Next Article in Journal
Round-the-Clock Adsorption–Degradation of Tetracycline Hydrochloride by Ag/Ni-TiO2
Previous Article in Journal
Mechanical Behavior of 3D-Printed Thickness Gradient Honeycomb Structures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Modulation of the Structural, Magnetic, and Dielectric Properties of YMnO3 by Cu Doping

College of Physics & Electronic Engineering, Xianyang Normal University, Xianyang 712000, China
*
Authors to whom correspondence should be addressed.
Materials 2024, 17(12), 2929; https://doi.org/10.3390/ma17122929
Submission received: 13 May 2024 / Revised: 7 June 2024 / Accepted: 12 June 2024 / Published: 14 June 2024

Abstract

:
The lower valence compensation of YMn1-xCuxO3 (x = 0.00, 0.05, and 0.10) is prepared by the solid-state reaction, and the effects of divalent cation Cu-doping on the construction and magnetic and dielectric attributes of multiferroic YMnO3 are systemically researched. Powder X-ray diffraction shows YMn1-xCuxO3 has a single-phase hexagonal construction with a P63cm space group as the parent YMnO3, and lattice parameters decrease systematically as Cu concentration increases. Using the scanning electric microscope, structure morphologies analysis shows that the mean grain size varies between 1.90 and 2.20 μm as Cu content increases. YMn1-xCuxO3 magnetization increases as Cu doping concentration increases, and the antiferromagnetic transition temperature declines from 71 K for x = 0.00 to 58 K for x = 0.10. The valence distributions of Mn ions conduce to the modified magnetic attributes. Due to Cu substitution, the dielectric loss and dielectric constant decline as frequency increases from 400 to 700 K, showing representative relaxation behaviors. Indeed, that is a thermally activated process. In addition, the peak of the dielectric loss complies with the Arrhenius law. The relaxation correlates to the dipole effect regarding carrier hopping between Mn3+ and Mn4+, and also correlates to oxygen vacancies generated by Mn2+.

1. Introduction

Hexagonal rare-earth manganites, YMnO3 (h-YMnO3) as a typical example of a multiferroic material, have attracted considerable interest owing to their promising physical properties and remarkable applications in spintronics and memory devices [1,2,3,4,5,6,7]. h-YMnO3 exhibits a ferroelectric transformation at TC~900 K as well as an antiferromagnetic order below TN~70 K. h-YMnO3 crystal consists of MnO5 trigonal bipyramids, separated by Y3+ ions along the c-axis [8,9]. MnO5 bipyramids are formed by Mn3+ ions, two apical oxygen atoms (O1 and O2), and three in-plane oxygen atoms (O3 and O4). MnO5 trigonal bipyramids tilt to the ab-plane, leading to the manganese ion displacing from the bipyramid center, which causes the Mn3+ triangular network lattice to distort [10,11]. The antiferromagnetic ordering originates from the super-exchange interactions between Mn3+ in the ab-plane. Simultaneously, to maintain the stability of the structure, the bond lengths of the Y-O at the top are not equal, which leads to polarization and ferroelectricity. The ferroelectricity and magnetism originating from various sources result in substantial variations in the ferroelectric transition temperature and spin-ordering temperature. Moreover, the application possibilities of YMnO3 are limited by the weak coupling of electricity and magnetic fields, which makes this material much more attractive [1,3,12,13].
To enhance magnetoelectric coupling and comprehend the multiferroic mechanism in YMnO3, many studies have valued the impact of substitution on the Mn or Y sites. Elemental doping is among the top crucial and common methods of altering physical attributes. The change in the tilting of the MnO5 trigonal bipyramid can be tuned by doping ions with distinct radii and variable valences at the Y or Mn sites. Many studies have been conducted on dielectric behaviors, electrical properties, and magnetic properties of YMnO3 by replacing the elements like Sr, Ca, Dy, Lu, Gd, or Zr at the Y site, or Mn, Mn, Al, Co, Ti, etc., at the Mn site [10,11,13,14,15,16,17,18,19,20,21,22,23,24,25,26]. For instance, when doping with Lu at the YMnO3 Y site, the dielectric constant achieved the pinnacle within the solid solution area, but the Curie temperature did not alter [27]. As reported by Rajesh, divalent Sr2+ ions substitute Y3+ ions, which introduces the ferromagnetic (FM) duel exchange of Mn3+-O2−-Mn4+ and reduces the magnetic frustration ratio f = |θCW|/TN. The influence of Sr doping on the magnetic properties of YMnO3 suggests the redirection of the Mn moment in the basal plane perpendicular to the initial orientation at ~38 K [20]. Dielectric research on Y1-xDyxMnO3 illustrates that the dielectric constant increases significantly as Dy doping increases, and double ionization oxygen vacancy relaxation occurs at 510 K [21]. Dielectric spectroscopy analysis of Ba-doped YMnO3 by Jyoth et al. showed that the dissipation factor and dielectric constant decreased as frequency increased, which denoted that space charge polarization plays an important role [17].
According to the literature survey, doping studies on the YMnO3 have mostly concentrated on the Mn sites. By substituting Mn3+ with magnetic ions like Ru3+, Fe3+, Ni3+, Co2+, Os2+, etc., the magnetic properties have been investigated intensively [10,11,12,21,22,23,24,25,26]. Perhaps, a new state emerges from the frustrated state, on account of diverse perturbations within a two-dimensional triangular network through doping [27]. Thus, it is interesting to investigate the modification of the ferroelectric and antiferromagnetic properties and magnetoelectric coupling in YMnO3 compounds. Recently, a large exchange bias effect in Fe-doped YMnO3 films at 2 K has been reported. This exchange bias is attributed to the interaction of the antiferromagnetic and spin glass states at low temperatures [28]. Olivera reported that the leakage currents were decreased by the partial substitution of Mn3+ with Ti4+ in YMnO3; however, the ferroelectric response was not improved remarkably. The increased magnetization and weak ferromagnetism indicated that the antiferromagnetic ordering between Mn3+ ions was suppressed by the nonmagnetic Ti4+ in YMnO3 [12]. However, there have been few reports on the substitution of Cu2+ ions in YMnO3. The orthogonal YMn2/3Cu1/3O3 electronic structure studied by Aoskan et al. using synchronous radiation X-ray absorption showed that doping with bivalent Cu2+ increased the valence state of Mn3+ ions and simultaneously enhanced the non-localization of eg electrons of Mn 3d orbitals, resulting in improved conductivity in the sample [29]. As observed by Jeuvrey et al., Cu doping affects the magnetic attributes of YMnO3, and the antiferromagnetic transition temperature of Cu-doped YMnO3 decreased significantly [30]. However, Xiao et al. suggested that substituting Mn3+ ions with Cu2+ ions had little effect on the antiferromagnetic transition temperature [31]. So far, research on the structure of Cu-doped hexagonal YMnO3 is insufficient, and the mechanism of the hybridization of Mn and O ions and the impact of varying electronic structures on the magnetic attributes of YMnO3 are still unclear. Gutierrez’s research indicates that when Cu doping exceeds 20%, the sample exhibits a perovskite structure of YMnO3 [29]. In this work, YMnO3 with a doping amount of less than 10% of Cu ions was selected for investigation. The Cu concentration is high enough to show the effect of Cu doping on the structural, magnetic, and dielectric behaviors and can avoid any instability in the crystal structure. The Cu-doped h-YMnO3 ceramic was synthesized through the solid-state reaction method. The structural, magnetic, as well as dielectric behaviors are systematically studied.

