Home Preparation and properties of dynamic crosslinked styrene butadiene rubber
Article Publicly Available

Preparation and properties of dynamic crosslinked styrene butadiene rubber

  • Hui Lu , Pingyin Wang , Yaozhu Tian EMAIL logo and Zhu Luo EMAIL logo
Published/Copyright: September 4, 2023
Become an author with De Gruyter Brill

Abstract

As the second largest synthetic rubber after styrene butadiene rubber, cis-butadiene rubber (BR) is one of the important raw materials for automobile tires and cold-resistant products. Herein, a traditional rubber preparation process was used to introduce dynamic reversible bonds into BR based on an “imitative” click reaction. Compared with traditional complex self-healing techniques, this method is undoubtedly simpler and more efficient. Dynamic reversible bonds are able to break and recombine under the stimulation of external conditions, which endow rubber with self-healing properties. We use the small biological molecule lipoic acid (LA) as a cross-linking agent and cross-link LA and BR through mechanical compounding and hot press vulcanization to obtain self-healing butadiene rubber (BLA). In addition, BLA-(Zn2+) was further prepared by introducing Zn2+ to form metal-oxygen coordination bonds with carboxyl groups. And systematically studied the effect of Zn2+ on the mechanical properties and self-healing properties of cross-linked BR. Through the combined action of disulfide bonds, hydrogen bonds and Zn2+-O coordination bonds, BLA-(Zn2+) has better properties than BLA, the tensile strength can reach 3.76 MPa, and the repair efficiency is about 82 %. This simple preparation process is certainly more cost effective.

1 Introduction

Rubber has been used in tires, gloves, seals, and many other fields because of its excellent solvent resistance, mechanical properties, and thermal stability. However, conventional rubber vulcanization is non-reversible cross-linking, where the rubber network chains are neither dissolved nor melted the vulcanization of traditional rubber is irreversible cross-linking, and the rubber network chain is insoluble and infusible. Once the rubber structure is destroyed, it cannot be reused, which limits its service life [1], [2], [3], [4], [5]. Not only that, once the rubber is scrapped, its using value will be lost, and subsequent processing is also very difficult, and it is difficult to recycle and reuse. Therefore, it makes sense to design reversible dynamic cross-linked rubbers so that they can be uncross-linked and re-cross-linked under certain stimuli. The reversible cross-linking bond is inserted into the rubber, and the rubber can be repaired automatically when it is damaged to increase the service life, thereby producing more valuable recycled rubber, improving the ecological environment and saving resources [6], [7], [8], [9]. The most commonly used intrinsic self-healing methods are roughly divided into two categories: reversible covalent bond type and reversible non-covalent bond type. Among them, the reversible non-covalent bond type realizes the self-healing of materials through the reversibility of intermolecular interactions. Compared with reversible covalent bonds, reversible non-covalent bonds can self-heal without external stimuli, including hydrogen bonds [10], metal bond coordination [11], host–guest interactions [12]. However, its tensile strength and self-healing efficiency are relatively low. On the contrary, the reversible covalent bond type with stronger bond energy mainly realizes the self-healing of the material through the dynamic properties of special covalent bonds, including Diels–Alder reaction [13, 14], boronateester bonds [15], disulfide bonds [16], imine bonds [17], and acylhydrazone bonds [18].

Smart materials with self-healing function have always been a hot topic in the field of materials. Most research has focused on the optimization of complex processes, the enhancement of mechanical properties and the design of molecular structures. For example, Bai Jing et al. [9] prepared a self-healing styrene-butadiene-styrene rubber (SBS). Through the thermally reversible cross-linking process of the Diels-Alder reaction, the prepared functionalized cross-linked SBS has a tensile strength of 14.96 MPa, which is higher than Pure SBS improves nearly 8 times, the elongation at break can reach 800 %, and it has excellent mechanical properties. Wang Dong et al. [19] prepared a cross-linked BR containing dynamic bonds, and 3-mercaptopropionic acid was grafted into the BR chain through the thiol-ene click reaction, so as to interact with the amino groups and side chains in the cross-linking agent. The carboxyl groups in the compound form ionic hydrogen bonds to obtain cross-linked BR with high toughness and ductility (1800 %). Guo Wenjin et al. [20] used click chemistry to select the coordination system of carboxyl group and Zn2+ to prepare polybutadiene rubber with dynamic cross-linked network. The study showed that the material has good self-healing properties and remodeling performance, the repair efficiency can reach 100 %. In particular, based on the self-repair of metal coordination bonds, the conditions for realizing self-repair are relatively mild and the repair efficiency is high.

