Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

polymers

Article
Effect of Epoxy Structure on Properties of Waterborne Coatings
and Electrical Steel Laminates
Cornelia Marchfelder 1, *, Robert Pugstaller 1 , Gernot M. Wallner 1 , Oliver Brüggemann 2 and
Maëlenn Aufray 3

1 Institute of Polymeric Materials and Testing & Christian Doppler Laboratory for Superimposed
Mechanical-Environmental Ageing of Polymeric Hybrid Laminates (CDL-AgePol), Johannes Kepler
University Linz, Altenbergerstraße 69, 4040 Linz, Austria; robert.pugstaller@jku.at (R.P.);
gernot.wallner@jku.at (G.M.W.)
2 Institute of Polymer Chemistry, Johannes Kepler University Linz, Altenbergerstraße 69, 4040 Linz, Austria;
oliver.brueggemann@jku.at
3 CIRIMAT, Université de Toulouse, CNRS, INP-ENSIACE, T 4 allée Emile Monso- BP44362, CEDEX 4,
31030 Toulouse, France; maelenn.aufray@toulouse-inp.fr
* Correspondence: cornelia.marchfelder@jku.at; Tel.: +43-732-2468-6624

Abstract: Epoxy varnishes are of high relevance to advanced steel laminates for the transformation of
electric energy. Structure–property correlations of epoxy varnishes, coil coatings and electrical steel
laminates are poorly described. Hence, the main objective of this paper was to develop, implement
and evaluate well-defined waterborne model epoxy varnishes for electrical steel laminates, and to
elucidate structure–property correlations. Adhesives with systematically varied equivalent epoxy
weight (EEW) based on bisphenol-A-diglycidyl ether (DGEBA) were investigated and used to
formulate waterborne varnishes. Crosslinking agent dicyandiamide (DICY) was added in an over-
stoichiometric ratio. The waterborne model varnishes were prepared by shear emulsification at
 elevated temperatures. The model varnishes in the A-stage were applied to electrical steel using

a doctoral blade. At a peak metal temperature of 210 ◦ C, the coatings were cured to the partly
Citation: Marchfelder, C.; Pugstaller,
crosslinked B-stage. Coated steel sheets were stacked, laminated and fully cured to C-stage at 180 ◦ C
R.; M. Wallner, G.; Brüggemann, O.;
for 2 h. For laminates with an epoxy adhesive layer in the C-stage, glass transition temperatures
Aufray, M. Effect of Epoxy Structure
on Properties of Waterborne Coatings
(TG ) in the range of 81 to 102 ◦ C were obtained by dynamic mechanical analysis in torsional mode.
and Electrical Steel Laminates. Within the investigated EEW range, a negative linear correlation of EEW and TG was ascertained.
Polymers 2022, 14, 1556. https:// Presumably, higher EEW of the varnish is associated with a less densely crosslinked network in
doi.org/10.3390/polym14081556 the fully cured state. Roll peel testing of laminates at ambient and elevated temperatures up to
140 ◦ C confirmed the effect of EEW. However, no clear correlation of roll peel strength and glass
Academic Editor: Michael Nase
transition temperature was discernible. In contrast, fatigue fracture mechanics investigations revealed
Received: 1 March 2022 that hydroxyl functionality and crosslinking density were affecting the crack growth resistance of
Accepted: 6 April 2022 laminates in a contrary manner. The energy-based fracture mechanics approach was much more
Published: 11 April 2022
sensitive than monotonic peel testing.
Publisher’s Note: MDPI stays neutral
with regard to jurisdictional claims in Keywords: epoxy equivalent weight; waterborne epoxy; thin adhesive layer; electrical steel laminates;
published maps and institutional affil- fatigue fracture mechanics
iations.

1. Introduction
Copyright: © 2022 by the authors.
¨υ
Licensee MDPI, Basel, Switzerland.
This article is an open access article
Electrical steel laminates based on epoxy adhesive layers are of high relevance for the
distributed under the terms and
transformation of electric energy [1–4]. By the stacking and bonding of thin electrical steel
conditions of the Creative Commons sheets, eddy current losses are reduced and the efficiency of electric machines is raised [3].
Attribution (CC BY) license (https:// Epoxy plies act as multifunctional separation layers providing electrical isolation and
creativecommons.org/licenses/by/ full-surface bonding. To maintain the high power density of electric engines or generators,
4.0/). the adhesive layer thickness should be uniform and restricted to a minimum. Lamination

Polymers 2022, 14, 1556. https://doi.org/10.3390/polym14081556 https://www.mdpi.com/journal/polymers


Polymers 2022, 14, 1556 2 of 13

and isolation of electrical steel with epoxy-based adhesives offers many advantages, but
also a few drawbacks. Compared to classical joining methods like welding or interlocking,
the high temperature performance of adhesively bonded laminates is limited [3].
Epoxy adhesives and coatings provide remarkable properties such as good adherence
to various substrates, thermal stability or electrical isolation. Waterborne epoxy varnishes,
especially, allow for multifunctional coatings with a thickness in the µm range. Therefore,
low viscosity varnishes are required. Both solvent-based and waterborne systems have
been developed. Due to environmental and legislative restrictions, waterborne epoxy
varnishes are preferred. Further advantages of waterborne epoxy varnishes are ease of
applicability by coil coating processes, low volatile organic content (VOC) and odor or
reduced flammability [5–7]. In order to enhance interfacial adherence, hydroxyl groups,
allowing for bonds to the oxidic passivation layer of the electrical steel substrate, are
essential. While secondary bonds, such as hydrogen and van der Waals forces, provide
good adherence at ambient conditions, chemical covalent or coordinative bonds are of
special relevance for long-term behavior under hot-humid conditions [8–11].
Liquid epoxy varnishes (A-stage) and coatings are applied in a two-stage process. First,
the electrical steel substrate is coated with the waterborne epoxy varnish system, which
is then dried and partly cured to the B-stage. The coated electrical steel substrates can be
stored for up to several months. In a second step, the sheets are cut to the desired geometry
and stacked. By heat press lamination, a high curing degree of the epoxy adhesive (in
C-stage) is ensured. Processing temperature, time and pressure are dependent on the epoxy
formulation but also on the geometry of the laminate [2,12,13].
Maximum service temperatures of electrical steel laminates are around 180 ◦ C [14].
The glass transition temperature of commonly used epoxy adhesive systems is lower.
Hence, the crosslinking density of epoxy networks, which is usually proportional to the
mechanical properties, is of utmost importance [15–17]. In addition to the bulk properties
of adhesives, interfacial bonding to the passivation layer of the substrate is essential. In
a few studies, structure–property correlations between the epoxy equivalent weight (or
hydroxyl group concentration) of non-waterborne epoxies and adhesive strength at ambient
temperatures have been established [8,10]. A higher EEW value, associated with increased
monomer chain length, results in a reduced crosslinking density and enhanced hydroxyl
group concentration [18]. So far in the literature, no specific attention has been given to the
structure–property correlations of waterborne epoxy varnishes allowing for micrometer
thick adhesive layers in electrical steel laminates.
Hence, the main objective of this paper was to investigate the effect of structural
parameters of waterborne epoxy varnishes on the thermomechanical and mechanical
properties of electrical steel laminates. Varnish systems with varying EEW values were
formulated and applied. Thermomechanical analysis was performed to determine the
glass transition. The adherence strength of laminates was characterized by roll peel testing
at ambient and elevated temperatures. Furthermore, fatigue crack growth kinetics were
examined at ambient conditions.

2. Materials and Methods


2.1. Materials
The investigated epoxy adhesives were based on bisphenol-A-diglycidyl ether (DGEBA).
Three different types of epoxides (Table 1) with varying repeating units n (Figure 1) were
used. Adhesives were classified by their epoxide equivalent weight (EEW) depending on
their repeating unit n. The EEW values ranged from 475 to 1900 g/mol. The letter “A”
indicates Araldite products supplied by Huntsman International LLC (Salt Lake City, UT,
USA). A-1800 represents Araldite GT 6097. In contrast, “D” indicates products from the
D.E.R. series manufactured by Olin Corporation (Clayton, MO, USA). D-900 was delivered
as D.E.R. 664UE and D-500 as D.E.R. 671. Epoxides with an EEW value of more than
400 g/mol are solid. The melting temperature range (TM ) according to the data sheets is
listed in Table 1.
Due to its latent hardening behavior, the curing agent dicyandiamide (DICY) was
selected. A solid micronized DICY powder, DYHARD 100SH (Alzchem, Trostberg,
Polymers 2022, 14, 1556 3 of 13
Germany), with a maximum particle size of 10 µ m was used. According to the data sheet,
the melting point was in the range from 209 to 212 °C. DICY reveals an amino-imino or a
diamino form (see Figure 1).
Table 1. Used epoxy base resins along with the designation, the supplier, the EEW value range and
Table
the 1. Usedmelting
estimated epoxy base resins along
temperature (TMwith
). the designation, the supplier, the EEW value range and
the estimated melting temperature (TM).
Designation Supplier EEW-Range, g/mol Melting Temp., ◦ C
Designation Supplier EEW-Range, g/mol Melting Temp., °C
A-1800 A 1695 to 1885 125 to 135
A-1800
D-900 DA 1695toto930
860 1885 125
100 toto 135
110
D-900
D-500 DD 860toto550
475 930 100
75 toto
85110
D-500 D 475 to 550 75 to 85

Figure 1. Chemical structure of DGEBA with repeating unit n (top) and DICY (bottom) with its two
Figure 1. Chemical
tautomeric structure
forms, diamino of DGEBA
(left) with repeating
and amino-imino unit n (top) and DICY (bottom) with its two
(right).
tautomeric forms, diamino (left) and amino-imino (right).
Due to its latent hardening behavior, the curing agent dicyandiamide (DICY) was se-
2.2. Preparation
lected. of Varnishes,DICY
A solid micronized Coatings and Laminates
powder, DYHARD 100SH (Alzchem, Trostberg, Germany),
with The
a maximum particleof
water content size
theof varnish
10 µm was used. According
formulations to thetodata
was fixed 45 sheet,
m% (of thethe
melting
total
point wasEmulsifier
in the rangePoloxamer
from 209 to407 ◦
212 (Croda,
C. DICYSnaith,
revealsUK)
an amino-imino
system). was used. or a diaminotoform
According the
(see Figure 1).
formulation guideline of the supplier, the emulsifier was 10 m% related to the epoxy base
resin. Depending on the EEW, two different formulation approaches were employed
2.2. Preparation
(Figure of Varnishes,
2 left). For Coatings
solid epoxies withand
anLaminates
EEW value up to 600 g/mol, homogenization of
The water content of the varnish formulations
the base resin and the emulsifier was possible by stirring was fixedat to 45 m% (of the
temperatures total
well system).
below the
Emulsifier Poloxamer 407 (Croda, Snaith, UK) was used. According
boiling point of water (~85 °C). After homogenization, distilled water was added to the formulation
guideline
dropwise at ofan
theelevated
supplier, the emulsifier
temperature under was 10 m%agitation.
constant related toAs the epoxy
such, base resin.
a dropping De-
funnel
pending
was used. onEpoxies
the EEW, twoEEW
with different formulation
values of more thanapproaches
600 g/molwere employed
were dissolved (Figure 2 left).
in acetone
For
priorsolid epoxies with
to addition of thean emulsifier.
EEW valueThis up toallowed
600 g/mol, homogenization of
for homogenization at the base resin
temperatures
and
belowthethe
emulsifier
boilingwas possible
point by stirring
of water. at temperatures
The emulsions were well belowusing
prepared the boiling
an DMX point160
of
water (~85 ◦ C). After homogenization, distilled water was added dropwise at an elevated
emulsifying shear head (Dynamic, Kehl, Germany). Hence, axial and radial flow forces
temperature
were ensuredunder constant agitation.
for optimized As such,
emulsification. For ahigher
dropping funnelweight
molecular was used. Epoxies
systems, with
acetone
EEW values of more than 600 g/mol were dissolved in acetone prior
was evaporated after emulsification. As the waterborne epoxy systems A-1800, D-900 and to addition of the
emulsifier. This allowed for homogenization at temperatures below
D-500 were prone to foaming, 0.75 m% of the defoaming agent BYK-016 (BYK, Wesel, the boiling point of
water.
Germany)The was
emulsions
added.were prepared using an DMX 160 emulsifying shear head (Dynamic,
Kehl,The
Germany).
amountHence, axial agent
of curing and radial
DICY flow forces
was were ensured
calculated for optimized
dependent on the emulsifica-
amine H
tion.
equivalent weight (eq wt), as well as the epoxy equivalent weight. The amine H eq As
For higher molecular weight systems, acetone was evaporated after emulsification. wt
the
valuewaterborne
of DICY was epoxy systems
assumed A-1800,
to be 21 basedD-900
on 4and D-500sites
reactive were prone
[19]. to foaming,
As the 0.75 m%
epoxy equivalent
of the defoaming agent BYK-016 (BYK, Wesel, Germany) was added.
weight, the lowest value in the given EEW range was considered. DICY content, in parts
per hundred (PHR), was calculated according to:
coated. Afterwards, the electric steel sheet with coatings on both surfaces was exposed in
an oven at PMT of 210 °C for 25 s. In this way, partly cured coatings in B-stage could be
applied to both sides. Subsequently, the coated sheets were cut to 200 × 25 mm2 stripes for
roll peel testing and 150 × 25 mm2 for fatigue crack growth characterization. Laminates
consisting of 4 steel layers (roll peel specimen) and 6 steel plies (fatigue crack growth
Polymers 2022, 14, 1556 4 of 13
specimen) were stacked. The laminates were fully crosslinked in a convection oven at a
temperature of 180 °C for 2 h with 0.3 MPa pressure applied.

Figure
Figure 2.
2. Schematic illustration of
Schematic illustration ofvarnish
varnishpreparation
preparation(left)
(left) and
and electrical
electrical steel
steel coating,
coating, stacking
stacking and
and lamination
lamination (right).
(right).

