Download as pdf or txt
Download as pdf or txt
You are on page 1of 5

Progress in Organic Coatings 100 (2016) 51–55

Contents lists available at ScienceDirect

Progress in Organic Coatings


journal homepage: www.elsevier.com/locate/porgcoat

Towards two-step photopolymerization using aza-Michael click


addition
Matthieu Retailleau, Ahmad Ibrahim, Céline Croutxé-Barghorn, Xavier Allonas ∗
Laboratory of Macromolecular Photochemistry and Engineering University of Haute Alsace, 3b rue Alfred Werner, 68093 Mulhouse, France

a r t i c l e i n f o a b s t r a c t

Article history: Click reactions involving thiols and acrylates can be performed either through one-step thiol-ene pho-
Received 21 January 2016 topolymerization or two-step thiol-Michael addition followed by photopolymerization. It is confirmed
Accepted 27 January 2016 that the homopolymerization of acrylates prevents full completion of thiol-ene click addition, therefore
Available online 2 February 2016
leading to high amount of unreacted functional groups in the final material. By contrast, a two-step pro-
cess consisting in off-stoichiometric thiol-Michael addition followed by photopolymerization of residual
Keywords:
acrylates leads to full conversion of thiols. As a consequence, the overall polymerization efficiency is
Photopolymerization
increased as well as the network homogeneity. Taking advantage of this two-step process, a third kind
Click chemistry
Michael addition
of polymer combining aza-Michael addition and photopolymerization was synthesized using amines
instead of thiols. To assess the validity of such process, kinetic profiles and thermomechanical properties
were investigated and compared to the ones obtained with thiols. It is found that the final network prop-
erties are very similar. This would afford more versatile polymer networks thanks to the large variety of
commercially available amines.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction mobility, higher degree of monomer conversion, more homoge-


neous cross-linked networks, and improved substrate adhesion can
UV-curable coatings are VOC-free systems leading to rapid be obtained [10]. Nevertheless, thiol-ene based coatings have poor
ambient crosslinking, reduced energy consumption, and lower mechanical properties (low stiffness and strength) and reduced
equipment space requirement. In most cases, the production of thermal properties (low Tg ) in comparison to UV-cured acrylates
UV-curable coatings proceeds via free-radical photopolymerization [10].
of unsaturated polyesters and of acrylate functionalized oligomers To improve thermo-mechanical properties, thiol-ene step
[1,2]. Through this process, curing occurs via a radical induced growth polymerization can be combined with radical chain poly-
chain growth polymerization which restrains the resin to undergo merization, acrylates standing for enes (one-step thiol/acrylate
a rapid transition from a liquid state to a vitreous solid state strategy) [11,12]. However, in such case, acrylates can react through
[3–6]. Due to this mechanism, the polymerization yields highly both chain growth mechanism (homopolymerization) and step
cross-linked networks at low conversion resulting in a vitrification growth mechanism by abstracting hydrogen from thiol monomers.
phenomenon [2]. This latter reduces molecular mobility, resulting This often results in incomplete thiols conversion [13]. To over-
in non-homogeneous cross-linked network, incomplete conversion come this effect and reach a better control of reaction kinetics and
and shrinkage [4,7,8]. properties [14–17], a two-step thiol/acrylate strategy [9,18,19] was
To overcome these limitations, UV-curable coatings based proposed and combined the catalyzed thiol-Michael addition (step
on thiol-ene click reaction have been developed via a radi- 1) and photopolymerization (step 2). The main thrust of this strat-
cal induced process involving step-growth polymerization which egy is that off-stoichiometric Michael addition leads to full thiols
allows reaching high molecular weights or highly cross-linked conversion and partial consumption of acrylates. Residual acry-
networks at high conversion [2,9]. Therefore, these systems are lates are then able to photopolymerized. The final photopolymer
able to stay in a viscoelastic state up to high conversion result- thus results from the formation of two interlocking networks. By
ing in a shrinkage reduction [2]. Due to this enhanced molecular tuning stoichiometry or by changing thiol and acrylate structures,
it is possible to access a wide variety of properties [15].
However, the limited choice of thiols restricts their use for a
∗ Corresponding author. large scale operation [9]. For this reason, heteroatomic nucleophiles
E-mail address: xavier.allonas@uha.fr (X. Allonas). involving nitrogen, oxygen and phosphorus have to be considered

http://dx.doi.org/10.1016/j.porgcoat.2016.01.023
0300-9440/© 2016 Elsevier B.V. All rights reserved.
52 M. Retailleau et al. / Progress in Organic Coatings 100 (2016) 51–55

aza-Michael addition), the mixture was finally photo-cured (step


2).

2.2.4. Film preparation


For DMA and DSC characterizations, mixtures were coated
onto a polypropylene film and were finally light-cured using a
395 nm Firejet FJ200 LED from Phoseon placed on a conveyor belt
(dose 7 J cm−2 , 395 nm). Regarding RT-FTIR investigation, mixtures
were first spread between two polypropylene films and laminated
between two CaF2 pellets using a Teflon spacer (50 ␮m). Then, sam-
ples were cured using a Roithner LaserTechnik LED395-66-60-110
(395 nm) at 70 mW cm−2 during 2 min.

2.3. Differential scanning calorimetry (DSC)

Thermal properties of the different components were evaluated


using Q200 from TA Instruments. Products were placed in hermet-
Scheme 1. Chemical structures of the materials used in this study. ically sealed aluminium pans (between 5 and 10 mg) and heated
from −80 to 30 ◦ C for Michael addition product (after step 1), and
from −20 to 70 ◦ C for photopolymers (10 ◦ C min−1 ramp).
as viable alternatives [20]. Herein, we suggest a strategy using aza-
Michael addition by replacing thiols by amines. Indeed, amines
2.4. Real time Fourier transform Infrared spectroscopy (RT-FTIR)
are widely marketed through epoxy and polyurethane technolo-
gies [21]. In addition, no catalyst is required as amines can act
Samples were laminated between two polypropylene foils and
as both nucleophiles and bases [22,23]. In this study, we will
CaF2 pellets using a Teflon spacer. The reactivity was monitored
first investigate the benefits of producing polymer via the two-
by RT-FTIR (Vertex 70, Bruker). Spectra were recorded with a time
step thiol/acrylate strategy compared to the one-step thiol/acrylate
resolution of 0.261 s and with a 4 cm−1 spectral resolution. Disap-
strategy. Then, two-step strategies involving either thiols or amines
pearance of the acrylate C C double bonds at 1640 cm−1 (stretching
will be compared in terms of polymerization kinetics and thermo-
of C C), thiol function at 2510 cm−1 (stretching of S H) and amine
mechanical properties.
function at 3380 cm−1 (stretching of N H) were followed with irra-
diation time. Conversion of acrylate, thiol and amine was then
2. Material and methods calculated, respectively, according to:
 At

2.1. Materials Conversion (%) = 1− × 100 (1)
A0
Ethoxylated Bisphenol A Diacrylate (SR349) and Irgacure where A0 and At stand for the absorbance of the considered vibra-
819 (phenyl-bis (2,4,6-trimethylbenzoyl)-phosphine oxide) tion band at time 0 and time t, respectively.
were obtained from Sartomer and BASF, respectively.
1,5-Diaminopentane (cadaverine), 1,5-pentanedithiol and tri- 2.5. Dynamic mechanical analysis (DMA)
ethylamine (TEA) were purchased from Sigma Aldrich (Scheme 1).
All compounds were used as received. The dynamic thermo-mechanical properties of the UV-cured
materials were investigated with a Q800 DMA (TA Instru-
2.2. Preparation of samples ments) in the tensile configuration. The samples were rectan-
gular (12.8 mm × 5.3 mm × 0.070 mm) free films removed from
2.2.1. One-step strategy polypropylene substrates. Temperatures ranged from −20 to 100 ◦ C
Formulations were prepared by mixing ethoxylated bisphenol A and the heating rate was set at 2 ◦ C·min−1 . The amplitude and fre-
diacrylate (SR349), 1,5-pentanedithiol and the radical photoinitia- quency of the oscillatory deformations were adjusted to 15 ␮m and
tor Irgacure 819 (2.5 wt% relative to amount of thiol and acrylate). 1 Hz, respectively.

2.2.2. Two-step strategy (thiol) 3. Results and discussion


Formulations were prepared by first mixing ethoxylated bisphe-
nol A diacrylate (SR349), the catalyst trimethylamine (0.8 wt% 3.1. One-step thiol/acrylate strategy
relative to amount of thiol and acrylate) and the radical photoinitia-
tor Irgacure 819 (2.5 wt% relative to amount of thiols and acrylates). The thiol-acrylate radical photopolymerization was first carried
Once it was dissolved, 1,5-pentanedithiol was added to the mix- out for different thiol/acrylate molar ratios under 395 nm light. The
ture according to the amount of acrylate C C double bonds. After initiating radicals formed from the irradiated photoinitiator are
300 min storage at room temperature in the dark (end of the ther- expected to initiate both the thiol-ene step-growth process and
mal thiol-Michael addition, i.e. step 1), the mixture was finally the acrylate free radical homopolymerization. As an example, 1:4
photo-cured (step 2). molar ratio FTIR kinetic is plotted in Fig. 1. Corresponding conver-
sions and polymer glass transition temperatures (Tg ) are shown in
2.2.3. Two-step strategy (amine) Table 1.
Formulations were prepared by first mixing ethoxylated bisphe- In agreement with the literature [13], thiol conversion was
nol A diacrylate (SR349) and the radical photoinitiator Irgacure found to be incomplete (between 39% and 53%) whatever the
819 (2.5 wt% relative to amount of amines and acrylates). Once it thiol/acrylate molar ratio used. This is merely attributed to the
was dissolved, 1,5-diaminopentane was added to the mixture. After high efficiency of the acrylate homopolymerization. Differences
80 min storage at room temperature in the dark (end of the thermal between the kinetic constant of the acrylate propagation and the
M. Retailleau et al. / Progress in Organic Coatings 100 (2016) 51–55 53

process). These results show that the thiol-ene step-growth mech-


anism is in competition with the homopolymerization of acrylate,
leading to incomplete conversion and loss of process control.

3.2. Two-step thiol/acrylate strategy: a thiol-Michael addition


followed by photopolymerization

This process implies two orthogonal reactions [14–16], the first


step involving a Michael addition between the acrylate (SR349)
and a thiol (1,5-pentanedithiol) in the presence of triethylamine
(0.8 wt%) as catalyst. Then, after full completion of step 1 (300 min),
a second step consisting in the radical photopolymerization of
residual acrylates can be triggered by light in the presence of
Irgacure 819 (2.5 wt%). Both step 1 and step 2 kinetics were fol-
lowed by FTIR and then DSC measurements were carried out
(Table 2). For a better understanding, 1:4 molar ratio FTIR kinetic
is plotted in Fig. 2.
Fig. 1. One-step thiol/acrylate strategy FTIR kinetics for 1:4 molar ratio.
As shown in Table 2, step 1 leads to final conversions which
are governed by the initial thiol/acrylate stoichiometry. Thiols are
Table 1 fully consumed during the reaction. The corresponding polymers
Conversion (%) and Tg (◦ C) obtained after the one-step photopolymerization for exhibit low Tg values, due to the presence of significant amounts
different thiol/acrylate molar ratios.
of unreacted acrylates. Then, these acrylate groups are photocured
SH/C C Thiol conversion (%) Acrylate conversion (%) Tg (◦ C) during the step 2 and consequently the Tg values were significantly
1:2 53 88 35 increased from +38 ◦ C for 1:2 to +68 ◦ C for 1:8 ratio. The extent of
1:4 48 76 60 this Tg increase merely depends on the content of acrylate func-
1:8 39 65 70 tional groups remaining after the first reaction [16]. Furthermore,
higher final acrylate conversions were obtained with this two-step
process compared to the one-step thiol/acrylate strategy. However,
kinetic constant of the chain transfer reaction from an acrylic radi- the fact that all thiols react with the acrylates during step 1 low-
cal to a thiol functional group by hydrogen abstraction account for ers the Tg due to the high flexibility introduced in the network.
this result [13]. Regarding acrylate conversion, it increased by 23% Therefore, the Michael addition strategy allows a full conversion
when increasing thiol content from 1:8 to 1:2. This effect is ascribed of thiols and delays the vitrification. To further compare both sys-
to the plasticizing effect of thiol chains which delays the gel point tems, network homogeneity can be characterized. For that purpose,
and therefore, the vitrification [24]. Concerning the Tg , a drastic DSC thermograms were used to determine the slope of the glass
drop from 70 ◦ C to 35 ◦ C is noticed due to the presence of thiols transition which is known to characterize the network homogene-
that give rise to the formation of a linear polymer (via a step growth ity [25,26]. As an example, DSC thermogramms of photopolymers

Table 2
Conversion (%) and Tg obtained after step 1 and step 2 for the two-step thiol/acrylate process based on thiol-Michael addition for different thiol/acrylate ratios.

SH/C C Thiol-Michael addition: step 1 Acrylate photopolymerization: step 2

Thiol conversion (%) Acrylate conversion (%) Tg (◦ C) Acrylate conversion (%) Tg (◦ C)

1:2 100 50 −25 *nm 13


1:4 100 25 −31 88 21
1:8 100 12.5 −36 80 32

*nm: not measurable.

Fig. 2. Two-step thiol/acrylate strategy FTIR kinetics for 1:4 molar ratio.
54 M. Retailleau et al. / Progress in Organic Coatings 100 (2016) 51–55

Table 4
Tg obtained via aza-Michael and thiol-Michael additions after 80 min and 300 min,
respectively, for different molar ratios.

Molar ratio Tg (◦ C)

Aza-Michael Thiol-Michael

1:1 −18 −21


1:2 −22 −25
1:4 −30 −31

Table 5
Acrylate conversion (%), glassy and rubbery modulus (MPa) and slope of the glass
transition (W·g−1 ·◦ C−1 ) obtained after step 2 of the two-step thiol/acrylate and
amine/acrylate strategy for the 1:4 molar ratio.

Two-step Two-step
amine/acrylate thiol/acrylate

Conversion (%) 79 ± 3 88 ± 3
Fig. 3. DSC thermogramms of photopolymers obtained by both two-step and one- Slope of the glass transition (W g−1 ◦ C−1 ) −3.6 × 10−3 −3.9 × 10−3
step thiol/acrylate strategies for the 1:4 molar ratio. Glassy modulus (MPa) 2305 ± 164 2160 ± 170
Rubbery modulus (MPa) 16 ± 1 22 ± 1

Table 3
Slope of the glass transition (W·g−1 ·◦ C−1 ) obtained after the one-step and two-step
thiol/acrylate photopolymerization for different thiol/acrylate molar ratios. show that the substitution of thiol by amine is quite relevant for
the first step.
Molar ratio Slope of the glass transition (W g−1 ◦ C−1 )
To check if the aza-system is still consistent for a two-step
One-step Two-step strategy, the photopolymerization was carried out for the molar
thiol/acrylate thiol/acrylate ratio 1:4, after full completion (80 min) of the aza-Michael addi-
strategy strategy
tion. Kinetics and thermomechanical properties of photopolymers
1:2 −5.6 × 10−3 −6.4 × 10−3 were hence compared with the two-step thiol/acrylate strategy
1:4 −2.9 × 10−3 −3.6 × 10−3 (1:4 molar ratio). Data are gathered in Table 5.
1:8 −1.5 × 10−3 −2.9 × 10−3
A first observation to be made is that the photopolymerization
process is more effective using the thiol-Michael process. Indeed,
the two-step thiol/acrylate strategy leads to a final conversion of
obtained by both two-step and one-step strategies for the 1:4 molar
88% whereas in the presence of amines the final conversion was
ratio are plotted in Fig. 3.
79%.
For all ratios, the slope of the glass transition in the two-step pro-
However, mechanical properties are quite close taking into
cess is higher compared to the one-step process (Table 3). These
account the standard deviation. Indeed glassy modulus is equal
data suggest that the transition is sharper for the two-step pro-
to 2305 MPa and 2160 MPa for the aza- and thiol-Michael pro-
cess and occurs in a narrower temperature range meaning that a
cesses, respectively. A slightly difference is noticed for the rubbery
more homogeneous polymer network has been formed. This is a
modulus which is lower with the aza-process compared to the
clear evidence that the two-step thiol/acrylate strategy allows more
thiol-process due to lower acrylate conversion. In addition, what-
homogeneous network formation while decreasing the possibil-
ever the nucleophile used, both two-step processes involving either
ity of monomers to migrates. Thus, using a two-step thiol/acrylate
aza-Michael or thiol-Michael exhibit comparable network homo-
strategy polymerization allows a better control of the process and
geneity.
properties.
Finally, even if a less efficient photopolymerization step occurs
with the amine, the aza-Michael alternative is promising since it
3.3. Comparison between two-step thiol/acrylate and can provide quickly a step 1 polymer without catalyst and a final
amine/acrylate strategies material with equivalent thermomechanical properties.

As discussed above, there is a limited choice of commercially


4. Conclusion
available thiols. The use of amines can thus appears to be suitable
nucleophiles for the first Michael addition step [25,26]. Amines
Through this study, it has been proven that using a two-step
have the benefits to be available in a wide variety [21]. In addi-
strategy to produce polymer networks affords a better control of
tion, it appears that amines can act as both nucleophiles and bases
the process and final properties. Nevertheless, this strategy requires
meaning that no additional catalyst is needed [20].
thiols which are difficult to use as starting materials. To overcome
To make a proper comparison, thermal aza-Michael addition
this issue, we found that using amines as nucleophile is possible
and thiol-Michael addition were first compared. This was done by
without significantly affecting the thermo-mechanical properties.
choosing appropriate amine and thiol which have a similar back-
It even, improves the processability. Thus, the use of amines opens
bone: 1,5 diaminopentane and 1,5 pentanedithiol. Aza-Michael
up a new field of possible starting materials and hence, new mate-
addition and thiol-Michael addition kinetics were followed by FTIR
rial properties.
for different molar ratios and showed that the time required to
complete the reaction with the amine is 4 times shorter than that
observed for the corresponding thiol-Michael reaction. Tg values, References
evaluated at the end of each reaction by DSC, are displayed in
[1] P. Glöckner, Radiation Curing: Coatings and Printing Inks; Technical Basics,
Table 4. Applications and Trouble Shooting, Vincentz Network, 2008.
It can be noted that there is a slight difference (3 ◦ C) between [2] R. Schwalm, Chapter 4 – Raw materials, in: UV Coatings, Elsevier, Amsterdam,
Tg obtained with the two different Michael additions. These results 2007, pp. 94–139.
M. Retailleau et al. / Progress in Organic Coatings 100 (2016) 51–55 55

[3] J.G. Drobny, Radiation Technology for Polymers, 2nd ed., Taylor & Francis, reactive thiol-acrylate composite polymer systems, Macromol. Symp. 329
2012. (2013) 101–107.
[4] R. Schwalm, Chapter 2 – The UV curing process, in: UV Coatings, Elsevier, [16] D.P. Nair, N.B. Cramer, J.C. Gaipa, M.K. McBride, E.M. Matherly, R.R. McLeod, R.
Amsterdam, 2007, pp. 19–61. Shandas, C.N. Bowman, Two-stage reactive polymer network forming
[5] D. Durand, C.-M. Bruneau, Reactivity and gelation. I. Intrinsic reactivity, J. systems, Adv. Funct. Mater. 22 (2012) 1502–1510.
Polym. Sci. Part B: Polym. Phys. 17 (1979) 273–294. [17] S. Chatani, T. Gong, B.A. Earle, M. Podgórski, C.N. Bowman, Visible-light
[6] D. Durand, C.-M. Bruneau, Reactivity and gelation. II. Substitution effect, J. initiated thiol-michael addition photopolymerization reactions, ACS Macro
Polym. Sci. Part B: Polym. Phys. 17 (1979) 295–303. Lett. 3 (2014) 315–318.
[7] J.G. Kloosterboer, G.F.C.M. Lijten, The influence of vitrification on the [18] A.B. Lowe, Thiol-ene “click” reactions and recent applications in polymer and
formation of densely crosslinked networks using photopolymerization, in: materials synthesis: a first update, Polym. Chem. 5 (2014) 4820–4870.
Biological and Synthetic Polymer Networks, Springer, Netherlands, 1988, pp. [19] M.J. Kade, D.J. Burke, C.J. Hawker, The power of thiol-ene chemistry, J. Polym.
345–355. Sci. Part A: Polym. Chem. 48 (2010) 743–750.
[8] J. Kloosterboer, Network formation by chain crosslinking [20] B.D. Mather, K. Viswanathan, K.M. Miller, T.E. Long, Michael addition
photopolymerization and its applications in electronics, in: Electronic reactions in macromolecular design for emerging technologies, Prog. Polym.
Applications, Springer, Berlin Heidelberg, 1988, pp. 1–61. Sci. 31 (2006) 487–531.
[9] C.E. Hoyle, C.N. Bowman, Thiol-ene click chemistry, Angew. Chem. Int. Ed. 49 [21] K.L. Forsdyke, T.F. Starr, Thermoset resins market report, Rapra Technol.
(2010) 1540–1573. (2002).
[10] C.E. Hoyle, T.Y. Lee, T. Roper, Thiol-enes: chemistry of the past with promise [22] G. Gonzalez, X. Fernandez-Francos, A. Serra, M. Sangermano, X. Ramis,
for the future, J. Polym. Sci. Part A: Polym. Chem. 42 (2004) 5301–5338. Environmentally-friendly processing of thermosets by two-stage sequential
[11] H. Wei, A.F. Senyurt, S. Jönsson, C.E. Hoyle, Photopolymerization of ternary aza-Michael addition and free-radical polymerization of amine-acrylate
thiol-ene/acrylate systems: film and network properties, J. Polym. Sci. Part A: mixtures, Polym. Chem. (2015).
Polym. Chem. 45 (2007) 822–829. [23] M. Retailleau, A. Ibrahim, C. Croutxé-Barghorn, X. Allonas, C. Ley, D. Le Nouen,
[12] A.F. Senyurt, H. Wei, C.E. Hoyle, S.G. Piland, T.E. Gould, Ternary One-pot three-step polymerization system using double click michael
thiol-ene/acrylate photopolymers: effect of acrylate structure on mechanical addition and radical photopolymerization, ACS Macro Lett. (2015)
properties, Macromolecules 40 (2007) 4901–4909. 1327–1331.
[13] N.B. Cramer, C.N. Bowman, Kinetics of thiol-ene and thiol-acrylate [24] A.B. Lowe, C.N. Bowman, Thiol-X Chemistries in Polymer and Materials
photopolymerizations with real-time fourier transform infrared, J. Polym. Sci. Science, Royal Society of Chemistry, 2013.
Part A: Polym. Chem. 39 (2001) 3311–3319. [25] C. Arends, Polymer Toughening, Taylor & Francis, 1996.
[14] D.P. Nair, N.B. Cramer, M.K. McBride, J.C. Gaipa, R. Shandas, C.N. Bowman, [26] A.J. Scott, A. Nabifar, A. Penlidis, Branched and crosslinked polymers
Enhanced two-stage reactive polymer network forming systems, Polymer 53 synthesized through NMRP: quantitative indicators for network
(2012) 2429–2434. homogeneity? Macromol. React. Eng. 8 (2014) 639–657.
[15] D.P. Nair, N.B. Cramer, M.K. McBride, J.C. Gaipa, N.C. Lee, R. Shandas, C.N.
Bowman, Fabrication and characterization of novel high modulus, two-stage

You might also like