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

thanks: To whom correspondence should be addressed; E-mail: [email protected]thanks: To whom correspondence should be addressed; E-mail: [email protected]

Localized-to-itinerant transition preceding antiferromagnetic quantum critical point and gapless superconductivity in CeRh0.5Ir0.5In5

Shinji Kawasaki Department of Physics, Okayama University, Okayama 700-8530, Japan    Toshihide Oka Department of Physics, Okayama University, Okayama 700-8530, Japan    Akira Sorime Department of Physics, Okayama University, Okayama 700-8530, Japan    Yuji Kogame Department of Physics, Okayama University, Okayama 700-8530, Japan    Kazuhiro Uemoto Department of Physics, Okayama University, Okayama 700-8530, Japan    Kazuaki Matano Department of Physics, Okayama University, Okayama 700-8530, Japan    Jing Guo Institute of Physics, Chinese Academy of Sciences, and Beijing National Laboratory for Condensed Matter Physics, Beijing 100190, China    Shu Cai Institute of Physics, Chinese Academy of Sciences, and Beijing National Laboratory for Condensed Matter Physics, Beijing 100190, China    Liling Sun Institute of Physics, Chinese Academy of Sciences, and Beijing National Laboratory for Condensed Matter Physics, Beijing 100190, China    John L. Sarrao Los Alamos National Laboratory, Los Alamos, New Mexico 87545, USA    Joe D. Thompson Los Alamos National Laboratory, Los Alamos, New Mexico 87545, USA    Guo-qing Zheng Department of Physics, Okayama University, Okayama 700-8530, Japan Institute of Physics, Chinese Academy of Sciences, and Beijing National Laboratory for Condensed Matter Physics, Beijing 100190, China
Abstract

A fundamental problem posed from the study of correlated electron compounds, of which heavy-fermion systems are prototypes, is the need to understand the physics of states near a quantum critical point (QCP). At a QCP, magnetic order is suppressed continuously to zero temperature and unconventional superconductivity often appears. Here, we report pressure (PP) -dependent 115In nuclear quadrupole resonance (NQR) measurements on heavy-fermion antiferromagnet CeRh0.5Ir0.5In5. These experiments reveal an antiferromagnetic (AF) QCP at PcAFP_{\rm c}^{\rm AF} = 1.2 GPa where a dome of superconductivity reaches a maximum transition temperature TcT_{\rm c}. Preceding PcAFP_{\rm c}^{\rm AF}, however, the NQR frequency νQ\nu_{\rm Q} undergoes an abrupt increase at PcP_{\rm c}^{\rm*} = 0.8 GPa in the zero-temperature limit, indicating a change from localized to itinerant character of cerium’s ff-electron and associated small-to-large change in the Fermi surface. At PcAFP_{\rm c}^{\rm AF} where TcT_{\rm c} is optimized, there is an unusually large fraction of gapless excitations well below TcT_{\rm c} that implicates spin-singlet, odd-frequency pairing symmetry.

pacs:

Understanding non-Fermi liquid behaviors Mathur ; Lohneysen due to a zero-temperature magnetic transition, a quantum critical point (QCP), and the unconventional superconductivity that emerges around it is one of the central issues in condensed matter physics. These phenomena are widely seen in strongly correlated electron systems such as heavy fermion systems Mathur , cuprates Schmalian , and iron pnictides Abraham ; Zhou . In cuprates, iron pnictides, or other compounds containing 3dd transition-metal elements JLuo , the quantum phase transition is described by itinerant spin-density wave (SDW) theories, where the QCP is due to an instability of the underlying large Fermi surfaces Moriya_SCR ; Hertz ; Millis ; SiSteglich . Cerium(Ce)-based heavy fermion systems are understood based on the Kondo lattice model in which localized Ce 4ff electron spins at high temperatures are screened below a characteristic temperature TKT_{\rm K} by the conduction electrons Doniach . At high temperatures, the ff electron spins are localized, and thus, the Fermi surface is small. With decreasing temperature, 4ff electrons couple with the conduction electrons through Kondo hybridization, and the Fermi-surface volume gradually increases with decreasing temperature DLFengCo ; DLFeng . In addition to the Kondo effect, there also is a long-range Ruderman-Kittel-Kasuya-Yosida (RKKY) interaction that is the indirect exchange interaction among weakly screened, nearly localized 4f4f electrons. If the RKKY interaction overcomes the Kondo effect, ff spins order antiferromagnetically below the Néel temperature TNT_{\rm N}. By tuning a non-thermal control parameter such as pressure and/or chemical substitution, TNT_{\rm N} can be suppressed completely to zero, and TKT_{\rm K} increases as the parameter increases Doniach . A crossover from the small (localized) to the large (itinerant) Fermi surfaces will occur well below TKT_{\rm K} in the Kondo lattice Burdin . Depending on the relative balance between Kondo hybridization and the RKKY interaction, magnetic order may be of the SDW-type or the RKKY-type antiferromagnetic (AF) order that is mediated by itinerant electrons of a small Fermi surface. Quantum criticality of the latter type of magnetic order is predicted theoretically to be qualitatively different from SDW criticality and involves a breakdown of Kondo screening and a transition from small-to-large Fermi surfaces at the QCP QSi ; ColemanJLTP ; SiSteglich . In practice, it can be difficult experimentally to distinguish unambiguously between these two scenarios, though distinctions have been inferred from combinations of thermodynamic, transport and inelastic neutron scattering measurements LohneysenRMP . Unfortunately, it has not been possible to perform neutron scattering experiments under high pressure conditions, the “cleanest” tuning parameter that does not introduce additional disorder or break symmetry, to shed light on the nature of criticality in Kondo lattice systems. In contrast, pressure-dependent nuclear quadrupole resonance (NQR) measurements, which probe the dynamic spin susceptibility as well as the influence of Kondo hybridization, are straightforward, even at very low temperatures, and, as we show, can be used as differentiating probe of quantum criticality.

The antiferromagnetic superconductor CeRh0.5Ir0.5In5 Pagliuso ; Yamaguchi is a good candidate to investigate this issue. CeRhIn5 has a small Fermi surface Shishido , orders antiferromagnetically below TNT_{\rm N} = 3.8 K, and exhibits pressure-induced superconductivity above PP = 1.6 GPa at the transition temperature TcT_{\rm c} = 2.1 K Hegger . TcT_{\rm c} increases to 2.3 K at 2.35 GPa where, in the limit of zero temperature, there is a magnetic-non-magnetic transition ParkNature ; Knebel accompanied by a change from small to large Fermi surface Shishido . The superconducting-gap symmetry is consistent with dd-wave symmetry Mito2001 ; ParkPRL . On the other hand, CeIrIn5 has a large Fermi surface ShishidoIr , and also shows dd-wave superconductivity below TcT_{\rm c} = 0.4 K Petrovic1 ; ZhengIr ; LuPRL , which increases to 0.8 K at PP = 2.1 GPa KawasakiIr . The alloyed compound CeRh0.5Ir0.5In5 orders antiferromagnetically below TNT_{\rm N} = 3.0 K at PP = 0 ADC and becomes superconducting below TcT_{\rm c} = 0.9 K Pagliuso ; Nicklas ; Yamaguchi . CeRh0.5Ir0.5In5 is closer to an antiferromagnetic QCP than CeRhIn5, suggesting that Ir substitution for Rh acts a positive chemical pressure. This suggestion can be understood by appreciating that underlying the evolution of ground states in the CeRh1-xIrxIn5 series is a systematic change in orbital anisotropy of the Γ72\Gamma_{7}^{2} crystal-electric field ground state wavefunction that produces progressively stronger hybridization with increasing xx Willers .

Here we report the results of 115In-NQR measurements on CeRh0.5Ir0.5In5 under pressure, crystal structure analysis and a first-principle calculation of NQR frequency νQ\nu_{\rm Q} (See Methods for details). From the temperature dependence of the nuclear spin-lattice relaxation rate (1/T11/T_{1}), we find that the antiferromagnetic QCP is at PcAFP_{\rm c}^{\rm AF} = 1.2 GPa, where Tc(P)T_{\rm c}(P) reaches its maximum. From the pressure dependence of νQ\nu_{\rm Q}, we find a localized-to-itinerant transition at PcP_{\rm c}^{\rm*} = 0.8 GPa before the antiferromagnetic QCP appears. Superconductivity is not only optimized at the antiferromagnetic QCP but also is realized with a remarkable proliferation of residual gapless excitations. Our results suggest that the large Fermi surface and antiferromagnetic instabilities in the presence of “impurity” scattering trigger unconventional gapless superconductivity in CeRh0.5Ir0.5In5.

Refer to caption
Fig. 1: Phase diagram of CeRh1-xIrxIn5. xx and pressure dependence of the Néel temperature TNT_{\rm{}_{N}} (open squares, triangles, and circles) and the superconducting transition temperature TcT_{\rm c} (solid squares, triangles, circles, and diamonds) for CeRh1-xIrxIn5 (PP = 0) Yamaguchi , CeRhIn5 Hegger ; ParkNature ; Knebel , and CeIrIn5 KawasakiIr under pressure. AFM and SC indicate antiferromagnetic metal and superconductivity, respectively.

Results

The hyperfine coupling constant at the In(1) site.

The hyperfine coupling between nuclear and electronic spins relates the measured 1/T11/T_{1} to the underlying dynamical spin susceptibility as discussed in Methods. Figure 1 shows the pressure-temperature phase diagram of the CeRh1-xIrxIn5 system. If substituting Ir for Rh acts as a positive chemical pressure, we would expect the hyperfine coupling constant [A115(1){}^{115}A(1)] at the In(1) site (Fig. 2a) of CeRh0.5Ir0.5In5 to be smaller than that of the host material CeRhIn5 because A115(1){}^{115}A(1) = 25 kOe μB1\mu_{\rm B}^{-1} in CeRhIn5 decreases with increasing pressure but becomes constant at A115(1){}^{115}A(1) \sim 7 kOe μB1\mu_{\rm B}^{-1} above PP = 1 GPa. Such a pressure dependent A115(1){}^{115}A(1) will affect the information inferred from 1/T1T_{1} Curro and, therefore, it is important to determine A115(1){}^{115}A(1) for CeRh0.5Ir0.5In5 before proceeding to details of its T1T_{1} results under pressure. Figure 2b shows the frequency-swept 115In-nuclear magnetic resonance (NMR) spectra at a constant field. The spectrum is consistent with previously reported spectra of CeRhIn5 Curro . From these data, we establish the temperature dependence of the 115In(1)-NMR center line plotted in Fig. 2c. With these results, we calculate (Methods) the temperature dependence of the Knight shift K115(1)c{}^{115}K(1)_{\rm c} (%) and compare it to dc susceptibility χc\chi_{\rm c} (emu mol-1) in Fig. 2d. As clearly shown in the figure, the relation K(T)K(T) \propto A115(1)χ(T){}^{115}A(1)\chi(T) holds above TT = 30 K. A breakdown of this linear relationship at low temperature is common in heavy electron systems CurroYoung , but a previous In-NMR study suggested that A115(1){}^{115}A(1) is temperature-independent in CeIrIn5 in such temperature regime Kambe1 . In the inset of Fig. 2d we plot K115(1)c{}^{115}K(1)_{\rm c} versus χc\chi_{\rm c} and obtain the hyperfine coupling constant as A115(1){}^{115}A(1) = 7.64 kOe μB1\mu_{\rm B}^{-1}. This closely corresponds to the value of A115(1){}^{115}A(1) in CeRhIn5 under a pressure of 1 GPa (Supplementary Note 1 and Supplementary Figure 1), and thus, it substantiates the suggestion that Ir substitution with xx = 0.5 (chemical pressure) is equivalent to the application of a physical pressure greater than PP = 1 GPa to CeRhIn5. Therefore, we reasonably can ignore any significant pressure dependence of A115(1){}^{115}A(1) in inferring the pressure evolution of physical properties from 1/T1T_{1} in CeRh0.5Ir0.5In5.

Refer to caption
Fig. 2: Hyperfine coupling constant. (a) Crystal structure of CeRh0.5Ir0.5In5. (b) Frequency-swept 115In-nuclear magnetic resonance (NMR) spectra of CeRh0.5Ir0.5In5. Solid [dotted] arrows indicate In(1)[In(2)] resonance peaks, respectively. (c) Temperature dependence of the 115In(1)-NMR center line. The curves are Gaussian fits. (d) Temperature dependence of the Knight shift 115K(1)cK(1)_{\rm c} (open circles) and dc susceptibility χc\chi_{\rm c} (solid curve). The inset shows 115K(1)cK(1)_{\rm c} versus χc\chi_{\rm c}. The solid straight line is a fit to the data above TT = 30 K which yields the hyperfine coupling constant 115A(1)A(1) = 7.64 kOe μB1\mu_{\rm B}^{-1}. Error bars are smaller than the size of the data points.
Refer to caption
Fig. 3: Antiferromagnetism and superconductivity under pressure. Temperature dependence of the In(1)-nuclear quadrupole resonance (NQR) nuclear spin-lattice relaxation rate 1/T1T_{1} (a) below and (b) above the antiferromagnetic (AF) quantum critical point PcAFP_{\rm c}^{\rm AF}. (Inset) Data for PP = 1.12 and 1.24 GPa just around the Néel temperature TNT_{\rm N} and the superconducting transition temperature TcT_{\rm c}. The dotted (solid) arrows indicate TN(Tc)T_{\rm N}(T_{\rm c}), while the dashed line for La(Ir,Rh)In5 indicates 1/T1TT_{1}T = constant. Data at PP = 0 Yamaguchi and for La(Ir,Rh)In5 are obtained from GQZ . Error bars are smaller than the size of the data points.
Refer to caption
Fig. 4: Localized to itinerant transition. Pressure dependence of the In(1) 115In-nuclear quadrupole resonance (NQR) spectrum (±\pm1/2 \leftrightarrow ±\pm3/2 transition line) at TT = 4.2 K (a) and 1.6 K (b). The curves are Gaussian fits. The dotted vertical lines indicate the peak positions at PP = 0. (c) Pressure dependence of the NQR frequency (νQ\nu_{\rm Q}) obtained at TT = 0.3, 1.6, and 4.2 K. For clarity, the vertical axis for TT = 1.6 K and 4.2 K are offset by the amount shown in the parenthesis. Solid arrows indicate localized-to-itinerant transition pressure (PP^{\rm*}) which is defined from the midpoint of νQ\nu_{\rm Q} jump. Solid straight lines are fits to the data which yield the slope dνQ/dPd\nu_{\rm Q}/dP = 0.008 (0.027) below (above) PP^{\rm*}. Error bars represent the uncertainty in estimating νQ\nu_{\rm Q}. (d) Temperature dependence of the In(2) 115In-NQR spectrum (±\pm3/2 \leftrightarrow ±\pm5/2 transition line) at PP = 1.12 GPa. The curves are Gaussian fits. Vertical line indicates the peak position at TT = 8 K. (e) Temperature dependence of the peak position of the In(2) spectrum (solid triangles) and In(1) νQ\nu_{\rm Q} (open circles) at PP = 1.12 GPa. Solid arrow indicates localized-to-itinerant transition temperature TT^{\rm*} which is defined from the midpoint of νQ\nu_{\rm Q} jump. The dashed line serves as visual guide. Error bars are smaller than the size of the data points except for those in (c).
Refer to caption
Fig. 5: Lattice parameters and band calculation. (a) Pressure dependence of aa- and cc-axis lengths for CeRh0.5Ir0.5In5 at room temperature. (b) Pressure dependence of the nuclear quardrupole resonance frequency νQHiLAPW\nu_{\rm Q}^{\rm HiLAPW} obtained from band calculations [Hiroshima Linear-Augmented-Plane-Wave (HiLAPW) code] for CeRh(Ir)In5 and LaRh(Ir)In5. LaRh(Ir)In5 corresponds to the 4ff-localized model of CeRh(Ir)In5. Both are calculated with the same lattice constant. Error bars are smaller than the size of the data points.
Refer to caption
Fig. 6: Antiferromagnetic (AF) quantum critical point (PcAFP_{\rm c}^{\rm AF}). Temperature dependence of the nuclear spin-lattice relaxation rate divided by temperature 1/T1T1/T_{1}T below (a) and above (b) PcAFP_{\rm c}^{\rm AF} with the fitting curves (see text). The dotted (solid) arrows indicate the Néel temperature TNT_{\rm N} (the superconducting transition temperature Tc)T_{\rm c}). (c) Summary of pressure dependence of the characteristic temperature θ\theta (see text), TNT_{\rm N}, and TcT_{\rm c}. The dotted straight line (filled diamonds) is a fit to the data which yields PcAFP_{\rm c}^{\rm AF} (θ=0\theta=0) = 1.2 GPa. PM and AFM indicate paramagnetic and antiferromagnetic metal, respectively. SC indicates superconductivity. Error bars are smaller than the size of the data points.
Refer to caption
Fig. 7: Low-lying excitations in the superconducting state. Temperature dependence of the nuclear spin-lattice relaxation rate divided by temperature 1/T1TT_{1}T below (a) and above (b) the antiferromagnetic (AF) quantum critical point (PcAFP_{\rm c}^{\rm AF}). Data at PP = 0 are obtained from the literature Yamaguchi . Horizontal and vertical axes are normalized by the superconducting transition temperature Tc(P)T_{\rm c}(P) and 1/T1T[Tc(P)]1/T_{1}T[T_{\rm c}(P)], respectively. Error bars are smaller than the size of the data points.
Refer to caption
Fig. 8: Phase diagram. (a) Pressure-temperature phase diagram for CeRh0.5Ir0.5In5 obtained at zero-magnetic field. The solid circles, squares, and stars indicate the Néel temperature TNT_{\rm N}, the superconducting transition temperature TcT_{\rm c}, and the localized-to-itinerant transition temperature TT^{*}, respectively. The broken curve is a phase boundary separating small and large Fermi surfaces. (b) Pressure dependence of the relative density of states (DOS) on the TT = 0.3 K plane. The dotted straight line indicates the antiferromagnetic (AF) quantum critical point (PcAFP_{\rm c}^{\rm AF} = 1.2 GPa). The solid and dashed curves serve as visual guides. AFM and SC indicate antiferromagnetic metal and superconductivity, respectively. Error bars are smaller than the size of the data points.

Pressure dependence of TNT_{\rm N} and TcT_{\rm c}. Figure 3a and 3b show the temperature dependence of 1/T11/T_{1}. The magnitude of 1/T11/T_{1} for CeRh0.5Ir0.5In5 is much greater than that for the nonmagnetic reference material La(Rh,Ir)In5 GQZ , due to ff electron spin correlations, that are reflected in the dynamical susceptibility χ′′\chi^{\prime\prime} to which 1/T11/T_{1} is proportional. At PP = 0, 1/T11/T_{1} exhibits a small peak at TNT_{\rm N} = 3.0 K and decreases below TcT_{\rm c} = 0.9 K Yamaguchi . As shown in Fig. 3a and the inset, TN(P)T_{\rm N}(P) can be identified up to PP = 1.12 GPa but disappears above PcAFP_{\rm c}^{\rm AF} = 1.2 GPa. A superconducting transition is observed at all pressures, as evidenced by an abrupt reduction of 1/T1T_{1} below TT = Tc(P)T_{\rm c}(P). TcT_{\rm c} increases with increasing pressure and exhibits a maximum TcmaxT_{\rm c}^{\rm max} = 1.4 K around PcAFP_{\rm c}^{\rm AF} = 1.2 GPa and then decreases with further increase of PP; TcmaxT_{\rm c}^{\rm max} is 1.6 times higher than TcT_{\rm c} = 0.9 K at PP = 0.

Pressure dependence of the Kondo hybridization. To probe the character of ff electrons as a function of pressure, we use the In(1) 115In-NQR frequency νQ\nu_{\rm Q}. In general, νQ\nu_{\rm Q} is determined by the surrounding lattice and on-site electrons with the latter being dominant in strongly correlated electron systems ZhengJPSJ ; in the current case the latter reflects ff-cc hybridization that generates an electric field gradient (EFG) at the In(1)-site, as was found in previous 115In-NQR and NMR studies on CeIn3 KawasakiCeIn3 and CeRhIn5 Curro under pressure. Figures 4a, 4b, and 4c show the pressure dependence of the In(1) NQR spectrum and νQ\nu_{\rm Q} (see Supplementary Note 2 and Supplementary Figure 2 for detail). In general, νQ\nu_{\rm Q} is expected to increase smoothly with decreasing volume KawasakiCeIn3 ; however, this is not the case here. At TT = 4.2 K, νQ\nu_{\rm Q} weakly increases up to PP = 1.24 GPa but jumps at PP^{\rm*} = 1.35 GPa above which the slope dνQ/dPd\nu_{\rm Q}/dP increases by more than a factor of three. The same trend is found at lower temperatures where we see that PP^{\rm*} decreases as TT is reduced. We denote the midpoint of the νQ\nu_{\rm Q} jump in the PP-TT plane as (PP^{\rm*}, TT^{\rm*}) = (1.35 GPa, 4.2 K), (1.13 GPa, 1.6 K), and (0.90 GPa, 0.3 K), respectively. The same result is also found in the In(2) site. Figures 4d and 4e show the temperature dependence of the In(2) NQR spectrum and its peak position together with In(1) νQ\nu_{\rm Q} at PP = 1.12 GPa. As found in the In(1) site, the In(2) νQ\nu_{\rm Q} increases below TT^{*} = 1.5 K. In a previous study on CeRhIn5, a similar change of EFG was found in the In(2) site at PP^{\rm*} = 1.75 GPa, although detailed pressure dependence for the In(1) site with a high accuracy was not conducted Curro .

We consider the origin for the νQ\nu_{\rm Q} jump. Figure 5a shows results of X-ray diffraction measurements that give the pressure dependence of aa- and cc-axis lengths; the lattice parameters decrease smoothly with increasing pressure to PP = 4 GPa, without any signature of a structural transition. From these data, we calculate the EFG using the first-principles Hiroshima Linear-Augmented-Plane-Wave (HiLAPW) codes Oguchi . As expected, the calculated νQHiLAPW\nu_{\rm Q}^{\rm HiLAPW} increases monotonically with applied pressure (Fig. 5b). This and the lack of any anomaly in 1/T1T_{1} at (P,TP^{\rm*},T^{\rm*}) rule out a change in lattice contribution to νQ\nu_{\rm Q} as the origin of the νQ\nu_{\rm Q} jump. Furthermore, Fig. 5b compares the pressure dependence of νQHiLAPW\nu_{\rm Q}^{\rm HiLAPW} for CeRh(Ir)In5 and LaRh(Ir)In5. νQHiLAPW\nu_{\rm Q}^{\rm HiLAPW} for CeRh(Ir)In5 is uniformly greater than that for LaRh(Ir)In5, because CeRh(Ir)In5 has the additional EFG from hybridized ff electrons, unlike non-magnetic LaRh(Ir)In5. This result is consistent with previous band calculations Harima ; CurroBandCalc .

We conclude from these results that the jump in νQ(P)\nu_{\rm Q}(P) at (PP^{\rm*}, TT^{\rm*}) is due to a pronounced increase in Kondo hybridization at (PP^{\rm*}, TT^{\rm*}) and that the larger dνQ/dPd\nu_{\rm Q}/dP above (PP^{\rm*}, TT^{\rm*}) reflects that increased hybridization. Because increased Kondo hybridization transfers ff spectral weight from localized to itinerant degrees of freedom, and hence an increase in Fermi surface volume, the pronounced jump in νQ\nu_{\rm Q} signals the experimental observation of a small (localized) to large (itinerant) Fermi surface. As the principal axis of the EFG at the In(2) site is perpendicular to that of the In(1) site, the νQ\nu_{\rm Q} change at both sites suggests that the entire Fermi-surface volume changes at (PP^{*}, TT^{*}).

Determination of the antiferromagnetic QCP. The temperature dependence of T1T_{1} just above TT \geq Tc(P)T_{\rm c}(P) can be described by the self-consistent renormalization (SCR) theory for spin fluctuations around an antiferromagnetic QCP Moriya . A three-dimensional antiferromagnetic spin fluctuation model is applicable also to the low temperature thermopower S/TS/T around the pressure-induced antiferromagnetic QCP of CeRh0.58Ir0.42In5 Luo . Near an antiferromagnetic QCP, the SCR model predicts 1/T1Tχ𝐐(T)=1/T+θ1/T_{1}T\propto\sqrt{\chi_{\rm{\bf Q}}(T)}=1/\sqrt{T+\theta} Moriya where χ𝐐(T)\chi_{\rm{\bf Q}}(T) is the Curie-Weiss staggered susceptibility and θ\theta is a measure of the distance to the antiferromagnetic QCP. At the QCP, θ=0\theta=0 and χ𝐐(T)\chi_{\rm{\bf Q}}(T) diverges toward 0 K. 1/T1T1/T_{1}T can be represented by the sum of magnetic and small non-magnetic contributions as 1/T1T1/T_{1}T = 1/(T1T)AF+1/(T1T)lattice1/(T_{1}T)_{\rm AF}+1/(T_{1}T)_{\rm lattice}. For the lattice term, we use the mean value of 1/T1T1/T_{1}T from reference materials LaRhIn5 and LaIrIn5 (Fig. 3b), which gives 1/(T1T)lattice1/(T_{1}T)_{\rm lattice} = 1.44 (s-1K-1).

Figures 6a and 6b are plots of 1/T1T1/T_{1}T vs. TT for the antiferromagnetic phase below PP = 1.12 GPa and for the paramagnetic phase above PP = 1.24 GPa, respectively. The solid curves in Fig. 6b are least-squares fits to 1/T1TT_{1}T = a/(T+θ)0.5a/(T+\theta)^{0.5} + 1.44 just above Tc(P)T_{\rm c}(P), with aa and θ\theta as parameters. Approaching the antiferromagnetic QCP from PP = 2.53 GPa, θ\theta decreases with decreasing pressure, as can be seen in Fig. 6b. The pressure dependences of θ\theta, TNT_{\rm N}, and TcT_{\rm c} are plotted in Fig. 6c. From a linear fit of θ(P)\theta(P), the antiferromagnetic QCP (θ\theta = 0 K) for CeRh0.5Ir0.5In5 is obtained at PcAFP_{\rm c}^{\rm AF} = 1.2 GPa, where the highest TcT_{\rm c} is realized. The present results clearly indicate that spin fluctuations play a significant role for superconductivity in CeRh0.5Ir0.5In5.

Unconventional superconductivity at PcAFP_{\rm c}^{\rm AF}. As seen in Figs. 6a and 6b, there is a strong pressure dependence of the magnitude of 1/T1TT_{1}T at the lowest temperatures of these measurements. To place these results in perspective, we normalize 1/T1TT_{1}T by its value at TcT_{\rm c}, 1/T1T(Tc)T_{1}T(T_{\rm c}), and plot the ratio in Figs. 7a and 7b as a function of reduced temperature T/Tc(P)T/T_{\rm c}(P). Deep in the superconducting state, this ratio is clearly largest at PP = 1.12 GPa near PcAFP_{\rm c}^{\rm AF} and depends only weakly on temperature below TcT_{\rm c}. This result contrasts with expectations for a fully gapped, e.g., ss-wave, superconductor where 1/T1T_{1} should decrease exponentially to a very small value well below TcT_{\rm c} and for a clean dd-wave superconductor where 1/T1T_{1} decreases as T3T^{3}. In CeRh0.5Ir0.5In5 at PP = 1.12 GPa there must be a substantial fraction of low-lying excitations in the normal state that remains ungapped below TcT_{\rm c}. Namely, [T1TT_{1}T(TT=0.3K)]0.5]^{-0.5}/[T1T(Tc)]0.5T_{1}T(T_{\rm c})]^{-0.5} = N(EF)residualN(E_{\rm F})^{\rm residual}/N(EF)normalN(E_{\rm F})^{\rm normal} is the relative density of state (DOS) at TT = 0.3 K, which is consistent with the fraction of ungapped quasiparticle DOS in the superconducting state.

To obtain the relative DOS, for simplicity we assume that T1T_{1} below TcT_{\rm c} is predominantly determined by itinerant quasi particles. Previously, an analysis within the context of two-fluid phenomenological theory has deduced that the 4ff local moments also contribute to relaxation PinesPNAS ; Curro2fluid . Nonetheless, as shown in Supplementary Note 3 and Supplementary Figures 3 and 4, such a model reproduces essentially the same result as we obtained here.

Phase Diagram. Figure 8a shows the pressure dependence of TNT_{\rm N} and TT^{\rm*} inferred from In(1) NQR, together with TcT_{\rm c} on the PP-TT plane. TT^{\rm*} inferred from In(2) NQR at PP = 1.12 GPa coincides with the result obtained from In(1). TT^{\rm*} extrapolates to zero at PcP_{\rm c}^{\rm*} = 0.8 GPa, which is distinctly smaller than PcAFP_{\rm c}^{\rm AF} = 1.2 GPa. In CeRhIn5 (xx = 0), a similar result was suggested with PcP_{\rm c}^{\rm*} = 1.75 GPa Curro and PcAFP_{\rm c}^{\rm AF} = 2.1 GPa Yashima2007 . Comparison of these critical pressure values shows that Ir substitution with xx = 0.5 (chemical pressure) is effectively equivalent to the application of a physical pressure of about PP = 1 GPa to CeRhIn5. This conclusion is consistent with that drawn from our measurement of the hyperfine coupling presented earlier (Supplementary Note 1 and Supplementary Figure 1). We emphasize again that, in the pressure regions we are interested for CeRhIn5 and CeRh0.5Ir0.5In5, A115(1){}^{115}A(1) is constant, and thus changes in physical properties are not related to a changing hyperfine coupling but to the quantum criticality.

As shown in Fig. 8b, in CeRh0.5Ir0.5In5, remarkably, the fraction of ungapped excitations strongly depends on pressure, reaching a maximum at the antiferromagnetic QCP. In the coexistent state at PP = 0 Yamaguchi , this fraction is 50 %, but increases to 96% at PcAFP_{\rm c}^{\rm AF} and then decreases to 55% with the increasing pressure at PP = 2.53 GPa. The highest TcT_{\rm c} around the antiferromagnetic QCP of CeRh0.5Ir0.5In5 is realized unexpectedly with the largest fraction of gapless excitations. The present observation is completely different from that for CeRhIn5 under pressure; in CeRhIn5, the relative fraction of gapless excitations (88%) in the coexistent state is rapidly suppressed to almost zero as it approaches the QCP Yashima2007 .

I Discussion

From the phase diagram shown in Fig. 8, it is clear that the localized to itinerant transition (T(P)T^{\rm*}(P)) does not occur exactly at the antiferromagnetic QCP; in the limit of zero temperature, PcP_{\rm c}^{\rm*} precedes PcAFP_{\rm c}^{\rm AF}. One possibility would be that the Fermi-surface change across the T(P)T^{\rm*}(P) boundary marks a line of abrupt changes of the Ce valence that terminates near TT = 0 in a critical end point. A model that considers this possibility, however, appears to exclude the T(P)T^{\rm*}(P) boundary from extending into the antiferromagnetic state MiyakeWatanabe ; MiyakeWatanabe2014 , contrary to our results. In contrast, a breakdown of the Kondo effect gives rise to a small to large Fermi surface change across T(P)T^{\rm*}(P) QSiT1 . This idea leads to a TT = 0 phase diagram QSiPhysB ; ColemanJLTP similar to the results of Fig. 8. Associated with the Kondo breakdown and development of a large Fermi surface, soft charge fluctuations can emerge without a change in formal valence of Ce ions Komijani . Within experimental uncertainty of ±\pm1.5 %, there is no detectable difference between CeRhIn5 and CeIrIn5 at 10 K in their spectroscopically determined Ce valence Sundermann , even though their Fermi volumes differ - a result that, together with the phase diagram of Fig. 8, supports a Kondo breakdown interpretation as do thermopower measurements on CeRh0.58Ir0.42In5 Luo . Notably, the TT^{\rm*} boundary has no notable effect on the evolution of Tc(P)T_{\rm c}(P). In passing, we mention a possibility of a more general case, a Lifshitz transition. Watanabe and Ogata Ogata , and Kuramoto etet alal Kuramoto pointed out that, even though the Kondo screening remains, a competition between the Kondo effect and the RKKY interaction can lead to a topological Fermi surface transition (Lifshitz transition) below the antiferromagnetic QCP Ogata ; Kuramoto , which is also consistent with our observation.

The large 1/T1T1/T_{1}T below TcT_{\rm c} at pressures near the antiferromagnetic QCP (Figs. 6 and 7) is quite remarkable, and its temperature independence implies a large DOS that remains ungapped in the superconducting state even though this is the pressure range where TcT_{\rm c} is a maximum. Such an anomalously high value of ungapped DOS in the superconducting state has never been found in other QCP materials such as high TcT_{\rm c} cuprates Asayama and the iron-pnictide superconductors Zhou ; Oka , even though magnetic fluctuations are also strong around their QCP. For the Ce115 family, no significant gapless excitations have been observed so far around a QCP Kohori .

In general, gapless excitations are expected from impurity scattering in dd-wave superconductors MiyakeVarma ; Hirschfeld ; Maebashi ; Haas . Though there are no intentionally added impurities in our crystal, the random replacement of 50 % Rh by Ir results in a broadening of the In(1) NQR line by a factor of \sim 5 compared to CeIrIn5 Yamaguchi ; ZhengIr . Such randomness increases the resistivity at 4 K (just above TNT_{\rm N}) from \sim 4 μΩ\mu\Omegacm in CeRhIn5 to over 20 μΩ\mu\Omegacm in CeRh0.5Ir0.5In5 Nicklas . Quantum critical fluctuations can further enhance that scattering Maebashi to make part of a multi-sheeted Fermi surface gapless Barzykin . Such scattering concomitantly leads to pair breaking, resulting in a large reduction of TcT_{\rm c} Abrikosov ; MiyakeVarma ; Hirschfeld ; Maebashi ; Haas ; Barzykin , which is inconsistent with our observations. The relative DOS at the pressure-induced QCP is almost zero in CeRhIn5 Yashima2007 but is enhanced to 96% in CeRh0.5Ir0.5In5 at PcAFP_{\rm c}^{\rm AF}. If we assume that the symmetry of CeRh0.5Ir0.5In5 is also dd-wave, TcT_{\rm c} should be reduced to zero with such a significant residual DOS at EFE_{\rm F} Maki . In contrast to this expectation, the maximum TcT_{\rm c} = 1.4 K for CeRh0.5Ir0.5In5 remains at 61% of TcT_{\rm c} = 2.3 K for CeRhIn5 at their respective QCPs. Hence, the present results suggest that superconductivity near PcAFP_{\rm c}^{\rm AF} in CeRh0.5Ir0.5In5 is more exotic than dd-wave.

Model calculations of superconductivity in a two-dimensional Kondo lattice show that near an antiferromagnetic QCP dd- and pp-wave spin-singlet superconducting states are nearly degenerate, with an odd frequency pp-wave spin-singlet state being favored when entering the large Fermi surfaces region to take advantage of the nesting condition with the vector 𝐐{\bf Q} = (π\pi,π\pi) Otsuki . A pp-orbital wave function with spin-singlet pairing symmetry satisfies Fermi statistics in the odd-frequency channel Berezinskii ; Coleman ; Balatsky ; Fuseya , and this odd-frequency pairing is more robust against non-magnetic impurity scattering than even-frequency pairing Yoshioka . Indeed, for a scattering strength that kills dd-wave superconductivity completely, spin-singlet odd-frequency pairing will survive with a TcT_{\rm c} that is approximately 60 % of that in the absence of scattering Yoshioka . Motivated by these theoretical results, we suggest that odd-frequency spin-singlet pairing is realized in CeRh0.5Ir0.5In5 in the vicinity of its critical pressures. Its robust TcT_{\rm c} in the presence of substantial disorder scattering that gives rise to a large residual density of states at PcAFP_{\rm c}^{\rm AF} where quantum critical fluctuations are strongest and the presence of a nearby change from small to large Fermi surface at PcP_{\rm c}^{*} are fully consistent with our proposal. We stress that the unique aspect of both strong fluctuations and large Fermi surface is not shared by the end members, CeIrIn5 or CeRhIn5. Knight shift and experiments that directly probe the gap symmetry will be useful to test this possibility.

In summary, we reported systematic 115In-NQR measurements on the heavy fermion antiferromagnetic superconductor CeRh0.5Ir0.5In5 under pressure and find that an antiferromagnetic QCP is located at PcAFP_{\rm c}^{\rm AF} = 1.2 GPa, at which Tc(P)T_{\rm c}(P) reaches its maximum. The pressure and temperature dependence of νQ\nu_{\rm Q} reveal a pronounced increase in hybridization that signals a change from small to large Fermi surface in the limit of zero temperature at PcP_{\rm c}^{\rm*} = 0.8 GPa which is notably lower than PcAFP_{\rm c}^{\rm AF}. Thus, our work sheds new light on the quantum phase transition in ff-electron systems. There is a strong enhancement of the quasiparticle DOS in the superconducting state around PcAFP_{\rm c}^{\rm AF} where the Fermi surface is large. The robustness of TcT_{\rm c} under these conditions can be understood if the superconductivity is odd-frequency pp-wave spin singlet. Traditionally, Hall coefficient and quantum oscillation experiments have been used to probe the Fermi surface change. Our work demonstrates that the NQR frequency can be used as a powerful tool to examine the change in Fermi surface volume for heavy electron systems. In particular, the NQR technique does not require single crystals and is not limited by sample quality or pressure, and thus will open a new venue to understand strongly correlated electron superconductivity.

Methods

Samples.

Single crystals of CeRh0.5Ir0.5In5 were grown from an In flux as reported in a previous study Pagliuso . All experiments were performed with the same batch of crystals used in the previous NQR study Yamaguchi . As documented in detail in Ref. Yamaguchi , there is no phase separation into Rh-rich and Ir-rich parts even in the coarsely crushed powder. In fact, no excess peaks in the NMR/NQR spectrum are found and the spectrum can be reproduced by a Gaussian function. Moreover, T1T_{1} is of single component, which also indicates that Ir is randomly distributed. For NMR Knight shift measurements, two single crystals, sized 2 mm - 4 mm - 0.5 mm and 2 mm - 3 mm - 0.4 mm, were used. For NQR measurements, the crystals were moderately crushed into grains to allow RF pulses to penetrate easily into the samples. Small and thin single-crystal platelets cleaved from an as-grown ingot were used for X ray and dc-susceptibility measurements.

NQR measurement.

For NQR, the nuclear spin Hamiltonian can be expressed as, Q\mathcal{H}_{\rm Q} = (hνQ/6)[3Iz2I(I+1)+η(Ix2Iy2)](h\nu_{\rm Q}/6)[3{I_{z}}^{2}-I(I+1)+\eta({I_{x}}^{2}-{I_{y}}^{2})], where hh is Planck’s constant and II = 9/2 for the In nucleus is the nuclear spin; νQ\nu_{\rm Q} and the asymmetry parameter η\eta are defined as νQ\nu_{\rm Q} = 3eQVzz2I(2I1)h\frac{3eQV_{zz}}{2I(2I-1)h} and, η\eta == VxxVyyVzz\frac{V_{xx}-V_{yy}}{V_{zz}}, respectively, and QQ and VαβV_{\alpha\beta} are the nuclear quadrupole moment and EFG tensor, respectively. In CeRh0.5Ir0.5In5, there are two inequivalent In sites, one in the CeIn [In(1)] layer and another in the (Rh,Ir)In4 [In(2)] layer (see Fig. 2a). The principle axis of the EFG at the In(1) [In(2)] site is parallel [perpendicular] to the cc axis. The 115In-NQR spectra for In(1) ±\pm1/2 \leftrightarrow ±\pm3/2 transition line (Supplementary Note 2 and Supplementary Figure 2), ±\pm7/2 \leftrightarrow ±\pm9/2 transition line (Supplementary Figure 5), the In(2) ±\pm3/2 \leftrightarrow ±\pm5/2 transition line and T1T_{1} for the In(1) site (η=0\eta=0) (Supplementary Note 4 and Supplementary Figures 6 and 7) were obtained as reported in an earlier study Yamaguchi ; recovery . Here, T1T_{1} probes the dynamic spin susceptibility through the hyperfine coupling constant A𝐪A_{\bf q} as 1/T1T𝐪|A𝐪|2χ′′(𝐪,ω0)/ω01/T_{1}\propto T\sum_{\bf q}\left|A_{\bf q}\right|^{2}\chi^{\prime\prime}({\bf q},\omega_{0})/\omega_{0}, where ω0\omega_{0} is the NQR frequency MoriyaJPSJ and 𝐪{\bf q} is a wave vector for antiferromagnetic order and/or quantum critical fluctuation in CeRh0.5Ir0.5In5. Meanwhile, 1/T11/T_{1} \propto N(EF)2N(E_{\rm F})^{2}kBTk_{\rm B}T holds in a Pauli paramagnetic metal, i.e., in a heavy Fermi liquid state (Korringa law). Here, N(EF)N(E_{\rm F}) is the density of states at EFE_{\rm F}.

A NiCrAl/BeCu piston-cylinder-type pressure-cell filled with Daphne (7474) oil was used. The TcT_{\rm c} of Sn was measured to determine the pressure. Measurements below 1 K were performed using a 3He refrigerator.

X-ray measurement.

A diamond anvil cell filled with a CeRh0.5Ir0.5In5 single crystal and Daphne oil were used for room temperature X-ray measurements under pressure; the pressure was determined by measuring the fluorescence of ruby. All measurements were made at zero magnetic field.

EFG calculation.

The EFG is calculated using the all electron full-potential linear augmented plane wave method implemented in the Hiroshima Linear-Augmented-Plane-Wave (HiLAPW) code with generalized gradient approximation including spin-orbit coupling Oguchi .

Hyperfine coupling constant.

To estimate the hyperfine coupling constant for CeRh0.5Ir0.5In5, A115(1){}^{115}A(1), we measure the 115In(1)-NMR spectrum and dc susceptibility. For NMR, the nuclear spin Hamiltonian is expressed as \mathcal{H} = 115-^{115}γ𝐈𝐇\gamma\hbar\bf{I}\cdot\bf{H}(1+K)+Q(1+K)+\mathcal{H}_{\rm Q}, where the gyromagnetic ratio γ115{}^{115}\gamma = 9.3295 MHz T-1, 𝐇\bf{H} is the external magnetic field, and KK is the Knight shift. 115In-NMR spectra have nine transitions from IzI_{\rm z} = (2mm++1)/2 to (2mm-1)/2 where mm = -4, -3, -2, -1, 0, 1, 2, 3, 4 for In(1) and In(2) sites, respectively, with KK, νQ\nu_{\rm Q}, and η\eta as parameters. At ambient pressure, νQ\nu_{\rm Q} and η\eta at the In(1) [In(2)] site are 6.35 MHz (17.37 MHz) and 0 (0.473), respectively Yamaguchi . The Knight shift for In(1), K115(1)c(T){}^{115}K(1)_{\rm c}(T), was calculated from the peak in the 115In(1)-NMR center line (mm = 0) taken by sweeping the RF frequency at a fixed field HH = 12.950 T and χc\chi_{\rm c} (emu mol-1) is obtained by dc susceptibility measurements at HH = 3 T using a superconducting quantum interference device (SQUID) with the vibrating sample magnetometer (VSM). The magnetic field HH is fixed perpendicular to the CeIn layer (cc axis).

References

  • (1) Mathur, N. D., Grosche, F. M., Julian, S. R., Walker, I. R., Freye, D. M., Haselwimmer, R. K. W. & Lonzarich, G. G. Magnetically mediated superconductivity in heavy fermion compounds. Nature 394, 39-43 (1998).
  • (2) Löhneysen, H. v., Pietrus, T., Portisch, G., Schlager, H. G., Schröder, A., Sieck, M. & Trappmann, T. Non-Fermi-liquid behavior in a heavy-fermion alloy at a magnetic instability. Phys. Rev. Lett. 72, 3262-3265 (1994).
  • (3) Abanov, Ar., Chubukov, A. V. & Schmalian, J. Quantum-critical theory of the spin-fermion model and its application to cuprates: Normal state analysis. Adv. Phys. 52, 119-218 (2003).
  • (4) Abrahams, E. & Si, Q. Quantum criticality in the iron pnictides and chalcogenides. J. Phys.: Condens. Matter 23 223201 (2011) .
  • (5) Zhou, R., Li, Z., Yang, J., Sun, D. L., Lin, C. T. & Zheng, Guo-qing. Quantum Criticality in Electron-Doped BaFe2-xNixAs2. Nat. Commun. 4, 2265 (2013).
  • (6) Luo, J., Yang, J., Zhou, R., Mu, Q. G., Liu, T., Ren, Zhi-an., Yi, C. J., Shi, Y. G. & Zheng, Guo-qing. Tuning the Distance to a Possible Ferromagnetic Quantum Critical Point in A2Cr3As3. Phys. Rev. Lett. 123, 047001 (2019).
  • (7) Hertz, J. A. Quantum Critical Phenomena. Phys. Rev. B 14, 1165-1184 (1976).
  • (8) Moriya, T. Spin Fluctuations in Itinerant Electron Magnetism, Springer Series in Solid-State Sciences (Springer, Berlin), Vol 56 (1985).
  • (9) Millis, A. J. Effect of a nonzero temperature on quantum critical points in itinerant fermion systems. Phys. Rev. B 48, 7183-7196 (1993).
  • (10) Si, Q. & Steglich, F. Heavy Fermions and Quantum Phase Transitions. Science 329, 1161-1166 (2010).
  • (11) Doniach, S. The Kondo lattice and weak antiferromagnetism. Phys. B Condens. Matter 91, 231–234 (1977).
  • (12) Chen, Q. Y., Xu, D. F., Niu, X. H., Jiang, J., Peng, R., Xu, H. C., Wen, C. H. P., Ding, Z. F., Huang, K., Shu, L., Zhang, Y. J., Lee, H., Strocov, V. N., Shi, M., Bisti, F., Schmitt, T., Huang, Y. B., Dudin, P., Lai, X. C., Kirchner, S., Yuan, H. Q. & Feng, D. L. Direct Observation of How the Heavy-Fermion State Develops in CeCoIn5. Phys. Rev. B 96, 045107 (2017).
  • (13) Chen, Q. Y., Xu, D. F., Niu, X. H., Peng, R., Xu, H. C., Wen, C. H. P., Liu, X., Shu, L., Tan, S. Y., Lai, X. C., Zhang, Y. J., Lee, H., Strocov, V. N., Bisti, F., Dudin, P., Zhu, J. -X., Yuan, H. Q., Kirchner, S. & Feng, D. L. Band Dependent Interlayer ff-Electron Hybridization in CeRhIn5. Phys. Rev. Lett. 120, 066403 (2018).
  • (14) Burdin, S., Georges, A. & Grempel, D. R. Coherence Scale of the Kondo Lattice. Phys. Rev. Lett. 85, 1048-1051 (2000).
  • (15) Si, Q., Rabello, S., Ingersent, K. & Smith, J. L. Locally critical quantum phase transitions in strongly correlated metals. Nature 413, 804-808 (2001).
  • (16) Coleman, P. & Nevidomskyy, A. H. Frustration and the Kondo Effect in Heavy Fermion Materials. J. Low Temp. Phys. 161, 182-202 (2010).
  • (17) Löhneysen, H. v., Rosch, A., Vojta, M. & Wölfle, P. Fermi-Liquid Instabilities at Magnetic Quantum Phase Transitions. Rev. Mod. Phys. 79, 1015-1075 (2007).
  • (18) Pagliuso, P. G., Petrovic, C., Movshovich, R., Hall, D., Hundley, M. F., Sarrao, J. L., Thompson, J. D. & Fisk, Z. Coexistence of magnetism and superconductivity in CeRh1-xIrxIn5. Phys. Rev. B 64, 100503(R) (2001).
  • (19) Zheng, G. -q., Yamaguchi, N., Kan, H., Kitaoka, Y., Sarrao, J. L., Pagliuso, P. G., Moreno, N. O. & Thompson, J. D. Coexistence of antiferromagnetic order and unconventional superconductivity in heavy-fermion CeRh1-xIrxIn5 compounds: Nuclear quadrupole resonance studies. Phys. Rev. B 70, 014511 (2004).
  • (20) Shishido, H., Settai, R., Harima, H. & Ōnuki, Y. A Drastic Change of the Fermi Surface at a Critical Pressure in CeRhIn5: dHvA Study under Pressure. J. Phys. Soc. Jpn 74, 1103-1106 (2005).
  • (21) Hegger, H., Petrovic, C., Moshopoulou, E. G., Hundley, M. F., Sarrao, J. L., Fisk, Z. & Thompson, J. D. Pressure-Induced Superconductivity in Quasi-2D CeRhIn5. Phys. Rev. Lett. 84, 4986-4989 (2001).
  • (22) Park, T., Ronning, F., Yuan, H. Q., Salamon, M. B., Movshovich, R., Sarrao, J. L. & Thompson, J. D. Hidden Magnetism and Quantum Criticality in the Heavy Fermion Superconductor CeRhIn5. Nature 440, 65-68 (2006).
  • (23) Knebel, G., Aoki, D., Braithwaite, D., Salce, B. & Flouquet, J. Coexistence of Antiferromagnetism and Superconductivity in CeRhIn5 under High Pressure and Magnetic Field. Phys. Rev. B 74, 020501(R) (2006).
  • (24) Mito, T., Kawasaki, S., Zheng, G. -q., Kawasaki, Y., Ishida, K., Kitaoka, Y., Aoki, D., Haga, Y. & Ōnuki, Y. Pressure-induced anomalous magnetism and unconventional superconductivity in CeRhIn5 : 115In-NQR study under pressure. Phys. Rev. B 63, 220507(R) (2001).
  • (25) Park, T., Bauer, E. D. & Thompson, J. D. Probing the Nodal Gap in the Pressure-Induced Heavy Fermion Superconductor CeRhIn5. Phys. Rev. Lett. 101, 177002 (2008).
  • (26) Shishido, H., Settai, R., Aoki, D., Ikeda, S., Nakawaki, H., Nakamura, N., Iizuka, T., Inada, Y., Sugiyama, K., Takeuchi, T., Kindo, K., Kobayashi, T. C., Haga, Y., Harima, H., Aoki, Y., Namiki, T., Sato, H. & Ōnuki, Y. Fermi Surface, Magnetic and Superconducting Properties of LaRhIn5 and CeTIn5 (T: Co, Rh and Ir). J. Phys. Soc. Jpn. 71, 162-173 (2002).
  • (27) Zheng, G. -q., Tanabe, K., Mito, T., Kawasaki, S., Kitaoka, Y., Aoki, D., Haga, Y. & Ōnuki, Y. Unique Spin Dynamics and Unconventional Superconductivity in the Layered Heavy Fermion Compound CeIrIn5: NQR Evidence. Phys. Rev. Lett. 86, 4664-4667 (2001).
  • (28) Lu, X., Lee, H., Park, T., Ronning, F., Bauer, E. D. & Thompson, J. D. Heat-Capacity Measurements of Energy-Gap Nodes of the Heavy-Fermion Superconductor CeIrIn5 Deep inside the Pressure-Dependent Dome Structure of Its Superconducting Phase Diagram. Phys. Rev. Lett. 108, 027001 (2012).
  • (29) Petrovic, C., Movshovich, R., Jaime, M., Pagliuso, P. G., Hundley, M. F., Sarrao, J. L., Fisk, Z. & Thompson, J. D. A new heavy-fermion superconductor CeIrIn5: A relative of the cuprates ? Euro. Phys. Lett. 53, 354-359 (2001).
  • (30) Kawasaki, S., Zheng, G. -q., Kan, H., Kitaoka, Y., Shishido, H. & Ōnuki, Y. Enhancing the Superconducting Transition Temperature of the Heavy Fermion Compound CeIrIn5 in the Absence of Spin Correlations. Phys. Rev. Lett. 94, 037007 (2005).
  • (31) Christianson, A. D., Llobet, A., Bao, W., Gardner, J. S., Swainson, I. P., Lynn, J. W., Mignot, J. -M., Prokes, K., Pagliuso, P. G., Moreno, N. O., Sarrao, J. L., Thompson, J. D. & Lacerda, A. H. Novel Coexistence of Superconductivity with Two Distinct Magnetic Orders. Phys. Rev. Lett. 95, 217002 (2005).
  • (32) Nicklas, M., Sidorov, V. A., Borges, H. A., Pagliuso, P. G., Sarrao, J. L. & Thompson, J. D. Two superconducting phases in CeRh1-xIrxIn5. Phys. Rev. B 70, 020505(R) (2004).
  • (33) Willers, T., Strigari, F., Hu, Z., Sessi, V., Brookes, N. B., Bauer, E. D., Sarrao, J. L., Thompson, J. D., Tanaka, A., Wirth, S., Tjeng, L. H. & Severing, A. Correlation Between Ground State and Orbital Anisotropy in Heavy Fermion Materials. Proc. Natl. Acad. Sci. (USA) 112, 2384-2388 (2015).
  • (34) Lin, C. H., Shirer, K. R., Crocker, J., Dioguardi, A. P., Lawson, M. M., Bush, B. T., Klavins, P. & Curro, N. J. Evolution of hyperfine parameters across a quantum critical point in CeRhIn5. Phys. Rev. B 92, 155147 (2015).
  • (35) Curro, N. J., Young, B. -L., Schmalian, J., & Pines, D. Scaling in the emergent behavior of heavy-electron materials. Phys. Rev. B 70, 235117 (2004).
  • (36) Kambe, S., Tokunaga, Y., Sakai, H., Chudo, H., Haga, Y., Matsuda, T. D., & Walstedt, R. E. One-component description of magnetic excitations in the heavy-fermion compound CeIrIn5. Phys. Rev. B 81, 140405(R) (2010).
  • (37) Zheng, G. -q., Kawasaki, S., Yamaguchi, N., Tanabe, K., Mito, T., Kan, H., Kitaoka, Y., Haga, Y., Aoki, D., Ōnuki, Y., Sarrao, J. L. & Thompson, J. D. Magnetism and superconductivity in CeRhIn5 under chemical and hydrostatic pressures. Phys. B 329-333, 450-451 (2003).
  • (38) Zheng, G.-q., Kitaoka, Y., Ishida, K., & Asayama, K. Local Hole Distribution in the CuO2 Plane of High-TcT_{\rm c} Cu-Oxides Studied by Cu and Oxygen NQR/NMR. J. Phys. Soc. Jpn. 64, 2524-2532 (1995).
  • (39) Kawasaki, S., Yashima, M., Kitaoka, Y., Takeda, K., Shimizu, K., Oishi, Y., Takata, M., Kobayashi, T. C., Harima, H., Araki, S., Shishido, H., Settai, R. & Ōnuki, Y. Pressure-induced unconventional superconductivity in the heavy-fermion antiferromagnet CeIn3: An 115In-NQR study under pressure. Phys. Rev. B 77, 064508 (2008).
  • (40) Oguchi, T. Electronic Band Structure and Structural Stability of LaBiPt. Phys. Rev. B 63, 125115 (2001).
  • (41) Betsuyaku, K. & Harima, H. Electronic structure and electric field gradient of RIn3 and RTIn5 (R=La and Ce, T= Co, Rh and Ir). J. Magn. Magn. Mater. 272-276, 187-188 (2004).
  • (42) Rusz, J., Oppeneer, P. M., Curro, N. J., Urbano, R. R., Young, B.-L., Lebègue, S., Pagliuso, P. G., Pham, L. D., Bauer, E. D., Sarrao, J. L. & Fisk, Z. Probing the electronic structure of pure and doped CeMMIn5 (MM=Co,Rh, Ir) crystals with nuclear quadrupolar resonance. Phys. Rev. B 77, 245124 (2008).
  • (43) Moriya, T. & Ueda, K. Antiferromagnetic spin fluctuation and superconductivity. Rep. Prog. Phys. 66, 1299-1341 (2003).
  • (44) Luo, Y. K., Lu, X., Dioguardi, A. P., Rosa, P. F. S., Bauer, E. D., Si, Q. & Thompson, J. D. Unconventional and conventional quantum criticalities in CeRh0.58Ir0.42In5. npj Quant. Mat. 3, 6 (2018).
  • (45) Yang, Y.-f., Urbano, R., Curro, N. J., Pines, D. & Bauer, E. D. Magnetic Excitations in the Kondo Liquid: Superconductivity and Hidden Magnetic Quantum Critical Fluctuations. Phys. Rev. Lett. 103, 197004 (2009).
  • (46) Yang, Y.-f. & Pines, D. Emergent states in heavy-electron materials. Proc. Natl. Acad. Sci. (USA) 109, E3060-E3066 (2012).
  • (47) Yashima, M., Kawasaki, S., Mukuda, H., Kitaoka, Y., Shishido, H., Settai, R. & Ōnuki, Y. Quantum phase diagram of antiferromagnetism and superconductivity with a tetracritical point in CeRhIn5 in zero magnetic field. Phys. Rev. B. 76, 020509(R) (2007).
  • (48) Watanabe, S. & Miyake, K. Origin of Drastic Change of Fermi Surface and Transport Anomalies in CeRhIn5 under Pressure. J. Phys. Soc. Jpn. 79, 033707 (2010).
  • (49) Miyake, K. & Watanabe, S. Unconventional Quantum Criticality Due to Critical Valence Transition. J. Phys. Soc. Jpn. 83, 061006 (2014).
  • (50) Si, Q., Rabello, S., Ingersent, K. & Smith, J. L. Local fluctuations in quantum critical metals. Phys. Rev. B 68, 115103 (2003).
  • (51) Si, Q. Global Magnetic Phase Diagram and Local Quantum Criticality in Heavy Fermion Metals. Phys. B 378-380, 23-27 (2006).
  • (52) Komijani, Y. & Coleman, P. Emergent Critical Charge Fluctuations at the Kondo Breakdown of Heavy Fermions. Phys. Rev. Lett. 122, 217001 (2019).
  • (53) Sundermann, M., Strigari, F., Willers, T., Weinen, J., Liao, Y. F., Tsuei, K.-D., Hiraoka, N., Ishii, H., Yamaoka, H., Mizuki, J., Zekko, Y., Bauer, E. D., Sarrao, J. L., Thompson, J. D., Lejay, P., Muro, Y., Yutani, K., Takabatake, T., Tanaka, A., Hollmann, N., Tjeng, L. H., Severing, A. Quantitative Study of the f Occupation in CeMIn5 and Other Cerium Compounds with Hard X-Rays. J. Elect. Spectr. and Rel. Phenom. 209, 1-8 (2016).
  • (54) Watanabe, H. & Ogata, M. Fermi-Surface Reconstruction without Breakdown of Kondo Screening at the Quantum Critical Point. Phys. Rev. Lett. 99, 136401 (2007).
  • (55) Kuramoto, Y. & Hoshino, S. Composite Orders and Lifshitz Transition of Heavy Electrons. J. Phys. Soc. Jpn. 83, 061007 (2014).
  • (56) Asayama, K., Kitaoka, Y., Zheng, Guo-qing. & Ishida, K. NMR studies of high TcT_{\rm c} superconductors. Prog. Nucl. Magn. Reson. Spectrosc. 28, 221-253 (1996).
  • (57) Oka, T., Li, Z., Kawasaki, S., Chen, G. F., Wang, N. L. & Zheng, Guo-qing. Antiferromagnetic Spin Fluctuations above the Dome-Shaped and Full-Gap Superconducting States of LaFeAsO1-xFx Revealed by 75As-Nuclear Quadrupole Resonance. Phys. Rev. Lett. 108, 047001 (2012).
  • (58) Kohori, Y., Yamato, Y., Iwamoto, Y., Kohara, T., Bauer, E. D., Maple, M. B. & Sarrao, J. L. NMR and NQR studies of the heavy fermion superconductors CeTTIn5 (TT = Co and Ir). Phys. Rev. B 64, 134526 (2001).
  • (59) Schmitt-Rink, S., Miyake, K. & Varma, C. M. Transport and Thermal Properties of Heavy-Fermion Superconductors: A Unified Picture. Phys. Rev. Lett. 57, 2575-2578 (1986).
  • (60) Hirschfeld, P., Vollhardt, D. & Wölfle, P. Resonant impurity scattering in heavy fermion superconductors. Solid. State. Comm. 59, 111-115 (1986).
  • (61) Miyake, K. & Maebashi, H. A route to small Drude weight in metals with nested spin fluctuations and enhanced impurity scattering associated with quantum critical phenomena. J. Phys. Chem. Solids. 62, 53-57 (2011).
  • (62) Haas, S., Balatsky, A. V., Sigrist, M. & Rice, T. M. Extended gapless regions in disordered dx2y2d_{x^{2}-y^{2}} wave superconductors. Phys. Rev. B 56, 5108-5111 (1997).
  • (63) Barzykin, V. & Gor’kov, L. P. Gapless Fermi surfaces in superconducting CeCoIn5. Phys. Rev. B 76, 014509 (2007).
  • (64) Abrikosov, A. & Gor’kov, L. P. Contribution to the theory of superconducting alloys with paramagnetic impurities. Sov. Phys. JETP 12, 1243 (1961).
  • (65) Puchkaryov, E. & Maki, K. Impurity scattering in d-wave superconductivity. Unitarity limit versus Born limit. Eur. Phys. J. B 4, 191-194 (1998).
  • (66) Otsuki, J. Competing dd-Wave and pp-Wave Spin-Singlet Superconductivities in the Two-Dimensional Kondo Lattice. Phys. Rev. Lett. 115, 036404 (2015).
  • (67) Balatsky, A. & Abrahams, E. New class of singlet superconductors which break time reversal and parity. Phys. Rev. B 45, 13125(R) (1992).
  • (68) Coleman, P., Miranda, E. & Tsvelik, A. Possible Realization of Odd-Frequency Pairing in Heavy Fermion Compounds. Phys. Rev. Lett. 70, 2960-2963 (1993).
  • (69) Fuseya, Y., Kohno, H. & Miyake, K. Realization of Odd-Frequency pp-Wave Spin-Singlet Superconductivity Coexisting with Antiferromagnetic Order near Quantum Critical Point. J. Phys. Soc. Jpn. 72, 2914-2923 (2003).
  • (70) Berezinskii, V. L. New model of the anisotropic phase of superfluid 3He. JETP Lett. 20, 287 (1974).
  • (71) Yoshioka, Y. & Miyake, K. Impurity Effect on Frequency Dependent Superconductivity: Odd-Frequency Pairing and Even-Frequency Pairing. J. Phys. Soc. Jpn. 81, 093702 (2012).
  • (72) MacLaughlin, D. E., Williamson, J. D. & Butterworth, J. Nuclear Spin-Lattice Relaxation in Pure and Impure Indium. I. Normal State. Phys. Rev. B 4, 60-70 (1971).
  • (73) Moriya, T. The Effect of Electron-Electron Interaction on the Nuclear Spin Relaxation in Metals. J. Phys. Soc. Jpn. 18, 516-520 (1963).

Acknowledgments

We would like to thank T. Kambe for help with the susceptibility measurement, Y. Fuseya, Y. Kuramoto, H. Kusunose, J. Otsuki, K. Miyake, S. Watanabe, T. Oguchi and Q. Si for discussion. This work was supported in part by research grants from MEXT (No. JP19K03747, JP23102717, and JP25400374), NSFC grant No. 11634015 and MOST Grant (No. 2016YFA0300300, No. 2017YFA0302904, and No. 2017YFA0303103) and, at Los Alamos, was performed under the auspices of the U.S. Department of Energy, Office of Basic Energy Sciences, Division of Materials Sciences and Engineering.

Authors contributions

G.-q.Z planned the project. J.L.S and J.D.T synthesized the single crystal. S.K, T.O, A.S, Y.K, and K.U performed NQR measurements. S.K and Y.K performed NMR and susceptibility measurements. J.G, S.C, and L.L.S performed X-ray measurement. K.M performed band structure calculation. G.-q.Z, S.K and J.D.T wrote the manuscript. All authors discussed the results and interpretation.