Kazuki Iida1,2, , Maiko Kofu1,3, Katsuhiro Suzuki4,, Naoki Murai3, Seiko Ohira-Kawamura3, Ryoichi Kajimoto3 , Yasuhiro Inamura3, Motoyuki Ishikado2, Shunsuke Hasegawa5, Takatsugu Masuda5, Yoshiyuki Yoshida6, Kazuhisa Kakurai2 , Kazushige Machida7, and Seunghun Lee1 + Affiliations1Department of Physics, University of Virginia, Charlottesville, VA 22904-4714, U.S.A.2Neutron Science and Technology Center, Comprehensive Research Organization for Science and Society (CROSS), Tokai, Ibaraki 319-1106, Japan3J-PARC Center, Japan Atomic Energy Agency (JAEA), Tokai, Ibaraki 319-1195, Japan4Research Organization of Science and Technology, Ritsumeikan University, Kusatsu, Shiga 525-8577, Japan5Neutron Science Laboratory, Institute for Solid State Physics, University of Tokyo, Kashiwa, Chiba 277-8581, Japan6National Institute of Advanced Industrial Science and Technology, Tsukuba, Ibaraki 305-8565, Japan7Department of Physics, Ritsumeikan University, Kusatsu, Shiga 525-8577, Japan
Received January 9, 2020; Accepted March 13, 2020; Published April 7, 2020
We investigated the low-energy incommensurate (IC) magnetic fluctuations in Sr2RuO4 by the high-resolution inelastic neutron scattering measurements and random phase approximation (RPA) calculations. We observed a spin resonance with energy of ħωres = 0.56 meV centered at a characteristic wavevector \(\mathbf{Q}_{\text{res}} = (0.3,0.3,0.5)\). The resonance energy corresponds well to the superconducting gap 2Δ = 0.56 meV estimated by the tunneling spectroscopy. The spin resonance shows the L modulation with a maximum at around L = 0.5. The L modulated intensity of the spin resonance and our RPA calculations indicate that the superconducting gaps regarding the quasi-one-dimensional α and β sheets at the Fermi surfaces have the horizontal line nodes. These results may set a strong constraint on the pairing symmetry of Sr2RuO4. We also discuss the implications on possible superconducting order parameters.
©2020 The Physical Society of Japan
Strontium ruthenate Sr2RuO4 with \(T_{\text{c}}=1.5\) K1) has attracted a great deal of interest as a prime candidate for the chiral p-wave superconductor;2–5) nuclear magnetic resonance (NMR)6) and polarized neutron scattering7) measurements reported spin-triplet superconductivity whereas muon spin rotation8) and Kerr effect9) measurements showed spontaneously time reversal symmetry breaking. On the other hand, there are some experimental results such as the absence of the chiral edge currents,10) first-order superconducting transition,11–14) and strong \(H_{c2}\) (\(\parallel ab\)) suppression,15) all of which challenge the chiral p-wave superconductivity. Recently, experimental and theoretical studies under an application of uniaxial pressure along \(\langle 100\rangle\) reported a factor of 2.3 enhancement of \(T_{\text{c}}\) owing to the Lifshitz transition when the Fermi level passes through a van Hove singularity, raising the possibility of an even-parity spin-singlet order parameter in Sr2RuO4.16–22) More recently, new NMR results demonstrated that the spin susceptibility substantially drops below \(T_{\text{c}}\) provided that the pulse energy is smaller than a threshold, ruling out the chiral-p spin-triplet superconductivity.23,24) As such, these recent works turn the research on Sr2RuO4 towards a fascinating new era.
So far, various experimental techniques reported that the superconducting gaps of Sr2RuO4 have line nodes,25,26) but the details of the line nodes, e.g., of the vertical or horizontal nature, are not uncovered yet. The thermal conductivity measurements reported vertical line nodes on the superconducting gaps,27) whereas field-angle-dependent specific heat capacity measurements are indicating the horizontal line nodes.28,29) Since the complete information of the superconducting gaps can shed light on the symmetry of the pairing, exclusive determination of the direction of the line nodes in Sr2RuO4 is highly desirable.
The inelastic neutron scattering (INS) technique can directly measure the imaginary part of generalized spin susceptibility (\(\chi''\)) as a function of momentum (\(\mathbf{Q}\)) and energy (\(\hbar\omega\)) transfers, yielding information on the Fermi surface topology. In addition, \(\mathbf{Q}\) dependence of a spin resonance as a consequence of the Bardeen–Cooper–Schrieffer (BCS) coherence factor can provide information on the symmetry of the superconducting gap. In Sr2RuO4, the most pronounced magnetic signal in the normal state is nearly two-dimensional incommensurate (IC) magnetic fluctuations at \(\mathbf{Q}_{\text{IC}}=(0.3, 0.3, L)\) owing to the Fermi surface nesting between (or within) the quasi-one-dimensional α and β sheets consisted of the \(d_{zx}\) and \(d_{yz}\) orbitals of Ru4+.30–33) In contrast to the marked signal from the IC magnetic fluctuations, only weak signals due to ferromagnetic fluctuations originating from the two-dimensional γ sheet (\(d_{xy}\)) are observed around the Γ point.34,35) Upon decreasing temperature below \(T_{\text{c}}\), however, no sizable spin resonance regarding the IC magnetic fluctuations36) was observed at \(\mathbf{Q}=(0.3, 0.3, 0)\) with energy close to the superconducting gap \(2\Delta=0.56\) meV33) estimated by the tunneling spectroscopy.37)
Based on the horizontal line nodes model, the spin resonance is supposed to emerge at \(\mathbf{Q}=(0.3, 0.3, L)\) with a finite L but not at \((0.3, 0.3, 0)\) because of the sign change of the superconducting gap along \(k_{z}\).29) The vertical line nodes model, on the other hand, expects the spin resonance at \(\mathbf{Q}=(0.3, 0.3, 0)\) because of the sign change within the \(k_{z}\) plane. Therefore, the horizontal line nodes model can naturally explain the absence of the spin resonance at \(\mathbf{Q}=(0.3, 0.3, 0)\), and search for the spin resonance at \((0.3, 0.3, L)\) is of particular interest. In this letter, we investigate in detail the low-energy IC magnetic fluctuations in Sr2RuO4 focusing on their L dependence using the high-resolution INS technique. The experimental results, especially the L dependence of neutron scattering intensities, are compared with the random phase approximation (RPA) calculations.
Three single crystals of Sr2RuO4 with a total mass of ∼10 g were grown by the floating-zone method,38,39) and each crystal shows the superconducting onset temperature of ∼1.35 K. They were co-aligned with the (\(HHL\)) plane being horizontal to the scattering plane and attached to a closed-cycle 3He refrigerator. We performed INS measurements at 0.3 and 1.8 K using the disk chopper time-of-flight (TOF) neutron spectrometer AMATERAS of the Materials and Life Science Experimental Facility (MLF) in Japan Proton Accelerator Research Complex (J-PARC).40) Frequency of the pulse shaping and velocity selecting disk choppers was 300 Hz, yielding the combinations of multiple incident neutron energies (\(E_{\text{i}}\)s) of 2.64, 5.93, and 23.7 meV with the energy resolutions of 0.046, 0.146, and 1.07 meV, respectively, at the elastic channel. The software suite Utsusemi41) was used to visualize the TOF neutron scattering data.
Figure 1 summarizes the overall features of high-energy (\(\hbar\omega\gg 2\Delta\)) IC magnetic fluctuations with \(L = 0.5\) in Sr2RuO4. Figure 1(a) depicts the constant-energy INS intensity map in the (\(HK0.5\)) plane with energy transfer of 2.5 meV at 0.3 K. IC magnetic fluctuations are observed at \(\mathbf{Q}_{\text{IC}}=(0.3, 0.3, 0.5)\), \((0.7, 0.3, 0.5)\), and \((0.7, 0.7, 0.5)\). In addition to the IC magnetic fluctuations, the Fermi surface nesting also induces the ridge scattering connecting equivalent \(\mathbf{Q}_{\text{IC}}\)s around \((0.5, 0.5, 0.5)\). To explore the energy evolution of the IC magnetic fluctuations with \(L=0.5\), the INS intensity is converted to the imaginary part of the spin susceptibility \(\chi''(\hbar\omega)\) via the fluctuation-dissipation theorem \(\chi''(\hbar\omega)=(1-e^{-\hbar\omega/k_{\text{B}}T})I(\hbar\omega)\) after subtracting the background. The \(\chi''(\hbar\omega)\) spectra at \(\mathbf{Q}_{\text{IC}}=(0.3, 0.3, 0.5)\) below and above \(T_{\text{c}}\) are plotted in Fig. 1(c). The \(\chi''(\hbar\omega)\) spectra above 1.0 meV are well fitted by the relaxation response model \(\chi''(\hbar\omega)=\chi'\Gamma\hbar\omega/[(\hbar\omega)^{2}+\Gamma^{2}]\) where \(\chi'\) is the static susceptibility and Γ the relaxation rate [or the peak position of \(\chi''(\hbar\omega)\)], yielding \(\Gamma=6.3(2)\) meV [6.5(2) meV] at 0.3 K (1.8 K). The observed IC magnetic fluctuations with \(L=0.5\) are quantitatively consistent with the IC magnetic fluctuations with \(L=0\) reported in the previous INS works.31–35) This is reasonable since the IC magnetic fluctuations along \((0.3, 0.3, L)\) monotonically decrease in intensity with increasing L [Fig. 1(b)], representing the quasi-two-dimensional nature of the IC magnetic fluctuations in this energy region. This is consistent with the quasi-one-dimensional band structures of the cylindrical α and β sheets.42)
Figure 1. (Color online) High-energy INS results of the IC magnetic fluctuations in Sr2RuO4. (a, b) Constant-energy INS intensity maps in the (a) (\(HK0.5\)) and (b) (\(HHL\)) planes with the energy window of \([1.5, 3.5]\) meV at 0.3 K. (c) \(\chi''(\hbar\omega)\) spectra at \(\mathbf{Q}_{\text{IC}}=(0.3, 0.3, 0.5)\) below and above \(T_{\text{c}}\). Data from different incident energies of neutrons \(E_{\text{i}}=2.64\), 5.93, and 23.7 meV (circles, triangles, and squares) are combined for each temperature after background estimated by \(\chi''(\hbar\omega)\) at \(\mathbf{Q}=(0.525, 0.525, 0.5)\) is subtracted from each spectrum. Solid lines are fitting results by the conventional relaxation response model.
Let us turn to the low-energy (\(\hbar\omega\simeq 2\Delta\)) IC magnetic fluctuations of Sr2RuO4. We found that a spin resonance appears in the superconducting state at \(\mathbf{Q}_{\text{res}}=(0.3, 0.3, 0.5)\) and \(\hbar\omega_{\text{res}} = 2\Delta\) [white arrow in Fig. 2(a)] but not at \(\mathbf{Q}=(0.3, 0.3, 0)\) [Fig. 2(c)]. The L modulated intensity of the spin resonance is observed only at \(\hbar\omega = 2\Delta\) [Fig. 2(e)]. The intensity modulation along L of the low-energy IC magnetic fluctuations indicates the presence of a weakly three-dimensional superconducting gaps. It is worth mentioning that the low-energy magnetic fluctuations at \(\mathbf{Q}=(0.7,0.7,L)\) is difficult to detect in this low-energy region because of the quickly decaying squared magnetic form factor at high Q [\(F^{2}_{\mathbf{Q}=(0.7,0.7,0.5)}/F^{2}_{\mathbf{Q}=(0.3,0.3,0.5)}\sim 1/6\)43)] and the weak intensity of the IC magnetic fluctuations at low energies [Fig. 1(c)]. For quantitative analysis on the spin resonance in Sr2RuO4, \(\hbar\omega\) and \(\mathbf{Q}\) dependencies of the INS intensity [\(I(\hbar\omega)\) and \(I(\mathbf{Q})\)] are analyzed in detail below.
Figure 2. (Color online) Low-energy INS intensity maps regarding the IC magnetic fluctuations in Sr2RuO4. (a–c) Low-energy IC magnetic fluctuations (a) along \(\mathbf{Q}=(H, H, 0.5)\) at 0.3 K, (b) along \((H, H, 0.5)\) at 1.8 K, and (c) along \((H, H, 0)\) at 0.3 K. White arrows indicate the energy of the superconducting gap \(2\Delta\) at \(\mathbf{Q}_{\text{IC}}\). (d–f) Constant-energy INS intensity maps in the (\(HHL\)) plane at 0.3 K with the energy windows of (d) \([0.23, 0.47]\) meV, (e) \([0.47, 0.65]\) meV, and (f) \([0.65, 0.83]\) meV.
\(I(\hbar\omega)\) cuts at \(\mathbf{Q}=(0.3, 0.3, 0.5)\) and \((0.3, 0.3, 0)\) are plotted in Figs. 3(a) and 3(b), respectively. Below \(T_{\text{c}}\), a clear increment in intensity at \(\hbar\omega\sim 0.56\) meV can be seen in the \(I(\hbar\omega)\) cut at \((0.3, 0.3, 0.5)\) [shaded area in Fig. 3(a)]. Meanwhile, such enhancement at \(\hbar\omega\sim 0.56\) meV below \(T_{\text{c}}\) is not observed in the \(I(\hbar\omega)\) cut at \((0.3, 0.3, 0)\) [Fig. 3(b)], consistent with the previous INS study.33) \(I(\mathbf{Q})\) cuts at \(\hbar\omega=0.56\) meV along \(\mathbf{Q}=(H, H, 0.5)\) and \((H, H, 0)\) also show the same trend [Figs. 3(c) and 3(d)]. The IC fluctuations at \(\hbar\omega=0.56\) meV along \(\mathbf{Q}=(H, H, 0.5)\) are enhanced below \(T_{\text{c}}\) [shaded area in Fig. 3(c)], while the IC fluctuations at \(\hbar\omega=0.56\) meV along \(\mathbf{Q}=(H, H, 0)\) do not change in intensity across \(T_{\text{c}}\) [Fig. 3(d)]. Figure 3(e) shows the \(I(\mathbf{Q})\) cuts along \(\mathbf{Q}=(H, H, 0.5)\) for three different energies, \(\hbar\omega=0.35\), 0.56, and 0.74 meV. The \(I(\mathbf{Q})\) cuts exhibit peaks centered at \(\mathbf{Q}_{\text{IC}}=(0.3, 0.3, 0.5)\) owing to the IC magnetic fluctuations in all the energy windows at both temperatures. Interestingly, only the INS intensity at \(\hbar\omega=0.56\) meV shows sizable enhancement at \(\mathbf{Q}_{\text{IC}}\) in the superconducting state compared to the normal state [shaded area in Fig. 3(e)]. We emphasize that the energy, \(\hbar\omega=0.56\) meV, corresponds well to the superconducting gap \(2\Delta\)37) [vertical bar in Fig. 3(a)], suggesting that the observed enhancement, localized in both \(\mathbf{Q}_{\text{res}}=(0.3, 0.3, 0.5)\) and \(\hbar\omega_{\text{res}}=0.56\) meV, is the spin resonance [see also white arrow in Fig. 2(a)]. The spin gap of the IC magnetic fluctuations at \(\mathbf{Q}_{\text{IC}}=(0.3, 0.3, 0.5)\) is smaller than 0.2 meV [Fig. 3(a)], which makes the spin resonance less pronounced. \(I(\mathbf{Q})\) cuts at 0.74 meV in Fig. 3(e) also shows a tiny increment at 0.3 K. A possible scenario is that the reported superconducting gap value37) is band averaged and the superconducting gaps regarding the α and β bands are slightly larger than \(2\Delta=0.56\) meV.
Figure 3. (Color online) \(I(\hbar\omega)\) and \(I(\mathbf{Q})\) cuts of the low-energy IC magnetic fluctuations in Sr2RuO4. (a, b) \(I(\hbar\omega)\) cuts at (a) \(\mathbf{Q}=(0.3, 0.3, 0.5)\) and (b) \((0.3, 0.3, 0)\) below and above \(T_{\text{c}}\). Solid lines are the guides for the eye. Vertical bars represent the superconducting gap \(2\Delta=0.56\) meV.37) (c, d) \(I(\mathbf{Q})\) cuts along (c) \(\mathbf{Q}=(H, H, 0.5)\) and (d) \((H, H, 0)\) at 0.3 and 1.8 K with the energy window of \([0.47, 0.65]\) meV. Solid lines are the fitting results by the Gaussian function with linear background. (e) \(I(\mathbf{Q})\) cuts along \(\mathbf{Q}=(H, H, 0.5)\) at 0.3 and 1.8 K with the energy windows of \([0.23, 0.47]\), \([0.47, 0.65]\), and \([0.65, 0.83]\) meV. Solid lines are the fitting results. (f) \(I(\mathbf{Q})\) cuts along \(\mathbf{Q}=(0.3, 0.3, L)\) at 0.3 K with the energy windows of \([0.23, 0.47]\), \([0.47, 0.65]\), and \([0.65, 0.83]\) meV. Background estimated by \(I(\mathbf{Q})\) along \(\mathbf{Q}=(0.525, 0.525, L)\) at 0.3 K is subtracted from each spectrum. Dashed lines are the squared magnetic form factor of Sr2RuO443) and therefore represent the components of the nearly two-dimensional IC magnetic fluctuations. The solid line is the guides for the eye. Shaded areas in panels (a, c, e, f) indicate the spin resonance.
L dependence of the low-energy IC magnetic fluctuations can provide us a compelling evidence to clarify the nodal nature of the superconducting gaps.29) Figures 2(d)–2(f) illustrate contour maps of the INS intensities in the (\(HHL\)) plane at 0.3 K with the energies 0.35, 0.56, and 0.74 meV, and \(I(\mathbf{Q})\) cuts along \(\mathbf{Q}=(0.3, 0.3, L)\) at 0.3 K with the corresponding energies are also plotted in Fig. 3(f). In contrast to the monotonically decreasing intensities of the IC magnetic fluctuations at 0.35 and 0.74 meV following the squared magnetic form factor of Sr2RuO443) [the dashed lines in Fig. 3(f)], the \(I(\mathbf{Q})\) cut at the spin resonance energy (0.56 meV) shows a maximum at \(L\sim 0.5\) besides the monotonically decreasing component of the IC magnetic fluctuations [the shaded area and the dashed line in Fig. 3(f)]. The L modulated spin resonance represents the gap symmetry as the feedback effect from the superconducting gaps. Also, the L modulated intensity explains the absence of the spin resonance at \(\mathbf{Q}=(0.3, 0.3, 0)\) [Figs. 2(c), 3(b), and 3(d)]. To theoretically elucidate the origin of the L modulated intensity of the spin resonance in Sr2RuO4, RPA calculations assuming the horizontal line nodes at the superconducting gaps were further performed.
To construct a realistic model, we perform density functional theory (DFT) calculations using the Wien2k package.44) We obtain an effective 3-orbital model considering the Ru \(d_{xz}\), \(d_{yz}\), \(d_{xy}\)-orbitals using the maximum localized Wannier functions.45) The generalized gradient approximation (GGA) exchange–correlation functional46) is adopted with the cut-off energy \(RK_{\text{max}}=7\) and 512 k-point mesh. We renormalize the bandwidth considering the effective mass \(m^{*}=3.5\), and the resulting renormalized bandwidth is \(W\sim 1.05\) eV. We consider the following gap function with horizontal line nodes: \begin{equation} \Delta(\mathbf{k}) = \Delta_{0}\cos ck_{z} \end{equation} (1) within the standard BCS framework. We take the gap amplitude \(\Delta_{0}=4.8\times 10^{-3}W\). In the body center tetragonal system, the period along to the \(k_{z}\) axis is \(4\pi/c\), and thus we take \(k_{z}\) as \(0\leq k_{z}<4\pi/c\). We obtain the dynamical spin susceptibility \(\chi^{\text{total}}_{s}(\mathbf{q},\omega)\) applying RPA as \begin{align} \chi^{\text{total}}_{s}(\mathbf{q},\omega) &= \sum_{l,m} \chi_{s}^{l,l,m,m}(\mathbf{q},\omega), \end{align} (2) \begin{align} \hat{\chi}_{s}(\mathbf{q},\omega) &= \hat{\chi}_{0}(\mathbf{q},\omega)[\hat{I}-\hat{S}\hat{\chi}_{0}(\mathbf{q},\omega)]^{-1}, \end{align} (3) \begin{align} \hat{\chi}_{0}(\mathbf{q},\omega) &= \hat{\chi}_{0,G}(\mathbf{q},\omega)+\hat{\chi}_{0,F}(\mathbf{q},\omega), \end{align} (4) \begin{align} \chi^{l_{1},l_{2},l_{3},l_{4}\\ \sigma_{1},\sigma_{2},\sigma_{3},\sigma_{4} }_{0,G(F)}(\mathbf{q},\omega) &= \sum_{k}\sum_{n,m}\frac{f(E^{n}_{\mathbf{k}+\mathbf{q}})-f(E^{m}_{\mathbf{k}})}{\omega+i\delta-E^{n}_{\mathbf{k}+\mathbf{q}}+E^{m}_{\mathbf{k}}} \notag \\ &\quad \times U_{l_{1},\sigma_{1},n}(\mathbf{k}+\mathbf{q})U_{l_{4},\sigma_{4},m}(\mathbf{k})\notag\\ &\quad \times U^{\dagger}_{m,l_{2},\sigma_{2}}(\mathbf{k})U^{\dagger}_{n,l_{3},\sigma_{3}}(\mathbf{k}+\mathbf{q}), \end{align} (5) where \(l_{1}\sim l_{4}\) and \(\sigma_{1}\sim\sigma_{4}\) are the orbital \((d_{xz}, d_{yz}, d_{xy})\) and spin (↑ and ↓) indices. \(\hat{\chi}_{0,G(F)}\) denotes the normal (anomalous) part of the irreducible bare susceptibility \(\hat{\chi}_{0}\) at \(\sigma_{1}=\sigma_{2}=\sigma_{3}=\sigma_{4}\) (\(\sigma_{1}=\sigma_{2}\neq\sigma_{3}=\sigma_{4}\)). \(E^{n}_{\mathbf{k}}\) and \(f(E^{n}_{\mathbf{k}})\) are the eigenvalue and Fermi distribution function of Bogoliubov quasi-particles. \(\hat{S}\) is the interaction vertex matrix.47) We consider the on-site intra- and inter-orbital Coulomb interactions \(U_{l}\) and \(U'\) as \(U_{dxz/dyz}=0.21W\), \(U_{d_{xy}}=0.7U_{dxz/dyz}\), \(U'=3U_{dxz/dyz}/4\). The Hund's coupling and pair hopping \(J=J'=U_{dxz/dyz}/8\). We take the temperature \(T=1.9\times 10^{-3}W\), the smearing factor \(\delta=1.9\times 10^{-3}W\) and \(1024\times 1024\times 32\) k-mesh. The resulting characteristic spectral features of the dynamical spin susceptibly are not much influenced by the detailed parameter values mentioned above.
Figures 4(a) and 4(b) illustrate the calculated dynamical spin susceptibilities for energies, \(\hbar\omega=2\Delta_{0}\) and \(4\Delta_{0}\). We note that our calculations do not include the squared magnetic form factor of Sr2RuO4 for clear visualization of the L dependence. For \(\hbar\omega=2\Delta_{0}\), the dynamical spin susceptibility shows the maximum at \(\mathbf{Q}=(1/3, 1/3, 1/2)\) and \((2/3, 2/3, 1/2)\) as illustrated in Fig. 4(a). The superconducting gaps with the horizontal line nodes described in Eq. (1,) give such the feature along L, which is indeed observed in our INS measurements [Fig. 2(e)]. We address that the observed L modulated intensity cannot be explained by the superconducting gap with the vertical line nodes since the superconducting gap changes its sign within the \(k_{z}\) plane.29) On the other hand, no pronounced L modulation at higher energy transfers is observed either in the experiment [Fig. 2(f)], except the monotonically decreasing intensities due to the magnetic form factor, or in the calculation [Fig. 4(b)]. Since the energy window is too high compared to the superconducting gaps in Sr2RuO4,37) the effect of the superconducting gaps smears out, and only the quasi-two-dimensional feature of the IC magnetic fluctuations show up in the higher energy windows (\(\hbar\omega\gg 2\Delta\)).
Figure 4. (Color online) Calculated density maps of the dynamical spin susceptibility assuming the horizontal line nodes in Sr2RuO4. (a) \(\hbar\omega=2\Delta_{0}\). (b) \(\hbar\omega=4\Delta_{0}\).
The observed spin resonance at \(\mathbf{Q}_{\text{res}}=(0.3, 0.3, 0.5)\) and \(\hbar\omega_{\text{res}} = 2\Delta\) indicates that the quasi-one-dimensional α and β sheets are active bands for the bulk superconductivity of Sr2RuO4. Our RPA calculations reveal that the observed L modulated intensity of the spin resonance originates from the horizonal line nodes at the superconducting gaps, in agreement with the field-angle-dependent specific heat capacity measurements.12) We believe that our results give a strong constraint on the superconducting pairing symmetry of Sr2RuO4.
Let us now discuss the possible pairing symmetry of Sr2RuO4. Among the pairing symmetries with the horizontal line nodes proposed by the recent uniaxial pressure16–22) and NMR works,23,24) the chiral d-wave \((k_{x} + ik_{y})k_{z}\)48) is compatible with the time-reversal symmetry breaking.8,9) However, its wave function is odd in the \(k_{z}\) plane giving rise to the spin resonance at \(\mathbf{Q}=(0.3, 0.3, 0)\), and thus such pairing symmetry can be excluded by our INS data. Recently, the Josephson effects measurements suggested the time-reversal invariant superconductivity in Sr2RuO4.49) Combined with the time-reversal invariant superconductivity and the horizontal line nodes, d-wave \(d_{3k_{z}^{2}-1}\)29) is proposed for the candidate for the pairing symmetry of Sr2RuO4. The single component order parameter is also in line with the absence of a split transition in the presence of uniaxial strain and \(T_{\text{c}}\) cusp in the limit of zero strain.16–18) We note that the horizontal line nodes assuming \(d_{3k_{z}^{2}-1}\) is not symmetry protected. However, the horizontal line nodes can be realized even if the d-wave \(d_{3k_{z}^{2}-1}\) belonging to the \(A_{1g}\) representation in the point group \(D_{4h}\) slightly mixes with the s wave. Here, we emphasize that the spin resonance at \(L=0.5\) is more sensitive to the gap modulation along \(k_{z}\) than the \(k_{z}\) position of horizontal line nodes. The candidate of the origin of the horizontal nodes is a competition of pairing mechanisms. Since spin fluctuations are highly two-dimensional, it would be difficult to induce three-dimensional gap structure only with pure spin fluctuation mechanism. The crystal structure of Sr2RuO4 is body-centered tetragonal. Therefore, it seems that the electron–phonon50) mediated mechanism would have three-dimensionality. The observed weakly three-dimensional spin resonance at \(L = 0.5\) may come from a competition between spin fluctuations and electron–phonon coupling.
In summary, we investigated the low-energy IC magnetic fluctuations in Sr2RuO4 below and above \(T_{\text{c}}\) with detailed analysis on the L dependence. Below \(T_{\text{c}}\), the spin resonance appears at \(\mathbf{Q}_{\text{res}}=(0.3, 0.3, 0.5)\) and \(\hbar\omega_{\text{res}}=0.56\) meV. The spin resonance shows the L modulated intensity with a maximum at \(L\sim 0.5\). The L modulated intensity of the spin resonance and our RPA calculations indicate that the superconducting gaps regarding the α and β sheets have horizontal line nodes.
Acknowledgements
The experiments at AMATERAS were conducted under the J-PARC MLF user program with the proposal numbers 2018A0060 and 2018AU1402. The present work was supported by JSPS KAKENHI Grant Numbers JP17K05553 and JP17K14349, and the Cooperative Research Program of “Network Joint Research Center for Materials and Devices” (20181072).
References
- 1 Y. Maeno, H. Hashimoto, K. Yoshida, S. Nishizaki, T. Fujita, J. G. Bednorz, and F. Lichtenberg, Nature (London) 372, 532 (1994). 10.1038/372532a0 Crossref, Google Scholar
- 2 T. M. Rice and M. Sigrist, J. Phys.: Condens. Matter 7, L643 (1995). 10.1088/0953-8984/7/47/002 Crossref, Google Scholar
- 3 G. Baskaran, Physica B 223–224, 490 (1996). 10.1016/0921-4526(96)00155-X Crossref, Google Scholar
- 4 A. P. Mackenzie and Y. Maeno, Rev. Mod. Phys. 75, 657 (2003). 10.1103/RevModPhys.75.657 Crossref, Google Scholar
- 5 Y. Maeno, S. Kitaoka, T. Nomura, S. Yonezawa, and K. Ishida, J. Phys. Soc. Jpn. 81, 011009 (2012). 10.1143/JPSJ.81.011009 Link, Google Scholar
- 6 K. Ishida, H. Mukuda, Y. Kitaoka, K. Asayama, Z. Q. Mao, Y. Mori, and Y. Maeno, Nature (London) 396, 658 (1998). 10.1038/25315 Crossref, Google Scholar
- 7 J. A. Duffy, S. M. Hayden, Y. Maeno, Z. Mao, J. Kulda, and G. J. McIntyre, Phys. Rev. Lett. 85, 5412 (2000). 10.1103/PhysRevLett.85.5412 Crossref, Google Scholar
- 8 G. M. Luke, Y. Fudamoto, K. M. Kojima, M. L. Larkin, J. Merrin, B. Nachumi, Y. J. Uemura, Y. Maeno, Z. Q. Mao, Y. Mori, H. Nakamura, and M. Sigrist, Nature (London) 394, 558 (1998). 10.1038/29038 Crossref, Google Scholar
- 9 J. Xia, Y. Maeno, P. T. Beyersdorf, M. M. Fejer, and A. Kapitulnik, Phys. Rev. Lett. 97, 167002 (2006). 10.1103/PhysRevLett.97.167002 Crossref, Google Scholar
- 10 J. R. Kirtley, C. Kallin, C. W. Hicks, E.-A. Kim, Y. Liu, K. A. Moler, Y. Maeno, and K. D. Nelson, Phys. Rev. B 76, 014526 (2007). 10.1103/PhysRevB.76.014526 Crossref, Google Scholar
- 11 S. Yonezawa, T. Kajikawa, and Y. Maeno, Phys. Rev. Lett. 110, 077003 (2013). 10.1103/PhysRevLett.110.077003 Crossref, Google Scholar
- 12 S. Kittaka, A. Kasahara, T. Sakakibara, D. Shibata, S. Yonezawa, Y. Maeno, K. Tenya, and K. Machida, Phys. Rev. B 90, 220502(R) (2014). 10.1103/PhysRevB.90.220502 Crossref, Google Scholar
- 13 Y. Amano, M. Ishihara, M. Ichioka, N. Nakai, and K. Machida, Phys. Rev. B 91, 144513 (2015). 10.1103/PhysRevB.91.144513 Crossref, Google Scholar
- 14 N. Kikugawa, T. Terashima, S. Uji, K. Sugii, Y. Maeno, D. Graf, R. Baumbach, and J. Brooks, Phys. Rev. B 93, 184513 (2016). 10.1103/PhysRevB.93.184513 Crossref, Google Scholar
- 15 K. Machida and M. Ichioka, Phys. Rev. B 77, 184515 (2008). 10.1103/PhysRevB.77.184515 Crossref, Google Scholar
- 16 C. W. Hicks, D. O. Brodsky, E. A. Yelland, A. S. Gibbs, J. A. N. Bruin, M. E. Barber, S. D. Edkins, K. Nishimura, S. Yonezawa, Y. Maeno, and A. P. Mackenzie, Science 344, 283 (2014). 10.1126/science.1248292 Crossref, Google Scholar
- 17 A. Steppke, L. Zhao, M. E. Barber, T. Scaffidi, F. Jerzembeck, H. Rosner, A. S. Gibbs, Y. Maeno, S. H. Simon, A. P. Mackenzie, and C. W. Hicks, Science 355, eaaf9398 (2017). 10.1126/science.aaf9398 Crossref, Google Scholar
- 18 C. A. Watson, A. S. Gibbs, A. P. Mackenzie, C. W. Hicks, and K. A. Moler, Phys. Rev. B 98, 094521 (2018). 10.1103/PhysRevB.98.094521 Crossref, Google Scholar
- 19 M. E. Barber, A. S. Gibbs, Y. Maeno, A. P. Mackenzie, and C. W. Hicks, Phys. Rev. Lett. 120, 076602 (2018). 10.1103/PhysRevLett.120.076602 Crossref, Google Scholar
- 20 Y. Luo, A. Pustogow, P. Guzman, A. P. Dioguardi, S. M. Thomas, F. Ronning, N. Kikugawa, D. A. Sokolov, F. Jerzembeck, A. P. Mackenzie, C. W. Hicks, E. D. Bauer, I. I. Mazin, and S. E. Brown, Phys. Rev. X 9, 021044 (2019). 10.1103/PhysRevX.9.021044 Crossref, Google Scholar
- 21 V. Sunko, E. A. Morales, I. Marković, M. E. Barber, D. Milosavljević, F. Mazzola, D. A. Sokolov, N. Kikugawa, C. Cacho, P. Dudin, H. Rosner, C. W. Hicks, P. D. C. King, and A. P. Mackenzie, npj Quantum Mater. 4, 46 (2019). 10.1038/s41535-019-0185-9 Crossref, Google Scholar
- 22 Y.-S. Li, N. Kikugawa, D. A. Sokolov, F. Jerzembeck, A. S. Gibbs, Y. Maeno, C. W. Hicks, M. Nicklas, and A. P. Mackenzie, arXiv:1906.07597. Google Scholar
- 23 A. Pustogow, Y. Luo, A. Chronister, Y.-S. Su, D. A. Sokolov, F. Jerzembeck, A. P. Mackenzie, C. W. Hicks, N. Kikugawa, S. Raghu, E. D. Bauer, and S. E. Brown, Nature (London) 574, 72 (2019). 10.1038/s41586-019-1596-2 Crossref, Google Scholar
- 24 K. Ishida, M. Manago, K. Kinjo, and Y. Maeno, J. Phys. Soc. Jpn. 89, 034712 (2020). 10.7566/JPSJ.89.034712 Link, Google Scholar
- 25 K. Ishida, H. Mukuda, Y. Kitaoka, Z. Q. Mao, Y. Mori, and Y. Maeno, Phys. Rev. Lett. 84, 5387 (2000). 10.1103/PhysRevLett.84.5387 Crossref, Google Scholar
- 26 I. Bonalde, B. D. Yanoff, M. B. Salamon, D. J. Van Harlingen, E. M. E. Chia, Z. Q. Mao, and Y. Maeno, Phys. Rev. Lett. 85, 4775 (2000). 10.1103/PhysRevLett.85.4775 Crossref, Google Scholar
- 27 E. Hassinger, P. Bourgeois-Hope, H. Taniguchi, S. René de Cotret, G. Grissonnanche, M. S. Anwar, Y. Maeno, N. Doiron-Leyraud, and L. Taillefer, Phys. Rev. X 7, 011032 (2017). 10.1103/PhysRevX.7.011032 Crossref, Google Scholar
- 28 S. Kittaka, S. Nakamura, T. Sakakibara, N. Kikugawa, T. Terashima, S. Uji, D. A. Sokolov, A. P. Mackenzie, K. Irie, Y. Tsutsumi, K. Suzuki, and K. Machida, J. Phys. Soc. Jpn. 87, 093703 (2018). 10.7566/JPSJ.87.093703 Link, Google Scholar
- 29 K. Machida, K. Irie, K. Suzuki, H. Ikeda, and Y. Tsutsumi, Phys. Rev. B 99, 064510 (2019). 10.1103/PhysRevB.99.064510 Crossref, Google Scholar
- 30 I. I. Mazin and D. J. Singh, Phys. Rev. Lett. 82, 4324 (1999). 10.1103/PhysRevLett.82.4324 Crossref, Google Scholar
- 31 Y. Sidis, M. Braden, P. Bourges, B. Hennion, S. NishiZaki, Y. Maeno, and Y. Mori, Phys. Rev. Lett. 83, 3320 (1999). 10.1103/PhysRevLett.83.3320 Crossref, Google Scholar
- 32 K. Iida, M. Kofu, N. Katayama, J. Lee, R. Kajimoto, Y. Inamura, M. Nakamura, M. Arai, Y. Yoshida, M. Fujita, K. Yamada, and S.-H. Lee, Phys. Rev. B 84, 060402(R) (2011). 10.1103/PhysRevB.84.060402 Crossref, Google Scholar
- 33 S. Kunkemöller, P. Steffens, P. Link, Y. Sidis, Z. Q. Mao, Y. Maeno, and M. Braden, Phys. Rev. Lett. 118, 147002 (2017). 10.1103/PhysRevLett.118.147002 Crossref, Google Scholar
- 34 M. Braden, Y. Sidis, P. Bourges, P. Pfeuty, J. Kulda, Z. Mao, and Y. Maeno, Phys. Rev. B 66, 064522 (2002). 10.1103/PhysRevB.66.064522 Crossref, Google Scholar
- 35 P. Steffens, Y. Sidis, J. Kulda, Z. Q. Mao, Y. Maeno, I. I. Mazin, and M. Braden, Phys. Rev. Lett. 122, 047004 (2019). 10.1103/PhysRevLett.122.047004 Crossref, Google Scholar
- 36 D. K. Morr, P. F. Trautman, and M. J. Graf, Phys. Rev. Lett. 86, 5978 (2001). 10.1103/PhysRevLett.86.5978 Crossref, Google Scholar
- 37 H. Suderow, V. Crespo, I. Guillamon, S. Vieira, F. Servant, P. Lejay, J. P. Brison, and J. Flouquet, New J. Phys. 11, 093004 (2009). 10.1088/1367-2630/11/9/093004 Crossref, Google Scholar
- 38 S. I. Ikeda, U. Azuma, N. Shirakawa, Y. Nishihara, and Y. Maeno, J. Cryst. Growth 237–239, 787 (2002). 10.1016/S0022-0248(01)02033-4 Crossref, Google Scholar
- 39 Z. Q. Mao, Y. Maeno, and H. Fukazawa, Mater. Res. Bull. 35, 1813 (2000). 10.1016/S0025-5408(00)00378-0 Crossref, Google Scholar
- 40 K. Nakajima, S. Ohira-Kawamura, T. Kikuchi, M. Nakamura, R. Kajimoto, Y. Inamura, N. Takahashi, K. Aizawa, K. Suzuya, K. Shibata, T. Nakatani, K. Soyama, R. Maruyama, H. Tanaka, W. Kambara, T. Iwahashi, Y. Itoh, T. Osakabe, S. Wakimoto, K. Kakurai, F. Maekawa, M. Harada, K. Oikawa, R. E. Lechner, F. Mezei, and M. Arai, J. Phys. Soc. Jpn. 80, SB028 (2011). 10.1143/JPSJS.80SB.SB028 Link, Google Scholar
- 41 Y. Inamura, T. Nakatani, J. Suzuki, and T. Otomo, J. Phys. Soc. Jpn. 82, SA031 (2013). 10.7566/JPSJS.82SA.SA031 Link, Google Scholar
- 42 H. Iwasawa, Y. Yoshida, I. Hase, S. Koikegami, H. Hayashi, J. Jiang, K. Shimada, H. Namatame, M. Taniguchi, and Y. Aiura, Phys. Rev. Lett. 105, 226406 (2010). 10.1103/PhysRevLett.105.226406 Crossref, Google Scholar
- 43 T. Nagata, M. Urata, H. Kawano-Furukawa, H. Yoshizawa, H. Kadowaki, and P. Dai, Phys. Rev. B 69, 174501 (2004). 10.1103/PhysRevB.69.174501 Crossref, Google Scholar
- 44 P. Blaha, K. Schwarz, G. K. H. Madsen, D. Kvasnicka, and J. Luitz, Wien2k: An Augmented PlaneWave + Local Orbitals Program for Calculating Crystal Properties (Vienna University of Technology, Vienna, 2001) (http://www.wien2k.at/). Google Scholar
- 45 N. Marzari, A. A. Mostofi, J. R. Yates, I. Souza, and D. Vanderbilt, Rev. Mod. Phys. 84, 1419 (2012). 10.1103/RevModPhys.84.1419 Crossref, Google Scholar
- 46 J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett. 77, 3865 (1996). 10.1103/PhysRevLett.77.3865 Crossref, Google Scholar
- 47 K. Suzuki, H. Usui, K. Kuroki, and H. Ikeda, Phys. Rev. B 96, 024513 (2017). 10.1103/PhysRevB.96.024513 Crossref, Google Scholar
- 48 I. Žutić and I. Mazin, Phys. Rev. Lett. 95, 217004 (2005). 10.1103/PhysRevLett.95.217004 Crossref, Google Scholar
- 49 S. Kashiwaya, K. Saitoh, H. Kashiwaya, M. Koyanagi, M. Sato, K. Yada, Y. Tanaka, and Y. Maeno, Phys. Rev. B 100, 094530 (2019). 10.1103/PhysRevB.100.094530 Crossref, Google Scholar
- 50 M. Braden, W. Reichardt, Y. Sidis, Z. Mao, and Y. Maeno, Phys. Rev. B 76, 014505 (2007). 10.1103/PhysRevB.76.014505 Crossref, Google Scholar
Cited by
View all 19 citing articles
Still Mystery after All These Years —Unconventional Superconductivity of Sr2RuO4—
062001, 10.7566/JPSJ.93.062001
A Method to Examine Likelihood of Candidate Theories beyond Parameter Optimization — Application to Specific Heat Data for the Multiband Superconductor Sr2RuO4
074707, 10.7566/JPSJ.92.074707