Next Article in Journal
Metallic Nanoparticle Integrated Ternary Polymer Blend of PVA/Starch/Glycerol: A Promising Antimicrobial Food Packaging Material
Previous Article in Journal
The Influence of Extraction Conditions on the Yield and Physico-Chemical Parameters of Pectin from Grape Pomace
Previous Article in Special Issue
Nanocomposites of Polyhydroxyurethane with POSS Microdomains: Synthesis via Non-Isocyanate Approach, Morphologies and Reprocessing Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Biodegradable Natural Rubber Based on Novel Double Dynamic Covalent Cross-Linking

1
School of Materials Science and Engineering, Hainan University, Haikou 570228, China
2
Guangdong Provincial Key Laboratory of Natural Rubber Processing, Agricultural Products Processing Research Institute of Chinese Academy of Tropical Agricultural Sciences, Zhanjiang 524001, China
*
Author to whom correspondence should be addressed.
Polymers 2022, 14(7), 1380; https://doi.org/10.3390/polym14071380
Submission received: 2 March 2022 / Revised: 24 March 2022 / Accepted: 25 March 2022 / Published: 29 March 2022
(This article belongs to the Special Issue Self-Healing Polymers and Vitrimers)

Abstract

:
In this paper, biodegradable epoxidized natural rubber containing cyclic carbonate groups (CNR) was prepared by the reaction between epoxidized natural rubber (ENR) and carbon dioxide. Dynamic disulfide bonds and a boronic ester structure were successfully constructed and then the cross-linking network was formed by the thermally initiated “click” reaction between thiol groups of the cross-linker and the residual epoxy groups of ENR. As a result of the exquisite double dynamic covalent structure, the material exhibits high self-healing efficiency. Moreover, by virtue of the cyclic carbonate structure of the CNR, the natural rubber was confirmed to be biodegradable according to the biodegradable measurement. To the best of our knowledge, natural rubber with biodegradable and self-healing characteristics was obtained for the first time.

1. Introduction

Natural rubber (NR), compared with non-renewable fossil-based materials, is also a sustainable polymer material. NR has been widely used in the tire industry, as well as the aerospace and medical fields, owing to its unique features of high tensile and tearing strength, high fatigue resistance, low heat built-up and so on [1]. After curing, the covalent cross-linked structure endows NR with excellent comprehensive properties; however, the crosslinked NR would be “insoluble” and “unmeltable”, and could not be recycled without impairment of performance, leading to crucial black pollution consequences for the environment [2]. It would be of significance to reuse the vulcanized rubber products for the purpose of alleviating the burden on the Earth. To achieve secondary utilization of NR, some technologies, such as a hydropress with a liquid atmosphere, microwave or ultrasonic waves assistance, and regenerating agents have been employed [2,3]. In these ways, waste rubber could be reprocessed with the sacrifice of mechanical properties and energy consumption. Recently, a novel and powerful self-healing strategy utilizing dynamic covalent bonds was proposed [4,5,6,7].
A dynamic covalent bond is a kind of chemical bond which can be physical cross-linked and is dynamically reversible. It can be dissociated and reconstructed under the stimulation of mild conditions or thermal-light conditions, as well as exchange reactions, which has recently been identified as an extremely important kind of chemical bond [8,9]. Reversible covalent bonds usually exhibit higher bond strength and better stability than non-covalent bonds, and therefore they have broad application prospects in dynamic polymers, especially in the structural design of materials [10,11]. A considerable number of exchange reactions, including disulfide exchange [12,13], boronic ester exchange [14,15,16,17], Diels-Alder chemistry [18,19], siloxane equilibration [20], imine exchange [21,22], transesterification [23,24], and silyl ether exchange [25,26] in the context of dynamic polymers have been developed to improve the efficiency of service.
This dazzling field has attracted the notice of many research groups. For example, Santana et al. [27] first introduced furan functional groups into natural rubber and used the Diels-Alder coupling reaction to react furfurylamine with maleic anhydride groups to form furan functional group imide, which was grafted onto the NR skeleton to form maleic anhydride grafted rubber (NR-g-MA). The material exhibited self-healing behavior at 130 °C, especially at low strain. After self-healing, the tensile strength of the samples can reach 80% of the original. Guo et al. [28] also achieved styrene-butadiene rubber (SBR) with a self-healing function. The vitrified rubber was synthesized by the thiol-ene click reaction of dithiol-containing boronic ester with SBR. Due to the transesterification of boronic ester bond, the network topology can be changed, and thus the material was provided with a self-healing ability and ductility. To improve self-healing efficiency, Yang et al. [6] reported epoxidized natural rubber cross-linked by a series of aromatic disulfides and double cross-linked by dynamic disulfide decomposition and thermal reversible hydrogen bonds. The self-healing efficiency of double cross-linked ENR was as high as 98%, and simultaneously fracture or cyclic fatigue damage had no negative effect on the self-healing property of the material.
Although a large number of articles have reported self-healing rubber with high self-healing efficiency, its self-healing ability will gradually weaken with the increase of the number of self-healing cycles, and became waste rubber. As far as we know, natural rubber that is biodegradable and self-healing has so far not been reported on. In this work, we used the epoxy group of ENR to react with carbon dioxide to obtain a weak bond of cyclic carbonate, which was intended to make the rubber (CNR) biodegradable. In addition, we constructed units of disulfide bond and boronic ester for the purpose of a dynamic reversible effect. Because of the dynamic characteristics of boronic esters and disulfide bonds, the cross-linked network of the rubber can be rearranged, which endowed CNR with the ability of self-healing. The experimental process was shown in Scheme 1. The ENR with both biodegradability and self-healing functions was thus produced, which greatly reduced the negative impact of tradition natural rubber materials on the environment.

2. Materials and Methods

2.1. Materials

ENR was obtained from the Agricultural Products Processing Research Institute, Chinese Academy of Tropical Agricultural Sciences, with an epoxy content of 25 mol%. Carbon dioxide (CO2, 99.99%) was purchased from Hainan Jiateng Chemical Gas Co., Ltd. (Haikou, China). Tetrabutylammonium iodide (TBAI, 98%); cystamine dihydrochloride and deuterium chloroform were supplied by J&K Scientific Co., Ltd. (Beijing China). N-hydroxysuccinimide (NHS), 1-thioglycerol, 1-ethyl-3-(3-(dimethylamino)propyl) carbodiimide hydrochloride (EDC), and phosphate buffer solution (PBS) were supplied by Shanghai Macklin Biochemical Co., Ltd (Shanghai China). Deionized water was produced with an ultrapure water machine. Solvents such as toluene were analytical reagent grade without further treatment. All other reagents were purchased from Xilong Technology Co., Ltd (Shantou, China).

2.2. Methods

2.2.1. Preparation of Carbonated Natural Rubber (CNR)

CNR was synthesized according to Scheme 2. A rubber solution was prepared according to Gan et al.’s work [29]. 10 g ENR was masticated by an open mixer (LN-6, Guangdong Lina Industrial Co., Ltd., Dongguang, China). The roll diameter was 16 cm, and the gap between the two rolls was adjusted to 0.1 mm. After grinding the ENR for 20 min, the ENR was dissolved in toluene in a flask for 7 days (30 g ENR dissolved in 100 mL toluene) to allow it to be dissolved sufficiently. The CNR was synthesized by coupling CO2 with 4 g ENR in a 25 mL autoclave (YZPR-500ML, Shanghai Yanzheng Experimental Instrument Co., Ltd, Shanghai, China). The amount of catalyst was 10% with respect to ENR. After 5 min of leakage test and purification with CO2, the ENR solution was stirred in a CO2 atmosphere and heated to 130 °C at a ramp rate of 10 °C/min. Reaction time parameters of 12 and 30 h and reaction pressure parameters of 0.5 and 1 MPa were adopted orthogonally for the sake of obtaining the CNR with high cyclic carbonate conversion. After the reaction, the viscous solution was transferred to a rotary evaporator in consideration of concentration. Then, the condensed solution was poured into a sufficient amount of methanol for precipitation. Afterward, the product was dried in vacuo at 70 °C for 24 h. The corresponding CNR was abbreviated as CNR-0.5 M12 h, CNR-1M12 h, CNR-0.5M30 h, and CNR-1M30 h, respectively.

2.2.2. Synthesis of N,N’-(Disulfanediylbis(ethane-2,1-diyl))bis(3–4-(mercaptomethyl)-1,3,2-dioaboroborolan-2-yl)benzaminde) (DEDB)

Firstly, as shown in Scheme 3a, the cross-linker Bis(phenylboronic acid carbamoyl) Cystamine (BPBAC) was prepared by a coupling reaction between 3-carboxylphenylboronic acid (CPBA) and cystamine dihydrochloride. NHS (1.40 g, 12.20 mmol), EDC (1.53 g, 8.00 mmol) and CPBA (1.65 g, 10.00 mmol) were dissolved in 20 mL PBS (pH = 6.80) and stirred for 2 h, after which cystamine dihydrochloride (0.90 g, 4.00 mmol) was added. The reaction mixture was stirred at room temperature for 12 h, and the obtained solution was filtered under vacuum. After that, BPBAC was rinsed with deionized water, recrystallized with methanol, and finally dried in a vacuum oven. Finally, the crosslinking agent DEDA, a disulfide-containing boronic ester, was obtained from the reaction of BPBAC and 1-thioglycerol, as shown in Scheme 3b. Briefly, BPBAC (1.35 g, 3.01 mmol) was dissolved in 15 mL deionized water until it was completely dissolved. Then, 1-thioglycerol (1.00 g, 9.28 mmol) was added and stirred for 5 min, after which magnesium sulfate (MgSO4, 1.25 g) was added to react at room temperature for 24 h. After the reaction was completed, the resulting product was filtered, and then the filtrate was concentrated. Finally, the product was kept drying in a 100 °C vacuum oven for 24 h.

2.2.3. Procedure for CNR Cross-Linking

Firstly, 600 mg of the cross-linker was added to the homogeneous solution containing 60 g CNR and 200 mL toluene, and the mixture was stirred continuously at room temperature for 8 h. The viscous solution was concentrated by a rotary evaporator, and then the concentrated solution was dropped into excessive methanol for precipitation. The product featured with a dynamic cross-linked polymer network was put in a vacuum oven and dried at 60 °C for 48 h to further remove the residual solvent. Then, the resultant rubber material was hot pressed. The vulcanizing temperature was 145 °C and the curing time was 21 min, which was obtained by a rubber processing analyzer (RPA).

2.2.4. Characterizations

FTIR. Fourier transform infrared spectra (FTIR) were collected with 32 scans and a resolution of 4 cm−1 in a FTIR spectrometer (Spectrum One FTIR, PerkinElmer, Shelton, CT, USA) with attenuated total reflectance (ATR) mode.
1H NMR. Nuclear magnetic resonance spectra (1H NMR) were collected on a spectrometer (AV 400 NMR, Bruker, Billerica, MA, USA) at 400 MHZ and 25 °C. Deuterated water (D2O) and deuterated chloroform (CDCl3) were used as solvents, and tetramethylsilane (TMS) was used as internal standard.
TGA. Thermal gravimetric analysis (TGA) was conducted by a thermogravimetric analysis (SDT Q600, TA Instruments, Delaware, USA) The specimen was heated from 25 °C to 500 °C under a nitrogen atmosphere at a rate of 10 °C/min.
DSC. Differential scanning calorimetry (DSC) was measured on a instrument (Toledo DSC 822e, Mettler, Zurich, Switzerland). A 5–10 mg sample was heated to 20 °C to remove the thermal history, then cooled to −20 °C, and then heated to 20 °C on an aluminum pan to evaluate the glass transition temperature (Tg) value. The heating rate was 10 K/min, and the whole process was carried out in a nitrogen atmosphere with a flow rate of 10 mL/min.
WCA. The water contact angle (WCA) was measured by a instrument (SDC-100, SINDIN, Dongguang, China) contact angle analyzer as averages of 5 measurements.
Tensile measurement. The tensile properties of CNR were tested according to ISO 37-2017 using a electronic universal testing machine (CMT 4104, SANS, Shenzhen, China) instrument at room temperature with a 500 mm/min strain rate.
Self-Healing tests. The self-healing measurements were carried out by cutting the test sample into two completely separated pieces with a blade and then keeping the two pieces connected gently to ensure the cut surfaces fitted entirely at 60 °C for 12 h, 24 h, and 48 h, respectively. The repaired specimens were measured using standard static uniaxial tensile tests, and the mechanical properties of each sample were measured three times. The self-healing efficiency was determined from the Equation (1):
H self - healing = σ healed σ pristine × 100 %
where Hself-healing is the efficiency of self-healing and σpristine and σhealed are the maximum stress reached during stretching of the original sample and self-healing sample, respectively.
Biodegradable tests. The biodegradable performance was analyzed by weight loss. CNR-DEDB samples in the shape of a square with the size of 1 cm × 1 cm and the weight of W0 were placed in 30 mL glass bottles containing PBS (pH = 7.4) for 28 days. PBS was replaced every 4 days and the weight Wt was recorded every 4 days. Biodegradation efficiency was evaluated from the Equation (2):
  H b i o d e g r a d a t i o n = W 0 W t W 0 × 100 %
where Hbiodegradation is the efficiency of biodegradation, W0 is the initial sample weight, and Wt is the sample weight at a given time.

3. Results and Discussion

3.1. Physicochemical Structure of CNR

From Figure 1, the spectra clearly showed the structural difference between ENR and CNR. The results showed that a new peak appeared at 1720 cm−1, which belonged to C=O stretching vibration of the carbonate units. The epoxy -O- signal peak at 870 cm−1 was not completely faded, which was caused by the fact that the conversion of this reaction for the sample of 1M30 h reached maximum but not 100%, or the in-plane flexural vibration overlapped that of the -CH3 groups [30]. It was proved by FTIR that the epoxy groups were partly converted into cyclic carbonate groups.
Figure 2 illustrated the 1H NMR spectra of ENR, CNR-0.5M30 h, and CNR-1M30 h. Two new signals were found in the CNR 1H NMR spectrum at 4.0 and 4.8 ppm, respectively, and these two signals did not appear in the ENR spectrum. This also verified the formation of cyclic carbonates in CNR, which was consistent with the results of FTIR. According to Furst’s [31,32] method, the 1H chemical shift of cyclic carbonate group of 4,4,5-trimethyl-1,3-dioxolan-2-one was 4.0 ppm. Compared with Kawahara’s [33] propylene carbonate model, the two signals of 4.0 and 4.6 ppm were derived from the equatorial methylene and axial methylene of the cyclic carbonate group interacting with -CH3 and -H, respectively. The signal of CNR at 4.1 ppm could be attributed to the methyl proton of the cyclic carbonate group of CNR. In addition, by using the internal standard method in the 1H NMR of CNR-0.5M30 h and CNR-1M30 h, the conversions from epoxy groups to cyclic carbonate groups were 35% and 51%, respectively.
The change of chemical structure was further investigated by DSC in Figure 3. According to the glass transition temperature measured by DSC, all the Tg values were still significantly lower than room temperature, indicating that CNR was still in a high elastic state at room temperature. In addition, the effect of reaction pressure or reaction time on the glass transition temperature of CNR seemed to be insensitive. In this sense, the five membered ring (cyclic carbonate structure) and the ternary ring (epoxy structure) had a similar influence on the segmental motion of the polymer chain, and the cyclic carbonate structure may be formed randomly across the ENR skeleton.

3.2. Preparation and Analysis of Cross-Linker DEDB

The synthetic procedure for the cross-linker DEDB is presented in Scheme 2. The peak value of 1H NMR of DEDB was shown in Figure 4: δ 2.84 ppm (t, 4H, -S-CH2- in position a), δ 3.50 ppm (t, 4H, -CH2-NH in position b), δ 8.23 ppm (s, 2H, -NH- in position c), δ 7.88 ppm (d, 4H, phenyl proton in position d), δ 8.00 ppm (d, 2H, phenyl proton in position e), δ 7.44 ppm (t, 2H, phenyl proton in position f), δ 4.10 ppm (m, 4H, methylene in position g), δ 3.87 ppm (t, 2H, methine in position h), δ 2.47 ppm (m, 2H, -S-CH2- in position i), δ 2.59 ppm (dd, 2H, -CH2-SH in position i′), and δ 1.45 ppm (s, 2H, -SH in position j). It can be confirmed that DEDB was successfully synthesized.

3.3. Preparation and Characterization of CNR-DEDB Elastomer

According to the FTIR and 1H NMR results of the reaction products obtained under different reaction conditions of press and time, the conversion of 1M30 h was the highest. For this reason, the sample of 1M30 h was investigated further.
For the FTIR of CNR, chemical structure information of unvulcanized and vulcanized CNR-DEDB was shown in Figure 5a. Compared with CNR, new peaks at 1209, 1735, and 3205 cm−1 appeared in CNR-DEDB, which were caused by the stretching vibration of -SH, -BO and -NH of DEDB, respectively. Moreover, by comparing the FTIR spectra of unvulcanized and vulcanized CNR-DEDB, there was a peak at 2337 cm−1 caused by -SH stretching vibration when unvulcanized. However, this peak disappeared after vulcanization. Therefore, it was validated that the epoxy group reacted with the -SH group successfully by click reaction, and DEDB was successfully cross-linked with CNR. In addition, the CNR-DEDB did not dissolve at all after being immersed in toluene solution for 72 h, which further proved the success of cross-linking as shown in Figure 5b.
In order to verify the effect of cyclic carbonate groups and DEDBon thermal stability of ENR after cross-linking, TGA was used to analyze this performance of ENR, CNR, and CNR-DEDB. It was found that the formation of cyclic carbonate groups decreased thermal stability, as shown in Figure 6. With the DEDB cross-linking with CNR, the thermal decomposition temperature of CNR-DEDB increased compared with CNR. As shown in Table 1, the thermal decomposition temperature of CNR-DEDB was the highest, which was 363.96 °C for Td-5% and 464.32 °C for Td-max (Td-5% is 5% weight loss temperature, and Td-max is the maximum weight loss temperature). The raise of thermal decomposition temperature was interpreted to mean that the cross-linking inhibited the thermal degradation of CNR-DEDB, and thus improved the thermal stability of CNR-DEDB.

3.4. Self-Healing Procedure

The self-healing efficiency could be affected by healing time [4,6]. By controlling the healing time at the same repair temperature, the relationship between the repair efficiency and the healing time of the samples was explored. As shown in Figure 7, CNR-DEDB was cut into two parts, and one part was blackened while the other was brown. After healing at 60 °C for 12 h, it was obvious that the black part and the brown part had been partially healed.
In order to quantitatively acquire the healing degree, the self-healing degrees of samples healed for 12 h, 24 h, and 48 h were tested at a constant temperature of 60 °C, respectively. Figure 8a showed the difference of CNR-DEDB with a self-healing time from 12 to 48 h, which the tensile characteristics of the repaired specimens were closer to the tensile properties of the original sample, indicating that the efficiency of self-healing increased with the increase of sample repair time. From Figure 8b, when the repair time was prolonged, the degree of healing increased accordingly. Due to the transesterification of boronic ester bonds and disulfide exchange, the bond reorganization and rearrangement would occur, and the covalent bond would be re-established at the interface of the fracture surface of rubber, as shown in Figure 9 [4,6]. Therefore, more new covalent bonds could be created at the interface with the extension of repairing time. The reaction of the transesterification of boronic ester bonds and disulfide exchange was more complete, and the damaged parts would be repaired fully. After 12 h, the self-healing efficiency reached 41.84%. With the extension of time, the self-healing efficiencies of the samples continued to increase, but the tend of increase was significantly weakened. This resulted from the fact that the disulfide exchange and boronic ester exchange reaction between the segments reached equilibrium after a certain time, and the extension time was not as efficient as before.

3.5. Biodegradation Test

In general, the water absorption of polymer materials in soil burial or soaking process plays an important role in biodegradability. The higher the water absorption, the stronger the biodegradability will be. The water absorption of the tested sample is usually correlated with hydrophilicity [34]. Here, we analyzed the hydrophilicity of ENR and CNR-DEDB by testing the WCA. The smaller WCA was, the stronger hydrophilicity and the stronger biodegradability would be. Figure 10 exhibited that the WCA of ENR was about 104.3 degrees, indicating that it was hydrophobic. After cross-linking, the water contact angle decreased to 36.2 degrees, indicating a significant increase in hydrophilicity. This might be due to the high hydrophilicity of the CNR-DEDB containing cyclic carbonate group. As a consequence, the biodegradation ability of CNR-DEDB should be improved in comparison to ENR.
In this study, the CNR-DEDB elastomers were immersed in water or PBS for 28 days to measure its weight loss. The results shown in Figure 11 demonstrate that after soaking for 28 days, the weight loss rates of ENR-DEDB in PBS, CNR-DEDB in water, and PBS were 1.25%, 2.17%, and 22.37%, respectively. The weight loss rate of CNR-DEDB in PBS was much faster than that of hydrolysis. In fact, weight loss evaluation reflects performance of water resistance as an indirect measure of biodegradability. By comparing the weight loss behavior of CNR-DEDB in PBS and water, we can deduce that the faster weight loss resulted from the combined effect of hydrolysis and biodegradation. It is generally accepted that hydrolysis is one of the initial processes of biodegradation [35]. In a typical enzymatic degradation measurement, trace amounts of lipase will be incorporated into PBS. In this work, ENR was derived from natural rubber, a natural product containing lipase. As a result, weight loss of CNR-DEDB in PBS was obviously faster than that in water. In order to study the degradation behavior under soaking time, the weight loss curve was exponentially fitted, as shown in Figure 11. Interestingly, the weight loss of CNR-DEDB showed a typical exponential process in the first 28 days, demonstrating that its long-term degradation could be controllable.

4. Conclusions

ENR was partially cyclically carbonated by the reaction of epoxy groups in the backbone of ENR with dioxide carbon under high pressure. The residual epoxy groups of the polymer were then clicked to the double terminal thiol groups of the novel cross-linking agent containing a boronic ester bond and a disulfide bond to introduce a self-healing function to the network. The physicochemical structure of cyclic carbonated natural rubber was characterized by means of FTIR, 1H NMR, and DSC. The formation of cyclic carbonate groups endowed the elastomer with biodegradability, which can be confirmed by the further biodegradation test. After the sample was immersed in PBS for 28 days, the biodegradation rate reached 22.37%. At the same time, the biodegradable rubber can be self-healed owing to the introduction of double dynamical covalent bonds. At a constant temperature of 60 °C, transesterification of boron ester bonds and disulfide exchange led to the movement of molecular chain segments, which enabled the cut samples to be re-crosslinked. The healing degree was as high as 82.87%.

5. Patents

Rentong Yu has the patent of “Preparation of biodegradable elastomers” issued by Hainan University; the patent number is CN113336930.

Author Contributions

Q.J.: Investigation, Validation, Formal analysis, Writing original draft. Y.G.: Investigation, Validation. L.L.: Funding acquisition, Supervision. R.Y.: Conceptualization, Supervision, Resources, Writing—review & editing. J.L.: Validation, Resources, Writing—review & editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Key Research and Development Project of Hainan Province, grant number ZDYF2021GXJS030, the Foundation of Guangdong Provincial Key Laboratory of Natural Rubber Processing, P.R. China, grant number 2019B121203004 and Research and Development Program in key area of Guangdong Province of China, grant number 2020B020217001.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

The authors are grateful to the Agricultural Products Processing Research Institute of Chinese Academy of Tropical Agricultural Sciences for providing epoxidized natural rubber.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Koenig, J.L. The Chemical Reactions of Network Structures in Elastomers. Acc. Chem. Res. 1999, 32, 1–8. [Google Scholar] [CrossRef]
  2. Imbernon, L.; Norvez, S. From landfilling to vitrimer chemistry in rubber life cycle. Eur. Polym. J. 2016, 82, 347–376. [Google Scholar] [CrossRef]
  3. Stevenson, K.; Stallwood, B.; Hart, A.G. Tire Rubber Recycling and Bioremediation: A Review. Bioremediation J. 2008, 12, 1–11. [Google Scholar] [CrossRef]
  4. Chen, Y.; Tang, Z.; Liu, Y.; Wu, S.; Guo, B. Mechanically Robust, Self-Healable, and Reprocessable Elastomers Enabled by Dynamic Dual Cross-Links. Macromolecules 2019, 52, 3805–3812. [Google Scholar] [CrossRef]
  5. Zhang, G.; Zhou, X.; Liang, K.; Guo, B.; Li, X.; Wang, Z.; Zhang, L. Mechanically Robust and Recyclable EPDM Rubber Composites by a Green Cross-Linking Strategy. ACS Sustain. Chem. Eng. 2019, 7, 11712–11720. [Google Scholar] [CrossRef]
  6. Cheng, B.; Lu, X.; Zhou, J.; Qin, R.; Yang, Y. Dual Cross-Linked Self-Healing and Recyclable Epoxidized Natural Rubber Based on Multiple Reversible Effects. ACS Sustain. Chem. Eng. 2019, 7, 4443–4455. [Google Scholar] [CrossRef]
  7. Wemyss, A.M.; Bowen, C.; Plesse, C.; Vancaeyzeele, C.; Nguyen, G.T.M.; Vidal, F.; Wan, C. Dynamic crosslinked rubbers for a green future: A material perspective. Mater. Sci. Eng. R Rep. 2020, 141, 100561. [Google Scholar] [CrossRef]
  8. Zou, W.; Dong, J.; Luo, Y.; Zhao, Q.; Xie, T. Dynamic Covalent Polymer Networks: From Old Chemistry to Modern Day Innovations. Adv. Mater. 2017, 29, 1606100. [Google Scholar] [CrossRef]
  9. Chakma, P.; Konkolewicz, D. Dynamic Covalent Bonds in Polymeric Materials. Angew. Chem. Int. Ed. Engl. 2019, 58, 9682–9695. [Google Scholar] [CrossRef]
  10. Kloxin, C.J.; Bowman, C.N. Covalent adaptable networks: Smart, reconfigurable and responsive network systems. Chem. Soc. Rev. 2013, 42, 7161–7173. [Google Scholar] [CrossRef] [Green Version]
  11. Neal, J.A.; Mozhdehi, D.; Guan, Z. Enhancing mechanical performance of a covalent self-healing material by sacrificial noncovalent bonds. J. Am. Chem. Soc. 2015, 137, 4846–4850. [Google Scholar] [CrossRef] [PubMed]
  12. Wen, J.; Jia, Z.; Zhang, X.; Pan, M.; Yuan, J.; Zhu, L. Tough, thermo-Responsive, biodegradable and fast self-healing polyurethane hydrogel based on microdomain-closed dynamic bonds design. Mater. Today Commun. 2020, 25, 101569. [Google Scholar] [CrossRef]
  13. Beaupre, D.M.; Weiss, R.G. Thiol- and Disulfide-Based Stimulus-Responsive Soft Materials and Self-Assembling Systems. Molecules 2021, 26, 3332. [Google Scholar] [CrossRef]
  14. Yang, G.-W.; Wu, G.-P. High-Efficiency Construction of CO2-Based Healable Thermoplastic Elastomers via a Tandem Synthetic Strategy. ACS Sustain. Chem. Eng. 2018, 7, 1372–1380. [Google Scholar] [CrossRef]
  15. Cash, J.J.; Kubo, T.; Bapat, A.P.; Sumerlin, B.S. Room-Temperature Self-Healing Polymers Based on Dynamic-Covalent Boronic Esters. Macromolecules 2015, 48, 2098–2106. [Google Scholar] [CrossRef]
  16. Zeng, Y.; Liu, S.; Xu, X.; Chen, Y.; Zhang, F. Fabrication and curing properties of o-cresol formaldehyde epoxy resin with reversible cross-links by dynamic boronic ester bonds. Polymer 2020, 211, 123116. [Google Scholar] [CrossRef]
  17. Zych, A.; Tellers, J.; Bertolacci, L.; Ceseracciu, L.; Marini, L.; Mancini, G.; Athanassiou, A. Biobased, Biodegradable, Self-Healing Boronic Ester Vitrimers from Epoxidized Soybean Oil Acrylate. ACS Appl. Polym. Mater. 2020, 3, 1135–1144. [Google Scholar] [CrossRef]
  18. Thiessen, M.; Abetz, V. Influence of the Glass Transition Temperature and the Density of Crosslinking Groups on the Reversibility of Diels-Alder Polymer Networks. Polym. Basel 2021, 13, 1189. [Google Scholar] [CrossRef]
  19. Raut, S.K.; Mondal, P.; Parameswaran, B.; Sarkar, S.; Dey, P.; Gilbert, R.; Bhadra, S.; Naskar, K.; Nair, S.; Singha, N.K. Self-healable ultrahydrophobic modified bio-based elastomer using Diels-Alder ‘click chemistry’. Eur. Polym. J. 2021, 146, 110204. [Google Scholar] [CrossRef]
  20. Zheng, P.; McCarthy, T.J. A surprise from 1954: Siloxane equilibration is a simple, robust, and obvious polymer self-healing mechanism. J. Am. Chem. Soc. 2012, 134, 2024–2027. [Google Scholar] [CrossRef]
  21. Taynton, P.; Yu, K.; Shoemaker, R.K.; Jin, Y.; Qi, H.J.; Zhang, W. Heat- or water-driven malleability in a highly recyclable covalent network polymer. Adv. Mater. 2014, 26, 3938–3942. [Google Scholar] [CrossRef] [PubMed]
  22. Memon, H.; Wei, Y.; Zhu, C. Correlating the thermomechanical properties of a novel bio-based epoxy vitrimer with its crosslink density. Mater. Today Commun. 2021, 29, 102814. [Google Scholar] [CrossRef]
  23. Fu, Q.; Yan, Q.; Jiang, X.; Fu, H. Heat driven self-healing isocyanate-based crosslinked three-arm Star-shaped polyglycolide based on dynamic transesterification. React. Funct. Polym. 2020, 146, 104440. [Google Scholar] [CrossRef]
  24. Capelot, M.; Montarnal, D.; Tournilhac, F.; Leibler, L. Metal-catalyzed transesterification for healing and assembling of thermosets. J. Am. Chem. Soc. 2012, 134, 7664–7667. [Google Scholar] [CrossRef]
  25. Nishimura, Y.; Chung, J.; Muradyan, H.; Guan, Z. Silyl Ether as a Robust and Thermally Stable Dynamic Covalent Motif for Malleable Polymer Design. J. Am. Chem. Soc. 2017, 139, 14881–14884. [Google Scholar] [CrossRef]
  26. Zych, A.; Pinalli, R.; Soliman, M.; Vachon, J.; Dalcanale, E. Polyethylene vitrimers via silyl ether exchange reaction. Polymer 2020, 199, 122567. [Google Scholar] [CrossRef]
  27. Tanasi, P.; Hernández Santana, M.; Carretero-González, J.; Verdejo, R.; López-Manchado, M.A. Thermo-reversible crosslinked natural rubber: A Diels-Alder route for reuse and self-healing properties in elastomers. Polymer 2019, 175, 15–24. [Google Scholar] [CrossRef]
  28. Chen, Y.; Tang, Z.; Zhang, X.; Liu, Y.; Wu, S.; Guo, B. Covalently Cross-Linked Elastomers with Self-Healing and Malleable Abilities Enabled by Boronic Ester Bonds. ACS Appl. Mater. Interfaces 2018, 10, 24224–24231. [Google Scholar] [CrossRef]
  29. Gan, S.N.; Yahya, R.; Rooshenass, P. Comparison of Three Different Degradation Methods To Produce Liquid Epoxidized Natural Rubber. Rubber Chem. Technol. 2016, 89, 177–198. [Google Scholar] [CrossRef] [Green Version]
  30. Ke, J.; Li, X.; Wang, F.; Kang, M.; Feng, Y.; Zhao, Y.; Wang, J. The hybrid polyhydroxyurethane materials synthesized by a prepolymerization method from CO2-sourced monomer and epoxy. J. CO2 Util. 2016, 16, 474–485. [Google Scholar] [CrossRef]
  31. Fuerst, A.; Pretsch, E. A computer program for the prediction of carbon-13-NMR chemical shifts of organic compounds. Anal. Chim. Acta 1990, 229, 17–25. [Google Scholar] [CrossRef]
  32. Pretsch, E.; Furst, A.; Badertscher, M.; Buergin, R.; Munk, M.E. C13Shift: A computer program for the prediction of carbon-13 NMR spectra based on an open set of additivity rules. J. Chem. Inf. Comput. Sci. 1992, 32, 291. [Google Scholar] [CrossRef]
  33. Kawahara, S.; Saito, T. Preparation of carbonated natural rubber. J. Polym. Sci. Part A Polym. Chem. 2006, 44, 1561–1567. [Google Scholar] [CrossRef]
  34. Li, F.; Su, Y.; Pi, G.; Ma, P.X.; Lei, B. Biodegradable, Biomimetic Elastomeric, Photoluminescent, and Broad-Spectrum Antibacterial Polycitrate-Polypeptide-based Membrane toward Multifunctional Biomedical Implants. ACS Biomater. Sci. Eng. 2018, 4, 3027–3035. [Google Scholar] [CrossRef] [PubMed]
  35. Kijchavengkul, T.; Auras, R.; Rubino, M.; Selke, S.; Ngouajio, M.; Fernandez, R.T. Biodegradation and hydrolysis rate of aliphatic aromatic polyester. Polym. Degrad. Stab. 2010, 95, 2641–2647. [Google Scholar] [CrossRef]
Scheme 1. (a) Formation of cyclic carbonate groups and the double dynamic covalent network; (b) Generation of CNR-DEDB linkage via thermally initiated thiols and epoxy groups of “click” reaction.
Scheme 1. (a) Formation of cyclic carbonate groups and the double dynamic covalent network; (b) Generation of CNR-DEDB linkage via thermally initiated thiols and epoxy groups of “click” reaction.
Polymers 14 01380 sch001
Scheme 2. Synthesis of CNR from the reaction of ENR and CO2 with the catalyst of TBAI.
Scheme 2. Synthesis of CNR from the reaction of ENR and CO2 with the catalyst of TBAI.
Polymers 14 01380 sch002
Scheme 3. (a) Synthesis of BPBAC via the EDC/NHS coupling reaction between CPBA and cystamine dihydrochloride; (b) Synthesis of DEDB from the reaction of BPBACwith 1-thioglycerol.
Scheme 3. (a) Synthesis of BPBAC via the EDC/NHS coupling reaction between CPBA and cystamine dihydrochloride; (b) Synthesis of DEDB from the reaction of BPBACwith 1-thioglycerol.
Polymers 14 01380 sch003
Figure 1. FTIR spectra of ENR and CNR under different reaction pressures and reaction time.
Figure 1. FTIR spectra of ENR and CNR under different reaction pressures and reaction time.
Polymers 14 01380 g001
Figure 2. 1H NMR spectra of ENR, CNR-0.5M30 h and 1M30 h.
Figure 2. 1H NMR spectra of ENR, CNR-0.5M30 h and 1M30 h.
Polymers 14 01380 g002
Figure 3. DSC curves of CNR with various reaction pressure and time compared with ENR.
Figure 3. DSC curves of CNR with various reaction pressure and time compared with ENR.
Polymers 14 01380 g003
Figure 4. 1H NMR spectrum of DEDB.
Figure 4. 1H NMR spectrum of DEDB.
Polymers 14 01380 g004
Figure 5. (a) FTIR spectra of CNR, uncured CNR-DEDB, and cured CNR-DEDB; (b) Photographs of CNR-DEDB after 72 h in toluene.
Figure 5. (a) FTIR spectra of CNR, uncured CNR-DEDB, and cured CNR-DEDB; (b) Photographs of CNR-DEDB after 72 h in toluene.
Polymers 14 01380 g005
Figure 6. TGA curves of ENR, CNR and CNR-DEDB.
Figure 6. TGA curves of ENR, CNR and CNR-DEDB.
Polymers 14 01380 g006
Figure 7. Photographs of self-healing of cut-off samples. (a) uncut; (b) cut; (c) self-healing.
Figure 7. Photographs of self-healing of cut-off samples. (a) uncut; (b) cut; (c) self-healing.
Polymers 14 01380 g007
Figure 8. (a) Tensile curves of uncut sample and the CNR-DEDB elastomers with the self-healing time of 12 h, 24 h, and 48 h; (b) Self-healing efficiencies of the CNR-DEDB healed for 12 h, 24 h, and 48 h.
Figure 8. (a) Tensile curves of uncut sample and the CNR-DEDB elastomers with the self-healing time of 12 h, 24 h, and 48 h; (b) Self-healing efficiencies of the CNR-DEDB healed for 12 h, 24 h, and 48 h.
Polymers 14 01380 g008
Figure 9. The mechanism of dynamic exchange of disulfide bonds and boronic ester.
Figure 9. The mechanism of dynamic exchange of disulfide bonds and boronic ester.
Polymers 14 01380 g009
Figure 10. WCA of ENR and CNR-DEDB.
Figure 10. WCA of ENR and CNR-DEDB.
Polymers 14 01380 g010
Figure 11. Weight loss ratio of CNR after soaking in PBS and water for 28 days, Weight loss ratio of ENR after soaking in PBS for 28 days and Exponential fitting curve of weight loss process of samples after soaking in PBS.
Figure 11. Weight loss ratio of CNR after soaking in PBS and water for 28 days, Weight loss ratio of ENR after soaking in PBS for 28 days and Exponential fitting curve of weight loss process of samples after soaking in PBS.
Polymers 14 01380 g011
Table 1. Td-5% and Td-max of ENR, CNR and CNR-DEDB.
Table 1. Td-5% and Td-max of ENR, CNR and CNR-DEDB.
SampleTd-5%(°C)Td-max(°C)
ENR336.06452.17
CNR322.16447.36
CNR-DEDB363.96464.32
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jiang, Q.; Gao, Y.; Liao, L.; Yu, R.; Liao, J. Biodegradable Natural Rubber Based on Novel Double Dynamic Covalent Cross-Linking. Polymers 2022, 14, 1380. https://doi.org/10.3390/polym14071380

AMA Style

Jiang Q, Gao Y, Liao L, Yu R, Liao J. Biodegradable Natural Rubber Based on Novel Double Dynamic Covalent Cross-Linking. Polymers. 2022; 14(7):1380. https://doi.org/10.3390/polym14071380

Chicago/Turabian Style

Jiang, Qinggeng, Yi Gao, Lusheng Liao, Rentong Yu, and Jianhe Liao. 2022. "Biodegradable Natural Rubber Based on Novel Double Dynamic Covalent Cross-Linking" Polymers 14, no. 7: 1380. https://doi.org/10.3390/polym14071380

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