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

thanks: [email protected]

Magnetism behavior of TT^{\prime}-type Eu2CuO4 revealed by muon spin rotation/relaxation measurements

M. Fujita Institute for Materials Research, Tohoku University, 2-1-1 Katahira, Aoba-ku, Sendai 980-8577, Japan    K. M. Suzuki Institute for Materials Research, Tohoku University, 2-1-1 Katahira, Aoba-ku, Sendai 980-8577, Japan    S. Asano Department of Physics, Graduate School of Science, Tohoku University, 6-3 Aoba, Aramaki, Aoba-ku, Sendai 980-8578, Japan    H. Okabe Muon Science Laboratory, Institute of Materials Structure Science, High Energy Accelerator Research Organization (KEK), Tsukuba, Ibaraki 305-0801, Japan    A. Koda Muon Science Laboratory, Institute of Materials Structure Science, High Energy Accelerator Research Organization (KEK), Tsukuba, Ibaraki 305-0801, Japan    R. Kadono Muon Science Laboratory, Institute of Materials Structure Science, High Energy Accelerator Research Organization (KEK), Tsukuba, Ibaraki 305-0801, Japan    I. Watanabe Advanced Meson Science Laboratory, Nishina Center for Accelerator-Based Science, The Institute of Physical and Chemical Research (RIKEN), Wako, Saitama 351-0198, Japan
Abstract

We performed muon spin rotation/relaxation measurements to investigate the magnetic behavior of TT^{\prime}-type Eu2CuO4 (ECO), which is the parent compound of electron-doped cuprate superconductors, and the effects of oxygen-reduction annealing on its magnetism. In as-sintered (AS) ECO, we clarified the development of magnetic correlations upon cooling below TN1T_{\mathrm{N1}} (= 265 K) as well as the coexistence of a dominant fluctuating spin state and partially ordered spin state in the temperature range between \sim150 K and TN1T_{\mathrm{N1}}. Upon further cooling, uniform long-range magnetic order was observed below TN2=110T_{\mathrm{N2}}=110 K, which is close to the ordering temperature of 115 K in TT^{\prime}-type La2CuO4 (LCO) [Phys. Rev. B 82, 180508(R) (2010)]. For oxygen-reduction-annealed ECO, a similar ordering sequence with the same TN2T_{\mathrm{N2}} was observed but without the partially ordered spin state. Therefore, the fluctuating spin state over a wide temperature range and a TN2T_{\mathrm{N2}} less than the Néel temperature (TN2TN1T_{\mathrm{N2}}\approx T_{\mathrm{N1}}) in TT-type LCO are common features of the TT^{\prime}-type parent R2R_{2}CuO4 (RRCO, RR: rare-earth ion). The origin of the partially ordered spin state in AS ECO is discussed from the viewpoint of chemical defect. Furthermore, we discuss the roles of electron doping and repairing defect in the observed effect of annealing on the magnetism of TT^{\prime}-type RRCO.

pacs:
PACS numbers:74.25.Ha, 74.62.Yb, 74.72.Dn, 74.90.+n

I Introduction

The discovery of superconductivity in TT^{\prime}-type R2R_{2}CuO4 (RRCO, RR: rare-earth) without Ce substitution stimulated research on the ground state of host materials in terms of the oxygen coordination around Cu Tsukada2005 ; Matsumoto2009 ; Asai2011 ; Takamatsu2012 . It is well known that high-temperature-synthesized standard samples of TT^{\prime}-type RRCO with square planar coordination exhibits Néel order at \sim250 KLuke1990 , similar to the case of TT-type La2CuO4 (LCO) with six-oxygen coordination Uemura1989 . However, Hord and et al. reported through muon spin rotation/relaxation (μ\muSR) measurements a lower ordering temperature in non-superconducting TT^{\prime}-type LCOHord2010 , which is synthesized using a soft chemical method. Spin correlations start to develop upon cooling below TN1200T_{\mathrm{N1}}\approx 200 K, and a long-range magnetic order emerges at TN2=115T_{\mathrm{N2}}=115 K, showing a successive ordering sequence. It is considered that the lower TN2T_{\mathrm{N2}} is due to enhanced two-dimensionality and related to the real ground state, namely, the metallic state of the ideal TT^{\prime}-type RRCO, which is suggested by previous theoretical studies Das2009a ; Weber2010a ; Weber2010b .

Subsequently, through μ\muSR measurements, Kojima et al.  Kojima2014 reported the absence of long-range magnetic order in thin films of superconducting (SC) TT^{\prime}-type RRCO. The reason for the distinct ground states in different samples is understood in terms of chemical defect Krockenberger2013 ; Naito2016 : Defects in a sample induce magnetic order due to strong electron correlations, but no such defects yield the real ground state of TT^{\prime}-type RRCO. Furthermore, it is considered that both as-prepared powder samples and thin films contain defects such as excess oxygen at the apical positions Radaelli1994c ; Schultz1996 and copper-ion deficiency Kang2007 . For thin films, the defects can be effectively removed through annealing because of their large surface-to-volume ratio. In contrast, the complete removal of defects from an entire sample is difficult for bulk compounds Naito2016 ; Adachi2013 . In this context, in the low-temperature-synthesized LCO, partial defect induces magnetic order with a lower TN2T_{\mathrm{N2}}. The ordered phase vanishes when the complete removal of the defects realizes a uniform electrostatic potential on the CuO2 plane. To support the above scenario, a suppression of magnetic order accompanied by the appearance of superconductivity was reported by annealing the low-temperature-synthesized TT^{\prime}-type La1.8Eu0.2CuO4 (LECO)  Adachi2016 ; Fukazawa2017 . However, many experiments, including recent X-ray absorption fine-structure measurements, have indicated electron doping as an alternative aspect of annealing treatment Jiang1993 ; Higgins2006 ; Asano2018 ; Asano2020 . Therefore, the common magnetism in TT^{\prime}-type RRCO and the effect of annealing on the physical properties are fundamental issues in the research on superconductivity in TT^{\prime}-type cuprates.

To address these issues, a comprehensive investigation of high-temperature-synthesized non-SC bulk samples of TT^{\prime}-type RRCO is necessary, but the precise thermal evolution of magnetism and the effect of annealing on magnetism have not yet been studied for the host compounds. Hence, we performed μ\muSR measurements on as-sintered (AS) and annealed (AN) TT^{\prime}-type Eu2CuO4 (ECO). Owing to the absence of a rare-earth magnetic moment in ECOTovar1989 ; Tanikawa1994 , we can extract the inherent magnetism of the CuO2 plane in the TT^{\prime}-type parent cuprates. We found that both AS and AN ECO have the same magnetic evolution as TT^{\prime}-type LCO. Thus, the sequential ordering process with a lower TN2T_{\mathrm{N2}} is a common feature of magnetism in TT^{\prime}-type R2R_{2}CuO4. In addition, in the AS sample, we found the coexistence of a fluctuating spin state and partially ordered spin state in the temperature range of \sim150 K (>> TN1T_{\mathrm{N1}}) to TN2T_{\mathrm{N2}}. The partially ordered state disappears with annealing, indicating that the coexistence of the two states originates from defects existing in AS ECO.

The remainder of this paper is organized as follows. Section II describes the sample preparation and experimental details. Section III presents the results of zero-field (ZF) and longitudinal-field (LF) μ\muSR measurements. Section IV discusses the common features of the ordering sequence and the effect of annealing on the magnetism of TT^{\prime}-type RRCO. Finally, Sec. V summarizes this paper.

II Experimental details

We prepared polycrystalline samples of ECO by using a standard solid-state reaction method. A pre-fired powder was pressed into pellets and sintered in air at 1050C. We estimated the amounts of removed oxygen (δ\delta) to be 0.02\approx 0.02 with an accuracy of 0.001 based on the weight loss of the samples through the reduction annealing process. Furthermore, we confirmed the phase purity of the samples by x-ray powder diffraction measurements and determined the lattice constants from the diffraction pattern. The evaluated lattice constants with I4/mmmI4/mmm crystal symmetry are summarized in Table 1. None of our samples showed superconductivity, even after the annealing procedure.

Table 1: Lattice constants and characterized magnetic ordering temperatures TN1T_{\mathrm{N1}} and TN2T_{\mathrm{N2}} for as-sintered (AS) and annealed (AN) Eu2CuO4.
a (Å) c (Å) TN1T_{\mathrm{N1}} (K) TN2T_{\mathrm{N2}} (K)
Eu2CuO4 AS 3.9049(1) 11.9121(3) 265(10) 110(10)
AN 3.9047(1) 11.9126(3) 250(10) 110(10)

ZF and LF μ\muSR measurements were performed at the Materials and Life Science Facility (MLF) at J-PARC, Japan and at the RIKEN-RAL Muon Facility of the Rutherford Appleton Laboratory (RAL), UK using single-pulsed positive surface muon beams. The samples were cooled to 10 K using an open-cycle 4He-flow cryostat. The μ\muSR time spectrum, i.e., the time evolution of the asymmetry of positron events, is given by A(t)=[F(t)αB(t)]/[F(t)+αB(t)]A(t)=[F(t)-\alpha B(t)]/[F(t)+\alpha B(t)], where F(t)F(t) and B(t)B(t) are histograms of the positron event at the detectors positioned upstream and downstream of the sample, respectively, along the muon momentum direction. Further, α\alpha is a calibration factor for the efficiency and solid angle of the detectors. We evaluated α\alpha from an oscillation spectrum under a weak transverse magnetic field at approximately 300 K and calibrated the spectra for each sample. The obtained μ\muSR time spectra were analyzed using the WiMDA program Pratt2000 .

