (2019) 024707. arXiv:1808.10142, doi:10.7566/JPSJ.88.024707.
  • [43] K. Suzuki, K. Hattori, Superconducting correlations in the one-dimensional kondo lattice models under magnetic fields, Journal of the Physical Society of Japan 89 (3) (2020) 034703. doi:10.7566/jpsj.89.034703.
  • [44] B. H. Bernhard, Metamagnetism and tricritical behavior in the Kondo lattice model, Physical Review B 106 (5) (2022) 054436. doi:10.1103/PhysRevB.106.054436.
  • [45] F. Lévy, I. Sheikin, B. Grenier, A. D. Huxley, Magnetic Field-Induced Superconductivity in the Ferromagnet URhGe, Science 309 (5739) (2005) 1343–1346. doi:10.1126/science.1115498.
  • [46] S. Burdin, V. Zlatić, Multiple temperature scales of the periodic Anderson model: Slave boson approach, Physical Review B 79 (11) (2009) 115139. doi:10.1103/PhysRevB.79.115139.
  • [47] V. P. Mineev, Metamagnetic phase transition in the ferromagnetic superconductor URhGe, Physical Review B 103 (14) (2021) 144508. doi:10.1103/PhysRevB.103.144508.
  • [48] F. Hardy, D. Aoki, C. Meingast, P. Schweiss, P. Burger, H. v. Löhneysen, J. Flouquet, Transverse and longitudinal magnetic-field responses in the Ising ferromagnets URhGe, UCoGe, and UGe 2, Physical Review B 83 (19) (2011) 195107. doi:10.1103/PhysRevB.83.195107.
  • [49] A. Gourgout, A. Pourret, G. Knebel, D. Aoki, G. Seyfarth, J. Flouquet, Collapse of Ferromagnetism and Fermi Surface Instability near Reentrant Superconductivity of URhGe, Physical Review Letters 117 (4) (2016) 046401. doi:10.1103/PhysRevLett.117.046401.
  • [50] A. Miyake, D. Aoki, J. Flouquet, Field Re-entrant Superconductivity Induced by the Enhancement of Effective Mass in URhGe, Journal of the Physical Society of Japan 77 (9) (2008) 094709. doi:10.1143/JPSJ.77.094709.
  • [51] E. A. Yelland, J. M. Barraclough, W. Wang, K. V. Kamenev, A. D. Huxley, High-field superconductivity at an electronic topological transition in URhGe, Nature Physics 7 (11) (2011) 890–894. doi:10.1038/nphys2073.
  • [52] D. Aoki, G. Knebel, J. Flouquet, Fermi Surface Instabilities in Ferromagnetic Superconductor URhGe, Journal of the Physical Society of Japan 83 (9) (2014) 094719. arXiv:1407.5799, doi:10.7566/JPSJ.83.094719.
  • [53] S.-i. Fujimori, I. Kawasaki, A. Yasui, Y. Takeda, T. Okane, Y. Saitoh, A. Fujimori, H. Yamagami, Y. Haga, E. Yamamoto, Y. Ōnuki, Itinerant magnetism in URhGe revealed by angle-resolved photoelectron spectroscopy, Physical Review B 89 (10) (2014) 104518. doi:10.1103/PhysRevB.89.104518.