As a kind of click chemistry, the thiol-ene click reaction not only has all the advantages of click chemistry, but also has its own remarkable characteristics, such as no need to add metal catalysts, a wide range of applications, almost all thiols and most olefins can react and the reaction conditions are simple. Based on these advantages, thiol-ene click reactions have been extensively studied in recent years. Bai et al. [21] reported a new type of cross-linked polybutadiene rubber. The polybutadiene rubber first undergoes a thiol-ene click reaction with furfuryl mercaptan to modify the rubber chain with furan groups. The dynamic cross-linked polybutadiene rubber obtained by forming a cross-linked network through the Diels–Alder reaction can be recycled and remodeled at high temperature, and can achieve good self-healing at 150 °C after cracks appear.

Lipoic acid (LA) is a naturally occurring biological small molecule [22, 23]. It is usually used as a necessary coenzyme for animal aerobic metabolism. It plays an important role in maintaining animal physiological metabolism, balancing blood sugar concentration, enhancing liver function, and improving energy metabolism rate. A natural small molecule with biosafety. Due to the existence of two special functional groups, namely disulfide bonds and carboxyl groups, adding them into polymers can form dynamic covalent disulfide bonds and dynamic non-covalent hydrogen bonds. By introducing LA, Yuetao Liu et al. [24] prepared a silicone elastomer network with three different types of reversible cross-linking bonds: disulfide bonds, hydrogen bonds and metal-ligand bonds. The material not only has good self-healing properties, but also has excellent mechanical properties. The “simulated” click reaction was first proposed by Bo Huang et al. [25] in the synthesis of P-LA to solve the problem of the deterioration of thiol intermediates. By ultraviolet light irradiation, the disulfide bond is broken to form a thiol radical, which avoids the oxidation of the thiol intermediate to generate by-products. The reaction mechanism is similar to the thiol-ene click reaction, so they named it “simulated” click reaction.

In this study, based on the thiol-ene click reaction mechanism, LA was introduced into commercial BR for the first time as a cross-linking agent. By designing a simple method, the reversible disulfide bond is introduced into the carbon-carbon double bond of BR to undergo a click reaction similar to thiol-ene (the reaction mechanism is shown in Figure 1), and the compound containing cross-linked BR with dynamic disulfide and hydrogen bonds. Due to the reversibility of disulfide bonds and hydrogen bonds, cross-linked BR has certain ductility and self-healing properties. In order to further improve its mechanical properties, we introduce metal coordination bonds, which can not only improve the mechanical properties of materials, but also greatly improve their repair efficiency through their synergistic effect with hydrogen bonds. Compared with other methods for preparing cross-linked rubber containing reversible bonds, the advantage of this preparation method is that it does not require complex synthesis routes, but prepares samples through simple blending and hot-pressing vulcanization. This preparation method is more in line with traditional rubber processing methods. And using LA as a cross-linking agent is safer and more environmentally friendly, and is an easier way to mass-produce.

Figure 1: 
Preparation process of LA cross-linked BR by “simulated click” reaction.
Figure 1:

Preparation process of LA cross-linked BR by “simulated click” reaction.

2 Materials and methods

2.1 Materials

Commercial BR were supplied by PetroChina Co., Ltd. SBR 1502 E (styrene content = 23.3 wt%, vinyl content = 76.7 wt%) was obtained from Sinopec-China Petroleum Co. Ltd. LA (LA, 98 %) were obtained from Shanghai Meryer Chemical Technology Co., Ltd. 2,2′-Azobis(2-methylpropionitrile) (AIBN, 98 %) were supplied by Shanghai Macklin Biochemical Technology Co., Ltd. Toluene was analytically pure and purchased from Chongqing Chuandong Chemical Co., Ltd. Dichloromethane, and anhydrous ethanol were also analytically pure, supplied by Chengdu Cologne Co., Ltd.

2.2 Preparation

The BR is properly mixed by a two-roller refiner, and added different content of LA in sequence (0.5 %, 1 %, 1.5 %, 2 %, 3 % compared to the molar mass of vinyl in the BR) and AIBN (16.7 % compared to the molar mass of LA) for 14 min. After mixing for 14 min, the compound is obtained. Leave it for 24 h to eliminate internal stress. Set the temperature to 130 °C, measure the positive curing time with a rotorless vulcanizer, and vulcanize with a plate vulcanizer to obtain cross-linked BR by thermally initiating a “simulated click” reaction. After 24 h at room temperature, experimental test strips were prepared as required. The preparation process is shown in Figure 1. The components are named: BLA0.5, BLA1, BLA1.5, BLA2, and BLA3.

By adding different contents of ZnCl2 to the uncured BLA2 (the molar ratios of LA and ZnCl2 were 80:1, 60:1, 40:1, and 20:1, respectively) and blending for 16 min, the Zn2+-O coordination bond was obtained. The network was placed for 24 h to relieve internal stress, and the temperature was set to 130 °C. The sample names are: BLA-(Zn2+)-(LA:Zn2+ = 80:1),BLA-(Zn2+)-(LA:Zn2+ = 60:1), BLA-(Zn2+)-(LA:Zn2+ = 40:1), BLA-(Zn2+)-(LA:Zn2+ = 20:1).

