Next Article in Journal
Microstructural and Morphological Characterization of the Cobalt-Nickel Thin Films Deposited by the Laser-Induced Thermionic Vacuum Arc Method
Next Article in Special Issue
Rheological Properties of Composite Inorganic Micropowder Asphalt Mastic
Previous Article in Journal
Performance-Based Expressway Asphalt Pavement Structural Surface Layer Modulus Matching Mode Research
Previous Article in Special Issue
Sprayed-Polyurea-Modified Asphalt: Optimal Preparation Parameters, Rheological Properties and Thermal Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mechanism of Sodium Dodecyl Diphenyl Ether Disulfonate Filled Hydrotalcite Inhibiting the Photo-Degradation of Polyvinyl Chloride under Different Ranges of Ultraviolet Wavelength Irradiation

1
Key Laboratory of Clean Chemistry Technology of Guangdong Regular Higher Education Institutions, Guangdong Engineering Technology Research Center of Modern Fine Chemical Engineering, School of Chemical Engineering and Light Industry, Guangdong University of Technology, Guangzhou 510006, China
2
Jieyang Branch of Chemistry and Chemical Engineering Guangdong Laboratory (Rongjiang Laboratory), Jieyang 515200, China
3
GCH Technology Co., Ltd., Guangzhou 510540, China
*
Authors to whom correspondence should be addressed.
Coatings 2023, 13(6), 985; https://doi.org/10.3390/coatings13060985
Submission received: 30 March 2023 / Revised: 17 May 2023 / Accepted: 19 May 2023 / Published: 25 May 2023
(This article belongs to the Special Issue Recent Development in Novel Green Asphalt Materials for Pavement)

Abstract

:
This content introduces a novel Ultraviolet (UV)-shielding material, Zn2Al-MADS-LDH (MADS-LDH), which was synthesized through co-precipitation method to insert sodium dodecyl diphenyl ether disulfonate (MADS) into the interlayer of Zn2Al-LDH layered double hydroxide (LDH), to improve the photoaging resistance of polyvinyl chloride (PVC). The characterization results indicated that MADS-LDH had a host-guest interaction between the LDH host layer and MADS guest anion, and it exhibited superior UV absorption capabilities than Zn2Al-CO3-LDH (CO3-LDH) and a broader absorption spectrum compared to MADS. A series of LDHs/PVC film composite materials containing LDHs nanosheets were prepared by incorporating the prepared LDHs into a PVC matrix via a solvent casting method. As expected, the MADS-LDH/PVC film composite materials exhibited enhanced photoaging resistance. The results of photoaging tests indicated that MADS-LDH inhibits the rate of carbonyl generation during photoaging of MADS-LDH/PVC film composite materials, resulting in a decrease in the carbonyl index (ΔCl) and relative degradation rate (RDR) compared to pristine PVC film and CO3-LDH/PVC film composite materials. Furthermore, the study evaluated the influence of different UV light wavelength ranges, such as UVB (280~315 nm), UVC (200~280 nm), and UV (200~400 nm), on the aging performance of PVC film and LDHs/PVC film composite materials. The results demonstrated that UV had the highest aging effect on PVC composite films, followed by UVC and UVB. Therefore, the MADS-LDH is a highly efficient and promising UV-shielding material with excellent potential for wide applications in the field of PVC.

1. Introduction

Polyvinyl chloride (PVC) is a highly versatile and commonly used polymer materials in modern industries [1,2]. However, as a polymer material, PVC is susceptible to degradation and aging under the influence of external factors such as light, water, heat, and oxygen. This degradation can lead to discoloration, brittleness, and hardening, which ultimately reduces its physical properties [3,4]. Research has shown that UV light from sunlight is the primary contributor to the aging of polymer materials [3,5]. Therefore, incorporating UV-shielding materials is crucial to improve the stability of PVC film composite materials.
In recent years, there has been a growing interest in the development and utilization of inorganic-organic hybrid nanomaterials, such as hydrotalcite, due to their unique properties that arise from the interaction between the inorganic matrix and the organic material. Layered double hydroxides (LDHs), also known as hydrotalcite-like compounds or anionic clays, are a diverse group of layered materials that can be either natural or synthetic. LDHs are characterized by the general formula: [M2+1−xM3+x(OH)2]x+(An−)x/n·mH2O, where M2+ and M3+ represent divalent and trivalent metal cations, respectively, and An− represents exchangeable anions with negative charge n, which can be organic or inorganic in nature [6].
By regulating its plate and interlayer structure, hydrotalcite can be customized for various functional purposes. It has important applications in various fields including catalysis, environmental protection [7,8,9,10], electrochemistry [11,12], and polymer additives [13,14].
The use of organic UV absorbers and inorganic nanoparticles to delay the photoaging degradation of polymeric materials, such as PVC, is common. However, these materials have certain limitations, such as high lipophilicity [15] and a narrow range of UV wavelengths shielding ability [16]. Moreover, the UV absorption ability of anionic surfactant intercalated hydrotalcite, containing a conjugated structure and dodecyl chain length as UV-shielding materials, is insufficiently studied. Existing studies mainly focus on the adsorption capacity of such anionic surfactants after intercalation of hydrotalcite. They can adsorb heavy metals or organic pollutants through hydrogen bonding and electrostatic interactions after anion insertion into hydrotalcite, and physical adsorption by increased layer spacing [9,17,18,19,20,21].
Although numerous studies have examined the effect of single UV wavelength range on the photoaging of PVC films [22,23], there is a lack of comparison between the results obtained and different wavelength ranges. Additionally, when adding intercalated hydrotalcite powders to polymers, it is important to consider their compatibility and interlayer crystalline water content, a factor that has not been thoroughly explored in previous studies.
To address these gaps in knowledge, this study investigates the mechanism of MADS-LDH fillers to inhibit the photoaging degradation of PVC film composite materials under different UV wavelength range irradiation. This is based on the existing oxidation reaction mechanism of photo-oxidation reaction generating carbonyl during PVC aging [24]. As a result, the objectives of this study are to: (1) study and compare the UV absorption range and absorption intensity between CO3-LDH and MADS-LDH; (2) analyze and discuss the desorption behavior of adsorbed water and crystalline water during the thermal decomposition of MADS-LDH; (3) analyze the morphology and pore size of MADS-LDH fillers, and their effect on structure and compatibility of PVC; (4) investigate the influence of different wavelength ranges of UV light on the MADS-LDH/PVC film composite materials’ irradiation aging behavior, and the mechanism by which MADS-LDH inhibits the photoaging degradation of MADS-LDH/PVC film composite materials.

2. Materials and Methods

2.1. Materials

ZnCl2 (98.0%), AlCl3·6H2O (97.0%), and NaOH (96.0%) were supplied by Shanghai Macklin Biochemical Co., Ltd. (Shanghai, China), Sodium dodecyl diphenyl ether disulfonate (45% aqueous solution) was purchased from Shanghai Aladdin Biochemical Technology Co., Ltd. (Shanghai, China), and ethanol (99.7%) was supplied by Anhui Zesheng Technology Co., Ltd. (Anqing, China). Further, 1, 2-dichloroethane (99.5%) was purchased from Meryer (Shanghai) Biochemical Technology Co., Ltd. (Shanghai, China). PVC (99.0%) and was supplied by Zhangmutou Ruixiang Polymer Materials Business Department, Dongguan City (Dongguan, China). Deionized water was decarbonized by boiling for 5 min before use.

2.2. Preparation of MADS Anion-Intercalated LDH

MADS-LDH was prepared by co-precipitation method as follows: a mixture of 0.625 M ZnCl2 and 0.313 M AlCl3·6H2O (Zn/Al molar ratio = 2:1) was slowly added to 100 mL of MADS solution (0.125 M) with stirring in N2 atmosphere. The pH of the reaction solution was adjusted to 9.0 by dropwise addition of 1 M NaOH alkaline solution. The resulting slurry was crystallized at 70 °C for 24 h. The white precipitates were filtered and separated. They were rinsed with hot decarbonated deionized water and anhydrous ethanol until reaching a state of neutrality. Finally, the sample was subjected to a 24 h drying process at 100 °C, followed by grinding and sieving. Instead of MADS, Na2CO3 was utilized as the precursor. Meanwhile, the preparation of CO3-LDH was exactly following the same method.

2.3. Preparation of LDHs/PVC Composite Materials

Prepared MADS-LDH of 1.0 g dry weight was added to a round bottom flask containing 100 mL of 1,2-dichloroethane for dispersion and stirring for 30 min to ensure uniform dispersion. Then, 10.0 g of PVC was added to the same round-bottom flask and stirred at 25 °C for 2 h. The resulting dispersion system of MADS-LDH and PVC was defoamed and poured onto a flat and smooth glass plate. The dispersion system solution was immediately scraped with a wet applicator and placed in a fume hood for static drying. The resulting film was then cut into dimensions of length × width × height= 30 mm × 20 mm × 0.1 mm was and used for subsequent testing. The reference samples of pure PVC and 10.0 wt% CO3-LDH/PVC film materials were also prepared using the same method described above.

2.4. UV Photoaging Test

The UV photoaging tests were performed in a custom-made UV-aging chamber that was equipped with individually controllable UVA (315~400 nm, 8 W), UVB (280~315 nm, 8 W), and UVC (200~280 nm, 8 W) lamps. The chamber was designed to have a cooling system to eliminate the influence of thermal aging during the photoaging test of PVC films. The PVC films were subjected to photoaging treatment three times with a total UV exposure time of 144 h.

2.5. Characterization

The X-ray diffraction (XRD) patterns were analyzed using a Shimadzu XRD-6000 diffractometer (Shimadzu Corporation, Japan) with monochromatic Cu Kα radiation (λ = 0.15406 nm, 40 kV, 40 mA) from 2 to 70° at a scan speed of 10°·min−1. Fourier transform infrared (FT-IR) spectra were recorded on a Bruker Vector 22 infrared spectrophotometer (Bruker Corporation, Billerica, MA, USA) with a scan range of 4000~400 cm−1 and a resolution of 1 cm−1. The surface and internal morphology of the sample were examined using scanning electron microscope (TESCAN MIRA LMS SEM, TESCAN Orsay Holding, a.s., Brno, Czech Republic) and transmission Electron Microscope (Hitachi HT7700 TEM, Hitachi High Technologies Corporation, Tokyo, Japan). Thermogravimetric analysis (TGA) was carried out on a PCT-IA instrument (Thermal Analysis Instruments, Inc., New Castle, DE, USA) in the range of 25~800 °C with a heating rate of 10 °C·min−1 under N2 atmosphere. BET analysis was performed on a Micromeritics ASAP 2460 instrument (Micromeritics Instrument Corporation, Norcross, GA, USA). The ultraviolet–visible (UV-vis) absorption spectra were obtained using a Shimadzu UV-2501 PC instrument (Shimadzu Corporation, Kyoto, Japan).

3. Results and Discussion

3.1. XRD Analysis

The XRD patterns of powder CO3-LDH and MADS-LDH are shown in Figure 1. These two samples have characteristic diffraction peaks of LDHs materials. From Table 1, the planar spacing of LDHs (003), (006), and (009) showed good multiplicative relationships, with (d003) = 2 (d006) = 3 (d009) [25], confirming the successful preparation of LDHs materials. The XRD spectrum LDH of MADS-LDH (Figure 1) shows only one set of typical (00l) (l = 3, 6, 9) Bragg reflections, indicating the successful intercalation of MADS into the interlayer of Zn2Al-LDH with only a single crystalline phase. The basal spacings (d003) of CO3-LDH and MADS-LDH are approximately 0.75 and 2.87 nm, respectively. Subtracting the LDH plate layer thickness of 0.48 nm [26], the interlayer spacings of CO3-LDH and MADS-LDH are approximately 0.27 and 2.39 nm, respectively. Additionally, we propose a possible orientation for MADS intercalated in Zn2Al-LDH interlayer (Figure 2).
It has been reported by Yanjun Lin et al. [27] that an increase in LDHs layer spacing can improve the stability of PVC films. This is because a larger interlayer spacing facilitates the entry of Cl into the LDHs interlayer channels, thereby inhibiting the autocatalytic dechlorination decomposition of PVC by Cl. Therefore, MADS-LDH, with a layer spacing of 2.39 nm, is expected to improve the UV aging resistance of PVC.

