Biobased Ester 2-(10-Undecenoyloxy)ethyl Methacrylate as an

Jul 16, 2019 - ... territory, which would enable income generation through family farming, ... Another interesting function of diene molecules is thei...
0 downloads 0 Views 4MB Size
Article pubs.acs.org/IECR

Cite This: Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Biobased Ester 2‑(10-Undecenoyloxy)ethyl Methacrylate as an Asymmetrical Diene Monomer in Thiol−Ene Polymerization Cristian de Oliveira Romera, Deb́ ora de Oliveira, Pedro Henrique Hermes de Araújo, and Claudia Sayer* Department of Chemical Engineering and Food Engineering, Federal University of Santa Catarina, P.O. Box 476, 88.040-900 Florianópolis, Santa Catarina, Brazil

Downloaded via BUFFALO STATE on July 30, 2019 at 01:30:13 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

S Supporting Information *

ABSTRACT: In this work, a diene ester monomer, 2-(10-undecenoyloxy)ethyl methacrylate (MHU), was synthesized by the enzymatic esterification of 10-undecenoic acid (derived from castor oil) and 2-hydroxyethyl methacrylate and employed in two different ways: (1) in thiol−ene polymerization with 1,4-butanedithiol (1,4-BDT) and (2) as a precursor in a Michael addition reaction with hexamethylenediamine. In thiol−ene bulk polymerization of the asymmetrical diene monomer MHU, which presents different reactivities for each double bond, higher molecular weights (up to 44.3 kDa) were reached when nonstoichiometric thiol−ene ratios were used. In the second approach, Michael addition allowed us to obtain a symmetrical diene monomer that was used to produce a polymer (11.3 kDa) containing both amine and ester groups by thiol−ene polymerization. Finally, for the same MHU/1,4-BDT molar ratio, miniemulsion thiol−ene polymerization resulted in higher molecular weights (up to 17 kDa) when compared with those from bulk polymerization (8.7 kDa).

1. INTRODUCTION The substitution of petroleum derivatives by renewable materials has been the focus of several lines of research in order to achieve more sustainable development, as a result of the encouragement of the use of natural resources and bioprocesses, especially since Agenda 21 of Rio 92.1,2 Plant oils are outstanding as raw materials because of their ability to generate new and renewable monomers and polymers even after a few modification reactions, presenting high availability and industrial viability.3 Thus, triglycerides have been used in the production of different polymers, such as polyesters, polyurethanes, polyamides, polyesteramides, and alkyd and epoxy resins, as versatile precursors.4 Among these plant oils, castor oil has been used in industrial and pharmacological applications, mainly because of its high concentration of ricinoleic acid (90%), which makes it the only source of hydroxylated fatty acid.5 This property makes this oil a natural polyol, conferring resistance to oxidation.6 From a Brazilian socioeconomic point of view, the cultivation area of castor bean has been estimated at 31 800 ha with production of 20 000 000 kg during the harvest of 2017−2018, and despite its potential to be produced throughout Brazilian territory, which would enable income generation through family farming, castor bean is still concentrated mainly in four states: Bahia, Mato Grosso, Ceará, and Minas Gerais.7 The potential of castor oil to be transformed into several unsaturated compounds through chemical modifications, such as 10-undecenoic acid, methyl-10-undecenoate, 10-undecenol, and 10-undecenoyl chloride, has been highlighted in order to apply it as the building blocks of biobased polymers.3,8−14 © XXXX American Chemical Society

In 1973, John Ugelstad, a visiting professor from Norway; postdoctoral researcher Mohamed S. El-Aasser; and Professor John Vanderhoff15 at Lehigh University showed that submicrometer monomer droplets, stabilized by a combination of sodium lauryl sulfate emulsifier and cetyl alcohol, which resulted in a liquid−liquid colloidal system that remained stable for 2 weeks, were loci for polymerization reactions with styrene occurring via droplet nucleation. This work gave birth to the now well-known miniemulsion polymerization technique. The important change provided by the long-chain fatty alcohol at the interface between the organic droplets and the continuous aqueous phase resulted in adequate stability, maintaining the system while simple mixing was applied.16,17 Miniemulsion polymerization involves the effective combination of a surfactant and a costabilizer to produce submicrometer (50 to 500 nm) monomer droplets.18 Currently, the use of high shear in order to create small droplets and the addition of a highly water insoluble costabilizer to prevent Ostwald ripening have been more frequently applied.17,19 Ostwald ripening is an effect from the chemical potential in small droplets being higher, causing monomer diffusion from small to large droplets.20 The stability provided by the costabilizer is a consequence of its low solubility in water, retarding the diffusion of monomers into the Special Issue: Mohamed El-Aasser Festschrift Received: Revised: Accepted: Published: A

April 30, 2019 July 13, 2019 July 16, 2019 July 16, 2019 DOI: 10.1021/acs.iecr.9b02386 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research

Figure 1. Mechanism of thiol−ene polymerization with an asymmetrical diene monomer with one alkene and one methacrylate.

hydrophilic system, whereas the polymerization is faster.21,22 Despite the extensive use of hexadecane as a costabilizer, the

main drawback has been its permanency in the polymer after polymerization, which could impair applications.22 One B

DOI: 10.1021/acs.iecr.9b02386 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research

basis of these perspectives, this work aims to investigate thiol− ene polymerization of the diene ester monomer 2-(10undecenoyloxy)ethyl methacrylate (MHU) in bulk and in miniemulsion, evaluate the effects of reaction temperature and molar ratio of thiol/ene, and compare it with free-radical polymerization. In addition, the asymmetrical monomer was modified by Michael addition in order to build a symmetrical amino diene ester monomer for thiol−ene polymerization.

alternative is the use of monomers such as stearyl methacrylate,22−26 lauryl methacrylate,24,25 and dodecyl methacrylate22 as reactive costabilizers. Miniemulsion polymerization has been used in several applications, such as the production of low viscosity latexes, controlled radicalar polymerization in dispersed media, catalytic polymerization, encapsulation of inorganic solids, incorporation of hydrophobic monomers, formation of hybrid polymer particles, and development of pressure-sensitive adhesives in continuous tubular reactors, among others.17,19,27 Thus, the miniemulsion process has become quite versatile, allowing researchers to obtain nanoparticles by various polymerization mechanisms, such as anionic, cationic, radicalar, polyaddition, and polycondensation mechanisms.28,29 However, its application in thiol−ene polymerization is quite recent, and few studies have been reported so far in this research field. Lobry et al.30 performed for the first time thiol−ene polymerization in miniemulsion. Applying diallyl adipate as diene and ethylene glycol dithiol as thiol, they obtained nanoparticles with diameters of 155 nm. Jasinski et al.31 reported the preparation of polymer nanoparticles from the same diene and thiol by photopolymerization using αhydroxyketone as the initiator. Amato et al.32 related thiol−ene miniemulsion photopolymerization of 1,3,5-triallyl-1,3,5-triazine-2,4,6, a triene, with pentaerythritol pentaerythritoltetra(3-mercaptopropionate), a tetrathiol, to the obtainment of primary amine; however, after the reaction, the formed secondary amines had lower reactivity than the primary amines. This behavior was attributed to the steric hindrance of the carbon-chain, which made the interaction of the nucleophile with a double bond more difficult.59 González et al.45 observed that a substitution of an acrylate with a methacrylate group led to a reduction in conversion from 87 to 9% in Michael addition with poly(ethylenimine) and bisphenol A glycerolate diacrylate. They attributed this effect to the reduction of the electrophilic behavior due to the steric hindrance caused by the methyl group. These studies indicated that primary amines were more reactive than secondary amines already reacted by Michael addition because of steric hindrance; in addition, methacrylate also conferred steric hindrance because of its methyl group, which suggested that reactions with MHU involved partial addition in the amine functional group, preventing two methacrylates from reacting with the same amine group of HMD and thus leading to the obtainment of the symmetrical monomer MHU-HMD expected in Figure 5. In order to test the reactivity of this new monomer, thiol−ene polymerization was performed using 1 mol % AIBN as the initiator at 80 °C and a molar MHU-HMD/1,4-BDT ratio of 1.0:1.0 to be comparable to reaction B03. Figure 6 presents the spectrum of polymer (PHU-HMD), which shows the disappearance of the peaks related to the hydrogens of the double bonds related to complete conversion. This polymer was not soluble in THF because of the insertion of the hydrophilic amine functional group, hindering GPC

Figure 7. DSC curves of the second heating run of polymers prepared with different MHU/1,4-BDT molar ratios: 1.0:1.00 (B03), 1.0:0.90 (B05), 1.0:0.75 (B06), and 1.0:000 (free-radical, B09). A MHU-HMD/ 1,4-BDT molar ratio of 1.0:1.00 (PHU-HMD) was also analyzed.

melting temperature (Tm) showed an increase as the amount of 1,4-BDT was decreased. As expected, the DSC curve for PHUHMD presented different behavior, possibly related to the new functionality. Sánchez-Soto et al.61 reported an increase in Tm with increasing Mw up to 4 kDa. Higher molecular weights presented stabilization in melting temperatures of poly(ethylene oxide) samples. Wu et al.,62 working with the synthesis of poly(βH