2.3 Characterization

  1. Fourier transform infrared spectroscopy (FTIR) analysis: the ATR transmission method of the American Nicolet Nexus670 FTIR instrument was used to perform infrared test and characterization on the surface of BR and BLA with different components and different content.

  2. Raman spectroscopy (Raman) analysis: the Raman spectrometer produced by HORIBA (HORIBA Scientific Lab RAM HR Evolution) was used to characterize the bond structure of the polymer chain. The laser wavelength was 785 nm, and the measurement range of the Raman spectrum was 300 cm−1–1100 cm−1.

  3. Curing time test: weigh about 4 g of the mixed uncrosslinked rubber, use a rotorless vulcanizer (MD-3000 A, Taiwan High Speed Rail), set the temperature to 130 °C, and test the curing time.

  4. Thermogravimetric analysis (TGA): the sample of rubber composite was tested by thermogravimetric analyzer (Q50, Waters, USA). Test conditions: nitrogen flow rate is 100 mL/min, scanning speed is 10 °C/min, and temperature is 20–1050 °C.

  5. Swelling test: calculate the sol fraction, swelling rate, and cross-linking density of SBR and BR by using equilibrium swelling experiment [26, 27]. First of all, the initial mass was taken as m 1 , the sample was soaked in toluene for 48 h. Then, the sample was taken out, the surface was dried quickly, and the swelling mass of the sample was m 2 . Finally, the sample was placed in a vacuum oven and dried at 60 °C until constant weight. After drying, the mass of the sample was measured as m 3 . Three samples were measured for each component.

    The crosslink density (VCLD) was calculated by Flory–Rehner formula [28, 29]:

    (1) V C L D = 1 V s [ ln ( 1 V R ) + V R + χ V R 2 V R 1 / 3 ]

    V R is the volume fraction of rubber in the swelling sample, and its formula is as follows:

    (2) V R = m 3 m 3 + ( m 2 m 3 ) ρ ρ s

    V S is the molar volume of the solvent (toluene is 105.9 mL/mol at room temperature), ρ is the density of the polymer (measured 1502 E SBR density: 0.901 g/cm³; BR density: 0.905 g/cm³), ρs is the solvent density (toluene density: 0.866 g/cm³).

    The swelling rate is calculated as follows:

    (3) m 2 m 3 m 3 × 100 %

    The sol fraction is calculated as follows:

    (4) m 1 m 3 m 1 × 100 %
  6. X-ray photoelectron spectroscopy (XPS) analysis: the elements in the rubber film were determined by X-ray photoelectron spectroscopy (Thermo Scientific K-Alpha, Thermo). The excitation source was Al Kα rays (0.6eV).

  7. Dynamic mechanical analysis (DMA): SLA and BLA were measured by dynamic mechanical analyzer (Q800 type, TA) under the following conditions: double cantilever deformation mode, frequency of 1 Hz; the amplitude is 8 μm; temperature scanning range is −125 °C–80 °C; the heating rate is 5 °C/min.

  8. Mechanical performance test: universal testing machine (TSE104B, Shenzhen Wance) was used to test the tensile properties, the test standard was GB/T 528–1998, the sample was dumbbell shape, the tensile speed was 100 mm/min. At least 5 samples were tested in each group and the arithmetic average was taken. The stress-strain cycle test was performed on the sample, the tensile strain was 100 %, and the tensile speed was 100 mm/min. After each cycle, the sample was placed for 15 min before the next cycle test.

  9. Polarizing microscope: the sample was cut open with a blade, and the fracture was photographed with a polarizing microscope (Axio Scope Al A, Zeiss, Germany). Then the fracture was fully contacted and placed in an oven with a set temperature for a certain time, and then the repair of the fracture after self-healing was photographed.

  10. Self-healing test: the samples were cut into two pieces with a blade, and then the sample sections were joined together, and applied a certain force for 6 min. Then the mechanical properties of the samples under different healing temperatures and healing times were tested. The formula for calculating rubber self-healing efficiency is as follows [25, 30]:

    (5) η = σ ( h e a l e d ) σ ( i n i t i a l ) × 100 %

    σ(initial) and σ(waken) are the tensile strength before and after healing, respectively.

  11. Scanning electron microscope (SEM): the surface of the prepared cross-linked BLA was sprayed with gold, and the surface morphology characteristics were observed with a scanning electron microscope (Zeiss Merlin, Zeiss, Germany). Then, the elemental distribution on the surface was analyzed with energy spectrum analyzer (Oxford Instruments, UK).

