Next Article in Journal
Assessing the Aesthetic Performance of CAD/CAM Provisional Restorative Materials: A Spectrophotometric Analysis
Next Article in Special Issue
Multifaceted Impact of Lipid Extraction on the Characteristics of Polymer-Based Sewage Sludge towards Sustainable Sludge Management
Previous Article in Journal
Hydrothermal Carbonization of Biomass for Electrochemical Energy Storage: Parameters, Mechanisms, Electrochemical Performance, and the Incorporation of Transition Metal Dichalcogenide Nanoparticles
Previous Article in Special Issue
Effects of Recycled Polyethylene on Natural Rubber Composite Blends Filled with Aluminum Trihydroxide and Polyurethane Waste: Mechanical and Dynamic Mechanical Properties, Flammability
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Facile Synthesis of Bis-Diphenylphosphine Oxide as a Flame Retardant for Epoxy Resins

by
Yan Li
1,
Chong Tian
2,
Guiqing Cheng
2,
Chunhui Li
2 and
Zhongwei Wang
2,*
1
School of Biological and Chemical Engineering, Qingdao Technical College, Qingdao 266555, China
2
College of Materials Science and Engineering, Shandong University of Science and Technology, Qingdao 266590, China
*
Author to whom correspondence should be addressed.
Polymers 2024, 16(18), 2635; https://doi.org/10.3390/polym16182635
Submission received: 2 July 2024 / Revised: 10 September 2024 / Accepted: 11 September 2024 / Published: 18 September 2024

Abstract

:
A phosphorus-containing compound, (oxybis(4,1-phenylene))bis(phenylphosphine oxide) (ODDPO), was successfully synthesized and used as a flame retardant for epoxy resin (EP). The results demonstrated that EP/ODDPO, containing 1.2 wt% phosphorus, achieved a vertical burning V-0 rating, with a limited oxygen index value of 29.2%, indicating excellent flame retardancy. Comprehensive evaluations revealed that ODDPO exhibited both gas-phase and condensed-phase flame-retardant effects on EP, with a particularly notable barrier effect. In addition, the incorporation of ODDPO had a minimal negative impact on the glass transition temperature (Tg) and thermal stability of the EP matrix. Compared to unmodified EP (EP-0), the Tg value and initial decomposition temperature of EP/ODDPO-1.2 decreased by only 7.6 °C and 10.0 °C, respectively. Moreover, the introduction of ODDPO significantly improved the hydrophobicity and water absorption resistance of epoxy materials, which is attributed to ODDPO’s rigidity and symmetric structure, reducing water molecule permeation. Furthermore, the dielectric properties of ODDPO-modified EP samples were strengthened compared to EP-0, due to the ODDPO’s symmetric structure reducing the polarity of the matrix. The above results indicated that ODDPO serves as an excellent flame retardant while enhancing other properties of the EP matrix, thereby contributing to the preparation and application of high-performance epoxy materials.

Graphical Abstract

1. Introduction

Epoxy resin (EP), one of the important thermosetting polymer materials, is valued for its excellent mechanical properties, chemical resistance, insulation performance, ease of processing, and thermal stability [1,2,3,4,5,6,7]. It has become essential in modern industry and is widely used in fields such as electronics [8,9,10], coatings [11,12,13,14], adhesives [15,16], construction [17,18], aerospace [19,20,21], and automotive industries [22,23,24,25]. However, EP has the drawback of high flammability, which limits its application in fireproof fields. Therefore, it is necessary and crucial to improve the flame retardancy of epoxy thermosets.
Halogenated flame retardants, widely used to enhance the flame resistance of polymer materials, pose potential risks to the environment and human health due to the release of toxic gases during combustion. Consequently, many industries have banned their use for the sake of environmental concerns. Today, researchers are focused on creating eco-friendly halogen-free flame retardants. Phosphorus-based flame retardants, known for their high efficiency, environmental friendliness, and low toxicity, have become the preferred flame-retardant solution in many industries [26]. Among them, 9,10-dihydro-9-oxa-10-phosphaphenanthrene-10-oxide (DOPO) and diphenylphosphine oxide (DPO) have proven suitable for flame-retardant research across various polymers. Nevertheless, DPO offers significant advantages over DOPO due to its structural benefits, including its higher phosphorus content and the absence of P–O bonds [27,28]. This allows for a lower dosage in polymer modification and improved thermal stability in the modified polymers. Thus, DPO is a highly advantageous and environmentally friendly flame retardant. Wei et al. reported that EP/DPO-0.9 (DPO-modified EP with 0.9 wt% phosphorus) reached a vertical burning (UL-94) V-0 rating, with a limiting oxygen index (LOI) value of 30.5%, showing a prominent flame retardancy [27]. However, compared to unmodified EP, the addition of DPO also led to a noticeable decrease in the glass transition temperature (Tg) and thermal stability of the EP matrix. Furthermore, DPO can form hydrogen bonds with water molecules due to the presence of hydrophilic functional groups in its chemical structure, making it prone to moisture absorption and inconvenient to use [29].
To address the moisture absorption of DPO, increasing focus has been placed on studying DPO derivatives that reduce hygroscopicity by incorporating hydrophobic, or larger, more complex, substituents into their molecular structure. This modification decreases the molecule’s polarity and increases steric hindrance, thereby reducing moisture absorption. Zhang et al. synthesized a phosphorus-containing flame retardant, diphenyl-(2,5-dihydroxyphenyl)-phosphine oxide (DPDHPP), by reacting DPO with benzoquinone and incorporating it into EP [30]. With a phosphorus content of 2.11 wt%, the DPDHPP-modified EP achieved a V-0 rating in the UL-94 test. However, the initial decomposition temperature decreased by 29 °C compared to the unmodified EP. Yuan et al. synthesized a phosphorus-nitrogen compound named DPO-PR from DPO, phenol, paraformaldehyde, and dicyandiamide, and investigated its flame-retardant application in epoxy [31]. The EP modified with DPO-PR demonstrated excellent self-extinguishing properties and achieved a V-0 rating. Nevertheless, the initial decomposition temperature of the modified EP was significantly reduced by 54 °C compared to the unmodified EP. In our previous work, we also extensively investigated the synthesis and application of DPO derivatives. Hydroxy(4-hydroxyphenyl)methyl)diphenylphosphine oxide (DPO-H) was synthesized and utilized as a flame retardant in EP [32]. With a phosphorus content of 0.5 wt%, the DPO-H-modified EP achieved a V-0 rating. However, compared to the unmodified sample, the modified EP exhibited a decrease of 31 °C in Tg and a decrease of 13.7 °C in initial decomposition temperature. Additionally, we explored the application of the commercial flame retardant 1,2-bis(diphenylphosphinoyl)ethane (EDPO) in epoxy [33]. The findings revealed that, at a phosphorus content of 0.9 wt%, the modified EP attained a UL-94 V-0 rating. However, compared to the unmodified EP, the Tg decreased by 16 °C, and the initial decomposition temperature decreased by 23 °C.
The studies above indicate that most DPO derivatives can address the problem of high hygroscopicity through structural optimization. However, their enhancements in Tg and thermal stability have been modest. It has been reported that incorporating a rigid skeleton into the molecular chain of fire retardants can significantly enhance the Tg value of the EP matrix by restricting thermoset chain movement [34]. Additionally, studies in the literature have demonstrated that flame retardants containing a high content of aromatic groups can improve the thermal stability of epoxy matrices [35]. Furthermore, the polarity of additives has been shown to impact the dielectric properties of EP, a critical factor in electronic applications, where higher polarity molecules typically lead to decreased dielectric properties [36,37].
In this work, (oxybis(4,1-phenylene))bis(phenylphosphine oxide) (ODDPO), a novel phosphorus-containing compound with a high content of aromatic groups, symmetrical structure, and specific rigidity, was successfully synthesized and applied for the flame-retardant modification of EP. Comprehensive properties, including Tg, thermal stability, flame retardancy, flame-retardant mechanism, hydrophobicity, water absorption, and dielectric properties, were investigated in detail.

2. Materials and Methods

2.1. Materials

Diphenyl ether, dichlorophenylphosphine (DCPP), and 4,4′-diaminodiphenyl sulfones (DDS) were purchased from Shanghai Aladdin Bio-Chem Technology Co., Ltd. (Shanghai, China). Aluminum chloride (AlCl3) and hydrochloric acid (HCl) were provided by Qingdao Zhengye Reagent Instrument Co., Ltd. (Qingdao, China). The diglycidyl ether of bisphenol A (DGEBA, grade E-44) was supplied by Zhenjiang Danbao Resin Co., Ltd. (Zhenjiang, China). The dichloromethane was obtained from Tianjin Beilian Fine Chemical Co., Ltd. (Tianjin, China).

2.2. Synthesis of ODDPO

The organophosphorus compound ODDPO was synthesized as follows: diphenyl ether (17.1 g, 0.1 mol) and DCPP (46.53 g, 0.26 mol) were mixed in a 250 mL three-necked flask and reacted at 20 °C for 15 h in an N2 atmosphere with AlCl3 (36.0 g, 0.27 mol) serving as a catalyst. Subsequently, the solution was cooled to room temperature and hydrolyzed with a 10 vol% HCl solution. The resulting mixture was extracted with dichloromethane, and the organic layer was washed with distilled water until neutral. Finally, the organic layer was subjected to vacuum distillation to obtain the target product with a yield of 74.1%. Scheme 1 shows the synthesis route of ODDPO.

2.3. Preparation of Epoxy Resin Thermosets

The ODDPO and DGEBA were placed in a 250 mL three-necked flask and stirred under an N2 atmosphere at 120 °C until ODDPO completely dissolved. Then, DDS was added to the above solution, and the temperature was raised to 180 °C with continued stirring until DDS dissolved completely. Subsequently, the mixture was subjected to vacuum pumping for 3–5 min to remove air from the solution, thereby preventing the formation of bubbles in the prepared epoxy samples. Finally, the hot solution was poured into the polytetrafluoroethylene mold for curing. Curing conditions were 2 h at 120 °C followed by 2 h at 180 °C. The resulting epoxy samples were labeled EP/ODDPO-x, where x denotes the mass percentage of phosphorus in the modified epoxy. Additionally, an epoxy sample without flame retardants was prepared as a control group under the same conditions and labeled EP-0. Table 1 lists the detailed formulations of epoxy resins with different phosphorus contents.

2.4. Characterization

Fourier transform infrared (FTIR) spectra were recorded using a Nicolet 380 infrared spectrometer (Thermo Fisher Scientific, Waltham, MA, USA) over the spectral range from 400 to 4000 cm−1.
The structure of the ODDPO was tested by a 400 M Bruker Avance (Bruker, Bremen, Germany) nuclear magnetic resonance (NMR) spectrometer using DMSO-d6 as a solvent.
Dynamic mechanical analysis (DMA) was conducted using a DMA1 analyzer (Mettler Toledo, Greifensee, Switzerland) at a frequency of 1 Hz. The specimen with a dimension of 50 mm × 5 mm × 3 mm was heated from room temperature to 250 °C at a rate of 3 °C/min.
Thermogravimetry analysis (TGA) was carried out on an HTG-1 thermogravimetric analyzer (Beijing Hengjiu Experimental Equipment Co., Ltd., Beijing, China). The sample (10–15 mg) was heated at a rate of 10 °C/min from 100 °C to 800 °C under an N2 atmosphere.
The LOI value was measured using a BG-5207 oxygen index meter (Suzhou Bengao Instrument Co., Ltd., Suzhou, China) in accordance with the ISO 4589-2017 standard [38], with a specimen size of 130 mm × 6.5 mm × 3 mm. The UL-94 test was conducted on a BG-5210 instrument (Suzhou Bengao Instrument Co., Ltd., China) in accordance with the GB/T 2408-2008 [39], with a specimen size of 130 mm × 13 mm × 3 mm.
The cone calorimeter test was performed on an FTT cone calorimeter (East Grinstead, UK) under an external heat flux of 50 kW/m2 according to ISO 5660-1:2015 standard [40], with a specimen size of 100 mm × 100 mm × 3 mm.
The morphology of carbon residue after the cone calorimeter test was observed by Apreo S HiVac scanning electron microscope (SEM, Thermo Fisher Scientific, USA) with an electron voltage of 2 kV. The samples were coated with gold in a vacuum before observation.
Thermogravimetric analysis/infrared spectrometry (TG-IR) was measured using a PerkinElmer 8000/Nicolet IS20 instrument (Shanghai PerkinElmer Enterprise Management Co., Ltd., Shanghai, China). The sample was heated at a rate of 10 °C/min from 50 °C to 800 °C under an N2 atmosphere with a flow rate of 100 mL/min.
The static contact angle was measured using a JC2000D1 contact angle meter (Shanghai Zhongchen Digital Technology Equipment Co., Ltd., Shanghai, China) at room temperature. A water droplet of approximately 3 μL was placed on the surface of the sample. After the droplet stabilized, the contact angle was calculated using a five-point fitting method.
The water absorption performance was evaluated through a boiling water experiment. A sample measuring 150 mm × 13 mm × 3 mm was immersed in boiling water. Every 6 h, the sample was taken out, dried, and weighed. The water absorption rate was calculated using the following equation:
Water   absorption   rate = ( W W 0 1 ) × 100 %
where W is the weight of the sample after being placed in boiling water for a certain time, and W0 is the initial weight of the sample.
Dielectric constant and dielectric loss were measured at frequencies of 1 MHz and 10 MHz at room temperature using an AS2853 digital display high-frequency dielectric measuring instrument (Shanghai Aiyi Electronic Equipment Co., Ltd., Shanghai, China).

3. Results and Discussion

3.1. Characterization of ODDPO

The chemical structure of ODDPO was characterized by 1H NMR, 13C NMR, 31P NMR, and FTIR spectra, as shown in Figure 1. In the 1H NMR spectrum, the chemical shifts at 8.67 ppm and 7.44 ppm (d, 2H) correspond to the hydrogen peaks at position c of P-H. It should be noted that due to the influence of the phosphorus atom, the hydrogen at position c undergoes coupling, splitting into a doublet with a coupling constant of 491.73 Hz. The chemical shifts at 7.75–7.67 ppm (m, 8H) can be assigned to the hydrogen peaks at positions a and b on the benzene ring. In addition, the chemical shifts at 7.62–7.50 ppm (m, 6H) and 7.22–7.18 ppm (m, 4H) are attributed to the hydrogen peaks at positions e, f, and d on the benzene ring, respectively. In the 13C NMR spectrum, eight chemical shifts appeared at 159.52, 133.42, 133.29, 133.00, 130.86, 129.62, 128.48, and 119.94 ppm, corresponding to the eight different carbon atoms on the benzene ring. In the 31P NMR spectrum, phosphorus peaks appear at 19.65 ppm and 16.61 ppm. It is reported that phosphinylidene compounds may display prototropic tautomerism [41]. We supposed that isomers of ODDPO are formed during synthesis, showing P(=O)H to P−OH tautomerism. Therefore, ODDPO exhibits two signals in the 31P NMR spectrum. The above results confirmed that the synthesized compound is the target product ODDPO.
The product structure was further analyzed by the FTIR spectrum (Figure 1d). The peak at 3485 cm−1 corresponds to the stretching vibration of the hydroxyl group [42], which results from P(=O)H to P−OH tautomerism during the synthesis of ODDPO, consistent with the NMR results. The peak at 3055 cm−1 was attributed to the C-H stretching vibration on the benzene ring [43]. The peaks at 1590 cm−1 and 1493 cm−1 were the C-C stretching vibrations on the benzene ring [43]. The P-H stretching vibration can be found at 2326 cm−1 [44]. The peaks of P=O, Ar-O-Ar, and P-Ph appeared at 1250 cm−1, 1180 cm−1, and 749 cm−1, respectively [42,45,46]. In summary, the synthesized compound exhibited the structural characteristics of the target product ODDPO.

3.2. Thermomechanical Properties of Epoxy Composites

DMA was used to test the thermomechanical properties of the cured epoxy samples. The results in Table 2 reveal that the crosslinking density ( v e ) values of ODDPO-modified epoxies are mostly reduced in comparison with the unmodified sample (EP-0). The ODDPO synthesized in this work is a rigid bifunctional flame retardant. The ve of the EP is influenced by the molecular structure of this flame retardant. At low concentrations, the rigid structure of ODDPO may reduce the mobility of the polymer chains and interfere with their crosslinking, resulting in a decrease in the ve of EP/ODDPO-0.9 [47]. As the ODDPO content increases, the two ends of the bifunctional additive gradually participate in the polymer’s crosslinking reaction, promoting the formation of new crosslinking points, which leads to a smaller reduction in the ve of EP/ODDPO-1.2. As a result, the effects of ODDPO’s rigid structure and functional groups on ve oppose each other. With a further increase in ODDPO concentration, the influence of its rigid structure becomes more dominant than that of the functional groups, leading to a further decrease in the ve of EP/ODDPO-1.5. As shown in Figure 2a, the Tg of the ODDPO-modified EP samples was lower than that of EP-0. The Tg of EP-0 was 184.3 °C, while the Tg values of EP/ODDPO-0.9 and EP/ODDPO-1.2 were 175.2 °C and 176.7 °C, respectively. The Tg value is influenced by the v e and the inherent rigidity of the EP thermosets. On the one hand, the rigid structure of ODDPO favored an increase in the Tg of the matrix with the addition of the flame retardant [48]. On the other hand, a reduction in ve of the polymer matrix lowers its Tg value. The combined effects of these factors led to a slight decrease in the Tg of the epoxy. With a further increase in ODDPO content, the significant decrease in v e led to a reduction in the Tg of EP/ODDPO-1.5 to 171.5 °C. Additionally, as shown in Figure 2b, the storage modulus (E′) of EP/ODDPO-1.2 and EP/ODDPO-1.5 was higher than that of EP-0, possibly because the flame retardant enhanced the rigidity within the epoxy to a certain extent. The v e can be calculated through the E′ and Tg described by Equation (2) [49], as follows:
v e = E / 3 R T
where E′ is the storage modulus at temperature T, R is 8.314 J mol−1 K−1, and T is defined as Tg + 40 K.
Although the addition of ODDPO slightly reduced the Tg of the epoxy matrix, the Tg value of the ODDPO-modified sample was still higher than that of the epoxy modified by commercial flame retardants (Table 2). Comparative studies reported that the Tg for EP/DOPO-1.2 (DOPO-modified EP with 1.2 wt% phosphorus) and EP/DPO-1.2 (DPO-modified EP with 1.2 wt% phosphorus) were 148 °C and 154 °C, respectively [27]. EP/ODDPO-1.2 exhibited a 19.4% enhancement in Tg compared to EP/DOPO-1.2 and a 14.7% increase compared to EP/DPO-1.2, indicating that ODDPO has a less pronounced effect on the Tg of the epoxy than both DOPO and DPO.

3.3. Thermal Stability of Epoxy Composites

The thermal stability of the cured epoxy samples was evaluated using TGA. Figure 3 presents the TGA and DTG curves for the EP-0 and ODDPO-modified EP samples. The char residue at 800 °C, the maximum mass loss rate (MMLR), and the temperatures corresponding to 5 wt% mass loss (T5%), 10 wt% mass loss (T10%), and maximum mass loss (Tmax) are summarized in Table 2. DTG curves show that all EP samples exhibited a single-stage thermal weight-loss process (Figure 3b), indicating that the introduction of the ODDPO did not alter the thermal degradation behavior of the epoxy matrix. Meanwhile, as the ODDPO content increases, the MMLR of the modified epoxies initially increases, then decreases, and subsequently rises again (Figure 3b and Table 2). This trend may be explained as follows: At low ODDPO content, the flame retardant promotes the decomposition of the EP, leading to rapid weight loss over a short period and resulting in a higher MMLR. At moderate content levels, the flame retardant exhibits a more effective charring effect, slowing down the material’s decomposition and reducing the MMLR. However, at higher content levels, the large amount of volatile products generated during flame-retardant decomposition intensifies the breakdown of the matrix, causing the MMLR to increase once again.
Additionally, compared to EP-0, the characteristic temperatures of EP/ODDPO-0.9 and EP/ODDPO-1.2 gradually decreased (Figure 3a and Table 2). Specifically, the T5%, T10% and Tmax of EP/ODDPO-1.2 were 360.4 °C, 370.5 °C, and 399.2 °C, respectively, which represent decreases of 10.0 °C, 9.1 °C, and 9.1 °C. The stability of the P–C bonds in the flame retardant is lower than that of the C–C bonds in the epoxy matrix. Consequently, the premature decomposition of ODDPO accelerated the breakdown of the epoxy matrix, thereby reducing the thermal stability of the EP samples. It is reported that the T5%, T10%, and Tmax for EP/DOPO-1.2 were 350 °C, 359 °C, and 374 °C, respectively, while for EP/DPO-1.2 they were 351 °C, 369 °C and 402 °C (Table 2) [27]. These findings suggest that, compared to EP samples modified with DOPO and DPO, ODDPO-modified epoxy exhibited superior thermal stability. With the increase in phosphorus content, the T5%, T10% and Tmax of EP/ODDPO-1.5 decreased to 366.5 °C, 375.1 °C, and 403.9 °C, respectively. This improvement in thermal stability may be attributed to the increased rigidity of the material due to the presence of ODDPO [50]. Furthermore, the char residue of the EP samples increased from 0.7% for EP-0 to 22.6% for EP/ODDPO-1.5, which is beneficial for condensed-phase flame retardancy.

3.4. Flame Retardancy of Epoxy Composites

The flame retardancy of the cured EP samples was assessed by the LOI and the UL-94 tests, with the results summarized in Table 3. The EP-0 showed a relatively low LOI value of 23.2% and did not meet the criteria for a UL-94 rating, indicating its inherent flammability. With the addition of ODDPO, the EP/ODDPO-0.9 achieved an enhanced LOI value of 28.1% and reached a V-1 classification in the UL-94 test. Further increasing the phosphorus content in the EP/ODDPO-1.2 and EP/ODDPO-1.5 samples resulted in higher LOI values of 29.2% and 29.9%, respectively, and successful passing of the UL-94 V-0 rating. It is reported that both EP/DPO-1.2 and EP/DOPO-1.2 achieve a UL-94 V-0 rating, with LOI values of 30.8% and 30.0%, respectively (Table 3) [27]. These results demonstrated that ODDPO, as a flame retardant, provides flame retardancy comparable to that of commercial flame retardants. The significant improvement in the flame resistance of EP confirmed ODDPO’s effectiveness as an excellent flame retardant.

3.5. Combustion Behavior of Epoxy Composites

The combustion behavior of EP samples was evaluated through a cone calorimeter test. The heat release rate (HRR) and total heat release (THR) curves of EP-0 and EP/ODDPO-1.2 are shown in Figure 4, and the detailed relevant data are summarized in Table 4. The time-to-ignition (TTI) for EP-0 was 38 s, whereas the TTI for EP/ODDPO-1.2 was 34 s. This reduction may be due to the premature decomposition of the flame-retardant ODDPO, which promoted the thermal degradation of the epoxy matrix, thereby reducing the TTI. The peak heat release rate (pHRR) for EP-0 was 891.85 kW/m2, and THR was 88.20 MJ/m2. In comparison, the EP/ODDPO-1.2 exhibited a pHRR of 567.60 kW/m2 and a THR of 83.94 MJ/m2, ca. 36.4% and 4.8% reduction, respectively. These results indicated that the introduction of ODDPO significantly enhances the flame retardancy of the epoxy matrix.
In general, the average effective heat of combustion (av-EHC) can indicate the combustion intensity of volatile substances in the gaseous flame [51]. The av-EHC for EP-0 was 22.12 MJ/kg, while it decreased to 20.82 MJ/kg for EP/ODDPO-1.2. This indicated that the introduction of the flame retardant reduces the combustion intensity of volatile gases in the gas phase. The changes in average CO yield (av-CO) and average CO2 yield (av-CO2) can also account for this effect. In comparison to EP-0, the av-CO of EP/ODDPO-1.2 increased by 20.0%, while the av-CO2 decreased by 13.1%. The increase in the av-CO to av-CO2 ratio indicated a rise in the incomplete combustion of the EP matrix. These changes may result from the release of phosphorus-containing free radicals during the combustion of the modified EP, which can quench H• and OH• free radicals, interrupting the combustion process and thereby exerting a flame-retardant effect in the gas phase.
Additionally, the char residue increased from 2.42% for EP-0 to 9.16% for EP/ODDPO-1.2, indicating that ODDPO exerts a flame-retardant effect in the condensed phase. This may be attributed to the instability of the P–C bonds in the ODDPO structure, which decomposed upon heating to form polyphosphoric acid, thereby catalyzing the dehydration and charring of the epoxy matrix, consistent with the TGA results.
To understand the flame-retardant mode of ODDPO for EP, its flame-inhibition effect, charring effect, and barrier effect were quantitatively evaluated. The calculations were based on Equations (3), (4) and (5), respectively, with the subscripts FR-EP referring to EP/ODDPO-1.2. Detailed results are summarized in Table 5. From the perspective of gas-phase flame retardancy, the flame-inhibition effect of EP/ODDPO-1.2 improved by 5.88% compared to EP-0. In terms of condensed-phase flame retardancy, the charring effect and barrier effect of EP/ODDPO-1.2 increased by 6.91% and 33.12%, respectively, in comparison with EP-0. This indicated that the barrier effect of the flame retardant is more pronounced, which can prevent or slow down the transfer of heat, oxygen, and flammable gases, effectively inhibiting the combustion process. Overall, the ODDPO exhibited both gas-phase and condensed-phase flame-retardant effects, consistent with the cone calorimeter results.
Flame   inhibition   effect = 1 EHC FR - EP EHC EP - 0
Charring   effect = 1 TML FR - EP TML EP - 0
Barrier   effect = 1 pHRR FR - EP / pHRR EP - 0 THR FR - EP / THR EP - 0

3.6. Flame Retardancy Mechanism

The morphology of the char residues of EP-0 and EP/ODDPO-1.2 after cone calorimeter tests was observed using SEM to study the flame-retardant effect of ODDPO in the condensed phase. The external surface of the char residue for EP-0 displayed numerous cracks and holes (Figure 5a), allowing combustible gases and heat to escape, thereby enhancing the combustion of the epoxy matrix. In contrast, the external surface of the char residue for EP/ODDPO-1.2 was compact and continuous, with fewer holes (Figure 5c), effectively acting as a barrier to heat and material exchange. Additionally, the inner surface of the char residue for EP-0 was continuous and without holes (Figure 5b), while that for EP/ODDPO-1.2 exhibited a dense porous structure capable of storing gases produced during the pyrolysis of the matrix (Figure 5d). Once the storage limit is reached, a significant release of these gases will occur, which may help to dissipate heat and reduce the concentrations of oxygen and flammable gases, thereby mitigating combustion [52]. The significant differences in the char structure between the modified and unmodified EP samples indicated that ODDPO plays a role in solid-phase flame retardancy.
To further elucidate the gas-phase flame retardancy of ODDPO on EP, the pyrolysis products of EP samples under a nitrogen atmosphere were characterized using TG-IR. Figure 6 presents the FTIR spectra of the pyrolysis products of EP-0 and EP/ODDPO-1.2 at different temperatures. EP-0 decomposed significantly at 430 °C, generating H2O (3945–3550 cm−1), hydroxyl group (3385 cm−1), C-H group of aliphatic hydrocarbons (3066 cm−1 and 2971 cm−1), CO2 (2365 cm−1), CO (2321 cm−1), aromatic compounds (1735–1348 cm−1), compounds containing ethers (1320–1058 cm−1), and N-H group from DDS (829 cm−1) [42,53,54]. EP/ODDPO-1.2 underwent significant decomposition at 370 °C, attributed to the influence of ODDPO on the thermal stability of the matrix. The gaseous products of EP/ODDPO-1.2 were similar to those of EP-0. Notably, the peak at 1253 cm−1 in EP/ODDPO-1.2 was intensified due to the overlap of P=O stretching vibration with the C-O vibration of bisphenol A [45]. Furthermore, EP/ODDPO-1.2 exhibited a new peak at 745 cm−1 corresponding to P-C complexes [45]. These two groups consist of phosphorus-containing entities released during the thermal decomposition of the flame-retardant ODDPO. They can combine with H• and OH• free radicals produced from the decomposition of EP, thereby disrupting the combustion reaction, and imparting a gas-phase flame-retardant effect.

3.7. Wettability and Water Absorption Properties of Epoxy Composites

The hydrophobic properties of EP materials offer unique advantages in many industrial applications, particularly in waterproofing, corrosion resistance, and electrical insulation [55]. It is necessary to study the impact of additives on the wettability of the polymer matrix. Figure 7a shows the water-contact angles of EP-0 and ODDPO-modified EP samples. The water-contact angle of EP-0 was 80.9°, indicating that it is hydrophilic. For EP/ODDPO-0.9, EP/ODDPO-1.2, and EP/ODDPO-1.5, the contact angles were 95.3°, 97.3°, and 101.3°, respectively, showing increases of 17.8%, 20.3%, and 25.2% compared to EP-0. These results indicated that all the ODDPO-modified EP samples are hydrophobic. This is likely due to the symmetric and rigid structure of the flame-retardant ODDPO, which reduced water molecule permeation and enhanced the material’s hydrophobicity [56,57].
The water absorption of EP materials impacts their mechanical properties, electrical insulation, and dimensional stability. Lower water absorption enhances the performance and stability of epoxy in humid environments. Figure 7b shows the water absorption of EP-0 and ODDPO-modified EP samples over time. The water absorption of all samples increased with time. Moreover, the ODDPO-modified EP samples showed lower water absorption compared to EP-0. For instance, the water absorption dropped from 3.20 for EP-0 to 3.06 for EP/ODDPO-1.5 at 60 h. This decrease is likely due to the flame-retardant ODDPO’s symmetric and rigid structure, which reduced the permeation paths for water molecules, thereby lowering the water absorption [56,57].

3.8. Dielectric Properties of Epoxy Composites

In electrical applications, a low dielectric constant and a low dielectric loss in insulating epoxy can effectively minimize signal delay and energy loss [58]. Therefore, studying the characteristics of these parameters at different frequencies is crucial for designing and optimizing high-performance electronic devices. Figure 8 illustrates the dielectric constant and dielectric loss of EP-0 and ODDPO-modified EP samples at 1 MHz and 10 MHz. The dielectric constant values for EP-0 at 1 MHz and 10 MHz were 3.26 and 3.18, respectively (Figure 8a). With increases in the flame-retardant content, the dielectric constant of the ODDPO-modified EP samples gradually decreased at both frequencies. For instance, the dielectric constant values of EP/ODDPO-1.5 at 1 MHz and 10 MHz were 2.97 and 2.80, respectively, representing decreases of 8.90% and 11.95% compared to EP-0. Additionally, the dielectric loss of EP-0 was 0.0344 at 1 MHz and 0.0317 at 10 MHz (Figure 8b). As the flame-retardant content increased, the dielectric loss of the ODDPO-modified EP samples also showed a decreasing trend at both frequencies. For instance, the dielectric loss of EP/ODDPO-1.5 was 0.0291 at 1 MHz and 0.0294 at 10 MHz, representing decreases of 15.41% and 7.26%, respectively, compared to EP-0. The modified epoxy samples exhibited lower dielectric constant and dielectric loss, likely due to the presence of ODDPO with its symmetrical structure and small dipole moment, which reduced the polarizability of the epoxy matrix. The above results indicated that the introduction of ODDPO can enhance the dielectric properties of the EP matrix, thereby promoting the application of epoxy materials in electronic and electrical packaging.

4. Conclusions

In this study, the successful synthesis of ODDPO and its application as a flame retardant in epoxy thermosets demonstrated significant advancements in flame retardancy while maintaining the material’s intrinsic thermal stability and Tg. With a phosphorus content of 1.2 wt%, the ODDPO-modified EP sample achieved a UL-94 V-0 rating and exhibited a notable increase in the LOI value, to 29.2%. These results, supported by comprehensive cone calorimeter tests, highlight the ability of ODDPO to enhance flame retardancy through both gas-phase and condensed-phase mechanisms, particularly due to its barrier effect, thereby preventing flame propagation. Additionally, the modified epoxies showed improved hydrophobicity, water absorption resistance, and dielectric properties, which could be attributed to the symmetric and rigid structure of ODDPO. These properties make ODDPO-modified epoxies particularly suitable for electronic and electrical packaging applications, where flame resistance, moisture protection, and stable dielectric performance are critical. While the primary focus of this study was on improving flame retardancy and other functional properties, the impact of ODDPO on the mechanical properties of the epoxy composites remains an essential consideration for broader industrial applications. As physical and mechanical properties, such as tensile strength, bending strength, compressive strength, and impact resistance, play a critical role in the selection of materials for structural and load-bearing applications, future research will systematically investigate the effect of ODDPO on these properties.

Author Contributions

Conceptualization, Y.L. and Z.W.; Methodology, Z.W.; Formal analysis, Y.L., C.T. and G.C.; Investigation, C.T. and C.L.; Writing—original draft preparation, Y.L.; Writing—review and editing, G.C. and Z.W.; Supervision, Z.W. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Key Research and Development Special Fund of Qingdao Technical College (2021ZDYF 04).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Memon, H.; Wei, Y.; Zhu, C. Recyclable and reformable epoxy resins based on dynamic covalent bonds–Present, past, and future. Polym. Test. 2022, 105, 107420. [Google Scholar] [CrossRef]
  2. Bekeshev, A.; Mostovoy, A.; Shcherbakov, A.; Zhumabekova, A.; Serikbayeva, G.; Vikulova, M.; Svitkina, V. Effect of phosphorus and chlorine containing plasticizers on the physicochemical and mechanical properties of epoxy composites. J. Compos. Sci. 2023, 7, 178. [Google Scholar] [CrossRef]
  3. Nabipour, H.; Wang, X.; Song, L.; Hu, Y. Facile synthesis of a novel zinc-triazole complex for simultaneous improvement in fire safety and mechanical properties of epoxy resins. Compos. Appl. Sci. Manuf. 2021, 143, 106284. [Google Scholar] [CrossRef]
  4. Memon, H.; Wei, Y.; Zhu, C. Correlating the thermomechanical properties of a novel bio-based epoxy vitrimer with its crosslink density. Mater. Today Commun. 2021, 29, 102814. [Google Scholar] [CrossRef]
  5. Xu, B.; Liu, Y.; Wei, S.; Zhao, S.; Qian, L.; Chen, Y.; Shan, H.; Zhang, Q. A phosphorous-based bi-functional flame retardant based on phosphaphenanthrene and aluminum hypophosphite for an epoxy thermoset. Int. J. Mol. Sci. 2022, 23, 11256. [Google Scholar] [CrossRef] [PubMed]
  6. Lu, J.-H.; Xu, Y.-J.; Chen, L.; Chen, J.-H.; He, J.-H.; Li, Z.; Li, S.-L.; Wang, Y.-Z. Facile fabrication of intrinsically fire-safety epoxy resin cured with phosphorus-containing transition metal complexes for flame retardation, smoke suppression, and latent curing behavior. Chem. Eng. J. 2022, 442, 136097. [Google Scholar] [CrossRef]
  7. Gao, T.-Y.; Wang, F.-D.; Xu, Y.; Wei, C.-X.; Zhu, S.-E.; Yang, W.; Lu, H.-D. Luteolin-based epoxy resin with exceptional heat resistance, mechanical and flame retardant properties. Chem. Eng. J. 2022, 428, 131173. [Google Scholar] [CrossRef]
  8. Tian, Y.; Wang, Q.; Shen, L.; Cui, Z.; Kou, L.; Cheng, J.; Zhang, J. A renewable resveratrol-based epoxy resin with high Tg, excellent mechanical properties and low flammability. Chem. Eng. J. 2020, 383, 123124. [Google Scholar] [CrossRef]
  9. Xie, Z.; Xiao, D.; Yu, Q.; Wang, Y.; Liao, H.; Zhang, T.; Liu, P.; Xu, L. Fabrication of multifunctional silylated GO/FeSiAl epoxy composites: A heat conducting microwave absorber for 5G base station packaging. Materials 2023, 16, 7511. [Google Scholar] [CrossRef]
  10. Jia, X.-W.; Mu, W.-L.; Shao, Z.-B.; Xu, Y.-J. Flame-retardant cycloaliphatic epoxy systems with high dielectric performance for electronic packaging materials. Int. J. Mol. Sci. 2023, 24, 2301. [Google Scholar] [CrossRef]
  11. Liu, T.; Zhao, Y.; Deng, Y.; Ge, H. Preparation of fully epoxy resin microcapsules and their application in self-healing epoxy anti-corrosion coatings. Prog. Org. Coat. 2024, 188, 108247. [Google Scholar] [CrossRef]
  12. Ai, D.; Mo, R.; Wang, H.; Lai, Y.; Jiang, X.; Zhang, X. Preparation of waterborne epoxy dispersion and its application in 2K waterborne epoxy coatings. Prog. Org. Coat. 2019, 136, 105258. [Google Scholar] [CrossRef]
  13. Langer, E.; Waskiewicz, S.; Kuczynska, H. Application of new modified Schiff base epoxy resins as organic coatings. J. Coat. Technol. Res. 2019, 16, 1109–1120. [Google Scholar] [CrossRef]
  14. Cardoso, A.P.; de Sa, S.C.; Beraldo, C.H.M.; Hidalgo, G.E.N.; Ferreira, C.A. Intumescent coatings using epoxy, alkyd, acrylic, silicone, and silicone-epoxy hybrid resins for steel fire protection. J. Coat. Technol. Res. 2020, 17, 1471–1488. [Google Scholar] [CrossRef]
  15. Fu, Y.-X.; He, Z.-X.; Mo, D.-C.; Lu, S.-S. Thermal conductivity enhancement with different fillers for epoxy resin adhesives. Appl. Therm. Eng. 2014, 66, 493–498. [Google Scholar] [CrossRef]
  16. Lee, W.-J.; Cha, S.-H.; Kim, D.-H. Preparation and characterization of cardanol-based flame retardant for enhancing the flame retardancy of epoxy adhesives. Polymers 2022, 14, 5205. [Google Scholar] [CrossRef]
  17. Xiang, Q.; Xiao, F. Applications of epoxy materials in pavement engineering. Constr. Build. Mater. 2020, 235, 117529. [Google Scholar] [CrossRef]
  18. Qi, B.; Gao, S.; Xu, P. The application of recycled epoxy plastic sheets waste to replace concrete in urban construction and building. Processes 2023, 11, 201. [Google Scholar] [CrossRef]
  19. Wang, L.; Ma, Z.; Qiu, H.; Zhang, Y.; Yu, Z.; Gu, J. Significantly enhanced electromagnetic interference shielding performances of epoxy nanocomposites with long-range aligned lamellar structures. Nano-Micro Lett. 2022, 14, 224. [Google Scholar] [CrossRef]
  20. Rong, L.; Su, J.; Li, Z.; Liu, X.; Zhang, D.; Zhu, J.; Li, X.; Zhao, Y.; Mi, C.; Kong, X.; et al. Silicon hybridization for the preparation of room-temperature curing and high-temperature-resistant epoxy resin. Polymers 2024, 16, 634. [Google Scholar] [CrossRef]
  21. Kim, H.J.; Yoon, M.; Seo, B.; Lim, C.-S. Effects of fatty acid modified epoxy resin on long-chain epoxy and its physical properties. J. Polym. Sci. 2023, 61, 2194–2202. [Google Scholar] [CrossRef]
  22. Wang, Y.; Liu, W.; Qiu, Y.; Wei, Y. A one-component, fast-cure, and economical epoxy resin system suitable for liquid molding of automotive composite parts. Materials 2018, 11, 685. [Google Scholar] [CrossRef]
  23. Ji, Y.; Qiao, H.; Liang, Y.; Xu, G.; Wang, Y.; Hu, J. Preparation of water-based epoxy resin and its application as an automotive air filter paper binder. BioResources 2019, 14, 7148–7156. [Google Scholar] [CrossRef]
  24. Yoon, M.; Yoo, K.; Seo, B.; Ko, S.H.; Lim, C.-S. Development of low-shrink epoxy putty to solve appearance-quality defects of carbon-fiber-reinforced plastic automotive exterior parts. Materials 2021, 14, 6419. [Google Scholar] [CrossRef] [PubMed]
  25. Zheng, J.; Liu, Y.; Wang, Q.; Cheng, L.; Zhang, C.; Zhang, T.; Shao, J.; Dai, F. Mechanical properties and thermal characteristics of three nano-filler/silk fiber reinforced hybrid composites: A comparative study using a ductile epoxy resin matrix. Polym. Test. 2024, 130, 108319. [Google Scholar] [CrossRef]
  26. Huo, S.; Song, P.; Yu, B.; Ran, S.; Chevali, V.S.; Liu, L.; Fang, Z.; Wang, H. Phosphorus-containing flame retardant epoxy thermosets: Recent advances and future perspectives. Prog. Polym. Sci. 2021, 114, 101366. [Google Scholar] [CrossRef]
  27. Wei, Z.; Gu, X.; Wu, J.; Wei, M.; Yu, Q.; Tian, J.; Wang, Z. Performance comparison of epoxy resins modified with diphenylphosphine oxide and DOPO. Fire Mater. 2019, 43, 892–902. [Google Scholar] [CrossRef]
  28. Zhao, J.; Dong, X.; Huang, S.; Tian, X.; Song, L.; Yu, Q.; Wang, Z. Performance comparison of flame retardant epoxy resins modified by DPO–PHE and DOPO–PHE. Polym. Degrad. Stab. 2018, 156, 89–99. [Google Scholar] [CrossRef]
  29. Zhang, J.-Q.; Gao, M.; Han, L.-B. 2-(Hydroxypropan-2-yl)diphenylphosphine oxide (HDPO): A convenient precursor to diphenylphosphine oxide (DPO). Tetrahedron Lett. 2024, 135, 154889. [Google Scholar] [CrossRef]
  30. Zhang, H.; Xu, M.; Li, B. Synthesis of a novel phosphorus-containing flame retardant curing agent and its application in epoxy resins. J. Nanosci. Nanotechnol. 2016, 16, 2811–2821. [Google Scholar] [CrossRef]
  31. Yuan, L.; Yang, Z.; Zeng, P.; Gao, T.; Lv, Y.; Liu, X.; Ou, Y.; Chen, L. Behavior and mechanism of flame retardant epoxy resins with intrinsic phosphorus-nitrogen phenolic resin curing agents. Eur. Polym. J. 2024, 210, 112959. [Google Scholar] [CrossRef]
  32. Tian, X.; Yin, Q.; Wang, Z. Synthesis of diphenylphosphine oxide derivative and its flame retardant application in epoxy resin. J. Photopolym. Sci. Technol. 2019, 32, 769–778. [Google Scholar] [CrossRef]
  33. Wei, Z.; Wu, J.; Liu, Z.; Gu, Y.; Luan, G.; Sun, H.; Yu, Q.; Zhang, S.; Wang, Z. Effect of ethyl-bridged diphenylphosphine oxide on flame retardancy and thermal properties of epoxy resin. Polym. Adv. Technol. 2020, 31, 1426–1436. [Google Scholar] [CrossRef]
  34. Han, Y.; Zhao, H.; Chen, J.; Wang, X.; Chen, L.; Ran, D.; Wang, Z.; Zeng, P. A new phosphorus flame-retard curing agent for epoxy resin/anhydride system. Polym. Adv. Technol. 2022, 33, 927–936. [Google Scholar] [CrossRef]
  35. Xiong, Y.-Q.; Zhang, X.-Y.; Liu, J.; Li, M.-M.; Guo, F.; Xia, X.-N.; Xu, W.-J. Synthesis of novel phosphorus-containing epoxy hardeners and thermal stability and flame-retardant properties of cured products. J. Appl. Polym. Sci. 2012, 125, 1219–1225. [Google Scholar] [CrossRef]
  36. Dong, J.; Sang, X.; Yin, W.; Chen, X. Preparation of fluorinated epoxy-phthalonitrile resins with excellent thermal stability and low dielectric constant. J. Appl. Polym. Sci. 2023, 140, e53641. [Google Scholar] [CrossRef]
  37. Zhang, Y.; Lin, H.; Dong, K.; Tang, S.; Zhao, C. Comparison of the properties of epoxy resins containing various trifluoromethyl groups with low dielectric constant. Polymers 2023, 15, 2853. [Google Scholar] [CrossRef]
  38. ISO 4589-2017; Plastics—Determination of Burning Behaviour by Oxygen Index. International Organization for Standardization: Geneva, Switzerland, 2017.
  39. GB/T 2408-2008; Plastics—Vertical Burning Test. Standardization Administration of China: Beijing, China, 2008.
  40. ISO 5660-1:2015; Reaction-to-Fire Tests—Heat Release, Smoke Production and Mass Loss Rate—Part 1: Heat Release Rate (Cone Calorimeter Method). International Organization for Standardization: Geneva, Switzerland, 2015.
  41. Janesko, B.G.; Fisher, H.C.; Bridle, M.J.; Montchamp, J.-L. P(=O)H to P−OH tautomerism: A theoretical and experimental study. J. Org. Chem. 2015, 80, 10025–10032. [Google Scholar] [CrossRef]
  42. 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. Stab. 2021, 190, 109629. [Google Scholar] [CrossRef]
  43. Tang, S.; Wachtendorf, V.; Klack, P.; Qian, L.; Dong, Y.; Schartel, B. Enhanced flame-retardant effect of a montmorillonite/phosphaphenanthrene compound in an epoxy thermoset. RSC Adv. 2017, 7, 720–728. [Google Scholar] [CrossRef]
  44. Wang, P.; Xia, L.; Jian, R.; Ai, Y.; Zheng, X.; Chen, G.; Wang, J. Flame-retarding epoxy resin with an efficient P/N/S-containing flame retardant: Preparation, thermal stability, and flame retardance. Polym. Degrad. Stab. 2018, 149, 69–77. [Google Scholar] [CrossRef]
  45. Huo, S.; Yang, S.; Wang, J.; Cheng, J.; Zhang, Q.; Hu, Y.; Ding, G.; Zhang, Q.; Song, P.; Wang, H. A liquid phosphaphenanthrene-derived imidazole for improved flame retardancy and smoke suppression of epoxy resin. ACS Appl. Polym. Mater. 2020, 2, 3566–3575. [Google Scholar] [CrossRef]
  46. Liu, J.; Dai, J.; Wang, S.; Peng, Y.; Cao, L.; Liu, X. Facile synthesis of bio-based reactive flame retardant from vanillin and guaiacol for epoxy resin. Compos. Part B-Eng. 2020, 190, 107926. [Google Scholar] [CrossRef]
  47. Chen, C.; Wang, X.; Luo, T.; Zhen, H.; Yang, X.; Yang, L.; Yan, Z. Synthesis of solid reactive organophosphorus-nitrogen flame retardant and its application in epoxy resin. J. Appl. Polym. Sci. 2023, 140, e54282. [Google Scholar] [CrossRef]
  48. Hu, G.; Zhang, X.; Bu, M.; Lei, C. Toughening and strengthening epoxy resins with a new bi-DOPO biphenyl reactive flame retardant. Eur. Polym. J. 2022, 178, 111488. [Google Scholar] [CrossRef]
  49. Meng, J.; Zeng, Y.; Chen, P.; Zhang, J.; Yao, C.; Fang, Z.; Guo, K. New ultrastiff bio-furan epoxy networks with high Tg: Facile synthesis to excellent properties. Eur. Polym. J. 2019, 121, 109292. [Google Scholar] [CrossRef]
  50. Dong, C.; Wirasaputra, A.; Luo, Q.; Liu, S.; Yuan, Y.; Zhao, J.; Fu, Y. Intrinsic flame-retardant and thermally stable epoxy endowed by a highly efficient, multifunctional curing agent. Materials 2016, 9, 1008. [Google Scholar] [CrossRef]
  51. Qian, L.; Qiu, Y.; Sun, N.; Xu, M.; Xu, G.; Xin, F.; Chen, Y. Pyrolysis route of a novel flame retardant constructed by phosphaphenanthrene and triazine-trione groups and its flame-retardant effect on epoxy resin. Polym. Degrad. Stab. 2014, 107, 98–105. [Google Scholar] [CrossRef]
  52. Wang, P.; Chen, L.; Xiao, H. Flame retardant effect and mechanism of a novel DOPO based tetrazole derivative on epoxy resin. J. Anal. Appl. Pyrol. 2019, 139, 104–113. [Google Scholar] [CrossRef]
  53. Huo, S.; Zhou, Z.; Jiang, J.; Sai, T.; Ran, S.; Fang, Z.; Song, P.; Wang, H. Flame-retardant, transparent, mechanically-strong and tough epoxy resin enabled by high-efficiency multifunctional boron-based polyphosphonamide. Chem. Eng. J. 2022, 427, 131578. [Google Scholar] [CrossRef]
  54. Yang, Y.; Chen, W.; Li, Z.; Huang, G.; Wu, G. Efficient flame retardancy, good thermal stability, mechanical enhancement, and transparency of DOPO-conjugated structure compound on epoxy resin. Chem. Eng. J. 2022, 450, 138424. [Google Scholar] [CrossRef]
  55. Guo, L.; Gu, C.; Feng, J.; Guo, Y.; Jin, Y.; Tu, J. Hydrophobic epoxy resin coating with ionic liquid conversion pretreatment on magnesium alloy for promoting corrosion resistance. J. Mater. Sci. Technol. 2020, 37, 9–18. [Google Scholar] [CrossRef]
  56. Duan, X.; Li, R.; Li, B.; Feng, Q.; Xu, P.; Wan, J. Thymol-derived trifunctional epoxy resin with outstanding thermomechanical properties and low dielectric constant: From high-throughput synthesis to high-performance thermoset. ACS Appl. Polym. Mater. 2024, 6, 3481–3492. [Google Scholar] [CrossRef]
  57. Fu, F.; Zhou, X.; Shen, M.; Wang, D.; Xu, X.; Shang, S.; Song, Z.; Song, J. Preparation of hydrophobic low-k epoxy resins with high adhesion using a benzocyclobutene-rosin modifier. ACS Sustain. Chem. Eng. 2023, 11, 5973–5985. [Google Scholar] [CrossRef]
  58. Zhang, Y.-C.; Li, W.-D.; Zhao, X.; Meng, F.-B.; Sun, P.; Wang, C.; Guo, G.-Z.; Li, W.-R.; Zhang, G.-J. Epoxy-based high-k composite vitrimer: With low dielectric loss, high breakdown strength and surface electrical damage repairability. Chem. Eng. J. 2023, 473, 145199. [Google Scholar] [CrossRef]
Scheme 1. Synthesis route of ODDPO.
Scheme 1. Synthesis route of ODDPO.
Polymers 16 02635 sch001
Figure 1. ODDPO characterization: (a) 1H NMR spectrum; (b) 13C NMR spectrum; (c) 31P NMR spectrum; and (d) FTIR spectrum. The asterisk (*) indicates impurities in the synthesized compound. The letters are used to represent hydrogen atoms at different positions within the synthesized compound.
Figure 1. ODDPO characterization: (a) 1H NMR spectrum; (b) 13C NMR spectrum; (c) 31P NMR spectrum; and (d) FTIR spectrum. The asterisk (*) indicates impurities in the synthesized compound. The letters are used to represent hydrogen atoms at different positions within the synthesized compound.
Polymers 16 02635 g001
Figure 2. (a) DMA; and (b) storage modulus of epoxy resins.
Figure 2. (a) DMA; and (b) storage modulus of epoxy resins.
Polymers 16 02635 g002
Figure 3. (a) TGA; and (b) DTG curves of epoxy resins.
Figure 3. (a) TGA; and (b) DTG curves of epoxy resins.
Polymers 16 02635 g003
Figure 4. (a) HRR; and (b) THR curves of epoxy resins.
Figure 4. (a) HRR; and (b) THR curves of epoxy resins.
Polymers 16 02635 g004
Figure 5. SEM images of the residues for: (a,b) EP-0; and (c,d) EP/ODDPO-1.2 after the cone calorimeter test.
Figure 5. SEM images of the residues for: (a,b) EP-0; and (c,d) EP/ODDPO-1.2 after the cone calorimeter test.
Polymers 16 02635 g005
Figure 6. FTIR spectra of gaseous products of: (a) EP-0; and (b) EP/ODDPO-1.2 at different temperatures.
Figure 6. FTIR spectra of gaseous products of: (a) EP-0; and (b) EP/ODDPO-1.2 at different temperatures.
Polymers 16 02635 g006
Figure 7. (a) Water contact angle; and (b) water absorption of epoxy resins.
Figure 7. (a) Water contact angle; and (b) water absorption of epoxy resins.
Polymers 16 02635 g007
Figure 8. (a) Dielectric constant; and (b) dielectric loss of epoxy resins.
Figure 8. (a) Dielectric constant; and (b) dielectric loss of epoxy resins.
Polymers 16 02635 g008
Table 1. The formulations of epoxy resins with different phosphorus contents.
Table 1. The formulations of epoxy resins with different phosphorus contents.
SamplesDGEBA (g)DDS (g)ODDPO (g)P (wt%)
EP-010028.5500
EP/ODDPO-0.910027.348.160.9
EP/ODDPO-1.210026.9111.041.2
EP/ODDPO-1.510026.4714.011.5
Table 2. The DMA and TGA data of epoxy composites.
Table 2. The DMA and TGA data of epoxy composites.
SamplesTg (°C)νe (×103 mol/m3)T5% (°C)T10% (°C)Tmax (°C)MMLR
(%/min)
Char Residue at 800 °C (wt%)References
EP-0184.35.06370.4379.6408.31.360.7This work
EP/ODDPO-0.9175.23.56367.3376.1402.71.3820.4
EP/ODDPO-1.2176.73.68360.4370.5399.21.3121.5
EP/ODDPO-1.5171.52.40366.5375.1403.91.4122.6
EP/DPO-1.2154-351369402--[27]
EP/DOPO-1.2148-350359374--
Table 3. LOI values and UL-94 grades of epoxy resins.
Table 3. LOI values and UL-94 grades of epoxy resins.
SamplesLOI (%)UL-94 (3 mm)References
DrippingIgniting CottonRating
EP-023.2YesYesNRThis work
EP/ODDPO-0.928.1NoNoV-1
EP/ODDPO-1.229.2NoNoV-0
EP/ODDPO-1.529.9NoNoV-0
EP/DPO-1.230.8--V-0[27]
EP/DOPO-1.230.0--V-0
Table 4. The combustion behavior data of epoxy resins.
Table 4. The combustion behavior data of epoxy resins.
SamplesTTI (s)pHRR (kW/m2)THR (MJ/m2)av-EHC (MJ/kg)av-CO (MJ/kg)av-CO2 (MJ/kg)TML
(%)
Residue (wt%)
EP-038891.8588.2022.120.101.6097.582.42
EP/ODDPO-1.234567.6083.9420.820.121.3990.849.16
Table 5. Quantitative evaluation of the action mode of the flame retardant.
Table 5. Quantitative evaluation of the action mode of the flame retardant.
SamplesFlame Inhibition EffectCharring EffectBarrier Effect
EP-0---
EP/ODDPO-1.25.88%6.91%33.12%
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

Li, Y.; Tian, C.; Cheng, G.; Li, C.; Wang, Z. Facile Synthesis of Bis-Diphenylphosphine Oxide as a Flame Retardant for Epoxy Resins. Polymers 2024, 16, 2635. https://doi.org/10.3390/polym16182635

AMA Style

Li Y, Tian C, Cheng G, Li C, Wang Z. Facile Synthesis of Bis-Diphenylphosphine Oxide as a Flame Retardant for Epoxy Resins. Polymers. 2024; 16(18):2635. https://doi.org/10.3390/polym16182635

Chicago/Turabian Style

Li, Yan, Chong Tian, Guiqing Cheng, Chunhui Li, and Zhongwei Wang. 2024. "Facile Synthesis of Bis-Diphenylphosphine Oxide as a Flame Retardant for Epoxy Resins" Polymers 16, no. 18: 2635. https://doi.org/10.3390/polym16182635

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