2. Experimental Details

Ceramic polycrystalline samples of YMn1-xCuxO3 (x = 0.00, 0.05, and 0.10) were synthesized by a typical solid-state reaction technique using stoichiometric quantities of high-purity Y2O3, Mn2O3, and CuO. The mixed raw materials were calcined in air at 1100 °C for 24 h by interval grinding. To improve the homogeneity, the precursor powders were compacted into tiny pellets under 10 tons of hydraulic pressure and 5% polyvinyl alcohol as an adhesive. The samples were heated up to 1035 °C and held for 12 h. The crystalline structures of all samples were characterized in the angular range from 10° to 80° with a step of 0.02° and scanning speed of 2°/min. through X-ray powder diffraction (XRD) with Cu Kα radiation (λ = 1.5460 Å). The XRD modes were validated through Rietveld analysis with the FullProf project. The surface topography of the samples was examined through scanning electron microscopy (SEM). Moreover, the distribution of grain size was gauged through Nano Measurer software (Nano Measurer 1.2). X-ray photoelectron spectroscopy (XPS) with Al Kα radiation was used to analyze the variation of ionic valence state for Cu-doped YMnO3 powder, and the obtained curves were fitted with XPS-PEAK4.1 software according to the Gauss–Lorentz line. The binding energy was corrected using the neutral carbon peak of C 1 s, which was assigned a value of 284.6 eV to compensate for surface charge effects. Magnetization was performed through a Superconducting Quantum Interference Device (SQUID) at 2 K-300 K, showing a magnetic field of H = 1 kOe in zero-field cooling (ZFC) and field-cooling (FC) patterns. Dielectric attributes were determined in the frequency of 1 kHz to 1 MHz at 300 K-900 K through an HP4284LCR meter.

3. Results and Discussion

The X-ray diffraction (XRD) patterns for synthesizing YMn1-xCuxO3 (x = 0.00, 0.05, and 0.10) powders sintered at 1350 °C for 30 h are shown in Figure 1. As verified through JCPDS card no. 25-1079, the total samples exhibit a hexagonal structure showing the space group P63cm. No other diffraction peaks of impurities are observed in the XRD patterns. It is noticeable that the diffraction peaks of the XRD pattern (x = 0.05 and 0.10) slightly shift to higher 2θ values compared with x = 0.00, which proves that Cu has been successfully doped into YMnO3. The (112) and (110) peaks of the x = 0.00 shift towards higher 2θ values are shown in Figure 1a. This suggests that the lattice parameter decreases with Cu doping. Rietveld refinement was implemented to obtain the explicit structural information of the total samples. As shown in Figure 1b, a schematic depiction of the atomic arrangements within the hexagonal lattice of YMnO3 exhibits a P63cm space group. In the XRD structure refinement, the Y ions occupied two specific sites: Site 2a, located at (0, 0, z), and Site 4b, located at (1/3, 2/3, z). Simultaneously, the Mn/Cu ion was situated at Site 6c, surrounded by five oxygen ions, designated by (x, 0, z). The five oxygen ions at Sites 2a, 4b, and 6c have coordinates of (0, 0, z), (1/3, 2/3, z), and (x, 0, z), respectively [32]. Rietveld refinement for the samples was finished through Fullprof software (FullProf 2020.6), and the peaks were fitted with the Pseudo–Voigt function. Refined patterns are shown in Figure 1c. In addition, refinement parameters, such as Rwp, Rp, and Rexp, were considered in terms of refinement course. Refined parameters are displayed in Table 1. The ionic radius of Cu2+ is 0.73 Å which is larger than Mn3+ (0.65 Å), Mn4+ (0.53 Å), and Mn2+ (0.67 Å). The decrease in lattice parameters with increasing Cu2+ amount, as shown in Table 1, suggested that partial substitution of Cu2+ ions with Mn3+ in YMnO3 will not only result in a change in the blended-valence state of Mn ions, but also cause a chemical stress effect. Compared with parent YMnO3, most Mn-O bond lengths increased with increasing Cu concentration. The Y-O bond length of parent YMnO3 was longer than that of Cu-doped YMnO3, which illustrates that Cu doping raises Mn trimerization in the ab-plane and decreases the c structure parameter [31]. The tolerance factor, which determined the stability of prepared perovskite ceramic, increased from 0.891 for x = 0.00 to 0.893 for x = 0.10. The addition of Cu ions can cause a decrease in the average ionic radius of Mn sites, which is the main reason for the increase in the tolerance factor. This leads to the formation of ferromagnetic properties in YMnO3 [33]. This validates that a stable perovskite structure can be formed for all prepared ceramic samples. Thus, the increase in the Cu2+ ions with x may exhibit that the distortion of the samples intensifies.
In view of the isotropic essence of the crystal, the microstrain and average crystallite size are achieved through the Williamson–Hall equation:
β h k l cos θ = K λ D v + 4 ε sin θ
where Dv means the volume-weighted crystallite size, K means the shape factor, and ε means lattice strain. As exhibited in Figure 2, a graph is drawn through 4sinθ along the X-axis and βhklcosθ along the Y-axis for the samples (x = 0.00, 0.05, and 0.1). The lattice strain and crystallite size were derived from a linear fit. The estimation values of the strain and size are depicted in Table 1. The crystallite size declines as Cu content increases. At the same time, the lattice strain varies between 6.03 × 10−4 for x = 0.00 and 8.73 × 10−4 for x = 0.10.
Figure 3 shows the scanning electron micrograph (SEM) of the morphology of YMn1-xCuxO3 powders annealed at 1350 °C and the distribution histogram of the particle size through Nano Measurer software. Figure 3a shows larger particles and cracks in the x = 0.00 sample. The generation of cracks may be related to the heat treatment temperature. As shown in Figure 3b,c, as the amount of Cu doping increases, the sample particles gradually decrease and most of the uniform particles are mutually necked. From the histogram corresponding to SEM, it can be seen that the grain size of 90% of the particles in each sample varies in the range of approximately 1–4 μm. For x = 0.00, most particles were distributed between 2 and 2.5 μm in the sample. With increasing Cu doping, particles of 1.2 to 2 μm appeared in the samples. Thus, the mean grain size decreases from 2.14 μm for undoped YMnO3 to 1.88 μm for x = 0.10, which corresponds to the results through the Williamson–Hall equation fitting in XRD. However, all sample grain sizes highly exceed the crystallite sizes estimated by the Williamson–Hall equation. This is because the grain size by XRD analysis is the coherent domain of microcrystals, whereas the average grain analysis by SEM is formulated by aggregating multiple crystallites in the sintering procedure [10,15].
X-ray photoelectron spectroscopy (XPS) has become an efficient tool for studying valence-change ions [22]. Figure 4 shows the XPS spectra of the Mn 2p and O 1s regions of YMn1-xCuxO3 (x = 0.00, 0.05, and 0.10) at ordinary temperature. The Mn 2p peak is divided into 2p3/2 and 2p1/2, on account of spin–orbit coupling. The Mn 2p3/2 peak can be deconvoluted into differentiated peaks through Gaussian–Lorentzian curve fitting, which indicates that there are various valence Mn ions in the samples. The peaks at a binding energy of 641.28 and 643.67 eV are consistent with Mn3+ and Mn4+ oxidation states for x = 0.00, as shown in Figure 4a. In light of the fitting outcomes, the Mn3+:Mn4+ ratio for the YMnO3 ceramic is 75:25. This is close to the values reported in the literature [22,34]. The energy corresponding to 639.12 eV is the peak of the Mn2+ oxidation state, which is detected in the samples of x = 0.05 and 0.10. By calculating the peak region, the proportions of Mn2+:Mn3+:Mn4+ are 14:56:30 and 10:47:43 for x = 0.05 and 0.10, respectively, which illustrates that the trivalent state predominates in the mixed-valence manganese ions. The Mn4+ concentration increases as Cu doping increases. On account of the charge compensation, the introduction of Cu2+ (3d9) into the YMnO3 system transforms a portion of Mn3+ (3d4) into Mn4+ (3d3). The existence of Mn2+ states is associated with oxygen deficiencies caused by doping or the sample fabrication process.
The O1s spectrum revealed the hybridization of O2p with Mn 3d ions. The oxygen vacancy concentration changes with Cu content, which is validated by XPS analysis. Figure 4d–f exhibit the O1s core-level XPS spectra of Cu-doped YMnO3 with x = 0.00, 0.05, and 0.10. The main peak at approximately 529.50 eV, labeled as O1s, is attributed to the lattice oxygen in its normal position, with a reported range of 528–530 eV. [22]. Close to the O1s peak, another peak around 531.5 eV is ascribed to oxygen vacancies originating from the heat treatment or lattice distortion in the YMnO3 structure due to Cu doping [34]. The high-energy peaks located at the banding energy of 534.5 eV are assigned to the chemisorption of oxygen or adsorbed oxygen species from OH on the surface. XPS analysis results show that the relatively large contribution of the peak corresponds to the banding energy at 531.5 eV as Cu doping concentration increases, which denotes fewer oxygen vacancies in YMn1-xCuxO3 series samples. This is most probably attributed to lattice distortion within the YMnO3 structure, which was induced by the incorporation of Cu. These oxygen vacancies are expected to produce free carriers (electrons), which will contribute to increasing the exchange interaction effects between magnetic impurities.
Figure 5 shows magnetization versus temperature curves at 1000 Oe for YMn1-xCuxO3 (x = 0.00, 0.05, and 0.10) with increasing temperature between 5 K and 300 K in zero-field-cooled (ZFC) and field-cooled (FC) patterns. As depicted in Figure 5a, the ZFC and FC magnetization for YMnO3 rises as temperature declines and there exists a divergence between FC and ZFC data below TN~71 K. By comparison with the ZFC curve, the FC data rose slightly, which should be ascribed to a weak ferromagnetic component in the YMnO3 ceramics [26], which aligns with former reports on magnetic measurements of YMnO3 [23,26]. A similar change in ZFC-FC magnetization is also examined in x = 0.05 and 0.10, while the transition temperature TN decreased to 58 K, rapidly. The antiferromagnetic transition temperature TN decreased from 71 K to 58 K, which is consistent with the findings reported by Xiao et al. [26]. Moreover, as the amount of Cu doping was increased, the magnetization intensity increased over the entire temperature range. This notable difference at low temperatures implies an increase in the size of ferromagnetic clusters in the Cu-substituted sample. According to the charge compensation, introducing Cu2+ at the Mn site will generate Mn4+ in the sample. Based on the XPS analysis of the O element, oxygen vacancies gradually decreased with an increase in the doping amount. The origin of Mn2+ ions in the YMnO3 ceramics is due to the presence of oxygen vacancies, which is common in perovskite oxides. Three of the four electrons in Mn3+ (3d4) occupy the t2g energy level, while the remaining one enters the eg level. For Mn4+ (3d3) ions, three electrons are in the t2g orbital. Since for Mn2+ (3d5), there is no crystal field splitting, instead of an eg and t2g orbital, now there is a d5 orbital with five electrons. Thus, the exchange interactions eg-t2g (Mn3+-O2−-Mn4+), eg-d5 (Mn3+-O2−-Mn2+), and t2g-d5 (Mn4+-O2−-Mn2+) exist in the Cu-doped sample. At low doping levels, Hunds coupling is strong and the eg spin is determined by both eg-t2g and eg-d5 interactions. XPS analysis revealed that the concentration of Mn3+ ions in the sample decreased, as the Cu content increased. This result indicates that as the Cu-doping content increases, and the ferromagnetic interactions between Mn3+-O2−-Mn4+ and Mn3+-O2−-Mn2+ in the sample increase. Then, the magnetization of the sample increases with the increase in the doping amount of Cu2+ ions.
Figure 6 shows the inverse susceptibility (χ−1) versus T at a magnetic field of 1000 Oe for the samples in the FC mode. The χ−1(T) curve is capable of being fitted through the Curie–Weiss law above 150 K (paramagnetic state).
χ = C / T θ C W
where C refers to the Cure constant and θCW refers to the Cure–Weiss temperature.
The lines represent the Curie–Weiss fit of the results. Using Curie constant values, the magnetic moment of each Mn atom (μeff) is computed as below:
μ e f f = 3 k B C N A μ B 2
in the above equation, x, N, and kB mean the Cu atom concentration, the Avogadro quantity, and the Boltzmann constant, respectively.
The magnetic features are listed in Table 2. The total samples possess negative θCW, which denotes the antiferromagnetic interactions are still predominant in Cu-doped YMnO3 ceramics. However, it was discovered that the |θCW| decreases from 456.714 K for x = 0.00 to 181.509 K with 10%-doped Cu2+ ions. This result indicates that fewer Mn3+ ions and Mn3+-O2−-Mn3+ antiferromagnetic super-exchange interactions in the triangular network were suppressed by Cu2+ doping. The apparent improvements in Mn3+-O2−-Mn4+ and Mn3+-O2−-Mn2+ ferromagnetic double-exchange are conducive to the improved weak ferromagnetic attributes with increasing concentration of Cu, even though weaker Cu2+-O2−-Mn3+ and Cu2+-O2−-Mn4+ antiferromagnetism are also introduced in Cu-doped samples.
The experimental values of the magnetic moment per Mn atom decreased from 5.249 μB to 4.859 μB with increasing Cu content. The decrease in μeff value in the experiment is consistent with that of the theoretical results. This decrease might have resulted from the introduction of electrons by Cu3+ (3d9). The effective magnetic moment of Cu2+ is 1.7 μB smaller than that of Mn ion. On the other hand, the substitution of Mn3+ ions by divalent Cu2+ ions introduces the Mn4+ ions to the crystalline structure according to the charge compensation theory. The XPS analysis exhibits Mn3+ and Mn4+ in Cu-doped YMnO3 compounds, showing that Mn4+ increased with doping, which is in agreement with the theory. The Y ions did not contribute magnetically to the system. Substitution of Cu2+ at the Mn3+ site results in one electron of the Mn3+ (3d4) ion in YMn1-xCuxO3 being captured by Cu2+ (3d9) to form a more stable Cu+ (3d10). The Mn4+ ‘carries’ a hole that can hop in the Mn ions by hybridizing Mn d-states with the O p-states [11]. The Mn3+ component simultaneously transforms into Mn4+. The efficient magnetic moments of Mn3+ (3d4) and Mn4+ (3d3) reach 4.9 and 3.87 μB, respectively. The theory’s effective magnetic moment (µeff) was 4.900, 4.741, and 4.580 μB for YMn1-xCuxO3 with x = 0.00, 0.05, and 0.10, respectively. For pure YMnO3, the experimental effective magnetic moment is smaller than the theoretical value mainly because of the formation of holes at the oxygen sites during the sintering process, captured by Mn3+ ions to form Mn4+ ions. The XPS results also confirmed the presence of Mn4+ ions. This result is consistent with those of other studies [18,22,34]. However, the experimental effective magnetic moment of Cu-doped YMnO3 exceeds the theoretical value, predominantly owing to the Mn2+ of the sample.
Chemical substitution is a key influencing factor of geometrical frustration. It is widely known that the inter-plane coupling neglects YMnO3, since the closet adjacent inter-plane coupling contributes less to 3 × 3 structures (the distance between neighboring Mn atoms reaches 3a), as exhibited in Figure 7a. As a measurement of geometrical frustration of the antiferromagnetic system, the frustration parameter (f) is regarded as the proportion of the Curie–Weiss constant to the Néel temperature:
