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

Corresponding Author: ]leishu@fudan.edu.cn

Superconducting Properties of La(2{}_{2}(Cu1-xNix)5As3O2: A μ\rm\muSR Study

Qiong Wu    Kaiwen Chen    Zihao Zhu    Cheng Tan    Yanxing Yang    Xin Li State Key Laboratory of Surface Physics, Department of Physics, Fudan University, Shanghai 200438, People’s Republic of China    Toni Shiroka Laboratory for Muon-Spin Spectroscopy, Paul Scherrer Institute, CH-5232 Villigen PSI, Switzerland Laboratorium für Festkörperphysik, ETH Zürich, CH-8093 Zürich, Switzerland    Xu Chen    Jiangang Guo    Xiaolong Chen Beijing National Laboratory for Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, P. O. Box 603, Beijing 100190, China    Lei Shu [ State Key Laboratory of Surface Physics, Department of Physics, Fudan University, Shanghai 200438, People’s Republic of China Collaborative Innovation Center of Advanced Microstructures, Nanjing 210093, People’s Republic of China Shanghai Research Center for Quantum Sciences, Shanghai 201315, People’s Republic of China
Abstract

We report the results of muon spin rotation and relaxation (μ\rm\muSR) measurements on the recently discovered layered Cu-based superconducting material La(2{}_{2}(Cu1-xNix)5As3O2 (x=x= 0.40, 0.45). Transverse-field μ\rm\muSR experiments on both samples show that the temperature dependence of superfluid density is best described by a two-band model. The absolute values of zero-temperature magnetic penetration depth λab(0)\lambda_{\rm ab}(0) were found to be 427(1.7) nm and 422(1.5) nm for x=x= 0.40 and 0.45, respectively. Both compounds are located between the unconventional and the standard BCS superconductors in the Uemura plot. No evidence of time-reversal symmetry (TRS) breaking in the superconducting state is suggested by zero-field μ\rm\muSR measurements.

I Introduction

The relation between magnetism and superconductivity is one of the most prominent issues in condensed matter physics [1, 2, 3, 4, 5, 6, 7, 8, 9, 10]. Since the discovery of cuprate high transition temperature (TcT_{\rm c}) superconductors, considerable efforts have been made to investigate the role of in-plane impurities in them. It is now well established that, in copper oxide superconductors, nonmagnetic Zn ions suppress TcT_{c} even more strongly than magnetic Ni ions [11, 12, 13, 14]. Such behavior is in sharp contrast to that of conventional BCS superconductors, in which magnetic impurities can act as pairing-breaking agents, rapidly suppressing superconductivity [15, 16]. Another interesting behavior of unconventional superconducting systems such as the heavy-fermion, high TcT_{c} cuprate, and iron-pnictide superconductors is the dome shape of the chemical doping dependence of TcT_{\rm c} [17, 18, 19, 20].

Recently, the first Cu-As superconductor was discovered in the layered La2Cu5As3O2 with Tc=0.63T_{\rm c}=0.63 K [21]. When Cu2+ is replaced by Ni2+, also in La(2{}_{2}(Cu1-xNix)5As3O2, the TcT_{\rm c} exhibits a dome-like structure. Remarkably, while the superconductivity in cuprate- and iron-based superconductors is completely suppressed when the substitution ratio of Cu or Fe exceeds 20% [22, 23], superconductivity in La(2{}_{2}(Cu1-xNix)5As3O2 persists until the substitution ratio exceeds 60% [21]. In this case, the robustness of superconductivity reveals the unexpected effect of impurities on inducing and enhancing superconductivity. Hence, La(2{}_{2}(Cu1-xNix)5As3O2  provides a broader platform for studying the doping effect in the superconducting phase diagram.

Specific heat measurements have revealed that the optimally doped La(2{}_{2}(Cu1-xNix)5As3O2 (x=x= 0.40) sample shows a sharp superconducting transition at Tc=T_{\rm c}= 2.05 K, with a dimensionless jump Ce/γsTc=1.42C_{\rm e}/\gamma_{\rm s}T_{\rm c}=1.42 [21], consistent with the BCS weak-coupling limit (1.43). In the superconducting state, the temperature dependence of the specific heat coefficient is described by a fully-gapped model Ce/TeΔ/kBTC_{\rm e}/T\propto e^{-\Delta/k_{\rm B}T} after subtracting the upturn of Ce/TC_{\rm e}/T below T<0.5T<0.5 K, which is attributed to the Schottky effect. These results suggest that La(2{}_{2}(Cu1-xNix)5As3O2 (x=x= 0.40) is a conventional BCS superconductor with a fully developed energy gap. However, the fit yields 2Δ/kBTc=2.582\Delta/k_{\rm B}T_{\rm c}=2.58 [21], much smaller than the BCS weak coupling limit.

μ\rm\muSR experiments have been widely utilized to probe superconductivity in type-II superconductors at the microscopic level [24], and they are free from the influence of the Schottky effect. Transverse-field (TF) μ\rm\muSR  measures the absolute value of the magnetic penetration depth λ\lambda, which is related to the density of superconducting carriers. The temperature dependence of λ\lambda is sensitive to the lowest-lying superconducting excitations and provides information on the symmetry of superconducting pairing [25, 26, 27, 28, 29, 30]. In addition, zero-field (ZF) μ\rm\muSR is a powerful method to detect small spontaneous internal magnetic fields due to the possible breaking of time-reversal symmetry (TRS) at the superconducting transition. These can be as small as 10 μ\muT, corresponding to about 10210^{-2} of Bohr magneton μB\rm{\mu_{B}} [31, 32, 33, 25, 34, 24].

Here, in order to study the doping effect on superconductivity, we perform μ\rm\muSR measurements on the polycrystalline samples of La(2{}_{2}(Cu1-xNix)5As3O2  for x=0.40x=0.40 (optimal doping) and x=0.45x=0.45 (overdoped). The temperature dependence of superfluid density determined from TF-μ\rm\muSR is best described by a two-band superconductivity model. dd-wave superconductivity possibly exists in one of the bands, with the fraction of dd-wave smaller in the overdoped sample than that in the optimally doped one. The superconducting energy gap is larger for optimal doping, suggesting that the coupling strength decreases with the increase of doping. Both compounds are located between the unconventional and the standard BCS superconductors in the Uemura plot. Meanwhile, no evidence of TRS breaking is suggested by ZF-μ\rm\muSR measurements.

II Experimental Details

Solid-state reactions were used to produce polycrystalline samples of La(2{}_{2}(Cu1-xNix)5As3O2 (xx = 0.40, 0.45) [21]. The performed X-ray diffraction (XRD) studies and the density functional theory (DFT) calculation shows the band structures of La2Cu5As3O2, indicating that this newly synthesized sample is a layered superconducting compound, with the superconducting atomic layers consisting of the [Cu5As3]2[\rm{Cu_{5}As_{3}}]^{2-} cage-like structure [21].

μ\rm\muSR experiments were carried out using nearly 100% spin-polarized positive muons (μ+\mu^{+}) on the M15 beam line at TRIUMF, Vancouver, Canada for x=0.40x=0.40, and the DOLLY spectrometer of the Sμ\muS muon source at Paul Scherrer Institute (PSI), Switzerland for x=0.45x=0.45, respectively. The samples were mounted on a silver sample holder at TRIUMF and a copper sample holder at PSI, respectively. Only a very limited amount of muons stopped in the extremely thin copper sample holder. In TF-μ\rm\muSR measurements, where the external field is applied perpendicular to the initial muon spin polarization, muons are implanted one at a time into a sample which is cooled (from above TcT_{c}) in an external magnetic field. Muon spins precesses around the local field at the implantation site, and the functional form of the muon spin polarization depends on the field distribution of the vortex state, including the magnetic penetration depth, the vortex core radius, and the structure of the flux-line lattice. For ZF-μ\rm\muSR measurements, the ambient magnetic field was actively compensated to better than 1 μ\muT. μ\rm\muSR data were analyzed using the musrfit software package [35].

