Open access peer-reviewed chapter

Dependence of pH Variation on the Structural, Morphological, and Magnetic Properties of Sol-Gel Synthesized Strontium Ferrite Nanoparticles

Written By

Muhammad Syazwan Mustaffa, Rabaah Syahidah Azis and Sakinah Sulaiman

Submitted: 12 April 2018 Reviewed: 02 August 2018 Published: 13 February 2019

DOI: 10.5772/intechopen.80667

Chapter metrics overview

1,337 Chapter Downloads

View Full Metrics

Abstract

In this research work, an attempt of regulating the pH as a sol-gel modification parameter during preparation of SrFe12O19 nanoparticles sintered at a low sintering temperature of 900°C has been presented. The relationship of varying pH (pH 1–14) on structural microstructures and magnetic behaviors of SrFe12O19 nanoparticles was characterized by X-ray diffraction (XRD), field emission scanning microscope (FESEM), thermogravimetric analysis (TGA), Fourier-transform infrared (FTIR), and vibrating-sample magnetometer (VSM). The single-phase SrFe2O19 with optimum magnetic properties can be obtained at pH 1 with a sintering temperature of 900°C. As pH values increase, the presence of impurity Fe2O3 was observed. TGA data-varying pH shows that the total weight loss of most samples was at 30.44% which corresponds to the decomposition process. The IR spectra showed three main absorption bands in the range of 400–600 cm−1 corresponding to strontium hexaferrite. SEM micrographs exhibit a circular crystal type of strontium ferrite with an average crystal size in the range of 53–133 nm. A higher saturation magnetization Ms, remanent magnetization Mr, and hysteresis Hc were recorded to have a large loop of 55.094 emu/g, 33.995 emu/g, and 5357.6 Oe, respectively, at pH 11, which make the synthesized materials useful for high-density recording media and permanent magnets.

Keywords

  • strontium hexaferrite (SrFe12O19)
  • sol-gel
  • pH
  • structural
  • magnetic properties

1. Introduction

Ferrite is a magnetic material in the form of ceramic like. Ferrite is usually brittle, hard, iron containing, and generally gray or black in color. It consisted of iron oxides and reacts with preferable high electrical resistivity of metal oxides. Ferrites have impressive properties such as high magnetic permeability and high electrical resistance [1]. Ferrite magnets have a low hysteresis loss and high intrinsic coercivity [2] which give greater effect in resistance demagnetization from external magnetic field. In addition, a low-cost ferrite magnet has good heat resistance and good corrosion resistance which are useful to many applications like permanent magnet [3, 4], solid-state devices, magnetic recording media [5, 6], microwave device [5], etc. A generic formula of magnetoplumbite structure of ferrite is MFe12O19, where M is divalent cations like Ba2+ [3, 4], Sr2+ [1, 2, 5, 7], and Pb2+ [8]. Pullar [9] has mentioned that the best known hexaferrite is those containing divalent cations, because it has preferable high electrical resistivity compared to other types of ferrite. SrFe12O19 has been chosen in order to produce a good quality of magnetic recording media due to high electrical resistivity of 108 Ω cm [9]. The high coercivity leads to high energy product BHmax behavior. Liu [10] has mentioned that a good quality of magnetic recording media should have possible high signal and low noise. In order to meet those criteria, the magnetic materials should have high magnetization; high coercivity but correlated with recording field; single-domain particles or grains; a smaller size of particles or grain size, thermally stable, and therefore a reduced thickness of the active magnetic film of the medium; and a good alignment of the particle or grain easy axis [10]. In recent years, higher levels of recording density have been achieved in the field of magnetic recording. Magnetic tapes employing hexagonal barium ferrite magnetic powder achieve a surface recording density of 29.5 bpsi (bits per square inch). However, when the size of hexagonal ferrite magnetic particles is reduced, the energy for maintaining the direction of magnetization of the magnetic particles (the magnetic energy) tends to become inadequate to counter thermal energy, and thermal fluctuation ends up compromising the retention of recording.

Various techniques are presented for the synthesis of strontium hexaferrite powders such as solid-state synthesis method [11, 12], chemical coprecipitation [13, 14, 15], ceramic method [16], and sol-gel [17, 18, 19] and hydrothermal methods [20]. The effect of pH variation in this research work via sol-gel method for producing SrFe12O19 is key factor for controlling hexaferrite nanostructure and magnetic properties. Other than that, this proposed method has not yet been reported elsewhere in producing SrFe12O19 nanoparticles. Recently, the sol-gel route has received considerable attention in the last few years because it has lower calcination temperature, the fact that it also enables smaller crystallites to grow [2]. Sol-gel method produces a better outcome than microemulsion and coprecipitation methods. The sol-gel hydrothermal method combines the advantages of the sol-gel method and the high pressure in the hydrothermal condition [7]. In the hydrothermal process, the particle size and particle morphology can be controlled. SrFe12O19 nanoparticles have high purity, ultrafine size, and high coercivity. Some efforts have been carried out to modify the sol-gel process parameters such as pH, basic agent, carboxylic acid, and starting metal salts for further decreasing the calcination temperature and achieving the finer crystallite size [1]. Optimizing the molar ratio of Fe to Sr is very important to produce a single-phase sample, ultrafine particle, and lower calcination temperatures [21]. This ratio varies with the change in starting materials and with the change in method of production [21]. The obtained products that have single-phase particles have a hexagonal shape, the right proportion, and high coercively. The prolonging annealing time has a significant effect on the high saturation magnetization (Ms), and the high annealing rate formed a highly percentage of pure strontium hexaferrite. Masoudpanah and Ebrahimi [2] state that the preferred molar ratio of Fe/Sr is 10, which is the lowest calcination temperature (800°C) on the formation of single phase of SrM thin films. In addition, XRD showed that the crystallite sizes at a range of 20–50 nm. The magnetic properties of this preferred molar ratio exhibit a good saturation magnetization (267 emu/cm3), high coercivity (4290 Oe), and a relatively high remanent magnetization (134 emu/cm3). Minh et al. [7] state that the preferred molar ratio is at 11. The obtained SrFe12O19 has high purity, ultrafine size, and high coercivity at Hc = 6315 Oe. This chapter discussed an attempt to employ water as the gel precursor to synthesize nano-sized M-type strontium ferrite (SrFe12O19) bulk sample at low sintering temperature 900°C by using a common laboratory chemical. A solution of metal nitrates and citric acid and ammonia has been used to prepare strontium hexaferrite at varying pH.

Advertisement

2. Brief overview of preparation methods

2.1. Raw materials

Strontium nitrate anhydrous granular Sr (NO3)2 (98%, Alfa Aesar), iron (III) nitrate Fe(NO3)3 (99%, HmbG), citric acid (CA) C6H8O7 (99%, Alfa Aesar), ammonia NH4OH (25%, SYSTERM), and deionized water were used as starting material in order to synthesize SrFe12O19 nanoparticles as listed in Table 1.

Chemical name Compound formula Molecular weight (g/mol) Weight ratio (g)
Strontium nitrate anhydrous (salt) Sr(NO3)2 211.63 0.4183
Iron (III) nitrate Fe(NO3)3 403.84 9.5817
Citric acid (powder) C6H8O7 191.12 11.8368
Ammonia NH4OH 35.04 Varied depend on pH
Measured by pH meter

Table 1.

Compound used for sol-gel synthesis.

The general chemical equation for desired SrFe12O19 samples is weighed according to the formula:

Sr NO 3 2 + 12 Fe NO 3 3 NH 4 OH C 6 H 8 O 7 SrFe 12 O 19 + volatile E1

The nitrates were calculated as one mole of Sr(NO3)3, and 12 moles of Fe(NO3)2 were needed in order to synthesize one mole of SrFe12O19 nanoparticles. In the process of reaction, CA was used as a chelating agent and fuel of combustion. The CA was then calculated according to the molar ratio of citrate to nitrate of 0.75 which first obtained each number of mole nitrate as below:

Sr NO 3 2 Sr 2 + + 2 NO 3 E2
Fe NO 3 3 Fe 3 + + 3 NO 3 E3

Then, mass of citrate was calculated as:

Mass CA = 0.75 × total nitrate × molar mass CA E4

In this study, NH4OH was used to vary the pH value of SrFe12O19 in order to study the effect of pH value in its morphology and magnetic properties.

2.2. Sample preparation and characterizations

An appropriate amount of Sr (NO3)2, Fe(NO3)3, and C6H8O7 was dissolved in 100 ml of deionized water for 30 min at 50°C with constant stirrer rotation of 250 rpm. The mixtures were continuously stirred, and NH4OH was added in order to vary the pH from pH 1–14 which is measured by HI 2211 pH/ORP meter (HANNA instruments). The solutions then were stirred on the hot plate for 24 h at 60°C. The solution was left in oven at temperature of 80°C for 2 days to turn the solution into a sticky gel. The sticky gel was stirred again stirred on hot plate, and the temperature was increased up to 150°C to dehydrate and form a powder. The powder formed were crushed by using mortar before sintering it at 900°C for 6 h with the heating rate of 3.5°C/min. The crystalline structural characterization of XRD was performed using a Philips X’Pert X-ray diffractometer model 7602 EA Almelo with Cu Kα radiation at 1.5418 Å. The range of diffraction angle used is from 20 to 80° at room temperature. The accelerating current and working voltage were 35 mA and 4.0 kV, respectively. The data are then analyzed by using X’Pert Highscore Plus software. The lattice constant, a, is obtained by Eq. (5):

a = d h 2 + k 2 + l 2 E5

Where d is the interatomic spacing and (h k l) are miller indices. The volume cell Vcell was calculated using Eq. (6):

V cell = 3 2 a 2 c E6

Where a and c are lattice constants. The theoretical density ρtheory of sample was calculated using Eq. (7):

E7

Where M is molecular weight of SrFe12O19, which is equal to 1061.765 g. The weight of two molecules in one unit cell is 2 × 1061.765 = 2123.53 g; NA is the Avogadro’s number (6.022 × 1023 mol−1). The porosity P of the samples can be calculated using Eq. (8):

E8

Where ρexp is the experimental density and ρtheory is the xrd density.