f = | θ C W | / T N
It reduces from 6.432 for x = 0.00 to 3.129 for x = 0.10., illustrating that Cu doping reduces the geometrical frustration of the triangular lattice. Accordingly, stringent antiferromagnetic ordering occurs.
The Mn ion in hexagonal YMnO3 is located at the center of the MnO5 triangular double pyramid. As exhibited in Figure 7b, the 3d orbitals of Mn ions are divided into two doublets, e1g (yz/zx) and e2g (xy↓/x2-y2), and a singlet, a1g (z2↓). The inter-plane AFM super-exchange interaction (Jnn) between Mn-O3-Mn and Mn-O4-Mn is formulated through the overlap between Mn 3d(x2-y2) or dxy and O(3, 4) 2pxy. The Mn-O1-O2-Mn AFM super-exchange interplay takes place through the overlap between the O(1, 2) 2pz orbital tails [35,36]. Rietveld results show that the bound distance between Mn and O along the c-axis is larger than that in the ab-plane. Thus, we consider the ab-plane to have a major impact. The bond length of Mn-O3 decreases with the increase in Cu doping. Due to orbital hybridization between Mn and O3, this result indicates a decrease in the number of electrons in the e2g (xy↓/x2-y2) orbital. Thus the Coulomb interaction between Mn and O3 is weakened. Furthermore, the inter-plane AFM super-exchange interaction weakened with increasing Cu. The TN are 71, 65, and 58 K for x = 0.00, 0.05, and 0.10, respectively, which can be ascribed to the decline in antiferromagnetic exchange interplay. Meanwhile, magnetization is enhanced because the Mn3+ transforms into Mn4+ ions.
According to the XXZ triangular antiferromagnetic model, the magnetic transition temperature is computed as follows:
T N = t J nn ( S + 1 / 2 )
where t aligns with the important temperature, and Jnn and S can be achieved from the following equations:
θ C W = 1 3 Z J nn S S + 1
S = 1 x y S M n 3 + 2 + x S M n 4 + 2 + y S Mn 2 + 2
where Z = 6 stands for the quantity of the closest neighbors, and S is equal to 3/2, 2, and 5/2 for Mn4+, Mn3+, and Mn2+, respectively. Additionally, x and y stand for the ratios of Mn4+ and Mn2+, respectively, and they are determined by the area under the XPS binding energy curve.
The exchange integral Jnn was −38.807, −26.298, and −16.414 for x = 0.00, 0.05, and 0.01. A negative value indicates that the antiferromagnetic interaction is predominant. The decreased value indicates a competition between ferromagnetic and antiferromagnetic interactions, owing to antiferromagnetic exchange Mn3+-O2−-Mn3+ interactions which are strong yet have a long range, whereas ferromagnetic exchange Mn3+-O2−-Mn4+ and Mn3+-O2−-Cu2+ interactions are strong yet have a short range [23]. The experiment shows that the t value increases from 0.731 to 1.469, indicating that copper substitution favors spin alignment in the ab-plane.
To explore the influence of Cu2+ on the dielectric behavior of YMnO3, dielectric measurements were implemented from 300 K to 900 K. Figure 8 shows the temperature dependence of dielectric constant and loss (x = 0.05 and x = 0.10) samples determined at discrepant frequencies from 1 KHz to 1 MHz, respectively. The overall dielectric constant values of x = 0.10 are higher than those of the x = 0.05. The dielectric constant and dielectric loss decrease as frequency rises at the identical temperature. An obvious dielectric relaxation effect can be examined from 400 K to 700 K; in addition, the dielectric relaxation peak shifts to more elevated temperatures as frequency rises. This result is consistent with the findings of Ma et al. on the dielectric relaxation of YMnO3, as well as other researchers’ findings on the impact of Co and Sr doping on dielectric properties [16,20,37]. To expand a deeper awareness of the dielectric relaxation of Cu-doped YMnO3, the activation energy Ea was calculated according to the Arrhenius law:
f = f 0 exp ( E a / k B T )
where f0 stands for the characteristic relaxation frequency at limited temperatures, Ea stands for the activation energy, kB stands for the Boltzmann parameter, and T stands for the peak temperature.
Figure 9 exhibits the peak temperature as a function of frequency, and experimental data are greatly fitted to the Arrhenius law. More specifically, the activation energies (x = 0.05 and x = 0.10) reach 0.76 eV and 1.23 eV, respectively. A rise in doping concentration triggers various compensatory mechanisms. As reported by Moure et al., 0.36 eV is achieved through DC conductivity measurements for the temperature of 330 to 500 K, and that conduction mechanism is thermally activated hopping of small polarons between local positions of Mn3+ and Mn4+ [38]. The oxide vacancy is an important factor for the activation energy is 0.76 eV between 450 and 720 K for Cr-doped YMnO3 [10]. Generally, ionic valence significantly affects the dielectric attributes of the material. For Cu-doped samples, the XPS outcomes indicate that Mn2+, Mn3+, and Mn4+ emerge simultaneously in Cu-doped samples, and that Mn4+ increases as Cu doping increases. The valence from Mn3+ to Mn4+ occurs, on account of the hole conduction mechanism. This suggests that Cu2+ might trigger an increased transition from Mn3+ to Mn4+ ions, and increase charge hopping. Therefore, the dielectric constant and activation energy emerge.
The activation energy of Cu-doped YMnO3 is much larger than that of Cr-doped samples reported in the literature, which may be mainly related to other factors. Generally, an increase in doping concentration induces a range of compensation mechanisms, such as electrical, B-site, and oxide vacancy compensation. These mechanisms can result in significant polarization under an applied electric field. XPS analysis showed that the content of oxygen vacancies in the sample decreased from 41% for x = 0.05 to 10% for x = 0.10. The presence of oxygen vacancies primarily results from lost traces of oxygen during sintering at higher temperatures, and electrons are created. It may be represented by the following equation:
V O × O O + 2 e ' + 1 2 O 2
The reduction in oxygen vacancies results in decreased ionic conductivity and increased activation energy, which weakens the dielectric response. Meanwhile, the generated electrons are captured by Mn3+ to form Mn2+.
Mn 3 + + e Mn 2 +
The ordering of Mn2+ and Mn3+ can produce the polar clusters and the thermally activated dielectric relaxation will be intensified. Thus, the primary factors responsible for the increase in dielectric constant and dielectric relaxation, as well as the relaxation activation energy increasing with Cu doping, are not only the strengthened dipole effect between Mn3+, Mn2+, and Mn4+ due to Cu doping, but also the decreased oxygen vacancies, which is a significant mechanism contributing to this outcome.

