Vaishaly
Duhan
a,
Shivani
Yadav
a,
Christophe
Len
bc and
Bimlesh
Lochab
*a
aMaterials Chemistry Laboratory, Department of Chemistry, School of Natural Sciences, Shiv Nadar Institution of Eminence, Gautam Buddha Nagar, Uttar Pradesh 201314, India. E-mail: bimlesh.lochab@snu.edu.in
bSchool of Chemistry, Xi'an Jiaotong University, 28 Xianning West Road, Xi'an, China
cInstitute of Chemistry for Life and Health Sciences, Chimie ParisTech, PSL Research University, 11 rue Pierre et Marie Curie, Paris, France
First published on 11th December 2023
Sustainable methodologies, such as mechanochemical mixers, have revolutionized the way compounds are synthesized in minimal time and with better yields. In this work, a series of latest 4th generation benzoxazine monomers were synthesized via a facile, highly scalable, efficient mechanochemical ball-milling process, making them more viable than those synthesized via traditional synthesis route. The 4th generation benzoxazine (BZ) monomers contained 100% biosynthons (vanillin, ortho-vanillin and furfuryl amine) and were purified without strenuous separation techniques. We found that the variation in the nature of aryl substitution (with and without phenolic–OH) at the reactive oxazine C2 centre governs the polymerization temperature, volatilization of monomers, kinetics of polymerization, thermal and mechanical properties. As expected, the control monomer without inherent phenolic–OH showed a high polymerization temperature and the advantages offered by vanillin regioisomers facilitated the polymerization with minimal volatile release. The latent catalytic effect in the monomer was confirmed by both temperature-dependent NMR and SCXRD studies. Interestingly, distal phenolic–OH was found to be more labile and polymerized easily at a lower temperature than when it was present at the ortho-position. Additionally, the former monomer structure led to a well-defined polymer network with an appreciable Tg (116 °C) and thermal stability (Tmax of 348 °C and char yield of 36%). Furthermore, it also revealed an excellent storage modulus and good adhesion properties compared to many classical petroleum-based polybenzoxazines. Overall, we showcased the viability of employing the benefits of copolymerizing two generations of benzoxazines at low temperature and explored this new class of latest benzoxazine monomers as greener adhesives with improved stability to enable future designing and open up the possibility of using them in several practical and innovative applications.
Currently, the practical applications of PBZs have remained elusive to some extent because of the requirement of a very high polymerization temperature, accounting for a concomitant significant mass loss, especially in earlier benzoxazine (BZ) monomer generations, during polymerization. To enable the polymerization of BZ at lower temperatures, various curing accelerators, catalysts, or initiators have been reported, ranging from inorganic salts to organic compounds. These external aids, including phenols,3 carboxylic acids,4p-toluenesulfonic acid and its derivatives,5 amines,6–8 methylimidazole,9 inorganic salts,10–12 organocatalysts, metal–organic frameworks,13 nanoparticles14 and polymers15 led to fast rates of polymerization. However, their use affects the nature of linkages in the polymer structure and often suffers from incompatibility issues, demanding additional processing steps. Alternatively, the introduction of active acidic functionalities, such as phenolic hydroxyls, thiols, and hydroxyethyl groups, in BZ monomers can also lower the ROP temperature.16–19
However, the issue of the short shelf life of BZs persists. To mitigate this limitation, the incorporation of functionalities with an inherent latent catalytic effect in the phenolic and amine counterparts of BZ has been attempted.20–24 Incorporating inactive structural motifs within the BZ monomer, which become activated upon an external trigger, is an interesting approach. In particular, functionalities involving intramolecular H-bonding lock the acidic labile hydrogen atoms and render them available once heat is applied to accelerate the polymerization reaction without compromising the shelf life of monomers.
The utilization of renewable feedstocks has led to a paradigm shift in reducing overdependence on the widespread use of petroleum-based feedstocks, aiming to maximize waste utilization and reduce environmental hazards arising from polymer usage. Several desirable functionalities are naturally integrated into renewable resources, and introducing them synthetically in petro-based starting materials is often challenging. Over the past few decades, the production of sustainable feedstock–derived polymeric materials including polybenzoxazines,25 gained significant interest to overcome the challenges associated with the depletion of fossil reserves, fluctuation in petroleum prices, and environmental concerns. A wide variety of renewable phenolics and amines are explored in polybenzoxazines, including cardanol,26,27 magnolol,28 urushiol,29 coumarin,30 eugenol,31 vanillin,32,33 guaiacol,34 isoeugenol, sesamol,35 bisguaiacol-F,36 resveratrol,37 catechol,38 catechin,23 chavicol,39 carvacrol,40 ferulic acid,41 peterostilbene,42 chrysin,43 anethole,44 furfurylamine, stearylamine,34 chitosan,15 isomannide,45 rosin,46 and tyramine.47 Biomass-derived derivatives of phenols and amines have proven successful as renewable resources, offering excellent opportunities to synthesize sustainable polymers capable of partially or fully replacing petroleum-based PBZs.22,25,30,34,48,49 Biobased PBZs have demonstrated their utility in several applications including high-performance anticorrosive coatings,50 superparamagnetic materials,51 cathode materials in next-generation batteries beyond Li-ions,52,53 antibacterials,54,55 and as potential flame-retardant materials.36,56,57 Recently, the replacement of formaldehyde with other aldehydes has also been employed to form 4th generation BZs.58–62 More recently, in this latest generation of BZs with oxazine-ring substitution found to empower structural tailoring close to the oxazine-ring reactive centre and accounted as a strategy to overcome mass loss issues and facilitated polymerization at lower temperature requirements than unsubstituted monomers.63,64 The progression of structural changes in benzoxazine monomers over the years is illustrated in Fig. 1.
The first generation is the most primitive and the monomer bears benign substituents. The second-generation monomers possess additional functionalities, which can self-/co-polymerize, and crosslink via other mechanisms, and may aid oxazine ring-opening reactions. In the third generation, oxazine groups are present in the main-chain, side-chain, and as terminal groups in the polymer to provide properties of both thermoplastics and thermosets. The fourth-eneration monomers possess smart architecture features and are subclassified into two types. This typically involves either in situ transformation of oxazine functionalities to others during polymerization65 or substitutions at the 2- and/or 4-position of the oxazine ring are explored.66
Vanillin, a biomass-derived feedstock phenolic aldehyde extracted from the vanilla bean, has been extensively explored as a phenol feedstock in unsubstituted BZs.32 Naturally, vanillin exists in two regioisomers, where the aldehyde is either in the ortho- or para-position relative to the phenolic–OH (Hphenolic), as shown in Fig. 2.
However, the exploration of a complete environmentally friendly process right from the synthetic methods, use of biosynthons, and latent catalytic effects at the oxazine-ring-substituted 4th generation benzoxazines have not been investigated and may provide a combined benefits in one go. The utility of the aldehyde functionality in vanillin as a replacement for formaldehyde has not been reported in BZ monomer synthesis. It is interesting to explore vanillin as a dual feedstock, serving as both phenolic and aldehyde starting materials, to synthesize oxazine-ring-substituted vanillin-based BZs. Furthermore, it is interesting to study the influence of this Hphenolic, at the 2- and 4-positions in vanillin, on polymerization characteristics in the latest generation of BZ. Still, although polybenzoxazines show great promise in many applications, their scalability is limited by their impractical industrially viable synthetic method. In the past few decades, there has been significant interest in alternative, cleaner, less energy-intensive, and sustainable chemical synthesis approaches compared to traditional conventional heating methods. Among the notable green techniques, both microwave-assisted and mechanochemical methods67 have emerged as highly efficient and environmentally friendly processes for the preparation of various molecules.68 They allowed accessibility to certain structures with a high product selectivity, which are difficult to obtain by other methods.69,70 Notably, the mechanical activation of molecules has enabled solvent-free, unprecedented mechanochemical “parallel synthesis”, with scalability ranging from milligrams to multikilogram, encouraging their usage for industrial polymer applications.
Current work presents a mechanochemical synthesis as a green, fast, and efficient process for producing fully biobased 4th generation BZs. Natural feedstock, vanillin, has been cleverly chosen to showcase an inbuilt latent catalytic effect, eliminating the need for the usual additional synthetic and processing steps. The polymerization behavior, mechanical properties, and thermal stability were studied and compared with the control structure.
LOI = 17.5 + 0.4 × CY | (1) |
Scheme 1 A generalized synthetic scheme of oxazine-ring-substituted BZ monomers using microwave (MW), mixer mill (MM) and planetary mill (PM). |
Monomer | Aldehyde | Conventional heating | Microwave reactor (MW) | Mechanochemical synthesis | ||||||
---|---|---|---|---|---|---|---|---|---|---|
Mixer mill (MM) | Planetary mill (PM)a | |||||||||
Time (h) | Yield (%) | Power (W) | Time (min) | Yield (%) | Time (min) | Yield (%) | Time (min) | Yield (%) | ||
a 12 mmol of aldehyde was used. | ||||||||||
oV-fa-[2]ph | Benzaldehyde | 6 | 75 | 100 | 30 | 76 | 15 | 72 | — | — |
oV-fa-[2]ov | o-Vanillin | 3 | 53 | — | 10 | 60 | 5 | 56 | — | — |
oV-fa-[2]v | Vanillin | 12 | 47 | 500 | 20 | 53 | 30 | 20 | 25 | 52 |
FTIR (cm−1): oxazine ring skeletal C–H bending out-of-plane vibration (918 cm−1), CAr–O–C symmetric (1003 cm−1) and anti-symmetric (1233 cm−1) stretching bands, and C–N–C symmetric stretching vibrations (1265 cm−1).
The polymerization of monomers and blends was performed in an air oven by sequential heating for 1 h each to obtain homopolymers and copolymers.
(2) |
The mechanical properties were evaluated using a Universal Testing Machine (UTM Instron-3366, USA) mechanical tester and rectangular polymeric specimens were made with a sample size of 65 × 1 × 0.1 mm3, at a strain rate of 0.5 mm s−1.
(3) |
Statistical analysis was performed using the GraphPad Prism Version 7.04 software. Student's two-tailed t-test was performed using the mean from the minimum of 3–6 independent observed values, and the respective p values are reported.
A telescoping mechanochemical procedure of reductive amination of o-hydroxy aldehyde derivative, o-vanillin, and furfuryl amine was performed involving an initial condensation reaction to form the Schiff base, followed by in situ reduction to form the aminophenol (AP) derivative. A series of C2-substituted BZ monomers were prepared in one pot by solvent-free condensation of AP with three different aldehydes (benzaldehyde, vanillin, and o-vanillin) for the ring closure reaction. The methodologies adopted for this step were conventional heating, microwave and mechanochemical. Without the need for bulk dissolution of reactants via typical heating of reactants, both milling and microwave irradiation assisted in achieving easy chemical transformation to form BZ monomers in good yields. Synthesis via conventional heating occurred in 3–12 h (47–75%), where the yield is the highest for benzaldehyde. Similarly, microwave reactions followed the same trend with the reaction occurring relatively fast (10–30 min) with good conversion observed in a mechanochemical mixer (MM) process (5–30 min). Despite several attempts, the condensation of AP with vanillin in MM resulted in a poor yield (20%, 30 min). An easier transformation with an improved yield (53%, 25 min) is achieved in planetary mill (PM) ascribed to the difference in the application of mechanical force, “planetary” motion of the jar with rotation and spinning around a central and own axis, compared with MM. Such creation of centrifugal forces provides the gravitational effects which is similar to roller mills with an industrial scale-up potential.
The structure of the synthesized intermediate and monomers was confirmed by FTIR, 1H and 13C NMR spectroscopy, and high-resolution electrospray ionization mass spectrometry. BZ monomers revealed the characteristic FTIR peaks due to the oxazine ring, CAr–O–C2 anti-symmetric (1212–1245 cm−1) and symmetric (1075 cm−1) stretching bands, and the oxazine ring skeletal C–H bending out-of-plane vibrations (912–922 cm−1) (Fig. S1†). A broad band at 3333 cm−1 was observed in the case of the oV-fa-[2]v monomer corresponding to the Hphenolic, while it remain unnoticed in oV-fa-[2]ov indicating involvement in intramolecular H-bonded formation. From 1H NMR, the number of protons clearly match the integral value of the NMR signals, confirming the successful formation of monomers (Fig. 3a, S2a, and S3a†). The characteristic methine proton O–C2(Ph)–N, “a” signal, at the C2 centre of the oxazine-ring-substituted BZ monomers, was observed in the range of 5.99–6.08 ppm as a singlet. The methylene protons due to N–CH2–Ar “b” and N–CH2–fa “c” are expected to appear as a doublet of doublet (dd) due to geminal coupling with a prochiral centre at C2.63
However in all the synthesized monomers, methylene proton signals N–CH2–Ar (b, 3.85–4.1 ppm) and N–CH2–fa (c, 3.72–3.94 ppm) appeared as multiplets due to the co-existence of –OCH3 signal(s) (d and e) in the similar range. The structural characterization of monomers was also confirmed by 13C NMR spectra (Fig. 3b, S2b, and S3b†). The characteristic carbons due to O–CH(Ph)–N, N–CH2–Ar, N–CH2–fa and –OCH3 are also successfully assigned as A (87.36 to 90.52 ppm), B (44.71 to 46.60 ppm), C (48.47 to 56.12 ppm) and D and E (56.12 ppm), confirming the successful formation of monomers. This assignment of protons and carbons was assisted by 2D and DEPT NMR analyses (Fig. S4–S6†). Mass spectrometry confirmed the formation of the monomers (Fig. S7†).
Fig. 4 The effect on H-bonding. Stacked zoomed-in 1H NMR spectra for oV-fa-[2]ov: (a) concentration-dependent study without and with D2O shake; recorded in CDCl3. (b) A temperature-dependent study with shifting of the phenolic proton resonance value; recorded in DMSO-d6. (c) A representative chemical structure of possible conformers of oV-fa-[2]ov mediated by temperature. An ORTEP diagram of oV-fa-[2]ov recorded at 16 °C (left) and 70 °C (right) [ORTEP diagram at 40 °C, is provided in Fig. S8†]. |
The key parameters to estimate the relative strength of hydrogen bonding is the difference in chemical shift [Δδ = δ(DMSO-d6) − δ(CDCl3)] and hydrogen bond acidity [ANMR = 0.0065 + 0.133 × (Δδ)] values, which are calculated as 0.52 and 0.069, respectively, which confirmed Hphenolic involvement in a very strong intramolecular H-bonding interaction.22,74 From a variable temperature NMR study (Fig. 4b), a well-defined Hphenolic signal upfield clearly shifted (lower ppm) with significant signal broadening upon increase in the temperature-supported temperature-driven weakening of intramolecular hydrogen bonding, and ultimately, the signal vanished at 70 °C. Undoubtedly, the hydrogen bonding becomes sufficiently weak with temperature to assist the release of the bonded proton “trapped-state” as in the six-membered ring to the “free-state”.
Learning from the above-mentioned solution-based NMR study and correlating with bulk polymerization, we were able to perform a similar study in solid state using single crystals obtained for one of the monomers along with the confirmation of the full 3D structure. The asymmetric unit, ORTEP diagram for oV-fa-[2]ov of formula C21H21NO5 is represented in Fig. 4c. The symmetry of space group P−1 generates a triclinic unit cell [a = 7.9414(6), α = 71.148°(3) b = 9.2292(7), β = 82.961°(3) c = 13.4539(11), γ = 81.232°(3)]. Complete data can be found in the CIF file, freely available with the CCDC number 2291158 (16 °C), 2291159 (40 °C) and 2291694 (70 °C)† at the Cambridge Crystallographic Data Centre. The successful claim of a strong 6-membered hydrogen bonding, with the bond distance between hydrogen and nitrogen (Hphenolic⋯Noxazine) determined as 1.994 Å lies within the range of the covalent interactions, as confirmed by SCXRD. The characteristic O–C2(Ph)–N angle which is prone to heterolytic cleavage to assist ROP got reduced from 112.06, which is less than that of the earlier unsubstituted and substituted BZ structures,45,64 supporting prudential role of functionalities at the 2-substituted oxazine-ring of the monomer. To gain better insights of the temperature dependency on the strength of hydrogen bonding for the same crystal, SCXRD data were recorded at three different temperatures (16, 40 and 70 °C). Due to the limitation of acquiring SCXRD data at elevated temperatures (>70 °C) for prolonged time and a low melting temperature of the monomer, the analysis was performed up to 70 °C. The ORTEP diagram for oV-fa-[2]ov at 40 and 70 °C is presented in Fig. S8† and Fig. 4c, respectively. Table S1† reports the main crystal data and refinement parameters for oV-fa-[2]ov at each temperature. As expected, with the increase in temperature, the bond distances Hphenolic⋯Noxazine, Hphenolic⋯Ooxazine, and Hphenolic⋯Ofuran were altered. Interestingly, the Hphenolic⋯Noxazine distance increased from 1.994 to 1.998 Å, while the Hphenolic⋯Ooxazine distance decreased from 3.853 to 3.850 Å and the characteristic O–C2(Ph)–N angle got reduced from 112.06 to 111.81°. Even though the changes are infinitesimal, they clearly indicate an increase in ring strain at the O–C2(Ph)–N angle and approach of Hphenolic to Ooxazine. Eventually, temperature-guided geometric conformations may offer a possibility of the C2–Cphenolic bond rotation, enabling the proximity of Hphenolic to the other neighbouring proton acceptor (Ooxazine) to create intramolecular O–Hphenolic⋯Ooxazine hydrogen bonds via a quasi-chelating ring to assist ROP in the dual mode, both due to the increase in angle strain and latent catalytic effects.62 As per earlier reported data, the latent catalyst can be activated by simply breaking the hydrogen bonding upon heating and can facilitate the acid-catalyzed polymerization of the BZs.
Fig. 5 (a–c) The DSC thermograms of the monomers at a heating rate of 10 °C min−1 under an N2 atmosphere. |
Monomer | Monomer | Polymer | Char yielda (%) | LOI | |||||
---|---|---|---|---|---|---|---|---|---|
T polymerization (°C) | ΔH (kJ mol−1) | E a modified Ozawa (kJ mol−1) | T decomposition (°C) | ||||||
T o | T p | T 5% | T 10% | T max | |||||
a At 800 °C. b Value in parenthesis is obtained from the Kissinger–Akahira–Sunose (KAS) method. | |||||||||
oV-fa-[2]ph | 149 | 214 | 28.41 | 100.7 (103.7)b | 299 | 318 | 357 | 40 | 33.5 |
oV-fa-[2]ov | 142 | 189 | 47.36 | 134.4 (135.84)b | 306 | 323 | 361 | 42 | 34.3 |
oV-fa-[2]v | 147 | 181 | 59.94 | 88.29 (99.17)b | 307 | 326 | 348 | 47 | 36.3 |
From the TGA analysis of monomers (Fig. 5), it was observed that both fully biobased monomers, oV-fa-[2]ov and oV-fa-[2]v, showed a lower overall mass and eventually reduced by three-fold than that observed in oV-fa-[2]ph. Thus, the introduction of a phenyl group in the oxazine ring at C2 is quiet promising to minimize such mass loss issues. In general, the higher the ΔH value, the higher the ease of polymerization. In our case, ΔH and %mass loss of monomers followed a reverse trend ascribed to the concomitant mass loss during polymerization. Thus, the obtained ΔH values require a careful interpretation. If we assume that the presence of –OH and –OMe groups in the phenyl ring at C2 in both oV-fa-[2]ov and oV-fa-[2]v reduces the number of available reaction sites for electrophilic aromatic substitution polymerization reactions at this benzene ring, then a reduced ΔH value is expected. Conversely, higher ΔH values are obtained in both than oV-fa-[2]ph, supporting the higher susceptibility of these monomers towards electrophilic aromatic substitution. Among oV-fa-[2]ov and oV-fa-[2]v, the lower ΔH value of the former could be an induction of steric interference due to substituents and neighbouring furan ring at the C2 reactive centre for polymerization, consequently making it prone to higher volatilization.
To gain mechanistic insights into the mass loss behaviour and to investigate the effect of a monomer structure on the nature of volatiles liberated during polymerization, GC-MS analysis was performed. The method applied and spectra obtained are presented in Fig. 6 and Table S2,† respectively. The monomers undergo dissociation, and the fragments liberated were captured and detected. Interestingly, the first chromatogram appeared at only 6 min and a much lower temperature of 97 °C for oV-fa-[2]ph, while it was observed at a delayed time of ∼11.5 min and a higher temperature of 150 °C for both oV-fa-[2]ov and oV-fa-[2]v. This indicates the appreciable thermal stability of the latter monomer structures in congruence to TGA of monomers. The volatiles are rich in phenolic and benzene fragments in oV-fa-[2]ph, while oV-fa-[2]ov liberated all possible different mass fragments, and oV-fa-[2]v showed predominance of substituted benzene species. Notably, a higher abundance of furan fragments was observed at 12 min in oV-fa-[2]ov than in oV-fa-[2]v, indicating the dangling nature and steric congestion towards the polymer network growth due to the proximity of –OH/OMe groups in the former monomer. At 222 °C, a broad vs. sharp elution chromatogram was obtained in phenyl vs. vanillin-based monomers, respectively.
Fig. 6 The GC/MS chromatogram of the monomers and the corresponding fragmentation pattern and m/z (Da) values. |
The liberation of furan-based imines in each monomer type is mass driven, 185 Da being lighter showcased relatively higher volatilization than the imine (231 Da) derivative from the other two monomers, i.e., oV-fa-[2]ov/v. Consequently the residual mass left to polymerize is seemingly rich in furan units, retain monomer features, furan and substituted benzene units for oV-fa-[2]ph, oV-fa-[2]ov, and oV-fa-[2]v, respectively. Clearly, this indicates the promising thermal stability and resistance to volatilization offered by fully biobased monomers.
FTIR kinetics was further performed to determine the temperature at which the characteristic oxazine absorption bands in the synthesized monomers are affected due to their participation in the ROP. Fig. S9† shows significant broadening of the absorption bands (1212–1245 cm−1, C–O–C antisymmetric stretching mode and 912–922 cm−1 BZ related mode) at 130 °C in both oV-fa-[2]ov and oV-fa-[2]v, while a similar effect is noticed at 160 °C in oV-fa-[2]ph, inferring the catalyzing effect induced in the ring-opening reaction due to Hphenolic. Temperature-dependent SCXRD studies supported the mapping of phenolic–OH proximity more to the oxazine N than to O; however, to investigate whether Hphenolic transfers fast intra- or intermolecularly, the calculations of activation energy (Ea) of polymerization were performed using the Kissinger–Akahira–Sunose (KAS)75 and modified Ozawa methods. A sequence of DSC traces at different heating rates of 5, 7.5, 10, 15, and 20 °C min−1 were recorded. From the slope of the plots of ln(β/Tp2) and ln(β) against 1/Tp, the activation energy of polymerization was calculated using eqn (4) and (5) for the KAS and modified Ozawa methods:
(4) |
(5) |
Fig. 7a–c shows the plots obtained for the KAS and modified Ozawa methods using respective DSC thermograms of the monomers (details are provided in Table S3†). The Ea values for polymerization were calculated from the slope, which follows the order oV-fa-[2]ov (134 kJ.mol−1) > oV-fa-[2]ph (101 kJ.mol−1) > oV-fa-[2]v (88 kJ.mol−1). From this trend, it can be interpreted that kinetically, the ROP is most favoured in the monomer having Hphenolic at the para-position than the one having no additional functional groups attached and followed by the monomer having Hphenolic at the ortho-position of the C2(Ph) entity. This is accounted to the availability of more propagation sites free from steric congestion at the reactive centre and resonance stabilization of the intermediate iminium ion (Fig. 7a′–c′) to dictate the extension of the polymer network.
Fig. 8 TGA thermograms for (a) poly(oV-fa-[2]ph), (b) poly(oV-fa-[2]ov) and (c) poly(oV-fa-[2]v) at a heating rate of 10 °C min−1 under a N2 atmosphere. |
Besides benzene rings, the participation of a furan ring in cross-linking reactions accounted for improved char yields.77,78 High decomposition temperatures, char yields and LOI values are indicative of the effectiveness of bio-synthons to form halogen-free flame-resistant sustainable polymers.79
To ensure crosslinking, we performed swelling studies of neat polymers and copolymers in different solvents (acetone, hexane, water, dimethyl sulfoxide, ethanol, toluene and dichloromethane). The variation in the mass of samples when soaked in different solvents was monitored after every 24 h for four consecutive days. The relative swelling ratios are shown in Fig. S11.† It was observed that the swelling ratio of neat polymers and copolymers remains nearly unaffected in water, hexane, and ethanol. A lower swelling ratio in water and ethanol followed by DMSO suggests the highly hydrophobic nature of the samples. The homopolymers of the 4th generation swelled up significantly in dichloromethane and revealed insignificant colouration of the solvent after 3 days, which is not evident in case of copolymers with the 1st generation monomer, confirming their improved solvent resistance.
The adhered coupons using various homo- and copolymers based on bare BZ monomers and their monomer blends (1:1 or 1:3 or 3:1 w/w) are analyzed using an Instron instrument. The pictorial representation of the assembly used to measure the adhesive strength along with the LSS values obtained is presented in Fig. 9a and b. Expectedly, all the neat 4th generation PBZs showed poor adhesion strength (LSS = 0.1 kg cm−2) ascribed to the formation of a rigid and brittle polymer network. On the contrary, C–a based PBZ showed LSS = 20 kg cm−2. The copolymers with C–a with oV-fa-[2]ph/ov/v showed a much higher adhesive strength of 51 kg cm−2 and 37 kg cm−2 at the respective optimum comonomer feed-in ratio, signifying the benefits gained due to the co-existence of hard and soft segments. The thermo-mechanical properties of 4th generation homopolymers and their copolymers was assessed by rheological studies. From Fig. S12,† it is clear that the storage modulus of poly(oV-fa-[2]v) is highest compared to poly(oV-fa-[2]ph) followed by poly(oV-fa-[2]ov). No rubbery plateau is observed in poly(oV-fa-[2]ph), indicating that the molecular weight of polymers is low involving an assemblage of units with broad molecular weight distribution. Usually, for a low crosslinked thermoset expected to show a low Tg value, and for such polymers, the storage modulus begins to decrease at much low temperature.
The observed poly(oV-fa-[2]ov) has low storage modulus than poly(oV-fa-[2]ph) and poly(oV-fa-[2]v), but instead of the decreasing storage modulus, it began to increase with the temperature, indicating the reorganization of a quasi-network to more stable network structure. Expectedly, poly(oV-fa-[2]v) showed a significant high storage modulus and a sharp tanδ peak indicating the formation of a well-defined polymer network. The reason why tanδ is high despite a broad peak observed in the case of poly(oV-fa-[2]ph) is a separate matter and not very well understood. Certainly, the reactivity and stability of iminium ion intermediates to propagate into a crosslinked network vs. the stability of the iminium ion, electrophilic centre, via the respective phenyl derivative (substituted vs. unsubstituted) is another governing parameter. However, a high network heterogeneity indicate “partial” cross-linking and broadening of the tanδ plot (Fig. 9c). An increase in the glass transition temperature (Tg), which is determined as the temperature at the tanδ peak, increases for each polymer in the following order copolymer poly(oV-fa-[2]ov) < poly(oV-fa-[2]v) < poly(oV-fa-[2]ph). The same trend is observed in their respective analyzed copolymers with C–a, with an increased Tg value (Fig. 9d).
The absence of independent polymerizing exothermic peaks due to individual monomers (4th generation and 1st generation C–a, Tp = 260 °C) in the DSC plot (Fig. S13†) confirmed the successful copolymerization reaction. Furthermore, the beneficial effect of the ring-opened structures of these C2-substituted monomers enabled a lowering in C–a copolymerization temperature. The knitting-in of C–a units within the 4th generation polymer network may have led to reinforcement of the copolymeric network, altering the molecular structure with restricting mobility of polymer chains, thus improving the overall thermal (Fig. S14†) and mechanical properties. Moreover, this is due to increased viscoelastic dissipation and simultaneously the highest and broad tanδ accounting for the strongest adhesion characteristics in case of oV-fa-[2]ph and its copolymer.
Following the same optimized composition of samples used for adhesive strength measurements, bulk mechanical properties were also determined. The samples were cured without any reinforcing fillers at atmospheric pressure in a Teflon mould in an air oven, which accounted for their low mechanical properties. The samples showed a nearly blister-free composite (Fig. S15†), indicating good processability and nearly zero mass loss during polymerization. The tensile stress–strain curves of polybenzoxazines, prepared from pure cardanol–aniline (C–a) and its different blends with other 4th generation benzoxazine monomers, revealed an elastic response with minimal plastic deformation and finally terminated by brittle failure (Fig. S16†), typical of a thermoset-associated rigid network. Except Ov1Ca3, which additionally showed a relatively sustained plateau of deformation at a high strain, likely a result of bucking plastic yielding before brittle fracture. The tensile strength, modulus, and elongation at the break of samples is presented in Fig. S16c and d.† C–a being unsubstituted mono-oxazine (compared to 4th generation monomers) produced a polymer with low cross-linking densities, and therefore, it showed overall inferior mechanical properties. The comonomer not only allowed low-temperature polymerization of C–a, but also provided densification of the network structure that empowered a steady increase in stress and stiffness. A more plasticizing effect of Oo1Ca3 than that of Ov1Ca3 due to dangling methoxy and hydroxy functionalities near the oxazine ring C2 centre in the former comonomer contributed to relatively high tensile strength and elongation at break.
Footnote |
† Electronic supplementary information (ESI) available: Synthesis of C-a, NMR (1H, 2D (HMBC, HSQC, COSY), 13C, and DEPT of monomers), FTIR and DSC data of monomers, FTIR kinetics of polymerization, ORTEP diagram of oV-fa-[2]ov at 40 °C, crystal data for compound at variable temperatures, GC-MS method, swelling studies, DMA analysis and tensile stress–strain plots. CCDC 2291158 (16 °C), 2291159 (40 °C) and 2291694 (70 °C). For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d3gc03522k |
This journal is © The Royal Society of Chemistry 2024 |