III RESULTS

III.1 Transverse-field μ\rm\muSR experiments

The μ\rm\muSR asymmetry spectrum usually consists of a signal from muons that stop in the sample and a slowly relaxing background signal from muons that miss the sample for example, stop in the sample holder. Fig. 1 (a) (b) show the typical TF-μ\rm\muSR muon spin precession signals at an applied field of μ0H=\mu_{\rm{0}}H= 30 mT in the normal (red squares) and superconducting states (blue circles) for La(2{}_{2}(Cu1-xNix)5As3O2 (x=x= 0.40 and 0.45) after subtracting the background signal. The superconducting volume fractions are estimated to be 60%\% and 70%\% from the TF μ\rm\muSR asymmetry at long times for x = 0.40 and 0.45, respectively. Fig. 1 (c) (d) show the Fourier transformations (FFT) of the total TF-μ\rm\muSR asymmetry. It can be inferred from the figure that vortex states are constructed in both samples at low temperatures [36]. Fig. 1 (e) (f) show the details of the FFT spectra. As shown in the figure, the magnetic fields in the superconducting state and the normal state are relatively close to each other. This is common in anisotropic powder superconducting samples [37, 38].

Refer to caption
Figure 1: (a)-(b) Representative TF-μ\rm\muSR asymmetry spectra A(t)A(t) for La(2{}_{2}(Cu1-xNix)5As3O2 (x=x= 0.40 (panel (a)) and 0.45 (panel (b))) in the normal (red squares) and superconducting (blue circles) states with an external magnetic field of μ0H=\rm{\mu_{0}}H= 30 mT after subtracting the background signal. Solid curves: fits to the data using Eq. (1). (c)-(f) Fourier transformation of the total TF-μ\rm\muSR time spectra. (e)-(f) show the details of the FFT spectra.

The TF-μ\rm\muSR time spectra after subtracting the background signal can be well fitted by the function

A(t)=A0exp(12σTF2t2)cos(γμBst+φs),A(t)=A_{0}\exp(-\frac{1}{2}{\sigma_{\rm TF}^{2}t}^{2})\cos(\gamma_{\rm\mu}B_{\rm s}t+\varphi_{s}), (1)

where A0 is the initial asymmetry of the muon spin in the sample. The Gaussian relaxation rate σTF\sigma_{\rm TF} due to the nuclear dipolar fields in the normal state is enhanced in the superconducting state due to the field broadening generated by the emergence of the flux-line lattice (FLL). γμ=8.516\gamma_{\rm\mu}=8.516 ×\times 108 s-1T-1 is the gyromagnetic ratio of the muon, and BsB_{\rm s} is the magnetic field at muon stopping sites.

Figure 2 shows the temperature dependence of σTF\sigma_{\rm TF} obtained from the fits using Eq. (1) for La(2{}_{2}(Cu1-xNix)5As3O2 (x=x= 0.40 and 0.45). There are noticeable upturns in σTF\sigma_{\rm TF} that develop below Tc=2.2T_{c}=2.2 K and 1.8 K, respectively. The internal field distribution in the vortex state is the convolution of the FLL field distribution and the nuclear dipolar field distribution of the host material:

σTF={σFLL2+σdip2(TTc)σdip(T>Tc).\sigma_{\rm TF}=\begin{cases}\sqrt{\sigma_{\rm FLL}^{2}+\sigma_{\rm dip}^{2}}&(T\leqslant T_{c})\\ \sigma_{\rm dip}&(T>T_{c})\end{cases}. (2)
Refer to caption
Figure 2: Temperature dependencies of TF-μ\rm\muSR Gaussian relaxation rate σTF\sigma_{\rm TF} in La(2{}_{2}(Cu1-xNix)5As3O2 (x=0.40x=0.40 and 0.45) for μ0\rm{\mu_{0}}H=30H=30 mT. The arrows indicate TcT_{\rm c} from resistivity measurements [21]. Inset: the normalized superfluid density σFLL(T)/σFLL(0)\sigma_{\rm FLL}(T)/\sigma_{\rm FLL}(0) from Eq. (2) vs. the reduced temperature T/TcT/T_{\rm c}. Solid curves: phenomenological two-fluid model fits (see text).

Above TcT_{\rm c}, σTF=σdip\sigma_{\rm TF}=\sigma_{\rm dip} generated from the nuclear dipolar fields is roughly independent of temperature. Therefore, it was fixed to the average value (0.1179(11) μs1\rm{\mu s^{-1}} for x=0.45x=0.45 and 0.127(2) μs1\rm{\mu s^{-1}} for x=0.40x=0.40). The value of σdip\sigma_{\rm dip} is smaller for larger xx, consistent with the fact that the nuclear magnetic moment of nickel is smaller than that of copper [39].

In the inset of Fig. 2, the normalized FLL relaxation rate σFLL(T)/σFLL(0)\sigma_{\rm FLL}(T)/\sigma_{\rm FLL}(0) is plotted vs. the reduced temperature T/TcT/T_{\rm c} for x=x= 0.40 and 0.45. The data can be fitted with the phenomenological two-fluid model [40, 41]

σFLL(T)=σFLL(0)[1(T/Tc)N](TTc).\sigma_{\rm FLL}(T)=\sigma_{\rm FLL}(0)[1-(T/T_{\rm c})^{N}]\quad(T\leqslant T_{\rm c}). (3)

The fitting parameters are listed in Table 1. There are several common predictions for the value of NN. For traditional BCS superconductors, N4N\sim 4. However, both values of NN for x=x= 0.40 and 0.45 are much lower than 4. The dirty-limit dd-wave model predicts a value of N=2N=2 [42, 43]. However, in the La(2{}_{2}(Cu1-xNix)5As3O2 system, the upper critical field Hc2(0)H_{\rm c2}(0) is estimated to be 3 T [21], giving an estimated value of the Ginzburg-Landau coherence length ξ(0)=[Φ0/2πHc2(0)]1/2=10.47\xi(0)=[\Phi_{0}/2\pi H_{c2}(0)]^{1/2}=10.47 nm, where Φ0=2.07×103\Phi_{0}=2.07\times{10}^{-3} Tm2 represents the magnetic flux quantum. Given the metallic behavior, the mean free path can be estimated by le=kF/ρ0ne2l_{e}=\hbar k_{\rm F}/\rho_{0}ne^{2} [44], with residual resistivity ρ0=\rho_{0}=0.34 mΩcm\rm{m\Omega cm} [21], yielding lel_{e} = 15 nm for x=0.40x=0.40. Thus, La(2{}_{2}(Cu1-xNix)5As3O2 may be a relatively clean superconductor with ξ/le1\xi/l_{e}\approx 1. The predicted clean-limit dd-wave model NN value is N=1N=1 [45]. As a result, our system may include both ss and dd waves.

Table 1: Parameters from fits to TF-μ\rm\muSR data using the phenomenological two-fluid model for La(2{}_{2}(Cu1-xNix)5As3O2.
Parameters x=0.40x=0.40 x=0.45x=0.45
σFLL(0)(μs1)\sigma_{\rm FLL}(0)~{}(\rm{\mu s^{-1}}) 0.287(2)0.287(2) 0.300(4)0.300(4)
Tc(K)T_{c}~{}(\rm{K}) 2.08(4)2.08(4) 1.70(2)1.70(2)
NN 2.72(17)2.72(17) 2.38(13)2.38(13)
λeff(0)(nm)\lambda_{\rm eff}(0)~{}(\rm nm) 559.4(1.9)559.4(1.9) 547(3)547(3)
λab(0)(nm)\lambda_{\rm ab}(0)~{}(\rm nm) 427.02(1.49)427.02(1.49) 418(3)418(3)
σdip(μs1)\sigma_{\rm dip}~{}(\rm{\mu s^{-1}}) 0.127(2)0.127(2) 0.1179(11)0.1179(11)
Adj. R2 0.98443 0.99172

For powder superconductor samples with 0.13/κ2H/Hc210.13/\kappa^{2}\ll H/H_{\rm c2}\ll 1  (κ=λ/ξ\kappa=\lambda/\xi is the Ginzburg-Laudau parameter, HH is the applied field), the Gaussian depolarization rate σFLL\sigma_{\rm FLL} is directly related to the magnetic penetration depth λeff\lambda_{\rm eff} by [46, 47]

σFLLγμ=0.172(1b)[1+1.21(1b)3]Φ02πλeff2,\displaystyle\frac{\sigma_{\rm FLL}}{\gamma_{\mu}}=\frac{0.172(1-b)[1+1.21(1-\sqrt{b})^{3}]\Phi_{0}}{2\pi\lambda_{\rm eff}^{2}}, (4)

where bb is the reduced applied field b=H/Hc2=0.011b=H/H_{\rm c2}=0.01\ll 1. Therefore, the absolute values of effective magnetic penetration depth λeff(0)\lambda_{\rm eff}(0) can be obtained and listed in Table 1. In addition, for layered superconductors, the in-plane magnetic penetration depth λab(0)\lambda_{\rm ab}(0) is also estimated by the relation λeff(0)=31/4λab(0)\lambda_{\rm eff}(0)=3^{1/4}\lambda_{ab}(0) [48, 49, 27].

The effective magnetic penetration depth λeff\lambda_{\rm eff} for isotropic superconductors is related to the density nsn_{\rm s} of superconducting carriers and mm^{*} by the general London equation [40]

λeff2=mμ0e2ns,\lambda_{\rm{eff}}^{2}=\frac{m^{*}}{\mu_{0}e^{2}n_{\rm s}}, (5)

where μ0\mu_{0} is the magnetic constant, mm^{*} is the effective electron mass, and ee is the elementary charge. Therefore, the FLL relaxation rate σFLL\sigma_{\rm FLL} is directly related to the superfluid density through σFLLλeff2ns\sigma_{\rm FLL}\propto\lambda_{\rm{eff}}^{-2}\propto n_{\rm s}. Then we can use microscopic models to investigate the gap symmetry of La(2{}_{2}(Cu1-xNix)5As3O2 in more detail, as shown in Fig. 3.

Refer to caption
Figure 3: The FLL relaxation rate σFLL(T)\sigma_{\rm FLL}(T) from Eq. (2) vs. temperature for La(2{}_{2}(Cu1-xNix)5As3O2 , x=0.40x=0.40 (a) and 0.45 (b). Curves correspond to the ss-wave, dd-wave, s+ss+s-wave, and s+ds+d-wave models.
Table 2: Parameters from the fits to the single-gap and two-gap models from TF-μ\rm\muSR data. The weighting factor of phenomenological two-gap α\alpha model fΔ1f_{\Delta_{1}}, zero-temperature superconducting gap Δ(0)\Delta(0), zero-temperature magnetic penetration depth λab(0)\lambda_{ab}(0), and reduced χred2\rm\chi_{red}^{2} from fits of Eq. (6) and Eq. (9).
Ni doping Model fΔ1f_{\Delta_{1}} 2Δ1/kBTc\Delta_{1}/k_{B}T_{c} 2Δ2/kBTc\Delta_{2}/k_{B}T_{c} λab(0)\lambda_{ab}(0) (nm) χred2\rm\chi^{2}_{red}
ss 1 3.69(17) 427.4(1.8) 1.74
dd 1 6.68(37) 411(3) 4.46
0.40 s+ss+s 0.95(3) 3.86(17) 0.20(13) 419(5) 1.31
s+ds+d 0.70(12) 4.03(21) 5.62(7) 422(3) 1.50
ss 1 3.34(7) 421.5(1.7) 2.19
dd 1 5.09(26) 392(5) 12.9
0.45 s+ss+s 0.988(154) 3.42(2) 0.262(2) 420(31) 2.06
s+ds+d 0.976(114) 3.43(3) 4.15(3) 422(5) 2.09

The single-gap model is defined within the local London approximation [50]:

σFLL(T)σFLL(0)=1+1π02πΔ(T,φ)f(E)EEdEdφE2Δ2(T,φ),\frac{\sigma_{\rm FLL}(T)}{\sigma_{\rm FLL}(0)}=1+\frac{1}{\pi}\int_{0}^{2\pi}\int_{\Delta(T,\varphi)}^{\infty}\frac{\partial f(E)}{\partial E}\frac{E\ \mathrm{d}E\ \rm{d}\varphi}{\sqrt{E^{2}-\Delta^{2}(T,\varphi)}}, (6)
Δ(T,φ)=Δ0δ(T/Tc)g(φ),\Delta(T,\varphi)=\Delta_{0}\delta\left(T/T_{\rm c}\right)g(\varphi), (7)
δ(T/Tc)=tanh{1.82[1.018(Tc/T1)]0.51},\delta\left(T/T_{\rm c}\right)=\tanh\left\{1.82\left[1.018\left(T_{\rm c}/T-1\right)\right]^{0.51}\right\}, (8)

where σFLL(0)\sigma_{\rm FLL}(0) is the FLL relaxation rate at zero temperature, f(E)f(E) is the Fermi-Dirac distribution function, φ\varphi is the angle along the Fermi surface, and Δ0\Delta_{0} is the maximum superconducting gap value at T=0T=0. g(φ)g\left(\varphi\right) in Eq. (7) describes the angular dependence of the superconducting gap. Here g(φ)=1g\left(\varphi\right)=1 and |cos(2φ)||\cos(2\varphi)| refers to the ss-wave and dd-wave model, respectively.

The fitting parameters are listed in Table 2. It is clear from Fig. 3 that the dd-wave model (the dashed yellow lines) does not fit the data, also evidenced by the largest χred2\chi^{2}_{\rm red} in Table 2. The solid red curves in Fig. 3 representing the ss-wave model seem to fit the data well giving the reduced chi-square χred2\rm\chi^{2}_{red} 1.74 and 2.19 for xx = 0.40 and 0.45, respectively. Thus our results preliminarily suggest the ss-wave pairing for both overdoped and underdoped samples for La(2{}_{2}(Cu1-xNix)5As3O2 (xx = 0.45, 0.40). Δ(0)\Delta(0) is larger for the optimal doping sample, suggesting the coupling strength decreases with the increase of doping concentration.

However, we notice that the superfluid density of x=x= 0.40 has a minor upturn at low temperatures. This may be due to nodal or multiband superconductivity. As a result, in addition to single-gap functions, we also employ the phenomenological two-gap α\alpha model with a weighting factor fΔ1f_{\Delta_{1}}  [51, 52, 53],

σFLL(T)σFLL(0)=fΔ1σFLL,Δ1(T)σFLL,Δ1(0)+(1fΔ1)σFLL,Δ2(T)σFLL,Δ2(0),\frac{\sigma_{\rm FLL}(T)}{\sigma_{\rm FLL}(0)}=f_{\Delta_{1}}\frac{\sigma_{\rm FLL,\Delta_{1}}(T)}{\sigma_{\rm FLL,\Delta_{1}}(0)}+(1-f_{\Delta_{1}})\frac{\sigma_{\rm FLL,\Delta_{2}}(T)}{\sigma_{\rm FLL,\Delta_{2}}(0)}, (9)

where σFLL,Δi2(T)/σFLL,Δi2(0)(i=1,2){\sigma_{\rm FLL,\Delta_{i}}^{-2}(T)}/{\sigma_{\rm FLL,\Delta_{i}}^{-2}(0)}~{}(i=1,~{}2) is the superfluid density contribution of one of the gaps.

While the smallest χred2\chi^{2}_{\rm red} in both compounds suggests that the s+ss+s model may be the best model, the resulting values of 2Δ2/kBTc\Delta_{2}/k_{B}T_{c} is unreasonably small. The s+ds+d model is also better than the ss-wave model according to the χred2\chi^{2}_{\rm red} values. Moreover, the gap-to-TcT_{c} ratios 2Δ1,2/kBTc\Delta_{1,2}/k_{B}T_{c} determined from the s+ds+d model are also close to the BCS theoretical predictions (2Δs/kBTc=\Delta_{s}/k_{B}T_{c}= 3.43(3) for ss-wave, and 2Δd/kBTc=\Delta_{d}/k_{B}T_{c}= 4.15(3) for dd-wave, respectively) for xx = 0.45. Both s+ss+s and s+ds+d models can well deal with the upwarping phenomenon of superfluid density at low temperatures in xx = 0.40. However, we also notice that the relaxation rate of xx = 0.45 does not show obvious upwarping at the current lowest temperature of 0.27 K. And also the obtained fΔdf_{\Delta_{d}} values for both compounds are extremely small. More studies are needed to determine whether dd-wave superconductivity exists in one of the bands.

III.2 Uemura Plot

An Uemura plot [54, 55, 56] is shown in Fig. 4, including elemental superconductors (such as Nb, Al, Sn, and Zn), cuprates, alkali-doped C60 (K3C60 and Rb3C60), heavy-fermion superconductors, and La(2{}_{2}(Cu1-xNix)5As3O2. In this classification, unconventional superconductors usually lie in the orange area for 1/100Tc/TF1/101/100\leqslant T_{\rm c}/T_{\rm F}\leqslant 1/10, conventional BCS superconductors fall in the blue region for Tc/TF1/1000T_{\rm c}/T_{\rm F}\leqslant 1/1000, where TFT_{\rm F} is the Fermi temperature [40, 55].

Refer to caption
Figure 4: The Uemura classification scheme [54, 55, 57]. The superconducting transition temperature TcT_{\rm c} versus the effective Fermi temperature, TFT_{\rm F}, where the position of La(2{}_{2}(Cu1-xNix)5As3O2 (x=x= 0.45, 0.40) is shown by the red stars. The unconventional superconductors fall within a band indicated by the red and orange lines. The BCS superconductors are in the lower right region with blue marks.

For the quasi-2D systems, TFT_{\rm F} can be estimated by the following relation [40, 58],

kBTF=π2n2Dsm,k_{\mathrm{B}}T_{\mathrm{F}}=\frac{\pi\hbar^{2}n_{\rm 2D}^{s}}{m^{\ast}}, (10)

where n2Dsn_{\rm{2D}}^{s} is the two dimensional superfluid density within the superconducting planes derived from the in-plane superfluid density via n2Ds=nabsdn_{\rm{2D}}^{s}=n_{\rm{ab}}^{s}d, where dd represents the interplanar distance. According to the general London equation Eq. (5), we can get the in-plane superfluid density through nabs=mμ0e2λab2n_{\rm{ab}}^{s}=\frac{m^{*}}{\mu_{0}e^{2}\lambda_{\rm{ab}}^{2}}.

Correspondingly,

TF=π2λab2dkBμ0e2,T_{\rm F}=\frac{\pi\hbar^{2}\lambda_{\rm{ab}}^{-2}d}{k_{\rm B}\mu_{0}e^{2}}, (11)

therefore, TFT_{\rm F} of La(2{}_{2}(Cu1-xNix)5As3O2 (x=x= 0.45 and 0.40) are estimated, and the results are listed in Table 3. As shown in Fig. 4, La(2{}_{2}(Cu1-xNix)5As3O2 (x=x= 0.40 and 0.45) fall in the crossover between the unconventional superconducting area and the conventional BCS region. The fact that Tc/TFT_{\rm c}/T_{\rm F} is larger in x=x= 0.40 than 0.45 may suggest a stronger pairing interaction in x=0.40x=0.40.

Table 3: Uemura plot parameters for La(2{}_{2}(Cu1-xNix)5As3O2
xx λab(0)\lambda_{\rm{ab}}(0) (nm) dd (Å\rm{\AA}) TF(K)T_{\rm F}(\rm{K}) Tc(K)T_{\rm c}(\rm{K}) Tc/TFT_{\rm c}/T_{\rm F}
0.40 427 22.44 964 2.11 \sim1/457
0.45 422 22.43 990 1.77 \sim1/560

III.3 ZF-μ\rm\muSR

To further investigate the superconductivity in La(2{}_{2}(Cu1-xNix)5As3O2, ZF-μ\rm\muSR experiments were performed. Fig. 5 (a) shows the time evolution of the decay positron count asymmetry, which is proportional to the muon spin polarization, at temperatures above and below TcT_{c} in x=x= 0.45. There is no noticeable difference between the superconducting and the normal state, suggesting the absence of a spontaneous magnetic field below TcT_{c}. Therefore, TRS is conserved below the superconducting transition. Over the entire temperature range, the ZF-μ\rm\muSR spectra can be well described by the following function,

Refer to caption
Figure 5: (a) ZF-μ\rm\muSR time spectra at representative temperatures for polycrystalline samples of La(2{}_{2}(Cu1-xNix)5As3O2 (xx = 0.45). Curves: Fits to the data by Eq. (12). The background signal from muons that stop in the copper sample holder has not been subtracted. (b) Temperature dependence of the Gaussian KT relaxation rate σZF\sigma_{\rm ZF} for La(2{}_{2}(Cu1-xNix)5As3O2 (xx = 0.45). The red dash line represents TcT_{c} determined from the transport measurements. The gray dashed line shows an average of σZF\sigma_{\rm ZF} = 0.1336 μ\mus-1.
A(t)A(0)=(1fbg)GKT(t)+fbgGbg(t),\frac{A\left(t\right)}{A(0)}=(1-f_{\rm bg})G_{\rm KT}(t)+f_{\rm bg}G_{\rm bg}(t), (12)

where the two relaxing terms are the signals of the sample and the background. A(0)A(0) is the initial total asymmetry at time t=t= 0, and fbgf_{\rm bg} is the proportion of the background signal, which has the same value as the one in the TF-μ\rm\muSR experiments. GKT(t)G_{\rm KT}(t) is the static Gaussian Kubo-Toyabe (KT) function [31]

GKT(σZF,t)=13+23(1σZF2t2)exp(12σZF2t2),G_{\rm KT}(\sigma_{\rm ZF},t)=\frac{1}{3}+\frac{2}{3}\left(1-\sigma^{2}_{\rm ZF}t^{2}\right)\exp\left(-\frac{1}{2}\sigma^{2}_{\rm ZF}t^{2}\right), (13)

where σZF\sigma_{\rm ZF} is the Gaussian KT relaxation rate which corresponds to the relaxation due to static, randomly oriented local fields associated with the nuclear moments at the muon site. Fig. 5 (b) shows that there is no significant change in relaxation rate σZF\sigma_{\rm ZF} down to base temperature T=0.27T=0.27 K. The average value of σZF\sigma_{\rm ZF} is 0.1336 μs1\mu s^{-1}, consistent with the typical nuclear dipolar moments of Ni and Cu [59, 60, 61].

