1 Introduction

Pyrochlore oxides have attracted a lot of scientific interest over the past two decades [1,2,3,4]. In pyrochlore oxides A2B2O7, the rare earth cations and the transition metal cations generally occupy A and B sites, respectively. And the A- and B-site cations form the three-dimension array of corner-shared tetrahedra, respectively. When magnetic metal cations with the inherent antiferromagnetic (AFM) interaction occupy the vertexes of tetrahedra, spins encounter the magnetic frustration. This spin frustration originating from the structure leads to highly degenerate ground states. In some geometrically frustrated antiferromagnets, the degenerate magnetic ground states are lifted by the internal or external factors such as the lattice distortion, the strain, the next-nearest-neighbor interaction, and the ionic doping, which generates a rich variety of magnetic phases including spin glass, spin liquid, spin ice, antiferromagnetism [5,6,7].

In Y2Mo2O7, nonmagnetic Y3+ and magnetic Mo4+ ions occupy A and B sites, respectively. The magnetic interaction between the nearest neighbor Mo4+ ions at B sites is AFM, so the magnetic Mo4+ ions in Y2Mo2O7 meet the geometrical spin frustration [8]. In the past years, physical properties of Y2Mo2O7 and its doping compounds have been investigated in some references [9,10,11]. It has been established that Y2Mo2O7 behaves as a nearly ideal spin glass from nonlinear susceptibility, heat capacity, AC susceptibility, elastic and inelastic neutron scattering, muon-spin relaxation, and neutron spin echo data [10,11,12,13]. The other metal ions doping in Y2Mo2O7 would change magnetic property due to the change of structure and generation of chemical disorder [14]. The magnetic property in the Ti4+-doped Y2Mo2O7 was investigated in reference [15]. In this work, the effect of Ti4+ doping in Y2Mo2O7 was investigated. The Ti4+ doping does not change the structure except slightly increasing the lattice constant in the system. Spin glass is confirmed to persist in the Ti4+-doped Y2Mo2O7, but the transition temperature is suppressed with the Ti4+ doping.

2 Experimental

Polycrystalline samples of Y2Mo2(1-x)Ti2xO7 were prepared by using the standard solid-state method. Firstly, the stoichiometric reactants of Y2O3 (99.99 wt%), MoO2 (99 wt%), and TiO2 (99.99 wt%) were thoroughly mixed. Then the mixtures were pressed into sheets and pre-fired at 1073 K under the flowing N2. Finally, they were sintered at 1523 K under the same atmosphere with the intermediate grinding. The structure was checked by using the x-ray diffractometer (XRD) with the Cu Kα radiation at room temperature. The XRD data were recorded in the 2θ range from 20° to 80°. The XRD data have been analyzed using Rietveld refinement program (GSAS). The FRIR and Raman spectra were employed to check the change of microstructure in the Ti4+-doped samples. The magnetic properties were measured by employing a superconducting quantum interference device magnetometer. Temperature dependence of magnetization was measured in the zero-field-cooling (ZFC) and the field-cooling (FC) modes under the magnetic field of 100 Oe. The AC susceptibilities at different frequencies were measured at the zero DC magnetic field and a constant AC field of 3.5 Oe from 4 to 200 K. The heat capacity at low temperatures was measured by using physical properties measurement system.

3 Results and Discussion

The experimental and the Rietveld refined XRD data of Y2Mo2(1-x)Ti2xO7 with x = 0.05, 0.10, 0.20, and 0.30 at room temperature were shown in Fig. 1. All Y2Mo2(1-x)Ti2xO7 samples present in the cubic phase with Fd3m symmetry and no impurity peaks are observed, which indicates that the Ti4+ doping does not change the crystal structure in this series of samples. Compared with the standard XRD spectra of Y2Mo2O7, the (222) peaks of Y2Mo2(1−x)Ti2xO7 with x = 0.05, 0.10, 0.20, and 0.30 shift − 0.004°, + 0.053°, + 0.144°, and + 0.155°. For x = 0.05, the (222) peak unexpectedly and slightly shifts to the low degree, which may ascribe to measurement inaccuracy from the large measurement step (0.033° per step). The Ti4+ doping shifts the (222) peak to the high degree, which means that the lattice constant becomes small. The experimental XRD data for all samples is well simulated by the Rietveld refinement. As shown in Table 1, the lattice constant (D) for Y2Mo2(1-x)Ti2xO7 decreases monotonically upon the substitution of Ti4+. This is due to that the ionic radius of Ti4+ (0.605 Å) is smaller than that of Mo4+ (0.65 Å). The bond lengths between different ions from Rietveld refinements for Y2Mo2(1−x)Ti2xO7 are shown in Table 2. When Ti4+ substitutes Mo4+, the bond lengths between most different ions in Y2Mo2(1−x)Ti2xO7 vary slightly (≤ 3.3%) except the Ti–O bond (up to 12%). The large change of the Ti–O bond length is related with the microstructure crystalline distortion resulting from the different ionic radii between Mo4+ and Ti4+.

Fig. 1
figure 1

The XRD data along with Rietveld analysis are shown for Y2Mo2(1−x)Ti2xO7 with x = 0.05, 0.10, 0.20, and 0.30

Table 1 Structural parameters from the Rietveld refinement of XRD data for Y2Mo2(1−x)Ti2xO7 with x = 0.05, 0.10, 0.20, and 0.30 at room temperature
Table 2 The bond lengths between different ions from Rietveld refinements for Y2Mo2(1−x)Ti2xO7 with x = 0.05, 0.10, 0.20, and 0.30

Figure 2 shows magnetization as a function of temperature [M(T)] for Y2Mo2(1-x)Ti2xO7 measured under 100-Oe magnetic field. The ZFC and FC M(T) curves are coincident in the high-temperature region and show irreversible at low temperatures. The irreversibility temperature (Tirr) of Y2Mo2(1-x)Ti2xO7 is 20.2, 18.1, 14.1, and 12.0 K for x = 0.05, 010, 0.20, and 0.30, respectively. The inset of Fig. 2 exhibits that Tirr decreases almost linearly with the doping content of Ti4+. Obviously, the doping of Ti4+ suppresses the irreversibility temperature. In Y2Mo2O7, the observance of magnetic irreversibility is considered to originate from the thermodynamic spin glass [16, 17]. The generation of spin glass in Y2Mo2O7 is also confirmed by neutron powder diffraction, the heat capacity, and density functional theory calculations, and suggested to be the results of the spin-orbital coupling and random fluctuations in the Mo environment at the local level [18]. However, magnetic irreversibility behavior is also found in the antiferromagnetic and ferromagnetic systems. So, it could not be concluded that spin glass occurs in the Ti4+-doped Y2Mo2(1-x)Ti2xO7 system. Temperature dependence of the inverse magnetic susceptibility [χ−1 = (M/H)−1] is shown in Fig. 3. The curves are fitted through the Curie-Weiss law χ = C/(T − θCW), where C is the Curie constant and θCW is the Curie-Weiss temperature. The obtained θCWs for all samples are all negative, which indicates that the main magnetic interaction in Y2Mo2(1-x)Ti2xO7 is antiferromagnetic.

Fig. 2
figure 2

Temperature-dependent magnetization of Y2Mo2(1-x)Ti2xO7 under the field of 100 Oe. Filled and unfilled symbols represent the ZFC and FC data, respectively. The inset shows the variation of irreversibility temperature with Ti4+ substitution

Fig. 3
figure 3

Temperature-dependent inverse susceptibility deduced from magnetization data of Y2Mo2(1−x)Ti2xO7 with x = 0.05, 0.10, 0.20, and 0.30. The solid lines show the fitting curves with the Curie-Weiss law

To further clarify the magnetic state in the Y2Mo2(1−x)Ti2xO7 system, we studied the isothermal magnetization as a function of a magnetic field [M(H)] measured at 5 K for Y2Mo2(1−x)Ti2xO7 (Fig. 4). The M(H) curves show almost linear behavior, which reflects the strong AFM interaction in this series of samples. But the M(H) curves also show a slight magnetic hysteresis in the low-field region. The typical M(H) curve of x = 0.20 as shown in the inset of Fig. 4. The magnetic hysteresis indicates that there is a weak ferromagnetic (FM) interaction in the system. Figure 5 shows the corresponding Arrott plot (M2 vs H/M) from the M(H) data [19]. For the materials with spontaneous magnetization, the Arrott plot will give a positive intercept in the high-field regime. On the other hand, a negative intercept in the Arrott plot implies an antiferromagnetic state [19, 20]. As seen in Fig. 5, the intercepts for all samples are negative, which indicates that the AFM interaction is predominated rather than the FM interaction in Y2Mo2(1-x)Ti2xO7. Therefore, the AFM interaction is predominated but the weak FM interaction also exists in Y2Mo2(1-x)Ti2xO7.

Fig. 4
figure 4

Magnetic field–dependent magnetization of Y2Mo2(1-x)Ti2xO7 at 5 K. The inset shows the detailed M(H) curves of Y2Mo1.40Ti0.60O7

Fig. 5
figure 5

The Arrott plot (M2 vs H/M) of Y2Mo2(1-x)Ti2xO7

The coexistence of the AFM and FM interactions in this series of samples would be possible to induce spin glass in the system. To demonstrate whether or not spin glass occurs in Y2Mo2(1-x)Ti2xO7, we further measured temperature dependence of AC susceptibility under zero DC magnetic field and a constant AC field of 3.5 Oe from 4 to 200 K with different frequencies (f) of 10 Hz and 100 Hz. The real part of AC susceptibility curves [χ’(T)] with different frequencies are shown in Fig. 6. A pronounced peak is observed in the χ’(T) curves for all samples, and the corresponding temperature is the transition temperature. When frequency change from 10 to 100 Hz, the χ’(T) curve at 100 Hz is coincident with that of 10 Hz and no peak shift is found for each sample. In a conventional spin glass, when increasing frequency, peak shifts to a higher temperature [8]. But in the Y2Mo2(1−x)Ti2xO7, no peak shift is found, which is unlike the conventional spin glass. In some spin glass system, the behavior of peak shift could not be detected due to that this shift is smaller than the measurement accuracy. Therefore, the spin glass could not be excluded. The transition temperature obtained from the AC susceptibility is 22, 20, 16, and 14 K for x = 0.05, 0.10, 0.20, and 0.30, respectively.

Fig. 6
figure 6

The real part of the AC susceptibility Y2Mo2(1−x)Ti2xO7 with x = a 0.05, b 0.10,c 0.20, and d 0.30 at the frequency of 10 Hz and 100 Hz

We further made the heat capacity measurements as a function of temperature [Cp(T)] for Y2Mo2(1−x)Ti2xO7 with x = 0.05, 0.10, 0.20, and 0.30, shown in Fig. 7. The measured Cp(T) curves of Y2Mo2(1−x)Ti2xO7 (x = 0.05, 0.10, 0.20, and 0.30) is similar to that of Y2Mo2O7 [18]. To get the magnetic specific heat, Y2Ti2O7 was used as a subtraction from the lattice contribution at low temperatures because Y2Mo2O7 presents a similar crystalline structure to Y2Ti2O7. Figure 8 shows temperature dependence of the magnetic specific heat [Cm(T)] of Y2Mo2(1−x)Ti2xO7 (x = 0.05, 0.10, 0.20, and 0.30). A broad peak is observed at the transition temperature for all samples. Especially, the magnetic specific heat in the low-temperature region exhibits linear behavior. This phenomenon suggests a constant density of states of the low-temperature magnetic excitations and it is also claimed to be a common feature of spin glass [11, 21, 22]. A broad magnetic specific heat also extends up to high temperatures, which indicates that short-range order contributions persist at the high temperatures [11]. This also confirms spin glass. The coexistence of AFM and FM interactions in the frustrated magnet may be the origin of spin glass in this series of samples.

Fig. 7
figure 7

The measured specific heat of Y2Mo2(1−x)Ti2xO7 with x = 0.05, 0.10, 0.20, and 0.30 at low temperatures

Fig. 8
figure 8

Magnetic specific heat of Y2Mo2(1−x)Ti2xO7 with x = 0.05, 0.10, 0.20, and 0.30 at low temperatures. The solid lines show the linear fitting curves in the low-temperature region

4 Conclusion

In summary, we have investigated the structural, heat capacity, and magnetic properties of the pyrochlore Y2Mo2(1-x)Ti2xO7. No structural phase transition is induced, but the lattice constant decreases continuously with Ti4+ doping. Spin glass persists in the Ti4+-doped Y2Mo2O7 in the low-temperature region. The Ti4+ doping suppresses the transition temperature. This is ascribed to the diluted effect of nonmagnetic Ti4+ substitution of magnetic Mo4+. The coexistence of AFM and FM interactions in the frustrated magnet may be the origin of spin glass.