Next Article in Journal
κ-Carrageenan Hydrogel as a Matrix for Therapeutic Enzyme Immobilization
Next Article in Special Issue
The Synthesis of Bio-Based Michael Donors from Tall Oil Fatty Acids for Polymer Development
Previous Article in Journal
Molecular Simulations and Network Analyses of Surface/Interface Effects in Epoxy Resins: How Bonding Adapts to Boundary Conditions
Previous Article in Special Issue
Comprehensive Investigation of Stoichiometry–Structure–Performance Relationships in Flexible Polyurethane Foams
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Bio-Based Polymer Developments from Tall Oil Fatty Acids by Exploiting Michael Addition

1
Polymer Laboratory, Latvian State Institute of Wood Chemistry, Dzerbenes St. 27, LV-1006 Riga, Latvia
2
Institute of Technology of Organic Chemistry, Faculty of Materials Science and Applied Chemistry, Riga Technical University, P. Valdena St. 3/7, LV-1048 Riga, Latvia
*
Author to whom correspondence should be addressed.
Polymers 2022, 14(19), 4068; https://doi.org/10.3390/polym14194068
Submission received: 31 August 2022 / Revised: 23 September 2022 / Accepted: 24 September 2022 / Published: 28 September 2022
(This article belongs to the Special Issue Recent Advances in Polyurethane Materials)

Abstract

:
In this study, previously developed acetoacetates of two tall-oil-based and two commercial polyols were used to obtain polymers by the Michael reaction. The development of polymer formulations with varying cross-link density was enabled by different bio-based monomers in combination with different acrylates—bisphenol A ethoxylate diacrylate, trimethylolpropane triacrylate, and pentaerythritol tetraacrylate. New polymer materials are based on the same polyols that are suitable for polyurethanes. The new polymers have qualities comparable to polyurethanes and are obtained without the drawbacks that come with polyurethane extractions, such as the use of hazardous isocyanates or reactions under harsh conditions in the case of non-isocyanate polyurethanes. Dynamic mechanical analysis, differential scanning calorimetry, thermal gravimetric analysis, and universal strength testing equipment were used to investigate the physical and thermal characteristics of the created polymers. Polymers with a wide range of thermal and mechanical properties were obtained (glass transition temperature from 21 to 63 °C; tensile modulus (Young’s) from 8 MPa to 2710 MPa and tensile strength from 4 to 52 MPa). The synthesized polymers are thermally stable up to 300 °C. The suggested method may be used to make two-component polymer foams, coatings, resins, and composite matrices.

1. Introduction

The European Commission has set up a “European Green Deal” to tackle climate and environmental challenges. It is a new growth strategy that aims to transform the European Union into a prosperous society with a modern, resource-efficient, and competitive economy, achieving climate neutrality and a balance in carbon emissions and removals by 2050, which means reducing net greenhouse gas emissions to zero [1]. Improving the polymer industry by introducing innovative, sustainable, and knowledge-intensive technologies for producing materials from renewable bio-resources is also an essential field for achieving this goal [2].
In the polyurethane industry, sustainability has emerged as a major concern. Conventionally, polyurethanes are produced from two components—A (polyol) and B (isocyanate)—that are derived from petrochemicals. Particular attention is given towards the synthesis of polyols from biological fatty acids, diglycerides, and triglycerides [3,4,5,6,7,8]. Such bio-polyols are suitable for the production of polyurethanes as a method to make polyurethanes more sustainable [9]. However, the use of isocyanates still remains problematic as isocyanates are considered as toxic substances relative to the environment [10,11,12,13,14].
Recently, one of the extensive research topics is the development of non-isocyanate polyurethanes (NIPUs) materials as an alternative to polyurethanes. This technology has significant drawbacks, such as high pressure, high temperature, and long reaction times [15,16]. The polymer industry requires alternative methods for producing polymers in relatively mild conditions with similar properties to polyurethanes. The Michael reaction can be used as a new synthesis pathway to yield bio-based polymers with a wide range of properties [2,17].
The Michael reaction is a versatile synthesis type for linking electron-poor olefins to a wide range of nucleophiles [18]. Enolates [19,20], nitrogen-containing [21,22,23,24,25,26] reagents, and thiols [26,27,28,29,30,31,32] are typical Michael donors, while alkyl acrylates are mainly used as Michael acceptors [33]. Adding an enolate anion to the α,β-unsaturated carbonyl compound to form a carbon–carbon bond is a Michael reaction [19,20,34]. A type of polymerization for the addition of hydroxyl monomers with electrophilic double or triple bonds is called the oxa-Michael reaction [35,36]. In the case of nitrogen and sulphur-containing donors, the reaction is called aza-Michael [21,22,23,24,25,26,37] and thio-Michael [26,27,28,29,30,31,32] addition reactions, respectively. In recent years, aza- and thio-Michael reactions have gained attention for polymerization [35,38]. However, carbon-Michael reactions have gained less attention.
The majority of the studies has been conducted on coating materials [39,40,41,42,43]. Recently, Naga et al. presented Michael-addition reactions of multi-functional acetoacetate and diacrylate monomers, yielding corresponding gels at room temperature in the presence of 1,8-diazabicyclo[5.5.0]undecane-7-ene as a catalyst [44]. Park reported how hydrophobic thin-film nanocomposite membranes could be successfully developed using only naturally occurring compounds by interfacial polymerization [45]. Sinha et al. has concluded that the nucleophilicity of the used enamines had a significant impact on the efficiency and total reaction yields. Moreover, photo-protecting amines with 2-(2-nitrophenyl)propyl chloroformate in the presence of a ketone create enamines in situ [46].
The Michael reaction can yield a variety of linear, graft, dendric, highly cross-linked, and hyperbranched polymer networks [18,33]. The mechanism of the reaction is relatively well-studied and straightforward. The thermodynamical mechanism depends on the base’s relative strengths and the type of acetoacetate. The base deprotonates the acetoacetate, resulting in an enolate anion. The enolate anion then reacts with the acrylate in a 1,4-conjugate addition, and the carbonyl of the acrylate stabilizes, allowing the base to regenerate (see Figure 1). The enthalpic transition that occurs when a π-bond is replaced with a σ-bond is the overall driving force for conjugate addition. As a result, 1,4-addition is preferred over 1,2-addition. The product of the first Michael reaction contains active methylene hydrogen that can be added to another acrylate in a second step [14,34], which means that the Michael donor reacts twice with two Michael acceptors [33], of course, if the ratio allows it. The polymeric material can only be formed if the molecule of the Michael acceptor component contains at least two functional groups.
The reaction between acrylates and the active methylene group of β-ketoesters is quick, and the system cures at room temperature after mixing it in the presence of catalysts. This reaction is a strong base-catalyzed addition [18,20,34]. For the Michael reaction amine, amidine- and guanidine-based catalysts may be the most suitable [47,48] for example 1,1,3,3-tetramethylguanidine, triethylamine, 1,4-diazabicyclo[2.2.2]octane, (1,8-diazabicyclo[5.4.0]undec-7-ene), 4,4′-methylenebiscycloheexanamine, (1,5-diazabicyclo[4.3.0]non-5-ene), and 7-methyl-1,5,7-triazabicyclo[0,4,4]dec-5-ene [18,20,34,39,42,48,49,50,51].
In this study, we used previously developed acetoacetates of two tall-oil-based and two commercial polyols to obtain polymers by the Michael reaction. The previously developed acetoacetates had different chemical structures and functionalities. In conjunction with the use of various acrylates in polymerization reactions, it allowed the development of polymer formulations with varied cross-link density. Three acrylates with different functionalities—bisphenol A ethoxylate diacrylate (BPAEDA), trimethylolpropane triacrylate (TMPTA), and pentaerythritol tetraacrylate (PETA)—were used. The physical and thermal properties of the developed polymers were studied using dynamic mechanical analysis (DMA), differential scanning calorimetry (DSC), thermal gravimetric analysis (TGA), and a universal strength testing machine. The chemical structure and functional groups present in the developed polymer were determined using Fourier-transform infrared spectroscopy (FTIR). The proposed technology has a potential in the production of two-component polymer foams, coatings, resins, and matrix for composites.

2. Materials and Methods

2.1. Materials

Epoxidized tall oil fatty acids 1,4-butanediol polyol acetoacetate (EIRTOFA_DB_AA) were obtained from polyol that was synthesized from tall oil fatty acids via epoxidation using ion exchange resin (EIRTOFA) followed by oxirane ring-opening and esterification with 1,4-butandiol and subsequent acetoacetylation with tert-butyl acetoacetate by a transesterification reaction, acid value < 5 mg KOH·g−1, hydroxyl value 36.2 mg KOH·g−1, moisture 0.025%, and acetoacetate groups 0.3307 mol·100 g−1. Epoxidized tall oil fatty acids trimethylolpropane polyol acetoacetate (EIRTOFA_TMP_AA) was obtained from EIRTOFA by oxirane ring-opening and esterification with trimethylolpropane and subsequent acetoacetylation with tert-butyl acetoacetate by a transesterification reaction, acid value < 5 mg KOH·g−1, hydroxyl value 41.6 mg KOH·g−1, moisture 0.037%, and acetoacetate groups 0.4562 mol·100 g−1. Acetoacetyled Neopolyol 380 (NEO380_AA) was synthesized from Neopolyol 380 (NEO380) by transesterification with tert-butyl acetoacetate from the NEO Group, acid value < 5 mg KOH·g−1, hydroxyl value 40.7 mg KOH·g−1, moisture 0.048%, and acetoacetate groups 0.4242 mol·100 g−1. Acetoacetyled Lupranol 3300 (L3300_AA) was synthesized from Lupranol 3300 (L3300) by transesterification with tert-butyl acetoacetate, from BASF, acid value < 5 mg KOH·g−1, hydroxyl value 26.2 mg KOH·g−1, and moisture 0.021%, acetoacetate groups 0.4456 mol·100 g−1. BPAEDA, average Mn~512, contains 1000 ppm monomethyl ether hydroquinone as an inhibitor from Sigma-Aldrich; TMPTA contains 600 ppm monomethyl ether hydroquinone as an inhibitor and is of technical grade from Sigma-Aldrich; PETA contains 350 ppm monomethyl ether hydroquinone as an inhibitor and is technical grade from Sigma-Aldrich (St. Louis, MO, USA). 1,1,3,3-tetramethylguanidine (TMG), assay 99%, was purchased from Sigma-Aldrich. Catalyst and acrylates were directly used as delivered without further purification.

2.2. Solid Polymer Development

The polymerization was carried out by mixing the acrylate with the catalyst (TMG), adding acetoacetate, and stirring the mixture rapidly until a homogeneous medium was obtained. Mixing was about 30 s. The ratio of acetoacetate to acrylic groups was 1:2 mol. The amount of TMG was 1% of the total mass of the mixture. After mixing, the mixture was quickly poured into a 15 mL centrifuge tube and centrifuged for 1 min to remove air bubbles and obtain a homogeneous polymer monolith. Samples were made with a weight of 14 g and solidified in a tube at room temperature and atmospheric pressure for 2 min to 3 min. Polymer syntheses were repeated twice. Samples were tested from both syntheses.
Tensile samples were prepared by immediately pouring the polymerization mixture into a mold after centrifugation. Samples were tested after 24 h curing at room temperature.

2.3. Characterization Methods

FTIR. Polymer chemical structure was analyzed using FTIR data, which was obtained with a Thermo Scientific Nicolet iS50 spectrometer (Thermo Fisher Scientific, Waltham, MA, USA) at a resolution of 4 cm−1 (32 scans). FTIR data were collected using the attenuated total reflectance technique with diamond crystals. A sheet of the obtained polymer sample was pressed against the prism and analyzed.
DSC. Metler Toledo DSC 823e (Mettler Toledo, Greifensee, Switzerland) was used to obtain DSC data. STARe Software ver. 9.00 (Mettler Toledo, Greifensee, Switzerland) and OriginPro 2021 9.8.0.200 programs (Northampton, MA, USA) were used for data processing. Specimens weighing approximately 0.2–2.0 mg were cut from the polymer sample. A total of 5 mg of samples were placed in an aluminium sample pan with an accuracy of ±0.5 mg. A Mettler Toledo sealing press for crucibles was used to crimp covers on DSC pans of aluminium. The test was organized in two cycles. The start temperature of the first cycle was 25 °C with heating up to 180 °C, followed by cooling to −50 °C. The second heating cycle was from −50 °C to 180 °C, and the temperature was reduced to 25 °C in the second cooling cycle. The heating rate used was 10 °C·min−1. Two samples from each series were tested for DSC.
TGA. Thermogravimetric data were obtained using Discovery TGA equipment (TA instruments, New Castle, DE, USA). Data processing was performed using the OriginPro 2021 9.8.0.200 and TA Instruments TRIOS version #5.0.0.44608 software. A 10 mg sample consisting of similar size pieces (0.5–2.0 mg) was placed on a platinum scale sample pan with an accuracy of ±0.5 mg. Then, the sample on the sample pan was heated in a nitrogen atmosphere at 10 °C·min−1 in a temperature range between 30 °C and 700 °C. Two samples from each series were tested for TGA.
DMA. DMA was carried out with Mettler Toledo DMA/SDTA861e (Mettler Toledo, Switzerland). The temperature range was from −100 °C to 180 °C, the ramp rate was 3 °C·min−1, the frequency was 1 Hz, amplitude was 30 μm, and maximal force was 5 N. The compression oscillation mode was used. Polymer monolith samples with a diameter of 13 mm and a height of 7 mm, with an accuracy of ±0.2 mm, were used for DMA tests. Two samples from each series were tested for DMA.
The cross-link density was calculated from the data obtained in the DMA test according to Equation (1):
ν e = E 3 R T
where νe is the cross-link density (mol·cm−3), E′ is the tensile storage modulus (MPa), R is the gas constant (8.314 m3·Pa·K−1·mol−1), and T is the temperature corresponding to the storage modulus value into the rubbery plateau (K) [52,53,54,55].
Knowing density ρ (g·cm−3), the molecular weight between cross-links (Mc (g·mol−1)) was calculated according to Equation (2) [56].
M c = 3 R T ρ E r u b b e r y
Tensile modulus (Young’s) and compressive strength were determined from tests carried out at room temperature (22 °C) using a universal machine, an ElectroPuls E3000 from Instron (Norwood, Norfolk, MA, USA), following the ISO 14125:1998 standard. The sample size was 2.5 mm × 4 mm × 25 mm. The deformation rate for all samples was 2 mm·min−1, and at least four samples from each series were tested.

3. Results and Discussion

A schematic overview of polymer syntheses performed, including information on the synthesis steps of the Michael donor monomers, is provided in Figure 2. Previously developed acetoacetates of two tall oil-based polyols (EIRTOFA_DB_AA and EIRTOFA_TMP_AA) and two commercial polyols (NEO380_AA and L3300_AA) were used to obtain polymers by the Michael reaction with three acrylates with different functionalities—BPAEDA, TMPTA, and PETA. Monomers of different degrees of functionalities were used to demonstrate that it was possible to obtain a polymeric material from tall oil polyol acetoacetates with various acrylates. It also allowed the investigation of the effect of acrylate functionality on the properties of the obtained polymeric materials.

3.1. FTIR Analysis of Monoliths

The FTIR spectra of the film of obtained thin polymeric materials are shown in Figure 3. The FTIR spectra did not show very intensive peaks at ~1630 cm−1 and ~810 cm−1, which are characteristic of the absorbance of acrylic group vibration. This indicated that the content of free acrylic groups in polymers was low [57]. Thus, a relatively high conversion occurred.
Polymers derived from BPAEDA have two intense peaks (at 830 cm−1 and 1510 cm−1) corresponding to aromatic bonds characterizing bisphenol A in the FTIR spectrum. In the case of NEO380_AA-based polymers, an intense peak at 730 cm−1 was observed in the FTIR spectra, indicating the out-of-plane C-H bending of an aromatic ring (Figure 3c).

3.2. DSC Analysis of Monoliths

DSC heating curves are shown in Figure 4. The glass transition temperature can be very clearly identified in almost all DSC curves of the new polymeric materials. The exception was in the case of polymers from PETA acrylate. The glass transition temperature was less pronounced in the curves of these polymers by the used DSC analysis mode. For polymers synthesized using BPAEDA, the glass transition was clearly visible. The enthalpy relaxation peak also accompanied it.
For polymers obtained from commercial polyol acetoacetate (L3300_AA and NEO380_AA), the glass transition temperature was mainly influenced by the used acrylate. The glass transition temperature differed by less than 2 °C when polymers obtained with the same acrylates were compared (Figure 4c,d). Depending on the used acrylate, the glass transition temperature increased as follows: BPAEDA < TMPTA < PETA. The functionality of the used acrylate impacted the glass’s transition temperature. When acrylate with higher functionality was used to obtain polymers, the polymer exhibited higher glass transition temperature due to a higher cross-link density in the polymer matrix [58].
The same trend was observed for polymers obtained from acetoacetylated tall oil polyols (Figure 4a,b). The used tall oil polyol acetoacetate had a significant effect on the glass transition temperature of the obtained polymer. Higher glass transition temperature was observed for EIRTOFA_TMP_AA, which was synthesized using a higher functionality polyol.
In the case of commercial polyol-based acetoacetate polymers, the difference in glass transition temperatures was relatively small when considering the effect of the polyol used. Of course, the previously established correlation between acrylate functionality and the glass transition temperature remained. A higher glass transition temperature was observed for polymers in which acrylate with higher functionalities was used (Figure 4c,d).

3.3. TGA Analysis of Monoliths

TGA curves and their derivatives of polymers obtained from tall oil-based and commercial polyol acetoacetates are shown in Figure 5. As shown in Figure 5a,b, no significant weight loss was observed up to almost 300 °C for polymers derived from tall oil polyols; they were thermally stable. This is an advantage because, for example, classic polyurethane materials without special thermal stabilizer additives decompose in the temperature range from 200 °C to 300 °C [59,60]. Moreover, organic glass poly(methyl methacrylate) (PMMA) onset decomposition temperatures are under 300 °C [61,62]. In Table 1, data of temperatures are summarized representing the temperatures at which the following mass losses were achieved: 5% (Tm5%), 10% (Tm10%), 25% (Tm25%), and 50% (Tm50%). If the influence of used polyol on thermal properties of polymer material was compared, then polymers derived from TMP polyols had higher thermal stability.
On the other hand, if polymer materials were compared according to the used acrylate, then the most thermally stable polymers were obtained from BPAEDA acrylate. This could be explained by the aromatic structure of bisphenol A. It was observed that the temperature at which a 5% mass loss occurred was lower when acrylate with higher functionality was used.
The highest Tm5% was observed for polymers derived using difunctional BPAEDA, followed by trifunctional TMPTA. The lowest Tm5% was for polymers derived from tetrafunctional PETA. Thermal degradation continued up to almost 500 °C for all samples, after which no significant weight loss was observed.
The peaks indicate the point of the greatest rate of change on the weight loss curve (Figure 5). Two distinct peaks were observed for polymers obtained using BPAEDA as the acrylate. For other polymers, the similarity of the curves was determined by the used polyol and less by the acrylate.
Polymers from commercial polyols also showed similar characteristics in TGA mass-loss curves and their derivative curves (Figure 5c,d). Polymeric materials derived from tall oil polyols with BPAEDA acrylate had higher thermal stability (when comparing the temperature for weight loss of 5%) than commercial polyols with the same acrylate. The temperature depends significantly on the used acrylate for a weight loss of 5%, but this effect was less pronounced for weight losses at higher temperatures.

3.4. DMA Analysis of Monoliths

DMA test was performed to determine the dynamic mechanical properties of polymer materials. The obtained curves are shown in Figure 6. Based on the tanδ data, all samples had a quite homogeneous network structure because there was only one relatively narrow peak with a clear maximum, which appears in the graphs. These relationships between peak width, symmetry and structure, and the homogeneity of the structure have been described in several studies [20,63,64,65].
Wider peaks and with a lower maximum of tanδ value were observed in the case of the tall oil acetoacetate polymer with a tetrafunctional acrylate ester PETA (Figure 6a,b). This indicated a lower level of homogeneity in cross-link density. A symmetrical peak with the highest tanδ value was reached for all polymer samples obtained using the difunctional acrylate BPAEDA, indicating higher homogeneity of the cross-linked network.
The glass transition temperatures determined by the DMA method are summarized in Table 2. The glass transition temperatures obtained from DMA were compared with the glass transition temperatures of materials from DSC analysis. Although the glass transition temperatures were different, same trends between the glass transition temperatures were observed. The glass transition temperatures determined using the DMA method were higher than those determined using DSC analysis. DMA method has a higher sensitivity than DSC for detecting the glass transition temperature [66]. DMA is beneficial for polymers with hard-to-find glass-transition temperatures and heavily cross-linked polymers [67].
Cross-link density and molecular weight between cross-links were calculated from the DMA data and are presented in Table 2. Results showed that cross-link density and molecular weight between cross-links of the polymers were highly dependent on the functionality of the used acrylate and acetoacetate. If higher functionalized monomers were used, the cross-link density of the obtained polymer was higher, and the molecular weight between cross-links was smaller, which significantly affects the polymer’s mechanical properties.
The highest cross-link density (3.17 × 10−3 moles·cm−3) and the lowest molecular weight between cross-links (315 g·mol−1) were obtained from EIRTOFA_TMP_AA and PETA (tetrafunctional acrylate). It can be explained by the higher functionality of both—the Michael donor and the Michael acceptor compound. The lowest cross-link density and molecular weight between cross-links were for the polymers derived from the lower functionalized acetoacetate and difunctional acrylate—BPAEDA.

3.5. Tensile Tests

Values of tensile modulus (Young’s) and tensile strength for the samples are shown in Figure 7. Polymer monolith samples obtained by the reaction of polyol acetoacetates with tetraacrylate PETA showed the highest tensile modulus values. The results of tensile strength showed the same trend. This could be explained by the higher functionality of the acrylate and, thus, the higher cross-link density (see Table 2), resulting in the highest module value. A correlation between the functionality of acrylate used in polymer formulations and the tensile module values was observed: the higher functionality of the used acrylate yielded polymers with higher tensile module values.
The tensile strength for NEO380_AA_PETA polymer was 51 MPa. Relatively high modulus values were also obtained for polymers from tall oil from EIRTOFA_TMP_AA_TMPTA and EIRTOFA_TMP_AA_PETA, respectively, at 2250 MPa and 2370 MPa, and tensile strength was 40 MPa and 44 MPa. Slightly higher average tensile strength and modulus values were for the PETA acrylate, but they were lower for the TMPTA acrylate. These materials can be compared to PMMA, which had a tensile strength value of at least 55 MPa and a tensile modulus of about 2700 MPa [68,69]. Fleischer et al. prepared non-filled NIPU from hexamethylene diamine and glycerol cyclic carbonates: trimethylolpropane. This NIPU showed a tensile strength of 68 MPa and Young’s modulus of 2100 MPa [70]. The mechanical properties of obtained polymers are also comparable to polyurethane materials [71,72]. This shows that tall-oil-based polymer materials obtained by the Michael addition reaction are competitive and promising alternatives.
The highest tensile modulus values were for polymer materials obtained from triacrylate or tetraacrylate and NEO380_AA. A maximum average value of the modulus exceeds 2710 MPa for a NEO380_AA_PETA polymer. The polymers derived using BPAEDA had the lowest modulus value and tensile strength. These polymers (EIRTOFA_BD_AA_BPAEDA, NEO380_AA_BPAEDA, and L3300_AA_BPAEDA) exhibited characteristics similar to rubber.
The elongation at break correlates with the modulus value and the tensile strength. The higher modulus and tensile strength values, the smaller the elongation. The smallest elongations at breaks were 1.9%, 2.7%, and 2.9% for polymers EIRTOFA_TMP_AA_PETA, NEO380_AA_PETA, and L3300_AA_PETA, and the highest elongations at breaks were 55%, 53% and 47% for polymers EIRTOFA_BD_AA_BPAEDA, NEO380_AA_BPAEDA, and L3300_AA_BPAEDA, respectively.
The relationship between mechanical properties and the molecular weight between cross-links was observed. The lower the molecular weight between cross-links, the higher the cross-link density (see Table 2) and the correspondingly higher tensile modulus (Young’s) and tensile strength.
Results indicate that it is possible to obtain very different polymeric materials by changing the chemical structure of the used acrylate. Polymeric materials that are similar in properties to rubbers can be obtained by using lower functionality monomers. It is also possible to obtain materials that are similar in their properties even to organic glass [68,69,73]. By compiling the literature on similar materials and their characteristics, the polymer materials obtained in this study are competitive with several other alternative materials. The production of polymers from fatty-acid-based monomers by Michael reactions is relatively simple and is carried out in milder conditions compared to other methods, such as NIPU synthesis. Michael addition polymerizations can be realized in a short period of time, at room temperature, and at an atmospheric pressure. The chemicals used to synthesize polymers via Michael additions are less toxic than isocyanates, which is one of the main raw materials for polyurethane production. The mechanical and thermal properties of synthesized polymers by the Michael reaction are equal to or even better than materials shown in the literature. Moreover, the properties of polymers can be tailored by varying the monomers used for polymerization. For more information on other similar materials and their characteristics from the literature, see Supplementary Materials Table S1 “Characteristics (glass transition temperatures (Tg) by DMA, tensile (Young’s) modulus (E1), tensile strength and cross-link density (ve)) of selected polymers for comparison”.

4. Conclusions

Polymer materials were successfully obtained from the synthesized tall oil acetoacetates. The properties of synthesized bio-based polymer monoliths were comparable to polymeric materials obtained from commercially available acrylates. The modulus, tensile strength, and glass transition temperatures were determined for all obtained samples. The properties of the polymer were strongly dependent on the functionality of the monomers used. It was shown that a wide range of polymeric materials with different properties could be obtained by varying the functionality of the monomer. This study demonstrated that polymers can be synthesized with a wide range of mechanical properties, from rubber-like (E1 = 8 MPa and σmax = 4 MPa for EIRTOFA_BD_AA_BPAEDA) to close to organic glass and NIPUs (E1 = 2370 MPa and σmax = 44 MPa for EIRTOFA_TMP_AA_PETA). Glass transition temperatures varied from 21.0 °C for EIRTOFA_BD_AA_BPAEDA to 63.4 °C for EIRTOFA_TMP_AA_PETA obtained by DMA. The synthesized polymers were thermally stable up to 300 °C. Tall-oil-based polymers via the Michael nucleophilic 1,4-addition are an attractive alternative to poly(methyl methacrylate), conventional polyurethanes, NIPUs, and other fossil-based polymer materials. The polymer formulations developed in this study may be suitable for two-component polymer foams, coatings, resins, and composite matrices.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/polym14194068/s1, Table S1: Characteristics (glass transition temperatures (Tg) by DMA, tensile (Young’s) modulus (E1), tensile strength, and cross-link density (ve)) of the selected polymers for comparison. References [5,15,39,43,52,53,56,68,74,75] are cited in Supplementary Materials.

Author Contributions

Conceptualization, M.K.; data curation, R.P. and I.M.; formal analysis, R.P.; writing—original draft preparation, R.P. and I.M.; writing—review and editing, R.P., I.M., M.K. and A.F.; project administration, A.F.; software, R.P. and I.M.; supervision, M.K.; visualization R.P. and I.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Latvian Council of Science, project “High bio-based content thermoset polymer foam development from plant origin oils (Bio-Mer)”, project No. lzp-2020/1-0385. This publication was supported by Riga Technical University’s Doctoral Grant programme (DOK.OĶTI/21).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. European Commission. The European Green Deal; Cambridge University Press: Cambridge, UK, 2019. [Google Scholar]
  2. Rosenboom, J.G.; Langer, R.; Traverso, G. Bioplastics for a circular economy. Nat. Rev. Mater. 2022, 7, 117–137. [Google Scholar] [CrossRef] [PubMed]
  3. Cifarelli, A.; Boggioni, L.; Vignali, A.; Tritto, I.; Bertini, F.; Losio, S. Flexible polyurethane foams from epoxidized vegetable oils and a bio-based diisocyanate. Polymers 2021, 13, 612. [Google Scholar] [CrossRef] [PubMed]
  4. Sittinun, A.; Pisitsak, P.; Manuspiya, H.; Thiangtham, S.; Chang, Y.H.; Ummartyotin, S. Utilization of Palm Olein-Based Polyol for Polyurethane Foam Sponge Synthesis: Potential as a Sorbent Material. J. Polym. Environ. 2020, 28, 3181–3191. [Google Scholar] [CrossRef]
  5. Chen, R.; Zhang, C.; Kessler, M.R. Polyols and polyurethanes prepared from epoxidized soybean oil ring-opened by polyhydroxy fatty acids with varying oh numbers. J. Appl. Polym. Sci. 2015, 132, 41213. [Google Scholar] [CrossRef]
  6. Zhang, J.; Tang, J.J.; Zhang, J.X. Polyols prepared from ring-opening epoxidized soybean oil by a castor oil-based fatty diol. Int. J. Polym. Sci. 2015, 2015, 529235. [Google Scholar] [CrossRef]
  7. Coman, A.E.; Peyrton, J.; Hubca, G.; Sarbu, A.; Gabor, A.R.; Nicolae, C.A.; Iordache, T.V.; Averous, L. Synthesis and characterization of renewable polyurethane foams using different biobased polyols from olive oil. Eur. Polym. J. 2021, 149, 110363. [Google Scholar] [CrossRef]
  8. Liao, Y.-H.; Su, Y.-L.; Chen, Y.-C. The Influence of Neem Oil and Its Glyceride on the Structure and Characterization of Castor Oil-Based Polyurethane Foam. Polymers 2021, 13, 2020. [Google Scholar] [CrossRef] [PubMed]
  9. Abolins, A.; Pomilovskis, R.; Vanags, E.; Mierina, I.; Michalowski, S.; Fridrihsone, A.; Kirpluks, M. Impact of different epoxidation approaches of tall oil fatty acids on rigid polyurethane foam thermal insulation. Materials 2021, 14, 894. [Google Scholar] [CrossRef] [PubMed]
  10. Dechent, S.E.; Kleij, A.W.; Luinstra, G.A. Fully bio-derived CO2 polymers for non-isocyanate based polyurethane synthesis. Green Chem. 2020, 22, 969–978. [Google Scholar] [CrossRef]
  11. Sternberg, J.; Pilla, S. Materials for the biorefinery: High bio-content, shape memory Kraft lignin-derived non-isocyanate polyurethane foams using a non-toxic protocol. Green Chem. 2020, 22, 6922–6935. [Google Scholar] [CrossRef]
  12. Lindsay, C.D.; Timperley, C.M. TRPA1 and issues relating to animal model selection for extrapolating toxicity data to humans. Hum. Exp. Toxicol. 2020, 39, 14–36. [Google Scholar] [CrossRef]
  13. Kathalewar, M.; Sabnis, A.; D’Mello, D. Isocyanate free polyurethanes from new CNSL based bis-cyclic carbonate and its application in coatings. Eur. Polym. J. 2014, 57, 99–108. [Google Scholar] [CrossRef]
  14. Kathalewar, M.S.; Joshi, P.B.; Sabnis, A.S.; Malshe, V.C. Non-isocyanate polyurethanes: From chemistry to applications. RSC Adv. 2013, 3, 4110–4129. [Google Scholar] [CrossRef]
  15. Bähr, M.; Xie, W.; Łukaszewska, I.; Hebda, E.; Pielichowski, K. Recent advances in fabrication of non-isocyanate polyurethane-based composite materials. Materials 2021, 14, 3497. [Google Scholar] [CrossRef]
  16. Gomez-Lopez, A.; Elizalde, F.; Sardon, H. Trends in non-isocyanate polyurethane (NIPU) development. Chem. Commun. 2021, 57, 12254–12265. [Google Scholar] [CrossRef]
  17. Omer, R.A.; Hama, J.R.; Rashid, R.S.M. The Effect of Dextran Molecular Weight on the Biodegradable Hydrogel with Oil, Synthesized by the Michael Addition Reaction. Adv. Polym. Technol. 2015, 36, 120–127. [Google Scholar] [CrossRef]
  18. Mather, B.D.; Viswanathan, K.; Miller, K.M.; Long, T.E. Michael addition reactions in macromolecular design for emerging technologies. Prog. Polym. Sci. 2006, 31, 487–531. [Google Scholar] [CrossRef]
  19. Noordover, B.; Liu, W.; McCracken, E.; DeGooyer, B.; Brinkhuis, R.; Lunzer, F. Michael addition curable coatings from renewable resources with enhanced adhesion performance. J. Coat. Technol. Res. 2020, 17, 1123–1130. [Google Scholar] [CrossRef]
  20. Sonnenschein, M.F.; Werness, J.B.; Patankar, K.A.; Jin, X.; Larive, M.Z. From rigid and flexible foams to elastomers via Michael addition chemistry. Polymer 2016, 106, 128–139. [Google Scholar] [CrossRef]
  21. Yang, X.; Liu, H.; Qian, L.; Meng, Q.; Wu, H.; Li, Z.; Zhou, H. Surface functionalization of cellulose fibers via aza-Michael addition for CO2-assisted water remediation. Appl. Surf. Sci. 2021, 554, 149593. [Google Scholar] [CrossRef]
  22. Bhattacharjee, S.; Shaikh, A.A.; Ahn, W.S. Heterogeneous Aza-Michael Addition Reaction by the Copper-Based Metal–Organic Framework (CuBTC). Catal. Lett. 2021, 151, 2011–2018. [Google Scholar] [CrossRef]
  23. Su, G.; Thomson, C.J.; Yamazaki, K.; Rozsar, D.; Christensen, K.E.; Hamlin, T.A.; Dixon, D.J. A bifunctional iminophosphorane squaramide catalyzed enantioselective synthesis of hydroquinazolines: Via intramolecular aza-Michael reaction to α,β-unsaturated esters. Chem. Sci. 2021, 12, 6064–6072. [Google Scholar] [CrossRef] [PubMed]
  24. Mushtaque, M.; Avecilla, F.; Ahmad, I.; Alharbi, A.M.; Khan, P.; Ahamad, S.; Hassan, M.I. 5-Fluorouracil (5-FU)-based Aza-Michael addition product: A selective carbonic anhydrase IX inhibitor. J. Mol. Struct. 2021, 1231, 129977. [Google Scholar] [CrossRef]
  25. Alaneed, R.; Golitsyn, Y.; Hauenschild, T.; Pietzsch, M.; Reichert, D.; Kressler, J. Network formation by aza-Michael addition of primary amines to vinyl end groups of enzymatically synthesized poly (glycerol adipate). Polym. Int. 2021, 70, 135–144. [Google Scholar] [CrossRef]
  26. Gunay, U.S.; Cetin, M.; Daglar, O.; Hizal, G.; Tunca, U.; Durmaz, H. Ultrafast and efficient aza- and thiol-Michael reactions on a polyester scaffold with internal electron deficient triple bonds. Polym. Chem. 2018, 9, 3037–3054. [Google Scholar] [CrossRef]
  27. Li, Q.F.; Chu, S.; Li, E.; Li, M.; Wang, J.T.; Wang, Z. Lanthanide-based hydrogels with adjustable luminescent properties synthesized by thiol-Michael addition. Dye. Pigment. 2020, 174, 108091. [Google Scholar] [CrossRef]
  28. Chen, J.; Ma, X.; Edgar, K.J. A Versatile Method for Preparing Polysaccharide Conjugates via Thiol-Michael Addition. Polymers 2021, 13, 1905. [Google Scholar] [CrossRef]
  29. Cespedes, S.; Hand, R.A.; Chmel, N.; Moad, G.; Keddie, D.J.; Schiller, T.L. Enhanced properties of well-defined polymer networks prepared by a sequential thiol-Michael-radical thiol-ene (STMRT) strategy. Eur. Polym. J. 2021, 151, 110440. [Google Scholar] [CrossRef]
  30. Piñeiro-García, A.; Vega-Díaz, S.M.; Tristán, F.; Meneses-Rodríguez, D.; Labrada-Delgado, G.J.; Semetey, V. New insights in the chemical functionalization of graphene oxide by thiol-ene Michael addition reaction. FlatChem 2021, 26, 100230. [Google Scholar] [CrossRef]
  31. Lenardão, E.J.; Trecha, D.O.; Ferreira, P.d.C.; Jacob, R.G.; Perin, G. Green Michael Addition of Thiols to Electron Deficient Alkenes using KF/Alumina and Recyclable Solvent or Solvent-free Conditions. J. Braz. Chem. Soc. 2009, 20, 93–99. [Google Scholar] [CrossRef]
  32. Chatani, S.; Nair, D.P.; Bowman, C.N. Relative reactivity and selectivity of vinyl sulfones and acrylates towards the thiol-Michael addition reaction and polymerization. Polym. Chem. 2013, 4, 1048–1055. [Google Scholar] [CrossRef]
  33. Liu, N. New Polymers Synthesis by Organocatalyzed Step-Growth Polymerization of Aldehydic Monomers: Polyaldols, Linear Polybenzoin and Hyperbranched Polyacetals. Ph.D. Thesis, Université Sciences et Technologies-Bordeaux I, Gradignan, France, 2014. [Google Scholar]
  34. Williams, S.R.; Miller, K.M.; Long, T.E. Michael addition reaction kinetics of acetoacetates and acrylates for the formation of polymeric networks. Prog. React. Kinet. Mech. 2007, 32, 165–194. [Google Scholar] [CrossRef]
  35. Jiang, Q.; Zhang, Y.L.; Du, Y.; Tang, M.; Jiang, L.; Huang, W.; Yang, H.; Xue, X.; Jiang, B. Preparation of hyperbranched polymers by oxa-Michael addition polymerization. Polym. Chem. 2020, 11, 1298–1306. [Google Scholar] [CrossRef]
  36. Nising, C.F.; Bräse, S. The oxa-Michael reaction: From recent developments to applications in natural product synthesis. Chem. Soc. Rev. 2008, 37, 1218–1228. [Google Scholar] [CrossRef]
  37. González, G.; Fernández-Francos, X.; Serra, À.; Sangermano, M.; Ramis, X. Environmentally-friendly processing of thermosets by two-stage sequential aza-Michael addition and free-radical polymerization of amine-acrylate mixtures. Polym. Chem. 2015, 6, 6987–6997. [Google Scholar] [CrossRef]
  38. Farmer, T.J.; Comerford, J.W.; Pellis, A.; Robert, T. Post-polymerization modification of bio-based polymers: Maximizing the high functionality of polymers derived from biomass. Polym. Int. 2018, 67, 775–789. [Google Scholar] [CrossRef]
  39. Cao, Z.; Gao, F.; Zhao, J.; Wei, X.; Cheng, Q.; Zhong, J.; Lin, C.; Shu, J.; Fu, C.; Shen, L. Bio-based coating materials derived from acetoacetylated soybean oil and aromatic dicarboxaldehydes. Polymers 2019, 11, 1809. [Google Scholar] [CrossRef]
  40. He, X.; Zhong, J.; Cao, Z.; Wang, J.; Gao, F.; Xu, D.; Shen, L. An exploration of the Knoevenagel condensation to create ambient curable coating materials based on acetoacetylated castor oil. Prog. Org. Coat. 2019, 129, 21–25. [Google Scholar] [CrossRef]
  41. Trevino, A.S.; Trumbo, D.L. Acetoacetylated castor oil in coatings applications. Prog. Org. Coat. 2002, 44, 49–54. [Google Scholar] [CrossRef]
  42. Xu, D.; Cao, Z.; Wang, T.; Zhong, J.; Zhao, J.; Gao, F.; Luo, X.; Fang, Z.; Cao, J.; Xu, S.; et al. An ambient-cured coating film obtained via a Knoevenagel and Michael addition reactions based on modified acetoacetylated castor oil prepared by a thiol-ene coupling reaction. Prog. Org. Coat. 2019, 135, 510–516. [Google Scholar] [CrossRef]
  43. Zuo, H.; Cao, Z.; Shu, J.; Xu, D.; Zhong, J.; Zhao, J.; Wang, T.; Chen, Y.; Gao, F.; Shen, L. Effect of structure on the properties of ambient-cured coating films prepared via a Michael addition reaction based on an acetoacetate-modified castor oil prepared by thiol-ene coupling. Prog. Org. Coat. 2019, 135, 27–33. [Google Scholar] [CrossRef]
  44. Naga, N.; Satoh, M.; Magara, T.; Ahmed, K.; Nakano, T. Synthesis of gels by means of Michael addition reaction of multi-functional acetoacetate and diacrylate compounds and their application to ionic conductive gels. J. Appl. Polym. Sci. 2021, 59, 2129–2139. [Google Scholar] [CrossRef]
  45. Park, S.-H.; Alammar, A.; Fulop, Z.; Pulido, B.A.; Nunes, S.P.; Szekely, G. Hydrophobic thin film composite nanofiltration membranes derived solely from sustainable sources. Green Chem. 2021, 23, 1175–1184. [Google Scholar] [CrossRef]
  46. Sinha, J.; Soars, S.; Bowman, C.N. Enamine Organocatalysts for the Thiol-Michael Addition Reaction and Cross-Linking Polymerizations. Macromolecules 2021, 54, 1693–1701. [Google Scholar] [CrossRef]
  47. Williams, S.R.; Mather, B.D.; Miller, K.M.; Long, T.E. Novel michael addition networks containing urethane hydrogen bonding. J. Polym. Sci. Part A Polym. Chem. 2007, 45, 4118–4128. [Google Scholar] [CrossRef]
  48. Bureau, I. Functionalized Oligomers, WO 2014/052081 A2, 4 April 2014.
  49. Ozturk, G.; Long, T.E. Michael addition for crosslinking of poly(caprolactone)s. J. Polym. Sci. Part A Polym. Chem. 2009, 47, 5437–5447. [Google Scholar] [CrossRef]
  50. Balotupi, T.; Lanta, T.; Rzwpk, D.A.N.P.K. Non-Isocyanate Rigid Polymer Foams By Carbon-Michael Addition, and Foaming Process. WO 2013/101682 Al, 4 July 2013. [Google Scholar]
  51. Zhaoo, M.Y.; Hsu, C.-P.; Voeks, S.L.; Landtiser, R. Acetoacetyl Thermosetting Resin for Zero Voc Gel Coat. WO 2013/132077, 12 September 2013. [Google Scholar]
  52. Kim, T.H.; Kim, M.; Lee, W.; Kim, H.G.; Lim, C.S.; Seo, B. Synthesis and characterization of a polyurethane phase separated to nano size in an epoxy polymer. Coatings 2019, 9, 319. [Google Scholar] [CrossRef]
  53. Jena, K.K.; Raju, K.V.S.N. Synthesis and characterization of hyperbranched polyurethane hybrids using tetraethoxysilane (TEOS) as cross-linker. Ind. Eng. Chem. Res. 2008, 47, 9214–9224. [Google Scholar] [CrossRef]
  54. Hill, L.W. Calculation of crosslink density in short chain networks. Prog. Org. Coat. 1997, 31, 235–243. [Google Scholar] [CrossRef]
  55. Zhang, C.; Madbouly, S.A.; Kessler, M.R. Biobased polyurethanes prepared from different vegetable oils. ACS Appl. Mater. Interfaces 2015, 7, 1226–1233. [Google Scholar] [CrossRef] [PubMed]
  56. Barszczewska-Rybarek, I.M.; Korytkowska-Wałach, A.; Kurcok, M.; Chladek, G.; Kasperski, J. DMA analysis of the structure of crosslinked poly(methyl methacrylate)s. Acta Bioeng. Biomech. 2017, 19, 47–53. [Google Scholar] [CrossRef]
  57. Omer, R.A.; Hughes, A.; Hama, J.R.; Wang, W.; Tai, H. Hydrogels from dextran and soybean oil by UV photo-polymerization. J. Appl. Polym. Sci. 2015, 132, 41446. [Google Scholar] [CrossRef]
  58. Bermejo, J.S.; Ugarte, C.M. Influence of cross-linking density on the glass transition and structure of chemically cross-linked PVA: A molecular dynamics study. Macromol. Theory Simul. 2009, 18, 317–327. [Google Scholar] [CrossRef]
  59. Amado, J.C.Q. Thermal Resistance Properties of Polyurethanes and its Composites: A Short Review. J. Res. Updat. Polym. Sci. 2019, 8, 66–84. [Google Scholar] [CrossRef]
  60. Shufen, L.; Zhi, J.; Kaijun, Y.; Shuqin, Y.; Chow, W.K. Studies on the thermal behavior of polyurethanes. Polym.-Plast. Technol. Eng. 2006, 45, 95–108. [Google Scholar] [CrossRef]
  61. Zhang, F.A.; Lee, D.K.; Pinnavaia, T.J. PMMA-mesocellular foam silica nanocomposites prepared through batch emulsion polymerization and compression molding. Polymer 2009, 50, 4768–4774. [Google Scholar] [CrossRef]
  62. Rajkumar, T.; Muthupandiyan, N.; Vijayakumar, C.T. Synthesis and investigation of thermal properties of PMMA-maleimide-functionalized reduced graphene oxide nanocomposites. J. Thermoplast. Compos. Mater. 2020, 33, 85–96. [Google Scholar] [CrossRef]
  63. Hassan, M.K.; Tucker, S.J.; Abukmail, A.; Wiggins, J.S.; Mauritz, K.A. Polymer chain dynamics in epoxy based composites as investigated by broadband dielectric spectroscopy. Arab. J. Chem. 2016, 9, 305–315. [Google Scholar] [CrossRef]
  64. Dave, V.J.; Patel, H.S. Synthesis and characterization of interpenetrating polymer networks from transesterified castor oil based polyurethane and polystyrene. J. Saudi Chem. Soc. 2013, 21, 18–24. [Google Scholar] [CrossRef]
  65. Chen, S.; Wang, Q.; Pei, X.; Wang, T. Dynamic mechanical properties of castor oil-based polyurethane/epoxy graft interpenetrating polymer network composites. J. Appl. Polym. Sci. 2010, 116, 1144–1151. [Google Scholar] [CrossRef]
  66. Hagen, R.; Salmén, L.; Lavebratt, H.; Stenberg, B. Comparison of dynamic mechanical measurements and Tg determinations with two different instruments. Polym. Test. 1994, 13, 113–128. [Google Scholar] [CrossRef]
  67. Gracia-Fernández, C.A.; Gómez-Barreiro, S.; López-Beceiro, J.; Tarrío Saavedra, J.; Naya, S.; Artiaga, R. Comparative study of the dynamic glass transition temperature by DMA and TMDSC. Polym. Test. 2010, 29, 1002–1006. [Google Scholar] [CrossRef]
  68. Xie, W.; Guo, S.; Liu, Y.; Chen, R.; Wang, Q. Organic-inorganic hybrid strategy based on ternary copolymerization to prepare flame retardant poly (methyl methacrylate) with high performance. Compos. Part B Eng. 2020, 203, 108437. [Google Scholar] [CrossRef]
  69. Ali, U.; Karim, K.J.B.A.; Buang, N.A. A Review of the Properties and Applications of Poly (Methyl Methacrylate) (PMMA). Polym. Rev. 2015, 55, 678–705. [Google Scholar] [CrossRef]
  70. Fleischer, M.; Blattmann, H.; Mülhaupt, R. Glycerol-, pentaerythritol- and trimethylolpropane-based polyurethanes and their cellulose carbonate composites prepared via the non-isocyanate route with catalytic carbon dioxide fixation. Green Chem. 2013, 15, 934–942. [Google Scholar] [CrossRef]
  71. Abolins, A.; Yakushin, V.; Vilsone, D. Properties of polyurethane coatings based on linseed oil phosphate ester polyol. J. Renew. Mater. 2018, 6, 737–745. [Google Scholar] [CrossRef]
  72. Yakushin, V.; Abolins, A.; Vilsone, D.; Sevastyanova, I. Polyurethane coatings based on linseed oil phosphate ester polyols with intumescent flame retardants. Fire Mater. 2019, 43, 92–100. [Google Scholar] [CrossRef]
  73. Ishigami, A.; Watanabe, K.; Kurose, T.; Ito, H. Physical and morphological properties of tough and transparent PMMA-based blends modified with polyrotaxane. Polymers 2020, 12, 1790. [Google Scholar] [CrossRef] [PubMed]
  74. Zhu, Y.; Gao, F.; Zhong, J.; Shen, L.; Lin, Y. Renewable Castor Oil and DL-Limonene Derived Fully Bio-Based Vinylogous Urethane Vitrimers. Eur. Polym. J. 2020, 135, 109865. [Google Scholar] [CrossRef]
  75. Bähr, M.; Mülhaupt, R. Linseed and Soybean Oil-Based Polyurethanes Prepared via the Non-Isocyanate Route and Catalytic Carbon Dioxide Conversion. Green Chem. 2012, 14, 483–489. [Google Scholar] [CrossRef]
Figure 1. Mechanism of carbon–carbon Michael nucleophilic 1,4-addition reactions.
Figure 1. Mechanism of carbon–carbon Michael nucleophilic 1,4-addition reactions.
Polymers 14 04068 g001
Figure 2. A schematic overview of polymer syntheses and synthesis of Michael donor monomers.
Figure 2. A schematic overview of polymer syntheses and synthesis of Michael donor monomers.
Polymers 14 04068 g002
Figure 3. FTIR spectra of polymer monoliths from synthesized acetoacetates and BPAEDA, TMPTA, and PETA acrylates. (a) EIRTOFA_BD_AA polymers; (b) EIRTOFA_TMP_AA polymers; (c) NEO380_AA polymers; (d) L3300_AA polymers.
Figure 3. FTIR spectra of polymer monoliths from synthesized acetoacetates and BPAEDA, TMPTA, and PETA acrylates. (a) EIRTOFA_BD_AA polymers; (b) EIRTOFA_TMP_AA polymers; (c) NEO380_AA polymers; (d) L3300_AA polymers.
Polymers 14 04068 g003
Figure 4. DSC curves of polymer monoliths from synthesized acetoacetates and BPAEDA, TMPTA, and PETA acrylates. (a) EIRTOFA_BD_AA polymers; (b) EIRTOFA_TMP_AA polymers; (c) NEO380_AA polymers; (d) L3300_AA polymers.
Figure 4. DSC curves of polymer monoliths from synthesized acetoacetates and BPAEDA, TMPTA, and PETA acrylates. (a) EIRTOFA_BD_AA polymers; (b) EIRTOFA_TMP_AA polymers; (c) NEO380_AA polymers; (d) L3300_AA polymers.
Polymers 14 04068 g004
Figure 5. TGA mass loss curve of polymer monoliths from synthesized acetoacetates and BPAEDA, TMPTA, PETA acrylates, and derivative of the mass loss. (a) EIRTOFA_BD_AA polymers; (b) EIRTOFA_TMP_AA polymers; (c) NEO380_AA polymers; (d) L3300_AA polymers.
Figure 5. TGA mass loss curve of polymer monoliths from synthesized acetoacetates and BPAEDA, TMPTA, PETA acrylates, and derivative of the mass loss. (a) EIRTOFA_BD_AA polymers; (b) EIRTOFA_TMP_AA polymers; (c) NEO380_AA polymers; (d) L3300_AA polymers.
Polymers 14 04068 g005
Figure 6. Tanδ curves of polymer monoliths from synthesized acetoacetates and BPAEDA, TMPTA, and PETA acrylates. (a) Polymers from EIRTOFA_BD_AA; (b) polymers from EIRTOFA_TMP_AA; (c) polymers from NEO380_AA_TMPTA; (d) polymers from L3300_AA.
Figure 6. Tanδ curves of polymer monoliths from synthesized acetoacetates and BPAEDA, TMPTA, and PETA acrylates. (a) Polymers from EIRTOFA_BD_AA; (b) polymers from EIRTOFA_TMP_AA; (c) polymers from NEO380_AA_TMPTA; (d) polymers from L3300_AA.
Polymers 14 04068 g006
Figure 7. Tensile modulus (Young’s) (E1) and tensile strength (σmax) for the polymers from tall oil polyol or commercial polyol-based acetoacetates and BPAEDA, TMPTA, and PETA acrylates.
Figure 7. Tensile modulus (Young’s) (E1) and tensile strength (σmax) for the polymers from tall oil polyol or commercial polyol-based acetoacetates and BPAEDA, TMPTA, and PETA acrylates.
Polymers 14 04068 g007
Table 1. Weight loss dependence on temperature of synthesized polymers from tall oil polyol or commercial polyol-based acetoacetates and BPAEDA, TMPTA, and PETA acrylates.
Table 1. Weight loss dependence on temperature of synthesized polymers from tall oil polyol or commercial polyol-based acetoacetates and BPAEDA, TMPTA, and PETA acrylates.
SampleFirst Onset, °CTm5%, °CTm10%, °CTm25%, °CTm50%, °CResidue, %
Tall oil-basedEIRTOFA_BD_AA_BPAEDA349.0342.0361.5388.1417.02.0
EIRTOFA_BD_AA_TMPTA342.2318.3341.8369.0396.21.8
EIRTOFA_BD_AA_PETA340.2297.2332.9365.2396.12.6
EIRTOFA_TMP_AA_BPAEDA360.9351.2370.2391.8416.73.0
EIRTOFA_TMP_AA_TMPTA354.1330.9356.3385.8415.32.6
EIRTOFA_TMP_AA_PETA344.7300.3337.7377.7410.02.8
Commercial
polyols-based
L3300_AA_BPAEDA342.5331.3349.2368.5394.20.1
L3300_AA_TMPTA418.8320.8339.0362.8397.42.1
L3300_AA_PETA332.3309.7331.2356.6393.02.7
NEO380_AA_BPAEDA347.1331.6352.7378.1406.32.1
NEO380_AA_TMPTA341.0321.0344.5374.0407.36.0
NEO380_AA_PETA347.5317.0345.4376.5411.38.2
Table 2. Glass transition temperatures (Tg), cross-link density (νe) and molecular weight between cross-links (Mc) from DMA tests and density (ρ) of polymer monoliths from tall oil polyol or commercial polyol-based acetoacetates and BPAEDA, TMPTA, and PETA acrylate.
Table 2. Glass transition temperatures (Tg), cross-link density (νe) and molecular weight between cross-links (Mc) from DMA tests and density (ρ) of polymer monoliths from tall oil polyol or commercial polyol-based acetoacetates and BPAEDA, TMPTA, and PETA acrylate.
SampleTemperature Range, °C
(Tanδ > 0.06 °C)
Max. TanδTg, °Cνe, moles·cm−3ρ,
g·cm−3
Mc, g·mol−1
Tall oil-basedEIRTOFA_BD_AA_BPAEDA2.9–36.41.2121.00.59 × 10−31.1671695
EIRTOFA_BD_AA_TMPTA2.0–50.70.7334.61.98 × 10−31.211505
EIRTOFA_BD_AA_PETA−20.0–64.50.4540.42.67 × 10−31.219375
EIRTOFA_TMP_AA_BPAEDA21.9–54.51.0740.51.89 × 10−31.116529
EIRTOFA_TMP_AA_TMPTA32.6–68.90.7053.02.04 × 10−31.177490
EIRTOFA_TMP_AA_PETA52.2–80.20.2763.43.17 × 10−31.235315
Commercial
polyols-based
NEO380_AA_BPAEDA9.9–43.91.3028.50.93 × 10−31.2571075
NEO380_AA_TMPTA22.1–56.80.8138.72.04 × 10−31.328490
NEO380_AA_PETA20.8–62.70.7245.32.55 × 10−31.330392
L3300_AA_BPAEDA1.0–36.51.1824.51.13 × 10−31.171885
L3300_AA_TMPTA22.5–55.10.8039.72.20 × 10−31.202455
L3300_AA_PETA28.5–57.50.4446.62.70 × 10−31.247369
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pomilovskis, R.; Mierina, I.; Fridrihsone, A.; Kirpluks, M. Bio-Based Polymer Developments from Tall Oil Fatty Acids by Exploiting Michael Addition. Polymers 2022, 14, 4068. https://doi.org/10.3390/polym14194068

AMA Style

Pomilovskis R, Mierina I, Fridrihsone A, Kirpluks M. Bio-Based Polymer Developments from Tall Oil Fatty Acids by Exploiting Michael Addition. Polymers. 2022; 14(19):4068. https://doi.org/10.3390/polym14194068

Chicago/Turabian Style

Pomilovskis, Ralfs, Inese Mierina, Anda Fridrihsone, and Mikelis Kirpluks. 2022. "Bio-Based Polymer Developments from Tall Oil Fatty Acids by Exploiting Michael Addition" Polymers 14, no. 19: 4068. https://doi.org/10.3390/polym14194068

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop