Next Article in Journal
Pembrolizumab Plus Axitinib for Metastatic Papillary and Chromophobe Renal Cell Carcinoma: NEMESIA (Non Clear MEtaStatic Renal Cell Carcinoma Pembrolizumab Axitinib) Study, a Subgroup Analysis of I-RARE Observational Study (Meet-URO 23a)
Next Article in Special Issue
Flame-Retardant Cycloaliphatic Epoxy Systems with High Dielectric Performance for Electronic Packaging Materials
Previous Article in Journal
Revealing the Tick Microbiome: Insights into Midgut and Salivary Gland Microbiota of Female Ixodes ricinus Ticks
Previous Article in Special Issue
Preparation, Properties, and Mechanism of Flame-Retardant Poly(vinyl alcohol) Aerogels Based on the Multi-Directional Freezing Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

High-Efficient Flame-Retardant Finishing of Cotton Fabrics Based on Phytic Acid

College of Textiles and Clothing, Institute of Functional Textiles and Advanced Materials, National Engineering Research Center for Advanced Fire-Safety Materials D & A (Shandong), State Key Laboratory of Bio-Fibers and Eco-Textiles, Qingdao University, Qingdao 266071, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(2), 1093; https://doi.org/10.3390/ijms24021093
Submission received: 20 December 2022 / Revised: 2 January 2023 / Accepted: 4 January 2023 / Published: 6 January 2023
(This article belongs to the Special Issue Advanced Flame Retardant Polymeric Materials)

Abstract

:
In this study, an efficient phosphorus-containing flame retardant, PAPBTCA, was synthesized from phytic acid, pentaerythritol, and 1,2,3,4-butane tetracarboxylic acid, and its structure was characterized. PAPBTCA was finished on cotton fabrics by the pad-dry-curing process, and the flame retardancy, flame-retardant durability, and wrinkle resistance of the obtained flame-retardant fabrics were investigated. It should be noted that the heat release rate value of the flame-retardant cotton fabrics treated with 200 g/L PAPBTCA decreased by 90% and its excellent flame retardancy was maintained after 5 washing cycles. Meanwhile, the wrinkle resistance of flame-retardant cotton fabrics has been significantly improved. In addition, compared with the control, the breaking force loss of PAPBTCA-200 in the warp and weft directions was 24% and 21%, respectively. This study provides a new way to utilize natural phosphorus-based flame retardants to establish multifunctional finishing for cotton fabrics.

1. Introduction

Cellulose, as one of the natural polymer materials, is widely used because it is green, renewable, and abundant in resources. Cotton fabrics are favored among natural cellulose fabrics because of their comfort, breathability, and softness [1,2,3]. Due to their composition unit of glucose, cotton fabrics have a low limiting oxygen index (LOI) value of only 18%, making them flammable, which limits the application of cotton fabrics in various fields [4]. It has been reported that the annual casualties and economic losses caused by fires are unimaginable, so people’s awareness of fire safety is increasing, and the development of fire safety textile research is also increasing [5,6]. However, considering the ongoing advancements in science and technology and the increasing needs of people, only the flame-retardant function of cotton fabrics cannot meet the expectations of people in many aspects [7]. Therefore, making cotton fabrics versatile on the basis of excellent flame retardancy is a hot spot for many researchers [8,9,10].
Currently, phosphorus-based flame retardants with good flame retardancy are a common choice for flame-retardant natural cellulose fabrics [11,12]. When cellulose is burned at higher temperatures, many phosphorus-based flame retardants dehydrate and decompose to produce metaphosphoric acid or polyphosphoric acid. Metaphosphoric acid or polyphosphoric acid promotes cellulose to dehydrate and carbonize into char residue [13,14]. However, partial phosphorus-based flame retardants also have problems, including a waste of resources and environmental pollution. As the world’s major environmental problems are becoming serious, environmentally friendly flame retardants free of formaldehyde and halogens are becoming increasingly popular [15,16,17]. Because of its excellent biocompatibility, environmental friendliness, renewability, and degradability, phytic acid (PA) containing 28% phosphorus in its structure, an organic phosphonic acid produced from plants, has become a popular flame retardant in recent years [18]. Ma et al. used PA and urea to finish cotton fabrics to provide them with flame retardancy, and the finished fabrics had excellent flame retardancy and durability [19]. Liu et al. also prepared flame retardants by reacting PA with urea and finishing them on lyocell fabrics. The LOI value of finished lyocell fabrics reached 39.2% and was maintained at 29.7% after 30 washing cycles. This study achieved a combination of highly efficient flame retardancy and good durability [20]. However, PA is highly acidic and can cause great damage to the mechanical properties of cellulose fabrics when used alone as a flame retardant [21]. To solve these problems, it is a feasible method to introduce a suitable carbon-forming agent into the PA system to reduce the damage that PA brings to cellulose fabrics. Pentaerythritol (PER) is a small molecule carbon-forming agent that was used in the early days, but its water solubility and easy precipitation made it ineffective when used in an intumescent flame-retardant system (IFR). Zhu’s team designed an organic–inorganic hybrid phosphorus–boron nitride synergistic flame retardant using PA, PER, and boric acid. This system provided excellent flame retardancy and durability to cotton fabrics with lower weight gain [22]. BTCA (1,2,3,4-butane tetracarboxylic acid) has been extensively studied as a polycarboxylic, acid-based, anti-wrinkle finishing agent [23]. Since it has four carboxyl groups, it can be used to couple fabrics and flame retardants, and the residual BTCA during the finishing process can provide an anti-wrinkle finishing effect to cotton fabrics.
Cheng et al. successfully synthesized a flame retardant through PA, PER, and BTCA for wool fabrics, with the flame retardant linked to wool through electrostatic interaction and ester bonding, resulting in good flame-retardant properties for wool fabrics after 20 washing cycles [24]. This work was excellent but the system was not applied to cotton fabrics. In order to investigate whether this system also has a bright effect on the durability of flame-retardant cotton fabrics, PA, PER, and BTCA were used to synthesize PAPBTCA. The hydroxyl group in PER is highly reactive, so it can be esterified with the phosphate group in PA to obtain the flame retardant PAP. PAP can be cross-linked with BTCA to form the cross-linkable flame retardant PAPBTCA, and it can react with cotton fabric’s hydroxyl groups to create long-lasting flame-retardant cotton fabrics. A series of characterizations of the synthesized PAPBTCA have been performed and the properties and mechanism of the flame-retardant cotton fabrics have been examined.

