Main

Although recent theories4,8,9,10 provide possible routes to an explanation of T-linear resistivity—found in both f-electron systems (for example, Y1−xUxPd3 (ref. 1), CeCu6−xAux (ref. 2), YbRh2Si2 (ref. 3) and CeCu2Si2 (ref. 4)) and the normal state of the cuprate superconductors5—a general interpretation awaits arrival7. Several paradoxical features regarding this anomalous scattering rate continue to defy understanding, such as its persistence over decades of energy scales1,3,5 and down to millikelvin temperatures in three-dimensional materials1,2,3,4,6, its coexistence with conventional (T2) Hall-angle scattering11,12 and its inconsistency with one-parameter scaling13. Most recently, its observation over three decades of T at the field-tuned quantum critical point (QCP) of CeCoIn5 has been linked to a violation of the Wiedemann–Franz law14, an indication that this scattering rate is associated with the failure of Fermi-liquid theory in its most basic form.

Here, we present a rigorous study of the effects of rare-earth substitution on three closely related features of the exotic metal CeCoIn5: unconventional superconductivity, Kondo-lattice coherence and anomalous charge-carrier scattering. By diluting the Ce lattice within high-quality single-crystal specimens of Ce1−xRxCoIn5 with both non-magnetic (full or empty 4f-shell) and stable-4f-moment substituent ions of varying size and electronic configuration, we are able to inject both ‘Kondo holes’ (isoelectronic ions without magnetic moments) and strongly localized magnetic moments into the coherent Kondo lattice. This has allowed us to probe the spin exchange between the Ce3+ localized magnetic moments and the spins of the conduction electrons involved in Cooper pairing, Kondo screening and anomalous transport in a controlled way, revealing a surprising contrast between the response of coherent phenomena and non-Fermi-liquid behaviour to this perturbation.

Figure 1 shows the evolution of both the superconducting transition temperature Tc (identified by the transition in resistivity, ρ) and Kondo-lattice coherence temperature Tcoh (identified by the maximum in ρ(T)) for all rare-earth substitutions made in Ce1−xRxCoIn5 through the complete range of concentrations where both features exist. As shown, the salient features are the same for all variants: as a function of residual resistivity (ρ0x—see the Methods section), both Tc and Tcoh are suppressed to zero temperature at rates irrespective of the nature of the rare-earth ion, which spans both magnetic (Pr3+, Gd3+, Dy3+, Er3+) and non-magnetic (Y3+, Yb2+, Lu3+) f-electron configurations. This highlights the insensitivity of two ‘coherent’ electronic properties of CeCoIn5, heavy-fermion superconductivity and Kondo-lattice screening, to the magnetic configuration of the substituted rare-earth ions, the implications of each we will consider in turn.

Figure 1: Dependence of superconducting transition temperature, Tc, and Kondo-lattice coherence temperature, Tcoh of Ce1−xRxCoIn5 on rare-earth concentration.
figure 1

Plotted as a function of residual resistivity (ρ0x—see the Methods section), this figure highlights the absence of any effect of the electronic configuration of replacement-ion R on either Tc (filled symbols) or Tcoh (open symbols) as they evolve from x=0 (grey triangles). Note the lack of contrast between two particular species that are similar in all respects except f-electron filling: both Y3+ (5s24d1) and Gd3+ (6s24f75d1) are isovalent with Ce3+ (6s24f15d1) and have nearly identical metallic radii of 1.801 Å, slightly smaller than that of Ce (1.825 Å) and yielding a similarly small (1%) change of the lattice parameters on substitution. The absence of an f-electron shell in Y3+ leaves it non-magnetic, whereas the half-filled f-shell of Gd3+ has the simplest configuration of the rare earths: a spherically symmetric f-shell with no orbital component (J=S=7/2, L=0) produces a large effective moment μeff2=g2J(J+1)=(7.9 μB)2 with minimal effects from crystalline electric field anisotropy and spin–orbit coupling. The trends in Tc and Tcoh are also consistent with those found for the Ce1−xLaxCoIn5 series15.

The pair-breaking effect in unconventional superconductors arises via both potential (non-magnetic) and spin-flip scattering mechanisms. Potential scattering was shown via La substitution in CeCoIn5 to follow the Abrikosov–Gor’kov (AG) model for an anisotropic order parameter15, where it is well known that superconductivity is destroyed once the mean free path, lmfp, approaches the superconducting coherence length, ξ. Here, we estimate this critical scattering length to be lcr180 Å at the point where Tc→0 (that is, at ρcr20 μΩ cm, Fig. 1), assuming that the proportionality between lmfp(x=0)1,200 Å (ref. 16) and ρ(x=0) near Tc is independent of doping. This value is roughly twice the in-plane coherence length ξa=80 Å (ref. 6) and consistent with previous work15. Interestingly, the value ρcr20 μΩ cm coincides with that found in the series CeCoIn5−xSnx (ref. 17), where Sn substitution for In preferentially occurs in the Ce–In layers18. In the absence of any dependence on replacement-ion size, as demonstrated by the contrast in metallic radii of Lu (1.735 Å) and Y (1.801 Å), pair-breaking in CeCoIn5 thus seems to be dominated by general disorder in the CeIn3 planes.

The spin-flip interaction imposed on Cooper pairs by magnetic impurities is characterized by a further pair-breaking term , which includes the exchange interaction parameter and the de Gennes factor DJ=(g−1)2J(J+1), with the latter reflecting the classic competition between superconductivity and magnetism19. The absence of a dependence of ΔTc on this term in Ce1−xRxCoIn5 is intriguing, but not unprecedented. In UPt3, the insensitivity of ΔTc to DJ is attributable to an odd-parity pairing state, where an equal Zeeman shift on parallel spin states renders the spin-flip process ineffective20. In the spin-singlet cuprates, Tc is insensitive to the flavour of the rare-earth ion, R, placed in RBa2Cu3O6−δ (ref. 21) owing to the large physical separation between the R ions and the CuO2 layers, and hence owing to negligible magnetic interaction. In CeCoIn5, evidence for even-parity pairing22 also suggests a small value of , given the drastic range of DJ values (from 0.80 for R=Pr to 15.75 for R=Gd, largest in the rare-earth series). However, in contrast to the case of the cuprates, the placement of R ions directly into the active pairing layer18 of CeCoIn5 provides the first example of Tc suppression in a spin-singlet superconductor that is truly independent of DJ. Assuming the AG model applies, this places stringent bounds on both the strength of the exchange interaction involved in pair-breaking and the nature of the pairing mechanism itself.

Interestingly, this insensitivity to DJ is mimicked in the suppression of Tcoh with rare-earth substitution, as shown in Fig. 1. The temperature Tcoh is a characteristic property of the Kondo lattice; associated with the single-ion Kondo temperature, TK (ref. 23), and hybridization gap24, it signifies the onset of Kondo singlet formation and marks the scale where single-site magnetic scatterers begin to dissolve into a coherent state. Interestingly, in the same way that superconductivity is destroyed when lmfplcrξ, Tcoh also disappears when lmfp approaches a characteristic coherence length ξcohvF/kBTcoh100 Å (using Tcoh=50 K and vF6.5×10−4 m s−1, where kB and vF are Boltzmann’s constant and the Fermi velocity, respectively)9, again with no dependence on the magnetism of the dopant ion R. Furthermore, note that Tcoh→0 near the 40% percolation limit for a two-dimensional lattice. Together these support the notion that, regardless of its internal structure, the Ce lattice vacancy, or ‘Kondo hole’, is the dominant contributor to coherence destruction, leading to a universal dilution of the Kondo lattice as expected by the periodic Anderson model25. Thus, both the superconducting electron pair-breaking effect and the suppression of coherent Kondo screening proceed in a manner that is insensitive to the magnetic configuration of the dopant atom, advancing a scenario where spin-independent disorder is the dominant perturbation in both phenomena.

In contrast, the evolution of the non-Fermi-liquid electronic transport in Ce1−xRxCoIn5 shows a striking sensitivity to the dopant atom’s f-moment configuration, with T-linear resistivity persisting only in the presence of strong local-moment exchange. This is introduced in Fig. 2 through a direct comparison of the evolution of ρ(T) as a function of both non-magnetic (Y3+) and magnetic (Gd3+) Ce-site substitution in Ce1−xRxCoIn5: an increasing Y concentration introduces strong downward curvature in ρ(T) below Tcoh (Fig. 2a), whereas T-linear scattering seems to be robust against magnetic Gd substitution (Fig. 2b). We further explore this duality by presenting resistivity data for several characteristic rare-earth substitutions in Fig. 3, fitting ρ(T) for each between Tc and 20 K with a simple power law (ρ=ρ0+A Tn) and plotting Δρ=ρρ0 versus T to emphasize the exponent n, which appears as the slope on a log–log scale. As shown explicitly in the inset of Fig. 3, n spans a range of sublinear values, with deviations from T-linear being strongest for non-magnetic substitutions.

Figure 2: Comparison of electrical resistivity evolution of Ce1−xRxCoIn5 with both magnetic and non-magnetic rare-earth substitution.
figure 2

a,b, Resistivity, ρ, plotted for Ce1−xYxCoIn5 (a) and Ce1−xGdxCoIn5 (b) as a function of nominal concentration of rare-earth substitution. Although both the superconducting transition and Kondo coherence temperature (maximum in ρ(T)) are suppressed at the same rate for both substitution series, the temperature dependence of ρ is strongly dependent on the magnetic nature of the substituent ion: Y-doping imposes a strong downward curvature on ρ(T) with increasing concentration, whereas Gd-doping elicits a negligible change in the T-linear resistivity present in pure CeCoIn5.

Figure 3: Effect of chemical substitution on T-linear resistivity power laws in CeCoIn5.
figure 3

The filled symbols represent various rare-earth substitutions in Ce1−xRxCoIn5; the open circles (shifted by ×2 for clarity) represent a single-crystal sample of CeRh0.15Co0.85In5 in its field-induced normal state at 9 T, showing the close connection between sublinear curvature in ρ(T) and the proximity of a spin-density wave instability. The dashed lines are guides showing slopes for various powers of temperature. Inset: Evolution of the temperature power-law exponent n (that is, in ΔρTn) with rare-earth substitution (plotted as residual resistivity ρ0—see the Methods section), highlighting the isolated behaviour of Gd substitution. Whereas the large effective moment (μeff=7.9 μB) of Gd3+ ions in Ce1−xRxCoIn5 sets it apart from its non-magnetic counterparts, the sublinear exponent observed for Er3+ substitution (red square)—with μeff=9.6 μB—rules out a simple correlation between the moment size and the sublinear power-law exponent, suggesting the importance of the spin configuration of the rare-earth ions and de Gennes factor scaling. The error bars reflect estimates of uncertainty in n on the basis of the temperature range and number of data points used in nonlinear least-squares fits of ρ(T). (The open symbols denote concentrations used in the main figure.)

A sub-T-linear transport scattering rate is highly anomalous, yet not unprecedented. For instance, the resistivity of the strongly correlated f-electron system Sc1−xUxPd3 was indeed observed to follow the form ρ(T)=ρ0A Tn with an exponent n0.5 (ref. 26), consistent with the n=1/2 expectation of the theoretical multichannel Kondo model for TTK (ref. 9). However, the n<1 curvature in Sc1−xUxPd3 is more likely due to quantum criticality associated with the suppression of spin-glass freezing to T=0 near xc0.3, rather than the multichannel Kondo effect26.

Likewise, the phenomenological trend of n<1 curvature in Ce1−xRxCoIn5 also hints at the proximity of a magnetic instability not unlike that found in CeRhIn5, where similar sublinear curvature is present in ρ(T) above the antiferromagnetic transition at TN=3.8 K (ref. 27). In CeRhIn5, this curvature is proportional to the magnetic entropy, a reflection of the fact that magnetic correlations dominate the transport scattering process27. In CeCoIn5 the same phenomenon was found to be dependent on the proximity to a field-tuned QCP28. A connection between the two was established via resistivity measurements of the alloy series CeRh1−yCoyIn5, where a crossover to sublinear behaviour in ρ(T) was shown to be intimately related to the antiferromagnetic QCP29. As shown in Fig. 3, ρ(T) of a single-crystal sample of CeRh1−yCoyIn5 with y=0.85 (close to the alloy-tuned QCP) indeed follows an n0.5 exponent over almost two decades in T in its field-induced normal state, indicating a strong connection between n<1 scattering and the proximity of a QCP related to the spin-density wave instability in CeRhIn5.

