Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Polymer 179 (2019) 121663

Contents lists available at ScienceDirect

Polymer
journal homepage: www.elsevier.com/locate/polymer

Design, synthesis, and properties of novel amino-ester and amino-ester- T


alcohol polymer backbones
Mark F. Sonnenschein*, Kshitish Patankar, Justin Virgili1, Thomas Collins, Benjamin Wendt
The Dow Chemical Company, Corporate Research and Development, Midland, MI, 48674, USA

H I GH L IG H T S

• Aza-Michael chemistry is utilized to prepare novel materials.


• Aza-Michael networks hybridized with epoxy networks tied through amine co-reactants create new amino-ester-alcohol backbones with interesting properties.
• Hybridized networks possess very good adhesive and compressive properties.
• Hybridized networks are able to form to very high conversion in the presence of large amounts of water or organic contamination.

A R T I C LE I N FO A B S T R A C T

Keywords: Simultaneous polymerization of polyacrylic, polyamine, and polyepoxy monomers offers the potential to gen-
Aza-Michael erate a novel polymer backbone with new and beneficial properties. The Aza-Michael addition of amines to
Polymer design acrylates is fast and exothermic and can thus increase the rate of amine epoxy reactions producing poly amino-
Hybrid polymer ester-alcohol polymer backbones. The polymerization is shown to be remarkably tolerant of organic and aqueous
solvent interactions. The resultant polymers are transparent and possess good compressive and thermal prop-
erties that are amenable to polymer design. The polymers of this article are predicated on efficient poly-
merization of the acrylate and amine functionalized monomers, so preliminary rules for their polymerization are
provided as a preamble to the hybrid polymers with polyepoxy monomers.

1. Introduction nucleophile (Michael donor) consists of an amine that adds to an acti-


vated α,β-unsaturated carbonyl-containing compound (Michael ac-
Aza-Michael addition chemistry provides a flexible route for mac- ceptor) [8,9]. In many cases aza-Michael chemistry has been used for
romolecular design [1–3]. The large number of available polyamines making linear and branched backbones in solution [10–12]. Acrylate
and polyacrylates, as well as the ability of this reaction to proceed to based Michael acceptors resulting in the formation of amino-ester
completion accompanied by an adequate exotherm, makes aza-Michael monomer, and the synthesis of linear and branched poly(amino-esters)
chemistry particularly suitable for developing new and useful polymer in solution has been reported [13–15]. While many of these reports
architectures [4]. In addition, this chemistry can often be accomplished have emphasized the potential for these backbones in biotechnology,
without needing a solvent making it an efficient chemistry as well [5,6]. the potential of amino-esters as building blocks for thermoset chemis-
This article describes a new class of polymer resins comprising the si- tries, similar to urethane and epoxy networks, has received less atten-
multaneous co-reaction of an aza-Michael network and an epoxy net- tion. In this article, the bulk synthesis, characterization and physical
work using co-reactant polyamines to force network hybridization to properties of aza- Michael networks are explored. In addition, hybrid
the nanometer scale. We have previously published research demon- networks of poly(amino-esters) and poly(amino-alcohols) are prepared
strating the application and structure-property relationships of carbon- and characterized by stimulating simultaneous aza-Michael and amine-
Michael chemistry applied to their utility for making foams and elas- epoxy polymerizations producing poly(amino-ester-alcohol) backbones.
tomers [7]. Relative to carbon-Michael chemistry, aza- Michael addi- Polymer physical properties as a function of reactant structure is ex-
tion refers to a subset of Michael addition chemistries in which the plored. Reaction kinetics and selectivity of the aza-Michael addition are

*
Corresponding author.
E-mail address: mfsonnenschein@dow.com (M.F. Sonnenschein).
1
Current address: Tactus Technology, Fremont CA 94538, USA.

https://doi.org/10.1016/j.polymer.2019.121663
Received 30 April 2019; Received in revised form 15 July 2019; Accepted 18 July 2019
Available online 19 July 2019
0032-3861/ © 2019 Published by Elsevier Ltd.
M.F. Sonnenschein, et al. Polymer 179 (2019) 121663

investigated as a function of temperature and shown to proceed rapidly testing. An Instron 5567A load frame was used to characterize the
over a broad temperature range when the Michael acceptor is an ac- tensile properties per ASTM D638. Type IV samples for compression
rylate, and an order of magnitude more slowly when a methacrylate. were prepared by mixing in a speedmixer and then pouring the mixture
The hybridization of the poly(meth)acrylates with epoxy monomers is into a syringe with internal diameter of 12.7 mm. After the sample had
accomplished by their common co-reactant amines, and the reaction cured, the syringe was cut from the sample, and the sample then sliced
exotherm which harmonizes reaction rates forces co-continuity of the into sections about 25.4 mm(height) for characterization per ASTM
network structure. D695.
Shear adhesion strength of resins for bonding metal specimens was
2. Experimental characterized by a single lap joint geometry per ASTM D1002-10.
Aluminum substrates were cleaned with soapy water, acetone and
1-(2-aminoethyl)piperazine (AEP) and isophorone diamine(IPDA) isopropanol and allowed to dry. Aluminum substrates were
were obtained from Sigma-Aldrich. 1,4-butanediol diacrylate (1,4-BDO 152 mm × 25.4 mm x 3 mm (LxWxT). A 25.4 × 25.4 mm area of the
diAc), 1,3 butanediol diacrylate (1,3 BDO-diAc), trimethylolpropane substrate was coated with resin, a second substrate pressed against it,
triacrylate (TMPTA), poly(propylene oxide) diacrylate (800 g/mol, and secured with binder clips. Consistent resin thickness was achieved
PPG800 diAc) poly(ethylene oxide) diacrylate (575 and 700 g/mol with standard 100 μm spacer spheres sprinkled into the resin. Excess
(PEG575 diAc and PEG700 diAc) ethylene glycol dimethacrylate (EG resin exuded from the overlap area was removed using a razor blade.
diMAc) and diethylene glycol dimethacrylate (DEG diMAc) were ob- The lap-shear samples were allowed to cure at ambient conditions for
tained from Sigma-Aldrich. Ditrimethylolpropane triacrylate (di- one and seven days, and then tested on an Instron 4201 test frame at
TMPTA) was obtained from Sartomer. All materials were used as re- 0.1 mm/min with a 10 kN load cell. Precautions were taken to ensure
ceived. In the case of PEG575 diAc, PEG 700 diAc, and PPG 800 diAc, that the sample long axis was parallel to the sample orientation, and
the molecular weight of the starting material was determined by through the centerline of the grip assembly. Ten lap joint specimens
quantitative 13C NMR and determined to be 532, 614, and 825 g/mol were tested for shear strength.
respectively. Epoxy resin (epoxy equivalent weight 187g/eq) was ob- The solvent stability of selected resins was characterized for the
tained from the Dow Chemical Co. following solvents: water, hexane and xylene. The cured resin samples
Gel permeation chromatography was performed on an Agilent 1100 were immersed in the solvent for 24 h, the sample blotted dry to the
series instrument using THF as the solvent. A PEG molecular weight touch, and the weight uptake measured.
calibration was utilized to convert retention time into molecular The effect of solvent contamination (water and diesel fuel) on cure
weight. were studied using the following method. The diesel fuel was an ad-
Differential scanning calorimetry (DSC) was performed on a TA ditive free formulation provided by British Petroleum. The auto-igni-
Q1000 series instrument. Samples were heated to 100–150 °C (de- tion temperature and closed-cup flash point of the fuel were 257 °C
pending on sample) at 10 °C/min, and cooled to −100 °C at the same and > 38 °C respectively. The water was obtained from a still producing
rate. The heating and cooling cycle was subsequently repeated. mega-ohm resistance water. All samples for solvent contaminated cure
Thermogravimetric analysis was performed on a TA TGA 5000 series studies were prepared by adding the solvent to a homogeneous mixture
instrument. Samples were equilibrated at 40 °C and then heated to of the acrylate and epoxy components. The AEP or IPDA was then
500 °C at 10 °C/min in an air environment. added to the contaminated amine reactive components and homo-
Temperature dependent viscosities were measure using a strain genized using a speedmixer. Viscosity increase was monitored on the
controlled Rheometric Scientific ARES II rheometer with parallel plate Ares II rheometer using a cone and plate geometry and otherwise the
geometry (25 mm) and a stress controlled TA Instruments G2 same conditions described earlier for cure-rate measurements.
Rheometer with cone and plate geometry (60 mm, 2°). Strain sweeps
were measured at 0 and 60 °C to ensure recorded viscosities are strain 3. Results and discussion
independent. Typical experimental conditions include: ω = 1–2 rad/s,
50–99% strain (strain controlled) or 1 Pa stress (stress controlled) and a 3.1. Synthesis and characterization of aza-Michael networks
temperature sweep from 0 to 60 °C (2 or 4 °C/min) followed by a
cooling and re-heating cycle. In all cases solid samples were pre-con- 3.1.1. Aza-Michael reaction of AEP with PPG800-diAc at a fixed
ditioned above their melting point prior to measurement and no sig- temperature
nificant difference in the viscosity profile was observed in the heating Preliminary attempts to react AEP with PPG800 diAc at room
versus cooling sample. temperature led to gelation with an accompanying high exotherm. This
Cure rate was measured by the rise in viscosity of the 2 K resin (vide led to standardizing the use of acrylate as a limiting reagent with active
infra). The resin components were first thoroughly mixed using a removal of heat in order to mitigate crosslinking and moderate reaction
speedmixer and transferred quickly into the TA ARES II rheometer. The rate. Therefore, PPG800 diAc was added dropwise (150 g, 0.188 mol,
isothermal viscosity build was monitored at 10, 25 and 65 °C as a 0.376 eq acrylate functionality) into 45.44 g AEP (0.352 mol, 1.06 eq
function of time. The test parameters were as mentioned above. The test NH functionality) in a 3-neck flask (−10 °C) with N2 purge (Scheme 1).
was stopped when the viscosity reached 104 Pa s (107 cP). The shelf-life The disappearance of the acrylate vinyl protons was monitored by 1H
of non-amine component mixtures was measured by monitoring its NMR (d-CHCl3) (δ = 6.3 (d, 1H), δ = 6.1 (q, 1H), δ = 5.7 (d, 1H)) with
viscosity as a function of temperature after aging at 40 °C for 30 days. respect to the proton on the polypropylene oxide end group
Samples for tensile testing were prepared by mixing the amines and (δ = 4.8–5.0 (m, 1H) (Fig. 1). The maximum recorded temperature
non-amine components in a speed mixer and then casting the reacting during reaction was 5 °C. Kinetics for the reaction between AEP and
mixture into a silicone mold and allowed to cure overnight under ASTM PPG800 diAc showed the reaction to be 88% complete in 15 min and
conditions. The tensile bars were removed from the mold at 24 h for 100% complete in less than 2.5 h (13C data can be found in

Scheme 1. Synthesis of AEP functionalized


PPG-diAc. Michael addition only through
the secondary amine of AEP is shown. See
Table 1 for actual distribution of primary vs.
secondary addition products.

2
M.F. Sonnenschein, et al. Polymer 179 (2019) 121663

Fig. 1. 1H NMR (d-CHCl3) of (a) PPG 800 diAc and (b) AEP functionalized PPG800 diAc product using a reaction temperature of −10 °C. Minimal vinyl acrylate
bonds (δ = 5.8–6.4) are evident in Fig. 1b.

3
M.F. Sonnenschein, et al. Polymer 179 (2019) 121663

Scheme 2. Synthesis of AEP functionalized


BDO-diAc. Michael addition through the
secondary and primary amine of AEP is
shown. See Table 2 for actual distribution of
primary vs. secondary addition products.

Fig. 2. 13
C NMR (0.02 M Cr(acac) in DMSO‑d6) of reaction product of 1,4-BDO diAc and AEP using a reactor temperature of −5 °C. Peak k is unresolved from e*.

supplemental on-line information). It is evident that the aza-Michael reactions) with a nitrogen purge. 50 g of 1,4 BDO diAc (0.25 mol, 0.50
addition proceeds rapidly during the addition of diacrylate due to in- eq. acrylate) was added at 0.6 ml/min using a peristaltic pump with
creased temperature and high concentration of reactive groups. constant stirring of the reaction. In all cases the reaction was allowed to
proceed for 1 h and disappearance of the acrylate vinyl protons were
3.1.2. Aza-Michael reactions of AEP with 1,4 BDO-diAc at various confirmed by 1H NMR (d = 6.3 (d, 1H), d = 6.1 (q, 1H), d = 5.7 (d,
temperatures 1H) Fig. 3). The maximum recorded temperature was measured using a
The Michael addition is known [1,9] to proceed more rapidly with a thermocouple immersed in the reaction mixture. Product was collected
secondary amine compared to a primary amine [1,8], and the reaction without purification and characterized by quantitative 13C NMR and
between AEP and 1,4 BDO-diAc was selected to study aza-Michael GPC. Characterization of the AEP-1,4 BDO-diAc product is summarized
addition as a function of temperature (Scheme 2). in Table 1. In Fig. 2 detailed 13C NMR assignments are provided (13C
All reactions were run by adding 67 g AEP (0.52 mol, 1.56 eq. NH) NMR of neat 1,4 BDO diAc is provided in SOM The vinyl carbons are at
to a 500 ml kettle (for reactions ≤ 20 °C or a 3-neck flask (for 40 °C 128 and 132 ppm). It is important to note that oligomeric poly(amino-

4
M.F. Sonnenschein, et al. Polymer 179 (2019) 121663

Table 2
Summary of reaction conditions and analytical product characterization of AEP
functionalized acrylates synthesized at constant stoichiometry (3.075:1 NH
(meth)acrylate). In all cases the starting substrate contains acrylate function-
ality unless denoted with † indicating methacrylate functionality. “n.d.” in-
dicates no data available. Mol% product determined by quantitative13C NMR.
Acrylate Functionality Control Maximum Mol% Wt% Mol%
substrate (theoretical) Temp recorded Product Product 1°
(1,4 C) Reaction amine
Temp (oC)

1,4- BDO 2 −10 7 83 95 97


1,3-BDO 2 −10 < 10 96 99 97
EG† 2 40 45 76 92 94
DEG† 2 40 47 45 76 n.d
PPG800 2 −10 <5 95 99 98
PEG575 2 23 37 88 98 85
PEG700 2 23 35 77 96 100
TMPTA 3 −10 3 60 89 100
di-TMPTA 4 5 2 64 93 n.d

3.1.3. Aza-Michael reaction of AEP with EG diMAc at a fixed temperature


Synthesis of AEP functionalized EG diMAc was executed according
to Scheme 3. Michael addition through both the primary and secondary
amine of AEP is observed; however; the substitution (determined by 13C
NMR analysis) is primarily through the secondary amine due to its
greater nucleophilicity. AEP (66.81 g, 0.517 mol (1.55 eq) was added to
a heated (40 °C) 3-neck round bottom flask and placed under a nitrogen
environment with stirring. EG diMAc (50.0 g, 0.252 mol, 0.50 eq) was
added drop-wise to the reaction. The highest recorded temperature was
45 °C during EG diMAc addition. The reaction between AEP and EG
diMAc was observed to proceed considerably more slowly than the
reaction of BDO diAc at −10 °C for instance (8 h vs 2 h for full con-
version) and thus the synthesis temperature of all methacrylate-based
systems were run at 40 °C without observation of gelation, homo-
polymerization, or other undesired side reactions. Spectra and assign-
ments are provided in SOM.
Fig. 3. Differential scanning calorimetry of aza-Michael products (A) the re-
action product of AEP and 1,4 BDO-diAc and (B) the reaction product of AEP
and TMP-triacryclate (TMPTA). 3.1.4. Properties and characteristics of aza-Michael networks with various
Michael acceptors
Reaction conditions and product composition of the functionalized
Table 1
acrylates are summarized in Table 2. In the case of fn = 2 (meth)ac-
The effect of reaction temperature on the product characteristics of the reaction
of AEP with 1,4 BDO diAc Scheme 2. Mol% 1° amine product indicates the rylates, final product compositions were observed to be ≥ 70 mol%
percentage of primary amine functionality remaining after reaction of AEP and (≥89 wt%) of the reaction product of AEP and the di(meth)acrylate
BDO diAc. based on quantitative 13C NMR analysis except for the aza-Michael
product of AEP and DEG diMAc (45 mol%). In all cases, except for the
Control Maximum Mol% Mol% 1° Mn Mw
Temperature (oC) recorded Temp product Amine of (g/ (G/
AEP-DEG diMAc, the remaining product composition is excess AEP.
(oC) product mol) mol) Increased variation in aza-Michael product is observed through addi-
tion of both the secondary and primary amine of AEP with the 2-
−5 8 77 98 132 175 functional (meth)acrylate samples (85–100 mol% 1° amine bearing
5 19 78 97 123 155
product), and is likely related to the varying electronic and steric en-
20 33 85 96 126 159
40 55 73 90 122 149 vironments present [8,9,16]. The presence of excess AEP in the final
product is related to both the AEP to (meth)acrylate stoichiometric
ratio of 3.075 to 1 (NH to (meth)acrylate) employed to mitigate un-
ester)s will also form in the reaction; however, due to the similarity in wanted homopolymerization and/or side reactions, uncertainty in the
13
C NMR peak positions of the different products it is not possible to true functionality, molecular weight of the starting materials, and some
quantify the amount. The result indicates that as temperature increases, uncharacterized oligomerization. In the case of higher functionality
selectivity of the secondary amine for the Michael acceptor decreases, acrylates such as TMPTA (fn = 3) and di-TMPTA (fn = 4) the true
as would be expected. functionality of the starting acrylate is lower than theoretical due to
incomplete conversion of hydroxyl groups to acrylate in the starting
materials.

Scheme 3. Synthesis of AEP functionalized


EG diMAc. Michael addition through both
the secondary and primary amine of AEP is
shown. See Table 2 for actual distribution of
primary vs. secondary addition products.

5
M.F. Sonnenschein, et al. Polymer 179 (2019) 121663

Table 3
Summary of structural, thermal and rheological characterization of AEP functionalized acrylates synthesized at constant stoichiometry (3.075:1 NH: (meth)acrylate).
In all case the reaction was with acrylate except where denoted with a “†” indicating methacrylate functionality. “__” indicates that no thermal transition was
detected, and “nd” signifies no available data. AHEW is amine hydrogen equivalent weight.
Acrylate Functionality (theo) AHEW (g/NH eq., theo) Tg (oC) Tm (oC) Td,100oC % η(23 °C) mPa.s η(40 °C) mPa.s