4. Conclusions

Hexagonal YMn1-xCuxO3 ceramics were compounded using a solid-state reaction. The magnetic and dielectric properties of YMnO3 varied with Cu doping. X-ray diffraction confirmed the formation of a single hexagonal structure in YMn1-xCuxO3. An increase in the weak ferromagnetic component was detected in the Cu-doped sample. The Curie–Weiss temperature decreased because of the weakened AFM exchange interaction of Mn3+-O2−-Mn3+ with Cu2+. The dielectric activation energies enhance with Cu concentration, indicating that the Cu-doped YMnO3 samples are controlled by the hopping charge carriers between Mn3+, Mn2+, and Mn4+. Oxygen vacancies are another important factor that causes an increase in the activation energy and dielectric constant with doping.

Author Contributions

Conceptualization, F.W.; data curation, F.W.; investigation, F.W. and Q.G.; validation, X.H. and Q.G.; writing—original draft, F.W. All authors have read and agreed to the published version of the manuscript.

Funding

Shaanxi Provincial Basic Research Program of Natural Science, China (2020JQ-898).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors have no relevant financial or non-financial interests to disclose.

References

  1. Ponet, L.; Artyukhin, S.; Kain, T.; Wettstein, J.; Pimenov, A.; Shuvaev, A.; Wang, X.; Cheong, S.W.; Mostovoy, M.; Pimenov, A. Topologically protected magnetoelectric switching in a multiferroic. Nature 2022, 607, 81–85. [Google Scholar] [CrossRef]
  2. Otomo, M.; Hasegawa, T.; Asakura, Y.; Yin, S. Remarkable effects of lanthanide substitution for the Y-site on the oxygen storage/release performance of YMnO3+δ. ACS Appl. Mater. Interfaces 2021, 13, 31691–31698. [Google Scholar] [CrossRef]
  3. Skjærvø, S.H.; Wefring, E.T.; Nesdal, S.K.; Gaukås, N.H.; Olsen, G.H.; Glaum, J.; Tybell, T.; Selbach, S.M. Interstitial oxygen as a source of p-type conductivity in hexagonal manganites. Nat. Commun. 2016, 7, 13745. [Google Scholar] [CrossRef]
  4. Park, P.; Park, K.; Oh, J.; Lee, K.H.; Leiner, J.C.; Sim, H.; Kim, T.; Jeong, J.; Rule, K.C.; Kamazawa, K.; et al. Spin texture induced by non-magnetic doping and spin dynamics in 2D triangular lattice antiferromagnet h-Y (Mn, Al) O3. Nat. Commun. 2021, 12, 2306. [Google Scholar] [CrossRef] [PubMed]
  5. Tzschaschel, C.; Satoh, T.; Fiebig, M. Tracking the ultrafast motion of an antiferromagnetic order parameter. Nat. Commun. 2019, 10, 3995. [Google Scholar] [CrossRef]
  6. Kumar, P.A.; Kumar, A.; Kumar, K.; Singh, P.; Babu, G.A.; Vijayakumar, P.; Ganesamoorthy, S.; Ramasamy, P.; Pandey, D. Growth and characterization of Dy1−xYxMnO3 single crystals by optical floating zone technique: A combined X-ray diffraction and DC magnetization study. J. Cryst. Growth 2021, 565, 126152. [Google Scholar] [CrossRef]
  7. Xing, Z.; Weng, J.; Ma, C. Improvement on microwave dielectric properties of Zn2GeO4 via gallium doping. J. Mater. Sci. Mater. Electron. 2022, 33, 11625–11631. [Google Scholar] [CrossRef]
  8. Martins, H.P.; Mossanek, R.J.O.; Martí, X.; Sanchez, F.; Fontcuberta, J.; Abbate, M. Mn 3d bands and Y-O hybridization of hexagonal and orthorhombic YMnO3 thin films. J. Phy. Condens. Mat. 2017, 29, 295501. [Google Scholar] [CrossRef]
  9. Cheng, S.; Deng, S.Q.; Zhao, Y.G.; Sun, X.F.; Zhu, J. Correlation between oxygen vacancies and sites of Mn ions in YMnO3. Appl. Phys. Lett. 2015, 106, 062905. [Google Scholar] [CrossRef]
  10. Saxena, P.; Mishra, A. Structural and electrical properties of YMnO3 manganites: Influence of Cr ion doping. J. Solid State Chem. 2021, 301, 122364. [Google Scholar] [CrossRef]
  11. Jiang, N.; Zhang, X. Atomistic simulation of Mn-site substitution in multiferroic h-YMnO3. J. Phy. Condens. Mat. 2012, 24, 235402. [Google Scholar] [CrossRef] [PubMed]
  12. Zemljak, O.; Golić, D.L.; Počuča-Nešić, M.; Dapčević, A.; Šenjug, P.; Pajić, D.; Radošević, T.; Branković, G.; Branković, Z. Titanium doped yttrium manganite: Improvement of microstructural properties and peculiarities of multiferroic properties. J. Sol-Gel Sci. Technol. 2022, 103, 807–819. [Google Scholar] [CrossRef]
  13. Wan, F.; Bai, X.; Song, K.; Han, X.; Zheng, J.; Lin, X.; Cao, C. Effect of Co substitution on magnetic and magnetocaloric properties in multiferroic hexagonal YMnO3. J. Mater. Sci. Mater. Electron. 2017, 28, 15819–15825. [Google Scholar] [CrossRef]
  14. Srivastava, S.K.; Kar, M.; Ravi, S.; Mishra, P.K.; Babu, P.D. Magnetic properties of electron-doped Y1−xCexMnO3 compounds. J. Magn. Magn. Mater. 2008, 320, 2382–2386. [Google Scholar] [CrossRef]
  15. Karoblis, D.; Zarkov, A.; Garskaite, E.; Mazeika, K.; Baltrunas, D.; Niaura, G.; Beganskiene, A.; Kareiva, A. Study of gadolinium substitution effects in hexagonal yttrium manganite YMnO3. Sci. Rep. 2021, 11, 2875. [Google Scholar] [CrossRef]
  16. Ma, Y.; Wang, X.; Wang, Z.; Liu, H.; Wang, Z.; Yang, C.; Zou, Q.; Zeng, W. Investigation of magnetic, dielectric and electrical properties of Ce-substituted YMn0.8Fe0.2O3 multiferroic ceramics. J. Mater. Sci. Mater. Electron. 2015, 26, 398–404. [Google Scholar] [CrossRef]
  17. Shukla, J.; Mishra, A. Influence of Ba2+ doping on structural and electrical transport properties of YMnO3 ceramics. J. Supercond. Nov. Magn. 2021, 34, 451–459. [Google Scholar] [CrossRef]
  18. Zou, Q.; Ma, Y.; Wang, X.; Wang, Z.; Liu, H.; Yang, C. Effect of zirconium substitution on the dielectric and magnetic properties of YMn0. 8Fe0.2O3 multiferroic ceramics. J. Mater. Sci. Mater Electron. 2017, 28, 2107–2112. [Google Scholar] [CrossRef]
  19. Muneeswaran, M.; Jang, J.W.; Choi, B.C.; Jeong, J.H.; Giridharan, N.V. Structural, optical and multiferroic properties of pure and Dy modified YMnO3. J. Mater. Sci. Mater Electron. 2017, 28, 16788–16796. [Google Scholar] [CrossRef]
  20. Thakur, R.K.; Thakur, R.; Shanmukharao Samatham, S.; Kaurav, N.; Ganesan, V.; Gaur, N.K.; Okram, G.S. Dielectric, magnetic, and thermodynamic properties of Y1−xSrxMnO3 (x = 0.1 and 0.2). J. Appl. Phys. 2012, 112, 104115. [Google Scholar]
  21. Quintana-Cilleruelo, J.Á.; Veerapandiyan, V.K.; Deluca, M.; Algueró, M.; Castro, A. Mechanosynthesis of the whole Y1-xBixMn1-xFexO3 perovskite system: Structural characterization and study of phase transitions. Materials 2019, 12, 1515. [Google Scholar] [CrossRef] [PubMed]
  22. Durán, A.; Herbert, C.; García-Guaderrama, M.; Mata, J.; Tavizón, G. Zr and Mo doped YMnO3: The role of dopants on the structural, microstructural, chemical state, and dielectric properties. Ceram. Int. 2022, 48, 17009–17019. [Google Scholar] [CrossRef]
  23. Paul, P.; Rajarajan, A.K.; Debnath, A.K.; Rao, R.; Rao, T.C. Valence fluctuation and magnetic frustration in Ga substituted YMnO3. J. Magn. Magn. Mater. 2020, 503, 166617. [Google Scholar] [CrossRef]
  24. Ma, Y.; Chen, X.M.; Wu, Y.J.; Lin, Y.Q. Dielectric relaxation and enhanced multiferroic properties in YMn0.8Fe0.2O3 ceramics prepared by in situ spark plasma sintering. Ceram. Int. 2010, 36, 727–731. [Google Scholar] [CrossRef]
  25. Polat, O.; Coskun, F.M.; Coskun, M.; Durmus, Z.; Caglar, Y.; Caglar, M.; Turut, A. Tailoring the band gap of ferroelectric YMnO3 through tuning the Os doping level. J. Mater. Sci. Mater Electron. 2019, 30, 3443–3451. [Google Scholar] [CrossRef]
  26. Zaghrioui, M.; Greneche, J.M.; Autret-Lambert, C.; Gervais, M. Effect of Fe substitution on multiferroic hexagonal YMnO3. J. Magn. Magn. Mater. 2011, 323, 509–514. [Google Scholar] [CrossRef]
  27. Paul, P.; Rajarajan, A.K.; Prajapat, C.L.; Debnath, A.K.; Rao, T.V. Structural and Magnetic Properties of YMn1−xGaxO3; AIP Publishing: College Park, MD, USA, 2020; Volume 2265. [Google Scholar]
  28. Apostolov, A.T.; Apostolova, I.N.; Wesselinowa, J.M. Size, external fields and ion doping effects on the multiferroic properties of hexagonal YMnO3 nanoparticles. Mater. Today Commun. 2022, 30, 103123. [Google Scholar]
  29. Gutierrez, D.; Peña, O.; Durän, P.; Moure, C. Crystalline structure and electrical properties of YCuxMn1−xO3 solid solutions. J. Eur. Ceram. Soc. 2002, 22, 2939–2944. [Google Scholar] [CrossRef]
  30. Jeuvrey, L.; Peña, O.; Moure, A.; Moure, C. Synthesis and magnetic properties of hexagonal Y (Mn, Cu) O3 multiferroic materials. J. Magn. Magn. Mater. 2012, 324, 717–722. [Google Scholar] [CrossRef]
  31. Xiao, L.X.; Xia, Z.C.; Wang, X.; Ni, Y.; Yu, W.; Shi, L.R.; Jin, Z.; Xiao, G.L. Doping effects on the relaxation of frustration and magnetic properties of YMn0.9Cu0.1O3. J. Magn. Magn. Mater. 2017, 444, 206–210. [Google Scholar] [CrossRef]
  32. de Lima, A.F. Exchange interactions in hexagonal YMnO3 and LuMnO3 multiferroic compounds. J. Solid State Chem. 2021, 299, 122175. [Google Scholar] [CrossRef]
  33. Nazarenko, A.V.; Pavlenko, A.V.; Abdulvakhidov, K.G. Structure, the grain structure, and the dielectric properties of ceramics of solid solutions YCuxMn1−xO3. Phys. Solid State 2021, 63, 492–497. [Google Scholar] [CrossRef]
  34. Mandal, P.R.; Nath, T.K. Oxygen-vacancy and charge hopping related dielectric relaxation and conduction process in orthorhombic Gd doped YFe0.6Mn0.4O3 multiferroics. J. Alloy Comp. 2015, 628, 379–389. [Google Scholar] [CrossRef]
  35. Kumar, V.; Gaur, A.; Gaur, U.K. Low temperature magnetic and anomalous high temperature dielectric response of Dy-Ni co-doped hexagonal YMnO3 ceramics. J. Magn. Magn. Mater. 2015, 384, 241–246. [Google Scholar] [CrossRef]
  36. Kumar, V.; Gaur, A. Low temperature magnetization and anomalous high temperature dielectric behaviour of (1−x) YMnO3/xZnFe2O4 composites. J. Magn. Magn. Mater. 2018, 451, 351–359. [Google Scholar] [CrossRef]
  37. Wan, F.; Li, L.; Bai, X.; Wang, Y.; Gao, L.; Li, J.; Cao, C. Structure and dielectric relaxation behaviors of Co-doped YMnO3 multiferroic ceramics. J. Mater. Sci. Mater Electron. 2022, 33, 17361–17371. [Google Scholar] [CrossRef]
  38. Moure, C.; Villegas, M.; Fernandez, J.F.; Tartaj, J.; Duran, P. Phase transition and electrical conductivity in the system YMnO3-CaMnO3. J. Mater. Sci. 1999, 34, 2565–2568. [Google Scholar] [CrossRef]
Figure 1. (a) Room-temperature XRD patterns for YMn1-xCuxO3 (x = 0.00, 0.05 and 0.10). (b) Schematic structure for hexagonal YMnO3. (c) Rietveld refinement plot for YMn1-xCuxO3 (x = 0.00, 0.05, and 0.10).
Figure 1. (a) Room-temperature XRD patterns for YMn1-xCuxO3 (x = 0.00, 0.05 and 0.10). (b) Schematic structure for hexagonal YMnO3. (c) Rietveld refinement plot for YMn1-xCuxO3 (x = 0.00, 0.05, and 0.10).
Materials 17 02929 g001
Figure 2. Williamson–Hall (W-H) plot of YMn1-xCuxO3 ((a) x = 0.00, (b) 0.05, and (c) 0.10) ceramics.
Figure 2. Williamson–Hall (W-H) plot of YMn1-xCuxO3 ((a) x = 0.00, (b) 0.05, and (c) 0.10) ceramics.
Materials 17 02929 g002
Figure 3. SEM micrographs of YMn1-xCuxO3 ((a) x = 0.00, (b) 0.05, and (c) 0.10) with the grain size histograms.
Figure 3. SEM micrographs of YMn1-xCuxO3 ((a) x = 0.00, (b) 0.05, and (c) 0.10) with the grain size histograms.
Materials 17 02929 g003
Figure 4. XPS spectra of Mn-2p and O-1s for YMn1-xCuxO3 ((a,d) x = 0.00, (b,e) 0.05, and (c,f) 0.10).
Figure 4. XPS spectra of Mn-2p and O-1s for YMn1-xCuxO3 ((a,d) x = 0.00, (b,e) 0.05, and (c,f) 0.10).
Materials 17 02929 g004
Figure 5. Temperature variation of magnetization with temperature for YMn1−xCuxO3.
Figure 5. Temperature variation of magnetization with temperature for YMn1−xCuxO3.
Materials 17 02929 g005
Figure 6. The temperature dependency of the FC inverse magnetic susceptibility for YMn1−xCuxO3. The straight lines represent the fitting curves through the Curie–Weiss law.
Figure 6. The temperature dependency of the FC inverse magnetic susceptibility for YMn1−xCuxO3. The straight lines represent the fitting curves through the Curie–Weiss law.
Materials 17 02929 g006
Figure 7. (a) The ab-plane of YMnO3 and the construction of Mn ions. Blue and red balls are integral to trimmers in c = 0 and c = 1/2, respectively. (b) The 3d orbital of Mn ions.
Figure 7. (a) The ab-plane of YMnO3 and the construction of Mn ions. Blue and red balls are integral to trimmers in c = 0 and c = 1/2, respectively. (b) The 3d orbital of Mn ions.
Materials 17 02929 g007
Figure 8. Temperature dependence of (a,b) dielectric constant and (c,d) dielectric loss under the frequencies between 1 kHz and 1 MHz for YMn1-xCuxO3 (x = 0.05 and x = 0.10).
Figure 8. Temperature dependence of (a,b) dielectric constant and (c,d) dielectric loss under the frequencies between 1 kHz and 1 MHz for YMn1-xCuxO3 (x = 0.05 and x = 0.10).
Materials 17 02929 g008
Figure 9. Frequency dependence of dielectric constant inflection point temperature for dielectric relaxation of YMn1-xCuxO3 ((a) x = 0.05 and (b) x = 0.10) ceramics. The solid line represents the fitting curves through Arrhenius's law.
Figure 9. Frequency dependence of dielectric constant inflection point temperature for dielectric relaxation of YMn1-xCuxO3 ((a) x = 0.05 and (b) x = 0.10) ceramics. The solid line represents the fitting curves through Arrhenius's law.
Materials 17 02929 g009
Table 1. Structural parameters, selected bond lengths, and bond angles of YMn1-xCuxO3 (x = 0.00, 0.05, and 0.10) ceramics obtained from Rietveld XRD refinement.
Table 1. Structural parameters, selected bond lengths, and bond angles of YMn1-xCuxO3 (x = 0.00, 0.05, and 0.10) ceramics obtained from Rietveld XRD refinement.
Samplex = 0.00x = 0.05x = 0.10
a (Å)6.150436.152926.15859
c (Å)11.4166511.3921311.38779
V (Å)374.007370.235374.052
Rwp (%)21.37.8112.6
Rp (%)14.35.647.57
Rexp (%)5.225.325.19
χ22.022.155.88
Y1-O1 (Å)2.262152.157252.18806
Y1-O2 (Å)2.319632.194712.17532
Y1-O3 (Å)2.364362.320772.25759
Y2-O1 (Å)2.312162.401082.43321
Y2-O2 (Å)2.324762.347212.30083
Y2-O4 (Å)2.414652.562552.65046
Mn-O1 (Å)1.826751.804491.79189
Mn-O2 (Å)1.865831.991532.00210
Mn-O3 (Å)2.071162.097122.09743
Mn-O4 (Å)2.058082.055432.05234
Tolerance factor0.8910.8920.893
Crystallite size from W-H plot (nm)52.18350.15449.686
Micro strain6.03 × 10−47.28 × 10−48.73 × 10−4
Table 2. Magnetic parameters through Curie–Weiss law fitting of χ-T for the total compounds.
Table 2. Magnetic parameters through Curie–Weiss law fitting of χ-T for the total compounds.
SampleTN (K)CqCW (K)f = |qCW|/TNµeff (exp) (μB)µeff (calc) (μB)
x = 0.00713.483−456.6846.4325.2494.900
x = 0.05653.321−303.2224.6655.1254.741
x = 0.10582.984−181.5093.1294.8594.580
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wan, F.; Hua, X.; Guo, Q. Modulation of the Structural, Magnetic, and Dielectric Properties of YMnO3 by Cu Doping. Materials 2024, 17, 2929. https://doi.org/10.3390/ma17122929

AMA Style

Wan F, Hua X, Guo Q. Modulation of the Structural, Magnetic, and Dielectric Properties of YMnO3 by Cu Doping. Materials. 2024; 17(12):2929. https://doi.org/10.3390/ma17122929

Chicago/Turabian Style

Wan, Feng, Xuexia Hua, and Qiufen Guo. 2024. "Modulation of the Structural, Magnetic, and Dielectric Properties of YMnO3 by Cu Doping" Materials 17, no. 12: 2929. https://doi.org/10.3390/ma17122929

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop