Open Access Article
Laihui Xiao
ab,
Zengshe Liuc,
Fangfang Huab,
Yigang Wanga,
Jinrui Huang
a,
Jie Chena and
Xiaoan Nie*a
aInstitute of Chemical Industry of Forestry Products, CAF, National Engineering Lab for Biomass Chemical Utilization, Key Lab on Forest Chemical Engineering, SFA, Key Lab of Biomass Energy and Material, Nanjing, Jiangsu 210042, PR China. E-mail: niexiaoan@126.com; Tel: +86-025-8548-2528
bCo-Innovation Center of Efficient Processing and Utilization of Forest Resources, Nanjing Forestry University, China
cUSDA, ARS, National Center for Agricultural Utilization Research, Bio-Oils Research Unit, 1815 N University St, Peoria, IL 61604, USA
First published on 19th August 2019
In this study, a modifier (CTMA) prepared by emulsion copolymerization of tung oil fatty acid, methyl esters of tung oil fatty acid and acrylonitrile was used to toughen epoxy resins. The structural characterization of the copolymer was carried out by Fourier transform infrared spectroscopy, 1H NMR spectroscopy and high-temperature gel permeation chromatography. Mechanical testing, thermal characterization and scanning electron microscopy were conducted to investigate the properties of epoxy resin modified by the copolymer and further reveal its toughening mechanism. The results indicated that the newly synthesized copolymer effectively toughened the epoxy resin because the elongation-at-break was increased to 89.48%, the maximum toughness calculated by work before break was nearly 4.6 times that of the neat epoxy resin, and apparent shear yields and plastic deformations were observed in the morphology of the fractured surfaces. CTMA, which acts as a flexible cross-linker in the epoxy thermoset, may decrease the cross-linking density.
Many kinds of modified resins, such as rubbers, thermoplastic particles, core–shell polymers and hyperbranched polymers, have been used to improve the flexibility of epoxy resins. Acting as stress concentrators, the modifiers in epoxy matrices terminate cracks by different mechanisms such as multiple crazing mechanism, multiple shear yielding mechanism, multiple crazing with multiple shear yielding mechanism and cavitation mechanism.2–5 However, because of the reaction-induced mechanism, the generation of phase separation depends on the choice of epoxy resin, curing agent and processing conditions. Therefore, a curing system toughened by modified resins will not always induce phase separation.6–9
Among the rubbers used as modifiers, the carboxyl-terminated copolymer of acrylonitrile and butadiene (CTBN) has been extensively used. The chemical reaction between carboxyl and epoxy groups is able to improve the compatibility between CTBN and epoxy resin, and the long flexible chains may toughen epoxy matrices by developing stress concentrations under stress.10–12 Laura et al. have studied CTBN-modified epoxy resin by laser confocal microscopy characterization and found that the fracture toughness increases, whereas the glass transition temperature and the tensile modulus decrease as the size and phase volume of CTBN increase;13 Huang et al. have prepared a flexible and transparent epoxy resin with tunable mechanical properties by the reaction between hydrazine hydrate and the epoxy functional group diglycidyl ether of bisphenol A (DGEBA). The curing systems incorporated with the modified resin demonstrated excellent flexibility; however, no phase separation was observed.14 Day et al. have toughened epoxy resin with a series of multilayer core–shell particles. Compared to acrylic toughening particles (ATP) and CTBN under the same conditions, the epoxy resin modified by these multilayer core–shell particles improved the toughness, which led to a significant decrease in Young's modulus.15 Fei et al. used hyperbranched polymers to toughen the epoxy resin that were synthesized from the diglycidyl ether of bisphenol A, succinic acid (SA) and 1,3,5-benzenetricarboxylic acid (BA). After being cured by methylhexahydrophthalic anhydride (MeHHPA), the modified epoxy thermosets exhibited better toughness as the elongation-at-break of 5.25% could be achieved as compared to 3% of the neat epoxy thermoset; in addition, the glass transition temperatures of the modified epoxy thermosets increased because of their hyperbranched structure.8
However, due to the growing concerns over sustainability and environmental friendliness, the use of renewable materials is necessary; therefore, traditional methods of toughening the epoxy resin by petroleum-based modified resins are being gradually replaced by those involving the use of renewable materials. Thus, the synthesis of modifiers partly or completely from renewable feedstock is urgent.
Vegetable oils (VOs) are excellent alternatives due to their abundance, low price and biodegradability.16,17 Many studies have been devoted towards the toughening of epoxy resin by epoxidized vegetable oils.18,19 These epoxidized vegetable oils are incorporated into epoxy thermosets as a flexible cross-linker, which toughen the epoxy resin by decreasing the cross-linking density. Ratna et al. toughened DGEBA by epoxidized soybean oil using an ambient-temperature hardener. When the addition amount of the modifier was increased from 5% to 30%, the maximum elongation-at-break obtained was 7.5% (which was 3.2% for the sample without the modifier); this led to a significant improvement in toughness;20 however, it still could not satisfy the requirement for application in certain special occasions.
Moreover, the conjugated structure in tung oil obtained via the Diels–Alder reaction rather than epoxidization has received significant research interest. By the Diels–Alder reaction, a number of tung oil-derived toughened epoxy resins have been produced.21–24 As is well-known, the products obtained via the Diels–Alder reaction contain a ring structure, which may reduce the toughening effect. Hence, to avoid the formation of a ring structure, an alternative strategy is to synthesize the tung oil-based toughener via radical polymerization; however, only few studies have been reported on the synthesis of an epoxy resin toughener by the radical polymerization of tung oil or tung oil-derives.
In this study, a renewable toughener was synthesized by the emulsion polymerization of tung oil fatty acid (TOFA), methyl esters of tung oil fatty acid (METOFA) and acrylonitrile. The long flexible aliphatic chains of tung oil-derives were able to toughen the rigid epoxy matrix, and the carboxyl groups of TOFA improved the compatibility of blends. Then, to study the effect of the addition content of the synthesized resin on the properties of the modified epoxy resin, the synthesized resin in different contents was incorporated into the epoxy matrix by a pre-reaction between the carboxyl groups and epoxy groups. After curing, the mechanical properties and thermal characterizations were investigated by mechanical testing, differential scanning calorimetry (DSC), dynamic mechanical analysis (DMA) and thermogravimetric analysis (TGA). In addition, the fracture surfaces were observed by a scanning electron microscope (SEM). The results of this study show that the synthesized copolymer resin can effectively toughen the epoxy resin.
:
1 v/v) were taken in a flask, followed by refluxing for 3 h at 75 °C. After decomposing the soaps by hydrochloric acid, the water layer was removed. The crude product was rinsed with water, and the trace amount of water was removed by a rotary evaporator under vacuum at 70 °C. Then, 161 g of a light brown product was obtained, whose acid value was 200.3 mg g−1 (theoretical: 201.8 mg g−1).The preparation method is shown in Fig. 1.
:
25, 100
:
35, and 100
:
45) were charged into a flask. After catalysis by triphenylphosphine (0.2 wt% of epoxy resin), the mixture was stirred at 130 °C for 3 h. After the product was cooled down to room temperature, polyetherimide was added (epoxy and curing agent were in the 1
:
1 equivalent ratio). The mixture was then charged into a stainless mold preheated at 80 °C. Curing was performed at 40 °C for 24 h and at 80 °C for 3 h. These samples were then coded as E-CTMA25, E-CTMA35 and E-CTMA45. Epoxidized soybean oil (ESO) was blended with epoxy resin in the mass ratio of 100
:
45 at room temperature, and the sample was coded as E-ESO45, which was then used for comparison purposes. Neat epoxy resin (EP) was also prepared in the same way as a control group.
C–H appeared at 3012 cm−1 as compared to that of tung oil-derives, which indicated the polymerization of the double bond. Furthermore, the peaks of –COOCH3 at 1739 cm−1, –COOH at 1707 cm−1 and –CN at 2228 cm−1 were observed in the CTMA spectrum, indicating that CTMA contained these three monomer units. Due to the reaction between the carboxyl groups and the epoxy groups, the peak of –COOH at 1707 cm−1 disappeared in the spectrum of E-CTMA45.
Fig. 3 shows the GPC spectrum of CTMA. Only one peak was observed in the spectrum, which confirmed that the main product was the copolymer of the three monomer units. The number-average molecular weight is about 28
740 g mol−1, and the Đ value is 5.43. The high Đ value may be attributed to the different monomer activities of acrylonitrile and tung oil-derives. Furthermore, the simultaneous addition of monomers contributed to high Đ value.
Fig. 4 and 5 show the 1H NMR spectra of tung oil-derives and CTMA, which have been used to determine the copolymer composition. In the 1H NMR spectrum of CTMA, the proton resonances of –COOCH3 (3.69 ppm) in METOFA and –CH3 (0.91 ppm) in TOFA and METOFA are well separated, whereas those for methylene protons and methine protons (1–2.5 ppm) are overlapped. Hence, their relative peak areas can be used to calculate the mole fraction of the monomer by the following formulas,
![]() | (1) |
![]() | (2) |
![]() | (3) |
| Part | A | B | C |
|---|---|---|---|
| Peak area (relative to C) | 1.75 | 14.28 | 1 |
| Monomer | AN | METOFA | TOFA |
| Mole fraction (%) | 45.26 | 31.28 | 24.06 |
However, no melting or crystallization peaks were observed in the curves of E-CTMA, and only one glass transition temperature could be detected, which indicated that no phase separation occurred. This was because the epoxy resin incorporated CTMA by the reaction between the carboxyl groups and the epoxy groups; this contributed to the molecular-level compatibility.
As flexible cross-linkers, many long flexible aliphatic chains were incorporated into E-CTMA thermosets. Segment movement in flexible aliphatic chains can be easily induced by increasing the temperature, which will contribute to a lower glass transition temperature; furthermore, the high cross-linking network system will be destroyed, and the less restriction from covalent bonds contributes to segment movement, which will also contribute to a lower glass transition temperature. All these results are in accordance with the decreasing Tg value shown in Fig. 6(b); moreover, further explanation about this point has been provided in the discussion of DMA.
Furthermore, the curing behavior was studied by DSC. In Fig. 7(a), all samples show a similar curing behavior, and only one curing process has been determined. The accurate parameters of the curing process are presented in Table 2. Obviously, the To and Tp values of E-CTMAs are significantly lower than that of the neat epoxy resin, and a decreasing tendency is observed as the content of CTMA is increased. Previous studies have demonstrated that the existence of strong hydrogen bonding species, such as acids and alcohols, is able to accelerate the epoxy–amine reaction.25 In the blending process, the carboxyl groups in CTMA react with the epoxy groups, and the same amount of hydroxyls are simultaneously produced. Therefore, the curing process is accelerated, which is in agreement with the decreased values of To and Tp. Furthermore, this positive effect on the curing process is shown in Fig. 7(b) in which it is observed that CTMA enhances the conversion at a specific temperature; however, higher content of CTMA results in lower enthalpy (ΔH). This is because the addition of CTMA reduces the content of epoxy groups and curing agent; this leads to fewer reactions. Moreover, the incorporation of CTMA increases the viscosity of epoxy resin (Table 2), which is attributed to the much higher molecular weight of CTMA. Furthermore, the pre-reaction between the carboxyl and the epoxy groups may produce hydroxyls, which may form hydrogen bonds with other polar groups and further increase the viscosity. Similar results have been obtained in previous studies, in which rubber and other oil-based polymers have been used to toughen the epoxy resin.12,21,26 The increased viscosity may restrict the movement of reaction groups and then reduce the enthalpy.
![]() | ||
| Fig. 7 DSC curves of the curing process: heat flow against temperature (a) and conversion against temperature (b). | ||
| Sample | Toa (°C) | Tpb (°C) | ΔH (J g−1) | Viscosityc (kPa s) |
|---|---|---|---|---|
| a To is the initial curing temperature.b Tp is the temperature at which the curing rate is maximum.c Viscosity is measured at room temperature. | ||||
| EP | 89.17 | 126.18 | 379.65 | 12.36 |
| E-CTMA25 | 78.37 | 119.43 | 297.89 | 19.16 |
| E-CTMA35 | 77.89 | 119.25 | 292.83 | 27.08 |
| E-CTMA45 | 74.43 | 116.78 | 244.34 | 51.22 |
δ of the neat epoxy resin and E-CTMA thermosets. The accurate statistics of thermal properties are presented in Table 3.
Tg decreases with an increase in the CTMA content; this is in accordance with the abovementioned DSC results provided in the DSC section. Furthermore, the explanation of the decreasing Tg can be confirmed by the results of half-peak width (HPW) in tan
δ curves, which is related to structural uniformity.27 The neat epoxy resin has the lowest HPW value, whereas E-CTMA45 has the highest HPW value of E-CTMA. This is because CTMA acts as a flexible crosslinker, which increases the molecular weight between cross-linking points. Furthermore, the random carboxyl distribution of CTMA extends the dispersity of molecular weight between cross-linking points. However, every tan
δ curve of E-CTMAs and neat epoxy resin has only one peak, which illustrates that the uniform network structure is unable to lead to phase separation.28,29
Based on the theory of rubber elasticity, the cross-linking density ve can be calculated from the storage modulus in the rubbery region:
![]() | (4) |
The value of ve decreases with the increasing content of CTMA; this indicates lower cross-linking density. With fewer cross-linking points, chain segments move more easily; this contributes to a lower glass transition temperature. Thus, the decreasing glass transition temperatures are presented as Tg value in Table 3.
![]() | ||
| Fig. 9 SEM images of the frozen fracture morphology for the thermosets: neat epoxy, E-CTMA25, E-CTMA35 and E-CTMA45. | ||
When CTMA was added as a modifier, the carboxyl groups reacted with the epoxy groups; this led to molecular level incorporation instead of phase separation. This phenomenon has also been observed by other researchers.31–34 Because of the long aliphatic chains of CTMA, external stresses can be redistributed, and plastic deformations will be induced by the flexible cross-linker. This is the reason why the fracture surfaces become rougher.
| Sample code | T5%a (°C) | T10%a (°C) | Tmb (°C) |
|---|---|---|---|
| a T5% and T10% are the temperatures of 5% and 10% weight loss, respectively.b Tm is the temperature at which the degradation rate is maximum. | |||
| EP | 372.06 | 378.92 | 392.7 |
| E-CTMA25 | 286.11 | 368.04 | 394.0 |
| E-CTMA35 | 285.91 | 365.29 | 400.2 |
| E-CTMA45 | 279.49 | 355.32 | 393.8 |
| CTMA | 204.94 | 227.84 | 263.9 |
The sample E-ESO45 was used for comparison. Obviously, although E-ESO45 has higher tensile strength and Young's modulus, its toughness value is much lower than that of E-CTMA45; this is mainly attributed to its low elongation-at-break. Moreover, many studies on the toughening of epoxy resin with ESO have been reported, which indicate that ESO shows limited improvement in the elongation-at-break.20,35,38,39 Hence, in terms of elongation-at-break and toughness value, CTMA exhibits excellent toughening ability, which is better than ESO. Moreover, the mechanical properties suggest that the epoxy resin modified by CTMA is a potential material that can be used as a coating, pouring sealant and other casting material and has higher requirement for toughness but less demand for strength.44 However, owing to the higher viscosity of some formulations, some solvents should be used to reduce the viscosity for applications.
Footnote |
| † Mention of trade names or commercial products in this publication is solely for the purpose of providing specific information and does not imply recommendation or endorsement by the U.S. Department of Agriculture. USDA is an equal opportunity provider and employer. |
| This journal is © The Royal Society of Chemistry 2019 |