Next Article in Journal
Nitrates Removal from Simulated Groundwater Using Nano Zerovalent Iron Supported by Polystyrenic Gel
Next Article in Special Issue
Research and Application of Biomass-Based Wood Flame Retardants: A Review
Previous Article in Journal
A Functionalized Polysaccharide from Sphingomonas sp. HL-1 for High-Performance Flocculation
Previous Article in Special Issue
Understanding the Flame Retardant Mechanism of Intumescent Flame Retardant on Improving the Fire Safety of Rigid Polyurethane Foam
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Improving the Heat Resistance and Flame Retardancy of Epoxy Resin Composites by Novel Multifunctional Cyclophosphazene Derivatives

1
School of Chemistry and Chemical Engineering, Jiangsu University of Technology, Changzhou 213001, China
2
School of Computer Engineering, Jiangsu University of Technology, Changzhou 213001, China
3
School of Materials Science and Engineering, Changzhou University, Changzhou 213164, China
*
Author to whom correspondence should be addressed.
Polymers 2023, 15(1), 59; https://doi.org/10.3390/polym15010059
Submission received: 27 November 2022 / Revised: 16 December 2022 / Accepted: 20 December 2022 / Published: 23 December 2022
(This article belongs to the Special Issue Heat-Resistant and Flame-Retardant Polymer Materials)

Abstract

:
A novel multiple-ring molecule containing P and N, called HCCP-SA, was successfully prepared by the nucleophilic substitution reaction of salicylamide (SA) and hexachlorocyclotriphosphazene (HCCP). Particularly, HCCP-SA possessed the dual functions of heat resistance and flame retardancy. The molecular structure of HCCP-SA was identified by Fourier transform infrared spectroscopy and nuclear magnetic resonance spectroscopy. HCCP-SA was bonded into the molecular chain of epoxy resin by the ring-opening curing reaction of epoxy resin, aiming to form a heat-resistant and flame-retardant composite (E-HS-x). In particular, the best-prepared E-HS-x composite with a 20 phr content of HCCP-SA (E-HS-20) presented excellent thermal stability, with an initial decomposition temperature of 267.94 °C and a max weight loss speed of only 0.95 mg·min−1. Moreover, E-HS-20 exhibited remarkable flame retardancy with a limiting oxygen index value of 27.1% and a V-2 rating in the UL94 flame retardancy test. The best-prepared E-HS-20 composite would be a suitable and potential candidate for heat-resistant and flame-retardant polymer materials.

Graphical Abstract

1. Introduction

As a kind of commercial thermosetting material, epoxy resins (EP) possess excellent mechanical strength, thermal stability, adhesion and electrical insulation [1,2,3]; therefore, they are widely used in the field of electronic information technology, such as packaging materials for electronic components, anti-corrosion coatings, etc. With the rapid development of semiconductor technology and integrated circuit integration, the heat generation of electronic components has increased dramatically in recent years. Accordingly, advanced high-density installation technology (especially large and super large-scale integrated circuits) and thin packaging technology urgently require that the heat resistance and flame resistance of epoxy resin be simultaneously improved, in order to meet the industrial demand in all walks of life.
The improvement of the heat resistance of electronic packaging materials can usually be achieved by increasing the crosslinking degree of epoxy resin [4,5,6]; however, the brittleness of epoxy resin will increase if the crosslinking degree is too high. A large number of investigations have shown that the introduction of additives containing phenyl and naphthyl groups in epoxy resin can also improve the heat resistance [7,8,9]; however, it is a challenge to build a new multi-ring structure in epoxy resin to improve its heat resistance.
Two main methods to obtain epoxy resin with high flame retardancy are to add flame retardants and/or prepare inherent flame retardants by embedding flame-retardant groups in the molecular chain of epoxy resin. The addition of flame retardants usually shows some limitations. For example, due to the compatibility between EP resin and additives, additives are easily separated from the EP resin matrix when the composite is used for a long time, resulting in the mechanical properties of the EP resin composite [10,11,12]; therefore, the development of efficient, inherently flame-retardant epoxy resin products with multi-ring structures has become a research focus.
Recently, some fillers available in the commercial market, such as carbon nanotubes [13], multiwalled carbon nanotubes [14], functionalized graphene [15,16] and hexachlorocyclotriphosphazene (HCCP) [17,18], have been used to improve the flame retardancy of epoxy resin, and excellent results have been obtained. Qu [16] synthesized reactive cyclophosphazene and grafted it onto graphene oxide to prepare flame-retardant graphene oxide (fGO) nano sheets for EP nanocomposites. The obtained fGO showed good compatibility and improved the thermal stability and dynamic mechanical properties of the composite. Among the fillers mentioned above, HCCP is a classic example of reinforcing elements for polymeric composites because of its high reactivity, low cost and nearly unlimited modifiability. More importantly, HCCP can improve the flame retardancy of epoxy resin by introducing flame-retardant groups while building a multi-ring structure to improve the heat resistance.
In this investigation, a novel HCCP derivative with a multiple-ring structure (named HCCP-SA) was successfully synthesized though bonding SA into the conjugated ring of HCCP. Then, HCCP-SA was embedded into the main chain of epoxy resin by the ring-opening curing reaction of epoxy resin to prepare a heat-resistant and flame-retardant EP composite (named E-HS-x: x denotes the parts of HCCP-SA used in per hundred parts of EP resins). The effects of HCCP-SA on thermal property and flame retardancy in the E-HS-x composite are discussed in detail.

2. Materials and Methods

2.1. Materials

HCCP (98%) and epoxy resin (E-51) were provided by Sigma-Aldrich Co. Ltd. (Shanghai, China); salicylamide (SA) was provided by Shanghai Tengzhun Biotechnology Co. Ltd. (Shanghai, China); n-heptane (A.R.), triethylamine (A.R.), diethylenetriamine (A.R.) and tetrahydrofuran (A.R.) were purchased from Changzhou Tongcheng Co., Ltd. (Changzhou, China).

2.2. Synthesis of HCCP-SA

HCCP (13.90 g, 0.04 mol) and tetrahydrofuran (120 mL) were placed in a three necked flask and stirred vigorously for 0.5 h at 65 °C in a nitrogen atmosphere. SA (41.90 mL) and triethylamine (10.00 mL) were mixed evenly and added drop by drop into the three necked flask with a constant pressure-dropping funnel within 20 min. The reaction was kept at 65 °C for approximately 8 h under a nitrogen atmosphere. Then, vacuum distillation was performed on a rotary evaporator, model RE520 (Yarong, Shanghai, China), at 80 °C, to remove the solvent (tetrahydrofuran). The crude product was washed with n-heptane and deionized water 3 times, and then dried at 60 °C for 36 h in a vacuum oven. Finally, white powder HCCP-SA was obtained after grinding. Figure 1a shows the synthesis route of HCCP-SA and its multiple-ring structure.

2.3. Preparation of E-HS-x Composites

According to the calculated mass, appropriate amounts of diethylenetriamine were added into a beaker containing EP resin and stirred quickly for 3 min at room temperature. Immediately after, HCCP-SA was added into the above precuring epoxy resin and stirred quickly for 3 min. The mixture was injected into the molds with a cavity size of 100 mm × 10 mm × 4 mm (comprising polytetrafluoroethylene) as soon as possible to prepare the E-HS-x composite. Detailed experimental procedures for the preparation of the E-HS-x composite are reported in our previous investigations [18]. The mass of HCCP-SA used in the preparation of E-HS-x composite is listed in Table 1, in which phr is defined as the dosage per hundred grams of EP.

2.4. Characterization

Infrared measurements were recorded on a Nicolet IS10 Fourier transform infrared (FTIR) spectrometer (ThermoFisher, Waltham, MA, USA) to obtain the FTIR spectra of the SA, HCCP-SA and E-HS-x composites. The 1H NMR (400 MHz) spectra of HCCP-SA were recorded on a AVANCE III 400M nuclear magnetic resonance spectrometer (NMR, Bruker, Romanshorn, Switzerland), using Hexadeuterated dimethyl sulfoxide (DMSO-d6) as the solvent. The morphology of the E-HS-x composites was recorded using a scanning electron microscope (SEM, Sigma 500, Carl Zeiss, Germany). The thermogravimetric analysis (TGA) data for the E-HS-x composites were investigated separately by a TA Instruments DTA7300 (Waltham, MA, USA) at a heating rate of 20 °C min−1, from 20 °C to 700 °C, under a N2 atmosphere. The limiting oxygen index (LOI) values of the E-HS-x samples were characterized by using an oxygen index meter (JF-3, Jiangning, China), according to GB/T2406.2-2009. The Raman spectra and X-ray photoelectron spectroscopy (XPS) of the char residues were measured by spectrometer models, Dxr2xi and ESCALAB 250Xi (ThermoFisher, Waltham, MA, USA), respectively. The combustion properties of the EP and E-HS-x composites were investigated by a cone calorimeter (CC), iCone Classic (FTT., East Grinstead, UK), according to ISO-5660-1 standard. The thermal radiation power was 50 kW/m2 in CC testing, and the size of the sample was 100 mm × 100 mm × 4 mm.

3. Results and Discussion

3.1. Morphologies and Element Analysis of HCCP-SA

The morphology of HCCP and HCCP-SA was recorded by SEM and is summarized in Figure 1b,c, respectively; furthermore, the insets of Figure 1b,c show the element content measurement result of HCCP and HCCP-SA by EDS. It can clearly be seen from Figure 1b that the N, Cl and P element amounts of HCCP were 18.7%, 30.6% and 42.9%, respectively, which are close to the theoretical values of those of HCCP. When the nucleophilic substitution reaction of SA and HCCP was completed, the amount of N and P in the HCCP-SA molecule was 17.2% and 15.9%, respectively, as shown in Figure 1c.
It can be clearly observed from Figure 1b that HCCP presented an irregular loose network structure with many significant large pores, which can increase the specific surface area of HCCP and enable more functional groups to appear on the surface, thus, speeding up the nucleophilic substitution reaction between HCCP and SA. As shown in Figure 1c, the reactant HCCP-SA exhibited an irregular massive structure with a diameter of about 300 μm. In the process of the nucleophilic substitution reaction, the -NH2 and -OH groups in the SA molecule lost H atoms and were bonded to the conjugated aromatic ring of HCCP through the N-P and O-P bond [19], respectively, and finally formed a larger HCCP-SA block.

3.2. Structural Characterization of HCCP-SA

Figure 2a shows the FTIR spectra of the HCCP-SA and SA specimens. In the SA spectrum, the -CONH- motion peak centered at 3170 cm−1, and the primary amino peaks centered at 3395 cm−1 and 3295 cm−1 are easily observed. The stretching vibration absorption peaks of the C=C bond in the aromatic rings appeared from 1450 cm−1 to 1600 cm−1. Two vibrational absorption peaks appeared at 1280 cm−1 and 1353 cm−1 and were assigned to C-N and C-O, respectively [16]. For HCCP-SA, the peaks near 1189 cm−1 were assigned to the motions of P=N. Two new vibrational absorption peaks appeared at 1008 cm−1 and 1036 cm−1, which were assigned to P-N and P-O, respectively. Wide bands at 3351 cm−1 were assigned to the secondary amino. The FTIR results, especially the appearance of the absorption peaks of three new chemical bonds (i.e., P-O, P-N and secondary amino -NH-), primarily prove the successful preparation of HCCP-SA [20,21].
An 1H NMR analysis was executed to further analyze the structure of the HCCP-SA, and the NMR spectrums are presented in Figure 2b. It can easily be observed in Figure 2b that the chemical shift at 2.50 ppm was apportioned to DMSO-d6, which is used as a solvent for HCCP-SA. The multiple absorption peaks centered at 6.94 ppm (m, 2H) can be attributed to the coupling of H atoms corresponding to Ha and Hc in the benzene ring structure. The peaks centered at 7.31 (t, J = 6.6, 1H) and 7.98 ppm (d, J = 2.8, 1H) can be attributed to Hb and Hd in the benzene ring structure, respectively. A single strong peak at 9.32 ppm (s, 1H) was assigned to He atoms, which was the only secondary amino hydrogen atom (i.e., -NH-) in the molecule HCCP-SA [22,23]. As evidenced by the FTIR analyses in the previous section, the results of the 1H NMR analysis further prove that HCCP-SA was successfully prepared.

3.3. Thermal Stability of EP and E-HS-x Composites

In order to understand the thermal stability of the EP and E-HS-x composites, TG curves were tested, and these are summarized in Figure 3a. Under the nitrogen (N2) atmosphere, EP resins without flame retardant showed the initial decomposition temperature (T5 wt.%) and maximum decomposition temperature (Tmax) at 367.07 and 390.04 °C, and the char yield of the EP was 8.64% at 700 °C, respectively, which was mainly ascribed to the degradation reaction of the EP molecular chain. Compared with the pure EP resin, the char yield increased from 9.79% to 25.30% at ~700 °C, but the T5 wt.% of the E-HS-x composites reduced from 355.20 °C to 267.94 °C, respectively, when the amount of HCCP-SA increased from 5 to 20 phr. Moreover, as shown in Figure 3b, the Tmax of the E-HS-x composites reduced from 387.43 °C to 382.21 °C, and the weight loss speed of the EP and E-HS-x composites reduced from 1.49 to 0.95 mg·min−1. This indicates that the added HCCP-SA can decompose and absorb the energy of the system at a lower temperature, reducing the decomposition rate and Tmax of the E-HS-x composites, and thus, preventing further degradation of the EP resin matrix. In a word, the multiple-ring cross-linking network structure composed of the benzene ring, P-N, and N-O heterocycles is the main reason why E-HS-x composites have outstanding heat resistance [24].

3.4. Flame Retardancy of EP and E-HS-x Composites

LOI and UL-94 tests were carried out in order to understand the flame retardancy of the EP and E-HS-x composites. In the UL-94 test, the E-HS-x composites were made into splines in the size of 125 mm × 13 mm × 13 mm, according to ASTMD635. Table 1 presents the composition of the EP and E-HS-x composites and the determination results of LOI and UL-94 testing, in which the amount of diethylenetriamine (used as a curing agent) in EP was higher than that in E-HS-x. The reason for this is that the -NH- and -OH groups in HCCP-SA participated in the ring-opening curing reaction of the epoxy resin, acting as a curing agent, which is verified by FTIR in Section 3.2. The LOI value of EP resins without flame retardant was 22.4%, and no rating was measured in UL-94 testing.
It is worth noting that the LOI value of the E-HS-x composite increased from 24.1% to 25.3%, but no rating was measured in the UL-94 tests, as the amount of HCCP-SA increased from 5 to 15 phr. The reason for this is that the “blowing-out effect” of the N element and the synergistic flame-retardant effect of N-P were not obvious when the content of the P and N element was small, which is insufficient to significantly improve the flame retardancy of the E-HS-x composites. Encouragingly, as expected, the E-HS-x composites showed excellent flame retardancy when the additional amount of HCCP-SA reached 20 phr. The LOI value of the E-HS-20 composite was as high as 27.1%, which is 20.9% higher than that of pure EP. Meanwhile, there was no dripping in UL-94 testing, even reaching the V-2 rating. This is because the EP molecular chain was entangled with the branch chain on the HCCP-SA conjugate ring to form a network structure, which increased the viscosity of the EP composites and prevented the generation of droplets. In addition, the thickness and density of the surface protective layer of the E-HS-x composites were increased by the functional groups (P=O, P-O and P-N) generated by the ring-opening curing reaction, which could effectively prevent heat and O2 from contacting the internal substrate of the epoxy resin.
Cone calorimetry tests were implemented to simulate the combustion behavior of a real fire hazard. In order to achieve efficient measurement and comparability, EP, E-HS-10 and E-HS-20 were selected to measure the following six main indicators: HRR (peak heat release rate), THR (total heat release), T-HRR (time to peak heat release rate), total smoke release, average CO and CO2 yield. The results are shown in Figure 4 and Table 2. EP without flame retardant burns violently and releases a large amount of heat immediately after ignition, reaching the HRR of 798 kW/m2 in just 28 s (see Table 2 and Figure 4a). However, the HRR value of E-HS-10 and E-HS-20 was 609 (taking 34 s) and 537 (taking 36 s) kW/m2, respectively, which indicate that the combustion process is delayed by HCCP-SA linked to the EP molecular chain. It can be clearly observed from Figure 4b that the THR of the EP, E-HS-10 and E-HS-10composites was 127 MJ/m2, 105 MJ/m2 and 77 MJ/m2, respectively, which indicates that HCCP-SA can terminate the combustion process of E-HS-x in advance. The E-HS-20 composites presented a distinctly reduced HRR (32.7.0%) and THR (39.4%) compared with those of the pure EP. Furthermore, total smoke release (50%), average CO yield (46.8%) and average CO2 yield (31.3%) both gradually reduced with an increasing amount of HCCP-SA (see Table 2). Thus, the cone calorimeter results confirmed the beneficial effect of the HCCP-SA in boosting the flame-retardant performance of the E-HS-x composites.
In order to estimate the fire hazard of the composites, the fire growth rate (FGR) was defined and calculated by the following equation:
FGR = HRR T-HRR
A large number of studies have proved that the lower the FGR value, the higher the fire safety of the corresponding material [25,26,27]. The FGR values of the EP, E-HS-10 and E-HS-20 specimens were calculated, and these are listed in Table 2. As expected, the FGR value showed a significant downward trend, reducing from 28.5 kW·m−2·s−1 (EP) to 14.9 kW·m−2·s−1 (E-HS-20), which significantly demonstrated that the fire safety and flame retardancy of EP could be significantly improved by HCCP-SA after being embedded into the EP molecular chain through a ring-opening curing reaction.

3.5. Char Residue Analysis

3.5.1. Morphologies Analysis

Figure 5a–e illustrates the SEM of the char residues (CRE) acquired by the EP and E-HS-x composites. Figure 5a shows that the CRE of EP without flame retardant was finer and more uniform, which could theoretically prevent EP from continuing to burn. However, the CRE on the surface of EP was not dense enough, even accompanied by many cracks, which means that most of the combustible gases could pass through the cracks during the combustion process and reach the surface of the EP to intensify the combustion. At the same time, the heat could be quickly released to the interior of EP, so that the EP matrix could be rapidly decomposed by heat.
Encouragingly, it is evident in Figure 5b–e that EP added with HCCP-SA formed a dense CRE protective layer on the surface of the E-HS-x composite after combustion. Moreover, the thickness of CRE increased with the increase in HCCP-SA addition. H4P2O7 and H3PO4, with a powerful dehydration effect during combustion, were produced by the presence of abundant N and P containing functional groups in HCCP-SA. In particular, a great deal of CRE was obtained after combustion, which interfered with the thermal degradation process of the EP composites, as well as improved their thermal stability. The above results are confirmed by the Raman spectroscopy analysis of CRE, which is demonstrated in Figure 6.

3.5.2. Raman Spectroscopy Analysis

Raman spectroscopy analysis is a powerful method that can characterize the graphitization degree of CREs acquired by the combustion of organic macromolecular composites. A higher degree of graphitization indicates a better effect of the char layer. Let ID/IG (the integral area ratio of peak D and G) represent the graphitization degree of the char layer. The smaller ID/IG value indicates the stronger organic macromolecular composites’ flame retardancy or ablation resistance, as well as the higher graphitization degree of CREs [28,29]. The Raman spectra of carbon signals typically exhibit a D band at 1360 cm−1 and a G band at 1580 cm−1, and the D and G bands have different intensities. The calculated ID/IG value of CRE of EP, E-HS-10 and E-HS-20 shown in Figure 6a–c is 2.76, 2.41 and 1.93, respectively, demonstrating that the graphitization degree of CREs, ablation resistance and flame retardancy of E-HS-x composites can be effectively improved by HCCP-SA.

3.5.3. XPS Spectra Analysis

To confirm the bonding form of CREs—which is acquired by the combustion of the EP and E-HS-20 composites (defined as CRE-EP and CRE-20, respectively)—the XPS spectra of CREs were measured, as shown in Figure 7a–d. There were five obvious characteristic peaks at 530.3, 400.8, 284.8, 186.5 and 134.1 eV in the full survey spectra of CRE-20, which correspond to O1s, N1s, C1s, P2s and P2p, respectively, and the tested values are in good agreement with the reported values in references [30,31,32]. In the C1s fitting spectrum of CRE-EP, four kinds of C element bonding forms are as follows: C=O (288.4 eV), C-O (286.4 eV), C-N (285.6 eV) and C-C (284.8 eV) (see Figure 7b) [33,34,35,36]. It is noteworthy that there was no characteristic peak at 286.4 eV in the C1s fitting spectrum of CRE-20 (see Figure 7c). Moreover, the content of carbon oxide in CRE-EP (integral areas of C-O and C=O characteristic peak) was much higher than that in CRE-20 (integral areas of C=O characteristic peak), indicating that it became more difficult for EP to form carbon oxide owing to the addition of HCCP-SA. On the contrary, it could only form a denser and thicker protective carbon layer, which could separate heat, oxygen and the epoxy resin matrix. In Figure 7d, the three kinds of bonding forms of P element in the P2p fitting spectrum of CR-EP and CRE-20 (see) are P2O5/P2O72− (135.2 eV), PO32− (134.1 eV) and PO43− (133.4 eV), respectively [28]. The formation of these phosphides or phosphates absorbs a significant amount of heat, thus, reducing the system temperature and causing the epoxy resin to self-extinguish. These results are consistent with the analysis of FTIR shown in Section 3.2.

3.5.4. FTIR Spectra Analysis

The FTIR spectra of the thermal decomposition product, obtained by TG at 30 °C, 200 °C, 300 °C and 400 °C, respectively, were utilized to evaluate the thermal degradation process of the E-HS-x composites and were illustrated in Figure 8. When the temperature increased from 30 °C to 200 °C, there was no significant change in the intensity of the absorption peak on the two infrared spectral curves. In fact, there was only a small amount of mass loss before 200 °C under an air atmosphere, revealing that E-HS-20 is thermally steady below 200 °C. It is noteworthy that the peaks centered at 2922 cm−1, 2851 cm−1 and 3250 cm−1 belong to the C-H bond and –O-H [15], respectively. This demonstrates that prepared HCCP-SA was involved in the ring-opening curing reaction of epoxy resin. However, the intensity of almost all the absorption peaks decreased significantly, as the temperature was higher than 300 °C, indicating that the decomposition of the E-HS-x composites mainly occurred in this phase. Moreover, in the infrared spectral curves corresponding to 300 °C and 400 °C, the appearance of the new peaks centered at 1105 cm−1 and 1225 cm−1 was allocated to the vibration of the P-O-P and P=O bond, respectively. This indicates that the E-HS-x composite generates phosphate, polyphosphate (P2O5 and P4O10), H4P2O7 or H3PO4 with excellent thermal stability after combustion [19,21]. The process of the char layer formation was revealed by the appearances and disappearances of all the characteristic peaks.

3.6. Flame-Retardant Mechanism

Based on the above characterization results, we proposed a feasible flame-retardant mechanism of E-HS-x composites as follows. Firstly, HCCP-SA decomposes to generate NH3 and H2O when the temperature is above 267 °C. The vaporization of water absorbs a significant amount of heat and lowers the temperature of the system. Meanwhile, steam and NH3 can dilute O2 and combustible gas, hence, slowing down the combustion rate of EP composites. Secondly, when the temperature continues to increase, P-containing volatiles in the gas phase can further decompose into PO· and PO2· to clear the combustible H· or OH· radicals and achieve a quenching effect. Finally, the char produced by combustion is foamed by steam and NH3, resulting in the formation of a compact, thermally stable, and continuous char residue layer, which covers the surface of the EP matrix. The dense char layer with P2O5, P2O72−, PO32− and PO43− prevents heat, oxygen and mass loss. The flame-retardant mechanisms were ascribed to barrier effects in the condensed phase and the quenching effect in the gas phase.

4. Conclusions

In this study, HCCP-SA with a multiple-ring molecule containing P, O and N elements was successfully synthesized and utilized to prepare EP composites with the dual functions of heat resistance and flame retardancy (i.e., E-HS-x). HCCP-SA possessed objective functional groups and a multiple-ring structure (i.e., benzene ring, P-N and N-O heterocycles) through systematic characterizations. Furthermore, the preparation of epoxy resin inherent flame retardants was realized by embedding HCCP-SA in the molecular chain of the resin in the progress of its ring-opening curing reaction.
The E-HS-20 composites presented outstanding heat resistance and flame retardancy. Their initial decomposition temperature (T5 wt.%) was 267.94 °C and the max weight loss speed was only 0.95 mg·min−1, which was reduced by 27.0% and 36.2% than that of pure EP, respectively. In the UL-94 tests, the E-HS-20 composites presented a V-2 rating and LOI at 27.1%, which was increased by 20.9% compared to pure EP. Moreover, the E-HS-20 composites presented a distinctly reduced HRR (32.7%), THR (39.4%), total smoke release (50.0%) and ID/IG value (30%) of CRE-20 compared with those of pure EP. Compact char residue layers and suppressed volatile products indicate that HCCP-SA presents N-P synergistic flame retardancy and a blowing-out effect in the gaseous phase. Meanwhile, the heat, oxygen and mass loss was avoided by the dense char layer, and the flame-retardant mechanisms were due to the barrier effects in the condensed phase, as well as the quenching effect in the gas phase. Therefore, the best-prepared E-HS-x composite would be a suitable and potential candidate for heat-resistant and flame-retardant polymer materials, especially electronic packaging and sealing materials.

Author Contributions

Conceptualization and project administration, W.F.; methodology, Q.L. and Z.Y.; investigation, L.Z., L.K. and W.F.; software and writing, Z.L.; funding acquisition, W.F. and M.X. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (grant No. 21908086), the Youth Science Foundation of Jiangsu University of Technology (grant No. KYY18019, and KYY19031) and the Innovation and Entrepreneurship Training Project of Jiangsu University of Technology (grant No. KYH22042, and KYX22027).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

List of Notations

Full NameNotation
hexachlorocyclotriphosphazeneHCCP
salicylamideSA
prepared flame retardant (i.e., the product of HCCP and SA reaction)HCCP-SA
flame-retardant graphene oxidefGO
epoxy resinEP
epoxy resin curing products with different quality rates of HCCP-SA* E-HS-x
initial decomposition temperature T5 wt.%
maximum decomposition temperatureTmax
peak heat release rateHRR
time-to-peak heat release rateT-HRR
total heat releaseTHR
fire growth rateFGR
char residuesCRE
char residue acquired by combustion of EPCRE-EP
char residue acquired by combustion of E-HS-x* CRE-x
* x denotes parts of HCCP-SA used in per hundred parts of EP resins (phr).

References

  1. Dagdag, O.; Gouri, M.E.; Safi, Z.S. Flame retardancy of an intumescent epoxy resin containing cyclotriphosphazene: Experimental, computational and statistical studies. Iran. Polym. 2021, 30, 1169–1179. [Google Scholar] [CrossRef]
  2. Ning, K.; Zhou, L.L.; Zhao, B. A novel aminothiazole-based cyclotriphosphazene derivate towards epoxy resins for high flame retardancy and smoke suppression. Polym. Degrad. Stabil. 2021, 190, 109–651. [Google Scholar] [CrossRef]
  3. Liu, Q.Y.; Zhao, Y.L.; Gao, S.S. Recent advances in the flame retardancy role of graphene and its derivatives in epoxy resin materials. Compos. A-Appl. 2021, 149, 106539. [Google Scholar] [CrossRef]
  4. Ai, L.; Chen, S.; Zeng, J.; Liu, P.; Liu, W.; Pan, Y.; Liu, D. Synthesis and flame retardant properties of cyclophosphazene derivatives containing boron. Polym. Degrad. Stabil. 2018, 155, 250–261. [Google Scholar] [CrossRef]
  5. Delucchi, M.; Castellano, M.; Vicini, S.; Vita, S.; Finocchio, E.; Ricotti, R. A methodological approach for monitoring the curing process of fairing compounds based on epoxy resins. Prog. Org. Coat. 2018, 123, 20–26. [Google Scholar] [CrossRef]
  6. Molero, G.; Sue, H.J. Scratch behavior of model epoxy resins with different crosslinking densities. Materi. Design 2019, 182, 107965. [Google Scholar] [CrossRef]
  7. Lqab, C.; Ys, A.; Cz, A.; Xd, A.; Pl, A.; Gs, A. Improved flame retardancy of epoxy resin composites modified with a low additive content of silica-microencapsulated phosphazene flame retardant. Reacti. Funct. Polym. 2020, 148, 104485. [Google Scholar]
  8. Battig, A.; Markwart, J.C.; Wurm, F.R.; Schartel, B. Hyperbranched phosphorus flame retardants: Multifunctional additives for epoxy resins. Polym. Chem. 2019, 10, 4346. [Google Scholar] [CrossRef] [Green Version]
  9. Kamalipour, J.; Beheshty, M.H.; Zohuriaan-Mehr, M.J. Novel phosphonated hardeners derived from diamino diphenyl sulfone for epoxy resins: Synthesis and one-pack flame-retardant formulation alongside dicyandiamide. Polym. Degrad. Stabil. 2022, 199, 109917. [Google Scholar] [CrossRef]
  10. Petrakova, V.V.; Kireev, V.V.; Onuchin, D.V. Benzoxazine monomers and polymers based on 3,3′-Dichloro-4,4′-Diaminodiphenylmethane: Synthesis and Characterization. Polymers 2021, 13, 1421. [Google Scholar] [CrossRef]
  11. Baskaran, M.; Hashim, R.; Leong, J.Y.; Ong, Y.N.; Yhaya, M.F.; Sulaiman, O. Flame retardant properties of oil palm trunk particleboard with addition of epoxy resin as a binder and aluminium hydroxide and magnesium hydroxide as additives. Mater. Sci. 2019, 12, 138. [Google Scholar] [CrossRef]
  12. Fang, Y.; Qian, L.; Huang, Z. Synergistic barrier flame-retardant effect of aluminium poly-hexamethylenephosphinate and bisphenol-A bis(diphenyl phosphate) in epoxy resin. Polym. Int. 2017, 66, 719–725. [Google Scholar] [CrossRef]
  13. Wang, J. Synthesis and characterization of flame-retardant-wrapped carbon nanotubes and its flame retardancy in epoxy nanocomposites. Polym. Polym. Compos. 2021, 29, 835–843. [Google Scholar] [CrossRef]
  14. Xu, Z.; Deng, N.; Yan, L.; Chu, Z. Functionalized multiwalled carbon nanotubes with monocomponent intumescent flame retardant for reducing the flammability and smoke emission characteristics of epoxy resins. Polym. Adv. Technol. 2018, 29, 3002–3013. [Google Scholar] [CrossRef]
  15. Rhili, K.; Chergui, S.; Eldouhaibi, A.S. Hexachlorocyclotriphosphazene functionalized graphene oxide as a highly efficient flame retardant. ACS Omega 2021, 6, 6252–6260. [Google Scholar] [CrossRef]
  16. Qu, L.; Sui, Y.; Zhang, C.; Li, P.; Xu, B. Compatible cyclophosphazene-functionalized graphene hybrids to improve flame retardancy for epoxy nanocomposites. React. Funct. Polym. 2020, 15, 104697. [Google Scholar] [CrossRef]
  17. Pang, Q.; Deng, J.; Kang, F.; Shao, S. Effect of expandable Graphite/Hexaphenoxycyclotriphosphazene beads on the flame retardancy of silicone rubber foam. Mater. Res. Express. 2020, 7, 055308. [Google Scholar] [CrossRef]
  18. Fan, W.X.; Li, Z.F.; Yang, Z.; Ou, J.F.; Xiang, M.; Qin, Z.L. Improving the curing and flame retardancy of epoxy resin composites by multifunctional Si-containing cyclophosphazene derivatives. RSC Adv. 2022, 12, 13756–13764. [Google Scholar] [CrossRef]
  19. Wang, Q.; Xiong, L.; Liang, H. Synthesis of a novel polysiloxane containing phosphorus, and boron and its effect on flame retardancy, mechanical, and thermal properties of epoxy resin. Polym. Compos. 2016, 39, 807–814. [Google Scholar] [CrossRef]
  20. Shiu, B.C.; Wu, K.; Lou, C.W.; Lin, Q.; Lin, J.H. Synthesis of a Compound Phosphorus-Nitrogen Intumescent Flame Retardant for Applications to Raw Lacquer. Polymers 2021, 13, 2858. [Google Scholar] [CrossRef]
  21. Xin, F.; Zhai, C.; Guo, C.; Chen, Y.; Qian, L.; Chen, X. A wrapped nano-flame retardant composed of carbon nanotubes and phosphorus-nitrogen containing polymer: Synthesis, properties and flame-retardant mechanism. J. Polym. Res. 2018, 25, 201. [Google Scholar] [CrossRef]
  22. Liu, D.; Ji, P.; Zhang, T.; Lv, J.; Cui, Y. A bi-DOPO type of flame retardancy epoxy prepolymer: Synthesis, properties and flame-retardant mechanism. Polym. Degrad. Stabil. 2021, 190, 109629. [Google Scholar] [CrossRef]
  23. Fulmer, G.R.; Miller, A.J.M.; Sherden, N.H. NMR chemical shifts of trace impurities: Common laboratory solvents, organics, and gases in deuterate fulmer. Organometallics 2010, 29, 2176–2179. [Google Scholar] [CrossRef] [Green Version]
  24. Wu, M.; Yang, K.; Li, Y.; Rong, J.; Jia, D.; Jia, Z.; Naito, K.; Yu, X.; Zhang, Q. A pyridazine-containing phthalonitrile resin for heat-resistant and flame-retardant polymer materials. Polymers 2022, 14, 4144. [Google Scholar] [CrossRef]
  25. Wang, S.; Wang, S.; Shen, M. Biobased phosphorus Siloxane-containing polyurethane foam with flame-retardant and smoke-suppressant performances. ACS Sustain. Chem. Eng. 2021, 9, 8623–8634. [Google Scholar] [CrossRef]
  26. Wang, S.; Yang, X.; Li, Z. Novel eco-friendly maleopimaric acid based polysiloxane flame retardant and application in rigid polyurethane foam. Compos. Sci. Technol. 2020, 198, 108272. [Google Scholar] [CrossRef]
  27. Mariappan, T.; You, Z.; Hao, J. Influence of oxidation state of phosphorus on the thermal and flammability of polyurea and epoxy resin. Eur. Polym. 2013, 49, 3171–3180. [Google Scholar] [CrossRef]
  28. Lee, S.H.; Lee, S.G.; Lee, J.S.; Ma, B.C. Understanding the flame retardant mechanism of intumescent flame retardant on improving the fire safety of rigid polyurethane foam. Polymers 2022, 14, 4904. [Google Scholar] [CrossRef]
  29. Hardis, R.; Jessop, J.; Peters, F. Cure kinetics characterization and monitoring of an epoxy resin using DSC, Raman spectroscopy, and DEA. Compos. A Appl. Sci. Manuf. 2013, 49, 100–108. [Google Scholar] [CrossRef] [Green Version]
  30. Shi, Y.M.; Hamsen, C.; Jia, X.T.; Kim, K.K.; Reina, A.; Hofmann, M.; Hsu, A.L.; Zhang, K.; Li, H.N.; Jiang, Z.Y. Synthesis of few-layer hexagonal boron nitride thin film by chemical vapor deposition. Nano Lett. 2010, 10, 4134–4139. [Google Scholar] [CrossRef]
  31. Ko, W.Y.; Chen, C.Y.; Chen, W.H.; Lin, K.J. Fabrication of hexagonal boron nitride nanosheets by using a simple thermal exfoliation process. Chin. Chem. Soc. 2016, 63, 303–307. [Google Scholar] [CrossRef]
  32. Zheng, S.; Wang, B.; Zhang, X.; Qu, X. Amino acid-assisted sand-milling exfoliation of boron nitride nanosheets for high thermally conductive thermoplastic polyurethane composites. Polymers 2022, 14, 4674. [Google Scholar] [CrossRef]
  33. Yao, Y.; Wang, M.; Wu, H. Synthesis of waterborne epoxy resin with diethanolamine-assisted succinimide for improving the strand integrity of polyimide filament. Ind. Text. 2021, 50, 115. [Google Scholar] [CrossRef]
  34. Lin, F.; Li, W.; Tang, Y. High-performance polyimide filaments and composites Improved by O2 plasma treatment. Polymers 2018, 10, 695. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Kiattipornpithak, K.; Thajai, N.; Kanthiya, T. Reaction mechanism and mechanical property improvement of poly(lactic acid) reactive blending with epoxy resin. Polymers 2021, 13, 2429. [Google Scholar] [CrossRef] [PubMed]
  36. Yang, Z.; Meng, X.; Zhu, Y. Single-atom platinum or ruthenium on C4N as 2D high-performance electrocatalysts for oxygen reduction reaction. Chem. Eng. 2021, 426, 131–347. [Google Scholar] [CrossRef]
Figure 1. Synthesis route (a) and SEM (b,c) of HCCP and HCCP-SA; the inset shows the element content measurement result by EDS.
Figure 1. Synthesis route (a) and SEM (b,c) of HCCP and HCCP-SA; the inset shows the element content measurement result by EDS.
Polymers 15 00059 g001
Figure 2. FTIR (a) and 1H NMR (b) of SA and HCCP-SA. In (b), a, b, c, d and e represents the peak of H atom at a, b, c, d and e in the molecular formula of HCCP-SA, respectively.
Figure 2. FTIR (a) and 1H NMR (b) of SA and HCCP-SA. In (b), a, b, c, d and e represents the peak of H atom at a, b, c, d and e in the molecular formula of HCCP-SA, respectively.
Polymers 15 00059 g002
Figure 3. TG curves (a) and weight loss rate (b) of EP and E-HS-x.
Figure 3. TG curves (a) and weight loss rate (b) of EP and E-HS-x.
Polymers 15 00059 g003
Figure 4. HRR curves (a) and THR curves (b) for samples under an external heat flux of 50 kW/m2.
Figure 4. HRR curves (a) and THR curves (b) for samples under an external heat flux of 50 kW/m2.
Polymers 15 00059 g004
Figure 5. SEM of char residues of: (a) EP, (b) E-HS-5, (c) E-HS-10, (d) E-HS-15 and (e) E-HS-20.
Figure 5. SEM of char residues of: (a) EP, (b) E-HS-5, (c) E-HS-10, (d) E-HS-15 and (e) E-HS-20.
Polymers 15 00059 g005
Figure 6. Raman spectra of char residues: (a) EP, (b) E-HS-10 and (c) E-HS-20.
Figure 6. Raman spectra of char residues: (a) EP, (b) E-HS-10 and (c) E-HS-20.
Polymers 15 00059 g006
Figure 7. XPS spectra of CREs: (a) high-magnification XPS survey spectra of CRE-EP and CRE-20; (b) peak-fitting curves of C1s in CRE-EP; (c) peak-fitting curves of C1s in CRE-20; (d) peak-fitting curves of P2p in CRE-20.
Figure 7. XPS spectra of CREs: (a) high-magnification XPS survey spectra of CRE-EP and CRE-20; (b) peak-fitting curves of C1s in CRE-EP; (c) peak-fitting curves of C1s in CRE-20; (d) peak-fitting curves of P2p in CRE-20.
Polymers 15 00059 g007
Figure 8. FTIR spectra of thermal decomposition product of E-HS-20 at several representative temperatures.
Figure 8. FTIR spectra of thermal decomposition product of E-HS-20 at several representative temperatures.
Polymers 15 00059 g008
Table 1. Composition of EP and E-HS-x and flame-retardancy data.
Table 1. Composition of EP and E-HS-x and flame-retardancy data.
SpecimenEP/gDiethylene-Triaminen/phrHCCP-SA
/phr
LOI/%UL-94
DrippingUL-94 Rating
EP10015022.4 ± 0.10yesNo rating
E-HS-510010524.1 ± 0.22yesNo rating
E-HS-10100101024.3 ± 0.13noNo rating
E-HS-15100101525.3 ± 0.08noNo rating
E-HS-20100102027.1 ± 0.14noV-2
Table 2. Data from cone calorimetry tests.
Table 2. Data from cone calorimetry tests.
SamplesHRR
(kW/m2)
THR
(MJ/m2)
T-HRR (s)Total Smoke Release (m2/m2)Average CO Yield (kg/kg)Average CO2 Yield (kg/kg)FGR (kW/m2·s−1)
EP79812728240.0321.628.5
E-HS-1060910534160.0211.217.9
E-HS-205377736120.0171.114.9
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

Fan, W.; Li, Z.; Liao, Q.; Zhang, L.; Kong, L.; Yang, Z.; Xiang, M. Improving the Heat Resistance and Flame Retardancy of Epoxy Resin Composites by Novel Multifunctional Cyclophosphazene Derivatives. Polymers 2023, 15, 59. https://doi.org/10.3390/polym15010059

AMA Style

Fan W, Li Z, Liao Q, Zhang L, Kong L, Yang Z, Xiang M. Improving the Heat Resistance and Flame Retardancy of Epoxy Resin Composites by Novel Multifunctional Cyclophosphazene Derivatives. Polymers. 2023; 15(1):59. https://doi.org/10.3390/polym15010059

Chicago/Turabian Style

Fan, Wangxi, Zefang Li, Qin Liao, Lintong Zhang, Longjie Kong, Zhou Yang, and Meng Xiang. 2023. "Improving the Heat Resistance and Flame Retardancy of Epoxy Resin Composites by Novel Multifunctional Cyclophosphazene Derivatives" Polymers 15, no. 1: 59. https://doi.org/10.3390/polym15010059

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop