Subscriber access provided by Massachusetts Institute of Technology
J. Phys. Soc. Jpn. 88, 114602 (2019) [6 Pages]
FULL PAPERS

Crystal Structures of Highly Hole-Doped Layered Perovskite Nickelate Pr2−xSrxNiO4 Studied by Neutron Diffraction

+ Affiliations
1Materials and Life Science Division, J-PARC Center, Tokai, Ibaraki 319-1195, Japan2Institute for Materials Research, Tohoku University, Sendai 980-8577, Japan3Neutron Science and Technology Center, Comprehensive Research Organization for Science and Society (CROSS), Tokai, Ibaraki 319-1106, Japan4Institute of Materials Structure Science, High Energy Accelerator Research Organization (KEK), Tsukuba, Ibaraki 305-0801, Japan

A high-resolution time-of-flight powder neutron diffraction study of the layered nickel oxide Pr2−xSrxNiO4 with x = 0.7 and 0.9 was performed to characterize the crystal structures of these highly hole-doped nickelates. For the sample with x = 0.7, the Ni–O bond lengths decrease uniformly with decreasing temperature, and the atomic displacement parameters are similar to those for x = 1/3. In contrast, for the sample with x = 0.9, the out-of-plane Ni–O bond length shows a sharp thermal contraction in the high-temperature region, which is suggestive of changes in the orbital occupation accompanied by the development of checkerboard-type charge correlations. Furthermore, the x = 0.9 sample is characterized by a large atomic displacement parameter for the apical O atoms along the out-of-plane direction, which is interpreted as the existence of two types of Ni3+ sites with different orbital occupancies. The distinct difference between the x = 0.9 sample and the lower-concentration compounds should be related to the development of checkerboard-type charge ordering in the metallic matrix and possible orbital ordering at the Ni3+ sites at x ∼ 1 in the hole-doped layered nickel oxides.

©2019 The Physical Society of Japan
1. Introduction

Hole-doped \(R_{2-x}\)SrxNiO4 (R = La, Nd, Pr, Eu, etc.) has a crystal structure that is isostructural with one of the typical high-temperature superconducting copper oxides, La\(_{2-x}\)SrxCuO4, and shows unique charge and spin orderings. Although they remain insulating over a wide hole concentration range, layered nickelates have attracted interest because of their close association with superconducting cuprates. The hole concentration per Ni site is nominally equal to x, and it is well known that the material shows stripe charge and spin ordering at \(x\leq 1/2\).114) In the stripe-ordered state, doped holes (Ni3+ sites) form one dimensional stripes in the NiO2 planes with a wave vector of \(2{\boldsymbol{{\epsilon}}}\), followed by long-period spin ordering with a wave vector of \({\boldsymbol{{\epsilon}}}\), where \(\epsilon\sim x\). Since the discovery of stripe ordering in cuprates,15) the relationship between stripe ordering and superconductivity has become one of the most fascinating topics in condensed matter physics, and \(R_{2-x}\)SrxNiO4 has been extensively studied in this regard for more than two decades.

The stripe-type charge ordering leads to checkerboard-type charge ordering at \(x = 1/2\). The latter type of charge and spin orderings is observed commonly over a wide x region, at least up to \(x = 0.7\), with almost constant wave vectors.16,17) According to crystal structure studies, the \(c/a\)1821) and Ni–O1/Ni–O2 ratios,20,21) where O1 and O2 are the apical and in-plane oxygen atoms in a NiO6 octahedron, respectively, decrease at \(x > 1/2\). These structural variations suggest that the doped holes enter the \(d_{3z^{2}-r^{2}}\) orbitals at \(x > 1/2\), which was recently confirmed in an x-ray absorption spectroscopy (XAS) study.22) These experimental findings pertaining to changes in the orbital character indicate that we should consider the orbital degree of freedom at \(x > 1/2\). In contrast, for the stripe ordering at \(x\leq 1/2\), it was naively believed that doped holes would enter the in-plane \(d_{x^{2}-y^{2}}\) orbitals of Ni2+ ions. Furthermore, \(R_{2-x}\)SrxNiO4 transitions from a Mott insulator to a metal at \(x\sim 1\).2325) This metallic state shows anomalous charge dynamics originating from considerable checkerboard-type charge correlations.22) The appearance of an anomalous metallic state of a multi-orbital nature in highly hole-doped nickelates contrasts with the behavior of cuprates, in which the orbital degree of freedom is less important, and renders nickelates unique in the field of condensed matter physics.

Because the orbital character is coupled with the crystal structure, a detailed analysis of the crystal structure is fundamental for evaluating it. In particular, the orbital states of the Ni ions are expected to be reflected in the shapes of the NiO6 octahedra. Therefore, a precise determination of the oxygen positions and the Ni–O bond lengths is indispensable. Neutron diffraction is the most suitable tool for this purpose because of its high sensitivity to light elements such as oxygen. Thus far, neutron diffraction studies on the hole concentration dependence at room temperature at \(x > 1/2\)1921) and on the temperature dependence at \(x = 1/4\) and 1/326,27) have been conducted. Unfortunately, however, there have been no detailed studies on the temperature dependence of the crystal structure at \(x > 1/2\). Therefore, in the present work, we performed a high-resolution neutron diffraction study on the crystal structure of highly hole-doped Pr\(_{2-x}\)SrxNiO4 with \(x=0.7\) and 0.9 to characterize the crystal structure parameters and their temperature dependences.

2. Experiments

4 g of polycrystalline samples of Pr1.3Sr0.7NiO4 and Pr1.1Sr0.9NiO4 were prepared by the solid-state reaction method. Dried powders of Pr6O11, SrCO3, and NiO were mixed at the nominal molar ratios and pressed into pellets. Subsequently, the pellets were calcined at 1400 °C for 24 h in air with intermediate grinding. The resistivities of the pelletized samples were measured as a function of temperature, and the results were consistent with previous works.19,22,23,25) High-resolution (\(\Delta d/d\approx 3.5\times 10^{-4}\)) time-of-flight neutron diffraction experiments were performed using the Super High Resolution Powder Diffractometer SuperHRPD28) installed at the BL08 beamline at the Materials and Life Science Experimental Facility (MLF) at the Japan Proton Accelerator Research Complex (J-PARC). The samples were sealed in vanadium cells with helium gas and were attached to the cold finger of a closed-cycle refrigerator. The temperature dependence of the powder diffraction patterns was studied in the temperature (T) range of \(9\,\text{K}\leq T\lesssim 290\,\text{K}\), and the obtained powder diffraction patterns were refined by following the Rietveld method by using Z-Rietveld (version 1.0.3.RC10).29,30) The crystal structure was illustrated using the Balls & Sticks program.31)

3. Results and Discussion
Structural parameters

Figures 1(a) and 1(b) show the powder diffraction patterns of Pr1.3Sr0.7NiO4 and Pr1.1Sr0.9NiO4, respectively, measured at 9 K. According to our Rietveld refinements, both diffraction patterns were well reproduced by tetragonal structures with the space group \(I4/mmm\). Because the magnetic Bragg peaks were too weak to observe in the powder diffraction measurements, they were not considered in the refinements. Even with our high-resolution diffraction measurements, we did not observe peak splitting of peaks in any of the diffraction patterns, indicating that the crystal structures of Pr1.3Sr0.7NiO4 and Pr1.1Sr0.9NiO4 remain tetragonal in the temperature range studied. For both samples, we found that the peaks whose indices are dominated by h and k have tails on their left sides, while the peaks characterized by dominant l indices have tails on their right sides. The tails cannot be reproduced by anisotropic peak broadening, but they can be interpreted as the existence of minority phases with larger c axes and shorter a axes. The differences in the lattice parameters suggest that the minority phases originate from domains with hole concentrations lower than the nominal values. The coexistence of minority phases has also been reported for La\(_{2-x}\)SrxNiO4 with \(x\sim 1\).20,21) We successfully fit the powder diffraction patterns to two (majority and minority) tetragonal phases, both characterized by \(I4/mmm\) symmetry [Fig. 1(d)]. The mass fractions of the minority phases were 24 and 13% for Pr1.3Sr0.7NiO4 and Pr1.1Sr0.9NiO4, respectively. We discuss the crystal structure of the majority phases hereinafter. We should note that the hole concentrations (or x values) for the majority phases are slightly larger than the nominal concentrations because of the existence of minority phases. We estimated the x values in the majority phases from the \(c/a\) ratios20) and found that their deviations from the nominal values are less than 0.01 for both samples.


Figure 1. (Color online) Powder diffraction patterns of (a) Pr1.3Sr0.7NiO4 and (b) Pr1.1Sr0.9NiO4 measured at 9 K and shown here as a function of lattice spacing d. The solid lines are calculated intensities obtained by Rietveld refinement. The vertical lines are calculated peak positions of the majority (top rows) phase and minority (bottom rows) phase. The solid lines below the peak positions represent the differences between the observed and calculated intensities. (c) Schematic of the crystal structure of Pr\(_{1-x}\)SrxNiO4. The arrows indicate the directions of the atomic displacements corresponding to the ADPs. (d) Enlarged view of the diffraction pattern of Pr1.1Sr0.9NiO4 near the 006 (\(d\sim 2.06\) Å) and 105 (\(d\sim 2.07\) Å) peaks.

Table I lists the crystal structure parameters for the majority phases of both samples at 9 K and ∼290 K.32) In Pr1.1Sr0.9NiO4, the lattice parameters a and c are 0.3% longer and 1% shorter, respectively, when compared with those of Pr1.3Sr0.7NiO4. This concentration dependence of the lattice parameters is similar to that of La\(_{2-x}\)SrxNiO4.20,23) Meanwhile, the lattice parameters of Pr\(_{2-x}\)SrxNiO4 are 1% smaller than those of La\(_{2-x}\)SrxNiO4 at the same concentrations.20,23) The latter result reflects the smaller ionic radius of Pr3+ compared with La3+.33)

Data table
Table I. Crystal structure parameters and bond lengths of Pr1.3Sr0.7NiO4 and Pr1.1Sr0.9NiO4 obtained by Rietveld refinement. The space group is \(I4/mmm\) with the following atomic positions: Pr/Sr at 4e \((0,0,z)\), Ni at 2a \((0,0,0)\), O1 at 4e \((0,0,z)\), and O2 4c \((0,1/2,0)\).
Lattice parameters

Figures 2(a) and 2(b) show the temperature dependences of the lattice parameters of Pr1.3Sr0.7NiO4 and Pr1.1Sr0.9NiO4, respectively. The lattice parameters for both materials decrease as the temperature decreases. For La\(_{3/2}\)Sr\(_{1/2}\)NiO4, it has been reported that the lattice parameter a increases below 120 K in accordance with a decrease in the optical gap,34) which was attributed to the development of stripe charge ordering.17) However, such an increase in a was not observed in Pr1.3Sr0.7NiO4 or Pr1.1Sr0.9NiO4, which suggests that the temperature-dependent variations in the hole concentration and orbital character accompanied by a transition from checkerboard to stripe charge ordering are less significant compared with those at \(x=1/2\). This finding is consistent with the fact that the loss of the optical-conductivity spectral intensity due to the appearance of stripe charge ordering is negligible at \(x=0.7\).17) At high temperatures, for both compounds, the c axis shrinks considerably faster than the a axis as the temperature decreases. The temperature dependence of c saturates at low temperatures, resulting in a minimum ratio of the two lattice parameters, \(c/a\). The minimum of \(c/a\) occurs at ∼60 K for Pr1.3Sr0.7NiO4, while the minimum occurs at ∼80 K for Pr1.1Sr0.9NiO4 [insets in Figs. 2(a) and 2(b)]. Because these temperatures are close to the spin ordering temperatures (\(T_{\text{s}}\)),16,17,35) it would be interesting if the minima of \(c/a\) correspond to the spin orderings. However, \(T_{\text{s}}\) is higher in the case of \(x=0.7\), which is inconsistent with the difference in the temperatures at the minimum \(c/a\). Furthermore, the minimum \(c/a\) value occurs at a similar temperature for \(x=1/4\) and 1/3 in La\(_{2-x}\)SrxNiO4,26) although their \(T_{\text{s}}\)s differ from those at \(x=0.7\) and 0.9. Therefore, the temperature dependence of \(c/a\) is not related to the magnetic ordering, but instead can be ascribed to the greater compressibility of the c axis compared with that of the a axis at high temperatures due to the layered perovskite structure. The absence of anomalous features in the temperature dependence of the lattice parameters can be confirmed by the fact that the temperature dependence of the unit cell volume (V) is well described by the Debye model. Figure 2(c) shows the temperature dependence of V, where the solid lines present a fit to the Debye model: \begin{equation} V \approx V(T = 0) + \frac{9\gamma N k_{\text{B}}}{B} T \left(\frac{T}{\Theta_{\text{D}}}\right)^{3} \int^{\Theta_{\text{D}}/T}_{0} \frac{x^{3}}{e^{x} - 1}\,dx. \end{equation} (1) Here γ, B, \(k_{\text{B}}\), \(\Theta_{\text{D}}\), and N are the Grüneisen parameter, bulk modulus, Boltzmann constant, Debye temperature, and number of atoms in the unit cell, respectively.36) The observed values show no significant deviations from the calculated values throughout the temperature range studied. This finding suggests that the temperature dependence of the unit cell volume is dominated by phonon contributions. The fitting parameters obtained from these fits are summarized in Table II, exhibiting values similar to those for La\(_{7/4}\)Sr\(_{1/4}\)NiO4 and La\(_{5/3}\)Sr\(_{1/3}\)NiO4.26)


Figure 2. (a, b) Temperature dependence of the lattice parameters of (a) Pr1.3Sr0.7NiO4 and (b) Pr1.1Sr0.9NiO4. The insets show the temperature dependence of the \(c/a\) ratio. (c) Temperature dependence of the unit cell volume of Pr1.3Sr0.7NiO4 (open circles) and Pr1.1Sr0.9NiO4 (closed circles). The solid lines represent fits to the Debye model [Eq. (1)].

Data table
Table II. Parameters obtained by fitting the unit cell volumes to the Debye function.
Bond lengths

The Ni–O bond lengths should more directly reflect the electronic state of the Ni ions than the lattice parameters. Figure 3 shows the temperature dependence of the two types of Ni–O bonds in Pr1.3Sr0.7NiO4 and Pr1.1Sr0.9NiO4. The in-plane Ni–O bond length (Ni–O2) is exactly half the lattice parameter a, and its error bars are very small. This parameter decreases smoothly with decreasing temperature for both compounds. In contrast, the out-of-plane Ni–O bond length (Ni–O1) depends on the z position of O1, and large error bars originating from the ambiguity in atomic position can be seen. Nevertheless, there is a qualitative difference between the behaviors of the Ni–O1 bonds of the two compounds: The Ni–O1 bond length for Pr1.3Sr0.7NiO4 decreases uniformly as the temperature decreases, whereas that of Pr1.1Sr0.9NiO4 decreases more rapidly at \(T\gtrsim 150\) K than at \(T\lesssim 150\) K. This difference can be more clearly observed in the ratios of the two Ni–O bonds (insets in Fig. 3) and should, at least partly, contribute to the difference in temperatures corresponding to the minimum \(c/a\) (Fig. 2). This change in the bond contraction rate of Ni–O1 at \(x=0.9\) also affects the atomic displacement parameter (ADP) of O1 along the out-of-plane direction (\(U_{33}\)). As shown in Fig. 4(a), the slope of the temperature dependence of O1 \(U_{33}\) changes at approximately 150 K.


Figure 3. Temperature dependence of the Ni–O bond length for (a) Pr1.3Sr0.7NiO4 and (b) Pr1.1Sr0.9NiO4. The open and solid symbols denote the out-of-plane (Ni–O1) and in-plane (Ni–O2) bonds, respectively. The insets show the temperature dependence of the ratio of the two Ni–O bond lengths, Ni–O1/Ni–O2. The solid lines serve as a guide for the eye.


Figure 4. (Color online) Temperature dependence of the ADP \(U_{ij}\) for (a) O1, (b) O2, (c) Pr/Sr, and (d) Ni. The open and closed symbols denote values for Pr1.3Sr0.7NiO4 and Pr1.1Sr0.9NiO4, respectively. The circles, diamonds, and squares denote \(U_{11}\), \(U_{22}\), and \(U_{33}\), respectively. In (a) and (b), the values of the O atoms for La\(_{5/3}\)Sr\(_{1/3}\)NiO4 reproduced from Ref. 26 are shown as broken (\(U_{11}\)), dashed-dotted (\(U_{22}\)), and solid (\(U_{33}\)) lines.38)

Notably, Ni–O1 in Pr1.3Sr0.7NiO4 exhibits a temperature dependence similar to the monotonic temperature dependence observed at \(x=1/3\) and the convex temperature dependence observed at \(x=1/4\).26) Therefore, only the \(x=0.9\) sample shows qualitatively different behavior compared with the lower concentrations. We suggest that this difference in the behavior of Ni–O1 is related to the difference in charge itinerancy. \(R_{2-x}\)SrxNiO4 shows insulating behavior with charge ordering at \(x < 1\). However, as the hole concentration increases, the itinerancy increases, and the material becomes an anomalous metal with checkerboard-type charge correlations at \(x\gtrsim 1\).22,25) The \(x=0.9\) sample is located near the boundary of the insulator–metal transition. Its electrical conductivity is metallic at high temperatures, and it transitions to an insulator at low temperatures.25) At this point, the development of checkerboard-type charge correlations in the NiO2 planes as a function of temperature should sweep out the excess holes into the out-of-plane \(d_{3z^{2}-r^{2}}\) orbitals, which should result in a more rapid decrease in the out-of-plane Ni–O bond lengths at higher temperatures. The bond contraction rate finally decreases and becomes similar to that of the insulating lower-concentration compounds at lower temperatures, where the checkerboard-type charge correlations are well developed.

Atomic displacement parameters

The lattice parameters and bond lengths described above were obtained by structural refinements based on a tetragonal structure with \(I4/mmm\) symmetry. The structural modulations associated with the doped holes and modified orbital occupations are averaged with respect to this symmetry and are obscured. In contrast, the difference between the modulated atomic positions and the atomic positions of the average crystal structure are incorporated in the ADPs in the structural refinement. Therefore, a precise determination of the ADPs, particularly those of the O atoms, provides useful information for investigating the local charge and orbital occupations of the Ni ions.27,37)

Figure 4 shows the temperature dependence of the ADPs for Pr1.3Sr0.7NiO4 (open symbols) and Pr1.1Sr0.9NiO4 (closed symbols). The displacement directions corresponding to the respective ADPs are drawn schematically in Fig. 1(c). In Figs. 4(a) and 4(b), the values for the O atoms in La\(_{5/3}\)Sr\(_{1/3}\)NiO4 reported in Ref. 26 are plotted as well (lines). The ADPs of the O atoms in Pr1.3Sr0.7NiO4 are similar to those of the O atoms in La\(_{5/3}\)Sr\(_{1/3}\)NiO4 [Figs. 4(a) and 4(b)]. \(U_{11}\) is larger than \(U_{33}\) for O1, whereas \(U_{33}\) is the largest and \(U_{22}\) is the smallest for O2. These ADP magnitudes suggest that it is difficult for the O atoms to vibrate along the Ni–O bonds, and the displacements of the O atoms are dominated by rotation of the NiO6 octahedra. For Pr1.1Sr0.9NiO4, the ADPs of O2 are similar to those for Pr1.3Sr0.7NiO4 and La\(_{5/3}\)Sr\(_{1/3}\)NiO4, although the values are slightly larger [Fig. 4(b)]. This result holds for the ADPs of Ni [Fig. 4(d)]. Abeykoon et al. reported that the ADPs of the in-plane O and Ni atoms along the Ni–O bond directions deviate from the normal Debye behavior due to the development of stripe charge correlations.27) In the present study, we did not observe any anomalies associated with spin or charge ordering in the ADPs of O2 or Ni. The absence of anomalies in the in-plane ADPs may originate from the fact that the two-dimensional character of the checkerboard-type charge ordering restricts the degrees of freedom of the associated atomic displacements compared with the one-dimensional stripe charge ordering. In contrast, the ADPs of O1 in Pr1.1Sr0.9NiO4 clearly show different behavior compared with those for Pr1.3Sr0.7NiO4 and La\(_{5/3}\)Sr\(_{1/3}\)NiO4 [Fig. 4(a)]. \(U_{33}\) is considerably larger than that for Pr1.3Sr0.7NiO4, while \(U_{11}\) possesses similar values to Pr1.3Sr0.7NiO4. As a result, the magnitude relationship between \(U_{11}\) and \(U_{33}\) is reversed with respect to that for Pr1.3Sr0.7NiO4 and La\(_{5/3}\)Sr\(_{1/3}\)NiO4. The ADPs of Pr/Sr for Pr1.1Sr0.9NiO4 show a similar anomalous behavior [Fig. 4(d)]; again, \(U_{33}\) is larger than \(U_{11}\), in contrast to the results for Pr1.3Sr0.7NiO4.

The significantly larger O1 \(U_{33}\) indicates exceptionally large displacements of the apical oxygen atoms along the out-of-plane direction around their average positions in Pr1.1Sr0.9NiO4. A recent XAS study22) revealed that the Ni3+ sites created by hole doping are divided into two sites at \(x > 1/2\), while the number of Ni2+ sites continues to decrease. The occupancy of Ni3+ sites with \(d_{x^{2}-y^{2}}\) orbitals (holes in the \(d_{3z^{2}-r^{2}}\) orbitals) increases as x increases, whereas that of Ni3+ sites with \(d_{3z^{2}-r^{2}}\) orbitals (holes in the \(d_{x^{2}-y^{2}}\) orbitals) remains constant at 0.5. If the occupancies of the Ni3+ sites with \(d_{x^{2}-y^{2}}\) orbitals and the Ni2+ sites change as functions of \(x-0.5\) and \(-x+1\), respectively, the Ni sites will consist of 50% \(d_{x^{2}-y^{2}}\) Ni3+ sites and 50% \(d_{3z^{2}-r^{2}}\) Ni3+ sites at \(x=1\). In other words, the checkerboard-type charge ordering is accompanied by orbital ordering of the \(d_{x^{2}-y^{2}}/d_{3z^{2}-r^{2}}\) orbitals. This orbital ordering results in a mixture of long and short out-of-plane Ni–O bonds and induces a large O1 \(U_{33}\). Because the Pr and Sr atoms are located above or below the O1 atoms [Fig. 1(d)], displacement of the O1 atoms along the out-of-plane direction should affect the positions of the Pr/Sr atoms. Therefore, the ADPs of Pr/Sr in Pr1.1Sr0.9NiO4 show tendencies similar to those of O1.

4. Conclusion

We performed a powder neutron diffraction study of the layered nickel oxide Pr\(_{2-x}\)SrxNiO4 with \(x=0.7\) and 0.9 to characterize the crystal structures of these highly hole-doped nickelates. For both samples, the powder diffraction patterns were refined based on tetragonal crystal structures at all measurement temperatures for \(T\lesssim 290\) K. The temperature dependences of the lattice parameters showed no observable anomalies associated with spin or charge ordering, and those of the unit cell volumes could be described by the Debye model, with parameter values similar to those in the literature for \(x=1/4\) and 1/3. As observed for the Ni–O bond lengths and ADPs, the sample with \(x=0.7\) showed behaviors similar to those of lower-concentration compounds. Conversely, for these parameters, the sample with \(x=0.9\) was distinct from the other compositions. A sharp thermal contraction of the out-of-plane Ni–O bond length was observed in the high-temperature region, where the compound shows metallic conductivity, suggesting a change in orbital occupation concomitant with the development of checkerboard-type charge correlations. Most notably, significantly large displacements of the apical O atoms along the out-of-plane direction were interpreted as the existence of two types of Ni3+ sites with different orbital occupancies. The distinct difference for the sample with \(x=0.9\), as determined by our high-resolution neutron diffraction measurements, should be related to the increase in itinerancy and possible orbital ordering of the Ni3+ sites at \(x\sim 1\) in \(R_{2-x}\)SrxNiO4. Further structural analysis at \(x > 1\) using single crystals would be useful to elucidate this anomalous metallic phase with an orbital degree of freedom, which is left for future work.

Acknowledgments

We thank Sanghyun Lee, Masato Hagihara, Seiko Ohira-Kawamura, Naoki Murai, and Maiko Kofu for valuable discussions. The resistivity measurements were performed by using the Physical Property Measurement System (PPMS, Quantum Design) at the CROSS user laboratory. The proposal number of the neutron diffraction experiment at MLF, J-PARC is 2012B0247. This work was supported by JSPS KAKENHI Grant Number JP15K04742.


References

  • 1 C. H. Chen, S.-W. Cheong, and A. S. Cooper, Phys. Rev. Lett. 71, 2461 (1993). 10.1103/PhysRevLett.71.2461 CrossrefGoogle Scholar
  • 2 J. M. Tranquada, D. J. Buttrey, V. Sachan, and J. E. Lorenzo, Phys. Rev. Lett. 73, 1003 (1994). 10.1103/PhysRevLett.73.1003 CrossrefGoogle Scholar
  • 3 V. Sachan, D. J. Buttrey, J. M. Tranquada, J. E. Lorenzo, and G. Shirane, Phys. Rev. B 51, 12742 (1995). 10.1103/PhysRevB.51.12742 CrossrefGoogle Scholar
  • 4 J. M. Tranquada, J. E. Lorenzo, D. J. Buttrey, and V. Sachan, Phys. Rev. B 52, 3581 (1995). 10.1103/PhysRevB.52.3581 CrossrefGoogle Scholar
  • 5 J. M. Tranquada, D. J. Buttrey, and V. Sachan, Phys. Rev. B 54, 12318 (1996). 10.1103/PhysRevB.54.12318 CrossrefGoogle Scholar
  • 6 S.-H. Lee and S.-W. Cheong, Phys. Rev. Lett. 79, 2514 (1997). 10.1103/PhysRevLett.79.2514 CrossrefGoogle Scholar
  • 7 P. Wochner, J. M. Tranquada, D. J. Buttrey, and V. Sachan, Phys. Rev. B 57, 1066 (1998). 10.1103/PhysRevB.57.1066 CrossrefGoogle Scholar
  • 8 H. Yoshizawa, T. Kakeshita, R. Kajimoto, T. Tanabe, T. Katsufuji, and Y. Tokura, Phys. Rev. B 61, R854 (2000). 10.1103/PhysRevB.61.R854 CrossrefGoogle Scholar
  • 9 S.-H. Lee, S.-W. Cheong, K. Yamada, and C. F. Majkrzak, Phys. Rev. B 63, 060405(R) (2001). 10.1103/PhysRevB.63.060405 CrossrefGoogle Scholar
  • 10 R. Kajimoto, T. Kakeshita, H. Yoshizawa, T. Tanabe, T. Katsufuji, and Y. Tokura, Phys. Rev. B 64, 144432 (2001). 10.1103/PhysRevB.64.144432 CrossrefGoogle Scholar
  • 11 S.-H. Lee, J. M. Tranquada, K. Yamada, D. J. Buttrey, Q. Li, and S.-W. Cheong, Phys. Rev. Lett. 88, 126401 (2002). 10.1103/PhysRevLett.88.126401 CrossrefGoogle Scholar
  • 12 J. Li, Y. Zhu, J. M. Tranquada, K. Yamada, and D. J. Buttrey, Phys. Rev. B 67, 012404 (2003). 10.1103/PhysRevB.67.012404 CrossrefGoogle Scholar
  • 13 K. Ishizaka, T. Arima, Y. Murakami, R. Kajimoto, H. Yoshizawa, N. Nagaosa, and Y. Tokura, Phys. Rev. Lett. 92, 196404 (2004). 10.1103/PhysRevLett.92.196404 CrossrefGoogle Scholar
  • 14 M. Hücker, M. v. Zimmermann, R. Klingeler, S. Kiele, J. Geck, S. N. Bakehe, J. Z. Zhang, J. P. Hill, A. Revcolevschi, D. J. Buttrey, B. Büchner, and J. M. Tranquada, Phys. Rev. B 74, 085112 (2006). 10.1103/PhysRevB.74.085112 CrossrefGoogle Scholar
  • 15 J. M. Tranquada, B. J. Sternlieb, J. D. Axe, Y. Nakamura, and S. Uchida, Nature 375, 561 (1995). 10.1038/375561a0 CrossrefGoogle Scholar
  • 16 R. Kajimoto, K. Ishizaka, H. Yoshizawa, and Y. Tokura, Phys. Rev. B 67, 014511 (2003). 10.1103/PhysRevB.67.014511 CrossrefGoogle Scholar
  • 17 K. Ishizaka, Y. Taguchi, R. Kajimoto, H. Yoshizawa, and Y. Tokura, Phys. Rev. B 67, 184418 (2003). 10.1103/PhysRevB.67.184418 CrossrefGoogle Scholar
  • 18 J. Gopalakrishnan, G. Colsmann, and B. Reuter, J. Solid State Chem. 22, 145 (1977). 10.1016/0022-4596(77)90031-7 CrossrefGoogle Scholar
  • 19 Y. Takeda, R. Kanno, M. Sakano, O. Yamamoto, M. Takano, Y. Bando, H. Akinaga, K. Takita, and J. B. Goodenough, Mater. Res. Bull. 25, 293 (1990). 10.1016/0025-5408(90)90100-G CrossrefGoogle Scholar
  • 20 J. E. Millburn, M. A. Green, D. A. Neumann, and M. J. Rosseinsky, J. Solid State Chem. 145, 401 (1999). 10.1006/jssc.1999.8111 CrossrefGoogle Scholar
  • 21 A. Aguadero, M. J. Escudero, M. Pérez, J. A. Alonso, V. Pomjakushin, and L. Daza, Dalton Trans. 2006, 4377 (2006). 10.1039/B606316K CrossrefGoogle Scholar
  • 22 M. Uchida, Y. Yamasaki, Y. Kaneko, K. Ishizaka, J. Okamoto, H. Nakao, Y. Murakami, and Y. Tokura, Phys. Rev. B 86, 165126 (2012). 10.1103/PhysRevB.86.165126 CrossrefGoogle Scholar
  • 23 R. J. Cava, B. Batlogg, T. T. Palstra, J. J. Krajewski, W. F. Peck, Jr., A. P. Ramirez, and L. W. Rupp, Jr., Phys. Rev. B 43, 1229(R) (1991). 10.1103/PhysRevB.43.1229 CrossrefGoogle Scholar
  • 24 S. Shinomori, Y. Okimoto, M. Kawasaki, and Y. Tokura, J. Phys. Soc. Jpn. 71, 705 (2002). 10.1143/JPSJ.71.705 LinkGoogle Scholar
  • 25 M. Uchida, K. Ishizaka, P. Hansmann, Y. Kaneko, Y. Ishida, X. Yang, R. Kumai, A. Toschi, Y. Onose, R. Arita, K. Held, O. K. Andersen, S. Shin, and Y. Tokura, Phys. Rev. Lett. 106, 027001 (2011). 10.1103/PhysRevLett.106.027001 CrossrefGoogle Scholar
  • 26 G. Wu, J. J. Neumeier, C. D. Ling, and D. N. Argyriou, Phys. Rev. B 65, 174113 (2002). 10.1103/PhysRevB.65.174113 CrossrefGoogle Scholar
  • 27 A. M. M. Abeykoon, E. S. Božin, W.-G. Yin, G. Gu, J. P. Hill, J. M. Tranquada, and S. J. L. Billinge, Phys. Rev. Lett. 111, 096404 (2013). 10.1103/PhysRevLett.111.096404 CrossrefGoogle Scholar
  • 28 S. Torii, M. Yonemura, T. Y. S. P. Putra, J. Zhang, P. Miao, T. Muroya, R. Tomiyasu, T. Morishima, S. Sato, H. Sagehashi, Y. Noda, and T. Kamiyama, J. Phys. Soc. Jpn. 80, SB020 (2011). 10.1143/JPSJS.80SB.SB020 LinkGoogle Scholar
  • 29 R. Oishi, M. Yonemura, Y. Nishimaki, S. Torii, A. Hoshikawa, T. Ishigaki, T. Morishima, K. Mori, and T. Kamiyama, Nucl. Instrum. Methods Phys. Res., Sect. A 600, 94 (2009). 10.1016/j.nima.2008.11.056 CrossrefGoogle Scholar
  • 30 R. Oishi-Tomiyasu, M. Yonemura, T. Morishima, A. Hoshikawa, S. Torii, T. Ishigaki, and T. Kamiyama, J. Appl. Crystallogr. 45, 299 (2012). 10.1107/S0021889812003998 CrossrefGoogle Scholar
  • 31 T. C. Ozawa and S. J. Kang, J. Appl. Crystallogr. 37, 679 (2004). 10.1107/S0021889804015456 CrossrefGoogle Scholar
  •   (32) The larger R factors for Pr1.1Sr0.9NiO4 at 291 K than the other data are due to accidentally low statistics of the data. Google Scholar
  • 33 R. D. Shannon, Acta Crystallogr., Sect. A 32, 751 (1976). 10.1107/S0567739476001551 CrossrefGoogle Scholar
  • 34 J. H. Jung, D.-W. Kim, T. W. Noh, H. C. Kim, H.-C. Ri, S. J. Levett, M. R. Lees, D. M. Paul, and G. Balakrishnan, Phys. Rev. B 64, 165106 (2001). 10.1103/PhysRevB.64.165106 CrossrefGoogle Scholar
  • 35 T. Jestädt, K. H. Chow, S. J. Blundell, W. Hayes, F. L. Pratt, B. W. Lovett, M. A. Green, J. E. Millburn, and M. J. Rosseinsky, Phys. Rev. B 59, 3775 (1999). 10.1103/PhysRevB.59.3775 CrossrefGoogle Scholar
  • 36 T. Kiyama, K. Yoshimura, K. Kosuge, Y. Ikeda, and Y. Bando, Phys. Rev. B 54, R756 (1996). 10.1103/PhysRevB.54.R756 CrossrefGoogle Scholar
  • 37 P. Dai, J. Zhang, H. A. Mook, S.-H. Liou, P. A. Dowben, and E. W. Plummer, Phys. Rev. B 54, R3694 (1996). 10.1103/PhysRevB.54.R3694 CrossrefGoogle Scholar
  •   (38) For the atomic displacement parameters of the O2 atom, the definitions of U11 and U22 are reversed with respect to those in Refs. 26 and 27 due to the difference in the definition of the atomic position. In Fig. 4(d), we plotted U11 and U22 in Ref. 26 as U22 and U11, respectively. Google Scholar