DOI: 10.1021/acs.iecr.9b02386 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research

polymerization with 9.0 μmol·cm−3 SDS (M02) and 13.5 μmol· cm−3 Lut50 (M03). Machado et al.13 compared particle sizes with 8.0 μmol·cm−3 SDS and Lut50, observing smaller particle diameters for SDS than for Lut50. Cardoso et al.14 observed similar behavior when SDS and Lutensol AT80 were applied. The MHU monomer enabled the formation of stable miniemulsions, leading to intensity average diameters of polymer particles of approximately 200 nm without requiring the presence of a costabilizer because of its hydrophobicity. Thus, the fact that MHU acted as a reactive costabilizer in the miniemulsion, avoiding the addition of another low molecular weight costabilizer that would remain in the final product and could negatively impact the polymer properties and limit its application, was an advantage. Table 2 presents the molecular weights after thiol−ene and free-radical polymerizations in miniemulsions using 1.0 mol % AIBN as the initiator. In terms of molecular weight, an increase in the temperature from 70 to 80 °C led to a small increase in Mw from 5.2 (M04) to 7.9 kDa (M03), with an MHU/1,4-BDT molar ratio of 1.0:1.0, as was also observed for bulk polymerizations. In addition, the stoichiometric imbalance of MHU/1,4-BDT also resulted in an increase in Mw in miniemulsion polymerizations. Although in these reactions had stoichiometric imbalances of MHU/1,4-BDT, the increase in temperature did not lead to a further increase in Mw. The poly(thioether ester) resulting from these reactions presented a Tm below the reaction temperatures; thus, similar molecular weights at this temperature range were expected because diffusion limitations were not observed. By comparing molecular weight results for bulk and miniemulsion polymerizations in Tables 1 and 2, it was possible to observe that for stoichiometric thiol−ene molar ratios, both polymerization techniques resulted in similar molecular weights, considering the Mw values of bulk (B03) and miniemulsion (M03) polymerizations were 7.7 and 7.9 kDa, respectively. Cardoso69 worked with thiol−ene polymerization of symmetric terminal diene esters (1,3-propylene dipent-4-enoate and 1,3propylene diundec-10-enoate) and observed similar molecular weights in bulk and miniemulsion polymerizations. The discrepancy in Mw values was pronounced when the MHU/1,4-BDT ratio was 1:0:0.9, because bulk polymerization (B05) resulted in Mw of 8.7 kDa, whereas miniemulsion polymerization reached 14.7 (M06) and 17.0 kDa (M07) when Lut50 and SDS were used, respectively, under the same polymerization conditions. Machado et al.13 observed an increase in molecular weight using a stoichiometric diene/1,4BDT molar ratio in miniemulsion polymerization and attributed this behavior to the radical compartmentalization effect, which led to a decrease in the termination of the free-radical mechanism.17,70 It has been shown that compartmentalization occurs even when oil soluble initiators are used, as single radicals may be efficiently formed in the polymer particles by radical desorption and reentry.71,72 In fact, because the stoichiometry between MHU and 1,4-BDT was modified, by reducing the 1,4BDT amount, the importance of the free-radical mechanism tended to increase for these reactions, making the radical compartmentalization effect more pronounced. Nanoparticles were characterized on the basis of the TEM images presented in Figure 9. Thiol−ene polymer nanoparticles resulted in low molecular weights and low Tm values, which caused problems during analysis because of the melting of samples M03 and M08. Thus, most nanoparticles from M03 and M08 presented deformation by melting as a result of the

thioether ester) and poly(β-thioether ester-co-lactone) from methyl 3-((2-hydroxiethyl)thio) propanoate (MHETP) and εcaprolactone, observed an increase in Tm from 5 to 47 °C with the enhancement of the molar ratio of ε-caprolactone/MHETP from 0.5:0.5 to 0.8:0.2, which caused an increase in Mn from 44.4 to 51.4 kDa. Furthermore, for a thiol−ene photopolymerization reaction between pentaerythritol tetra(3-mercaptopropionate) and 1,6-hexanedioldiacrylate, the reduction of thiol molar concentration caused an increase in the glass transition temperature of the polymeric film.49 Multiple melting peaks, as those observed in Figure 7, are common in semicrystalline polymers, being reported, for example, in poly(ethylene terephthalate),63 poly(ether ether ketone),64 propylene-ethylene copolymers,65 poly(butylene2,6-naphthalate),66 and the polymer from methyl 10-undecenoate.67 Recently, studies with poly(thioether esters) have shown double melting points13 and multiple melting temperatures during the first heating.68 In these cases, the lower peaks were related to partial melting of crystals, with continuous recrystallization and melting during intervals, whereas the higher peak represented the melting of the organized crystallite.64 3.4. Thiol−Ene and Free-Radical Miniemulsion Polymerization of MHU. Nanoparticles were obtained by thiol− ene and free-radical miniemulsion polymerizations over 4 h at 80 °C with 1 mol % AIBN and different types and amounts of surfactants: 9.0 and 13.5 μmol·cm−3 Lut50 (M01) and 9.0 μmol· cm−3 SDS (M02). For the concentration of 9.0 μmol·cm−3, the mass percentage of Lut50 (20 wt %) was 10 times that used for SDS (2 wt %), considering the molecular weight of Lut50 was 2460 g·mol−1, and that of SDS was 288.68 g·mol−1. Higher stability was provided by SDS, as big monomer droplets were visible on the surface of the miniemulsion prepared with 9.0 μmol·cm−3 Lut50. This behavior can be associated with the different mechanisms for colloidal stabilization, because SDS presents electrostatic coverage, whereas Lut50 stabilizes by steric hindrance. Thus, the concentration of Lut50 was increased to 13.5 μmol·cm−3 in order to improve the stability by steric hindrance, and thus the appearance of macroscopic droplets could be avoided. Figure 8 shows the intensity percent size distributions of particles prepared by thiol−ene miniemulsion

Figure 8. Particle size distribution after thiol−ene miniemulsion polymerization with different surfactant types. Reactions were performed at 80 °C for 4 h with 1 mol % AIBN. I

DOI: 10.1021/acs.iecr.9b02386 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research Table 2. Weight Average Molecular Weights (Mw), Dispersities (Đ), Intensity Average Particle Diameters (Dp), and Polydispersity Indices (PDI) of Polymers from MHU under Different Thiol−Ene and Free-Radical Miniemulsion Polymerization Conditions entry

MHU/1,4-BDT

surfactant (wt %a)

T (°C)

Dp0b

PDI0b

Dp

PDI

Mw (kDa)

Đ

M04 M03 M05 M06 M07 M08 M09c

1.0:1.0 1.0:1.0 1.0:0.9 1.0:0.9 1.0:0.9 1.0:0.9 1.0:0.0

Lut50 (30) Lut50 (30) Lut50 (30) Lut50 (30) SDS (2) SDS (2) Lut50 (30)

70 80 70 80 70 80 80

159 ± 1 159 ± 1 184 ± 1 184 ± 1 226 ± 5 226 ± 5 139 ± 1

0.12 0.12 0.20 0.20 0.15 0.15 0.19

161 ± 1 153 ± 1 240 ± 4 181 ± 1 201 ± 1 243 ± 4 140 ± 1

0.17 0.21 0.24 0.25 0.25 0.15 0.19

5.2 7.9 15.4 14.7 16.6 17.0 

2.3 3.1 3.3 3.4 4.0 3.5 

a

Related to the monomer. bInitial monomer droplets. cFree-radical polymerization (without BDT).

Figure 9. (a,b) TEM images of samples after thiol−ene miniemulsion polymerization with (a) Lut50 (M03) and (b) SDS (M08). (c) TEM image of sample after free-radical miniemulsion polymerization using Lut50 (M09).

miniemulsion, resulting in higher molecular weights (up to 17 kDa) when compared with those from bulk polymerization (8.7 kDa) with the same MHU/dithiol molar ratio.

intensity of the electron beam. The melting of poly(thioether esters) due to exposure to the electron beam of the TEM equipment has already been reported, even when operating at the lowest current exposure possible.13,14 On the other hand, the cross-linked nanoparticles in Figure 9c from free-radical polymerization (M09) showed well-defined spherical morphologies with similar diameters when compared to those from DLS analysis, demonstrating thermal resistance against the electron beam in TEM. The fact that MHU acts as its own reactive costabilizer in miniemulsion polymerization increases the range of potential applications of the nanoparticles, as they do not contain an additional low molecular weight costabilizer that could negatively impact their properties. In addition, the cytocompatibility for HELA and L929 cells and blood biocompatibility have already been studied for other poly(thioether ester) nanoparticles from dienes based on 10-undecenoic acid.13,14 Thus, this system presents high potential as drug delivery nanoparticles or for transport of antimicrobial33 and antioxidant34 compounds, although more studies on the polymers synthesized in this work are required.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.iecr.9b02386. HPLC chromatogram of purified MHU and Debye plot for the determination of the weight average molecular weight of PHU-HMD (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Débora de Oliveira: 0000-0001-7544-5045 Pedro Henrique Hermes de Araújo: 0000-0001-5905-0158 Claudia Sayer: 0000-0003-1044-2905



Notes

The authors declare no competing financial interest.



CONCLUSIONS The asymmetric α,ω-diene ester monomer MHU, obtained from enzymatic esterification and partially derived from plant oil, was properly polymerized by thiol−ene polymerization to obtain poly(thioether ester). MHU was also modified by Aza− Michael addition by reacting the methacrylate group of MHU with a diamine, forming a symmetric α,ω-diene ester amino monomer, which was polymerized by thiol−ene polymerization to obtain a poly(amino-thioether ester). The thiol−ene bulk polymerization of MHU showed that the presence of a methacrylate and an alkene group with different reactivities led to a combination of step- and chain-growth polymerization, which could be more pronounced through the stoichiometric imbalance of the diene monomer and dithiol, leading to higher molecular weights. Finally, the diene ester monomer MHU was also successfully polymerized by thiol−ene polymerization in

ACKNOWLEDGMENTS The authors would like to thank Coordenaçaõ de Aperfeiçoá Superior (CAPES) and Conselho mento de Pessoal de Nivel ́ Nacional de Desenvolvimento Cientifico e Tecnológico (CNPq) for financial support; Novozymes for kindly donating the enzymes used in this work; and Central de Análises at the Department of Chemical and Food Engineering and Laboratório Central de Microscopia Eletrônica (LCME/UFSC), both at the Federal University of Santa Catarina, for DSC and TEM analyses, respectively.



REFERENCES

(1) Eissen, M.; Metzger, J. O.; Schmidt, E.; Schneidewind, U. 10 Years after Rio−Concepts on the Contribution of Chemistry to a Sustainable Development. Angew. Chem., Int. Ed. 2002, 41, 414−436.

J

DOI: 10.1021/acs.iecr.9b02386 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research

taining Monomer and Reactive Costabilizer. J. Dispersion Sci. Technol. 2010, 31, 1568−1573. (25) Kusdianto, C. T. L.; Yu, F. E.; Chern, C. S. Miniemulsion Copolymerizations of Styrene and Reactive Alkyl Methacrylate Costabilizers. J. Appl. Polym. Sci. 2011, 119, 620−628. (26) Ballard, N.; Aguirre, M.; Simula, A.; Leiza, J. R.; Van Es, S.; Asua, J. M. High Solids Content Nitroxide Mediated Miniemulsion Polymerization of: N-Butyl Methacrylate. Polym. Chem. 2017, 8, 1628−1635. (27) Asua, J. M. Challenges and Opportunities in Continuous Production of Emulsion Polymers: A Review. Macromol. React. Eng. 2016, 10, 311−323. (28) Landfester, K.; Musyanovych, A.; Mailänder, V. From Polymeric Particles to Multifunctional Nanocapsules for Biomedical Applications Using the Miniemulsion Process. J. Polym. Sci., Part A: Polym. Chem. 2010, 48, 493−515. (29) Crespy, D.; Landfester, K. Miniemulsion Polymerization as a Versatile Tool for the Synthesis of Functionalized Polymers. Beilstein J. Org. Chem. 2010, 6, 1132−1148. (30) Lobry, E.; Jasinski, F.; Penconi, M.; Chemtob, A.; CroutxéBarghorn, C.; Oliveros, E.; Braun, A. M.; Criqui, A. Continuous-Flow Synthesis of Polymer Nanoparticles in a Microreactor via Miniemulsion Photopolymerization. RSC Adv. 2014, 4, 43756−43759. (31) Jasinski, F.; Lobry, E.; Tarablsi, B.; Chemtob, A.; CroutxéBarghorn, C.; Le Nouen, D.; Criqui, A. Light-Mediated Thiol-Ene Polymerization in Miniemulsion: A Fast Route to Semicrystalline Polysul Fide Nanoparticles. ACS Macro Lett. 2014, 3, 958−962. (32) Amato, D. V.; Amato, D. N.; Flynt, A. S.; Patton, D. L. Functional, Sub-100 Nm Polymer Nanoparticles via Thiol-Ene Miniemulsion Photopolymerization. Polym. Chem. 2015, 6, 5625− 5632. (33) Amato, D. N.; Amato, D. V.; Mavrodi, O. V.; Braasch, D. A.; Walley, S. E.; Douglas, J. R.; Mavrodi, D. V.; Patton, D. L. Destruction of Opportunistic Pathogens via Polymer Nanoparticle-Mediated Release of Plant-Based Antimicrobial Payloads. Adv. Healthcare Mater. 2016, 5, 1094−1103. (34) de Meneses, A. C.; dos Santos, P. C. M.; Machado, T. O.; Sayer, C.; de Oliveira, D.; de Araújo, P. H. H. Poly(Thioether-Ester) Nanoparticles Entrapping Clove Oil for Antioxidant Activity Improvement. J. Polym. Res. 2017, 24, 202. (35) Hoelscher, F.; Machado, T. O.; de Oliveira, D.; Hermes de Araújo, P. H.; Sayer, C. Enzymatically Catalyzed Degradation of Poly (Thioether-Ester) Nanoparticles. Polym. Degrad. Stab. 2018, 156, 211− 217. (36) Hoyle, C. E.; Cole, M.; Bachemin, M.; Kuang, W.; Kalyanaraman, V.; Jönsson, S. Photoinitiated Polymerization of Selected Thiol-Ene Systems. In Photoinitiated Polymerization; Belfield, K. D., Crivello, J. V., Eds.; ACS Symposium Series 847; American Chemical Society: Washington, DC, 2003; pp 52−64. (37) Cramer, N. B.; Davies, T.; O’Brien, A. K.; Bowman, C. N. Mechanism and Modeling of a Thiol-Ene Photopolymerization. Macromolecules 2003, 36, 4631−4636. (38) Lecamp, L.; Houllier, F.; Youssef, B.; Bunel, C. Photoinitiated Cross-Linking of a Thiol-Methacrylate System. Polymer 2001, 42, 2727−2736. (39) Moszner, N.; Völkel, T.; Rheinberger, V. Synthesis, Characterization and Polymerization of Dendrimers with Methacrylic End Groups. Macromol. Chem. Phys. 1996, 197, 621−631. (40) Klee, J. E.; Neidhart, F.; Flammersheim, H. J.; Mülhaupt, R. Monomers for Low Shrinking Composites, 2: Synthesis of Branched Methacrylates and Their Application in Dental Composites. Macromol. Chem. Phys. 1999, 200, 517−523. (41) Müh, E.; Marquardt, J.; Klee, J. E.; Frey, H.; Mülhaupt, R. Bismethacrylate-Based Hybrid Monomers via Michael-Addition Reactions. Macromolecules 2001, 34, 5778−5785. (42) Mathias, L. J.; Shemper, B. S.; Alirol, M.; Morizur, J. F. Synthesis of New Hydroxylated Monomers Based on Methacrylate, Dimethacrylate, and Tetramethacrylate Michael Adducts and Photopolymeriza-

(2) Meier, M. A. R.; Metzger, J. O.; Schubert, U. S. Plant Oil Renewable Resources as Green Alternatives in Polymer Science. Chem. Soc. Rev. 2007, 36, 1788−1802. (3) Montero de Espinosa, L.; Meier, M. A. R. Plant Oils: The Perfect Renewable Resource for Polymer Science?! Eur. Polym. J. 2011, 47, 837−852. (4) Güner, F. S.; Yaǧci, Y.; Erciyes, A. T. Polymers from Triglyceride Oils. Prog. Polym. Sci. 2006, 31, 633−670. (5) Severino, L. S.; Auld, D. L.; Baldanzi, M.; Cândido, M. J. D.; Chen, G.; Crosby, W.; Tan, D.; He, X.; Lakshmamma, P.; Lavanya, C.; et al. A Review on the Challenges for Increased Production of Castor. Agron. J. 2012, 104, 853−880. (6) Patel, V. R.; Dumancas, G. G.; Viswanath, L. C. K.; Maples, R.; Subong, B. J. J. Castor Oil: Properties, Uses, and Optimization of Processing Parameters in Commercial Production. Lipid Insights 2016, 9, 1−12. (7) Acompanhamento da Safra Brasileira de Grãos; V. 6 - Safra 2018/19 ́ - N. 3; Companhia Nacional de Abastecimento, Observatório Agricola, 2018. (8) Mutlu, H.; De Espinosa, L. M.; Meier, M. A. R. Acyclic Diene Metathesis: A Versatile Tool for the Construction of Defined Polymer Architectures. Chem. Soc. Rev. 2011, 40, 1404−1445. (9) Türünç, O.; Meier, M. A. R. Fatty Acid Derived Monomers and Related Polymers via Thiol-Ene (Click) Additions. Macromol. Rapid Commun. 2010, 31, 1822−1826. (10) Warwel, S.; Brüse, F.; Demes, C.; Kunz, M.; Klaas, M. R. G. Polymers and Surfactants on the Basis of Renewable Resources. Chemosphere 2001, 43, 39−48. (11) Fokou, P. A.; Meier, M. A. R. Studying and Suppressing Olefin Isomerization Side Reactions during Admet Polymerizations. Macromol. Rapid Commun. 2010, 31, 368−373. (12) Cardoso, P. B.; Musyanovych, A.; Landfester, K.; Sayer, C.; De Araújo, P. H. H.; Meier, M. A. R. ADMET Reactions in Miniemulsion. J. Polym. Sci., Part A: Polym. Chem. 2014, 52, 1300−1305. (13) Machado, T. O.; Cardoso, P. B.; Feuser, P. E.; Sayer, C.; Araújo, P. H. H. Thiol-Ene Miniemulsion Polymerization of a Biobased Monomer for Biomedical Applications. Colloids Surf., B 2017, 159, 509−517. (14) Cardoso, P. B.; Machado, T. O.; Feuser, P. E.; Sayer, C.; Meier, M. A. R.; Araújo, P. H. H. Biocompatible Polymeric Nanoparticles From Castor Oil Derivatives via Thiol-Ene Miniemulsion Polymerization. Eur. J. Lipid Sci. Technol. 2018, 120, 1700212. (15) Ugelstad, J.; El-Aasser, M. S.; Vanderhoff, J. W. Emulsion Polymerization: Initiation of Polymerization in Monomer Droplets. J. Polym. Sci., Polym. Lett. Ed. 1973, 11, 503−513. (16) El-Aasser, M. S.; Lack, C. D.; Choi, Y. T.; Min, T. I.; Vanderhoff, J. W.; Fowkes, F. M. Interfacial Aspects of Miniemulsions and Miniemulsion Polymers. Colloids Surf. 1984, 12, 79−97. (17) El-Aasser, M. S.; Sudol, E. D. Miniemulsions: Overview of Research and Applications. J. Coatings Technol. Res. 2004, 1, 20−31. (18) Schork, F. J.; Luo, Y.; Smulders, W.; Russum, J. O.; Butté, A.; Fontenot, K. Miniemulsion Polymerization. Adv. Polym. Sci. 2005, 175, 129−255. (19) Asua, J. M. Miniemulsion Polymerization. Prog. Polym. Sci. 2002, 27, 1283−1346. (20) Kabalnov, A. S.; Shchukin, E. D. Ostwald Ripening Theory: Applications to Fluorocarbon Emulsion Stability. Adv. Colloid Interface Sci. 1992, 38, 69−97. (21) Erdem, B.; Sudol, E. D.; Dimonie, V. L.; El-Aasser, M. S. Encapsulation of Inorganic Particles via Miniemulsion Polymerization. III. Characterization of Encapsulation. J. Polym. Sci., Part A: Polym. Chem. 2000, 38, 4441−4450. (22) Asua, J. M. Ostwald Ripening of Reactive Costabilizers in Miniemulsion Polymerization. Eur. Polym. J. 2018, 106, 30−41. (23) Chern, C. S.; Lim, H.; Cala, N. A. Miniemulsion Copolymerizations of Styrene and Methyl Methacrylate in the Presence of Reactive Costabilizer. J. Appl. Polym. Sci. 2009, 112, 173−180. (24) Meliana, Y.; Cala, N. A.; Lin, C. T.; Chern, C. S. Ostwald Ripening of Two-Component Disperse Phase Miniemulsions ConK

DOI: 10.1021/acs.iecr.9b02386 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research tion Kinetics of Bulk Cross-Linkers. Macromolecules 2004, 37, 3231− 3238. (43) Weickmann, H.; Gurr, M.; Meincke, O.; Thomann, R.; Mülhaupt, R. A Versatile Solvent-Free “One-Pot” Route to Polymer Nanocomposites and the in Situ Formation of Calcium Phosphate/ Layered Silicate Hybrid Nanoparticles. Adv. Funct. Mater. 2010, 20, 1778−1786. (44) Bilgici, Z. S.; Ordu, O. D.; Isik, M.; Avci, D. Synthesis and Polymerizations of Six Aminophosphonate-Containing Methacrylates. J. Polym. Sci., Part A: Polym. Chem. 2011, 49, 5042−5048. (45) González, G.; Fernández-Francos, X.; Serra, À .; Sangermano, M.; Ramis, X. Environmentally-Friendly Processing of Thermosets by TwoStage Sequential Aza-Michael Addition and Free-Radical Polymerization of Amine-Acrylate Mixtures. Polym. Chem. 2015, 6, 6987−6997. (46) Navarro, R.; Monterde, C.; Molina, S.; Pérez-Perrino, M.; Reviriego, F.; del Prado, A.; Gallardo, A.; Reinecke, H. Understanding the Regioselectivity of Michael Addition Reactions to Asymmetric Divinylic Compounds. RSC Adv. 2017, 7, 56157−56165. (47) de Oliveira Romera, C.; de Oliveira, D.; Sayer, C.; de Araújo, P. H. H. Enzymatic Synthesis of a Diene Ester Monomer Derived from Renewable Resource. Appl. Biochem. Biotechnol., in press, 2019, DOI: 10.1007/s12010-019-03043-9 (48) Romera, C. de O.; Cardoso, P. B.; Meier, M. A. R.; Sayer, C.; Araújo, P. H. H. Acyclic Triene Metathesis (ATMET) Miniemulsion Polymerization of Linseed Oil Produces Polymer Nanoparticles with Comparable Molecular Weight to That of Bulk Reactions. Eur. J. Lipid Sci. Technol. 2015, 117, 235−241. (49) Hoyle, C. E.; Lee, T. Y.; Roper, T. Thiol-Enes: Chemistry of the Past with Promise for the Future. J. Polym. Sci., Part A: Polym. Chem. 2004, 42, 5301−5338. (50) Cramer, N. B.; Reddy, S. K.; O’Brien, A. K.; Bowman, C. N. Thiol-Ene Photopolymerization Mechanism and Rate Limiting Step Changes for Various Vinyl Functional Group Chemistries. Macromolecules 2003, 36, 7964−7969. (51) Northrop, B. H.; Coffey, R. N. Thiol-Ene Click Chemistry: Computational and Kinetic Analysis of the Influence of Alkene Functionality. J. Am. Chem. Soc. 2012, 134, 13804−13817. (52) Reddy, S. K.; Cramer, N. B.; Kalvaitas, M.; Lee, T. Y.; Bowman, C. N. Mechanistic Modelling and Network Properties of Ternary Thiol - Vinyl Photopolymerizations. Aust. J. Chem. 2006, 59, 586−593. (53) Cramer, N. B.; Couch, C. L.; Schreck, K. M.; Carioscia, J. A.; Boulden, J. E.; Stansbury, J. W.; Bowman, C. N. Investigation of ThiolEne and Thiol-Ene-Methacrylate Based Resins as Dental Restorative Materials. Dent. Mater. 2010, 26, 21−28. (54) Yao, B.; Mei, J.; Li, J.; Wang, J.; Wu, H.; Sun, J. Z.; Qin, A.; Tang, B. Z. Catalyst-Free Thiol-Yne Click Polymerization: A Powerful and Facile Tool for Preparation of Functional Poly(Vinylene Sulfide)s. Macromolecules 2014, 47, 1325−1333. (55) Okay, O.; Reddy, S. K.; Bowman, C. N. Molecular Weight Development during Thiol-Ene Photopolymerizations. Macromolecules 2005, 38, 4501−4511. (56) Cramer, N. B.; Bowman, C. N. Kinetics of Thiol-Ene and ThiolAcrylate Photopolymerizations with Real-Time Fourier Transform Infrared. J. Polym. Sci., Part A: Polym. Chem. 2001, 39, 3311−3319. (57) Cramer, N. B.; Couch, C. L.; Schreck, K. M.; Boulden, J. E.; Wydra, R.; Stansbury, J. W.; Bowman, C. N. Properties of MethacrylateThiol-Ene Formulations as Dental Restorative Materials. Dent. Mater. 2010, 26, 799−806. (58) Lee, T. Y.; Smith, Z.; Reddy, S. K.; Cramer, N. B.; Bowman, C. N. Thiol-Allyl Ether-Methacrylate Ternary Systems. Polymerization Mechanism. Macromolecules 2007, 40, 1466−1472. (59) Wu, D.; Liu, Y.; He, C.; Chung, T.; Goh, S. Effects of Chemistries of Trifunctional Amines on Mechanisms of Michael Addition Polymerizations with Diacrylates. Macromolecules 2004, 37, 6763− 6770. (60) Kroeger, A.; Zhang, B.; Rosenauer, C.; Schlüter, A. D.; Wegner, G. Solvent Induced Phenomena in a Dendronized Linear Polymer. Colloid Polym. Sci. 2013, 291, 2879−2892.

(61) Sánchez-Soto, P. J.; Ginés, J. M.; Arias, M. J.; Novák, C.; RuizConde, A. Effect of Molecular Mass on the Melting Temperature, Enthalpy and Entropy of Hydroxy-Terminated PEO. J. Therm. Anal. Calorim. 2002, 67, 189−197. (62) Wu, W. X.; Qu, L.; Liu, B. Y.; Zhang, W. W.; Wang, N.; Yu, X. Q. Lipase-Catalyzed Synthesis of Acid-Degradable Poly(β-Thioether Ester) and Poly(β-Thioether Ester-Co-Lactone) Copolymers. Polymer 2015, 59, 187−193. (63) Hiramatsu, N.; Hirakawa, S. Melting Behavior of Poly(Ethylene Terephthalate) Crystallized and Annealed under Elevated Pressure. Polym. J. 1980, 12, 105−111. (64) Blundell, D. J. On the Interpretation of Multiple Melting Peaks in Poly(Ether Ether Ketone). Polymer 1987, 28, 2248−2251. (65) Feng, Y.; Jin, X.; Hay, J. N. Evaluation of Multiple Melting Peaks of Propylene-Ethylene Copolymers. Polym. J. 1998, 30, 215−221. (66) Ju, M.-Y.; Chang, F.-C. Multiple Melting Behavior of Poly(Butylene-2,6-Naphthalate). Polymer 2001, 42, 5037−5045. (67) Türünç, O.; Meier, M. A. R. Fatty Acid Derived Monomers and Related Polymers via Thiol-Ene (Click) Additions. Macromol. Rapid Commun. 2010, 31, 1822−1826. (68) Zhang, D.; Dumont, M. J. Synthesis, Characterization and Potential Applications of 5-Hydroxymethylfurfural Derivative Based Poly(β-Thioether Esters) Synthesized via Thiol-Michael Addition Polymerization. Polym. Chem. 2018, 9, 743−756. (69) Cardoso, P. B. Biobased Polymeric Nanoparticles from Castor Oil Derivatives by Admet and Thiol-Ene Miniemulsion Polymerizations. Ph.D. Thesis, Federal University of Santa Catarina, Florianópolis, Brazil, June 2016. (70) Landfester, K. Miniemulsions for Nanoparticle Synthesis. Top. Curr. Chem. 2003, 227, 75−123. (71) Autran, C.; De la Cal, J. C.; Asua, J. M. (Mini)emulsion Polymerization Kinetics Using Oil-Soluble Initiators. Macromolecules 2007, 40, 6233−6238. (72) Costa, C.; Timmermann, S. A. S.; Pinto, J. C.; Araujo, P. H. H.; Sayer, C. Compartmentalization Effects on Miniemulsion Polymerization with Oil-Soluble Initiator. Macromol. React. Eng. 2013, 7, 221− 231.

L

DOI: 10.1021/acs.iecr.9b02386 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX