Next Article in Journal
Valorisation of Waste Glasses for the Development of Geopolymer Mortar—Properties and Applications: An Appraisal
Next Article in Special Issue
Static and Fatigue Debond Resistance between the Composite Facesheet and Al Cores under Mode-1 in Sandwich Beams
Previous Article in Journal
Preparation of a Photosensitive Composite Carbon Fiber for Spilled Oil Cleaning
Previous Article in Special Issue
Reducing the Weakening Effect in Fibre-Reinforced Polymers Caused by Integrated Film Sensors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Cure Kinetics of Samarium-Doped Fe3O4/Epoxy Nanocomposites

by
Maryam Jouyandeh
1,
Mohammad Reza Ganjali
1,2,3,
Mehdi Mehrpooya
4,
Otman Abida
5,
Karam Jabbour
5,
Navid Rabiee
6,
Sajjad Habibzadeh
7,
Amin Hamed Mashahdzadeh
8,
Alberto García-Peñas
9,10,
Florian J. Stadler
9,* and
Mohammad Reza Saeb
11,*
1
Center of Excellence in Electrochemistry, School of Chemistry, University of Tehran, Tehran 14176-14411, Iran
2
School of Resources and Environment, University of Electronic Science and Technology of China, Chengdu 611731, China
3
Biosensor Research Center, Endocrinology and Metabolism Molecular-Cellular Sciences Institute, Tehran University of Medical Sciences, Tehran 14117-13137, Iran
4
Department of Renewable Energies and Environment, Faculty of New Sciences and Technologies, University of Tehran, Tehran 14155-6619, Iran
5
College of Engineering and Technology, American University of the Middle East, P.O. Box 220, Dasman 15453, Kuwait
6
Department of Physics, Sharif University of Technology, Tehran 11155-9161, Iran
7
Department of Chemical Engineering, Amirkabir University of Technology (Tehran Polytechnic), Tehran 15916-34311, Iran
8
Mechanical and Aerospace Engineering, School of Engineering and Digital Sciences, Nazarbayev University, Nur-Sultan 010000, Kazakhstan
9
College of Materials Science and Engineering, Guangdong Research Center for Interfacial Engineering of Functional Materials, Nanshan District Key Laboratory for Biopolymers and Safety Evaluation, Shenzhen Key Laboratory of Polymer Science and Technology, Lihu Campus, Shenzhen University, Shenzhen 518055, China
10
Departamento de Ciencia e Ingeniería de Materiales e Ingeniería Química (IAAB), Universidad Carlos III de Madrid, 28911 Leganés, Spain
11
Department of Polymer Technology, Gdańsk University of Technology, G. Narutowicza 11/12, 80-233 Gdańsk, Poland
*
Authors to whom correspondence should be addressed.
J. Compos. Sci. 2022, 6(1), 29; https://doi.org/10.3390/jcs6010029
Submission received: 6 December 2021 / Revised: 5 January 2022 / Accepted: 8 January 2022 / Published: 17 January 2022
(This article belongs to the Special Issue Polymer Composites and Fibers)

Abstract

:
To answer the question “How does lanthanide doping in iron oxide affect cure kinetics of epoxy-based nanocomposites?”, we synthesized samarium (Sm)-doped Fe3O4 nanoparticles electrochemically and characterized it using Fourier-transform infrared spectroscopy (FTIR), X-ray powder diffraction (XRD), field emission scanning electron microscopy (FE-SEM), energy dispersive X-Ray analysis (EDX), transmission electron microscopy (TEM), and X-ray photoelectron spectroscopy analyses (XPS). The magnetic particles were uniformly dispersed in epoxy resin to increase the curability of the epoxy/amine system. The effect of the lanthanide dopant on the curing reaction of epoxy with amine was explored by analyzing differential scanning calorimetry (DSC) experimental data based on a model-free methodology. It was found that Sm3+ in the structure of Fe3O4 crystal participates in cross-linking epoxy by catalyzing the reaction between epoxide rings and amine groups of curing agents. In addition, the etherification reaction of active OH groups on the surface of nanoparticles reacts with epoxy rings, which prolong the reaction time at the late stage of reaction where diffusion is the dominant mechanism.

1. Introduction

The subject of magnetic nanoparticles has attracted significant interest, in particular, in the medical field and other high-tech applications [1,2,3]. Magnetite nanoparticle-filled polymers have been widely used in various applications such as electronic devices, nonlinear optic systems, sensors, magnetic filters, and photovoltaic solar cells [4]. Owing to their good properties and low price, magnetite (Fe3O4) nanoparticles have been studied most among the various kinds of magnetic nanoparticles [5,6,7].
Fe3O4 nanoparticles have been used in epoxy matrices to improve their final properties [8]. Dispersion state and interfacial interaction between Fe3O4 nanoparticles and epoxy matrix are two important parameters that affect the cross-linking reaction and final properties of epoxy nanocomposites [9,10]. Homogenous dispersion of nanoparticles in an epoxy matrix and strong interfacial interaction between Fe3O4 nanoparticles and polymer matrix are required to achieve a dense crosslinked network with improved final properties, in particular, in terms of mechanical performance [11,12]. However, magnetite nanoparticles have a high tendency to aggregate in the epoxy matrix [13]. It was shown that the success in curing the reaction of epoxy is highly dependent on the dispersibility of Fe3O4 nanoparticles in the resin matrix. However, the possibility of unmodified Fe3O4 nanoparticles to form agglomerates is high, hindering cross-linking reactions [14]. Many research works revealed that surface functionalization of Fe3O4 nanoparticles results in well-dispersed nanoparticles in the epoxy matrix [15,16]. Recently, it was shown that bulk modification of Fe3O4 nanoparticles by doping other metal ions can strongly affect cross-linking reaction of epoxy [17,18,19]. Furthermore, with an eye to possibly to enhance the properties of the Fe3O4 nanoparticles, doping of magnetite nanoparticles with other metal ions has been explored [20,21].
Lanthanides feature an f-electron shell, a property that only very heavy atoms have, which gives them unique optical and magnetic properties, making them interesting as dopants for other nanoparticles [22]. Besides introducing new properties, the doping of magnetite particles with small amounts of lanthanides leads to only minor alterations of physical and chemical characteristics. The doping of other metals/metal ions, such as zinc, manganese, copper, nickel, and cobalt, into the Fe3O4 nanoparticles enhances the availability of its surface sites [23,24,25]. Among lanthanides, samarium (Sm) has high magnetic properties and is relatively stable at room temperature. This rare-earth element is used in many applications such as microwave, electronic devices, optical lasers, and neutron absorbers in nuclear reactors [26,27]. Doping Fe3O4 nanoparticles with Sm atoms can increase the number of surface active sites and surface area of base nanoparticles, which may enhance the curability of epoxy resin. To the best of our knowledge, the effect of Sm-doped Fe3O4 nanoparticles on the curing reaction of epoxy resin has not been studied until now.
In the present study, Sm-doped Fe3O4 nanoparticles were fabricated via an electrochemical method. The synthesized samples were well characterized by Fourier-transform infrared spectroscopy (FTIR), X-ray powder diffraction (XRD), field emission scanning electron microscopy (FE-SEM), energy dispersive X-ray analysis (EDX), transmission electron microscopy (TEM), and X-ray photoelectron spectroscopy analyses (XPS). Then, the epoxy-based film was reinforced with Sm-doped Fe3O4 nanoparticles to obtain an excellent corrosion protection coating. The cure potential of the epoxy-containing Sm-doped Fe3O4 nanoparticles was evaluated with dynamic differential scanning calorimetry (DSC) at different heating rates of 2.5, 5, 7.5, and 10 °C/min.

