Next Article in Journal
Effect of Carbon and Cooling Rate on the Structure of Hypereutectic High Chromium Cast Iron in the Cast State and after Heat Treatment
Previous Article in Journal
Nonwoven Electrospun Membranes as Tissue Scaffolds: Practices, Problems, and Future Directions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Influence of Pristine and Aminoacetic Acid-Treated Aluminum Nitride on the Structure, Curing Processes, and Properties of Epoxy Nanocomposites

1
Laboratory of Polymer Composites, K. Zhubanov Aktobe Regional State University, Aliya Moldagulova Avenue 34, Aktobe 030000, Kazakhstan
2
Laboratory of Modern Methods of Research of Functional Materials and Systems, Yuri Gagarin State Technical University of Saratov, Polytechnichskaya St., 77, 410054 Saratov, Russia
3
Laboratory of Support and Maintenance of the Educational Process, Yuri Gagarin State Technical University of Saratov, Polytechnichskaya St., 77, 410054 Saratov, Russia
4
Department “Chemistry and Chemical Technology”, K. Zhubanov Aktobe Regional State University, Aliya Moldagulova Avenue 34, Aktobe 030000, Kazakhstan
5
Department “Physics”, K. Zhubanov Aktobe Regional State University, Aliya Moldagulova Avenue 34, Aktobe 030000, Kazakhstan
6
Department of Economics and Humanitarian Sciences, Yuri Gagarin State Technical University of Saratov, Polytechnichskaya St., 77, 410054 Saratov, Russia
*
Author to whom correspondence should be addressed.
J. Compos. Sci. 2023, 7(12), 482; https://doi.org/10.3390/jcs7120482
Submission received: 19 October 2023 / Revised: 10 November 2023 / Accepted: 20 November 2023 / Published: 21 November 2023
(This article belongs to the Section Nanocomposites)

Abstract

:
This article describes the preparation of aluminum nitride (AlN) functionalized with amino groups by chemically modifying the surface of AlN with aminoacetic acid and their dispersion in the epoxy composition. As a result of the studies, it was found that the functionalization of AlN particles with aminoacetic acid ensured their better chemical compatibility with the epoxy composition, which facilitated the dispersibility of AlN particles; it was also found that the functionalization of the nanofiller has a significant effect on the structure, curing kinetics, physicochemical and mechanical properties of epoxy nanocomposites. The optimal concentration of the functionalizing agent (aminoacetic acid) has been determined, which is necessary for the chemical binding of the functional groups of aminoacetic acid, the epoxy oligomer and AlN, which best affects the dispersion and the size distribution of AlN particles, and also ensures maximum strength characteristics of epoxy composites containing AlN treated with aminoacetic acid: bending stress and bending modulus increase by 35% and 80%, respectively, while tensile strength and tensile elastic modulus increase by 74% and 36%, respectively. Furthermore, the impact strength shows a remarkable boost of 133% compared to plasticized epoxy composite without AlN.

1. Introduction

Dramatic technological progress in recent decades, which has created a demand for new materials with a unique set of properties necessary to achieve the assigned tasks, has become a response to the growing requirements in construction, automotive industry, shipbuilding, aerospace industry, etc. [1,2,3]. Polymers based on epoxy resin are widely used; their advantages are low shrinkage, corrosion resistance, and ease of molding, as well as the ability to use a wide range of modifiers, giving them unique properties. A special place among the modifiers of epoxy materials is occupied by such nanosized particles as carbon nanotubes (CNTs) and multi-walled carbon nanotubes (MWCNTs) [4,5], various graphene derivatives [6,7,8], Mxene [9,10], potassium polytitanates [11,12,13], and nanosized metal particles [14,15], as well as nitrides [16,17] and oxides of various elements [18,19].
In [20], the authors study the influence of SiO2 and boron nitride (BN) on the strength of the epoxy matrix. When adding up to 2 wt. % nanofillers, the system filled with BN has the best mechanical properties. SiO2 particles are less prone to agglomeration due to the phenomenon of self-lubrication and lower van der Waals forces between particles, which makes it possible to achieve higher strength values even with high filler content. The elastic modulus tends to increase with increasing the filler amount. Le Sanchez et al. [21] used a hybrid filler consisting of aluminum nitride (AlN) and BN to produce a protective epoxy coating for electronic devices. The addition of 75 wt. % made it possible to achieve a thermal conductivity of 10.18 W/(m × K), which is 46 times more than that of the original matrix, combined with a low coefficient of thermal expansion of 22.56 ppm/°C.
The reinforcement of polymers using fillers, including nano-dispersed ones, in some cases, can lead to some strengthening, but in most cases, the strengthening effect is low, which is explained by the tendency of nanomaterials to aggregate and low adhesive ability to the polymer matrix. To reduce the tendency to aggregation and to increase the adhesive ability of nanomaterials, it is necessary to functionalize them [22,23] and use effective methods for modifying composites, such as plasma treatment [24,25], microwave modification [26,27], exposure to magnetic and electric fields [28,29], etc. The method of surface functionalization is most widespread since it allows both to reduce an average particle size and to establish the chemical interaction at the phase interface, which will improve the deformation-strength characteristics of composites based on them [22,23,30].
In work [31], the thermal conductivity is increased by modifying the matrix with graphene quantum dots modified with polyetheramine with a large number of amine groups at the edges of the molecule and AlN. With the addition of 20 wt. % thermal conductivity index of the composite is 1.31 W·m−1·K−1, which is 6.89 times more than that of the original matrix. The functionalization of Si3N4 particles with polyaniline [32] leads to their better distribution in the volume of the epoxy matrix, which allows for increased corrosion resistance and adhesion to the metal substrate. The effect of glycidyl silane ether as a functionalizing agent for aramid nanofibers on the properties of an epoxy composite was studied in [33]. The addition of 1–1.5 wt. % of this filler made it possible to increase tensile strength and elastic modulus by 14 and 16% and impact strength by more than four times. DMA data showed that storage modulus increased by 13.4% by adding 1 wt. % functionalized aramid nanofibers.
In this work, the influence of pristine and aminofunctionalized (using aminoacetic acid) AlN on the processes of the structure formation during curing and the structure and properties of epoxy nanocomposites was studied. The results indicate that the incorporation of aminoacetic acid enhanced the chemical compatibility between AlN particles and the epoxy composition, resulting in the improved dispersibility of AlN particles. Furthermore, the functionalization of the nanofiller significantly influenced the structure and kinetics of curing, as well as the physicochemical and mechanical properties of the epoxy nanocomposites. This work is important in the field of modification and optimization of properties in the creation of highly efficient, strengthened AlN/epoxy nanocomposites.

2. Materials and Methods

2.1. Materials

In this study, epoxy resin ED-20 and polyethylene polyamine (PEPA) hardener produced by CHIMEX Limited, St. Petersburg, Russia, were used. We used tris(1-chloro-2-propyl) phosphate (TCPP) with purity of 95–99%, manufactured by Taizhou Ruishite New Material Co., Ltd. (Taizhou, China), providing plasticization of the polymer matrix and reducing its flammability. The TCPP molecule contains phosphorus (9.4%) and chlorine (32.4%), which allows this compound to be used as a fire retardant. During the thermal destruction of a composite containing TCPP, an increase in the yield of carbonized structures was noted, which prevents the release of volatile pyrolysis products into the gas phase, which generally reduces the flammability of the epoxy composite [34].
As a nanostructuring additive, we used AlN obtained from Xuzhou Jiechuang New Material Technology Co., Ltd. (Guangzhou, China).

2.2. Functionalization of the AlN Surface

The AlN surface was treated with aminoacetic acid. To carry this out, 0.25 g of AlN was dispersed in 50 mL of H2O–aminoacetic acid solution for 15 min using an ultrasonic homogenizer; the concentration of aminoacetic acid was 2.5%, 5.0%, and 7.5%. The suspension was then refluxed at 80 °C for 12 h with constant low-speed stirring at 100 rpm. To remove excess aminoacetic acid around the AlN particles, the suspension was centrifuged and washed twice with distilled water. Then, the resulting product was dried at 80 °C for 5 h.

2.3. Characterization of AlN

A Tescan VEGA 3 SBH scanning electron microscope (Brno, Czech Republic) was used to study the morphology of AlN particles. A Zetasizer Nano S (Malvern, Worcestershire, UK) was used to determine the size distribution of AlN particles. A Shimadzu IRTracer-100 (Tokyo, Japan) was used to perform FT-IR spectroscopy of AlN samples. An ARL X’TRA diffractometer was used to perform X-ray phase analysis (CuKα radiation, λ = 0.15412 nm, 2θ angle range 5–60°). A Quantachrome Nova 2200 surface area and porosity analyzer were used to determine the specific surface area of AlN particles.

2.4. Preparation of Epoxy Composites

A previously developed composition consisting of ED-20 (100 parts by mass), polyethylene polyamine (15 parts by mass), and TCPP (40 parts by mass) was used as a polymer matrix [34].
As a nanostructuring additive, AlN was introduced into the epoxy composition in an amount of 0.01–0.50 parts by mass. Ultrasonic treatment of the epoxy composition at a frequency of 22 ± 2 kHz, power of 400 W, and duration of 60 min ensured uniform dispersion of AlN particles.
Curing of the epoxy composition was carried out at a temperature of 25 ± 2 °C for 24 ± 1 h, followed by stepwise heat treatment at 90 ± 5 °C and 120 ± 5 °C for 2 h, respectively.

2.5. Testing of the Composites

A WDW-5E testing machine manufactured by Time Group Inc. (Beijing, China) was used to determine the tensile and flexural strength at a testing speed of 5 mm/min in tension and 50 mm/min in flexure. To determine the flexural strength and flexural modulus of elasticity, samples were tested in accordance with ISO 178:2019. To determine the tensile strength and tensile modulus of elasticity, samples were tested in accordance with ISO 527-2:2012. An LCT-50D impact tester (Beijing United Test Co., Ltd., Beijing, China) was used to determine the impact strength in accordance with ISO 179-1:2010. The heat resistance, according to Vicat’s method, was determined by ISO 306:2004. The curing kinetics was determined by the temperature method in accordance with the procedure described in [35]. The “DTAS-1300” (Samara, Russia) thermal analyzer was utilized to perform differential scanning calorimetry (DSC) on epoxy compositions (sample weight—20 mg, heating—up to 400 °C, heating rate—16 degrees per minute). Heat flow, which is the time derivative of heat, was used to determine the heat.

3. Results

The structure of AlN nanoparticles was studied by scanning electron microscopy (SEM). AlN particles are large aggregates (500–2000 nm) consisting of smaller particles (100–300 nm) of a scaly shape with irregular edges, Figure 1.
The fractional composition of AlN fractions determined in the distilled water, which is used to modify epoxy composites, are monodisperse and contain agglomerates with sizes in the range from 410 to 560 nm, in which particles with a size of 470 nm predominate, Figure 2.
Using the low-temperature nitrogen adsorption method on a Quantachrome Nova 2200 device, the specific surface area of AlN particles was determined to be 55 m2/g.
The study of the structure, size distribution, and specific surface area of AlN particles suggests that they can have a nanostructuring effect when introduced into an epoxy composition, which should improve the performance properties of composite materials [36,37].
AlN was added into the epoxy composition in an amount of 0.01–0.50 parts by mass. The results presented in Table 1 show that the optimal content of AlN as a nanomodifying additive is 0.05 parts by mass since it is at this content that the maximum values of physico-mechanical characteristics are achieved: bending stress and bending modulus increase by 13% and 58%, respectively, impact strength increases by 56%, while tensile strength and tensile elastic modulus increase by 56% and 28%, respectively.
A decrease in strength at less or more than the optimal AlN content is the result of ineffective interaction of the polymer matrix with the filler particles and is also caused by the aggregation of AlN particles [37,38,39].
The functionalization of nanofillers is an effective method that reduces the tendency to aggregation and increases the adhesive ability of nanomaterials to the polymer matrix. According to the literature, the most promising method of functionalization is treating the surface of nanoparticles with compounds that can ensure chemical interaction between nanomaterial particles and the polymer matrix, as well as reducing the polydispersity of the nanofiller, which will increase the deformation-strength characteristics of epoxy composites [40,41]. In this work, aminoacetic acid was used as such a compound, which contains functional groups capable of interacting with both the polymer matrix (amino groups) [42,43] and AlN particles (carboxyl groups).
We previously proved the chemical interaction of the functional groups of aminoacetic acid with the epoxy oligomer [44]. The formation of strong bonds between AlN particles and aminoacetic acid was proven by Fourier transform infrared spectroscopy, Figure 3. As can be seen from the spectra of the samples after modification, the vibration intensity of hydroxo-groups (3400 cm−1), which participate in the process of formation of the aminoacetic acid layer, decreases significantly. Moreover, vibration peaks corresponding to aminoacetic acid appear in the spectra of modified AlN, these groups being non-hydrolyzable because they are preserved after washing the particles with distilled water. In addition, it was found that an increase in the concentration of aminoacetic acid results in a significant increase in the intensity of the corresponding functional groups on the surface of nanoparticles. After undergoing treatment with aminoacetic acid, the spectrum of the sample exhibits a noticeable peak separation within the range of plane stretching vibrations of the Al—N bond (at 708 cm−1). This separation suggests an improved ordering of the AlN structure, attributable to the surface treatment of the material [45].
The most important characteristic of nanofillers is their morphology and specific surface area, on which the efficiency of interaction with the polymer matrix depends, especially when nanofillers are treated with various functionalizing agents [40,41].
Treatment with aminoacetic acid ensured delamination and fracture of AlN particles with a large lateral size. The most optimal concentration of aminoacetic acid for the modification of AlN is 5%, while the fractional composition of AlN particles modified with a 5% solution of aminoacetic acid is characterized by a unimodal distribution of particles and is represented by particles from 120 to 200 nm, particles with sizes of 160 nm predominating, line c in Figure 4, which is also confirmed by SEM data, Figure 5.
One of the determining factors in the influence of nanoparticles on the properties of composites is the filler wettability with the binder [46]. The results of the study of the AlN nanoparticles wetting with the epoxy composition (ED-20 + TCPP) using the sessile drop method are presented in Figure 6. Treatment of AlN particles with aminoacetic acid ensures their better wetting with the epoxy composition, which is proven by a decrease in the contact angle from 29° to 24°, which is achieved due to the active interaction of the reactive groups of the modifier and a significant increase in the specific surface area of AlN nanoparticles from 55 to 70 m2/g, Figure 6.
Based on the data obtained, a possible mechanism of interaction between the epoxy oligomer, aminoacetic acid, and aluminum nitride was proposed, Figure 7.
Treatment of a nanofiller with functionalizing agents can, in many cases, significantly improve its interaction with the polymer. This treatment reduces the free surface energy at the polymer–nanofiller interface and increases the work of adhesion, which improves the deformation–strength characteristics of the composite material [47,48]. Chemical interaction at the epoxy matrix/AlN interface provides an increase in the physico-mechanical properties of epoxy nanocomposites. The greatest strengthening effect is achieved by adding AlN treated with a 5% solution of aminoacetic acid; thus, the bending stress and bending modulus increase by 20% and 14%, respectively, impact strength increases by 50%, while tensile strength and tensile elastic modulus increase by 11% and 6%, respectively, and Vicat heat resistance increases by 14 °C, relative to the epoxy composite containing the pristine AlN, Figure 8, Figure 9 and Figure 10.
The study of the structure of epoxy composites showed that the fracture structure of the original epoxy polymer has a fairly smooth surface, which is characterized by low crack resistance [49,50], Figure 11a. The introduction of pristine AlN into the epoxy composite affects the fracture structure; an increase in the number of defects on the cleavage surface was found, which indicates that the destruction of the polymer composite requires more energy than the destruction of the original epoxy polymer, Figure 11b.
The presence of functional groups on the AlN surface led to an increase in the depth of defects on the cleavage surface; the number of defects increased significantly, and they became deeper and more elongated, Figure 11c, which indicates an increase in the energy required to destroy this system [51,52]. Besides a brittle fracture, the epoxy composite contains local areas indicating the flow of the material during its destruction. Moreover, in some places of plastic destruction, pronounced fibrous structures are observed, which are formed due to intense stretching of the polymer matrix, Figure 11d, which is explained by the formation of trans-boundary layers at the epoxy matrix–nanoparticle interface. This effect confirms the theory of the interaction of functional groups of the aminoacetic acid sizing additive with the epoxy oligomer, increasing the required energy for destruction, which is confirmed by an increase in the strength of the composite [51,52].
The addition of nanoparticles with a developed surface into an epoxy composition usually affects the polymerization processes that occur during the curing of the epoxy composite, which manifests itself in a change in the kinetic parameters of curing [53,54,55].
The influence of pristine and modified AlN on the processes of the structure formation during the curing has been established in Figure 12 and Figure 13. The addition of pristine AlN into the epoxy composition accelerates the polymerization process, resulting in a decreased gelation time from 104 to 75 min and a curing time from 146 to 105 min. Moreover, the maximum self-heating temperature of the composition increases from 88 to 105 °C. The addition of AlN functionalized with aminoacetic acid into the epoxy composition initiates the polymerization process due to the participation of reactive amino groups in the polymerization reaction, resulting in a reduction in gelation time from 75 to 53–70 min and curing time from 105 to 79–97 min. Simultaneously, there is an observed increase in the maximum self-heating temperature of the composition from 105 to 120–128 °C was noted, Table 2. Analysis of the data obtained shows that an increase in the concentration of aminoacetic acid accelerates the process of the structure formation of the epoxy composite containing modified AlN [56].
The results of studying the kinetics of curing epoxy composites by the thermometric method are confirmed by data obtained by differential scanning calorimetry, Figure 13. The addition of AlN treated with aminoacetic acid into the epoxy composition ensures an increase in the enthalpy of the curing reaction from 660 to 671–700 J/g and initiation of the curing process, which is confirmed by a decrease in the onset temperature of curing from 57 to 43–52 °C, Table 3.
Thus, the analysis of the results of differential scanning calorimetry of epoxy compositions showed that the addition of AlN nanoparticles functionalized with aminoacetic acid into the epoxy composition additionally initiates the polymerization process due to the participation of reactive amino groups in the polymerization reaction, while a decrease in the initial curing temperature and an increase in thermal effects of the polymerization reaction. Analysis of the data obtained showed that increasing the concentration of aminoacetic acid accelerates the process of structure formation of the epoxy composite containing modified nanoparticles. The results obtained are in good agreement with the literature data of previously published studies on the effect of functionalized nanoparticles on the curing kinetics of epoxy composites [57,58].

4. Conclusions

The conducted studies of the structure, fractional composition, and wettability of AlN nanoparticles have shown that AlN nanoparticles tend to agglomerate, which makes it difficult to distribute them uniformly in the epoxy composition and does not provide the maximum effect by their addition into the polymer composite. The data obtained show the need for the functionalization of nanoparticles, which made it possible to reduce the agglomeration of AlN nanoparticles and also ensure chemical interaction at the polymer matrix/nanoparticle interface. The optimal concentration of the functionalizing agent (aminoacetic acid) necessary for the chemical binding of the functional groups of aminoacetic acid, epoxy oligomer, and AlN was determined, which best affects the dispersion and size distribution of AlN particles and also ensures maximum strength characteristics of epoxy composites, containing AlN treated with aminoacetic acid: bending stress and bending modulus increase by 35% and 80%, respectively, while tensile strength and tensile elastic modulus increase by 74% and 36%, respectively. Furthermore, the impact strength shows a remarkable boost of 133% compared to plasticized epoxy composite without AlN.
The addition of AlN nanoparticles functionalized with aminoacetic acid into the epoxy composition additionally initiates the polymerization process due to the participation of reactive amino groups in the polymerization reaction, while a decrease in the initial curing temperature and an increase in thermal effects of the polymerization reaction. Analysis of the data obtained showed that increasing the concentration of aminoacetic acid accelerates the process of structure formation of the epoxy composite containing modified nanoparticles.
In addition, it was found that the addition of AlN nanoparticles treated with aminoacetic acid has a significant effect on the structure of the epoxy composite, which explains the significant strengthening of the epoxy composite.

Author Contributions

Conceptualization, A.B., A.M., A.S., L.T., M.A., A.A., R.O. and M.L.; Data curation, A.B., A.M., A.S., L.T., M.A., A.A. and R.O.; Formal analysis, A.B., A.M., A.S., L.T., M.A., A.A., R.O. and M.L.; Funding acquisition, A.B.; Investigation, A.M. and A.S.; Methodology, A.B., A.M., A.S., L.T., M.A., A.A. and R.O.; Project administration, A.B.; Resources, A.B., L.T., M.A., A.A. and R.O.; Software, A.B., A.M., M.A., A.A. and R.O.; Supervision, A.M.; Validation, A.M., A.S., M.A., A.A. and R.O.; Visualization, A.M., L.T., M.A. and A.A.; Writing—original draft, A.B. and A.M.; Writing—review and editing, A.M., A.S. and M.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research has been funded by the Science Committee of the Ministry of Science and Higher Education of the Republic of Kazakhstan (grant no. BR18574094).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hou, W.; Xiao, Y.; Han, G.; Lin, J.-Y. The Applications of Polymers in Solar Cells: A Review. Polymers 2019, 11, 143. [Google Scholar] [CrossRef]
  2. Janani, R.; Majumder, D.; Scrimshire, A.; Stone, A.; Wakelin, E.; Jones, A.H.; Wheeler, N.V.; Brooks, W.; Bingham, P.A. From Acrylates to Silicones: A Review of Common Optical Fibre Coatings Used for Normal to Harsh Environments. Prog. Org. Coat. 2023, 180, 107557. [Google Scholar] [CrossRef]
  3. Namsheer, K.; Rout, C.S. Conducting Polymers: A Comprehensive Review on Recent Advances in Synthesis, Properties and Applications. RSC Adv. 2021, 11, 5659–5697. [Google Scholar] [CrossRef]
  4. Bansal, S.A.; Khanna, V.; Twinkle; Singh, A.P.; Kumar, S. Small Percentage Reinforcement of Carbon Nanotubes (CNTs) in Epoxy(Bisphenol-A) for Enhanced Mechanical Performance. Mater. Today Proc. 2022, 61, 275–279. [Google Scholar] [CrossRef]
  5. Shcherbakov, A.S.; Mostovoy, A.S.; Yakovlev, N.A.; Arzamastsev, S.V. Effect of Carbon Nanotube Functionalization on the Physicochemical and Mechanical Properties of Modified Fiber-Reinforced Composites Based on an Epoxy Resin. Russ. J. Appl. Chem. 2021, 94, 1080–1087. [Google Scholar] [CrossRef]
  6. Akkalamattam Maitheenkunju, R.; Gopalakrishnan, J. Reduced Graphene Oxide/Epoxy Nanocomposites with Enhanced Dielectric, Mechanical, Thermomechanical and Thermal Properties. J. Elastomers Plast. 2022, 54, 407–428. [Google Scholar] [CrossRef]
  7. Du, X.; Zhou, H.; Sun, W.; Liu, H.-Y.; Zhou, G.; Zhou, H.; Mai, Y.-W. Graphene/Epoxy Interleaves for Delamination Toughening and Monitoring of Crack Damage in Carbon Fibre/Epoxy Composite Laminates. Compos. Sci. Technol. 2017, 140, 123–133. [Google Scholar] [CrossRef]
  8. Fang, F.; Ran, S.; Fang, Z.; Song, P.; Wang, H. Improved Flame Resistance and Thermo-Mechanical Properties of Epoxy Resin Nanocomposites from Functionalized Graphene Oxide via Self-Assembly in Water. Compos. Part B Eng. 2019, 165, 406–416. [Google Scholar] [CrossRef]
  9. Feng, A.; Hou, T.; Jia, Z.; Zhang, Y.; Zhang, F.; Wu, G. Preparation and Characterization of Epoxy Resin Filled with Ti3C2Tx MXene Nanosheets with Excellent Electric Conductivity. Nanomaterials 2020, 10, 162. [Google Scholar] [CrossRef]
  10. Song, P.; Qiu, H.; Wang, L.; Liu, X.; Zhang, Y.; Zhang, J.; Kong, J.; Gu, J. Honeycomb Structural rGO-MXene/Epoxy Nanocomposites for Superior Electromagnetic Interference Shielding Performance. Sustain. Mater. Technol. 2020, 24, e00153. [Google Scholar] [CrossRef]
  11. Mostovoi, A.S.; Plakunova, E.V.; Panova, L.G. New Epoxy Composites Based on Potassium Polytitanates. Int. Polym. Sci. Technol. 2013, 40, 49–51. [Google Scholar] [CrossRef]
  12. Mostovoy, A.S.; Kadykova, Y.A.; Bekeshev, A.Z.; Tastanova, L.K. Epoxy Composites Modified with Microfibers of Potassium Polytitanates. J. Appl. Polym. Sci. 2018, 135, 46651. [Google Scholar] [CrossRef]
  13. Mostovoi, A.S.; Yakovlev, E.A.; Burmistrov, I.N.; Panova, L.G. Use of Modified Nanoparticles of Potassium Polytitanate and Physical Methods of Modification of Epoxy Compositions for Improving Their Operational Properties. Russ. J. Appl. Chem. 2015, 88, 129–137. [Google Scholar] [CrossRef]
  14. Gao, W.; Yu, Y.; Chen, T.; Zhang, Q.; Chen, Z.; Chen, Z.; Jiang, J. Enhanced Flame Retardancy of Unsaturated Polyester Resin Composites Containing Ammonium Polyphosphate and Metal Oxides. J. Appl. Polym. Sci. 2020, 137, 49148. [Google Scholar] [CrossRef]
  15. Prabhu, S.; Bubbly, S.G.; Gudennavar, S.B. Bismuth (III) Oxide Decorated Graphene Oxide Filled Epoxy Nanocomposites: Thermo-Mechanical and Photon Attenuation Properties. Adv. Compos. Mater. 2023, 32, 602–628. [Google Scholar] [CrossRef]
  16. Pan, D.; Zhang, X.; Yang, G.; Shang, Y.; Su, F.; Hu, Q.; Patil, R.R.; Liu, H.; Liu, C.; Guo, Z. Thermally Conductive Anticorrosive Epoxy Nanocomposites with Tannic Acid-Modified Boron Nitride Nanosheets. Ind. Eng. Chem. Res. 2020, 59, 20371–20381. [Google Scholar] [CrossRef]
  17. Wu, Y.; Zhang, X.; Negi, A.; He, J.; Hu, G.; Tian, S.; Liu, J. Synergistic Effects of Boron Nitride (BN) Nanosheets and Silver (Ag) Nanoparticles on Thermal Conductivity and Electrical Properties of Epoxy Nanocomposites. Polymers 2020, 12, 426. [Google Scholar] [CrossRef]
  18. Liang, X.; Dai, F. Epoxy Nanocomposites with Reduced Graphene Oxide-Constructed Three-Dimensional Networks of Single Wall Carbon Nanotube for Enhanced Thermal Management Capability with Low Filler Loading. ACS Appl. Mater. Interfaces 2020, 12, 3051–3058. [Google Scholar] [CrossRef]
  19. Liu, Y.; Lu, M.; Wu, K.; Yao, S.; Du, X.; Chen, G.; Zhang, Q.; Liang, L.; Lu, M. Anisotropic Thermal Conductivity and Electromagnetic Interference Shielding of Epoxy Nanocomposites Based on Magnetic Driving Reduced Graphene oxide@Fe3O4. Compos. Sci. Technol. 2019, 174, 1–10. [Google Scholar] [CrossRef]
  20. Li, X.; Wang, Q.; Cui, X.; Feng, X.; Teng, F.; Xu, M.; Su, W.; He, J. Study on the Mechanical and Toughness Behavior of Epoxy Nano-Composites with Zero-Dimensional and Two-Dimensional Nano-Fillers. Polymers 2022, 14, 3618. [Google Scholar] [CrossRef]
  21. Lee Sanchez, W.A.; Li, J.-W.; Chiu, H.-T.; Cheng, C.-C.; Chiou, K.-C.; Lee, T.-M.; Chiu, C.-W. Highly Thermally Conductive Epoxy Composites with AlN/BN Hybrid Filler as Underfill Encapsulation Material for Electronic Packaging. Polymers 2022, 14, 2950. [Google Scholar] [CrossRef]
  22. Malekshahinezhad, K.; Ahmadi-khaneghah, A.; Behniafar, H. Amine-Functionalized TiO2 Nanoparticles Covalently Loaded into Epoxy Networks via Thermal and Microwave Curing Processes. Macromol. Res. 2020, 28, 567–572. [Google Scholar] [CrossRef]
  23. Mozaffarinasab, H.; Jamshidi, M. Surface Modification of Carbon Nanotubes by a Bifunctional Amine Silane; Effects on Physical/Mechanical/Thermal Properties of Epoxy Nanocomposite. Prog. Org. Coat. 2023, 179, 107521. [Google Scholar] [CrossRef]
  24. Jung, H.; Choi, H.K.; Oh, Y.; Hong, H.; Yu, J. Enhancement of Thermo-Mechanical Stability for Nanocomposites Containing Plasma Treated Carbon Nanotubes with an Experimental Study and Molecular Dynamics Simulations. Sci. Rep. 2020, 10, 405. [Google Scholar] [CrossRef] [PubMed]
  25. Seghini, M.C.; Touchard, F.; Sarasini, F.; Cech, V.; Chocinski-Arnault, L.; Mellier, D.; Tirillò, J.; Bracciale, M.P.; Zvonek, M. Engineering the Interfacial Adhesion in Basalt/Epoxy Composites by Plasma Polymerization. Compos. Part A Appl. Sci. Manuf. 2019, 122, 67–76. [Google Scholar] [CrossRef]
  26. Sun, Y.; Zhang, P.; Hu, J.; Liu, B.; Yang, J.; Liang, S.; Xiao, K.; Hou, H. A Review on Microwave Irradiation to the Properties of Geopolymers: Mechanisms and Challenges. Constr. Build. Mater. 2021, 294, 123491. [Google Scholar] [CrossRef]
  27. Shcherbakov, A.; Mostovoy, A.; Bekeshev, A.; Burmistrov, I.; Arzamastsev, S.; Lopukhova, M. Effect of Microwave Irradiation at Different Stages of Manufacturing Unsaturated Polyester Nanocomposite. Polymers 2022, 14, 4594. [Google Scholar] [CrossRef] [PubMed]
  28. Domingues, D.; Logakis, E.; Skordos, A.A. The Use of an Electric Field in the Preparation of Glass Fibre/Epoxy Composites Containing Carbon Nanotubes. Carbon 2012, 50, 2493–2503. [Google Scholar] [CrossRef]
  29. Miękoś, E.; Cichomski, M.; Zieliński, M.; Klepka, T.; Sroczyński, D.; Fenyk, A. Tests of Physicochemical and Mechanical Strength Properties of Polymer Composites on an Epoxy Resin Matrix, Modified by a Constant Magnetic Field. Materials 2022, 15, 6730. [Google Scholar] [CrossRef]
  30. Hetemi, D.; Pinson, J. Surface Functionalisation of Polymers. Chem. Soc. Rev. 2017, 46, 5701–5713. [Google Scholar] [CrossRef]
  31. Zhang, D.; Liu, F.; Wang, S.; Yan, M.; Hu, X.; Xu, M. D-GQDs Modified Epoxy Resin Enhances the Thermal Conductivity of AlN/Epoxy Resin Thermally Conductive Composites. Polymers 2021, 13, 4074. [Google Scholar] [CrossRef]
  32. Wang, H.; Qi, Q.; Zhang, Y.; Chen, S.; Dong, B.; Zhu, S.; Hu, Q.; Guo, Z. Anticorrosive Epoxy Nanocomposite Coatings Filled with Polyaniline-Functionalized Silicon Nitride Particles. Ind. Eng. Chem. Res. 2020, 59, 16649–16659. [Google Scholar] [CrossRef]
  33. Jung, J.; Sodano, H.A. High Strength Epoxy Nanocomposites Reinforced by Epoxy Functionalized Aramid Nanofibers. Polymer 2020, 195, 122438. [Google Scholar] [CrossRef]
  34. Bekeshev, A.; Mostovoy, A.; Shcherbakov, A.; Zhumabekova, A.; Serikbayeva, G.; Vikulova, M.; Svitkina, V. Effect of Phosphorus and Chlorine Containing Plasticizers on the Physicochemical and Mechanical Properties of Epoxy Composites. J. Compos. Sci. 2023, 7, 178. [Google Scholar] [CrossRef]
  35. Bekeshev, A.; Mostovoy, A.; Tastanova, L.; Kadykova, Y.; Kalganova, S.; Lopukhova, M. Reinforcement of Epoxy Composites with Application of Finely-Ground Ochre and Electrophysical Method of the Composition Modification. Polymers 2020, 12, 1437. [Google Scholar] [CrossRef] [PubMed]
  36. Demir, K.; Gavgali, E.; Yetim, A.F.; Akpinar, S. The Effects of Nanostructure Additive on Fracture Strength in Adhesively Bonded Joints Subjected to Fully Reversed Four-Point Bending Fatigue Load. Int. J. Adhes. Adhes. 2021, 110, 102943. [Google Scholar] [CrossRef]
  37. Rodríguez, H.A.; Kriven, W.M.; Casanova, H. Development of Mechanical Properties in Dental Resin Composite: Effect of Filler Size and Filler Aggregation State. Mater. Sci. Eng. C 2019, 101, 274–282. [Google Scholar] [CrossRef]
  38. Wang, Y.; Yao, D.; He, Z.; Wang, D.; Zheng, Y. Enhanced Mechanical and Damping Properties of Epoxy Using Aggregated Nanoparticles Organic-Inorganic Hybrid as a Filler. Compos. Interfaces 2022, 29, 523–536. [Google Scholar] [CrossRef]
  39. Baek, K.; Shin, H.; Cho, M. Multiscale Modeling of Mechanical Behaviors of Nano-SiC/Epoxy Nanocomposites with Modified Interphase Model: Effect of Nanoparticle Clustering. Compos. Sci. Technol. 2021, 203, 108572. [Google Scholar] [CrossRef]
  40. Prasad, T.; Halder, S.; Dhar, S.S. Imidazole-Supported Silica One-Pot Processed Nanoparticles to Enhance Toughness of Epoxy Based Nanocomposites. Mater. Chem. Phys. 2019, 231, 75–86. [Google Scholar] [CrossRef]
  41. Feichtenschlager, B.; Pabisch, S.; Svehla, J.; Peterlik, H.; Sajjad, M.; Koch, T.; Kickelbick, G. Epoxy Resin Nanocomposites: The Influence of Interface Modification on the Dispersion Structure—A Small-Angle-X-Ray-Scattering Study. Surfaces 2020, 3, 664–682. [Google Scholar] [CrossRef]
  42. Bao, T.; Wang, Z.; Zhao, Y.; Wang, Y.; Yi, X. Improving Tribological Performance of Epoxy Composite by Reinforcing with Polyetheramine-Functionalized Graphene Oxide. J. Mater. Res. Technol. 2021, 12, 1516–1529. [Google Scholar] [CrossRef]
  43. Tiwari, N.; Shaikh, A.A.; Malek, N.I. Modification of the Multiphase Shape Memory Composites with Functionalized Graphene Nanoplatelets: Enhancement of Thermomechanical and Interfacial Properties. Mater. Today Chem. 2022, 24, 100826. [Google Scholar] [CrossRef]
  44. Mostovoy, A.; Shcherbakov, A.; Yakovlev, A.; Arzamastsev, S.; Lopukhova, M. Reinforced Epoxy Composites Modified with Functionalized Graphene Oxide. Polymers 2022, 14, 338. [Google Scholar] [CrossRef] [PubMed]
  45. Teng, C.-C.; Ma, C.-C.M.; Chiou, K.-C.; Lee, T.-M. Synergetic Effect of Thermal Conductive Properties of Epoxy Composites Containing Functionalized Multi-Walled Carbon Nanotubes and Aluminum Nitride. Compos. Part B Eng. 2012, 43, 265–271. [Google Scholar] [CrossRef]
  46. Zhang, Y.; Park, M.; Park, S.-J. Implication of Thermally Conductive Nanodiamond-Interspersed Graphite Nanoplatelet Hybrids in Thermoset Composites with Superior Thermal Management Capability. Sci. Rep. 2019, 9, 2893. [Google Scholar] [CrossRef]
  47. Zappalorto, M.; Pontefisso, A.; Fabrizi, A.; Quaresimin, M. Mechanical Behaviour of Epoxy/Silica Nanocomposites: Experiments and Modelling. Compos. Part A Appl. Sci. Manuf. 2015, 72, 58–64. [Google Scholar] [CrossRef]
  48. Esmaeili, A.; Ma, D.; Manes, A.; Oggioni, T.; Jiménez-Suárez, A.; Ureña, A.; Hamouda, A.M.S.; Sbarufatti, C. An Experimental and Numerical Investigation of Highly Strong and Tough Epoxy Based Nanocomposite by Addition of MWCNTs: Tensile and Mode I Fracture Tests. Compos. Struct. 2020, 252, 112692. [Google Scholar] [CrossRef]
  49. Goyat, M.S.; Hooda, A.; Gupta, T.K.; Kumar, K.; Halder, S.; Ghosh, P.K.; Dehiya, B.S. Role of Non-Functionalized Oxide Nanoparticles on Mechanical Properties and Toughening Mechanisms of Epoxy Nanocomposites. Ceram. Int. 2021, 47, 22316–22344. [Google Scholar] [CrossRef]
  50. Zhang, Q.; Bai, G.; Xiao, W.; Sui, G.; Yang, X. Effect of Amine Functionalized MWCNT-Epoxy Interfacial Interaction on MWCNT Dispersion and Mechanical Properties of Epoxy-Amine Composites. Polym. Compos. 2018, 39, E2552–E2561. [Google Scholar] [CrossRef]
  51. Sari, M.G.; Ramezanzadeh, B. Epoxy Composite Coating Corrosion Protection Properties Reinforcement through the Addition of Hydroxyl-Terminated Hyperbranched Polyamide Non-Covalently Assembled Graphene Oxide Platforms. Constr. Build. Mater. 2020, 234, 117421. [Google Scholar] [CrossRef]
  52. Ali, F.; Ishfaq, N.; Said, A.; Nawaz, Z.; Ali, Z.; Ali, N.; Afzal, A.; Bilal, M. Fabrication, Characterization, Morphological and Thermal Investigations of Functionalized Multi-Walled Carbon Nanotubes Reinforced Epoxy Nanocomposites. Prog. Org. Coat. 2021, 150, 105962. [Google Scholar] [CrossRef]
  53. 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]
  54. Sul, J.-H.; Prusty, B.G.; Crosky, A. Effect of the Addition of Multi-Walled Carbon Nanotubes on the Thermomechanical Properties of Epoxy Resin. Polym. Compos. 2017, 38, 1873–1880. [Google Scholar] [CrossRef]
  55. Hameed, A.; Islam, M.; ahmad, I.; Mahmood, N.; Saeed, S.; Javed, H. Thermal and Mechanical Properties of Carbon Nanotube/Epoxy Nanocomposites Reinforced with Pristine and Functionalized Multiwalled Carbon Nanotubes. Polym. Compos. 2015, 36, 1891–1898. [Google Scholar] [CrossRef]
  56. Shahrestanaki, A.A.K.; Mehrshad, M.; Akhlaghi, S.H. Preparation and Non-Isothermal Cure Kinetics Study of Epoxy Resin Nanocomposites with Amine and Epoxy Functionalized Magnetic Nanoparticles. High Perform. Polym. 2021, 33, 1025–1034. [Google Scholar] [CrossRef]
  57. Konuray, O.; Fernández-Francos, X.; Ramis, X.; Serra, À. State of the Art in Dual-Curing Acrylate Systems. Polymers 2018, 10, 178. [Google Scholar] [CrossRef] [PubMed]
  58. Ribas-Massonis, A.; Cicujano, M.; Duran, J.; Besalú, E.; Poater, A. Free-Radical Photopolymerization for Curing Products for Refinish Coatings Market. Polymers 2022, 14, 2856. [Google Scholar] [CrossRef] [PubMed]
Figure 1. SEM data of AlN particles: (a) ×10,000; (b,c) ×100,000.
Figure 1. SEM data of AlN particles: (a) ×10,000; (b,c) ×100,000.
Jcs 07 00482 g001
Figure 2. Fractional composition of AlN particles.
Figure 2. Fractional composition of AlN particles.
Jcs 07 00482 g002
Figure 3. FT-IR spectroscopy of pristine AlN, aminoacetic acid, and AlN samples treated with aminoacetic acid: 1—pristine AlN; 2—aminoacetic acid; 3—AlN samples treated with 2.5% aminoacetic acid solution; 4—AlN samples treated with 5.0% aminoacetic acid solution; 5—AlN samples treated with 7.5% aminoacetic acid solution.
Figure 3. FT-IR spectroscopy of pristine AlN, aminoacetic acid, and AlN samples treated with aminoacetic acid: 1—pristine AlN; 2—aminoacetic acid; 3—AlN samples treated with 2.5% aminoacetic acid solution; 4—AlN samples treated with 5.0% aminoacetic acid solution; 5—AlN samples treated with 7.5% aminoacetic acid solution.
Jcs 07 00482 g003
Figure 4. Fractional composition of AlN particles: a—pristine AlN; b—AlN, modified with 2.5% aminoacetic acid; c—AlN, modified with 5.0% aminoacetic acid; d—AlN, modified with 7.5% aminoacetic acid.
Figure 4. Fractional composition of AlN particles: a—pristine AlN; b—AlN, modified with 2.5% aminoacetic acid; c—AlN, modified with 5.0% aminoacetic acid; d—AlN, modified with 7.5% aminoacetic acid.
Jcs 07 00482 g004
Figure 5. SEM data of AlN particles treated with aminoacetic acid.
Figure 5. SEM data of AlN particles treated with aminoacetic acid.
Jcs 07 00482 g005
Figure 6. Wetting angle and specific surface area of ND particles: (a) pristine ND; (b) after aminoacetic acid treatment.
Figure 6. Wetting angle and specific surface area of ND particles: (a) pristine ND; (b) after aminoacetic acid treatment.
Jcs 07 00482 g006
Figure 7. Possible chemistry of the interaction of epoxy oligomer, aminoacetic acid, and aluminum nitride: (a)—interaction of aminoacetic acid with aluminum nitrides; (b)—interaction of aluminum nitride treated with aminoacetic acid with functional groups of epoxy oligomer.
Figure 7. Possible chemistry of the interaction of epoxy oligomer, aminoacetic acid, and aluminum nitride: (a)—interaction of aminoacetic acid with aluminum nitrides; (b)—interaction of aluminum nitride treated with aminoacetic acid with functional groups of epoxy oligomer.
Jcs 07 00482 g007
Figure 8. Dependence of bending stress (1) and modulus of elasticity in bending (2) of an epoxy composite containing 0.05 parts by mass of AlN on the concentration of aminoacetic acid used to modify AlN.
Figure 8. Dependence of bending stress (1) and modulus of elasticity in bending (2) of an epoxy composite containing 0.05 parts by mass of AlN on the concentration of aminoacetic acid used to modify AlN.
Jcs 07 00482 g008
Figure 9. Dependence of the strength (1) and tensile modulus (2) of an epoxy composite containing 0.05 parts by mass of AlN on the concentration of aminoacetic acid used to modify AlN.
Figure 9. Dependence of the strength (1) and tensile modulus (2) of an epoxy composite containing 0.05 parts by mass of AlN on the concentration of aminoacetic acid used to modify AlN.
Jcs 07 00482 g009
Figure 10. Dependence of impact strength (1) and Vicat heat resistance (2) of an epoxy composite containing 0.05 parts by mass of AlN on the concentration of aminoacetic acid used to modify AlN.
Figure 10. Dependence of impact strength (1) and Vicat heat resistance (2) of an epoxy composite containing 0.05 parts by mass of AlN on the concentration of aminoacetic acid used to modify AlN.
Jcs 07 00482 g010
Figure 11. SEM data for epoxy composite samples, parts by mass: (a)—100 ED-20 + 40 TCPP + 15 PEPA; (b)—100 ED-20 + 40 TCPP + 0.05 AlN + 15 PEPA; (c,d)—100 ED-20 + 40 TCPP + 0.05 AlN (aminoacetic acid) + 15 PEPA.
Figure 11. SEM data for epoxy composite samples, parts by mass: (a)—100 ED-20 + 40 TCPP + 15 PEPA; (b)—100 ED-20 + 40 TCPP + 0.05 AlN + 15 PEPA; (c,d)—100 ED-20 + 40 TCPP + 0.05 AlN (aminoacetic acid) + 15 PEPA.
Jcs 07 00482 g011
Figure 12. Dependence of sample temperature over time during curing of epoxy compositions containing pristine and modified AlN: (1)—100 ED-20 + 40 TCPP + 0.05 AlN + 15 PEPA; (2)—100 ED-20 + 40 TCPP + 0.05 AlN(2.5% aminoacetic acid) + 15 PEPA; (3)—100 ED-20 + 40 TCPP + 0.05 AlN(5.0% aminoacetic acid) + 15 PEPA; (4)—100 ED-20 + 40 TCPP + 0.05 AlN(7.5% aminoacetic acid) + 15 PEPA; (5)—100 ED-20 + 40 TCPP + 15 PEPA.
Figure 12. Dependence of sample temperature over time during curing of epoxy compositions containing pristine and modified AlN: (1)—100 ED-20 + 40 TCPP + 0.05 AlN + 15 PEPA; (2)—100 ED-20 + 40 TCPP + 0.05 AlN(2.5% aminoacetic acid) + 15 PEPA; (3)—100 ED-20 + 40 TCPP + 0.05 AlN(5.0% aminoacetic acid) + 15 PEPA; (4)—100 ED-20 + 40 TCPP + 0.05 AlN(7.5% aminoacetic acid) + 15 PEPA; (5)—100 ED-20 + 40 TCPP + 15 PEPA.
Jcs 07 00482 g012
Figure 13. DSC results: (1)—100 ED-20 + 40 TCPP + 0.05 AlN + 15 PEPA; (2)—100 ED-20 + 40 TCPP + 0.05 AlN (2.5% aminoacetic acid) + 15 PEPA; (3)—100 ED-20 + 40 TCPP + 0.05 AlN (5.0% aminoacetic acid) + 15 PEPA; (4)—100 ED-20 + 40 TCPP + 0.05 AlN (7.5% aminoacetic acid) + 15 PEPA.
Figure 13. DSC results: (1)—100 ED-20 + 40 TCPP + 0.05 AlN + 15 PEPA; (2)—100 ED-20 + 40 TCPP + 0.05 AlN (2.5% aminoacetic acid) + 15 PEPA; (3)—100 ED-20 + 40 TCPP + 0.05 AlN (5.0% aminoacetic acid) + 15 PEPA; (4)—100 ED-20 + 40 TCPP + 0.05 AlN (7.5% aminoacetic acid) + 15 PEPA.
Jcs 07 00482 g013
Table 1. Properties of epoxy nanocomposites.
Table 1. Properties of epoxy nanocomposites.
Samplesσben, MPaEben,
MPa
σten, MPaEten,
MPa
aim,
kJ/m2
100 ED-20 + 15 PEPA + 40 TCPP85 ± 2.82077 ± 6234 ± 1.71634 ± 509.0 ± 0.35
100 ED-20 + 15 PEPA + 40 TCPP + 0.01 AlN91 ± 2.52503 ± 7542 ± 2.01980 ± 7913.0 ± 0.52
100 ED-20 + 15 PEPA + 40 TCPP + 0.05 AlN96 ± 2.83286 ± 9153 ± 2.32091 ± 8214.0 ± 0.56
100 ED-20 + 15 PEPA + 40 TCPP + 0.10 AlN94 ± 2.63422 ± 9951 ± 2.22096 ± 8312.6 ± 0.50
100 ED-20 + 15 PEPA + 40 TCPP + 0.50 AlN92 ± 2.53537 ± 10540 ± 2.02102 ± 8410.8 ± 0.43
Note: σben—bending stress; Eben—bending modulus of elasticity; σten—tensile strength; Eten—tensile modulus of elasticity; aim—impact strength.
Table 2. Values of curing indicators of epoxy compositions.
Table 2. Values of curing indicators of epoxy compositions.
Compositionτgel, minτcur, minTmax, °C
ED-20 + TCPP + PEPA10414688
ED-20 + TCPP + AlN + PEPA75105105
ED-20 + TCPP + AlN(2.5% aminoacetic acid) + PEPA7097120
ED-20 + TCPP + AlN(5.0% aminoacetic acid) + PEPA6691122
ED-20 + TCPP + AlN(7.5% aminoacetic acid) + PEPA5379128
Note: τgel—gelation time, τcur—curing time, Tmax—maximum self-heating temperature.
Table 3. Results of differential scanning calorimetry of epoxy compositions.
Table 3. Results of differential scanning calorimetry of epoxy compositions.
CompositionTstart–Tend
Tmax
°C
H, J/g
ED-20 + TCPP + AlN + PEPA57–176
108
660
ED-20 + TCPP + AlN (2.5% aminoacetic acid) + PEPA52–175
108
671
ED-20 + TCPP + AlN (5.0% aminoacetic acid) + PEPA50–176
108
683
ED-20 + TCPP + AlN (7.5% aminoacetic acid) + PEPA43–173
108
700
Note: Tstart, Tend—starting and ending curing temperatures, Tmax—the temperature of the maximum heat release during curing, H—enthalpy of the curing reaction.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Bekeshev, A.; Mostovoy, A.; Shcherbakov, A.; Tastanova, L.; Akhmetova, M.; Apendina, A.; Orynbassar, R.; Lopukhova, M. The Influence of Pristine and Aminoacetic Acid-Treated Aluminum Nitride on the Structure, Curing Processes, and Properties of Epoxy Nanocomposites. J. Compos. Sci. 2023, 7, 482. https://doi.org/10.3390/jcs7120482

AMA Style

Bekeshev A, Mostovoy A, Shcherbakov A, Tastanova L, Akhmetova M, Apendina A, Orynbassar R, Lopukhova M. The Influence of Pristine and Aminoacetic Acid-Treated Aluminum Nitride on the Structure, Curing Processes, and Properties of Epoxy Nanocomposites. Journal of Composites Science. 2023; 7(12):482. https://doi.org/10.3390/jcs7120482

Chicago/Turabian Style

Bekeshev, Amirbek, Anton Mostovoy, Andrey Shcherbakov, Lyazzat Tastanova, Marzhan Akhmetova, Ainagul Apendina, Raigul Orynbassar, and Marina Lopukhova. 2023. "The Influence of Pristine and Aminoacetic Acid-Treated Aluminum Nitride on the Structure, Curing Processes, and Properties of Epoxy Nanocomposites" Journal of Composites Science 7, no. 12: 482. https://doi.org/10.3390/jcs7120482

Article Metrics

Back to TopTop