3.2. FT-IR Analysis

The resulting FT-IR spectra of CO3-LDH, MADS-LDH, and MADS are shown in Figure 3. The broad absorption band near 3430 cm−1 in the spectra of CO3-LDH and MADS-LDH can be attributed to the O-H stretching vibration of the LDH plate layer and interlayer water molecules [18,20]. The intense band at very low wavenumber (428 cm−1) was only observed for CO3-LDH and MADS-LDH, which is attributed to the Zn-O-bound for LDH structures [28,29]. The absence of CO32− groups at 1360 cm−1 indicates that there were no CO32− anions in the interlayer of MADS-LDH [30]. The C-H of alkyl chains formed strong peaks at 2958, 2928, and 2858 cm−1, while the C-C of diphenoxy produced strong peaks at 1590 and 1482 cm−1, and the -O- of diphenoxy resulted in a peak at 1245 cm−1 [31,32].
In the FT-IR spectra of MADS-LDH, most of the characteristic absorption bands of MADS were observed, indicating that MADS was successfully intercalated into the interlayer of Zn2Al-LDH. However, some subtle differences were observed between MADS-LDH and MADS. The asymmetric and symmetric vibrational absorption peaks of the -SO3 groups of MADS were shifted from 1190 and 1092 cm−1 to 1186 and 1088 cm−1, respectively, due to the host-guest interaction between the Zn2Al-LDH layer and interlayer MADS anion [31]. These shifts suggest that the environment experienced by MADS anion in the interlayer gallery of LDHs is distinct from that of UV absorber. The successful preparation of a new organic-inorganic UV shielding material, MADS-LDH, is confirmed by XRD and FT-IR analyses. Similarly, the XRD and FT-IR spectra of CO3-LDH (Figure 1 and Figure 3) verified the successful intercalation of CO32− into Zn2Al-LDH.

3.3. SEM and TEM Analysis

Figure 4 illustrates the SEM images of CO3-LDH and MADS-LDH. The obtained CO3-LDH exhibited flaky morphological features, while the SEM image of MADS-LDH showed aggregation of irregular flaky particles. This aggregation is attributed to the intercalation of MADS anions, which leads to the aggregation of Zn2Al-LDH crystals [33]. To further confirm the successful intercalation of MADS anion into Zn2Al-LDH, transmission electron microscopy (TEM) was used, and the obtained TEM image of MADS-LDH was displayed in Figure 5. Clusters with irregularly shaped MADS-LDH nanosheets stacked in the TEM images were observed. The surfactant intercalated samples exhibit a rather different morphology. The agglomeration of MADS-LDH resulted in a decrease in its crystallinity, which is consistent with the XRD results.

3.4. BET Analysis

Figure 6 shows the N2 adsorption–desorption isotherms of CO3-LDH and MADS-LDH. The nitrogen adsorption method revealed that both CO3-LDH and MADS-LDH are mesoporous materials with type IV isotherms and H3 hysteresis loops [34]. The BET surface areas were calculated to be 30.430 m2/g for CO3-LDH and 23.927 m2/g for MADS-LDH [35]. The decrease in the specific surface area of MADS-LDH after MADS anion intercalation is consistent with the increase in the unit mass of MADS-LDH due to agglomeration [36], which further confirms the successful intercalation of MADS anion into Zn2Al-LDH.
Figure 7 presents the pore diameter distribution of CO3-LDH and MADS-LDH. The results show that MADS-LDH has a narrower pore diameter distribution with a maximum effective pore diameter of 4.6 nm compared to CO3-LDH. These findings are consistent with the adsorption isotherm results, which indicate that both CO3-LDH and MADS-LDH have a mesoporous structure with pore diameters ranging from 2 nm to 50 nm. These findings are crucial in the development of new UV-shielding materials based on MADS-LDH, as they provide insights into the specific surface area and porosity of the materials.
Table 2 show Constitutive properties of CO3-LDH and MADS-LDH. As shown in Table 2, the specific surface area of CO3-LDH (SBET = 30.430 m2/g) is much larger than that of MADS-LDH (SBET = 23.927 m2/g). Meanwhile, the pore volume Vp and the pore diameter Dp of both were almost equal. This indicates that the MADS anion intercalation into the interlayer of Zn2Al-LDH has the greatest effect on its specific surface area.

3.5. TGA Analysis

Thermogravimetric analysis (TGA) was utilized to assess the thermal stability of the prepared LDHs. Figure 8 displays the TG curves of CO3-LDH, MADS-LDH, and MADS. The mass loss stage for MADS occurred within the range of 25~765 °C, with 10% mass loss temperature (T0.1) and 50% mass loss temperature (T0.5) at 58 and 156 °C, respectively. In comparison, the weight loss rate of LDHs was substantially lower than that of MADS within the thermal decomposition temperature of 25~800 °C, indicating a better thermal stability.
Specifically, the weight loss rate of MADS-LDH between 120~452 °C was lower than that of CO3-LDH. At this stage, the weight loss for LDHs is primarily caused by the desorption of water of crystallization between its layers [18]. The hydrophobic environment created by the alkyl chain of the MADS anion within the MADS-LDH interlayer made it difficult for water molecules to enter and form crystalline water [36]. Furthermore, the weight loss rate of MADS-LDH between 452~800 °C was higher than that of CO3-LDH due to the thermal decomposition of organic carbon constituted of the MADS anion present in the MADS-LDH interlayer [37]. The detailed T0.1 and T0.5 for MADS and LDHs are listed in Table 3, where ΔT represents the difference between MADS and LDHs at the temperature of T0.1 or T0.5. Notably, the T0.1 and T0.5 values of MADS-LDH were at least 41 °C higher than those of MADS, indicating the superior thermal stability of MADS-LDH over MADS alone.

3.6. UV Absorption of CO3-LDH, MADS-LDH and MADS

The UV-vis absorption spectra of CO3-LDH, MADS-LDH, and MADS are presented in Figure 9. CO3-LDH displayed moderate UV absorption in the wavelength range of 200~228 nm and 300~400 nm, with low UV absorption in the range of 228~300 nm. In comparison with CO3-LDH, MADS exhibited stronger absorption in the range of 228~300 nm. After intercalation, MADS-LDH showed a wider absorption band in the range of 200~400 nm and stronger UV absorption in the range of 228~320 nm. The absorbance of MADS-LDH reached a maximum of 0.835, which was about 28% higher than that of MADS. These results could be attributed to supramolecular interactions, such as hydrogen bonding, electrostatic and gravitational forces, and van der Waals forces, between the cations on the LDHs plate layer and the interlayer MADS anion [38]. The π-π conjugation interactions between the aromatic groups of the MADS-LDH interlayer MADS anion resulted in a lower excitation energy, leading to a red-shifted and broadened absorption band [39,40]. Therefore, the Zn2Al-LDH not only inherits the strong UV absorbing capabilities of MADS but also broadens the absorption band through π-π conjugate interactions between the interlayer guest anions, resulting in a new material with superior UV shielding performance.

3.7. Effect of LDHs on the Structure of PVC Film Composite Materials

The structural and dispersion effects of CO3-LDH and MADS-LDH on PVC film composite materials were illustrated using X-ray diffraction (XRD) patterns. The XRD patterns of PVC, CO3-LDH/PVC, and MADS-LDH/PVC film composite materials are presented in Figure 10. For the MADS-LDH/PVC film composite materials, the absence of characteristic diffraction peaks of MADS-LDH in the range of 2~10° on the test scale, as well as the increased half-peak width of the XRD diffraction peaks, indicated that the PVC macromolecular chains were intercalated within the pore channels of MADS-LDH [41]. This is related to the formation of a hydrophobic environment within the interlayer space of hydrotalcite by the alkyl chains on the interlayer MADS anion after intercalation [41,42], and the excellent compatibility of PVC with the alkyl chains on the MADS anion. During film preparation, hydrophobic solvated PVC molecules, which are soluble in 1,2-dichloroethane, preferred to enter the MADS-LDH interlayer [43], leading to the disappearance of the (003) characteristic diffraction peak of MADS-LDH. Conversely, the interlayer of CO3-LDH constitutes a hydrophilic environment that is not conducive to the entry of hydrophobic PVC molecules into the interlayer. Consequently, all the characteristic X-ray diffraction peaks of CO3-LDH were present in the XRD pattern of CO3-LDH/PVC.

3.8. Film Color Change of LDHs/PVC Film Composite Materials during Photoaging

Table 4 presents a comparison of the photostability of PVC film, CO3-LDH/PVC, and MADS-LDH/PVC film composite materials. During UVB band irradiation aging, no significant color changes were observed in any of the films.
However, under UVC band irradiation aging, PVC and CO3-LDH/PVC film composite materials experienced significant color changes, turning orange-red after 96 h of aging and finally brown after 144 h. In contrast, the MADS-LDH/PVC film composite materials had a longer aging time before showing significant color changes and did not undergo an abrupt jump in color as observed in the other films.
During the aging process of UV band irradiation, the PVC film changed color rapidly. It turned brown after 48 h of aging, turns dark brown after aging to 96 h, and finally photoages to black after 144 h. Similarly, CO3-LDH/PVC films underwent a similar change process during the aging process, but with relatively light aging. The color change during aging of CO3-LDH/PVC film was more evident than that of MADS-LDH/PVC film. The color change process of MADS-LDH/PVC films when aged under UV band irradiation for 144 h was similar to the change process when aged under UVC band irradiation conditions. These findings suggest that the MADS-LDH/PVC films exhibit better photostability under UVC and UV band conditions.
The improved photostability of the MADS-LDH/PVC film composite materials can be further verified by the increase in carbonyl index values and relative degradation rates, as shown in Figure 11 and Table 5. Specifically, PVC film composite materials containing MADS-LDH exhibited good photostability.
Moreover, by comparing the color changes during the aging process of the films, it could be inferred that the degree of influence of UVB, UVC, and UV band on the aging of MADS-LDH/PVC film composite materials was in the order of UV > UVC > UVB, which was consistent with the results obtained in Figure 11d and Figure 12d below.

3.9. Light Stability of LDHs/PVC Film Composite Materials

To evaluate the UV-light aging resistance of MADS-LDH/PVC film composite materials, the surface chemical properties of pristine PVC, CO3-LDH/PVC, and MADS-LDH/PVC film composite materials were analyzed and compared during UV photoaging. Previous studies have shown that carbonyl compounds are the primary molecules produced on the surface of PVC films during photo-oxidative aging [44]. PVC undergoes de-hydrogen chloride and photo-oxidative aging under UV-light, which results in the production of C=O compounds [45].
UV accelerated photoaging tests were conducted to assess the photostability of the pristine PVC and LDHs/PVC film composite materials. The degree of aging during photoaging was determined by analyzing the variation of ΔCl values with irradiation time [42,46]. When exposed to UV light radiation, PVC undergoes photoaging through oxidation reactions, which leads to the production of numerous carbonyl groups [44]. The ΔCl value represents the increase in the number of carbonyl groups induced by UV light [28]. Figure 11a–c display the curves of irradiation time vs. carbonyl index increase for pristine PVC, CO3-LDH/PVC, and MADS-LDH/PVC film composite materials under UVB, UVC, and UV band ultraviolet light conditions, respectively. The increase in carbonyl index (ΔCl) during aging was calculated using the peak area method, with the absorption peak area A 1430 at 1430 cm−1 as the internal standard. This method provided a more intuitive way of evaluating the degree of photoaging. The calculation equations were as follows:
Δ Cl = ( A 1730 , t A 1730 , 0 ) A 1430
The absorption peak areas at 1730 cm−1 at the aging times t and 0 were denoted as A 1730 , t and A 1730 , 0 , respectively, to calculate the carbonyl index increase. Figure 11a–c shows the ΔCl values of both the pristine PVC and LDHs/PVC film composite materials increased with increasing photoaging time, indicating the occurrence of photooxidation reactions with different oxidation rates. The ΔCl value of the LDHs/PVC film composite materials was smaller during aging than that of the pristine PVC, suggesting that the LDH samples could reduce the formation of peroxides during PVC aging and improve its UV stability. Moreover, the ΔCl curves of the MADS-LDH films had the smallest values during the aging process, indicating a slower oxidation rate and better UV shielding properties compared to CO3-LDH.
The relative degradation rates (RDR) were calculated using the following equations, and the results are presented in Table 5. After 144 h of UVB, UVC, and UV band irradiation aging, the MADS-LDH/PVC films showed RDR values of 48.975%, 42.785%, and 11.014%, respectively, compared to the original PVC. Furthermore, analysis of Table 5 and Figure 11d indicated that the magnitude of the aging effect on all three films follows the order of UV > UVC > UVB.
Relative   degradation   rate = ( Δ Cl LDH / PVC Δ Cl PVC ) × 100 %
where Δ Cl PVC and Δ Cl LDH / PVC are the increase values of carbonyl index of the original PVC and LDHs/PVC film composite materials after irradiation aging process for 144 h, respectively.
Figure 12a–c present the FT-IR spectra of MADS-LDH/PVC film composite materials subjected to different photoaging times under UVB, UVC, and UV wavelength aging conditions. It is known that photoaging of pristine PVC under UVB irradiation produced a considerable number of carbonyl groups through oxidation reactions. The C-H anti-stretching vibrational absorption peak at 1430 cm−1 in the R-CH2-R’ structure of the PVC molecular chain was accompanied by the peaks in the range of 1620 cm−1 and 1730 cm−1, which were associated with the C=C stretching vibrational absorption peaks in the R-C=C-R’ polyene structure and C=O in the R-C=O-R’ carbonyl structure [47]. Both C=C and C=O stretching vibration absorption peaks in the sample were produced by the UV photoaging process. Therefore, the carbonyl group’s absorption intensity at 1730 cm−1 was selected to represent the degree of photoaging. As shown in Figure 12a–c, the intensity of the carbonyl absorption peak at 1730 cm−1 increased to different degrees with increasing photoaging time, indicating different levels of photoaging degradation of the MADS-LDH/PVC film composite materials. Comparing with Figure 12d, it can be seen that the effect of UV, UVC, and UVB band irradiation conditions on the aging of MADS-LDH/PVC film composite materials followed the order of UV > UVC > UVB, which was consistent with the results obtained in Figure 11d above.

4. Conclusions

This study focused on the preparation and characterization of MADS-LDH and its potential to improve the photoaging resistance of PVC films. The mechanism of MADS-LDH inhibition of photoaging degradation of MADS-LDH/PVC film composite materials was also investigated. The results indicate that MADS-LDH effectively inhibits the production rate of carbonyl groups during photoaging, thereby improving the photoaging resistance of MADS-LDH/PVC film composite materials. Based on the experimental results obtained in this study, the following conclusions can be drawn:
(1)
The successful preparation of MADS-LDH was confirmed through XRD and FT-IR analysis, which presented a d003 peak position at 3.08°, a proportional relationship between planar spacing in (d003) = 2 (d006) = 3 (d009), and a single set of Bragg reflections (00l) (l = 3, 6, 9). Most of the characteristic absorption bands of MADS were detected in the FT-IR spectra of MADS-LDH.
(2)
The agglomeration of MADS-LDH nanosheet particles through stacking was observed using SEM and TEM. BET and TGA analysis indicated that the pore size of MADS-LDH is mesoporous with a homogeneous pore size distribution, and that the MADS anion intercalation created a hydrophobic environment within the layer, reducing the formation of interlayer crystalline water in MADS-LDH.
(3)
MADS-LDH/PVC film composite materials exhibited the least color change and the lowest degree of aging under both UVC and UV band irradiation compared to pristine PVC and CO3-LDH/PVC films. The color of PVC films changed significantly under both UVC and UV band irradiation, with similar changes in CO3-LDH/PVC films, but the degree of aging was relatively low.
(4)
The addition of MADS-LDH inhibited the generation of carbonyl groups in the MADS-LDH/PVC film composite materials during photoaging, resulting in lower values of both carbonyl index (ΔCl) and relative degradation rate (RDR) compared to pristine PVC and CO3-LDH/PVC films. The degree of influence of UVB, UVC, and UV bands on the photoaging of PVC film and LDHs/PVC film composite materials was found to be UV > UVC > UVB.

Author Contributions

Conceptualization, E.Z. and Y.L.; Methodology, E.Z. and X.C.; Validation, J.H. and X.L.; Data curation, H.Y. and Y.Z.; Writing—original draft preparation, E.Z.; writing—review and editing, E.Z.; Visualization, E.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This paper describes research activities mainly requested and supported by the Natural Science Foundation of China under grant numbers 22278086, supported by the Science and Technology Program of Guangzhou under grant numbers 202102021043, supported by the Demonstration Base for Joint Training of Postgraduates between Guangdong University of Technology and GCH Technology Co., Ltd., supported by the GCH Technology Co., Ltd., under grant numbers 607210383.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kann, Y.; Billingham, N.C. Chemiluminescence is shedding light on degradation and stabilisation of plasticised poly (vinyl chloride). Polym. Degrad. Stab. 2004, 85, 957–966. [Google Scholar] [CrossRef]
  2. Belhaneche-Bensemra, N.; Ouazene, N. Study of the influence of atmospheric pollutants on the natural ageing of rigid polyvinyl chloride. Macromol. Symp. 2002, 180, 181–190. [Google Scholar] [CrossRef]
  3. Decker, C. Photostabilization of poly (vinyl chloride) by protective coatings. J. Vinyl Addit. Technol. 2001, 7, 235–243. [Google Scholar] [CrossRef]
  4. Belhaneche-Bensemra, N. Influence of atmospheric pollutants on the natural and artificial aging of rigid poly (vinyl chloride). J. Vinyl Addit. Technol. 2002, 8, 45–54. [Google Scholar] [CrossRef]
  5. Benavides, R.; Castillo, B.M.; Castañeda, A.O.; López, G.M.; Arias, G. Different thermo-oxidative degradation routes in poly (vinyl chloride). Polym. Degrad. Stab. 2001, 73, 417–423. [Google Scholar] [CrossRef]
  6. Mohapi, M.; Sefadi, J.S.; Mochane, M.J.; Magagula, S.I.; Lebelo, K. Effect of LDHs and Other Clays on Polymer Composite in Adsorptive Removal of Contaminants: A Review. Crystals 2020, 10, 957. [Google Scholar] [CrossRef]
  7. Nazir, M.A.; Khan, N.A.; Cheng, C.; Shah, S.S.A.; Najam, T.; Arshad, M.; Sharif, A.; Akhtar, S.; Rehman, A.u. Surface induced growth of ZIF-67 at Co-layered double hydroxide: Removal of methylene blue and methyl orange from water. Appl. Clay Sci. 2020, 190, 105564. [Google Scholar] [CrossRef]
  8. Nazir, M.A.; Najam, T.; Jabeen, S.; Wattoo, M.A.; Bashir, M.S.; Shah, S.S.A.; ur Rehman, A. Facile synthesis of Tri-metallic layered double hydroxides (NiZnAl-LDHs): Adsorption of Rhodamine-B and methyl orange from water. Inorg. Chem. Commun. 2022, 145, 110008. [Google Scholar] [CrossRef]
  9. Haleem, A.; Shafiq, A.; Chen, S.Q.; Nazar, M. A Comprehensive Review on Adsorption, Photocatalytic and Chemical Degradation of Dyes and Nitro-Compounds over Different Kinds of Porous and Composite Materials. Molecules 2023, 28, 1081. [Google Scholar] [CrossRef]
  10. Jamshaid, M.; Nazir, M.A.; Najam, T.; Shah, S.S.A.; Khan, H.M.; Rehman, A.u. Facile synthesis of Yb3+-Zn2+ substituted M type hexaferrites: Structural, electric and photocatalytic properties under visible light for methylene blue removal. Chem. Phys. Lett. 2022, 805, 139939. [Google Scholar] [CrossRef]
  11. Balbin Tamayo, A.I.; Esteva Guas, A.M.; Pupim Ferreira, A.A.; Aucélio, R.Q.; Huertas Flores, J.O. Influence of molar fraction of Mg/Al, on the electrochemical behavior of hycrotalcite-epoxy-graphite composites. Mater. Chem. Phys. 2020, 253, 123392. [Google Scholar] [CrossRef]
  12. Scavetta, E.; Berrettoni, M.; Giorgetti, M.; Tonelli, D. Electrochemical characterisation of Ni/Al X hydrotalcites and their electrocatalytic behaviour. Electrochim. Acta 2002, 47, 2451–2461. [Google Scholar] [CrossRef]
  13. Shen, L.; Chen, Z.; Kou, J. High-quality modification of general polypropylene by the synergistic effect of zinc adipate, hydrotalcite, and polypropylene (SP179). J. Appl. Polym. Sci. 2023, 140, e53704. [Google Scholar] [CrossRef]
  14. Yan, J.; Yang, Z. Intercalated hydrotalcite-like materials and their application as thermal stabilizers in poly (vinyl chloride). J. Appl. Polym. Sci. 2017, 134, 44896. [Google Scholar] [CrossRef]
  15. Huang, Y.; Law, J.C.-F.; Lam, T.-K.; Leung, K.S.-Y. Risks of organic UV filters: A review of environmental and human health concern studies. Sci. Total Environ. 2021, 755, 142486. [Google Scholar] [CrossRef] [PubMed]
  16. Gao, H.; Yao, A.; Shi, Y.; Noor, N.; Zeb, A.; Li, M.; Li, H. Preparation and properties of hierarchical Al–Mg layered double hydroxides as UV resistant hydrotalcite. Mater. Chem. Phys. 2020, 256, 123630. [Google Scholar] [CrossRef]
  17. Zeng, R.; Tang, W.; Zhou, Q.; Liu, X.; Liu, Y.; Wang, S.; Chen, Z.; Yi, N.; Wang, Z.; Chen, J. Efficient adsorption of Pb (II) by sodium dodecyl benzene sulfonate intercalated calcium aluminum hydrotalcites: Kinetic, isotherm, and mechanisms. Environ. Sci. Pollut. Res. Int. 2022, 29, 46161–46173. [Google Scholar] [CrossRef]
  18. Bouraada, M.; Lafjah, M.; Ouali, M.; Demenorval, L. Basic dye removal from aqueous solutions by dodecylsulfate- and dodecyl benzene sulfonate-intercalated hydrotalcite. J. Hazard. Mater. 2008, 153, 911–918. [Google Scholar] [CrossRef]
  19. Li, Y.; Bi, H.Y.; Shen, S.L. Removal of bisphenol A from aqueous solution by a dodecylsulfate ion-intercalated hydrotalcite-like compound. Environ. Technol. 2012, 33, 1367–1373. [Google Scholar] [CrossRef]
  20. Bouraada, M.; Ouali, M.S.; de Ménorval, L.C. Dodecylsulfate and dodecybenzenesulfonate intercalated hydrotalcites as adsorbent materials for the removal of BBR acid dye from aqueous solutions. J. Saudi Chem. Soc. 2016, 20, 397–404. [Google Scholar] [CrossRef]
  21. Milagres, J.L.; Bellato, C.R.; Ferreira, S.O.; de Moura Guimaraes, L. Preparation and evaluation of hydrocalumite-iron oxide magnetic intercalated with dodecyl sulfate for removal of agrichemicals. J. Environ. Manag. 2020, 255, 109845. [Google Scholar] [CrossRef] [PubMed]
  22. Zhang, X.; Pi, H.; Guo, S. The mechanism for inorganic fillers accelerating and inhibiting the UV irradiation aging behaviors of rigid poly (vinyl chloride). J. Appl. Polym. Sci. 2011, 122, 2869–2875. [Google Scholar] [CrossRef]
  23. Ma, X.; Gao, H.; Lu, Y.; Liu, X.; Dang, L.; Xu, S. Application of β-diketone boron complex as an ultraviolet absorber in polyvinyl chloride film. Mater. Res. Express 2020, 7, 076403. [Google Scholar] [CrossRef]
  24. Yousufzai, A.; Zafar, M.J.; Hasan, S.U. Radical degradation of polyvinyl chloride. Eur. Polym. J. 1972, 8, 1231–1236. [Google Scholar] [CrossRef]
  25. Yu, J.J.; Jiang, Z.; Zhu, L.; Hao, Z.P.; Xu, Z.P. Adsorption/Desorption Studies of NOx on Well-Mixed Oxides Derived from Co−Mg/Al Hydrotalcite-like Compounds. J. Phys. Chem. B 2006, 110, 4291–4300. [Google Scholar] [CrossRef]
  26. Li, L.; Ma, R.; Ebina, Y.; Fukuda, K.; Takada, K.; Sasaki, T. Layer-by-Layer Assembly and Spontaneous Flocculation of Oppositely Charged Oxide and Hydroxide Nanosheets into Inorganic Sandwich Layered Materials. J. Am. Chem. Soc. 2007, 129, 8000–8007. [Google Scholar] [CrossRef]
  27. Lin, Y.; Wang, J.; Evans, D.G.; Li, D. Layered and intercalated hydrotalcite-like materials as thermal stabilizers in PVC resin. J. Phys. Chem. Solids 2006, 67, 998–1001. [Google Scholar] [CrossRef]
  28. Zhang, Q.; Song, K.; Zhao, J.; Kong, X.; Sun, Y.; Liu, X.; Zhang, Y.; Zeng, Q.; Zhang, H. Hexanedioic acid mediated surface–ligand-exchange process for transferring NaYF4: Yb/Er (or Yb/Tm) up-converting nanoparticles from hydrophobic to hydrophilic. J. Colloid Interface Sci. 2009, 336, 171–175. [Google Scholar] [CrossRef]
  29. Sampath, S.K.; Cordaro, J.F. Optical properties of zinc aluminate, zinc gallate, and zinc aluminogallate spinels. J. Am. Ceram. Soc. 1998, 81, 649–654. [Google Scholar] [CrossRef]
  30. Mannepalli, L.K.; Dupati, V.; Vallabha, S.J.; Sunkara, V.M. Synthesis of substituted guanidines using Zn–Al hydrotalcite catalyst. J. Chem. Sci. 2013, 125, 1339–1345. [Google Scholar] [CrossRef]
  31. Bai, L.; Liu, X.; Jiao, T.; Wang, Y.; Huo, Y.; Niu, J. Surface and Interfacial Properties of Mono and Didodecyl Diphenyl Ether Disulfonates. Tenside Surfactants Deterg. 2018, 55, 302–311. [Google Scholar] [CrossRef]
  32. Liu, X.; Niu, J.; Wang, X. Synthesis, Surface and Interfacial Properties of Dodecyl Diphenyl Oxide Disulfonate with Different Counterions. J. Surfactants Deterg. 2015, 18, 675–679. [Google Scholar] [CrossRef]
  33. Tao, Q.; Yuan, J.; Frost, R.L.; He, H.; Yuan, P.; Zhu, J. Effect of surfactant concentration on the stacking modes of organo-silylated layered double hydroxides. Appl. Clay Sci. 2009, 45, 262–269. [Google Scholar] [CrossRef]
  34. Wang, X.; Wu, P.; Huang, Z.; Zhu, N.; Wu, J.; Li, P.; Dang, Z. Solar photocatalytic degradation of methylene blue by mixed metal oxide catalysts derived from ZnAlTi layered double hydroxides. Appl. Clay Sci. 2014, 95, 95–103. [Google Scholar] [CrossRef]
  35. Li, Z.-X.; Zeng, H.-Y.; Gohi, B.F.C.A.; Ding, P.-X. Preparation of CeO2-decorated organic-pillared hydrotalcites for the UV resistance of polymer. Appl. Surf. Sci. 2020, 507, 145110. [Google Scholar] [CrossRef]
  36. Sharma, D.; Sakthivel, A.; Michelraj, S.; Muthurasu, A.; Ganesh, V. Surfactant Intercalated Mono-metallic Cobalt Hydrotalcite: Preparation, Characterization, and its Bi-functional Electrocatalytic Application. ChemistrySelect 2020, 5, 9615–9622. [Google Scholar] [CrossRef]
  37. Mahjoubi, F.Z.; Khalidi, A.; Elhalil, A.; Barka, N. Characteristics and mechanisms of methyl orange sorption onto Zn/Al layered double hydroxide intercalated by dodecyl sulfate anion. Sci. Afr. 2019, 6, e00216. [Google Scholar] [CrossRef]
  38. Wang, C.-X.; Pu, M.; Zhang, P.-H.; Gao, Y.; Yang, Z.-Y.; Lei, M. Structure Simulation and Host–Guest Interaction of Histidine-Intercalated Hydrotalcite–Montmorillonite Complex. Minerals 2018, 8, 198. [Google Scholar] [CrossRef]
  39. Ma, R.; Tang, P.; Feng, Y.; Li, D. UV absorber co-intercalated layered double hydroxides as efficient hybrid UV-shielding materials for polypropylene. Dalton Trans. 2019, 48, 2750–2759. [Google Scholar] [CrossRef]
  40. Feng, Y.; Li, D.; Wang, Y.; Evans, D.G.; Duan, X. Synthesis and characterization of a UV absorbent-intercalated Zn–Al layered double hydroxide. Polym. Degrad. Stab. 2006, 91, 789–794. [Google Scholar] [CrossRef]
  41. Kohno, Y.; Asai, S.; Shibata, M.; Fukuhara, C.; Maeda, Y.; Tomita, Y.; Kobayashi, K. Improved photostability of hydrophobic natural dye incorporated in organo-modified hydrotalcite. J. Phys. Chem. Solids 2014, 75, 945–950. [Google Scholar] [CrossRef]
  42. Miotto Menino, N.; da Silveira Salla, J.; do Nascimento, M.S.; Dallago, R.M.; Peralta, R.A.; Moreira, R.F. High-performance hydrophobic magnetic hydrotalcite for selective treatment of oily wastewater. Environ. Technol. 2021, 44, 1426–1437. [Google Scholar] [CrossRef] [PubMed]
  43. Zhang, Y.; Yuan, Z.-P.; Qin, Y.; Dai, J.; Zhang, T. Comparative Studies on Hydrophilic and Hydrophobic Segments Grafted Poly (vinyl chloride). Chin. J. Polym. Sci. 2018, 36, 604–611. [Google Scholar] [CrossRef]
  44. Owen, E.D. Photodegradation of Polyvinyl chloride. In Ultraviolet Light Induced Reactions in Polymers; American Chemical Society: Washington, DC, USA, 1976; Volume 25, pp. 208–219. [Google Scholar]
  45. Owen, J. Degradation and Stabilisation of PVC; Springer Science & Business Media: Cham, Switzerland, 2012. [Google Scholar]
  46. Zhang, X.; Pi, H.; Guo, S.; Fu, J. Influence of ultraviolet absorbers on the ultraviolet-irradiation behaviour of rigid poly vinyl chloride. Plast. Rubber Compos. 2016, 45, 352–361. [Google Scholar] [CrossRef]
  47. Liu, H.; Dong, L.; Xie, H.; Wan, L.; Liu, Z.; Xiong, C. Ultraviolet light aging properties of PVC/CaCO3 composites. J. Appl. Polym. Sci. 2013, 127, 2749–2756. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of CO3-LDH and MADS-LDH.
Figure 1. XRD patterns of CO3-LDH and MADS-LDH.
Coatings 13 00985 g001
Figure 2. The probable orientation of MADS intercalated in Zn2Al-LDH interlayer.
Figure 2. The probable orientation of MADS intercalated in Zn2Al-LDH interlayer.
Coatings 13 00985 g002
Figure 3. FT-IR spectra of CO3-LDH,MADS-LDH and MADS.
Figure 3. FT-IR spectra of CO3-LDH,MADS-LDH and MADS.
Coatings 13 00985 g003
Figure 4. SEM images of (a,b) CO3-LDH and (c,d) MADS-LDH.
Figure 4. SEM images of (a,b) CO3-LDH and (c,d) MADS-LDH.
Coatings 13 00985 g004
Figure 5. TEM image of MADS-LDH.
Figure 5. TEM image of MADS-LDH.
Coatings 13 00985 g005
Figure 6. N2 adsorption–desorption isotherms of CO3-LDH and MADS-LDH.
Figure 6. N2 adsorption–desorption isotherms of CO3-LDH and MADS-LDH.
Coatings 13 00985 g006
Figure 7. Pore diameter distribution curves of the (a,b) MADS-LDH and (c,d) CO3-LDH.
Figure 7. Pore diameter distribution curves of the (a,b) MADS-LDH and (c,d) CO3-LDH.
Coatings 13 00985 g007
Figure 8. TG curves of CO3-LDH, MADS-LDH, and MADS.
Figure 8. TG curves of CO3-LDH, MADS-LDH, and MADS.
Coatings 13 00985 g008
Figure 9. UV absorbance curves of CO3-LDH, MADS-LDH and MADS.
Figure 9. UV absorbance curves of CO3-LDH, MADS-LDH and MADS.
Coatings 13 00985 g009
Figure 10. XRD patterns of PVC, CO3-LDH/PVC, and MADS-LDH/PVC film composite materials.
Figure 10. XRD patterns of PVC, CO3-LDH/PVC, and MADS-LDH/PVC film composite materials.
Coatings 13 00985 g010
Figure 11. (ac) The increase in carbonyl index (ΔCl) with aging time for PVC, CO3-LDH/PVC, and MADS-LDH/PVC films during UVB (a), UVC (b), and UV (c) aging, respectively; (d) ΔCl of UVB, UVC and UV on PVC, CO3-LDH/PVC, and MADS-LDH/PVC films after aging for 144 h, respectively.
Figure 11. (ac) The increase in carbonyl index (ΔCl) with aging time for PVC, CO3-LDH/PVC, and MADS-LDH/PVC films during UVB (a), UVC (b), and UV (c) aging, respectively; (d) ΔCl of UVB, UVC and UV on PVC, CO3-LDH/PVC, and MADS-LDH/PVC films after aging for 144 h, respectively.
Coatings 13 00985 g011
Figure 12. (ac) FT-IR patterns of MADS-LDH/PVC films at different aging times under UVB, UVC, and UV photoaging conditions, respectively; (d) FT-IR patterns of UVB, UVC, and UV on MADS-LDH/PVC films after 144 h aging, respectively.
Figure 12. (ac) FT-IR patterns of MADS-LDH/PVC films at different aging times under UVB, UVC, and UV photoaging conditions, respectively; (d) FT-IR patterns of UVB, UVC, and UV on MADS-LDH/PVC films after 144 h aging, respectively.
Coatings 13 00985 g012
Table 1. Crystal spacing for CO3-LDH and MADS-LDH.
Table 1. Crystal spacing for CO3-LDH and MADS-LDH.
Parameter (nm)CO3-LDH MADS-LDH
d0030.75 2.87
d0060.38 1.41
d0090.26 0.93
d1100.150.15
Table 2. Constitutive properties of CO3-LDH and MADS-LDH.
Table 2. Constitutive properties of CO3-LDH and MADS-LDH.
SampleSBET (m2/g)Vp (cm3/g)Dp (nm)
CO3-LDH30.4300.11916.610
MADS-LDH23.9270.11716.187
Table 3. Summary of TG analysis results (unit: °C).
Table 3. Summary of TG analysis results (unit: °C).
SampleT0.1ΔT0.1T0.5ΔT0.5
MADS58156
CO3-LDH10446>800>644
MADS-LDH9941685529
Table 4. Film color change process of PVC, CO3-LDH/PVC, and MADS-LDH/PVC films during UVB, UVC, and UV ultraviolet light aging, respectively.
Table 4. Film color change process of PVC, CO3-LDH/PVC, and MADS-LDH/PVC films during UVB, UVC, and UV ultraviolet light aging, respectively.
UVBUVCUV
Time0 h48 h96 h144 h48 h96 h144 h48 h96 h144 h
Sample
PVCCoatings 13 00985 i001Coatings 13 00985 i002Coatings 13 00985 i003Coatings 13 00985 i004Coatings 13 00985 i005Coatings 13 00985 i006Coatings 13 00985 i007Coatings 13 00985 i008Coatings 13 00985 i009Coatings 13 00985 i010
MADS-LDH/PVCCoatings 13 00985 i011Coatings 13 00985 i012Coatings 13 00985 i013Coatings 13 00985 i014Coatings 13 00985 i015Coatings 13 00985 i016Coatings 13 00985 i017Coatings 13 00985 i018Coatings 13 00985 i019Coatings 13 00985 i020
CO3-LDH/PVCCoatings 13 00985 i021Coatings 13 00985 i022Coatings 13 00985 i023Coatings 13 00985 i024Coatings 13 00985 i025Coatings 13 00985 i026Coatings 13 00985 i027Coatings 13 00985 i028Coatings 13 00985 i029Coatings 13 00985 i030
Table 5. Carbonyl index (ΔCl) and relative degradation rate (RDR) of pristine PVC, CO3-LDH/PVC, and MADS-LDH/PVC films after irradiation aging for 144 h.
Table 5. Carbonyl index (ΔCl) and relative degradation rate (RDR) of pristine PVC, CO3-LDH/PVC, and MADS-LDH/PVC films after irradiation aging for 144 h.
UVBUVCUV
SampleΔClRDRΔClRDRΔClRDR
PVC0.439100%1.185100%6.401100%
CO3-LDH/PVC0.39389.522%0.69758.819%2.68741.978%
MADS-LDH/PVC0.21548.975%0.50742.785%0.70511.014%
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

Zhou, E.; Liu, Y.; Yuan, H.; Cheng, X.; Zhong, Y.; He, J.; Lu, X. Mechanism of Sodium Dodecyl Diphenyl Ether Disulfonate Filled Hydrotalcite Inhibiting the Photo-Degradation of Polyvinyl Chloride under Different Ranges of Ultraviolet Wavelength Irradiation. Coatings 2023, 13, 985. https://doi.org/10.3390/coatings13060985

AMA Style

Zhou E, Liu Y, Yuan H, Cheng X, Zhong Y, He J, Lu X. Mechanism of Sodium Dodecyl Diphenyl Ether Disulfonate Filled Hydrotalcite Inhibiting the Photo-Degradation of Polyvinyl Chloride under Different Ranges of Ultraviolet Wavelength Irradiation. Coatings. 2023; 13(6):985. https://doi.org/10.3390/coatings13060985

Chicago/Turabian Style

Zhou, Enguo, Yuan Liu, Huajin Yuan, Xiaoling Cheng, Yuanhong Zhong, Jiebing He, and Xi Lu. 2023. "Mechanism of Sodium Dodecyl Diphenyl Ether Disulfonate Filled Hydrotalcite Inhibiting the Photo-Degradation of Polyvinyl Chloride under Different Ranges of Ultraviolet Wavelength Irradiation" Coatings 13, no. 6: 985. https://doi.org/10.3390/coatings13060985

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