1,4 BDO 2 114 −45 72 97 2120 530


1,3 BDO 2 114 −58 __ n.d n.d. n.d.
EG† 2 114 −40 __ 90 22,000 8000
DEG† 2 125 −33 __ 92 114,000 13,560
PPG800 2 218 −54 __ 96 1757 536
PEG575 2 198 −49 5 94 8000 2300
PEG700 2 271 −52 7 90 3600 1000
TMPTA 3 114 −20 __ 97 59,000 9500
di-TMPTA 4 123 −15 __ 96 72,000 9800

term maintenance of a liquid product appears to be governed by crys-


tallization.
Thermo-oxidative stability was assessed using TGA (10 °C/min) in
an air environment (Fig. 4). Td,100oC is defined as the weight percent of
product retained upon heating to 100 °C and provides a measure of
volatility (Table 3, AEP(neat) b.p. = 219–222 °C). Volatilized species
were not identified. In Fig. 4 it is evident that both the absolute mass
loss and rate of mass loss of neat AEP at 100 °C is significantly greater
than any of the aza-Michael products characterized. The expected
correspondence between the weight percent of residual AEP and Td,100oC
of the various products presented in Table 3 is confirmed and can be
explained by AEP volatilization and/or degradation. The high mole-
cular weight (≥500 g/mol theor.) aza-Michael products from fn = 2
substrates presented in Fig. 4 possess both lower absolute mass loss, and
lower rate of mass loss at 250 °C compared to lower molecular weight
(≤500 g/mol theor.) aza-Michael products. This suggests that thermo-
oxidative degradation and volatilization processes may proceed
through a reverse aza-Michael reaction [18,19]. The anomalously low
Fig. 4. Thermogravimetric analysis of AEP and the reaction products of viscosity for the AEP-PEG700-diAc suggests that conversion of this
PPG800-diAc and AEP and trimethylolpropane triacrylate and AEP. The pro- substrate to the diamine telechelic was incomplete, possibly due to
ducts are thermally stable relative to the incorporated AEP. Differences in incomplete conversion of the PEG700 to the diacrylate.
product stability may reflect reverse Michael reaction or intrinsic factors.
For many applications, product viscosity is a critical characteristic
of practical utilization [20,21]. In Fig. 5, the temperature dependence
Equivalent weights, thermal properties, and viscosities of the AEP of the viscosity of aza-Michael products is presented. In all cases, no
functionalized acrylates are summarized in Table 3. GPC traces of some significant difference in the viscosity profile is observed in the heating
of the aza-Michael products are shown in supplemental on-line in- versus the cooling cycle and so described as thermo-rheologically
formation. Apparent GPC molecular weights are based on poly(ethylene simple [22]. All the di-functional products exhibit viscosity less than
oxide) calibration standards. Peak breadth from ca. 16–20 min was about 20,000 mPa-s while the fn = 3 and fn = 4 products exhibit much
influenced by strong column interaction with AEP based on the pure greater room temperature viscosity. This is presumably due the bran-
material elutions. In many other samples a high intensity peak is evi- ched structure of the higher functionality products. The 10× higher
dent and is hypothesized to correspond to the aza-Michael product. viscosity of the EG dimAc compared to 1,4 BDO diAc is perhaps un-
Furthermore, in no sample is signal discernible at t < 16 min sug- expected given the similarity in molecular weight and excess AEP
gesting minimal oligomeric poly(amino-ester) formation. (Fig. 5A). The disparity suggests that there may be a critical difference
The Tg of all products derived from di-functional polyether (meth) in the number of oligomeric species, that pendant methyl groups of the
acrylates (fn = 2) (Fig. 3A, Table 3) were observed to be < -30 °C dimAc end groups create additional hydrodynamic drag as the chains
consistent with the polyether backbones present in the (meth)acrylate slip past one another [23,24], or that the closer proximity of the ac-
starting materials. Higher functionality, lower molecular weight pro- rylate functionality in the EG-dmAc based material creates a higher
ducts (fn ≥ 3) possess Tg's in the vicinity of −15 to −30 °C. (Fig. 3b). cohesive energy density within the fluid [24]. This is consistent with
The presence of pendant methyl groups either along the backbone of the corresponding states of these materials wherein the material with
the spacer (1,3 BDO diAc or PPG800 diAc), or in the form of a me- the lower glass transition temperature (1,4 BDO diAc-diAEP adduct)
thacrylate end group (EG diMAc or DEG diMAc) was observed to inhibit has the lower viscosity [24–26], but in this case the viscosities are not
crystallization of the aza-Michael product(data not shown)[17]. made the same by a simple temperature shift of the data by an amount
Melting transitions are observed in products following long storage equal to about the difference in Tg. Also surprising is the higher visc-
(months) in which the spacer is known to crystallize (1,4-BDO diAc, osity exhibited by PEG575 diAc compared to PEG700 diAc (Fig. 5B),
PEG575 diAc, and PEG700 diAc). In the case of TMPTA and di-TMPTA, but would also be consistent with a model whereby the number and
transitions were observed via DSC at greater than 100 °C however; due proximity of acrylate end-groups in these materials plays a significant
to the breadth and noisiness of the transition it is unclear whether the role in product viscosity. Additional viscosity data are provided in
transition is caused by melting, or by AEP volatilization. This set of supplemental online information.
observations suggests that crystallization of the aza-Michael products is
driven by the spacer length and backbone type between the acrylate
end groups. Shelf-life stability which is defined in this context as long-

6
M.F. Sonnenschein, et al. Polymer 179 (2019) 121663

Fig. 5. Viscosity temperature dependence of aza-Michael reaction product of AEP with (A) 1,4 BDO-diAc and EG diMAc and (B) PEG 575 –diAc and PEG 700 diAc.

Table 4
Properties of elastomers with amino-ester backbones prepared via aza-Michael
reaction of polyacrylates and polyamines as a function of polyacrylate func-
tionality. In all cases the amine was aminoethyl piperazine. See supplemental
on-line information for tensile properties of benchmark TPU materials.
Stoichiometry of N-H to acrylate is 1:1.

Scheme 4. Components of an aza-Michael composition to improve material BDO-diAc/ Functionality Tensile Elong. @ Ultimate
TMPTA-diAc modulus MPa break % tensile
properties by introducing crosslinks via introduction of TMPTA.
(mole ratio) strength MPa

100/0 2.0 liquid NA NA


3.1.5. Polymer networks based on aza-Michael chemistry
96/4 2 0.4 ± 0.2 353 ± 177 0.7 ± 0.2
Amino-ester thermoplastics and thermosets have been prepared 87.5/12.5 2.125 0.65 ± 0.4 231 ± 15 0.8 ± 0.1
before [8,27–29]. The application of these materials has been limited, 67/33 2.33 2.2 ± 1.0 74 ± 2 1.2 ± 0.5
because amino-ester properties are relatively poor compared to epoxy 50/50 2.5 3.8 ± 0.7 74 ± 2 1.8 ± 0.5
and urethane based materials. We found that substantially linear ma- TPU (Shore A 2 3.5 > 2000 34
70)
terials based on 1,4 BDO diacrylate and aminoethyl piperazine pro-
TPU (Shore A 3 2 > 1500 14
duced poor materials, and progressive substitution of the 1,4 BDO diAc 60)
with trimethylolpropane triacrylate (Scheme 4) improved initial tensile
modulus slightly, but did not result in materials with useful properties
(Table 4). monomer such as bisphenol-A glycerolate-diAc. Such a substitution can
A modification of the aliphatic amino-ester polymer presented increase glass transition temperature with concomitant increase in
above is to substitute some of the BDO-diAc with a more rigid co- tensile modulus, and tensile strength. Scheme 5 presents the general

7
M.F. Sonnenschein, et al. Polymer 179 (2019) 121663

Scheme 5. Formulation of aza-Michael product to improve mechanical properties through addition of crosslinks (TMPTA) and by stiffening the backbone via
addition of aromatic epoxy diacrylate.

