Biobased Thermosets Prepared from Rigid Isosorbide and Flexible

Nov 7, 2016 - *Phone: 86-591-83726495. ... Component for Tailoring Biobased Unsaturated Polyester Thermosets for a Wide Structure–Property Window ...
0 downloads 0 Views 2MB Size
Research Article pubs.acs.org/journal/ascecg

Biobased Thermosets Prepared from Rigid Isosorbide and Flexible Soybean Oil Derivatives Wendi Liu,† Tianshun Xie,† and Renhui Qiu*,‡ †

College of Material Engineering and ‡College of Transportation and Civil Engineering, Fujian Agriculture and Forestry University, 15 Shangxiadian Road, Jianshan, Fuzhou, Fujian 350002, P. R. China

Downloaded via TUFTS UNIV on July 17, 2018 at 19:44:54 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

S Supporting Information *

ABSTRACT: A rigid monomer, isosorbide-methacrylate (IM), was synthesized from isosorbide with methacrylate anhydride (MAA) via a solvent-free, ultrasonic-assisted method and then was used to copolymerize with acrylated epoxidized soybean oil (AESO) to formulate a biobased thermosetting resin (IM-AESO). The synthesis of IM was monitored by attenuated total reflectance Fourier transform infrared (ATR-FTIR) by tracking the changes in the functional groups of the reaction system. The AESO was further modified with MAA to replace the hydroxyl groups of AESO with methacrylate groups, generating a resin (IM-MAESO) with an improved degree of unsaturation. The chemical structure of IM and modification of AESO with MAA were characterized using 1H NMR, 13C NMR, and ATR-FTIR analyses. The miscibility of IM with AESO was predicted according to Hansen’s solubility theory and evaluated experimentally. The formulated IM-AESO and IM-MAESO resins were compared with the pure AESO and IM resins in terms of their rheological behaviors, curing kinetic characteristics, flexural properties, dynamic mechanical properties, and thermal stabilities. The results indicated that both the IM-AESO resin and the IM-MAESO resin have much lower viscosities, activation energies, and curing temperatures as well as higher polymerization rates and curing degrees than pure AESO due to the incorporation of IM as a reactive diluent. The combination of stiff IM and flexible AESO results in biobased networks with superior flexural strength, flexural modulus, flexural strain, storage modulus, glass transition temperature, and thermal stability. Furthermore, the MAA modification gives rise to the cross-linking degree and hence stiffness of the IM-MAESO resin as a result of the increase in the unsaturation degree of the MAESO. KEYWORDS: Isosorbide, Biobased reactive diluent, Acrylated epoxidized soybean oil (AESO), Thermosetting resin, Processability, Mechanical properties



INTRODUCTION Thermosetting resins are important matrices in fiber-reinforced composites due to their attractive qualities such as good physical and mechanical properties, thermal and chemical resistances, and processabilities. Currently, most commercially available thermosets are produced from petroleum chemicals; these techniques consume large amounts of oil and gas. Driven by increasing environmental concerns and depleting petroleum resources, there is increasing interest in the exploitation of biobased monomers and polymers from renewable resources.1−3 Vegetable oil, which consists of esters of glycerol with three long-chain fatty acids, is a promising feedstock for producing biobased thermosetting polymers.4,5 Acrylated epoxidized soybean oil (AESO) is derived from soybean oil via epoxidation followed by acrylation. AESO has been commercialized and utilized in the development of thermosetting resins for use in composites.6 However, AESO is highly viscous at room temperature and has a low cross-linking capacity due to its long aliphatic chains and low degree of unsaturation. Therefore, a reactive diluent (RD), i.e., a © 2016 American Chemical Society

comonomer, is highly desired for AESO-based matrices because the RD could facilitate the synthesis of a resin with low viscosity to wet reinforcing fibers well and the formation of a three-dimensional network with high cross-linking density after the resin is cured. AESO-based thermosets with superior processability and physical−mechanical performance were extensively prepared by using styrene as an RD because styrene is miscible with AESO and can cross-link efficiently with it.7−10 However, styrene is listed as an anticipated carcinogen.11 Other vinyl monomers including vinyl toluene, methyl methacrylate, and divinylbenzene were used to replace styrene in maleinated AESO resins for the fabrication of biobased composites with glass/flax fibers.12 Hybridization of AESO with petroleum-based polymers including poly(methyl methacrylate),13 vinyl ester, and vinyl ester−urethane14,15 could facilitate the formulation of Received: September 2, 2016 Revised: October 18, 2016 Published: November 7, 2016 774

DOI: 10.1021/acssuschemeng.6b02117 ACS Sustainable Chem. Eng. 2017, 5, 774−783

ACS Sustainable Chemistry & Engineering



resins with an interpenetrating network structure. However, they are petrochemical-based. To obtain soybean oil-based polymers with high biobased contents, considerable attention has been paid to developing RDs from renewable resources such as fatty acids,16−18 cellulose,19 lignin,19,20 sucrose,21 rosin acid,22 and itaconic acid.23 However, there are two drawbacks in impeding the application of these thermosets in fiber-reinforced composites: (1) Some of these renewable resources including lignin and vegetable oil have high molecular weights, which gives the resulting RDs high viscosities and therefore negatively affects the processability of the resulting AESO resins. The high viscosities of the resins also reduce the polymerization efficiencies of RDs with AESO, thus generating cured resins with low cross-linking densities. For instance, methacrylated vanillin is a solid at room temperature, and both methacrylated guaiacol and eugenol have much higher viscosities (17 and 28 mPa·s) than styrene (0.7 mPa·s) at room temperature.24 (2) These biobased RDs are most commonly synthesized from renewable materials via esterification with petroleum-based acids or oils, which generally requires the use of catalysts,21,25 organic solvents,26,27 elevated temperatures,23 and/or long reaction times.25,28 These harsh reaction conditions might generate side-reactions at high temperature, produce considerable amounts of solvent waste, have high energy consumption, and introduce residual metal catalysts that are difficult to remove. Isosorbide is one of the top 12 most promising renewable building blocks. It is derived from glucose, which is widely available from the depolymerization of biomass-based materials including cellulose and starch.29 Isosorbide is a chiral diol with a unique bicyclic ring structure, i.e., a V-shaped molecule composed of two cis-connected tetrahydrofuran rings and two secondary hydroxyl (−OH) groups in the 2- and 5- positions.30 These features have motivated numerous works on the preparation of isosorbide-based polymers with superior mechanical and thermal properties.31−34 Isosorbide is crystalline at room temperature and is immiscible with many acids and anhydrides; therefore, during the synthesis of isosorbidebased monomers or polymers without organic solvents, esterification occurs only at the phase boundary between the isosorbide and the acid with a low reaction rate. The present work aimed to synthesize a rigid and biobased monomer, i.e., isosorbide-methacrylate (IM), from isosorbide with methacrylic anhydride (MAA) and then to formulate soybean oil-based thermosets with high performance and renewable contents by using IM as a main structural component. The ultrasonic treatment was used for increasing the reaction rate between isosorbide and MAA. The synthesis process of IM was monitored by attenuated total reflectance Fourier transform infrared (ATR-FTIR) spectroscopy. The obtained IM was characterized by proton (1H) nuclear magnetic resonance (NMR), carbon (13C) NMR, and ATRFTIR spectroscopies as well as solubility evaluation. The IM was blended with AESO to prepare the IM-AESO resin, and the modified IM-AESO (IM-MAESO) resin was obtained by directly blending AESO with the reaction products that resulted from the IM synthesis; the excess MAA in the mixture would react with the −OH groups of AESO, thus increasing the unsaturation degree of the resin. Both the IM-AESO and IMMAESO resins were compared with the resins composed of pure AESO and IM monomers in terms of their rheological and curing behaviors, flexural properties, thermal stabilities, and dynamic mechanical properties.

Research Article

EXPERIMENTAL SECTION

Materials. AESO (average molecular weight 1200 g/mol; viscosity (25 °C) 18 000−32 000 mPa·s; acid value ≤10 mg KOH/g; inhibitor 3500−4500 ppm monomethyl ether hydroquinone), MAA (94%, inhibitor 2000 ppm topanol A), 4-(dimethylamino)pyridine (DMAP, 99%), tert-butyl peroxybenzoate (TBPB, 98%), magnesium sulfate (MgSO4), sodium bicarbonate (NaHCO3), and sodium hydroxide (NaOH) were purchased from Sigma-Aldrich. Isosorbide (98%) was purchased from Fisher Scientific. All chemicals were used as received. Synthesis of Isosorbide-Methacrylate. Isosorbide (14.61 g, 0.1 mol) and DMAP (0.61 g, 0.05 mol) were suspended in MAA (46.25 g, 0.3 mol) in a 150 mL round-bottom flask; the molar ratio of MAA to isosorbide was 3:1. A molar ratio of MAA to isosorbide of 2:1 was also used for the comparison of reaction production yield. The flask was placed in an ultrasonic water bath (ultrasound power 80 W) and then stirred magnetically at a rate of 500 rpm at 60 °C for 12 h. Crystalline isosorbide was first suspended in MAA and then gradually dissolved into the liquid phase as the reaction progressed. After the reaction, the flask was cooled to room temperature, and the resulting mixture was sequentially washed with 0.5 mol/L NaOH, saturated NaHCO3 and water, dried over MgSO4, and concentrated under reduced pressure to generate a pale yellow oil, i.e., isosorbide-methacrylate (IM). The reaction between isosorbide and MAA is given in Scheme 1. The 1H

Scheme 1. Synthesis of IM from Isosorbide and MAA

NMR, 13C NMR, and FTIR spectra of the obtained IM are provided in the Supporting Information (SI1 and SI2). The IM yields were 92.3% and 66.5% for the 3:1 and 2:1 MAA/isosorbide systems, respectively. Preparation and Modification of Soybean Oil-Based Resins. AESO was mixed with IM at a weight ratio of 1:1 in a beaker at 70 °C using a magnetic stirrer at 500 rpm for 10 min. The resulting mixture was cooled to room temperature and designated IM-AESO. To increase the unsaturation degree and hence the cross-linking sites of the AESO-based resins, the MAA-isosorbide reaction product from Synthesis of Isosorbide-Methacrylate without purification (MAA:isosorbide = 3:1) was directly mixed with AESO to prepare the MAA-modified AESO resin (IM-MAESO). The mixture contains the generated IM and unreacted MAA, and their weights are theoretically 28.23 g (0.1 mol) and 15.42 g (0.1 mol), respectively. To keep the weight ratio of AESO to IM at 1:1, AESO (28.23 g, 235.25 mmol) was added to the mixture, where the molar ratio of MAA to AESO was 4.26:1. The flask was placed in a water bath (without ultrasound), and the mixture continued to react at 60 °C for 6 h while being stirred at 500 rpm. With DMAP catalysis, the anhydride groups of MAA would react with the −OH groups of AESO to induce the grafting of methacrylic groups onto AESO molecules through ester linkages (Scheme 2), which are evidenced by FTIR and 13 C NMR analyses (Supporting Information, SI3). After the reaction was finished, the mixture was purified using the same procedure as in Synthesis of Isosorbide-Methacrylate, and thus, the IM-MAESO resin containing IM, unmodified AESO, and modified AESO was obtained. Polymerization of Biobased Thermosetting Resins. Four kinds of resin, i.e., pure AESO, IM-AESO, IM-MAESO, and pure IM, were blended with 2 wt % TBPB (based on the total mixture) and then stirred for 2 min, respectively. The resulting mixtures were degassed under vacuum and transferred to silicon molds to prepare the cured resin samples. The curing reaction was performed at 120 °C for 2 h, followed by 160 °C for another 4 h in an oven. After being cooled 775

DOI: 10.1021/acssuschemeng.6b02117 ACS Sustainable Chem. Eng. 2017, 5, 774−783

Research Article

ACS Sustainable Chemistry & Engineering

Characterization. The chemical structures of the IM monomer and the IM-AESO and IM-MAESO resins were characterized by ATRFTIR, 1H NMR, and 13C NMR analyses. During the reaction of isosorbide with MAA, aliquots of the reaction mixture were collected at intervals and used instantly for FTIR tests. The FTIR tests were conducted on a PerkinElmer Spectrum One Spectrometer (PerkinElmer, USA) equipped with a 3× bounce diamond crystal and an incident angle of 45°. Spectra were collected under the following conditions: 4000−650 cm−1 range; 4 cm−1 resolution; 16 scans. 1H NMR and 13C NMR spectra were recorded on a JEOL 600 MHz NMR spectrometer (JEOL, Japan) with CDCl3 as a solvent. The processabilities of the resins (pure AESO, IM-AESO, IMMAESO, and pure IM) were investigated based on their rheological and curing behaviors. The rheological behavior tests were carried out on an HAAKE MARS III rotational rheometer (Thermo Electron, USA) using a PP35Ti parallel plate (gap 0.105 mm) at a shear rate of 10 γ/s and a heating rate of 10 °C/min from 25 to 100 °C. The curing kinetics were obtained from nonisothermal differential scanning calorimetry (DSC) scans at four different heating rates (5, 10, 15, and 20 °C/min) from 25 to 250 °C. The scans were conducted on an STA 449 F3 Jupiter Simultaneous Thermal Analyzer (NETZSCH, Germany), and the sample (5−10 mg) was placed in a standard porcelain crucible with a lid and tested under N2 (flow rate: 30 mL/ min) protection. The flexural properties of the four cured resins were evaluated in compliance with ASTM D 790-10. Ten rectangular specimens (80 × 10 mm2) were tested using a CMT6104 microcomputer controlled electronic universal testing machine (MTS Systems, USA) at a crosshead rate of 10 mm/min. Dynamic mechanical analysis (DMA) of the resins was conducted with rectangular samples (55 × 10 mm2) using a DMA 242 (NETZSCH, Germany) machine. The measurements ranged from 25 to 250 °C under single cantilever bending mode. The heating rate was 5 °C/min, and the frequency was 1 Hz. Thermogravimetric analysis (TGA) of the resins was performed on the same testing equipment as the DSC analysis at a heating rate of 10 °C/ min from 25 to 600 °C.

Scheme 2. Proposed Reaction during the Modification of IM-AESO

to room temperature, the samples were removed from the mold for testing. Evaluation of the Solubility of IM. The solubility of IM was measured by mixing IM with ethanol, furan, styrene, and N-vinyl-2pyrrolidone (NVP), respectively. A predetermined 5 g of IM was added into a vial containing 5 g of the respective solvent. The mixture was stirred for 5 min by a glass rod and then kept at room temperature for 24 h. Unsaturated polyester (UPE, 5 g), synthesized from propylene glycol, isophthalic acid, and fumaric acid,35 was also dissolved in IM (5 g) in a beaker at 70 °C. The results of the dissolution tests were recorded with a digital camera and divided into two categories, i.e., soluble and insoluble, based on visual inspection. The miscibilities of IM with water, methacrylic acid, MAA, AESO, and MAESO were confirmed during the formulation of the IM and soybean oil-based resins. According to Hansen’s solubility theory,36,37 a RED value was proposed to evaluate the theoretical solubility of a substance in IM, i.e., IM can dissolve the substance if the RED value is less than 1, whereas it is insoluble when this value is higher than 1.38 A detailed calculation of the solubility parameters and RED values of the substances can be found in the Supporting Information (SI4).



RESULTS AND DISCUSSION ATR-FTIR Monitoring of the IM Synthesis. The reaction of isosorbide with MAA at different molar ratios was monitored via ATR-FTIR to determine the effect of the MAA

Figure 1. Characteristic ATR-FTIR spectra obtained from pure MAA, IM, and the reaction products during the synthesis of IM from isosorbide and MAA (molar ratio = 1:3) after 30, 120, and 300 min. (a) Full spectra and (b) carbonyl region from 1850 to 1575 cm−1 deconvoluted with Gaussian− Lorentzian functions. 776

DOI: 10.1021/acssuschemeng.6b02117 ACS Sustainable Chem. Eng. 2017, 5, 774−783

Research Article

ACS Sustainable Chemistry & Engineering

Figure 2. (a) Yield of IM and (b) conversions of MAA (xMAA) and methacrylate acid (xacid) calculated from the ATR-FTIR spectra of the reaction products at different MAA/isosorbide molar ratios as functions of reaction time.

site first and then with that at C5 site in a later stage of the reaction. These two effects would reduce the reaction rate of esterification significantly. As shown in Figure 2, the reaction system with an MAA/ isosorbide molar ratio of 3:1 shows a slightly higher increased rate in IM yield, decreased rate in xMAA, and generation rate in xacid than the 2:1 system in the initial stage of the reaction. The IM yield of the 3:1 system at equilibrium is close to 100%, i.e., full conversion of isosorbide, which is much higher than that of the 2:1 system (approximately 77%) (Figure 2a). This is in agreement with the IM yields from the experimental results, although the latter (92.3% and 66.5%) is much lower than the calculated values from the FTIR spectra because of the weight loss in the IM purification process. Therefore, an excess MAA system, i.e., 3:1 of MAA/isosorbide, was chosen for the esterification to obtain the IM at a full conversion of isosorbide, where the reaction was almost finished after 300 min. For the conversion of MAA and methacrylic acid, the xMAA and xacid of the 2:1 system (89%) at the final stage are much higher than those of the 3:1 system (78%) because the 3:1 system has a very high and excess MAA concentration (Figure 2b). For both systems, the xMAA is almost equal to the xacid during the reaction process, which confirms the assumption that an anhydride group reacts only with an −OH group of isosorbide to generate an ester and an acid (Scheme 1). This is in line with the reported fact that isosorbide hardly reacts with acids without unique conditions due to the low reactivity of the secondary alcohols at the C2 and C5 sites.28 Solubility of IM Monomer. The predicted and experimental solubility of IM with various substances are presented in Table 1. Pictures of the test results obtained using a digital camera are given in the Supporting Information (SI6). The experimental results are completely consistent with the predicted ones, indicating that IM is miscible with all of the tested substances except water. As one of the precursors for IM, isosorbide exhibits hydrogen bonding between the two secondary −OH groups; therefore, it is easy to dissolve in water but not in MAA due to its high polarity. However, as the esterification between isosorbide and MAA proceeds, the −OH groups on isosorbide are converted into methacrylic groups, producing IM with greatly reduced polarity. Hence, IM can dissolve in most common organic solvents, such as ethanol and furan, but not in water. This is also consistent with the high

concentration on the IM yield and the reaction rate. The FTIR spectra of the reaction mixture (MAA:isosorbide = 3:1) after different reaction intervals and the deconvoluted carbonyl region (1850−1575 cm−1) with Gaussian−Lorentzian functions are given in Figure 1. The following features of the isosorbide− MAA reaction were observed: (1) the band at 1780 cm−1 corresponding to the CO stretch vibration of MAA decreased, and the peak at 1160 cm−1, assigned to the stretch vibration of new C−O−C ester groups, increased with the progress of the reaction; (2) the appearance and growth of a shoulder at 1696 cm−1 is assigned to the formation of carboxylic acid. The esterification of the anhydride groups of MAA with the −OH groups of isosorbide generates ester groups and byproducts, i.e., methacrylic acid (Scheme 1). The characteristic FTIR peaks at 1780, 1696, and 1160 cm−1 of the reaction system were used to monitor the conversion of anhydride, acid, and ester with respect to reaction time. The peak at 1637 cm−1, associated with the CC bonds, was used for the normalization of the absorbance peaks. The detailed calculation methods are provided in the Supporting Information (SI5). The isosorbide−MAA esterification with the assistance of ultrasound can be divided into two stages (Figure 2): (1) For both systems with different MAA/isosorbide molar ratios, a sharp increase in the IM yield is observed during the initial stage of the reaction (reaction time shorter than 60 min), which indicates that the esterification of the −OH groups and anhydride groups is favored in the presence of the ultrasound assistance, resulting in significant consumption of MAA (xMAA) and the generation of methacrylic acid (xacid). In contrast to conventional chemical reactions, acoustic cavitation occurs in the ultrasonic reaction with steps of formation, growth, and implosive collapse of bubbles, which strongly disturbs the phase boundary between the solid isosorbide and the liquid MAA, thereby causing emulsification, enhancing the mass transfer process, and increasing the reaction rate.39,40 (2) After 120 min of reaction, the reaction takes place slowly as indicated by the gradual changes in the IM yield, xMAA and xacid. As the reaction progresses, MAA is continuously consumed, and thus, the reaction is prevented due to the lack of MAA in the system. In addition, the −OH group at the C2 site of isosorbide is much less reactive than that at the C5 site due to the internal hydrogen bonds with the oxygen atom of the ether ring.28 Hence, the MAA tends to react with the −OH groups at the C2 777

DOI: 10.1021/acssuschemeng.6b02117 ACS Sustainable Chem. Eng. 2017, 5, 774−783

Research Article

ACS Sustainable Chemistry & Engineering η = η0exp[Eη /(RT )]

Table 1. Miscibility of IM with Various Substances and Their Predicted Solubility Based on Hansen’s Theory

where η is the apparent viscosity, η0 is the prefactor, Eη is the activation energy, R is the ideal gas constant, and T is the absolute temperature. As presented in Table 2, the neat AESO

Miscibility of IM with Various Substances monomers

RED valuea

predicted

experimental

IM MAA methacrylic acid waterb ethanol furan styrene NVP UPEc AESO MAESO

/ 0.17 0.12 1.87 0.57 0.23 0.51 0.36 0.49 0.18 0.17

/ + + − + + + + + + +

/ + + − + + + + + + +

(1)

Table 2. Calculated Arrhenius Parameters of AESO, IM, IMAESO, and IM-MAESO Resins resins AESO styrene-UPEa IM-AESO IM-MAESO IM a

a

RED values were calculated according to the Hansen’s theory and group contribution method (Supporting Information SI4). bThe value of water was reported by Zhang et al.38 cUPE is synthesized from propylene glycol, isophthalic acid, and fumaric acid.35

η (mPa·s) (30 °C) 4789 ± 69 375 151 ± 1 186 ± 7 12 ± 1

η0 (mPa·s) 2.1 1.0 4.0 3.5 4.4

× × × × ×

−5

10 10−4 10−4 10−4 10−1

Eη (kJ/mol)

R2

48.3 32.2 33.2 34.1 8.5

0.9938 0.9854 0.9624 0.9631 0.9485

Typical UPE resin with 40 wt % styrene as an RD.35

resin has the highest viscosity and Eη value because of the restricted molecular mobility as provided by the long polymer chain and high molecular weight of the AESO monomer. By contrast, the lowest viscosity and Eη value are observed for IM monomer, indicating that IM flows easily and the influence of temperature on its viscosity is insignificant. Therefore, after the incorporation of IM into AESO resins, IM-AESO resin has a considerably decreased viscosity and Eη value because IM is a small molecule, which can alleviate the entanglement of AESO molecular chains and hence increase flowability of the resulting resin. The viscosity of IM-MAESO is 22.6% higher than that of IM-AESO due to the increased molecular weight of MAESO compared to AESO as a result of the grafting of methacrylic groups on AESO; however, this does not significantly influence the Eη values of the two resins. In addition, the viscosities of both IM-AESO and IM-MAESO are much lower than that of a typical UPE resin with 40 wt % styrene as an RD, indicating that the soybean oil-based resins have more superior processability than the traditional UPE resins in the fabrication of fiber-reinforced composites.35 Curing Behaviors of the Resins. The curing processes of the four kinds of resins initiated by TBPB were monitored by DSC scans at a heating rate of 10 °C/min (Figure 4a). The DSC curves of all the resins display two distinct exothermic peaks associated with the free-radical polymerization of the main components in the resin systems. TBPB decomposes slowly without a promoter at low temperature because it is a high temperature initiator, and the thermal decomposition of TBPB starts at 65−100 °C and reaches a peak at 135−165 °C.43,44 Therefore, all of the resins start to polymerize at 90− 105 °C with the generation of free-radicals from the initial decomposition of TBPB, which contributes to the first exothermic peak (TP1) at 115−125 °C in the DSC curves (Figure 4a). However, the polymerization is incomplete, and some CC bonds remain due to the lack of sufficient freeradicals and the confined molecular mobility of the AESO and IM oligomers. Furthermore, the increase in temperature would increase the molecular mobility of the polymer chains and accelerate the decomposition of TBPB, which would further facilitate the polymerization of the resin systems. This explains the second exothermic peak (TP2) in the DSC curves of the resins because the peak temperatures (160−190 °C) are almost higher than the maximum thermal decomposition temperature of TBPB (Table 3). The lower temperature and higher intensity of the peak in the DSC curve indicate the higher reactivity of the compounds

RED value (1.87) of the water−IM pair. The miscibilities of IM with MAA and methacrylic acid are much higher than those with the others due to the lower RED values of the MAA-IM and methacrylic acid−IM systems. Additionally, IM is capable of mixing with both styrene and NVP, and thus, it has the potential for forming copolymers with these two monomers; NVP has been used as an RD for preparing styrene-free UPE and soybean oil-based resins in hemp fiber composites.35,41 The dissolving capacity of IM in UPE is similar to that of styrene because of their close RED values, indicating that IM has the potential to replace styrene in UPE resins. The IM-AESO and IM-MAESO systems possess similar RED values that are less than 0.2, indicating the superior miscibility of IM with unmodified and MAA-modified AESO. Rheological Analysis of the Resins. The main concern of developing soybean oil-based thermosets for fiber-reinforced composites is their viscosities, especially in the liquid molding technique (resin viscosity below 500 mPa·s).42 Due to the improved mobility of polymer chains at higher temperatures, the viscosities of all resins decrease exponentially with increasing temperature (Figure 3), which can be fitted with the Arrhenius equation:16

Figure 3. Change in the viscosity of IM, AESO, IM-AESO, and IMMAESO versus temperature. 778

DOI: 10.1021/acssuschemeng.6b02117 ACS Sustainable Chem. Eng. 2017, 5, 774−783

Research Article

ACS Sustainable Chemistry & Engineering

Figure 4. DSC curves of (a) IM, AESO, IM-AESO, and IM-MAESO systems at a heating rate of 10 °C/min and (b) IM-AESO resin at different heating rates.

Table 3. Characteristic Curing Temperatures Obtained from DSC Curves and Calculated Curing Kinetic Parameters Based on Kissinger’s Theory TP (°C) at different heating rates (°C/min) resins AESO IM IM-AESO IM-MAESO

peaks

5

10

15

20

E (kJ/mol)

ln A

R2

P1 P2 P1 P2 P1 P2 P1 P2

115.7 161.4 116.5 176.4 109.8 163.6 111.1 163.7

124.2 167.8 121.0 183.3 116.6 172.3 117.5 172.1

131.0 175.2 127.0 190.4 122.5 179.1 123.6 178.9

136.0 177.6 128.9 193.9 125.7 183.9 127.8 183.2

83.4 124.6 129.6 127.2 102.0 105.6 98.5 109.0

17.83 26.71 32.52 26.20 24.29 21.10 23.06 22.06

0.9902 0.9664 0.9530 0.9790 0.9899 0.9930 0.9780 0.9931

in the curing reaction.45 Taking the resins cured at a heating rate of 10 °C/min as an example (Figure 4a), the peak height in the DSC curve of pure AESO is much lower than those of other resins, showing a lower polymerization efficiency of AESO resin due to its high viscosity and low unsaturation level. As shown in Table 3, the TP1 of IM is slightly lower than that of AESO, whereas the DSC curve of IM has a higher peak intensity than that of AESO, indicating that IM has a higher reactivity than AESO due to the lower viscosity of IM with respect to AESO. However, the higher reactivity of IM results in an IM oligomer with a higher cross-linking degree, which would restrict the molecular mobility and thus increase the TP2 of IM when compared to that of AESO in the second stage (Table 3). For the blends of IM and AESO, a significantly decreased TP1 is observed for IM-AESO when compared to the pure IM and AESO; however, the TP2 of IM-AESO is between those of pure IM and AESO (Table 3). This is probably because of the conversion of resin curing from homopolymerization to copolymerization. Additionally, the MAA modification for AESO does not significantly affect the TP1 and TP2 but results in an reduced peak intensity in the DSC curve, which is most likely due to the increased molecular weight and thus increased viscosity. The nonisothermal curing kinetics of resins was further investigated by DSC scans with different heating rates. The DSC curves of IM-AESO show that the resin curing characteristic temperatures (TP1 and TP2) move obviously to high temperatures as the heating rate increases due to the thermal lag effect (Figure 4b). Based on Kissinger’s theory, the activation energy of the curing reaction was determined from

the peak temperatures at different heating rates using the following equation:46,47 ln(β /TP 2) = ln[(AR )/E] − E /(RTP)

(2)

where β is the heating rate, TP is the exothermic peak temperature, A is the pre-exponential factor, E is the activation energy, and R is the gas constant. The calculated kinetic parameters of the resins are given in Table 3. E presents how much energy is needed for the reaction to proceed in the forward direction.48 The E values of AESO at TP1 and TP2 are 83.4 and 124.6 kJ/mol, respectively (Table 3), which are considerably higher than those of similar acrylated epoxidized hemp oil-based thermosets (45−59 kJ/mol) due to the lack of styrene as an RD.49 The EP1 value of AESO is much lower than that of IM as a result of the higher reactivity of acrylic groups than methacrylic groups; however, a comparable EP2 was noticed between AESO and IM because of the limited molecular mobility. Furthermore, it was reported that in a methacrylate/acrylate system, the reactivity ratios (r) of the comonomer pairs usually follow r1 (methacrylate) > 1 > r2 (acrylate),50 which indicates that the IM-AESO resin has a tendency toward copolymerization with a higher incorporation rate of IM (methacrylate) into the copolymers than AESO (acrylate). This is responsible for the fact that the IM-AESO has a significantly higher EP1 than AESO but lower than IM. Comparing IM-MAESO to IM-AESO, an increased EP1 and a decreased EP2 are obtained because the IM-MAESO with a higher functionality would promote the cross-linking initially 779

DOI: 10.1021/acssuschemeng.6b02117 ACS Sustainable Chem. Eng. 2017, 5, 774−783

Research Article

ACS Sustainable Chemistry & Engineering Table 4. Flexural and Dynamic Mechanical Properties of the Crosslinked Thermosets samples AESO IM-AESO IM-MAESO IM

flexural strength (MPa)

flexural modulus (MPa)

± ± ± ±

62 ± 17 1373 ± 153 1546 ± 162 2060 ± 168

4.8 28.4 33.9 23.5

1.1 4.5 2.7 4.3

flexural strain at break (%)

E′ at 30 °C (MPa)

± ± ± ±

303 1961 2133 2736

10.4 2.1 1.9 1.1

0.9 0.3 0.2 0.4

Tg (°C)

ve × 103 (mol m3)

degree of cureb (%)

75.6 94.9 142.2

16.0a 75.2 83.3 59.7

81.0 97.6 88.3 85.4

The cross-linking density (ve) of pure AESO was calculated from the E′ at 70 °C according to the reported Tg (14−35 °C) value of the neat AESO network, which was not detected in our work;14,23,27 bDegree of cure for the resins was calculated from their FTIR spectra before and after curing based on the consumption of vinyl groups in the resin systems (Supporting Information SI7). a

Figure 5. DMA curves of IM, AESO, IM-AESO, and IM-MAESO thermosets. (a) Storage modulus. (b) Damping parameter.

network in the cured resin, which will be further discussed in Dynamic Mechanical Properties of the Cured Thermosets. However, in contrast to the soft fatty acids from AESO, the unique bicyclic ring structure from isosorbide predominately provides stiffness for the cured IM resin. Therefore, a higher flexural modulus and lower flexural strain are seen in the pure IM resin when compared to the IM-AESO and IM-MAESO resins. Dynamic Mechanical Properties of the Cured Thermosets. The storage modulus (E′) and loss factor (tan δ) of the cured networks are plotted as functions of temperature (Figure 5). E′ indicates the energy stored during cyclic deformation, and tan δ shows the damping behavior under vibrating conditions.51 The glass transition temperature (Tg) was determined from the peak position of the tan δ curve, and the cross-linking density (ve) was further estimated according to the rubbery elastic theory, eq 3:51,52

but hinder the process in the second stage due to the higher cross-linking level in the first stage. Flexural Properties of Cured Thermosets. The flexural properties of the cross-linked biobased thermosets are given in Table 4. Pure AESO has the lowest flexural strength and modulus but the highest flexural strain among the four resins, which is attributed to the low cross-linking degree of the neat AESO network as a result of the low unsaturation degree and long aliphatic molecular chains of the AESO monomer. The flexural strength and modulus of the AESO networks increase considerably, and the flexural strain dramatically decreases after the incorporation of IM. The incorporated IM would participate in the free-radical polymerization of the AESO system such that a cured AESO resin with a high cross-linking density is obtained. A higher cross-linking density would result in improvements in the strength, modulus, and brittleness. The introduced IM functions as a cross-linking agent, giving rise to the cross-linking density and rigidity while reducing the toughness of the resulting resin, i.e., the improved cross-linking density results in the cured resin fracturing via a fragile failure with low energy absorption. Furthermore, a 19.4% enhancement in flexural strength of the resin is observed after the modification of AESO with MAA, while no appreciable difference in flexural modulus and flexural strain are measured between IM-AESO and IM-MAESO. The MAA modification increases the number of CC bonds of the AESO monomer, thus generating IM-MAESO with an increased cross-linking density. Additionally, IM is a monomer with two CC bonds, and therefore, it can self-polymerize to form a thermoset with a three-dimensional network. The flexural strength of neat IM resin is lower than those of IM blends with modified soybean oil. This is most likely due to the incomplete cross-linking of CC bonds of IM monomers resulting in a highly defective

ve = E′/(3RT )

(3)

where E′ is the storage modulus above Tg in the rubbery plateau region (E′s at Tg + 40 °C for all networks were chosen in this study), R is the gas constant, and T is the absolute temperature. The E′ at 30 °C, Tg, and ve of the resins are listed in Table 4. Pure AESO has the lowest E′, indicating the weak interaction in the cured network due to its low cross-linking density (16.0 mol/m3) as restricted by its long molecular chains. This is in agreement with the calculated cure degree of the AESO network (81.0%) from its FTIR spectra before and after curing (Supporting Information, SI7). However, the addition of IM into AESO resin results in the generation of a cured IM-AESO network with a high degree of cure (97.6%) and cross-linking density (75.2 mol/m3). This contributes to the substantial increases in the E′ and Tg of IM-AESO 780

DOI: 10.1021/acssuschemeng.6b02117 ACS Sustainable Chem. Eng. 2017, 5, 774−783

Research Article

ACS Sustainable Chemistry & Engineering

Figure 6. (a) TG and (b) DTG curves of the IM, AESO, IM-AESO, and IM-MAESO thermosets.

compared to those of pure AESO. As mentioned in Curing Behaviors of the Resins, the reactivity ratio of AESO in the IMAESO system was lower than 1, indicating that AESO tends to copolymerize with IM, which facilitates the conversion of the CC bonds of AESO and hence increases the cure degree of the IM-AESO resin. The Tg of IM-AESO (75.6 °C) is comparable with that (71−78 °C) of the AESO resin with styrene as an RD,16,17 which is also because the incorporated IM has a peculiar bicyclic ring structure and thus superior thermal stability. After modifying AESO with MAA, the E′ at 30 °C and Tg increase from 1961 MPa and 75.6 °C for IM-AESO to 2133 MPa and 94.9 °C for IM-MAESO, respectively, which is due to the enhanced cross-linking density from 75.2 to 83.3 mol/m3. However, the increased unsaturated sites of IMMAESO did not completely participate in the free-radical polymerization, hence forming an imperfect network with a low degree of cure (88.3%). The neat IM has much higher E′ and Tg but lower cross-linking density and curing extent than IMAESO and IM-MAESO. The former is attributed to the unique structure of isosorbide, whereas the latter is associated with the formation of an imperfect network in the cured IM resin. As a small molecule and divinyl monomer, IM is extremely difficult to form a perfect network with full conversion of CC bonds. During the polymerization of IM monomer, once one methacrylic group of IM molecule polymerizes, the molecule with an unreacted CC bond would be locked in the network.53 Therefore, the mobility of the attached CC bond is dramatically reduced and the network with high degree of unsaturated sites (i.e., incomplete network) would form as the curing proceeds. In addition, the broader tan δ curve and a shoulder peak at high temperature in the curve of the cured IM resin confirm the formation of a wide range of segmental units that resulted from the incomplete three-dimensional network with different degrees of cross-linking. Thermal Stabilities of the Cured Thermosets. The thermal degradation behavior of the cured thermosets was evaluated by TGA (Figure 6 and Table 5). The IM-AESO, IMMAESO, and IM networks show considerably higher maximum weight loss temperatures (Tmax) than the pure AESO network due to the enhanced cross-linking density and the introduced isosorbide structure, which has superior thermal stability. However, an obvious degradation in the initial stage occurs in IM and IM-MAESO resins; in particular, IM has a wide and significant decomposition at 250−350 °C. To specify the

Table 5. Thermal Stability Data of the Cured Thermosets Obtained from the TGA Results characteristic weight loss temperature (°C) thermoset

Tmax

T5

T30

Ts

AESO IM-AESO IM-MAESO IM

385.7 424.0 417.3 432.2

318.5 312.0 289.8 271.4

375.5 387.1 381.1 393.4

172.8 175.0 168.8 168.9

overall thermal stability of the cured thermosets, the heatresistant index (Ts) was determined from the 5% and 30% weight loss temperatures (T5 and T30) according to the following equation:27,54 Ts = 0.49[T5 + 0.6(T30 − T5)]

(4)

As given in Table 5, the IM-AESO has the highest calculated Ts among all the networks because of its highest degree of cure. The thermal stability is not only associated with the crosslinking density of the cured resins but also has a close relationship with the chemical structure of the resins. As discussed in Dynamic Mechanical Properties of the Cured Thermosets, both IM-MAESO and IM networks are incomplete and have a number of unsaturated chain-ends inside the networks, which easily decompose with increased temperature. This is evidenced by the significantly lower T5 and hence reduced Ts of the thermosets when compared to AESO and IM-AESO.



CONCLUSION IM was successfully synthesized from isosorbide and MAA via a benign route and used to cross-link AESO for the formulation of soybean oil-based thermosets with high performances and renewable contents. ATR-FTIR analysis monitoring the synthesis of IM indicated that ultrasonic assistance was an effective method to accelerate the esterification between isosorbide and MAA in the absence of organic solvents; and the reaction rate and IM yield were greatly influenced by the molar ratio of MAA to isosorbide. The predicted solubility and experimental tests revealed that IM had superior miscibility with various substances and could effectively replace carcinogenic styrene in soybean oil- and UPE-based resins. 781

DOI: 10.1021/acssuschemeng.6b02117 ACS Sustainable Chem. Eng. 2017, 5, 774−783

Research Article

ACS Sustainable Chemistry & Engineering

(3) Gandini, A. Polymers from renewable resources: A challenge for the future of macromolecular materials. Macromolecules 2008, 41, 9491−9504. (4) Meier, M. A.; Metzger, J. O.; Schubert, U. S. Plant oil renewable resources as green alternatives in polymer science. Chem. Soc. Rev. 2007, 36, 1788−1802. (5) Montero de Espinosa, L.; Meier, M. A. Plant oils: The perfect renewable resource for polymer science? Eur. Polym. J. 2011, 47, 837− 852. (6) Lu, Y.; Larock, R. C. Novel polymeric materials from vegetable oils and vinyl monomers: Preparation, properties, and applications. ChemSusChem 2009, 2, 136−147. (7) Qiu, J. F.; Zhang, M. Q.; Rong, M. Z.; Wu, S. P.; Karger-Kocsis, J. Rigid bio-foam plastics with intrinsic flame retardancy derived from soybean oil. J. Mater. Chem. A 2013, 1, 2533−2542. (8) Lu, J.; Wool, R. P. Sheet molding compound resins from soybean oil: Thickening behavior and mechanical properties. Polym. Eng. Sci. 2007, 47, 1469−1479. (9) Colak, S.; Küsefoğlu, S. H. Synthesis and interfacial properties of aminosilane derivative of acrylated epoxidized soybean oil. J. Appl. Polym. Sci. 2007, 104, 2244−2253. (10) Zhang, P.; Xin, J.; Zhang, J. Effects of catalyst type and reaction parameters on one-step acrylation of soybean oil. ACS Sustainable Chem. Eng. 2014, 2, 181−187. (11) U. S. Department of Health and Human Services, Public Health Services. National Toxicology Program 12th Report on Carcinogens; 2011. (12) Campanella, A.; Zhan, M.; Watt, P.; Grous, A. T.; Shen, C.; Wool, R. P. Triglyceride-based thermosetting resins with different reactive diluents and fiber reinforced composite applications. Composites, Part A 2015, 72, 192−199. (13) Saithai, P.; Lecomte, J.; Dubreucq, E.; Tanrattanakul, V. Effects of different epoxidation methods of soybean oil on the characteristics of acrylated epoxidized soybean oil-co-poly (methyl methacrylate) copolymer. eXPRESS Polym. Lett. 2013, 7, 910−924. (14) Grishchuk, S.; Karger-Kocsis, J. Hybrid thermosets from vinyl ester resin and acrylated epoxidized soybean oil (AESO). eXPRESS Polym. Lett. 2011, 5, 2−11. (15) Grishchuk, S.; Karger-Kocsis, J. Modification of vinyl ester and vinyl ester-urethane resin-based bulk molding compounds (BMC) with acrylated epoxidized soybean and linseed oils. J. Mater. Sci. 2012, 47, 3391−3399. (16) Campanella, A.; La Scala, J. J.; Wool, R. P. The use of acrylated fatty acid methyl esters as styrene replacements in triglyceride-based thermosetting polymers. Polym. Eng. Sci. 2009, 49, 2384−2392. (17) Campanella, A.; La Scala, J. J.; Wool, R. P. Fatty acid-based comonomers as styrene replacements in soybean and castor oil-based thermosetting polymers. J. Appl. Polym. Sci. 2011, 119, 1000−1010. (18) Meiorin, C.; Aranguren, M. I.; Mosiewicki, M. A. Polymeric networks based on tung oil: Reaction and modification with green oil monomers. Eur. Polym. J. 2015, 67, 551−560. (19) Beach, E. S.; Cui, Z.; Anastas, P. T.; Zhan, M.; Wool, R. P. Properties of thermosets derived from chemically modified triglycerides and bio-based comonomers. Appl. Sci. 2013, 3, 684−693. (20) Liu, K.; Madbouly, S. A.; Kessler, M. R. Biorenewable thermosetting copolymer based on soybean oil and eugenol. Eur. Polym. J. 2015, 69, 16−28. (21) Chen, Z.; Wu, J. F.; Fernando, S.; Jagodzinski, K. Soy-based, high biorenewable content UV curable coatings. Prog. Org. Coat. 2011, 71, 98−109. (22) Ma, Q.; Liu, X.; Zhang, R.; Zhu, J.; Jiang, Y. Synthesis and properties of full bio-based thermosetting resins from rosin acid and soybean oil: the role of rosin acid derivatives. Green Chem. 2013, 15, 1300−1310. (23) Dai, J.; Ma, S.; Wu, Y.; Han, L.; Zhang, L.; Zhu, J.; Liu, X. Polyesters derived from itaconic acid for the properties and bio-based content enhancement of soybean oil-based thermosets. Green Chem. 2015, 17, 2383−2392.

Furthermore, the formulated IM-AESO blend was further modified with MAA, which resulted in a modified resin (IMMAESO) with a higher level of CC bonds. The soybean oilbased resins blended with IM as an RD (IM-AESO and IMMAESO) had superior processability due to the low viscosity. DSC analysis demonstrated a significantly increased curing efficiency in the blends of modified soybean oil with IM when compared to the pure IM and AESO resins. In addition, both the IM-AESO and IM-MAESO thermosets showed much higher flexural strengths, flexural moduli, storage moduli, and Tg but lower flexural strains than the neat AESO because of the increased cross-linking density and cure degree. A slight increase in strength and modulus was observed in the biobased thermoset after the modification of AESO with MAA. The neat IM resin also had higher flexural modulus, storage modulus and Tg than those of both IM-AESO and IM-MAESO due to the structural characteristic of IM; however, a much lower flexural strength was obtained for the IM resin because of the incomplete network formation. In summary, IM exhibits promise as a new biobased cross-linking agent for formulating soybean oil-based thermosets with superior processability and high performance for potential application in fiber-reinforced composites.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acssuschemeng.6b02117. ATR-FTIR, 1H NMR, and 13C NMR spectra of IM; ATR-FTIR and 13C NMR spectra of IM-AESO and IMMAESO resins; calculation of solubility parameters and RED values of IM and various chemicals according to the Hansen solubility theory; pictures of solubility evaluation results of IM with various chemicals obtained from digital camera; calculation of conversion of MAA, acid, and ester during the synthesis of IM from the ATR-FTIR spectra of reaction systems; calculation of cure degree for resins from ATR-FTIR spectra (PDF)



AUTHOR INFORMATION

Corresponding Author

*Phone: 86-591-83726495. Fax: 86-591-83715175. E-mail: [email protected] (R.Q.). Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We gratefully thank the Fujian Agriculture and Forestry University (Grant No. 1122YB019, KXB16007A), the National Natural Science Foundation of China (Grant No. 31670568), the Ministry of Education, China (Grant No. 220133515110015), and the State Administration of Forestry, China (Grant No. 2014-4-41) for funding.



REFERENCES

(1) Xia, Y.; Larock, R. C. Vegetable oil-based polymeric materials: Synthesis, properties, and applications. Green Chem. 2010, 12, 1893− 1909. (2) Gandini, A.; Lacerda, T. M. From monomers to polymers from renewable resources: Recent advances. Prog. Polym. Sci. 2015, 48, 1− 39. 782

DOI: 10.1021/acssuschemeng.6b02117 ACS Sustainable Chem. Eng. 2017, 5, 774−783

Research Article

ACS Sustainable Chemistry & Engineering (24) Stanzione, J. F., III; Sadler, J. M.; La Scala, J. J.; Wool, R. P. Lignin model compounds as bio-based reactive diluents for liquid molding resins. ChemSusChem 2012, 5, 1291−1297. (25) Liu, R.; Zhu, J.; Luo, J.; Liu, X. Synthesis and application of novel UV-curable hyperbranched methacrylates from renewable natural tannic acid. Prog. Org. Coat. 2014, 77, 30−37. (26) Jang, N. R.; Kim, H. R.; Hou, C. T.; Kim, B. S. Novel biobased photo-crosslinked polymer networks prepared from vegetable oil and 2, 5-furan diacrylate. Polym. Adv. Technol. 2013, 24, 814−818. (27) Ma, S.; Jiang, Y.; Liu, X.; Fan, L.; Zhu, J. Bio-based tetrafunctional crosslink agent from gallic acid and its enhanced soybean oil-based UV-cured coatings with high performance. RSC Adv. 2014, 4, 23036−23042. (28) Shin, S.; Kim, B.-C.; Chang, E.; Cho, J. K.; Suh, D. H. A biobased photocurable binder for composites with transparency and thermal stability from biomass-derived isosorbide. RSC Adv. 2014, 4, 6226−6231. (29) Wang, B.-T.; Lu, F.-D.; Xu, F.; Li, Y.-Z.; Kessler, M. R. Synthesis of renewable isosorbide-based monomer and preparation of the corresponding thermosets. Chin. Chem. Lett. 2016, 27, 875−878. (30) Rose, M.; Palkovits, R. Isosorbide as a renewable platform chemical for versatile applicationsQuo vadis? ChemSusChem 2012, 5, 167−176. (31) Sadler, J. M.; Toulan, F. R.; Palmese, G. R.; La Scala, J. J. Unsaturated polyester resins for thermoset applications using renewable isosorbide as a component for property improvement. J. Appl. Polym. Sci. 2015, 132, 42315. (32) Chrysanthos, M.; Galy, J.; Pascault, J.-P. Preparation and properties of bio-based epoxy networks derived from isosorbide diglycidyl ether. Polymer 2011, 52, 3611−3620. (33) Li, C.; Dai, J.; Liu, X.; Jiang, Y.; Ma, S.; Zhu, J. Green synthesis of a bio-based epoxy curing agent from isosorbide in aqueous condition and shape memory properties investigation of the cured resin. Macromol. Chem. Phys. 2016, 217, 1439−1447. (34) Niedermann, P.; Szebényi, G.; Toldy, A. Novel high glass temperature sugar-based epoxy resins: Characterization and comparison to mineral oil-based aliphatic and aromatic resins. eXPRESS Polym. Lett. 2015, 9, 85−94. (35) Liu, W.; Xie, T.; Qiu, R. Styrene-free unsaturated polyesters for hemp fibre composites. Compos. Sci. Technol. 2015, 120, 66−72. (36) Van Krevelen, D. W.; Te Nijenhuis, K. In Properties of polymers: Their correlation with chemical structure; their numerical estimation and prediction from additive group contributions; Elsevier: Oxford, 2009; Chapter 7, pp 189−227. (37) Hansen, C. M. Hansen solubility parameters: A user’s handbook, 2nd ed.; CRC press: London, 2007. (38) Zhang, C.; Kessler, M. R. Bio-based polyurethane foam made from compatible blends of vegetable-oil-based polyol and petroleumbased polyol. ACS Sustainable Chem. Eng. 2015, 3, 743−749. (39) Zou, H.; Lei, M. Optimum process and kinetic study of Jatropha curcas oil pre-esterification in ultrasonical field. J. Taiw. Inst. J. Taiwan Inst. Chem. Eng. 2012, 43, 730−735. (40) Zhang, F.; Pang, Z.; Dong, C.; Liu, Z. Preparing cationic cotton linter cellulose with high substitution degree by ultrasonic treatment. Carbohydr. Polym. 2015, 132, 214−220. (41) Liu, W.; Chen, T.; Xie, T.; Qiu, R. Soybean oil-based thermosets with N-vinyl-2-pyrrolidone as crosslinking agent for hemp fiber composites. Composites, Part A 2016, 82, 1−7. (42) La Scala, J. J.; Sands, J. M.; Orlicki, J. A.; Robinette, E. J.; Palmese, G. R. Fatty acid-based monomers as styrene replacements for liquid molding resins. Polymer 2004, 45, 7729−7737. (43) Lin, C.-P.; Tseng, J.-M.; Chang, Y.-M.; Cheng, Y.-C.; Lin, H.-Y.; Chien, C.-Y. Green thermal analysis for predicting thermal hazard of storage and transportation safety for tert-butyl peroxybenzoate. J. Loss Prev. Process Ind. 2012, 25, 1−7. (44) Cheng, S.-Y.; Tseng, J.-M.; Lin, S.-Y.; Gupta, J. P.; Shu, C.-M. Runaway reaction on tert-butyl peroxybenzoate by DSC tests. J. Therm. Anal. Calorim. 2008, 93, 121−126.

(45) Liu, X.; Xin, W.; Zhang, J. Rosin-based acid anhydrides as alternatives to petrochemical curing agents. Green Chem. 2009, 11, 1018−1025. (46) Vyazovkin, S.; Burnham, A. K.; Criado, J. M.; Pérez-Maqueda, L. A.; Popescu, C.; Sbirrazzuoli, N. ICTAC Kinetics Committee recommendations for performing kinetic computations on thermal analysis data. Thermochim. Acta 2011, 520, 1−19. (47) Hong, C.; Wang, X.; Pan, Z.; Zhang, Y. Curing thermodynamics and kinetics of unsaturated polyester resin with different chain length of saturated aliphatic binary carboxylic acid. J. Therm. Anal. Calorim. 2015, 122, 427−436. (48) Shao, F.; Yang, Q.; Li, L.; Lu, D. Self-cross-linking kinetics of unsaturated polyester poly (fumaric-co-itaconic-co-butanediol). J. Elastomers Plast. 2015, 47, 293−305. (49) Francucci, G.; Cardona, F.; Manthey, N. W. Cure kinetics of an acrylated epoxidized hemp oil-based bioresin system. J. Appl. Polym. Sci. 2012, 128, 2030−2037. (50) Siołek, M.; Matlengiewicz, M. Reactivity ratios of butyl acrylates in radical copolymerization with methacrylates. Int. J. Polym. Anal. Charact. 2014, 19, 222−233. (51) Wan, J.; Gan, B.; Li, C.; Molina-Aldareguia, J.; Kalali, E. N.; Wang, X.; Wang, D.-Y. A sustainable, eugenol-derived epoxy resin with high biobased content, modulus, hardness and low flammability: Synthesis, curing kinetics and structure−property relationship. Chem. Eng. J. 2016, 284, 1080−1093. (52) Zhang, C.; Madbouly, S. A.; Kessler, M. R. Biobased polyurethanes prepared from different vegetable oils. ACS Appl. Mater. Interfaces 2015, 7, 1226−1233. (53) Sadler, J. M.; Nguyen, A.-P. T.; Toulan, F. R.; Szabo, J. P.; Palmese, G. R.; Scheck, C.; Lutgen, S.; La Scala, J. J. Isosorbidemethacrylate as a bio-based low viscosity resin for high performance thermosetting applications. J. Mater. Chem. A 2013, 1, 12579−12586. (54) Nouailhas, H.; Aouf, C.; Le Guerneve, C.; Caillol, S.; Boutevin, B.; Fulcrand, H. Synthesis and properties of biobased epoxy resins. Part 1. Glycidylation of flavonoids by epichlorohydrin. J. Polym. Sci., Part A: Polym. Chem. 2011, 49, 2261−2270.

783

DOI: 10.1021/acssuschemeng.6b02117 ACS Sustainable Chem. Eng. 2017, 5, 774−783