Next Article in Journal
Synthesis of Gentamicin-Immobilized Agar with Improved Antibacterial Activity
Previous Article in Journal
Study on the Impact and Water Absorption Performance of Prosopis juliflora & Glass Fibre Reinforced Epoxy Composite Laminates
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Photo-Curing Kinetics of 3D-Printing Photo-Inks Based on Urethane-Acrylates

1
Department of Life Science and Bioprocesses, Fraunhofer Institute for Applied Polymer Research IAP, Geiselbergstraße 69, 14476 Potsdam, Germany
2
Department of Functional Polymer Systems, Fraunhofer Institute for Applied Polymer Research IAP, Geiselbergstraße 69, 14476 Potsdam, Germany
*
Author to whom correspondence should be addressed.
Polymers 2022, 14(15), 2974; https://doi.org/10.3390/polym14152974
Submission received: 5 July 2022 / Revised: 20 July 2022 / Accepted: 21 July 2022 / Published: 22 July 2022
(This article belongs to the Special Issue Advances in Photocurable Resins and Elastomers)

Abstract

:
In this study, photo-curing kinetics for urethane-acrylate-based photo-inks for 3D printing were evaluated using a photo-differential scanning calorimetry analysis. Initially, the photopolymerization kinetics of di- and monofunctional monomers were separately studied at different temperatures (5–85 °C). Later, the photo-curing kinetics and mechanical properties of photo-inks based on different monomer mixtures (40/60–20/80) were evaluated. The results showed that urethane-dimethacrylate (UrDMA) and urethane-acrylate (UrA) had no light absorption in the region of 280–700 nm, making them a proper crosslinker and a reactive diluent, respectively, for the formulation of 3D-printing photo-inks. The kinetics investigations showed a temperature dependency for the photo-curing of UrDMA, where a higher photopolymerization rate (Rp,max: from 5.25 × 10−2 to 8.42 × 10−2 1/s) and double-bound conversion (DBCtotal: from 63.8% to 92.2%) were observed at elevated temperatures (5–85 °C), while the photo-curing of UrA was independent of the temperature (25–85 °C). Enhancing the UrA content from 60% to 80% in the UrDMA/UrA mixtures initially increased and later decreased the photopolymerization rate and conversion, where the mixtures of 30/70 and 25/75 presented the highest values. Meanwhile, increasing the UrA content led to lower glass transition temperatures (Tg) and mechanical strength for the photo-cured samples, where the mixture of 30/70 presented the highest maximum elongation (εmax: 73%).

Graphical Abstract

1. Introduction

Photo-curable inks (photo-inks) play a key role in light-based, 3D-printing technologies, e.g., stereolithography (SLA), digital light processing (DLP), liquid crystal precision (LCP), and digital light synthesis (DLS) or continuous liquid interface production (CLIP) [1]. Under irradiation, liquid photo-ink consisting of mono- and multifunctional monomers and a photoinitiator undergoes a photopolymerization reaction and transforms into a solid polymeric network within a very short time. Recently, urethane-acrylates have been used to formulate 3D-printing photo-inks [2,3,4,5,6,7,8,9,10]. The intermolecular hydrogen bonding between the urethane groups allows the pre-association of urethane-acrylate molecules and, consequently, 3–6 times faster photo-curing in comparison to their corresponding nonhydrogen-bonding acrylates [11,12]. Moreover, hydrogen bonding improves the mechanical properties of 3D-printed constructs [13,14].
The main research trend for light-based, 3D-printing technologies is focused on the development of advanced photo-ink formulations and innovative processes to obtain high-performance constructs for cutting-edge applications [1,15]. Meanwhile, the fundamental investigation of the photopolymerization kinetics of photo-inks has been less reported [10,16,17,18]. To better understand the characteristics of photo-inks and, consequently, to improve the quality and performance of 3D-printed constructs, studying photo-curing kinetics is essential. Recently, Jiang et al. [17] reported that higher UV intensity improved the photo-curing rate (Rp) and the double-bound conversion (DBC) of a photo-ink based on 1,6-hexamethylene diacrylate, while longer exposure time partially enhanced the DBC. Deng et al. [10] reported that enhancing the concentration of the photoinitiator improved the Rp and DBC values for photo-ink formulations based on urethane-acrylates. According to Kim et al. [18], higher temperature led to higher Rp, DBC, and glass transition temperature (Tg) values for a commercial UV-curable, 3D-printing acrylic resin (VeroWhitePlus RGD 835, Stratasys Ltd., Minneapolis, MN, USA). Bachmann et al. [16] studied the photo-curing kinetics of three commercial photo-inks for DLS 3D printing under different wavelengths, light intensities, temperatures, relative humidities, and atmospheric oxygen concentrations. One of their observations was the improvement of Rp by decreasing the atmospheric oxygen concentration, which is the reason behind a shorter printing time under an inert atmosphere.
In this work, initially, the photopolymerization kinetics of a difunctional and a monofunctional urethane-acrylate monomer, respectively, as a crosslinker and a reactive diluent in 3D-printing photo-ink formulations are studied separately at different temperatures (5–85 °C). For this purpose, the released heat during the exothermic photopolymerization reactions is measured through photo-differential scanning calorimetry (photo-DSC) analysis. Later, the photo-curing of photo-inks based on different mixtures of these monomers (40/60–20/80), as well as the thermomechanical properties of corresponding photo-cured samples, are investigated.

2. Experimental

2.1. Materials

7,7,9(or 7,9,9)-Trimethyl-4,13-dioxo-3,14-dioxa-5,12-diazahexadecane-1,16-diol dimethacrylate (UrDMA) as a crosslinker, 2-[[(butylamino)carbonyl]oxy]ethyl 2-propenoate (UrA) as a reactive diluent, and ethyl (2,4,6-trimethylbenzoyl)phenylphosphinate (TPOL) as a photoinitiator were received from German industrial suppliers and used without any treatment.

2.2. Instruments

UV–Vis spectroscopy of the monomers and photoinitiator in an acetonitrile solution (0.8 g/L) was performed using a PerkinElmer instrument (Lambda 950, Norwalk, CT, USA) operating in the range of 200–700 cm−1. Acetonitrile was used as a blank solution.
A DSC from Netzsch (DSC 204 F1 Phoenix, Mannheim, Germany) equipped with a UV-curing spot system (OmniCure® S2000, Lumen Dynamics, Toronto, ON, Canada) was employed to study the photo-curing of the samples. The experiments were conducted in an isothermal mode (5, 25, 45, 65, or 85 °C) with a UV intensity of 1 W/m2 under N2 atmosphere. The UV intensity was measured with a radiometer (R2000 Radiometer, Lumen Dynamics, Toronto, ON, Canada). The samples were weighted in uncovered aluminum pans in darkness and analyzed immediately. An empty, uncovered aluminum pan was employed as a reference. The irradiation step was prolonged up to 3 min for the full photo-curing of the samples and reached a plateau region, which was employed as a baseline for the peak integrations. The Tg of the photo-cured samples was determined using DSC measurements in the temperature range of −50 to 200 °C with the heating rate of 10 °C/min under N2 atmosphere. The Tg values were extracted from the middle point of the baseline change in the first heating cycle.
The tensile properties of the photo-cured specimens were measured with a Zwick/Roell machine (model Z020, Ulm, Germany) equipped with a load cell of 100 N at room temperature. Photo-cured samples with a thickness of 300–500 µm were cut into 8 × 2 cm2 pieces and analyzed with a clamping length of 34 mm and a crosshead speed of 17 mm/min. The reported values were the averages of at least three specimens for each sample.

2.3. Methods

The Rp (1/s) and DBC (%) values as a function of time were calculated from the photo-DSC data using Equations (1)–(3) [7]:
Δ H t h e o r = f · n · Δ H 0 M w
R p = d H p d t / Δ H t h e o r
D B C = Δ H p Δ H t h e o r × 100
where ΔHtheor (J/g) is the theoretical total heat released during the complete polymerization of the monomers within the sample, f is the mass fraction of each monomer within the sample, n is the number of double bonds in each monomer, ΔH0 (J/mol) is the standard heat of polymerization for either methacrylate (54.8 kJ/mol [19,20]) or acrylate (86.2 kJ/mol [20,21]), Mw (g/mol) is the molecular weight of each monomer, dHp/dt (J/s·g or W/g) is the normalized heat flow per second, and ΔHp (J/g) is the heat generated from the start of photo-curing up to a certain time that is obtained from the integration of the thermograms.
To prepare tensile test specimens, formulations based on UrDMA/UrA mixtures (40/60–20/80) containing TPOL (1 wt.%) were cast with a thickness of 700 µm on a PET substrate using an adjustable applicator and cured under a UV lamp (145 mW/m2) for 60 s. After washing in isopropanol, the backs of the films were post-cured for 60 s.

3. Results and Discussion

The chemical structures of the used chemicals are depicted in Scheme 1. UrDMA as a viscous, difunctional monomer (η = 9500 mPa·s) was used in the photo-ink formulations to improve the dimensional stability of the 3D constructs by generating a thermoset network structure [22,23]. Monofunctional UrA (η = 35 mPa·s) was employed as a reactive diluent to decrease the viscosity of the formulations, while resulting in flexibility for the 3D constructs [9,24]. TPOL as a photoinitiator was added to the formulations at 1 wt.% regarding the total weight of the monomers. TPOL has previously shown excellent biocompatibility, color stability, and 3D-printing accuracy compared to common photoinitiators, i.e., bis(2,4,6-trimethylbenzoyl)phenylphosphine oxide (BAPO) and (2,4,6-Trimethylbenzoyl)diphenylphosphine oxide (TPO) [25,26,27].

3.1. UV–vis Absorption

Under a light, photoinitiator molecules absorb energy, convert to an excited state, and generate free radicals to initiate the photopolymerization of monomers. Before studying the photopolymerization kinetics of monomers, their UV–vis spectra in an acetonitrile solution were recorded (Figure 1). Both monomers (UrDMA and UrA) were transparent in the region of 280–700 nm, which is ideal for the photo-curing process since the whole irradiated light is accessible to the photoinitiator molecules. Meanwhile, TPOL showed absorption in the range of 200–420 nm, with two maximums at 273 and 371 nm. The absorption at 371 nm makes TPOL active for commercial DLP 3D printers and post-curing units, which mainly function at 365, 385, or 405 nm.

3.2. Photo-Curing

The most common type of photopolymerization for 3D-printing applications is through the free-radical mechanism, which involves three main steps: (1) initiation, (2) propagation, and (3) termination [15]. In the initiation step, the photoinitiator molecules are disintegrated under light irradiation and produce reactive radical fragments. In the propagation step, these radical species react with the double bonds (C=C) of monomer molecules to generate growing polymer chains. The growth of these polymer chains continues simultaneously with the addition of further monomers and the formation of macroradicals. In the termination step, the growing polymer chains are deactivated through combination, radical disproportionation, or chain transfer reactions [15].
Photo-DSC analyses have been widely utilized to evaluate the photo-activity of 3D-printing photo-inks [7,16,17]. Primarily, the photopolymerization kinetics of difunctional UrDMA and monofunctional UrA were studied separately in the isothermal mode at different temperatures, i.e., at 5, 25, 45, 65, or 85 °C (Figure 2a and Figure 3a). The Rp and DBC values as a function of irradiation time were calculated from the photo-DSC data according to Equations (1)–(3) (Figure 2b–d and Figure 3b–d). For both UrDMA and UrA, the Rp value, an overall value of the propagation and termination rates, initially increased and later decreased during the photo-curing process (Figure 2b and Figure 3b), which are known as the auto-acceleration and auto-deceleration phenomena, respectively. This observation was due to the gradual increase in viscosity with conversion and, consequently, the decrease in molecular mobility of the monomers and macroradicals within the photo-curing mixture over time [28,29]. Initially, at a low viscosity, the Rp of the liquid monomers was constant and dependent on the double-bond activity (chemistry-controlled). Later, at medium viscosity, the coupling of macroradicals in the photo-curing mixture for termination was diffusion-limited when the monomers were still mobile for propagation, and thus, the Rp increased (auto-acceleration). Finally, at a high viscosity, when the photo-curing mixture was transforming into a rubbery, glassy network (gel point), the diffusion of monomers to reach the macroradicals was significantly restricted, and therefore, the Rp decreased (auto-deceleration) [28,29]. For both UrDMA and UrA, the DBC sharply increased over the first 20–30 s of the photo-curing process and remained constant after 40 s (Figure 2c and Figure 3c). The total double-bound conversion (DBCtotal) values obtained from the total generated photo-curing heat (ΔHp,total) were less than unity due to the gelation or vitrification phenomenon [28,29], which meant that all the double bonds did not react in the course of the photo-curing process (Table 1 and Table 2).
For difunctional UrDMA, the maximum photo-curing rate (Rp,max) gradually increased from 5.25 × 10−2 to 8.42 × 10−2 1/s by raising the photo-curing temperature (Tp) from 5 to 85 °C (Table 1). Meanwhile, the gel-point time (tGP), the time to reach Rp,max, decreased from 7.0 to 3.3 s, and the gel-point conversion (DBCGP), the conversion at Rp,max, increased from 10.9% to 12.3%. These observations showed higher photo-activity leading to faster photopolymerization for UrDMA at elevated temperatures, which is in agreement with previous reports on multifunctional monomers [28,30,31,32]. The increase in Rp,max values by raising the temperature can also be explained according to Arrhenius’ law (Equation (4)) [31]:
k = k p = A × e E a / R T
where k is the overall photopolymerization rate constant equal to the propagation rate constant (kp) at low conversions, Ea is the activation energy, A is the frequency factor, R is the ideal gas constant, and T is the absolute temperature. According to Arrhenius’ law, a rise in temperature results in an increase in k and, consequently, Rp,max.
On the other hand, the DBCtotal of UrDMA was improved from 63.8% to 92.2% (Table 1) due to lower viscosity and higher molecular mobility within the photo-curing mixture at higher temperatures, which is in agreement with previous reports [28,30,31,32]. Meanwhile, the times to reach 50%, 90%, and 95% of DBCtotal (t50%, t90%, and t95%, respectively) were shortened, which meant faster photo-curing processes at elevated temperatures. The Tg of the photo-cured samples obtained from DSC measurements was improved from 63 to 122 °C, which was expected according to the DBCtotal trend, where more conversion for difunctional UrDMA resulted in networks with higher crosslinking densities, lower polymer chain flexibility, and consequently, higher Tg values. It is worth mentioning that, for all the photo-cured samples, the Tg values were higher than the Tp, which meant that the conversion was restricted due to the transformation of the liquid monomer to a glassy network with insufficient molecular mobility [31].
For monofunctional UrA, the Rp,max remained constant in the range of 5.08 × 10−2–5.47 × 10−2 1/s by increasing the Tp from 5 to 85 °C (Table 2). In the Tp range of 25–85 °C, the tGP, DBCGP, and DBCtotal remained constant in the ranges of 5.2–5.7 s, 16.5–18.5%, and 83.3–83.8%, respectively. These phenomena meant that raising the temperature did not affect the photo-activity and conversion of UrA, which is in contrast with previous results about 2-hydroxyethyl acrylate (HEA) [11], undecyl amide N-ethyl acrylate [12], and 2-carboxyethyl acrylate [33], where the Rp,max has decreased with increasing Tp from 25 °C to 90 °C as a result of the dissociation of the intermolecular hydrogen bonding and disappearance of the pre-association of monomers. At 5°C, UrA displayed a shorter tGP (4.4 s) and lower DBCGP (11.9%) and DBCtotal (78.7%) values, which was attributed to the crystallization of UrA molecules with a melting point (Tm) of 5 °C, leading to lower molecular mobility within the photo-curing mixture. This observation is in agreement with a previous report on crystalizable monomers [7].
In the Tp range of 25–85 °C, t50%, t90%, and t95% were shortened by increasing the temperature, which meant faster photo-curing processes at higher temperatures. The Tg of the photo-cured samples remained constant at −11 °C in the Tp range of 5–65 °C and then partially increased to −8 °C at a Tp of 85 °C, which was expected due to almost the same DBCtotal values. For all the photo-cured samples, the Tg values were lower than the Tp, meaning the photo-curing mixture transformed from a low-viscosity liquid monomer into a rubbery network with flexible polymer chains. It is worth mentioning that, although the photo-cured monofunctional UrA was supposed to be a thermoplastic, the DSC analyses up to 200 °C did not show any exothermic melting peak.
Later, the photopolymerization kinetics of different UrDMA/UrA mixtures (40/60–20/80) as bases for photo-inks were evaluated in isothermal mode at 25 °C (Figure 4 and Table 3). The Rp,max and DBCtotal for the UrDMA/UrA mixtures were higher (7.35 × 10−2–9.50 × 10−2 1/s and 90–92%, respectively) compared with pure UrDMA and UrA at 25 °C, while t50%, t90%, and t95% were shorter. These observations showed higher photo-activity, leading to faster photopolymerization and higher conversion for the UrDMA/UrA mixtures.
By enhancing the content of the monofunctional reactive diluent (UrA) from 60% to 80% in the UrDMA/UrA mixtures, the Rp,max, tGP, DBCGP, DBCtotal initially increased and then decreased, where the mixtures of 30/70 and 25/75 presented the highest values (Table 3). Meanwhile, t50%, t90%, and t95% initially decreased and then increased. Since UrA has a lower viscosity (η = 35 mPa·s) than UrDMA (η = 9500 mPa·s), increasing the UrA content led to photo-curing mixtures with lower viscosity. At the same time, it resulted in lower average double-bound functionality within the photo-curing mixture and, consequently, later gelation. Both phenomena enhanced the molecular mobility within the photo-curing mixture, which explained the initial increasing trend for the Rp,max, tGP, DBCGP, DBCtotal values, as well as the decreasing trend for the t50%, t90%, and t95% values. Therefore, faster photopolymerization and higher conversion were observed for the mixtures of 30/70 and 25/75. The later decreasing trend for the Rp,max, tGP, DBCGP, DBCtotal values and the increasing trend for the t50%, t90%, and t95% values, especially for the mixture of 20/80, could be attributed to the higher deactivation of macroradicals through the combination of macroradicals and, thus, the limited auto-acceleration at a very low viscosity.
The Tg of the photo-cured samples decreased from 23 to −11 °C by increasing the UrA content from 60% to 80%, which was expected due to the reduction in the average double-bound functionality, the crosslinking density of the polymeric network, and thus, the polymer chain flexibility. For all the photo-cured samples, the Tg values were lower than 25 °C, meaning the photo-curing mixture transformed from a viscous liquid monomer into a rubbery network with flexible polymer chains.

3.3. Mechanical Properties

Enhancing the content of the monofunctional reactive diluent in photo-inks, which decreased the average double-bound functionality, could influence the mechanical properties of 3D-printed objects. The mechanical properties of the photo-cured samples were investigated with tensile tests (Figure 5). The results presented the typical stress–strain curves of a thermoset polymer (Figure 5a). Enhancing the UrA content from 60% to 80% led to a decrease in the tensile module (E) from 248.8 to 5.4 MPa (Figure 5b) and tensile strength (σmax) from 13.6 to 1.9 MPa (Figure 5c), which was expected due to the reduction in the crosslinking density of the polymeric network as a result of the fall in the average double-bound functionality within corresponding photo-curing mixtures. Meanwhile, the maximum elongation (εmax) initially increased from 41% to 73% (for the mixture of 30/70) and then decreased to 50% (Figure 5d). These observations are in agreement with a previous report about photo-cured elastomers based on mono- and difunctional urethane-acrylates [10].

4. Conclusions

UrDMA and UrA with no light absorption in the region of 280–700 nm were presented as a proper crosslinker and a reactive diluent, respectively, for the formulation of 3D-printing photo-inks. Kinetics investigations showed a temperature dependency for the photo-curing of UrDMA, where higher photopolymerization rates, conversion, and Tg were observed at elevated temperatures, while the photo-curing of UrA was independent of temperature (25–85 °C). Enhancing the UrA content from 60% to 80% in the UrDMA/UrA mixtures initially increased and later decreased the photopolymerization rate and conversion, where the mixtures of 30/70 and 25/75 presented the highest values. Meanwhile, increasing the UrA content led to lower Tg, E, and σmax values for the photo-cured samples, where the mixture of 30/70 presented the highest εmax value. For future investigations, studying the effects of other parameters, such as photoinitiator type and concentration, not only on photo-curing kinetics but also on the 3D printability of photo-inks may be interesting. Employing other instruments, such as online FTIR or a rheometer, may be helpful to better understand the chemical changes during the photo-curing process.

Author Contributions

H.B. designed the research, performed the experiments, analyzed the data, and wrote the manuscript. G.K. and F.W. helped with performing the experiments. W.M. acquired the funding, managed the project, and reviewed the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Federal Ministry of Education and Research of Germany in the framework of “ProMatLeben–Polymere” (project number 13XP5087E, PolyKARD). The APC was funded by Fraunhofer’s Publication Fund.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The work is dedicated to Hartmut Krüger, who established the topic of photo-curing 3D printing at Fraunhofer IAP and now continues to pursue it in his well-deserved retirement.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Quan, H.; Zhang, T.; Xu, H.; Luo, S.; Nie, J.; Zhu, X. Photo-curing 3D printing technique and its challenges. Bioact. Mater. 2020, 5, 110–115. [Google Scholar] [CrossRef] [PubMed]
  2. Tzeng, J.-J.; Yang, T.-S.; Lee, W.-F.; Chen, H.; Chang, H.-M. Mechanical Properties and Biocompatibility of Urethane Acrylate-Based 3D-Printed Denture Base Resin. Polymers 2021, 13, 822. [Google Scholar] [CrossRef] [PubMed]
  3. Buchheit, H.; Bruchmann, B.; Stoll, K.; Mülhaupt, R. Functionalized acrylic polyhydroxy urethanes as molecular tool box for photocurable thermosets and 3D printing. J. Appl. Polym. Sci. 2021, 59, 882–892. [Google Scholar] [CrossRef]
  4. Peng, S.; Li, Y.; Wu, L.; Zhong, J.; Weng, Z.; Zheng, L.; Yang, Z.; Miao, J.-T. 3D Printing Mechanically Robust and Transparent Polyurethane Elastomers for Stretchable Electronic Sensors. ACS Appl. Mater. Interfaces 2020, 12, 6479–6488. [Google Scholar] [CrossRef]
  5. Chen, H.; Lee, S.-Y.; Lin, Y.-M. Synthesis and Formulation of PCL-Based Urethane Acrylates for DLP 3D Printers. Polymers 2020, 12, 1500. [Google Scholar] [CrossRef]
  6. Kim, S.; Lee, J.; Han, H. Synthesis of UV Curable, Highly Stretchable, Transparent Poly(urethane-acrylate) Elastomer and Applications Toward Next Generation Technology. Macromol. Res. 2020, 28, 896–902. [Google Scholar] [CrossRef]
  7. Singh, N.; Bakhshi, H.; Meyer, W. Developing non-isocyanate urethane-methacrylate photo-monomers for 3D printing application. RSC Adv. 2020, 10, 44103–44110. [Google Scholar] [CrossRef]
  8. Baker, M.; Wang, R.; Damanik, F.; Kuhnt, T.; Ippel, H.; Dijkstra, P.; ten Cate, T.; Dias, A.; Moroni, L. Biodegradable Poly(Ester) Urethane Acrylate Resins for Digital Light Processing: From Polymer Synthesis to 3D Printed Tissue Engineering Constructs. ChemRxiv 2020. [Google Scholar]
  9. Soreni-Harari, M.; Pierre, R.S.; McCue, C.; Moreno, K.; Bergbreiter, S. Multimaterial 3D Printing for Microrobotic Mechanisms. Soft Robot. 2020, 7, 59–67. [Google Scholar] [CrossRef]
  10. Deng, Y.; Li, J.; He, Z.; Hong, J.; Bao, J. Urethane acrylate-based photosensitive resin for three-dimensional printing of stereolithographic elastomer. J. Appl. Polym. Sci. 2020, 137, 49294. [Google Scholar] [CrossRef]
  11. Lee, T.Y.; Roper, T.M.; Jönsson, E.S.; Guymon, C.A.; Hoyle, C.E. Influence of Hydrogen Bonding on Photopolymerization Rate of Hydroxyalkyl Acrylates. Macromolecules 2004, 37, 3659–3665. [Google Scholar] [CrossRef]
  12. Jansen, J.F.G.A.; Dias, A.A.; Dorschu, M.; Coussens, B. Fast Monomers: Factors Affecting the Inherent Reactivity of Acrylate Monomers in Photoinitiated Acrylate Polymerization. Macromolecules 2003, 36, 3861–3873. [Google Scholar] [CrossRef]
  13. Barszczewska-Rybarek, I.M. Structure–property relationships in dimethacrylate networks based on Bis-GMA, UDMA and TEGDMA. Dent. Mater. 2009, 25, 1082–1089. [Google Scholar] [CrossRef]
  14. Lemon, M.T.; Jones, M.S.; Stansbury, J.W. Hydrogen bonding interactions in methacrylate monomers and polymers. J. Biomed. Mater. Res. Part A 2007, 83, 734–746. [Google Scholar] [CrossRef]
  15. Bagheri, A.; Jin, J. Photopolymerization in 3D Printing. ACS Appl. Polym. Mater. 2019, 1, 593–611. [Google Scholar] [CrossRef] [Green Version]
  16. Bachmann, J.; Schmölzer, S.; Ruderer, M.A.; Fruhmann, G.; Hinrichsen, O. Photo-differential scanning calorimetry parameter study of photopolymers used in digital light synthesis. SPE Polym. 2021, 3, 41–53. [Google Scholar] [CrossRef]
  17. Jiang, F.; Drummer, D. Curing Kinetic Analysis of Acrylate Photopolymer for Additive Manufacturing by Photo-DSC. Polymers 2020, 12, 1080. [Google Scholar] [CrossRef]
  18. Kim, Y.C.; Hong, S.; Sun, H.; Kim, M.G.; Choi, K.; Cho, J.; Choi, H.R.; Koo, J.C.; Moon, H.; Byun, D.; et al. UV-curing kinetics and performance development of in situ curable 3D printing materials. Eur. Polym. J. 2017, 93, 140–147. [Google Scholar] [CrossRef]
  19. Assumption, H.J.; Mathias, L.J. Photopolymerization of urethane dimethacrylates synthesized via a non-isocyanate route. Polymer 2003, 44, 5131–5136. [Google Scholar] [CrossRef]
  20. Anseth, K.S.; Wang, C.M.; Bowman, C.N. Kinetic evidence of reaction diffusion during the polymerization of multi(meth)acrylate monomers. Macromolecules 1994, 27, 650–655. [Google Scholar] [CrossRef]
  21. Harikrishna, R.; Ponrathnam, S.; Rajan, C.R.; Tambe, S.S. Photopolymerization of bis-aromatic and alicyclic based solid urethane acrylate macromonomer in the presence of large excess of reactive diluent. J. Therm. Anal. 2012, 112, 805–813. [Google Scholar] [CrossRef]
  22. Xu, Y.; Wang, H.; Xie, D. Preparation of new low viscosity urethane dimethacrylates for dental composites. J. Biomater. Sci. Polym. Ed. 2017, 29, 1011–1025. [Google Scholar] [CrossRef]
  23. Barszczewska-Rybarek, I. The role of molecular structure on impact resistance and bending strength of photocured urethane-dimethacrylate polymer networks. Polym. Bull. 2017, 74, 4023–4040. [Google Scholar] [CrossRef]
  24. Weigand, J.J.; Miller, C.I.; Janisse, A.P.; McNair, O.D.; Kim, K.; Wiggins, J.S. 3D printing of dual-cure benzoxazine networks. Polymer 2020, 189, 122193. [Google Scholar] [CrossRef]
  25. Kim, G.-T.; Go, H.-B.; Yu, J.-H.; Yang, S.-Y.; Kim, K.-M.; Choi, S.-H.; Kwon, J.-S. Cytotoxicity, Colour Stability and Dimensional Accuracy of 3D Printing Resin with Three Different Photoinitiators. Polymers 2022, 14, 979. [Google Scholar] [CrossRef]
  26. Zeng, B.; Cai, Z.; Lalevée, J.; Yang, Q.; Lai, H.; Xiao, P.; Liu, J.; Xing, F. Cytotoxic and cytocompatible comparison among seven photoinitiators-triggered polymers in different tissue cells. Toxicol. Vitr. 2021, 72, 105103. [Google Scholar] [CrossRef]
  27. Steyrer, B.; Neubauer, P.; Liska, R.; Stampfl, J. Visible Light Photoinitiator for 3D-Printing of Tough Methacrylate Resins. Materials 2017, 10, 1445. [Google Scholar] [CrossRef] [Green Version]
  28. Kardar, P.; Ebrahimi, M.; Bastani, S. Influence of temperature and light intensity on the photocuring process and kinetics parameters of a pigmented UV curable system. J. Therm. Anal. 2014, 118, 541–549. [Google Scholar] [CrossRef]
  29. Sideridou, I.; Tserki, V.; Papanastasiou, G. Effect of chemical structure on degree of conversion in light-cured dimethacrylate-based dental resins. Biomaterials 2002, 23, 1819–1829. [Google Scholar] [CrossRef]
  30. Kousaalya, A.B.; Ayalew, B.; Pilla, S. Photopolymerization of Acrylated Epoxidized Soybean Oil: A Photocalorimetry-Based Kinetic Study. ACS Omega 2019, 4, 21799–21808. [Google Scholar] [CrossRef] [Green Version]
  31. Rusu, M.C.; Block, C.; Van Assche, G.; Van Mele, B. Influence of temperature and UV intensity on photo-polymerization reaction studied by photo-DSC. J. Therm. Anal. 2012, 110, 287–294. [Google Scholar] [CrossRef]
  32. Macarie, L.; Ilia, G. The influence of temperature and photoinitiator concentration on photoinitiated polymerization of diacrylate monomer. Open Chem. 2005, 3, 721–730. [Google Scholar] [CrossRef]
  33. Zhou, H.; Li, Q.; Lee, T.Y.; Guymon, C.A.; Jönsson, E.S.; Hoyle, C.E. Photopolymerization of Acid Containing Monomers: Real-Time Monitoring of Polymerization Rates. Macromolecules 2006, 39, 8269–8273. [Google Scholar] [CrossRef]
Scheme 1. The chemical structures of UrDMA, UrA, and TPOL.
Scheme 1. The chemical structures of UrDMA, UrA, and TPOL.
Polymers 14 02974 sch001
Figure 1. UV–vis spectra for UrDMA, UrA, and TPOL in acetonitrile solution (0.8 g/L).
Figure 1. UV–vis spectra for UrDMA, UrA, and TPOL in acetonitrile solution (0.8 g/L).
Polymers 14 02974 g001
Figure 2. Photo-curing rates and conversions for UrDMA containing TPOL (1 wt.%) under N2 atmosphere at different temperatures obtained from photo-DSC data. The intensity of the UV lamp was 1 W/m2. (a) DSC vs. time; (b) Rp vs. time; (c) DBC vs. time; (d) Rp vs. DBC.
Figure 2. Photo-curing rates and conversions for UrDMA containing TPOL (1 wt.%) under N2 atmosphere at different temperatures obtained from photo-DSC data. The intensity of the UV lamp was 1 W/m2. (a) DSC vs. time; (b) Rp vs. time; (c) DBC vs. time; (d) Rp vs. DBC.
Polymers 14 02974 g002
Figure 3. Photo-curing rates and conversions for UrA containing TPOL (1 wt.%) under N2 atmosphere at different temperatures obtained from photo-DSC data. The intensity of the UV lamp was 1 W/m2. (a) DSC vs. time; (b) Rp vs. time; (c) DBC vs. time; (d) Rp vs. DBC.
Figure 3. Photo-curing rates and conversions for UrA containing TPOL (1 wt.%) under N2 atmosphere at different temperatures obtained from photo-DSC data. The intensity of the UV lamp was 1 W/m2. (a) DSC vs. time; (b) Rp vs. time; (c) DBC vs. time; (d) Rp vs. DBC.
Polymers 14 02974 g003
Figure 4. Photo-curing rates and conversions for UrDMA/UrA mixtures containing TPOL (1 wt.%) under N2 atmosphere at 25 °C obtained from photo-DSC data. The intensity of the UV lamp was 1 W/m2. (a) DSC vs. time; (b) Rp vs. time; (c) DBC vs. time; (d) Rp vs. DBC.
Figure 4. Photo-curing rates and conversions for UrDMA/UrA mixtures containing TPOL (1 wt.%) under N2 atmosphere at 25 °C obtained from photo-DSC data. The intensity of the UV lamp was 1 W/m2. (a) DSC vs. time; (b) Rp vs. time; (c) DBC vs. time; (d) Rp vs. DBC.
Polymers 14 02974 g004
Figure 5. Stress–strain curves (a), tensile modulus (E, (b)), maximum tensile strength (σmax, (c)), and maximum elongation (εmax, (d)) of the photo-cured samples based on UrDMA/UrA mixtures containing TPOL (1 wt.%).
Figure 5. Stress–strain curves (a), tensile modulus (E, (b)), maximum tensile strength (σmax, (c)), and maximum elongation (εmax, (d)) of the photo-cured samples based on UrDMA/UrA mixtures containing TPOL (1 wt.%).
Polymers 14 02974 g005
Table 1. Photo reactivity of UrDMA containing 1 wt.% TPOL under N2 atmosphere at different temperatures obtained from photo-DSC data.
Table 1. Photo reactivity of UrDMA containing 1 wt.% TPOL under N2 atmosphere at different temperatures obtained from photo-DSC data.
Tp
(°C)
Rp,max
(1/s)
tGP
(s)
DBCGP
(%)
ΔHp,total
(J/g)
DBCtotal
(%)
t50%
(s)
t90%
(s)
t95%
(s)
Tg
(°C)
55.25 × 10−27.010.9148.663.812.429.237.863
256.70 × 10−24.712.1180.877.69.926.334.577
456.52 × 10−24.512.7192.082.410.228.538.2101
658.40 × 10−23.612.3199.985.88.826.238.4106
858.42 × 10−23.312.3214.892.28.523.730.9122
Table 2. Photo reactivity of UrA containing 1 wt.% TPOL under N2 atmosphere at different temperatures obtained from photo-DSC data.
Table 2. Photo reactivity of UrA containing 1 wt.% TPOL under N2 atmosphere at different temperatures obtained from photo-DSC data.
Tp
(°C)
Rp,max
(1/s)
tGP
(s)
DBCGP
(%)
ΔHp,total
(J/g)
DBCtotal
(%)
t50%
(s)
t90%
(s)
t95%
(s)
Tg
(°C)
55.42 × 10−24.411.9315.278.710.030.439.2−11
255.08 × 10−25.717.8334.883.611.230.439.3−11
455.19 × 10−25.718.5335.783.811.128.836.3−11
655.47 × 10−25.217.1335.283.710.627.333.9−11
855.31 × 10−25.316.5335.583.310.927.734.3−8
Table 3. Photo reactivity of UrDMA/UrA mixtures containing 1 wt.% TPOL under N2 atmosphere at 25 °C obtained from photo-DSC data.
Table 3. Photo reactivity of UrDMA/UrA mixtures containing 1 wt.% TPOL under N2 atmosphere at 25 °C obtained from photo-DSC data.
UrDMA/UrARp,max
(1/s)
tGP
(s)
DBCGP
(%)
ΔHp,total
(J/g)
DBCtotal
(%)
t50%
(s)
t90%
(s)
t95%
(s)
Tg
(°C)
40/608.20 × 10−24.212.9303.191.89.324.632.823
35/657.35 × 10−25.016.7302.989.59.624.832.117
30/709.10 × 10−25.223.5324.393.58.422.230.49
25/759.50 × 10−24.921.3327.592.38.120.527.19
20/807.87 × 10−24.615.6331.491.29.024.433.34
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bakhshi, H.; Kuang, G.; Wieland, F.; Meyer, W. Photo-Curing Kinetics of 3D-Printing Photo-Inks Based on Urethane-Acrylates. Polymers 2022, 14, 2974. https://doi.org/10.3390/polym14152974

AMA Style

Bakhshi H, Kuang G, Wieland F, Meyer W. Photo-Curing Kinetics of 3D-Printing Photo-Inks Based on Urethane-Acrylates. Polymers. 2022; 14(15):2974. https://doi.org/10.3390/polym14152974

Chicago/Turabian Style

Bakhshi, Hadi, Guanxing Kuang, Franziska Wieland, and Wolfdietrich Meyer. 2022. "Photo-Curing Kinetics of 3D-Printing Photo-Inks Based on Urethane-Acrylates" Polymers 14, no. 15: 2974. https://doi.org/10.3390/polym14152974

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