IV DISCUSSION

Low-temperature investigations are crucial for determining the superconducting pairing mechanism. Our TF-μ\rm\muSR measurements were performed down to 20 mK, and the temperature dependence of superfluid density suggests that the two-band model best fits the data for both measured samples, one of which is dominated by the ss wave and the other in such a small proportion that we cannot confirm whether it is the ss wave or the dd wave under the current data conditions. The two-band superconductivity scenario is also supported by the density functional theory (DFT) calculations showing that two bands crossing the Fermi energy contribute to the Fermi surface in La2Cu5As3O2 [21]. Furthermore, the temperature dependence of the upper critical field Hc2(T)H_{c2}(T) of La(2{}_{2}(Cu1-xNix)5As3O2 exhibits a upward curvature [21], significantly different from the Werthamer-Helfand-Hohenberg relation [62], which may also proposes multi-band superconductivity [63].

BCS superconductivity and Bose-Einstein condensation (BEC) are two asymptotic limits of a fermionic superfluid. Systems with a small Tc/TFT_{\rm c}/T_{\rm F} < 0.001 are usually considered to be BCS-like, while large Tc/TFT_{\rm c}/T_{\rm F} values are expected only in the BEC-like picture and are considered to be a hallmark feature of unconventional superconductivity [54, 64, 40]. As shown in Fig. 4, La(2{}_{2}(Cu1-xNix)5As3O2 falls in the crossover area between BCS and exotic superconductors regions, like many other superconductors such as TRS breaking superconductors La7Ir3 and LaNiC2 [60, 65, 57], and multi-band iron-based superconductors BaFe2(As1-xPx)2 [66]. In the Uemura plot classification scheme, these superconductors may not be traditional electron-phonon coupled BCS superconductors but more like exotic unconventional superconductors.

More experimental evidence is required to investigate whether dd-wave superconductivity exists in one of the bands. Table 2 shows that the proportion of dd-wave decreases as Ni concentration increases if it does exist. Such a behavior is rare but was observed in layered cuprates La2-xCexCuO4-y and Pr2-xCexCuO4-y, where a transition from dd- to ss-wave pairing occurs near the optimal doping [67]. Also in LaFeAs1-xPx[68], the superconducting order parameter evolves from nodal to nodeless as the doping concentration exceeds 50%. Moreover, a crossover from a nodal to nodeless superconducting energy gap was also suggested in skutterudite PrPt4Ge12 through Ce substitution  [69, 70], although the possibility of a transition from multi-band to single-band superconductivity can not be excluded. In addition, a molecular pairing scenario [45] was proposed to explain the transition from nodal to nodeless superconductivity in Yb-substituted CeCoIn5 [71]. The Yb doping increases the chemical potential and drives a Lifshitz transition of the nodal Fermi surface, forming a fully gapped molecular superfluid of composite pairs. For La(2{}_{2}(Cu1-xNix)5As3O2, the Fermi pocket around the Γ\Gamma point is relatively small according to the DFT calculations [21], and detailed electronic structure study is required to investigate whether the molecular pairing scenario can be applied.

V Conclusions

In summary, we performed ZF and TF -μ\rm\muSR measurements on the recently discovered layered superconductor La(2{}_{2}(Cu1-xNix)5As3O2 (x=x= 0.40, 0.45). The preservation of TRS is suggested by the ZF-μ\rm\muSR  measurements. In combination with the Hc2(T)H_{c2}(T) data and DFT calculations [21], the temperature dependence of the superfluid density of La(2{}_{2}(Cu1-xNix)5As3O2 measured by TF-μ\rm\muSR is best described by the two-band model with the dominant ss wave superconducting energy gap larger for optimal doping, suggesting the coupling strength decreases with the increase of doping concentration. Both samples are classified between unusual superconductors and traditional BCS superconductors in the Uemura plot. More experimental evidence is required to investigate whether dd-wave superconductivity exists in one of the bands.

Acknowledgements.
This work is based on experiments performed at the Swiss Muon Source Sμ\muS, Paul Scherrer Institute, Villigen, Switzerland, and TRIUMF, Vancouver, Canada. The research performed in this work was supported by the National Key Research and Development Program of China, No. 2022YFA1402203, the National Natural Science Foundations of China, No. 12174065 and No.51922105, and the Shanghai Municipal Science and Technology (Major Project Grant No. 2019SHZDZX01 and No. 20ZR1405300).

References

  • Luke et al. [1993] G. M. Luke, A. Keren, L. P. Le, W. D. Wu, Y. J. Uemura, D. A. Bonn, L. Taillefer, and J. D. Garrett, Muon spin relaxation in UPt3\mathrm{UPt}_{3}Phys. Rev. Lett. 71, 1466 (1993).
  • Tsuei and Kirtley [2000] C. C. Tsuei and J. R. Kirtley, Pairing symmetry in cuprate superconductors, Rev. Mod. Phys. 72, 969 (2000).
  • Kivelson et al. [2003] S. A. Kivelson, I. P. Bindloss, E. Fradkin, V. Oganesyan, J. M. Tranquada, A. Kapitulnik, and C. Howald, How to detect fluctuating stripes in the high-temperature superconductors, Rev. Mod. Phys. 75, 1201 (2003).
  • Balatsky et al. [2006] A. V. Balatsky, I. Vekhter, and J.-X. Zhu, Impurity-induced states in conventional and unconventional superconductors, Rev. Mod. Phys. 78, 373 (2006).
  • Onari and Kontani [2009] S. Onari and H. Kontani, Violation of Anderson’s theorem for the sign-reversing ss-wave state of iron-pnictide superconductors, Phys. Rev. Lett. 103, 177001 (2009).
  • Hanaguri et al. [2010] T. Hanaguri, S. Niitaka, K. Kuroki, and H. Takagi, Unconventional ss-wave superconductivity in Fe(Se,Te), Science 328, 474 (2010).
  • Wu et al. [2011] T. Wu, H. Mayaffre, S. Krämer, M. Horvatić, C. Berthier, W. N. Hardy, R. Liang, D. A. Bonn, and M.-H. Julien, Magnetic-field-induced charge-stripe order in the high-temperature superconductor YBa2Cu3Oy\mathrm{YBa}_{2}\mathrm{Cu}_{3}\mathrm{O}_{y}Nature 477, 191 (2011).
  • Guguchia et al. [2014] Z. Guguchia, R. Khasanov, M. Bendele, E. Pomjakushina, K. Conder, A. Shengelaya, and H. Keller, Negative oxygen isotope effect on the static spin stripe order in superconducting La2xBaxCuO4(x=1/8){\mathrm{La}}_{2-x}{\mathrm{Ba}}_{x}{\mathrm{CuO}}_{4}(x=1/8) observed by muon-spin rotation, Phys. Rev. Lett. 113, 057002 (2014).
  • Sarkar et al. [2020] T. Sarkar, D. S. Wei, J. Zhang, N. R. Poniatowski, P. R. Mandal, A. Kapitulnik, and R. L. Greene, Ferromagnetic order beyond the superconducting dome in a cuprate superconductor, Science 368, 532 (2020).
  • Hayes et al. [2021] I. M. Hayes, D. S. Wei, T. Metz, J. Zhang, Y. S. Eo, S. Ran, S. R. Saha, J. Collini, N. P. Butch, D. F. Agterberg, A. Kapitulnik, and J. Paglione, Multicomponent superconducting order parameter in UTe2Science 373, 797 (2021).
  • Nachumi et al. [1996] B. Nachumi, A. Keren, K. Kojima, M. Larkin, G. M. Luke, J. Merrin, O. Tchernyshöv, Y. J. Uemura, N. Ichikawa, M. Goto, and S. Uchida, Muon spin relaxation studies of Zn-substitution effects in high- Tc{T}_{\mathit{c}} cuprate superconductors, Phys. Rev. Lett. 77, 5421 (1996).
  • Smith et al. [2001] C. M. Smith, A. H. Castro Neto, and A. V. Balatsky, Tc{T}_{c} suppression in Co-doped striped cuprates, Phys. Rev. Lett. 87, 177010 (2001).
  • Yang et al. [2013] H. Yang, Z. Wang, D. Fang, Q. Deng, Q. H. Wang, Y. Y. Xiang, Y. Yang, and H. H. Wen, In-gap quasiparticle excitations induced by non-magnetic Cu impurities in Na(Fe(0.96)Co(0.03)Cu(0.01))As revealed by scanning tunnelling spectroscopy, Nat. Commun. 4, 2749 (2013).
  • Guguchia et al. [2017] Z. Guguchia, B. Roessli, R. Khasanov, A. Amato, E. Pomjakushina, K. Conder, Y. J. Uemura, J. M. Tranquada, H. Keller, and A. Shengelaya, Complementary response of static spin-stripe order and superconductivity to nonmagnetic impurities in cuprates, Phys. Rev. Lett. 119, 087002 (2017).
  • Suhl and Matthias [1959] H. Suhl and B. T. Matthias, Impurity scattering in superconductors, Phys. Rev. 114, 977 (1959).
  • Anderson [1959] P. Anderson, Theory of dirty superconductors, J. Phys. Chem. Solids 11, 26 (1959).
  • Li et al. [2009] L. J. Li, Y. K. Luo, Q. B. Wang, H. Chen, Z. Ren, Q. Tao, Y. K. Li, X. Lin, M. He, Z. W. Zhu, G. H. Cao, and Z. A. Xu, Superconductivity induced by Ni doping in BaFe2As2 single crystals, New J. Phys. 11, 025008 (2009).
  • Stewart [2011] G. R. Stewart, Superconductivity in iron compounds, Rev. Mod. Phys. 83, 1589 (2011).
  • Ideta et al. [2013] S. Ideta, T. Yoshida, I. Nishi, A. Fujimori, Y. Kotani, K. Ono, Y. Nakashima, S. Yamaichi, T. Sasagawa, M. Nakajima, K. Kihou, Y. Tomioka, C. H. Lee, A. Iyo, H. Eisaki, T. Ito, S. Uchida, and R. Arita, Dependence of carrier doping on the impurity potential in transition-metal-substituted FeAs-based superconductors, Phys. Rev. Lett. 110, 107007 (2013).
  • Dai [2015] P. Dai, Antiferromagnetic order and spin dynamics in iron-based superconductors, Rev. Mod. Phys. 87, 855 (2015).
  • Chen et al. [2019] X. Chen, J. Guo, C. Gong, E. Cheng, C. Le, N. Liu, T. Ying, Q. Zhang, J. Hu, S. Li, and X. Chen, Anomalous dome-like superconductivity in RE2(Cu1-xNix)5As3O2 (RE = La, Pr, Nd), iScience 14, 171 (2019).
  • Wu et al. [1996] X. Wu, S. Jiang, F. Pan, J. Lin, N. Xu, Mao Zhiqiang, Xu Gaoji, and Zhang Yuheng, Microstructures of La1.85Sr0.15CuO4 doped with Ni at high doping level, Physica C Supercond. 271, 331 (1996).
  • Itoh et al. [2002] Y. Itoh, S. Adachi, T. Machi, Y. Ohashi, and N. Koshizuka, Ni-substituted sites and the effect on Cu electron spin dynamics of YBa2Cu3xNixO7δ{\mathrm{YBa}}_{2}{\mathrm{Cu}}_{3-x}{\mathrm{Ni}}_{x}{\mathrm{O}}_{7-\delta}Phys. Rev. B 66, 134511 (2002).
  • Hillier et al. [2022] A. D. Hillier, S. J. Blundell, I. McKenzie, I. Umegaki, L. Shu, J. A. Wright, T. Prokscha, F. Bert, K. Shimomura, A. Berlie, H. Alberto, and I. Watanabe, Muon spin spectroscopy, Nat. Rev. Methods Primers 2, 5 (2022).
  • Alain Yaouanc [2010] P. D. d. R. Alain Yaouanc, Muon spin rotation, relaxation and resonance, Oxford University Press  (2010).
  • Sonier et al. [2011] J. E. Sonier, W. Huang, C. V. Kaiser, C. Cochrane, V. Pacradouni, S. A. Sabok-Sayr, M. D. Lumsden, B. C. Sales, M. A. McGuire, A. S. Sefat, and D. Mandrus, Magnetism and disorder effects on muon spin rotation measurements of the magnetic penetration depth in iron-arsenic superconductors, Phys. Rev. Lett. 106, 127002 (2011).
  • Zhang et al. [2016] J. Zhang, K. Huang, Z. F. Ding, D. E. MacLaughlin, O. O. Bernal, P.-C. Ho, C. Tan, X. Liu, D. Yazici, M. B. Maple, and L. Shu, Superconducting gap structure in ambient-pressure-grown LaO0.5F0.5BiS2{\mathrm{LaO}}_{0.5}{\mathrm{F}}_{0.5}{\mathrm{BiS}}_{2}Phys. Rev. B 94, 224502 (2016).
  • Zhang et al. [2018] J. Zhang, Z. Ding, C. Tan, K. Huang, O. O. Bernal, P.-C. Ho, G. D. Morris, A. D. Hillier, P. K. Biswas, S. P. Cottrell, H. Xiang, X. Yao, D. E. MacLaughlin, and L. Shu, Discovery of slow magnetic fluctuations and critical slowing down in the pseudogap phase of YBa2Cu3OySci. Adv. 4, eaao5235 (2018).
  • Tan et al. [2018] C. Tan, T. P. Ying, Z. F. Ding, J. Zhang, D. E. MacLaughlin, O. O. Bernal, P. C. Ho, K. Huang, I. Watanabe, S. Y. Li, and L. Shu, Nodal superconductivity coexists with low-moment static magnetism in single-crystalline tetragonal FeS: A muon spin relaxation and rotation study, Phys. Rev. B 97, 174524 (2018).
  • Zhu et al. [2022] Z. H. Zhu, C. Tan, J. Zhang, P. K. Biswas, A. D. Hillier, M. X. Wang, Y. X. Yang, C. S. Chen, Z. F. Ding, S. Y. Li, and L. Shu, Muon spin rotation and relaxation study on topological noncentrosymmetric superconductor PbTaSe2New J. Phys. 24, 023002 (2022).
  • Hayano et al. [1979] R. S. Hayano, Y. J. Uemura, J. Imazato, N. Nishida, T. Yamazaki, and R. Kubo, Zero-and low-field spin relaxation studied by positive muons, Phys. Rev. B 20, 850 (1979).
  • Amato [1997] A. Amato, Heavy-fermion systems studied by μ\muSR technique, Rev. Mod. Phys. 69, 1119 (1997).
  • Aoki et al. [2003] Y. Aoki, A. Tsuchiya, T. Kanayama, S. R. Saha, H. Sugawara, H. Sato, W. Higemoto, A. Koda, K. Ohishi, K. Nishiyama, and R. Kadono, Time-reversal symmetry-breaking superconductivity in heavy-fermion PrOs4Sb12{\mathrm{P}\mathrm{r}\mathrm{O}\mathrm{s}}_{4}{\mathrm{S}\mathrm{b}}_{12} detected by muon-spin relaxation, Phys. Rev. Lett. 91, 067003 (2003).
  • Neha et al. [2019] P. Neha, P. K. Biswas, T. Das, and S. Patnaik, Time-reversal symmetry breaking in topological superconductor Sr0.1Bi2Se3{\mathrm{Sr}}_{0.1}{\mathrm{Bi}}_{2}{\mathrm{Se}}_{3}Phys. Rev. Materials 3, 074201 (2019).
  • Suter and Wojek [2012] A. Suter and B. Wojek, Musrfit: A free platform-independent framework for μ\muSR data analysis, Physics Procedia 30, 69 (2012), 12th International Conference on Muon Spin Rotation, Relaxation and Resonance (μ\muSR2011).
  • Karl et al. [2019] R. Karl, F. Burri, A. Amato, M. Donegà, S. Gvasaliya, H. Luetkens, E. Morenzoni, and R. Khasanov, Muon spin rotation study of type-I superconductivity: Elemental β\beta-Sn, Phys. Rev. B 99, 184515 (2019).
  • Weber et al. [1993] M. Weber, A. Amato, F. N. Gygax, A. Schenck, H. Maletta, V. N. Duginov, V. G. Grebinnik, A. B. Lazarev, V. G. Olshevsky, V. Y. Pomjakushin, S. N. Shilov, V. A. Zhukov, B. F. Kirillov, A. V. Pirogov, A. N. Ponomarev, V. G. Storchak, S. Kapusta, and J. Bock, Magnetic-flux distribution and the magnetic penetration depth in superconducting polycrystalline Bi2{\mathrm{Bi}}_{2}Sr2{\mathrm{Sr}}_{2}Ca1x{\mathrm{Ca}}_{1\mathrm{-}\mathit{x}}Yx{\mathrm{Y}}_{\mathit{x}}Cu2{\mathrm{Cu}}_{2}O8+δ{\mathrm{O}}_{8+\mathrm{\delta}} and Bi2x{\mathrm{Bi}}_{2\mathrm{-}\mathit{x}}Pbx{\mathrm{Pb}}_{\mathit{x}}Sr2{\mathrm{Sr}}_{2}CaCu2{\mathrm{CaCu}}_{2}O8+δ{\mathrm{O}}_{8+\mathrm{\delta}}Phys. Rev. B 48, 13022 (1993).
  • Maisuradze et al. [2009a] A. Maisuradze, R. Khasanov, A. Shengelaya, and H. Keller, Comparison of different methods for analyzing μ\muSR line shapes in the vortex state of type-II superconductors, Journal of Physics: Condensed Matter 21, 075701 (2009a).
  • N.W.Ashcroft [1976] N. M. N.W.Ashcroft, Solid state physics (Saunders College Publishing, 1976) pp. 657–658.
  • Hillier and Cywinski [1997] A. D. Hillier and R. Cywinski, The classification of superconductors using muon spin rotation, Appl. Magn. Reson. 13, 95 (1997).
  • Ding et al. [2019] Z. F. Ding, J. Zhang, C. Tan, K. Huang, Q. Y. Chen, I. Lum, O. O. Bernal, P.-C. Ho, D. E. MacLaughlin, M. B. Maple, and L. Shu, Renormalizations in unconventional superconducting states of Ce1xYbxCoIn5{\mathrm{Ce}}_{1-x}{\mathrm{Yb}}_{x}{\mathrm{CoIn}}_{5}Phys. Rev. B 99, 035136 (2019).
  • Luetkens et al. [2008] H. Luetkens, H.-H. Klauss, R. Khasanov, A. Amato, R. Klingeler, I. Hellmann, N. Leps, A. Kondrat, C. Hess, A. Köhler, G. Behr, J. Werner, and B. Büchner, Field and temperature dependence of the superfluid density in LaFeAsO1xFx{\mathrm{LaFeAsO}}_{1-x}{\mathrm{F}}_{x} superconductors: A muon spin relaxation study, Phys. Rev. Lett. 101, 097009 (2008).
  • Hirschfeld et al. [1994] P. J. Hirschfeld, W. O. Putikka, and D. J. Scalapino, dd-wave model for microwave response of high-Tc{T}_{c} superconductors, Phys. Rev. B 50, 10250 (1994).
  • Wang et al. [2022] Y. Wang, M. Li, C. Pei, L. Gao, K. Bu, D. Wang, X. Liu, L. Yan, J. Qu, N. Li, B. Wang, Y. Fang, Y. Qi, and W. Yang, Critical current density and vortex phase diagram in the superconductor Sn0.55In0.45Te{\mathrm{Sn}}_{0.55}{\mathrm{In}}_{0.45}\mathrm{Te}Phys. Rev. B 106, 054506 (2022).
  • Erten et al. [2015] O. Erten, R. Flint, and P. Coleman, Molecular pairing and fully gapped superconductivity in Yb-doped CeCoIn5{\mathrm{CeCoIn}}_{5}Phys. Rev. Lett. 114, 027002 (2015).
  • Brandt [1988] E. H. Brandt, Flux distribution and penetration depth measured by muon spin rotation in high-Tc{T}_{c} superconductors, Phys. Rev. B 37, 2349 (1988).
  • Brandt [2003] E. H. Brandt, Properties of the ideal Ginzburg-Landau vortex lattice, Phys. Rev. B 6810.1103/PhysRevB.68.054506 (2003).
  • Barford and Gunn [1988] W. Barford and J. Gunn, The theory of the measurement of the London penetration depth in uniaxial type II superconductors by muon spin rotation, Physica C Supercond. 156, 515 (1988).
  • Fesenko et al. [1991] V. Fesenko, V. Gorbunov, and V. Smilga, Analytical properties of muon polarization spectra in type-II superconductors and experimental data interpretation for mono- and polycrystalline HTSCs, Physica C Supercond. 176, 551 (1991).
