Next Article in Journal
Imprinting the Polytype Structure of Silicon Carbide by Rapid Thermal Processing
Next Article in Special Issue
Effect of Various Nanoparticles (GaF3, ZnF2, Zn(BF4)2 and Ga2O3) Additions on the Activity of CsF-RbF-AlF3 Flux and Mechanical Behavior of Al/Steel Brazed Joints
Previous Article in Journal
Investigation of the Blistering and Exfoliation Mechanism of GaAs Wafers and SiO2/Si3N4/GaAs Wafers by He+ and H+ Implantation
Previous Article in Special Issue
Simulating the Hysteretic Characteristics of Hard Magnetic Materials Based on Nd2Fe14B and Ce2Fe14B Intermetallics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Bi Substitution on Structural and AC Magnetic Susceptibility Properties of Nd1−xBixMnO3

by
Nurul Atiqah Azhar
1,
Intan Solehah Ismail
1,
Nur Baizura Mohamed
2,
Azhan Hashim
2 and
Zakiah Mohamed
1,*
1
Faculty of Applied Sciences, Universiti Teknologi MARA, UiTM, Shah Alam 40450, Malaysia
2
Faculty of Applied Sciences, Universiti Teknologi MARA, UiTM, Pahang 26400, Malaysia
*
Author to whom correspondence should be addressed.
Crystals 2020, 10(6), 521; https://doi.org/10.3390/cryst10060521
Submission received: 1 June 2020 / Revised: 14 June 2020 / Accepted: 16 June 2020 / Published: 17 June 2020
(This article belongs to the Special Issue Intermetallic Compound)

Abstract

:
This study synthesizes the neodymium-based manganites with Bi doping, Nd1−xBixMnO3 (x = 0, 0.25 and 0.50) using the solid-state reaction route. The crystal structural, morphological and magnetic properties were determined using X-ray diffraction (XRD), fourier transform infrared spectroscopy (FTIR), field emission scanning electron microscopy (FESEM) and AC magnetic susceptibility. The Rietveld refinement confirmed that the compounds were in the single-phase orthorhombic structure of the NdMnO3 with Pbnm space group and lattice parameter b increased with doping from 5.5571 (x = 0) to 5.6787 (x = 0.5). FTIR spectra showed that absorption bands were located within the range of 550–600 cm−1, which corresponded to the Mn–O stretching vibration. FESEM exhibited homogenous compound. The AC magnetic susceptibility measurement studies showed a strong antiferromagnetic (AFM) to paramagnetic (PM) transition existed at 76 K, 77 K and 67 K for samples (x = 0, 0.25 and 0.50), respectively.

Graphical Abstract

1. Introduction

Existing studies on the family of manganites perovskites, RMnO3 as a parent of mixed valence perovskites, where R is made up of rare earth elements, such as La, Pr and Nd, were found a few decades ago due to their exotic behavior in electrical and magnetic properties [1,2]. Manganites perovskites were the favorite amongst researchers compared with other families of oxides because of their various possible substitutions that may change their crystal structure and properties [3]. Since the rediscovery of colossal magnetoresistance (CMR), many researchers have been focusing on the doped compound manganites due to their fascinating basis in physics and significance in technological importance, such as magnetic sensors [4]. Few theories have surfaced to explain the CMR mechanisms, such as double exchange (DE), superexchange (SE), Jahn-Teller (JT) effect, polaronic effects and so on.
Primarily, undoped NdMnO3 is an antiferromagnetic (AFM) insulator with Néel temperature 67.2 K due to the interplay between the ferromagnetic (FM) DE and AFM SE interactions present in the compound [5]. Chatterji et al. confirmed that NdMnO3 has an A-type AFM structure of the Mn magnetic sublattice below Néel temperature, TN ≈ 81.7 K [6]. When doped with divalent cations on the Nd sites, for example in Nd1−xCaxMnO3, FM semiconducting behavior was observed with x ≤ 0.25. However, when x > 0.80, the compound exhibited AFM semiconducting at low temperatures [7]. For Nd0.5Ca0.5MnO3, the compound undergoes a charge ordering (CO) transition at TCO = 250 K and the AFM ordering transition at TN = 160 K, which is consistent with the study done by Wu et al. [7,8]. In this case, both authors suggested that the presence of CO was due to the JT distortion, resulting in the distortion of MnO6 octahedral and localization of charge carrier, by decreasing its mobility, which seemed to weaken the DE interaction [7,8,9]. Meanwhile, when NdMnO3 was doped with monovalent cations such as Na+ and K+, it showed an FM insulator behavior and CO transition was observed for Nd1−xNaxMnO3 at −250 K when x > 0.1 [5,10]. However, a lack of knowledge still exists on Nd series doped with trivalent cations such as Nd1−xBixMnO3.
Previous studies reported that although Bi3+ (1.24 Å) and La3+ (1.22 Å) have almost similar ionic radii, both systems exhibit different structures and properties [11]. LaMnO3 exhibited an AFM insulator [12], while BiMnO3 exhibited an FM insulator with TC of 105 K–110 K [11,13,14]. Recent studies on bismuth-doped manganites, such as La0.7−xBixAg0.3MnO3 [12], La1−xBixMnO3 [13] and La0.67−xBixSr0.33MnO3 [15], showed a decreased metal–insulator transition temperature, TMI, with an increase of Bi content. The author suggested that the behavior of the latter was related to possible hybridization of Bi 6s2 lone pairs and eg electron of Mn3+, which could have affected the localization of carriers and weakened the DE interaction. Moreover, Bi substitution may also reduce the Mn–O–Mn angle, which can block the movement on eg electron from Mn3+ to Mn4+, which decreases the mobility of conduction electron and weaken the DE process [2,12].
Several works were reported on Bi substitution, especially in La-based system, but a lack of study on Bi substitution in Nd-based system still exists, especially for optical and magnetic properties. In this study, we focus on reporting the structural, optical and magnetic properties of Nd1−xBixMnO3 to investigate the unique role of Bi in the system.

2. Materials and Methods

Polycrystalline samples for Nd1−xBixMnO3 (x = 0, 0.25 and 0.50) were synthesized by using conventional solid state reaction method. High purity (≥99.9%) neodynium oxide (Nd2O3), bismuth oxide (Bi2O3) and manganese oxide (Mn2O3) powders were weighed in stoichiometric ratio using electronic balance and mixed together. Agate mortar and pestle were used to grind the powders for about 2 h before calcination in air at 700 °C for 24 h, repeating twice in order to eliminate volatile foreign particles. Finally, the obtained powders were then pressed into pellet under a pressure of 5 tons and were sintered at 1100 °C for 24 h.
The structure and phase purity of the samples were observed by using the X–ray diffraction technique (X’pert PRO MPD) at room temperature with Cu–Kα (λ = 1.5418 Å) radiation by using a PANanalytical diffractometer. The XRD data were analyzed using the Rietveld refinement method using several software programs, such as General Structure Analysis System (GSAS) and Graphical User Interfaces (EXP–GUI) for the refinement and CMPR to convert data from ASC file to DAT file. Fourier transform infrared spectroscopy (FTIR) results were recorded using FTIR–Raman Drift Nicolet 6700 in the range of 450–2000 cm−1 to directly probe with the functional group present in perovskite samples. Samples were thoroughly mixed in KBr before the characterization. LEO model 982 Gemini Field Emission Scanning Electron Microscopy (FESEM) with energy-dispersive spectroscopy (EDX) were used to determine the surface morphology and homogeneity of the perovskite samples. AC susceptibility (χ’) measurements were carried out using a Stanford research model SR-7265 lock-in amplifier and were performed in the temperature range from 30 K to 300 K.

3. Results

3.1. Structural Analysis

This study obtained all the samples of NdMnO3 doped with Bi as black and well-crystallized powders. Figure 1 shows the XRD patterns for manganites Nd1−xBixMnO3 (x = 0, 0.25 and 0.50) that were conducted at room temperature to validate the structure and sample purities and were then analyzed by using the Rietveld refinement technique (Figure 2). All samples were indexed in an orthorhombic structure (a ≠ b ≠ c and α = β = γ = 90°) with Pbnm space group. The refined patterns were quite acceptable as the obtained value of agreement factor, χ2, showed a value of ∼ 1, which indicated a good fit. Rietveld refinement showed the samples x = 0 and x = 0.25 crystallized in a single phase and were chemically pure. However, a small amount of Bi2O3 phase was detected in Nd0.5Bi0.5MnO3 sample. The appearance of a secondary phase and redistribution of peak intensities in the Nd0.5Bi0.5MnO3 sample was an indication of incomplete crystallization of the sample at the sintering temperature. The assumed structure for NdMnO3 was consistent with most previous studies [6,10,16,17] that reported the same structure with Pbna or Pbnm space group (different setting), ICSD No. 153214, 247143. The sharpening of the peaks was due to the better crystallinity of the nanoparticles and only minimal impurities were detected in the XRD pattern of the x = 0.50 sample. In addition, the main diffraction peaks corresponded to (121) hkl plane and matched with the previous study [18]. The crystallite size (D) of the samples estimated by the Debye–Scherer equations, D = K λ / ( β ( θ ) cos θ ) , where D is the crystallite size (nm), K is constant with 0.94, λ is wavelength of XRD that is 0.1541 nm for CuKα radiation Å and β is full-width at half-maximum (FWHM) and θ is the angle of peak of XRD. The average crystallite size was found to be in the range 74–128 nm.
Table 1 presents the corresponding data of lattice parameters obtained from Rietveld refinement technique. For the undoped sample, the lattice parameters obtained, a = 5.424 Å, b = 5.5570 Å and c = 7.6648 Å, were found to be quite consistent with the value reported by the previous study [10]. The structural parameter b of the samples increased as the Bi doping increased up to x = 0.50. However, for both lattice parameters, a and c showed no systematic trend due to Bi substitution. The unit cell volume of the samples also increased with the increase of the Bi3+ substitution. The increases in unit cell volume were suggested due to the Bi substitution at Nd site, where the Bi3+ had a slightly larger ionic radius compared with Nd3+ and could replace it, which led to the expansion of cell volume [12,13]. Both authors stated that the radius of Bi3+ may depend on the character of 6s2 lone pairs; it can be seen that when Bi 6s2 lone pair is dominant, the ionic radius of Bi3+ is 1.24 Å, while when not, the ionic radius of Bi3+ is approximately 1.16 Å [12]. Hence, the increase in structural parameter, b, and cell volume, V, with Bi substitution suggests that the Bi 6s2 lone pair is dominant in the Nd1−xBixMnO3 system. From Table 2, the Mn–O bond length decreased and the average ionic radius, rA, increased as Bi concentration level increased. However, Mn–O–Mn bond angle showed no trend with the increased Bi-doped. The tolerance factor (τ) was calculated using formula τ = (< rA > + < rO >)/ √2 ((< rB > + < rO >), where < rA >, < rB > and < rO > represent the average ionic radii of A, B and O sites, respectively [9]. The values of τ showed an increase from 0.515 (x = 0) to 0.529 (x = 0.50), as seen in Table 2. The increase in τ indicates that a slight reduction of MnO6 distortion exists, thus the lattice reduced the mismatch between A–O and B–O layer as the Bi concentration increased [18,19].
Figure 3 illustrates the 3D representation of the compound of Nd1−xBixMnO3, which shows an octahedral MnO6, constructed using Visualisation for Electronic and Structural Analysis (VESTA) software. In this structure, the compounds Nd3+, Mn3+ cations and O2− anions occupy the corner, body center and face center positions, respectively. The A site in the ABO3− type perovskite structure occupied by Nd/Bi cations was surrounded by 12 oxygen ions, while the octahedral MnO6 formed by the position of Mn ions at the B site was surrounded by six oxygen ions.

3.2. FTIR Spectra

Figure 4 shows the spectra for the compound of NdMnO3 doped with Bi in the frequency range from 450 cm−1–2000 cm−1. From this figure, we can see that the most significant absorption bands are located at the range of 600 cm−1–550 cm−1. The absorption bands at this range correspond to the Mn–O stretching vibration, linked to the internal movement of a length change of the bounds Mn–O–Mn, associated with the octahedron MnO6 attributed to a vibration characteristics of the perovskite-type ABO3 [3,19]. The absorption peak band at the range of 600 cm−1–550 cm−1 seem to shift towards lower wave number values for a structure having lower symmetry with decreasing Bi3+ ion content, according to the increase of eg electrons in the anti-bonding orbital. Hence, the bond order would decrease with an increase in the number of eg electrons. The lower frequency band can be related to a higher deformation mode of the MnO6 octahedron [20,21]. The FTIR spectra confirm the formation of the NdBiMnO3 sample.

3.3. Morphology

To understand the morphology of the samples, we performed FESEM with EDX and obtained the results, as shown in Figure 5. The images in Figure 5 reveal that the surface of the sample is non-homogeneous and the grains are agglomerated. The particles are also observed as non-spherical in shape. According to the image, the grain size of all samples are in the range of ~3.7 μm to ~5.6 μm, respectively. The grain sizes observed by FESEM are larger than those calculated by the Debye–Scherer formula. This can be explained by the fact that each particle observed by FESEM is formed by several crystallized grains. The results of EDX plot matches with a standard peak position of Nd, Bi, Mn and O from the previous study [19]. The EDX spectrum confirms the homogeneity of the samples.

3.4. AC Susceptibility Measurement

Figure 6 shows the temperature dependence of the AC susceptibility of the Nd1−xBixMnO3 system on the real part χ’ with the inset dχ’/dT versus the temperature range of 30–300 K. For x = 0 and x = 0.25, a definite and sharp peak was observed at 70 K and 72 K, respectively, associated to the ordering temperature of Mn spins [22]. The observed cusp depicts the PM to AFM transition, which is consistent with the results from previous studies [5,8,23]. For x = 0.50, two sharp peaks were observed, which were at 42 K and 57 K, respectively. The existence of two peaks indicates the magnetic inhomogeneity that may be induced by Bi substitution [12]. The TC was determined using the minimum, while TN was determined using the maximum point of differentiation, as shown in the inset of the graph in Figure 6. In the figure, all the three compounds exhibit a strong AFM with a competing weak FM, where TN and TC decreased with Bi substitution. The decreasing value of TC shows the weakening of the FM behavior due to the Bi substitution. Temperature dependence of inverse susceptibility, 1/χ’, for all samples is plotted in Figure 7. The experimental data were fitted by using Curie–Weiss (C–W) equation, χ = C / ( T θ ) , where C is Curie constant and θ is Curie–Weiss temperature. The red line is the best fit for C–W law in the paramagnetic region. The inverse susceptibility curve follows the C–W law at the higher temperature and starts to deviate from the linear fitting of C–W law at the lower temperature above TC, which shows the existence of Griffiths phase (GP) in the samples [24,25]. The deviation shows that the C–W law is inapplicable in the temperature range between TC and Griffiths temperature, TG (TC < T < TG). This anomaly indicates the possible existence of ferromagnetic cluster within the paramagnetic region in the system which would enhance the total magnetic susceptibility of the samples [24,25,26,27]. The temperature ranges of Griffiths phase normalized with TC were evaluated using the equation GP = (TG − TC)/TC and were found to increase as the substitution of Bi increased. The increasing values were due to an increase of magnetic inhomogeneity near TC. Therefore, the substitution of Bi in Nd1−xBixMnO3 system could enhance the appearance of Griffiths phase [24]. The values of TC, TN, TG and GP are summarized in Table 3.

4. Conclusions

In summary, this study investigates the effect of Bi3+ substitution on structural, optical, morphological and magnetic properties of Nd1−xBixMnO3 samples. The XRD analysis and Rietveld refinement show the pattern corresponding to the perovskite-type NdMnO3, which crystallizes in the orthorhombic system with main peak at (121) hkl plane with the Pbnm space group. The structural parameter and cell volume of the samples increase as the Bi3+ content increases. In addition, the average ionic radii of A site, < rA >, increased, hence τ also increased. The FTIR spectra show that active vibrational bands present at 600 cm−1–550 cm−1, which correspond to the metal oxide Mn–O stretching vibration of perovkite NdMnO3. FESEM images show that the sample is non-spherical and somewhat agglomerated. The samples show a strong AFM behavior at low temperature with decreasing TC due to the Bi substitution. Deviation of temperature dependence of inverse susceptibility curves shows the existence of Griffiths phase in these materials.

Author Contributions

Z.M. conceived and designed the experiment; I.S.I. performed the experiments supervised by Z.M. The manuscript was written by N.A.A. and Z.M. with contributions from N.B.M. and A.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was financially supported by the Ministry of Education Malaysia (MOE) and Universiti Teknologi MARA, grant number 600-IRMI/FRGS-RACER 5/3 (042/2019).

Acknowledgments

The authors would like to thanks the Faculty of Applied Sciences, Universiti Teknologi MARA for FESEM and FTIR characterization, and Center for Superconducting & Magnetic Novel Oxides of Universiti Teknologi MARA for ACS characterization.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Massa, N.E.; Campo, L.; Meneses, D.D.S.; Martínez-lope, M.J.; Alonso, J.A. High temperature far infrared dynamics of orthorhombic NdMnO3; emissivity and reflectivity. J. Phys. Condens. Matter. 2013, 25, 235603. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Renwen, Q.; Zhe, L.; Jun, F. Influence of Bi3+ doping on electronic transport properties of La0.5xBixCa0.5MnO3 manganites. Phys. B Phys. Condens. Matter. 2011, 406, 1312–1316. [Google Scholar]
  3. Hernández, E.; Sagredo, V.; Delgado, G.E. Synthesis and magnetic characterization of LaMnO3 nanoparticles. Rev. Mex. Fis. 2015, 61, 166–169. [Google Scholar]
  4. Ahmadvand, H.; Salamati, H.; Kameli, P. The effect of grain boundaries on the domain wall dynamics in Pr1−xAgxMnO3 manganites. J. Appl. Phys. 2010, 107, 083913. [Google Scholar] [CrossRef] [Green Version]
  5. Nandy, A.; Roychowdhury, A.; Das, D.; Pradhan, S.K. Structural and magnetic characterizations of undoped and K-doped NdMnO3 single crystals synthesized by sol-gel route: A comparative study. Powder Technol. 2014, 254, 538–547. [Google Scholar] [CrossRef]
  6. Chatterji, T.; Ouladdiaf, B.; Bhattacharya, D. Neutron diffraction investigation of the magnetic structure and magnetoelastic effects in NdMnO3. J. Phys. Condens. Matter. 2009, 21, 1–6. [Google Scholar] [CrossRef]
  7. Liu, K.; Wu, X.W.; Ahn, K.H.; Sulchek, T.; Chien, C.L.; Xiao, J.Q. Charge ordering and magnetoresistance in Nd1−xCaxMnO3 due to reduced double exchange. Phys. Rev. B 1996, 54, 3007. [Google Scholar] [CrossRef] [Green Version]
  8. Wu, S.Y.; Hwang, S.R.; Li, W.; Lee, K.C.; Lynn, J.W.; Liu, R.S. Jahn-Teller distortion, charge ordering, and magnetic transitions in Ca-Doped NdMnO3. J. Phys. 2000, 38, 354–359. [Google Scholar]
  9. Arifin, M.; Ibrahim, N.; Mohamed, Z.; Yahya, A.K.; Khan, N.A.; Khan, M.N. Revival of metal-insulator and ferromagnetic-paramagnetic transitions by Ni substitution at Mn site in charge-ordered monovalent doped Nd0.75Na0.25MnO3 manganites. J. Supercond. Nov. Magn. 2018, 31, 2851–2868. [Google Scholar] [CrossRef]
  10. Tang, T.; Tien, C.; Hou, B.Y. Electrical transport and magnetic properties of Nd1−xNaxMnO3 manganites. J. Alloys Compd. 2008, 461, 42–47. [Google Scholar] [CrossRef]
  11. Chiba, H.; Atou, T.; Syono, Y. Magnetic and Electrical Properties of Bi1−xSrxMnO3: Hole-Doping Effect on Ferromagnetic Perovskite BiMnO3. J. Solid State Chem. 1997, 132, 139–143. [Google Scholar] [CrossRef]
  12. Ghani, M.A.; Mohamed, Z.; Yahya, A.K. Effects of Bi substitution on magnetic and transport properties of La0.7−xBixAg0.3MnO3 ceramics. J. Supercond. Nov. Magn. 2012, 25, 2395–2402. [Google Scholar] [CrossRef]
  13. Zhao, Y.D.; Park, J.; Jung, R.J.; Noh, H.J.; Oh, S.J. Structure, magnetic and transport properties of La1−xBixMnO3. J. Magn. Magn. Matter. 2004, 280, 404–411. [Google Scholar] [CrossRef]
  14. Sugawara, F.; Iiida, S. Magnetic properties and crystal distortions of BiMnO3 and BiCrO3. J. Phys. Soc. Jpn. 1968, 25, 1553–1558. [Google Scholar] [CrossRef]
  15. Lalitha, G.; Reddy, P.V. Low temperature resistivity anomalies in bismuth doped manganites. J. Alloys Compd. 2010, 494, 476–482. [Google Scholar] [CrossRef]
  16. Alonso, J.A.; Martínez-Lope, M.J.; Casais, M.T.; Fernández-Díaz, M.T. Evolution of the Jahn-Teller distortion of MnO6 octahedra in RMnO3 perovskites (R = Pr, Nd, Dy, Tb, Ho, Er, Y): A neutron diffraction study. Inorg. Chem. 2000, 39, 917–923. [Google Scholar] [CrossRef]
  17. Srivastava, S.R.S.K. Magnetic properties of Nd1−xAgxMnO3 compounds. J. Phys. Condens. Matter. 2008, 20, 505212. [Google Scholar] [CrossRef]
  18. Sangale, M.D.; Gaikwad, D.N.; Sonawane, D.V.; Suryavanshi, D.M. Synthesis and characteristic properties of perovskite-type NdMnO3 nanocrystal materials via a co-precipitation method. J. Chem. Phys. Sci. 2018, 7, 291–296. [Google Scholar]
  19. Somvanshi, A.; Husain, S. Study of structural, dielectric and optical properties of NdMnO3. AIP Conf. Proc. 2018, 1953, 1–5. [Google Scholar]
  20. Turky, A.O.; Rashad, M.M.; Hassan, A.M.; Elnaggar, E.M.; Bechelany, M. Optical, electrical and magnetic properties of lanthanum strontium manganite La1−xSrxMnO3 synthesized through the citrate combustion method. Phys. Chem. Chem. Phys. 2017, 19, 6878–6886. [Google Scholar] [CrossRef]
  21. Mohamed, Z.; Shahron, I.S.; Ibrahim, N.; Maulud, M.F. Influence of ruthenium doping on the crystal structure and magnetic properties of Pr0.67Ba0.33Mn1−xRuxO3 Manganites. Crystal 2020, 10, 295. [Google Scholar] [CrossRef] [Green Version]
  22. Wu, S.Y.; Kuo, C.M.; Wang, H.Y.; Li, W.-H.; Lee, K.C. Magnetic structure and spin reorientation of the Mn ions in NdMnO3. J. Appl. Phys. 2000, 87, 5822–5824. [Google Scholar] [CrossRef]
  23. Asmira, N.; Ibrahim, N.; Mohamed, Z.; Yahya, A.K. Effect of Cr3+ substitution at Mn-site on electrical and magnetic properties of charge ordered Bi0.3Pr0.3Ca0.4MnO3 manganites. Phys. B Condens. Matter. 2018, 544, 34–46. [Google Scholar] [CrossRef]
  24. Manjunatha, S.O.; Rao, A.; Poornesh, P.; Lin, W.J.; Kuo, Y.K. Magnetic inhomogeneity and Griffiths phase in Bi substituted La0.65−xBixCa0.35MnO3 manganites. Phys. B Condens. Matter. 2016, 498, 82–91. [Google Scholar] [CrossRef]
  25. Zheng, X.; Gao, T.; Jing, W.; Wang, X.Y.; Liu, Y.S.; Chen, B.; Dong, H.L.; Chen, Z.Q.; Cao, S.X.; Cai, C.B.; et al. Evolution of Griffiths phase and spin reorientation in perovskite manganites. J. Magn. Magn. Mater. 2019, 491, 165611. [Google Scholar] [CrossRef]
  26. Pȩkała, M.; Szydłowska, J.; Pȩkała, K.; Drozd, V. Griffiths like phase in nanocrystalline manganite La0.50Ca0.50MnO3 studied by magnetic susceptibility and electron spin resonance. J. Alloys Compd. 2016, 685, 237–241. [Google Scholar] [CrossRef]
  27. Elyana, E.; Mohamed, Z.; Kamil, S.A.; Supardan, S.N.; Chen, S.K.; Yahya, A.K. Revival of ferromagnetic behavior in charge-ordered Pr0. 75Na0. 25MnO3 manganite by ruthenium doping at Mn site and its MR effect. J. Solid State Chem. 2018, 258, 191–200. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction of Nd1−xBixMnO3 (x = 0, 0.25 and 0.50) samples.
Figure 1. X-ray diffraction of Nd1−xBixMnO3 (x = 0, 0.25 and 0.50) samples.
Crystals 10 00521 g001
Figure 2. Rietveld refinement for (a) NdMnO3, (b) Nd0.75Bi0.25MnO3 and (c) Nd0.5Bi0.5MnO3. Black solid lines are observed data, the solid red line is the calculated patterns and the pink line is the difference. Tick marks indicate the allowed Bragg reflections. Bi2O3 impurity phase was detected for Nd0.5Bi0.5MnO3 and indicated by green tick marks.
Figure 2. Rietveld refinement for (a) NdMnO3, (b) Nd0.75Bi0.25MnO3 and (c) Nd0.5Bi0.5MnO3. Black solid lines are observed data, the solid red line is the calculated patterns and the pink line is the difference. Tick marks indicate the allowed Bragg reflections. Bi2O3 impurity phase was detected for Nd0.5Bi0.5MnO3 and indicated by green tick marks.
Crystals 10 00521 g002
Figure 3. Crystallographic structure of NdMnO3. Purple, orange and red-colored balls represent the Mn, Nd and O, respectively.
Figure 3. Crystallographic structure of NdMnO3. Purple, orange and red-colored balls represent the Mn, Nd and O, respectively.
Crystals 10 00521 g003
Figure 4. Fourier transform infrared spectroscopy (FTIR) spectra of Nd1−xBixMnO3 (x = 0, 0.25 and 0.50).
Figure 4. Fourier transform infrared spectroscopy (FTIR) spectra of Nd1−xBixMnO3 (x = 0, 0.25 and 0.50).
Crystals 10 00521 g004
Figure 5. Field emission scanning electron microscopy (FESEM) micrograph and EDX spectra of Nd1−xBixMnO3 (a) x = 0 (b) x = 0.25 and (c) x = 0.5.
Figure 5. Field emission scanning electron microscopy (FESEM) micrograph and EDX spectra of Nd1−xBixMnO3 (a) x = 0 (b) x = 0.25 and (c) x = 0.5.
Crystals 10 00521 g005
Figure 6. AC magnetic susceptibility against temperature curves for Nd1−xBixMnO3.
Figure 6. AC magnetic susceptibility against temperature curves for Nd1−xBixMnO3.
Crystals 10 00521 g006
Figure 7. Temperature dependence of inverse susceptibility spectra of Nd1−xBixMnO3 (a) x = 0, (b) x = 0.25 and (c) x = 0.5.
Figure 7. Temperature dependence of inverse susceptibility spectra of Nd1−xBixMnO3 (a) x = 0, (b) x = 0.25 and (c) x = 0.5.
Crystals 10 00521 g007
Table 1. Lattice parameters a, b and c, and cell volume, V, of Nd1−xBixMnO3 samples.
Table 1. Lattice parameters a, b and c, and cell volume, V, of Nd1−xBixMnO3 samples.
Doping Content (x)0.00.250.50
Space groupPbnmPbnmPbnm
SymmetryOrthorhombicOrthorhombicOrthorhombic
a (Å)5.42405.42925.4321
b (Å)5.55705.65345.7712
c (Å)7.66487.63237.8536
α = β = γ909090
Volume (Å3)231.03234.27236.25
Table 2. Bond length and bond angle between Mn–O–Mn and goodness of fit for Nd1−xBixMnO3 samples.
Table 2. Bond length and bond angle between Mn–O–Mn and goodness of fit for Nd1−xBixMnO3 samples.
Doping Content (x)00.250.50
Mn–O1 (Å)1.9903 (5)2.0407 (7)1.9874 (4)
Mn–O2 (Å)2.0256 (5)1.9685 (6)1.9289 (3)
< Mn–O > (Å)2.0080 (5)2.0046 (6)1.9581 (2)
Mn–O1–Mn (°)148.62(1)138.45 (3)148.58 (9)
Mn–O2–Mn (°)154.61(1)149.36 (12)154.08 (4)
< Mn–O–Mn > (°)151.62 (1)143.91 (2)151.33 (6)
χ21.0382.5813.812
RP (%)8.2716.7614.37
RWP (%)10.9821.1019.87
< rA > (Å)1.1091.1421.175
τ0.5150.5220.529
Note that < rA > are calculated from ionic radii r(Nd) = 1.109 Å and r(Bi) = 1.24 Å.
Table 3. Parameters obtained from AC magnetic susceptibility studies of Nd1−xBixMnO3 samples.
Table 3. Parameters obtained from AC magnetic susceptibility studies of Nd1−xBixMnO3 samples.
x00.250.50
TN (K)646455
TC (K)767567
TG (K)181198182
GP (K)138164171

Share and Cite

MDPI and ACS Style

Azhar, N.A.; Ismail, I.S.; Mohamed, N.B.; Hashim, A.; Mohamed, Z. Effect of Bi Substitution on Structural and AC Magnetic Susceptibility Properties of Nd1−xBixMnO3. Crystals 2020, 10, 521. https://doi.org/10.3390/cryst10060521

AMA Style

Azhar NA, Ismail IS, Mohamed NB, Hashim A, Mohamed Z. Effect of Bi Substitution on Structural and AC Magnetic Susceptibility Properties of Nd1−xBixMnO3. Crystals. 2020; 10(6):521. https://doi.org/10.3390/cryst10060521

Chicago/Turabian Style

Azhar, Nurul Atiqah, Intan Solehah Ismail, Nur Baizura Mohamed, Azhan Hashim, and Zakiah Mohamed. 2020. "Effect of Bi Substitution on Structural and AC Magnetic Susceptibility Properties of Nd1−xBixMnO3" Crystals 10, no. 6: 521. https://doi.org/10.3390/cryst10060521

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