Following previous studies Kubo2002 ; Fujita2003 , we define TN1T_{\mathrm{N1}} (TN2T_{\mathrm{N2}}) as the temperature at which the ZF μ\muSR spectra changes from Gaussian-type decay to exponential-type decay (the temperature at which the bulk oscillation component appears) upon cooling. Table 1 summarizes TN1T_{\mathrm{N1}} and TN2T_{\mathrm{N2}}.

III Results

III.1 Zero-field μ\muSR measurements

Figure 1(a) shows the time spectra of ZF μ\muSR for AS ECO at representative temperatures. A Gaussian-type slow decay due to the nuclear dipole fields is observed at 275 K. Upon cooling to 264 K, an exponential-type spectrum appears, indicating a slowdown of electron spin fluctuations. With further cooling, the spectra show faster decay at 235 K and, subsequently, another phase of slower decay at 164 K. This trend is not a typical ordering sequence; that is, the spin fluctuations slow down monotonically toward the freezing temperature. Furthermore, in the spectra at 164 K and 235 K, an oscillation component with a small amplitude is superimposed on the dominant exponential-type decay. Therefore, a partially ordered spin state coexists with a fluctuating spin state and/or spin-glass-like state at these temperatures. The small oscillation component disappears at 126 K, resulting in a simple exponential-type decay spectrum. A well-defined oscillation spectrum appears at even lower temperatures. Thus, the true magnetic order occurs at a much lower temperature than the previously reported temperature of 265 K Chattopadhyay1994 , and the ordering temperature is comparable to the TN2T_{\mathrm{N2}} of 110 K for TT^{\prime}-type LCO Hord2010 . Here, we emphasize that the existence of the small oscillation component and the detailed thermal evolution of magnetism on the CuO2 plane were revealed by the high-quality data with fine temperature steps.

Figure 1(b) shows the ZF μ\muSR spectra of AN ECO. Similar to the results for AS ECO, exponential-type depolarization appears at TN1=250T_{\mathrm{N1}}=250 K upon cooling (see the spectra at 200, 171, and 132 K). The spectrum at 171 K shows a slower time decay than that at 200 K. It transforms into a faster decay spectrum again at 132 K, following which oscillation spectra appear below TN2=110T_{\mathrm{N2}}=110 K. Thus, the sequential development of magnetism with a low TN2T_{\mathrm{N2}} is the same in AS and AN ECO. One of the qualitative differences between the spectra of AS and AN ECO is the small oscillation component around 200 K, which is absent in the AS ECO spectra. Another difference is the considerably weaker depolarization in AN ECO around this temperature. These results indicate that the magnetism at higher temperatures is varied considerably by annealing.

Refer to caption
Figure 1: Time spectra of zero-field μ\muSR for (a) as-sintered and (b) annealed Eu2CuO4. Solid curves indicate the results of least-squares fitting based on Eq. (1)–(3). All the data were acquired at MLF.

For quantitative evaluation of the thermal evolution of magnetism, we analyzed the time spectra using the following functions:

A(t)=A0{fsee(λset)β+fosceλd1tcos(ω1t+ϕ1)+(1fsefosc)eλst}+ABG(T>TN2),\displaystyle\begin{split}A(t)=A_{0}\{f_{\mathrm{se}}e^{-(\lambda_{\mathrm{se}}t)^{\beta}}+f_{\mathrm{osc}}e^{-\lambda_{\mathrm{d1}}t}\cos(\omega_{1}t+\phi_{1})\\ +(1-f_{\mathrm{se}}-f_{\mathrm{osc}})e^{-\lambda_{\mathrm{s}}t}\}+A_{\mathrm{BG}}\hskip 14.22636pt(T>T_{\mathrm{N2}}),\end{split} (1)
A(t)=A0{foscGT(t)+(1fosc)eλst}+ABG(TTN2),\displaystyle\begin{split}A(t)=A_{0}\{f_{\mathrm{osc}}G_{\mathrm{T}}(t)+(1-f_{\mathrm{osc}})e^{-\lambda_{s}t}\}+A_{\mathrm{BG}}\\ (T\leq T_{\mathrm{N2}}),\end{split} (2)

where

GT(t)=r1eλd1tcos(ω1t+ϕ1)+(1r1)eλd2t.\displaystyle\begin{split}G_{\mathrm{T}}(t)=r_{1}e^{-\lambda_{\mathrm{d1}}t}\cos(\omega_{1}t+\phi_{1})+(1-r_{1})e^{-\lambda_{\mathrm{d2}}t}.\end{split} (3)

The first (third) and second terms in Eq. 1 represent the fast (slow) decay and the small oscillation components, respectively. Similarly, the first and second terms in Eq. 2 represent the oscillation and decay components, respectively. A0A_{0} is the initial asymmetry at t=0t=0, and ABGA_{\mathrm{BG}} is the time-independent background. Both A0A_{0} and ABGA_{\mathrm{BG}} were fixed throughout the analysis. fsef_{\mathrm{se}} (foscf_{\mathrm{osc}}) is the fraction of the stretched exponential (oscillation) component. λse\lambda_{\mathrm{se}} and β\beta are the relaxation rate and power of the stretched exponential component, respectively. λd1\lambda_{\mathrm{d1}} and λd2\lambda_{\mathrm{d2}} are the damping rates of the oscillations, and r1r_{1} is the ratio of the damping components. ω1\omega_{1} and ϕ1\phi_{1} are the frequency and phase of the oscillation, respectively.

Refer to caption
Figure 2: Temperature dependencies of (a) the depolarization rate of the stretched exponential term λse\lambda_{\mathrm{se}}, (b) the asymmetry fraction of the oscillation component foscf_{\mathrm{osc}}, and (c) the oscillation frequency ω1\omega_{1} for as-sintered and annealed Eu2CuO4. The spectra, acquired at MLF and RAL, were analyzed using Eq. (1)–(3) to obtain the parameters.
Refer to caption
Figure 3: Field dependencies of μ\muSR time spectra in longitudinal magnetic fields for as-sintered Eu2CuO4 at (a) 187 and (b) 138 K. All the data for Eu2CuO4 were acquired at MLF.

As shown in Fig. 2(a), AS and AN ECO exhibit similar temperature dependencies of λse\lambda_{\mathrm{se}}, i.e., with cooling, λse\lambda_{\mathrm{se}} increases sharply at TN1T_{\mathrm{N1}}, maintains a constant value until \sim150 K, and increases further below this temperature. However, there are apparent differences between the two samples; in AS ECO, the magnetism around TN1T_{\mathrm{N1}} develops in a rather broader temperature range, and the magnitude of λse\lambda_{\mathrm{se}} below TN1T_{\mathrm{N1}} is much smaller. These results indicate the enhancement of spin fluctuations by annealing. Upon cooling, the small oscillation component in AS ECO disappears with a reduction in foscf_{\mathrm{osc}}, as shown in Fig. 2(b). This disappearance proceeds concomitantly with the development of λse\lambda_{\mathrm{se}} toward TN2T_{\mathrm{N2}}, suggesting an interrelation between the partially ordered and fluctuating spin states. In both AS and AN ECO, bulk long-range order is observed below TN2T_{\mathrm{N2}} of 110 K, which is comparable to TN2T_{\mathrm{N2}} for TT^{\prime}-type LCO Hord2010 . Therefore, the static order with a low TN2T_{\mathrm{N2}} is a robust feature of TT^{\prime}-type RRCO. Furthermore, ω1\omega_{1} does not exist for 110 K T\lesssim T\lesssim 150 K owing to the disappearance of the small oscillation component. The extrapolation of ω1\omega_{1} at 150K<T<TN1150\>\mathrm{K}<T<T_{\mathrm{N1}} toward lower temperatures does not seem to connect smoothly with the values of ω1\omega_{1} below TN2T_{\mathrm{N2}}. This discontinuity and the disappearance of the small oscillation component immediately above TN2T_{\mathrm{N2}} suggest that the partial order at high temperatures is not a direct precursor of the static magnetic order below TN2T_{\mathrm{N2}}.

III.2 Longitudinal-field μ\muSR measurements

As mentioned in Section III.1, we found an anomalous ordering sequence above TN2T_{\rm N2} in AS ECO. Therefore, we performed LF μ\muSR measurements to obtain further information on the spin states above TN2T_{\mathrm{N2}}. Figures 3(a) and (b) show the field dependencies of the μ\muSR spectra for AS ECO at 187 and 138 K, respectively. At 187 K, where the small oscillation component was observed, the spectra in the longer time region exhibited a parallel upward shift under the magnetic fields. This shift can be observed in the decoupling process of the interaction between muon spins and spontaneous local magnetic fields at the muon sites, providing the magnitude of the spontaneous magnetic fields. In AS ECO at 187 K, an external field of 20 mT is sufficient to shift the overall spectra over the entire time range, suggesting that the static magnetic fields at the muon sites are less than 20 mT in magnitude. Therefore, the partially and weakly ordered spin state exists at 187 K. At 138 K, where the small oscillation component was absent, the spectra slightly changed with an external field of 50 mT, and relaxation remained at 395 mT, as shown in Fig. 3(b). These results indicate that a dynamical nature indeed develops at a lower temperature of 138 K. This anomalous behavior is discussed in relation to the inhomogeneous spin state in Sec. IV.3.

