Next Article in Journal
Influence of Different Hot Runner-Systems in the Injection Molding Process on the Structural and Mechanical Properties of Regenerated Cellulose Fiber Reinforced Polypropylene
Next Article in Special Issue
Melanin Nanoparticles Obtained from Preformed Recombinant Melanin by Bottom-Up and Top-Down Approaches
Previous Article in Journal
An Investigative Study on the Structural, Thermal and Mechanical Properties of Clay-Based PVC Polymer Composite Films
Previous Article in Special Issue
The Role of Mucoadhesion and Mucopenetration in the Immune Response Induced by Polymer-Based Mucosal Adjuvants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Wood−Derived Polymers from Olefin−Functionalized Lignin and Ethyl Cellulose via Thiol–Ene Click Chemistry

1
School of Geographic and Biologic Information, Nanjing University of Posts and Telecommunications, Nanjing 210023, China
2
Institute of Chemical Industry of Forest Products, Chinese Academy of Forestry, 16 Suojin North Road, Nanjing 210042, China
*
Authors to whom correspondence should be addressed.
Polymers 2023, 15(8), 1923; https://doi.org/10.3390/polym15081923
Submission received: 3 April 2023 / Revised: 14 April 2023 / Accepted: 14 April 2023 / Published: 18 April 2023

Abstract

:
Lignin and cellulose derivatives have vast potential to be applied in polymer materials. The preparation of cellulose and lignin derivatives through esterification modification is an important method to endow cellulose and lignin with good reactivity, processability and functionality. In this study, ethyl cellulose and lignin are modified via esterification to prepare olefin−functionalized ethyl cellulose and lignin, which are further used to prepare cellulose and lignin cross−linker polymers via thiol–ene click chemistry. The results show that the olefin group concentration in olefin−functionalized ethyl cellulose and lignin reached 2.8096 mmol/g and 3.7000 mmol/g. The tensile stress at break of the cellulose cross−linked polymers reached 23.59 MPa. The gradual enhancement in mechanical properties is positively correlated with the olefin group concentration. The existence of ester groups in the cross−linked polymers and degradation products makes them more thermally stable. In addition, the microstructure and pyrolysis gas composition are also investigated in this paper. This research is of vast significance to the chemical modification and practical application of lignin and cellulose.

1. Introduction

With the increasing depletion of fossil resources and the increasing environmental problems of “white pollution” and “microplastics” caused by the extensive use of non−degradable petroleum−based polymers, the use of bio−based raw materials, instead of petroleum−based compound raw materials, to prepare polymer materials has attracted widespread attention [1]. Cellulose has become an ideal raw material for bio−based polymer materials due to its rich sources, low price, excellent biodegradability, easy modification and many other advantages. Cellulose and its derivatives have been widely used in the fiber, paper, film, plastics, coatings and other industrial fields [2,3,4]. However, due to the existence of a large number of hydrogen bonds within and between the molecules of natural cellulose, the complexity of cellulose aggregation structure and high crystallinity, cellulose is insoluble in water and general organic solvents, in addition to not being able of being melted or processed as traditional plastics, which severely limits the application of cellulose materials. The chemical modification of cellulose includes oxidation, esterification, etherification and other grafting methods [5,6,7]. At present, the esterification derivatives of cellulose include cellulose acetate, cellulose propionate, cellulose butyrate and various cellulose−mixed esters, which are widely used in plastics, coatings, separation membranes, cigarette filters and other daily necessities [8,9,10].
Lignin is a rich natural resource. Compared with other biomass products, lignin has a relatively complex structure and contains a variety of functional groups [11]. It has broad research prospects for the development of appropriate methods for separating and extracting lignin, and then prepare functional composites. At present, lignin has been applied to the preparation of high−value materials, such as porous carbon materials, adsorption materials, capacitor electrode materials, graphene materials, surfactants and hydrogels, and has broad application prospects in the energy, medical, construction, agriculture and other fields [12,13,14,15,16]. The structural unit of lignin is similar to that of phenol. Preparing lignin−based phenolic resin by partially replacing phenol with lignin is the most feasible method [16,17,18,19,20]. Lignin and its derivatives can be used to synthesize bipolar plate materials for phenolic resin fuel cells, phenolic resin catalysts, phenolic resin foams, phenolic resin adhesives and other phenolic resin materials [21,22,23]. Since the 20th century, researchers have been exploring the preparation of lignin−based phenolic resin by replacing phenol with structurally modified lignin derivatives. After the 1990s, the preparation technology of lignin−based phenolic resin developed rapidly because many kinds of structurally modified lignin derivatives with flame retardancy were produced. However, due to the shortcomings of high pollution, high energy consumption, and a complex preparation process, lignin−based phenolic resin cannot be produced on a large scale.
As a simple C–S bonding reaction, “thiol–ene” reaction was discovered over 100 years ago [24]. This reaction has very attractive advantages: First, the C–S bonding reaction can be conducted under a variety of conditions. Secondly, various olefins can be used as suitable substrates, including activated and inactive polysubstituted olefins. Third, almost all mercaptans can be used for reactions, including highly functional substances. Finally, this reaction is very rapid, and the reaction conditions are mild, which can be conducted in the air environment [25,26,27]. The “thiol–ene “click reaction can be initiated by free radicals, initiated by ultraviolet light and free radicals using natural light, red light and redox system. The click reaction is fast and can be performed under normal temperature and pressure. If it is initiated by ultraviolet light, the reaction yield can reach more than 90% in a few seconds. As a means of constructing new materials, the “thiol–ene” click reaction has many advantages: olefin compounds are very rich, with high selectivity, and can synthesize a large number of compounds with various structures [28,29,30,31]. Recently, Jawerth reported that the ethanol−soluble fraction of Lignoboost Kraft lignin was selectively allylated using allyl chloride by means of a mild and industrially scalable procedure. The obtained modified lignin was then subsequently cross−linked to prepare thermosetting resin via thermally induced thiol–ene chemistry [29]. Cao et al. prepared thermosetting lignin−based polyurethane coatings with superior corrosion resistance and a high content of lignin by the polymerization of lignin−based polyol. Firstly, the phenolic hydroxyls of enzymatic hydrolysis lignin were, firstly, selectively converted to primary aliphatic hydroxyls by an allylation reaction. Subsequently, the thermal radical initiated thiol–ene click reaction was applied to efficiently prepare the lignin−based polyol [30]. Zeng et al. developed durable a superhydrophobic and oleophobic coating based on perfluorodecanethiol fluorosilicone polyurethane (PFDT−FSPU) and thiol−modified cellulose substrate. The cross−linked network structure was formed by the radical polymerization of double bonds in PFDT−FSPU when the polyurethane was irradiated with ultraviolet light, and it was anchored on the surface of the cotton fibers by click reaction between the thiol−modified cellulose substrate and PFDT−FSPU. The coated fabric showed excellent durability and can still maintain superhydrophobicity and oleophobicity even after 600 cycles of abrasion or 30 times of washing cycles or 168 h of accelerated aging test [31].
The chemical modification of ethyl cellulose and lignin to prepare olefin−functionalized lignin and ethyl cellulose can effectively increase reaction activity and enrich their application performance. The chemical modification of ethyl cellulose and lignin to construct new functional materials has become a major strategy to increase their added value. In this study, ethyl cellulose (EC) and lignin are modified via esterification to prepare olefin−functionalized ethyl cellulose and lignin, which are further used to prepare cellulose and lignin cross−linker polymers via thiol–ene click chemistry. The chemical structure of olefin−functionalized EC and lignin, as well as ethyl cellulose and lignin cross−linked polymers are characterized. In addition, the thermal stability, microstructure, mechanical property and pyrolysis gas composition are also investigated.

2. Experimental Procedure

2.1. Materials

Lignin (Hydroxyl content: 5.6 mmol/g), EC (Mw = 151,237, Mw unit = 454.5 g/mol, DP = 332, Hydroxyl content: 1.57–1.82 mmole/g), 1−Hydroxycyclohexyl phenyl ketone, undecylenic acid, 4−Dimethylaminopyridine (DMAP), trimethylacetic anhydride, dichloromethane (DCM), methanol, tetrahydrofuran (THF), ketone, azodiisobutyronitrile (AIBN), 3,6−Dioxa−1,8−octanedithiol and pentaerythritol tetra(3−mercaptopropionate) were obtained from commercial resources and used as received, unless otherwise mentioned.

2.2. Preparation of Olefin−Functionalized Ethyl Cellulose (OFE)

EC (3.0 g, 0.01975 mol), undecylenic acid (1.82 g, 0.009878 mol), DMAP (12 mg, 0.09878 mmol) and trimethylacetic anhydride (1.83 g, 0.009878 mol) were dissolved in 50 mL dry THF and stirred at 60 °C under a nitrogen atmosphere for 48 h. In order to remove the unreacted undecylenic acid, the purification process was conducted by repeating dissolution into dry tetrahydrofuran and precipitation from methanol. The solid product was dried under vacuum to obtain OFE and label as EC−1−0.5. When the mole ratios of EC, undecylenic acid, DMAP and trimethylacetic anhydride were 1:1:0.001:1 and 1:1.3:0.001:1, the obtained OFEs were labeled as EC−1−1 and EC−1−1.3, respectively. Scheme 1 shows the Synthesis of OFE.

2.3. Preparation of Olefin−Functionalized Lignin (OFL)

Lignin (5.0 g, 28.75 mmol), undecylenic acid (2.645 g, 14.375 mmol), DMAP (17.53 mg, 0.14375 mmol) and trimethylacetic anhydride (2.677 g, 14.375 mmol) were dissolved in 50 mL dry THF. The mixture was stirred at 60 °C under a nitrogen atmosphere for 48 h. In order to remove the unreacted undecylenic acid, the purification process was conducted by repeating dissolution into dry tetrahydrofuran and precipitation from methanol. The solid product was dried under vacuum to obtain OFLs and labeled lignin−1−0.5. When the mole ratios of lignin, undecylenic acid, DMAP and trimethylacetic anhydride were 1:1:0.001:1 and 1:1.3:0.001:1, the obtained OFLs were labeled as lignin−1−1 and lignin−1−1.3, respectively. Scheme 2 shows the synthesis of OFL.

2.4. Determination of Olefin Group Concentration in OFE and OFL

The olefin group concentration in OFE and OFL was investigated with the internal standard method using 1,2,4,5−tetrachloro−3−nitrobenzene as the interior label. OFE, OFL and 1,2,4,5−tetrachloro−3−nitrobenzene were dissolved in CDCl3, and the 1H NMR of the mixture was tested. The signals at 5.0 ppm and 5.82 ppm from the protons of the olefin group were also integrated. The signal at 7.75 ppm from the protons of 1,2,4,5−tetrachloro−3−nitrobenzene was 100. A detailed calculation is presented in [32].

2.5. Preparation of Wood−Derived Polymers through the Chemical Cross−Linking of OFE and OFL via Thiol–Ene Click Chemistry

OFL or OFL polymers, thiol cross−linkers and 1−Hydroxycyclohexyl phenyl ketone (5% of total mass) were dissolved in dry THF. The mixture was stirred for 30 min and ultrasound for 20 min at room temperature. Then, the mixture was poured into a polytetrafluoroethylene mold, dried under tin foil for 24 h and vacuumed for 24 h at room temperature; then, a UV light was used for 10 min (two 10 W UV lamp tubes). The reactants and obtained wood−derived polymers are shown in Table 1. A schematic diagram of the chemical structure of the wood−derived polymers is presented in Scheme 3.

2.6. Characterization

Fourier transform infrared spectrometry (FTIR) spectra were obtained using a PerkinElmer spectrum 100 FTIR spectrometer. 1H NMR spectra were recorded on a Bruker Avance III HD 300 spectrometer using CDCl3 as the solvent with tetramethylsilane (TMS) as the internal reference. The gel content (Cgel) was determined by Soxhlet extraction. First, approximately 0.5 g of polymer (recorded as m0) was precisely weighed, extracted with acetone for 24 h, dried in a vacuum oven at 50 °C for 24 h and weighed again (recorded as m1). The Cgel was calculated using (m1/m0) × 100%, and the three samples were tested for average. TG−FTIR spectra were analyzed by using a 409PC thermal analyzer coupled with a Nicolet IS10 FTIR instrument. The temperature was increased from 40 to 800 °C at a rate of 10 °C min−1 under a N2 atmosphere. Tensile tests were conducted on an E43.104 Universal Testing Machine (MTS Instrument Crop., Shenzhen, China). Dog−bone shaped polymers with a length of 20 mm and a width of 5.0 mm were tested at room temperature with a crosshead speed of 20 mm min−1. Five replicate samples were used to obtain an average value for each. The microstructure of the cross−linked polymer was determined using German Leica microscope DM750M.

3. Results and Discussion

The esterification products of lignin and EC with undecenoic acid were characterized with FT−IR and 1H NMR, As seen in Figure 1, there are two strong proton signals at 5.0 ppm and 5.82 ppm in the 1H NMR of undecenoic acid, which are attributed to the protons of olefin [33,34]. There are no proton signals at 5.0 ppm and 5.82 ppm in the 1H NMR of EC. After esterification, proton signals at 5.0 ppm and 5.82 ppm gradually increased in the 1H NMR of EC−1−0.5, EC−1−1.0 and EC−1−1.3. The results indicate that OFE products were obtained. As seen in Figure 2, the strong peak at 1725 cm−1 is attributed to the infrared absorption peak of the carbonyl group in the FT−IR of undecenoic acid [35,36,37]. There is no peak at 1725 cm−1 in FT−IR of EC. After esterification, infrared absorption peaks of the carbonyl group gradually increased in the FT−IR of EC−1−0.5, EC−1−1.0 and EC−1−1.3, which indicates that the carbonyl groups were formed after esterification. The peak at 1675 cm−1 in the FT−IR of EC−1−0.5, EC−1−1.0 and EC−1−1.3 is attributed to the olefin groups [35,36,37]. The formation of carbonyl groups and the appearance of olefin groups suggests that OFE products were obtained.
The 1H NMR of OFL showed a similar characteristic absorption peak compared with that of OFE. As seen in Figure 3, there are no proton signals at 5.0 ppm and 5.82 ppm in the 1H NMR of lignin. After esterification, proton signals gradually increased at 5.0 ppm and 5.82 ppm in the 1H NMR of lignin−1−0.5, lignin−1−1.0 and lignin−1−1.3. The results indicate that OFL products were obtained. Similarly, the FT−IR of OFL showed a similar characteristic absorption peak compared that of with OFE. As seen from Figure 4, the strong peak at 1725 cm−1 is attributed to the infrared absorption peak of the carbonyl group in the FT−IR of undecenoic acid [33,34]. There is no peak at 1725 cm−1 in the FT−IR of lignin. After esterification, infrared absorption peaks of the carbonyl group gradually increased in the FT−IR of lignin−1−0.5, lignin−1−1.0 and lignin−1−1.3, which indicates that the carbonyl groups were formed after esterification. The peak at 1675 cm−1 in the FT−IR of lignin−1−0.5, lignin−1−1.0 and lignin−1−1.3 is attributed to olefin groups [35,36,37]. The formation of carbonyl groups and the appearance of olefin groups suggests that OFL products were obtained.
The olefin group concentration in OFL and OFE was determined with the internal standard method using 2,3,5,6−Tetrachloro−3−nitrobenzene as the interior label. The 1H NMR of OFL and OFE with the interior label was detected and the results are shown in Figure 5 and Figure 6. The strong signal at 7.75 ppm in Figure 5 and Figure 6 is attributed to the protons of 2,3,5,6−Tetrachloro−3−nitrobenzene [38,39]. The olefin group concentration in OFL and OFE was calculated according to the procedure presented in a recent study [32]. The results are shown in Table 2, revealing that the olefin group concentration in OFL increased from 1.8060 to 2.8096 mmol/g when the n(−COOH):n(−OH) increased from 0.5 to 1.3. The olefin group concentration in OFE increased from 2.9200 to 3.7000 mmol/g when the n(−COOH):n(−OH) increased from 0.5 to 1.3.
The chemical structure of EC cross−linked polymers and lignin cross−linked polymers were investigated with FT−IR and 1H NMR. As seen from Figure 7, the FT−IR of EC−1−1−2SH and EC−1−1−4SH was compared with EC−1−1. The peak at 1675 cm−1 in the FT−IR of EC−1−1 is attributed to the olefin groups [35,36,37]. After the thiol–ene click reaction, there is no peak at 1675 cm−1 in the FT−IR of EC−1−1−2SH and EC−1−1−24H. The chemical structure of the dissolved part for EC cross−linked polymers was investigated with 1H NMR. As seen from Figure 8, the 1H NMR of EC cross−linked polymer was compared with that of EC−1−1. The two strong proton signals at 5.0 ppm and 5.82 ppm in the 1H NMR of EC−1−1 are attributed to the protons of olefin [33,34]. When the thiol–ene click reaction finished, there were no protons at 5.0 ppm and 5.82 ppm in the 1H NMR of EC−1−1−2SH and EC−1−1−4SH, which indicates that there was a thiol–ene click reaction between OFL and the cross−linker (3,6−Dioxa−1,8−octanedithiol, pentaerythritol tetra(3−mercaptopropionate)), and that the EC cross−linked polymers were obtained.
Figure 9 shows the FT−IR of ligin−1−1−2SH and lignin−1−1−4SH compared with that of lignin−1−1. The peak at 1675 cm−1 in the FT−IR of ligin−1−1 is attributed to olefin groups [35,36,37]. When the thiol–ene click reaction finished, the peak at 1675 cm−1 in the FT−IR of ligin−1−1−2SH and ligin−1−1−4SH decreased obviously, which shows that there is still a small amount of the olefin group in the lignin cross−linked polymer. Figure 10 shows the 1H NMR of the chemical structure of the dissolved part for the lignin cross−linked polymers. The 1H NMR of the lignin cross−linked polymer was compared with that of lignin−1−1. The two strong proton signals at 5.0 ppm and 5.82 ppm in the 1H NMR of ligin−1−1 are attributed to the protons of olefin [34,34]. After the thiol–ene click reaction, the proton signals at 5.0 ppm and 5.82 ppm in the 1H NMR of lignin−1−1−2SH and lignin−1−1−4SH appeared weak and almost disappeared, which indicates that the thiol–ene click reaction between OFL and cross−linker (3,6−Dioxa−1,8−octanedithiol, pentaerythritol tetra(3−mercaptopropionate)) is incomplete, because the precise proportion of olefin and sulfhydryl is difficult to control.
Figure 11 shows the FT−IR of the EC–lignin cross−linked polymer compared with that of lignin−1−1, EC−1−1 and undecenoic acid. When the thiol–ene click reaction occurred, the peak at 1675 cm−1 in the FT−IR of EC−ligin−2SH and EC−ligin−4SH was weaker than those of lignin−1−1 and EC−1−1, which shows that there is still a small amount of the olefin group in the EC–lignin cross−linked polymer. Figure 12 shows the 1H NMR of the chemical structure of the dissolved part for the EC–ligin cross−linked polymers, and the 1H NMR of the lignin cross−linked polymers compared with that of lignin−1−1 and EC−1−1. The proton signals at 5.0 ppm and 5.82 ppm in the 1H NMR of EC−ligin−2SH and EC−ligin−4SH were low, which was caused by the thiol–ene click reaction.
The tensile stress and strain at break of the EC and cross−linked polymers were investigated and the results are shown in Figure 13, Figure 14, Figure 15 and Figure 16. As seen in Figure 13, the tensile stress and strain at break of EC were 146.8 MPa and 1.32%, respectively. For the cross−linked polymers, the tensile stress and strain at break increased when the olefin group concentration increased in OFE. When the olefin group concentration increased from 2.9200 to 3.7000 mmol/g, the tensile stress at break increased from 16.41 MPa to 23.59 MPa, as seen in Table 3, while the tensile strain at break firstly increased from 17.25% to 19.41% and then decreased to 18.12%. When the pentaerythritol tetra(3−mercaptopropionate) was used as cross−linker, as seen in Figure 15, the tensile stress at break increased from 6.86 MPa to 13.94 MPa, and the tensile strain increased from 14.85% to 22.38. EC−lignin−4SH showed excellent tensile properties compared to EC−lignin−2SH. The tensile stress at break for EC−lignin−4SH was 15.19 MPa, which is higher than that of EC−lignin−2SH at 13.72 MPa, and the tensile strain at break increased from 15.12% to 17.25%. The gradual enhancement in mechanical properties is positively correlated with the olefin group concentration. The mechanical properties of the lignin cross−linked polymers (lignin−1−1−2SH and lignin−1−1−4SH) were not tested because the polymers were in powder form. Compared with EC, the tensile strength of the cross−linked polymers decreased significantly, but the tensile strain increased sharply, which is caused that the flexible aliphatic hydrocarbon chain from undecylenic acid as the branched chains of EC and lignin increased the distance among the main chains, weakened the interaction force among the main chains and reduced the hydrogen bond interaction in the matrix of EC and lignin. The flexible aliphatic hydrocarbon chain from undecylenic acid contributed to plasticization by functioning as an internal plasticizer.
The Cgel of the cross−linked polymers were detected and the results are shown in Table 3. The results show that Cgel is positively correlated to the tensile stress for the same type of polymers. For the EC−n−2SH cross−linked polymers, when the tensile stress increased from 16.41 MPa to 23.59 MPa, the Cgel increased from 94.9% to 98.0%. The other cross−linked polymers showed the same relationship between the gel content and the tensile strength.
The microstructure of the cross−linked polymers were investigated using a Leica DM750M optical microscope. As seen in Figure 17, the EC cross−linked polymers showed a uniform surface structure without large cracks or micropores, while the EC−lignin polymers showed an uneven surface structure with many large cracks and micropores, which may be due to solvent evaporation.
Figure 18 shows the thermal stability of the EC and cross−linked polymers. Only one thermal degradation stage occurred at 360–370 °C for EC and all cross−linked polymers. It has been reported that, when cellulose−based polymers are degraded at high temperatures, their polymerization degree is reduced, the chemical composition also changes and the carbonyl groups increase [40,41]. When cellulose−based polymers are fully degraded, carbon monoxide, carbon dioxide, ethylene, water and carbon are produced [42,43]. Table 4 shows the thermal degradation temperature (Td), the peak value of the thermal degradation temperature (TP) and the char residue of the cross−linked polymers. The thermal stability of the cross−linked polymers (EC−1−1.0−2SH, EC−1−1.0−4SH and EC−1−1.3−4SH) was higher than that of EC. The Td, TP and char residue of the cross−linked polymers increased compared with those of EC. This is because OFE and the cross−linkers, such as 3,6−Dioxa−1,8−octanedithiol and pentaerythritol tetra(3−mercaptopropionate), contain thermostable ester groups, which makes the cross−linked polymers difficult degrade.
In order to further investigate the composition of the thermal degradation products, TGA−FTIR was carried out. Figure 19 and Figure 20 show the 3D and 2D FT−IR, respectively, of EC (a), EC−1−1.0−2SH (b), EC−1−1.0−4SH (c) and EC−1−1.3−4SH (d). The infrared characteristic absorption peak of the gas phase of the thermal degradation products can be clearly observed in Figure 20. The data were collected at the fastest decomposition temperature of 340 °C. The infrared characteristic absorption peaks at 3684, 2979, 2306 and 1747, 1391, 1057 cm−1 were attributed to H2O, aliphatic hydrocarbon segments, CO2 and degradation products containing ester groups, respectively [44,45,46]. The existence of ester groups in the cross−linked polymers and degradation products make them more thermally stable.

4. Conclusions

In this study, ethyl cellulose and lignin were modified via esterification to prepare olefin−functionalized ethyl cellulose and lignin, which were further used to prepare cellulose and lignin cross−link polymers via thiol–ene click chemistry. The results show that the olefin group concentration in the olefin−functionalized ethyl cellulose and lignin reached 2.8096 mmol/g and 3.7000 mmol/g, respectively. The tensile stress at break of the cellulose cross−linked polymers reached 23.59 MPa. The gradual enhancement in the mechanical properties was positively correlated with the olefin group concentration. Compared with EC, the tensile strength of the cross−linked polymers decreased significantly, but the tensile strain increased sharply. The flexible aliphatic hydrocarbon chain from undecylenic acid contributed to plasticization by functioning as an internal plasticizer. The changes in mechanical properties make the cross−linked polymers easier to process, expanding their application range. The EC–lignin polymers had an uneven surface structure and there were many large cracks and micropores due to the rigid benzene ring structure of lignin. The infrared characteristic absorption peaks at 3684, 2979, 2306 and 1747, 1391, 1057 cm−1 indicated that H2O, aliphatic hydrocarbon segments, CO2 and degradation products containing ester groups were released. The existence of ester groups in the cross−linked polymers and degradation products makes them more thermally stable. The obtained ethyl cellulose and lignin cross−link polymers are of great significance for practical applications and contribute towards the high−value−added utilization of lignin and cellulose.

Author Contributions

Conceptualization, P.J.; Writing—original draft, investigation, R.A.; Writing—review & editing, P.J. and J.W.; Supervision, C.L.; Funding acquisition, P.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Forestry Science and Technology Innovation and Extension Project of Jiangsu Province (LYKJ[2021]04).

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. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Qin, M.; Chen, C.; Song, B.; Shen, M.; Cao, W.; Yang, H.; Zeng, G.; Gong, J. A review of biodegradable plastics to biodegradable microplastics: Another ecological threat to soil environments? J. Clean. Prod. 2021, 312, 127816. [Google Scholar] [CrossRef]
  2. Shaghaleh, H.; Xu, X.; Wang, S. Current progress in production of biopolymeric materials based on cellulose, cellulose nanofibers, and cellulose derivatives. RSC Adv. 2018, 8, 825–842. [Google Scholar] [CrossRef] [PubMed]
  3. Platnieks, O.; Gaidukovs, S.; Barkane, A.; Sereda, A.; Gaidukova, G.; Grase, L.; Thakur, V.K.; Filipova, I.; Fridrihsone, V.; Skute, M.; et al. Bio−Based Poly(butylene succinate)/Microcrystalline Cellulose/Nanofibrillated Cellulose−Based Sustainable Polymer Composites: Thermo−Mechanical and Biodegradation Studies. Polymers 2020, 12, 1472. [Google Scholar] [CrossRef] [PubMed]
  4. Calvino, C.; Macke, N.; Kato, R.; Rowan, S.J. Development, processing and applications of bio−sourced cellulose nanocrystal composites. Prog. Polym. Sci. 2020, 103, 101221. Available online: https://www.sciencedirect.com/science/article/pii/S0079670020300149 (accessed on 10 April 2023). [CrossRef]
  5. Li, C.; Wu, J.; Shi, H.; Xia, Z.; Sahoo, J.K.; Yeo, J.; Kaplan, D.L. Fiber-Based Biopolymer Processing as a Route toward Sustainability. Adv. Mater. 2022, 34, 2105196. [Google Scholar] [CrossRef]
  6. Azimi, B.; Maleki, H.; Gigante, V.; Bagherzadeh, R.; Mezzetta, A.; Milazzo, M.; Guazzelli, L.; Cinelli, P.; Lazzeri, A.; Danti, S. Cellulose−based fiber spinning processes using ionic liquids. Cellulose 2022, 29, 3079–3129. [Google Scholar] [CrossRef]
  7. Kumari, S.V.G.; Pakshirajan, K.; Pugazhenthi, G. Recent advances and future prospects of cellulose, starch, chitosan, polylactic acid and polyhydroxyalkanoates for sustainable food packaging applications. Int. J. Biol. Macromol. 2022, 221, 163–182. Available online: https://www.sciencedirect.com/science/article/pii/S0141813022019134 (accessed on 10 April 2023). [CrossRef]
  8. Fox, S.C.; Li, B.; Xu, D.; Edgar, K.J. Regioselective Esterification and Etherification of Cellulose: A Review. Biomacromolecules 2011, 12, 1956–1972. [Google Scholar] [CrossRef]
  9. Wang, Y.; Wang, X.; Xie, Y.; Zhang, K. Functional nanomaterials through esterification of cellulose: A review of chemistry and application. Cellulose 2018, 25, 3703–3731. [Google Scholar] [CrossRef]
  10. Shokri, Z.; Seidi, F.; Saeb, M.R.; Jin, Y.; Li, C.; Xiao, H. Elucidating the impact of enzymatic modifications on the structure, properties, and applications of cellulose, chitosan, starch and their derivatives: A review. Mater. Today Chem. 2022, 24, 100780. Available online: https://www.sciencedirect.com/science/article/pii/S246851942200009X (accessed on 10 April 2023). [CrossRef]
  11. Svinterikos, E.; Zuburtikudis, I.; Al−Marzouqi, M.H. Electrospun Lignin−Derived Carbon Micro− and Nanofibers: A Review on Precursors, Properties, and Applications. ACS Sustain. Chem. Eng. 2020, 8, 13868–13893. [Google Scholar] [CrossRef]
  12. Lee, D.-W.; Jin, M.-H.; Park, J.-H.; Lee, Y.-J.; Choi, Y.-C. Flexible Synthetic Strategies for Lignin−Derived Hierarchically Porous Carbon Materials. ACS Sustain. Chem. Eng. 2018, 6, 10454–10462. [Google Scholar] [CrossRef]
  13. Zhang, B.; Yang, D.; Qiu, X.; Qian, Y.; Yan, M.; Li, Q. Influences of aggregation behavior of lignin on the microstructure and adsorptive properties of lignin−derived porous carbons by potassium compound activation. J. Ind. Eng. Chem. 2020, 82, 220–227. Available online: https://www.sciencedirect.com/science/article/pii/S1226086X19305532 (accessed on 10 April 2023). [CrossRef]
  14. Tian, Y.; Zhou, H. A novel nitrogen−doped porous carbon derived from black liquor for efficient removal of Cr(VI) and tetracycline: Comparison with lignin porous carbon. J. Clean. Prod. 2022, 333, 130106. Available online: https://www.sciencedirect.com/science/article/pii/S0959652621042724 (accessed on 10 April 2023). [CrossRef]
  15. Xi, Y.; Yang, D.; Lou, H.; Gong, Y.; Yi, C.; Lyu, G.; Han, W.; Kong, F.; Qiu, X. Designing the effective microstructure of lignin−based porous carbon substrate to inhibit the capacity decline for SnO2 anode. Ind. Crop. Prod. 2021, 161, 113179. Available online: https://www.sciencedirect.com/science/article/pii/S0926669020310967 (accessed on 10 April 2023). [CrossRef]
  16. Gong, X.; Meng, Y.; Lu, J.; Tao, Y.; Cheng, Y.; Wang, H. A Review on Lignin-Based Phenolic Resin Adhesive. Macromol. Chem. Phys. 2022, 223, 2100434. [Google Scholar] [CrossRef]
  17. Chen, Y.; Gong, X.; Yang, G.; Li, Q.; Zhou, N. Preparation and characterization of a nanolignin phenol formaldehyde resin by replacing phenol partially with lignin nanoparticles. RSC Adv. 2019, 9, 29255–29262. [Google Scholar] [CrossRef]
  18. Huang, C.; Peng, Z.; Li, J.; Li, X.; Jiang, X.; Dong, Y. Unlocking the role of lignin for preparing the lignin−based wood adhesive: A review. Ind. Crop. Prod. 2022, 187, 115388. Available online: https://www.sciencedirect.com/science/article/pii/S0926669022008718 (accessed on 10 April 2023). [CrossRef]
  19. Weng, S.; Li, Z.; Bo, C.; Song, F.; Xu, Y.; Hu, L.; Zhou, Y.; Jia, P. Design lignin doped with nitrogen and phosphorus for flame retardant phenolic foam materials. React. Funct. Polym. 2023, 185, 105535. Available online: https://www.sciencedirect.com/science/article/pii/S138151482300038X (accessed on 10 April 2023). [CrossRef]
  20. Sarika, P.R.; Nancarrow, P.; Khansaheb, A.; Ibrahim, T. Bio−Based Alternatives to Phenol and Formaldehyde for the Production of Resins. Polymers 2020, 12, 2237. [Google Scholar] [CrossRef] [PubMed]
  21. Asgher, M.; Qamar, S.A.; Bilal, M.; Iqbal, H.M.N. Bio−based active food packaging materials: Sustainable alternative to conventional petrochemical−based packaging materials. Food Res. Int. 2020, 137, 109625. Available online: https://www.sciencedirect.com/science/article/pii/S0963996920306505 (accessed on 10 April 2023). [CrossRef] [PubMed]
  22. Yang, W.; Rallini, M.; Natali, M.; Kenny, J.; Ma, P.; Dong, W.; Torre, L.; Puglia, D. Preparation and properties of adhesives based on phenolic resin containing lignin micro and nanoparticles: A comparative study. Mater. Des. 2019, 161, 55–63. Available online: https://www.sciencedirect.com/science/article/pii/S0264127518308396 (accessed on 10 April 2023). [CrossRef]
  23. Xu, Y.; Guo, L.; Zhang, H.; Zhai, H.; Ren, H. Research status, industrial application demand and prospects of phenolic resin. RSC Adv. 2019, 9, 28924–28935. [Google Scholar] [CrossRef]
  24. Firdaus, M. Thiol−Ene (Click) Reactions as Efficient Tools for Terpene Modification. Asian J. Org. Chem. 2017, 6, 1702–1714. [Google Scholar] [CrossRef]
  25. Firdaus, M.; Montero de Espinosa, L.; Meier, M.A.R. Terpene−Based Renewable Monomers and Polymers via Thiol–Ene Additions. Macromolecules 2011, 44, 7253–7262. [Google Scholar] [CrossRef]
  26. Ahangarpour, M.; Kavianinia, I.; Harris, P.W.R.; Brimble, M.A. Photo−induced radical thiol–ene chemistry: A versatile toolbox for peptide−based drug design. Chem. Soc. Rev. 2021, 50, 898–944. [Google Scholar] [CrossRef] [PubMed]
  27. Kazybayeva, D.S.; Irmukhametova, G.S.; Khutoryanskiy, V.V. Thiol−Ene “Click Reactions” as a Promising Approach to Polymer Materials. Polym. Sci. Ser. B 2022, 64, 1–16. [Google Scholar] [CrossRef]
  28. Hoyle, C.E.; Bowman, C.N. Thiol−Ene Click Chemistry. Angew. Chem. Int. Ed. 2010, 49, 1540–1573. [Google Scholar] [CrossRef]
  29. Jawerth, M.; Johansson, M.; Lundmark, S.; Gioia, C.; Lawoko, M. Renewable Thiol–Ene Thermosets Based on Refined and Selectively Allylated Industrial Lignin. ACS Sustain. Chem. Eng. 2017, 5, 10918–10925. [Google Scholar] [CrossRef]
  30. Cao, Y.; Liu, Z.; Zheng, B.; Ou, R.; Fan, Q.; Li, L.; Guo, C.; Liu, T.; Wang, Q. Synthesis of lignin−based polyols via thiol−ene chemistry for high−performance polyurethane anticorrosive coating. Compos. Part B Eng. 2020, 200, 108295. [Google Scholar] [CrossRef]
  31. Zeng, T.; Zhang, P.; Li, X.; Yin, Y.; Chen, K.; Wang, C. Facile fabrication of durable superhydrophobic and oleophobic surface on cellulose substrate via thiol−ene click modification. Appl. Surf. Sci. 2019, 493, 1004–1012. [Google Scholar] [CrossRef]
  32. Yuan, L.; Zhang, Y.; Wang, Z.; Han, Y.; Tang, C. Plant Oil and Lignin−Derived Elastomers via Thermal Azide–Alkyne Cycloaddition Click Chemistry. ACS Sustain. Chem. Eng. 2018, 7, 2593–2601. [Google Scholar] [CrossRef]
  33. Yuan, L.; Wang, Z.; Trenor, N.M.; Tang, C. Preparation and Applications of Polymers with Pendant Fatty Chains from Plant Oils. In Sustainable Polymers from Biomass; Wiley: Hoboken, NJ, USA, 2017; pp. 181–207. [Google Scholar] [CrossRef]
  34. Yan, K.; Wang, J.; Wang, Z.; Yuan, L. Bio−based monomers for amide−containing sustainable polymers. Chem. Commun. 2023, 59, 382–400. [Google Scholar] [CrossRef] [PubMed]
  35. Yasa, S.R.; Cheguru, S.; Krishnasamy, S.; Korlipara, P.V.; Rajak, A.K.; Penumarthy, V. Synthesis of 10−undecenoic acid based C22−dimer acid esters and their evaluation as potential lubricant basestocks. Ind. Crop. Prod. 2017, 103, 141–151. Available online: https://www.sciencedirect.com/science/article/pii/S092666901730225X (accessed on 10 April 2023). [CrossRef]
  36. Kontham, V.; Ansari, K.R.; Padmaja, K.V. Tribological Properties of 10−Undecenoic Acid−Derived Schiff Base Lubricant Additives. Arab. J. Sci. Eng. 2021, 46, 5593–5603. [Google Scholar] [CrossRef]
  37. Lluch, C.; Lligadas, G.; Ronda, J.C.; Galià, M.; Cadiz, V. “Click” Synthesis of Fatty Acid Derivatives as Fast−Degrading Polyanhydride Precursors. Macromol. Rapid Commun. 2011, 32, 1343–1351. [Google Scholar] [CrossRef]
  38. Çiçek, S.S.; Girreser, U.; Zidorn, C. Quantification of the total amount of black cohosh cycloartanoids by integration of one specific 1 H NMR signal. J. Pharm. Biomed. Anal. 2018, 155, 109–115. Available online: https://www.sciencedirect.com/science/article/pii/S0731708518303376 (accessed on 10 April 2023). [CrossRef]
  39. Crenshaw, M.D.; Tefft, M.E.; Buehler, S.S.; Brinkman, M.C.; Clark, P.I.; Gordon, S.M. Determination of nicotine, glycerol, propylene glycol and water in electronic cigarette fluids using quantitative 1 H NMR. Magn. Reson. Chem. 2016, 54, 901–904. [Google Scholar] [CrossRef]
  40. Puls, J.; Wilson, S.A.; Hölter, D. Degradation of Cellulose Acetate−Based Materials: A Review. J. Polym. Environ. 2011, 19, 152–165. [Google Scholar] [CrossRef]
  41. Lucena, M.d.C.C.; de Alencar, A.E.V.; Mazzeto, S.E.; Soares, S.d.A. The effect of additives on the thermal degradation of cellulose acetate. Polym. Degrad. Stab. 2003, 80, 149–155. Available online: https://www.sciencedirect.com/science/article/pii/S0141391002003968 (accessed on 10 April 2023). [CrossRef]
  42. Mülhaupt, R. Green Polymer Chemistry and Bio−based Plastics: Dreams and Reality. Macromol. Chem. Phys. 2013, 214, 159–174. [Google Scholar] [CrossRef]
  43. Williams, P.T.; Onwudili, J. Subcritical and Supercritical Water Gasification of Cellulose, Starch, Glucose, and Biomass Waste. Energy Fuels 2006, 20, 1259–1265. [Google Scholar] [CrossRef]
  44. Jia, P.; Zhang, M.; Liu, C.; Hu, L.; Feng, G.; Bo, C.; Zhou, Y. Effect of chlorinated phosphate ester based on castor oil on thermal degradation of poly (vinyl chloride) blends and its flame retardant mechanism as secondary plasticizer. RSC Adv. 2015, 5, 41169–41178. [Google Scholar] [CrossRef]
  45. Jia, P.; Hu, L.; Zhang, M.; Zhou, Y.-H. TG−FTIR and TG−MS analysis applied to study the flame retardancy of PVC–castor oil−based chlorinated phosphate ester blends. J. Therm. Anal. Calorim. 2016, 124, 1331–1339. [Google Scholar] [CrossRef]
  46. Jia, P.; Zhang, M.; Hu, L.; Liu, C.; Feng, G.; Yang, X.; Bo, C.; Zhou, Y. Development of a vegetable oil based plasticizer for preparing flame retardant poly(vinyl chloride) materials. RSC Adv. 2015, 5, 76392–76400. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of OFE.
Scheme 1. Synthesis of OFE.
Polymers 15 01923 sch001
Scheme 2. Synthesis of OFL.
Scheme 2. Synthesis of OFL.
Polymers 15 01923 sch002
Scheme 3. Schematic diagram of the chemical structure of the wood−derived polymers through the chemical cross−linking of OFE and OFL via thiol–ene click chemistry.
Scheme 3. Schematic diagram of the chemical structure of the wood−derived polymers through the chemical cross−linking of OFE and OFL via thiol–ene click chemistry.
Polymers 15 01923 sch003
Figure 1. 1H NMR of OFE, EC and undecenoic acid.
Figure 1. 1H NMR of OFE, EC and undecenoic acid.
Polymers 15 01923 g001
Figure 2. FT−IR of OFE, EC and undecenoic acid.
Figure 2. FT−IR of OFE, EC and undecenoic acid.
Polymers 15 01923 g002
Figure 3. 1H NMR of OFL, lignin and undecenoic acid.
Figure 3. 1H NMR of OFL, lignin and undecenoic acid.
Polymers 15 01923 g003
Figure 4. FT−IR of OFL, lignin and undecenoic acid.
Figure 4. FT−IR of OFL, lignin and undecenoic acid.
Polymers 15 01923 g004
Figure 5. 1H NMR of OFL with the interior label compared with lignin.
Figure 5. 1H NMR of OFL with the interior label compared with lignin.
Polymers 15 01923 g005
Figure 6. 1H NMR of OFE with the interior label compared with EC.
Figure 6. 1H NMR of OFE with the interior label compared with EC.
Polymers 15 01923 g006
Figure 7. FT−IR of EC cross−linked polymers compared with that of EC−1−1.
Figure 7. FT−IR of EC cross−linked polymers compared with that of EC−1−1.
Polymers 15 01923 g007
Figure 8. 1H NMR of EC cross−linked polymer compared with that of EC−1−1.
Figure 8. 1H NMR of EC cross−linked polymer compared with that of EC−1−1.
Polymers 15 01923 g008
Figure 9. FT−IR of lignin cross−linked polymer compared with that of lignin.
Figure 9. FT−IR of lignin cross−linked polymer compared with that of lignin.
Polymers 15 01923 g009
Figure 10. 1H NMR of lignin cross−linked polymer compared with that of lignin.
Figure 10. 1H NMR of lignin cross−linked polymer compared with that of lignin.
Polymers 15 01923 g010
Figure 11. FT−IR of the EC–lignin cross−linked polymers compared with that of lignin−1−1, EC−1−1 and undecenoic acid.
Figure 11. FT−IR of the EC–lignin cross−linked polymers compared with that of lignin−1−1, EC−1−1 and undecenoic acid.
Polymers 15 01923 g011
Figure 12. 1H NMR of EC–lignin cross−linked polymer compared with those of lignin−1−1 and EC−1−1.
Figure 12. 1H NMR of EC–lignin cross−linked polymer compared with those of lignin−1−1 and EC−1−1.
Polymers 15 01923 g012
Figure 13. Tensile stress and strain of EC.
Figure 13. Tensile stress and strain of EC.
Polymers 15 01923 g013
Figure 14. Tensile stress and strain of the EC cross−linked polymers using 3,6−Dioxa−1,8−octanedithiol as the cross−linker.
Figure 14. Tensile stress and strain of the EC cross−linked polymers using 3,6−Dioxa−1,8−octanedithiol as the cross−linker.
Polymers 15 01923 g014
Figure 15. Tensile stress and strain of the EC cross−linked polymers using pentaerythritol tetra(3−mercaptopropionate) as the cross−linker.
Figure 15. Tensile stress and strain of the EC cross−linked polymers using pentaerythritol tetra(3−mercaptopropionate) as the cross−linker.
Polymers 15 01923 g015
Figure 16. Tensile stress and strain of the EC–lignin cross−linked polymers.
Figure 16. Tensile stress and strain of the EC–lignin cross−linked polymers.
Polymers 15 01923 g016
Figure 17. Microstructure of the cross−linked polymers.
Figure 17. Microstructure of the cross−linked polymers.
Polymers 15 01923 g017
Figure 18. TGA (a) and DTG (b) curves of the cross−linked polymers.
Figure 18. TGA (a) and DTG (b) curves of the cross−linked polymers.
Polymers 15 01923 g018
Figure 19. Three−dimensional FT−IR of EC (a), EC−1−1.0−2SH (b), EC−1−1.0 − 4SH (c) and EC−1−1.3−4SH (d).
Figure 19. Three−dimensional FT−IR of EC (a), EC−1−1.0−2SH (b), EC−1−1.0 − 4SH (c) and EC−1−1.3−4SH (d).
Polymers 15 01923 g019
Figure 20. Two−dimensional FT−IR of EC, EC−1−1.0−2SH, EC−1−1.0−4SH and EC−1−1.3−4SH.
Figure 20. Two−dimensional FT−IR of EC, EC−1−1.0−2SH, EC−1−1.0−4SH and EC−1−1.3−4SH.
Polymers 15 01923 g020
Table 1. Reactants and the obtained wood−derived polymers.
Table 1. Reactants and the obtained wood−derived polymers.
Cross−linked PolymerReactants
EC−1−0.5−2SHEC−1−0.5 and 3,6−Dioxa−1,8−octanedithiol,
n(−CH=CH−) = n(−SH)
EC−1−1.0−2SHEC−1−1.0 and 3,6−Dioxa−1,8−octanedithiol,
n(−CH=CH−) = n(−SH)
EC−1−1.3−2SHEC−1−1.3 and 3,6−Dioxa−1,8−octanedithiol,
n(−CH=CH−) = n(−SH)
EC−1−0.5−4SHEC−1−0.5 and pentaerythritol tetra(3−mercaptopropionate),
n(−CH=CH−) = n(−SH)
EC−1−1.0−4SHEC−1−1.0 and pentaerythritol tetra(3−mercaptopropionate),
n(−CH=CH−) = n(−SH)
EC−1−1.3−4SHEC−1−1.3 and pentaerythritol tetra(3−mercaptopropionate),
n(−CH=CH−) = n(−SH)
EC−Lignin−2SHEC−1−1, Lignin−1−1, and 3,6−Dioxa−1,8−octanedithiol,
n(−CH=CH−)/n(−CH=CH−)/n(−SH) = 0.5:0.5:1
EC−Lignin−4SHEC−1−1, lignin−1−1, and pentaerythritol tetra(3−mercaptopropionate), n(−CH=CH−)/n(−CH=CH−)/n(−SH) = 0.5:0.5:1
Lignin−1−1−2SHLignin−1−1 and 3,6−Dioxa−1,8−octanedithiol, n(−CH=CH−) = n(−SH)
Lignin−1−1−4SHLignin−1−1 and pentaerythritol tetra(3−mercaptopropionate), n(−CH=CH−) = n(−SH)
Table 2. Olefin group concentration in OFL and OFE.
Table 2. Olefin group concentration in OFL and OFE.
OFL and OFE SamplesOlefin Group Concentration in OFL and OFE (mmol/g)
Lignin−1−0.51.8060
Lignin−1−12.3818
Lignin−1−1.32.8096
EC−1−0.52.9200
EC−1−13.5873
EC−1−1.33.7000
Table 3. Tensile stress, tensile strain and Cgel of the cross−linked polymers.
Table 3. Tensile stress, tensile strain and Cgel of the cross−linked polymers.
Polymer FilmsTensile Stress (MPa)Tensile Strain (%)Cgel (%)
EC146.89 ± 10.241.32 ± 0.02
EC−1−0.5−2SH16.41 ± 1.2217.25 ± 0.1994.9 ± 0.8
EC−1−1.0−2SH16.53 ± 1.1019.41 ± 2.1896.4 ± 0.4
EC−1−1.3−2SH23.59 ± 3.0118.12 ± 0.1398.0 ± 0.4
EC−1−0.5−4SH6.86 ± 0.2114.85 ± 1.1192.7 ± 0.6
EC−1−1.0−4SH10.55 ± 0.8916.92 ± 1.7694.3 ± 0.7
EC−1−1.3−4SH13.94 ± 1.0022.38 ± 3.2095.2 ± 0.9
EC−Lignin−2SH13.72 ± 0.0215.12 ± 1.1994.8 ± 0.2
EC−Lignin−4SH15.19 ± 1.0317.25 ± 2.1097.2 ± 0.5
Lignin−1−1−2SH
Lignin−1−1−4SH
Table 4. Td, TP and char residue of the cross−linked polymers.
Table 4. Td, TP and char residue of the cross−linked polymers.
Cross−linked PolymersTd (°C)TP (°C)Char Residue (%)
EC339.8364.10.9
EC−1−1.0−2SH348.2370.26.6
EC−1−1.0−4SH346.8371.27.2
EC−1−1.3−4SH344.0374.15.7
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

An, R.; Liu, C.; Wang, J.; Jia, P. Wood−Derived Polymers from Olefin−Functionalized Lignin and Ethyl Cellulose via Thiol–Ene Click Chemistry. Polymers 2023, 15, 1923. https://doi.org/10.3390/polym15081923

AMA Style

An R, Liu C, Wang J, Jia P. Wood−Derived Polymers from Olefin−Functionalized Lignin and Ethyl Cellulose via Thiol–Ene Click Chemistry. Polymers. 2023; 15(8):1923. https://doi.org/10.3390/polym15081923

Chicago/Turabian Style

An, Rongrong, Chengguo Liu, Jun Wang, and Puyou Jia. 2023. "Wood−Derived Polymers from Olefin−Functionalized Lignin and Ethyl Cellulose via Thiol–Ene Click Chemistry" Polymers 15, no. 8: 1923. https://doi.org/10.3390/polym15081923

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