  • Si et al. [2010] W. Si, Q. Jie, L. Wu, J. Zhou, G. Gu, P. D. Johnson, and Q. Li, Superconductivity in epitaxial thin films of Fe1.08Te:OxPhys. Rev. B 81, 092506 (2010).
  • Padamsee et al. [1973] H. Padamsee, J. E. Neighbor, and C. A. Shiffman, Quasiparticle phenomenology for thermodynamics of strong-coupling superconductors, J. Low Temp. Phys. 12, 387 (1973).
  • Carrington and Manzano [2003] A. Carrington and F. Manzano, Magnetic penetration depth of MgB2Physica C Supercond. 385, 205 (2003).
  • Khasanov et al. [2007] R. Khasanov, A. Shengelaya, A. Maisuradze, F. L. Mattina, A. Bussmann-Holder, H. Keller, and K. A. Müller, Experimental evidence for two gaps in the high-temperature La1.83Sr0.17CuO4 superconductor, Phys. Rev. Lett. 98, 057007 (2007).
  • Uemura et al. [1989] Y. J. Uemura, G. M. Luke, B. J. Sternlieb, J. H. Brewer, J. F. Carolan, W. N. Hardy, R. Kadono, J. R. Kempton, R. F. Kiefl, S. R. Kreitzman, P. Mulhern, T. M. Riseman, D. L. Williams, B. X. Yang, S. Uchida, H. Takagi, J. Gopalakrishnan, A. W. Sleight, M. A. Subramanian, C. L. Chien, M. Z. Cieplak, G. Xiao, V. Y. Lee, B. W. Statt, C. E. Stronach, W. J. Kossler, and X. H. Yu, Universal correlations between Tc{T}_{c} and nsm\frac{{n}_{s}}{{m}^{*}} (carrier density over effective mass) in high-Tc{T}_{c} cuprate superconductors, Phys. Rev. Lett. 62, 2317 (1989).
  • Uemura [2004] Y. J. Uemura, Condensation, excitation, pairing, and superfluid density in high-Tc{T}_{c} superconductors: the magnetic resonance mode as a roton analogue and a possible spin-mediated pairing, J. Phys. Condens. Matter 16, S4515 (2004).
  • Nakagawa et al. [2021] Y. Nakagawa, Y. Kasahara, T. Nomoto, R. Arita, T. Nojima, and Y. Iwasa, Gate-controlled BCS-BEC crossover in a two-dimensional superconductor, Science 372, 190 (2021).
  • Barker et al. [2018] J. A. T. Barker, B. D. Breen, R. Hanson, A. D. Hillier, M. R. Lees, G. Balakrishnan, D. M. Paul, and R. P. Singh, Superconducting and normal-state properties of the noncentrosymmetric superconductor Re3Ta\mathrm{Re}_{3}\mathrm{Ta}Phys. Rev. B 98, 104506 (2018).
  • Benfatto et al. [2001] L. Benfatto, S. Caprara, C. Castellani, A. Paramekanti, and M. Randeria, Phase fluctuations, dissipation, and superfluid stiffness in dd-wave superconductors, Phys. Rev. B 63, 174513 (2001).
  • Luke et al. [1991] G. M. Luke, J. H. Brewer, S. R. Kreitzman, D. R. Noakes, M. Celio, R. Kadono, and E. J. Ansaldo, Muon diffusion and spin dynamics in copper, Phys. Rev. B 43, 3284 (1991).
  • Hillier et al. [2009] A. D. Hillier, J. Quintanilla, and R. Cywinski, Evidence for time-reversal symmetry breaking in the noncentrosymmetric superconductor LaNiC2{\mathrm{LaNiC}}_{2}Phys. Rev. Lett. 102, 117007 (2009).
  • K. P. et al. [2020] S. K. P., D. Singh, A. D. Hillier, and R. P. Singh, Probing nodeless superconductivity in LaMSi\mathrm{La}{M}\mathrm{Si} (M=Ni{M}=\mathrm{Ni}, Pt) using muon-spin rotation and relaxation, Phys. Rev. B 102, 094515 (2020).
  • Werthamer et al. [1966] N. R. Werthamer, E. Helfand, and P. C. Hohenberg, Temperature and purity dependence of the superconducting critical field, Hc2{H}_{c2}. III. Electron spin and spin-orbit effects, Phys. Rev. 147, 295 (1966).
  • Hunte et al. [2008] F. Hunte, J. Jaroszynski, A. Gurevich, D. C. Larbalestier, R. Jin, A. S. Sefat, M. A. McGuire, B. C. Sales, D. K. Christen, and D. Mandrus, Two-band superconductivity in LaFeAsO0.89F0.11 at very high magnetic fields, Nature 453, 903 (2008).
  • Uemura et al. [1991] Y. J. Uemura, L. P. Le, G. M. Luke, B. J. Sternlieb, W. D. Wu, J. H. Brewer, T. M. Riseman, C. L. Seaman, M. B. Maple, M. Ishikawa, D. G. Hinks, J. D. Jorgensen, G. Saito, and H. Yamochi, Basic similarities among cuprate, bismuthate, organic, chevrel-phase, and heavy-fermion superconductors shown by penetration-depth measurements, Phys. Rev. Lett. 66, 2665 (1991).
  • Barker et al. [2015] J. A. T. Barker, D. Singh, A. Thamizhavel, A. D. Hillier, M. R. Lees, G. Balakrishnan, D. M. Paul, and R. P. Singh, Unconventional superconductivity in La7Ir3 revealed by muon spin relaxation: Introducing a new family of noncentrosymmetric superconductor that breaks time-reversal symmetry, Phys. Rev. Lett. 115, 267001 (2015).
  • Hashimoto et al. [2012] K. Hashimoto, K. Cho, T. Shibauchi, S. Kasahara, Y. Mizukami, R. Katsumata, Y. Tsuruhara, T. Terashima, H. Ikeda, M. A. Tanatar, H. Kitano, N. Salovich, R. W. Giannetta, P. Walmsley, A. Carrington, R. Prozorov, and Y. Matsuda, A sharp peak of the zero-temperature penetration depth at optimal composition in BaFe2(As1-xPx)2Science 336, 1554 (2012).
  • Skinta et al. [2002] J. A. Skinta, M.-S. Kim, T. R. Lemberger, T. Greibe, and M. Naito, Evidence for a transition in the pairing symmetry of the electron-doped cuprates La2-xCexCuO4-y and Pr2-xCexCuO4-yPhys. Rev. Lett. 88, 207005 (2002).
  • Shiroka et al. [2018] T. Shiroka, N. Barbero, R. Khasanov, N. D. Zhigadlo, H. R. Ott, and J. Mesot, Nodal-to-nodeless superconducting order parameter in LaFeAs1-xPxO synthesized under high pressure, npj Quant Mater 3, 25 (2018).
  • Maisuradze et al. [2009b] A. Maisuradze, M. Nicklas, R. Gumeniuk, C. Baines, W. Schnelle, H. Rosner, A. Leithe-Jasper, Y. Grin, and R. Khasanov, Superfluid density and energy gap function of superconducting PrPt4Ge12\mathrm{PrPt}_{4}\mathrm{Ge}_{12}Phys. Rev. Lett. 103, 147002 (2009b).
  • Huang et al. [2014] K. Huang, L. Shu, I. K. Lum, B. D. White, M. Janoschek, D. Yazici, J. J. Hamlin, D. A. Zocco, P.-C. Ho, R. E. Baumbach, and M. B. Maple, Probing the superconductivity of PrPt4Ge12\mathrm{PrPt}_{4}\mathrm{Ge}_{12} through Ce substitution, Phys. Rev. B 89, 035145 (2014).
  • Kim et al. [2015] H. Kim, M. A. Tanatar, R. Flint, C. Petrovic, R. Hu, B. D. White, I. K. Lum, M. B. Maple, and R. Prozorov, Nodal to nodeless superconducting energy-gap structure change concomitant with fermi-surface reconstruction in the heavy-fermion compound CeCoIn5\mathrm{CeCoIn}_{5}Phys. Rev. Lett. 114, 027003 (2015).