3 Results and discussion

3.1 Structural characterization

Figure 2A is a photograph of the sample before and after swelling with toluene for 48 h. When immersed in toluene for 48 h, BR was completely dissolved to form a viscous polymer solution, while BLA only swelled after soaking for 48 h, indicating that LA and BR can also be cross-linked by hot pressing after blending. This is also demonstrated by the vulcanization time-torque (S´) curves of the BLA samples (Supplementary Figure S1). As shown in Supplementary Figure S2, BLA had the same results after immersion in other common solvents. The crosslink density gradually increased with increasing LA content (Figure 2B). Further ATR-FTIR spectroscopy, Raman spectroscopy and XPS tests were applied to confirm the structural features of BLA and BLA-Zn.

Figure 2: 
Structural characterization of BLA and BLA-(Zn2+). A) Photos of samples before and after swelling for 48 h; B) crosslinking density (VCLD) of BLA; C) IR absorption spectra of BR, BR + LA, BLA; D) IR absorption spectra of BLA, BLA-(Zn2+); E) Raman spectrum of BLA; F) SEM image and element distribution of BLA-(Zn2+).
Figure 2:

Structural characterization of BLA and BLA-(Zn2+). A) Photos of samples before and after swelling for 48 h; B) crosslinking density (VCLD) of BLA; C) IR absorption spectra of BR, BR + LA, BLA; D) IR absorption spectra of BLA, BLA-(Zn2+); E) Raman spectrum of BLA; F) SEM image and element distribution of BLA-(Zn2+).

In the infrared spectrum (Figure 2C), about 726 cm−1 is attributed to the out-of-plane bending vibration of = CH, and after cross-linking, the peak here does not change significantly, and the characteristic peak of C–S bond may be related to the surface of = CH The external bending vibration peaks overlap each other, so the existence of C–S bonds cannot be judged, but the existence of C–S bonds is detected in the subsequent XPS analysis (Figure 3). It can be clearly seen from the BLA infrared spectrum of Figure 2C that a characteristic peak appears at 1699 cm−1, which is the stretching vibration peak of C=O, which proves that LA is successfully inserted into the double bond of BR. Supplementary Figure S3 shows the decrease of the C=O characteristic peak from 1710 cm−1 to 1690 cm−1 for different LA contents. It indicates that the C=O characteristic peak is red-shifted with the increase of LA content, which is due to the hydrogen bonding of C=O with –OH. Therefore, the addition of LA can generate reversible hydrogen bond in BLA and confer BR self-healing function (Supplementary Figure S5).

Figure 3: 
XPS survey spectra A) and S2p spectra B) of BLA.
Figure 3:

XPS survey spectra A) and S2p spectra B) of BLA.

Supplementary Figure S3 shows the wavenumber change of the C=O characteristic peak under different LA contents. It can be seen from the figure that when the LA content is 0.5 %, the C=O characteristic peak is at 1710 cm−1, while the LA content is 0.5 %. When the content increased to 3 %, the wavenumber decreased to 1690 cm−1, indicating that with the increase of LA content, the characteristic peak of C=O shifted to red, which was caused by the hydrogen bond between C=O and –OH.

Figure 2D shows the infrared absorption spectra of BLA and BLA-(Zn2+) under the same LA content. It can be seen from the figure that 1699 cm−1 is the stretching vibration peak of C=O in BLA, and 1690 cm−1 is the stretching vibration peak of C=O in BLA-(Zn2+) and the characteristic peak of C=O in BLA-(Zn2+) are shifted to lower wavenumbers compared with the characteristic peak of C=O in BLA, and have a red shift, while 1264 cm−1 The stretching vibration peak of C–OH on the carboxyl group at basically disappeared in BLA-(Zn2+), indicating that Zn2+ successfully formed a coordination bond with O in the carboxyl group.

As shown in Figure 2E, the Raman spectrum of BLA has the absorption peak of S–S bond at 510 cm−1 and the absorption peak of C–C bond at 991 cm−1. The S–S bond is generated because a part of the LA in the reaction opens the five-membered ring through thermal initiation, and the formed sulfur radical reacts with the sulfur radical of another LA to form a reversible covalent S–S bond. The hydrogen bond interaction, through the introduction of double reversible bonds, enables BLA to obtain self-healing function, which indicates that self-healing rubber can still be prepared from LA in BR.

Figure 2F is the SEM image of BLA-(Zn2+) and the energy spectrum of surface scanning. It can be seen from the figure that the surface of BLA-(Zn2+) is smooth and smooth, and there is no Zn2+ aggregate, which is also measured from the energy spectrometer. The existence of three elements, S, O, and Zn, and the distribution of the three elements is uniform, which also shows that Zn2+ exists in the BLA-(Zn2+) material after forming a coordination bond with O.

The DMA of BLA shows the relationship between the storage modulus, tan δ and temperature dependence of BLA under different LA contents (Figure 4). From the storage modulus-temperature curve of Figure 4A, it can be seen that when the temperature is lower than about −120 °C, BLA is in a glass state, showing a higher modulus, while as the temperature increases, the rubber molecular chain The exercise capacity is enhanced, so the modulus decreases, and with the increase of LA content, the modulus increases from 4813 MPa to 6272 MPa. Furthermore, it is observed from Figure 4B that the peak of the tan δ-temperature curve decreases from 0.17 to 0.16 with the increase of LA, and the temperature corresponding to the peak increases from −98 °C to −95 °C, resulting in an increase in Tg. Since the crosslinking density increases with the increase of LA content, the movement of the rubber molecular chains is hindered, which is consistent with the swelling experimental results.

Figure 4: 
Storage modulus curves A) and tan δ curves B) of BLA with different content LA.
Figure 4:

Storage modulus curves A) and tan δ curves B) of BLA with different content LA.

BLA was tested by TGA to analyze its thermal stability. Figure 5 shows the thermal weight loss curves of BLA under different LA contents. It can be seen from the figure that the initial decomposition temperature of SLA is around 330 °C, and the temperature of the maximum weight loss rate is around 450 °C. Still have excellent thermal stability.

Figure 5: 
TGA curves of BLA with different content LA.
Figure 5:

TGA curves of BLA with different content LA.

3.2 Mechanical properties

Figure 6A shows the stress-strain curves of BLA with different LA contents. It can be seen from the figure that with the increase of LA content, the tensile strength of BLA increases and the elongation at break decreases. When the LA content is 0.5 %, the tensile strength of BLA is 1.12 MPa, and the elongation at break is 937 %. With the increase of LA content, the tensile strength can reach up to 2.21 MPa, and the corresponding elongation at break is 519 %. The increase in tensile strength is due to the increase in LA content, which provides more disulfide bonds to interact with hydrogen bonds, which increases the cross-linking density of BLA, thereby forming more cross-linking points, which restricts the movement of the rubber molecular chain. As a result, its tensile strength increases and its elongation at break decreases.

Figure 6: 
Mechanical properties of BLA and BLA-(Zn2+). A) Stress–strain curve of BLA with different content LA; B) the first cyclic loading–unloading stress–strain curves of BLA with different content LA at room temperature; C) the first, the second and the third circle stress–strain curves of BLA2; D) stress–strain curve of BLA2 with different content Zn2+; E) the first cyclic loading–unloading stress–strain curves of BLA2 with different content Zn2+ at room temperature; F) the first, the second and the third circle stress–strain curves of BLA-(Zn2+)-LA:Zn2+ = 40:1.
Figure 6:

Mechanical properties of BLA and BLA-(Zn2+). A) Stress–strain curve of BLA with different content LA; B) the first cyclic loading–unloading stress–strain curves of BLA with different content LA at room temperature; C) the first, the second and the third circle stress–strain curves of BLA2; D) stress–strain curve of BLA2 with different content Zn2+; E) the first cyclic loading–unloading stress–strain curves of BLA2 with different content Zn2+ at room temperature; F) the first, the second and the third circle stress–strain curves of BLA-(Zn2+)-LA:Zn2+ = 40:1.

It can be observed from Figure 6B that the stress during stretching is greater than that during unloading, resulting in a hysteresis phenomenon, which is due to the disentanglement of some BR molecular chains during the stretching process. As the LA content increases, the strain recovery is greater, which is due to the increase in the degree of cross-linking and the better the resilience of the rubber. From Figure 6C, it can be seen that the stress-strain curve of the three cycles does not change much, indicating that in more BLA still has good resilience in the secondary stretching, which indicates that it has the characteristics of permanently cross-linked rubber.

Figure 6D shows the stress-strain curves of BLA-(Zn2+) with different Zn2+ contents. It can be observed that the addition of Zn2+ makes the mechanical properties higher than that of BLA. With the increase of Zn2+ content, the tensile strength of BLA-(Zn2+) increases and the elongation at break decreases. When the molar ratio of LA to Zn2+ was 80:1, the tensile strength was 2.52 MPa and the elongation at break was 543 %, while when the molar ratio was 20:1, the tensile strength increased to 3.76 MPa, and the elongation at break was 3.76 MPa. reduced to 484 %. This is because the metal coordination bond formed by Zn2+ with oxygen in the carboxyl group of LA has high bond energy and good kinetics.

Figure 6E is the first loading–unloading stress-strain curve of BLA-(Zn2+) material with different molar ratios of LA to Zn2+, (Figure 6F) is the first three times that the molar ratio of LA to Zn2+ is 20:1 stress–strain curves during cyclic stretching. It can be seen from the figure that BLA-(Zn2+) also has excellent resilience, and its stress–strain curves still basically overlap during repeated loading–unloading processes.

3.3 Self-healing properties

Through the above test analysis, BLA contains dynamic disulfide bonds and hydrogen bonds, which will lead to excellent self-healing properties. Figure 7B is the optical microscope image of BLA-(Zn2+) before and after self-healing. It can be seen that the cracks generated in BLA-(Zn2+) can be obviously repaired after being treated at 65 °C for 24 h. From the stress–strain curves of BLA and BLA-(Zn2+) at different healing temperatures and times (Figure 7D and E), it can be seen that the healing efficiency increases as the healing temperature and healing time increase. The healing efficiencies of BLA and BLA-(Zn2+) reached only 37 % and 48 %, respectively, at a temperature of 40 °C. When the temperature was raised to 85 °C, the healing efficiencies reached 73 % and 82 %, respectively (Figure 7F). As the temperature increases, the movement of the molecular chain becomes easier, and the material undergoes self-healing through the synergistic effect of dynamic bond breaking and recombination, and as time goes on, more dynamic bonds participate in the reaction, so it has a higher repair rate. Comparing the self-healing efficiency of BLA and BLA-(Zn2+) at different temperatures (Figure 6D), BLA-(Zn2+) has better repairing effect than BLA, which is due to the coordination of metal ions with oxygen atoms as well as disulfide and hydrogen bonds synergy between them.

Figure 7: 
Self-repair mechanism of BLA-(Zn2+). A) Photos of cut BLA-(Zn2+) sample and stretched after healing; B) the optical microscopy images of cut BLA-(Zn2+) sample after healing at 65 °C for 24 h; C) schematic illustration of BLA-(Zn2+) healing process; D) the stress-strain curves of BLA2 after healing at various temperatures for 24 h; E) the stress-strain curves of BLA-(Zn2+) after healing at various temperatures for 24 h; F) healing efficiency of BLA and BLA-(Zn2+) after healing at various temperatures for 24 h.
Figure 7:

Self-repair mechanism of BLA-(Zn2+). A) Photos of cut BLA-(Zn2+) sample and stretched after healing; B) the optical microscopy images of cut BLA-(Zn2+) sample after healing at 65 °C for 24 h; C) schematic illustration of BLA-(Zn2+) healing process; D) the stress-strain curves of BLA2 after healing at various temperatures for 24 h; E) the stress-strain curves of BLA-(Zn2+) after healing at various temperatures for 24 h; F) healing efficiency of BLA and BLA-(Zn2+) after healing at various temperatures for 24 h.

4 Conclusions

Based on the disulfide five-membered ring structure of LA, the self-healing BR was successfully prepared by using a “simulated click” reaction under thermally induced conditions. Dynamic disulfide bonds, hydrogen bonds, and metal ion coordination bonds were simultaneously introduced into commercial BR by simple open-kneading blending and hot pressing. The obtained rubber has good self-healing properties and is regulated by the content of LA and Zn2+. At 80 °C, the self-healing efficiency of BLA-(Zn2+) can reach 82 %. The introduction of Zn2+-O coordination bonds helps to enhance the stiffness and rearrangement of the material at higher temperatures, thereby improving the mechanical properties without compromising its self-healing ability. The preparation method is simple and easy to implement, and is different from the traditional method of preparing self-healing rubber. It can be obtained without a complex synthetic preparation process, and is more in line with the traditional rubber processing method. The obtained material has high toughness, excellent tensile properties, and good self-healing performance.


Corresponding authors: Yaozhu Tian and Zhu Luo, College of Materials and Metallurgy, Guizhou University, Guiyang 550025, China, E-mail: ,

Funding source: Guizhou Province Science and Technology Plan Project

Award Identifier / Grant number: Qiankehe Support [2018] 2184

  1. Research ethics: The local Institutional Review Board deemed the study exempt from review.

  2. Author contributions: The author has accepted responsibility for the entire content of this manuscript and approved its submission.

  3. Competing interests: The authors declare that they have no conflict of interest.

  4. Research funding: This work was supported by Guizhou Province Science and Technology Plan Project (Qiankehe Support [2018] 2184).

  5. Data availability: The data that support the findings of this study are available from the corresponding author upon reasonable request.

References

1. Hager, M. D., Greil, P., Leyens, C., van der Zwaag, S., Schubert, U. S. Self-healing materials. Adv. Mater. 2010, 22, 5424–5430; https://doi.org/10.1002/adma.201003036.Search in Google Scholar PubMed

2. Yang, Y., Urban, M. W. Self-healing polymeric materials. Chem. Soc. Rev. 2013, 42, 7446–7467; https://doi.org/10.1039/c3cs60109a.Search in Google Scholar PubMed

3. Bekas, D. G., Tsirka, K., Baltzis, D., Paipetis, A. S. Self-healing materials: a review of advances in materials, evaluation, characterization and monitoring techniques. Compos. B. 2016, 87, 92–119; https://doi.org/10.1016/j.compositesb.2015.09.057.Search in Google Scholar

4. Dong, Y. W., Meure, S., Solomon, D. Self-healing polymeric materials: a review of recent developments. Prog. Polym. Sci. 2008, 33, 479–522; https://doi.org/10.1016/j.progpolymsci.2008.02.001.Search in Google Scholar

5. Tan, Y. J., Wu, J., Li, H., Tee, B. C. K. Self-healing electronic materials for a smart and sustainable future. ACS Appl. Mater. Interfaces 2018, 10, 15331–15345; https://doi.org/10.1021/acsami.7b19511.Search in Google Scholar PubMed

6. Yanagisawa, Y., Nan, Y., Okuro, K., Aida, T. Mechanically robust, readily repairable polymers via tailored noncovalent cross-linking. Science 2018, 359, 72–76; https://doi.org/10.1126/science.aam7588.Search in Google Scholar PubMed

7. Zhang, H., Wang, D., Liu, W., Li, P., Liu, J., Liu, C., Zhang, J., Zhao, N., Xu, J. Recyclable polybutadiene elastomer based on dynamic imine bond. J. Polym. Sci. A: Polym. Chem. 2017, 55, 2011–2018; https://doi.org/10.1002/pola.28577.Search in Google Scholar

8. Bai, J., Shi, Z. Dynamically crosslinked elastomer hybrids with light-induced rapid and efficient self-healing ability and reprogrammable shape memory behavior. ACS Appl. Mater. Interfaces 2017, 9, 27213–27222; https://doi.org/10.1021/acsami.7b06407.Search in Google Scholar PubMed

9. Bai, J., Li, H., Shi, Z., Tian, M., Yin, J. Dynamic crosslinked poly(styrene-block-butadiene-block-styrene) via diels–alder chemistry: an ideal method to improve solvent resistance and mechanical properties without losing its thermal plastic behavior. RSC Adv. 2015, 5, 45376–45383; https://doi.org/10.1039/c5ra08719h.Search in Google Scholar

10. Cai, G., Wang, J., Qian, K., Chen, J., Li, S., Lee, P. S. Extremely stretchable strain sensors based on conductive self‐healing dynamic cross‐links hydrogels for human‐motion detection. Adv. Sci. 2017, 4, 1600190; https://doi.org/10.1002/advs.201600190.Search in Google Scholar PubMed PubMed Central

11. Liu, X., Su, G., Guo, Q., Lu, C., Zhou, T., Zhou, C., Zhang, X. Hierarchically structured self-healing sensors with tunable positive/negative piezoresistivity. Adv. Funct. Mater. 2018, 28, 1706658; https://doi.org/10.1002/adfm.201706658.Search in Google Scholar

12. Li, Y., Li, J., Zhao, X., Yan, Q., Gao, Y., Hao, J., Hu, J., Ju, Y. Triterpenoid‐based self‐healing supramolecular polymer hydrogels formed by host–guest interactions. Chemistry 2016, 22, 18435–18441; https://doi.org/10.1002/chem.201603753.Search in Google Scholar PubMed

13. Liu, Y. L., Chuo, T. W. Self-healing polymers based on thermally reversible diels–alder chemistry. Polym. Chem. 2013, 4, 2194; https://doi.org/10.1039/c2py20957h.Search in Google Scholar

14. Watanabe, M., Yoshie, N. Synthesis and properties of readily recyclable polymers from bisfuranic terminated poly(ethylene adipate) and multi-maleimide linkers. Polymer 2006, 47, 4946–4952; https://doi.org/10.1016/j.polymer.2006.05.036.Search in Google Scholar

15. Pettignano, A., Grijalvo, S., Häring, M., Eritja, R., Tanchoux, N., Quignard, F., Díaz Díaz, D. Boronic acid-modified alginate enables direct formation of injectable, self-healing and multistimuli-responsive hydrogels. Chem. Commun. 2017, 53, 3350–3353; https://doi.org/10.1039/c7cc00765e.Search in Google Scholar PubMed

16. Wen, J. Y., Xi, T., Zhao, T., Weng, L., Wang, L. Antifouling and antibacterial hydrogel coatings with self-healing properties based on a dynamic disulfide exchange reaction. Polym. Chem. 2015, 6, 7027–7035; https://doi.org/10.1039/c5py00936g.Search in Google Scholar

17. Huang, X., Bolen, M. J., Zacharia, N. S. Silver nanoparticle aided self-healing of polyelectrolyte multilayers. Phys. Chem. Chem. Phys. 2014, 16, 10267–10273; https://doi.org/10.1039/c4cp00349g.Search in Google Scholar PubMed