2. Materials and Methods

Iron (II) chloride (FeCl2·4H2O), iron(III) nitrate nonahydrate 99.9% (Fe(NO3)3·9H2O), and samarium(III) nitrate (Sm(NO3)3·(H2O)2) were supplied by Sigma-Aldrich. Araldite LY 5052 epoxy resin and HY 5052 curing agent were purchased from MIS Hindustan Ciba-Geigy.

2.1. Synthesis of Sm-Doped Fe3O4 Nanoparticles

Sm3+-doped Fe3O4 nanoparticles were prepared through the cathodic electrodeposition (CED) procedure using a stainless steel cathode (316 L, 5 cm × 5 cm × 0.5 mm) inside two graphite anodes. The electrolyte 0.005 molar solution of iron(III) nitrate nonahydrate (2 g), iron(II) chloride (1 g), and samarium(III) nitrate (0.6 g) was prepared in water. Then, deposition occurred using a Potentiostat/Galvanostat, Model: NCF-PGS 2012 (Metrohm Autolab) at 25 °C and current density of 10 mA cm−2 for 30 min followed by rinsing with deionized water several times. Finally, the dispersed Sm-Fe3O4 deposit in deionized water was centrifuged at 6000 rpm for 20 min, separated, and dried at 70 °C for 1 h.

2.2. Preparation of Epoxy/Sm-Doped Fe3O4 Nanocomposite

Epoxy nanocomposites were obtained by mixing 0.1 wt.% of Sm-doped Fe3O4 using a mechanical mixer at 2500 rpm for 15 min. Then, Sm-doped Fe3O4 in epoxy was further mixed by sonication for 5 min. Finally, the curing agent was added to EP/Sm-Fe3O4 nanocomposite in the stoichiometric ratio of 38/100 (curing agent/epoxy).

2.3. Characterization

The FTIR spectrum of Sm-doped Fe3O4 nanoparticles was obtained by Bruker Vector spectrometer, Coventry, UK, between 4000–400 cm−1 wavelength. X-ray diffraction (XRD) of nanoparticles was performed by a Philips PW-1800 apparatus (Amsterdam, Netherlands) with Co Kα radiation. The micro- and nano-images of Sm-Fe3O4 nanoparticles were obtained by FESEM and EDX-Mapping (Mira 3-XMU, TESCAN, Brno-Kohoutovice, Czech Republic) at the voltage of 100 kV and TEM (Zeiss-EM10C-80 kV, Jena, Germany). A Thermo Fisher Scientific instrument XPS elemental analyzer (Waltham, MA, USA) determined the X-ray Photoelectron Spectroscopic properties of the nanoparticles.
The cure reaction of neat epoxy and EP/Sm-Fe3O4 nanocomposite was investigated by analyzing 12 mg of each sample using DSC (Perkin Elmer, DSC 4000, Waltham, MA, USA) at four different heating rates (2.5, 5, 7.5 and 10 °C·min−1) in the temperature range of 15–300 °C under nitrogen atmosphere.

3. Results and Discussion

3.1. Characterization of Sm-Fe3O4 Nanoparticles

The FTIR spectrum of the prepared Sm–Fe3O4 nanoparticle is shown in Figure 1a. Two sharp bands can be observed at 562 cm−1 and 628 cm−1, which are ascribed to the splitting of the ν1 band of the Fe–O. A wide peak in the range of 415–443 cm−1 is due to ν2 bands of the Fe–O and Sm–O [26]. The appearance of bands at 1648 and 3325 cm−1 are attributed to the stretching and deformation vibrations of O–H groups on the surface of Sm–Fe3O4 nanoparticles.
Figure 1b shows the XRD pattern of Sm-Fe3O4 nanoparticles, which is the cubic spinel structure of Magnetite Fe3O4 [Joint Committee on Powder Diffraction Standards (JCPDS) 76-1849 and Inorganic Crystal Structural Database (ICSD) 28664]. A significant change was not observed in the XRD patterns of doped nanoparticles. The XRD pattern shows that doping Sm does not change the crystal structure of Fe3O4 nanoparticles.
The oxygen atoms in magnetite (Fe3O4) form an inverse Spinel structure—a close-packed face-centered cubic sublattice, where Fe(II) are located in the octahedral sites and Fe(III) are occupying octahedral and tetrahedral sites [28], which can be described by (Fe83+)tetr[Fe3+Fe2+]8octO32. The close-packed plane is along the (111) axis, occupied by oxygen atoms. Those sites can also be occupied by the transition metal atoms [29]. Samarium ions, similarly to other lanthanide(III) ions, have six coordination sites and seven octahedral coordination sites, when crystallizing as Sm2O3 [30], making them suitable for being included in the inverse spinel structure at the octahedral sites.
Figure 2 shows the FESEM, EDS, mapping, and TEM images of Sm-Fe3O4 nanoparticles. The average particle size is about 20 nm. The EDX experiments proved the presence of nanoparticles, containing both Fe and Sm at approximately equal amounts (Figure 2b). Additionally, O (25.4%) and Au were found, the latter stemming from the TEM grids used.
The presence of Fe and Sm, O, and C in the Sm-Fe3O4 nanoparticles is confirmed by XPS analysis (Figure 3a), as the Fe2p peaks (Figure 3b) are clearly visible in terms of Fe2p1/2 (710 eV) and Fe2p3/2 peaks (720 eV), confirming the Fe(III)-oxidation state [31]. The Fe(II) in Fe3O4 is confirmed by the shoulder at 708 eV (Fe2p1/2 peak) and at 721 eV (Fe2p3/2 peak) [32].
Sm-3d5/2 regions of Sm-Fe3O4 nanoparticles are shown in Figure 3c. The binding energy values (3d5/2) of Sm(III) (1079 and 1107 eV) observed in the nanoparticles are in accordance with the standard values [Sm2O3 (1082 and 1108 eV) [33], thus this lanthanide is in its +3 oxidation state in Sm-doped Fe3O4. The varied coordination conditions in the crystal structure could explain the subtle differences observed.

3.2. Curing Analysis

Figure 4 displays nonisothermal DSC thermographs of neat epoxy and EP/Sm-Fe3O4 cured with a stoichiometric amount of amine curing agent at heating rates of 2.5, 5, 7.5, and 10 °C/min. One exothermic peak can be observed for both samples at different heating rates, which revealed that the presence of Sm-Fe3O4 nanoparticles in the epoxy matrix does not change the domination of the chemically controlled reaction mechanism [34,35].
The cure characteristics of EP and EP/Sm-Fe3O4 nanocomposite include TOnset, TEndset, Tp, ΔT, and ΔH, which are the onset, endset, the exothermal peak temperature, temperature interval, and the enthalpy of complete cure, respectively, and are reported in Table 1. TOnset, TEndset, and Tp shifted towards higher temperatures by increasing heating rates from 2.5 to 10 °C/min to compensate for reducing curing time [36,37].
The addition of Sm-Fe3O4 nanoparticles decreased TOnset and Tp of epoxy/amine reaction, indicating that Sm doped magnetic nanoparticles accelerate cross-linking reaction. The surface activity of Sm-Fe3O4 nanoparticles can ascribe this increment in the system’s reactivity due to the presence of Sm3+ in the crystal structure of nanoparticles that catalyze the reaction between epoxy and amine curing agents [23,38]. However, TEndset, ΔT increased for EP/Sm-Fe3O4 nanocomposite compared to neat epoxy, which means that at the late stage of cure reaction, the OH groups on the surface of nanoparticles participate in etherification reaction and prolong the cross-linking of epoxy reaction.
The effect of the etherification reaction of OH groups on the surface of Sm-Fe3O4 nanoparticles as well as the catalyzing effect of Sm3+, which acts as Lewis acid, increase total heat of cure (ΔH) of EP/Sm-Fe3O4 nanocomposite in comparison to neat epoxy (Figure 5) [39].
Figure 6 shows the conversion (α) of curing reaction as a function of temperature, which is obtained from Equation (1) [40]:
α = Δ H T Δ H ,
where ∆HT is the enthalpy of reaction at a specific temperature.
In the initial stage of the curing reaction, cross-linking occurs rapidly until reaching gel point under the control of chemical reaction between the epoxy ring and amine groups of curing agent. In contrast, cross-linking occurs slowly at the late stage of cure, where diffusion is dominant. Additionally, Sm-Fe3O4 nanoparticles accelerate cross-linking of epoxy after vitrification, which indicated an acceleration of diffusion mechanisms due to the presence of OH groups on the nanoparticle surface [41].

Cure State of EP/Sm-Fe3O4

The effect of curability of Sm-Fe3O4 nanoparticles in epoxy/amine system are specified by Cure Index [42,43,44]:
C I = Δ T * × Δ H * ,   Δ T * = Δ T nanocomposite Δ T Reference   a n d   Δ H * = Δ H nanocomposite Δ H Reference ,
T is temperature window, within which curing occurs, with subscripts of “nanocomposite” and “Reference” for nanocomposite and blank epoxy systems, respectively. Similarly, ΔH of such systems is defined. The terms with asterisks are dimensionless in each case. Good, Poor, and Excellent curing reactions of nanocomposites occur at CI > ∆H*, CI < ∆T*, and ∆T* < CI < ∆H*, respectively [45,46,47]. The addition of Sm-Fe3O4 nanoparticles in the epoxy matrix resulted in a Good cure reaction, which means that Sm3+ participates in cross-linking of epoxy by catalyzing the reaction between epoxide rings and amine groups of curing agents. In addition, the active OH groups on the surface of nanoparticles react with epoxy polar groups that increase both ∆T and ∆H and result in Good CI [48,49,50].

3.3. Cure Kinetics of EP/Sm-Fe3O4

3.3.1. Calculation of Activation Energy

Isoconversional model-free Friedman (differential) and Kissinger–Akahira–Sunose (KAS, integral) were employed to obtain the apparent activation energies (Eα) of the curing reactions [51,52].
By plotting ln [ β i ( d α / d T ) α , i ] vs. 1/Tα from Equation (3), the value of Eα is obtained from the slope of Figure 7a,b [53,54].
ln [ β i ( d α d T ) α , i ] α = ln [ f ( α ) A α ] E α R T α , i ,
Additionally, plotting l n ( β i / T α , i 2 ) vs. 1/Tα from Equation (4) for each α value gives a value for the Eα (Figure 7c,d).
ln [ β i T α , i 1.92 ( d α d T ) α , i ] α = Constant 1.0008 ( E α R T α , i ) ,
The apparent activation energy of neat epoxy and EP/Sm-Fe3O4 nanocomposite as a function of α based on both Friedman and KAS are shown in Figure 8. Eα was reduced for neat epoxy and its nanocomposite in α higher than 0.5 due to the participation of OH groups in epoxide ring-opening at a later stage of curing reaction revealing the autocatalytic mechanism of epoxy cure reaction [55,56]. The higher Eα values for EP/Sm-Fe3O4 nanocomposite compared to neat epoxy can be attributed to the higher viscosity of the epoxy system in the presence of Sm-Fe3O4 nanoparticles [57].

3.3.2. Determination of Reaction Model

Friedman Model

The Friedman method can be used to show the non-catalytic or autocatalytic reaction model using Equation (5). The plot of ln[Af(α)] vs. ln(1 − α) shows a maximum in α between 0.2 and 0.4 that suggests an autocatalytic reaction mechanism (Figure 9) [58].
l n [ A f ( α ) ] = l n ( d α d t ) + E R T = l n A + n l n ( 1 α ) ,

Malek Model

A more accurate Malek method was applied to determine the cross-linking reaction model through the shape and maximum points of y(α) and z(α) Malek master plots (Equations (6) and (7)) [59]. As can be observed in Figure 10 and Table 2, αm = Max (y(α)) is lower than αp = Max (z(α)) and αp (the conversion at the maximum point of DSC curves) is smaller than 0.632, which revealed that the two-parameter autocatalytic model could be used for both EP and EP/Sm-Fe3O4 nanocomposite [60,61].
y ( α ) = ( d α d t ) α e x p ( E 0 R T α ) = A f ( α )
z ( α ) = ( d α d t ) α T α 2 ,
Therefore, the autocatalytic reaction model (f(α), Equation (8)) of neat epoxy and its nanocomposite were obtained by Friedman and Malek models.
f ( α ) = α m ( 1 α ) n ,
The reaction model parameters, including the pre-exponential factor (lnA), non-catalytic (n), and autocatalytic (m) reaction orders, were determined from Equations (9) and (10) and Figure 11 and reported in Table 3.
V a l u e I = l n ( d α d t ) + E α R T l n [ d ( 1 α ) d t ] E α R T = ( n m ) l n ( 1 α α ) ,
V a l u e II = l n ( d α d t ) + E α R T + l n [ d ( 1 α ) d t ] + E α R T = ( n + m ) l n ( α α 2 ) + 2 l n A ,
As can be observed, both n and m increased for EP/Sm-Fe3O4 nanocomposite compared to neat epoxy. An increment of n indicated the catalyzing effect of Sm3+ as a Lewis acid in the reaction between the epoxy ring and amine curing agent, and the enhancement of m is because of the reaction of OH groups on the surface of the Sm-Fe3O4 nanoparticles with epoxide rings. At a low heating rate (2.5 °C/min), the curing moieties have sufficient time to take part in cross-linking reactions at the early stage of curing reactions through chemically controlled reactions. At a high heating rate of 10 °C/min, the kinetic energy of the curing moieties is sufficiently high to facilitate a chemically control reaction as confirmed by a higher value of n. However, at medium heating rates (β = 5 and 7.5 °C/min), the n value may be decreased because of the absence of high kinetic energy and appropriate reaction time. The retardation effect of Sm-Fe3O4 nanoparticles on the cure reaction of epoxy is reflected in higher lnA values.

Validation of Isoconversional Methods

The validation of isoconversional methods (Friedman and KAS) is obtained by comparison with the experimental data and shown in Figure 12. Clearly, both KAS and Friedman approaches can predict the curing rate of the cross-linking reaction for neat epoxy and Sm-Fe3O4 nanoparticles incorporated epoxy systems.

4. Conclusions

Sm-doped Fe3O4 nanoparticle was synthesized through an electrochemical method to investigate their effect on the curability of epoxy/amine system. XPS results indicated that samarium is present in its +3 oxidation states in the structure of Fe3O4 lattice. The XRD pattern revealed that Sm3+ ions occupy the octahedral sites in the Fe3O4 crystal structure. DSC analysis at a different heating rate was performed to indicate Sm doping effect on the curing reaction of EP/Sm-Fe3O4 nanocomposites. It was found that the addition of Sm-Fe3O4 nanoparticles accelerates cross-linking reaction due to the catalyzing effect of Sm3+ in the crystal structure of Fe3O4 nanoparticles on the reaction between epoxy and amine curing agent, which is reflected in lower TOnset and Tp. Obtaining Good CI by addition of Sm-Fe3O4 nanoparticles in epoxy matrix showed that Sm3+ participate in cross-linking of epoxy by catalyzing the reaction between epoxide rings and amine groups of the curing agent, and the etherification reaction of active OH groups on the surface of nanoparticles react with epoxy rings. The apparent activation energy determined by isoconversional Friedman and KAS methods indicated a complex curing reaction of epoxy in the presence of Sm-Fe3O4 nanoparticles, which caused an increase in average Eα value from 47.3 for neat epoxy to 52.6 kJ/mol. The autocatalytic reaction model was validated by experimental data. It can be concluded that doping Fe3O4 nanoparticles with Sm rare-earth element can enhance the surface active site of parent magnetic nanoparticles and enhance their dispersibility in the epoxy matrix and facilitate the cross-linking reaction of epoxy.

Author Contributions

Conceptualization, M.R.S.; methodology, M.J. and M.R.G.; software, M.J.; validation, M.J., and M.R.S.; formal analysis, M.J., M.M., and O.A.; investigation, K.J. and N.R.; data curation, S.H. and A.H.M.; writing—original draft preparation, M.J.; writing—review and editing, A.G.-P., F.J.S. and M.R.S.; visualization, M.J. and M.R.S.; supervision, M.R.S.; project administration, M.R.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are available upon request from the authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yavuz, C.T.; Mayo, J.; William, W.Y.; Prakash, A.; Falkner, J.C.; Yean, S.; Cong, L.; Shipley, H.J.; Kan, A.; Tomson, M. Low-field magnetic separation of monodisperse Fe3O4 nanocrystals. Science 2006, 314, 964–967. [Google Scholar] [CrossRef]
  2. Samal, S.; Kolinova, M.; Blanco, I. The magneto-mechanical behavior of active components in iron-elastomer composite. J. Compos. Sci. 2018, 2, 54. [Google Scholar] [CrossRef] [Green Version]
  3. Yesmin, N.; Chalivendra, V. Electromagnetic Shielding Effectiveness of Glass Fiber/Epoxy Laminated Composites with Multi-Scale Reinforcements. J. Compos. Sci. 2021, 5, 204. [Google Scholar] [CrossRef]
  4. Long, Y.; Chen, Z.; Duvail, J.L.; Zhang, Z.; Wan, M. Electrical and magnetic properties of polyaniline/Fe3O4 nanostructures. Phys. B Condens. Matter 2005, 370, 121–130. [Google Scholar] [CrossRef]
  5. Roca, A.; Morales, M.; O’Grady, K.; Serna, C. Structural and magnetic properties of uniform magnetite nanoparticles prepared by high temperature decomposition of organic precursors. Nanotechnology 2006, 17, 2783. [Google Scholar] [CrossRef]
  6. Samal, S.; Škodová, M.; Blanco, I. Effects of filler distribution on magnetorheological silicon-based composites. Materials 2019, 12, 3017. [Google Scholar] [CrossRef] [Green Version]
  7. Bayat, M.; Yang, H.; Ko, F. Effect of iron oxide nanoparticle size on electromagnetic properties of composite nanofibers. J. Compos. Mater. 2018, 52, 1723–1736. [Google Scholar] [CrossRef]
  8. Yao, D.; Peng, N.; Zheng, Y. Enhanced mechanical and thermal performances of epoxy resin by oriented solvent-free graphene/carbon nanotube/Fe3O4 composite nanofluid. Compos. Sci. Technol. 2018, 167, 234–242. [Google Scholar] [CrossRef]
  9. Pathak, A.K.; Borah, M.; Gupta, A.; Yokozeki, T.; Dhakate, S.R. Improved mechanical properties of carbon fiber/graphene oxide-epoxy hybrid composites. Compos. Sci. Technol. 2016, 135, 28–38. [Google Scholar] [CrossRef]
  10. Alexopoulos, N.D.; Paragkamian, Z.; Poulin, P.; Kourkoulis, S.K. Fracture related mechanical properties of low and high graphene reinforcement of epoxy nanocomposites. Compos. Sci. Technol. 2017, 150, 194–204. [Google Scholar] [CrossRef]
  11. Yao, H.; Hawkins, S.A.; Sue, H.-J. Preparation of epoxy nanocomposites containing well-dispersed graphene nanosheets. Compos. Sci. Technol. 2017, 146, 161–168. [Google Scholar] [CrossRef]
  12. Liu, X.; Xu, F.; Zhang, K.; Wei, B.; Gao, Z.; Qiu, Y. Characterization of enhanced interfacial bonding between epoxy and plasma functionalized carbon nanotube films. Compos. Sci. Technol. 2017, 145, 114–121. [Google Scholar] [CrossRef]
  13. Sangermano, M.; Allia, P.; Tiberto, P.; Barrera, G.; Bondioli, F.; Florini, N.; Messori, M. Photo-Cured Epoxy Networks Functionalized With Fe3O4 Generated by Non-hydrolytic Sol–Gel Process. Macromol. Chem. Phys. 2013, 214, 508–516. [Google Scholar] [CrossRef]
  14. Jouyandeh, M.; Rahmati, N.; Movahedifar, E.; Hadavand, B.S.; Karami, Z.; Ghaffari, M.; Taheri, P.; Bakhshandeh, E.; Vahabi, H.; Ganjali, M.R. Properties of nano-Fe3O4 incorporated epoxy coatings from Cure Index perspective. Prog. Org. Coat. 2019, 133, 220–228. [Google Scholar] [CrossRef]
  15. Lakouraj, M.M.; Rahpaima, G.; Zare, E.N. Effect of functionalized magnetite nanoparticles and diaminoxanthone on the curing, thermal degradation kinetic and corrosion property of diglycidyl ether of bisphenol A-based epoxy resin. Chin. J. Polym. Sci. 2014, 32, 1489–1499. [Google Scholar] [CrossRef]
  16. Nguyen, T.V.; Do, T.V.; Ha, M.H.; Le, H.K.; Le, T.T.; Nguyen, T.N.L.; Dam, X.T.; Vu, Q.T.; Dinh, D.A.; Dang, T.C. Crosslinking process, mechanical and antibacterial properties of UV-curable acrylate/Fe3O4-Ag nanocomposite coating. Prog. Org. Coat. 2020, 139, 105325. [Google Scholar] [CrossRef]
  17. Jouyandeh, M.; Paran, S.M.R.; Khadem, S.S.M.; Ganjali, M.R.; Akbari, V.; Vahabi, H.; Saeb, M.R. Nonisothermal cure kinetics of epoxy/MnxFe3-xO4 nanocomposites. Prog. Org. Coat. 2020, 140, 105505. [Google Scholar] [CrossRef]
  18. Jouyandeh, M.; Karami, Z.; Hamad, S.M.; Ganjali, M.R.; Akbari, V.; Vahabi, H.; Kim, S.-J.; Zarrintaj, P.; Saeb, M.R. Nonisothermal cure kinetics of epoxy/ZnxFe3-xO4 nanocomposites. Prog. Org. Coat. 2019, 136, 105290. [Google Scholar] [CrossRef]
  19. Jouyandeh, M.; Ganjali, M.R.; Ali, J.A.; Akbari, V.; Karami, Z.; Aghazadeh, M.; Zarrintaj, P.; Saeb, M.R. Curing epoxy with polyethylene glycol (PEG) surface-functionalized GdxFe3-xO4 magnetic nanoparticles. Prog. Org. Coat. 2019, 137, 105283. [Google Scholar] [CrossRef]
  20. Liu, J.; Bin, Y.; Matsuo, M. Magnetic behavior of Zn-doped Fe3O4 nanoparticles estimated in terms of crystal domain size. J. Phys. Chem. C 2012, 116, 134–143. [Google Scholar] [CrossRef]
  21. Guo, M.; Balamurugan, J.; Li, X.; Kim, N.H.; Lee, J.H. Hierarchical 3D cobalt-doped Fe3O4 nanospheres@ NG hybrid as an advanced anode material for high-performance asymmetric supercapacitors. Small 2017, 13, 1701275. [Google Scholar] [CrossRef] [PubMed]
  22. Groman, E.V.; Bouchard, J.C.; Reinhardt, C.P.; Vaccaro, D.E. Ultrasmall mixed ferrite colloids as multidimensional magnetic resonance imaging, cell labeling, and cell sorting agents. Bioconjug. Chem. 2007, 18, 1763–1771. [Google Scholar] [CrossRef]
  23. Jouyandeh, M.; Ali, J.A.; Aghazadeh, M.; Formela, K.; Saeb, M.R.; Ranjbar, Z.; Ganjali, M.R. Curing epoxy with electrochemically synthesized ZnxFe3-xO4 magnetic nanoparticles. Prog. Org. Coat. 2019, 136, 105246. [Google Scholar] [CrossRef]
  24. Jouyandeh, M.; Ali, J.A.; Akbari, V.; Aghazadeh, M.; Paran, S.M.R.; Naderi, G.; Saeb, M.R.; Ranjbar, Z.; Ganjali, M.R. Curing epoxy with polyvinylpyrrolidone (PVP) surface-functionalized MnxFe3-xO4 magnetic nanoparticles. Prog. Org. Coat. 2019, 136, 105247. [Google Scholar] [CrossRef]
  25. Jouyandeh, M.; Ganjali, M.R.; Ali, J.A.; Aghazadeh, M.; Karimzadeh, I.; Formela, K.; Colom, X.; Cañavate, J.; Saeb, M.R. Curing epoxy with ethylenediaminetetraacetic acid (EDTA) surface-functionalized CoxFe3-xO4 magnetic nanoparticles. Prog. Org. Coat. 2019, 136, 105248. [Google Scholar] [CrossRef]
  26. Aghazadeh, M.; Ganjali, M.R. Samarium-doped Fe3O4 nanoparticles with improved magnetic and supercapacitive performance: A novel preparation strategy and characterization. J. Mater. Sci. 2018, 53, 295–308. [Google Scholar] [CrossRef]
  27. Hoz, S. Samarium Iodide Showcase: Unraveling the Mechanistic Puzzle. Acc. Chem. Res. 2020, 53, 2680–2691. [Google Scholar] [CrossRef]
  28. Ketteler, G.; Weiss, W.; Ranke, W.; Schlögl, R. Bulk and surface phases of iron oxides in an oxygen and water atmosphere at low pressure. Phys. Chem. Chem. Phys. 2001, 3, 1114–1122. [Google Scholar] [CrossRef]
  29. Kim, T.Y.; Lee, M.S.; Kim, Y.I.; Lee, C.-S.; Park, J.C.; Kim, D. The enhanced anisotropic properties of the Fe3− xMxO4 (M= Fe, Co, Mn) films deposited on glass surface from aqueous solutions at low temperature. J. Phys. D Appl. Phys. 2003, 36, 1451. [Google Scholar] [CrossRef]
  30. Cotton, S. Lanthanide and Actinide Chemistry; John Wiley & Sons: Hoboken, NJ, USA, 2013. [Google Scholar]
  31. Kataby, G.; Ulman, A.; Cojocaru, M.; Gedanken, A. Coating a bola-amphiphile on amorphous iron nanoparticles. J. Mater. Chem. 1999, 9, 1501–1506. [Google Scholar] [CrossRef]
  32. Tie, S.-L.; Lee, H.-C.; Bae, Y.-S.; Kim, M.-B.; Lee, K.; Lee, C.-H. Monodisperse Fe3O4/Fe@ SiO2 core/shell nanoparticles with enhanced magnetic property. Colloids Surf. A Physicochem. Eng. Asp. 2007, 293, 278–285. [Google Scholar] [CrossRef]
  33. De Silva, C.R.; Smith, S.; Shim, I.; Pyun, J.; Gutu, T.; Jiao, J.; Zheng, Z. Lanthanide (III)-doped magnetite nanoparticles. J. Am. Chem. Soc. 2009, 131, 6336–6337. [Google Scholar] [CrossRef]
  34. Jouyandeh, M.; Tikhani, F.; Shabanian, M.; Movahedi, F.; Moghari, S.; Akbari, V.; Gabrion, X.; Laheurte, P.; Vahabi, H.; Saeb, M.R. Synthesis, characterization, and high potential of 3D metal–organic framework (MOF) nanoparticles for curing with epoxy. J. Alloys Compd. 2020, 829, 154547. [Google Scholar] [CrossRef]
  35. Karami, Z.; Jouyandeh, M.; Ali, J.A.; Ganjali, M.R.; Aghazadeh, M.; Maadani, M.; Rallini, M.; Luzi, F.; Torre, L.; Puglia, D.; et al. Cure Index for labeling curing potential of epoxy/LDH nanocomposites: A case study on nitrate anion intercalated Ni-Al-LDH. Prog. Org. Coat. 2019, 136, 105228. [Google Scholar] [CrossRef]
  36. Jouyandeh, M.; Ganjali, M.R.; Ali, J.A.; Aghazadeh, M.; Stadler, F.J.; Saeb, M.R. Curing epoxy with electrochemically synthesized NixFe3-xO4 magnetic nanoparticles. Prog. Org. Coat. 2019, 136, 105198. [Google Scholar] [CrossRef]
  37. Jouyandeh, M.; Zarrintaj, P.; Ganjali, M.R.; Ali, J.A.; Karimzadeh, I.; Aghazadeh, M.; Ghaffari, M.; Saeb, M.R. Curing epoxy with electrochemically synthesized GdxFe3-xO4 magnetic nanoparticles. Prog. Org. Coat. 2019, 136, 105245. [Google Scholar] [CrossRef]
  38. Jouyandeh, M.; Karami, Z.; Ali, J.A.; Karimzadeh, I.; Aghazadeh, M.; Laoutid, F.; Vahabi, H.; Saeb, M.R.; Ganjali, M.R.; Dubois, P. Curing epoxy with polyethylene glycol (PEG) surface-functionalized NixFe3-xO4magnetic nanoparticles. Prog. Org. Coat. 2019, 136, 105250. [Google Scholar] [CrossRef]
  39. Jouyandeh, M.; Ganjali, M.R.; Ali, J.A.; Aghazadeh, M.; Stadler, F.J.; Saeb, M.R. Curing epoxy with electrochemically synthesized MnxFe3-xO4 magnetic nanoparticles. Prog. Org. Coat. 2019, 136, 105199. [Google Scholar] [CrossRef]
  40. Akbari, V.; Jouyandeh, M.; Paran, S.M.R.; Ganjali, M.R.; Abdollahi, H.; Vahabi, H.; Ahmadi, Z.; Formela, K.; Esmaeili, A.; Mohaddespour, A.; et al. Effect of Surface Treatment of Halloysite Nanotubes (HNTs) on the Kinetics of Epoxy Resin Cure with Amines. Polymers 2020, 12, 930. [Google Scholar] [CrossRef] [Green Version]
  41. Tikhani, F.; Moghari, S.; Jouyandeh, M.; Laoutid, F.; Vahabi, H.; Saeb, M.R.; Dubois, P. Curing Kinetics and Thermal Stability of Epoxy Composites Containing Newly Obtained Nano-Scale Aluminum Hypophosphite (AlPO2). Polymers 2020, 12, 644. [Google Scholar] [CrossRef] [Green Version]
  42. Jouyandeh, M.; Paran, S.M.R.; Jannesari, A.; Saeb, M.R. ‘Cure Index’ for thermoset composites. Prog. Org. Coat. 2019, 127, 429–434. [Google Scholar] [CrossRef]
  43. Jouyandeh, M.; Ganjali, M.R.; Ali, J.A.; Aghazadeh, M.; Paran, S.M.R.; Naderi, G.; Saeb, M.R.; Thomas, S. Curing epoxy with polyvinylpyrrolidone (PVP) surface-functionalized ZnxFe3-xO4 magnetic nanoparticles. Prog. Org. Coat. 2019, 136, 105227. [Google Scholar] [CrossRef]
  44. Karami, Z.; Jouyandeh, M.; Ali, J.A.; Ganjali, M.R.; Aghazadeh, M.; Maadani, M.; Rallini, M.; Luzi, F.; Torre, L.; Puglia, D.; et al. Development of Mg-Zn-Al-CO3 ternary LDH and its curability in epoxy/amine system. Prog. Org. Coat. 2019, 136, 105264. [Google Scholar] [CrossRef]
  45. Karami, Z.; Jouyandeh, M.; Hamad, S.M.; Ganjali, M.R.; Aghazadeh, M.; Torre, L.; Puglia, D.; Saeb, M.R. Curing epoxy with Mg-Al LDH nanoplatelets intercalated with carbonate ion. Prog. Org. Coat. 2019, 136, 105278. [Google Scholar] [CrossRef]
  46. Seidi, F.; Jouyandeh, M.; Akbari, V.; Paran, S.M.R.; Livi, S.; Ducos, F.; Vahabi, H.; Ganjali, M.R.; Saeb, M.R. Super-crosslinked ionic liquid-intercalated montmorillonite/epoxy nanocomposites: Cure kinetics, viscoelastic behavior and thermal degradation mechanism. Polym. Eng. Sci. 2020, 60, 1940–1957. [Google Scholar] [CrossRef]
  47. Karami, Z.; Aghazadeh, M.; Jouyandeh, M.; Zarrintaj, P.; Vahabi, H.; Ganjali, M.R.; Torre, L.; Puglia, D.; Saeb, M.R. Epoxy/Zn-Al-CO3 LDH nanocomposites: Curability assessment. Prog. Org. Coat. 2020, 138, 105355. [Google Scholar] [CrossRef]
  48. Jouyandeh, M.; Ganjali, M.R.; Ali, J.A.; Aghazadeh, M.; Saeb, M.R.; Ray, S.S. Curing epoxy with polyvinylpyrrolidone (PVP) surface-functionalized NixFe3-xO4 magnetic nanoparticles. Prog. Org. Coat. 2019, 136, 105259. [Google Scholar] [CrossRef]
  49. Seidi, F.; Jouyandeh, M.; Paran, S.M.R.; Esmaeili, A.; Karami, Z.; Livi, S.; Habibzadeh, S.; Vahabi, H.; Ganjali, M.R.; Saeb, M.R. Imidazole-functionalized nitrogen-rich Mg-Al-CO3 layered double hydroxide for developing highly crosslinkable epoxy with high thermal and mechanical properties. Colloids Surf. A Physicochem. Eng. Asp. 2021, 611, 125826. [Google Scholar] [CrossRef]
  50. Jouyandeh, M.; Hamad, S.M.; Karimzadeh, I.; Aghazadeh, M.; Karami, Z.; Akbari, V.; Shammiry, F.; Formela, K.; Saeb, M.R.; Ranjbar, Z.; et al. Unconditionally blue: Curing epoxy with polyethylene glycol (PEG) surface-functionalized ZnxFe3-xO4 magnetic nanoparticles. Prog. Org. Coat. 2019, 137, 105285. [Google Scholar] [CrossRef]
  51. Calvino, M.M.; Lisuzzo, L.; Cavallaro, G.; Lazzara, G.; Milioto, S. Non-isothermal thermogravimetry as an accelerated tool for the shelf-life prediction of paracetamol formulations. Thermochim. Acta 2021, 700, 178940. [Google Scholar] [CrossRef]
  52. Cavallaro, G.; Gallitto, A.A.; Lisuzzo, L.; Lazzara, G. Comparative study of historical woods from XIX century by thermogravimetry coupled with FTIR spectroscopy. Cellulose 2019, 26, 8853–8865. [Google Scholar] [CrossRef]
  53. Sbirrazzuoli, N.; Vyazovkin, S. Learning about epoxy cure mechanisms from isoconversional analysis of DSC data. Thermochim. Acta 2002, 388, 289–298. [Google Scholar] [CrossRef]
  54. Sbirrazzuoli, N.; Vyazovkin, S.; Mititelu, A.; Sladic, C.; Vincent, L. A study of epoxy-amine cure kinetics by combining isoconversional analysis with temperature modulated DSC and dynamic rheometry. Macromol. Chem. Phys. 2003, 204, 1815–1821. [Google Scholar] [CrossRef]
  55. Jouyandeh, M.; Jazani, O.M.; Navarchian, A.H.; Shabanian, M.; Vahabi, H.; Saeb, M.R. Surface engineering of nanoparticles with macromolecules for epoxy curing: Development of super-reactive nitrogen-rich nanosilica through surface chemistry manipulation. Appl. Surf. Sci. 2018, 447, 152–164. [Google Scholar] [CrossRef]
  56. Jouyandeh, M.; Jazani, O.M.; Navarchian, A.H.; Shabanian, M.; Vahabi, H.; Saeb, M.R. Bushy-surface hybrid nanoparticles for developing epoxy superadhesives. Appl. Surf. Sci. 2019, 479, 1148–1160. [Google Scholar] [CrossRef]
  57. Jouyandeh, M.; Paran, S.M.R.; Shabanian, M.; Ghiyasi, S.; Vahabi, H.; Badawi, M.; Formela, K.; Puglia, D.; Saeb, M.R. Curing behavior of epoxy/Fe3O4 nanocomposites: A comparison between the effects of bare Fe3O4, Fe3O4/SiO2/chitosan and Fe3O4/SiO2/chitosan/imide/phenylalanine-modified nanofillers. Prog. Org. Coat. 2018, 123, 10–19. [Google Scholar] [CrossRef]
  58. Karami, Z.; Ganjali, M.R.; Dehaghani, M.Z.; Aghazadeh, M.; Jouyandeh, M.; Esmaeili, A.; Habibzadeh, S.; Mohaddespour, A.; Inamuddin; Formela, K.; et al. Kinetics of Cross-Linking Reaction of Epoxy Resin with Hydroxyapatite-Functionalized Layered Double Hydroxides. Polymers 2020, 12, 1157. [Google Scholar] [CrossRef]
  59. Jouyandeh, M.; Ganjali, M.R.; Seidi, F.; Xiao, H.; Saeb, M.R. Nonisothermal Cure Kinetics of Epoxy/Polyvinylpyrrolidone Functionalized Superparamagnetic Nano-Fe3O4 Composites: Effect of Zn and Mn Doping. J. Compos. Sci. 2020, 4, 55. [Google Scholar] [CrossRef]
  60. Jouyandeh, M.; Karami, Z.; Paran, S.M.R.; Mashhadzadeh, A.H.; Ganjali, M.R.; Bagheri, B.; Zarrintaj, P.; Habibzadeh, S.; Vijayan P., P.; Saeb, M.R. Effect of Nickel Doping on the Cure Kinetics of Epoxy/Fe3O4 Nanocomposites. J. Compos. Sci. 2020, 4, 102. [Google Scholar] [CrossRef]
  61. Moghari, S.; Jafari, S.H.; Yazdi, M.K.; Jouyandeh, M.; Hejna, A.; Zarrintaj, P.; Saeb, M.R. In-Out Surface Modification of Halloysite Nanotubes (HNTs) for Excellent Cure of Epoxy: Chemistry and Kinetics Modeling. Nanomaterials 2021, 11, 3078. [Google Scholar] [CrossRef]
Figure 1. (a) FTIR spectra of Ce-doped Fe3O4 and (b) XRD pattern of Sm-doped Fe3O4.
Figure 1. (a) FTIR spectra of Ce-doped Fe3O4 and (b) XRD pattern of Sm-doped Fe3O4.
Jcs 06 00029 g001
Figure 2. (a) FESEM, (b) EDX, (c) elemental mapping and (d) TEM of Sm-doped Fe3O4.
Figure 2. (a) FESEM, (b) EDX, (c) elemental mapping and (d) TEM of Sm-doped Fe3O4.
Jcs 06 00029 g002
Figure 3. XPS spectra of the Sm-doped Fe3O4 (a) Survey, (b) Fe2p, and (c) Sm3d. Colored lines refer to the individual peak fits as explained in the text.
Figure 3. XPS spectra of the Sm-doped Fe3O4 (a) Survey, (b) Fe2p, and (c) Sm3d. Colored lines refer to the individual peak fits as explained in the text.
Jcs 06 00029 g003
Figure 4. Dynamic DSC thermograms of EP and EP/Sm-Fe3O4 at different heating rates.
Figure 4. Dynamic DSC thermograms of EP and EP/Sm-Fe3O4 at different heating rates.
Jcs 06 00029 g004
Figure 5. The schematic representation of synthesis Sm-Fe3O4 nanoparticles and its effect on the curing reaction of epoxy.
Figure 5. The schematic representation of synthesis Sm-Fe3O4 nanoparticles and its effect on the curing reaction of epoxy.
Jcs 06 00029 g005
Figure 6. The fractional extent of conversion as a function of reaction time for EP and EP/Sm-Fe3O4 nanocomposite at heating rates of 2.5, 5, 7.5, and 10 °C/min.
Figure 6. The fractional extent of conversion as a function of reaction time for EP and EP/Sm-Fe3O4 nanocomposite at heating rates of 2.5, 5, 7.5, and 10 °C/min.
Jcs 06 00029 g006
Figure 7. Friedman model plots of ln(dα/dt) vs. 1/T for (a) EP and (b) EP/Sm-Fe3O4 nanocomposite and KAS model plots of ln(β/T2) vs. 1/T for (c) EP and (d) EP/Sm-Fe3O4 nanocomposite at β of 2.5 °C/min.
Figure 7. Friedman model plots of ln(dα/dt) vs. 1/T for (a) EP and (b) EP/Sm-Fe3O4 nanocomposite and KAS model plots of ln(β/T2) vs. 1/T for (c) EP and (d) EP/Sm-Fe3O4 nanocomposite at β of 2.5 °C/min.
Jcs 06 00029 g007
Figure 8. Evolution of activation energy for EP and EP/Sm-Fe3O4 nanocomposite estimated by (a) differential Friedman model and (b) integral KAS model.
Figure 8. Evolution of activation energy for EP and EP/Sm-Fe3O4 nanocomposite estimated by (a) differential Friedman model and (b) integral KAS model.
Jcs 06 00029 g008
Figure 9. Plots of ln[Af(α)] vs. ln(1 − α) using Friedman model for EP and EP/Sm-Fe3O4 nanocomposite.
Figure 9. Plots of ln[Af(α)] vs. ln(1 − α) using Friedman model for EP and EP/Sm-Fe3O4 nanocomposite.
Jcs 06 00029 g009
Figure 10. The shape and alteration pattern of y(α) and z(α) versus the extent of reaction captured by the Malek model.
Figure 10. The shape and alteration pattern of y(α) and z(α) versus the extent of reaction captured by the Malek model.
Jcs 06 00029 g010
Figure 11. Value I and Value II for EP and EP/Sm-Fe3O4 nanocomposite.
Figure 11. Value I and Value II for EP and EP/Sm-Fe3O4 nanocomposite.
Jcs 06 00029 g011
Figure 12. Comparison of experimental data with the kinetic models for EP and EP/Sm-Fe3O4 nanocomposite based on Friedman and KAS model.
Figure 12. Comparison of experimental data with the kinetic models for EP and EP/Sm-Fe3O4 nanocomposite based on Friedman and KAS model.
Jcs 06 00029 g012
Table 1. Cure characteristics of EP and EP/Sm-Fe3O4 nanocomposite as a function of heating rate.
Table 1. Cure characteristics of EP and EP/Sm-Fe3O4 nanocomposite as a function of heating rate.
SampleHeating Rate (°C/min)TOnset (°C)Tp (°C)TEndset (°C)ΔT (°C)ΔH (J/g)ΔT*ΔH*CI
EP2.525.790.4159.9134.2329.6N.a.N.a.N.a.
5.037.4101.7171.7134.4336.2N.a.N.a.N.a.
7.540.7110.3180.7140.0344.1N.a.N.a.N.a.
1041.4119.2225.3183.9404.3N.a.N.a.N.a.
EP/Sm-Fe3O42.530.889.4170.6139.8385.81.041.171.22
5.035.7101.5175.2139.5363.71.041.081.12
7.539.1107.7207.0167.9296.31.20.861.03
1040.1115.1226.0185.9403.41.011.011.02
N.a.: Not applicable.
Table 2. Malek parameters of EP and EP/Sm-Fe3O4.
Table 2. Malek parameters of EP and EP/Sm-Fe3O4.
SampleHeating Rate (°C/min)αpαmαp
EP2.50.4470.0750.497
50.440.0810.545
7.50.4820.0870.553
100.3190.0920.484
EP/Sm-Fe3O42.50.4560.0510.487
50.5820.0490.551
7.50.4290.050.51
100.4180.0410.505
Table 3. The kinetic parameters evaluated for the curing of EP and EP/Sm-Fe3O4 nanocomposite based on Friedman and KAS models at different heating rates.
Table 3. The kinetic parameters evaluated for the curing of EP and EP/Sm-Fe3O4 nanocomposite based on Friedman and KAS models at different heating rates.
DesignationHeating Rate (°C/min)FriedmanKAS
mnlnA (s−1)mnlnA (s−1)
EP2.50.141.3212.090.091.3613.3
5.00.291.3812.710.241.4213.9
7.50.291.3612.760.241.413.92
100.251.6912.570.21.7413.71
EP/Sm-Fe3O42.50.161.5814.50.131.6215.39
5.00.231.4714.80.201.515.66
7.50.341.7215.050.301.7515.9
100.311.8514.90.281.8915.73
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jouyandeh, M.; Ganjali, M.R.; Mehrpooya, M.; Abida, O.; Jabbour, K.; Rabiee, N.; Habibzadeh, S.; Mashahdzadeh, A.H.; García-Peñas, A.; Stadler, F.J.; et al. Cure Kinetics of Samarium-Doped Fe3O4/Epoxy Nanocomposites. J. Compos. Sci. 2022, 6, 29. https://doi.org/10.3390/jcs6010029

AMA Style

Jouyandeh M, Ganjali MR, Mehrpooya M, Abida O, Jabbour K, Rabiee N, Habibzadeh S, Mashahdzadeh AH, García-Peñas A, Stadler FJ, et al. Cure Kinetics of Samarium-Doped Fe3O4/Epoxy Nanocomposites. Journal of Composites Science. 2022; 6(1):29. https://doi.org/10.3390/jcs6010029

Chicago/Turabian Style

Jouyandeh, Maryam, Mohammad Reza Ganjali, Mehdi Mehrpooya, Otman Abida, Karam Jabbour, Navid Rabiee, Sajjad Habibzadeh, Amin Hamed Mashahdzadeh, Alberto García-Peñas, Florian J. Stadler, and et al. 2022. "Cure Kinetics of Samarium-Doped Fe3O4/Epoxy Nanocomposites" Journal of Composites Science 6, no. 1: 29. https://doi.org/10.3390/jcs6010029

Article Metrics

Back to TopTop