Table 5 Table 6
Compressive mechanical properties of amino-ester alcohol polymers from 1,4 Formulation and tensile properties of amino-ester-alcohol polymer backbones
BDO diAc, TMPTA, epoxy acrylate (see Scheme 5) with AEP. Stoichiometry of from reaction of epoxy resin, PPG 800 diAc, TMPTA with AEP. Stoichiometry of
N-H to acrylate is 1:1. N-H to acrylate and epoxy is 1:1.
BPA diAc/1,4-BDO Max Stress Max Young's Water Epoxy resin/PPG800 DiAc/ Max Stress Max Strain Young's modulus
diAc/TMPTAin (MPa) Strain (%) modulus Resistance TMPTA in equiv. (MPa) (%) (MPa)
equiv. (MPa)
30/15/31 3.2 489 1
1/1/0.4 18 1382 68 Extremely poor 34/8/34 15.2 529 8.7
1/1/0.6 20 955 52 36/3/36 Brittle
1/1/1.5 8 518 41 23/5/47 10.3 462 8.8
1/1/1 Too brittle to test 44/11/22 21 450 87
51/12/12 22 406 93
56/14/7 16 258 102
59/14/3 11.5 315 78
reaction and the expected topology. Topologically there is no difference
between the use of the acrylate functionalized bisphenol-A and BDO-
diAc, but the difference in chain dynamics is shown to be significant in
diAc. Another simple modification would be to substitute bisphenol-A
defining tensile properties (Table 5).
diglycidyl ether for the corresponding glycerolate. The epoxy resin re-
The thermoset properties described in Table 5 do not represent a
acts into the amino ester backbone by reaction of the oxirane ring with
simple linear effect of substitution of the acrylate functionalized bi-
either amine of the aminoethyl piperazine (or adducts with acrylate).
sphenol-A for the TMPTA. Increasing the triacrylate has the expected
While the reactivity of the amines with acrylates vs. the amines and
effect of reducing the elongation at break, and the maximum stress by a
oxiranes are different [32], as are their activation energies [9,33–37],
corresponding amount. However; the Young's modulus also decreased
the exotherm of the system and the preferred reaction of acrylates with
substantially suggesting that changes in the glass transition temperature
the more nucleophilic secondary amines, and the oxiranes with the
due to the bisphenol-A adduct had an effect on the solid state properties
more reactive and less hindered primary amines, made resulting
[30] apart from the effect of crosslinks and the molecular weight be-
backbones homogeneous possessing a single glass transition tempera-
tween them [31]. One possibility is that the TMPTA is the most reaction
ture (vide infra – Fig. 8). As such, we propose the following network
limited and its over-incorporation inhibits complete network formation.
structure for this backbone in Scheme 6. Formulations screening the
performance of this structural design were created and are detailed in
3.1.6. Hybrid aza-Michael and epoxy networks Table 6.
In Table 5 the poor water resistance of aza-Michael networks is The formulations of Table 6 primarily probe the relative contribu-
noted. This is no doubt due to the very large number of polar groups tions of substitution of rigid epoxy resin for “soft” PPG diAc, and sub-
along the backbone with very strong hydrogen bond potential. When stitution of di-functional rigid epoxy resin for tri-functional crosslinking
these materials were allowed to sit quiescently in water they became TMPTA.
very highly swollen. Design of backbone structure to mitigate unwanted Statistical analysis of the data in Table 6 (detailed analysis provided
factors could include for instance reducing the frequency of backbone in on-line supplemental information) suggests that the compositional
carbonyls by substituting polypropylene oxide diacrylate for the BDO-

Scheme 6. Polymer design to improve mechanical properties of aza-Michael derived networks by including bisphenol-A materials to create amino-ester-alcohol
polymer backbone networks.

8
M.F. Sonnenschein, et al. Polymer 179 (2019) 121663

Fig. 6. Model predictions compared to measured results for (A) strength (maximum stress) and (B) modulus of amino-ester alcohol backbones of material in Table 6
based on composition and cross term of epoxy substitution with other amine reacting components. See supplemental on-line information for model details and
arithmetic expression.

changes have strong and intuitively predictable effects on polymer


properties. Prediction of maximum stress (Fig. 6A) based on formula-
tion variables (i.e. equivalence ratios) resulted in a square of the cor-
relation coefficient (r2) of 0.97 and adjusted r2 of 0.80 due to the large
number of variables (which included cross terms) relative to data points
(also reflected in the root mean square error “RMSE”). Incorporation of
epoxy resin had twice the positive effect on maximum stress (also
termed “strength”) that TMPTA evidenced. Substitution of epoxy resin
for PEG800 diAc within the boundaries of these formulations also in-
creased measured strength. Analysis of the material property compo-
nents of strength include elongation at break, and the material's tensile
modulus. Individually, the correlation of polymer formulation to break
elongation yielded r2 of 0.96 adjusted to 0.85 for the large number of
fitting variables. Substitution of epoxy resin for PPG800 diAc had the
largest effect on the elongation at break, but the effect was weaker than
might have been expected. This may be due to the relative low mole-
cular weight of the PPG based soft segment, and the improved network
structure offered by epoxy resin substitution owing to the possibly in-
complete acrylate substitution of the PPG 800 component. Statistical
analysis of the dependence of Young's modulus (Fig. 6B) on formulation
variables yields a fair model with r2 of 0.92 and adjusted r2 of 0.5. This
low adjusted correlation coefficient is due to the small amount of data
Fig. 7. Comparison of mechanical properties for amino-ester-alcohol backbones relative to the number of variables (also reflected in the substantial
for a formulation where IPDA and AEP are interchanged. All formulations RMSE), and the poor distribution of data essentially fitting into two
consisted of 10.6 gr epoxy resin, 5.63 gr PPG800-diAc, and 0.69 gr TMPTA with regions [38]. Bimodality of the modulus data may reflect non-con-
stoichiometric equivalence of either AEP or IPDA. tinuous vs co-continuous network formation of rigid components within
the polymer bulk [39]. This is validated by non-linear analyses (pre-
sented in supplemental on-line information) showing that while the
concentration of bis-epoxy has a very strong and linear effect on
modulus, the other components have weaker and less systematic re-
lationships to the data.
While the prior exposition of structure based properties explored the
effects of varying components producing ester and alcohol backbone
functionalities, the effect of the amine curing reagent can also result in
significant modification of polymer properties. Fig. 7 presents data on a
controlled formulation in which the only variable was the amine,
comparing the properties using AEP (trifunctional) versus isophorone
diamine (IPDA, tetra-functional).
In comparison to AEP, the properties of IPDA formulated backbones
exhibit a higher glass transition temperature resulting from either a
more highly crosslinked structure, or simply impeded chain motions
from the bulkier IPDA [40,41]. Fig. 8 compares dynamic mechanical
Fig. 8. DMA analyses of amino-ester-alcohol polymer networks from formula-
tion in Fig. 7. The formulation cured with IPDA exhibits a higher storage analyses (DMA) for these formulations revealing the significant effect
modulus and a higher glass transition temperature. on chain mobility that can be affected by straightforward polymer de-
sign. Fig. 8 shows the AEP containing polymer possesses a Tg of about

9
M.F. Sonnenschein, et al. Polymer 179 (2019) 121663

Fig. 9. Compressive stress-strain curve for amino-ester-alcohol polymer networks. The higher yield stress is consistent for the formulation containing IPDA is
consistent with the higher glass transition temperature of this polymer.

38 °C while the IPDA material Tg is about 50 °C. We propose that this provided in supplemental on-line materials.
difference in glass transition temperature is the primary feature con-
trolling the material property differences shown in Fig. 7 [42–45]. This 3.2. Adhesion
difference is further dramatized by inspection of the compressive stress
strain curve shown in Fig. 9. Adhesive properties of the poly(amino-ester-alcohol) backbones was
Cure kinetics for amino-ester-alcohol backbones produced by hybrid evaluated for their adhesion to aluminum and nylon 6 substrates.
aza-Michael-epoxy components are shown in Fig. 10 for curing by AEP Table 7 shows the room temperature adhesion of AEP and IPDA cured
and IPDA. It is clear that backbones cured by AEP are made faster than resins in a lap shear geometry for 1 day and 7 days. These results are
those cured by IPDA. This is presumably due to the greater nucleo- furthermore compared to a generic 2-part epoxy cured at room tem-
philicity of the AEP secondary amine, and the subsequent greater re- perature for 7 days (ostensibly the recommended cure schedule), and
activity with acrylate functional components of the aza-Michael reac- following 60 °C cure for 24 h. The epoxy formulation consisted of equal
tion. The aza-Michael reaction is anticipated to occur first with the mass ratios of epoxy resin and a room temperature curing agent CAP-
accompanying exotherm stimulating the epoxy reaction. The tempera- CURE™ 3830-81 (a mercaptan based curative, trademark of Gabriel
ture dependence of the time-to- gelation presents an Arrhenius form as Performance Products) as per manufacturer's recommendation. The
shown in Fig. 11. Given the complications of this system comprising data shows that the room temperature amino ester alcohol polymer
two different addition reactions (aza-Michael and epoxy-amine) in adhesive generates equivalent adhesion to aluminum as the commercial
which the different reactions compete for one reagent (the amine), it is epoxy formulation. This is perhaps unsurprising since all formulations
not intuitive that the temperature dependence should be well fit to an are about half epoxy resin. However; the aza-Michael hybrid formula-
Arrhenius form. However; given the preponderance of epoxy functional tions of this article also possess a much lower mass ratio of curative, and
material (~75 mol% of amine reactive moiety) it is perhaps reasonable a significant amount (~25% by mass) of relatively soft acrylic func-
that in as much as the dependence is linear, the calculated activation tional components that cures without the need of added catalyst. The
energy is close to previous measures of the epoxy-amine reaction acti- adhesive failure was cohesive in the adhesive for bonding to aluminum,
vation energy -about 10 kcal/mol (~40 kJ/mol) [36]. This is also and adhesive at the interface for the nylon 6 bonded substrates. The
consistent with the observation that while the rates of polymerization presence of ester linkages in the aza-Michael epoxy hybrid did not
are dependent on the amine reactive component, the activation energy degrade metal adhesion and adhesion to nylon was in keeping with
is somewhat insensitive to these particular amine co-reactants. reported epoxy acrylate hybrids [46]. The difference in adhesion of the
The amino-ester-alcohol polymer backbone presents a relatively experimental polymers to nylon and aluminum is related to the factor of
diverse chemical environment with which to interact with solvents. 20 difference in surface energy between aluminum and nylon. Epoxy
Designs cured with AEP and IPDA were studied for comparison. As resins (surface energy 45 dyne/cm) are about the same as nylon resins
indicated in Table 5, the aza-Michael networks exhibited very poor suggesting a similar affinity of the surface for the bulk as the surface
water resistance, and so differences between the epoxy hybrid and the [47,48].
aza-Michael network result from the presence of the epoxy amine ad-
dition product, and accompanying crosslinked network properties. 3.3. Solvent/aqueous effects on polymerization
Fig. 12 shows the relative solvent uptake of IPDA and AEP cured ma-
terials to water, hexane (aliphatic) and xylene (aromatic) following a In addition to tunable properties, solvent resistance, and good ad-
24 h room temperature immersion. The IPDA cured material exhibited hesion properties, the amino-ester-alcohol networks described in this
better water and hexane resistance, and slightly worse xylenes re- article also show the unusual property of proceeding to completion
sistance, however; the trends for both structures were similar. It is even in the presence of a significant amount of organic or aqueous di-
possible that the difference in water resistance is a result of the in- lution. Prior work has indicated that small amounts of water (up to 6 wt
creased crosslink density of IPDA concomitant with its higher glass percent) can accelerate epoxy-amine curing kinetics [49,50], while
transition temperature. Pictures of the solvent exposed samples are other publications have discussed the problems of such contamination

10
M.F. Sonnenschein, et al. Polymer 179 (2019) 121663

Fig. 10. Curing kinetics for amino-ester-alcohol network polymers emplying AEP or IPDA amine curing agent in the formulation. (A) Reaction at 25 °C (B) Reaction at
65 °C. The more rapid kinetics with AEP probably reflect the faster Michael reaction with the AEP secondary amine.

with amine transport and availability. Other articles have discussed the backbones such as epoxy and amides [51–54]. In the case of the current
problems of curing epoxy networks and amide-imide networks in the amino-ester-alcohol networks, it appears that diesel fuel (a possible
presence of organic solvents [51–54]. We have observed that the ester- contaminant of conceivable field application processes), while modestly
alcohol-amine networks of this publication form quite readily in the slowing the reaction, does not significantly alter the attainment of cure.
presence of a significant amount (up to 33%) of water or diesel fuel In addition, it is notable that polymer cured with AEP and diluted with
(typically a mix of aliphatic, cycloaliphatic and aromatic components). 20% diesel fuel, still achieves a faster cure rate than the same for-
As cited in previous publications, we observed a significant rate en- mulation cured with IPDA. This is a dramatic illustration of the driving
hancement of the formation of the ester-amine-alcohol networks using force of the secondary amine reaction with acrylic moiety in the overall
the current formulations, even at 20% addition of water to the for- formation of these amino-ester alcohol polymers. Fig. 14 provides a
mulation, with no appearance of amine bloom at the surface. However; photograph of the appearance of the amino-ester-alcohol networks of
as shown in Fig. 13, there was little increase in rate for water additions this research in their diluted state. Additional data on the diluted curing
greater than 5% addition. Organic solvent additions have been reported and viscosities of the polymers of this article can be found in on-line
to be detrimental in many cases to network formation of polar polymer supplemental information.

11
M.F. Sonnenschein, et al. Polymer 179 (2019) 121663

Fig. 11. Pseudo-Arrhenius behavior of amino-ester-alchohol networks achieving the polymer gel point. The slopes are nearly the same probably reflecting the
preponderance of epoxy reaction that must occur with primary amine functionality regardless of the amine employed.

Fig. 12. Solvent resistance of amino-ester-alcohol networks described in Fig. 8 to various solvents.

Table 7
Single lap-shear data for hybrid polymers from co-reaction of aza-Michael and epoxy backbone components compared to a room temperature cured epoxy for-
mulation on aluminum and nylon 6. The aza-Michael-epoxy hybrids are nearly independent of amine curing agent.
Amino-ester-alcohol AEP cured (MPa) Amino-ester-alcohol IPDA cured (MPa) Epoxy (MPa)

aluminum, 1-day cure 13 ± 1 12 ± 1 12 ± 1


aluminum, 7 day cure 17 ± 0.5 18 ± 2 19 ± 1
aluminum, 1 day cure 60 °C 12 ± 0.5 10 ± 1
nylon 6, 1 day cure 4 + 0.5 3±1
Nylon 6, 7 day cure 5+1 5±1

4. Conclusion and selective through a secondary amine when available. When the
reaction is between an acrylate moiety and a secondary amine, com-
Aza-Michael products based on the reactions of amines and (meth) plete conversion is reached in less than 1 h from between −10 °C to
acrylates have been synthesized in bulk, and characterized using a 40 °C. Rate of reaction was seen to proceed an order of magnitude more
variety of techniques. Reaction kinetics and selectivity of AEP func- slowly when the moiety was methacrylate instead of acrylate. Viscous
tionalized acrylates show that the addition is generally very efficient, properties of the aza-Michael products were shown to be strongly

12
M.F. Sonnenschein, et al. Polymer 179 (2019) 121663

Fig. 13. Cure of amino-ester-alcohol networks in the presence of large dilutions


of (A) water at 40 °C and (b) commercial diesel fuel at 65 °C. In (A) the for-
mulation in the absence of water is the slowest curing, while in the presence of
diesel fuel each formulation cures slower, however; the diesel dilute AEP for-
mulation cures faster than the uncontaminated IPDA formulation indicating the
importance of the secondary amine in the aza-Michael cure kinetics.

dependent on the molecular weight and functionality of the acrylate


spacer. The backbone of the aza-Michael product is very amenable to
Fig. 14. From top to bottom: Cured resin, cured resin + diesel 1:1 (w:w) resin:
modification by changing amine or acrylate components, but in general
diesel, cured resin + diesel 4:1 by volume, cured resin + water 4:1 by volume.
the networks formed possess inferior mechanical properties to other
networked polymers such as polyurethanes or epoxies. The amino-ester
backbone produced by the aza-Michael reaction can be greatly modified Appendix A. Supplementary data
by hybridization with amine reactive components such as epoxies. The
reaction with epoxies can be achieved without need of a catalyst taking Supplementary data to this article can be found online at https://
advantage of the exotherm from the aza-Michael reaction to accelerate doi.org/10.1016/j.polymer.2019.121663.
the amine/oxirane addition. When hybridized with the diglycidylether
of bisphenol-A, the mechanical and chemical properties of the network References
improve greatly. This change in network properties is then reflected in
the hybrid polymer adhesive properties for instance. Conversion of [1] B. Mather, K. Viswanathan, K. Miller, T. Long, Prog. Polym. Sci. 31 (2006) 487-431.
[2] S. Williams, B. Mather, K. Miller, T. Long, J. Polym. Sci. A Polym. Chem. 45 (2007)
monomers to hybrid polymer formation is enhanced in the presence of 4118–4128.
up to about 30 vol percent water. Conversion of monomer to hybrid [3] D. Wu, Y. Liu, C. He, T. Chung, S. Goh, Macromolecules 37 (2004) 6763–6770.
polymer in the presence of up to about 30 vol percent organic solvent [4] K. Jin, N. Wilmot, W. Heath, J. Torkelson, Macromol. 49 (2016) 4115–4123.
[5] B. Lin, A Computational Mechanistic Study of an Aza-Michael Addition, Auburn
such as diesel fuel is slowed down by dilution, but conversion never- University, Dissertation, May 2015.
theless goes to completion. [6] J. Deng, Z. Dai, J. Yan, M. Sandru, R. Spontak, L. Deng, J. Membr. Sci. 570–571
(2019) 455–463.
[7] M. Sonnenschein, J. Werness, K. Patankar, X. Jin, M. Larive, Polymer 106 (2016)
128–139.
Acknowledgements [8] D. Wu, Y. Liu, C. He, T. Chung, S. Goh, Macromolecules 37 (2004) 6763–6770.
[9] G. Desmet, D. D’hooge, P. Omurtag, P. Espeel, G. Marin, F. Du Prez, M.-F. Reyniers,
The authors would like to thank Dr. Cassie Kopke Marker for help J. Org. Chem. 81 (2016) 12291–12302.
[10] A. Pellis, P. Hanson, J. Comerford, J. Clark, T. Farmer, Polym. Chem. 10 (2019)
with some of the experiments, and express appreciation to The Dow 843–851.
Chemical Company for its support of this work. Dr. Michael Desanker is [11] J. Wang, R. Cooper, H. He, B. Li, H. Yang, Macromolecules 51 (2018) 6111–6118.
thanked for his generous help with the manuscript. [12] A. Genest, D. Portinha, E. Fluery, F. Ganachaud, Prog. Polym. Sci. 72 (2017)
61–110.
[13] Y. Yu, Z. Wei, X. Leng, Y. Li, Polym. Chem. 9 (2018) 5426–5441.
[14] S. Patil, V. Shinde, R. Misra, K. Devesh, J. Polym. Sci. A Polym. Chem. 56 (2018)

13
M.F. Sonnenschein, et al. Polymer 179 (2019) 121663

2080–2095. [36] M. Pramanik, E. Fowler, J. Rawlins, Polym. Eng. Sci. 54 (2014) 1990–2004.
[15] G. Ozturk, T. Long, J. Polym. Sci. A Polym. Chem. 47 (2009) 5437–5447. [37] M. Kour, R. Gupta, R. Bansal, Chem. Sel. 2 (2017) 8465–8470.
[16] R. Connor, W. McClellen, J. Org. Chem. 3 (1938) 570–577. [38] J. Mijovic, J. Wijaya, Macromolecules 27 (1994) 7589–7600.
[17] M. Sonnenschein, M. Greaves, B. Bell, B. Wendt, Ind. Eng. Chem. Res. 51 (2012) [39] P. Goos, D. Meintrup, Statistics with JMP: Hypothesis Tests, ANOVA and
8386–8393. Regression, first ed., John Wiley and Sons Pubs W. Susse, U.K., 2016, pp. 494–552.
[18] Y. Cail, M.-X. Luo, K.-F. Yang, G.-Q. Lai, J.-X. Jiang, L.-W. Xu, Chirality 23 (2011) [40] M. Sonnenschein, Polyurethanes: Science, Technology, Markets, and Trends, John
397–403. Wiley and Sons Pub., Hoboken NJ, 2014, pp. 138–143.
[19] M. Rosello, J. Acena, A. Fuentes, C. delPozo, Chem. Soc. Rev. 43 (2014) 7430–7453. [41] C. Estridge, Polymer 141 (2018) 12–20.
[20] J. Dealy, J. Wang, Melt Rheology and its Application in the Plastics Industry, second [42] M. Perry, T. Ebrahimi, E. Morgan, P. Edwards, S. Hatzikriakos, L. Schafer,
ed., Springer Pubs, New York, 2013 (Chapter 1). Macromolecules 49 (2016) 4423–4430.
[21] T. Davenport (Ed.), Phys, vol. 3, 1968, pp. 139–143. [43] D. Anokhin, M. Gorbunova, Y. Estrin, V. Komratova, E. Badamshina, Phys. Chem.
[22] H.-C. Tseng, J.-S. Wu, R.-Y. Chang Phys, Chem. Chem. Phys. 12 (2010) 4051–4965. Chem. Phys. 18 (2016) 31769–31776.
[23] L. Fetters, D. Lohse, T. Witten, A. Zirkel, Macromolecules 27 (1994) 4639–4647. [44] V. Tereshatov, M. Makarova, V. Senichev, E. Volkova, Z. Vnutskikh,
[24] J. Kim, C. Biag, J. Chem. Phys. 144 (2016) 1–5. A. Slobodinyuk, Colloid Polym. Sci. 293 (2015) 153–164.
[25] 3rd, J. Bicerano (Ed.), Prediction of Polymer Properties, Marcel Dekker Pub., New [45] B. Silva, R. Bello, C. Ferreira, L. Coelho, Polym. Int. 67 (2018) 1248–1255.
York, 2002. [46] A. Lesser, E. Crawford, Jour, Appl. Polym. Sci. 66 (1997) 387–395.
[26] V. Nanda, R. Simha, T. Somcynsky, J. Polym. Sci. Part C: Polym Symp. 12 (1966) [47] M. Sonnenschein, S. Webb, R. Cieslinski, B. Wendt, J. Polym. Sci. A Polym. Chem.
277–295. 43 (2007) 989–998.
[27] I. Sanchez, T. Truskett, Jour Phys. Chem.B 24 (1999) 5106–5116. [48] J. Bardis, K. Kedward, J. Compos. Technol. Res. 24 (2002) 30–37.
[28] S. Anantharaj, M. Jayakannan, Biomacromolecules 13 (2012) 2446–2455. [49] J. Bikerman, J. Adhes. 1 (1969) 160–167.
[29] L. Cutlar, Y. Gao, D. Zhou, T. Zhao, U. Greiser, W. Wang, W. Wang, [50] J. Chen, T. Nakamura, K. Aoki, Y. Aoki, T. Utsunomiya, J. Appl. Polym. Sci. 79
Biomacromolecules 16 (2016). (2001) 214–220.
[30] F. Sanda, T. Endo, Macromol. Chem. Phys. 200 (1999) 2651–2661. [51] N. Sharp, C. Li, A. Strachan, D. Adams, R. Pipes, J. Polym. Sci. B Polym. Phys. 55
[31] D. Droste, A. Dibenedetto, J. Appl. Polym. Sci. 13 (1969) 2149–2168. (2017) 1150–1159.
[32] L. Nielsen, J. Macromol. Sci. C Polym. Rev. 3 (1969) 69–103. [52] M. Loos, L. Coelho, S. Pezzzin, Polímeros - Ciência Tecnol. 18 (2008) 76–80.
[33] L.-M. Tang, J. Feng, Y. Wu, Chin. J. Polym. Sci. 25 (2007) 545–553. [53] S. Hong, C. Wu, Thermochim. Acta 316 (1998) 167–175.
[34] M. Zoubir, M. Eldrissi, A. Elhaib, M. Oumou, R. Hammal, N. Mazoir, A. Benharref, [54] Z. Rasheva, L. Sorochynska, S. Grishchuk, K. Friedrich, Xpress Polym. Letts. 9
AElHajbi Jour Mater. Environ Sci. 8 (2017) 990–996. (2015) 196–210.
[35] M. Pramanik, E. Fowler, J. Rawlins, J. Coat. Tech. Res. 11 (2014) 143–157.

14

You might also like