2.3. Characterization
The amount ofofcuring
Raw Materials, Coatings
agent DICY and Laminates
was calculated dependent on the amine H equiv-
alentThe
weight (eq wt), as well as the epoxy equivalent
structure of the epoxides and the emulsifier wasweight. Theassessed
amine Hby eq wt value
infrared
of DICY was assumed to be 21 based on 4 reactive sites [19]. As the
spectroscopy. A Spectrum 100 FTIR (Perkin Elmer, Waltham, MA, USA) in ATR-modeepoxy equivalent
weight,
with the lowest
a MCT value
detector in the
was given
used. TheEEW range wasrange
wavenumber considered. DICYtocontent,
from 4000 600 cmin
−1 parts
was
per hundred (PHR), was calculated according to:
evaluated with 16 scans and a resolution of 4 cm−1. The applicability of the varnish was
examined after coating by optical microscopy. Therefore, a VHX 7000 (Keyence,
amine H eq wt
Mechelen, Belgium) digital PHR microscope
(curing agents was )=employed. Images ×100
were recorded using the (1)
epoxy eq wt
VH-ZST objective at a magnification of 20×.
The thermomechanical behavior
An over-stoichiometric amount of of the
thecuring
laminates
agentwaswascharacterized
added to an by Dynamic
excess of 1.1
Mechanical Analysis (DMA) according to DIN EN ISO 6721 on the rheometer
to 1 (DICY to epoxy) to ensure full crosslinking, as topological restrictions may limit Physica
MCR 502 (Anton
the maximum Paar, Graz,
achievable Austria) in
conversion torsional
[20]. DICY mode. The specimen
was dissolved size was
in methoxy 50 × 10 to
propanol ×
2avoid
mm .agglomeration.
3 DMA experiments were performed in a temperature range from 30 to 220 °C with
a heating
Usingrate of 3 K/min
a suitable and ablade,
doctoral straintheamplitude of 0.1%.were
model varnishes The applied
frequency on was set to 1grade
M800-50A Hz.
Storage modulus
electrical and(according
steel sheets loss factortowereDINevaluated
EN 10106)aswithfunctions of temperature.
a thickness Thecoated
of 0.5 mm. The glass
transition
steel sheetstemperature was deduced
were pre-crosslinked frommetal
at a peak the maximum
temperature of (PMT)
the loss 210 ◦ CA
offactor. schematic
(see Figure 2,
drawing of the
right). Both DMA
sides testelectrical
of the setup issteel
depicted
sheetsinwere
Figure 3 (left).
subsequently coated. The coating on the
top side of the steel sheet was dried for 10 s at 210 ◦ C. Then, the bottom side was coated.
Afterwards, the electric steel sheet with coatings on both surfaces was exposed in an oven
at PMT of 210 ◦ C for 25 s. In this way, partly cured coatings in B-stage could be applied to
both sides. Subsequently, the coated sheets were cut to 200 × 25 mm2 stripes for roll peel
testing and 150 × 25 mm2 for fatigue crack growth characterization. Laminates consisting
of 4 steel layers (roll peel specimen) and 6 steel plies (fatigue crack growth specimen) were
stacked. The laminates were fully crosslinked in a convection oven at a temperature of
180 ◦ C for 2 h with 0.3 MPa pressure applied.

2.3. Characterization of Raw Materials, Coatings and Laminates


The structure of the epoxides and the emulsifier was assessed by infrared spectroscopy.
A Spectrum 100 FTIR (Perkin Elmer, Waltham, MA, USA) in ATR-mode with a MCT
detector was used. The wavenumber range from 4000 to 600 cm−1 was evaluated with
16 scans and a resolution of 4 cm−1 . The applicability of the varnish was examined after
coating by optical microscopy. Therefore, a VHX 7000 (Keyence, Mechelen, Belgium)
digital microscope was employed. Images were recorded using the VH-ZST objective at a
magnification of 20×.
The thermomechanical behavior of the laminates was characterized by Dynamic
Mechanical Analysis (DMA) according to DIN EN ISO 6721 on the rheometer Physica MCR
502 (Anton Paar, Graz, Austria) in torsional mode. The specimen size was 50 × 10 × 2 mm3 .
conditions, three repetitions for each system were carried out, while at 100 °C and 140 °C
the number of repetitions was set to two. Roll peel strength was deduced by averaging
the force normalized by specimen width over the displacement, cutting off the first and
last 25 mm of displacement.
Fatigue crack growth behavior was characterized by an electro dynamic machine
Polymers 2022, 14, 1556 5 of 13
ElectroPlus E3000 (Instron, Norwood, MA, USA) according to ISO 15024. Therefore,
double cantilever beam (DCB) specimens were used. A schematic DCB test configuration
is depicted in Figure 3 (right). Displacement controlled measurements in mode I were
DMA experiments ◦ C with a heating
carried out with awere performed
maximum in a temperature
amplitude of 3 mm.range from 30 to 220
The sinusoidal displacement had a
rate of 3 K/min
frequency of 5 and a strain
Hz and amplitude
a R-ratio of 0.1%. to
(minimum Themaximum
frequencydisplacement)
was set to 1 Hz.of Storage
0.1. A
modulus and loss method
compliance-based factor were
wasevaluated as functions
used to determine of temperature.
the crack length as a The glassoftransition
function number
temperature was deduced from the maximum of the loss factor. A schematic
of cycles. The debonding energy or strain energy release rate was calculated considering drawing of
athe DMAbeam
simple test setup
theoryis approach.
depicted inSpecific
Figure details
3 (left).are given in [3].

Figure
Figure 3.
3. Schematic
Schematic configurations
configurations of
of test
test setups
setups for
for dynamic
dynamic mechanical
mechanical analysis
analysis (DMA,
(DMA, left),
left),
monotonic roll peel strength testing (RPS, middle) and fatigue crack growth assessment (FCG) using
monotonic roll peel strength testing (RPS, middle) and fatigue crack growth assessment (FCG) using
a double cantilever beam (DCB) specimen (right).
a double cantilever beam (DCB) specimen (right).
3. Results and Discussion
Adhesive strength was determined by roll peel testing on a universal testing machine
Z020The chemical Ulm,
(ZwickRoell, structure of theaccording
Germany) epoxy resins
to ENwas 1464.evaluated
A schematic by roll
FTIR spectroscopy.
peel test configu-
IR-spectra are depicted
ration is illustrated in Figure
in Figure 4. Depending
3 (middle). The specimen on the molar
size mass
was 200 × 25 the 2epoxides,
of mm consistingtheof
oxirane ring
4 electrical intensity
steel layers.atOne915ply
cmof−1 varied. The lower the number of repeating units (n), the
the laminate was peeled off from the remaining plies. The
more pronounced
test rate was the oxirane
was 100 mm/min. Specimens ringwere
peak. The oxirane
tested at ambient ringconditions (23 ◦of
functionality C, a50%rh)
DGEBA as
molecule
well as 100 is and
equal140 ◦
to two.
C in aThe intensity of
temperature this peak
chamber. Atleveled
ambientoffconditions,
at EEW valuesthree of 900 and
repetitions
1800 g/mol.
for each Hence,
system were the highest
carried intensity
out, the◦oxirane
while atof100 C and 140 ring◦ Cwas
the detected
number of forrepetitions
epoxy D-500,
was
set toatwo.
with Roll peel
decrease strengthtowards
in intensity was deduced by averaging
the highest the forceA-1800.
EEW adhesive normalized by specimen
By weighting the
width over
oxirane peak the
atdisplacement,
915 cm−1 withcutting
the C-O-Coff the
etherfirststretching
and last 25 mmatof1035
peak displacement.
cm−1 [21] (vertical
dashedFatigue
lines),crack growth
this trend wasbehavior
confirmed. wasThecharacterized
ratio was 0.36 by anfor electro dynamic
D-500 and machine
0.33 for D-900
ElectroPlus
and A-1800. E3000 (Instron, Norwood, MA, USA) according to ISO 15024. Therefore, double
cantilever
Due tobeamwater(DCB)
uptake specimens were no
at the surface, used.clearAtrend
schematic DCB
as to the test configuration
OH-stretching peak atis
depicted
3400 cm inwas
−1 Figure 3 (right). Displacement
discernible. With increasing controlled measurements
n, the intensity of theinOH-absorption
mode I were carriedpeak
out withbecome
should a maximummore amplitude
pronounced. of 3By
mm. The sinusoidal
weighting the peak displacement
at 3400 cm−1 had a frequency
with the CH3
of 5 Hz and
stretching a R-ratio
peak at 2965(minimum to maximum
cm−1 (vertical displacement)
dashed lines), of 0.1. Afor
a clear indication compliance-based
higher amounts
method was used to determine the crack length as a
of OH-groups was ascertained for A-1800 or D-900, compared to D-500. The function of number of cycles. The
peak ratios
debonding energy or strain energy release rate was calculated considering
were 0.36 for A-1800 and D-900, and 0.32 for D-500. These findings are in good agreement a simple beam
theory
with approach.
data Specific details
in the literature [22,23].are given in [3].
3. Results and Discussion
The chemical structure of the epoxy resins was evaluated by FTIR spectroscopy. IR-
spectra are depicted in Figure 4. Depending on the molar mass of the epoxides, the oxirane
ring intensity at 915 cm−1 varied. The lower the number of repeating units (n), the more
pronounced was the oxirane ring peak. The oxirane ring functionality of a DGEBA molecule
is equal to two. The intensity of this peak leveled off at EEW values of 900 and 1800 g/mol.
Hence, the highest intensity of the oxirane ring was detected for epoxy D-500, with a
decrease in intensity towards the highest EEW adhesive A-1800. By weighting the oxirane
peak at 915 cm−1 with the C-O-C ether stretching peak at 1035 cm−1 [21] (vertical dashed
lines), this trend was confirmed. The ratio was 0.36 for D-500 and 0.33 for D-900 and A-1800.
Polymers 2022,
Polymers 14, x
2022, 14, 1556
FOR PEER REVIEW 6 6of
of 13
13

Figure 4. FTIR-ATR spectra in the range from 3700 to 2700 cm−1 (left) and 1100 to 850 cm−1 (right)
Figure 4. FTIR-ATR
of epoxies spectra
with average EEW invalues
the range from
of 500, 3700
900, to1800
and 2700 g/mol.
cm−1 (left) and 1100 to 850 cm−1 (right) of
epoxies with average EEW values of 500, 900, and 1800 g/mol.
Due to water uptake at the surface, no clear trend as to the OH-stretching peak at
3400The −1 was discernible.
cmformulated waterborne Withepoxy varnishes
increasing were
n, the applied of
intensity to the
electrical steel, driedpeak
OH-absorption and
should become moreinpronounced. By weighting theatpeak at metal
3400 cm − 1 with theofCH
pre-cured to coatings B-stage in a convection oven a peak temperature 2103
stretching
°C. The coatingpeakthickness
at 2965 cm in−B-stage
1 (vertical dashed lines), a clear indication for higher amounts
was 5 µ m for all epoxy thermosets. Figure 5 shows an
of OH-groups
exemplary was ascertained
comparison of the epoxyfor A-1800 or D-900,
resin D-500 withcompared
the coating to in
D-500. TheThe
B-stage. peakspectra
ratios
were normalized
were 0.36 for A-1800 andpeak
to the D-900,at and
15090.32
cmfor
−1 asD-500. These findings
the intensity of the are in good agreement
phenylene stretching
with datadid
vibration in the
notliterature [22,23]. crosslinking. Peaks relevant to the curing agent DICY
change during
were found at 1638 cm and 1100epoxy
The formulated waterborne
−1 varnishes
cm−1 due were applied
to resonant states of tothe
electrical
bending steel, dried and
vibration of
pre-cured to coatings in B-stage in a convection oven at a peak
NH2. The oligomer formation of DGEBA and DICY was proven by asymmetric stretching metal temperature of 210 ◦ C.
The coating
peaks of N-C≡N thickness
at 2187in cmB-stage
−1, which was 5 µm for all epoxy
is characteristic for linearthermosets.
oligomers.Figure 5 showsthe
Furthermore, an
exemplary
peak at 1560 comparison of the epoxy
cm−1 (ν (N-C≡N)) wasresin D-500 with
attributed to the theguanidine
coating in unit
B-stage. The spectra
resulting from
were normalized
oligomers [23]. Thetopeaks
the peakmarkedat 1509 cm−1 as
in Figure thecharacteristic
5 are intensity of the for phenylene stretching
DICY and oligomers
with DGEBA. Hence, partial crosslinking of the investigated coatings was confirmed.DICY
vibration did not change during crosslinking. Peaks relevant to the curing agent
were found at 1638 cm−1 and 1100 cm−1 due to resonant states of the bending vibration of
NH2 . The oligomer formation of DGEBA and DICY was proven by asymmetric stretching
peaks of N-C≡N at 2187 cm−1 , which is characteristic for linear oligomers. Furthermore,
the peak at 1560 cm−1 (ν (N-C≡N)) was attributed to the guanidine unit resulting from
oligomers [23]. The peaks marked in Figure 5 are characteristic for DICY and oligomers
with DGEBA. Hence, partial crosslinking of the investigated coatings was confirmed.
The topography of the applied coatings in B-stage is depicted in Figure 6. Full surface
wetting was ascertained optically and confirmed by FTIR-spectroscopy for all investigated
varnish formulations. However, most of the coatings exhibited thickness variations and a
wavelike topography. The wavelength of the pattern was finer for coatings based on epoxy
adhesives of higher EEW. Most likely, the waviness is dependent on the viscosity of the
varnish. This effect is described in the literature as orange peel. A main reason for orange
peel effects is non-optimized solvent content of the varnish system [24].
Storage modulus and loss factor thermograms of the investigated laminates are de-
picted in Figure 7. While storage modulus of the laminate was about 70 GPa in the glassy
state below glass transition temperature of the adhesive, it dropped by one magnitude to
7.3 GPa above TG in the rubbery state of the adhesive. The TG value was deduced from
Figure 5. FTIR-ATR
the maximum spectra
of the of the epoxy
loss factor. resin D-500
The highest valueandofthe ◦ C was
102coating in obtained
B-stage. for the laminate
based on varnish D-500. In contrast, lower TG values of 94 C and 81 ◦ C were detected for

The topography
laminates with coatings of based
the applied coatings
on higher molar inmass
B-stage is depicted
epoxies D-900 andin Figure
A-1800,6. Full surface
respectively.
wetting
Hence, inwas ascertained
agreement optically
with findings andliterature
in the confirmed by FTIR-spectroscopy
[16], glass transition temperature for was
all
investigated varnish formulations.
indirectly proportional to the EEW ofHowever, most of theThis
the epoxy thermoset. coatings exhibited
relationship thickness
was attributed
variations and a wavelike
to a lower crosslinking density topography.
of high EEW The wavelength
adhesives. of the pattern
The theoretical was finer
crosslinking for
density
coatings based on epoxy adhesives of higher EEW. Most likely, the waviness is dependent
Figure 4. FTIR-ATR spectra in the range from 3700 to 2700 cm−1 (left) and 1100 to 850 cm−1 (right) of
epoxies with average EEW values of 500, 900, and 1800 g/mol.
Polymers 2022, 14, 1556 7 of 13
The formulated waterborne epoxy varnishes were applied to electrical steel, dried and
pre-cured to coatings in B-stage in a convection oven at a peak metal temperature of 210
°C.
forThe coating
adhesive thickness
A-1800 was in B-stage
a factor ofwas 5 µ mthan
3 lower for all
thatepoxy thermosets.
of D-500. Figurevalues
The absolute 5 shows an
of the
exemplary comparison
storage modulus in theofglassy
the epoxy resin D-500
and rubbery withdid
plateaus thenot
coating
reflectinthe
B-stage. The spectra
pronounced effect
were normalized
of EEW. to the
The storage peak at
modulus was1509 cm −1affected
mainly as the byintensity of the steel
the electrical phenylene stretching
substrate. Due to
vibration
an adhesivedidlayer
not change
thickness during crosslinking.
well below Peaksfluctuations
10 µm, small relevant tointhe curing agent
thickness DICY
had a strong
were
impactfound at 1638 cm
on scattering and 1100values,
of−1modulus cm−1 due to resonant
especially in thestates of state.
rubbery the bending vibration
For laminates of
based
NH . The oligomeravailable
on 2commercially formationvarnishes,
of DGEBA and DICY
higher glasswas proventemperatures,
transition by asymmetric instretching
the range
from of
peaks 110N-C≡N ◦ C,
to 117 at were
2187 cmreported
−1, whichinis [25], slightly dependent
characteristic on the electrical
for linear oligomers. steel grade.
Furthermore, the
As described
peak at 1560 bycmAufray
−1 and Rochewas
(ν (N-C≡N)) [26],attributed
chemical interactions of amine-based
to the guanidine curing agents
unit resulting from
and metal[23].
oligomers oxide passive
The peakslayers
marked could also have
in Figure an characteristic
5 are impact on glass fortransition
DICY and temperature
oligomers
values
with and the
DGEBA. macromolecular
Hence, structureof
partial crosslinking ofthe
theinvestigated
thermoset atcoatings
the interphase.
was confirmed.

