Abstract
We obtain the conditions necessary for the emergence of various low-temperature ordered states (local-moment antiferromagnetism, unconventional superconductivity, quantum criticality, and Landau Fermi liquid behavior) in Kondo lattice materials by extending the two-fluid phenomenological theory of heavy-electron behavior to incorporate the concept of hybridization effectiveness. We use this expanded framework to present a new phase diagram and consistent physical explanation and quantitative description of measured emergent behaviors such as the pressure variation of the onset of local-moment antiferromagnetic ordering at TN, the magnitude of the ordered moment, the growth of superconductivity within that ordered state, the location of a quantum critical point, and of a delocalization line in the pressure/temperature phase diagram at which local moments have disappeared and the heavy-electron Fermi surface has grown to its maximum size. We apply our model to CeRhIn5 and a number of other heavy-electron materials and find good agreement with experiment.
Keywords: heavy fermion, kondo liquid, spin liquid
Heavy-electron materials provide a unique f-electron laboratory for the study of the interplay between localized and itinerant behavior. At comparatively high temperatures, itinerancy emerges as the localized f-electrons collectively reduce their entropy by hybridizing with the conduction electrons to form a new state of matter, an itinerant heavy-electron Kondo liquid that displays scaling (non-Landau Fermi liquid) behavior (1–8). The emergent Kondo liquid coexists with the hybridized spin liquid that describes the lattice of local moments whose magnitude has been reduced by hybridization until one reaches the comparatively low temperatures at which unconventional superconductivity and hybridized local-moment antiferromagnetism compete to determine the ground state of the system over a wide regime of pressures. In practice, the phases often coexist, as experiments on the 115 [CeRhIn5 (9–11) and CeCoIn5 (12)] and 127 [CePt2In7 (13)] Kondo lattice materials demonstrate. A major challenge for the heavy-electron community has been finding a consistent framework and simple phenomenological description of their measured emergent behaviors that include the pressure variation of the onset of antiferromagnetic ordering at TN, the growth of superconductivity within that ordered state, the onset of quantum critical behavior, and the growth of the heavy-electron Fermi surface.
Although progress has been made on developing a phenomenological theory of Kondo lattice materials (6), we do not yet possess a microscopic theory of the emergence of the Kondo liquid as a new quantum state of matter that displays universal behavior below a characteristic temperature (1–3, 5), T∗, or a satisfactory general framework for characterizing the conditions necessary for the emergence of the competing low-temperature ordered states, because experiments (4) have shown that the seminal phase diagram proposed by Doniach (14) does not apply to most materials. In this paper we introduce the concept of “hybridization effectiveness” as the organizing principle responsible for the emergence of low-temperature order and show it leads to a new phase diagram that provides a realistic explanation for the emergence of different low-temperature phases for Kondo lattice materials. The role of hybridization effectiveness in reducing the entropy of the lattice of local moments present at high temperatures can be described quantitatively through the pressure dependence of a simple parameter present in the phenomenological two-fluid description (1–8) of heavy-electron materials whose magnitude can be determined from fits to experimental data. We show that our approach makes possible a consistent physical explanation and quantitative description of the low-temperature behaviors of a number of heavy-electron materials and predicts the presence of a delocalization line in the pressure/temperature phase diagram. It offers promise of serving as a standard model of heavy-electron emergent behaviors and providing the basis for the development of a microscopic theory that explains these.
Model
Hybridization Effectiveness.
In the two-fluid description, the heavy-electron Kondo liquid emerges below a characteristic temperature, T∗, as a collective hybridization-induced instability of the spin liquid that describes the lattice of local moments coupled to background conduction electrons. T∗ is determined by the effective Ruderman–Kittel–Kasuya–Yosida (RKKY) interaction between the nearest-neighbor local moments (4); the emergent scaling behavior of the Kondo liquid is characterized by the heavy-electron order parameter (3),
[1] |
which extends from T∗ to a comparatively low temperature, T0, at which ordering begins to set in (7). Over this temperature range it coexists with the “hybridized” quantum spin liquid, whose individual components (that still interact through the intersite RKKY coupling) possess a spectral weight that diminishes with lowering temperature as fl(T) = 1 - fh(T).
The pressure-dependent parameter, f0, in Eq. 1 provides a direct measure of the ultimate effectiveness of the underlying collective hybridization in reducing the magnitude of strength of each localized f-moment. Importantly, as may be seen in the schematic summary of our findings in Fig. 1, it constrains the nature of the ground state. As long as f0 < 1, hybridization is always incomplete, so antiferromagnetic local-moment order becomes possible at a temperature reduced by hybridization: It takes place in the presence of itinerant heavy electrons that can become superconducting, so that over a wide range of pressures local-moment order coexists with superconductivity of magnetic origin. At f0 = 1, the collective hybridization of local moments becomes complete at T = 0; hence, if superconductivity is suppressed by application of a strong enough magnetic field, one expects to find a quantum critical point, denoting a T = 0 transition between f-electron local order and a fully itinerant heavy-electron state. For f0 > 1, the hybridization of local moments is complete and the Fermi surface will have grown to its maximum size at some finite temperature, TL; Kondo liquid scaling behavior is expected between T∗ and TL, whereas TL is fixed by f0 and T∗. From Eq. 1, we have:
[2] |
For f0 close to 1, because of the proximity of a quantum critical point, immediately below TL one does not expect to find Landau Fermi liquid (FL) behavior; instead, the properties of the heavy electrons will be determined by their coupling to quantum critical fluctuations; anomalous behavior, including further mass enhancement, continues until one reaches the still lower characteristic temperature, TFL, shown in Fig. 1, at which transport and other properties are no longer dominated by the coupling of quasiparticles to quantum critical fluctuations, and one finds Landau Fermi liquid behavior.
The Hybridized Quantum Spin Liquid.
The magnetic properties of the hybridized spin liquid can be studied by assuming its dynamic spin susceptibility takes a mean-field form,
[3] |
where χ0 is the local susceptibility of an individual f-moment, z is the coordination number, γl is the local relaxation rate, and Jq is the RKKY exchange coupling. For simplicity, the magnetic moment of the f-spins and the Boltzmann constant are set to unity. The hybridized spin liquid will begin to order at a Néel temperature that is, according to Eq. 3, determined by:
[4] |
All local physics can be included in the local-moment static susceptibility χ0. For example, crystal field effects become important when the crystal field splitting from the ground-state doublet is smaller than or comparable to T∗; if the crystal field configuration is known and plays a role, it can be easily taken into account in χ0. In this paper, we take for simplicity χ0 = C/T, where C is the Curie constant. For T > T∗, Eq. 3 reduces to the Curie–Weiss form of the static susceptibility, χl(0,0) = C/(T + θ), where θ = CzJq=0. We have then
[5] |
where the parameter η = CzJQ/T∗ reflects the effect of the crystal lattice and magnetic frustration on TN, whereas fl(TN) reflects the role played by collective hybridization in reducing the Néel temperature. Because of the RKKY nature of both T∗ and JQ (4), η is pressure-independent.
The magnitude of the ordered moment μ2 is determined by the local-moment order parameter:
[6] |
where μ0 is the local-moment strength above T∗. Where accurate measurements of the strength of the ordered moment in the antiferromagnetic state exist, a good test of our model is to combine Eqs. 5 and 6 to obtain an expression that relates three experimentally measurable quantities at a fixed value of η,
[7] |
Because T∗ increases with increasing pressure, whereas μ decreases, it follows that in general the TN versus pressure curve should exhibit the maximum shown in Fig. 1.
In applying Eqs. 5 and 6 to experiment, one needs to take relocalization effects into account. As discussed in apRoberts-Warren et al. (7) and Shirer et al. (8), Knight shift experiments demonstrate that Kondo liquid scaling ends at a temperature T0, below which local-moment ordering begins to win out over collective hybridization in determining the temperature evolution of fl(T). As the temperature is lowered, the rate of collective hybridization is first reduced and then reversed—i.e., the spectral weight of fl(T) begins to increase as the temperature is further lowered. For the two materials for which this relocalization has been explored in detail, it turns out that fl(TN) is approximately fl(T0) and T0 is approximately 2TN.
Determining f0.
We can use Eq. 3 to determine f0 from measurements of the static spin susceptibility if we make the physically reasonable assumption that the hybridized spin liquid contribution is dominant for a broad range of temperatures below T∗. Such approximation is reasonable because of the large magnetic response of the local moments seen at temperatures above T∗. Fig. 2 shows the evolution of χl(0,0) with f0 and the fit to experimental data in several compounds—from the antiferromagnet CeRhIn5 (15) to the superconductor CeCoIn5 (16) to the 5f compound URu2Si2 that exhibits hidden order (17). We find f0 = 0.95 for CeRhIn5, 1.0 for CeCoIn5, and 1.6 for URu2Si2. With increasing f0, we find a crossover from peak to plateau to peak structure in the uniform susceptibility of exactly the kind that has been observed in many heavy-electron materials. However, the origin of the peak seen in materials with f0 < 1, such as CeRhIn5, and those in which f0 > 1, such as URu2Si2, is different. For f0 > 1, the peak comes from rapid delocalization of the f-moments at approximately T∗; for f0 < 1, it is related to the onset of relocalization of the Kondo liquid as a precursor of the antiferromagnetic ordering at T0 ≪ T∗ (7, 8).
The Kondo Liquid.
The Kondo liquid contributes to the magnetic entropy and other material properties that display anomalous behavior, as well as physical quantities, such as transport and superconductivity, that originate uniquely in the itinerant f-electrons. The Kondo liquid entropy Sh can be obtained by integrating over the universal scaling expression for its specific heat C/T ∝ [1 + ln(T∗/T)] (3) and recognizing that continuity at T∗ requires that Sh(T∗) = Sl(T∗) = R ln 2 (R is the gas constant). When the scaling behavior of the Kondo liquid ends at a temperature Tx (= TL if f0 > 1), the two-fluid model predicts that the heavy-electron entropy will be given by (6):
[8] |
Below TL, no local moments are present; only heavy electrons are responsible for the measured system behavior and a one-fluid picture is recovered, as may be verified in Knight shift measurement. The Fermi liquid-specific heat coefficient, γh, at lower temperatures can then be estimated:
[9] |
In general, even when f0 < 1 or T > TL, the specific heat coefficient can still be calculated using Eq. 11.
Results
Application to CeRhIn5.
CeRhIn5 provides an excellent proving ground for the approach we have described. Our proposed phase diagram for CeRhIn5 is given in Fig. 3, and resembles closely the experimental pressure/temperature phase diagram derived from the resistivity exponent (18). The Knight shift experiments of Shirer et al. (8) show that fl(TN) ≈ fl(2TN) at ambient pressure; we assume this to be true in the entire pressure range in the following analysis, with fl(2TN) given by the scaling formula in Eq. 1.
Antiferromagnetism.
On applying Eqs. 5, 6, and 7 to CeRhIn5, we obtain the results for the variation of the Néel temperature with pressure, and of the magnitude of the ordered moment found in the Néel state that is given in Fig. 4. Despite its simplicity, our model yields results in good agreement with experiment (19–24). Fig. 4A compares the experimental data of TN and ηT∗fl(2TN). The best fit to the Néel temperature yields η = 0.36 for f0 = 0.95. On taking the strength of the local moments above T∗ to be 0.92 μB (25), η = 0.36 gives the right magnitude of the ordered moment when Eq. 7 is used together with the pressure variation of TN. A corresponding best fit to the results of Aso et al. (23) and Llobet et al. (24) requires that η = 0.64 and 0.32, respectively. Our results are seen to be more consistent with those of Llobet et al. (24).
In our mean-field approach, the influence of the crystalline lattice and magnetic frustration are incorporated in the coordination number z and η = CzJQ/T∗. To illustrate a reasonable range for these parameters, we consider a 2D square lattice for which the RKKY coupling has the usual form Jq = Jn[cos(qx) + cos(qy)] + 2Jnn cos(qx) cos(qy), where Jn and Jnn are the nearest- and next-nearest neighbor couplings, respectively. We have θ = 2Cz|Jn + Jnn| and ηT∗ = 2Cz|Jn - Jnn| for Q = (π,π). If we take the values η = 0.36 and θ/T∗ = 1.4, obtained in the susceptibility fit in Fig. 2B, we find a frustration ratio Jnn/Jn ≈ 0.6. This indicates T∗/2Cz|Jn| of approximately 1.1, consistent with its origin in nearest-neighbor RKKY coupling (4).
Itinerant Electrons and Superconductivity Within the Néel State.
Our model provides a natural explanation for the appearance of superconductivity within the Néel state and the pressure dependence of its transition temperature, Tc. Quite generally, local-moment ordering takes place before the relocalization process is complete. Below TN, a substantial number of heavy electrons coexist with the ordered local moments in the antiferromagnetic phase, a fraction that increases as the pressure increases. For example, in CeRhIn5, despite relocalization, fh(TN) is approximately 0.37 at ambient pressure. How do these heavy electrons lose their entropy within the Néel state? There are two possibilities: (i) The relocalization process continues within the Néel state; or (ii) the heavy-electrons order, most likely in an unconventional superconducting state because there are ample magnetic fluctuations present that can bring this about (26). Both likely occur in practice.
Where superconductivity emerges, it is appealing to argue that, to a first approximation, Tc scales with the density of heavy electrons present, an argument familiar from the well-known Uemura plot for the cuprate superconductors (27). We test this hypothesis by comparing the growth in Tc with pressure with the corresponding growth in the density of heavy electrons. As may be seen in Fig. 4A, the increase in Tc seen from 1 GPa to 1.75 GPa mirrors the increase in the density of heavy electrons able to participate in superconductivity, fh(Tc) = fh(2TN) assuming no further relocalization below TN. Further confirmation of this view of Kondo liquid behavior comes from the scaling of the Knight shift anomaly above and below Tc seen in CeCoIn5 (5), and the observation that Tc is maximum in the vicinity of the critical pressure at which some part of the heavy-electron Fermi surface can begin to localize.
Fermi Liquid Regime.
At pressures p > 2.35 GPa, the hybridization parameter is f0 > 1, so the ground state of CeRhIn5 should be a Fermi liquid if one applies a magnetic field strong enough to suppress superconductivity. At , when hybridization is complete, our model predicts a heavy electron-specific heat that depends only on T∗ and f0 and is given by Eq. 9. TL can be determined from a measurement of the Fermi surface, which should reach its maximum size there. Our analysis at low pressures suggests that f0 may scale with the Kondo coupling, J, in the vicinity of the quantum critical point. If this scaling extends to higher pressures, it leads to the results shown for TL in Fig. 3. Below TL, the f-electrons form a Fermi liquid, but its quasiparticles, being scattered by quantum critical spin fluctuations, do not exhibit the classic properties proposed by Landau until one reaches a lower-temperature TFL.
Spin-Lattice Relaxation Rate.
Nuclear magnetic/quadrupole resonance (NMR/NQR) experiments on the pressure dependence of the spin-lattice relaxation rate provide a further confirmation of the approach developed here. We assume that the spin-lattice relaxation rate between T∗ and T0 is dominated by local-moment behavior and calculate it using the Moriya formula (28),
[10] |
where γ is the gyromagnetic ratio and F(q)2 is the form factor. Following ref. 29, we use , where x, the only free parameter, defines the anisotropy of the hyperfine coupling. On using the values of f0(p) given in Fig. 3 (Inset) and assuming that Jq is close to its value for a 2D square lattice, we obtain the fit to the NQR data in CeRhIn5 shown in Fig. 5. We see there the expected change in slope caused by hybridization for and that our model provides a fit to experiment above T0, with T∗ = 17 K at 0 GPa, 20 K at 0.46 GPa, and 25 K at 1.23 GPa, respectively, and an anisotropy parameter x at ambient pressure of 0.08 that increases to 0.12 at 0.46 GPa and 0.2 at 1.23 GPa, in agreement with earlier estimates (29). The behavior found above T∗ reflects the comparatively weak local-moment interaction; the temperature dependence seen between T0 and TN originates from the precursor antiferromagnetic fluctuations (22) and Kondo liquid relocalization (7), and cannot be calculated using the present mean-field approximation. The local relaxation rate γl has been assumed to be a constant in the fit; a temperature-dependent γl may yield improved agreement between theory and experiment.
Magnetic Entropy.
The magnetic entropy is another important quantity of experimental interest. In many heavy-electron antiferromagnets, one finds a small entropy release at TN despite the presence of large ordered moment. As will be shown below, our expanded two-fluid model explains these apparently contradictory results as a direct consequence of hybridization and yields agreement with experiment for the magnetic entropy of a number of heavy-electron materials.
Above T∗, the magnetic entropy is that of localized f-electron moments in a ground-state doublet that is assumed to have the full magnetic entropy R ln 2. At lower temperatures, both the local moments and the Kondo liquid contribute and the total magnetic entropy is given by ref. 6 as:
[11] |
We expect Eq. 11 to be valid down to a temperature T0, at which low-temperature order begins to compete with hybridization in reducing the local-moment entropy. Above T0 and for arbitrary f0, where Eq. 9 does not apply, the specific heat coefficient can still be estimated from the temperature derivative of Eq. 11. For CeRhIn5, as shown in Fig. 3, a number of different experimental probes suggest that T0 is of order of 2TN; importantly, the measurements by Shirer et al. of the Knight shift anomaly show that Kondo liquid behavior is strongly suppressed below T0, indicating a rapid relocalization of itinerant heavy electrons (7, 8).
The physics below T0 require that Eq. 11 be modified. In addition to a deviation from scaling in the order parameters fh and fl, the contribution of each component to the magnetic entropy is expected to be reduced from its value at T0 by the approach of antiferromagnetic order. On comparing the magnetic entropy calculated from Eq. 11 with experiment for CeRhIn5 (19), we find that a factor of rN of approximately 0.35 has to be introduced at ambient pressure in order to obtain agreement with experiment; a reduction of this size is justified by the similar ratio S(TN)/S(T0) of approximately 0.4 found in experiment (19). As may be seen in Fig. 6, with this factor one obtains good agreement with experiment for pressures up to 1.75 GPa. In contrast, at higher pressures, where superconductivity takes over and the Kondo liquid scaling persists down to Tc, no corrections are needed to Eq. 11, which yields good agreement with experiment for the entropy at Tc.
In calculating the entropy at the superconducting transition deep within the antiferromagnetic phase (1.0 < p < 1.75 GPa) shown in Fig. 6, we have excluded the contribution from the ordered moments, whose fraction, a(Tc) = (1 - Tc/TN)0.5, has been measured (24), and have assumed the Kondo liquid-specific heat does not change between TN and Tc. The entropy release of the coexisting state is then given by
[12] |
and agrees well with experiment.
At first sight, given the comparatively large magnitude of the ordered moment in CeRhIn5 (0.6–0.8 μB), the entropy released by the antiferromagnetic ordering of the local moments appears unexpectedly small, being only approximately 30% R ln 2 (15). A little thought shows that the reduction comes in three sequential stages, corresponding to the entropy release at TN, T0, and T∗, respectively. When antiferromagnetism is destroyed at TN = 3.8 K, only a fraction (approximately 63%) of the spectral weight of the strongly hybridized f-moments is available to become “free”; the rest (approximately 37%) of that spectral weight has already gone to the Kondo liquid, whose entropy is already much reduced because of strong entanglement between conduction and f-electrons. It follows that above TN, the entropy of the freed moments is then only released at a higher temperature T0 of approximately 8 K because of strong magnetic correlations close to the phase transition (8), whereas the total spin entropy will not be fully recovered until all the correlated Kondo liquid has been transformed into uncorrelated f-moments above T∗ = 17 K. The continuous and slight change of the magnetic entropy at the ordering temperature as one moves from antiferromagnetism to superconductivity with increasing pressure range requires a gradual change of character of f-moments and argues against a dramatic change of hybridization across the quantum critical region.
Application to Other Materials.
In applying our model to other prototypical heavy-electron materials, we focus on the relation between f0 and the magnetic/nonmagnetic ground state, and the entropy release of the ordered state (when f0 < 1) or the residual specific heat coefficient of the Fermi liquid (if f0 > 1), because these only require experimental input at ambient pressure. Our results are summarized in Fig. 7 and Table 1, and discussed in the subsections below. Because they are strong hybridizers (f0 > 1), we conclude that UNi2Al3 must be a spin-density wave antiferromagnet in which antiferromagnetic order originates from Fermi surface nesting of the heavy-electron Kondo liquid (30), whereas the hidden-order transition in URu2Si2 must be caused by a Kondo liquid instability (31). As may be seen in Fig. 8, our model [Eq. 9] also yields results for the low-temperature Fermi liquid-specific heat of a number of materials that are surprisingly close to their experimental values (17, 32–35), which demonstrates that, for these materials, the average heavy-electron effective mass is determined by the onset of collective hybridization at T∗ and its effectiveness, f0.
Table 1.
Materials | T∗ (K) | TN/L/c (K) | f0 | Order |
Local-moment order | ||||
CePt2In7 | 41 | 5.6 | 0.4 | AFM |
CePb3 | 16 | 1.1 | 0.5 | AFM |
UPd2Al3 | 60 | 14.3 | 0.8 | AFM |
CeRhIn5 | 17 | 3.8 | 0.95 | AFM |
Kondo liquid order | ||||
CeCoIn5 | 50 | 2.3 | 1.0 | SC |
UPt3 | 25 | 5 | 1.4 | SC |
YbAl3 | 160 | 38 | 1.5 | FL |
URu2Si2 | 65 | 17.5 | 1.6 | HO |
UNi2Al3 | 120 | 39 | 1.8 | SDW |
AFM, antiferromagnetism; SC, superconductivity; FL, Fermi liquid; HO, hidden order; SDW, spin-density wave antiferromagnetism.
CeCoIn5.
For this material, the Kondo liquid temperature T∗ = 50 K has been determined using a number of experimental probes (3, 4). On using Eq. 3, we show in Fig. 2B that f0 of approximately 1 provides a good fit to the susceptibility data (16, 35), a conclusion that is consistent with the expectation that CeCoIn5 is close to a magnetic quantum critical point (36). We find S(Tc) = 0.18 R ln 2, close to the experimental value of 0.2 R ln 2 at Tc = 2.3 K (35). Taking Tc as the cutoff temperature, the predicted specific heat coefficient at Tc and zero field using Eq. 11 is approximately 368 mJ/mol K2, comparable to the experimental value of approximately 290 mJ/mol K2 (17).
URu2Si2.
As may be seen in Fig. 2B, our best fit to the c-axis susceptibility of URu2Si2 gives the hybridization parameter f0 = 1.6, and TL of approximately THO. Below TL, all f-electrons become itinerant so that a single-component picture is recovered and the Knight shift becomes once again proportional to the susceptibility. This allows us to subtract the spin susceptibility of the two components without uncertainty, and the detailed analysis given in ref. 8 yields similar results for f0 and TL. These suggest that most local moments have disappeared by the time one reaches THO, so that the hidden-order transition must have an itinerant and strongly correlated nature, consistent with experimental observation of gap opening (37, 38) and Fermi surface reconstruction at THO (39). Our predicted entropy release at THO is 0.44 R ln 2, somewhat larger than the experimental value of 0.3 mJ/mol K2 (17). The predicted specific heat coefficient without hidden order is approximately 203 mJ/mol K2, close to the experimental extrapolated value of 180 mJ/mol K2.
CePt2In7.
Using θ = T∗ = 41 K, we find f0 = 0.4 from a fit to the susceptibility, in which only the local-moment contribution is included. In an earlier analysis that included the Kondo liquid contribution and assumed a Wilson ratio RW = 2 (7), f0 was found to be approximately 1, a result in obvious contradiction with the magnetic ground state. The contradiction leads us to conclude that one should not in general assume that RW = 2, but, rather, determine it on a case-by-case basis. For this material, NMR measurements point to a relocalization temperature T0 of approximately 14 K, below which the magnetic entropy is strongly suppressed. Our scaling result yields S(TN) = 0.76 R ln 2, which is very much larger than the experimental value of approximately 0.2 R ln 2 (12): Just as was the case with CeRhIn5, relocalization effect is large between T0 = 14 K and TN = 5.6 K, a result that is consistent with the NMR experimental results (7).
CePb3.
The susceptibility of CePb3 has a plateau below T∗ = 16 K, followed by a peak before the magnetic transition at TN = 1.1 K (40). Our best fit gives f0 = 0.5 and θ = 25 K, and yields an entropy release of approximately 0.62 R ln 2 at TN, which is large compared to the experimental value of 0.32 R ln 2, a discrepancy that is likely caused by the rapid suppression of the magnetic entropy between approximately 3 K and TN (see figure 8 in ref. 40).
UPt3.
Our fit to the susceptibility of UPt3 gives f0 = 1.4, θ = 71 K, and T∗ = 25 K. We find TL of approximately 5 K, consistent with the deviation from scaling seen in the Knight shift anomaly (3), and γh = 415 mJ/mol K2, in good agreement with the experimental value of 422 mJ/mol K2 (33).
UM2Al3.
UPd2Al3 and UNi2Al3 are both antiferromagnetic at low temperatures but, as our fit to the susceptibility shows, their physical origins are different. Because of crystal field effects, for UPd2Al3 (4, 41) we can only give a rough estimate of θ = T∗ = 60 K and f0 = 0.8, whereas for UNi2Al3 (34) we find T∗ = 120 K, θ = 420 K, and f0 = 1.8. These quite different values of f0 strongly suggest that the magnetic ground state of UPd2Al3 originates in ordering of its local moments, whereas that of UNi2Al3 must reflect spin-density wave ordering that originates in the Fermi surface nesting of the itinerant Kondo liquid.
Because for both materials the specific heat is enhanced by antiferromagnetic fluctuations only in a narrow range above TN, we can safely neglect the relocalization correction; on doing so we obtain S(TN) of approximately 0.69 R ln 2 for UPd2Al3 and 0.05 R ln 2 for UNi2Al3, in good agreement with the experimental values of 0.67 R ln 2 for UPd2Al3 and 0.12 R ln 2 for UNi2Al3 (30). For UNi2Al3, we find TL = 39 K and γh = 75 mJ/mol K2, somewhat smaller than the experimental value of 150 mJ/mol K2 (16).
YbAl3.
Our fit to the susceptibility of YbAl3 (32) gives f0 = 1.5, θ = 340 K, and T∗ = 160 K. This is consistent with the nonmagnetic ground state of this compound, for which we find TL = 38 K. The specific heat coefficient is estimated to be γh = 62 mJ/mol K2, consistent with the experimental value of 45 mJ/mol K2.
For materials for which f0 cannot be easily obtained, one can still estimate the entropy release by making the approximation f0 = 1 in Eq. 11. For antiferromagnetic compounds with f0 < 1, a larger f0 results in a smaller S(TN), which may cancel the effect of relocalization. Taking CePd2Si2 as an example, we predict S(TN) = 0.62 R ln 2 using TN = 9.9 K and T∗ = 40 K (4), quite close to the experimental value of S(TN) = 0.68 R ln 2 (42, 43). This suggests a rough relation between S(To) and To/T∗ (To = Tc,TN, TL, etc.) that can be easily applied to other materials in the future.
Conclusions
In summary, our introduction of hybridization effectiveness into the two-fluid description yields a viable alternative to the Doniach phase diagram (14) for heavy electrons, and its determination from experiment leads to a unified phenomenological explanation of heavy-electron magnetic properties. It will be important to extend the present analysis to a number of other heavy-electron materials and to determine the expected magnetic-field dependence of the hybridization-effectiveness parameter, f0. An analysis of future experiments should yield a hybridization-based 3D pressure/magnetic-field phase diagram, tell us whether it is possible to place all heavy-electron materials on it, and establish the localization line for those materials with f0 > 1. Experiments may also provide further information on the physical origin of f0—e.g., does it scale with Jρ and so mirror the Kondo coupling? At this stage, the improved two-fluid description presented here shows considerable promise as a candidate phenomenological standard model of heavy-electron emergent behaviors and should provide the basis for the development of a microscopic theory that explains these behaviors as a natural consequence of the coupling of the local-moment spin liquid to the conduction electron sea in which it is immersed.
ACKNOWLEDGMENTS.
D.P. thanks colleagues at the Aspen Center for Physics for useful discussions during the writing of this paper. Y.-f.Y. is supported by National Science Foundation–China (Grant 11174339) and Chinese Academy of Sciences, and D.P. has been supported by branch-member contributions to the Institute for Complex Adaptive Matter.
Footnotes
References
- 1.Nakatsuji S, Pines D, Fisk Z. Two fluid description of the Kondo lattice. Phys Rev Lett. 2004;92:016401. doi: 10.1103/PhysRevLett.92.016401. [DOI] [PubMed] [Google Scholar]
- 2.Curro NJ, Young B-L, Schmalian J, Pines D. Scaling in the emergent behavior of heavy-electron materials. Phys Rev B Condens Matter Mater Phys. 2004;70:235117. [Google Scholar]
- 3.Yang Y-F, Pines D. Universal behavior in heavy-electron materials. Phys Rev Lett. 2008;100:096404. doi: 10.1103/PhysRevLett.100.096404. [DOI] [PubMed] [Google Scholar]
- 4.Yang Y-F, Fisk Z, Lee H-O, Thompson JD, Pines D. Scaling the Kondo lattice. Nature. 2008;454:611–613. doi: 10.1038/nature07157. [DOI] [PubMed] [Google Scholar]
- 5.Yang Y-F, Urbano R, Curro NJ, Pines D, Bauer ED. Magnetic excitations in the Kondo liquid: Superconductivity and hidden magnetic quantum critical fluctuations. Phys Rev Lett. 2009;103:197004. doi: 10.1103/PhysRevLett.103.197004. [DOI] [PubMed] [Google Scholar]
- 6.Yang Y-F, Curro NJ, Fisk Z, Pines D, Thompson JD. A predictive standard model for heavy electron systems. J Phys Conf Ser. 2011;273:012066. [Google Scholar]
- 7.apRoberts-Warren N, et al. Kondo liquid emergence and relocalization in the approach to antiferromagnetic ordering in CePt2In7. Phys Rev B Condens Matter Mater Phys. 2011;83:060408. [Google Scholar]
- 8.Shirer KR, et al. Long range order and two-fluid behavior in heavy electron materials. Proc Natl Acad Sci USA. 2012;45:3067–3073. doi: 10.1073/pnas.1209609109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Mito T, et al. Coexistence of antiferromagnetism and superconductivity near the quantum criticality of the heavy-fermion compound CeRhIn5. Phys Rev Lett. 2003;90:077004. doi: 10.1103/PhysRevLett.90.077004. [DOI] [PubMed] [Google Scholar]
- 10.Kawasaki S, et al. Gapless magnetic and quasiparticle excitations due to the coexistence of antiferromagnetism and superconductivity in CeRhIn5: A study of 115In NQR under pressure. Phys Rev Lett. 2003;91:137001. doi: 10.1103/PhysRevLett.91.137001. [DOI] [PubMed] [Google Scholar]
- 11.Park T, et al. Hidden magnetism and quantum criticality in the heavy fermion superconductor CeRhIn5. Nature. 2006;440:65–68. doi: 10.1038/nature04571. [DOI] [PubMed] [Google Scholar]
- 12.Kenzelmann M, et al. Coupled superconducting and magnetic order in CeCoIn5. Science. 2008;321:1652–1654. doi: 10.1126/science.1161818. [DOI] [PubMed] [Google Scholar]
- 13.Bauer ED, et al. Pressure-induced superconducting state and effective mass enhancement near the antiferromagnetic quantum critical point of CePt2In7. Phys Rev B Condens Matter Mater Phys. 2010;81:180507. [Google Scholar]
- 14.Doniach S. The Kondo lattice and weak antiferromagnetism. Physica B Condens Matter. 1977;91:231–234. [Google Scholar]
- 15.Hegger H, et al. Pressure-induced superconductivity in quasi-2D CeRhIn5. Phys Rev Lett. 2000;84:4986–4989. doi: 10.1103/PhysRevLett.84.4986. [DOI] [PubMed] [Google Scholar]
- 16.Curro NJ, et al. Anomalous NMR magnetic shifts in CeCoIn5. Phys Rev B Condens Matter Mater Phys. 2001;64:180514. [Google Scholar]
- 17.Yokoyama M, Tenya K, Amitsuka A. Thorium dilution effects of the heavy electron compound URu2Si2. Physica B Condensed matter. 2002;312–313:498–500. [Google Scholar]
- 18.Park T, et al. Isotropic quantum scattering and unconventional superconductivity. Nature. 2008;456:366–368. doi: 10.1038/nature07431. [DOI] [PubMed] [Google Scholar]
- 19.Park T, Thompson JD. Magnetism and superconductivity in strongly correlated CeRhIn5. New J Phys. 2009;11:055062. [Google Scholar]
- 20.Kawasaki S, et al. Evolution of the magnetism and superconductivity in CeRhIn5 under pressure: 115In-NQR study. J Phys Soc Jpn. 2002;71:282–284. [Google Scholar]
- 21.Nakajima Y, et al. Non-Fermi liquid behavior in the magnetotransport of CeMIn5 (M: Co and Rh): Striking similarity between quasi two-dimensional heavy fermion and high-Tc cuprates. J Phys Soc Jpn. 2007;76:024703. [Google Scholar]
- 22.Bao W, et al. Anisotropic three-dimensional magnetic fluctuations in heavy fermion CeRhIn5. Phys Rev B Condens Matter Mater Phys. 2002;65:100505. [Google Scholar]
- 23.Aso N, et al. Switching of magnetic ordering in CeRhIn5 under hydrostatic pressure. J Phys Soc Jap. 2009;78:073703. [Google Scholar]
- 24.Llobet A, et al. Magnetic structure of CeRhIn5 as a function of pressure and temperature. Phys Rev B Condens Matter Mater Phys. 2004;69:024403. [Google Scholar]
- 25.Christianson AD, et al. Neutron scattering study of crystal fields in CeRhIn5. Phys Rev B Condens Matter Mater Phys. 2002;66:193102. [Google Scholar]
- 26.Monthoux P, Pines D, Lonzarich GG. Superconductivity without phonons. Nature. 2007;450:1177–1183. doi: 10.1038/nature06480. [DOI] [PubMed] [Google Scholar]
- 27.Uemura YJ, et al. Universal correlations between Tc and ns/m∗ (carrier density over effective mass) in high-Tc cuprate superconductors. Phys Rev Lett. 1989;62:2317–2320. doi: 10.1103/PhysRevLett.62.2317. [DOI] [PubMed] [Google Scholar]
- 28.Moriya T. Nuclear magnetic relaxation in antiferromagnetics. Prog Theor Phys. 1956;16:23–44. [Google Scholar]
- 29.Curro NJ, et al. Low-frequency spin dynamics in the CeMIn5 materials. Phys Rev Lett. 2003;90:227202. doi: 10.1103/PhysRevLett.90.227202. [DOI] [PubMed] [Google Scholar]
- 30.Tateiwa N, Sato N, Komatsubara T. Heat-capacity investigation on magnetism in UNi2Al3. Phys Rev B Condens Matter Mater Phys. 1998;58:11131–11133. [Google Scholar]
- 31.Mydosh JA, Oppeneer PM. Hidden order, superconductivity, and magnetism: The unsolved case of URu2Si2. Rev Mod Phys. 2011;83:1301–1322. [Google Scholar]
- 32.Havinga EE, Buschow KHJ, van Daal HJ. The ambivalence of Yb in YbAl2 and YbAl3. Solid State Commun. 1973;13:621–627. [Google Scholar]
- 33.De Visser A, et al. Specific heat, susceptibility and high-field magnetisation experiments on heavy fermion UPt3 alloyed with Pd. Phys Lett. 1986;113:489–494. [Google Scholar]
- 34.Geibel C, et al. A new heavy-fermion superconductor: UNi2Al3. Z Phys B Condensed Matter. 1991;83:305–306. [Google Scholar]
- 35.Petrovic C, et al. Heavy-fermion superconductivity in CeCoIn5 at 2.3 K. J Phys Condens Matter. 2001;13:L337–L342. [Google Scholar]
- 36.Sidorov VA, et al. Superconductivity and quantum criticality in CeCoIn5. Phys Rev Lett. 2002;89:157004. doi: 10.1103/PhysRevLett.89.157004. [DOI] [PubMed] [Google Scholar]
- 37.Bonn DA, Garrett JD, Timusk T. Far-infrared properties of URu2Si2. Phys Rev Lett. 1988;61:1305–1308. doi: 10.1103/PhysRevLett.61.1305. [DOI] [PubMed] [Google Scholar]
- 38.Aynajian P, et al. Visualizing the formation of the Kondo lattice and the hidden order in URu2Si2. Proc Natl Acad Sci USA. 2010;107:10383–10388. doi: 10.1073/pnas.1005892107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Santander-Syro AF, et al. Fermi-surface instability at the “hidden-order” transition of URu2Si2. Nat Phys. 2009;5:637–641. [Google Scholar]
- 40.Dürkop D, et al. Antiferromagnetic order and 4f-instability of CePb3. Z Phys B Condens Matter. 1986;63:55–61. [Google Scholar]
- 41.Grauel A, et al. Tetravalency and magnetic phase diagram in the heavy-fermion superconductor UPd2Al3. Phys Rev B Condens Matter Mater Phys. 1992;46:5818–5821. doi: 10.1103/physrevb.46.5818. [DOI] [PubMed] [Google Scholar]
- 42.Besnus MJ, Braghta A, Meyer A. Kondo behavior in magnetic (Ce-La)Pd2Si2. Z Phys B Condens Matter. 1991;83:207–211. [Google Scholar]
- 43.Sheikin I, Wang Y, Bouquet F, Lejay P, Junod A. Specific heat of heavy-fermion CePd2Si2 in high magnetic fields. J Phys Condens Matter. 2002;14:L543–L549. [Google Scholar]