Skip to content
BY 4.0 license Open Access Published by De Gruyter June 16, 2020

Synthesis of a novel modified chitosan as an intumescent flame retardant for epoxy resin

  • Haihua Yu , Xiuhang Xu , Yunfei Xia , Mingzhen Pan , Nighat Zarshad , Bo Pang , Anis Ur Rahman , Min Wu and Henmei Ni EMAIL logo
From the journal e-Polymers

Abstract

According to the concept of fire life cycle assessment (LCA), a new type of intumescent flame retardant was designed and synthesized by chemically bonding chitosan, phosphorus pentoxide and melamine. The resultant compound, chitosan ethoxyl melamine phosphate (CEMP), was characterized by FTIR, 1H NMR, 31P NMR, XRD and SEM. The performance of CEMP and organic montmorillonite (OMMT) was evaluated in the substrate of epoxy resin (EP) with limited oxygen index (LOI), UL-94, cone calorimetric test (CCT), TGA and TG-IR. As a result, intumescent flame retardant EP (EP3) containing 30.6% LOI and V-0 rating was prepared by adding 3 wt% OMMT and 15 wt% CEMP. The CCT results indicated that CEMP and OMMT reduced the peak of heat release rate (PHRR) to about one fourth that of pure EP and total heat release (THR), 1/2. Decomposition of EP and EP3 was traced from 100 to 600°C by TG-IR.

1 Introduction

It is undoubtable that we have been living in the times of synthetic materials since the first synthetic fiber, Nylon, was born in Dupont lab in 1940s. Various synthetic materials that mostly originated from the paraffin have improved our life greatly to be more comfortable and convenient. However, meanwhile, the flammability of synthetic materials has caused lot of trouble to the human being. Every year, millions of lives are lost in the fire incidents. Hence, great efforts and endeavors have been made to develop new technology and new compounds to improve the flammability of synthetic materials. However, nowadays the main issues in the development of a new flame retardant are not limited in the range of technology, but rather in the field of environmental protection, namely how to relieve the environmental stress (1,2,3,4,5). The reason may date back to the last century. Since the 1960s, flame retardants (FRs) are invented and applied in synthetic materials, i.e., chemicals enable to both prevent combustion and delay the spread of fire after ignition. Among them, halogen-based fire retardants are very effective in reducing fire risk, i.e., the probability of occurrence of a fire, but there is a high fire hazard, that is, the probability of producing toxic, corrosive, obscuring smokes while performing the fire retardant action or when involved in a fully developed fire, beyond flashover, when the fire can no longer be extinguished (6,7,8,9,10,11). The halogen radicals access the vapor of combustion zone to quench the active radicals of H and OH, thereby stopping the oxidation reaction, and in turn preventing the heat generation (8,12).

Therefore, at the end of last century, the consumption of halogenated FRs occupied 50% of total FRs. However, in fact, the amount of FRs burnt in the incident fire is negligible. A large portion of FRs is finally disposed of as wastes accompanying its host goods of synthetic materials. Leachate of FRs from wastes and its derivatives created by various reactions in the environment are hazardous to animals and plants. Therefore, poly FR became a research hot spot, which was launched by the Dow Chemical Co. in 2011. It was a butadiene styrene brominated copolymer (13,14), a stable high molecular weight, non-PBT additive that can be used in XPS and expanded polystyrene (EPS) foams (12,15,16,17). Anyway, a ban on the halogenated FRs boosts the dramatic increase in organophosphate FR consumption. Alerting the potential hazards to the environment bore a new concept of fire life cycle assessment (LCA). LCA is applied to assess the overall environmental impact throughout the life of products containing flame retardants (18,19,20,21). Besides taking into account detection and extinction issues, smoke opacity and toxicity of effluents, it focuses on safety risks induced by flame retardants during process. These flame retardants migrated from goods are induced by the exposure via skin contact, risks by inhalation of substances emitted into the atmosphere during the life of product, by emission of toxic gases during incineration and by emanation of gaseous effluents during combustion in the event of accidental fire. These risks have now become the main issues in the development of new FRs in order to achieve the required high level of protection for health and safety.

Nevertheless, phosphorus is the unique element that is comprehensively evaluated as the best composition of FRs (12,22). Radicals of phosphorus oxides may act as the halogen radicals to quench the high-energy radicals of H and OH in the gas of combustion zone. Moreover, phosphorus oxides may form polyphosphate at high temperature, which is a nonflammable shell segregating the substrate and fire (23,24,25,26,27). However, according to LCA requirements, FRs with a high level of protection for health and safety should be highly efficient, eco-friendly, show no migration and certainly be nontoxic. Hence, in terms of migration, the small molecules of organophosphate compounds are first excluded from the recipes of FRs due to their migration property. Additionally, the use of inorganic phosphate salts, e.g., ammonium phosphate, should be avoided due to their water solubility, which will be leached in the life of products, thereby lowering the efficiency of FRs.

In terms of high efficiency, the technique of intumescence is a plausible one. The intumescent agent produces nonflammable and nontoxic gas, usually nitrogen, at high temperature. Nitrogen gas expands and volatizes at the expense of heat to lower the temperature. Meanwhile, nitrogen gas may dilute the concentration of flammable reagents. Besides, an additional effect is that nitrogen gas can foam the char and polyphosphate layer. Acidic phosphorus compounds created at high temperature catalyze the dehydrogenation reaction of polyolefin to form the nonflammable char. Sponge like char and polyphosphate layer are more effective at segregating substrate and fire, and weaken the transportation of mass and energy between the substrate and fire. Certainly, adding mineral particles is always an economic, eco-friendly and effective way to improve the efficiency of FRs. The main function of mineral particles in FRs is to dilute the concentration of polyolefin in the substrate. The lower concentration of decomposable polyolefin, the lower the concentration, or the slower the supply rate of flammable substances is. Therefore, the mineral particles are effective at suppressing the occurrence of smoke due to the full combustion. At last, a readily neglected factor is the compatibility of composition in the intumescent FR (IFRs), which severely impacts the efficiency of IFRs (28,29). It is said, if the intumescent agent and phosphorus compounds were simply mixed into substrate, the efficiency of FRs would be diminished due to the nonuniformity of components’ dispersion. Finally, in terms of eco-friendly aspects, the natural polymers are the best, e.g., cellulose, chitosan and alginate.

By far, a lot of intumescent flame retardants have been explored. However, few of them are satisfactory to the LCA requirements. For example, Fang et al. (30) synthesized polyhexamethylene guanidine phosphate and then physically mixed ammonium polyphosphate (PHMGP-APP) as the IFRs for cotton fabric. Chen et al. (31) used polyethylenimine (PEI) as intumescent agent to fabricate IFRs by physically mixing ammonium polyphosphate (APP) and fluorinated-decyl polyhedral oligomeric silsesquioxane (F-POSS). In these cases, the small molecule of phosphate compound was used. Hassan et al. (32) explored a chitosan-based IFR. The chitosan was first phosphate esterified, and then utilizing the surplus P-OH to form melamine phosphate salt. In this case, melamine ions will be easily leached during the life of products. Most recently, Borreguero et al. (33) developed a new method to synthesize IFR for the polyurethane foam. In this method, diethyl-N,N-bis(2-hydroxyethyl)aminomethyl phosphonate (Fyrol-6) was employed as initiator to conduct the ring opening polymerization of epoxy resin. The aminophosphate polyol exhibited good performance in the flame retardant polyurethane (PU) foam. The char residue increased from 7.79 to 22.13 wt% with the increase of polyol from 25 to 50 pph. Amazingly, Malucelli et al. (34) utilized the phosphoric acid in DNA to synthesize SiO2@DNA as IFR for cotton fabric. Reviewing these literatures, it can be found that LCA is gradually guiding the development of IFRs.

Chitosan is a natural material abundant in the nature. It has been widely used as a carbon resource in IFRs because it is a type of polyol (35,36,37). However, so far to our best knowledge, the chitosan-based IFRs satisfying LCA requirements have not been found. Therefore, in this paper, a new type of chitosan-based IFR will be prepared by covalently bonding melamine into chitosan via phosphoric covalent bridge, chitosan ethanol melamine phosphate (CEMP). The synthetic route of CEMP is shown in Scheme 1. Ethyl phosphoric ester (EPA) is also introduced to hydrophobilize the polymer, so as to increase the compatibility with polyolefins. Such a chemical structure is anticipated to be highly efficient due to the integrity of functional groups, no migration in the life of products, no toxicity and eco-friendly nature. The flame retardancy will be evaluated with epoxy resin (EP), which is the main material of glass composites.

Scheme 1 Synthetic route of chitosan ethoxyl melamine phosphate (CEMP).
Scheme 1

Synthetic route of chitosan ethoxyl melamine phosphate (CEMP).

2 Experimental

2.1 Materials

Chitosan (CS, viscosity 50–800 mPa s, degree of deacetylation 80–95%) was supplied by Zhengzhou Wanbo Chemical Products Co. Ltd (Henan, China). Epoxy resin (EP, DGEBA, commercial name: E-51, epoxy value: 0.51) and curing agent (polyether amine) were purchased from Hefei Jiangfeng Chemical Industry Co. Ltd (Anhui, China). Methyl sulfonic acid was bought from Aladdin Co. Ltd (Shanghai, China). Organic montmorillonite (OMMT, average particle size 70–80 nm) was purchased from Zhejiang Fenghong New Material Co. Ltd. Melamine was supplied by Maclean Biotechnology Co. Ltd. Ethanol, methanol, diethyl ether, acetone, toluene, pyridine and N,N-dimethylformamide were all provided by Sinopharm Chemical Reagent Co. Ltd (Shanghai, China). These reagents were of analytical grade and directly used without further purification.

2.2 Synthesis of CEMP

CEMP was synthesized according to the literature (38,39,40,41,42). At first, chitosan (CS, 2.0 g) was immersed in methyl sulfonic acid (12 mL) in a 100 mL four-necked round-bottomed flask equipped with a magnetic stirrer and a reflux condenser. The flask was placed in an ice-water bath under an inert atmosphere of N2. Then, phosphorus pentoxide (P2O5, 2.0 g) was slowly added under stirring for 3 h. The obtained phosphorylated chitosan (PCS) was washed successively with diethyl ether, acetone and methanol in order to remove the unreacted P2O5 and dried in a vacuum oven at 60°C for 12 h. Second, PCS (2.0 g) was dispersed in about 5 mL toluene in a 100 mL four-necked round-bottomed flask and subsequently thionyl chloride (1.0 mL) was added dropwise for 2 h under nitrogen atmosphere at 40°C. After ethanol was added, the reaction took place for 2 h. Washed with toluene and ethanol, the product was then dried in a vacuum oven at 60°C for 12 h. Third, acetylated chitosan (ACS) (2.0 g) and melamine (2.0 g) were charged into a 100 mL four-necked round-bottomed flask containing 20 mL of N,N-dimethylformamide (DMF) and 10 mL of the base pyridine. The reaction was carried out at 140°C for 4 h under nitrogen atmosphere. CEMP was obtained by washing the product with DMF and methanol.

2.3 Preparation of flame retardant EP

The recipe of flame retardant EP is shown in Table 1. It should be remarked that OMMT is usually used as inherent flame retardant additive because of its good fire resistant features (42,43), so OMMT was also used in this paper. Normally, the steps for preparing a flame retardant epoxy resin were as follows (44,45): at room temperature and pressure, the epoxy resin after adding a curing agent, CEMP and OMMT were completely blended with a small mixer according to the mass ratio of each component and then were injected into a custom-sized polymethyl methacrylate (PMMA) mold. Then the samples were put in a vacuum oven to defoam and polymerize at 60°C for 24 h. Finally, the obtained EP spline after drying was corrected by removing the corner burrs.

Table 1

Mass ratio of each component

SamplesEP (%)CEMP (%)OMMT (%)
EP10000
EP19253
EP287103
EP382153

2.4 Characterizations

CEMP was characterized by FTIR spectroscopy (Nicolet 6700, USA) using KBr disk and the scanning speed was 32 s−1. 1H and 31P NMR was conducted on the NMR spectrometer (AVANCE III HD 600 MHz, Bruker Instrument Corp., Germany) using dimethyl sulfoxide (DMSO-d6) as the solvent. Scanning electron microscopy (SEM, FEI Inspect F50, USA) was used to analyze its surface morphology and structural features of the modified product after the sample was sprayed with gold. The X-ray diffraction (XRD, Hitachi Instrument Corp., Japan) was performed on Ultima IV to analyze the crystal form of the product. The thermal stability of the sample was analyzed by thermogravimetry (TG, TG209 F3, Netzsch Instrument Corp., Germany) and the test temperature range was 25–800°C with a heating rate of 20°C/min under nitrogen atmosphere. The limiting oxygen index (LOI) values were measured on a JF-3 oxygen index meter from Jiangning Analysis Instrument Company. According to GB/T 2406-2008, the specimen dimensions were 100 × 6.5 × 3 mm. Vertical burning test (UL-94) was carried out on a CFZ-2-type instrument (Jiangning Analysis Instrument Company, China) and the specimens used for UL-94 were of dimensions 130 × 12.7 × 3 mm. Cone calorimetry (FTT2000, Suzhou Phoenix Instrument Company, China) was used to investigate the combustion behavior of flame retarded EP. According to ISO 5660, the size of samples was 100 × 100 × 3 mm and the heat flux was 35 kW/m. Thermogravimetry-infrared spectrometry (TG-IR) of flame retarded samples was measured on TG 209 F3 thermogravimeter (Netzsch Instrument Corp., Germany) which was interfaced to the Nicolet WAS10 FTIR spectrophotometer (Nicolet 6700, USA) to track and detect small molecules released during thermal degradation and the test temperature range was 25–600°C with a heating rate 20°C/min in an inert atmosphere. The morphology of char residue was analyzed by the FEI Inspect F50 SEM instrument (FEI, USA).

3 Results and discussion

3.1 FTIR analysis

In view of the previous studies on the attribution of the infrared spectrum of chitosan, which was a product of the de-N-acetylation of chitin and its degree of de-N-acetylation and crystallization was different, the infrared spectrum was slightly different (46,47). The FTIR mainly has two large spectral regions: the high frequency region with the wavenumber between 4,000 and 1,330 cm−1 is the characteristic frequency region. The low frequency region with a wavenumber of 1,330–500 cm−1 is referred to as a fingerprint region. In this paper, the infrared spectra of CS, PCS, ACS and CEMP were studied. It can be seen from Figure 1 that the CS had a strong absorption peak at 3,421 cm−1, which can be classified as the overlapping stretching vibration absorption peaks of O–H and N–H. The peak at 2,940 cm−1 belonged to the asymmetric stretching vibration absorption of C–H. The peak at 1,630 cm−1 is ascribed to the amide I stretching absorption. There were various peaks in the fingerprint area and the absorption peaks from 1,046 to 1,160 cm−1 coincided with the C–O stretching vibration absorption. There was a ring stretching vibration absorption peak at around 854 cm−1 and the out-of-plane bending of O–H was shown at 645 cm−1. After phosphorylation, the spectrum of PCS hardly changed. There was a new peak that appeared at 1,647 cm−1 due to the formation of NH3+. An absorption peak that appeared at 1,208 cm−1 may be a P═O stretching vibration absorption peak. It may be caused by P–O single bond stretching vibration absorption at 1,070 cm−1. There was a strong absorption peak at 778 cm−1 due to the in-plane bending of the P–O single bond. For the ACS after further acid chloride chlorination, it can be seen from Figure 1 that the absorption peaks of PCS and ACS showed no significant change. The absorption peak became stronger at 568 cm−1, indicating that it may be a stretching vibration absorption peak of P–Cl. As for CEMP, there was a stretching vibration absorption peak of N–H on melamine at 3,135–3,498 cm−1. The peaks at 1,627, 1,561 and 1,417 cm−1 were characteristic absorption peaks of the triazine ring. The absorption peak of P–Cl disappeared at 568 cm−1 and the above infrared characterization results can preliminarily determine that the target product probably has been synthesized.

Figure 1 FTIR spectra of CS, PCS, ACS and CEMP.
Figure 1

FTIR spectra of CS, PCS, ACS and CEMP.

3.2 1H and 31P NMR

In order to analyze the structure of the modified product, the 1H and 31P NMR of CS, CEMP was performed, respectively. As shown in Figure 2, there are many peaks corresponding to the various protons of hydroxyl groups and methylene groups in CS. However, when CS was esterified, only a broad peak was left in CEMP. The peaks at 2.50 and 3.36 ppm belonged to the absorption peak of solvent d6-DMSO. Two peaks at 6–7 ppm were characterized by the protons of NH2. It is clear that in CS, two peaks were ascribed to the coupling of adjacent methylene proton, whereas in CEMP, the two peaks of CS disappeared and a new peak at 6.35 ppm appeared. The peak at 6.35 ppm should be assigned to melamine. In order to further confirm the structure of reaction products, 31P NMR of PCS and CEMP was carried out, respectively (32,47,48,49). As shown in Figure 3, for PCS, the peak positions are negative, i.e., −0.17 and −1.07 ppm. These peaks are the characteristic peaks of phosphorylate esters. The peak at −1.07 ppm indicated the pure alkyl group linked to –O–P, while the peak at −0.17 ppm referred to the polarized alkyl group, thus showing that the chemical shift moved to the low field. According to the molecular structure of CS, it became clear that the esterified points were C-3 and C-6, respectively. In the 31P NMR of CEMP, those two peaks moved to −0.68 and 0.46 ppm. This result implied that there were strong polar groups attached on phosphorus atoms. All these results implied that CEMP was prepared successfully.

Figure 2 1H NMR spectra of CS and CEMP.
Figure 2

1H NMR spectra of CS and CEMP.

Figure 3 31P NMR spectra of PCS and CEMP.
Figure 3

31P NMR spectra of PCS and CEMP.

3.3 SEM analysis of CEMP

In a heterophase system, the specific area of additives is fairly important to the effective mixture. The big specific area is favorable to enlarge the contact area, in turn to improve the performance of additives. As shown in Figure 4, the surfaces of various compositions used in this paper were observed with SEM. The surface of pure chitosan was relatively smooth, whereas melamine had a regular crystal form. As the phosphorylation proceeded, the porous surface gradually appeared. The surface of CEMP became rougher, which was beneficial to generate noncombustible gas and stop the combustion further. Moreover, the porous network structure facilitated the formation of a porous carbon layer in the condensed phase and showed distinct flame retardant effect (50,51).

Figure 4 SEM photographs of CS, melamine, and CEMP.
Figure 4

SEM photographs of CS, melamine, and CEMP.

3.4 XRD analysis

It is well known that a large number of hydroxyl functional groups inside the chitosan resulted in the crystallization of CS. Therefore, the degree of crystallization is another index indirectly indicating the degree of phosphorylation (52,53). As shown in Figure 5, there were lots of sharp diffraction peaks indicating the high crystallinity of chitosan. However, for the target product CEMP, only a broad peak appeared at 2θ = 20°. It indicated that the crystalline structure of CS was completely destroyed due to the phosphorylation. It further proved that CEMP was successfully prepared.

Figure 5 XRD graphs of CS and CEMP.
Figure 5

XRD graphs of CS and CEMP.

3.5 TG analysis of CS and CEMP

It was analyzed by TGA in order to characterize the thermal behaviors of CS and CEMP, the relative experimental data of 5%, 50% and the maximum weight loss is shown in Table 2. It was clearly observed from Figure 6 that there was a platform at about 161°C because of the water evaporation process. For pure chitosan, its mass decreased sharply with increasing temperature. At 311°C the thermal degradation weight of CS reached 50%. The thermal degradation process of CS tended to be gentle at 528°C. However, there was a slight downward trend at about 800°C, at this time, the residual carbon content was about 26.8%. As for the modified chitosan, the weight of CEMP was also gradually reduced to a stable stage at 161°C like CS due to insufficient drying, and then, as the temperature increased, the magnitude of the sharp decrease in mass was smaller than that of unmodified chitosan. The temperature of 50% mass loss was 369°C, which was 58°C higher than that of CS. The residual carbon content of CEMP was about 31.2% at 800°C, which was 4.4% higher than that of CS. The possible reason was that the grafted phosphate and melamine groups were first cleaved, releasing HPO and PO radical groups during the thermal degradation of CEMP, and thus the chain reaction was interrupted by combining the H and OH released from the pyrolysis of the sample. In summary, it was certified that the overall thermal stability of chitosan after the modification was improved.

Table 2

TG data of CS and CEMP

SamplesT5% (°C)T50% (°C)Tmax (°C)Char residue at 800°C (%)
CS16131152826.8
CEMP16136960231.2
Figure 6 TGA curves of CS and CEMP.
Figure 6

TGA curves of CS and CEMP.

3.6 LOI and UL-94 vertical burning tests

The performance of EP in the presence of CEMP and OMMT or not was evaluated by LOI and UL-94 vertical burning tests (44,45,54). The results are summarized in Table 3. In the absence of flame retardant, flame retardant grade of EP fell off the standard, accompanied with a heavy black smoke and severe deformation in combustion. The cotton wool was ignited by the droplets of EP. The LOI was only 20.5%. However, the addition of CEMP and OMMT greatly improved the performance of EP in combustion. As shown in Table 3, first of all, the dripping was stopped. The LOI increased from 23.8 to 30.6%, as the amount of flame retardant additives increased from 8 to 18 wt%. Meanwhile, the flame retardant grade increased from V-1 to V-0 and the time of flame combustion of EP3 after two ignitions was both less than 2 s, far below that of pure EP showing that the reaction of burning was interrupted in time and had a better self-extinguishing effect in comparison to EP without flame retardant additives.

Table 3

LOI and UL-94 test results for flame retarded EP

SamplesFlame retardant ratio (%)LOI (%)t1 (s)t2 (s)UL-94 ratingDripping
EP020.5 ± 0.1>30>30NRYes
EP1823.8 ± 0.216.419.8V-1No
EP21326.4 ± 0.214.816.5V-1No
EP31830.6 ± 0.20.61.4V-0No

t1 is the first ignition and t2 is for the second. NR is no flame retardant rating.

Figure 7 shows the actual burning situation of samples after LOI and UL-94 tests. It was observed that, without CEMP and OMMT, the residue of EP completely disappeared due to dripping in the LOI test. The residue of EP containing CEMP and OMMT remained as porous char. The situation of samples was illustrated in the post UL-94 test. As shown in Figure 8, the sample of pure EP was seriously deformed and owing to smoke smudging, the surface became black. However, with the increase of CEMP, the combustion situation significantly improved. For example, sample deformation of EP3 was hardly observed after burning. It suggested that CEMP and OMMT interrupted the combustion reaction in a quick and effective manner. Many intumescent flame retardants have been reported. Qian et al. (55) prepared organic/inorganic IFRs with different phosphorus/nitrogen ratios through the sol–gel process and investigated the performance of the flame retardancy of epoxy resin, where the highest LOI value was only 29.5% with 10% portion of additives and it decreased with the addition of flame retardant. Moreover, the salt easily leaked into the substrate materials, in turn to result in the difficulty of distribution, again emphasizing the importance of the polymer type. Yan et al. (45) synthesized pentaerythritol phosphate melamine salt (PPMS)-functionalized expandable graphite (EG), a novel mono-component, as intumescent flame retardant for epoxy resin. However, the highest LOI value was only 27.5% with 20% PPMS-EG, in addition to inferior mechanical properties. Utilizing the physical blends of various FR components as IFR was advantageous for the sake of convenience. However, it was reported that the size and its distribution of each component in the substrate significantly affected the performance of both FR and mechanical property (56). The main reason was that the functional compositions could not come into contact with each other sufficiently, which was the disadvantage compared with the chemical bonding process. Meanwhile, it was proved that the flame retardant effect after adding OMMT and CEMP was greatly enhanced, thus reducing the fire risk.

Figure 7 Photographs of the samples after LOI and UL-94 tests (a: EP, b: EP1, c: EP2, and d: EP3).
Figure 7

Photographs of the samples after LOI and UL-94 tests (a: EP, b: EP1, c: EP2, and d: EP3).

Figure 8 Digital photographs of the char residues of flame retarded EP after the cone calorimetric test (a: EP, b: EP1, c: EP2, and d: EP3).
Figure 8

Digital photographs of the char residues of flame retarded EP after the cone calorimetric test (a: EP, b: EP1, c: EP2, and d: EP3).

3.7 Cone calorimetric test

Digital photographs of EP residues after the cone calorimetric tests are shown in Figure 9. It is clear that pure EP was completely burned without any swelling phenomenon and there were black and oily substances (Figure 8a) (41,57,58). With the addition of CEMP and OMMT, different degrees of expanded char layer appeared. As the amount of CEMP attained 15%, the expanded char layer increased significantly (Figure 8c and d). On the other hand, the smoke was also significantly suppressed. These results, at least apparently, indicated the flame retardant effect exhibited on EP.

Figure 9 Cone calorimetry curves (a: HRR, b: THR, c: SPR, and d: TSP).
Figure 9

Cone calorimetry curves (a: HRR, b: THR, c: SPR, and d: TSP).

The combustion process was evaluated by a cone calorimeter. Figure 9 depicts the heat release (HRR), total heat release (THR), smoke release (SPR) and the total smoke emission (TSP) during the combustion of flame retardant modified EP. The relevant parameters including the average yields of CO (av-COY) and CO2 (av-CO2Y) as well are listed in Table 4. It is obvious that, as shown in Table 4, the overall parameters of pure EP were significantly higher than those of composite EP. Figure 9a and b indicate that, once the pure epoxy resin was ignited, a lot of heat was released in a very short time, ca. 200 s. PHRR was 603.1 kW m−2 and THR, 86.4 MJ m−2. It implied that the temperature was high, which accelerated the decomposition of PE. In contrast, PHRR of samples, EP1, EP2 and EP3, were 159.7, 152.5 and 133.0 kW m−2, almost one fourth of HRR of EP. Correspondingly, THRs were 49.9, 44.6 and 35.1 MJ m−2, nearly half of THR of EP. Besides, the heat release of composite EP lasted for about 600 s. These results indicated that the temperature of composite EP was much lower than that of EP during the combustion, thus retarding the decomposition of EP. In principle, the polymer is inflammable. However, it decomposes to the smaller organic molecules due to the high temperature. The small organic molecules evaporate and mix with oxygen in air. Hence, the dramatic oxidation reactions take place and release a huge amount of heat. The heat once again induces the decomposition of polymer, thereby accelerating the combustion. One of the targets of an intumescent flame retardant is to absorb the heat by expansion of the char layer, and the char layer expansion slows down the heat transmission from the gas phase to the decomposing polymer. Accordingly, the above results indicated that the CEMP did work during the combustion.

Table 4

Cone calorimeter data of flame retarded EP

SamplesTTI (s)PHRR (kw m−2)THR (MJ m−2)av-COY (kg/kg)av-CO2Y (kg/kg)TSP (m2)
EP108 ± 5603.1 ± 2886.4 ± 7.00.01 ± 0.010.84 ± 0.0115.4 ± 0.2
EP122 ± 2159.7 ± 2549.9 ± 6.50.05 ± 0.010.74 ± 0.019.3 ± 0.2
EP221 ± 2152.5 ± 2544.6 ± 6.80.06 ± 0.010.59 ± 0.016.8 ± 0.2
EP320 ± 2133.0 ± 2535.1 ± 6.90.08 ± 0.010.58 ± 0.015.0 ± 0.2

t1 is the first ignition and t2 is for the second. NR is no flame retardant rating.

Addition of the inorganic OMMT is to reflect radiant heat from the flame and act as a catalyst for charring to accumulate on the surface of substrate. The char layer created by the CEMP further blocks the heat transfer in the polymer. Figure 9c and d show SPR and TSP. It is clear that, without OMMT and CEMP, SPR of pure EP reached 15 m2, whereas TSP of EP3 with the amount of 15% CEMP and 3% OMMT was only 5.0 m2, one third of the pure EP. It indicated that the effect of smoke suppression by OMMT and CEMP was significant. In addition, Table 4 also lists the ignition time (TTI) of EP and its composites. The ignition time of pure EP was 108 s, in contrast, after adding the flame retardant, the ignition time was drastically reduced to 20–30 s. One of the functions of CEMP was to accelerate the thermal decomposition of CS and EP, so as to create an expanded char layer as early as possible. The short TTI indicated that the thermal decomposition reaction of CS and EP took place readily under the catalysis of phosphoric acid. The high av-COY and low av-CO2Y supported the above results. As shown in Table 4, the values of av-COY and av-CO2Y of pure EP were 0.01 and 0.84 kg/kg, respectively, whereas av-COY of EP composite increased as the amount of CEMP increased. It indicated that the combustion of EP composite was incomplete due to the low temperature of combustion. All these results of experiment suggested that CEMP were effective at suppressing the combustion further.

3.8 TG analysis

The thermal stability of EP composites was evaluated by TGA. Figure 10 shows TG curves with the amount of intumescent flame retardant at 8, 13, and 18 wt%. The temperatures of 5%, 50% and the maximum weight loss are listed in Table 5 (59). It is clear that the temperature of both 5% and the maximum weight loss of pure EP were lower than those of EP composites, whereas the temperature of 50% weight loss inverted. The temperature of 5% weight loss is mainly ascribed to the dehydration process, while the temperature of 50% weight loss is corresponding to the pyrolysis. The high pyrolysis temperature of pure EP was expected, completely consistent with the initial anticipation to CEMP.

Figure 10 TGA curves of flame retardant EP.
Figure 10

TGA curves of flame retardant EP.

Table 5

TG data of flame retarded EP

SamplesT5% (°C)T50% (°C)Tmax (°C)Char residue at 800°C (%)
EP1203704363.6
EP113635247514.1
EP214135849017.1
EP315336250822.2

The temperature of maximum weight loss refers to the difficulty in polymer degradation. As shown in Table 5, the temperature of EP3 was 508°C, 72°C higher than that of pure EP. It indicated that heat transfer was effectively blocked by the char layer created by the CEMP and EP matrix. The amount of char residues supported this result. At the temperature of 800°C, the content of residual char increased, as the amount of CEMP increased. The char residue of EP3 was 22.2%, almost more than five times as much as that of pure EP.

3.9 TG-IR analysis

In order to further trace the pyrolysis process and elucidate the flame retardant mechanism, the volatile substances of decomposition were simultaneously evaluated with IR during TG measurement. The IR spectra of pure EP and flame retardant modification at different temperatures are shown in Figure 12. It should be noticed that the spectra were automatically calibrated by machine, subsiding the peak of CO2 absorption that inevitably existed in air, while the peaks of other absorption were going up and down due to the volatile substances including a lot of CO and CO2, even though it did not affect the results of the analysis. The assignment of main peaks was listed as follows (60): O–H and N–H stretch vibrations (3,400–4,000 cm−1), –CH3 and –CH2 (2,800–3,000 cm−1), CO2 symmetric vibration (2,300–2,400 cm−1), CO stretch vibration (2,100–2,200 cm−1), carbonyl stretch vibration C═O (1,640–1,840 cm−1), C–O ether bond (1,140 cm−1) and aromatics (1,300–1,800 cm−1, typically 1,600 cm−1). For OMMT and CEMP, the noticeable peaks are: Si–O (664 cm−1), P═O double bond (1,252 cm−1) and the asymmetric stretching vibration of P–O–P (1,180 and 899 cm−1). Characteristic expanding peak of triazine ring in melamine (1,400–1,630 cm−1) overlapped with those of benzene in EP (61).

As shown in Figure 11, the peaks at 3,400–4,000 cm−1 disappeared at the lower temperature, <300°C, together with the disappearance of 2,300–2,400 cm−1 (CO2 and CO). Typically, the peaks reflecting the existence of alkyl group at 2,879 and 2,918 cm−1 also disappeared. This result is coincident with the above results that EP combusted quickly. In contrast, as for EP3, those peaks appeared even at 600°C. Especially, the peaks in the vicinity of 2,357 cm−1 were very high during the decomposition and moved to 2,200 cm−1 as the temperature increased. It implies that the concentration of CO increased as the temperature increased. CEMP decomposed at 100–200°C because the P–O–P band at 1,180 cm−1 appeared at that temperature. It abated as the temperature increased, while the peak at 1,252 cm−1 (P═O) gradually appeared. The peak at 899 cm−1 reflects the substituent benzene or triazine. It is the highest at 300°C, indicating that the benzene and triazine rings appeared in the volatile gas at 300°C (11,62,63). Multiple peaks (1,300–2,000 cm−1) reflected the alkyl, benzyl, melamine, etc. of various substances. These peaks attained the maximum value accompanying the appearance of the highest peak at 899 cm−1. It implies that the sublimation of melamine took place at 300°C.

Figure 11 TG–IR curves of pure EP and EP3.
Figure 11

TG–IR curves of pure EP and EP3.

Comprehensively, the above results indicated that (1) the pyrolysis of pure EP took place completely because little CO was observed. Decomposition of pure EP gave rise to black smoke that consisted of short chains. (2) Decomposition of pure EP was completed at 300°C because the detectable volatile substances disappeared at that temperature. (3) Phosphoric ester of CEMP decomposed at first to produce phosphoric acid at 100°C. (4) Sublimation of melamine occurred at 300°C. It not only absorbed heat, but also decreased the concentration of combustible gas. (5) The thermal degradation of EP containing CEMP and OMMT was incomplete because a huge amount of CO2 was released. Hence, it was favorable to form a char layer so as to suppress the release of smoke . These results further confirmed that the addition of CEMP increased the thermal stability of EP and suppressed the occurrence of smoke.

3.10 SEM analysis of char residue

Intumescent effect was exhibited in char residues at post-combustion. Hence, the morphology of char residues was observed by SEM (64,65). The micrographs of char residues are shown in Figure 12, where the char residues of pure EP exhibited smooth surfaces, whereas there were many minipores that appeared on the surfaces of EP3 char residues. It confirmed the intumescent effect of CEMP.

Figure 12 SEM images of EP and EP3 char residues.
Figure 12

SEM images of EP and EP3 char residues.

4 Conclusions

According to the concept of fire life cycle assessment (LCA), i.e., assess the overall environmental impact throughout the life of products containing flame retardants, a new type of intumescent flame retardant was designed and synthesized by chemically bonding chitosan, phosphorus pentoxide and melamine to prevent the flame retardant components from leaking into the environment in the application. The structure of chitosan ethoxyl melamine phosphate was characterized by FT-IR, 1H NMR and 31P NMR. 1H NMR indicated that many sharp OH proton peaks disappeared in the CEMP 1H NMR spectrum, instead, a big broad peak appeared. A similar phenomenon was observed in the XRD diagrams of chitosan and CEMP, indicating that the crystalline structure of chitosan was destroyed during the phosphorylation reaction. In the 31P NMR spectrum, two peaks, i.e., 0.46 and −0.46 ppm, appeared in CEMP. Both of them reflected that the polar group was adjacent to the P–O bond. All results indicated that CEMP was successfully prepared. Due to the hardly soluble substrates, the surface of CEMP beads was observed with SEM. The porous surface was confirmed, implying the good miscibility with substrate.

The performance of intumescent flame retardant chitosan ethoxyl melamine phosphate (CEMP) and OMMT was evaluated in the epoxy resin (EP) substrate. As a result, with the addition of 3 wt% OMMT and 15 wt% CEMP (EP3), LOI and UL-94 tests evaluated the flame retarded EP to possess smoke suppression with LOI value, 30.6%, and rank V-0 rating without dripping. The results of cone calorimetric test (CCT) indicated that adding CEMP and OMMT significantly reduced PHRR, THR, TSP and CO emission. For example, with EP3, PHRR was one fourth that of pure EP, while THR was just 1/2. These results implied that the intumescent effect did work in the combustion. TGA showed that the char residue yield of EP3 at 800°C was 22.2%, whereas EP gave 3.6%. TG-IR traced the release of volatile substances stemming from the decomposition of EP. Accordingly, it was suggestive that CEMP decomposition commenced at 100°C to create phosphoric acid and sublimation of melamine occurred at 300°C. Intensive peaks of alkyl groups and a faint peak of CO2 implied that incomplete combustion of pure EP gave rise to smoke that consisted of decomposed short chains. In contrast to the completed combustion of EP below 300°C, EP3 exhibited a strong thermal stability, namely even at 600°C, the volatile substances were detectable. This result was consistent with TGA results that Tmax of EP3 was 508°C, much higher than 436°C of EP. The intumescent effect of CEMP was confirmed by SEM of char residues at post CCT. The surface of EP3 char residues clearly showed the porous structure, in contrast, the surface of the EP char residue was smooth.

Acknowledgments

This work was supported by the Fundamental Research Funds for the Central Universities (2242017k1G006) and financially assisted by the Nanjing University of Science and Technology Committee (Platform No. 2014-030002).

References

(1) Krokosz A, Jarosiewicz M, Marczak A, Bukowska B. Changes in the activities of antioxidant enzymes and reduced glutathione level in human erythrocytes exposed to selected brominated flame retardants. Chemosphere. 2019;227:93–99.10.1016/j.chemosphere.2019.04.008Search in Google Scholar PubMed

(2) Lv K, Cao D, Gao W, Fu J, Wu J, Fu J, et al. Presence and human exposure assessment of organophosphate flame retardants (OPEs) in indoor dust and air in Beijing, China. Ecotoxi Env Saf. 2019;169:383–91.10.1016/j.ecoenv.2018.11.038Search in Google Scholar PubMed

(3) Lohmann R, Khairy MA. Organophosphate flame retardants in the indoor and outdoor dust and gas-phase of Alexandria, Egypt. Chemosphere. 2019;220:275–85.10.1016/j.chemosphere.2018.12.140Search in Google Scholar PubMed

(4) Stapleton HM, Meeker JD. House dust concentrations of organophosphate flame retardants in relation to hormone levels and semen quality parameters. Env Health Perspect. 2010;118:318–23.10.1289/ehp.0901332Search in Google Scholar PubMed PubMed Central

(5) Wang Y, Kim UJ, Lia W, Kannan K. Occurrence of and human exposure to organophosphate flame retardants/plasticizers in indoor air and dust from various microenvironments in the United States. Env Intl. 2019;125:342–9.10.1016/j.envint.2019.01.065Search in Google Scholar PubMed

(6) Zaikov GE, Lomakin SM. Ecological Aspects of Flame Retardancy. Zeist, Holland: VSP International Science Publishers; 1999. p. 170.10.1533/9781845698584.5.419Search in Google Scholar

(7) Hahn O, Kemmlein S, Jann O. Emissions of organophosphate and brominated flame retardants from selected consumer products and building materials. Atmos Env. 2003;37:5485–93.10.1016/j.atmosenv.2003.09.025Search in Google Scholar

(8) Morgan AB, Wilkie CA, editors. Fire retardancy of polymeric materials. Boca Raton, FL: CRC Press; 2009. p. 91.10.1201/9781420084009Search in Google Scholar

(9) Zhuo AA, Huang GB, Wang LQ, Wang X. MaterChemPhys. 2011;130:714–20.10.1016/j.matchemphys.2011.07.047Search in Google Scholar

(10) Liu Y, Wang Q. The investigation on the flame retardancy mechanism of nitrogen flame retardant melamine cyanurate in polyamide 6. J Polym Res. 2009;16:583–9.10.1007/s10965-008-9263-6Search in Google Scholar

(11) Zheng Z, Zhang L, Liu Y, Wang H. A facile and novel modification method of β-cyclodextrin and its application in intumescent flame-retarding polypropylene with melamine phosphate and expandable graphite. J Polym Res. 2016;23:1–17.10.1007/s10965-015-0905-1Search in Google Scholar

(12) Battig A, Velencoso MM, Markwart JC, Schartel B, Wurm FR. Molecular firefighting-how modern phosphorus chemistry can help solve the challenge of flame retardanc. Angew Chem Int Ed. 2018;57:10450–67.10.1002/anie.201711735Search in Google Scholar

(13) Bhaskar T, Murai K, Matsui T, Brebu MA, Uddin MA, Muto A, et al. Studies on thermal degradation of acrylonitrile-butadiene-styrene copolymer (ABS-Br) containing brominated flame retardant. J Anal Appl Pyrolysis. 2003;70:369–81.10.1016/S0165-2370(02)00183-3Search in Google Scholar

(14) Brebu M, Bhaskar T, Murai K, Muto A, Sakata Y, Uddin MA. The individual and cumulative effect of brominated flame retardant and polyvinylchloride (PVC) on thermal degradation of acrylonitrile-butadiene-styrene (ABS) copolymer. Chemosphere. 2004;56:433–40.10.1016/j.chemosphere.2004.04.002Search in Google Scholar PubMed

(15) SM, Lomakin, Zaikov GE. Ecological aspects of polymer flame retardation. Int J Polym Mum. 2000;47:61–78.10.1080/00914030008033335Search in Google Scholar

(16) Lomakin SM, Zaikov GE. Ecological aspects of polymer flame retardation. J Vinyl Addit Techn. 1999;5:12–20.10.1002/vnl.10301Search in Google Scholar

(17) Song L, Qian XD, Wang BB, Hu Y, Yuen RKK. Synthesis of organophosphorus modified nanoparticles and their reinforcements on the fire safety and mechanical properties of polyurea. Mater Chem Phys. 2013;139:443–9.10.1016/j.matchemphys.2013.01.018Search in Google Scholar

(18) Betts KS. New thinking on flame retardants, fire-LCA model: furniture case study. 2008;116:210–3.10.1289/ehp.116-a210Search in Google Scholar PubMed PubMed Central

(19) Guillaume E, Chivas C, Sainrat A, Barbosa V. Assessment of risks and benefits in the use of flame retardants in upholstered furniture in continental Europe. Fire Saf J. 2009;44:801–7.10.1016/j.firesaf.2009.03.009Search in Google Scholar

(20) Simonson M, Andersson P, Blomqvist P. Fire-LCA model: furniture case study. Proc Flame Retardants. 2004;116:15–26.Search in Google Scholar

(21) Simonson M, Andersson P, Berg MVD. Cost benefit analysis model for fire safety methodology and TV (DecaBDE) case study. SP Rep. 2006;28:35–52. ISBN: 91-85533-85513-85530.Search in Google Scholar

(22) Han Z, Alongia J, Bourbigot S. Intumescence: tradition versus novelty. A comprehensive review. Prog Polym Sci. 2015;5:28–73.10.1016/j.progpolymsci.2015.04.010Search in Google Scholar

(23) Huo S, Yang S, Wang J, Cheng J, Zhang Q, et al. A liquid phosphorus-containing imidazole derivative as flame-retardant curing agent for epoxy resin with enhanced thermal latency, mechanical, and flame-retardant performances. J Hazard Mater. 2020;386:121984.10.1016/j.jhazmat.2019.121984Search in Google Scholar PubMed

(24) Huo S, Liu Z, Li C, Wang X, Cai H. Synthesis of a phosphaphenanthrene/benzimidazole-based curing agent and its application in flame-retardant epoxy resin. Polym Degrad Stab. 2019;163:100–9.10.1016/j.polymdegradstab.2019.03.003Search in Google Scholar

(25) Fang F, Huo S, Shen H, Ran S, Wang H, Song P, et al. A bio-based ionic complex with different oxidation states of phosphorus for reducing flammability and smoke release of epoxy resins. Compos Commun. 2020;17:104–8.10.1016/j.coco.2019.11.011Search in Google Scholar

(26) Cheng J, Wang J, Yang S, Zhang Q, Huo S, Zhang Q, et al. Benzimidazolyl-substituted cyclotriphosphazene derivative as latent flame-retardant curing agent for one-component epoxy resin system with excellent comprehensive performance. Compos Part B: Eng. 2019;177:1–11.10.1016/j.compositesb.2019.107440Search in Google Scholar

(27) Fang F, Ran SY, Fang ZP, Song PG, Wang H. Improved flame resistance and thermo-mechanical properties of epoxy resin nanocomposites from functionalized graphene oxide via self-assembly in water. Compos Pt B-Eng. 2019;165:406–16.10.1016/j.compositesb.2019.01.086Search in Google Scholar

(28) Zhan J, Pan Y, Pan H, Wang W, Tang G, Song L, et al. Effect of fully biobased coatings constructed via layer-by-layer assembly of chitosan and lignosulfonate on the thermal, flame retardant, and mechanical properties of flexible polyurethane foam. ACS Sustain Chem Eng. 2016;4:1431–8.10.1021/acssuschemeng.5b01423Search in Google Scholar

(29) Song L, Wang X, Xing WY, Lu HD, Hu YA. Mater Chem Phys. 2011;125:536–41.10.1016/j.matchemphys.2010.10.020Search in Google Scholar

(30) Fang F, Xiao D, Zhang X, Meng Y, Cheng C, Bao C, et al. Construction of intumescent flame retardant and antimicrobial coating on cotton fabric via layer-by-layer assembly technology. Surf Coat Technol. 2015;276:726–34.10.1016/j.surfcoat.2015.05.023Search in Google Scholar

(31) Li X, Chen S, Li Y, Sun J. Intumescent flame-retardant and self-healing superhydrophobic coatings on cotton fabric. ACS Nano. 2015;9:4070–6.10.1021/acsnano.5b00121Search in Google Scholar PubMed

(32) Hassan M, Nour M, Abdelmonem Y, Makhlouf G, Abdelkhalik A. Synergistic effect of chitosan-based flame retardant and modified clay on the flammability properties of LLDPE. Polym Degrad Stab. 2016;133:8–15.10.1016/j.polymdegradstab.2016.07.011Search in Google Scholar

(33) Velencoso MM, Borreguero AM, Rodríguez JF, Serrano A, Carrero MJ, Ramos MJ. Synthesis of aminophosphonate polyols and polyurethane foams with improved fire retardant properties. J Appl Polym Sci. 2019;44780:47780.10.1002/app.47780Search in Google Scholar

(34) Malucelli G, Ortelli S, Blosi M, Zanoni I, Costa AL. NanoTiO2@DNA complex: a novel eco, durable, fire retardant design strategy for cotton textiles. J Colloid Interfaces Sci. 2019;546:174–83.10.1016/j.jcis.2019.03.055Search in Google Scholar PubMed

(35) Kirkland C, Laufer G, Cain AA, Grunlan JC. Clay–chitosan nanobrick walls: completely renewable gas barrier and flame-retardant nanocoatings. ACS Appl Mater Interfaces. 2012;4:1643–9.10.1021/am2017915Search in Google Scholar PubMed

(36) Krain T, Bocz K, Marosi G. Effect of particle size of additives on the flammability and mechanical properties of intumescent flame retarded polypropylene compounds. Int J Polym Sci. 2015;2015:7.Search in Google Scholar

(37) El-Tahlawy K, Eid R, Sherif F, Hudson S. Chitosan: a new route for increasing the efficiency of stannate/phosphate flame retardants on cotton. J Text Inst. 2008;99:157–64.10.1080/00405000701570773Search in Google Scholar

(38) Gospodinova N, Grelard A, Jeannin M, Chitanu GC, Carpov A, Thiéry V, et al. Efficient solvent-free microwave phosphorylation of microcrystalline cellulose. Green Chem. 2002;4:220–2.10.1039/b200295gSearch in Google Scholar

(39) Zhan J, Song L, Nie S, Hu Y. Combustion properties and thermal degradation behavior of polylactide with an effective intumescent flame retardant. Polym Degrad Stab. 2009;94:291–6.10.1016/j.polymdegradstab.2008.12.015Search in Google Scholar

(40) Liang S, Neisius M, Mispreuve H, Naescher R, Gaan S. Flame retardancy and thermal decomposition of flexible polyurethane foams: Structural influence of organophosphorus compounds. Polym Degrad Stab. 2012;97:2428–40.10.1016/j.polymdegradstab.2012.07.019Search in Google Scholar

(41) Jiang S, Shi Y, Qian X, Zhou K, Xu H, Lo S, et al. Synthesis of a novel phosphorus- and nitrogen-containing acrylate and its performance as an intumescent flame retardant for epoxy acrylate. Ind & Eng Chem Res. 2013;52:17442–50.10.1021/ie4028439Search in Google Scholar

(42) Liu X, Guo J, Tang W, Li H, Gu X, Sun J. Enhancing the flame retardancy of thermoplastic polyurethane by introducing montmorillonite nanosheets modified with phosphorylated chitosan. Compos Part A: Appl Sci Manuf. 2019;119:291–8.10.1016/j.compositesa.2019.02.009Search in Google Scholar

(43) Fang K, Li J, Ke C, Zhu Q, Tao K, Zhu J, et al. Intumescent flame retardation of melamine-modified montmorillonite on polyamide 6: enhancement of condense phase and flame retardance. Polym Eng & Sci. 2011;51:377–85.10.1002/pen.21804Search in Google Scholar

(44) Ding H, Wang J, Liu J, Xu Y, Chen R, Wang C, et al. Preparation and properties of a novel flame retardant polyurethane quasi-prepolymer for toughening phenolic foam. J Appl Polym Sci. 2015;132:1–8.10.1002/app.42424Search in Google Scholar

(45) Yan L, Xu Z, Wang X, Deng N, Chu Z. Preparation of a novel mono-component intumescent flame retardant for enhancing the flame retardancy and smoke suppression properties of epoxy resin. J Therm Anal Calorim. 2018;134:1505–19.10.1007/s10973-018-7810-xSearch in Google Scholar

(46) Hu S, Song L, Pan H, Hu Y, Gong X. Thermal properties and combustion behaviors of flame retarded epoxy acrylate with a chitosan based flame retardant containing phosphorus and acrylate structure. J Anal Appl Pyrolysis. 2012;97:109–15.10.1016/j.jaap.2012.06.003Search in Google Scholar

(47) Shi X, Jiang S, Hu Y, Peng X, Yang H, Qian X. Phosphorylated chitosan-cobalt complex: a novel green flame retardant for polylactic acid. Polym Adv Technol. 2018;29:860–6.10.1002/pat.4196Search in Google Scholar

(48) Hu S, Song L, Pan H, Hu Y. Thermal properties and combustion behaviors of chitosan based flame retardant combining phosphorus and nickel. Ind Eng Chem Res. 2012;51:3663–9.10.1021/ie2022527Search in Google Scholar

(49) L, Si, Chan SY, Lee KI, Ng PF. A novel boron-nitrogen intumescent flame retardant coating on cotton with improved washing durability. Cellulose. 2017;25:843–57.10.1007/s10570-017-1577-2Search in Google Scholar

(50) Dong L, Hu C, Song L, Huang X, Chen N, Qu L. A large-area, flexible, and flame-retardant graphene paper. Adv Funct Mater. 2016;26:1470–6.10.1002/adfm.201504470Search in Google Scholar

(51) Oshima T, Taguchi S, Ohe K, Baba Y. Phosphorylated bacterial cellulose for adsorption of proteins. Carbohydr Polym. 2011;83:953–8.10.1016/j.carbpol.2010.09.005Search in Google Scholar

(52) Wu W, Zhifan Z, Weihong Z, Jiao M. Hydrothermal synthesis of 4ZnO·B2O3·H2O/RGO hybrid material and its flame retardant behavior in flexible PVC and magnesium hydroxide composites. Appl Surf Sci. 2017;425:896–904.10.1016/j.apsusc.2017.07.101Search in Google Scholar

(53) Yuan B, Hu Y, Chen X, Shi Y, Niu Y, Zhang Y, et al. Dual modification of graphene by polymeric flame retardant and Ni(OH)2 nanosheets for improving flame retardancy of polypropylene. Compos Part A: Appl Sci Manuf. 2017;100:106–17.10.1016/j.compositesa.2017.04.012Search in Google Scholar

(54) Wang X, Pang H, Chen W, Lin Y, Zong L, Ning G. Controllable fabrication of zinc borate hierarchical nanostructure on brucite surface for enhanced mechanical properties and flame retardant behaviors. ACS Appl Mater Interfaces. 2014;6:7223–35.10.1021/am500380nSearch in Google Scholar PubMed

(55) Qian X, Song L, Bihe Y, Yu B, Shi Y, Hu Y. Organic/inorganic flame retardants containing phosphorus, nitrogen and silicon: preparation and their performance on the flame retardancy of epoxy resins as a novel intumescent flame retardant system. Mater Chem Phys. 2014;143:1243–52.10.1016/j.matchemphys.2013.11.029Search in Google Scholar

(56) Bocz K, Krain T, Marosi G. Effect of particle size of additives on the flammability and mechanical properties of intumescent flame retarded polypropylene compounds. Int J Polym Sci. 2015;1–7.10.1155/2015/493710Search in Google Scholar

(57) Xu J, Liu C, Qu H, Ma H, Jiao Y, Xie J. Investigation on the thermal degradation of flexible poly(vinyl chloride) filled with ferrites as flame retardant and smoke suppressant using TGA-FTIR and TGA-MS. Polym Degrad Stab. 2013;98:1506–14.10.1016/j.polymdegradstab.2013.04.016Search in Google Scholar

(58) Ming G, Chen S, Sun Y, Wang Y. Flame retardancy and thermal properties of flexible polyurethane foam containing expanded graphite. Combust Sci Technol. 2016;189:793–805.10.1080/00102202.2016.1251910Search in Google Scholar

(59) Si M, Feng J, Hao J, Xu L, Du J. Synergistic flame retardant effects and mechanisms of nano-Sb2O3 in combination with aluminum phosphinate in poly(ethylene terephthalate). Polym Degrad Stab. 2014;100:70–78.10.1016/j.polymdegradstab.2013.12.023Search in Google Scholar

(60) Jia P, Zhang M, Hu L, Zhou J, Feng G, Zhou Y. Thermal degradation behavior and flame retardant mechanism of poly(vinyl chloride) plasticized with a soybean-oil-based plasticizer containing phosphaphenanthrene groups. Polym Degrad Stab. 2015;121:292–302.10.1016/j.polymdegradstab.2015.09.020Search in Google Scholar

(61) Brehme S, Köppl T, Schartel B, Altstädt V. Competition in aluminium phosphinate-based halogen-free flame retardancy of poly(butylene terephthalate) and its glass-fibre composites. e-Polymers. 2014;14:193–208.10.1515/epoly-2014-0029Search in Google Scholar

(62) Ni P, Fang Y, Qian L, Qiu Y. Flame-retardant behavior of a phosphorus/silicon compound on polycarbonate. J Appl Polym Sci. 2018;135:1–8.10.1002/app.45815Search in Google Scholar

(63) Ding H, Xia C, Wang J, Wang C, Chu F. Inherently flame-retardant flexible bio-based polyurethane sealant with phosphorus and nitrogen-containing polyurethane prepolymer. J Mater Sci. 2016;51:5008–18.10.1007/s10853-016-9805-ySearch in Google Scholar

(64) Feng C, Zhang Y, Liang D, Liu S, Chi Z, Xu J. Influence of zinc borate on the flame retardancy and thermal stability of intumescent flame retardant polypropylene composites. J Anal Appl Pyrolysis. 2015;115:224–32.10.1016/j.jaap.2015.07.019Search in Google Scholar

(65) Zhang S, Yan Y, Wang W, Gu X, Li H, Li J. Intercalation of phosphotungstic acid into layered double hydroxides by reconstruction method and its application in intumescent flame retardant poly (lactic acid) composites. Polym Degrad Stab. 2018;147:142–50.10.1016/j.polymdegradstab.2017.12.004Search in Google Scholar

Received: 2020-01-23
Revised: 2020-03-31
Accepted: 2020-04-01
Published Online: 2020-06-16

© 2020 Haihua Yu et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 22.5.2024 from https://www.degruyter.com/document/doi/10.1515/epoly-2020-0036/html
Scroll to top button