Next Article in Journal
Approaches in Sustainable, Biobased Multilayer Packaging Solutions
Previous Article in Journal
Intelligent Eucommia ulmoides Rubber/Ionomer Blends with Thermally Activated Shape Memory and Self-Healing Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Flame Retardancy and Mechanical Properties of Melt-Spun PA66 Fibers Prepared by End-Group Blocking Technology

State Key Laboratory for Modification of Chemical Fibers and Polymer Materials, College of Materials Science and Engineering, Donghua University, Shanghai 201620, China
*
Authors to whom correspondence should be addressed.
Polymers 2023, 15(5), 1183; https://doi.org/10.3390/polym15051183
Submission received: 14 February 2023 / Accepted: 22 February 2023 / Published: 26 February 2023
(This article belongs to the Section Polymer Fibers)

Abstract

:
Preparing flame-retardant polyamide 66 (PA66) fibers through melt spinning remains one of the biggest challenges nowadays. In this work, dipentaerythritol (Di−PE), an eco-friendly flame retardant, was blended into PA66 to prepare PA66/Di−PE composites and fibers. It was confirmed that Di−PE could significantly improve the flame-retardant properties of PA66 by blocking the terminal carboxyl groups, which was conducive to the formation of a continuous and compact char layer and the reduced production of combustible gas. The combustion results of the composites showed that the limiting oxygen index (LOI) increased from 23.5% to 29.4%, and underwriter laboratories 94 (UL-94) passed the V-0 grade. The peak of heat release rate (PHRR), total heat release (THR), and total smoke production (TSP) decreased by 47.3%, 47.8%, and 44.8%, respectively, for the PA66/6 wt% Di−PE composite compared to those recorded for pure PA66. More importantly, the PA66/Di−PE composites possessed excellent spinnability. The prepared fibers still had good mechanical properties (tensile strength: 5.7 ± 0.2 cN/dtex), while maintaining good flame-retardant properties (LOI: 28.6%). This study provides an outstanding industrial production strategy for fabricating flame-retardant PA66 plastics and fibers.

Graphical Abstract

1. Introduction

A survey from the Fire and Rescue Department Ministry of Emergency Management of China reports that 1.324 million fire accidents, 11,634 deaths, and 6738 injuries resulting from residential fires occurred nationwide over the period of 2012–2022. Polymeric materials are widely used in everyday day life, which increases the fire hazards [1]. The fire problem caused by textiles has always plagued people all over the world and seriously threatens people’s lives and property, public safety, and social development. Polyamide 66 (PA66), also known as nylon 66, is a thermoplastic textile material used in many fields including apparels and industrial textiles due to its excellent properties such as good mechanical behavior, low cost, resistance to shrinkage and pleasant aesthetics [2,3]. However, PA66 is flammable and also present a melt dripping problem during combustion, which limits its wide application [4]. Therefore, it is necessary and remains a huge challenge to prepare PA66 fabrics with excellent flame retardance properties.
Conventional approaches that can be considered to improve the flame retardance property of PA66 fabric include post-finishing, copolymerization, and blending. Among them, the flame-retardant finishing of PA66 fabric is the most widely used due to its simple processing [5,6,7]. Thiourea–formaldehyde is one of the most successful durable flame-retardance finishing systems for polyamide fabrics, but it releases formaldehyde during processing and use, which has known toxic effects on the environment and human health [8,9]. Although some research teams have developed relatively green flame-retardant finishing systems, their flame-retardant efficiency and durability appeared reduced [10,11]. More importantly, the biggest common problem of the post-finishing method is that the wear comfort of the fabric, including air permeability, hygroscopicity, flexibility, etc., is often sacrificed during the finishing process.
Compared with post-finishing, copolymerization can impart PA66 intrinsic flame-retardancy and enhanced durability [12,13,14,15,16]. Li et al. [17] prepared flame-retardant PA66 composites by condensation polymerization of a PA66 salt with 5.5 wt% of a 9,10-dihydro-9-oxa-10-phosphaphenanthrene-10-oxide (DOPO)-based flame retardant. The limiting oxygen index (LOI) value and underwriter laboratories 94 (UL-94) of the PA66 composites reached 32.9% and V-0 rating, respectively. It is worth noting that, although the authors prepared PA66 fibers by using the flame-retardant PA66 composites through melt spinning, unfortunately, the fiber strength was decreased by 37.4% compared with that of bare PA66, and more importantly, the flame-retardant performance of the composite fiber was not evaluated. As for the blending method, it has more advantages than the copolymerization method because of the simple preparation required and the easiness to scale up production to industrial levels. However, as far as the authors know, there are few reports about flame-retardant PA66 fibers or fabrics which are prepared by the blending method, and almost all of them are plastic products [18,19]. The reasons behind this are as follows: (1) the addition of flame retardants often results in poor spinnability of PA66 and even failure to spin due to the poor dispersion and compatibility; (2) there is a lack of high-performance flame retardants suitable for the melt spinning of nylon 66, which needs to meet lack of toxicity, high thermal stability, and good spinnability and to achieve high flame retardant efficiency with the addition of a low amount.
Herein, a series of PA66 flame retardant composites were prepared by the simple blending method using dipentaerythritol (Di−PE) as a flame retardant. The effect of different mass ratios of Di−PE on the flame retardancy of the PA66/Di−PE composites was investigated by limited oxygen index, underwriter laboratories 94, and cone calorimeter tests, and the possible flame retardancy mechanism was also explored. More importantly, PA66 fibers with excellent flame retardance performance were finally obtained via melt spinning and their mechanical properties were also evaluated.

2. Materials and Methods

2.1. Materials and Machining Equipment

PA66 was purchased from Shen Ma Industry Co., Ltd., China. Di−PE (90%) was purchased from Shanghai Aladdin Bio-Chem Technology Co., Ltd., Shanghai, China. The materials were used as received. The twin-screw extruder used (D: 21.7 mm, L/D: 40, model, SHJ-20) was purchased from Nanjing GIANT Machinery Co., Ltd., Nanjing, China. The vertical injection molding machine (TY-200) was purchased from TAYU Machinery Co., Ltd., China. The two-component composite spinning machine (Abe Φ25∗2, D = 25 mm, L/D = 28) was purchased from Abe Corporation, Japan.

2.2. Preparation of Flame-Retardant PA66 Composites

Di−PE and PA66 were dried overnight in a vacuum oven at 110 °C. The flame retardant PA66 composites were obtained by extruding the mixtures of PA66 and Di−PE in a twin-screw extruder. The temperature of the extruder was maintained at 230 °C, 245 °C, 260 °C, 270 °C, 265 °C, and 265 °C from hopper to die. The dried pellets of the composites were molded by the vertical injection molding machine at a temperature in three heating zones (240 °C, 270 °C, and 267 °C).

2.3. Melt Spinning of PA66 and PA66/Di−PE Composites

The pellets of PA66 and PA66/Di−PE composites were melt-spun with the two-component composite spinning machine. The main spinning parameters are summarized in Table 1.

2.4. Characterization

2.4.1. Limiting Oxygen Index

The LOI test was determined on an oxygen index meter (TTech-GBT2406-2, TES Tech Instrument Technologies Co., Ltd., Suzhou, China) with 80 mm × 10 mm × 4.0 mm specimens according to GB/T 2406.2-2009.

2.4.2. UL-94 Vertical Burning

The UL-94 tests were performed based on a horizontal and vertical burning test instrument (ZR-02 vertical burning test instrument, Qingdao Shanfang Instrument Co., Ltd., Qingdao, China) according to the GBT2408-2008 testing procedure, using samples with dimensions of 127 mm × 12.7 mm × 3.2 mm.

2.4.3. Fourier Transform Infrared Spectroscopy (FTIR)

The ATR-FTIR spectra of the PA66/Di−PE composites and char residues were obtained with a Nicolet 6700 FTIR instrument (Thermo Fisher Scientific, Waltham, Massachusetts, USA). The wavenumber range was set from 4000 to 600 cm−1.

2.4.4. Thermogravimetry Analysis

The thermal properties of the samples were studied with a thermogravimeter (TG 209 F1, Netzsch Netzsch, Selb, Germany). Fir this, 5–10 mg samples were heated at 20 °C/min from 25 to 700 °C under air atmosphere.
Differential scanning calorimetry (DSC) was performed on a Netzsch DSC 204 F1 (Selb, Germany), with a heating rate of 20 °C/min up to 280 °C under nitrogen and held for 2 min to completely remove the previous thermal history. Then, the samples were cooled down to 25 °C and finally heated to 280 °C at the cooling and heating rates of 10 °C/min.

2.4.5. Cone Calorimeter Test

The combustion performance was evaluated with a cone calorimeter (i-Cone, Fire Testing Technology Ltd., East Grinstead, UK) in accordance with ISO 5660-1 standard, using specimen with dimensions of 100 mm × 100 mm × 3 mm with 50 kW/m2.

2.4.6. Scanning Electron Microscope (SEM) Analysis

The residual chars were obtained in a muffle furnace at 500 °C, then observed by a scanning electron microscope (FlexSEM 1000II, Hitachi, Tokyo, Japan) at the accelerating voltage of 10 kV.

2.4.7. Raman Analysis

The Raman spectra of PA66 and the PA66/Di−PE composites residual chars were obtained with a laser Raman spectrometer (Renishaw inVia Reflex, Renishaw Co., Miskin, UK) using a 532 nm helium–neon laser line focused on a micrometer spot on the sample surface and scanning in the 2000–800 cm−1 region.

2.4.8. X-ray Photoelectron Spectroscopy

The XPS data were carried obtained with an Escalab 250Xi electron analyzer (Thermo Fisher Scientific Co., Waltham, MA, USA).

2.4.9. Thermogravimetry–Fourier Transform Infrared Spectrometry

The TG–FTIR was conducted on a thermogravimeter instrument (TG 209 F1, Netzsch Netzsch, Selb, Germany) with a Fourier transform infrared spectrometer (Tensor 27, Bruker Optics, Ettlingen, Germany). About 10 mg of the samples was heated at 20 °C/min from 25 to 800 °C under air atmosphere. The pyrolysis gaseous products were tunneled into the FTIR device and scanned in the range from 4000 to 600 cm−1.

3. Results

3.1. Thermal Stability of the PA66/Di−PE Composites

Thermogravimetric analysis (TGA) and derivative thermogravimetry (DTG) curves in the air atmosphere were utilized to assess the thermal-oxidative degradation behavior of PA66 and its composites (Figure 1). The summary of the results is presented in Table 2. T5% is the temperature corresponding to a 5 wt% weight loss and is indispensable to evaluate the thermal decomposition behavior of a polymer in the initial stage. The pure PA66 presented almost no char residue above 600 °C, and the T5% and Tmax (the temperature at the maximum weight loss rate) were about 396.3 °C and 477.7 °C, respectively.
It is obvious that in the presence of Di−PE, both the T5% and the Tmax of the PA66/Di−PE composites decreased, while the char yields of the PA66/4 wt% Di−PE and PA66/6 wt% Di−PE composites at 600 °C or 700 °C were much higher than that of pure PA66. In order to more intuitively verify that Di−PE promoted char formation from PA66, PA66 and PA66/6 wt% Di−PE were placed in muffle furnaces at different temperatures, and it was found that PA66/6 wt% Di−PE formed carbon layers earlier than PA66 at the same high temperature (Figure S1). Such a phenomenon confirmed that the presence of Di−PE promoted the decomposition and the charring process of the PA66/Di−PE composites in air atmosphere, which might contribute to the flame-retardant performance.

3.2. Flame-Retardant and Combustion Behavior

The LOI and UL-94 vertical burning tests were used to study the flame-retardant performance of the PA66/Di−PE composites. The corresponding results are listed in Table 3. The pure PA66 can be ignited and burnt with dripping, showing flammable characteristics. For the PA66/Di−PE composite, the LOI value increased with the Di−PE content. The LOI value of the PA66/Di−PE composite with a relatively low loading of 2 wt% could reach 28.80%, but the UL-94 rating was still V-2. Interestingly, a maximum LOI value of 29.4% and a UL-94 V-0 rating could be reached by further increasing the Di−PE content to 6 wt%. Moreover, the color of the char layer, which was formed in the PA66/Di−PE after the LOI test gradually, turned black with the increasing content of Di−PE, indicating that a higher content of the char layer and a higher degree of carbonization were obtained (Figure S2). These results showed that Di−PE was effective in improving the flame retardance of PA66.
In order to further evaluate the combustion process of the PA66/Di−PE composites in real combustion situations, the cone calorimetry test (CCT) was implemented. The CCT is a significant method to evaluate the fire behavior of polymers. It simulates fire hazard conditions and provides a series of key flammability parameters such as heat release rate (HRR), peak of heat release rate (PHRR), total heat release (THR), effective heat of combustion (EHC), mass loss rate (MLR), total smoke production (TSP), peak of smoke production rate (PSPR), average carbon monoxide yield (av-COY), average carbon dioxide yield (av−CO2Y) [20,21], and non-dimensional flame retardancy index (FRI) [22,23]. In general, lower values of PHRR, av-HRR, and THR indicate a better flame-retardant effect and are used to evaluate the fire safety of the materials in a real fire. It can be seen in Figure 2a,b and Table 4 that the PHRR and THR of PA66/2 wt% Di−PE were slightly higher than those of pure PA66. This was caused by the violent thermal degradation of the composite at a high temperature and the insufficient flame-retardant properties of the PA66/2 wt% Di−PE composite. In addition, with the increasing content of Di−PE, the values of PHRR, av-HRR, and THR decreased. Specifically, the PHRR, av-HRR, and THR of PA66/6 wt% Di−PE sharply decreased to 558.7 kW/m2, 112.8 kW/m2, and 60.1 MJ/m2, which represent a reduction by 47.3%, 68.6%, and 47.8%, respectively, compared to pure PA66. Figure 2c,d also presents the SPR and TSP curves of the PA66/Di−PE composites, which are considered other important parameters related to the hazard that threatens the life and environment in a real fire. Similar to the THR and HRR curves, the PA66/Di−PE composites showed significantly reduced values of TSP and SPR, while the Di−PE content was greater than or equal to 4 wt%, as compared to that in pure PA66. The PSPR and TSP of pure PA66/6 wt% decreased by 43.7% and 44.8%, respectively. Comprehensively, all the above data and curves (HRR, THR, SPR, and TSP) indicated the marked improvement of the flame retardancy and smoke suppression performance of PA66 with the incorporation of Di−PE. It acquired satisfactory flame-retardant properties, comparing with other works in the Table S1 (Supplementary Materials).
Additionally, the flame retardancy index (FRI), a non-dimensional criterion to evaluate the flame-retardant properties of composites, was calculated according to the following formula:
FRI = [ THR × ( PHRR TTI ) ] Neat   polymer [ THR × ( PHRR TTI ) ] Composite
The FRI is an important parameter for evaluating the flame-retardant properties of a composite. More specifically, the FRI is classified as Poor (FRI ≤ 1), Good (1 < FRI ≤ 10), and Excellent (FRI > 10) [24]. With the incorporation of Di−PE, the FRI of the PA66/Di−PE composites improved when the concentration of Di−PE increased. As shown in Table 4, when the concentration of Di−PE in PA66 was 4 wt% and 6 wt%, the FRI of PA66/4 wt% Di−PE and PA66/6 wt% Di−PE composites reached 1.53 and 2.42, respectively, which correspond to a “Good” level of fire safety. The higher FRI for the PA66/4 wt% Di−PE and PA66/6 wt% Di−PE composites proved their remarkable fire retardancy ability.

3.3. Analysis of the Flame-Retardant Mechanism

The Cone test not only describes the flame retardance of samples but also can be used to study flame-retardant mechanisms [25]. The TTI of the PA66/Di−PE composites decreased compared to that of pure PA66, indicating that Di−PE promoted the catalytic decomposition of PA66 to form char layers, which provided an effective shield protection to restrain the intensity of smoke emission, thus reducing the SPR and TSP. The mass loss curves (Figure 3a) also confirmed that the thermal decomposition time of PA66/Di−PE was shorter than that of pure PA66, which is in accordance with the TG result, while the peak MLR and av-MLR values of the PA66/Di−PE composites were lower than those of pure PA66, thus resulting in increased amounts of char residuals, indicating that Di−PE played a flame retardant role in the condensed phase.
EHC is an important parameter to demonstrate the degree of combustion of volatile gas in the gas phase [26]. Figure 3c shows that the EHC of the PA66/Di−PE composites was lower compared with that of PA66, indicating that the incorporation of Di−PE led to incomplete combustion in the gas phase, which was also confirmed by the av-COY and av-CO2Y shown in Figure 3d–f. Clearly, the av-COY increased, whereas the av-CO2Y decreased with the increase of Di−PE loading. These results showed that Di−PE also played a flame-retardant role in the gas phase.
In order to fully understand the flame retardance mechanism of the PA66/Di−PE composites, ATR-FTIR, SEM, Raman spectroscopy, XPS, and TG-IR were utilized to investigate the properties of the char residues and the volatile components. Figure 4 shows the microstructures of the residual chars for pure PA66 and the PA66/Di−PE composites. Clearly, it can be observed that the char layer of pure PA66 was fragile with many cracks and holes (Figure 4a,e, the red triangle), which would provide pathways for inflammable volatile products to mix with oxygen, then leading to the flame propagation. Therefore, the char layer cannot provide good flame protection for the underlying material from further burning. Interestingly, the char layers of PA66/2 wt% Di−PE only showed a few microcracks and were more continuous compared with those of pure PA66 (Figure 4b,f, the red triangle). Further increasing the content of Di−PE to 4 wt% (Figure 4c,g, green triangle) and 6 wt% (Figure 4d,h, green triangle), uniform and compact char layers with wrinkled morphologies formed instead of cracks and holes. As a result, the outstanding barrier effect of this kind of char layer can effectively prevent the transmission of heat, oxygen, and combustible gases, thus protecting the internal PA66 matrix from flame.
Moreover, besides the morphology of the char, the graphitization degree of the char also played an important role in improving the flame-retardant properties. The higher the graphitization degree in the char structure, the better the protection against thermal degradation of the material. Thus, Raman spectroscopy was applied to analyze the carbonaceous structures and graphitization of residual char (Figure 5). The graphitization degree of residual char was assessed by the relative intensity ratio of the D and G bands (ID/IG); the fitting equation is Gaussian, and the lower the ID/IG value, the higher the graphitization degree [27,28,29]. It is obvious that the ID/IG value decreased with the increasing content of Di−PE. Therefore, both the microstructure and the graphitization degree of the char residuals further indicated that Di−PE played a positive effect on enhancing the flame retardance in the condensed phase, which could promote the formation of a compact and continuous char layer with a high graphitization degree, thus protecting the matrix effectively.
The mechanism by which Di−PE could promote char formation of nylon 66 was further explored by FTIR analysis, as shown in Figure 6a,b. It can be seen that with the addition of Di−PE, the characteristic peak of the carboxyl group at 3080 cm−1 for the samples PA66/4 wt% Di−PE and PA66/6 wt% gradually decreased. XPS can determine precisely the reaction between a polymer and an additive by the change in chemical bonding. In Figure 6c, the O1 s XPS spectrum of PA66 was divided into two peaks at 532.0 and 531.0 eV, which corresponded to the C=O and COOH bonding states, respectively [30]. Compared with the O1 s XPS spectrum of PA66, the O1 s spectrum of PA66/6 wt% Di−PE was divided into four peaks, and those of the COOH and C=O groups shifted to 531.9 eV and 530.6 eV, respectively [31]. The new peaks at 532.7 and 533.4 eV corresponding to the C-OH and C-O-C bonding states, indicated that the terminal carboxyl group in nylon 66 was blocked, which could lead to a local cross-linked structure, thus helping to promote the formation of a more stable char layer during the combustion process. The FTIR of the char residues also confirmed the above conjecture (Figure 6b). The char residues of PA66 and the PA66/Di−PE composites revealed some typical characteristic peaks at 2918 cm−1 and 2856 cm−1 attributed to the -CH2- stretching vibration. Meanwhile, some new characteristic peaks were observed at 1648, 1531, 1456, 1375, and 723 cm−1. The peak at 1648 cm−1 corresponded to the stretching vibration of the aromatic structure of the benzene ring, and the characteristic peak at 723 cm−1 was attributed to the Carom-H rocking [32], showing that Di−PE could promote the formation in PA66 of a more stable benzene-containing char layer [32,33]. In addition, it was observed that the characteristic peak at 1531 cm−1 (-CONH-) of PA66 was not observed for the char residues, while the char residues of the PA66/Di−PE composites retain a visible characteristic peak here, proving that Di−PE caused the incomplete degradation of PA66 [34]. The peaks at 1456 and 1375 cm−1 correspond to the -CN- stretching vibration [35,36,37]. Char residues with such structures were more conducive to the improvement of the flame-retardant properties.
TG-IR is used to characterize volatile components during heating in the air atmosphere. It could help us to investigate the flame-retardance mechanism of Di−PE in the gas phase. It can be observed in Figure 7a that the pyrolysis behavior of PA66/6 wt% Di−PE shifted to an earlier time, and the absorption intensity of the volatilized components decreased significantly compared to pure PA66, which is in accordance with the CCT results. The characteristic peaks of major pyrolysis volatiles, such as NH3 (930–960 cm−1), hydrocarbons with the C-H functionality (2850–2950 cm−1), carbonyl compounds (aldehydes, ketones, carboxylic acids, 1633–1839 cm−1), CO (2150–2180 cm−1), and CO2 (2300–2400 cm−1), were further investigated [38]. As depicted in Figure 7b, the absorption intensity of NH3 for PA66/6 wt% Di−PE slightly increased compared to that of pure PA66, indicating a possible dilution of the flammable gas and oxygen, thus inhibiting gas combustion in the gas phase [39]. Furthermore, it was seen that the release of hydrocarbons (Figure 7c) and carbonyl compounds (Figure 7d) during the thermal decomposition of PA66/6 wt% Di−PE decreased significantly compared to that observed for pure PA66, which is beneficial for char formation and ensures the presence of less fuel into the fire zone [40]. In addition, the increased absorption intensity of CO and the decreased absorption intensity of CO2 confirmed that Di−PE promoted the incomplete combustion of the PA66/Di−PE composites, which is in accordance with the CCT results. All these observations confirmed that Di−PE also has a positive effect in the gas phase during the combustion process.

3.4. Tensile Properties of the PA66/Di−PE Composites

The tensile properties of pure PA66 and the PA66/Di−PE composite are shown in Figure 8 and Table 5. It can be seen that the addition of Di−PE had no significant effect on the tensile strength and the tensile modulus of PA66, but significantly reduced its elongation at break, from 71.7 ± 8.7% for pure PA66 to 18.0 ± 10.2% for PA66/6 wt% Di−PE. This could be due to the partial cross-linking between the hydroxyl group of Di−PE and the terminal carboxyl group of PA66.

3.5. Crystallization Behavior of PA66 and the PA66/Di−PE Composites

The influence of Di−PE on the crystallization behavior of the PA66/Di−PE composites was investigated by DSC. The DSC cooling and second heating curves of PA66 and the PA66/Di−PE composites are shown in Figure 9. As clearly observed, the PA66/Di−PE composites showed similar heating and cooling curves to those of pure PA66, indicating that the incorporation of Di−PE had little influence on the crystal form of the PA66/Di−PE composites. However, the crystallization temperature of the PA66/Di−PE composites decreased by nearly 7 °C compared with that of pure PA66 (Figure 9a), which was probably due to the fact that the incorporation of Di−PE hindered the movement of the PA66 chains and made the crystal structure less perfect, leading to a decrease in the melting point (Figure 9b). Nevertheless, the crystallinity of the PA66/Di−PE composite was slightly improved (Table 6).

3.6. Mechanical Properties of PA66 and PA66/Di−PE Fibers

Thanks to the excellent melt processing performance and high thermal stability of the PA66/Di−PE composites, PA66/Di−PE fibers could be obtained via melt spinning. Figure 10 shows the tensile strength test process of PA66 fibers and the PA66/Di−PE composite fibers, and Table 7 shows the mechanical properties of PA66 and PA66/4 wt% Di−PE fibers. It is clearly seen that, although the tensile strength and modulus of the flame-retardant fiber were slightly decreased, they could still reach 5.7 cN/dtex and thus still meet the requirements of most applications.

3.7. Burning Behavior of PA66 and PA66/Di−PE Composites Fabric

In addition to the mechanical properties, the maintenance of the flame-retardant properties of fibers is very important. Figure 11a,b shows the digital photos of knitted PA66 and PA66 4 wt% Di−PE fabrics, which were used for LOI testing. As can be seen, there was no difference in color, indicating that the addition of Di−PE had little effect on the appearance of the fibers. More excitingly, the LOI of the PA66/4 wt% Di−PE fabrics was still as high as 28.6%, and a UL-94 V-0 rating could also be reached. In addition, the flame-retardant fibers could self-extinguish within 3 s, with an average damaged length of 11.1 cm, largely better than that measured for pure PA66 fabrics (Table 8). All these results indicated that Di−PE is a very effective flame retardant for PA66 and has little effect on the fiber forming process; more valuably, the flame-retardant properties of the PA66/Di−PE composites can be highly maintained in the fibers.

4. Conclusions

Flame retardant PA66/Di−PE composites and fibers were prepared by blending and the melt spinning technology. It was found that Di−PE could significantly improve the flame-retardant properties of PA66 through the end-group blocking effect. The PA66/6 wt% Di−PE composite achieved the most outstanding LOI value of 29.4% and passed the V-0 rating of UL-94. The CCT results of PA66/6 wt% Di−PE showed that the PHRR and THR were reduced by 47.3% and 47.8%, respectively, compared to those of pure PA66. The FRI of the PA66/4 wt% Di−PE and PA66/6 wt% Di−PE composites reached 1.53 and 2.42, respectively. Moreover, the addition of Di−PE was confirmed to lead to the formation of a continuous and compact char layer, which could cut off the transfer of heat, combustible gas, and oxygen, as well as decrease the release of total gas and flammable hydrocarbons and increase the release of incombustible gas NH3. More importantly, the flame retardant PA66/4 wt% Di−PE composites could be made into fibers by melt spinning. The LOI, after-flame time, and damaged length of the PA66/4 wt% Di−PE fabric were significantly improved compared with those of pure PA66, and the absorbent cotton was not ignited. This work provides an outstanding industrial production strategy for fabricating flame-retardant PA66 plastics and fibers and shows that Di−PE has a great potential to reinforce the flame retardancy of PA66.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/polym15051183/s1, Figure S1: Pure PA66 and PA66/6% Di-PE composites in muffle furnace at different temperature; Figure S2: Digital photos of Pure PA66 (a), PA66/2% Di-PE (b), PA66/4% Di-PE (c) and PA66/6% Di-PE (d) after LOI test; Table S1: Representative information of PA66 flame retardant modified from recent years [41,42,43,44,45,46,47,48].

Author Contributions

Conceptualization, Y.W.; methodology, T.Y. and Y.C.; software, Y.C. and B.Y.; validation, Y.W., H.Y. and Y.W.; formal analysis, B.Y.; investigation, T.Y. and B.Y.; resources, T.H.; data curation, H.Y.; writing—original draft preparation, Y.W.; writing—review and editing, T.H.; visualization, Q.W.; supervision, M.Z.; project administration, Y.W. and H.Y.; funding acquisition, T.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Natural Science Foundation of China (No. 52203309, 51703024), the Fundamental Research Funds for the Central Universities (No. 2232022D-09), the Shanghai Sailing Program (No. 22YF1400400), the Shanghai Education Development Foundation, and Shanghai Municipal Education Commission (18CG37).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We would like to thank all the editors and reviewers for correcting the manuscript very carefully.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Morgan, A.B. The future of flame retardant polymers—Unmet needs and likely new approaches. Polym. Rev. 2019, 59, 25–54. [Google Scholar] [CrossRef]
  2. Kundu, C.K.; Wang, W.; Zhou, S.; Wang, X.; Sheng, H.B.; Pan, Y.; Song, L.; Hu, Y.A. A green approach to constructing multilayered nanocoating for flame retardant treatment of polyamide 66 fabric from chitosan and sodium alginate. Carbohydr. Polym. 2017, 166, 131–138. [Google Scholar] [CrossRef] [PubMed]
  3. Tamura, K.; Ohyama, S.; Umeyama, K.; Kitazawa, T.; Yamagishi, A. Preparation and properties of halogen-free flame-retardant layered silicate-polyamide 66 nanocomposites. Appl. Clay Sci. 2016, 126, 107–112. [Google Scholar] [CrossRef]
  4. Rahman, M.Z.; Kundu, C.K.; Nabipour, H.; Wang, X.; Song, L.; Hu, Y. Hybrid coatings for durable flame retardant and hydrophilic treatment of polyamide 6.6 fabrics. Prog. Org. Coat. 2020, 144, 105640. [Google Scholar] [CrossRef]
  5. Feldman, D. Polymer nanocomposites: Flammability. J. Macromol. Sci. Part A-Pure Appl. Chem. 2013, 50, 1241–1249. [Google Scholar] [CrossRef]
  6. Meng, D.; Guo, J.; Wang, A.; Gu, X.; Wang, Z.; Jiang, S.; Zhang, S. The fire performance of polyamide66 fabric coated with soybean protein isolation. Prog. Org. Coat. 2020, 148, 105835. [Google Scholar] [CrossRef]
  7. Norouzi, M.; Zare, Y.; Kiany, P. Nanoparticles as effective flame retardants for natural and synthetic textile polymers: Application, mechanism, and optimization. Polym. Rev. 2015, 55, 531–560. [Google Scholar] [CrossRef]
  8. Liu, W.; Zhang, S.; Sun, J.; Gu, X.Y.; Ge, X.G. An improved method for the durability of the flame retardant PA66 fabric. J. Therm. Anal. Calorim. 2017, 128, 193–199. [Google Scholar] [CrossRef]
  9. Sun, J.; Gu, X.Y.; Dong, Q.L.; Zhang, S.; Liu, W.; Li, L.Y.; Chen, X.S. Durable flame-retardant finishing for polyamide 66 fabrics by surface hydroxymethylation and crosslinking. Polym. Adv. Technol. 2013, 24, 10–14. [Google Scholar] [CrossRef]
  10. Kundu, C.K.; Wang, X.; Song, L.; Hu, Y. Chitosan-based flame retardant coatings for polyamide 66 textiles: One-pot deposition versus layer-by-layer assembly. Int. J. Biol. Macromol. 2020, 143, 1–10. [Google Scholar] [CrossRef]
  11. Xiao, L.H.; Xu, L.L.; Yang, Y.Y.; Zhang, S.; Huang, Y.; Bielawski, C.W.; Geng, J.X. Core-shell structured polyamide 66 nanofibers with enhanced flame retardancy. Acs Omega 2017, 2, 2665–2671. [Google Scholar] [CrossRef] [Green Version]
  12. Colovic, M.; Vasiljevic, J.; Strin, Z.; Korosin, N.C.; Sobak, M.; Simoncic, B.; Demsar, A.; Malucelli, G.; Jerman, I. New sustainable flame retardant DOPO-NH-functionalized polyamide 6 and filament yarn. Chem. Eng. J. 2021, 426, 130760. [Google Scholar] [CrossRef]
  13. Fan, S.; Peng, B.; Yuan, R.C.; Wu, D.Q.; Wang, X.L.; Yu, J.Y.; Li, F.X. A novel Schiff base-containing branched polysiloxane as a self-crosslinking flame retardant for PA6 with low heat release and excellent anti-dripping performance. Compos. Part B-Eng. 2020, 183, 107684. [Google Scholar] [CrossRef]
  14. Lyu, W.Y.; Chen, X.; Li, Y.B.; Cao, S.; Han, Y.M. Thermal stability and heat release effect of flame retarded PA66 prepared by end-pieces capping technology. Compos. Part B-Eng. 2019, 167, 34–43. [Google Scholar] [CrossRef]
  15. Lyu, W.Y.; Cui, Y.H.; Zhang, X.J.; Yuan, J.Y.; Zhang, W. Synthesis, thermal stability, and flame retardancy of PA66, treated with dichlorophenylphsophine derivatives. Des. Monomers Polym. 2016, 19, 420–428. [Google Scholar] [CrossRef] [Green Version]
  16. Zhang, J.; Lian, S.; He, Y.; Cao, X.; Shang, J.; Liu, Q.; Ye, G.; Zheng, K.; Ma, Y. Intrinsically flame-retardant polyamide 66 with high flame retardancy and mechanical properties. RSC Adv. 2021, 11, 433–441. [Google Scholar] [CrossRef] [PubMed]
  17. Li, Y.Y.; Liu, K.; Xiao, R. Preparation and characterizations of flame retardant polyamide 66 fiber. IOP Conf. Ser. Mater. Sci. Eng. 2017, 213, 012040. [Google Scholar] [CrossRef]
  18. Fang, Y.Z.; Miao, J.S.; Yang, X.W.; Zhu, Y.; Wang, G.Y. Fabrication of polyphosphazene covalent triazine polymer with excellent flame retardancy and smoke suppression for epoxy resin. Chem. Eng. J. 2020, 385, 123830. [Google Scholar] [CrossRef]
  19. Holdsworth, A.F.; Horrocks, A.R.; Kandola, B.K. Potential synergism between novel metal complexes and polymeric brominated flame retardants in Polyamide 6.6. Polymers 2020, 12, 1543. [Google Scholar] [CrossRef]
  20. Feng, H.S.; Li, D.H.; Cheng, B.; Song, T.L.; Yang, R.J. A cross-linked charring strategy for mitigating the hazards of smoke and heat of aluminum diethylphosphonate/polyamide 6 by caged octaphenyl polyhedral oligomeric silsesquioxanes. J. Hazard. Mater. 2022, 424, 127420. [Google Scholar] [CrossRef] [PubMed]
  21. Luo, H.; Rao, W.; Zhao, P.; Wang, L.; Liu, Y.; Yu, C. An efficient organic/inorganic phosphorus-nitrogen-silicon flame retardant towards low-flammability epoxy resin. Polym. Degrad. Stab. 2020, 178, 109195. [Google Scholar] [CrossRef]
  22. Rao, W.; Zhao, P.; Yu, C.; Zhao, H.-B.; Wang, Y.-Z. High strength, low flammability, and smoke suppression for epoxy thermoset enabled by a low-loading phosphorus-nitrogen-silicon compound. Compos. Part B Eng. 2021, 211, 108640. [Google Scholar] [CrossRef]
  23. Xu, Y.; Qiu, Y.; Yan, C.; Liu, L.; Xu, M.; Xu, B.; Li, B. A novel and multifunctional flame retardant nucleating agent towards superior fire safety and crystallization properties for biodegradable poly (lactic acid). Adv. Powder Technol. 2021, 32, 4210–4221. [Google Scholar] [CrossRef]
  24. Yang, T.; Wu, Y.; Cheng, Y.; Huang, T.; Yu, B.; Zhu, M.; Yu, H. Synthesis of a charring agent containing triazine and benzene groups and its intumescent flame retardant performance for polypropylene. Polym. Degrad. Stab. 2022, 204, 110107. [Google Scholar] [CrossRef]
  25. Hu, X.; Yang, H.; Jiang, Y.; He, H.; Liu, H.; Huang, H.; Wan, C. Facile synthesis of a novel transparent hyperbranched phosphorous/nitrogen-containing flame retardant and its application in reducing the fire hazard of epoxy resin. J. Hazard. Mater. 2019, 379, 120793. [Google Scholar] [CrossRef]
  26. Liu, X.; Zhou, Y.; Peng, H.; Hao, J. Catalyzing charring effect of solid acid boron phosphate on dipentaerythritol during the thermal degradation and combustion. Polym. Degrad. Stab. 2015, 119, 242–250. [Google Scholar] [CrossRef]
  27. Levchik, S.; Piotrowski, A.; Weil, E.; Yao, Q. New developments in flame retardancy of epoxy resins. Polym. Degrad. Stab. 2005, 88, 57–62. [Google Scholar] [CrossRef]
  28. Wang, H.W.; Qiao, H.; Guo, J.; Sun, J.; Li, H.F.; Zhang, S.; Gu, X.Y. Preparation of cobalt-based metal organic framework and its application as synergistic flame retardant in thermoplastic polyurethane (TPU). Compos. Part B-Eng. 2020, 182, 107498. [Google Scholar] [CrossRef]
  29. Zou, J.H.; Duan, H.J.; Chen, Y.S.; Ji, S.; Cao, J.F.; Ma, H.R. A P/N/S-containing high-efficiency flame retardant endowing epoxy resin with excellent flame retardance, mechanical properties and heat resistance. Compos. Part B-Eng. 2020, 199, 108228. [Google Scholar] [CrossRef]
  30. Ding, J.; Tang, S.; Chen, X.; Ding, M.; Kang, J.; Wu, R.; Fu, Z.; Jin, Y.; Li, L.; Feng, X.; et al. Introduction of benzotriazole into graphene oxide for highly selective coadsorption of An and Ln: Facile synthesis and theoretical study. Chem. Eng. J. 2018, 344, 594–603. [Google Scholar] [CrossRef]
  31. Liao, C.; Zhao, X.-R.; Jiang, X.-Y.; Teng, J.; Yu, J.-G. Hydrothermal fabrication of novel three-dimensional graphene oxide-pentaerythritol composites with abundant oxygen-containing groups as efficient adsorbents. Microchem. J. 2020, 152, 104288. [Google Scholar] [CrossRef]
  32. Yuan, Y.; Shi, Y.; Yu, B.; Zhan, J.; Zhang, Y.; Song, L.; Ma, C.; Hu, Y. Facile synthesis of aluminum branched oligo(phenylphosphonate) submicro-particles with enhanced flame retardance and smoke toxicity suppression for epoxy resin composites. J. Hazard. Mater. 2020, 381, 121233. [Google Scholar] [CrossRef] [PubMed]
  33. Zhan, Z.; Shi, J.; Zhang, Y.; Zhang, Y.; Zhang, B.; Liu, W. The study on flame retardancy synergetic mechanism of magnesium oxide for PA66/AlPi composite. Mater. Res. Express 2019, 6, 115317. [Google Scholar] [CrossRef]
  34. Kundu, C.K.; Song, L.; Hu, Y. Sol-gel coatings from DOPO-alkoxysilanes: Efficacy in fire protection of polyamide 66 textiles. Eur. Polym. J. 2020, 125, 109483. [Google Scholar] [CrossRef]
  35. Kundu, C.K.; Wang, X.; Rahman, M.Z.; Song, L.; Hu, Y. Application of chitosan and DOPO derivatives in fire protection of polyamide 66 textiles: Towards a combined gas phase and condensed phase activity. Polym. Degrad. Stab. 2020, 176, 109158. [Google Scholar] [CrossRef]
  36. Liu, W.; Shi, R.; Ge, X.G.; Huang, H.; Chen, X.S.; Mu, M. A bio-based flame retardant coating used for polyamide 66 fabric. Prog. Org. Coat. 2021, 156, 106271. [Google Scholar] [CrossRef]
  37. Rahman, M.Z.; Kundu, C.K.; Wang, X.; Song, L.; Hu, Y. Bio-inspired and dual interaction-based layer-by-layer assembled coatings for superior flame retardancy and hydrophilicity of polyamide 6.6 textiles. Eur. Polym. J. 2021, 147, 110320. [Google Scholar] [CrossRef]
  38. Kundu, C.K.; Song, L.; Hu, Y. Multi elements-based hybrid flame retardants for the superior fire performance of polyamide 66 textiles. J. Taiwan Inst. Chem. Eng. 2021, 118, 284–293. [Google Scholar] [CrossRef]
  39. Feng, X.; Wang, X.; Cai, W.; Hong, N.; Hu, Y.; Liew, K.M. Integrated effect of supramolecular self-assembled sandwich-like melamine cyanurate/MoS2 hybrid sheets on reducing fire hazards of polyamide 6 composites. J. Hazard. Mater. 2016, 320, 252–264. [Google Scholar] [CrossRef]
  40. Lu, P.; Zhao, Z.Y.; Xu, B.R.; Li, Y.M.; Deng, C.; Wang, Y.Z. A novel inherently flame-retardant thermoplastic polyamide elastomer. Chem. Eng. J. 2020, 379, 122278. [Google Scholar] [CrossRef]
  41. Chen, X.; Xu, D.; Zhang, H.; Feng, X.; Deng, J.; Pan, K. In situ polymerization of flame retardant modification polyamide 6,6 with 2-carboxy ethyl (phenyl) phosphinic acid. J. Appl. Polym. Sci. 2020, 137, 48687. [Google Scholar] [CrossRef]
  42. Li, Y.; Liu, K.; Zhang, J.; Xiao, R. Preparation and characterizations of inherent flame retarded polyamide 66 containing the phosphorus linking pendent group. Polym. Adv. Technol. 2018, 29, 951–960. [Google Scholar] [CrossRef]
  43. Lyu, W.; Cui, Y.; Zhang, X.; Yuan, J.; Zhang, W. Thermal stability, flame retardance, and mechanical properties of polyamide 66 modified by a nitrogen–phosphorous reacting flame retardant. J. Appl. Polym. Sci. 2016, 133, 43538. [Google Scholar] [CrossRef]
  44. Lyu, W.; Cui, Y.; Zhang, X.; Yuan, J.; Zhang, W. Fire and thermal properties of PA 66 resin treated with poly-N-aniline-phenyl phosphamide as a flame retardant. Fire Mater. 2017, 41, 349–361. [Google Scholar] [CrossRef]
  45. Zhang, H.; Lu, J.; Yang, H.; Yang, H.; Lang, J.; Zhang, Q. Synergistic flame-retardant mechanism of dicyclohexenyl aluminum hypophosphite and nano-silica. Polymers 2019, 11, 1211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Luo, D.; Liu, Y.; Shi, Y.; Wang, Q. The properties of flame retardant and heat conduction polyamide 66 based on melamine cyanurate/aluminum diethylphosphinate/graphene. J. Polym. Res. 2019, 26, 216. [Google Scholar] [CrossRef]
  47. Zhang, H.; Lu, J.; Yang, H.; Lang, J.; Yang, H. Comparative study on the flame-retardant properties and mechanical properties of PA66 with different dicyclohexyl hypophosphite acid metal salts. Polymers 2019, 11, 1956. [Google Scholar] [CrossRef] [Green Version]
  48. Guo, S.; Bao, M.; Ni, X. The synthesis of meltable and highly thermostable triazine-DOPO flame retardant and its application in PA66. Polym. Adv. Technol. 2021, 32, 815–828. [Google Scholar] [CrossRef]
Figure 1. TG (a) and DTG (b) curves for PA66 and the PA66/Di−PE composites in air conditions.
Figure 1. TG (a) and DTG (b) curves for PA66 and the PA66/Di−PE composites in air conditions.
Polymers 15 01183 g001
Figure 2. Curves of HRR (a); THR (b); SPR (c); and TSP (d).
Figure 2. Curves of HRR (a); THR (b); SPR (c); and TSP (d).
Polymers 15 01183 g002
Figure 3. Mass (a), peak MLR, and av−MLR (b); av-HRR and av-EHC (c); CO production rate (d); CO2 production rate (e); av-CO-CO2 (f) for PA66 and the PA66/Di−PE composites.
Figure 3. Mass (a), peak MLR, and av−MLR (b); av-HRR and av-EHC (c); CO production rate (d); CO2 production rate (e); av-CO-CO2 (f) for PA66 and the PA66/Di−PE composites.
Polymers 15 01183 g003
Figure 4. SEM images of the carbon layer on the surface of PA66 (a,e); PA66/2 wt% Di−PE (b,f); PA66/4 wt% Di−PE (c,g); and PA66/6 wt% Di−PE (d,h) at different magnifications. The red triangle refers to the fractured layer, and the green triangle refers to the dense layer.
Figure 4. SEM images of the carbon layer on the surface of PA66 (a,e); PA66/2 wt% Di−PE (b,f); PA66/4 wt% Di−PE (c,g); and PA66/6 wt% Di−PE (d,h) at different magnifications. The red triangle refers to the fractured layer, and the green triangle refers to the dense layer.
Polymers 15 01183 g004
Figure 5. Raman spectra of char residues for PA66 (a); PA66/2 wt% Di−PE (b); PA66/4% Di−PE (c); and PA66/6% Di−PE (d); the fitting equation was Gaussian.
Figure 5. Raman spectra of char residues for PA66 (a); PA66/2 wt% Di−PE (b); PA66/4% Di−PE (c); and PA66/6% Di−PE (d); the fitting equation was Gaussian.
Polymers 15 01183 g005
Figure 6. ATR-FTIR analyses of PA66 and PA66/Di−PE composites (a); FTIR spectra of the PA66/Di−PE composites’ char residues (b); O1 s XPS spectra of PA66 (c), and O1 s XPS spectra of PA66/6% Di−PE (d).
Figure 6. ATR-FTIR analyses of PA66 and PA66/Di−PE composites (a); FTIR spectra of the PA66/Di−PE composites’ char residues (b); O1 s XPS spectra of PA66 (c), and O1 s XPS spectra of PA66/6% Di−PE (d).
Polymers 15 01183 g006
Figure 7. Relationship between absorbance and time for typical evolved products from pure PA66 and PA66/6 wt% Di−PE: (a) total absorption; (b) NH3; (c) hydrocarbons; (d) carbonyl compounds; (e) CO; (f) CO2.
Figure 7. Relationship between absorbance and time for typical evolved products from pure PA66 and PA66/6 wt% Di−PE: (a) total absorption; (b) NH3; (c) hydrocarbons; (d) carbonyl compounds; (e) CO; (f) CO2.
Polymers 15 01183 g007
Figure 8. Tensile strength (a); tensile modulus (b), and elongation at break (c) of PA66 and the PA66/Di−PE composites samples.
Figure 8. Tensile strength (a); tensile modulus (b), and elongation at break (c) of PA66 and the PA66/Di−PE composites samples.
Polymers 15 01183 g008
Figure 9. DSC cooling curves (a) and heating curves (b) of PA66 and the PA66/Di−PE composites.
Figure 9. DSC cooling curves (a) and heating curves (b) of PA66 and the PA66/Di−PE composites.
Polymers 15 01183 g009
Figure 10. Tensile curves for PA66 fibers (a) and PA66/4 wt% Di−PE fibers (b).
Figure 10. Tensile curves for PA66 fibers (a) and PA66/4 wt% Di−PE fibers (b).
Polymers 15 01183 g010
Figure 11. Photographs of PA66 (a) and PA66/4 wt% Di−PE (b) before and after the vertical burning test of the fabric samples.
Figure 11. Photographs of PA66 (a) and PA66/4 wt% Di−PE (b) before and after the vertical burning test of the fabric samples.
Polymers 15 01183 g011
Table 1. Melt spinning parameters for PA66 and the PA66/Di−PE composites.
Table 1. Melt spinning parameters for PA66 and the PA66/Di−PE composites.
Melt Spinning ParametersValue
Barrel temperatures of zone one (°C)280
Barrel temperatures of zone two (°C)300
Barrel temperatures of zone three (°C)295
Barrel temperatures of zone four (°C)295
Temperature of transfer pipe (°C)300
Temperature of metering pump (°C)300
Temperature of spinneret (°C)300
Head pressure of extruder (bar)50
Speed (r/min)22
Speed of metering pump (r/min)25
Number of spinneret orifices36
Diameter of spinneret orifices (mm)0.3
L/D of spinneret orifices3.0
Take-up velocity, VL (m/min)800
Distance from the spinneret to the first roller (m)3.5
Table 2. TGA results for PA66 and the PA66/Di−PE composites under air.
Table 2. TGA results for PA66 and the PA66/Di−PE composites under air.
SpecimenT5% (°C)Tmax (°C)Char Yields at 600 °C (wt%)Char Yields at 700 °C (wt%)
Pure PA66396.3477.72.050
PA66/2 wt% Di−PE376.8465.71.930
PA66/4 wt% Di−PE362.8473.83.820.53
PA66/6 wt% Di−PE352.7467.03.601.31
Table 3. The results of LOI and UL-94 tests of PA66 and PA66/Di−PE composite.
Table 3. The results of LOI and UL-94 tests of PA66 and PA66/Di−PE composite.
SamplePA66Di−PELOI (%)UL-94 Rating
1100023.5V-2
298228.8V-2
396429.3V-0
494629.4V-0
Table 4. Relevant parameters for PA66 and FRPA66 obtained from the calorimetric test.
Table 4. Relevant parameters for PA66 and FRPA66 obtained from the calorimetric test.
SamplesPA66PA66/2 wt% Di−PEPA66/4 wt% Di−PEPA66/6 wt% Di−PE
TTI (s)75615150
PHRR (kW/m2)1059.81130.1685.4558.7
av-HRR (kW/m2)359.6311.3262.0112.8
THR (MJ/m2)115.1121.478.960.1
EHC (MJ/kg)30.429.719.722.1
PMLR (g/s)0.3410.3260.3070.233
av-MLR (g/s)0.1050.0920.0980.045
COY (kg/kg)0.0160.0250.0230.022
CO2Y (kg/kg)2.3152.2891.1941.223
PSPR (m2/s)0.1030.1140.0960.058
TSP (m2)8.409.226.244.64
FRI10.721.532.42
Table 5. Tensile properties of PA66 and the PA66/Di−PE composites samples.
Table 5. Tensile properties of PA66 and the PA66/Di−PE composites samples.
SampleTensile Strength (MPa)Tensile Modulus (MPa)Elongation at Break (%)
PA6677.7 ± 0.91216.3 ± 26.771.7 ± 8.7
PA66/2 wt% Di−PE79.5 ± 1.71212.2 ± 72.634.4 ± 5.7
PA66/4 wt% Di−PE77.4 ± 0.71187.5 ± 22.932.1 ± 9.1
PA66/6 wt% Di−PE76.6 ± 0.91129.7 ± 12.918.0 ± 10.2
Table 6. DSC data for PA66 and the PA66/Di−PE composites.
Table 6. DSC data for PA66 and the PA66/Di−PE composites.
SamplesTc (°C)Tm (°C)Hm (J/g)Xc (%)
PA66233.7262.771.831.8
PA66/2 wt% Di−PE226.7257.080.135.4
PA66/4 wt% Di−PE227.3257.375.533.4
PA66/6 wt% Di−PE227.0255.772.131.9
Table 7. Tensile properties of PA66 and PA66/Di−PE fibers.
Table 7. Tensile properties of PA66 and PA66/Di−PE fibers.
SampleLinear Density (dtex/36F)Draft RatioElongation at Break (%)Tensile Strength (cN/dtex)Modulus (cN/dtex)
PA6691.94.412.0 ± 0.66.4 ± 0.139.6 ± 2.1
PA66/4 wt% Di−PE115.24.816.1 ± 1.85.7 ± 0.231.7 ± 1.5
Table 8. Flame retardance of PA66 and PA66/Di−PE composites fabric samples.
Table 8. Flame retardance of PA66 and PA66/Di−PE composites fabric samples.
SampleLOI (%)After-Flame Time (s)Damaged Length (cm)Ignite Absorbent CottonBurn Behavior
PA66 fabric23.15.0 ± 0.2615.6 ± 0.6yesMelts, drips, and shrinks away from the flame
PA66/4 wt% Di−PE fabric28.62.0 ± 0.1911.1 ± 0.6noSelf-extinguishing within 3 s, minimum damaged length.
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

Wu, Y.; Yang, T.; Cheng, Y.; Huang, T.; Yu, B.; Wu, Q.; Zhu, M.; Yu, H. Flame Retardancy and Mechanical Properties of Melt-Spun PA66 Fibers Prepared by End-Group Blocking Technology. Polymers 2023, 15, 1183. https://doi.org/10.3390/polym15051183

AMA Style

Wu Y, Yang T, Cheng Y, Huang T, Yu B, Wu Q, Zhu M, Yu H. Flame Retardancy and Mechanical Properties of Melt-Spun PA66 Fibers Prepared by End-Group Blocking Technology. Polymers. 2023; 15(5):1183. https://doi.org/10.3390/polym15051183

Chicago/Turabian Style

Wu, Yanpeng, Tonghui Yang, Yongchang Cheng, Tao Huang, Bin Yu, Qilin Wu, Meifang Zhu, and Hao Yu. 2023. "Flame Retardancy and Mechanical Properties of Melt-Spun PA66 Fibers Prepared by End-Group Blocking Technology" Polymers 15, no. 5: 1183. https://doi.org/10.3390/polym15051183

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