Polymers 2022, 14, x FOR PEER REVIEW 7 of 13

on the viscosity of the varnish. This effect is described in the literature as orange peel. A
main reason for orange peel effects is non-optimized solvent content of the varnish system
Figure
Figure 5. FTIR-ATR spectra
spectra of
of the
the epoxy resin D-500 and the coating in B-stage.
[24]. 5. FTIR-ATR
The topography of the applied coatings in B-stage is depicted in Figure 6. Full surface
wetting was ascertained optically and confirmed by FTIR-spectroscopy for all
investigated varnish formulations. However, most of the coatings exhibited thickness
variations and a wavelike topography. The wavelength of the pattern was finer for
coatings based on epoxy adhesives of higher EEW. Most likely, the waviness is dependent

Figure6.6.Topographic
Figure Topographicimages
imagesof
ofB-stage
B-stagecoated
coatedelectrical
electricalsteel
steelsheets
sheetsbased
basedon
onthe
theepoxy
epoxyadhesives
adhesives
A-1800, D-900 and D-500.
A-1800, D-900 and D-500.
The width
Storage of the loss
modulus andfactor
losspeak at 50%
factor of the maximum
thermograms of the value was highest
investigated for lami-
laminates are
nates based on D-500, and lowest for D-900. In agreement with the findings
depicted in Figure 7. While storage modulus of the laminate was about 70 GPa in theof Misra [16],
the crosslinking
glassy network
state below glass of coatings temperature
transition made from adhesives of lowerit EEW
of the adhesive, values
dropped bywas,
one
presumably, less perfectly associated with wider loss factor peaks.
magnitude to 7.3 GPa above TG in the rubbery state of the adhesive. The TG value was
deduced from the maximum of the loss factor. The highest value of 102 °C was obtained
for the laminate based on varnish D-500. In contrast, lower TG values of 94 °C and 81 °C
were detected for laminates with coatings based on higher molar mass epoxies D-900 and
A-1800, respectively. Hence, in agreement with findings in the literature [16], glass
transition temperature was indirectly proportional to the EEW of the epoxy thermoset.
This relationship was attributed to a lower crosslinking density of high EEW adhesives.
The theoretical crosslinking density for adhesive A-1800 was a factor of 3 lower than that
Polymers 2022,
Polymers 14, 1556
2022, 14,
14, xx FOR
FOR PEER
PEER REVIEW
REVIEW 888of
of
of 13
13

Figure 7. Storage
Figure 7.
Figure 7. Storagemodulus
Storage modulusand
modulus andloss
and loss
loss factor
factor
factor curves
curves
curves of electrical
of electrical
of electrical steel
steelsteel laminates
laminates adhesively
adhesively
laminates bonded
bonded
adhesively with
bonded
with
with epoxy
epoxyepoxy varnishes
varnishes with
with with
varnishes EEW
EEW EEW
valuesvalues of
of 500,
values 500,
of900 900
500,and and
9001800 1800
and g/mol. g/mol.
1800 g/mol.

Adherence strength
Adherence strength was was determined
determined
determined by by roll peel
by roll peel testing
testing atat 23,
23, 100
100 and 140 ◦°C.
and 140 °C.
C. The
The
temperature-dependent roll
roll peel
peel strength
strength values
values of of
the the investigated
investigated
temperature-dependent roll peel strength values of the investigated laminates with laminates
laminates with with
adhe-
sives of differing
adhesives
adhesives of EEW EEW
of differing
differing valuesvalues
EEW are plotted
values are in Figure
are plotted
plotted 8. At ambient
in Figure
in Figure 8. At
8. temperature,
At ambient
ambient the lowest
temperature,
temperature, the
the
roll peel
lowest roll
lowest strength,
roll peel of
peel strength,7.5
strength, ofN/mm,
of 7.5 was
7.5 N/mm, determined
N/mm, waswas determined for
determined forD-500. Adhesives
for D-500.
D-500. Adhesives of higher
Adhesives of EEW
of higher
higher
(A-1800
EEW and
(A-1800 D-900)
and exhibited
D-900) increased
exhibited roll
increasedpeel strength,
roll peel higher
strength,
EEW (A-1800 and D-900) exhibited increased roll peel strength, higher than 8 N/mm. than 8
higher N/mm.
than 8 Presum-
N/mm.
ably, a lower crosslinking
Presumably,
Presumably, aa lower density isdensity
lower crosslinking
crosslinking beneficial
density for roll peel
is beneficial
is beneficial forstrength
for peelatstrength
roll peel
roll ambient at
strength conditions.
at ambient
ambient
Slightly lower
conditions. crosslinking
Slightly lower density
crosslinkingwould allow
density for
would less
allow brittle
for fracture
less
conditions. Slightly lower crosslinking density would allow for less brittle fracture and, brittleand, therefore,
fracture and,
higher rollhigher
therefore, peel strength.
roll peel Astrength.
similar trend
A was trend
similar proven wasbyproven
Pugstallerby et al. [25] for
Pugstaller et laminates
al. [25] for
therefore, higher roll peel strength. A similar trend was proven by Pugstaller et al. [25] for
made from
laminates madea commercially
made from available
from aa commercially varnish.
commercially available The roll
available varnish. peel
varnish. The strength
The roll
roll peelvalues
peel strengthwere in
strength values a similar
values were
were
laminates
range
in a of about
similar range8 N/mm.
of about 8 N/mm.
in a similar range of about 8 N/mm.

Figure 8.
Figure 8. Roll
Roll peel
Roll peel strength
peel strength at
strength at 23,
at 23, 100
23, 100 and
100 and 140
and 140 °C of
◦ Cof
140 °C laminates
oflaminates made
laminatesmade from
madefrom epoxy
from epoxy varnishes
epoxy varnishes with
varnishes with EEW
with EEW
EEW
values
values of
of 500,
500, 900
900 and
and 1800
1800 g/mol.
g/mol.
values of 500, 900 and 1800 g/mol.

At aaatest
At testtemperature
test temperatureof of
temperature of 100
◦ C, °C,
100100 which
°C, which was
was was
which withinwithin
the glass
within the transition
the glass transition
glass transition of the
the
of the investi-
of
investigated
investigated adhesives,
gated adhesives, a
a pronounced
adhesives, pronounced
decrease
a pronounced decrease in
in roll in
decrease roll
peel peel
rollstrength strength was ascertained.
was ascertained.
peel strength This
This effect
was ascertained. This
effect
could could
be be attributed
attributed to an to an
enhancedenhanced
segmentalsegmental
mobility mobility
of the of the
polymericpolymeric
effect could be attributed to an enhanced segmental mobility of the polymeric adhesive at adhesiveadhesive
at at
increas-
increasing
ing temperatures
temperatures [27]. The [27].
dropThewasdrop
less was less
pronounced pronounced
for for
laminates
increasing temperatures [27]. The drop was less pronounced for laminates based on D- laminates
based onbased
D-500, on D-
with
500, with
higher higher
storage storage modulus
modulus modulus
values at values
100 ◦ C at
values due100 °C
to°C due toto aahigher
a slightly slightly higher
glass glass transition
transition transition
tempera-
500, with higher storage at 100 due slightly higher glass
temperature. ◦ At 140 °C, well above T , a further drop of roll peel strength, to about
ture. At 140 C,
temperature. Atwell
140 above TG ,above
°C, well a further
TG, drop
G of rolldrop
a further peelof strength,
roll peelto strength,
about 3 N/mm,
to about was33
Polymers 2022, 14, 1556 9 of 13

discernible. The laminate based on D-500, with the highest TG , showed inferior adherence
at 140 ◦ C and could not be tested. Better adherence performance at 140 ◦ C was obtained
for thermosets with higher OH functionality (i.e., D-900 and A-1800). By comparing TG or
OH functionality with roll peel strength at 100 ◦ C or 140 ◦ C, a much higher impact of OH
functionality was deduced. Hence, OH functionality is presumably much more important
than crosslinking density. Due to hydroxyl groups, different bonds such as secondary or
covalent bonds to the metal oxide surface are established [8,10,28,29].
As confirmed by IR spectroscopy in reflection mode using an Ulbricht sphere, the
fractured surfaces of the investigated laminates, roll peel tested at 23, 100 and 140 ◦ C,
revealed, predominately, interfacial failure. This supports the hypothesis of high relevance
of OH functionality for improvement of interfacial debonding performance.
In order to test adherence performance under more service-relevant conditions, fatigue
fracture mechanics tests were performed. In roll peel testing, most of the peeling energy is
attributable to plastic deformation of the substrate [30]. This effect could be excluded by
fracture mechanics delamination testing using a double cantilever beam specimen (DCB).
In Figure 9, crack growth kinetics curves of the laminates tested at ambient temperature
are depicted. The curves can be divided into three different regions: the threshold, stable
crack growth and unstable crack growth regime. In the threshold regime (ultra-low crack
growth rates), the adhesive based on A-1800 showed lower strain energy release rate
threshold values (Gth ) than laminates prepared from D-900 or D-500 epoxies. Presumably,
slow crack propagation in the threshold regime of laminates based on epoxies of low and
intermediate EEW values was less pronounced due to better adherence to the passive layer
of the substrate. Additionally, at low strain energy release rate values, the plastic zone size
at the crack tip is less- or non-constrained by the substrate. If the plastic zone is constrained,
the crack tends to propagate within the adhesive layer, associated with failure in cohesive
Polymers 2022, 14, x FOR PEER REVIEW
manner [13]. Therefore, at slow crack propagation rates, better cohesive strength10ofofthe 13
epoxy layer would result in enhanced crack growth resistance.

Figure 9.
Figure 9. Crack
Crack kinetic
kinetic curves
curves at
at ambient
ambient temperature
temperature of
of laminates
laminates made
made from
from epoxy
epoxy varnishes with
varnishes with
EEW values of 500, 900 and 1800 g/mol compared to values from the literature [13], marked in
EEW values of 500, 900 and 1800 g/mol compared to values from the literature [13], marked in grey. grey.

Figure
In the 10,crack
stable structure–property relationships,
growth regime (II), of molar
the lowest crack mass or rate
propagation epoxy equivalent
was obtained
weight
for and the
laminates basedinvestigated properties,toare
on D-900. Compared depicted
fatigue crackschematically.
growth data forAslaminates
confirmedbasedby
DMA,
on raising EEW
commercially causes
available a decrease
varnish systems in[13]
crosslinking density
(see grey area (CLD).
in Figure 9), aInsimilar
contrast, the
fatigue
hydroxyl
crack growthfunctionality
resistance affecting the adherence
was confirmed in regimetoII.the
Forsubstrate
monotonic raises withmechanic
fracture higher molar
tests
massaofhigh
with the crack
epoxytipresin. Both
radius, roll peel
a trend strength (RPS)
of maximum and
fracture fatigue crack
toughness growth kinetics
at intermediate EEW
values
(FCG) was also reported
at ambient by Levita
temperature et al.
peak [31]. Furthermore,
at intermediate EEW for amorphous
values polymeric
of 900 g/mol. mate-
At low or
rials,
high EEWa higher crosslinking
values, roll peeldensity
strengthwasandrelated to increased
fatigue yieldkinetics
crack growth stress and
arereduced strain
lower, due to
at break
poor [32]. These
adherence twoEEW)
(at low contrary effects
or less necessitate
crosslinking an optimization
density and aHowever,
(at high EEW). compromise the
effect of hydroxyl groups on the epoxy-metal adherence is still under discussion [33–36].
The postulated schematic relationships do not account for chemical interactions of amine-
based hardeners and metal substrates [26], which have been investigated and ascertained
in a recent study [37]. Future research will focus on the potential effects of interfacial
Polymers 2022, 14, 1556 10 of 13

between crosslinking density and ductility. A better stress distribution at the crack tip
could be achieved by higher inner mobility (e.g., at higher temperatures). Nevertheless,
a certain level of yield stress has to be exceeded. In addition to bulk adhesive properties,
the plastic zone at the crack tip is constrained at higher strain energy release rates by the
substrate. Subsequently, the crack tends to propagate in the substrate-adhesive interface.
Under such conditions, substrate-adhesive interactions (eg., OH bonds or covalent bonds)
are of utmost relevance to ensure sufficient fatigue crack growth resistance [13].
The behavior of high EEW laminates (i.e., A-1800) was slightly better in the unstable
crack growth regime. Interestingly, the ranking of roll peel strength values and crack
growth9.resistance
Figure Crack kineticwas equivalent
curves in the
at ambient unstableof
temperature crack growth
laminates maderegime. However,
from epoxy contrary
varnishes with
to rollvalues
EEW peel testing, theand
of 500, 900 DCB specimen
1800 failed, mainly
g/mol compared to valuesin afrom
cohesive manner.
the literature This
[13], difference
marked is
in grey.
most likely related to excessive plastic deformation of the substrate during roll peel testing.
In Figure 10, structure–property relationships, relationships, of of molar mass or epoxy equivalent
weight andandthe theinvestigated
investigated properties,
properties, areare
depicted
depictedschematically.
schematically.As confirmed
As confirmedby DMA, by
raising raising
DMA, EEW causesEEW acausesdecrease in crosslinking
a decrease density (CLD).
in crosslinking densityIn(CLD).
contrast,
In the hydroxyl
contrast, the
functionality
hydroxyl affecting affecting
functionality the adherence to the substrate
the adherence raises with
to the substrate higher
raises withmolar
highermass
molar of
the epoxy
mass of theresin.
epoxyBoth resin.roll peelroll
Both strength (RPS) and
peel strength (RPS)fatigue crack growth
and fatigue kinetics
crack growth (FCG)
kinetics
at ambient
(FCG) temperature
at ambient peak atpeak
temperature intermediate EEW values
at intermediate EEW valuesof 900 ofg/mol. At low
900 g/mol. Atorlowhigh
or
EEW values, roll peel strength and fatigue crack growth kinetics
high EEW values, roll peel strength and fatigue crack growth kinetics are lower, due to are lower, due to poor
adherence
poor (at low
adherence (atEEW)
low EEW)or lessorcrosslinking density
less crosslinking (at high
density (at EEW). However,
high EEW). the effect
However, the
of hydroxyl groups on the epoxy-metal adherence is still under
effect of hydroxyl groups on the epoxy-metal adherence is still under discussion [33–36].discussion [33–36]. The
postulated
The schematic
postulated relationships
schematic do not
relationships doaccount for chemical
not account interactions
for chemical of amine-based
interactions of amine-
hardeners
based and metal
hardeners substrates
and metal [26], which
substrates have have
[26], which been been
investigated
investigatedand ascertained
and ascertained in a
recent study [37]. Future research will focus on the potential effects
in a recent study [37]. Future research will focus on the potential effects of interfacial of interfacial interac-
tions and interphase
interactions formation
and interphase in the coating
formation in the and lamination
coating process onprocess
and lamination the adherence
on the
performance of steel/epoxy laminates.
adherence performance of steel/epoxy laminates.

Figure
Figure 10.
10. Schematic
Schematic relationship
relationship of
of OH
OH functionality,
functionality, crosslinking
crosslinking density
density (CLD),
(CLD), roll
roll peel
peel strength
strength
(RPS)
(RPS) and
and fatigue
fatigue crack
crack growth
growth kinetics
kinetics (FCG)
(FCG) at ambient temperature
at ambient temperature and
and the
the epoxy
epoxy equivalent
equivalent
weight (EEW) of the adhesive in electrical steel laminates.
weight (EEW) of the adhesive in electrical steel laminates.

4. Conclusions
Waterborne varnishes for electrical steel laminates based on bisphenol-A-diglycidiyl
ether (DGEBA) resins with varying epoxy equivalent weights (EEW) were prepared and
investigated. The EEW values of the epoxy resins ranged from 500 to 1800 g/mol. 10 m%
of poloxamer emulsifier was added to the formulation. As curing agent, dicyandiamide
(DICY) was used in an over-stoichiometric amount. Electrical steel sheets were coated with
the investigated varnish systems. After pre-curing, the coated steel sheets were stacked and
Polymers 2022, 14, 1556 11 of 13

hot press-cured. The investigated laminates consisted of 4 or 6 electrical steel plies (roll peel
or fatigue crack growth tests) and epoxy adhesive layers with a thickness of about 5 µm.
By infrared spectroscopy, a higher amount of oxirane ring groups, but a lower OH
functionality, was confirmed for lower molar mass epoxy. The orange peel effect of partly
cured coatings on electrical steel substrates was also dependent on the epoxy equivalent
weight (EEW) of the adhesives. For higher EEW values of 1800 g/mol, a much finer
orange peel structure was discernible, most likely attributable to a higher viscosity in the
coating process.
The thermomechanical properties of laminates with fully cured adhesive layers was
examined by DMA. The lowest glass transition temperature was obtained at 81 ◦ C for
the highest EEW adhesive, with 1800 g/mol. With decreasing EEW value, the glass
transition values rose up to 102 ◦ C for the adhesive based on an EEW of 500 g/mol.
Hence, crosslinking density was correlated in a negative manner with the EEW value of
the adhesive.
The adhesive strength of the laminates was assessed by roll peel testing at 23, 100 and
140 ◦ C. Laminates based on adhesives of higher EEW values exhibited better adherence
performance below, and above, glass transition. A maximum strength of about 9 N/mm at
23 ◦ C was obtained, which was in good agreement with data from the literature [3]. At 100
and 140 ◦ C, roll peel strength dropped by 50 and 75%, respectively. This drop was attributed
to enhanced main chain mobility within or above the glass transition regime of the epoxy
adhesives. Especially at 140 ◦ C, the laminates based on the adhesive of lowest EEW failed
prior to testing. In contrast, laminates with coatings of intermediate or high EEW value
revealed roll peel strength values of 3 N/mm. Hence, it was clearly deduced that higher
EEW, associated with more OH functional groups, is of utmost importance to mechanical
strength at temperatures around glass transition or higher. Most likely, the OH functionality
and associated interfacial bonds are of higher relevance than crosslinking density. Due to
mainly interfacial failure of the investigated laminates at testing temperatures below and
above glass transition, there is significant potential for further improvement of the varnish
formulations. Special attention should be given to adhesion promoters or adequate surface
treatment procedures.
Fatigue crack growth resistance was characterized at ambient conditions. The best
performance in the threshold and stable crack growth regime was deduced for laminates
based on the adhesive of intermediate EEW of 900 g/mol. At slow crack propagation rates,
the plastic zone size was less constrained by the substrate. Therefore, higher cohesive
strength of the epoxy layer resulted in increased crack growth resistance. In the stable crack
growth regime, the higher crosslinking density was related to enhanced yield stress and
reduced strain at break. Overall, it was deduced that a well-balanced ratio of crosslinking
density and OH functionality is of high relevance to ensure good performance under
near-service conditions.
Finally, structure–property correlations were established and deduced for waterborne
epoxy varnish systems with systematically varied EEW values and electrical steel laminates.
It was clearly shown that enhancement of glass transition temperature and associated
crosslinking density did not necessarily lead to better mechanical performance of the
laminates. Especially at elevated temperatures and under fatigue loading conditions, the
adherence was dependent on both cohesive strength of the epoxy layer and interfacial
bonding to the metal oxide passivation layer of the substrate. Hence, a well-balanced
compromise of inner mobility and ductility of the epoxy layer and interfacial bonding to the
substrate is essential for the optimization of monotonic and fatigue delamination resistance.
The model varnish-based approach allowed for a clear separation of opposing or reinforcing
effects. The investigations confirmed that the energy-based fatigue testing approach was
much more sensitive for elucidation of material structure effects than conventional force-
based, monotonic testing.
To also assess the effect of ageing on the adherence of electrical steel laminates, future
research will focus on fatigue testing under superimposed mechanical stresses and envi-
Polymers 2022, 14, 1556 12 of 13

ronmental influences. Furthermore, potential interactions of the micrometer-thin epoxy


coating with the passive layer of electrical steel substrates, and their consequences on the
thermoset network structure, will be investigated.

Author Contributions: Conceptualization, C.M. and R.P.; methodology, C.M.; validation, G.M.W.,
M.A. and O.B.; investigation, C.M.; writing—original draft preparation, C.M. and G.M.W.; writing—
review and editing, O.B. and M.A.; supervision, G.M.W.; project administration, G.M.W. All authors
have read and agreed to the published version of the manuscript.
Funding: The financial support by the Austrian Federal Ministry for Digital and Economic Affairs,
the National Foundation for Research, Technology and Development and the Christian Doppler
Research Association are gratefully acknowledged.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Acknowledgments: Open Access Funding by the University of Linz.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Ellis, B. (Ed.) Introduction to the Chemistry, Synthesis, Manufacture and Characterization of Epoxy Resins. In Chemistry and
Technology of Epoxy Resins; Springer Netherlands: Dordrecht, The Netherlands, 1993; pp. 1–36. ISBN 978-94-011-2932-9.
2. Lamprecht, E. Der Einfluss der Fertigungsverfahren auf die Wirbelstromverluste von Stator-Einzelzahnblechpaketen für den Einsatz in
Hybrid- und Elektrofahrzeugen; Meisenbach: Bamberg, Germany, 2014; ISBN 978-3-87525-362-7.
3. Pugstaller, R.; Wallner, G. Development of a Fracture-Mechanics Based Fatigue Testing Method for Epoxy/Electrical Steel
Laminates with Thin Adhesive Layer. Eng. Fract. Mech. 2021, 258, 108045. [CrossRef]
4. Wimmer, M.; Dopler, A. Process for the Procution of Electrical Steel Sheet Cores. U.S. Patent US20050189067A1, 1 September 2004.
5. Motawie, A.M.; Sherif, M.H.; Badr, M.; Amer, A.A.; Shehat, A.S. Synthesis and Characterization of Waterborne Epoxy Resins for
Coating Application. Aust. J. Basic Appl. Sci. 2010, 4, 1376–1382.
6. Pollock, M. Beyond High-Solids Baking Enamels. Met. Finish. 1997, 95, 10–15. [CrossRef]
7. Yin, H.; Wan, Y.; Zhou, J.; Sun, D.; Li, B.; Ran, Q. Self-Emulsified Waterborne Epoxy Hardener without Acid Neutralizers and Its
Emulsifying and Curing Properties. Pigment Resin Technol. 2019, 48, 223–228. [CrossRef]
8. Bruyne, N. The Adhesive Properties of Epoxy Resins. J. Appl. Chem. 2007, 6, 303–310. [CrossRef]
9. Pascault, J.-P.; Williams, R.J.J. (Eds.) General Concepts about Epoxy Polymers. In Epoxy Polymers; Wiley-VCH: Weinheim,
Germany, 2010; pp. 1–12. ISBN 978-3-527-62870-4.
10. Wei, H.; Xia, J.; Zhou, W.; Zhou, L.; Hussain, G.; Li, Q.; Ostrikov, K. Adhesion and Cohesion of Epoxy-Based Industrial Composite
Coatings. Compos. Part B Eng. 2020, 193, 108035. [CrossRef]
11. Packham, D.E. Theories of Fundamental Adhesion. In Handbook of Adhesion Technology; da Silva, L.F.M., Öchsner, A., Adams, R.D.,
Eds.; Springer Berlin Heidelberg: Berlin/Heidelberg, Germany, 2011; pp. 9–38. ISBN 978-3-642-01169-6.
12. Remisol EB 548 Technisches Merkblatt, Rembrandtin Coatings GmbH: Wien, AUT, 1 January 2008. Available online: http:
//www.polarislaserlaminations.com/548Specs.pdf (accessed on 15 February 2022).
13. Pugstaller, R.; Wallner, G.M.; Strauß, B.; Fluch, R. Temperature Dependent Fatigue Crack Growth in Electrical Steel Laminates
with µm-Thick Epoxy Layers. Eng. Fract. Mech. 2022, 108416. [CrossRef]
14. Wößner, W.; Heim, M.; Walter, M.; Fleischer, J. Challenges and Potential Solutions for Reduced Unbalance of Permanent Magnet
Rotors for Electric Traction Motors. In Proceedings of the 2021 11th International Electric Drives Production Conference (EDPC),
Erlangen, Germany, 7 December 2021; pp. 1–9. [CrossRef]
15. Chen, J.-S.; Ober, C.; Poliks, M.; Zhang, Y.; Wiesner, U.; Cohen, C. Controlled Degradation of Epoxy Networks: Analysis of
Crosslink Density and Glass Transition Temperature Changes in Thermally Reworkable Thermosets. Polymer 2004, 45, 1939–1950.
[CrossRef]
16. Misra, S.C.; Manson, J.A.; Sperling, L.H. Effect of Cross-Link Density Distribution on the Engineering Behavior of Epoxies.
In Epoxy Resin Chemistry; Bauer, R., Ed.; American Chemical Society: Washington, DC, USA, 1979; Volume 114, pp. 137–156.
ISBN 978-0-8412-0525-3.
17. Niederwimmer, K.; Kieberger, B.; Hassel, A.W. Analysis of Main Failure Pattern Using Discrete Fourier Transform on Metal-
Epoxy-Based Adherent-Laminates: Analysis of Main Failure Pattern. Phys. Status Solidi A 2015, 212, 1242–1248. [CrossRef]
18. Pruksawan, S.; Lambard, G.; Samitsu, S.; Sodeyama, K.; Naito, M. Prediction and Optimization of Epoxy Adhesive Strength from
a Small Dataset through Active Learning. Sci. Technol. Adv. Mater. 2019, 20, 1010–1021. [CrossRef]
Polymers 2022, 14, 1556 13 of 13

19. Hayaty, M.; Honarkar, H.; Beheshty, M.H. Curing Behavior of Dicyandiamide/Epoxy Resin System Using Different Accelerators.
Iran. Polym. J. 2013, 22, 591–598. [CrossRef]
20. Pascault, J.-P.; Sautereau, H.; Verdu, J.; Williams, R.J. Thermosetting Polymers, 1st ed.; CRC Press: Boca Raton, FL, USA, 2002;
ISBN 978-0-8247-0670-8.
21. Saba, N.; Jawaid, M.; Alothman, O.Y.; Paridah, M.; Hassan, A. Recent Advances in Epoxy Resin, Natural Fiber-Reinforced Epoxy
Composites and Their Applications. J. Reinf. Plast. Compos. 2016, 35, 447–470. [CrossRef]
22. Tudorachi, N.; Mustata, F. Curing and Thermal Degradation of Diglycidyl Ether of Bisphenol A Epoxy Resin Crosslinked with
Natural Hydroxy Acids as Environmentally Friendly Hardeners. Arab. J. Chem. 2020, 13, 671–682. [CrossRef]
23. Gaukler, J.C. Oligomer Formation in Epoxy-Dicyandiamide Systems. J. Adhes. 2012, 88, 720–750. [CrossRef]
24. Liu, Y.; Jehanathan, N.; Yang, H.; Abdullah, J. SEM Observation of the “Orange Peel Effect” of Materials. Mater. Lett. 2007, 61,
1433–1435. [CrossRef]
25. Pugstaller, R.; Wallner, G.M.; Strauß, B.; Fluch, R. Advanced Characterization of Laminated Electrical Steel Structures under Shear
Loading. J. Adhes. 2019, 95, 834–848. [CrossRef]
26. Aufray, M.; Roche, A. Properties of the Interphase Epoxy–Amine/Metal: Influences from the Nature of the Amine and the
Metal. In Adhesion: Current Research and Applications; Possart, W., Ed.; Wiley-VCH: Weinheim, Germany, 2006; pp. 89–102.
ISBN 978-3-527-60730-3.
27. Gent, A.N.; Schultz, J. Effect of Wetting Liquids on the Strength of Adhesion of Viscoelastic Material. J. Adhes. 1972, 3, 281–294.
[CrossRef]
28. Rider, A.N.; Brack, N.; Andres, S.; Pigram, P.J. The Influence of Hydroxyl Group Concentration on Epoxy–Aluminium Bond
Durability. J. Adhes. Sci. Technol. 2004, 18, 1123–1152. [CrossRef]
29. Semoto, T.; Tsuji, Y.; Yoshizawa, K. Molecular Understanding of the Adhesive Force between a Metal Oxide Surface and an Epoxy
Resin. J. Phys. Chem. C 2011, 115, 11701–11708. [CrossRef]
30. Kinloch, A.; Koay, H.; Lee, S.H.; Ng, L. Using the Simple Peel Test to Measure the Adhesive Fracture Energy. In Proceedings of
the 35th Annual Meeting of the US Adhesion Society, New Orleans, FL, USA, 1 February 2012; ISBN 978-1-62276-131-9.
31. Levita, G.; De Petris, S.; Marchetti, A.; Lazzeri, A. Crosslink Density and Fracture Toughness of Epoxy Resins. J. Mater. Sci. 1991,
26, 2348–2352. [CrossRef]
32. Zhao, J.; Yu, P.; Dong, S. The Influence of Crosslink Density on the Failure Behavior in Amorphous Polymers by Molecular
Dynamics Simulations. Materials 2016, 9, 234. [CrossRef]
33. Fritah, Z.; Drouet, C.; Thouron, C.; Aufray, M. Direct Evidence of Amine-Metal Reaction in Epoxy Systems: An in Situ Calorimetry
Study of the Interphase Formation. Prog. Org. Coat. 2020, 148, 105769. [CrossRef]
34. Kotb, Y.; Cagnard, A.; Houston, K.R.; Khan, S.A.; Hsiao, L.C.; Velev, O.D. What Makes Epoxy-Phenolic Coatings on Metals
Ubiquitous: Surface Energetics and Molecular Adhesion Characteristics. J. Colloid Interface Sci. 2022, 608, 634–643. [CrossRef]
35. Nakazawa, M. Mechanism of Adhesion of Epoxy Resin to Steel Surface. Nippon Steel Technical Report, Nippon Steel, Tokio, JPN,
1994. Available online: https://www.nipponsteel.com/en/tech/report/nsc/pdf/6303.pdf (accessed on 15 February 2022).
36. Nakazawa, M.; Somorjai, G. Coadsorption of Water and Selected Aromatic Molecules to Model the Adhesion of Epoxy Resins on
Hydrated Surfaces of Zinc Oxide and Iron Oxide. Appl. Surf. Sci. 1995, 84, 309–323. [CrossRef]
37. Säckl, G.; Duchoslav, J.; Pugstaller, R.; Marchfelder, C.; Haselgrübler, K.; Aufray, M.; Wallner, G.M. The Interaction of Waterborne
Epoxy/Dicyandiamide Varnishes with Metal Oxides. Polymers, 2022; submitted.

You might also like