Next Article in Journal
A Dichroic Beamsplitter for the Laser Protection of Infrared Detectors
Previous Article in Journal
Streptomyces-Derived Bioactive Pigments: Ecofriendly Source of Bioactive Compounds
Previous Article in Special Issue
Marginal Adaptation and Internal Fit of 3D-Printed Provisional Crowns and Fixed Dental Prosthesis Resins Compared to CAD/CAM-Milled and Conventional Provisional Resins: A Systematic Review and Meta-Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Characterization of Urethane Acrylate Resin Based on 1,3-Propanediol for Coating Applications

Department of Chemical Organic Technology and Polymeric Materials, Faculty of Chemical Technology and Engineering, West Pomeranian University of Technology in Szczecin, Piastów Ave. 42, 71-065 Szczecin, Poland
*
Authors to whom correspondence should be addressed.
Coatings 2022, 12(12), 1860; https://doi.org/10.3390/coatings12121860
Submission received: 9 November 2022 / Revised: 24 November 2022 / Accepted: 28 November 2022 / Published: 30 November 2022
(This article belongs to the Special Issue Surface Modifications and Coatings for Implantable Biomaterials)

Abstract

:
The significant development of industry and the growing demand for renewable fuels lead to the accumulation of massive amounts of glycerol as a by-product. Scientists have been trying to use this product as a raw material for several years. One of its uses is in the acquirement of 1,3-propanediol (PDO). This work presents studies on the synthesis of two new urethane acrylate resins obtained from 1,3-propanediol and urethane acrylate oligomers containing isocyanate groups in each molecule. The method for obtaining the resins was presented, considering various conditions of synthesis, i.e., the structure of the carbon chain of the oligomer used, the molar ratio of the reactants, or the use of solvents. The reactions were monitored in real time by FTIR until the disappearance of the NCO groups. Then, polymer films were prepared from the obtained products and cured using UV radiation or thermally. The obtained coatings were tested in terms of the photopolymerization kinetics and properties of the cured coatings. Resultantly, the obtained bio-sourced coatings were observed to be characterized by good functional properties and a short curing duration, both with the use of UV radiation-based or thermal curing. These types of resins obtained using a bioproduct can be employed as eco-friendly film-forming products in the coating industry for many applications. In particular, due to their potential for dual curing methods (UV or thermal), these resins can be used on three-dimensional surfaces, i.e., those in which there is a possibility of the insufficient availability of UV radiation.

Graphical Abstract

1. Introduction

Due to the negative impact of the fossil fuel economy on the environment, the production of products from renewable energy sources is growing rapidly. Scientists and industry are also trying to care for the environment by looking for new solutions and new ways to obtain compounds that we know to be products made from fossil fuels [1,2]. The same is the case with acrylic acid and its esters. They can be obtained from both petrochemical and renewable resources [3,4]. Acrylic acid and its esters are the main building blocks for many industrial oligomers and polymers [5]. Despite scientists’ successes in the laboratory described in the literature, the actual implementation of bio-raw materials is very limited due to the high costs and limited availability of bio-raw materials [6,7].
In recent years, scientists have synthesized bio-based polyurethane resins [8,9,10] or acrylate resins based on bioproducts for use in the coating industry using ultraviolet radiation technology. A major limitation for the industry is the low availability of bio acrylic acid, but quite promising routes for the synthesis and polymerization of such monomers from biomass have already been developed [11]. Triglycerides from vegetable oils as well as furfural from lignocellulosic biomass are common routes for acrylate monomer synthesis from bio-based raw materials [12]. The modification of the coatings’ properties is dependent on the type of biomass: when vegetable oils are used, more flexible structures are obtained, while those made of furan provide compact structures with higher mechanical strength and stiffness [13]. For example, the combination of lauryl acrylate, methyl acrylate, and isobornyl acrylate (bio-based) resulted in a copolymer capable of being photopolymerized in thin films by UV cyclodimerization [14]. The literature also describes alkoxy-butenolide monomers obtained from furfuryl alcohol with different side chain lengths, which resulted in thermal resistance and high hardness [15]. Scientists are now also comparing acrylates obtained from fossil resources and those obtained from raw materials of a natural origin. The latter show higher mechanical values and lower viscosity [16]. In the case of the acrylate monomers obtained from plant monomers, they are characterized by worse thermal stability and weaker mechanical properties [17], but the final properties depend on the number of double bonds leading to different cross-linking densities [17,18]. It is also known from the literature that diluents with higher functionality and/or with short monomeric chains improve mechanical properties [13]. For longer bio-based chains with unsatisfactory mechanical resistance, scientists add acrylated sucrose cross-linkers [19] or diluents made of cardanol oil [20,21,22], linseed oil [23,24], or castor oil [25].
The significant development of industry and the growing demand for renewable fuels are accumulating massive amounts of glycerol as a by-product. Scientists have been trying to use this product as a raw material for several years [26]. One of its uses is for the acquirement of 1,3-propanediol (PDO). Due to its numerous advantages, but above all, the great possibility of its modification, this compound is gaining increasing import both in scientific research and industry [27]. This monomer can be synthetically obtained from acrolein or ethylene oxide, but these methods have significant production costs. This diol can also be obtained microbiologically from the glucose fermentation process, which lowers the negative impacts on the environment [28].
The compound 1,3-propanediol is highly modifiable and is highly in reactive in polycondensation reactions (due to the presence of hydroxyl groups), which makes it conducive for obtaining polyurethanes, polyesters, or polyethers. It dissolves well in polar solvents (e.g., water, esters, ethers, alcohols, and others) [29,30]. It is popularly used both as an extension and the base compound of a chain, and its task is to improve the mechanical properties and increase thermal resistance. It can also be used as an additive to the final product that improves the physical and mechanical properties of adhesives, laminates, cosmetics, resins, powder coatings, and detergents [31].
In this work, the modification of two isocyanate-bearing urethane acrylate oligomers with bio-sourced 1,3-propanediol, thereby enabling the acquirement of sustainable urethane acrylate resins containing one or two acrylic groups, has been reported. First, the optimization of the modification procedure was followed by a kinetics study, where the molar ratio and solvent of reaction were studied with regard to the chemical structure of the urethane acrylate oligomer. Subsequently, coatings were prepared from the obtained resins and cured using UV radiation or thermally. The samples were tested for the kinetics of the curing process and the properties of the cured coatings. This work is a continuation of the research on photo-curable coatings based on acrylate oligomers obtained from bio-based raw materials. Studies on acrylated epoxidized soybean oil (AESO) have previously been published [32], while this paper presents novel urethane acrylate resins using a bio-based raw material.

2. Materials and Methods

2.1. Materials

The following raw materials were used for the studies: two isocyanate-bearing urethane acrylates, namely, Ebecryl 4141 (2DB-2NCO; Allnex, The Netherlands) and Ebecryl 4396 (1DB-2NCO; Allnex, The Netherlands); 1,3-propanediol (PDO; TILAMAR® PDO with NØØVISTATM, 100% bio-sourced); dibutyltin dilaurate (DBTDL; 95%, Alfa Aesar, Germany) as a catalyst for the reaction; ethyl(2,4,6-trimethylbenzoyl)-phenyl phosphinate (Omnirad TPOL, IGM Resins, The Netherlands) as a photoinitiator (PI); and benzoyl peroxide (Merck, Darmstadt, Germany) as an initiator (I). Table 1 shows the characteristics of the reagents used in the synthesis of the bio-resin and considers their functionality. All chemicals were employed as received.

2.2. Synthesis of Sustainable Urethane Acrylate Resins

In this work, two urethane acrylate (UA) resins were prepared. The sustainable UA resins were obtained by the addition of 1,3-propanediol (PDO) to urethane acrylate oligomers containing two or one acrylate group and two isocyanate groups (2DB-2NCO or 1DB-2NCO). The synthesis was carried out in a 250 mL 3-neck glass reactor (equipped with a thermometer, a condenser, a nitrogen inlet, and a mechanical stirrer), into which isocyanate-bearing UA in a form without or with a solvent (acetone or DCM, 60% in relation to the weight of the entire charge) was introduced. Then, hydroquinone (0.0075 wt.% based on total batch weight) was transferred into the reactor as a radical scavenger at room temperature. Lastly, 1,3-propanediol (0.08 mol relative to 0.04 moles of UA used) and catalyst in the amount of 0.0005 wt.% (relative to the mass of resin and PDO) were added. All chemicals were used as received, in a single dose, and without dissolving. The reaction mixture was heated to 30 °C with vigorous stirring (120 rpm) using an oil bath. Once a homogenous mixture was obtained, the reaction was monitored over time by Fourier-transformed infrared (FTIR) spectroscopy and was completed when the NCO groups observed at length of 2268 cm−1 disappeared. The general structures of the obtained UA resins, together with the expected products, are presented in Figure 1.

2.3. The FTIR Spectroscopy

The Fourier-transformed infrared (FTIR) spectroscopy in the attenuated total reflectance mode (ATR) was carried out on a Nicolet iS5 instrument (Thermo Electron Corporation, Waltham, MA, USA). Sixteen scans were averaged for each sample in the range of 4000–400 cm−1 at room temperature. This spectroscopic technique permits in situ monitoring of the chemical processes via mimicking the disappearance of the characteristic bands of the reactive compounds. Thus, this method was used to monitor the addition of 1,3-propanediol (PDO) to urethane acrylate oligomers containing two or one acrylate group and two isocyanate groups in real time by observing the disappearance of the peak at 2268 cm−1 responsible for the presence of the NCO groups. For this purpose, samples were taken from the reactor during the course of the reaction at intervals of half an hour. In addition, the series real-time IR (RT-IR) were used to determine the conversion of the double bonds at 810 cm−1 or OH groups at 3366 cm−1. The photoreactive mixture was placed in the mold made from glass slides and spacers of 15 mm in diameter and 0.5 in thickness. The samples were placed in the compartment of the Fourier transform infrared spectrometer and were simultaneously exposed to a UV radiation source (mercury UV lamp, Indigo, Poland, 36 W, 280–400 nm, 10 mW/cm2) or heated using a spectrophotometer attachment at 100 °C (GladiATR ™ with diamond ATR, PIKE Technologies, Fitchburg, WI, USA) and the IR analysis light beam. The absorbance change of the acrylate double bond (C=C) peak area at 810 cm−1 [33] was correlated to the extent of polymerization. The following relation can express the degree of conversion (DC): DC (%) = (A0 − At)·100/A0, where A0 is the initial peak area before the irradiation and At is the peak area at time t. The photopolymerization rate (Rp) was calculated by the following relation: Rp = dDC/dt, where t is the time of the irradiation [34].

2.4. Viscosity

The viscosity of the obtained resins was determined with a Bohlin Visco 88 (Malvern Panalytical, Malvern, UK) viscometer. The measurement was carried out at the temperature of 20 °C using the C14 geometry at the speed of 20 rpm. The presented viscosity values represent the average values made from two measurements, in which the error did not exceed 3%.

2.5. Preparation of Coating Compositions and Curing Conditions

The coating compositions have been formulated using synthesized sustainable urethane acrylates and 3 wt% of photo initiator (ethyl(2,4,6-trimethylbenzoyl)-phenyl phosphinate) and/or thermal initiator (benzoyl peroxide). The components were stirred together under dark conditions until a homogeneous mixture was obtained (approximately 5 min). Subsequently, the curing solution was applied to the glass substrates employing a gap applicator (120 μm). The polymer films were cured using UV radiation (at room temperature and irradiated under UV light with an intensity of 200 mW/cm2, under UV lamp, Aktiprint-mini 18-2, type: UN50029, Technigraf GmbH, Grävenwiesbach, Germany) and/or elevated temperature (100 °C, Natural air circulation dryer ED 56 Binder, BINDER GmbH, Tuttlingen, Germany). The curing process was monitored in real time for 60 min.

2.6. Characterization of the Cured Coatings

The following tests and indices were carried out and assessed in order to evaluate the properties of the cured coatings: pendulum hardness test, cross-cut adhesion test, gloss, and yellowness index. The hardness was tested using the Persoz pendulum hardness tester according to ASTM D 4366 with respect to pendulum oscillation times on the coatings on the glass substrate. The presented values represent the average values determined from seven measurements, wherein the error did not exceed 5%. Adhesion of the coatings to the glass substrate was measured using the cross-cut adhesion tester according to the national and international standards, including ASTM D3359 and ISO 2409. The presented values represent the average values determined from seven measurements, wherein the error did not exceed 10%. Gloss was measured by spectrometer GLS (SADT Development Technology Co. Ltd., Beijing, China) according to ASTM D523. The presented values represent the average values determined from three measurements, wherein the error did not exceed 5%. Yellowness index is the number calculated from spectrophotometric data that describes the change in color of the tested samples. This parameter was measured according to ASTM E313 and using precision colorimeter NH-145 (3NH Technology Co., Ltd., Shenzhen, China). The presented values represent the average values determined from three measurements, wherein the error did not exceed 5%.

3. Results and Discussion

3.1. Approach to the Synthesis of Sustainable Urethane Acrylate Resins

In this study, urethane acrylate resins were obtained while considering several factors, such as the feed ratio of PDO to isocyanate-bearing urethane acrylate resin, the use of resins with different amounts of unsaturated functional groups, and, finally, the use of various solvents, which affected the course of the PDO’s addition to the resins. The remaining synthesis parameters, i.e., the reaction catalysts, were selected based on a literature review [35] and several preliminary runs were carried out at different doses. The best catalyst dose turned out to be 0.0005 wt.% based on the weight of the reactants, and 30 °C was assumed as the optimal reaction temperature. However, employing higher temperatures or an increased amount of catalyst appears to accelerate the reactions considerably; nevertheless, this rendered the process hard to control. That is to say, partially gelled, adhesive-like products were obtained. A possible explanation for this phenomenon may be found in the chemical structure of resin, i.e., the presence of two hydroxyl groups per molecule.
UA oligomers containing two NCO groups and two (2DB-2NCO) or one (1DB-2NCO) unsaturated group(s) were used to determine the influence of the chemical structure on the reaction extent, the conversion of the substrates, and, finally, the functional properties of the UA prepolymers. The influence of the molar ratio of the reagents on the rate of reaction progress and the conversion of NCO groups was tested by carrying out UA syntheses using reagents in a stoichiometric molar ratio (2DB-UA-A or 1DB-UA-A) or with a slight excess of alcohol (10% based on the molar ratio; 2DB-UA-B or 1DB-UA-B). The progress of the reaction over time was presented via the FTIR method through the monitoring of the disappearance of the peak characteristic for the NCO group at 2268 cm−1. The FTIR spectra collected throughout the reaction time concerning the model syntheses carried out in acetone for the resin containing two unsaturated groups, one in Figure 2 and one in Figure 3, are presented below. The reaction with the urethane acrylate oligomer containing two NCO groups and two acrylate groups (2DB-UA), which have a lower molecular weight, was carried out for a shorter period of time. After 80 min, the complete disappearance of the peak at 2268 cm−1 responsible for the presence of the NCO groups was observed. The reaction is different in the case of the reaction of PDO with a UA oligomer containing two NCO groups and only one acrylate group (1DB-UA). For this reaction, the disappearance of the NCO peak is observed after about 150 min. Thus, the reaction is slower than in the previous synthesis. The lower reactivity of this substrate compared to the previous one may be caused by a higher molecular weight and chemical structure architecture in the form of a longer carbon chain.
Figure 4 shows the progress of the reaction over time with respect to the PDO’s addition to the isocyanate-bearing urethane acrylates, considering the stoichiometric molar ratio of the substances (A) and the ratio with an excess of PDO (B). It was shown that an excess of alcohol causes a slight change in the reaction progress, so the stoichiometric ratio was used for further research. The use of a stoichiometric ratio also has its advantages; namely, the product does not need to be purified of excess propanediol. As shown in Figure 4, the synthesis of the 2DB-UA product proceeded faster and was completed after 80 min, while the course of the synthesis of the 1DB-UA product was milder and was completed after 150 min.
In the next stage, syntheses were carried out with the various solvents, including acetone (AC), dichloromethane (DCM), and bulk (M). The study of the progress of these reactions (Figure 5) showed that the reaction was fastest in acetone and slowest in dichloromethane, regardless of the UA resin used. The reason for the superior performance of acetone compared to dichloromethane is likely due to the better solubility of the substrates in this solvent. It should also be emphasized that the method of conducting the reaction in acetone had advantages over performing it in bulk, not only in terms of the rate but also in terms of controlling the viscosity of the system.
The above results are of the utmost importance because they can guide methods for adjusting reactions to obtain resins with desired characteristics. Table 2 summarizes the conditions and results related to the synthesis of the resins. Based on the obtained values, one can state that the optimum synthesis time to prepare 2DB-UA resin is about 1.5 h and about 2.5 h in the case of 1DB-UA resin, regardless of the molar ratio used and assuming that the reaction is carried out in acetone. Coating tests were performed on the synthesized resins in an equimolar ratio and in an acetone environment.

3.2. Examination of the Curing Process and Properties of Cured Coatings

The kinetics of photopolymerization are useful in understanding the impacts of the rate and degree of curing on the properties of cured coatings. In the first stage, research on the chemical structure of the obtained resins was presented, i.e., regarding the amount of reactive functional groups, particularly acrylate groups, as well as the molecular weight and viscosity. For this purpose, two obtained products were tested, including a resin containing two acrylate groups per molecule (2DB-UA) or one acrylate group (1DB-UA). For this test, the resins obtained in a stoichiometric ratio and in an acetone medium were used. The coatings were prepared with the use of resins from which acetone had previously been distilled. UA containing more acrylate groups, i.e., 2DB-UA, polymerized faster, resulting in the higher conversion of acrylate groups (Figure 6, Table 3). The higher rate of photopolymerization of this resin may also be related to its lower molecular weight and lower viscosity compared to the 1DB-UA resin. Shorter polymer chains had easier mobility. It is evident that higher double bond conversion will result in the higher hardness of the coatings. It also turns out that the 2DB-UA resin not only has a higher conversion and hardness, but also has advantages in terms of other properties of cured coatings, i.e., adhesion, gloss, and yellowness. Hence, it was selected in the next stage to study the impact of the curing method.
Due to the presence of the different functional groups that the obtained UA resin contains (OH and acrylate groups), various curing methods were tested, including without (none) and with the use of radiation (UV), as well as thermal methods (T). The methods were based on the curing of the resins in the following form: (i) without the use of initiators (none-none), (ii) without or with a photo initiator and cured by exposure to UV radiation (UV-none and UV-PI, respectively), (iii) without or with a thermal initiator and cured by the use of elevated temperature (T-none and T-I, respectively), and (iv) using a hybrid method, i.e., using UV radiation, elevated temperature, and initiators (UV/T-PI/I). Based on the FTIR spectra, the curing process was monitored by the disappearance of the peak at 810 cm−1 responsible for double bonds from the acrylate groups. Real-time studies were also performed to monitor the curing process of the coatings more closely, and the results were summarized in the form of a collective plot of double bond conversion and the cure rate (Figure 7). On the basis of the conducted research, it was concluded that the obtained resins form cured coatings only in a composition with a photo initiator or a thermal initiator, because the exposure of the coatings to air or to UV radiation did not change the number of unsaturated bonds. The exception was the exposure of the coatings to increased temperature, which led to a slight change in the number of unsaturated bonds with approximately 5% conversion. However, based on the organoleptic tests, this coating is also considered to be uncured. The possibility of the dual curing of the obtained resins was presented in order to demonstrate a method for eliminating the disadvantages of using photo-curable resins on three-dimensional surfaces, i.e., in places where the access of UV radiation is difficult and where there is a possibility of insufficient post-curing. In this case, it is worth taking advantage of the thermal curing of these resins. The use of initiators achieved a higher degree of conversion for the use of (i) a photo initiator and UV curing (UV-PI), which was 44%; (ii) a photo initiator and an initiator as well as hybrid curing (UV/T-PI/I), which was 34%; and in the presence of a thermal initiator and thermal curing (T-I), which was only 13%. The highest process rate was observed in the case of the photocuring method of the coating obtained from a composition of resin and a photo initiator (above 100 %/min). This coating also had the best properties of the cured coatings (Table 4), i.e., it had the highest hardness and gloss and lowest yellowness. Regardless of the method of curing, all coatings showed excellent adhesion to the substrate, which can probably be attributed to the large number of hydroxyl groups.

4. Conclusions

In this paper, the synthesis of sustainable urethane acrylate resins by the modification of two isocyanate-bearing urethane acrylate oligomers with bio-sourced 1,3-propanediol has been demonstrated. This topic provides an approach for the synthesis of products with different reaction conditions. Therefore, kinetic parameters were presented to assess the effect of changing the reaction conditions on the progress of the reaction. The best results regarding the progress of the reaction were obtained when a stoichiometric molar ratio of substrates to the reaction was used and when the reactions were carried out in acetone. Moreover, the preparation of coatings based on synthesized resin with the implementation of UV radiation and/or elevated temperature to cure the coatings has been presented. The curing process was also monitored in real time. Double bond peaks related to the reaction kinetics of UA resins decreased as the reaction progressed, which was due to a decrease in the number of acrylate groups in the prepolymer chain and the curing of the coatings. The obtained resin, which contains both acrylate and hydroxyl groups in its structure, made it possible to produce coatings in a short time with excellent properties, particularly in terms of adhesion and gloss. This type of resin, which was obtained using a bioproduct and can be dual-cured (photo- or thermo-), can be employed as an eco-friendly film-forming product for use in various applications in the coating industry. In particular, it can be used on three-dimensional surfaces, i.e., those in which there is a possibility of the insufficient availability of UV radiation.

Author Contributions

Conceptualization, P.B.; methodology, P.B.; validation, P.B. and Z.C.; formal analysis, P.B.; investigation, P.B., M.N. and K.M.; resources, P.B.; data curation, P.B.; writing—original draft preparation, P.B. and K.M.; writing—review and editing, Z.C.; supervision, P.B. and Z.C.; funding acquisition, M.N. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Rector of the West Pomeranian University of Technology in Szczecin for PhD students of the Doctoral School, grant number: 540/106.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to express their appreciation to Allnex, The Netherlands, for supplying Ebecryl 4141 and Ebecryl 4396.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. George, A.; Sanjay, M.R.; Srisuk, R.; Parameswaranpillai, J.; Siengchin, S. A Comprehensive Review on Chemical Properties and Applications of Biopolymers and Their Composites. Int. J. Biol. Macromol. 2020, 154, 329–338. [Google Scholar] [CrossRef] [PubMed]
  2. Mohanty, A.K.; Misra, M.; Drzal, L.T. Sustainable Bio-Composites from Renewable Resources: Opportunities and Challenges in the Green Materials World. J. Polym. Environ. 2002, 10, 19–26. [Google Scholar] [CrossRef]
  3. Sharma, B.; Sauraj, S.; Kumar, B.; Pandey, A.; Dutt, D.; Negi, Y.S.; Maji, P.K.; Kulshreshtha, A. Synthesis of Waterborne Acrylic Copolymer Resin as a Binding Agent for the Development of Water-based Inks in the Printing Application. Polym. Eng. Sci. 2021, 61, 1569–1580. [Google Scholar] [CrossRef]
  4. Kumar, B.; Negi, Y.S. Synthesis and Thermal Properties of Novel Poly(Potassium 1-Hydroxy Acrylate-Co-Potassium Acrylate) Based Copolymer. Mater. Lett. 2018, 232, 196–201. [Google Scholar] [CrossRef]
  5. Kumar, B.; Deeba, F.; Priyadarshi, R.; Bano, S.; Kumar, A.; Negi, Y.S. Development of Novel Cross-Linked Carboxymethyl Cellulose/Poly(Potassium 1-Hydroxy Acrylate): Synthesis, Characterization and Properties. Polym. Bull. 2020, 77, 4555–4570. [Google Scholar] [CrossRef]
  6. Niesbach, A.; Adams, T.A.; Lutze, P. Semicontinuous Distillation of Impurities for the Production of Butyl Acrylate from Bio-Butanol and Bio-Acrylic Acid. Chem. Eng. Process. Process Intensif. 2013, 74, 165–177. [Google Scholar] [CrossRef] [Green Version]
  7. Nagengast, J. Development and Optimization of a One-Step Liquid Phase Process for the Sustainable Production of Bio Acrylic Acid from Lactic Acid and Its Oligomers, Friedrich-Alexander-Universität Erlangen-Nürnberg (FAU). 2021. Available online: https://d-nb.info/1237499100/34 (accessed on 1 October 2022).
  8. Lubis, M.A.R.; Handika, S.O.; Sari, R.K.; Iswanto, A.H.; Antov, P.; Kristak, L.; Lee, S.H.; Pizzi, A. Modification of Ramie Fiber via Impregnation with Low Viscosity Bio-Polyurethane Resins Derived from Lignin. Polymers 2022, 14, 2165. [Google Scholar] [CrossRef]
  9. Handika, S.O.; Lubis, M.A.R.; Sari, R.K.; Laksana, R.P.B.; Antov, P.; Savov, V.; Gajtanska, M.; Iswanto, A.H. Enhancing Thermal and Mechanical Properties of Ramie Fiber via Impregnation by Lignin-Based Polyurethane Resin. Materials 2021, 14, 6850. [Google Scholar] [CrossRef]
  10. Aristri, M.A.; Lubis, M.A.R.; Iswanto, A.H.; Fatriasari, W.; Sari, R.K.; Antov, P.; Gajtanska, M.; Papadopoulos, A.N.; Pizzi, A. Bio-Based Polyurethane Resins Derived from Tannin: Source, Synthesis, Characterisation, and Application. Forests 2021, 12, 1516. [Google Scholar] [CrossRef]
  11. Fouilloux, H.; Thomas, C.M. Production and Polymerization of Biobased Acrylates and Analogs. Macromol. Rapid Commun. 2021, 42, 2000530. [Google Scholar] [CrossRef]
  12. Hernández, N.; Williams, R.C.; Cochran, E.W. The Battle for the “Green” Polymer. Different Approaches for Biopolymer Synthesis: Bioadvantaged vs. Bioreplacement. Org. Biomol. Chem. 2014, 12, 2834–2849. [Google Scholar] [CrossRef]
  13. Samyn, P.; Bosmans, J.; Cosemans, P. Comparative Study on Mechanical Performance of Photocurable Acrylate Coatings with Bio-Based versus Fossil-Based Components. Mater. Today Commun. 2022, 32, 104002. [Google Scholar] [CrossRef]
  14. Stouten, J.; Vanpoucke, D.E.P.; Van Assche, G.; Bernaerts, K.V. UV-Curable Biobased Polyacrylates Based on a Multifunctional Monomer Derived from Furfural. Macromolecules 2020, 53, 1388–1404. [Google Scholar] [CrossRef] [Green Version]
  15. Hermens, J.G.H.; Freese, T.; van den Berg, K.J.; van Gemert, R.; Feringa, B.L. A Coating from Nature. Sci. Adv. 2020, 6, eabe0026. [Google Scholar] [CrossRef]
  16. Su, Y.; Lin, H.; Zhang, S.; Yang, Z.; Yuan, T. One-Step Synthesis of Novel Renewable Vegetable Oil-Based Acrylate Prepolymers and Their Application in UV-Curable Coatings. Polymers 2020, 12, 1165. [Google Scholar] [CrossRef]
  17. Tang, J.; Zhang, J.; Lu, J.; Huang, J.; Zhang, F.; Hu, Y.; Liu, C.; An, R.; Miao, H.; Chen, Y.; et al. Preparation and Properties of Plant-Oil-Based Epoxy Acrylate-Like Resins for UV-Curable Coatings. Polymers 2020, 12, 2165. [Google Scholar] [CrossRef]
  18. Zhang, H.; Zhang, H.; Tang, L.; Zhang, Z.; Gu, L.; Xu, Y.; Eger, C. Wear-Resistant and Transparent Acrylate-Based Coating with Highly Filled Nanosilica Particles. Tribol. Int. 2010, 43, 83–91. [Google Scholar] [CrossRef]
  19. Chen, Z.; Wu, J.F.; Fernando, S.; Jagodzinski, K. Soy-Based, High Biorenewable Content UV Curable Coatings. Prog. Org. Coat. 2011, 71, 98–109. [Google Scholar] [CrossRef]
  20. Hu, Y.; Shang, Q.; Tang, J.; Wang, C.; Zhang, F.; Jia, P.; Feng, G.; Wu, Q.; Liu, C.; Hu, L.; et al. Use of Cardanol-Based Acrylate as Reactive Diluent in UV-Curable Castor Oil-Based Polyurethane Acrylate Resins. Ind. Crops Prod. 2018, 117, 295–302. [Google Scholar] [CrossRef]
  21. Su, Y.; Zhang, S.; Zhou, X.; Yang, Z.; Yuan, T. A Novel Multi-Functional Bio-Based Reactive Diluent Derived from Cardanol for High Bio-Content UV-Curable Coatings Application. Prog. Org. Coat. 2020, 148, 105880. [Google Scholar] [CrossRef]
  22. Liu, R.; Zhang, X.; Zhu, J.; Liu, X.; Wang, Z.; Yan, J. UV-Curable Coatings from Multiarmed Cardanol-Based Acrylate Oligomers. ACS Sustain. Chem. Eng. 2015, 3, 1313–1320. [Google Scholar] [CrossRef]
  23. Sung, J.; Sun, X.S. Cardanol Modified Fatty Acids from Camelina Oils for Flexible Bio-Based Acrylates Coatings. Prog. Org. Coat. 2018, 123, 242–253. [Google Scholar] [CrossRef]
  24. Li, Y.; Wang, D.; Sun, X.S. Epoxidized and Acrylated Epoxidized Camelina Oils for Ultraviolet-Curable Wood Coatings. J. Am. Oil Chem. Soc. 2018, 95, 1307–1318. [Google Scholar] [CrossRef]
  25. Tathe, D.S.; Jagtap, R.N. Biobased Reactive Diluent for UV-Curable Urethane Acrylate Oligomers for Wood Coating. J. Coat. Technol. Res. 2015, 12, 187–196. [Google Scholar] [CrossRef]
  26. Rezania, S.; Oryani, B.; Park, J.; Hashemi, B.; Yadav, K.K.; Kwon, E.E.; Hur, J.; Cho, J. Review on Transesterification of Non-Edible Sources for Biodiesel Production with a Focus on Economic Aspects, Fuel Properties and by-Product Applications. Energy Convers. Manag. 2019, 201, 112155. [Google Scholar] [CrossRef]
  27. Haas, T.; Jaeger, B.; Weber, R.; Mitchell, S.F.; King, C.F. New Diol Processes: 1,3-Propanediol and 1,4-Butanediol. Appl. Catal. A Gen. 2005, 280, 83–88. [Google Scholar] [CrossRef]
  28. Casali, S.; Gungormusler, M.; Bertin, L.; Fava, F.; Azbar, N. Development of a Biofilm Technology for the Production of 1,3-Propanediol (1,3-PDO) from Crude Glycerol. Biochem. Eng. J. 2012, 64, 84–90. [Google Scholar] [CrossRef]
  29. Ghozali, M.; Triwulandari, E.; Haryono, A. Preparation and Characterization of Polyurethane-Modified Epoxy with Various Types of Polyol: Preparation and Characterization of Polyurethane-Modified. Macromol. Symp. 2015, 353, 154–160. [Google Scholar] [CrossRef]
  30. Acevedo, M.; De La Campa, J.G.; Abajo, J.D. Synthesis and Characterization of Imide End-Capped Oligoesters of Terephthalic Acid and 2-Methyl-2-Propyl-1,3-Propanediol. J. Appl. Polym. Sci. 1990, 41, 163–176. [Google Scholar] [CrossRef]
  31. Tuan Ismail, T.N.M.; Ibrahim, N.A.; Sendijarevic, V.; Sendijarevic, I.; Schiffman, C.M.; Hoong, S.S.; Mohd Noor, M.A.; Poo Palam, K.D.; Yeong, S.K.; Idris, Z.; et al. Thermal and Mechanical Properties of Thermoplastic Urethanes Made from Crystalline and Amorphous Azelate Polyols. J. Appl. Polym. Sci. 2019, 136, 47890. [Google Scholar] [CrossRef]
  32. Bednarczyk, P.; Nowak, M.; Mozelewska, K.; Czech, Z. Photocurable Coatings Based on Bio-Renewable Oligomers and Monomers. Materials 2021, 14, 7731. [Google Scholar] [CrossRef]
  33. Deng, L.; Tang, L.; Qu, J. Synthesis and Photopolymerization of Novel UV-Curable Macro-Photoinitiators. Prog. Org. Coat. 2020, 141, 105546. [Google Scholar] [CrossRef]
  34. Esen, D.S.; Karasu, F.; Arsu, N. The Investigation of Photoinitiated Polymerization of Multifunctional Acrylates with TX-BT by Photo-DSC and RT-FTIR. Prog. Org. Coat. 2011, 70, 102–107. [Google Scholar] [CrossRef]
  35. Zhang, D.; Liu, J.; Li, Z.; Shen, Y.; Wang, P.; Wang, D.; Wang, X.; Hu, X. Preparation and Properties of UV-Curable Waterborne Silicon-Containing Polyurethane Acrylate Emulsion. Prog. Org. Coat. 2021, 160, 106503. [Google Scholar] [CrossRef]
Figure 1. General structure of the obtained UA resins.
Figure 1. General structure of the obtained UA resins.
Coatings 12 01860 g001
Figure 2. FTIR spectra of 2DB-UA-A prepolymers collected throughout the reaction time (A) and monitoring of the disappearance of the peak corresponding to the isocyanate groups (B).
Figure 2. FTIR spectra of 2DB-UA-A prepolymers collected throughout the reaction time (A) and monitoring of the disappearance of the peak corresponding to the isocyanate groups (B).
Coatings 12 01860 g002
Figure 3. FTIR spectra of 1DB-UA-A prepolymers collected throughout the reaction time (A) and monitoring of the disappearance of the peak corresponding to the isocyanate groups (B).
Figure 3. FTIR spectra of 1DB-UA-A prepolymers collected throughout the reaction time (A) and monitoring of the disappearance of the peak corresponding to the isocyanate groups (B).
Coatings 12 01860 g003
Figure 4. Monitoring of disappearance of NCO peak with the progression of the addition of PDO to the isocyanate-bearing urethane acrylates considering the use of stoichiometric molar ratio of substances (A) and the ratio with excess of PDO (B).
Figure 4. Monitoring of disappearance of NCO peak with the progression of the addition of PDO to the isocyanate-bearing urethane acrylates considering the use of stoichiometric molar ratio of substances (A) and the ratio with excess of PDO (B).
Coatings 12 01860 g004
Figure 5. Monitoring of disappearance of NCO peak with the progressive addition of PDO to the isocyanate-bearing urethane acrylates while considering the use of solvent.
Figure 5. Monitoring of disappearance of NCO peak with the progressive addition of PDO to the isocyanate-bearing urethane acrylates while considering the use of solvent.
Coatings 12 01860 g005
Figure 6. The unsaturated double bond conversion (DC) and the photopolymerization rate (Rp) of various UA resins.
Figure 6. The unsaturated double bond conversion (DC) and the photopolymerization rate (Rp) of various UA resins.
Coatings 12 01860 g006
Figure 7. The unsaturated double bond conversion (DC) and the polymerization rate (Rp) of 2DB-UA resin with respect to the curing method.
Figure 7. The unsaturated double bond conversion (DC) and the polymerization rate (Rp) of 2DB-UA resin with respect to the curing method.
Coatings 12 01860 g007
Table 1. Characteristics of the reagents used in the resin synthesis.
Table 1. Characteristics of the reagents used in the resin synthesis.
Raw Material SymbolChemical NameFunctionalityMmol (g/mol)Viscosity (20 °C, mPa·s)
Double BondsNCO GroupsOH Groups
2DB-2NCOisocyanate-bearing urethane diacrylate22-70010,000
1DB-2NCOisocyanate-bearing urethane acrylate12-120020,000
PDO1,3-propanediol- 27655
Table 2. Conditions and results for resin synthesis.
Table 2. Conditions and results for resin synthesis.
SamplenDB-NCOnPDOSolventR901 (%)t1002 (min)
2DB-UA-A0.040.08A10080
1DB-UA-A57150
2DB-UA-B0.088A10080
1DB-UA-B61150
2DB-UA-AC0.040.08A10080
1DB-UA-AC57150
2DB-UA-DCMDCM97120
1DB-UA-DCM35210
2DB-UA-Mnone10080
1DB-UA-M50180
n—mole number; solvents: A—acetone, DCM—dichloromethane; R90—reaction progress rate after 90 min; t100—time after which 100% conversion of NCO groups was obtained.
Table 3. The basic properties of the cured coatings based on various UA resins.
Table 3. The basic properties of the cured coatings based on various UA resins.
SampleDCmax (1) (%)Rpmax (2) (%/min)HardnessAdhesionGloss (GU)Yellowness Index
2DB-UA-
AC-UV-PI
441016501003.70
1DB-UA-
AC-UV-PI
122254853.99
(1) Conversion determined by RT-IR method; (2) maximum polymerization rate determined by RT-IR method.
Table 4. The basic properties of the cured coatings based on 2DB-UA resin with respect to the curing method.
Table 4. The basic properties of the cured coatings based on 2DB-UA resin with respect to the curing method.
SampleDCmax (1) (%)Rpmax (2) (%/min)HardnessAdhesionGloss (GU)Yellowness Index
2DB-UA-
UV-PI
441016501003.70
2DB-UA-
T-I
1313200853.87
2DB-UA-
UV/T-PI/I
3440450923.95
(1) conversion determined by RT-IR method; (2) maximum polymerization rate determined by RT-IR method.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nowak, M.; Bednarczyk, P.; Mozelewska, K.; Czech, Z. Synthesis and Characterization of Urethane Acrylate Resin Based on 1,3-Propanediol for Coating Applications. Coatings 2022, 12, 1860. https://doi.org/10.3390/coatings12121860

AMA Style

Nowak M, Bednarczyk P, Mozelewska K, Czech Z. Synthesis and Characterization of Urethane Acrylate Resin Based on 1,3-Propanediol for Coating Applications. Coatings. 2022; 12(12):1860. https://doi.org/10.3390/coatings12121860

Chicago/Turabian Style

Nowak, Małgorzata, Paulina Bednarczyk, Karolina Mozelewska, and Zbigniew Czech. 2022. "Synthesis and Characterization of Urethane Acrylate Resin Based on 1,3-Propanediol for Coating Applications" Coatings 12, no. 12: 1860. https://doi.org/10.3390/coatings12121860

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