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

Phenomenological model for the third-harmonic magnetic response of superconductors: application to Sr2RuO4

Fei Chen School of Physics and Astronomy, University of Minnesota, Minneapolis, 55455, USA    Damjan Pelc School of Physics and Astronomy, University of Minnesota, Minneapolis, 55455, USA Department of Physics, Faculty of Science, University of Zagreb, Bijenička 32, HR-10000 Zagreb, Croatia    Martin Greven School of Physics and Astronomy, University of Minnesota, Minneapolis, 55455, USA    Rafael M. Fernandes School of Physics and Astronomy, University of Minnesota, Minneapolis, 55455, USA
Abstract

We employ the phenomenological Lawrence-Doniach model to compute the contributions of the superconducting fluctuations to the third-harmonic magnetic response, denoted here by M3¯\overline{M_{3}}, which can be measured in a precise way using ac magnetic fields and lock-in techniques. We show that, in an intermediate temperature regime, this quantity behaves as the third-order nonlinear susceptibility, which shows a power-law dependence with the reduced temperature ϵ=TTcTc\epsilon=\frac{T-T_{c}}{T_{c}} as ϵ5/2\epsilon^{-5/2}. Very close to TcT_{c}, however, M3¯\overline{M_{3}} saturates due to the nonzero amplitude of the ac field. We compare our theoretical results with experimental data for three conventional superconductors – lead, niobium, and vanadium – and for the unconventional superconductor Sr2RuO4 (SRO). We find good agreement between theory and experiment for the elemental superconductors, although the theoretical values for the critical field systematically deviate from the experimental ones. In the case of SRO, however, the phenomenological model completely fails to describe the data, as the third-harmonic response remains sizable over a much wider reduced temperature range compared to Pb, Nb, and V. We show that an inhomogeneous distribution of TcT_{c} across the sample can partially account for this discrepancy, since regions with a locally higher TcT_{c} contribute to the fluctuation M3¯\overline{M_{3}} significantly more than regions with the “nominal” TcT_{c} of the clean system. However, the exponential temperature dependence of M3¯\overline{M_{3}} first reported in Ref. [D. Pelc et. al., Nature Comm. 10, 2729 (2019)] is not captured by the model with inhomogeneity. We conclude that, while inhomogeneity is an important ingredient to understand the superconducting fluctuations of SRO and other perovskite superconductors, additional effects may be at play, such as non-Gaussian fluctuations or rare-region effects.

I Introduction

In unconventional superconductors, not only the gap function, but also the superconducting fluctuations can be quite different from their conventional counterparts (for reviews, see Ref. (Varlamov et al., 2018; Larkin and Varlamov, 2005)). Indeed, several high-TcT_{c} superconductors have strongly anisotropic properties and small coherence lengths, suggestive of a wider temperature range in which fluctuations are important. Moreover, the magnitude of these fluctuations as well as their temperature dependence can also display unusual behaviors (Pelc et al., 2019). Signatures of superconducting fluctuations have been widely probed in both conventional and unconventional superconductors, in observables as diverse as specific heat (Suzuki and Tsuboi, 1977; Tsuboi and Suzuki, 1977; Tallon et al., 2011), linear and nonlinear conductivity (Glover, 1967; Strongin et al., 1968; Ruggiero et al., 1980; Mircea et al., 2009; Rullier-Albenque et al., 2011; Carbillet et al., 2016; Popčević et al., 2018; Pelc et al., 2018), microwave and THz response (Corson et al., 1999; Orenstein et al., 2006; Grbić et al., 2011; Bilbro et al., 2011; Pelc et al., 2018), susceptibility (Geballe et al., 1971; Gollub et al., 1973; Li et al., 2010; Kokanović et al., 2013; Kasahara et al., 2016; Yu et al., 2019), and the Nernst coefficient (Wang et al., 2006; Rullier-Albenque et al., 2006; Chang et al., 2012; Tafti et al., 2014; Behnia and Aubin, 2016).

Experimentally, one of the main difficulties is to unambiguously identify contributions that can be uniquely attributed to superconducting fluctuations, since these are usually small compared to the regular normal-state contributions (Geballe et al., 1971). Theoretically, modeling contributions of superconducting fluctuations to the magnetic susceptibility and to the conductivity, both phenomenologically and microscopically, dates back several decades (Schmidt, 1968; Shmidt, 1968; Maki, 1968; Thompson, 1970; Schmid, 1969; Prange, 1970; Abrahams et al., 1970; Kurkijärvi et al., 1972; Aslamazov and Larkin, 1974). More recent studies on superconducting fluctuations have focused on the role of phase fluctuations (Emery and Kivelson, 1995), on disordered 2D superconductors (Glatz et al., 2011), and on thermal and electric transport properties above TcT_{c} in cuprates (Ullah and Dorsey, 1991; Fisher et al., 1991; Ioffe et al., 1993; Ussishkin et al., 2002; Serbyn et al., 2009; Michaeli and Finkel'stein, 2009; Li and Levchenko, 2020).

Recently, a method to probe superconducting fluctuations based on the third-harmonic magnetic response was put forward in Ref. (Pelc et al., 2019). Specifically, an ac magnetic field H(t)=H0cos(ωt)H(t)=H_{0}\cos(\omega t) is applied and the magnetization is measured at a frequency 3ω3\omega. This observable, which we hereafter denote by M3¯\overline{M_{3}}, is related to, but not identical to the standard nonlinear susceptibility χ3\chi_{3}. The key point is that the third-harmonic response M3¯\overline{M_{3}} is vanishingly small in the normal state. As a result, its magnitude and temperature dependence near the superconducting transition temperature TcT_{c} should be dominated by superconducting fluctuations. In Ref. (Pelc et al., 2019), it was empirically found that M3¯\overline{M_{3}} displays an unusual exponential temperature dependence in perovskite-based superconductors such as cuprates, Sr2RuO4 (SRO) and SrTiO3, as opposed to a power-law temperature dependence in standard electron-phonon superconductors. However, the implications of these observations for the nature of superconducting fluctuations in unconventional superconductors remain unsettled.

In this paper, we employ a phenomenological approach based on the Lawrence-Doniach (LD) free-energy to compute the contributions to the experimentally-measured quantity M3¯\overline{M_{3}} of Ref. (Pelc et al., 2019) arising from Gaussian superconducting fluctuations. The main appeal of such an approach is that, being phenomenological, it is potentially applicable to both conventional and unconventional superconductors. In particular, we perform a quantitative comparison between the theoretical results predicted by the LD formalism and the data on several elemental superconductors (Pb, Nb, V) and on the unconventional superconductor SRO. We find that the LD result provides a good description of the data for elemental superconductors over a wide range of reduced temperature values, ϵTTcTc\epsilon\equiv\frac{T-T_{c}}{T_{c}}, and correctly captures the observed 5/25/2 power-law behavior of M3¯\overline{M_{3}} for intermediate values of ϵ\epsilon. The theoretically extracted values for the zero-temperature upper critical field Hc2(0)H_{c2}(0) differ by factors of 22 to 66 from the experimental ones; we argue that this difference could be an artifact of the LD model, which was developed for layered superconductors rather than cubic systems. Overall, the results demonstrate that measurements of the third-harmonic magnetic response are indeed a powerful probe of superconducting fluctuations.

However, in the case of Sr2RuO4, we find a sharp disagreement between the LD theoretical results and the data for M3¯\overline{M_{3}}. Not only is the temperature dependence qualitatively different, but the observed magnitude of M3¯\overline{M_{3}} near TcT_{c} is strongly underestimated by the theoretical model. Motivated by the evidence for significant inhomogeneity in several perovskite-based superconductors (Pelc et al., 2018, 2019, 2021), we modify our LD model for M3¯\overline{M_{3}} and include a distribution of TcT_{c} values. We find that even a modest width of this TcT_{c} distribution is capable of capturing the typical values of M3¯\overline{M_{3}} observed experimentally. However, this modification is not sufficient to explain the exponential temperature dependence reported in Ref. (Pelc et al., 2019). We thus conclude that while inhomogeneity at the mean-field level is important to elucidate the behavior of superconducting fluctuations in Sr2RuO4, it is likely not the sole reason for the observed exponential temperature dependence. One possibility is that such behavior arises from rare-region contributions (Dodaro and Kivelson, 2018; Pelc et al., 2019, 2021) or from non-Gaussian fluctuations, which are absent in the LD model employed here.

The paper is organized as follows: in Sec. II, we employ the LD model to derive an expression for the third-harmonic magnetic response M3¯\overline{M_{3}}, and discuss the temperature dependence of this quantity in different regimes. Sec. III presents a quantitative comparison between the theoretical and experimental results for three conventional superconductors (Pb, Nb, and V) and the unconventional superconductor Sr2RuO4. We note that some of the data were previously published in Ref. (Pelc et al., 2019). An extension of the model presented in Sec. II that includes the role of inhomogeneity is also introduced. Our conclusions are presented in Sec. IV.

II Phenomenological model for the third-harmonic magnetic response

In this section, we derive an expression for the third-harmonic magnetic response M3¯\overline{M_{3}}, measured in the experiments of Ref. (Pelc et al., 2019), based on the Lawrence-Doniach (LD) approach. We first review the contribution of the superconducting fluctuations to the magnetization in the presence of a static magnetic field within the LD approach. Here we only quote the LD results, which are well known from the literature – for their derivations, see for instance Refs. (Larkin and Varlamov, 2005; Mishonov and Penev, 2000). Using the LD results, we then proceed to include an ac field to explicitly calculate M3¯\overline{M_{3}}, and discuss its temperature dependence in different regimes.

II.1 Linear and nonlinear susceptibilities in the Lawrence-Doniach model

Fluctuations of a superconductor in the presence of an external magnetic field can be modeled within the phenomenological Ginzburg-Landau framework. In a regime close to TcT_{c}, the general superconducting Ginzburg-Landau free-energy functional takes the form:

Δ[Ψ(𝐱)]\displaystyle\Delta\mathcal{F}\left[\Psi\left(\mathbf{x}\right)\right] =ddx(a|Ψ|2+b2|Ψ|4\displaystyle=\int d^{d}x\left(a\left|\Psi\right|^{2}+\frac{b}{2}\left|\Psi\right|^{4}\right. (1)
+12m|(ie𝐀)Ψ|2+18π|×𝐀|2)\displaystyle\left.+\frac{1}{2m^{*}}\left|\left(\frac{\nabla}{i}-e^{*}\mathbf{A}\right)\Psi\right|^{2}+\frac{1}{8\pi}\left|\nabla\times\mathbf{A}\right|^{2}\right)

Here, Ψ(𝐱)\Psi\left(\mathbf{x}\right) is the superconducting order parameter, m=2mm^{*}=2m and e=2ee^{*}=2e are the mass and charge of a Cooper pair, 𝐀\mathbf{A} is the vector potential, and b>0b>0 is a Ginzburg-Landau parameter. The coefficient aa is parametrized as a=α(TTc)=αTcϵa=\alpha\left(T-T_{c}\right)=\alpha T_{c}\epsilon, where ϵ=TTcTc\epsilon=\frac{T-T_{c}}{T_{c}} is the reduced temperature and α\alpha a positive constant. Near TcT_{c}, but above the temperature range where critical fluctuations become important, as set by the Ginzburg-Levanyuk parameter, one assumes that the order parameter is small and slowly-varying. As a result, the quartic term in Eq. (1) can be neglected, and only Gaussian fluctuations are considered:

Δ[Ψ(𝐱)]=ddx(a|Ψ|2+14m|(i2e𝐀)Ψ|2)\Delta\mathcal{F}\left[\Psi\left(\mathbf{x}\right)\right]=\int{d^{d}x\left(a\left|\Psi\right|^{2}+\frac{1}{4m}\left|\left(\frac{\nabla}{i}-2e\mathbf{A}\right)\Psi\right|^{2}\right)} (2)

To obtain the Lawrence-Doniach (LD) free-energy expression, one assumes a layered superconductor and considers a magnetic field HH applied perpendicular to the layers. A detailed derivation can be found in standard textbooks and review papers, see for instance Refs. (Larkin and Varlamov, 2005; Mishonov and Penev, 2000). For completeness, we only highlight the main steps of the derivation and quote the results from Ref. (Larkin and Varlamov, 2005). Because of the layered nature of the system, there is a difference between in-plane and out-of-plane kinetic terms. While the former assumes the same form as in Eq. (1), the latter is described by δz|Ψl+1Ψl|2\delta_{z}\left|\Psi_{l+1}-\Psi_{l}\right|^{2}, where δz\delta_{z} is the inter-layer coupling constant and the subscript ll is a layer index. It is also convenient to introduce two dimensionless quantities, hh and rr. By using the result Hc2(0)=2mαTceH_{c2}\left(0\right)=\frac{2m\alpha T_{c}}{e} for the zero-temperature critical field, we define the dimensionless applied field hHHc2(0)h\equiv\frac{H}{H_{c2}(0)}. Moreover, we define the dimensionless anisotropy parameter r2δzαTcr\equiv\frac{2\delta_{z}}{\alpha T_{c}}, which can also be expressed in terms of the ratio between the correlation length along the zz direction, ξz(0)\xi_{z}(0), and the inter-layer separation ss, r=4ξz2(0)s2r=\frac{4\xi_{z}^{2}\left(0\right)}{s^{2}}. Writing the order parameter in a product form between in-plane Landau-level wave functions and plane waves propagating along the zz direction, one can evaluate the partition function Z=DΨDΨeΔ[Ψ(𝐱)]TZ=\int{D\Psi D\Psi^{*}e^{-\frac{\Delta\mathcal{F}\left[\Psi\left(\mathbf{x}\right)\right]}{T}}} and then obtain the LD free-energy expression (up to a constant) (Larkin and Varlamov, 2005; Mishonov and Penev, 2000):

F(ϵ)MHc2(0)\displaystyle\frac{F\left(\epsilon\right)}{M_{\infty}H_{c2}(0)} =2(ϵ+1)hln2[(ϵ+r2)lnh2h12ln2π\displaystyle=-\frac{2\left(\epsilon+1\right)h}{\ln 2}\left[\left(\epsilon+\frac{r}{2}\right)\frac{\ln h}{2h}-\frac{1}{2}\ln 2\pi\right.
+0π/2dϕπ/2lnΓ(12+ϵ+rsin2ϕ2h)]\displaystyle\left.+\int_{0}^{\pi/2}\frac{d\phi}{\pi/2}\,\ln\Gamma\left(\frac{1}{2}+\frac{\epsilon+r\sin^{2}\phi}{2h}\right)\right] (3)

Here, Γ(x)\Gamma(x) is the gamma function, the integration over the variable ϕ\phi effectively sums over the layers, vv is the volume, and MTcΦ0sln22M_{\infty}\equiv\frac{T_{c}}{\Phi_{0}s}\frac{\ln 2}{2} is the absolute value of the saturation magnetization at TcT_{c}, with Φ0\Phi_{0} denoting the flux quantum. Similarly, the LD expression for the magnetization is given by (Larkin and Varlamov, 2005; Mishonov and Penev, 2000):

M(ϵ)M=\displaystyle\frac{M\left(\epsilon\right)}{M_{\infty}}= 2(ϵ+1)ln20π/2dϕπ/2{ϵ+rsin2ϕ2h×\displaystyle-\frac{2\left(\epsilon+1\right)}{\ln 2}\int_{0}^{\pi/2}\frac{d\phi}{\pi/2}\left\{\frac{\epsilon+r\sin^{2}{\phi}}{2h}\times\right. (4)
[ψ(ϵ+rsin2ϕ2h+12)1]\displaystyle\left[\psi\left(\frac{\epsilon+r\sin^{2}{\phi}}{2h}+\frac{1}{2}\right)-1\right]
lnΓ(ϵ+rsin2ϕ2h+12)+12ln2π}\displaystyle\left.-\ln{\Gamma}\left(\frac{\epsilon+r\sin^{2}{\phi}}{2h}+\frac{1}{2}\right)+\frac{1}{2}\ln{2\pi}\right\}

where ψ(x)=dlnΓ(x)dx\psi(x)=\frac{d\ln\Gamma(x)}{dx} is the digamma function. By taking hϵ,rh\gg\epsilon,r in Eq. (4), the right-hand side gives 1-1 at ϵ=0\epsilon=0, confirming that MM_{\infty} is the saturation magnetization at TcT_{c}. Note that this expression is valid for h>0h>0; in the case of h<0h<0, symmetry implies F(h)=F(h)F\left(-h\right)=F\left(h\right) and M(h)=M(h)M\left(-h\right)=-M\left(h\right). For future reference, we list the three dimensionless parameters that will be employed throughout this work:

ϵ=TTcTc\displaystyle\epsilon=\frac{T-T_{c}}{T_{c}} (5)
r=[2ξz(0)s]2\displaystyle r=\left[\frac{2\xi_{z}\left(0\right)}{s}\right]^{2}
h=HHc2(0)\displaystyle h=\frac{H}{H_{c2}\left(0\right)}

While the anisotropy parameter rr is fixed, its impact on the magnetization depends on the temperature range probed. In a regime sufficently far from TcT_{c}, rϵr\ll\epsilon, the system essentially behaves as decoupled layers (r0r\rightarrow 0) and Eq.(4) becomes (Larkin and Varlamov, 2005; Mishonov and Penev, 2000)

M(ϵr)M=\displaystyle\frac{M\left(\epsilon\gg r\right)}{M_{\infty}}= 2(ϵ+1)ln2{ϵ2h[ψ(12+ϵ2h)1]\displaystyle-\frac{2\left(\epsilon+1\right)}{\ln 2}\left\{\frac{\epsilon}{2h}\left[\psi\left(\frac{1}{2}+\frac{\epsilon}{2h}\right)-1\right]\right. (6)
lnΓ(12+ϵ2h)2π}.\displaystyle\left.-\ln\frac{{\Gamma}\left(\frac{1}{2}+\frac{\epsilon}{2h}\right)}{\sqrt{2\pi}}\right\}.

On the other hand, as TcT_{c} is approached, the system will eventually cross over to the regime rϵr\gg\epsilon. Then, the three-dimensional nature of the system cannot be neglected, and the magnetization becomes (Larkin and Varlamov, 2005; Mishonov and Penev, 2000; Kurkijärvi et al., 1972):

M(ϵr)M=\displaystyle\frac{M\left(\epsilon\ll r\right)}{M_{\infty}}= 6(ϵ+1)ln2(2r)1/2h[ζ(12,12+ϵ2h)\displaystyle-\frac{6\left(\epsilon+1\right)}{\ln 2}\left(\frac{2}{r}\right)^{1/2}\sqrt{h}\left[\zeta\left(-\frac{1}{2},\frac{1}{2}+\frac{\epsilon}{2h}\right)\right. (7)
13ζ(12,12+ϵ2h)ϵ2h]\displaystyle\left.-\frac{1}{3}\zeta\left(\frac{1}{2},\frac{1}{2}+\frac{\epsilon}{2h}\right)\frac{\epsilon}{2h}\right]

where ζ(ν,x)\zeta(\nu,x) is Hurwitz zeta function.

Refer to caption
Figure 1: Magnetization (red curve, in units of MM_{\infty}) induced by superconducting fluctuations, in the presence of a dc field hh, as a function of the reduced temperature ϵ\epsilon according to Eq. (4). We also include for comparison the asymptotic expressions for M(ϵr)M(\epsilon\gg r) (green dashed curve) and M(ϵr)M(\epsilon\ll r) (blue dotted curve), Eqs. (6) and (7), respectively. A crossover clearly takes place when ϵr\epsilon\sim r. The dimensionless parameters chosen here were h=0.01,r=0.5h=0.01,\,r=0.5. The insets are zooms on different temperature ranges.

Therefore, as TcT_{c} is approached from above, we expect a crossover of the temperature-dependent magnetization from 2D-like behavior to 3D-like behavior, with the crossover temperature corresponding to ϵr\epsilon\sim r. This general behavior is illustrated in Fig. 1, where MM given by Eq. (4) is plotted as a function of the reduced temperature ϵ\epsilon together with the asymptotic expressions in Eqs. (6)-(7) for a fixed field value. As expected, the contribution of the superconducting fluctuations to the magnetization are negative.

It will be useful later to contrast the temperature dependence of the third-harmonic response M3¯\overline{M_{3}} with that of the nonlinear magnetic susceptibility. To derive the latter, we consider the limit of small fields, i.e. when the dimensionless magnetic field is the smallest parameter of the problem, hϵ,rh\ll\epsilon,r. Going back to the main expression for the magnetization in Eq. (4), it is convenient to define y=ϵ+rsin2ϕ2hy=\frac{\epsilon+r\sin^{2}\phi}{2h}. Since hϵ,rh\ll\epsilon,r, it follows that y1y\gg 1 and the integrand can be expanded as:

y[ψ(y+12)1]lnΓ(y+12)+12ln(2π)\displaystyle y\left[\psi\left(y+\frac{1}{2}\right)-1\right]-\ln\Gamma\left(y+\frac{1}{2}\right)+\frac{1}{2}\ln(2\pi) =\displaystyle=
112y7720y3+316720y5+𝒪(x7)\displaystyle\frac{1}{12y}-\frac{7}{720y^{3}}+\frac{31}{6720y^{5}}+\mathcal{O}\left(x^{-7}\right) (8)

The integrals over ϕ\phi can be analytically evaluated. Expanding the magnetization in odd powers of hh,

MM=χ1h+χ3h3+χ5h5+𝒪(h7)\frac{M}{M_{\infty}}=\chi_{1}h+\chi_{3}h^{3}+\chi_{5}h^{5}+\mathcal{O}(h^{7}) (9)

we find the following expressions for the linear and nonlinear susceptibilities (see also Refs. (Tsuzuki and Koyanagi, 1969; Mishonov and Penev, 2000)):

χ1\displaystyle\chi_{1} =(1+ϵ)3ln21ϵ1/2ϵ+r\displaystyle=-\frac{\left(1+\epsilon\right)}{3\ln 2}\frac{1}{\epsilon^{1/2}\sqrt{\epsilon+r}} (10)
χ3\displaystyle\chi_{3} =7(1+ϵ)360ln2(3r2+8rϵ+8ϵ2)ϵ5/2(ϵ+r)5/2\displaystyle=\frac{7\left(1+\epsilon\right)}{360\ln 2}\frac{\left(3r^{2}+8r\epsilon+8\epsilon^{2}\right)}{\epsilon^{5/2}\left(\epsilon+r\right)^{5/2}} (11)
χ5\displaystyle\chi_{5} =31(1+ϵ)13440ln2×\displaystyle=-\frac{31\left(1+\epsilon\right)}{13440\ln 2}\times
(35r4+160r3ϵ+288r2ϵ2+256rϵ3+128ϵ4)ϵ9/2(ϵ+r)9/2\displaystyle\frac{\left(35r^{4}+160r^{3}\epsilon+288r^{2}\epsilon^{2}+256r\epsilon^{3}+128\epsilon^{4}\right)}{\epsilon^{9/2}\left(\epsilon+r\right)^{9/2}} (12)

Close enough to TcT_{c}, when ϵr\epsilon\ll r, we find the following power-law behaviors

χ1\displaystyle\chi_{1} ϵ1/2r\displaystyle\sim-\frac{\epsilon^{-1/2}}{\sqrt{r}} (13)
χ3\displaystyle\chi_{3} ϵ5/2r\displaystyle\sim\frac{\epsilon^{-5/2}}{\sqrt{r}} (14)
χ5\displaystyle\chi_{5} ϵ9/2r\displaystyle\sim-\frac{\epsilon^{-9/2}}{\sqrt{r}} (15)

II.2 The third-harmonic magnetic response M3¯\overline{M_{3}}: experimental setup and theory

One of the most common experimental probes of superconducting fluctuations is to apply a dc magnetic field and measure the magnetic response, see Eq. (9). The key issue with measuring the linear susceptibility χ1\chi_{1} is that the diamagnetic contribution due to the superconducting fluctuations is typically much smaller than the paramagnetic contributions from other normal-state degrees of freedom. For the nonlinear susceptibility χ3\chi_{3}, however, one generally expects that the intrinsic normal-state contribution is negligible in most cases, which could in principle allow one to assess the contribution from the superconducting fluctuations in a more unambiguous fashion. Note that, while in principle the susceptibilities χ1\chi_{1} and χ3\chi_{3} are tensor quantities, our experimental setup is designed in such a way that both the excitation and detection coils are along the same axis. We therefore only measure in-plane diagonal components, which are equivalent for a tetragonal or cubic system. Hereafter we refer only to a scalar χ3\chi_{3}.

Instead of applying a dc magnetic field, the experimental technique presented in Ref. (Pelc et al., 2019) and utilized here employs an ac field (of the form H0cosωtH_{0}\cos{\omega}t) and a system of coils to measure the oscillating sample magnetization. In order to determine the third-order response, a lock-in amplifier is used at the third harmonic of the fundamental frequency ω\omega, which is typically in the kHz range. If the fifth-order susceptibility is significantly smaller than the third-order susceptibility, the third harmonic response is a good measure of the third-order susceptibility. This condition was experimentally verified by measuring at the fifth harmonic, where the signal was found to be vanishingly small except extremely close to TcT_{c}, where it was still an order of magnitude smaller than the third harmonic. We can thus safely ignore the higher-order contributions. Most of the data presented here were published in Ref. (Pelc et al., 2019), and were obtained in two separate experimental setups. Low-temperature measurements on strontium ruthenate were performed in a 3He evaporation refrigerator with a custom-made set of coils. Samples of conventional superconductors were measured in a modified Quantum Design MPMS, where we used the built-in AC susceptibility coil to generate the excitation magnetic field, and a custom-made probe with small detection coils to maximize the filling factor. We estimate that the magnetization sensitivity of both setups is better than 1 nanoemu, an improvement of 1-2 orders of magnitude over standard SQUID-based instruments. This is made possible by lock-in detection, matching the impedance of the detection coils and lock-in amplifier inputs, and large filling factors of the detection coils (Drobac et al., 2013).

Refer to caption
Figure 2: Absolute value of the third-harmonic response, |M3¯||\overline{M_{3}}| in Eq. (17), in units of MM_{\infty}, as a function of the reduced temperature ϵTTcTc\epsilon\equiv\frac{T-T_{c}}{T_{c}}, plotted on a log-log scale (red curve). The dashed black line corresponds to the analytical approximation in Eq. (19), which gives a ϵ5/2\epsilon^{-5/2} power-law behavior. The dimensionless parameters used here are h0=103h_{0}=10^{-3} and r=1r=1.

Although we expect the third-harmonic response to exhibit behavior similar to the third-order nonlinear susceptibility χ3\chi_{3}, there are important differences, since the amplitude of the oscillating field, albeit small (H01H_{0}\sim 1 Oe), is nonzero. Thus, to provide a more direct comparison between the LD model and experiments, we directly compute the third-harmonic response, which we denote by M3¯\overline{M_{3}}. In our experimental setup, the signal corresponds to the Fourier transform of Mt\frac{\partial M}{\partial t} at 3ω3\omega,

M3¯(ϵ)=πωπωM(ϵ,h(t))te3iωt𝑑t,\overline{M_{3}}\left(\epsilon\right)=\int_{-\frac{\pi}{\omega}}^{\frac{\pi}{\omega}}\frac{\partial M\left(\epsilon,h(t)\right)}{\partial t}\,e^{3i\omega t}dt, (16)

where M(ϵ,h(t))M\left(\epsilon,h(t)\right) is obtained from Eq.(4) by substituting h=h0cosωth=h_{0}\cos{\omega}t. Integration by parts gives M3¯(ϵ)=3iππM(ϵ,h0cosθ)e3iθ𝑑θ\overline{M_{3}}\left(\epsilon\right)=-3i\int_{-\pi}^{\pi}M(\epsilon,h_{0}\cos\theta)e^{3i\theta}d\theta with θ=ωt\theta=\omega t. Using the fact that M(ϵ,h)=M(ϵ,h)M\left(\epsilon,-h\right)=-M\left(\epsilon,h\right), we have ππ/2M(ϵ,h0cosθ)ei3θ𝑑θ=0π/2M(ϵ,h0cosθ)ei3θ𝑑θ\int_{-\pi}^{-\pi/2}M(\epsilon,h_{0}\cos\theta)e^{i3\theta}d\theta=\int_{0}^{\pi/2}M(\epsilon,h_{0}\cos\theta)e^{i3\theta}d\theta and π/2πM(ϵ,h0cosθ)ei3θ𝑑θ=π/20M(ϵ,h0cosθ)ei3θ𝑑θ\int_{\pi/2}^{\pi}M(\epsilon,h_{0}\cos\theta)e^{i3\theta}d\theta=\int_{-\pi/2}^{0}M(\epsilon,h_{0}\cos\theta)e^{i3\theta}d\theta, which yields

M3¯(ϵ)=6iπ/2π/2𝑑θM(ϵ,h0cosθ)cos3θ,\overline{M_{3}}\left(\epsilon\right)=-6i\int_{-\pi/2}^{\pi/2}d\theta\,M\left(\epsilon,h_{0}\cos\theta\right)\cos 3\theta, (17)

where the field h0cosθh_{0}\cos\theta remains positive between the integration limits. Experimentally, both the imaginary and real parts can be measured. However, due to issues with lock-in phase determination in third-harmonic measurements (Drobac et al., 2013), we simply use the absolute value of M3¯\overline{M_{3}} for comparison between the experimental and theoretical results.

Refer to caption
Figure 3: Absolute value of the third-harmonic response |M3¯||\overline{M_{3}}| (in units of MM_{\infty}) as a function of the reduced temperature ϵ\epsilon for varying h0h_{0} values (fixed r=1r=1, panel (a)) and varying rr values (fixed h0=103h_{0}=10^{-3}, panel (b)). The dashed lines mark the power-law behavior ϵ5/2\epsilon^{-5/2} displayed by the curves with larger values of rr.

In the temperature range where h0ϵh_{0}\ll\epsilon, we can substitute the series expansion (9) in Eq. (17) and find:

|M3¯|M3π4χ3h03+15π16χ5h05.\frac{\left|\overline{M_{3}}\right|}{M_{\infty}}\approx\frac{3\pi}{4}\chi_{3}h_{0}^{3}+\frac{15\pi}{16}\chi_{5}h_{0}^{5}. (18)

Now, in the relevant regime rϵr\gg\epsilon, according to Eqs. (14), we have χ3ϵ5/2\chi_{3}\sim\epsilon^{-5/2} and χ5ϵ9/2\chi_{5}\sim\epsilon^{-9/2}. Therefore, as long as we remain in the regime h0ϵh_{0}\ll\epsilon, the contribution from the fifth-order nonlinear susceptibility χ5\chi_{5} can be neglected. Using Eq. (11) we obtain:

|M3¯|M(7π160ln2)h03(1+ϵ)ϵ5/2r\frac{\left|\overline{M_{3}}\right|}{M_{\infty}}\approx\left(\frac{7\pi}{160\ln 2}\right)\frac{h_{0}^{3}\left(1+\epsilon\right)\epsilon^{-5/2}}{\sqrt{r}} (19)

Therefore, we expect that, in the temperature range h0ϵrh_{0}\ll\epsilon\ll r, the third-harmonic response |M3¯|\left|\overline{M_{3}}\right| displays the power-law behavior (TTc)5/2\left(T-T_{c}\right)^{-5/2} characteristic of the third-order nonlinear susceptibility χ3\chi_{3}. To verify this behavior explicitly, in Fig. 2 we present the numerically calculated |M3¯||\overline{M_{3}}| for h0=103h_{0}=10^{-3} and r=1r=1, and compare it with the analytical approximation in Eq. (19). It is clear that the expected power-law behavior appears over a rather wide temperature range. As one approaches TcT_{c} from above and reaches the temperature scale ϵh0\epsilon\sim h_{0}, deviations from the power-law are observed, and |M3¯|\left|\overline{M_{3}}\right| saturates to a constant value. This is a direct consequence of the fact that we are not computing the dc susceptibility, but the ac third-harmonic response at a fixed field amplitude h0h_{0}. Figs. 3(a)-(b) depict how the temperature window in which power-law behavior is observed is affected by changing rr and h0h_{0}. As expected, increasing h0h_{0} significantly suppresses the window of power-law behavior, as the temperature scale ϵh0\epsilon\sim h_{0} is moved up. On the other hand, the anisotropy parameter rr has a rather minor impact on the temperature range in which ϵ5/2\epsilon^{-5/2} behavior is observed.

III Comparison with experimental data

III.1 Conventional Superconductors (Pb, Nb, and V)

In order to validate the LD approach for the third-harmonic response, we first compare the theoretical results for M3¯\overline{M_{3}} from Eq.(17) with the experimental third-harmonic data for three conventional elemental superconductors: lead (Pb), niobium (Nb), and vanadium (V). Besides an overall pre-factor, there are three fitting parameters in our formalism: the upper critical field Hc2H_{c2}, the critical temperature TcT_{c}, and the anisotropy ratio rr. The field H0H_{0} is 1.3 Oe as generated by the excitation coil, but the true value could be modified by demagnetization factors (especially very close and below TcT_{c}) by up to a factor of 2\sim 2. Hereafter, for concreteness, we will use H0=1.3H_{0}=1.3 Oe for all cases. Since these materials are rather three-dimensional, we expect the zz-axis correlation length ξz\xi_{z} to be larger than the layer distance ss in the LD model, i.e. r>4r>4. Thus, because the reduced temperatures probed are very small (ϵmax102\epsilon_{\mathrm{max}}\sim 10^{-2}), the precise value of rr does not significantly affect the temperature dependence of |M3¯||\overline{M_{3}}| in the experimentally relevant temperature regime (as shown above in Fig. 3(b)). Therefore, to minimize the number of fitting parameters, we set r=10r=10 in all cases. This leaves only two free parameters, Hc2H_{c2} and TcT_{c}.

Refer to caption
Figure 4: Comparison between the measured third-harmonic response |M3¯|\left|\overline{M_{3}}\right| (circle and square black symbols, in arbitrary units) for Pb and the theoretical results obtained from Eq. (17) (dashed and solid red lines). Panels (a) and (c) ((b) and (d)) show the data on a linear (logarithmic) scale. Fit parameters are shown in Table 1. In panels (a)-(b), the fit parameter is the critical field H~c2\tilde{H}_{c2} in Table 1, whereas the critical temperature is set to its experimental value Tc(exp)T_{c}^{(\mathrm{exp})}. In panels (c)-(d), the fit parameters are Hc2H_{c2} and TcT_{c}. The anisotropy parameter is set to r=10r=10.
Refer to caption
Figure 5: Comparison between the measured third-harmonic response |M3¯|\left|\overline{M_{3}}\right| (circle and square blue symbols, in arbitrary units) for Nb and the theoretical results obtained from Eq. (17) (dashed and solid red lines). Panels (a) and (c) ((b) and (d)) show the data in linear (logarithmic) scale. Fit parameters are shown in Table 1. In panels (a)-(b), the fit parameter is the critical field H~c2\tilde{H}_{c2} in Table 1, whereas the critical temperature is set to its experimental value Tc(exp)T_{c}^{(\mathrm{exp})}. In panels (c)-(d), the fitting parameters are Hc2H_{c2} and TcT_{c}. The anisotropy parameter is set to r=10r=10.
Refer to caption
Figure 6: Comparison between the measured third-harmonic response |M3¯|\left|\overline{M_{3}}\right| (circle and square green symbols, in arbitrary units) for V and the theoretical results obtained from Eq. (17) (dashed and solid red lines). Data from two different samples are presented (light green and dark green symbols). Panels (a) and (c) ((b) and (d)) show the data in linear (logarithmic) scale. Fit parameters are shown in Table 1. In panels (a)-(b), the fit parameter is the critical field H~c2\tilde{H}_{c2} in Table 1, whereas the critical temperature is set to its experimental value Tc(exp)T_{c}^{(\mathrm{exp})}. In panels (c)-(d), the fit parameters are Hc2H_{c2} and TcT_{c}. The anisotropy parameter is set to r=10r=10.

The comparison between theoretical and experimental results is shown Figs. 4, 5, and 6 for Pb, Nb, and V, respectively. In all figures, the circle and square symbols correspond to data, whereas dashed and solid lines correspond to theoretical results. Experimental measurements of |M3¯|\left|\overline{M_{3}}\right| become challenging below ϵ104\epsilon\sim 10^{-4} due to thermometry resolution issues, and the signal typically decays below the noise level around ϵ102\epsilon\sim 10^{-2}, indicating a small temperature regime of significant superconducting fluctuations. In the case of V, a kink is observed in one sample (light green symbols), which is possibly a spurious signal due to solder superconductivity or the result of a slight macroscopic sample inhomogeneity. For this reason, we also include results from a second sample (dark green symbols). Because the overall magnitude of the experimental |M3¯|\left|\overline{M_{3}}\right| is arbitrary and changes with modifications of the set-up, we rescaled the |M3¯|\left|\overline{M_{3}}\right| values of the second sample (dark green symbols) by an overall constant to better match the behavior of |M3¯|\left|\overline{M_{3}}\right| of the first sample (light green symbols) at larger ϵ\epsilon values.

In order to obtain the best fit, we considered two slightly different procedures. In panels (a)-(b) of each figure (dashed lines), we fixed TcT_{c} to be the temperature at which the third-harmonic response displays a maximum; we refer to this value as Tc(exp)T_{c}^{(\mathrm{exp})}. It is important to note, however, that this value is not necessarily the exact temperature of zero resistance onset. For this reason, and given the intrinsic experimental uncertainties in the precise absolute determination of TcT_{c}, in panels (c)-(d) (solid lines) we allowed TcT_{c} to vary from Tc(exp)T_{c}^{(\mathrm{exp})}, but by no more than 0.5%0.5\%. The fit parameters are shown in Table 1, together with the experimental values for Tc(exp)T_{c}^{(\mathrm{exp})} and Hc2(exp)H_{c2}^{(\mathrm{exp})}, the latter taken from Ref. (Lide, 2004). Note that, to distinguish between the two fitting procedures, we denote by H~c2\tilde{H}_{c2} the value used in panels (a)-(b) of the figures. Moreover, since Pb is a type-I superconductor, Hc2(exp)H_{c2}^{(\mathrm{exp})} was estimated through 2κHc\sqrt{2}\kappa H_{c} (Tinkham, 2004), with κ=0.24\kappa=0.24 (Smith, 1969; Farrell et al., 1969) and Hc=803H_{c}=803 Oe (Smith, 1969; Martienssen and Warlimont, 2006).

Panels (a)-(b) of Figs. 4, 5, and 6 show that the theoretical curves obtained by fixing Tc=Tc(exp)T_{c}=T_{c}^{(\mathrm{exp})} provide a reasonable description of the third-harmonic data in the region not too close to TcT_{c} for Pb and V (Figs. 4 and 6), and in the region close to TcT_{c} for Nb (Fig. 5). In particular, the latter does not seem to display the characteristic ϵ5/2\epsilon^{-5/2} power-law behavior observed in the former two in the regime of intermediate ϵ\epsilon values. However, because of the definition of the reduced temperature, ϵ=TTcTc\epsilon=\frac{T-T_{c}}{T_{c}}, even small changes in TcT_{c} within typical experimental uncertainty could account for these deviations between theory and experiment. As noted above, to address this issue we performed a second fit procedure allowing TcT_{c} to be slightly different than Tc(exp)T_{c}^{(\mathrm{exp})}. As shown in panels (c)-(d) of the same figures, we find a better agreement between the theoretical and experimental results over a wider temperature range, including in the case of Nb in the intermediate ϵ\epsilon range. Comparing the theoretical TcT_{c} values in Table 1 with the Tc(exp)T_{c}^{(\mathrm{exp})} values, we note that in all cases TcT_{c} is slightly larger than Tc(exp)T_{c}^{(\mathrm{exp})}. This is the reason why in panels (c)-(d) the theoretical curves stop at ϵ=0\epsilon=0 whereas the data extend to the region ϵ<0\epsilon<0.

On the other hand, there is a more significant difference between Hc2H_{c2} and the experimental value Hc2(exp)H_{c2}^{(\mathrm{exp})} taken from the literature, with the former being a factor of approximately 22 to 66 smaller or larger than the latter. We note that the intrinsic uncertainty in the precise value of H0H_{0} in our experiment may explain at least part of this discrepancy. Moreover, the value of Hc2(exp)H_{c2}^{(\mathrm{exp})} strongly depends on material preparation details, especially for polycrystalline samples where significant internal strains can be present (Van Gurp, 1967). In principle, the critical fields are lower in more pristine materials, and it is therefore meaningful to take the lowest known experimental values (taken from Ref. (Lide, 2004)) for our comparison. Finally, while the LD model employed here to calculate |M3¯|\left|\overline{M_{3}}\right| assumes a layered system, the bulk elemental superconductors are cubic. On top of that, the LD approach of including only Gaussian fluctuations is expected to break down below a very small ϵcrit\epsilon_{\mathrm{crit}}, whose precise value is likely different for distinct materials. Despite these drawbacks, this comparison shows that the LD model for the third-harmonic response |M3¯|\left|\overline{M_{3}}\right| due to contributions from superconducting fluctuations provides a satisfactory description of the experimental results.

Tc(exp)T_{c}^{(\mathrm{exp})}(K) Hc2(exp)H_{c2}^{(\mathrm{exp})}(G) H~c2\tilde{H}_{c2} (G) Hc2H_{c2} (G) Tc(exp)/TcT_{c}^{(\mathrm{exp})}/T_{c}
Pb 7.18 273 2170 1083 0.9996
Nb 9.31 1710 166 286 0.9955
V 5.29 1200 1300 520 0.9980
Table 1: Experimental critical temperature and critical field values, Tc(exp)T_{c}^{(\mathrm{exp})} and Hc2(exp)H_{c2}^{(\mathrm{exp})}, compared to the theoretical fitting parameters TcT_{c}, H~c2\tilde{H}_{c2} and Hc2H_{c2}. H~c2\tilde{H}_{c2} corresponds to the fits in panels (a)-(b) of Figs. 4, 5, and 6, where TcT_{c} is forced to be equal to the temperature where the experimental third-harmonic response displays a maximum (denoted here by Tc(exp)T_{c}^{(\mathrm{exp})}). On the other hand, Hc2H_{c2} corresponds to the fits in panels (c)-(d) of the same figures, where TcT_{c} is allowed to be different from the experimental value. The Hc2(exp)H_{c2}^{(\mathrm{exp})} values for Nb and V are the smallest ones reported in Ref. (Lide, 2004), whereas Hc2(exp)H_{c2}^{(\mathrm{exp})} for Pb was estimated as explained in the text.

III.2 Strontium Ruthenate (Sr2RuO4)

Having validated our theoretical approach to compute the third-harmonic response |M3¯|\left|\overline{M_{3}}\right| by comparison with data for elemental superconductors, we now perform the same comparison with the lamellar perovskite-derived superconductor Sr2RuO4 (SRO). The main advantage of our LD calculation of |M3¯|\left|\overline{M_{3}}\right| is that it is entirely phenomenological and independent of microscopic details. In fact, the main assumption is that the superconducting fluctuations can be described by a Gaussian approximation. Consequently, the calculation could in principle be applicable to unconventional superconductors as well.

SRO is believed to host an unconventional superconducting state that breaks time-reversal symmetry (Luke et al., 1998; Xia et al., 2006; Grinenko et al., 2021). Whereas for a long time SRO was considered a promising candidate for pp-wave triplet superconductivity (Mackenzie and Maeno, 2003; Kallin, 2012), recent experiments have revealed problems with this interpretation (Mackenzie et al., 2017; Pustogow et al., 2019; Chronister et al., 2020). This has motivated alternative proposals involving e.g. dd-wave and gg-wave superconductivity (Røising et al., 2019; Ramires and Sigrist, 2019; Rømer et al., 2019; Suh et al., 2020; Kivelson et al., 2020; Willa et al., 2020). As mentioned above, the data presented here are the same as in Ref. (Pelc et al., 2019). As shown there, the third-harmonic response of other perovskite-based superconductors like strontium titanate and the cuprates display a similar unusual temperature dependence.

The data for SRO are shown by the orange symbols in Fig. 7 on linear scale (panel (a)), logarithmic scale (panel (b)), and semi-logarithmic scale (panel (c)). The theoretical results for |M3¯|\left|\overline{M_{3}}\right| are plotted in the same panels using the experimental critical temperature value, Tc=1.51K=Tc(exp)T_{c}=1.51\>\mathrm{K}=T_{c}^{(\mathrm{exp})}, and two different critical field values: Hc2=750G=Hc2(exp)H_{c2}=750\>\mathrm{G}=H_{c2}^{(\mathrm{exp})} (dashed lines) and Hc2=7.6G0.01Hc2(exp)H_{c2}=7.6\>\mathrm{G}\approx 0.01H_{c2}^{(\mathrm{exp})} (dotted lines). Here, Tc(exp)T_{c}^{(\mathrm{exp})} corresponds to the temperature at which the third-harmonic response is maximum, and Hc2(exp)H_{c2}^{(\mathrm{exp})} is the experimental value reported in the literature (Mackenzie and Maeno, 2003; Kittaka et al., 2009). The key observation is that the theoretical |M3¯|\left|\overline{M_{3}}\right| curve with Hc2=Hc2(exp)H_{c2}=H_{c2}^{(\mathrm{exp})} grossly underestimates the data. It is necessary to reduce Hc2H_{c2} by two orders of magnitude to obtain values that are comparable between theory and experiment. In contrast, for the elemental superconductors, the difference in the theoretical and experimental Hc2H_{c2} values was at most a factor of 66. More importantly, even by changing Hc2H_{c2} by such a large amount, the temperature dependence of the data is not captured by the theoretical |M3¯|\left|\overline{M_{3}}\right| curve, in contrast again to the case of conventional superconductors. Indeed, while the theoretical |M3¯|\left|\overline{M_{3}}\right| curve shows a power-law for intermediate reduced temperatures, the data display an accurately exponential temperature dependence, as discussed in Ref. (Pelc et al., 2019) and shown in panel (c) of Fig. 7. We note that the experimental Hc2H_{c2} value depends very strongly on the orientation of the field with respect to the crystalline cc-axis, such that a small misalignment can lead to sizable variation (Kittaka et al., 2009). However, the discrepancy between the theoretical and experimental results cannot be explained by sample misalignment, since the critical field increases with increasing angle between the field direction and the crystalline cc-axis, whereas our theoretical results require smaller Hc2H_{c2} values.

Refer to caption
Figure 7: Comparison between the experimentally measured third-harmonic response |M3¯|\left|\overline{M_{3}}\right| (orange symbols, in arbitrary units) for SRO and the theoretical results obtained from Eq. (17) (dashed and dotted red lines). Panels (a), (b), and (c) show the data on linear, logarithmic, and semi-logarithmic scale, respectively. For the theoretical curves, the critical temperature is set to its experimental value Tc(exp)T_{c}^{(\mathrm{exp})} whereas the critical field is set to Hc2(exp)H_{c2}^{(\mathrm{exp})} (dashed lines) and to 0.01Hc2(exp)0.01H_{c2}^{(\mathrm{exp})} (dotted lines). The anisotropy parameter is set to r=10r=10.
Refer to caption
Figure 8: Comparison between the normalized third-harmonic response data and the theoretical |M3¯|\left|\overline{M_{3}}\right| results for Pb, Nb, V and SRO on a logarithmic scale. The solid lines correspond to the best fits in Figs. 4, 5, and 6, which refer to the conventional superconductors, whereas the dashed and dotted lines correspond to the fits for SRO in Fig. 7.

Fig. 8 summarizes the third-harmonic response |M3¯|\left|\overline{M_{3}}\right| of the three conventional superconductors studied here (Pb, Nb, V), as well as of the unconventional superconductor SRO. The differences between SRO and the conventional superconductors are not only on the temperature dependence of |M3¯|\left|\overline{M_{3}}\right|, but also on the fact that |M3¯|\left|\overline{M_{3}}\right| is larger and extends over a much wider relative temperature range in SRO. Indeed, while superconducting fluctuations are detected up to ϵ102\epsilon\sim 10^{-2} in conventional superconductors, they extend all the way up to ϵ1\epsilon\sim 1 in SRO.

To attempt to address the discrepancy between the theoretical and experimental results for SRO, we revisit the assumptions behind the LD model, from which we derived the expression for |M3¯|\left|\overline{M_{3}}\right|. As discussed above, the LD model makes no reference to the microscopic pairing mechanism. However, it does assume a homogeneous system. In contrast, perovskites are known for their intrinsic inhomogeneity, arising from e.g. oxygen vacancies and local structural distortions that deviate strongly from the average lattice structure (see (Pelc et al., 2021) and references therein). Indeed, the experiments of Ref. (Pelc et al., 2019) indicate that universal structural inhomogeneity is present in perovskite-based superconductors such as SRO. It has also been argued that dislocations can have a strong impact on the superconducting state properties of several perovskites (Ying et al., 2013; Hameed et al., 2020; Willa et al., 2020). In the particular case of SRO, muon spin-rotation measurements find a rather inhomogeneous signature of time-reversal symmetry-breaking below TcT_{c} (Grinenko et al., 2021). It is also known that the TcT_{c} of SRO is strongly dependent on stress (Steppke et al., 2017; Grinenko et al., 2021), implying that inhomogeneous internal stresses would lead to regions with locally modified TcT_{c}. Simple point disorder also leads to a variation of the local critical temperature (Mackenzie et al., 1998). Indeed, scanning SQUID measurements have directly detected TcT_{c} inhomogeneity on the micron scale (Watson et al., 2018).

Refer to caption
Figure 9: (a) Normalized probability distribution function of the critical temperature tct_{c} for different values of the parameter σ\sigma in Eq. (20). Here, the parameter μ\mu is fixed by the condition vF(Tc(exp))=0.3v_{F}\left(T_{c}^{(\mathrm{exp})}\right)=0.3, with Tc(exp)=1.51T_{c}^{(\mathrm{exp})}=1.51 K (indicated by the dashed gray vertical line) and the temperature-dependent superconducting volume fraction vFv_{F} defined by Eq. (22). (b) Averaged third-harmonic response |M3¯|\left\langle\left|\overline{M_{3}}\right|\right\rangle calculated from the distribution functions of panel (a), compared to the data for SRO, as a function of ϵ=TTc1\epsilon=\frac{T}{T_{c}}-1. In this calculation, we used the experimental values Tc(exp)=1.51KT_{c}^{(\mathrm{exp})}=1.51\>\mathrm{K} and Hc2(exp)=750GH_{c2}^{(\mathrm{exp})}=750\>\mathrm{G}, and set r=10r=10.

The impact of inhomogeneity on superconducting properties has been studied by a variety of approaches (Coleman et al., 1995; Andersen et al., 2006; Swanson et al., 2014; Pelc et al., 2018; Dodaro and Kivelson, 2018). Here, we consider a phenomenological approach that introduces a probability distribution of the local TcT_{c} (see also Ref. (Mayoh and García-García, 2015)). Such an inhomogeneous TcT_{c} distribution may explain why the superconducting fluctuations in SRO are stronger and extend to higher reduced temperatures as compared to conventional superconductors, since regions with a locally higher TcT_{c} are expected to result in a much larger contribution to |M3¯|\left|\overline{M_{3}}\right| than that arising from the rest of the sample. To test this idea, we include a distribution function for TcT_{c} into our LD-based phenomenological model. We denote the “transition temperature variable” as tct_{c}, and reserve the notation TcT_{c} for the actual transition temperature of the system to avoid confusion. The form of the distribution function P(tc)P(t_{c}) depends on several sources of inhomogeneity in the system, see for instance Ref. (Dodaro and Kivelson, 2018). A microscopic derivation is thus very challenging, and beyond the scope of this work. Instead, here we opt for a simple phenomenological modeling of P(tc)P(t_{c}). In particular, we employ a normalized log-normal distribution:

P(tc)=1tc2πσ2exp[(lntcμ)22σ2]P\left(t_{c}\right)=\frac{1}{t_{c}\sqrt{2\pi\sigma^{2}}}\exp\left[-\frac{\left(\ln\frac{t_{c}}{\mu}\right)^{2}}{2\sigma^{2}}\right] (20)

where μ\mu and σ\sigma are positive parameters that determine the mean value and variance of the distribution. The choice of this distribution is motivated by its properties of only allowing non-zero values of tct_{c} and of having long tails toward larger values of tct_{c}. We note that a log-normal distribution for the local gap – and consequently of the local TcT_{c} – was previously derived theoretically in Ref. (Mayoh and García-García, 2015) for disordered quasi-two-dimensional superconductors in the limit of weak multifractality, and observed experimentally in weakly disordered monolayer NbSe2 (Rubio-Verdú et al., 2020). The averaged fluctuation magnetization in Eq.(4) acquires the following form:

M(ϵ)=0Tdtctc2πσ2exp[(lntcμ)22σ2]M(Ttc1)\left\langle M\right\rangle\left(\epsilon\right)=\int_{0}^{T}\,\frac{dt_{c}}{t_{c}\sqrt{2\pi\sigma^{2}}}\exp\left[-\frac{\left(\ln\frac{t_{c}}{\mu}\right)^{2}}{2\sigma^{2}}\right]M\left(\frac{T}{t_{c}}-1\right) (21)

with M(ϵ)M\left(\epsilon\right) given by Eq. (4). We can then compute the averaged third-harmonic response |M3¯|\left\langle\left|\overline{M_{3}}\right|\right\rangle from Eq. (17). We assume that |M3¯|\left\langle\left|\overline{M_{3}}\right|\right\rangle is dominated by superconducting fluctuations contributions, which appear only in regions that are locally non-superconducting (i.e. for which ϵ=Ttc1\epsilon=\frac{T}{t_{c}}-1 is positive). For this reason, the limits of the tct_{c} integration are such that 0<tc<T0<t_{c}<T.

Refer to caption
Figure 10: Averaged third-harmonic response |M3¯|\left\langle\left|\overline{M_{3}}\right|\right\rangle as a function of the reduced temperature ϵ=TTc1\epsilon=\frac{T}{T_{c}}-1 calculated using the parameters Tc=1.41K0.93Tc(exp)T_{c}=1.41\>\mathrm{K}\approx 0.93T_{c}^{(\mathrm{exp})} and σ=0.08\sigma=0.08, while keeping Hc2=Hc2(exp)=750GH_{c2}=H_{c2}^{(\mathrm{exp})}=750\>\mathrm{G} and r=10r=10 (solid red line). The orange symbols are the experimental results, and the dashed red line reproduces the theoretical third-harmonic response |M3¯|\left|\overline{M_{3}}\right| of the clean system with Tc=Tc(exp)T_{c}=T_{c}^{(\mathrm{exp})} and Hc2=Hc2(exp)H_{c2}=H_{c2}^{(\mathrm{exp})}.

The two parameters characterizing the distribution function, μ\mu and σ\sigma, are not independent, since they are related by the value of TcT_{c}. To see that, we first define the temperature-dependent superconducting volume fraction vF(T)v_{F}\left(T\right), which is given by

vF(T)=10TP(tc)𝑑tc=1212erf(lntcμ2σ),v_{F}\left(T\right)=1-\int_{0}^{T}{P\left(t_{c}\right)dt_{c}}=\frac{1}{2}-\frac{1}{2}\mathrm{erf}\left(\frac{\ln\frac{t_{c}}{\mu}}{\sqrt{2}\sigma}\right), (22)

since the integral on the right-hand side gives the non-superconducting volume fraction (T>tcT>t_{c}). When the volume fraction becomes larger than a threshold value vFv_{F}^{*}, the local superconducting regions are expected to percolate and the whole sample becomes superconducting. Note that a similar criterion was used in the analysis of Ref. (Mayoh and García-García, 2015). TcT_{c} is then obtained by solving the equation vF(Tc)=vFv_{F}\left(T_{c}\right)=v_{F}^{*},

μTc=exp[2σerf1(12vF)],\frac{\mu}{T_{c}}=\exp\left[-\sqrt{2}\sigma\mathrm{erf}^{-1}\left(1-2v_{F}^{*}\right)\right], (23)

where erf1(x)\mathrm{erf}^{-1}(x) is the inverse error function. For simplicity, we use for vFv_{F}^{*} the site percolation threshold value for a cubic lattice, vF=0.3v_{F}^{*}=0.3. While vFv_{F}^{*} itself could be considered a free parameter, we opt to fix it to avoid increasing the number of fitting parameters. As a result, the only additional parameter needed to compute |M3¯|\left\langle\left|\overline{M_{3}}\right|\right\rangle, as compared to the “clean” system |M3¯|\left|\overline{M_{3}}\right|, is the dimensionless σ\sigma, which determines the width of the distribution. In Fig. 9(a), we illustrate the profile of P(tc)P\left(t_{c}\right) for different values of σ\sigma under the constraint vF(Tc(exp))=0.3v_{F}\left(T_{c}^{(\mathrm{exp})}\right)=0.3. The full expression for |M3¯|\left\langle\left|\overline{M_{3}}\right|\right\rangle then becomes:

|M3¯|(ϵ)M=24(ϵ+1)πln201dxx2πσ2exp{[ln(xϵ+x)2σ+erf1(12vF)]2}π/2π/2(x,h0cosθ)cos3θdθ\frac{\left\langle\left|\overline{M_{3}}\right|\right\rangle\left(\epsilon\right)}{M_{\infty}}=\frac{24\left(\epsilon+1\right)}{\pi\ln 2}\int_{0}^{1}\frac{dx}{x\sqrt{2\pi\sigma^{2}}}\exp\left\{-\left[\frac{\ln\left(x\epsilon+x\right)}{\sqrt{2}\sigma}+\mathrm{erf}^{-1}\left(1-2v_{F}^{*}\right)\right]^{2}\right\}\int_{-\pi/2}^{\pi/2}\mathcal{M}\left(x,h_{0}\cos\theta\right)\cos 3\theta\,d\theta (24)

with:

(x,h)=0π2𝑑ϕ{1x1+rsin2ϕ2h[ψ(1x1+rsin2ϕ2h+12)1]lnΓ(1x1+rsin2ϕ2h+12)+12ln(2π)}\mathcal{M}\left(x,h\right)=-\int_{0}^{\frac{\pi}{2}}d\phi\left\{\frac{\frac{1}{x}-1+r\sin^{2}{\phi}}{2h}\left[\psi\left(\frac{\frac{1}{x}-1+r\sin^{2}{\phi}}{2h}+\frac{1}{2}\right)-1\right]-\ln{\Gamma}\left(\frac{\frac{1}{x}-1+r\sin^{2}{\phi}}{2h}+\frac{1}{2}\right)+\frac{1}{2}\ln{\left(2\pi\right)}\right\} (25)

Using the distribution functions of Fig. 9(a), in Fig. 9(b) we present the calculated averaged third-harmonic response |M3¯|\left\langle\left|\overline{M_{3}}\right|\right\rangle (solid red line) using the experimentally determined values for TcT_{c} and Hc2H_{c2}. The comparison with the data shows that even a relatively mild width of the distribution of tct_{c} values, with σ0.1\sigma\apprle 0.1, is capable of capturing the extended temperature window for which the third-harmonic response is sizable. As anticipated, this behavior is a consequence of the fact that regions with a local higher TcT_{c} value, although occupying a small volume, provide a sizable contribution to the third-harmonic response.

The temperature dependence of the third-harmonic response data, however, is not very well captured by the theoretical curves in Fig. 9(b). To try to address this issue, we promote TcT_{c} to a free parameter and allow it to deviate slightly from the experimental value Tc(exp)=1.51T_{c}^{(\mathrm{exp})}=1.51 K. Fig. 10 shows the results for |M3¯|\left\langle\left|\overline{M_{3}}\right|\right\rangle in the case of Tc=1.41K0.93Tc(exp)T_{c}=1.41\>\mathrm{K}\approx 0.93T_{c}^{(\mathrm{exp})} and σ=0.08\sigma=0.08. Clearly, the temperature dependence of the calculated |M3¯|\left\langle\left|\overline{M_{3}}\right|\right\rangle becomes more similar to the experimentally measured one, but still fails to capture it completely. Thus, our conclusion is that while TcT_{c} inhomogeneity may explain the extended temperature range where the third-harmonic response is sizable, it is unlikely to explain the exponential tail of |M3¯|\left|\overline{M_{3}}\right| observed experimentally in Ref. (Pelc et al., 2019).

IV Concluding remarks

In this work, we used the LD model to compute the third-harmonic magnetic response |M3¯|\left|\overline{M_{3}}\right| due to Gaussian superconducting fluctuations. Due to its phenomenological nature, the LD model could in principle be applicable to both conventional and unconventional superconductors. Our detailed comparison with measurements of |M3¯|\left|\overline{M_{3}}\right| found that the theoretical modeling provides a good description of the data in the case of Pb, Nb, and V – provided that the critical field is properly modified from its experimental value – but a rather poor account of the data for SRO. Inclusion of TcT_{c} inhomogeneity, which is intrinsically present in SRO, improved significantly the agreement between theoretical model and experimental data, although the model could not properly capture the experimentally observed exponential temperature dependence of |M3¯|\left|\overline{M_{3}}\right| (see Ref. (Pelc et al., 2019)).

Further investigation is thus required to elucidate the origin of this exponential behavior of |M3¯|\left|\overline{M_{3}}\right|, which was also seen in other perovskite superconductors such as STO and the cuprates, and appears to be quite robust (Pelc et al., 2019). One cannot completely discard simple TcT_{c} inhomogeneity as the source of this effect, since here we only focused on a very specific and particularly simple distribution function for TcT_{c}. While this choice allowed us to argue on a more quantitative basis that TcT_{c} inhomogeneity can explain why |M3¯|\left|\overline{M_{3}}\right| remains large over a wide temperature window in SRO, the actual TcT_{c} distribution is certainly more complicated and likely material-dependent. A phenomenological TcT_{c} distribution will likely require fine tuning to give an exponential temperature dependence of the third-harmonic response. Nevertheless, if rare regions are present, they might give rise to specific tails in the distribution function that may be common to different materials; these types of effects have been explored in more detail in Refs. (Dodaro and Kivelson, 2018; Pelc et al., 2021). We also note that, in the particular case of the cuprates, an exponential temperature-dependent behavior associated with superconducting fluctuations was also observed in other observables such as linear/nonlinear conductivity and specific heat, and described in terms of a Gaussian TcT_{c} distribution (Popčević et al., 2018; Pelc et al., 2018). It would be interesting to investigate whether the exponential temperature dependence observed in the third-harmonic response of SRO is also manifested in these other observables in the case of SRO. In fact, as shown in Ref. (Pelc et al., 2019), prior specific heat data (Nishizaki et al., 1999) are consistent with this possibility.

Besides inhomogeneity, another effect, more specific to SRO, is that if it is indeed a two-component superconductor, as proposed by different models (Rømer et al., 2019; Suh et al., 2020; Kivelson et al., 2020; Willa et al., 2020), the superconducting fluctuation spectrum will likely be more complicated than that of the LD model. However, the fact that the same exponential temperature-dependence of |M3¯|\left|\overline{M_{3}}\right| is seen in STO and cuprates, the latter being single-component superconductors, renders this scenario less likely. Finally, a crucial approximation of the LD model is that it solely focuses on Gaussian superconducting fluctuations. This raises the interesting question of whether non-Gaussian fluctuations may also play an important role in the fluctuation spectra of perovskite superconductors.

Acknowledgements.
We thank Z. Anderson and S. Griffitt for assistance in ac susceptibility probe design and construction, and A. Mackenzie and C. Hicks for providing the SRO samples. This work was supported by the U. S. Department of Energy through the University of Minnesota Center for Quantum Materials, under Award No. DE-SC-0016371.

References

  • Varlamov et al. (2018) A. A. Varlamov, A. Galda, and A. Glatz, Rev. Mod. Phys. 90, 015009 (2018).
  • Larkin and Varlamov (2005) A. Larkin and A. Varlamov, Theory of fluctuations in superconductors (Clarendon Press, 2005).
  • Pelc et al. (2019) D. Pelc, Z. Anderson, B. Yu, C. Leighton, and M. Greven, Nature Communications 10, 2729 (2019).
  • Suzuki and Tsuboi (1977) T. Suzuki and T. Tsuboi, Journal of the Physical Society of Japan 43, 444 (1977).
  • Tsuboi and Suzuki (1977) T. Tsuboi and T. Suzuki, Journal of the Physical Society of Japan 42, 437 (1977).
  • Tallon et al. (2011) J. L. Tallon, J. G. Storey, and J. W. Loram, Phys. Rev. B 83, 092502 (2011).
  • Glover (1967) R. Glover, Physics Letters A 25, 542 (1967).
  • Strongin et al. (1968) M. Strongin, O. F. Kammerer, J. Crow, R. S. Thompson, and H. L. Fine, Phys. Rev. Lett. 20, 922 (1968).
  • Ruggiero et al. (1980) S. T. Ruggiero, T. W. Barbee, and M. R. Beasley, Phys. Rev. Lett. 45, 1299 (1980).
  • Mircea et al. (2009) D. I. Mircea, H. Xu, and S. M. Anlage, Phys. Rev. B 80, 144505 (2009).
  • Rullier-Albenque et al. (2011) F. Rullier-Albenque, H. Alloul, and G. Rikken, Phys. Rev. B 84, 014522 (2011).
  • Carbillet et al. (2016) C. Carbillet, S. Caprara, M. Grilli, C. Brun, T. Cren, F. Debontridder, B. Vignolle, W. Tabis, D. Demaille, L. Largeau, et al., Phys. Rev. B 93, 144509 (2016).
  • Popčević et al. (2018) P. Popčević, D. Pelc, Y. Tang, K. Velebit, Z. W. Anderson, V. Nagarajan, G. Yu, M. Požek, N. Barišić, and M. Greven, npj Quant. Mat. 3, 42 (2018).
  • Pelc et al. (2018) D. Pelc, M. Vučković, M. S. Grbić, M. Požek, G. Yu, T. Sasagawa, M. Greven, and N. Barišić, Nature Communications 9, 4327 (2018).
  • Corson et al. (1999) R. Corson, L. Mallozzi, J. Orenstein, J. N. Eckstein, and Božović, Nature 398, 221 (1999).
  • Orenstein et al. (2006) J. Orenstein, J. Corson, S. Oh, and J. N. Eckstein, Ann. Phys. 15, 596 (2006).
  • Grbić et al. (2011) M. S. Grbić, M. Požek, D. Paar, V. Hinkov, M. Raichle, D. Haug, B. Keimer, N. Barišić, and A. Dulčić, Phys. Rev. B 83, 144508 (2011).
  • Bilbro et al. (2011) L. S. Bilbro, L. Valdes Aguilar, G. Logvenov, O. Pelleg, I. Božović, and N. P. Armitage, Nat. Phys. 7, 298 (2011).
  • Geballe et al. (1971) T. H. Geballe, A. Menth, F. J. Di Salvo, and F. R. Gamble, Phys. Rev. Lett. 27, 314 (1971).
  • Gollub et al. (1973) J. P. Gollub, M. R. Beasley, R. Callarotti, and M. Tinkham, Phys. Rev. B 7, 3039 (1973).
  • Li et al. (2010) L. Li, Y. Wang, S. Komiya, S. Ono, Y. Ando, G. D. Gu, and N. P. Ong, Phys. Rev. B 81, 054510 (2010).
  • Kokanović et al. (2013) I. Kokanović, D. J. Hills, M. L. Sutherland, R. Liang, and J. R. Cooper, Phys. Rev. B 88, 060505 (2013).
  • Kasahara et al. (2016) S. Kasahara, T. Yamashita, A. Shi, R. Kobayashi, Y. Shimoyama, T. Watashige, K. Ishida, T. Terashima, T. Wolf, F. Hardy, et al., Nat. Commun. 7, 12843 (2016).
  • Yu et al. (2019) G. Yu, D.-D. Xia, D. Pelc, R.-H. He, N.-H. Kaneko, T. Sasagawa, Y. Li, X. Zhao, N. Barišić, A. Shekhter, et al., Phys. Rev. B 99, 214502 (2019).
  • Wang et al. (2006) Y. Wang, L. Li, and N. P. Ong, Phys. Rev. B 73, 024510 (2006).
  • Rullier-Albenque et al. (2006) F. Rullier-Albenque, R. Tourbot, H. Alloul, P. Lejay, D. Colson, and A. Forget, Phys. Rev. Lett. 96, 067002 (2006).
  • Chang et al. (2012) J. Chang, N. Doiron-Leyraud, O. Cyr-Choiniere, G. Grissonnanche, F. Laliberté, E. Hassinger, J.-P. Reid, R. Daou, S. Pyon, T. Takayama, et al., Nature Physics 8, 751 (2012).
  • Tafti et al. (2014) F. F. Tafti, F. Laliberté, M. Dion, J. Gaudet, P. Fournier, and L. Taillefer, Phys. Rev. B 90, 024519 (2014).
  • Behnia and Aubin (2016) K. Behnia and H. Aubin, Reports on Progress in Physics 79, 046502 (2016).
  • Schmidt (1968) H. Schmidt, Zeitschrift für Physik A 216, 336 (1968).
  • Shmidt (1968) V. V. Shmidt, JETP 27, 142 (1968).
  • Maki (1968) K. Maki, Progress of Theoretical Physics 39, 897 (1968).
  • Thompson (1970) R. S. Thompson, Phys. Rev. B 1, 327 (1970).
  • Schmid (1969) A. Schmid, Phys. Rev. 180, 527 (1969).
  • Prange (1970) R. E. Prange, Phys. Rev. B 1, 2349 (1970).
  • Abrahams et al. (1970) E. Abrahams, M. Redi, and J. W. F. Woo, Phys. Rev. B 1, 208 (1970).
  • Kurkijärvi et al. (1972) J. Kurkijärvi, V. Ambegaokar, and G. Eilenberger, Phys. Rev. B 5, 868 (1972).
  • Aslamazov and Larkin (1974) L. G. Aslamazov and A. I. Larkin, Sov. Phys. JETP 40, 321 (1974).
  • Emery and Kivelson (1995) V. Emery and S. Kivelson, Nature 374, 434 (1995).
  • Glatz et al. (2011) A. Glatz, A. A. Varlamov, and V. M. Vinokur, Phys. Rev. B 84, 104510 (2011).
  • Ullah and Dorsey (1991) S. Ullah and A. T. Dorsey, Phys. Rev. B 44, 262 (1991).
  • Fisher et al. (1991) D. S. Fisher, M. P. A. Fisher, and D. A. Huse, Phys. Rev. B 43, 130 (1991).
  • Ioffe et al. (1993) L. B. Ioffe, A. I. Larkin, A. A. Varlamov, and L. Yu, Phys. Rev. B 47, 8936 (1993).
  • Ussishkin et al. (2002) I. Ussishkin, S. L. Sondhi, and D. A. Huse, Phys. Rev. Lett. 89, 287001 (2002).
  • Serbyn et al. (2009) M. N. Serbyn, M. A. Skvortsov, A. A. Varlamov, and V. Galitski, Phys. Rev. Lett. 102, 067001 (2009).
  • Michaeli and Finkel'stein (2009) K. Michaeli and A. M. Finkel'stein, EPL (Europhysics Letters) 86, 27007 (2009).
  • Li and Levchenko (2020) S. Li and A. Levchenko, Annals of Physics 417, 168137 (2020).
  • Pelc et al. (2021) D. Pelc, R. J. Spieker, Z. W. Anderson, M. J. Krogstad, N. Biniskos, N. G. Bielinski, B. Yu, T. Sasagawa, L. Chauviere, P. Dosanjh, et al., arxiv:2103.05482 (2021).
  • Dodaro and Kivelson (2018) J. F. Dodaro and S. A. Kivelson, Phys. Rev. B 98, 174503 (2018).
  • Mishonov and Penev (2000) T. Mishonov and E. Penev, International Journal of Modern Physics B 14, 3831 (2000).
  • Tsuzuki and Koyanagi (1969) T. Tsuzuki and M. Koyanagi, Physics Letters A 30, 545 (1969).
  • Drobac et al. (2013) D. Drobac, Z. Marohnic, I. Zivkovic, and M. Prester, Review of Scientific Instruments 84, 054708 (2013).
  • Lide (2004) D. R. Lide, CRC Handbook of Chemistry and Physics, vol. 84 (CRC press, 2004).
  • Tinkham (2004) M. Tinkham, Introduction to superconductivity (Courier Corporation, 2004).
  • Smith (1969) F. W. Smith, Ph. D. Thesis (1969).
  • Farrell et al. (1969) D. E. Farrell, B. S. Chandrasekhar, and H. V. Culbert, Phys. Rev. 177, 694 (1969).
  • Martienssen and Warlimont (2006) W. Martienssen and H. Warlimont, Springer handbook of condensed matter and materials data (Springer Science & Business Media, 2006).
  • Van Gurp (1967) G. J. Van Gurp, Philips Res. Rept. 22, 10 (1967).
  • Luke et al. (1998) G. M. Luke, Y. Fudamoto, K. Kojima, M. Larkin, J. Merrin, B. Nachumi, Y. Uemura, Y. Maeno, Z. Mao, Y. Mori, et al., Nature 394, 558 (1998).
  • Xia et al. (2006) J. Xia, Y. Maeno, P. T. Beyersdorf, M. M. Fejer, and A. Kapitulnik, Phys. Rev. Lett. 97, 167002 (2006).
  • Grinenko et al. (2021) V. Grinenko, S. Ghosh, R. Sarkar, J.-C. Orain, A. Nikitin, M. Elender, D. Das, Z. Guguchia, F. Brückner, M. E. Barber, et al., Nature Physics (2021), URL https://doi.org/10.1038/s41567-021-01182-7.
  • Mackenzie and Maeno (2003) A. P. Mackenzie and Y. Maeno, Rev. Mod. Phys. 75, 657 (2003).
  • Kallin (2012) C. Kallin, Reports on Progress in Physics 75, 042501 (2012).
  • Mackenzie et al. (2017) A. P. Mackenzie, T. Scaffidi, C. W. Hicks, and Y. Maeno, npj Quantum Materials 2, 1 (2017).
  • Pustogow et al. (2019) A. Pustogow, Y. Luo, A. Chronister, Y.-S. Su, D. Sokolov, F. Jerzembeck, A. P. Mackenzie, C. W. Hicks, N. Kikugawa, S. Raghu, et al., Nature 574, 72 (2019).
  • Chronister et al. (2020) A. Chronister, A. Pustogow, N. Kikugawa, D. A. Sokolov, F. Jerzembeck, C. W. Hicks, A. P. Mackenzie, E. D. Bauer, and S. E. Brown, arXiv:2007.13730 (2020).
  • Røising et al. (2019) H. S. Røising, T. Scaffidi, F. Flicker, G. F. Lange, and S. H. Simon, Phys. Rev. Research 1, 033108 (2019).
  • Ramires and Sigrist (2019) A. Ramires and M. Sigrist, Phys. Rev. B 100, 104501 (2019).
  • Rømer et al. (2019) A. T. Rømer, D. D. Scherer, I. M. Eremin, P. J. Hirschfeld, and B. M. Andersen, Phys. Rev. Lett. 123, 247001 (2019).
  • Suh et al. (2020) H. G. Suh, H. Menke, P. M. R. Brydon, C. Timm, A. Ramires, and D. F. Agterberg, Phys. Rev. Research 2, 032023 (2020).
  • Kivelson et al. (2020) S. A. Kivelson, A. C. Yuan, B. Ramshaw, and R. Thomale, npj Quantum Materials 5, 1 (2020).
  • Willa et al. (2020) R. Willa, M. Hecker, R. M. Fernandes, and J. Schmalian, arXiv:2011.01941 (2020).
  • Kittaka et al. (2009) S. Kittaka, T. Nakamura, Y. Aono, S. Yonezawa, K. Ishida, and Y. Maeno, Phys. Rev. B 80, 174514 (2009).
  • Ying et al. (2013) Y. Ying, N. Staley, Y. Xin, K. Sun, X. Cai, D. Fobes, T. Liu, Z. Mao, and Y. Liu, Nature Comm. 4, 2596 (2013).
  • Hameed et al. (2020) S. Hameed, D. Pelc, Z. Anderson, R. Spieker, M. Lukas, Y. Liu, M. Krogstad, R. Osborn, C. Leighton, and M. Greven, arXiv:2005.00514 (2020).
  • Steppke et al. (2017) A. Steppke, L. Zaho, M. E. Barber, T. Scaffidi, F. Jerzembeck, H. Rosner, A. S. Gibbs, Y. Maeno, S. H. Simon, A. P. Mackenzie, et al., Science 355, eaaf9398 (2017).
  • Mackenzie et al. (1998) A. P. Mackenzie, R. K. W. Haselwimmer, A. W. Tyler, G. G. Lonzarich, Y. Mori, S. Nishizaki, and Y. Maeno, Phys. Rev. Lett. 80, 161 (1998).
  • Watson et al. (2018) C. A. Watson, A. S. Gibbs, A. P. Mackenzie, C. W. Hicks, and K. A. Moler, Phys. Rev. B 98, 094521 (2018).
  • Coleman et al. (1995) A. Coleman, E. Yukalova, and V. Yukalov, Physica C: Superconductivity 243, 76 (1995).
  • Andersen et al. (2006) B. M. Andersen, A. Melikyan, T. S. Nunner, and P. J. Hirschfeld, Phys. Rev. B 74, 060501 (2006).
  • Swanson et al. (2014) M. Swanson, Y. L. Loh, M. Randeria, and N. Trivedi, Phys. Rev. X 4, 021007 (2014).
  • Mayoh and García-García (2015) J. Mayoh and A. M. García-García, Phys. Rev. B 92, 174526 (2015).
  • Rubio-Verdú et al. (2020) C. Rubio-Verdú, A. M. García-García, H. Ryu, D.-J. Choi, J. Zaldívar, S. Tang, B. Fan, Z.-X. Shen, S.-K. Mo, J. I. Pascual, et al., Nano Letters 20, 5111 (2020).
  • Nishizaki et al. (1999) S. Nishizaki, Y. Maeno, and Z. Mao, Journal of Low Temperature Physics 117, 1581 (1999).