2. Results and Discussion

2.1. FTIR of the Control, PAP-200, and PAPBTCA-200

The structure of the control, PAP-200, PAPBTCA-200, and PAPBTCA-200 after 5 washing cycles (PAPBTCA-200-5Ls) was analyzed by FTIR. The results in Figure 1 show that a new characteristic C=O peak at 1724 cm−1 appeared in PAPBTCA-200, which was attributed to the ester bond in PAPBTCA and the ester bond formed by the esterification reaction between PAPBTCA and the fabrics [25]. It is noteworthy that PAPBTCA-200-5Ls can still observe a significant C=O peak, and this indicates that there was still PAPBTCA deposited on the surface of the fabrics after 5 washing cycles. The results of FTIR show that PAPBTCA successfully deposited on the surface of cotton fabrics.

2.2. Surface Morphology of the Control, PAP-200, and PAPBTCA-200

SEM was performed on the control, PAP-200, and PAPBTCA-200 to determine their surface morphology. The resulting pictures are displayed in Figure 2. It is clear that the control has a natural twist, and also an irregular spiral twist in the longitudinal plane. Compared with the control, the surface morphology of PAP-200 and PAPBTCA-200 did not change much, but there was a significant deposition on the surface of the fibers. Compared with PAP-200, PAPBTCA-200 was deposited more obviously on the fabrics, and the cotton fibers became thicker. SEM micrographs proved that the exterior of the fabrics had been successfully coated with flame retardants.

2.3. Flame Retardancy

To investigate the flame retardancy of cotton fabrics treated with varying concentrations of PAP and PAPBTCA, respectively, VFT and LOI tests were used. The images and data obtained from the test are presented in Figure 3 and Table 1. The results show that the addition of BTCA had no discernible impact on the flame-retardant effect of PAP. Moreover, PAP-200 lost its flame-retardant property after the washing durability test. PAPBTCA-100 had a good flame-retardant effect with 17.3% weight gain and the LOI value reached 29.5%, while the LOI value of PAPBTCA-200 reached 31.5% when the weight gain was 20.3%. After 5 washing cycles, the weight gain was still 16.8% and the LOI value was 29.3%, which also passed the VFT test. This indicates that PAPBTCA-200 improves the washing durability of flame-retardant finishing due to the cross-linking effect of BTCA.

2.4. Thermal Stabilities

TG analysis was performed to examine the effect of PAP and PAPBTCA on the thermal degradation process of cotton fabrics. Figure 4 displays the samples’ TG and DTG curves. The data derived from these curves are summarized in Table 2.
The samples were significantly degraded between 250 and 450 °C with a maximum loss of weight, and this was brought on by the dehydration and carbonization of cotton fibers, which resulted in the production of volatiles and aliphatic char residues [26,27]. As the temperature increased, the cotton fabrics experienced further thermal degradation, releasing CO2 and CO, etc. In a N2 atmosphere, Rmax2 was observed at 345 °C and 409 °C for PAP-200 and PAPBTCA-200, respectively, which may be due to the thermal degradation of unstable intermediates. The char residues of PAP-200 and PAPBTCA-200 increased to 37.68% and 21.28%, respectively, at 700 °C compared with 3.65% char residue for the control. The altered thermal degradation pathway of cotton fabrics facilitated the formation of protective and stable char residue. Phosphoric acid or polyphosphate produced from the thermal degradation of PA aided in the dehydration and carbonization of cotton fabrics to produce stable char residue. The results show that the addition of flame retardant effectively inhibited the thermal decomposition process of cotton fabrics and improved the char formation ability of cotton fabrics in the high-temperature range [28]. The initial decomposition temperatures (T5%) of PAP-200 and PAPBTCA-200 were also lower than the control during the thermal oxidative degradation in air atmosphere. Compared with the control, the earlier Rmax1 of PAP-200 and PAPBTCA-200 was caused by the earlier decomposition of PAP and PAPBTCA, while the later Rmax2 reflected that the formed intermediates were not thermal stable to produce the more thermal stable char residue [28]. However, they had fewer char residues at 700 °C, with values of only 7.24% and 2.86%, and those were related to the subsequent thermal oxidation of the generated char residues in air.
In summary, the addition of PAPBTCA reduced the thermal stability of cotton fabrics at low temperatures, but increased the number of char residues at high temperatures. These results indicated that PAP and PAPBTCA enhanced the production of stable char residues during the thermal degradation of cotton fabrics, and that increased their thermal stability in the high-temperature range. From Figure 4 and Table 2, the results show that the maximum thermal decomposition rate (Rmax) values of both PAP-200 and PAPBTCA-200 were reduced, and that the amount of char residue was much higher than that of the control; however, the advantage of PAPBTCA-200 was smaller and the amount of char residue was less, compared with those of PAP-200. This may be due to the lower P content of PAPBTCA compared with that of PAP in the case of similar weight gain, resulting in a relatively weaker flame-retardant efficiency.

2.5. Burning Behaviors

CCT can simulate the real situation of fires and examine the burning characteristics of the control, PAP-200, and PAPBTCA-200. The samples’ curves from CCT are displayed in Figure 5, while Table 3 displays the CCT data.
The time to ignition (TTI) of the control was approximately 18 s, while PAP-200 and PAPBTCA-200 were not ignited to produce a flame at the tested heat flux. This situation showed that the treated cotton fabrics were hard to ignite. The peak heat release rate (PHRR) of the control was 110 kW/m2, comparatively, and PHRR values of PAP-200 and PAPBTCA-200 were decreased by 88% and 90%, respectively. Furthermore, the average heat release rate (Av-HRR) values of PAP-200 and PAPBTCA-200 were also reduced by 54% and 57%, respectively. This indicates that PAP and PAPBTCA effectively inhibited heat release during the burning process, while the lower PHRR value means that the burning process is more likely to be interrupted. The total heat release (THR) value of the control was 4.5 MJ/m2, and those of PAP-200 and PAPBTCA-200 were 2.3 and 2.1 MJ/m2, respectively. In comparison, they dropped by 49% and 53%, respectively. In summary, PAP and PAPBTCA reduced the heat-release properties of cotton fabrics.
One of the main factors that contributes to fire deaths is smoke [29,30]. Current studies also focus on the smoke suppression capabilities of flame-retardant materials [31]. As can be observed from Figure 5 and Table 3, the total smoke production (TSP) values of PAP-200 and PAPBTCA-200 decreased from 2.8 m2 to 0.4 and 0.5 m2, respectively. This indicates that the addition of PAP and PAPBTCA has a good smoke suppression effect, and PA facilitates the carbonization and dehydration of cellulose fibers to create a barrier, resulting in the inhibited effect on the smoke release at the condensed phase [32]. This also indicates that PAP and PAPBTCA play a key role in suppressing smoke released during burning. In addition, the information presented by the CO production rate (COP) and CO2 production rate (CO2P) in Figure 5 cannot be ignored. It is obvious that the COP values of PAP-200 and PAPBTCA-200 were significantly increased, while the CO2P values were seriously decreased. Those phenomena might be caused by the flame retardants promoting the decomposition of cotton fabrics into char residue, and the dense and stable char layers isolating the external heat and air, resulting in the incomplete burning of the fabrics. This was also attributed to the significantly higher char residue of PAP-200 and PAPBTCA-200. The fire growth index (FIGRA) indicates the degree of danger of a material in a fire, which can be found by PHRR/TPHRR. A decrease in FIGRA can indicate that people have a greater chance to escape in the event of a fire. From the test results, it can be seen that the FIGRA values of PAP-200 and PAPBTCA-200 are significantly lower compared with the control. The above results show that PAP-200 and PAPBTCA-200 have higher flame retardancy and smoke suppression performance.
The char residues after CCT for the control, PAP-200, and PAPBTCA-200 were imaged by SEM, as presented in Figure 6. As observed in Figure 6(A1–A3), after CCT, the control was turned into messy and fragile ash that did not maintain its fabric structure. By comparison, the char residue of PAP-200 and PAPBTCA-200 showed a significant difference. In Figure 6(B1–C3), the char residue of treated cotton fabrics had only some fractures and cracks caused by heat, but they generally maintained their original fiber structure and relatively complete shape. Moreover, the char residual morphology of treated cotton fabrics was relatively similar. The relatively undamaged char residue can prevent further burning by blocking heat and oxygen, which also reduces THR derived from CCT.

2.6. Flame-Retardant Mechanism

2.6.1. Analysis of the Structure of Char Residues

The elemental composition and different types of chemical bonds of the char residue after CCT were examined using XPS. C and P elements in char residues of the control, PAP-200, and PAPBTCA-200 are shown in Figure 7. In Figure 7A, compared with the control, PAP-200 and PAPBTCA-200 showed a new P2p peak at 130.0 eV. This indicates that the char residue contained P element. They were the characteristic peaks of P-O-C groups in the cross-linked char residue produced by the thermal degradation of the flame retardant and the P=O in the phosphate structure [33]. In Figure 7(C1–C3), the peaks at 284.6, 285.8, and 288.3 eV were responsible for C-C, C-O-P/C-OH, and C=O groups, respectively [34,35]. C elements in char residue of treated cotton fabrics mainly contained C-C, C-O-P, and C=O bonds, meaning that the flame retardant participated in the creation of char residue. Based on the previous research, PAP and PAPBTCA formed phosphoric acid or polyphosphate acid during the thermal degradation process. These two substances stimulated the carbonization and dehydration of cellulose, and also the formation of char residues. These char residue can not only isolate oxygen and heat sources but also restrain the further burning of cotton fabrics.
The char residue after CCT of PAP-200 and PAPBTCA-200 was also examined by Raman spectra, shown in Figure 8. The signature D and G peaks in Raman spectra were positioned at approximately 1300 cm−1 and 1580 cm−1, respectively. The ID/IG value was calculated by the area ratio of peak D to peak G. The density of graphite imperfections is represented by the ID/IG value, which is related to the severity of structural flaws in graphite. The ID/IG value of char residue from the control was 8.46, while those of PAP-200 and PAPBTCA-200 were 2.63 and 2.67, respectively. The ID/IG values of PAP-200 and PAPBTCA-200 were much lower compared with that of the control, demonstrating that the completed char residues had a higher degree of graphitization. In addition, the char residue with higher degrees of graphitization improved the thermal stability to stop heat transmission, thus enhancing the flame retardancy.

2.6.2. TG-FTIR

In order to analyze the flame-retardant mechanism of this system, the FTIR 3D spectra of the control, PAP-200, and PAPBTCA-200 obtained from TG-FTIR are displayed in Figure 9. From Figure 9A–C, the absorption peaks of the control, PAP-200, and PAPBTCA-200 are similar. However, most absorption peaks of PAP-200 and PAPBTCA-200 were weaker than those of the control. The –OH groups were responsible for the 3589 cm−1 peaks, which can be found in gaseous water and other substances containing –OH groups. The absorbance of the C-H groups appeared at approximately 2914 cm−1 [21], with the absorption intensities of the C-H groups lower than those of the control. Hydrocarbons are very common flammable chemicals, and these results showed that fewer hydrocarbons were produced during the thermal decomposition of PAP-200 and PAPBTCA-200. The peak at 2360 cm−1 was attributed to CO2 stretching vibration, the peak at 1745 cm−1 was due to the stretching vibration of C=O groups, and the peak at 1121 cm−1 was ascribed to the stretching vibration of C-O-C groups. Figure 9 clearly demonstrates that the absorption intensities of CO2 for PAP-200 and PAPBTCA-200 were considerably lower than that of the control. The above results indicate that more carbon remained in the condensed phase. However, the absorption intensities of the C=O and C-O-C groups for PAP-200 and PAPBTCA-200 were also slightly lower than those of the control, clearly showing that PAP and PAPBTCA promote the dehydration and carbonization of cellulose to create char residue rather than the production of flammable volatile gases and CO2 [36].
Six spectra from the aforementioned groups are shown in Figure 9a–f, including H2O (3589 cm−1), C–H (2914 cm−1), CO2 (2360 cm−1), C–O (1745 cm−1), CO (2182 cm−1), and C–O–C (1121 cm−1). Among them, the absorption intensities of CO2 for PAP-200 and PAPBTCA-200 were lower than that of the control, which is compatible with the above conclusion of the FTIR 3D spectra. The lower absorption intensities of PAP-200 and PAPBTCA-200 indicate that flammable gases were reduced to prevent the burning. In addition, compared with the control, the degraded product peaks of PAP-200 and PAPBTCA-200 occurred earlier, indicating that the addition of flame retardants catalyzed and accelerated the dehydration and carbonization of cellulose [37,38]. In summary, PAP and PAPBTCA primarily achieved an outstanding flame-retardant effect by inhibiting the production of flammable gases within the gas phase and encouraging the formation of char residue in the condensed phase.

2.7. Anti-Wrinkle Performance

The crease recovery angle test was conducted on the control and treated cotton fabrics to analyze whether the crease resistance was improved or not, with Table 4 displaying the pertinent information. As is evident from the crease recovery angle data, in comparison with the control, there were no noticeable changes in the crease phenomenon of PAP-200. However, PAPBTCA-200 showed an increase in crease recovery angle from 160.8° to 184.7°, an increase of 14.9% compared with that of the control. The slow crease recovery angle increased from 168.7° to 230.9°, an increase of 36.9%, indicating that PAPBTCA resulted in an improvement in the crease resistance of the cotton fabrics [39,40]. The improved anti-wrinkle property may be due to the cross-linking of excess BTCA with the –OH groups of the cotton fabrics, which acted as an anti-wrinkle agent [41]. From the results, it is evident that the anti-wrinkle performance of PAPBTCA-200 was significantly improved compared with those of the control and PAP-200.

2.8. Breaking Force

Breaking force was tested for the control, PAP-200, and PAPBTCA-200. The results are shown in Figure 10. Compared with the control, the breaking force of PAP-200 was significantly reduced, with 62% reduction in the warp direction and 61% reduction in the weft direction. However, the breaking force loss of PAPBTCA-200 was alleviated. Compared with the control, the breaking force loss of PAPBTCA-200 in the warp and weft directions was 24% and 21%, respectively. This suggests that the addition of BTCA can further reduce the acidity of PA compared to PAP, thus having a greater certain protective effect on the fabrics’ strength [42]. There was also another reason for this phenomenon: the films formed by PAPBTCA-200 were thicker than the films formed by PAP-200 (shown in Figure 2), making the cotton fibers less prone to breakage.

3. Materials and Methods

3.1. Materials

Cotton fabrics with 100 g/m2 were offered by Qingdao No. 6 Printing and Dyeing Co., Ltd. (Qingdao, China); ammonia and NaOH were supplied by Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China); PA (70% aqueous solution), PER, and BTCA were offered by Macklin Chemical Reagent Co., Ltd. (Shanghai, China). All the chemicals were used as received.

3.2. Preparation

3.2.1. Synthesis of PAP and PAPBTCA

PA (0.01 mol, 6.6 g) and PER (0.03 mol, 4.1 g) were mixed in a three-necked flask (250 mL) and heated for 2 h at 130 °C under magnetic stirring to obtain a viscous and little yellow liquid (PAP). The reaction was then performed at 130 °C with magnetic stirring for a further 1 h after the addition of BTCA (0.03 mol, 7.0 g) to the solution. Finally, the transparent liquid became more elastic and yellow. The raw product was freeze-dried and then refined using ethanol. PAPBTCA was prepared according to [24]. Figure S1 shows the relevant characterization of PAPBTCA.

3.2.2. Preparation of Flame-Retardant Cotton Fabrics

Cotton fabrics with suitable sizes were submerged into 1% NaOH solution at 100 °C for 1 h, washed, and dried at 80 °C. The unpurified PAPBTCA was dissolved in distilled water to obtain 100 g/L and 200 g/L solutions for treating cotton fabrics, which were named PAPBTCA-100 and PAPBTCA-200, respectively. The cotton fabrics were also treated with 100 g/L and 200 g/L of PAP solutions as a comparison and referred to as PAP-100 and PAP-200, respectively. Then, 3 wt% sodium hypophosphite was added to the flame-retardant solutions as a catalyst. The cotton fabrics were soaked in the finishing solution for 30 min, with two paddings, then dried at 80 °C for 3 min and, lastly, the treated cotton fabrics were cured at 170 °C for 3 min. The samples treated with PAPBTCA were thoroughly washed with deionized water to eliminate unbound PAPBTCA. In the last step, the samples were dried at 80 °C to determine the weight gain. Figure 11 shows the preparation process. Figure S2 shows the reactions that occur in the preparation of flame retardant cotton fabrics.

3.3. Characterizations

3.3.1. FTIR

Attenuated total reflection Fourier transform infrared spectroscopy (ATR-FTIR) was collected on a Nicolet IS50 FTIR spectrometer (Thermo Fisher Scientific, Waltham, MA, USA) to measure the chemical structure of PA, PAP, PAPBTCA, the control, PAP-200, and PAPBTCA-200 in a spectrum range of 4000∼500 cm−1 and a resolution of 2 cm−1. The detailed results of characterized PAPBTCA are presented in supporting information.

3.3.2. NMR

A Bruker AVANCE III HD 400 MHz spectrometer (Bruker, Bremen, Germany) was used to measure the 31P nuclear magnetic resonance (31P NMR) with DMSO as the solvent. The detailed results of characterized PAPBTCA are presented in supporting information.

3.3.3. ICP-OES

An inductively coupled plasma emission spectrometer (ICP-OES) (PE Avio 200, USA) was used to detect the P element content of PAPBTCA. Quantitative analysis of elements was obtained by accepting emission spectra at different wavelengths.

3.3.4. SEM

A scanning electron microscope (SEM) analyzer (Tescan Vega 3, TESCAN, Brno, Czech Republic) with an accelerating voltage of 10 kV was used to image the surface morphology of the control, PAP-200, and PAPBTCA-200, before and after the cone calorimeter test (CCT). Due to the SEM test, the sample surface received a gold sputtering treatment.

3.3.5. Vertical Flame Test

The vertical flame test (VFT) was conducted using an LFY-601A vertical flame tester (Shandong Textile Science Research Institute, Qingdao, China). Based on GB/T 5455-2014, the samples were cut into 300 × 80 mm2 segments and ignited for 12 s by a 40 ± 2 mm flame. Each sample was examined at least five times.

3.3.6. LOI

A JF-5 oxygen index gauge (Beijing Avic Times Instrument Equipment Co., Ltd., Beijing, China) was used to conduct the limiting oxygen index (LOI) test. According to GB/T 5454-1997, the samples were processed into 150 × 58 mm2 segments.

3.3.7. Washing Durability Test

The washing procedure of all samples was performed according to AATCC Test Method 61-2003 test NO.1A. The treated fabrics were washed using the oscillating bath method. The fabrics to be washed were impregnated in conical bottles containing 0.37 wt% detergent. Next, the conical flasks were transferred to an oscillating water bath (40 °C) and every 45 min of oscillation was determined as 5 washing cycles.

3.3.8. TG

The thermal stability of the samples was tested using a STA6000 TG (Perkin-Elmer Ltd., Waltham, MA, USA). Under nitrogen and air atmosphere, respectively, the temperature range was 50–700 °C, the heating rate was 10 °C/min, and the flow rate was 25 mL/min.

3.3.9. CCT

A cone calorimeter (Fire Testing Technology, East Grinstead, UK) with a heat flux of 35 kW/m2 was used to examine the burning behaviors of the samples. The samples were cut into 10 × 10 mm2 segments according to the ISO 5660-1. At least three evaluations of each sample were performed.

3.3.10. XPS

X-ray photoelectron spectroscopy (XPS) was conducted through an XSAM 800 spectrometer (Kratos Co., Stretford, UK), with 1486.6 eV Al Kα excitation radiation, 12 kV operating voltage, and 15 mA working current selected for this test. The C 1s line was located at the binding energy of 284.0 eV.

3.3.11. Raman Spectra

A Laser Confocal Raman Spectrometer (Thermo Scientific DXR2, Waltham, MA, USA) was used to obtain the Raman spectra of the remaining char residue after CCT.

3.3.12. TG-FTIR

A Perkin-Elmer STA 6000 thermogravimetric analyzer (TG) (Perkin-Elmer Ltd., Waltham, MA, USA) fitted with a Perkin-Elmer Fourier transform infrared spectrometer (FTIR) (Perkin-Elmer Ltd., Waltham, MA, USA) was used. The temperature range for the TG-FTIR test was 40 to 750 °C, with a heating rate of 10 °C/min and a flow rate of 50 mL/min, respectively, in nitrogen atmosphere. FTIR was calibrated to have a spectrum range of 4000 to 500 cm−1 and a resolution of 2 cm−1.

3.3.13. Anti-Wrinkle Performance Test

The crease recovery property of samples was tested by an LFY-210A/B Fabric crease recovery tester (Qingdao Shanfang Instrument Co., Ltd., Qingdao, China).

3.3.14. Breaking Force

The breaking force was tested on an Instron 5967 electromechanical universal testing machine (Instron Limited, Co., Norwood, MA, USA). The samples were prepared into 300 × 60 mm2 segments based on GB/T 3923.1-2013. Each sample was evaluated at least five times.

4. Conclusions

In this study, an environmentally friendly, reactive, and highly efficient flame retardant, PAPBTCA, was successfully synthesized by PA, PER, and BTCA. The flame retardant was grafted onto the surface of the cotton fabrics by reacting with the hydroxyl groups of the cotton fabrics and had certain washing durability. Compared with that of the control, the HRR value of PAPBTCA-200 was reduced by 90%, while the smoke release was also significantly reduced. XPS and TG-FTIR tests showed that PAPBTCA had a condensed-phase flame-retardant mechanism. In addition, the crease recovery angle of cotton fabrics after PAPBTCA treatment was significantly improved. Compared with PAP-200, the PAPBTCA-200 breaking force retention has significantly improved. This study shows that biomass containing phosphorus flame retardants has great potential for multifunctional finishing applications.

Supplementary Materials

The supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms24021093/s1. References [43,44,45] were cited in Supplementary Materials.

Author Contributions

Formal analysis, W.-M.S. and L.-Y.Z.; Investigation, W.-M.S.; Resources, W.-M.S.; Data curation, W.-M.S.; Writing—original draft, W.-M.S. and L.-Y.Z.; Writing—review & editing, P.L. and Y.L.; Supervision, Y.L.; Project administration, Y.L.; Funding acquisition, Y.L. All authors have read and agreed to the published version of the manuscript.

Funding

The authors are grateful for the financial support from the National Natural Science Foundation of China (Grant Nos. 51973098 and 51991354). This work was also partly supported by the State Key Laboratory of Bio-Fibers and Eco-Textiles (Qingdao University), No. ZDKT202107.

Data Availability Statement

The data are available upon request.

Conflicts of Interest

The authors declare that they have no conflict of interest.

References

  1. Dong, C.; Lu, Z.; Zhang, F.; Zhu, P.; Zhang, L.; Sui, S. Preparation and Properties of Cotton Fabrics Treated with a Novel Polysiloxane Water Repellent and Flame Retardant. Mater. Lett. 2015, 152, 276–279. [Google Scholar] [CrossRef]
  2. El-Shafei, A.; Al-Shemy, M.; Abou-Okeil, A. Eco-Friendly Finishing Agent for Cotton Fabrics to Improve Flame Retardant and Antibacterial Properties. Carbohydr. Polym. 2015, 118, 83–90. [Google Scholar] [CrossRef] [PubMed]
  3. Dong, C.; Lu, Z.; Zhang, F.; Zhu, P.; Wang, P.; Che, Y.; Sui, S. Combustion Behaviors of Cotton Fabrics Treated by a Novel Nitrogen- and Phosphorus-Containing Polysiloxane Flame Retardant. J. Therm. Anal. Calorim. 2016, 123, 535–544. [Google Scholar] [CrossRef]
  4. Gao, W.-W.; Zhang, G.-X.; Zhang, F.-X. Enhancement of Flame Retardancy of Cotton Fabrics by Grafting a Novel Organic Phosphorous-Based Flame Retardant. Cellulose 2015, 22, 2787–2796. [Google Scholar] [CrossRef]
  5. Liu, W.; Chen, L.; Wang, Y.-Z. A Novel Phosphorus-Containing Flame Retardant for the Formaldehyde-Free Treatment of Cotton Fabrics. Polym. Degrad. Stab. 2012, 97, 2487–2491. [Google Scholar] [CrossRef]
  6. Zhu, W.; Hao, S.; Yang, M.; Cheng, B.; Zhang, J. A Synergistic Flame Retardant of Glycosyl Cross-Linking Boron Acid and Ammonium Salt of Phytic Acid to Enhance Durable Flame Retardancy of Cotton Fabrics. Cellulose 2020, 27, 9699–9710. [Google Scholar] [CrossRef]
  7. Wang, Z.-H.; Liu, B.-W.; Zeng, F.-R.; Lin, X.-C.; Zhang, J.-Y.; Wang, X.-L.; Wang, Y.-Z.; Zhao, H.-B. Fully Recyclable Multifunctional Adhesive with High Durability, Transparency, Flame Retardancy, and Harsh-Environment Resistance. Sci. Adv. 2022, 8, eadd8527. [Google Scholar] [CrossRef]
  8. Birnbaum, L.S.; Staskal, D.F. Brominated Flame Retardants: Cause for Concern? Environ. Health Perspect. 2004, 112, 9–17. [Google Scholar] [CrossRef] [Green Version]
  9. Horrocks, A.R.; Kandola, B.K.; Davies, P.J.; Zhang, S.; Padbury, S.A. Developments in Flame Retardant Textiles—A Review. Polym. Degrad. Stab. 2005, 88, 3–12. [Google Scholar] [CrossRef]
  10. Liao, C.; Li, Y.; Gao, M.; Xia, Y.; Chai, W.; Su, X.; Zheng, Z.; Liu, Y. Bio-Inspired Construction of Super-Hydrophobic, Eco-Friendly Multifunctional and Bio-Based Cotton Fabrics via Impregnation Method. Colloids Surf. A-Physicochem. Eng. Asp. 2022, 651, 129647. [Google Scholar] [CrossRef]
  11. Xia, X.; Zhang, X.; Chen, Y.; Xiong, Y.; Weijian, X.W. Synthesis and Characterization of Phosphorus-Containing Epoxy Resin for Flame Retardance. Chem. Ind. Eng. Prog. 2007, 26, 56–59. [Google Scholar]
  12. Liao, Y.; Chen, Y.; Zhang, F. A Biological Reactive Flame Retardant for Flame Retardant Modification of Cotton Fabric. Colloids Surf. A-Physicochem. Eng. Asp. 2021, 630, 127601. [Google Scholar] [CrossRef]
  13. Li, Z.-F.; Zhang, C.-J.; Cui, L.; Zhu, P.; Yan, C.; Liu, Y. Fire Retardant and Thermal Degradation Properties of Cotton Fabrics Based on Aptes and Sodium Phytate through Layer-by-Layer Assembly. J. Anal. Appl. Pyrolysis 2016, 123, 216–223. [Google Scholar] [CrossRef]
  14. Schartel, B. Phosphorus-Based Flame Retardancy Mechanisms-Old Hat or a Starting Point for Future Development? Materials 2010, 3, 4710–4745. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Li, P.; Wang, B.; Xu, Y.J.; Jiang, Z.; Dong, C.; Liu, Y.; Zhu, P. Ecofriendly Flame-Retardant Cotton Fabrics: Preparation, Flame Retardancy, Thermal Degradation Properties, and Mechanism. ACS Sustain. Chem. Eng. 2019, 7, 19246–19256. [Google Scholar] [CrossRef]
  16. Alongi, J.; Carletto, R.A.; Di Blasio, A.; Carosio, F.; Bosco, F.; Malucelli, G. DNA: A Novel, Green, Natural Flame Retardant and Suppressant for Cotton. J. Mater. Chem. A 2013, 1, 4779–4785. [Google Scholar] [CrossRef]
  17. Zhang, Y.; Tian, W.; Liu, L.; Cheng, W.; Wang, W.; Liew, K.M.; Wang, B.; Hu, Y. Eco-Friendly Flame Retardant and Electromagnetic Interference Shielding Cotton Fabrics with Multi-Layered Coatings. Chem. Eng. J. 2019, 372, 1077–1090. [Google Scholar] [CrossRef]
  18. Angel, R.; Tamim, N.M.; Applegate, T.J.; Dhandu, A.S.; Ellestad, L.E. Phytic Acid Chemistry: Influence on Phytin-Phosphorus Availability and Phytase Efficacy. J. Appl. Poult. Res. 2002, 11, 471–480. [Google Scholar] [CrossRef]
  19. Ma, Y.; Luo, X.; Liu, L.; Zhang, C.; Shang, X.; Yao, J. Eco-Friendly, Efficient and Durable Fireproof Cotton Fabric Prepared by a Feasible Phytic Acid Grafting Route. Cellulose 2021, 28, 3887–3899. [Google Scholar] [CrossRef]
  20. Liu, X.-H.; Zhang, Q.-Y.; Cheng, B.-W.; Ren, Y.-L.; Zhang, Y.-G.; Ding, C. Durable Flame Retardant Cellulosic Fibers Modified with Novel, Facile and Efficient Phytic Acid-Based Finishing Agent. Cellulose 2018, 25, 799–811. [Google Scholar] [CrossRef]
  21. Li, Z.; Expósito, D.F.; González, A.J.; Wang, D.Y. Natural Halloysite Nanotube Based Functionalized Nanohybrid Assembled via Phosphorus-Containing Slow Release Method: A Highly Efficient Way to Impart Flame Retardancy to Polylactide. Eur. Polym. J. 2017, 93, 458–470. [Google Scholar] [CrossRef]
  22. Zhu, W.; Yang, M.; Huang, H.; Dai, Z.; Cheng, B.; Hao, S. A Phytic Acid-Based Chelating Coordination Embedding Structure of Phosphorus-Boron-Nitride Synergistic Flame Retardant to Enhance Durability and Flame Retardancy of Cotton. Cellulose 2020, 27, 4817–4829. [Google Scholar] [CrossRef]
  23. Xu, W.; Li, Y. Cotton Fabric Strength Loss from Treatment with Polycarboxylic Acids for Durable Press Performance. Text. Res. J. 2000, 70, 957–961. [Google Scholar] [CrossRef]
  24. Cheng, X.W.; Guan, J.P.; Kiekens, P.; Yang, X.H.; Tang, R.C. Preparation and Evaluation of an Eco-Friendly, Reactive, and Phytic Acid-Based Flame Retardant for Wool. React. Funct. Polym. 2019, 134, 58–66. [Google Scholar] [CrossRef]
  25. Yang, C.Q.; Wang, X. Formation of Cyclic Anhydride Intermediates and Esterification of Cotton Cellulose by Multifunctional Carboxylic Acids: An Infrared Spectroscopy Study. Text. Res. J. 1996, 66, 595–603. [Google Scholar] [CrossRef]
  26. Zhang, K.; Zong, L.; Tan, Y.; Ji, Q.; Yun, W.; Shi, R.; Xia, Y. Improve the Flame Retardancy of Cellulose Fibers by Grafting Zinc Ion. Carbohydr. Polym. 2016, 136, 121–127. [Google Scholar] [CrossRef]
  27. Shi, Y.; Yu, B.; Duan, L.; Gui, Z.; Wang, B.; Hu, Y.; Yuen, R.K.K. Graphitic Carbon Nitride/Phosphorus-Rich Aluminum Phosphinates Hybrids as Smoke Suppressants and Flame Retardants for Polystyrene. J. Hazard. Mater. 2017, 332, 87–96. [Google Scholar] [CrossRef]
  28. Cheng, X.-W.; Tang, R.-C.; Guan, J.-P.; Zhou, S.-Q. An Eco-Friendly and Effective Flame Retardant Coating for Cotton Fabric Based on Phytic Acid Doped Silica Sol Approach. Prog. Org. Coat. 2020, 141, 105539. [Google Scholar] [CrossRef]
  29. Zhou, K.; Gui, Z.; Hu, Y. The Influence of Graphene Based Smoke Suppression Agents on Reduced Fire Hazards of Polystyrene Composites. Compos. Part A-Appl. Sci. Manuf. 2016, 80, 217–227. [Google Scholar] [CrossRef]
  30. Jiang, Q.; Li, P.; Wang, B.; She, J.-H.; Liu, Y.; Zhu, P. Inorganic-Organic Hybrid Coatings from Tea Polyphenols and Laponite to Improve the Fire Safety of Flexible Polyurethane Foams. Colloids Surf. A-Physicochem. Eng. Asp. 2022, 655, 105539. [Google Scholar] [CrossRef]
  31. Zhou, L.L.; Li, W.X.; Zhao, H.B.; Wang, J.S.; Zhao, B. Niti-Layered Double Hydroxide Nanosheets toward High-Efficiency Flame Retardancy and Smoke Suppression for Silicone Foam. Polym. Degrad. Stab. 2022, 204, 105539. [Google Scholar] [CrossRef]
  32. Chen, X.; Wang, W.; Jiao, C. A Recycled Environmental Friendly Flame Retardant by Modifying Para-Aramid Fiber with Phosphorus Acid for Thermoplastic Polyurethane Elastomer. J. Hazard. Mater. 2017, 331, 257–264. [Google Scholar] [CrossRef]
  33. Pan, Y.; Liu, L.; Zhang, Y.; Song, L.; Hu, Y.; Jiang, S.; Zhao, H. Effect of Genipin Crosslinked Layer-by-Layer Self-Assembled Coating on the Thermal Stability, Flammability and Wash Durability of Cotton Fabric. Carbohydr. Polym. 2019, 206, 396–402. [Google Scholar] [CrossRef]
  34. Wang, P.; Zhang, J.; Chen, Z.G. Surface Structure of Pyrolytic Char of Cellulose Based on Xps Analysis. J. Combust. Sci. Technol. 2015, 21, 378–381. [Google Scholar]
  35. 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]
  36. Li, P.; Wang, B.; Liu, Y.-Y.; Xu, Y.-J.; Jiang, Z.-M.; Dong, C.-H.; Zhang, L.; Liu, Y.; Zhu, P. Fully Bio-Based Coating from Chitosan and Phytate for Fire-Safety and Antibacterial Cotton Fabrics. Carbohydr. Polym. 2020, 237, 109483. [Google Scholar] [CrossRef]
  37. Feng, Y.; Zhou, Y.; Li, D.; He, S.; Zhang, F.; Zhang, G. A Plant-Based Reactive Ammonium Phytate for Use as a Flame-Retardant for Cotton Fabric. Carbohydr. Polym. 2017, 175, 636–644. [Google Scholar] [CrossRef]
  38. Wang, S.; Xu, D.; Zhu, P.; Jiang, Z. Facile Fabrication of Antibacterial and Fire-Safety Multifunctional Cotton Fabric with a Triazine-Phosphonate N-Halamine. Ind. Crops Prod. 2022, 186, 115261. [Google Scholar] [CrossRef]
  39. Lou, J.; Wang, D.; Yuan, J.; Xu, J.; Fan, X. Improving the Anti-Wrinkle and Hydrophilicity Performance of Cotton Fabric via Crosslinking Cellulose with Carboxylated Polyaldehyde Trehalose. Cellulose 2021, 28, 5135–5149. [Google Scholar] [CrossRef]
  40. Lou, J.; Wang, D.; Fan, X. Study on the Cross-Linking Process of Carboxylated Polyaldehyde Sucrose as an Anti-Wrinkle Finishing Agent for Cotton Fabric. Sci. Rep. 2022, 12, 5379. [Google Scholar] [CrossRef]
  41. Huang, K.-S.; Hwang, M.-C.; Chen, J.-S.; Lin, S.-J.; Wang, S.-P. Application of Mixed Gel Solution in the Anti-Wrinkle Finishing of Cotton Fabrics. J. Text. Inst. 2007, 98, 169–176. [Google Scholar] [CrossRef]
  42. Ghanadpour, M.; Carosio, F.; Larsson, P.T.; Wågberg, L. Phosphorylated Cellulose Nanofibrils: A Renewable Nanomaterial for the Preparation of Intrinsically Flame-Retardant Materials. Biomacromolecules 2015, 16, 3399–3410. [Google Scholar] [CrossRef] [PubMed]
  43. Jiang, G.; Qiao, J.; Hong, F. Application of Phosphoric Acid and Phytic Acid-Doped Bacterial Cellulose as Novel Proton-Conducting Membranes to PEMFC. Int. J. Hydrogen Energy 2012, 37, 9182–9192. [Google Scholar] [CrossRef]
  44. Guo, X.; Miao, Y.; Ye, P.; Wen, Y.; Yang, H. Multi-Walled Carbon Nanotubes in Aqueous Phytic Acid for Enhancing Biosensor. Mater. Res. Express 2014, 1, 025403. [Google Scholar] [CrossRef]
  45. Bebot-Brigaud, A.; Dange, C.; Fauconnier, N.; Gérard, C. P-31 Nmr, Potentiometric and Spectrophotometric Studies of Phytic Acid Ionization and Complexation Properties toward Co2+, Ni2+, Cu2+, Zn2+ and Cd2+. J. Inorg. Biochem. 1999, 75, 71–78. [Google Scholar] [CrossRef]
Figure 1. FTIR spectra of the treated and untreated cotton fabrics.
Figure 1. FTIR spectra of the treated and untreated cotton fabrics.
Ijms 24 01093 g001
Figure 2. SEM micrographs (×400 and ×2000) of untreated and treated cotton fabrics: Control (A1,A2), PAP-200 (B1,B2), PAPBTCA-200 (C1,C2).
Figure 2. SEM micrographs (×400 and ×2000) of untreated and treated cotton fabrics: Control (A1,A2), PAP-200 (B1,B2), PAPBTCA-200 (C1,C2).
Ijms 24 01093 g002
Figure 3. VFT photos of the samples: PAP-100 (A), PAPBTCA-100 (B), PAP-200 (C), PAPBTCA-200 (D), PAP-200-5Ls (E), PAPBTCA-200-5Ls (F).
Figure 3. VFT photos of the samples: PAP-100 (A), PAPBTCA-100 (B), PAP-200 (C), PAPBTCA-200 (D), PAP-200-5Ls (E), PAPBTCA-200-5Ls (F).
Ijms 24 01093 g003
Figure 4. TG and DTG curves of the control, PAP-200, and PAPBTCA-200 in N2 (A1,A2) and air (B1,B2).
Figure 4. TG and DTG curves of the control, PAP-200, and PAPBTCA-200 in N2 (A1,A2) and air (B1,B2).
Ijms 24 01093 g004
Figure 5. HRR (a), COP (b), CO2P (c), THR (d), TSP (e), and SPR (f) curves of the control, PAP-200, and PAPBTCA-200.
Figure 5. HRR (a), COP (b), CO2P (c), THR (d), TSP (e), and SPR (f) curves of the control, PAP-200, and PAPBTCA-200.
Ijms 24 01093 g005
Figure 6. SEM micrographs (×400 and ×2000) of the control (A1A3), PAP-200 (B1B3), and PAPBTCA-200 (C1C3) after CCT.
Figure 6. SEM micrographs (×400 and ×2000) of the control (A1A3), PAP-200 (B1B3), and PAPBTCA-200 (C1C3) after CCT.
Ijms 24 01093 g006
Figure 7. XPS spectra of char residue (A), P2p of PAP-200 (B1), P2p of PAPBTCA-200 (B2), C1s of control cotton (C1), C1s of PAP-200 (C2), and C1s of PAPBTCA-200 (C3).
Figure 7. XPS spectra of char residue (A), P2p of PAP-200 (B1), P2p of PAPBTCA-200 (B2), C1s of control cotton (C1), C1s of PAP-200 (C2), and C1s of PAPBTCA-200 (C3).
Ijms 24 01093 g007
Figure 8. Raman spectra of char residues of control cotton (A), PAP-200 (B), and PAPBTCA-200 (C) after CCT.
Figure 8. Raman spectra of char residues of control cotton (A), PAP-200 (B), and PAPBTCA-200 (C) after CCT.
Ijms 24 01093 g008
Figure 9. Three-dimensional TG-FTIR spectra of the control cotton (A), PAP-200 (B), and PAPBTCA-200 (C), and the absorption intensities of selected peaks during the thermal degradation process of samples (af).
Figure 9. Three-dimensional TG-FTIR spectra of the control cotton (A), PAP-200 (B), and PAPBTCA-200 (C), and the absorption intensities of selected peaks during the thermal degradation process of samples (af).
Ijms 24 01093 g009
Figure 10. Breaking force of the control, PAP-200, and PAPBTCA-200.
Figure 10. Breaking force of the control, PAP-200, and PAPBTCA-200.
Ijms 24 01093 g010
Figure 11. The preparation process of flame-retardant cotton fabrics.
Figure 11. The preparation process of flame-retardant cotton fabrics.
Ijms 24 01093 g011
Table 1. Flame-retardant performance of the samples.
Table 1. Flame-retardant performance of the samples.
SamplesConcentration of Flame Retardants (g/L)Weight Gain (%)Afterflame Time (s)Afterglow Time (s)Damaged Length (mm)LOI
(%)
Control/0.0173430018.0
PAP-10010015.9 ± 1.10077 ± 529.7
PAPBTCA-10010017.3 ± 1.50079 ± 629.5
PAP-20020021.2 ± 0.40071 ± 432.6
PAPBTCA-20020020.3 ± 1.40062 ± 531.5
PAP-200-5Ls2003.2 ± 0.360300 ± 020.5
PAPBTCA-200-5Ls20016.8 ± 0.70098 ± 629.3
Table 2. Thermogravimetric analysis data of the control, PAP-200, and PAPBTCA-200 in N2 and air.
Table 2. Thermogravimetric analysis data of the control, PAP-200, and PAPBTCA-200 in N2 and air.
AtmosphereSamplesT5%
(°C)
T10%
(°C)
Tmax1
(°C)
Rmax1 (%/°C)Tmax2
(°C)
Rmax2
(%/°C)
Residue at 700 °C (%)
N2Control cotton3223673591.36--3.65
PAP-2002132332490.483450.1037.68
PAPBTCA-2002883153170.744090.1621.82
AirControl cotton3113313431.324620.170.48
PAP-2002102322480.485210.177.24
PAPBTCA-2002833103120.715000.242.86
Table 3. CCT data of the control, PAP-200, and PAPBTCA-200.
Table 3. CCT data of the control, PAP-200, and PAPBTCA-200.
SamplesTTI
(s)
PHRR
(kW/m2)
Av-HRR
(kW/m2)
THR
(MJ/m2)
TSP
(m2)
FIGRA
kW/(m2·s)
Residue
(wt%)
Control cotton18 ± 311015.94.52.82.765.5
PAP-200- *147.32.30.40.1213.1
PAPBTCA-200- *116.82.10.50.0720.7
Note: * PAP-200 and PAPBTCA-200 were not ignited.
Table 4. Crease recovery angle test of the control, PAP-200, and PAPBTCA-200.
Table 4. Crease recovery angle test of the control, PAP-200, and PAPBTCA-200.
SamplesCrease Sharp Return Angle (°)Total (°)Crease Slow Return Angle (°)Total (°)
WarpWeftWarpWeft
Control cotton79.980.9160.883.585.2168.7
PAP-20081.286.1167.386.691.1177.7
PAPBTCA-20091.693.1184.7114.7116.2230.9
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Song, W.-M.; Zhang, L.-Y.; Li, P.; Liu, Y. High-Efficient Flame-Retardant Finishing of Cotton Fabrics Based on Phytic Acid. Int. J. Mol. Sci. 2023, 24, 1093. https://doi.org/10.3390/ijms24021093

AMA Style

Song W-M, Zhang L-Y, Li P, Liu Y. High-Efficient Flame-Retardant Finishing of Cotton Fabrics Based on Phytic Acid. International Journal of Molecular Sciences. 2023; 24(2):1093. https://doi.org/10.3390/ijms24021093

Chicago/Turabian Style

Song, Wan-Meng, Li-Yao Zhang, Ping Li, and Yun Liu. 2023. "High-Efficient Flame-Retardant Finishing of Cotton Fabrics Based on Phytic Acid" International Journal of Molecular Sciences 24, no. 2: 1093. https://doi.org/10.3390/ijms24021093

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