Main

Measurements of the Nernst effect provide a unique opportunity to study the superconducting fluctuations deep inside the normal state above Tc (refs 2, 3, 8, 9, 10, 11, 12). The Nernst signal N is the electric field Ey (y) response to a transverse temperature gradient xT (x) in the presence of a magnetic field H (z), and is given by NEy/(−xT). The Nernst coefficient, defined as νN/μ0H above Tc, consists of two contributions generated by different mechanisms: ν = νS + νN. The first term, νS, represents the contribution of superconducting fluctuations of either amplitude or phase of the order parameter, which is always positive13. The second term, νN, represents the contribution from the normal quasiparticles, which can be either positive or negative14. The second contribution is usually small in conventional metals. In almost all superconductors the superconducting fluctuation contribution to the Nernst effect can be accounted for by the Gaussian-type fluctuations13,15,16,17,18. Recently, a large Nernst signal has been reported in the pseudogap state of the hole-doped underdoped high-Tc cuprates, which has been discussed in terms of possible vortex-like excitations of phase disordered superconductors2,3,8. Although its origin is still controversial, these results imply that the fluctuation-induced Nernst signal above Tc is intimately related to the exotic superconducting state below Tc.

The heavy-fermion compound URu2Si2 exhibits unconventional superconductivity (Tc = 1.45 K). This compound is distinguished from the other heavy-fermion compounds by the fact that the mysterious hidden-order transition takes place at THO = 17.5 K, and no evidence of magnetic order has been found below THO (ref. 19). This system has been suggested to be a candidate as a chiral d-wave superconductor that spontaneously breaks time-reversal symmetry (TRS) in the superconducting state4,5,6,7. Indeed, the angular dependence of the thermal conductivity and specific heat in magnetic fields indicate the presence of point nodes in the order parameter, and a chiral d-wave pairing symmetry in a complex form of kz(kx ± iky) has been proposed4,5. Very recently, the broken TRS has also been reported as a result of polar Kerr effect measurements (A. Kapitulnik, private communications). On the basis of these results, possible Weyl-type topological superconducting states have been discussed20. It is therefore highly intriguing to examine the superconducting fluctuations in URu2Si2.

Figure 1a shows ν(T) in the zero-field limit (see also Supplementary Fig. 2a) and Fig. 1b the in-plane resistivity ρxx of ultraclean URu2Si2 single crystals (Tc = 1.45 K) with residual resistivity ratios (RRR) of 1,080 (#1) and 620 (#2). It has been reported that for RRR 30, both THO and Tc are almost independent of the RRR value21. Above THO, ν(T) is negligibly small and exhibits a marked increase on entering the hidden-order state. Below T 5 K, ν(T) shows a further enhancement and increases divergently on approaching Tc (Fig. 1a, b). The inset of Fig. 1b shows the T-dependence of the ratios of Nernst coefficient and conductivity σ = 1/ρxx in the two crystals, rνν(#1)/ν(#2) and rσρxx(#2)/ρxx(#1), respectively. In contrast to the nearly T-independent rσ, rν increases steeply below T, suggesting the appearance of an additional mechanism that generates the enhanced Nernst effect in the cleaner crystal. These results indicate that the superconducting fluctuation effect sets in below T. As discussed later, this is supported by the H-dependence of the Nernst effect. In magnetic fields, ν(T) vanishes just below the vortex lattice melting temperature, Tmelt (Fig. 2a)6.

Figure 1: Transverse thermoelectric response in URu2Si2.
figure 1

a, T-dependence of the Nernst coefficient ν in the zero-field limit (H c) for single crystals #1 (RRR = 1,080) and #2 (RRR = 620). The RRR values are determined from ρ(300 K)/ρ0 by assuming the T-dependence of the in-plane resistivity ρxx as ρxx(T) = ρ0 + ATn, with n = 1.5 and 1.7 for #1 and #2, respectively, below 6 K. In both crystals, Tc defined by the point of zero resistivity is 1.45 K. The upper inset illustrates the crystal structure of URu2Si2 and the lower inset is a schematic of the measurement set-up. b, Low-temperature data of ν(T) and ρxx(T) for crystals #1 and #2. Below T, ν rises sharply above the T-linear dependence extrapolated from higher temperatures (dashed lines). The inset shows the T-dependence of the ratios of the Nernst coefficient and conductivity of the two crystals, rν = ν(#1)/ν(#2) and rσ = ρxx(#2)/ρxx(#1).

Figure 2: Anomalously large Nernst signal and thermomagnetic figure of merit.
figure 2

a, The T-dependence of ν (left scale) and ρxx (right scale) measured at μ0H = 1 T near the superconducting transition. Both ν and ρxx vanish at the vortex lattice melting transition temperature Tmelt. b, Comparison of the ν(T) data at μ0H = 1 T between samples with different scattering rates (RRR = 1,080, 620 and 30). The data for RRR 30 (expanded in the inset) is taken from ref. 23. c, Thermomagnetic figure of merit ZTε = N2σT/κ at 1.5 K as a function of field in crystal #1 of URu2Si2 (red diamonds), compared with previous data in the semimetals PrFe4P12 (blue line) and Bi (black line) at 1.2 K taken from ref. 24.

First, we discuss the Nernst signal in the range of T that is free from the superconducting fluctuations (T T THO). As shown in Fig. 2b, increasing RRR or the scattering time τ leads to an enhancement of νN. Within the Boltzmann theory, when τ is weakly dependent on energy, νN can be expressed as νN = (π2/3)(kB2T/m)(τ/ɛF) (refs 10, 22), where kB is the Boltzmann constant, m is the effective mass and ɛF is the Fermi energy. The striking enhancement of ν below THO is attributed to the strong reduction of ɛF associated with the disappearance of carriers and concomitant enhancement of τ, both of which have been reported previously4. The fact that rν above T coincides well with rσ (inset of Fig. 1b) provides quantitative support of νN τ.

At lower temperatures below T, ν of clean crystals becomes huge, especially in the vicinity of Tc. Indeed, ν of the cleanest crystal #1 is comparable to that of pure semimetal Bi, with the largest Nernst coefficient reported so far22. Moreover, the combination of the large Nernst signal and high conductivity in this system leads to a sizeable thermomagnetic figure of merit ZTε = N2σT/κ (κ is the thermal conductivity), which exceeds by far the values of previously studied materials (Fig. 2c)23,24.

Now we discuss the fluctuation-induced Nernst coefficient νS. The enhancement of rν(T) below T (inset of Fig. 1b) and no discernible enhancement of ν(T) near Tc for RRR 30 (inset of Fig. 2b) indicate that νS is greatly enhanced with τ. We stress that this τ-dependence of νS is opposite to that expected in the conventional Gaussian fluctuation theories, which predict νS ρxx 1/τ (refs 13, 16, 17, 18). It has also been reported that in underdoped cuprates the introduction of impurities by irradiation enhances νS (ref. 8), which is again opposite to the URu2Si2 case. It is intriguing to compare the present results with CeCoIn5, which shares several common features with URu2Si2, such as heavy-fermion unconventional superconductivity with a nodal gap and similar Tc and upper critical fields. It should be stressed that in very pure CeCoIn5, with ρ(Tc+) ≈ 4 μΩ cm, which is of the same order as that of our URu2Si2 crystals, no discernible νS is observed25,26; νS is at least two orders of magnitude smaller in pure CeCOIn5 than in URu2Si2. These results thus highlight an essential difference in the superconducting fluctuations between URu2Si2 and the other unconventional superconductors.

The unusual nature of the superconducting fluctuations in URu2Si2 is further revealed by the off-diagonal component of the thermoelectric tensor (Peltier coefficient) αxy, which is a more fundamental quantity associated with the fluctuations than the Nernst coefficient. The relation between ν and other coefficients is given as ν = 1/μ0H(αxyρxxS tanθH), where μ0 is the vacuum permeability, S is the Seebeck coefficient and θH is the Hall angle. Throughout the T- and H-regions in the present study, αxyρxx S tanθH—that is, ναxyρxx/μ0H (Supplementary Figs 1 and 2b). Figure 3 shows the T-dependence of the fluctuation-induced Peltier coefficient αxyS divided by μ0H in the zero-field limit. To determine αxyS, the deviation from T-linear behaviour in ν(T) (dashed lines in Fig. 1b) is attributed to νS. The Peltier coefficient that results from the Gaussian (Aslamazov–Larkin) fluctuations is given by

where are the fluctuation coherence lengths parallel to the ab plane and c axis, and is the magnetic length, with e being the electronic charge and Planck’s constant. The blue line in Fig. 3 shows the T-dependence of αxyAL/μ0H, calculated using ξab(0) = 10 nm and ξc(0) = 3.3 nm, which are determined by the initial slope of upper critical fields at Tc. What is most spectacular is that the observed αxyS/μ0H is four to six orders of magnitude greater than αxyAL/μ0H given by equation (1). Moreover, the T-dependence of αxyS(T) is much steeper than αxyAL(T).

