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

Present address: ]Institute for Materials Research, Tohoku University, Sendai 980‑8577, Japan

Enhanced Seebeck coefficient through the magnetic fluctuations in Sr2RuMx1x{}_{1-x}M_{x}O4 (M=M= Co, Mn)

Takayoshi Yamanaka [email protected] [    Ryuji Okazaki [email protected]    Hiroshi Yaguchi Department of Physics, Faculty of Science and Technology, Tokyo University of Science, Noda, Chiba 278-8510, Japan
Abstract

The layered perovskite Sr2RuO4 is a most intensively studied superconductor, but its pairing mechanism, which is often coupled intimately with magnetic fluctuations in correlated materials, is still an open question. Here we present a systematic evolution of the Seebeck coefficient in Co- and Mn-substituted Sr2RuO4 single crystals, in which ferromagnetic and antiferromagnetic glassy states respectively emerge in proximity to the superconducting phase of the parent compound. We find that the Seebeck coefficient SS divided by temperature TT, S/TS/T, shows a maximum near characteristic temperatures seen in the irreversible magnetization MirM_{\rm ir} in both of the Co- and Mn-substituted crystals, demonstrating both of the ferromagnetic and antiferromagnetic fluctuations to enhance the Seebeck coefficient. Interestingly, S/TS/T increases with lowering temperature in the parent compound, reminiscent of non-Fermi-liquid behavior, indicating an essential role of coexisting ferromagnetic and antiferromagnetic fluctuations for the itinerant electrons in Sr2RuO4.

I I. Introduction

Unconventional superconductivity in the layered perovskite Sr2RuO4 has long attracted interest [1, 2, 3, 4, 5, 6]. In particular, recent progress of nuclear magnetic resonance (NMR) experiments has posed strong constraints on the spin sector of the superconducting order parameter [7, 8, 9], offering a clue for the consistent understanding of several experimental results [10, 11, 12] that were difficult to reconcile with earlier NMR results [13, 14, 15, 16], as well as stimulating further symmetry-based experiments to elucidate the order parameter [17, 18, 19]. As a result, an exotic two-component order parameter with broken time reversal symmetry has been proposed, which is unique as suggested in several examples [20, 21], and various theoretical attempts have also been made [22, 23, 24, 25, 26, 27], opening an avenue for exploring unconventional pairing interaction to realize such order parameters.

To address this underlying issue, it is crucially important to establish the electronic phase diagram as a function of external parameters such as pressure and chemical substitutions, as is widely discussed in correlated matters [28, 29]. Indeed, although the superconducting state in Sr2RuO4 is extremely sensitive to impurity [30], dramatic changes in the electronic state with elemental substitutions have been reported. In the isovalent systems, for instance, a spin-glass state develops over a wide range of the Ca content in (Sr, Ca)2RuO4 [31, 32], which may originate from the degree of freedom in the RuO6 octahedra. On the other hand, in Sr2(Ru, Ti)O4, an incommensurate spin-density-wave ordering due to the Fermi-surface nesting appears with glassy behavior [33]. In contrast to the antiferromagnetic (AFM) coupling seen in the isovalent systems, La3+ substitution to Sr2+ sites results in electron doping to expand the electron-like γ\gamma Fermi surface, leading to ferromagnetic (FM) fluctuation owing to an enhancement of the density of states (DOS) at the Fermi energy N(εF)N(\varepsilon_{\rm F}) near the van Hove singularity (vHs) [34, 35, 36]. These results clearly show the complicated magnetic instabilities existing in the parent compound Sr2RuO4, which involve complex structural and electronic origins. Such instabilities are also discussed in several studies including NMR and neutron experiments [37, 38, 39, 40, 41].

Among many substituted systems, Co- and Mn-substituted Sr2RuO4 serve as a fascinating platform to investigate how magnetic fluctuations mediate the emergence of superconductivity, because slight substitutions of Co and Mn drastically vary the superconducting ground state into the FM and AFM glassy states, respectively [42]. In the Co-substituted system, the FM cluster glass characterized by an exponential relaxation of the remanent magnetization appears at low temperatures. Also, the electronic specific heat γe\gamma_{\rm e} increases with increasing Co contents, indicating an increased N(εF)N(\varepsilon_{\rm F}) similar to La-substitution [34]. It is noteworthy that the increase of N(εF)N(\varepsilon_{\rm F}) in a 1.5%1.5\% Co-substituted sample estimated from γe\gamma_{\rm e} is comparable to that in a 5%5\% La-substituted sample, possibly implying an effective electron doping effect by the Co substitution [42]. Similarly, the AFM transition temperature in Mn-substituted samples is much higher than that in Ti-substituted ones, demonstrating enigmatic roles of the Co and Mn substitutions to strengthen the magnetic couplings in Sr2RuO4.

The aim of this study is to examine such magnetic fluctuations in Sr2RuMx1x{}_{1-x}M_{x}O4 (M=M= Co, Mn) by means of Seebeck coefficient measurement, known as a powerful tool for fluctuations as it is a measure of the entropy per charge carrier [43]. The observed temperature dependence of S/TS/T (SS and TT being the Seebeck coefficient and temperature, respectively) has a maximum near characteristic temperatures in the magnetization in both Co- and Mn-substituted samples, indicating that both FM and AFM fluctuations are responsible for the enhancement of the Seebeck coefficient. Moreover, in sharp contrast to the typical metallic behavior in which S/TS/T remains constant, S/TS/T increases with cooling in the parent compound, implying both FM and AFM fluctuations remaining in the itinerant electrons in Sr2RuO4.

II II. Experimental details

Single crystals of Sr2RuMx1x{}_{1-x}M_{x}O4 (M=M= Co, Mn) were grown by a floating-zone method using an image furnace with a pair of halogen lamps and elliptical mirrors [44, 42]. We used high-purity SrCO3(99.99%+), RuO2(99.9%), MnO2(99.99%), and CoO(99.9%) as starting materials. An excess 15% amount of RuO2 was weighed to compensate for evaporation of Ru during single crystal growth. The concentration of Mn and Co in obtained crystals were determined by electron probe micro analyzer. The measured Mn concentration was 2.9%, well corresponding to the nominal concentration of 3%. In contrast, the measured Co concentrations were 1.8% and 0.8% for nominally 3% and 1% Co-substituted samples, respectively. The relation between the nominal and measured concentrations is consistent with the trend reported in Ref. 42.

Refer to caption
Figure 1: Temperature dependence of the Seebeck coefficient SS of Sr2RuMx1x{}_{1-x}M_{x}O4 (M=M= Co, Mn) single crystals.

The typical dimension of the measured crystals was 3×0.2×0.2\approx 3\times 0.2\times 0.2 mm3. The Seebeck coefficient was measured by a steady-state technique using a manganin-constantan differential thermocouple in a closed-cycle refrigerator. The thermoelectric voltage of the sample was measured with Keithley 2182A nanovoltmeter. The temperature gradient with a typical temperature gradient of 0.5 K/mm was applied along the abab-plane direction using a resistive heater. The thermoelectric voltage from the wire leads was subtracted. Magnetization was measured using a superconducting quantum interference device magnetometer (Quantum Design, MPMS) under field-cooled (FC) and zero-field-cooled (ZFC) processes with the external field of μ0H=0.1\mu_{0}H=0.1 T applied along the cc axis.

III III. results and discussions

Figure 1 summarizes the temperature variations of the Seebeck coefficient SS in Sr2RuMx1x{}_{1-x}M_{x}O4 (M=M= Co, Mn) single crystals. Overall behavior of SS in the parent compound Sr2RuO4 is consistent with earlier results [45, 46], and is discussed as an intriguing example to study the internal degrees of freedom in correlated metals [47]. In the substituted compounds, SS increases (decreases) with Co (Mn) substitutions near room temperature probably because of the electron (hole) doping effect as suggested in Ref. 42.

Refer to caption
Figure 2: Comparison of the temperature dependence of (a) dS/dTdS/dT and (b) S/TS/T of Sr2RuO4 single crystals. The red circles show the present data, and the others are data extracted from Refs. 45, 46.

In Refs. 45, 46, the Seebeck coefficient in the parent compound was analyzed in the differential form dS/dTdS/dT, and an anomaly was found near 202520-25 K. Xu et al. have suggested that a band-dependent coherence is developed below the anomaly temperature on the basis of the results of the Seebeck and the Nernst measurements [46]. Indeed, such a coherency seems to be vital in this system [48]. To see this anomaly at 202520-25 K, we compare the present data for Sr2RuO4 with the results extracted from Refs. 45, 46 in Fig. 2(a). Although there is a sample dependence in magnitude, all the data exhibit a kink around 202520-25 K, in good agreement with the present result. Also note that the effect from the impurity phase of SrRuO3 is negligible, because the Seebeck coefficient SS in a composite sample with the conductivity σ\sigma is given as σS=kαkσkSk\sigma S=\sum_{k}\alpha_{k}\sigma_{k}S_{k} in a parallel-circuit model, where αk,σk,Sk\alpha_{k},\sigma_{k},S_{k} are the volume fraction, the conductivity, and the Seebeck coefficient for the material kk [49], and the volume fraction of the impurity phase in the present crystal was negligibly small.

Here we discuss the Seebeck coefficient in the form of S/TS/T instead of dS/dTdS/dT, since the Seebeck coefficient in the free-electron model (oversimplified model to multiband Sr2RuO4) is expressed as SkB2TeN(μ)nS\approx-\frac{k_{\rm{B}}^{2}T}{e}\frac{N(\mu)}{n}, where kBk_{\rm{B}}, ee, nn, μ\mu, and NN are the Boltzmann constant, elementary charge, carrier density, chemical potential, and the DOS, respectively [50]. In this form, one can follow the temperature dependence of the DOS or carrier density, as is widely analyzed in correlated electron systems, similar to the case of the pseudogap state in transition-metal oxides [51, 52, 53] and heavy-fermion formation in rare-earth compounds [54]. Note that the differential dS/dTdS/dT includes the temperature derivatives of N(μ)N(\mu) and nn in complicated forms. Figure 2(b) represents the temperature dependence of S/TS/T in Sr2RuO4 in the present crystals and the calculated S/TS/T from the extracted data from Refs. 45, 46. Interestingly, all the data are temperature-dependent; in sharp contrast to conventional metals, in which S/TS/T remains constant [43], S/TS/T significantly increases with decreasing temperature, the origin of which will be discussed later.

Refer to caption
Figure 3: (a) Temperature dependence of S/TS/T of Sr2RuMx1x{}_{1-x}M_{x}O4 (M=M= Co, Mn) single crystals. The arrows show the peak temperature TpeakT_{\rm peak}. (b) Temperature dependence of the irreversible magnetization MirMFCMZFCM_{\rm ir}\equiv M_{\rm FC}-M_{\rm ZFC}. The arrows represent the characteristic temperatures TT^{\dagger}, TsT_{s}, and TwT_{w}. (See text for details.) In the inset, MFCM_{\rm FC}, MZFCM_{\rm ZFC}, (left axis) and MirM_{\rm ir} (right axis) of the 0.8% Co-substituted sample are shown together with the definitions of TsT_{\rm s} and TwT_{\rm w}.

We then focus on the temperature variation of S/TS/T in the substituted systems. Figure 3(a) represents the temperature dependence of S/TS/T in Sr2RuMx1x{}_{1-x}M_{x}O4 (M=M= Co, Mn), in which we find a drastic change due to substitutions. At low temperatures, S/TS/T exhibits a prominent peak structure in the temperature dependence for both substituted systems. The peak temperature TpeakT_{\rm peak} defined as the temperature at which S/TS/T exhibits a maximum is shown by arrows in Fig. 3(a). Note that the phonon-drag effect is unlikely to enhance the Seebeck coefficient in the substituted systems, because the substitutions generally suppress the phonon mean free path [55].

To shed light on the relation to the magnetism, we plot the irreversible magnetization MirM_{\rm ir} defined as MirMFCMZFCM_{\rm ir}\equiv M_{\rm FC}-M_{\rm ZFC}, in Fig. 3(b). (MFCM_{\rm FC} and MZFCM_{\rm ZFC} being the magnetization measured under FC and ZFC processes, respectively). While the onset of the irreversibility may correspond to a transition to the FM and AFM glassy states for the Co- and Mn-substituted systems, respectively [42], a close look of the MirM_{\rm ir} data reveals that the onset of the irreversibility is accompanied by a double step. The inset of Fig. 3(b) shows the temperature variations of MFCM_{\rm FC}, MZFCM_{\rm ZFC}, and MirM_{\rm ir} for x=0.008x=0.008 (Co) at low temperatures. The MZFCM_{\rm ZFC} data more steeply decrease than that in Ref. 42, but this may be due to the low applied field in our measurement. From the plots of MirM_{\rm ir} in the inset, one can see that a weak irreversibility sets in at Tw14T_{\rm w}\approx 14 K and subsequently the irreversibility is strongly enhanced below Ts4T_{\rm s}\approx 4 K. The weak and strong irreversibility temperatures TwT_{\rm w} and TsT_{\rm s} are defined as the points of intersection of the two linear lines as drawn in the dotted line in the inset. Importantly, such a coexistence of weak and strong irreversibility is often observed in ferromagnetic spin-glass systems [56, 57, 58, 59], and theoretically interpreted in terms of a multi component vector spin model given by Gabay and Toulouse [60], in which the transverse spin components are first frozen at TwT_{\rm w} on cooling and then followed by the freezing of the longitudinal spin components at TsT_{\rm s}, although it is still under debate whether it represents a true thermodynamic phase transition.

On the other hand, the temperature dependence of MirM_{\rm ir} in the Mn-substituted sample is different from that in the Co-substituted samples: On cooling, the irreversibility sets in at T27T\approx 27 K, which is almost identical to the onset temperature TirT_{\mathrm{ir}} to the static order [42], and then shows weak temperature dependence below T18T^{\dagger}\approx 18 K at which MirM_{\rm ir} displays a kink structure as shown in Fig. 3(b). In the Mn-substituted sample, TpeakT_{\mathrm{peak}} in S/TS/T is close to this anomaly temperature TT^{\dagger}, although the origin of TT^{\dagger} is unclear at present. Note that the remanent magnetization is also distinct among the Co- and Mn-substituted systems: While the temperature dependence of remanent magnetization is monotonic in the Co-substituted samples, it exhibits an anomalous peak structure below the onset temperature TirT_{\mathrm{ir}} for the Mn-substituted systems [42], implying the existence of a characteristic temperature below TirT_{\mathrm{ir}}. The nature of the magnetic glassy state is an issue for the future study using the microscopic NMR and neutron measurements.

Refer to caption
Figure 4: Phase diagram of Sr2RuMx1x{}_{1-x}M_{x}O4 (M=M= Co, Mn). The closed circles shows the peak temperature TpeakT_{\rm peak} at which S/TS/T exhibits maximum. The weak and strong irreversibility temperatures TwT_{\rm w} (stars) and TsT_{\rm s} (triangles) and the anomaly temperature TT^{\dagger} (square) are determined from the irreversibility magnetization MirM_{\rm ir}. The onset temperature TirT_{\rm ir} (open circles) and the regions of FM cluster glass (CG) and spin glass state with short-range incommensurate AFM order (SG-IC) are extracted from Ref. 42.

In Fig. 4, we plot the peak temperature TpeakT_{\rm peak} in S/TS/T together with TsT_{\mathrm{s}}, TwT_{\mathrm{w}}, and TT^{\dagger} on the phase diagram reported in Ref. 42. The peak temperature TpeakT_{\mathrm{peak}} in S/TS/T well coincides with the weak irreversibility temperature TwT_{\mathrm{w}} and the anomaly temperature TT^{\dagger} for the Co- and Mn-substituted samples, respectively, indicating an intimate relationship among the Seebeck coefficient and the magnetic fluctuations. Here, since the Co and Mn substitutions induce the FM cluster glass and the spin glass states with short-range AFM order, respectively [42], the present results indicate that both FM and AFM fluctuations are substantial for the enhancement of the Seebeck coefficient at TpeakT_{\rm peak}. Indeed, it bears a striking resemblance to that in the itinerant magnets such as the perovskite ruthenate CaRu0.8Sc0.2O3 [61] and the doped Heusler alloy Fe2VAl [62]; in these compounds, the Seebeck coefficient is enhanced near the ferromagnetic transition temperature through a sort of the magnon-drag effect, and is reduced by applying magnetic field owing to the field-induced suppression of the magnetic fluctuation. In the antiferro-quadrupole (AFQ) PrIr2Zn20, moreover, S/TS/T exhibits a peak structure near the AFQ transition temperature in high magnetic field where the quadrupolar fluctuation could be enhanced[63].

Now let us recall the S/TS/T behavior in Sr2RuO4 (Fig. 2), which increases with cooling down to the lowest temperature of the present measurement (3\sim 3 K). While Seebeck coefficients are not strictly linear in temperature even in simple metals [64], the non-linearity in Sr2RuO4 is unusual. As mentioned above, we show the suggestive evidence that both FM and AFM fluctuations drive the enhanced Seebeck coefficient in the vicinity of the parent compound. It is therefore reasonable to consider that, even in the parent compound Sr2RuO4, S/TS/T is enhanced similarly down to zero temperature, at which both FM and AFM glassy states tend to terminate for Sr2RuO4[42]. Thus this is a kind of non-Fermi-liquid (NFL) behavior close to the quantum critical point [65], which has also been observed in various correlated matters such as heavy fermions [66, 67, 68, 69, 70] and oxides [71]. Quite intriguingly, in contrast to the aforementioned systems, both FM and AFM fluctuations seem to be substantial for the itinerant electrons in Sr2RuO4 as seen in the pronounced peak of S/TS/T observed in both Co- and Mn-substituted compounds.

One may, however, pose a simple question regarding the specific heat. In Sr2RuO4, the electronic specific heat shows conventional Fermi-liquid (FL) behavior [72], in distinction to the NFL behavior observed in the Seebeck coefficient; since the Seebeck coefficient is also given as the specific heat per carrier [50], both quantities are expected to show similar anomalies [73]. On the other hand, an AFM quantum criticality may affect the ratio of these quantities in zero-temperature limit [74]. It is interesting to note that this effect of the AFM quantum criticality may be enhanced by the multiband nature; in Sr2RuO4, the Fermi surfaces are composed of three cylindrical sheets: hole-like α\alpha and electron-like β\beta and γ\gamma sheets with the order of magnitudes of the effective mass of mα<mβ<mγm_{\alpha}<m_{\beta}<m_{\gamma} [75, 76, 77, 78]. Since the lighter band contributes to the transport more significantly in general [20, 79], the lighter α\alpha and β\beta bands are essential here, and importantly, theseα\alpha and β\beta sheets possess the AFM instability due to nesting [38, 39, 40, 41]. This band-dependent magnetic fluctuation may give origin to the anomalous low-temperature increase in S/TS/T. On the other hand, it is unclear at present how the thermoelectric transport is affected by the FM fluctuation, the importance of which is clearly demonstrated in the Co-substituted systems; the existence of FM fluctuation is indicated by NMR [37], while the neutron experiments have revealed that it is very weak [38, 39, 40, 41].

It is also known that the electrical resistivity of Sr2RuO4 is well described within the FL scheme [72], in which the resistivity is given as ρ(T)=ρ0+ATν\rho(T)=\rho_{0}+AT^{\nu} with the exponent ν=2\nu=2. On the other hand, the determination of ν\nu is delicate [80], and as seen in other correlated materials, ν\nu apparently depends on the residual resistivity even in high-purity level [81]. It may be useful to examine the exponent in high-purity crystals [82]. We also mention that the thermal conductivity in the normal state is interesting in the sense that it also mirrors the entropy flow, while the earlier studies are mainly devoted to the superconducting state [83, 84, 85]. Also, the Seebeck coefficient in the system with the Fermi level at the vHs, which is not achieved with the present Co substitution, is worth exploring because the topological change in the Fermi surface at vHs may lead to the significant enhancement of the Seebeck coefficient [86, 87]. This effect of the vHs could be investigated in electron-doped Sr2-xLaxRuO4.

IV IV. Conclusion

To summarize, we performed Seebeck coefficient measurement in Sr2RuMx1x{}_{1-x}M_{x}O4 (M=M= Co, Mn). Although the S/TS/T of the parent compound Sr2RuO4 increases with cooling down to the lowest temperature 3\sim 3 K similar to previous reports, for Co- and Mn-substituted systems, S/TS/T is enhanced near TwT_{\rm w} and TT^{\dagger} in the irreversible magnetization. The emergence of the peak structure in S/TS/T can be related to glassy FM and AFM fluctuations in the Co- and Mn-substituted system, respectively, and therefore the increase of S/TS/T in Sr2RuO4 persisting at least down to 3\sim 3 K suggests that both FM and AFM fluctuations seem to be substantial in Sr2RuO4.

Acknowledgements.
The authors acknowledge Y. Maeno and K. Ishida for fruitful discussions. This work was supported by JSPS KAKENHI Grants No. JP17H06136 and No. JP18K13504.

References

  • [1] Y. Maeno, H. Hashimoto, K. Yoshida, S. Nishizaki, T. Fujita, J. G. Bednorz, and F. Lichtenberg, Superconductivity in a layered perovskite without copper, Nature 372, 532 (1994).
  • [2] T. M. Rice and M. Sigrist, Sr2RuO4: an electronic analogue of 3He?, J. Phys.: Condens. Matter 7, L643 (1995).
  • [3] A. P. Mackenzie and Y. Maeno, The superconductivity of Sr2RuO4 and the physics of spin-triplet pairing, Rev. Mod. Phys. 75, 657 (2003).
  • [4] C. Kallin, Chiral p-wave order in Sr2RuO4, Rep. Prog. Phys. 75, 042501 (2012).
  • [5] A. P. Mackenzie, T. Scaffidi, C. W. Hicks, and Y. Maeno, Even odder after twenty-three years: The superconducting order parameter puzzle of Sr2RuO4, npj Quant. Mater. 2, 40 (2017).
  • [6] S. A. Kivelson, A. C. Yuan, B. Ramshaw, and R. Thomale, A proposal for reconciling diverse experiments on the superconducting state in Sr2RuO4, npj Quant. Mater. 5, 43 (2020).
  • [7] A. Pustogow, Y. Luo, A. Chronister, Y.-S. Su, D. A. Sokolov, F. Jerzembeck, A. P. Mackenzie, C. W. Hicks, N. Kikugawa, S. Raghu, E. D. Bauer, and S. E. Brown Constraints on the superconducting order parameter in Sr2RuO4 from oxygen-17 nuclear magnetic resonance, Nature 574, 72 (2019).
  • [8] K. Ishida, M. Manago, K. Kinjo, and Y. Maeno, Reduction of the 17O Knight Shift in the Superconducting State and the Heat-up Effect by NMR Pulses on Sr2RuO4, J. Phys. Soc. Jpn. 89, 034712 (2020).
  • [9] A. Chronister, A. Pustogow, N. Kikugawa, D. A. Sokolov, F. Jerzembeck, C. W. Hicks, A. P. Mackenzie, E. D. Bauer, and S. E. Brown, Evidence for even parity unconventional superconductivity in Sr2RuO4, Proc. Natl. Acad. Sci. U.S.A. 118, e2025313118 (2021).
  • [10] S. Kittaka, T. Nakamura, Y. Aono, S. Yonezawa, K. Ishida, and Y. Maeno, Angular dependence of the upper critical field of Sr2RuO4, Phys. Rev. B 80, 174514 (2009).
  • [11] S. Yonezawa, T. Kajikawa, and Y. Maeno, First-Order Superconducting Transition of Sr2RuO4, Phys. Rev. Lett. 110, 077003 (2013).
  • [12] C. W. Hicks, D. O. Brodsky, E. A. Yelland, A. S. Gibbs, Jan A. N. Bruin, M. E. Barber, S. D. Edkins, K. Nishimura, S. Yonezawa, Y. Maeno, and A. P. Mackenzie, Strong Increase of TcT_{c} of Sr2RuO4 Under Both Tensile and Compressive Strain, Science 344, 283 (2014).
  • [13] K. Ishida, H. Mukuda, Y. Kitaoka, K. Asayama, Z. Q. Mao, Y. Mori, and Y. Maeno, Spin-triplet superconductivity in Sr2RuO4 identified by 17O Knight shift, Nature 396, 658 (1998).
  • [14] H. Murakawa, K. Ishida, K. Kitagawa, H. Ikeda, Z. Q. Mao, and Y. Maeno, 101Ru Knight Shift Measurement of Superconducting Sr2RuO4 under Small Magnetic Fields Parallel to the RuO2 Plane, J. Phys. Soc. Jpn. 76, 024716 (2007).
  • [15] K. Ishida, H. Murakawa, H. Mukuda, Y. Kitaoka, Z. Q. Mao, and Y. Maeno, NMR and NQR studies on superconducting Sr2RuO4, J. Phys. Soc. Jpn. 76, 024716 (2007).
  • [16] K. Ishida, M. Manago, T. Yamanaka, H. Fukazawa, Z. Q. Mao, Y. Maeno, and K. Miyake, Spin polarization enhanced by spin-triplet pairing in Sr2RuO4 probed by NMR, Phys. Rev. B 92, 100502(R) (2015).
  • [17] S. Benhabib, C. Lupien, I. Paul, L. Berges, M. Dion, M. Nardone, A. Zitouni, Z. Q. Mao, Y. Maeno, A. Georges, L. Taillefer, and C. Proust Ultrasound evidence for a two-component superconducting order parameter in Sr2RuO4, Nat. Phys. 17, 194 (2021).
  • [18] S. Ghosh, A. Shekhter, F. Jerzembeck, N. Kikugawa, D. A. Sokolov, M. Brando, A. P. Mackenzie, C, W. Hicks, and B. J. Ramshaw, Thermodynamic evidence for a two-component superconducting order parameter in Sr2RuO4, Nat. Phys. 17, 199 (2021).
  • [19] V. Grinenko, S. Ghosh, R. Sarkar, J.-C. Orain, A. Nikitin, M. Elender, D. Das, Z. Guguchia, F. Brückner, M. E. Barber, J. Park, N. Kikugawa, D. A. Sokolov, J. S. Bobowski, T. Miyoshi, Y. Maeno, A. P. Mackenzie, H. Luetkens, C. W. Hicks, and H.-H. Klauss, Split superconducting and time-reversal symmetry-breaking transitions in Sr2RuO4 under stress, Nat. Phys. 17, 748 (2021).
  • [20] Y. Kasahara, T. Iwasawa, H. Shishido, T. Shibauchi, K. Behnia, Y. Haga, T. D. Matsuda, Y. Onuki, M. Sigrist, and Y. Matsuda, Exotic Superconducting Properties in the Electron-Hole-Compensated Heavy-Fermion “Semimetal” URu2Si2, Phys. Rev. Lett. 99, 116402 (2007).
  • [21] M. Smidman, M. B. Salamon, H. Q. Yuan, and D. F. Agterberg, Superconductivity and spin–orbit coupling in non-centrosymmetric materials: A review, Rep. Prog. Phys. 80, 036501 (2017).
  • [22] O. Gingras, R. Nourafkan, A.-M. S. Tremblay, and M. Côté Superconducting Symmetries of Sr2RuO4 from First-Principles Electronic Structure, Phys. Rev. Lett. 123, 217005 (2019).
  • [23] A. T. Rømer, D. D. Scherer, I. M. Eremin, P. J. Hirschfeld, and B. M. Andersen, Knight Shift and Leading Superconducting Instability from Spin Fluctuations in Sr2RuO4, Phys. Rev. Lett. 123, 247001 (2019).
  • [24] H. G. Suh, H. Menke, P. M. R. Brydon, C. Timm, A. Ramires, and D. F. Agterberg, Stabilizing even-parity chiral superconductivity in Sr2RuO4, Phys. Rev. Research 2, 032023(R) (2020).
  • [25] R. Willa, M. Hecker, R. M. Fernandes, and J. Schmalian, Inhomogeneous time-reversal symmetry breaking in Sr2RuO4, Phys. Rev. B 104, 024511 (2021).
  • [26] A. T. Rømer, P. J. Hirschfeld, and B. M. Andersen, Superconducting state of Sr2RuO4 in the presence of longer-range Coulomb interactions, Phys. Rev. B 104, 064507 (2021).
  • [27] S.-J. Zhang, D. Wang, and Q.-H. Wang, Possible two-component spin-singlet pairings in Sr2RuO4, Phys. Rev. B 104, 094504 (2021).
  • [28] T. Das and C. Panagopoulos, Two types of superconducting domes in unconventional superconductors, New J. Phys. 18, 103033 (2016).
  • [29] S. Paschen and Q. Si, Quantum phases driven by strong correlations., Nat. Rev. Phys. 3, 9-26 (2021).
  • [30] A. P. Mackenzie, R. K. W. Haselwimmer, A. W. Tyler, G. G. Lonzarich, Y. Mori, S. Nishizaki, and Y. Maeno, Extremely Strong Dependence of Superconductivity on Disorder in Sr2RuO4, Phys. Rev. Lett. 80, 161 (1998).
  • [31] S. Nakatsuji and Y. Maeno, Quasi-Two-Dimensional Mott Transition System Sr2-xCaxRuO4, Phys. Rev. Lett. 84, 2666 (2000).
  • [32] J. P. Carlo, T. Goko, I. M. Gat-Malureanu, P. L. Russo, A. T. Savici, A. A. Aczel, G. J. MacDougall, J. A. Rodriguez, T. J. Williams, G. M. Luke, C. R. Wiebe, Y. Yoshida, S. Nakatsuji, Y. Maeno, T. Taniguchi, and Y. J. Uemura, New magnetic phase diagram of (Sr,Ca)2RuO4, Nat. Mater. 11, 323 (2012).
  • [33] M. Minakata and Y. Maeno, Magnetic ordering in Sr2RuO4 induced by nonmagnetic impurities, Phys. Rev. B 63, 180504(R) (2001).
  • [34] N. Kikugawa, A. P. Mackenzie, C. Bergemann, R. A. Borzi, S. A. Grigera, and Y. Maeno, Rigid-band shift of the Fermi level in the strongly correlated metal: Sr2-yLayRuO4, Phys. Rev. B 70, 060508(R) (2004).
  • [35] N. Kikugawa, C. Bergemann, A. P. Mackenzie, and Y. Maeno, Band-selective modification of the magnetic fluctuations in Sr2RuO4: A study of substitution effects, Phys. Rev. B 70, 134520 (2004).
  • [36] K. M. Shen, N. Kikugawa, C. Bergemann, L. Balicas, F. Baumberger, W. Meevasana, N. J. C. Ingle, Y. Maeno, Z.-X. Shen, and A. P. Mackenzie, Evolution of the Fermi Surface and Quasiparticle Renormalization through a van Hove Singularity in Sr2-yLayRuO4, Phys. Rev. Lett. 99, 187001 (2007).
  • [37] T. Imai, A. W. Hunt, K. R. Thurber, and F. C. Chou, 17O NMR Evidence for Orbital Dependent Ferromagnetic Correlations in Sr2RuO4, Phys. Rev. Lett. 81, 3006 (1998).
  • [38] Y. Sidis, M. Braden, P. Bourges, B. Hennion, S. NishiZaki, Y. Maeno, and Y. Mori, Evidence for Incommensurate Spin Fluctuations in Sr2RuO4, Phys. Rev. Lett. 83, 3320 (1999).
  • [39] M. Braden, Y. Sidis, P. Bourges, P. Pfeuty, J. Kulda, Z. Mao, and Y. Maeno, Inelastic neutron scattering study of magnetic excitations in Sr2RuO4, Phys. Rev. B 66, 064522 (2002).
  • [40] P. Steffens, Y. Sidis, J. Kulda, Z. Q. Mao, Y. Maeno, I. I. Mazin, and M. Braden, Spin Fluctuations in Sr2RuO4 from Polarized Neutron Scattering: Implications for Superconductivity, Phys. Rev. Lett. 122, 047004 (2019).
  • [41] K. Jenni, S. Kunkemöller, P. Steffens, Y. Sidis, R. Bewley, Z. Q. Mao, Y. Maeno, and M. Braden, Neutron scattering studies on spin fluctuations in Sr2RuO4, Phys. Rev. B 103, 104511 (2021).
  • [42] J. E. Ortmann, J. Y. Liu, J. Hu, M. Zhu, J. Peng, M. Matsuda, X. Ke, and Z. Q. Mao, Competition Between Antiferromagnetism and Ferromagnetism in Sr2RuO4 Probed by Mn and Co Doping, Sci. Rep. 3, 2950 (2013).
  • [43] K. Behnia, Fundamentals of Thermoelectricity, Oxford University Press (2015).
  • [44] Z. Q. Mao, Y. Maeno, and H. Fukazawa, Crystal growth of Sr2RuO4, Mater. Res. Bull. 35, 1813 (2000).
  • [45] H. Yoshino, K. Murata, N. Shirakawa, Y. Nishihara, Y. Maeno, and T. Fujita, Thermopower of a Layered Perovskite Superconductor, Sr2RuO4, J. Phys. Soc. Jpn. 65, 1548 (1996).
  • [46] X. F. Xu, Z. A. Xu, T. J. Liu, D. Fobes, Z. Q. Mao, J. L. Luo, and Y. Liu, Band-Dependent Normal-State Coherence in Sr2RuO4: Evidence from Nernst Effect and Thermopower Measurements, Phys. Rev. Lett. 101, 057002 (2008).
  • [47] J. Mravlje and A. Georges, Thermopower and Entropy: Lessons from Sr2RuO4, Phys. Rev. Lett. 117, 036401 (2016).
  • [48] J. Mravlje, M. Aichhorn, T. Miyake, K. Haule, G. Kotliar, and A. Georges, Coherence-Incoherence Crossover and the Mass-Renormalization Puzzles in Sr2RuO4, Phys. Rev. Lett. 106, 096401 (2011).
  • [49] R. Okazaki, A. Horikawa, Y. Yasui, and I. Terasaki, Photo-Seebeck Effect in ZnO, J. Phys. Soc. Jpn. 81, 114722 (2012).
  • [50] K. Behnia, D. Jaccard, and J. Flouquet, On the thermoelectricity of correlated electrons in the zero-temperature limit, J. Phys.: Condens. Matter 16, 5187 (2004).
  • [51] I. Terasaki, Cobalt Oxides and Kondo Semiconductors: A Pseudogap System as a Thermoelectric Material, Mater. Trans. 42, 951 (2001).
  • [52] Y. Ikeda, K. Saito, and R. Okazaki, Thermoelectric transport in the layered Ca3Co4-xRhxO9 single crystals, J. Appl. Phys. 119, 225105 (2016).
  • [53] C. Collignon, A. Ataei, A. Gourgout, S. Badoux, M. Lizaire, A. Legros, S. Licciardello, S. Wiedmann, J.-Q. Yan, J.-S. Zhou, Q. Ma, B. D. Gaulin, N. Doiron-Leyraud, and L. Taillefer, Thermopower across the phase diagram of the cuprate La1.6-xNd0.4SrxCuO4: Signatures of the pseudogap and charge density wave phases, Phys. Rev. B 103, 155102 (2021).
  • [54] V. Zlatić, B. Horvatić, I. Milat, B. Coqblin, G. Czycholl, and C. Grenzebach, Thermoelectric power of cerium and ytterbium intermetallics, Phys. Rev. B 68, 104432 (2003).
  • [55] K. Kurita, H. Sakabayashi, and R. Okazaki, Correlation in transport coefficients of hole-doped CuRhO2 single crystals, Phys. Rev. B 99, 115103 (2019).
  • [56] C. Pappa, J. Hammann and C. Jacoboni, Spin-glass-like H-T phase diagram for the frustrated insulator CsNiFeF6, J. Phys. C: Solid State Phys. 17, 1303 (1984).
  • [57] G. C. DeFotis, G. S. Coker, J. W. Jones, C. S. Branch, H. A. King, J. S. Bergman, S. Lee, and J. R. Goodey, Static magnetic properties and relaxation of the insulating spin glass Co1-xMnxCl2{}_{2}\cdot H2O, Phys. Rev. B 58, 12178 (1998).
  • [58] S. Dhar, O. Brandt, A. Trampert, K. J. Friedland, Y. J. Sun, and K. H. Ploog, Observation of spin-glass behavior in homogeneous (Ga,Mn)N layers grown by reactive molecular-beam epitaxy, Phys. Rev. B 67, 165205 (2003).
  • [59] J. Lago, S. J. Blundell, A. Eguia, M. Jansen, and T. Rojo, Three-dimensional Heisenberg spin-glass behavior in SrFe0.90Co0.10O3.0, Phys. Rev. B 86, 064412 (2012).
  • [60] M. Gabay, and G. Toulouse, Coexistence of Spin-Glass and Ferromagnetic Orderings, Phys. Rev. Lett. 47, 201 (1981).
  • [61] T. D. Yamamoto, H. Taniguchi, Y. Yasui, S. Iguchi, T. Sasaki, and I. Terasaki, Magneto-thermopower in the Weak Ferromagnetic Oxide CaRu0.8Sc0.2O3: An Experimental Test for the Kelvin Formula in a Magnetic Material, J. Phys. Soc. Jpn. 86, 104707 (2017).
  • [62] N. Tsujii, A. Nishide, J. Hayakawa, and T. Mori, Observation of enhanced thermopower due to spin fluctuation in weak itinerant ferromagnet, Sci. Adv. 5, eaat5935 (2019).
  • [63] T. Ikeura, T. Matsubara, Y. Machida, K. Izawa, N. Nagasawa, K. T. Matsumoto, T. Onimaru, and T. Takabatake, Anomalous Enhancement of Seebeck Coefficient in PrIr2Zn02{}_{2}0, JPS Conf. Proc. 3, 011091 (2014).
  • [64] D. K. C. Macdonald, W. B. Pearson, and I. M. Templeton, Thermo-electricity at low temperatures VIII. Thermo-electricity of the alkali metals below 2 K, Proc. Royal Soc. London A, 256, 334 (1960).
  • [65] I. Paul and G. Kotliar, Thermoelectric behavior near the magnetic quantum critical point, Phys. Rev. B 64, 184414 (2001).
  • [66] K. Izawa, K. Behnia, Y. Matsuda, H. Shishido, R. Settai, Y. Onuki, and J. Flouquet, Thermoelectric Response Near a Quantum Critical Point: The Case of CeCoIn5, Phys. Rev. Lett. 99, 147005 (2007).
  • [67] S. Hartmann, N. Oeschler, C. Krellner, C. Geibel, S. Paschen, and F. Steglich, Thermopower Evidence for an Abrupt Fermi Surface Change at the Quantum Critical Point of YbRh2Si2, Phys. Rev. Lett. 104, 096401 (2010).
  • [68] L. Malone, L. Howald, A. Pourret, D. Aoki, V. Taufour, G. Knebel, and J. Flouquet, Thermoelectricity of the ferromagnetic superconductor UCoGe, Phys. Rev. B 85, 024526 (2012).
  • [69] H. Pfau, R. Daou, M. Brando, and F. Steglich, Thermoelectric transport across the metamagnetic transition of CeRu2Si2, Phys. Rev. B 85, 035127 (2012).
  • [70] Y. Shimizu, A. Pourret, G. Knebel, A. Palacio-Morales, and D. Aoki, Non-Fermi-liquid nature and exotic thermoelectric power in the heavy-fermion superconductor UBe13, Phys. Rev. B 92, 241101(R) (2015).
  • [71] P. Limelette, W. Saulquin, H. Muguerra, and D. Grebille, From quantum criticality to enhanced thermopower in strongly correlated layered cobalt oxide, Phys. Rev. B 81, 115113 (2010).
  • [72] Y. Maeno, K. Yoshida, H. Hashimoto, S. Nishizaki, S. Ikeda, M. Nohara, T. Fujita, A. P. Mackenzie, N. E. Hussey, J. G. Bednorz, and F. Lichtenberg, Two-Dimensional Fermi Liquid Behavior of the Superconductor Sr2RuO4, J. Phys. Soc. Jpn. 66, 1405 (1997).
  • [73] K.-S. Kim and C. Pépin, Thermopower as a signature of quantum criticality in heavy fermions, Phys. Rev. B 81, 205108 (2010).
  • [74] K. Miyake and H. Kohno, Theory of Quasi-Universal Ratio of Seebeck Coefficient to Specific Heat in Zero-Temperature Limit in Correlated Metals, J. Phys. Soc. Jpn. 74, 254 (2005).
  • [75] T. Oguchi, Electronic band structure of the superconductor Sr2RuO4, Phys. Rev. B 51, 1385 (1995).
  • [76] D. J. Singh, Relationship of Sr2RuO4 to the superconducting layered cuprates, Phys. Rev. B 52, 1358 (1995).
  • [77] A. P. Mackenzie, S. R. Julian, A. J. Diver, G. J. McMullan, M. P. Ray, G. G. Lonzarich, Y. Maeno, S. Nishizaki, and T. Fujita, Quantum Oscillations in the Layered Perovskite Superconductor Sr2RuO4, Phys. Rev. Lett. 76, 3786 (1996).
  • [78] S. Hill, J. S. Brooks, Z. Q. Mao, and Y. Maeno, Cyclotron resonance and effective mass renormalizations in Sr2RuO4, Physica B 280, 283 (2000).
  • [79] K. Yano, T. Sakakibara, T. Tayama, M. Yokoyama, H. Amitsuka, Y. Homma, P. Miranović, M. Ichioka, Y. Tsutsumi, and K. Machida, Field-Angle-Dependent Specific Heat Measurements and Gap Determination of a Heavy Fermion Superconductor URu2Si2, Phys. Rev. Lett. 100, 017004 (2008).
  • [80] N. Kikugawa and Y. Maeno, Non-Fermi-Liquid Behavior in Sr2RuO4 with Nonmagnetic Impurities, Phys. Rev. Lett. 89, 117001 (2002).
  • [81] N. Tateiwa, T. D. Matsuda, Y. Ōnuki, Y. Haga, and Z. Fisk, Strong correlation between anomalous quasiparticle scattering and unconventional superconductivity in the hidden-order phase of URu2Si2, Phys. Rev. B 85, 054516 (2012).
  • [82] J. S. Bobowski, N. Kikugawa, T. Miyoshi, H. Suwa, H.-s. Xu, S. Yonezawa, D. A. Sokolov, A. P. Mackenzie, and Y. Maeno, Improved Single-Crystal Growth of Sr2RuO4, Condens. Matter 4, 6 (2019).
  • [83] M. A. Tanatar, M. Suzuki, S. Nagai, Z. Q. Mao, Y. Maeno, and T. Ishiguro, Anisotropy of Magnetothermal Conductivity in Sr2RuO4, Phys. Rev. Lett. 86, 2649 (2001).
  • [84] K. Izawa, H. Takahashi, H. Yamaguchi, Yuji Matsuda, M. Suzuki, T. Sasaki, T. Fukase, Y. Yoshida, R. Settai, and Y. Onuki, Superconducting Gap Structure of Spin-Triplet Superconductor Sr2RuO4 Studied by Thermal Conductivity, Phys. Rev. Lett. 86, 2653 (2001).
  • [85] M. Suzuki, M. A. Tanatar, N. Kikugawa, Z. Q. Mao, Y. Maeno, and T. Ishiguro, Universal Heat Transport in Sr2RuO4, Phys. Rev. Lett. 88, 227004 (2002).
  • [86] A. A. Varlamov, V. S. Egorov, and A. V. Pantsulaya, Kinetic properties of metals near electronic topological transitions (2 1/2-order transitions), Adv. Phys. 38, 469 (1989).
  • [87] N. Ito, M. Ishii, and R. Okazaki, Enhanced Seebeck coefficient by a filling-induced Lifshitz transition in KxRhO2, Phys. Rev. B 99, 041112(R) (2019).