In stark contrast, Gd substitution in Ce1−xRxCoIn5 fails to disrupt the mechanism of T-linear scattering: as shown in the inset of Fig. 3, the exponent n experiences an almost negligible change, decreasing at a rate at least five times slower than for non-magnetic substitutions. Because the zero-field magnetic entropy in CeCoIn5 also grows linearly with temperature above Tc (ref. 6), it is suspected that, like CeRhIn5, magnetic correlations are what shape this anomalous scattering rate. In Ce1−xGdxCoIn5, this must involve a Ruderman–Kittel–Kasuya–Yosida (RKKY)-type exchange, as demonstrated by both a linear increase with x of the effective moment (up to μeff=7.0 μB at x=1) and long-range antiferromagnetic order (TN32 K at x=1), which is in line with the proportionality between TN and DJ found in other magnetic RCoIn5 compounds30.

But what is the underlying property of Gd3+ magnetism that is amenable to T-linear scattering? As shown in Fig. 3, the curvature in ρ(T) of a sample doped with 25% Er3+—with an even larger moment (μeff=9.6 μB) than Gd3+—surprisingly exhibits a sublinear power law (n0.6) much closer to that of the non-magnetic samples. Furthermore, samples doped with Dy3+ (μeff=10.6 μB) exhibit intermediate behaviour, suggesting that the important parameter is not simply moment size itself, but rather involves details of the f-moment configuration. In particular, the wide range spanned by the de Gennes factors of Gd3+, Dy3+ and Er3+ (with DJ values of 15.75, 7.08 and 2.55, respectively) is the only aspect of the magnetic configuration that follows the evolution of n(x) suggested by our data set, with a phenomenological form n≈1+α(DJD0)ρ0 where D018 and α is a positive constant. Despite the peculiar position of DJ in the exponent (rather than as a coefficient) of T, its presence highlights the important role of the spin degrees of freedom in the scattering process that gives rise to T-linear resistivity, promoting the notion that the ‘control parameter’ may indeed be the projected spin of the scattering centres.

What remains highly anomalous, and more generic, is that the relatively strong relation between n and DJ must comply with the extremely weak exchange coupling between localized 4f-states and conduction-band states, as demonstrated by the insensitivity of both ΔTc and ΔTcoh to the magnetic configuration of R. This contrast provides evidence for a separation between the physics of the Kondo lattice and that of the non-Fermi-liquid transport in CeCoIn5, with the latter necessarily arising from ‘incoherent’ scattering processes. But how can this interaction coexist with the seemingly different long-range interactions that mediate superconductivity and resonant Kondo-lattice screening? One possibility is that the hybridization between f-states and conduction-electron states is incomplete, leaving a fraction of incoherent scatterers that conspire to cause such a dichotomy. Evidence for such two-fluid behaviour does indeed take form in CeCoIn5, where an ‘incoherent’ fraction of Kondo moments was shown to survive down to Tc (ref. 31). Another scenario is of a more profound nature: recent evidence for (1) a group of conduction electrons that remains unpaired in the T→0 limit32 and (2) a direction-dependent violation of the Wiedemann–Franz law14 points to a decoupled character of conduction electrons in CeCoIn5, suggesting that the separation between the mechanisms behind the coherent properties of CeCoIn5 and its T-linear resistivity is of a very fundamental nature.

Methods

Single-crystal platelets of Ce1−xRxCoIn5 (including R=Y, Pr, Gd, Dy, Er, Yb and Lu) were grown by the self-flux method6. Samples for measurements of electrical resistivity were prepared with typical dimensions 2×0.5×0.2 mm and measured with an a.c. resistance bridge by applying 0.1 mA excitation current, directed parallel to the basal plane of the tetragonal crystal structure. The data in Figs 1 and 2 are plotted as a function of residual resistivity to eliminate the uncertainty in nominal concentration values. However, note that ρ0x to within error as found previously15,32. The d.c. magnetization was measured using a superconducting quantum interference device magnetometer in a 50 mT field, and analysed using standard Curie–Weiss fits to data between approximately 25 and 300 K to extract effective moments for the magnetic Ce1−xRxCoIn5 series.