Figure 3: Comparison with the standard theory of superconducting fluctuations.
figure 3

T-dependence of the fluctuation-induced Peltier coefficient divided by the magnetic field, αxyS/μ0H, for crystals #1 and #2. The blue line represents the Peltier coefficient that results from Gaussian-type (Aslamazov–Larkin) fluctuations, given by equation (1).

We stress that the observed αxyS(T) far exceeding αxyAL(T) indeed originates from the superconducting fluctuations. This is evidenced by the steep enhancement of both ν(T) and rν(T) below T. The H-dependence of αxy(H) provides quantitative support for this. It has been pointed out that the size of superconducting fluctuation is set by the coherence length at low fields, whereas it is set by the magnetic length at high fields. As a result, αxy(H) peaks at a characteristic field H, where ξab(T) = H(H), so that μ0H = Φ0/(2πξab2(0))ln(T/Tc), where Φ0 is the flux quantum. The peak field H is called the ‘ghost critical field’, and has been reported both in conventional and unconventional superconductors11,12. As shown in Fig. 4, at all temperatures of interest, αxy(H) exhibits a peak. The inset of Fig. 4 shows the peak field plotted as a function of ln(T/Tc). The solid line, which represents H(T) calculated by using ξab(0) = 10 nm, is quantitatively consistent with the peak field.

Figure 4: Superconducting fluctuations and the ghost critical field.
figure 4

H-dependence of the Peltier coefficient αxy = νμ0H/ρxx of crystal #2 at various temperatures. Arrows mark the peak fields. Inset: peak field plotted as a function of ln(T/Tc). Error bars are due to the uncertainties in determining the peak positions. The solid line represents μ0H = Φ0/(2πξab2(0))ln(T/Tc), with ξab(0) = 10 nm, which is the so-called ghost critical field.

We discuss several possible origins for the observed colossal Nernst signal. First, equation (1) assumes the diffusive limit, kBT /τ, whereas the present URu2Si2 seems to be in the ballistic regime, kBT /τ. Although an extension of equation (1) to the clean case (see Supplementary Information) reveals the enhancement of αxyS with τ, this enhancement is slower than the reduction of ρxx, so that νS αxySρxx is still suppressed for larger τ, which is inconsistent with Fig. 2b. Second, in the multiband system, each band with different effective coherence lengths contributes differently to the total αxyS. To explain the observed αxyS, however, small bands with extremely large effective coherence lengths, whose effective Hc2 corresponds to less than 1 mOe, are required. The multiband effect is, therefore, highly unlikely to explain the observed αxyS. Third, the characteristic temperature scale of phase fluctuations is given as TΘ = A(2a/4μ0kBe2λab2(0)), where , λab is the in-plane penetration depth and A is a dimensionless number of the order of unity27. Using A = 2 and λab = 0.8 μm (ref. 6), we obtain TΘ/Tc 100, suggesting that the phase fluctuations are not important.

The colossal thermomagnetic response in URu2Si2 seems to point to a new type of superconducting fluctuation generated by a degree of freedom which has not been hitherto taken into account. It is tempting to consider that such a degree of freedom is intimately related to the superconducting state with broken TRS, because one of the most essential differences between URu2Si2 and CeCoIn5, with no discernible νS, is that TRS is broken in the former whereas it is not broken in the latter28. In fact, a chirality-induced anomalous Nernst signal has been pointed out in ref. 29. Very recently, it has been shown that the chirality or Berry-phase associated with the superconducting state with broken TRS gives rise to a new type of fluctuation30, showing that αxyS(T) is strikingly enhanced with τ and its T-dependence is different from that predicted by the Gaussian fluctuation theories (Supplementary Information). As shown in Supplementary Fig. 4, the experimental results are reproduced well within the framework of the model of Berry-phase fluctuations. The present results suggest that superconducting fluctuations contain a key ingredient for the topological nature of superconductors, which is a new frontier of condensed matter physics.

Methods

The ultraclean single crystals of URu2Si2 were grown by the Czochralski pulling method in a tetra-arc furnace21. The well-defined superconducting transition was confirmed by the specific heat measurements. The Nernst and Seebeck coefficients were measured by the standard d.c. method with one resistive heater, two Cernox thermometers and two lateral contacts (lower inset of Fig. 1a). For the accurate determination of these coefficients, the crystals are cut in a rectangular shape, with the long direction corresponding to the a crystalline axis. The dimensions are 1.9 × 0.56 × 0.13 mm3 (#1) and 2.0 × 0.60 × 0.15 mm3 (#2). Magnetic field is applied perpendicular to the cleavage ab-plane with a misalignment less than 1°.