IV Discussion

IV.1 Common feature of magnetism

The present μ\muSR measurements demonstrate that a uniform static magnetic order exists only below TN2T_{\mathrm{N2}} (= 110 K) and a fluctuating spin state exists for TN2<T<TN1T_{\mathrm{N2}}<T<T_{\mathrm{N1}} in AS and AN ECO. These results are consistent with previous μ\muSR studies for TT^{\prime}-type LCO Luke1990 ; Hord2010 and contradictory to the absence of a fluctuating state with a higher Néel temperature (TN2TN1T_{\mathrm{N2}}\approx T_{\mathrm{N1}} = 325 K) in TT-type LCO Uemura1989 . Therefore, the sequential ordering process with a lower TN2T_{\mathrm{N2}} is a common feature of magnetism of TT^{\prime}-type RRCO. The variation of the magnetic order in the parent RRCO with different structures most likely originates from differences in oxygen coordination. The spin direction can be perturbed by the contribution of the Cu 3d3z2r2d_{3z^{2}-r^{2}} orbital to the spin correlations, causing the magnetic order to be more stabilized in TT-type RRCO having apical oxygen atoms.

The fact that AS TT^{\prime}-type ECO (aa = 3.9049 Å) and LCO (aa = 4.0102 Å Hord2010 ) with different lattice constants aa have similar magnetic behaviors is in contrast to the variation of SC properties in AN RRCO depending on the lattice constant aa. In TT^{\prime}-type R2xR_{2-x}CexCuO4 with a larger aa, the critical Ce concentration for the appearance of the SC phase is smallerNaito2002 ; Fujita2003 ; Krockenberger2008 ; Krockenberger2014 and the optimal SC transition temperature is higherNaito2002 ; Maple1990 . Thus, the appearance of non-doped superconductivity is not directly related to the lower TN2T_{\mathrm{N2}}.

We note that the neutron diffraction measurements of TT^{\prime}-type RRCO showed magnetic Bragg peaks below approximately TN1T_{\mathrm{N1}} but no evidence of an anomaly around TN2T_{\mathrm{N2}} Chattopadhyay1994 . The different ordering temperatures determined using the μ\muSR and neutron diffraction measurements originate from the different time scales to observe the spin dynamics, which are typically \sim10-4 s to \sim10-11 s and \sim10-9 s to \sim10-14 s, respectively. Therefore, a fluctuating spin state with a time scale of \sim10-4 to \sim10-9 s would be realized for TN2<TTN1T_{\mathrm{N2}}<T\leq T_{\mathrm{N1}}. (See Fig. 5.)

IV.2 Effect of annealing on long-range magnetic order

The negligible annealing effect on TN2T_{\mathrm{N2}} in TT^{\prime}-type ECO appears to be different from the annealing-induced suppression of AFM order reported for TT^{\prime}-type LECO. As mentioned in the Introduction, reduction annealing plays two roles: electron dopingAsano2018 and the removal of chemical defect Radaelli1994c ; Schultz1996 ; Kang2007 . Because the δ\delta values for the present ECO and LECO samples are comparable Asano2020 ; Takagi1989 ; Armitage2010 , the overall structural change due to annealing is expected to be the same for both samples.

Recent x-ray absorption near-edge structure measurements revealed a Mott insulating state in both AS LECO and NCO but a significant increase of electron density due to annealing (nANn_{\rm AN}) in LECO (0.4 electrons per Cu), which is much larger than that in NCO (0.05 electrons per Cu) Asano2020 . On this basis, a distinct electron-doping process was proposed: the self-doping of electrons and holes associated with the collapse of the charge-transfer gap for the former with a smaller gap size (ΔCT\Delta_{\rm CT}) and electron doping into the upper Hubbard band for the latter with a larger ΔCT\Delta_{\rm CT}. According to this scenario, the present ECO sample, which has a relatively large ΔCT\Delta_{\rm CT} Arima1991 ; Braden2005 , follows the nAN=2δn_{\rm AN}=2\delta relation Asano2018 , as is the case for NCO. Therefore, the distinct effects of annealing on the long-range magnetic order in ECO and LECO Adachi2016 can be interpreted as due to different induced carrier densities. The suppression of the ordered state in LECO is attributed to the generated charges with a large nANn_{\rm AN} value, while the negligible variation in TN2T_{\mathrm{N2}} and the small reduction of ω1\omega_{\rm 1} in ECO are likely to have originated from electron doping with a low nANn_{\rm AN} value (= 2δ\delta \sim0.04 electrons per Cu).

Refer to caption
Figure 4: Normalized time spectra of zero-field μ\muSR for as-sintered Eu2-xCexCuO4 with xx = 0, 0.07, 0.10, and 0.14 measured at approximately 180 K. All the data were acquired at MLF. The arrows indicate the small oscillation component. The solid lines indicate the best-fitting results based on Eq. (1).
Refer to caption
Figure 5: Schematics of the spin states on the CuO2 plane in the as-sintered sample for (a) higher temperatures in 150K<T<TN1150\>\mathrm{K}<T<T_{\mathrm{N1}}, (b) lower temperatures in 150K<T<TN1150\>\mathrm{K}<T<T_{\mathrm{N1}}, and (c) TN2<T<150T_{\mathrm{N2}}<T<150 K. The open circles represent the chemical defect that can function as pinning centers of spin fluctuations. The figures show the case with \sim3%\% defect. The right column shows the corresponding position dependence of the frequency of spin fluctuations along the red line. The typical time scales of neutron scattering and muon spin rotation/relaxation measurements are indicated by color bars in the top figure.

IV.3 Partially ordered spin state

Next, we discuss the partially ordered spin state in AS ECO. Because this state disappears with annealing, its origin is argued from the viewpoint of the annealing effect. To exclude the possibility of electron doping aspect of annealing treatment, we performed additional ZF μ\muSR measurements on AS Eu2-xCexCuO4 (ECCO) with xx = 0.07, 0.10, and 0.14. If the short-range magnetic order is related to the doping level, the small rotation component should vanish in AS ECCO with xx \geq 0.04, the electron density in which is greater than that in AN ECO. Figure 4 shows normalized time spectra at \sim180 K. As indicated by arrows, the small oscillation component can be observed on the dominant exponential-type decay component for xx \leq 0.10. Therefore, in contrast to long-range magnetic order, which is varied by the electron-doping effects of the annealing process, the partial order for 150 << T << TN1T_{\mathrm{N1}} is an intrinsic property of AS compounds independent of the doping level. Note that in xx = 0.14, no evidence of oscillation was detected at 181 K, because this temperature is higher than TN1T_{\mathrm{N1}}.

To understand the origin of the partial order, the appearance of oscillation immediately below TN1T_{\mathrm{N1}} is noteworthy because it indicates a close connection between the partially ordered state and the dominant fluctuating spin state. In the slowly fluctuating spin state, chemical defect could become pinning centers for the fluctuating spins and induce local magnetic order Adachi2013 . The structural feature in AS samples containing chemical defects is qualitatively consistent with the appearance of the small oscillation component below TN1T_{\mathrm{N1}}. Reduction annealing can remove such defectsRadaelli1994c ; Schultz1996 ; Kang2007 , resulting in the suppression of local magnetism and the enhancement of the dynamical spin nature. In AN ECO, the non-monotonic evolution of the muon spin relaxation suggests that a tiny amount of defects remains.

Finally, we discuss the disappearance of the small oscillation component and the anomalous enhancement in the dynamical spin nature of AS ECO for TN2<T<150T_{\mathrm{N2}}<T<150 K with cooling. When the correlation length of the pristine fluctuating spin state develops toward static order, the partially and weakly ordered region contributes to the dominant state owing to the increase of magnetic energy. Consequently, the partial order could disappear immediately above TN2T_{\mathrm{N2}}, where the correlation length of the fluctuating state diverges Endoh1989 ; Birgeneau1999 ; Motoyama2007 . Based on the present results, this thermal evolution of the magnetic behavior is schematically drawn in Fig. 5. The figure depicts the case in which \sim3%\% of pinning centers (open circle) exists on the CuO2 plane. At high temperatures below TN1T_{\mathrm{N1}}, static order is locally induced around the pinning center owing to the slowing down of spin fluctuations in the time scale of the μ\muSR measurement. The ordered region shrinks upon cooling and vanishes immediately above TN2T_{\mathrm{N2}}. The figures in the right column show the corresponding position dependence of the frequency of spin fluctuations along the red line near the pinning centers. All these states are observed as static order in the neutron scattering measurements.

V Summary

We have performed ZF and LF μ\muSR measurements on TT^{\prime}-type ECO to study the common features of magnetism in the parent compounds of electron-doped superconductors with the TT^{\prime}-structure. The effects of reduction annealing on the magnetism were also investigated. We clarified the development of spin correlations below a TN1T_{\mathrm{N1}} of 265 K and the formation of true static magnetic order in both AS and AN ECO below a TN2T_{\mathrm{N2}} of 110 K. TN2T_{\mathrm{N2}} is comparable to the ordering temperature in TT^{\prime}-type LCO and much lower than that in TT-type LCO. Thus, the sequential ordering process and low TN2T_{\mathrm{N2}} are common features of magnetism in non-superconducting TT^{\prime}-type RRCO. Furthermore, we found in AS ECO the coexistence of a dominant fluctuating spin state and partially ordered spin state for \sim150K<TTN1150\>\mathrm{K}<T\leq T_{\mathrm{N1}}. The partially ordered spin state is absent in AN ECO, indicating that this state originates from chemical defect in the AS sample.

Acknowledgments

The μ\muSR measurements at the Materials and Life Science Experimental Facility of J-PARC (Proposals No. 2013A0084, No. 2014A0205, No. 2015MP001, and No. 2017B0270) and at the RIKEN-RAL Muon Facility in the Rutherford Appleton Laboratory (Proposal No. RB1670580) were performed under user programs. We thank the staff at J-PARC and RIKEN-RAL for their technical support during the experiments. We are also thankful for helpful discussions with T. Adachi, Y. Ikeda, Y. Koike, K. M. Kojima, Y. Nambu, and K. Yamada. This work was supported by MEXT KAKENHI under Grants No. 15K17689 and No. 16H02125 as well as the IMSS Multi-probe Research Grant Program.

References

  • (1) A. Tsukada, Y. Krockenberger, M. Noda, H. Yamamoto, D. Manske, L. Alff, and M. Naito, Solid State Commun. 133, 427 (2005).
  • (2) O. Matsumoto, A. Utsuki, A. Tsukada, H. Yamamoto, T. Manabe, and M. Naito, Physica C 469, 924 (2009).
  • (3) S. Asai, S. Ueda, and M. Naito, Physica C 471, 682 (2011).
  • (4) T. Takamatsu, M. Kato, T. Noji, and Y. Koike, Appl. Phys. Express 5, 073101 (2012).
  • (5) G. M. Luke, L. P. Le, B. J. Sternlieb, Y. J. Uemura, J. H. Brewer, R. Kadono, R. F. Kiefl, S. R. Kreitzman, T. M. Riseman, C. E. Stronach, M. R. Davis, S. Uchida, H. Takagi, Y. Tokura, Y. Hidaka, T. Murakami, J. Gopalakrishnan, A. W. Sleight, M. A. Subramanian, E. A. Early, J. T. Markert, M. B. Maple, and C. L. Seaman, Phys. Rev. B 42, 7981 (1990).
  • (6) Y. J. Uemura, W. J. Kossler, X. H. Yu, J. R. Kempton, H. E. Schone, D. Opie, C. E. Stronach, D. C. Johnston, M. S. Alvarez, and D. P. Goshorn, Phys. Rev. Lett. 59, 1045 (1987).
  • (7) R. Hord, H. Luetkens, G. Pascua, A. Buckow, K. Hofmann, Y. Krockenberger, J. Kurian, H. Maeter, H.-H. Klauss, V. Pomjakushin, A. Suter, B. Albert, and L. Alff, Phys. Rev. B 82, 180508(R) (2010).
  • (8) H. Das and T. Saha-Dasgupta, Phys. Rev. B 79, 134522 (2009).
  • (9) C. Weber, K. Haule, and G. Kotliar, Nat. Phys. 6, 574 (2010).
  • (10) C. Weber, K. Haule, and G. Kotliar, Phys. Rev. B 82, 125107 (2010).
  • (11) K. M. Kojima, Y. Krockenberger, I. Yamauchi, M. Miyazaki, M. Hiraishi, A. Koda, R. Kadono, R. Kumai, H. Yamamoto, A. Ikeda, and M. Naito, Phys. Rev. B 89, 180508(R) (2014).
  • (12) Y. Krockenberger, H. Irie, O. Matsumoto, K. Y., M. Mitsuhashi, A. Tsukada, Mi. Naito, and H Yamamoto, Scientific Reports, 3, 2235 (2013).
  • (13) M. Naito, Y. Krockenberger, A. Ikeda, and H. Yamamoto, Physica C 523, 28 (2016).
  • (14) P. G. Radaelli, J. D. Jorgensen, A. J. Schultz, J. L. Peng, and R. L. Greene, Phys. Rev. B 49, 15322 (1994).
  • (15) A. J. Schultz, J. D. Jorgensen, J. L. Peng, and R. L. Greene, Phys. Rev. B 53, 5157 (1996).
  • (16) H. J. Kang, P. Dai, B. J. Campbell, P. J. Chupas, S. Rosenkranz, P. L. Lee, Q. Huang, S. Li, S. Komiya, and Y. Ando, Nat. Mater. 6, 224 (2007).
  • (17) T. Adachi, Y. Mori, A. Takahashi, M. Kato, T. Nishizaki, T. Sasaki, N. Kobayashi, and Y. Koike, J. Phys. Soc. Jpn. 82, 063713 (2013).
  • (18) T. Adachi, A. Takahashi, K. M. Suzuki, M. A. Baqiya, T. Konno, T. Takamatsu, M. Kato, I. Watanabe, A. Koda, M. Miyazaki, R. Kadono, and Y. Koike, J. Phys. Soc. Jpn. 85, 114716 (2016).
  • (19) H. Fukazawa, S. Ishiyama, M. Goto, S. Kanamaru, K. Ohashi, T. Kawamata, T. Adachi, M. Hirata, T. Sasaki, Y. Koike, and Y. Kohori, Physica C 541, 30 (2017).
  • (20) W. Jiang, J. L. Peng, Z. Y. Li, and R. L. Greene, Phys. Rev. B 47, 8151 (1993).
  • (21) J. S. Higgins, Y. Dagan, M. C. Barr, B. D. Weaver, and R. L. Greene, Phys. Rev. B 73, 104510 (2006).
  • (22) S, Asano, K. Ishii, D. Matsumura, T. Tsuji, T. Ina, K. M. Suzuki, and M. Fujita, J. Phys. Soc. Jpn. 87, 094710 (2018).
  • (23) S. Asano, K. Ishii, D. Matsumura, T. Tsuji, K. Kudo, T. Taniguchi, S. Saito, T. Sunohara, T. Kawamata, Y. Koike, and M. Fujita, arXiv:2005.10681
  • (24) M. Tovar, D. Rao, J. Barnett, S. B. Oseroff, J. D. Thompson, S-W. Cheong, Z. Fisk, D. C. Vier, and S. Schultz, Phys. Rev. B 39, 2661 (1989).
  • (25) M. Tanikawa, H. Itoh, and M. Kusuhashi, (1994) T. Fujita, Y. Shiohara. (eds) Advances in Superconductivity VI. Springer, Tokyo
  • (26) F. L. Pratt, Physica B 289-290, 710 (2000).
  • (27) T. Kubo, T. Uefuji, M. Fujita, K. Yamada, I. Watanabe, and K. Nagamine, Physica C 378-381, 354 (2002).
  • (28) M. Fujita, T. Kubo, S. Kuroshima, T. Uefuji, K. Kawashima, K. Yamada, I. Watanabe, and K. Nagamine, Phys. Rev. B 67, 014514 (2003).
  • (29) T. Chattopadhyay, J. W. Lynn, N. Rosov, T. E. Grigereit, S. N. Barilo, and D. I. Zhigunov, Phys. Rev. B 49, 9944 (1994).
  • (30) Y. Endoh, M. Matsuda, K. Yamada, K. Kakurai, Y. Hidaka, G. Shirane, and R. J. Birgeneau, Phys. Rev. B 40, 7023 (1989).
  • (31) F. Borsa, P. Carretta, J. H. Cho, F. C. Chou, Q. Hu, D. C. Johnston, A. Lascialfari, D. R. Torgeson, R. J. Gooding, N. M. Salem, and K. J. E. Vos, Phys. Rev. B 52, 7334 (1995).
  • (32) M. Naito, S. Karimoto, and A. Tsukada, Supercond. Sci. Technol. 15, 1663 (2002).
  • (33) Y. Krockenberger, J. Kurian, A. Winkler, A. Tsukada, M. Naito, and L. Alff, Phys. Rev. B 77, 060505(R) (2008).
  • (34) Y. Krockenberger, B. Eleazer, H. Irie, and H. Yamamoto, J. Phys. Soc. Jpn. 83, 114602 (2014).
  • (35) M. B, Maple, 1996, Mater. Res. Bull. 15, 60 (1990).
  • (36) H. Takagi, S. Uchida, and Y. Tokura, Phys. Rev. Lett. 62, 1197 (1989).
  • (37) N. P. Armitage, P. Fournier, and R. L. Greene, Rev. Mod. Phys. 82, 2421 (2010) and references therein.
  • (38) T. Arima, T. Kikuchi, M. Kasuya, S. Koshihara, Y. Tokura, T. Ido, and S. Uchida, Phys. Rev. B 44, 917 (1991).
  • (39) M. Braden, L. Pintschovius, T. Uefuji, and K. Yamada, Phys. Rev. B 72, 184517 (2005).
  • (40) R. J. Birgeneau, M. Greven, M. A. Kastner, Y. S. Lee, B. O. Wells, Y. Endoh, K. Yamada, and G. Shirane, Phys. Rev. B 59, 13788 (1999).
  • (41) E. M. Motoyama, G. Yu, I. M. Vishik, O. P. Vajk, P. K. Mang, and M. Greven, Nature 445, 186 (2007).