Meanwhile, the crystallite size can be measured by using the Scherrer equation (Eq. 9):

D = β cos θ E9

Where D is crystallite size, k is the Scherrer constant value of 0.94, λ is Cu Kα radiation wavelength of 1.542 Å, β is half-peak width of diffraction band, and θ is the Bragg angle corresponding to the planes.

The thermal stability of these samples was obtained by using TGA/SDTA 851 of Mettler Toledo thermogravimetric analyzer. The sample weighted about 10 mg was used at operating temperature range from 0 to 1000°C with heating rate 5°C/min. Fourier-transform infrared by Perkin Elmer model 1650 was used to determine the infrared spectrum of absorption and emission bands of sample. It was performed between infrared spectra of 280–4000 cm−1 with resolution of 4 cm−1. The micrograph of microstructure was observed using a FEI Nova NanoSEM 230 machine to study the morphology and microstructure of solid material. The sample was prepared in bulk pallet at a diameter of 1 cm and coated with gold in order to avoid charge buildup as the electron beams are scanned over the samples’ surface. The distribution of grain size image was fixed at magnification of 100,000X with 5.0 kV. The distribution of average grain size of microstructure was calculated by using these images. The distributions of grain sizes were obtained by taking at least 200 different grain images for the sample and estimating the mean diameters of individual grains by using the J-image software. The magnetic properties of samples were measured by VSM Model 7404 LakeShore. The measurement was carried out in the room temperature with sample weight about 0.2 g. The external field applied was 12 kOe parallel to the sample. From this analysis, saturation magnetization, Ms; remanent magnetization, Mr; and coercivity, Hc, were recorded, and the hysteresis loop was plotted.

Advertisement

3. Research findings and outcomes

3.1. Structural analysis

Figure 1 shows the XRD spectra of the samples sintered at 900°C with different pH values (pH 1–14). The XRD spectrum shows the formation of a single phase of SrFe12O19 nanoparticles. The structure of XRD peaks was referred to standard strontium hexaferrite (SrFe12O19) with JCPDS reference code of 98002-9041 [22], with hexagonal crystal system belonging to space group of P63/mmc that proved the hexagonal crystal structure system formation. The SrFe12O19 phase formed with miller indices shown as [110], [008], [017], [114], [021], [018], [023], [116], [025], [026], [127], [034], [0211], [0115], [0214], and [137], respectively. The highest intensity can be observed at (34.218°) with miller indices of [114] of reference index code of 98-002-9041 [22]. However, as pH increases, the amount of ammonia required increases. This factor leads to the formation of hematite (Fe2O3) in pH 6, pH 8, pH 13, and pH 14 due to excess ammonia that could not completely vanish during reaction. The formation of Fe2O3 occurred at  = 33.139 and 49.673° with miller indices of [104] and [024]. The hematite Fe2O3 patterns were indexed to ICSD reference code of 98-005-3678 [23]. It was explained by Masoudpanah and Ebrahimi [2] that the increasing pH of the sol results in the absorption of positively charged Sr ions on iron gels and the formation of negatively charged iron gels. A single-phase SrFe12O19 was obtained at a low sintering temperature of 900°C for powder pH which proves the benefit of using sol-gel method in this SrFe12O19 reaction. It was agreed that obtained single phase of SrFe12O19 at lower temperature was due to the solubility of Sr (NO3)2 that decreases at elevated temperatures [24]. Hence, more Sr2+ ions are needed for the formation of the strontium hexaferrite [2]. The diffusion rates increased in the nonstoichiometric mixtures because of the induced lattice defects which could be observed from lower lattice parameter [2].

Figure 1.

The X-ray diffraction spectra of SrFe12O19 nanoparticles for pH 1–14 sintered at 900°C.

The average crystallite size (Table 2) determined from the full width at the half maximum (FWHM) of the XRD patterns was calculated using the Scherrer formula provided from X’Pert Highscore Plus software. From the plotted crystallite size relationship with pH of samples (Figure 2), it shows two groups of crystallite size distribution: acidic group (1) for pH 1–8 and alkaline group (2) for pH 9–14. Both groups show an improvement of crystallinity that gives out smaller crystallite size as the pH increases, which results in smaller grain size as the crystallite size increases.

pH Peak pos. 2θ (°) Miller indices (hkl) Peak width (°) Space group Lattice constant Vcell (nm3) ρxrd (g cm−3) ρexp (g cm−3) P (%) Calculated crystalline size, D (nm)
a (Å) c (Å)
1 34.20 [114] 0.13 P63/mmc 5.883 23.018 5.11 4.634 13.899 0.690 63.226
2 34.21 [114] 0.13 P63/mmc 5.882 23.051 5.11 4.399 11.217 0.691 63.228
3 34.22 [114] 0.16 P63/mmc 5.882 23.051 5.11 3.832 8.077 0.691 51.372
4 34.20 [114] 0.16 P63/mmc 5.884 23.058 5.10 4.693 13.237 0.691 51.369
5 34.25 [114] 0.16 P63/mmc 5.880 23.040 5.11 4.200 12.114 0.690 51.376
6 34.18 [114] 0.16 P63/mmc 5.884 23.060 5.10 4.492 11.831 0.691 51.366
7 34.18 [114] 0.18 P63/mmc 5.884 23.057 5.10 4.497 12.368 0.691 45.661
8 34.17 [114] 0.18 P63/mmc 5.885 23.058 5.10 3.419 9.633 0.691 45.660
9 34.12 [114] 0.14 P63/mmc 5.889 23.025 5.10 4.633 9.153 0.691 57.266
10 34.19 [114] 0.15 P63/mmc 5.884 23.047 5.10 4.784 6.254 0.691 54.685
11 34.18 [114] 0.15 P63/mmc 5.885 23.053 5.10 4.721 7.435 0.691 53.231
12 34.22 [114] 0.17 P63/mmc 5.880 23.030 5.12 4.699 8.125 0.689 47.762
13 34.14 [114] 0.18 P63/mmc 5.884 23.066 5.10 4.705 7.736 0.691 46.693
14 34.20 [114] 0.18 P63/mmc 5.882 23.023 5.11 4.612 9.782 0.690 46.941

Table 2.

The summary of the of SrFe12O19 nanoparticles for pH 1–14.

Figure 2.

Relationship of crystallite size versus pH.

The lattice constant a and c values (Table 2) observed were not far different from the theoretical SrFe12O19 lattice constant, where a = 5.8820 Å and c = 23.0230 Å [25], as similar as reported by Masoudpanah et al. [2, 26] and Dang et al. [27]. There is a slight increment in lattice constant c as pH increases and fluctuated data of lattice constant a. It is shown that, at pH 10, the lattice constant a of 5.884 Å was the highest peak with a lower peak of lattice constant c of 23.047 Å. The standard strontium hexaferrite (SrFe12O19) with JCPDS reference code of 98-002-9041 [22] has theoretical density of 5.11 g cm−3 [25]. Theoretically, the density of the sample, ρEXP, is affected by the lattice constants a and c. The lattice parameter a and c values observed were not far different from the theoretical SrFe12O19 lattice constant, where a = 5.8820 Å and c = 23.0230 Å (Figure 3) [25]. The a and c parameters observed are similar to Masoudpanah et al. [2] and Dang et al. [27].

Figure 3.

The TGA and DTA traces for dried powder of SrFe12O19 sintered up to 1000°C.

The lattice constant was fluctuated around the theoretical lattice constant. However, in the experiment, the density was more affected by the preparation of the sample which results in porosity of the sample. The distant the difference of density of XRD (ρXRD) and experimental density (ρEXP), the higher the number of porosity, which results in reducing the mass of the pallet sample by pores. The highest density value for ρXRD is at pH 12 (5.1148 gcm−3), and the highest density value for ρEXP is at pH 10 (4.784 gcm−3). The porosity occurs because of the presence of pores in the samples as a result after sintering of bulk samples. The pores occur due to an error from preparing sample and the loosen powder while pressing the sample using hydraulic presser. As the ρEXP approaches to the ρXRD, the pores’ percentage becomes lower. The highest porosity of 13.24% was found at pH 4 with ρXRD of 5.1001 g cm−3 and ρEXP of 4.425 g cm−3. Meanwhile, pH 10 exhibits a lower porosity of 6.254% with ρXRD of 5.1032 g cm−3 and ρEXP of 4.784 g cm−3 (Table 2).

The powder was synthesized using a control molar ratio of 1:12 with respect to strontium and nitrate. However, the sample with various pH was prepared with an addition of nitrate into the solution. The sample ratio of Sr (NO3)3 and Fe(NO3)2 is (Fe/Sr) = 12:1, and the samples were sintered at 900°C. From previous work reported, a single phase of strontium ferrite (SrFe12O19) was obtained for samples sintered at 850°C with Fe/Sr molar ratio of 11.5 via sol-gel route [28]. Masoudpanah and Ebrahimi [2] found a single phase of SrFe12O19 at sintering temperature of 900°C prepared using sol-gel technique. In general, the lowest sintering temperature of SrFe12O19 is around 800–1000°C. Hence, the raw powder (non-sintered) was tested by TGA to identify the best temperature by sintering up to 1000°C. The TGA curves as plotted in Figure 3 show a decreasing amount of weight as the powder sintered up to 1000°C in 20 min with a starting weight of 8.3609 mg. Meanwhile, the DTA diagrams reveal three peaks shown at range 86.80–100, 399.55, and 740.40°C due to decomposition process. At a constant heating rate, the endothermic peak at 86.80–100°C had −7.91% of weight loss due to the dehydration of the absorbed water as the powder slowly turns into burnt gel [2]. The first exothermic peak at 399.55°C with a weight loss of −11.53% is due to the elimination of the organic compound which tends to the decomposition of NH4NO3 that liberates NO, O2, and H2O [2]. Meanwhile, at stage 740.40°C, the exothermic peak with a weight loss of −10.49% shows the decomposition of citric acid and the breakdown of the Fe2O3 to Fe as reported [17]. The stable temperature is at 880°C which permits the completeness of reaction. Hence, sintering temperature at 900°C was used in this work.

Figure 4 shows the FTIR spectra of SrFe12O19 nanoparticles for pH variation (pH 1–14), with IR range of 400–4000 cm−1. It is noticeable that spectrum appeared in the range of 430, 583, 904, and 1446 cm−1 of IR characteristic band. The stretching band of CH2 appeared at 436 cm−1 attributed to the presence of CH saturated compound, which has been agreed by [29]. The vibration of CH bond could be due to the chemical reaction in a process of hexagonal structure form, where the CH bond of citric acid loses their CH bond. The spectrum of metal-oxygen vibration of Sr–O Fe–O was found at 583 cm−1 [26]. Masoudpanah and Ebrahimi [2] explained that an occurred reaction between citric acid and ferric ions is attributed to the stretching mode of Fe–O, which confirms the formation of chelate in sol-gel route. It is proven by many researchers who claim that the absorption bands at range 443–600 cm−1 were results of the formation of strontium ferrite as the stretching vibration of metal-oxygen bond of Sr–O Fe–O occurs [30, 31, 32, 33]. All pH reveals these two bond bands. However, there were some reducing and vanished bands in the next bond bands at 904 and 1460 cm−1. It is due to the purity of SrFe12O19 nanoparticles, as there was some interruption of Fe2O3 in the sample as shown in Figure 1 and Table 2. In this study, pH 8, pH 10, pH 13, and pH 14 come out with a percentage of hematite, Fe2O3. First, the pure SrFe12O19 (pH 1–7, pH 9, pH 11–12) had a relatively strong and broad bands at peak 904 cm−1, which revealed that there was amine functional group for N–H vibration due to decomposition of NH3. Pereira et al. [32] stated that this broad vibration of Sr–O stretching indicates the formation of strontium nanoferrites. It is agreed by Sivakumar et al. [34] that the strontium ferrite was formed and the iron oxide vanished at 900 cm−1. Meanwhile, pH 6, pH 8, pH 13, and pH 14 show a relative small vibration band at 904 cm−1 due to the presence of Fe2O3. As pH increases up to pH 14, the amount of ammonia increases gradually. Excess amount of ammonia failed to completely decompose the NH3 bonds and break down the N–H vibration. Lastly, the absorption bands at 1446 cm−1 found in pure pH sample were attributed to the vibrating bands of Fe-O-Fe due to the decomposition of metal with oxide band [29]. There was some significant data that show in pH 9, pH 11, and pH 12, as a single phase of samples is formed as pH increased.

Figure 4.

FTIR spectra of SrFe12O19 for pH variation sintered at 900°C.

3.2. Microstructural analysis

The microstructure and grain size distribution of bulk SrFe12O19 nanoparticles are shown in Figure 5. The grain size seems to have agglomerated and charged nanoparticles when increasing the pH value. The grain size was found in the range of 53.22–133.25 nm. The pH 4 produces pores of 13.24%. Meanwhile, the most packed grains are for sample at pH 10, with porosity of 6.25% (Table 2). The microstructure shows that some of the samples have a large porosity due to the presence of polyvinyl alcohol during the preparation of pellet bulk SrFe12O19 nanoferrites. The histogram of the grain distribution was shifted from small grain sizes to exhibiting larger grains from pH 1 to 8. Nevertheless, the grain size was observed to be decreasing as the pH is reaching 9–14 (Table 3).

Figure 5.

The micrograph image and grain size distribution of SrFe12O19 sintered at 900°C by varying pH value.

pH Grain size (nm)
1 108
2 114
3 115
4 96
5 111
6 120
7 116
8 133
9 61
10 79
11 75
12 62
13 57
14 53

Table 3.

Grain size of SrFe12O19 sintered at 900°C by varying pH value.

3.3. Magnetic behaviors

The development of M-H hysteresis loop at various pH is illustrated in Figure 6. The magnetic saturation, Ms; remanent magnetization, Mr; coercivity, Hc; grain size; and porosity of SrFe12O19 nanopowder are shown in Table 4. An obvious erect, larger, and well-defined hysteresis loop can be observed. It is probably due to the strong ferromagnetic behavior, indicating the formation of SrFe12O19 nanoparticles with high volume fraction of the complete crystalline SrFe12O19 phase. Thus a strong interaction of magnetic moments within domains occurred due to exchange forces. This observed phenomenon can be considered as ordered magnetism in the sample. In fact, in order to obtain an ordered magnetism and well-formed M-H hysteresis loop, there must exist a significant domain formation, a sufficiently strong anisotropy field (Ha), and optional addition contributions, which come from defects such as grain boundaries and pores [35]. The saturation magnetization (Ms), remnant magnetization (Mr), and coercivity (Hc) are found to decrease with increasing pH by addition of ammonia in the sol-gel precursor. From the previous study, the Hc is 4290 Oe, obtained at pH 7 [2, 26]. The microstructures of nanoparticles were affected by the increase of pH value. This is in agreement with findings reported by Yang et al. [36], where the formation of particles became larger [37] with the increase of pH from 5 to 11. This is due to the aggregation of small particle that occurs when there is a strong magnetic interaction between magnetic atoms (Fe or Co) containing in Co-Fe-Al grains as the composition of Co increases and the composition of Al decreases [38]. In this work, it is noticeable that pH 11 has the largest hysteresis loops as well as high magnetic properties. Moreover, the remaining pH exhibit almost the same hysteresis loop with a slight change in Ms and Mr. Meanwhile, the presence of Fe2O3 impurity in the samples of pH 6, 8, 13, and 14 shows a decrease in Hc, which affects the crystalline and grain boundary. The Hc is observed to reduce as pH increased. The presence of intragranular trapped pores in the grains was due to rapid grain growth of sample. The presence of intragranular pores would pin down the magnetic moment in grains, thus reducing the Ms and also the Hc. The decrease in Hc as pH increases can be attributed to the decrement of magnetocrystalline anisotropy with anisotropic Fe2+ ions located in a 2A site, and the enlargement of the grain size is evident in FESEM micrographs (Figure 5). The Ms and Mr are also observed to decrease as pH increases. The decrement of magnetic parameters as pH increases could be due to the existence of large amount of diamagnetic phases as the amount of ammonia NH3 increases. It seems that the main roles of the diamagnetic NH3 are to isolate Sr-ferrite nanoparticles from each other, thus reducing exchange interaction between them, and are known to have a detrimental effect on Ms and Mr.

Figure 6.

The M-H hysteresis loop SrFe12O19 of pH 1–14 sintered at 900°C.

pH Saturation magnetization, Ms (emu/g) Remnant magnetization, Mr (emu/g) Coercivity, Hc (Gs)
1 4.776 3.001 6094.7
2 7.822 4.870 6005.8
3 2.168 1.373 5966.1
4 3.006 1.929 5808.6
5 2.016 1.309 6074.8
6 7.022 4.416 5377.0
7 4.028 2.554 5461.2
8 31.342 19.363 5058.3
9 20.488 12.776 5422.2
10 25.471 15.825 5663.1
11 55.094 33.995 5357.6
12 25.114 15.674 5532.7
13 26.849 16.885 5185.9
14 14.239 9.1325 5520.7

Table 4.

Ms, Mr, and Hc of SrFe12O19 as a function of pH.

Advertisement

4. Conclusions

Single-phase SrFe2O19 ferrite nanoparticles were successfully synthesized by sol-gel citrate-nitrate method. From the discussion presented earlier, the influence of pH variation on the SrFe2O19 ferrite nanoparticles on the structural, microstructural, and magnetic properties was discussed. An increment amount of ammonia has changed the purity, average grain size, density, and its porosity, which affected the magnetic properties of the samples. Those characteristics reveal an understanding on how important effects of pH study (linear effect of pH and acidic-alkaline effect) underlining on SrFe12O19 nanoparticles, as most researchers neglect it.

Advertisement

Acknowledgments

The authors would like to thank the Ministry of Education Malaysia for providing funds; MyBrain15, Research University Grants Vot No. 9541600 and 5524942; and the Department of Physics, Faculty of Science and the Materials, Synthesis and Characterization Laboratories (MSCL) ITMA, UPM, for the measurement facilities.

Advertisement

Conflict of interest

The authors declare that they have no competing interest.

References

  1. 1. Masoudpanah SM, Ebrahimi SAS, Ong CK. Preparation of strontium hexaferrite film by pulsed laser deposition with in situ heating and post annealing. Journal of Magnetism and Magnetic Materials. 2012;324(18):2894-2898. DOI: 10.1016/j.jmmm.2012.04.034
  2. 2. Masoudpanah SM, Ebrahimi SAS. Synthesis and characterization of nanostructured strontium hexaferrite thin films by the sol-gel method. Journal of Magnetism and Magnetic Materials. 2012;324(14):2239-2244. DOI: 10.1016/j.jmmm.2012.02.109
  3. 3. Perdamean S, Muljadi M, Siregar RT, Tomi BW. Ferrite-based material as a permanent magnet for components of electrical generators. Advances in Natural Sciences: Nanoscience and Nanotechnology. 2011;2:045016. DOI: 10.1088/2043-6262/2/4/045016
  4. 4. Topal U, Bakan HI. Permanently magnetic BaFe12O19 foams: Synthesis and characterization. Materials Chemistry and Physics. 2010;123(1):121-124. DOI: 10.1016/j.matchemphys.2010.03.070
  5. 5. Ashiq MN, Iqbal MJ, Najam-Ul-Haq M, Hernandez Gomez P, Qureshi AM. Synthesis, magnetic and dielectric properties of ErNi doped Sr-hexaferrite nanomaterials for applications in high density recording media and microwave devices. Journal of Magnetism and Magnetic Materials. 2012;324(1):15-19. DOI: 10.1016/j.jmmm.2011.07.016
  6. 6. Hibst H, Schwab E. Magnetic recording materials. Materials Science and Technology. Federal Republic of Germany: WILEY-VCH Verlag GmbH & CoKGaA; 2006. DOI: 10.1002/9783527603978.mst0041
  7. 7. Minh T, Dang H, Trinh VD, Bui DH. Sol-gel hydrothermal synthesis of strontium hexaferrite nanoparticles and the relation between their crystal structure and high coercivity properties. Advances in Natural Science and Nanotechnology. 2012;3(2):1-7. DOI:10.1088/2043-6262/3/2/025015
  8. 8. Singhal S, Namgyal T, Singh J, Chandra K, Bansal S. A comparative study on the magnetic properties of MFe12O19 and MAlFe11O19 (M = Sr, Ba and Pb) hexaferrites with different morphologies. Ceramics International. 2011;37(6):1833-1837. DOI: 10.1016/j.ceramint.2011.02.001
  9. 9. Pullar RC. Hexagonal ferrites: A review of the synthesis, properties and applications of hexaferrite ceramics. Progress in Materials Science. 2012;57(7):1191-1334. DOI: 10.1016/j.pmatsci.2012.04.001
  10. 10. Liu JP, Fullerton E, Gutfleisch O, Sellmyer DJ. Nanoscale Magnetic Materials and Applications. London, New York: Springer Dordrecht Heidelberg; 2010. DOI: 10.1007/978-0-387-85600-1
  11. 11. Sharma P, Rocha RA, Medeiros SN, Hallouche B, Paesano A. Structural and magnetic studies on mechanosynthesized BaFe12−xMnxO19. Journal of Magnetism and Magnetic Materials. 2007;316:29-33. DOI: 10.1016/j.jmmm.2007.03.207
  12. 12. Sánchez-DeJesús F, Bolarín-Miró AM, Cortés-Escobedo CA, Valenzuela R, Ammar S. Mechanosynthesis, crystal structure and magnetic characterization of M-type SrFe12O19. Ceramics International. 2014;40:4033-4038
  13. 13. Luo J. Structural and magnetic properties of Nd-doped strontium ferrite nanoparticles. Materials Letters. 2012;8:162-164
  14. 14. Faloh-Gandarilla JC, Dıaz-Castanon S, Watts BE. Magnetization reversal and interactions in SrFe12O19. Physica Status Solidi B. 2016;254(4):1-7. https://doi.org/10.1002/pssb.201600393
  15. 15. Ashiq MN, Iqbal MJ, Gul IH. Structural, magnetic and dielectric properties of Zr-Cd substituted strontium hexaferrite (SrFe12O19) nanoparticles. Journal of Alloys and Compounds. 2009;487:341-345
  16. 16. Kostishyn VG, Panina LV, Kozhitov LV, Timofeev AV, Kovalev AN. Synthesis and multiferroic properties of M-type SrFe12O19 hexaferrite ceramics. Journal of Alloys and Compounds. 2015;645:297-300
  17. 17. Wong YC, Wang J, Teh GB. Structural and magnetic studies of SrFe12O19 by sol-gel method. Procedia Engineering. 2014;76:45-52. DOI: 10.1016/j.proeng.2013.09.246
  18. 18. Zhang Z, Liu X, Wang X, Wu Y, Li R. Effect of Nd-Co substitution on magnetic and microwave absorption properties of SrFe12O19 hexaferrites. Journal of Alloys and Compounds. 2012;525:114-119
  19. 19. Afghahi SSS, Peymanfar R, Javanshir S, Atassi Y, Jafarian M. Synthesis, characterization and microwave characteristics of ternary nanocomposite of MWCNTs/doped Sr-hexaferrite/PANI. Journal of Magnetism and Magnetic Materials. 2016;423:152-157
  20. 20. Xia A, Zuo C, Chen L, Jin C, Lv Y. Hexagonal SrFe12O19 ferrites: Hydrothermal synthesis and their sintering properties. Journal of Magnetism and Magnetic Materials. 2013;332(4):186-191
  21. 21. Mirkazemi SM, Alamolhoda S, Ghiami Z. Erratum to: Microstructure and magnetic properties of SrFe12O19 nano-sized powders prepared by sol-gel auto-combustion method with CTAB surfactant. Journal of Superconductivity and Novel Magnetism. 2015;28(5):1551-1558. DOI: 10.1007/s10948-014-2872-x
  22. 22. Liu Y, Drew MG, Liu Y. Preparation and magnetic properties of barium ferrites substituted with manganese, cobalt, and tin. Journal of Magnetism and Magnetic Materials. 2007;323:945-953. DOI: 10.1016/j.jmmm.2010.11.075
  23. 23. Nazlan R, Hashim M, Ibrahim IR, Mohd Idris F, Wan Ab Rahman WN, Abdullah NH, et al. Influence of indium substitution and microstructure changes on the magnetic properties evolution of Y3Fe5−xInxO12 (x = 0.0 − 0.4). Journal of Materials Science: Materials in Electronics. 2015;26(6):3596-3609. DOI: 10.1007/s10854-015-2874-x
  24. 24. Fatemeh B, Masoud R, Mehdi MP. Synthesis of SrFe12O19/SiO2/TiO2 composites with core/shell/shell nano-structure and evaluation of their photo-catalytic efficiency for degradation of methylene blue. Journal of Materials Science: Materials in Electronics. 2018;29(3):1877-1887
  25. 25. Mangai KA, Sureshkumar P, Priya M, Rathnakumari M. Structural and magnetic properties of strontium hexa-ferrites for permenant magnets. International Journal of Scientific & Engineering Research. 2014;5(3):65-69
  26. 26. Masoudpanah SM, Seyyed Ebrahimi SA. Effect of pH value on the structural and magnetic properties of nanocrystalline strontium hexaferrite thin films. Journal of Magnetism and Magnetic Materials. 2011;323(21):2643-2647. DOI: 10.1016/j.jmmm.2011.05.055
  27. 27. Dang TMH, Trinh VD, Bui DH, Phan MH, Huynh DC. Sol-gel hydrothermal synthesis of strontium hexaferrite nanoparticles and the relation between their crystal structure and high coercivity properties. Advances in Natural Sciences: Nanoscience and Nanotechnology. 2012;3(2):25015
  28. 28. Wang Y, Li Q, Zhang C, Li B. Effect of Fe/Sr mole ratios on the formation and magnetic properties of SrFe12O19 microtubules prepared by sol-gel method. Journal of Magnetism and Magnetic Materials. 2009;321(19):3368-3372. DOI: 10.1016/j.jmmm.2009.05.066
  29. 29. Reza G, Ghasemi A, Saidi A. Enhanced magnetic properties of substituted Sr-hexaferrite nanoparticles synthesized by co-precipitation method. Ceramics International. 2014;40(3):4945-4952. DOI: 10.1016/j.ceramint.2013.10.096
  30. 30. Malhotra S, Chitkara M, Sandhu IS. Microwave absorption study of nano synthesized strontium ferrite particles in X band. International Journal of Signal Processing, Image Processing and Pattern Recognition. 2015;8(10):115-120
  31. 31. Maswadeh Y, Mahmood SH, Awadallah A, Aloqaily AN. Synthesis and structural characterization of nonstoichiometric barium hexaferrite materials with Fe:Ba ratio of 11.5–16.16. IOP Conference Series: Materials Science and Engineering. 2015;92:012019. DOI: 10.1088/1757-899X/92/1/012019
  32. 32. Pereira FMM, Junior CAR, Santos MRP, Sohn RSTM, Freire FNA, ki JM, et al. Structural and dielectric spectroscopy studies of the M-type barium strontium hexaferrite alloys (BaxSr1−xFe12O19). Journal of Materials Science: Materials in Electronics. 2008;19(7):627-638. DOI: 10.1007/s10854-007-9411-5
  33. 33. Song F, Shen X, Xiang J, Song H. Formation and magnetic properties of M-Sr ferrite hollow fibers via organic gel-precursor transformation process. Materials Chemistry and Physics. 2010;120(1):213-216. DOI: 10.1016/j.matchemphys.2009.10.048
  34. 34. Sivakumar M, Gedanken A, Zhong W, Du YW, Bhattacharya D. Nanophase formation of strontium hexaferrite fine powder by the sonochemical method using Fe(CO)5. Journal of Magnetism and Magnetic Materials. 2004;268:95-104. DOI: 10.1016/S0304-8853(03)00479-7
  35. 35. Nazlan R, Hashim M, Abdullah NH, Ibrahim IR, Ismail I. Influence of milling time on the crystallization, morphology and magnetic properties of polycrystalline yttrium iron garnet. Advanced Materials Research. 2012;501:324-328. DOI: 10.4028/www.scientific.net/AMR.501.324
  36. 36. Yang FJ, Min JJ, Kang ZW, Tu SY, Chen HB, Liu DG, et al. The influence of pH value and composition on the microstructure, magnetic properties of Co-Fe-Al Heusler nanoparticles. Chemical Physics Letters. 2017;670(3):1-4. DOI: 10.1016/j.cplett.2016.12.062
  37. 37. Tian L, Xu J, Xiao S. The influence of pH and bath composition on the properties of Ni-Co coatings synthesized by electrodeposition. Vacuum. 2011;86(1):27-33. DOI: 10.1016/j.vacuum.2011.03.027
  38. 38. Hesani M, Yazdani a, Abedi Ravan B, Ghazanfari M. The effect of particle size on the characteristics of FeCo nanoparticles. Solid State Communications. 2010;150(13–14):594-597. DOI: 10.1016/j.ssc.2009.12.043

Written By

Muhammad Syazwan Mustaffa, Rabaah Syahidah Azis and Sakinah Sulaiman

Submitted: 12 April 2018 Reviewed: 02 August 2018 Published: 13 February 2019