Next Article in Journal
Impact of PP Impurities on ABS Tensile Properties: Computational Mechanical Modelling Aspects
Next Article in Special Issue
Viscoelasticity Modeling of Dielectric Elastomers by Kelvin Voigt-Generalized Maxwell Model
Previous Article in Journal
Determining the Optimal Conditions for the Production by Supercritical CO2 of Biodegradable PLGA Foams for the Controlled Release of Rutin as a Medical Treatment
Previous Article in Special Issue
Flexible Methyl Cellulose/Polyaniline/Silver Composite Films with Enhanced Linear and Nonlinear Optical Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Novel Polyimide/Copper-Nickel Ferrite Composites with Tunable Magnetic and Dielectric Properties

1
“Petru Poni” Institute of Macromolecular Chemistry, 41A Aleea Gr. Ghica Voda, 700487 Iasi, Romania
2
Faculty of Chemistry, “Al. I. Cuza” University of Iasi, 11 Bd. Carol I, 700506 Iasi, Romania
3
Integrated Centre of Environmental Science Studies in the North-Eastern Region—CERNESIM, “Al. I. Cuza” University of Iasi, 11 Bd. Carol I, 700506 Iasi, Romania
4
Prosupport Consulting SRL, 29 Peter Culianu Street, Valea Lupului, 707410 Iasi, Romania
5
Faculty of Electrical Engineering, “Gh. Asachi” Technical University, B-Dul D. Mangeron 67, 700050 Iasi, Romania
*
Author to whom correspondence should be addressed.
Polymers 2021, 13(10), 1646; https://doi.org/10.3390/polym13101646
Submission received: 20 April 2021 / Revised: 14 May 2021 / Accepted: 14 May 2021 / Published: 19 May 2021
(This article belongs to the Special Issue Polymer-Based Dielectric Materials)

Abstract

:
Heat-resistant magnetic polymer composites were prepared by incorporating cerium-doped copper-nickel ferrite particles, having the general formula Ni1-xCuxFe1.92Ce0.08O4 (x: 0.0, 0.3, 0.6, 1.0), into a polyimide matrix. The effects of particle type and concentration on the thermal, magnetic, and electrical properties of the resulting composites were investigated. The samples were characterized by FTIR, scanning electron microscopy, X-ray diffractometry, thermogravimetric analysis, differential scanning calorimetry, vibrating sample magnetometer, and broadband dielectric spectroscopy. The composites exhibited high thermal stability, having initial decomposition temperatures between 495 and 509 °C. Saturation magnetization (Ms), magnetic remanence (Mr), and coercivity (Hc) were found in range of 2.37–10.90 emu g−1, 0.45–2.84 emu g−1, and 32–244 Oe, respectively. The study of dielectric properties revealed dielectric constant values of 3.0–4.3 and low dielectric losses of 0.016–0.197 at room temperature and a frequency of 1 Hz.

Graphical Abstract

1. Introduction

With exceptional properties resulting from a unique combination of organic and inorganic components, polymer nanocomposites are of great importance in current material chemistry research. In this regard, the polymeric phase confers flexibility and good processability, while the inorganic phase gives them thermal, optical, electrical, or magnetic characteristics [1]. Magnetic polymer nanocomposites are a particularly well-studied subgroup, since they can easily be prepared by incorporating magnetic inorganic nanoparticles into a polymer matrix, and have been extensively explored in a wide variety of high-performance applications, including in vivo imaging, photocatalysis, and heavy metal absorption [2,3,4,5].
Cobalt ferrite (CoFe2O4) is one of the most important spinel ferrites, having a high Curie temperature, relatively large magnetic hysteresis, moderate magnetic saturation, high coercivity, and high thermal stability. CoFe2O4 found its application in many fields including energy conversion and storage, magnetic recording and high-density digital recording disks, gas sensing, catalysis, biotechnology, magnetic bulk cores, and microwave absorbers [6,7]. It has also found use in biomedical applications such as magnetic thermo-drug delivery, hyperthermia, biosensors, and magnetic resonance imaging [8,9,10,11]. Besides Co2+, various other metallic cations such as Ni2+, Cu2+, Zn2+, Mg2+, Mn2+, have been partially or even totally introduced into the ferritic structure (M2+Fe2O4), resulting in nanoparticles with specific morphology and properties [12,13,14,15,16,17]. For example, it was found that for Co1-xNixFe2O4 nanofibers, saturation magnetization, and coercivity decreased with increasing nickel concentration due to the smaller magnetic moment of Ni2+ ions compared to Co2+ ions [12]. At the same time, saturation magnetization decreased at Co1-xCuxFe2O4 nanoparticles by increasing Cu2+ content, due to the smaller magnetic moment of Cu2+ ions in comparison with Co2+ ions [13]. Several studies reported that cerium insertion into the ferrite structure leads to an enhancement of its electrical and magnetic properties, which makes these nanoparticles more suitable for use in microwave adsorption and photocatalytic applications [18,19]. Moreover, calcination at 600 °C with only a small quantity of cerium slightly increases the ferrites’ magnetic properties, as our previous work demonstrates for CuFe1.92Ce0.08O4 [20], NiFe1.92Ce0.08O4, Ni0.7Cu0.3Fe1.92Ce0.08O4, and Ni0.4Cu0.6Fe1.92Ce0.08O4 complex systems [21].
Aromatic polyimides represent the most important class of thermostable heterocyclic polymers due to their outstanding properties (e.g., high thermo-oxidative resistance, high glass transition temperatures, very good mechanical, and electrical properties) and their use in various technical applications (e.g., aerospace and microelectronics) [22,23,24,25]. Due to their aromatic rigid structure, polyimides are often insoluble and infusible and thus can be processed with difficulty. The incorporation of flexible linkages into their main chains, such as aromatic ether groups, can substantially improve the processability of these polymers. The introduction of magnetic inorganic particles into a polyimide matrix can produce materials with improved properties exhibiting high thermal stability, excellent mechanical behavior, magnetic properties, and flexibility [26,27,28,29,30,31,32,33,34]. Flexible magnetic polyimide films are important for the fabrication of flexible electronics, devices which have been developed rapidly in the last years.
In this study, the preparation and characterization of novel heat-resistant magnetic polyimide composites based on cerium-doped copper-nickel ferrite particles are reported. An imide-copolymer having pendant carboxylic groups and ether linkages was used as the polymer matrix. The effect of ferrite particles on the thermal, magnetic, and electrical properties of the resulting composites was also investigated. We show that the content and type of ferrite particles allow the fine tuning of magnetic and dielectric characteristics of these composites. Finally, we discuss the sub-glass and glass-rubber transitions, the conductivity as well as the changes produced in the relaxation processes.

2. Experimental

2.1. Materials and Methods

2.1.1. Materials

4,4′-Oxydiphthalic anhydride, 2,2-bis [4-(4-aminophenoxy)phenyl]propane, 3,5-diaminobenzoic acid, and N-methyl-2-pyrrolidone (NMP) were purchased from Aldrich and used as received. Cerium-doped copper-nickel ferrite particles were prepared as previously described [20,21]. Table 1 illustrates the chemical structure and main magnetic characteristics of the particles: saturation magnetization (Ms), remanence magnetization (Mr), and coercivity (Hc).

2.1.2. Preparation of Polyimide Composites

Magnetic polyimide composites were prepared by incorporating magnetic particles into a poly(amic acid) solution, followed by film casting and thermal imidization, in accordance with our previous work (Figure 1) [35]. Magnetic particles and NMP (11 mL) were introduced into a flask and the resulting mixture was sonicated for 2 h, after which it was mechanically stirred for 6 h until a suspension was obtained. Afterwards, 2,2-bis[4-(4-aminophenoxy)phenyl]propane (0.738 g, 0.0018 mol) and 3,5-diaminobenzoic acid (0.0104 g, 0.0002 mol) were added, and the mixture was stirred under nitrogen to complete dissolution. Furthermore, 4,4′-oxydiphthalic anhydride (0.620 g, 0.002 mol) was introduced and stirring was continued at room temperature for 12 h. Entrapped air bubbles were removed under vacuum. The suspension was cast onto glass plates and heated at 50, 100, 150, 200 °C, for 1 h each, and at 250 °C for 2 h. The inorganic content of the composites was controlled by the portion ratio between ferrite particles and poly(amic acid). The concentration of particles was calculated considering the appropriate amount of poly(amic acid) corresponding to final imide structure of the polymer matrix. Four polyimide composites PI-A-10, PI-A-20, PI-A-30, and PI-A-40 containing ferrite particles of type A in concentration of 10, 20, 30, and 40 wt%, respectively, and three polyimide composites PI-B-30, PI-C-30, and PI-D-30 with 30 wt% ferrite particles of type B, C, and D, respectively, were prepared (Table 2). A reference polyimide film without magnetic particles PI-0 was prepared as well.

2.1.3. Measurements

FTIR spectra were recorded on a Bruker Vertex 70 Spectrometer at frequencies ranging from 400 to 4000 cm−1.
Microscopic investigations were performed on an environmental scanning electron microscope type Quanta 200 operating at 10 kV with secondary and backscattering electrons in low vacuum mode. The Quanta 200 microscope is equipped with an energy dispersive X-ray (EDX) system for qualitative and quantitative analysis, as well as elemental mapping.
X-ray diffraction measurements (XRD) were carried out on a XpertPanalytical Pro MPD diffractometer equipped with a copper source radiation (Cu-Kα1 radiation, λ = 1.5406 Å) in 10–80° over 2θ range at room temperature.
Thermogravimetric (TG) curves and thermogravimetric derivative (DTG) curves were recorded with a Mettler Toledo TGA-SDTA851e equipment, in nitrogen atmosphere, under dynamic conditions with a flow rate of 20 mL min−1 and a heating rate of 10 °C min−1, in the temperature range of 25–900 °C and with a sample mass between 2.6 and 4.7 mg.
Differential scanning calorimetry (DSC) curves were obtained using a Mettler Toledo DSC1 type device in an inert atmosphere, with a heating rate of 10 °C min−1 and nitrogen purge at 100 mL min−1. Scanning was performed in the temperature range of 25–350 °C. The mass of samples encapsulated in aluminum pans having pierced lids to allow the evaporation of the volatile components was between 2.4 and 4.7 mg. The evaluation of TG and DSC curves was performed with STARe software, version 9, of Mettler-Toledo.
Room temperature magnetic properties were investigated in a ±10 kOe range using a vibrating sample magnetometer (VSM) (AGM & VSM Magnetometer, MicroMag 3900 series from former Princeton Measurement Co., currently Lake Shore Cryotronics, Inc., Westerville, OH, USA).
Broadband dielectric spectroscopy measurements were carried out on a dielectric spectrometer (Novocontrol Technologies, GmbH@Co., Montabaur, Germany) in a broad range of frequency (from 1 to 106 Hz) and temperature (from −150 to 250 °C). The alternating electrical field was applied with an Alpha-A High Performance Frequency Analyzer device and the temperature was controlled with 0.1 °C stability and high reproducibility using a Novocontrol Quatro Cryo system device. Composite samples were sandwiched between two gold-plated flat electrodes, and the measurements were performed in dry nitrogen.

3. Results and Discussions

3.1. Structural and Morphological Characterization

The chemical structure of the samples was confirmed by FTIR spectroscopy. Characteristic absorption bands of imide rings were observed in all spectra at around 1770 cm−1 (imide carbonyl asymmetric stretching), 1715 cm−1 (imide carbonyl symmetric stretching and C=O stretching of carboxylic pendant groups), 1370 cm−1 (C-N stretching), and 742 cm−1 (C=O bending). All the spectra showed absorption bands at 2970 and 2865 cm−1 (C-H of isopropylidene groups), 3060 cm−1 (C-H aromatic), and 1603 and 1501 cm−1 (-C=C- aromatic). The characteristic band for amide carbonyl at 1650 cm−1 was not observed in the spectra, suggesting that the imidization reaction took place. The FTIR spectra of composites were similar to that of reference polyimide PI-0, indicating that there is no chemical bonding between particles and polymer, as well as no degradation of the polymer matrix (Figure S1 from the Electronic Supplementary Data, ESD).
The morphology of the samples was investigated by SEM analysis, micrographs of PI-0, PI-A-10, PI-A-30, and PI-A-40 being illustrated in Figure 2. PI-0 had a homogeneous microstructure, while the surfaces of the composites were significantly rougher in comparison with the polyimide matrix PI-0, since the motion of the polymer chains was restricted by the inorganic filler. Some particle aggregations were observed, even in the case of PI-A-10. The dispersed particles had irregular shape and showed agglomerations attributed to the magnetic dipole–dipole interactions between inorganic particles. The semi-quantitative EDX analysis of PI-A-30 confirmed the presence of metal atoms Ce, Fe, Ni, and Cu. A mapping technique was used in order to illustrate the atom distribution on the composite surface. The EDX mapping of Ce, Fe, Ni, and Cu atoms showed a relatively uniform distribution of the particles (Figure S2 from ESD).

3.2. X-ray Analysis

The XRD patterns of the ferrite particles A, B, C, and D are depicted in Figure S3 from ESD. All ferrite samples used for those currently studied polyimide composites were previously reported [20,21]. Sample A, nominally Ni0.7Cu0.3Fe1.92Ce0.08O4 contains a 0.5 wt% of secondary cubic-CeO2 identified with ICSD card no. 01-073-6318. B sample (NiFe1.92Ce0.08O4) is considered 99.9% pure and C ferrite sample (Ni0.4Cu0.6Fe1.92Ce0.08O4) contains 1.6 wt% monoclinic CuO identified as ICSD card no. 01-070-6828. The XRD analysis and Rietveld refinement for D ferrite (CuFe1.92Ce0.08O4) confirmed the formation of cubic spinel ferrite with secondary 4.71 wt% tetragonal ferrite indexed as ICSD card no. 01-072-1174, 10.82 wt% monoclinic copper oxide, and 0.52 wt% cubic cerium oxide.
X-ray results of PI-0 and polyimide composites are illustrated in Figure 3. The diffractogram of PI-0 indicates an amorphous phase with two wide Bragg reflections, one around 12° and the second one around 21° (Figure 3a). The diffractograms of composites based on A-type ferrite (PI-A-10, PI-A-20, PI-A-30, and PI-A-40) have specific reflections corresponding to the cubic structure of Ni0.7Cu0.3Fe1.92Ce0.08O4 (A) particles. A noticeable increase in reflection intensity can be observed with increasing ferrite filling into the polyimide composite due to component ratio modification. At the same time, when compared to pure PI-0, the broad reflections of polyimide decreased in intensity with increasing magnetic particle content. This is due to a decrease in polymer content, coupled with the reduction in amorphization due to particle presence.
The diffractograms of composites with 30% ferrite (Figure 3b), PI-A-30, PI-B-30, PI-C-30, and PI-D-30, showed specific reflections corresponding to inorganic phases A, B, C, and D, respectively. The crystallinity of particles in composites was similar to that of particles in bulk form, suggesting that the particles are not modified by the thermal treatment during sample preparation.

3.3. Thermal and Magnetic Properties

Thermogravimetric analysis (TGA) was used to evaluate the thermal stability of the polyimide composites (Table 3). Figure S4 from ESD shows TG and DTG curves of PI-0, PI-A-10, PI-A-20, and PI-A-40. All samples exhibited high thermal stability as the initial decomposition temperature was found in 495–509 °C range. The temperature corresponding to 10% weight loss, T10, was 508–523 °C, while the temperature corresponding to maximum polymer decomposition, Tmax, was 509–527 °C. The char yield at 900 °C ranged between 42% and 53%, except PI-A-10 containing 10% particles, which showed a char yield of 26%, suggesting that the particles had a higher catalytic effect on the thermal decomposition reactions. A slight decrease in T5 and T10 values with increasing particle content could also be observed.
The glass transition temperature (Tg) of the samples varied from 233 to 241 °C and slightly decreased by the incorporation of particles in the polymer matrix, suggesting that the presence of particles reduced the interactions between the macromolecular chains. Thus, Tg of PI-0 was 241 °C, while Tg of the composites varied in the 233–235 °C range (Table 3). However, the increase in particles content or the introduction of various type of particles have not had influence on Tg values.
All composites exhibited a ferrimagnetic behavior due to particle interaction. The characteristic magnetic parameters including Ms, Mr, and Hc, determined from the hysteresis loops, ranged in the intervals 2.37–10.90 emu g−1, 0.45–2.84 emu g−1, and 32–244 Oe, respectively (Table 3). The hysteresis loops of the samples containing 10%, 20%, 30%, and 40% A-type filler, PI-A-10, PI-A-20, PI-A-30, and PI-A-40, are shown in Figure 4a, and those corresponding to the samples with different types of particles at 30% concentration, PI-A-30, PI-B-30, PI-C-30, and PI-D-30, are illustrated in Figure 4b. As can be seen from Figure 4a and Table 3, increasing ferrite insertion ratio led to the progressive modification of Ms and Mr, while coercivity stayed in the same range (79–90 Oe). Ms and Mr increased in a linear fashion for 10–30% filler concentration, as was expected, since an increase in particle concentration leads to an increase in the magnetic moment per unit volume of the sample, and hence its magnetization increases [4]. Between insertion ratio of 30% and 40%, Ms and Mr exhibited a lower increase, indicating a tendency of the composites to reach magnetic saturation by canceling the magnetic moments due to interparticle interactions. This effect limited the efficiency of ratio modification up to 30% particles. It was also evident that the Ms values of composites were lower than those of bulk magnetic particles.
The composites’ coercivity depended on particle type and exhibited lower variation with particle content. The magnetic values of composites suggest a different magnetic behavior induced by nickel substitution. For samples A, B, and C, the magnetic value changes are proportionally dropping, indicating that the sample magnetization and coercivity may decrease as a consequence of reduced interactions between FeTd-NiOh, which are substituted with weaker FeTd-CuOh interactions, as suggested by Patange et al. [36].

3.4. Dielectric Constant and Dielectric Loss

Broadband dielectric spectroscopy was employed for determining the behavior of electrical properties, namely, dielectric constant (ε′) and dissipation factor, also known as dielectric loss (ε″). The evolutions of dielectric constant and dielectric loss with frequency at 25 °C are represented in Figure 5 for the samples for which the ferrite loading varied from 0% to 30%. The dielectric constant decreased with increasing frequency (Figure 5a) since the molecular dipoles need a longer time than the alternating period to orient in the direction of the external electric field. As shown in Figure 5a, the isothermal plots displayed two different regions: a low-frequency region and a high-frequency region. In the low-frequency region, PI-0 and PI-A-10 exhibited a slight reduction in ε′, suggesting low dipolar activity. On the other hand, samples with superior ferrite content exhibited a higher value for ε′ and consequently, an elevated polarization ability. In the high-frequency region, the magnitude of ε′ was not significantly affected by the incorporation of ferrite in the PI backbone. Supplementary, the ε′(f) dependences of PI-A-30, PI-B-30, PI-C-30, and PI-D-30 are presented in Figure S5a from ESD. The magnitude of ε′ is diminished with the change of substitution ratio from x = 0 to 0.6, especially in the low-frequency region.
According to Figure 5b and Figure S5b from ESD, the magnitude of dielectric loss was considerably affected by the ferrite content and the nickel substitution ratio, respectively. At the same time, the ε″(f) trends showed specific dielectric signals for secondary γ- and β-relaxations. Therefore, at high frequencies, γ-relaxation appeared as a well-defined dielectric peak while the head-part of β-relaxation raised at low frequencies.
The values of dielectric parameters of the composites of different ferrite content are presented in Table 4. The lowest values of ε′ were observed for PI-0 throughout the considered frequency range (e.g., for f = 1 Hz, ε′ = 3), which were close to other reported polyimides [37,38]. Addition of ferrite particles in the polymer matrix increased the dielectric constant. Moreover, the values of the dielectric constant were influenced by the type of magnetic ferrite, especially due to nickel substitution (Table 4). At room temperature, ε′ of composites with various ferrite loadings ranged from 3.3 to 4.3, accompanied by low dielectric losses, between 10−1 and 10−2, and low conductivities, around 10−13–10−14 S cm−1 (1 Hz). From these data, it is clear that ferrite particle embedding in the polyimide matrix induces dielectric constant and conductivity enrichment, the greatest enhancement being observed at the highest ferrite concentration.
Figure 6 depicts the temperature behavior of dielectric constant and loss factor of the samples at a frequency of 1 Hz. First of all, it can be observed that from −150 to 200 °C, ε′ increased slowly with temperature, while the loss factor values were below 0.1, allowing the use of this type of composites as stable dielectrics in a broad range of temperature. Second, the magnitude of temperature dependences varied substantially with the ferrite content. Furthermore, the isochronal plots of dielectric loss revealed the presence of some important dipolar relaxation processes. At low temperatures, around −100 °C, the secondary γ-relaxation appeared as a narrow dielectric peak, which is generally activated by small groups of macromolecules that have attached dipolar water molecules from atmosphere. As temperature increased, β-relaxation process became visible (e.g., for PI-A-10, β-relaxation appeared around 30 °C) as a broad band with reduced intensity. According to already published literature, the β-relaxation is assigned to the cooperative rotational motions involving segmental motions from the main chain [39]. At higher temperatures, the dielectric loss increased considerably due to raised mobility of charge carriers.
Figure 7 depicts the evolution of dielectric loss with temperature at selected frequencies for the sample PI-A-30. The secondary γ- and β-relaxations were clearly recognized as intensive dielectric peaks that shifted gradually to higher temperatures as frequency increased from 1 to 106 Hz. The dipolar processes corresponding to PI-A-30 were easier visualized on the 3D version of dielectric loss as function of both the applied frequency and the temperature (Figure 7b).

3.5. Activation Energy

Detailed ε″(f) dependencies recorded at various temperatures are shown for PI-A-10 in Figure S6 from ESD. The secondary γ-relaxation appeared as one intense peak at −80 °C that shifted to higher frequencies and lowered in intensity as temperature increased (Figure S6a from ESD). The β-relaxation appeared as a broad peak at low frequencies (e.g., at 80 °C, the peak maximum appeared at around 102 Hz) and its signal was strongly affected by the conductivity signal. The remaining composites with various filling ratios provided analogous dipolar signals.
As previously discussed, the dipolar relaxation peaks are more or less overlapped with the conductivity signal. As a consequence, the ε″(f) dependencies were fitted considering the Havriliak–Negami (HN) expression:
ε * = ε i ε = ε u + ε r ε u [ 1 + ( i ω τ H N ) a ] b ,
where εr and εu are the relaxed and unrelaxed values of the dielectric constant, ω represents the angular frequency, τHN is the HN relaxation time of the process, while a and b represent the broadening and skewing parameters, respectively [40]. Deconvolution was performed with the WinFit software provided by Novocontrol and the relaxation time of each dielectric peak maxima, τmax, was calculated with the following relation:
τ m a x = τ H N [ s i n π a b 2 + 2 b s i n π a 2 + 2 b ] 1 a .
Figure S7 from ESD depicts an example of HN fitting for the secondary relaxation processes. The HN fit from Figure S8 from ESD encloses a HN term attributed to γ-relaxation and a supplementary term for both the β-relaxation and conductivity. The dependency of ε″(f) on β-relaxation was further processed with two HN functions (one for β-relaxation at around 102 Hz and one for γ-relaxation limited at high frequencies), as well as an additional function for conductivity.
The numerical values of broadening and skewing parameters from Equation (1) are represented as a function of temperature in Figure 8. The skewing parameter of PI-0 was around 0.3, indicating an asymmetry of the γ-relaxation. This parameter increased in value with the addition of ferrite, as well as with rising temperature, suggesting an improved symmetry. In addition, for composites with 30% ferrite loadings and different types of magnetic ferrite, the symmetry of the relaxation peaks increased considerably (see the Figure S8 from ESD). On the other hand, the width parameter was substantially reduced with increasing loadings of ferrite, suggesting complex motions of the side chain groups. Regarding the HN fitting process of β-relaxation, the dipolar peaks were symmetrical and, as consequence, the dielectric spectra were processed with b = 1 in the entire temperature range.
The Arrhenius equation was used to determine the activation energy, Ea, of the relaxation processes:
τ m a x ( T ) = τ 0 exp ( E a R T ) ,
where τmax was previously obtained from Equation (2), τ0 is a pre-exponential factor and R is the gas constant. The relaxation times associated with the peak maxima are plotted as a function of inverse temperature in Figure 9, and the numerical values of Equation (3) can be consulted in Table 5. The secondary γ-relaxation is characterized by small activation energy values, between 27 and 43 kJ mol−1, revealing that this relaxation is non-cooperative and can only be stimulated by limited fluctuations of small groups. The activation energy values were lower for composites than that of neat polyimide, indicating that the inorganic particles led to fewer restrictions to this dipolar relaxation. On the other hand, the values of Ea for β-relaxation were higher than that for γ-relaxation, confirming the existence of cooperative rotational motions. The proportional increase in Ea with the concentration of magnetic ferrite indicates the reinforcement of the polymer chains by the nanoparticles, and as a consequence, the hindered movement of chemical dipoles. It is important to note that β-relaxation could not be accurately determined for composites with different types of magnetic ferrite due to the signal overlap with conductivity.

3.6. Conductivity

The evolution of measured conductivity as a function of alternating field frequency at a moderate temperature (25 °C) and at a high temperature (250 °C) is presented in Figure 10. At 25 °C, the measured conductivity increased almost linearly with frequency for all considered samples (Figure 10a). The conductivity values were low (e.g., for PI-A-30, at f = 1 Hz, the measured conductivity was ∼10−13 S cm−1), due to the high insulator-type characteristic of the polyimide matrix (Table 4). At high temperatures (Figure 10b), the conductivity substantially increased, especially for samples containing a high amount of ferrite. It is clear that the addition of ferrite progressively increased the conductivity of the samples. The increase in conductivity with increasing temperature is a normal effect for polymer-based composites. For polyimides, it is known that the cooperative segment motions of the polymer backbone are activated at temperatures above glass transition, and consequently, the polymer matrix favors the movements of charge carriers. On the other hand, the incorporation of ferrite furnishes a proportional concentration of Fe3+ ions into the polyimide matrix. Thus, a higher amount of ferrite leads to the increased conductivity of PI-A-30 and PI-A-40.

4. Conclusions

Novel magnetic polyimide/copper-nickel ferrite composites were developed using cerium-doped ferrite particles, namely, NiFe1.92Ce0.08O4, Ni0.7Cu0.3Fe1.92Ce0.08O4, Ni0.4Cu0.6Fe1.92Ce0.08O4, and CuFe1.92Ce0.08O4, as inorganic fillers. An aromatic polyimide containing pendant carboxylic groups was used as polymer matrix in order to ensure a good dispersion of the filler particles. The influence of filler content on the composite properties was evidenced through specific ferrimagnetic behavior in the case of all the samples. Ms and Hc of composite materials strongly depended on ferrite composition and progressively increased up to a 40% filler ratio, allowing an efficient control over the polymer-ferrite composite magnetic values. The tendency to reach maximum Ms and Hc for the polymer-ferrite material can be explained by particle agglomeration at high ferrite content, as revealed by SEM investigations. According to TGA analysis, the composites showed high heat resistance, with initial decomposition temperatures in the range of 495–509 °C. The dielectric properties, as determined by broadband dielectric spectroscopy measurements, were related to ferrite structure and loading level. The observed low values for the dielectric parameters in a broad temperature range (between −150 and 200 °C) demonstrate a high thermal stability of these composites, allowing their use as stable dielectrics in a broad temperature range.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/polym13101646/s1, Figure S1: FTIR spectra of PI-0 and PI-A-30., Figure S2: EDX diagram (a) and EDX mapping (Ce, Fe, Ni and Cu atoms) (b) of PI-A-30., Figure S3: XRD patterns of inorganic fillers Ni1-xCuxFe1.92Ce0.08O4 (with x = 0.0, 0.3, 0.6, 1.0), A, B, C and D., Figure S4: TG (a) and DTG (b) curves of PI-0, PI-A-10, PI-A-20 and PI-A-40., Figure S5: Dielectric constant (a) and dielectric loss (b) evolution with alternating frequency for samples with 30% ferrite content at 25°C., Figure S6: Representative ε”(f) dependences for isothermal γ-relaxation (a) and β-relaxation (b) of PI-A-10., Figure S7: Representative deconvolution process for γ (a) and β (b) relaxations of PI-A-10., Figure S8: The evolution of broadening and skewing parameters as function of temperature for γ-relaxation of composites with different types of magnetic ferrite.

Author Contributions

Conceptualization, C.H. and A.P.; methodology, C.H., M.A., E.H., T.R. and M.A.O.; software, M.A. and T.R.; validation, C.H., M.A., E.H., T.R. and M.A.O.; formal analysis, E.H. and M.A.O.; investigation, M.A., E.H., T.R. and M.A.O.; resources, C.H. and A.P.; data curation, M.A.O.; writing—original draft preparation, C.H.; writing—review and editing, M.A., and T.R.; visualization, C.H. and T.R.; supervision, C.H. and A.P.; project administration, C.H. and A.P.; funding acquisition, C.H. and A.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by a grant of the Romanian Ministry of Education and Research, CNCS-UEFISCDI, project number PN-III-P4-ID-PCE-2020-1385, within PNCDI III.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

This work was supported by a grant of the Romanian Ministry of Education and Research, CNCS-UEFISCDI, project number PN-III-P4-ID-PCE-2020-1385, within PNCDI III.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Anandhan, S.; Bandyopadhyay, S. Polymer Nanocomposites: From Synthesis to Applications. In Nanocomposites and Polymers with Analytical Methods; Cuppoletti, J., Ed.; InTech Europe: Rijeka, Croatia, 2011. [Google Scholar] [CrossRef] [Green Version]
  2. Thévenot, J.; Oliveira, H.; Sandre, O.; Lecommandoux, S. Magnetic responsive polymer composite materials. Chem. Soc. Rev. 2013, 42, 7099–7116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Kalia, S.; Kango, S.; Kumar, A.; Haldorai, Y.; Kumari, B.; Kumar, R. Magnetic polymer nanocomposites for environmental and biomedical applications. Colloid Polym. Sci. 2014, 292, 2025–2052. [Google Scholar] [CrossRef]
  4. Agami, W.R.; El-Sayed, H.M. Enhancement of the magnetic and dielectric properties of cobalt nanoferrite/polymethyl methacrylate composites. J. Mater. Sci. Mater. Electron. 2015, 26, 3163–3167. [Google Scholar] [CrossRef]
  5. Antonel, P.; Berhó, F.; Jorge, G.; Molina, F. Magnetic composites of CoFe2O4 nanoparticles in a poly(aniline) matrix: Enhancement of remanence ratio and coercivity. Synth. Met. 2015, 199, 292–302. [Google Scholar] [CrossRef]
  6. Lalwani, S.; Marichi, R.B.; Mishra, M.; Gupta, G.; Singh, G.; Sharma, R.K. Edge enriched cobalt ferrite nanorods for symmetric/asymmetric supercapacitive charge storage. Electrochim. Acta 2018, 283, 708–717. [Google Scholar] [CrossRef]
  7. Routray, K.L.; Saha, S.; Behera, D.; Tripathy, N.; Ghosh, S.P. High dielectric constant, low loss and efficient visible photoluminescence properties of porous rose- flower shaped CoFe2O4 for photovoltaic application. Mater. Lett. 2019, 242, 62–65. [Google Scholar] [CrossRef]
  8. Stanciu, L.; Won, Y.-H.; Ganesana, M.; Andreescu, S. Magnetic Particle-Based Hybrid Platforms for Bioanalytical Sensors. Sensors 2009, 9, 2976–2999. [Google Scholar] [CrossRef] [PubMed]
  9. Amiri, S.; Shokrollahi, H. The role of cobalt ferrite magnetic nanoparticles in medical science. Mater. Sci. Eng. C 2013, 33, 1–8. [Google Scholar] [CrossRef] [PubMed]
  10. Mushtaq, M.W.; Kanwal, F.; Batool, A.; Jamil, T.; Zia-Ul-Haq, M.; Ijaz, B.; Huang, Q.; Ullah, Z. Polymer-coated CoFe2O4 nanoassemblies as biocompatible magnetic nanocarriers for anticancer drug delivery. J. Mater. Sci. 2017, 52, 9282–9293. [Google Scholar] [CrossRef]
  11. Hashim, M.; Raghasudha, M.; Meena, S.S.; Shah, J.; Shirsath, S.E.; Kumar, S.; Ravinder, D.; Bhatt, P.; Alimuddin; Kumar, R.; et al. Influence of rare earth ion doping (Ce and Dy) on electrical and magnetic properties of cobalt ferrites. J. Magn. Magn. Mater. 2018, 449, 319–327. [Google Scholar] [CrossRef]
  12. Xiang, J.; Chu, Y.; Shen, X.; Zhou, G.; Guo, Y. Electrospinning preparation, characterization and magnetic properties of cobalt–nickel ferrite (Co1−xNixFe2O4) nanofibers. J. Colloid Interface Sci. 2012, 376, 57–61. [Google Scholar] [CrossRef]
  13. Hashim, M.; Alimuddin; Shirsath, S.E.; Meena, S.S.; Kotnala, R.; Kumar, S.; Bhatt, P.; Jotania, R.; Kumar, R. Study of structural and magnetic properties of (Co–Cu)Fe2O4/PANI composites. Mater. Chem. Phys. 2013, 141, 406–415. [Google Scholar] [CrossRef]
  14. Doulabi, F.S.M.; Mohsen-Nia, M. Magnetic cobalt-zinc ferrite/PVAc nanocomposite: Synthesis and characterization. Iran. Polym. J. 2013, 22, 9–14. [Google Scholar] [CrossRef]
  15. Gherca, D.; Pui, A.; Cornei, N.; Cojocariu, A.; Nica, V.; Caltun, O. Synthesis, characterization and magnetic properties of MFe2O4 (M=Co, Mg, Mn, Ni) nanoparticles using ricin oil as capping agent. J. Magn. Magn. Mater. 2012, 324, 3906–3911. [Google Scholar] [CrossRef]
  16. Gherca, D.; Pui, A.; Nica, V.; Caltun, O.; Cornei, N. Eco-environmental synthesis and characterization of nanophase powders of Co, Mg, Mn and Ni ferrites. Ceram. Int. 2014, 40, 9599–9607. [Google Scholar] [CrossRef]
  17. Ciocarlan, R.G.; Pui, A.; Gherca, D.; Virlan, C.; Dobromir, M.; Nica, V.; Craus, M.L.; Gostin, I.N.; Caltun, O.; Hempelman, R.; et al. Quaternary M0.25Cu0.25Mg0.5Fe2O4 (M=Ni, Zn, Co, Mn) ferrite oxides: Synthesis, characterization and magnetic properties. Mater. Res. Bull. 2016, 81, 63–70. [Google Scholar] [CrossRef]
  18. Mosleh, Z.; Kameli, P.; Poorbaferani, A.; Ranjbar, M.; Salamati, H. Structural, magnetic and microwave absorption properties of Ce-doped barium hexaferrite. J. Magn. Magn. Mater. 2016, 397, 101–107. [Google Scholar] [CrossRef]
  19. Rahimi-Nasrabadi, M.; Behpour, M.; Sobhani-Nasab, A.; Jeddy, M.R. Nanocrystalline Ce-doped copper ferrite: Synthesis, characterization, and its photocatalyst application. J. Mater. Sci. Mater. Electron. 2016, 27, 11691–11697. [Google Scholar] [CrossRef]
  20. Roman, T.; Pui, A.; Lukacs, A.V.; Cimpoesu, N.; Lupescu, S.; Borhan, A.I.; Kordatos, K.; Ntziouni, A.; Postolache, P.; Zaharia, M.; et al. Structural changes of cerium doped copper ferrites during sintering process and magneto-electrical properties assessment. Ceram. Int. 2019, 45, 17243–17251. [Google Scholar] [CrossRef]
  21. Roman, T.; Ghercă, D.; Borhan, A.-I.; Grigoraș, M.; Stoian, G.; Lupu, N.; Turcan, I.; Cimpoesu, N.; Istrate, B.; Radu, I.; et al. Nanostructured quaternary Ni1-xCuxFe2-yCeyO4 complex system: Cerium content and copper substitution dependence of cation distribution and magnetic-electric properties in spinel ferrites. Ceram. Int. 2021. [Google Scholar] [CrossRef]
  22. Liaw, D.-J.; Wang, K.-L.; Huang, Y.-C.; Lee, K.-R.; Lai, J.-Y.; Ha, C.-S. Advanced polyimide materials: Syntheses, physical properties and applications. Prog. Polym. Sci. 2012, 37, 907–974. [Google Scholar] [CrossRef]
  23. Vanherck, K.; Koeckelberghs, G.; Vankelecom, I.F. Crosslinking polyimides for membrane applications: A review. Prog. Polym. Sci. 2013, 38, 874–896. [Google Scholar] [CrossRef]
  24. Wang, Y.; Gao, Q.; You, Q.; Liao, G.; Xia, H.; Wang, D. Porous polyimide framework: A novel versatile adsorbent for highly efficient removals of azo dye and antibiotic. React. Funct. Polym. 2016, 103, 9–16. [Google Scholar] [CrossRef]
  25. Yi, L.; Huang, W.; Yan, D. Polyimides with side groups: Synthesis and effects of side groups on their properties. J. Polym. Sci. Part A Polym. Chem. 2017, 55, 533–559. [Google Scholar] [CrossRef]
  26. Mazuera, D.; Perales, Ó.; Suarez, M.; Singh, S.P. Synthesis, characterization and thermal analysis of polyimide–cobalt ferrite nanocomposites. Mater. Sci. Eng. A 2010, 527, 6393–6399. [Google Scholar] [CrossRef]
  27. Zhan, J.; Wu, D.; Chen, C.; Li, J.; Pan, F.; Chen, J.; Zhan, X. Fabrication of polyimide composite film with both magnetic and surface conductive properties. Desalin. Water Treat. 2011, 34, 344–348. [Google Scholar] [CrossRef]
  28. Jia, H.; Jing, J.; Zhao, X.; Wang, W.; Wang, D.; Liu, C.; Zhou, H. Preparation and characterization of polyimide magnetic hollow nanospheres. Mater. Lett. 2012, 68, 86–89. [Google Scholar] [CrossRef]
  29. Moghanian, H.; Mobinikhaledi, A.; Baharangiz, Z. Synthesis, characterization and magnetic properties of novel heat resistant polyimide nanocomposites derived from 14H-dibenzo [a,j] xanthene. J. Polym. Res. 2014, 21. [Google Scholar] [CrossRef]
  30. Ding, D.; Yan, X.; Zhang, X.; He, Q.; Qiu, B.; Jiang, D.; Wei, H.; Guo, J.; Umar, A.; Sun, L.; et al. Preparation and enhanced properties of Fe3O4 nanoparticles reinforced polyimide nanocomposites. Superlattices Microstruct. 2015, 85, 305–320. [Google Scholar] [CrossRef] [Green Version]
  31. Liu, C.; Qu, C.; Wang, D.; Feng, H.; Liu, P.; Zhang, Y. Preparation and characterization of magnetic polyimide composite films copolymerized with aminophthalocyanine-coated Fe3O4 nanocrystals. J. Mater. Sci. Mater. Electron. 2015, 26, 4005–4014. [Google Scholar] [CrossRef]
  32. Luo, C.; Wang, X.; Wang, J.; Pan, K. One-pot preparation of polyimide/Fe3O4 magnetic nanofibers with solvent resistant properties. Compos. Sci. Technol. 2016, 133, 97–103. [Google Scholar] [CrossRef]
  33. Rybak, A.; Rybak, A.; Kaszuwara, W.; Awietjan, S.; Molak, R.; Sysel, P.; Grzywna, Z.J. The magnetic inorganic-organic hybrid membranes based on polyimide matrices for gas separation. Compos. Part B Eng. 2017, 110, 161–170. [Google Scholar] [CrossRef]
  34. Faghihi, K.; Moghanian, H.; Mozafari, F.; Shabanian, M. Magnetic and Heat Resistant Poly(imide-ether) Nanocomposites Derived from Methyl Rich 9H-xanthene: Synthesis and Characterization. Chin. J. Polym. Sci. 2018, 36, 822–834. [Google Scholar] [CrossRef]
  35. Hamciuc, C.; Hamciuc, E.; Asandulesa, M.; Kalvachev, Y.; Dimitrov, L.; Ignat, M. Study on dielectric behavior and harvesting properties of new Ba0.5Sr0.4Ca0.1TiO3/poly(ether imide) composite films. Mater. Res. Bull. 2018, 102, 70–78. [Google Scholar] [CrossRef]
  36. Patange, S.M.; Shirsath, S.E.; Toksha, B.G.; Jadhav, S.S.; Jadhav, K.M. Electrical and magnetic properties of Cr3+ substituted nanocrystalline nickel ferrite. J. Appl. Phys. 2009, 106, 023914. [Google Scholar] [CrossRef]
  37. Chern, Y.-T.; Shiue, H.-C. Low Dielectric Constants of Soluble Polyimides Based on Adamantane. Macromolecules 1997, 30, 4646–4651. [Google Scholar] [CrossRef]
  38. Hamciuc, C.; Hamciuc, E.; Vlad-Bubulac, T.; Vlad, S.; Asandulesa, M.; Wolinska-Grabczyk, A. Silica-containing polyetherimide hybrid films based on methyltriethoxysilane as precursor of inorganic network. Polym. Test. 2016, 52, 94–103. [Google Scholar] [CrossRef]
  39. Sava, I.; Damaceanu, M.-D.; Constantin, C.-P.; Asandulesa, M.; Wolińska-Grabczyk, A.; Jankowski, A. Structure–promoted high performance properties of triphenylmethane-containing polyimides and copolyimides. Eur. Polym. J. 2018, 108, 554–569. [Google Scholar] [CrossRef]
  40. Havriliak, S.; Negami, S. A complex plane representation of dielectric and mechanical relaxation processes in some polymers. Polymer 1967, 8, 161–210. [Google Scholar] [CrossRef]
Figure 1. Preparation of magnetic polyimide composites.
Figure 1. Preparation of magnetic polyimide composites.
Polymers 13 01646 g001
Figure 2. SEM micrographs of PI-0, PI-A-10, PI-A-30, and PI-A-40.
Figure 2. SEM micrographs of PI-0, PI-A-10, PI-A-30, and PI-A-40.
Polymers 13 01646 g002
Figure 3. XRD patterns of PI-0, PI-A-10, PI-A-20, PI-A-30, and PI-A-40 (a) and PI-A-30, PI-B-30, PI-C-30, and PI-D-30 (b).
Figure 3. XRD patterns of PI-0, PI-A-10, PI-A-20, PI-A-30, and PI-A-40 (a) and PI-A-30, PI-B-30, PI-C-30, and PI-D-30 (b).
Polymers 13 01646 g003
Figure 4. Hysteresis curves of PI-A-10, PI-A-20, PI-A-30, and PI-A-40, ±10 kOe representation—main figure, ±1.5 kOe—bottom right plot (a) and of PI-A-30, PI-B-30, PI-C-30, and PI-D-30, ±10 kOe representation—main figure, ±1.5 kOe—bottom right plot (b).
Figure 4. Hysteresis curves of PI-A-10, PI-A-20, PI-A-30, and PI-A-40, ±10 kOe representation—main figure, ±1.5 kOe—bottom right plot (a) and of PI-A-30, PI-B-30, PI-C-30, and PI-D-30, ±10 kOe representation—main figure, ±1.5 kOe—bottom right plot (b).
Polymers 13 01646 g004
Figure 5. Dielectric constant (a) and dielectric loss (b) evolution with alternating frequency for PI-0, PI-A-10, PI-A-20, and PI-A-30, at 25 °C.
Figure 5. Dielectric constant (a) and dielectric loss (b) evolution with alternating frequency for PI-0, PI-A-10, PI-A-20, and PI-A-30, at 25 °C.
Polymers 13 01646 g005
Figure 6. Dielectric constant (a) and dielectric loss (b) evolution with temperature for PI-0, PI-A-10, PI-A-20, and PI-A-30, at a frequency of 1 Hz.
Figure 6. Dielectric constant (a) and dielectric loss (b) evolution with temperature for PI-0, PI-A-10, PI-A-20, and PI-A-30, at a frequency of 1 Hz.
Polymers 13 01646 g006
Figure 7. Dielectric loss evolution with temperature at selected frequencies (a) and 3D presentation of ε″ as function of frequency and temperature (b) for PI-A-30.
Figure 7. Dielectric loss evolution with temperature at selected frequencies (a) and 3D presentation of ε″ as function of frequency and temperature (b) for PI-A-30.
Polymers 13 01646 g007
Figure 8. Broadening and skewing parameters as function of temperature for γ (a) and β (b) relaxations of PI-0, PI-A-10, PI-A-20, and PI-A-30.
Figure 8. Broadening and skewing parameters as function of temperature for γ (a) and β (b) relaxations of PI-0, PI-A-10, PI-A-20, and PI-A-30.
Polymers 13 01646 g008
Figure 9. Arrhenius plots of γ- and β- relaxations.
Figure 9. Arrhenius plots of γ- and β- relaxations.
Polymers 13 01646 g009
Figure 10. The frequency evolution of measured conductivity at 25 °C (a) and 250 °C (b) for PI-0, PI-A-10, PI-A-20, PI-A-30, and PI-A-40.
Figure 10. The frequency evolution of measured conductivity at 25 °C (a) and 250 °C (b) for PI-0, PI-A-10, PI-A-20, PI-A-30, and PI-A-40.
Polymers 13 01646 g010
Table 1. Magnetic values of Ni1-xCuxFe1.92Ce0.08O4 ferrite particles.
Table 1. Magnetic values of Ni1-xCuxFe1.92Ce0.08O4 ferrite particles.
Ferrite SampleChemical FormulaMs
(emu g−1)
Mr
(emu g−1)
Hc
(Oe)
A: [x = 0.3]Ni0.7 Cu0.3 Fe1.92 Ce0.08 O4 25.43.3453.45
B: [x = 0.0]NiFe1.92 Ce0.08 O4 33.865.2365.75
C: [x = 0.6]Ni0.4Cu0.6 Fe1.92 Ce0.08 O4 20.771.7332.32
D: [x = 1.0]CuFe1.92 Ce0.08 O4 19.446.47246.2
Table 2. Type and concentration of ferrite particles for each composite sample.
Table 2. Type and concentration of ferrite particles for each composite sample.
Polyimide CompositeType of Ferrite ParticlesConcentration of Particles in Polyimide Composite
(wt%)
PI-0
PI-A-10Ni0.7 Cu0.3 Fe1.92 Ce0.08 O4 10
PI-A-20Ni0.7 Cu0.3 Fe1.92 Ce0.08 O4 20
PI-A-30Ni0.7 Cu0.3 Fe1.92 Ce0.08 O4 30
PI-A-40Ni0.7 Cu0.3 Fe1.92 Ce0.08 O4 40
PI-B-30NiFe1.92 Ce0.08 O4 30
PI-C-30Ni0.4Cu0.6 Fe1.92 Ce0.08 O4 30
PI-D-30CuFe1.92 Ce0.08 O4 30
Table 3. The main thermal and magnetic characteristics of the samples.
Table 3. The main thermal and magnetic characteristics of the samples.
SampleT5a
(°C)
T10b
(°C)
Tmax c
(°C)
Char Yield at
900 °C
(%)
Tgd
(°C)
Ms
(emu g−1)
Mr
(emu g−1)
Hc
(Oe)
PI-050952252745241---
PI-A-10507523522262342.370.4584
PI-A-20506520515432355.521.1190
PI-A-30501513509532349.051.7679
PI-A-404985115114923310.281.9883
PI-B-305005125124223510.902.84119
PI-C-30502515513472345.530.6332
PI-D-30495508510482338.022.61244
a Initial decomposition temperature = temperature of 5% weight loss; b temperature of 10% weight loss; c temperature of maximum polymer decomposition, d glass transition temperature (the inflection point).
Table 4. Dielectric constant, dielectric loss, and conductivity values, at 25 °C and selected frequencies.
Table 4. Dielectric constant, dielectric loss, and conductivity values, at 25 °C and selected frequencies.
SampleDielectric ConstantDielectric LossConductivity (S cm−1)
1 Hz104 Hz1 Hz104 Hz1 Hz104 Hz
PI-03.03.01.6 × 10−20.5 × 10−28.8 × 10−152.9 × 10−11
PI-A-103.33.14.1 × 10−22.7 × 10−22.3 × 10−141.5 × 10−10
PI-A-203.52.913.9 × 10−27.9 × 10−27.8 × 10−144.4 × 10−10
PI-A-303.93.119.6 × 10−211.5 × 10−21.1 × 10−136.4 × 10−10
PI-B-304.33.619.7 × 10−213.3 × 10−21.1 × 10−137.4 × 10−10
PI-C-303.43.12.7 × 10−210.8 × 10−21.5 × 10−146.0 × 10−10
PI-D-303.73.118.2 × 10−214.8 × 10−21.0 × 10−138.2 × 10−10
Table 5. The numerical values of Equation (3) for the samples.
Table 5. The numerical values of Equation (3) for the samples.
Sampleγ-Relaxationβ-Relaxation
τ0 (s) (kJ mol−1)τ0 (s)Eβ (kJ mol−1)
PI-06.0 × 10−1543.33.9 × 10−1253.1
PI-A-101.1 × 10−1340.31.4 × 10−1157.7
PI-A-206.1 × 10−1030.01.7 × 10−1159.2
PI-A-301.9 × 10−1031.61.1 × 10−1266.3
PI-B-301.8 × 10−1129.0
PI-C-308.7 × 10−1027.0
PI-D-301.9 × 10−1027.3
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hamciuc, C.; Asandulesa, M.; Hamciuc, E.; Roman, T.; Olariu, M.A.; Pui, A. Novel Polyimide/Copper-Nickel Ferrite Composites with Tunable Magnetic and Dielectric Properties. Polymers 2021, 13, 1646. https://doi.org/10.3390/polym13101646

AMA Style

Hamciuc C, Asandulesa M, Hamciuc E, Roman T, Olariu MA, Pui A. Novel Polyimide/Copper-Nickel Ferrite Composites with Tunable Magnetic and Dielectric Properties. Polymers. 2021; 13(10):1646. https://doi.org/10.3390/polym13101646

Chicago/Turabian Style

Hamciuc, Corneliu, Mihai Asandulesa, Elena Hamciuc, Tiberiu Roman, Marius Andrei Olariu, and Aurel Pui. 2021. "Novel Polyimide/Copper-Nickel Ferrite Composites with Tunable Magnetic and Dielectric Properties" Polymers 13, no. 10: 1646. https://doi.org/10.3390/polym13101646

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