18. Sun, C., Jia, H., Lei, K., Zhu, D., Gao, Y., Zheng, Z., Wang, X. Self-healing hydrogels with stimuli responsiveness based on acylhydrazone bonds. Polymer 2019, 160, 246–253; https://doi.org/10.1016/j.polymer.2018.11.051.Search in Google Scholar

19. Wang, D., Zhang, H., Cheng, B., Qian, Z., Liu, W., Zhao, N., Xu, J. Dynamic cross‐links to facilitate recyclable polybutadiene elastomer with excellent toughness and stretchability. J. Polym. Sci. A: Polym. Chem. 2016, 54, 1357–1366; https://doi.org/10.1002/pola.27983.Search in Google Scholar

20. Guo, W., Chen, Y., Tang, Z., Zhang, X., Liu, Y., Wu, S., Guo, B. Synthesis and properties of self-repairing and remolding polybutadiene rubber based on metal coordination bonds. J. Chem. Educ. 2020, 41, 2090–2098.Search in Google Scholar

21. Bai, J., Li, H., Shi, Z., Yin, J. An eco-friendly scheme for the cross-linked polybutadiene elastomer via thiol–ene and diels–alder click chemistry. Macromolecules 2015, 48, 150515121429004; https://doi.org/10.1021/acs.macromol.5b00389.Search in Google Scholar

22. Endo, K., Yamanaka, T. Copolymerization of lipoic acid with 1,2-dithiane and characterization of the copolymer as an interlocked cyclic polymer. Macromolecules 2006, 39, 4038–4043; https://doi.org/10.1021/ma060063n.Search in Google Scholar

23. Zhang, X., Waymouth, R. M. 1,2-dithiolane-derived dynamic, covalent materials: cooperative self-assembly and reversible cross-linking. J. Am. Chem. Soc. 2017, 139, 3822–3833; https://doi.org/10.1021/jacs.7b00039.Search in Google Scholar PubMed

24. Liu, Y., Yuan, J., Zhang, K., Guo, K., Gao, C., Wu, Y. A novel type of self-healing silicone elastomers with reversible cross-linked network based on the disulfide, hydrogen and metal-ligand bonds. Prog. Org. Coat. 2020, 144, 105661; https://doi.org/10.1016/j.porgcoat.2020.105661.Search in Google Scholar

25. Huang, B., Liu, X., Tan, L., Cui, Z., Yang, X., Jing, D., Zheng, D., Li, Z., Liang, Y., Zhu, S., Yeung, K. W. K., Wang, X., Zheng, Y., Wu, S. “Imitative” click chemistry to form a sticking xerogel for the portable therapy of bacteria-infected wounds. Biomater. Sci. 2019, 7, 5383–5387. https://doi.org/10.1039/c9bm01417a.Search in Google Scholar PubMed

26. Flory, P. J., Rehner, J. Statistical mechanics of cross‐linked polymer networks I. Rubberlike elasticity. J. Chem. Phys. 1943, 11, 512–520; https://doi.org/10.1063/1.1723791.Search in Google Scholar

27. Flory, P. J. Statistical mechanics of swelling of network structures. J. Chem. Phys. 1950, 18, 108–111; https://doi.org/10.1063/1.1747424.Search in Google Scholar

28. Feng, Z., Hu, J., Yu, B., Tian, H., Zuo, H., Ning, N., Tian, M., Zhang, L. Environmentally friendly method to prepare thermo-reversible, self-healable biobased elastomers by one-step melt processing. ACS Appl. Polym. Mater. 2019, 1, 169–177; https://doi.org/10.1021/acsapm.8b00040.Search in Google Scholar

29. Sarkar, S., Banerjee, S. L., Singha, N. K. Dual‐responsive self‐healable carboxylated acrylonitrile butadiene rubber based on dynamic diels–alder “click chemistry” and disulfide metathesis reaction. Macromol. Mater. Eng. 2021, 306, 2000626; https://doi.org/10.1002/mame.202000626.Search in Google Scholar

30. Qi, Z., Chen-Yu, S., Da-Hui, Q., Yi-Tao, L., Feringa, B. L., He, T. Exploring a naturally tailored small molecule for stretchable, self-healing, and adhesive supramolecular polymers. Sci. Adv. 2018, 4, eaat8192, https://doi.org/10.1126/sciadv.aat8192.Search in Google Scholar PubMed PubMed Central


Supplementary Material

This article contains supplementary material (https://doi.org/10.1515/polyeng-2023-0036).


Received: 2023-02-09
Accepted: 2023-06-23
Published Online: 2023-09-04
Published in Print: 2023-10-26

© 2023 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 10.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/polyeng-2023-0036/html
Scroll to top button