ARTICLES
Efficiency comparison of hyperbranched polymers as toughening
agents for a one-part epoxy resin
Thatiane Brocksa)
UNESP – Univ. Estadual Paulista – Guaratinguetá Faculty of Engineering, Guaratinguetá (SP) 12516-410, Brazil
Laura Ascione and Veronica Ambrogi
UNINA – University of Naples “Federico II” – Department of Chemical, Materials, and Production Engineering,
Napoli (NA) 80125, Italy
Maria O. H. Cioffi
UNESP – Univ. Estadual Paulista – Guaratinguetá Faculty of Engineering, Guaratinguetá (SP) 12516-410, Brazil
Paola Persico
CNR – National Research Council - Institute for Macromolecular Studies (ISMAC), Milano 20133, Italy
(Received 5 September 2014; accepted 30 January 2015)
A previously synthesized hyperbranched poly(butylene adipate) (HPBA) polymer was compared
with a commercial dendritic polyol (HPOH) as a toughening agent for a commercial one-part
epoxy resin. Both modifiers were added in weight percentages of 1, 3, 5, and 10%. The modified
epoxies were characterized using differential scanning calorimetry (DSC), dynamic mechanical
analysis (DMA), thermogravimetric analysis (TGA), melt rheological tests, and linear elastic
fracture mechanics. Blend morphology and matrix–modifier interactions were evaluated using
scanning electron microscopy (SEM) and Fourier transform infrared spectroscopy (FTIR) analysis,
respectively. The toughness-improvement effect was achieved without substantial impairment of
thermomechanical properties or degradation of the thermal stability of the epoxy resin.
A meaningful decrease in viscosity was achieved with both modifiers, contributing to an easier
infusion processability. No evidence of new chemical linking was found although phase separation
was observed by SEM, leading to the conclusion that only interfacial linkage occurs between
modifiers and epoxy chains. SEM analysis also clearly shows the fracture mode changing from
brittle to ductile by addition of modifiers, which was more evident for blends of HPBA.
I. INTRODUCTION
Polymers have been widely used as conventional
material substitutes due to their range of thermal and
mechanical properties and low cost. The crosslinked
epoxy resin system, an outstanding material, is considered to be the most attractive and commonly applied
one, as it offers a large range of formulations and
structural characteristics that allow different applications varying from simple objects to aerospace parts.1,2
Although epoxy resins exhibit some desired features
for engineering applications, a significant restriction is
caused by the high crosslink density that affects the
toughness characteristics. This latter property has a
strong influence on properties such as impact, fatigue,
and damage tolerance, among others.3,4
Efforts to improve toughness have involved the use of
rubbers, thermoplastics, and inorganic materials, such as
silica nanoparticles.3,5 Reactive liquid rubbers have been
Contributing Editor: Linda S. Schadler
a)
Address all correspondence to this author.
e-mail: thati_bd@yahoo.com.br
DOI: 10.1557/jmr.2015.41
J. Mater. Res., 2015
http://journals.cambridge.org
Downloaded: 11 Mar 2015
the most widely used materials because they form a
homogeneous solution during preparation, beyond a second
phase precipitation after curing.6 Unfortunately, the
precipitated rubber particles can reduce both the
Young’s modulus and glass transition temperatures
(Tg ) in addition to impairing the processability due to
increasing viscosity.7 This fact has encouraged the
search for new materials that can improve the epoxy
resin toughness without significant influence on thermal
and mechanical properties.
Studies have shown that an efficient toughening agent
promotes phase separation because two phases have a
better effect on absorbing and deflecting cracks through
an increase in plastic shear yielding and rubber particle
cavitation, which is an additional mechanism contributing
to toughness improvement due to increasing energy
absorption by matrix plastic deformation.3,8–11 The effect
is also dependent on interfacial compatibility and chain
interaction, which are influenced by particle size, distribution, and shape.3,12
Very recently, the systems used as epoxy toughening
agents to overcome the conventional limitations have
been dendrimers and hyperbranched polymers (HBPs).
Ó Materials Research Society 2015
1
IP address: 200.145.17.118
T. Brocks et al.: Efficiency comparison of hyperbranched polymers as toughening agents for a one-part epoxy resin
Their Newtonian behavior, low viscosity, and small
dimensions, despite their high molecular weight, make
them very attractive as toughening agents for epoxy
resins designed for liquid infusion processing.13,14
The compatibility and reactivity as well as the phaseseparation between epoxy and HBP are functions of
the bulk structure and the nature of the HBP chemical
end-groups. The bulk structure of HBP contributes to
resin toughness improvement through absorption of
more energy when subjected to impact testing, while
the end-groups contribute to compatibility and reactivity
with the surrounding matrix materials.13,15
In this work, the comparison of the toughness
improvement between a commercial dendrimer and a
synthesized HBP upon addition to a one-part epoxy RTM
resin was analyzed. The thermal, thermomechanical, and
viscosity behaviors were monitored to define a formulation
that can improve the toughness while minimally affecting
the processability and thermal properties.
II. EXPERIMENTAL
A. Materials
A commercial one-part liquid epoxy resin system
(Cycom 890 RTM—Cytec Industries), Tg 5 210 °C,
was used as the matrix. This resin system is a blend
of multifunctional epoxy resins, 4,49-methylenebis(N,
N-diglycidylaniline) (30–60 wt%), N,N-diglycidyl4-glycidyloxyaniline (10–30 wt%), and an aromatic
diamine (30–60 wt%) acting as a curing agent.
A commercial hyperbranched polyol (HPOH) (Perstorp Boltorn H311, Mn 5 5300 g/mol, Tg 5 5 °C)
and a synthesized linear-hyperbranched poly(butylene
adipate) copolymer (HPBA), Tg 5 50 °C, were used as
toughening agents. This latter product was obtained
through a branching reaction of the linear tailored
OCH3-terminated prepolymer 1,1,1-tris(hydroxymethyl)propane as a branching agent, according to
the procedure described in detail in a previous work. 16
The chemical structure of HPBA is illustrated in
Fig. 1.
The HPOH has a very high hydroxyl functionality and
a highly branched flexible backbone while the HPBA
presents complex architectures with hydroxyl-terminated
branches.
B. Epoxy blend preparation
The blends were prepared by mixing the HPBA and
HPOH with the epoxy resin. The mixture was heated at
90 °C for 1 h under stirring. The cure was conducted at
180 °C for 2 h under vacuum. The percentages of a
toughening agent added to epoxy resin and the sample
codes are summarized in Table I.
C. Characterization
1. FTIR spectroscopy
A Thermo Fischer Scientific Nicolet 6700 FTIR
(Waltham, MA) was used to collect the attenuated total
reflectance (ATR) spectra of the cured and uncured
epoxy resins, the toughening agents, and their blends
with commercial RTM epoxy resin. ATR was used at
a resolution of 4 cm1, and 16 scans were averaged for
each spectrum in a range between 4000–650 cm1.
2. Differential scanning calorimeter (DSC)
The blend curing process was studied in a TA
Instrument Q20 DSC (New Castle, DE) under a nitrogen atmosphere (flow rate of 50 mL/min), at 10 °C/min
in a temperature range from 30 to 310 °C, with the
samples sealed in aluminum crucibles.
3. Thermogravimetric analysis (TGA)
TGA was performed in a TA Q5000 analyzer (New
Castle, DE) to investigate the thermal stabilities of the
cured samples, tested from 40 up to 600 °C at 10 °C/min
in a nitrogen atmosphere.
4. Dynamic mechanical analysis (DMA)
Samples were analyzed using a Triton Technology
DMA model Tritec 2000 (Granthan, UK) according to
ASTM D7028. Analysis was performed in a singlecantilever mode with the amplitude of 10 lm, frequency
of 1 Hz, at a heating rate of 3 °C/min from 50 up to
250 °C. The storage and loss moduli and tand peak of
neat RTM epoxy and blends were determined.
5. Viscosity measurements
The viscosities of the blends and neat epoxy were
measured with a TA Instruments model AR-G2, using
TABLE I. Composition and codes of RTM epoxy resin blend samples.
Toughening agent contenta
Toughening HPOH (commercial)
agent
HPBA (synthesized)
1 phr
(%)
3 phr
(%)
5 phr
(%)
10 phr
(%)
C1B
C1H
C3B
C3H
C5B
C5H
C10B
C10H
a
Toughening agent percentages have been calculated according to the
equation: % (phr) 5 (toughening agent mass)/(epoxy mass) 100.
FIG. 1. Schematic chemical structure of HPBA.
2
J. Mater. Res., 2015
http://journals.cambridge.org
Downloaded: 11 Mar 2015
IP address: 200.145.17.118
T. Brocks et al.: Efficiency comparison of hyperbranched polymers as toughening agents for a one-part epoxy resin
a steel plane-spindle ETC 25 mm in an oscillatory shear at 1
rad/s with the gap of 1 mm, strain percentage of 70%, and
strain amplitude of 40%. Using these parameters, a linear
regime preserving an adequate strain value was obtained.
The viscosity was analyzed at 80 and 100 °C for 60 min.
6. Fracture toughness
Fracture-toughness tests were performed using the
single-edge-notch bending (SENB) method according to
ASTM D 5045-99 using an Instron 4505 universal tester.
Samples with 60 6 4 mm and a cross-head speed of
1 mm/min were used for testing. The critical stressintensity factor, KIC, was determined.
7. Scanning electron microscopy (SEM)
The fracture areas of the samples were observed with a
JEOL JSM-5310 scanning electron microscope (Tokyo,
Japan) after metallization to understand the fracture
mechanisms involved in failure.
III. RESULTS AND DISCUSSION
HPOH and HPBA have different chemical structures, and
their introduction into the epoxy resin is expected to have
a different influence on both the crosslinking mechanism
and the final performance of the modified epoxy resin.
A. Neat epoxy and blend curing mechanism
Among the experimental methods used to evaluate
the polymer-to-polymer interactions, infrared spectroscopy (IR) is one of the most effective because it is highly
sensitive to the presence of noncovalent bonds, such as
hydrogen bonds. The FTIR-ATR spectra of cured epoxy
resin and blends are shown in Figs. 2(a) and 2(b).
During the epoxy curing reaction, the nucleophilic
attack by the amine epoxy unsubstituted nitrogen atom
causes epoxy ring opening. The amine-to-epoxy addition
reaction was monitored through spectroscopic analysis by
observing the disappearance of the band at 909 cm1
related to the oxirane C–O stretching absorption bond
deformation and the simultaneous appearance of C–N
stretch at approximately 3300 cm1, indicating the successful curing process. Moreover, a second band located at
approximately 3050 cm1 was also evident, which might
be attributed to the C–H tension of the methylene group of
the epoxy ring, although this band could not contribute
significantly due to its low intensity and closeness to the
strong O–H absorptions.17
The multicomponent shape of the O–H band indicates
a complex system in which multiple H-bonding species
coexist. In particular, two spectral ranges are of particular
interest for analysis: the hydroxyl-stretching associated
with the broad band in the range 3300–3450 cm1 and
the carbonyl stretching region from 1660 to 1800 cm1.
Additionally, in Fig. 1, it is evident that the peak relative
to the C5O group stretching typical of the polyester
sample increases as the percentage of modifier increases.
The interaction between the epoxy resin and the
toughening agents is a factor of paramount importance
in the mechanical behavior of the modified epoxy
systems. When the epoxy resin crosslinking reaction
occurs in the presence of a polyester, the possible
reactions are not only those between the amine group
of the epoxy hardener resin and the hydroxyl groups
generated within the resin during crosslinking but also
those forming ether linkages between the hydroxyl
groups of the hyperbranched polyester and the epoxy
resin. In addition to chemical reactions, the formation of
intermolecular hydrogen bonding between the toughening
agent polar groups and the epoxy resin-OH is also possible,
FIG. 2. FTIR-ATR spectra of (a) RTM epoxy/HPBA and (b) RTM epoxy/HPOH cured blends.
J. Mater. Res., 2015
http://journals.cambridge.org
Downloaded: 11 Mar 2015
3
IP address: 200.145.17.118
T. Brocks et al.: Efficiency comparison of hyperbranched polymers as toughening agents for a one-part epoxy resin
allowing the formation of miscible blends through the
establishment of interassociated hydrogen bonds.
From Fig. 2, the different nature of mutual interactions
between the epoxy resin and both modifiers is evident.
The infrared absorbance band at 3300–3450 cm1 associated with a hydroxyl-stretching shift to lower frequencies (red shift) increases in its intensity upon the addition
of increasing amounts of HPOH [Fig. 2(b)]. Conversely,
no meaningful change is observed upon HPBA addition
[Fig. 2(a)]. This result is justified by the hydroxyl
functionalized systems of HPOH compared to HPBA,
which can affect both the crosslinking density and the
start of curing in the blends, as shown in Table II.
However, the polyester band intensity at 1730 cm 1 is
higher for HPBA and RTM epoxy/HPBA blends.
Observing the calorimetric data reported in Table II
and Fig. 3(b), the Tonset and Tmax observed for RTM
epoxy/HPOH blends decrease with HPOH addition,
indicating a quantity-dependent effect and facilitation of
the curing process.3,18 Figure 3(a) shows a well-defined
cure onset and endset, the same baseline, and a similar
shape of the curing curve. For the latter, no significant
variation is visible, showing that the HPBA causes no
TABLE II. Dynamic DSC analysis results for neat RTM epoxy and
blends.
Formulation
Tonset (°C)
Tmax (°C)
DHnorm. (J/g)
Neat epoxy
C1H
C3H
C5H
C10H
C1B
C3B
C5B
C10B
242.0
241.0
241.0
241.0
242.0
238.0
232.0
230.0
220.0
276.0
277.0
278.0
278.0
281.0
275.1
272.5
269.5
265.0
433.9
416.5
423.2
444.5
459.0
420.2
437.0
439.2
441.0
change in the curing process.19,20 The different chemical
structures are the main reason for curing behavior differences because the HPOH presents a large number of
hydroxyl functional groups with respect to HPBA and,
consequently, is more prone to catalyze/accelerate the
epoxy ring-opening reaction.
The modifier influence on ∆H is a function of modifier
amount (Table II). Until 3% of added modifiers, the
curing enthalpies decrease, gradually increasing to
amounts equal and up to 5%. When two polymers are
blended, the overall enthalpy is a balance among contributions related to the breaking of hydrogen bonds in
the self-associating polymer and hydrogen bonds forming
between the two polymers. In most cases, the mixing is
endothermic, but in thermoset/thermoplastic blends, the
curing enthalpies can also be explained based on phase
separation. When curing the blends, the phase separation
can hinder the curing reaction through a decrease in the
reaction probability because thermoset monomer contact
is delayed by the thermoplastic phases.21 Therefore, the
mixing phenomenon likely contributed to lowering the
curing enthalpies in the blends; however, for high
percentages (i.e., C5H, C10H, C5B, and C10B), the
increase in mixing enthalpies was lost due to the
difficulty of the epoxy monomer reaction caused by
increasing phase separation.
Although HPBA shows partial crystallinity, no melting
peak was observed during curing because crystallinity is
lost during the blend preparation, which is performed at
the higher HPBA melting point temperature. The crystallinity is not recovered even when the uncured blend is
cooled to room temperature, keeping the HPBA amorphous in the presence of epoxy resin as verified by a
further DSC heating scan performed on the HPBAcontaining systems. The HPOH, however, is already an
amorphous polymer.
FIG. 3. Dynamic DSC scans of uncured (a) RTM epoxy/HPBA and (b) RTM epoxy/HPOH blends.
4
J. Mater. Res., 2015
http://journals.cambridge.org
Downloaded: 11 Mar 2015
IP address: 200.145.17.118
T. Brocks et al.: Efficiency comparison of hyperbranched polymers as toughening agents for a one-part epoxy resin
B. Thermal and thermomechanical properties
The E9 glassy plateau observed in Figs. 4(a) and 4(b) is
slightly high for blends up to 5 phr, lowering for those
with 10 phr of modifiers in relation to neat epoxy,
irrespective of the modifiers used, thereby suggesting
a crosslinking density reduction due to the higher
free volume induced by this amount of added
modifiers.22,23
Considering the rubbery plateau, the E9rp values of
blends are lower than that of neat RTM epoxy, suggesting
a crosslinking density reduction caused by modifier
addition, except for C1H and C1B blends. These higher
observed E9rp values can be attributed to lower network
structural interference and also thermal stress relief promoted by a reduced amount of modifier.21,22,24
As for Tgs [Figs. 4(c) and 4(d)], decreasing values were
observed for increasing amounts of toughening agents,
being more intense and proportional for the blends
containing HPBA. Moreover, the Tgs values were found
to be intermediate between those of their precursors,
suggesting complete miscibility, verified through
Gordon–Taylor (GT), and mixture rule (MR) presented
in Fig. 6.23
The Gordon–Taylor (GT) equation [Eq. (1)] is often
used for the prediction of Tg in miscible blends while
a simple mixture rule (MR) [Eq. (2)] can be applied if
both precursors have the same contribution to the
blend T g. 23,25
w1 Tg1 þ kw2 Tg2
Tg ¼
;
ð1Þ
ðw1 þ kw2 Þ
Tg ¼ w1 Tg1 þ w2 Tg2
;
ð2Þ
where w is the proportional weight and k is an adjustable
parameter estimated through constituent properties via
Simba–Boyer rules [Eq. (3)]. Indices 1 and 2 refer to the
precursors.
FIG. 4. E9 curves of (a) HPBA and (b) HPOH and tand curves of (c) HPBA and (d) HPOH cured blends comparison with neat RTM epoxy.
J. Mater. Res., 2015
http://journals.cambridge.org
Downloaded: 11 Mar 2015
5
IP address: 200.145.17.118
T. Brocks et al.: Efficiency comparison of hyperbranched polymers as toughening agents for a one-part epoxy resin
q Tg
kffi 1 1
q2 Tg2
:
ð3Þ
The Tg prediction and experimental data plotted in
Fig. 5 suggest an equal blend precursor contribution to
the Tg of this binary system because the experimental
data are nearer to the M-R than to the GT curve
prediction, leading to the traditional interpretation of
good miscibility of different species. Although this
FIG. 5. Tg curve prediction and experimental data for CH and CB
blends.
behavior can be observed in both blends, suggesting a
single phase, the epoxy Tg drop is considerably smaller
than that predicted by the GT model showing a partial
miscibility, as such observed for carboxyl-terminated
butadiene-acrylonitrile (CTBN)-modified epoxies, which
are also two phases. SEM images of Fig. 6, obtained from
fracture section, clearly show dual-phase formation.
Fracture pattern differences also corroborate with the
existence of two phases.
As observed by Bussi and Ishida, when the modifier
acts as a plasticizer and flexibilizer, phase separation
virtually does not occur. The blend has just one Tg value,
and the experimental data suggest a miscible blend due to
the proximity with the predicted Tg curves using M-R and
GT. The second phase becomes discernible when the Tg
value departs from the prediction curve.11,24 In this case,
the 10 phr showed a greater decrease in E0, Tg and, for
RTM epoxy/HPBA blends, as well as E9.
Although the Tg values decreased to a lesser extent,
neither tested modifier prevents the application of these
blends as advanced composite matrix. The influence of a
modifier physical structure on Tg and E9 of blends can be
attributed to the modifier relaxation dynamics, explained by
accounting for epoxy-network structure distortions and also
the free volume induced in epoxy by these structures.21,26
Figure 7 shows similar degradation behavior among neat
RTM epoxy and the blends although both modifiers present
lower thermal stability than neat RTM epoxy. Considering
that degradation behavior is a function of the type of bonds
FIG. 6. HPBA and HPOH dispersed in RTM epoxy blends (a) C1B, (b) C1H, (c) C10B, and (d) C10H.
6
J. Mater. Res., 2015
http://journals.cambridge.org
Downloaded: 11 Mar 2015
IP address: 200.145.17.118
T. Brocks et al.: Efficiency comparison of hyperbranched polymers as toughening agents for a one-part epoxy resin
present in the network structure, the insignificant influence
of toughening agents on thermal stability indirectly supports
the hypothesis that only physical interactions between
polymers occur, thereby corroborating the other results
presented.24 The values presented in Table III were obtained
at 5% weight loss. The residue obtained is lower for blends
but comparable for all (Table III), suggesting an advantage
of the environment.
modifiers added can be attributed to instantaneous
hydrogen-bond linkage and delinkage between hydroxyl
groups of the modifier and the epoxy resin functional
groups during the test, which was also observed by
Parzuchowiski et al.28
The results indicated advantages for polymer injection
processes, as well as the filtering phenomenon reduction
during fiber impregnation process compared with other
modifiers, such as rubber particles.28
C. Blend viscosity behavior
D. Toughening mechanism
Figure 8 shows the relative viscosity of the blends
changing with modifier addition. By analyzing the
values, it is clear that both contribute to viscosity
reduction in an epoxy system, which is an effect
attributed to the hydrodynamic volumes of the branched
polymers.27
The highest reduction was achieved using HPBA
molecules, most likely due to its highest free volume.
An increasing tendency observed with the amount of
Relevant results have been found for addition of
hyperbranched modifiers until an upper limit. Figure 9
shows the modifier effect on epoxy toughening, which
could be improved up to 90% using 5% HPBA, while
among HPOH blends, a higher value was achieved at 1%
loading.
FIG. 8. Relative viscosity values of HPBA (CH) and HPOH-based
blends (CB) at T 5 80 °C and T 5 100 °C.
FIG. 7. TGA comparison between neat RTM epoxy and HPBA and
HPOH cured blends under a N2 atmosphere.
TABLE III. DMA and TGA analysis results for cured neat epoxy and
blends.
Formulation
Tg, tand
(°C)
Neat epoxy
C1H
C3H
C5H
C10H
C1B
C3B
C5B
C10B
220
217
208
204
182
216
213
208
193
E9
(GPa)
E9rp
(MPa)
T5%lwa
(°C)
Degradation
residue (%)
1.89
1.94
2.09
2.00
1.74
2.00
2.10
2.03
1.90
20.5
22.9
16.0
18.8
15.3
21.1
19.2
19.8
15.9
347.9
345.0
346.2
350.0
345.5
345.9
344.4
343.5
340.4
21.2
17.5
17.5
17.6
17.6
19.4
19.5
20.1
19.7
a
Erp 5 E9 rubbery plateau; T5%lw 5 degradation temperature at 5% loss
weight.
FIG. 9. KIC values as a function of HPBA and HPOH weight
percentages in the blends.
J. Mater. Res., 2015
http://journals.cambridge.org
Downloaded: 11 Mar 2015
7
IP address: 200.145.17.118
T. Brocks et al.: Efficiency comparison of hyperbranched polymers as toughening agents for a one-part epoxy resin
HPOH addition above 1% most likely promotes
particle aggregation and excessive phase separation,
and, consequently, an impaired effect on toughness.
This effect was also observed by Dhevi et al., 29 who
showed a property decrease caused by particle aggregation and particle-size influence on the toughening
effect.
Better average results were obtained using HPBA
as the modifier, suggesting greater homogeneity in
distribution and better modifier/epoxy interaction, but
considering the error bars only the RTM epoxy/HPBA
5% formulation is tougher than the RTM epoxy/HPOH
system.28 Matrix/modifier interaction through interfaces
is classified as either chemical or physical and must
guarantee satisfactory stress transfer from matrix to
particles, allowing a property gradient around each
particle due to a large degree of chain interaction
between both30 ; this interaction is also a reason for the
toughness improvement.31
The neat RTM epoxy matrix fracture observed in
Figs. 10(a) and 10(b) is brittle. Matrix cracks are
attributed to poor toughness.
The matrix plastic shear yielding is the main cause
of polymer toughness improvement. Considering the presence of a second phase, shown in Fig. 5, the plastic
deformation starts from this phase because its acts as a stress
concentrator. This mechanism is able to dissipate a great
amount of fracture energy by changing the original crack
plane, which increases the surface roughness, i.e., the crack
surface area. The surface roughness indicates ductile fracture,
cause by a reduction in the crosslink density, but can
also be attributed to the presence of the second phase.10,11
The second phase stops crack propagation via segmented cracking, reduction of the crack propagation rate,
and bowing out around particles, thereby improving
tension around this second phase due to plastic zone
enhancement. A triaxial tension zone change can lead to
cavitation, i.e., detachment of added particles, which is
an additional toughness mechanism that allows higher
energy dissipation and consequently a toughening
improvement in the epoxy system.10,11,29,32
The shear deformation of the epoxy matrix promotes fibril formation due to high tensile stress in the
crack-tip region, which is corroborated by the ductile
behavior promoted by epoxy modifier addition. 30
Although toughness dimples can be observed for all
blends, revealing microvoid coalescence during crack
propagation, the RTM epoxy/HPBA [Figs. 10(d), 10(f),
10(h), and 10(j)] blends display fibrils, while RTM
epoxy/HPOH [Figs. 10(c), 10(e), 10(g), and 10(i)]
blends do not show the same features.
SEM images corroborate the higher RTM
epoxy/HPBA KIC values observed in Fig. 9, with these
superior values justified by the greatest matrix deformation and consequent higher energy absorption.
8
FIG. 10. SEM surface fracture micrographs of the (a) 50x, neat RTM
epoxy; (b) 1000x, neat RTM epoxy; (c) 1000x, C1B; (d) 1000x, C1H;
(e) 1000x, C3B; (f) 1000x, C3H; (g) 1000x, C5B; (h) 1000x, C5H;
(i) 1000x, C10B; and (j) 1000x, C10H.
IV. CONCLUSIONS
In this work, the effect of two different dendritic
polymers on a one-part RTM epoxy resin was studied.
Although both HPBA and HPOH showed KIC values
higher than neat RTM epoxy due to the influence of their
shell structure on energy absorption, the addition of 5%
HPBA yielded the best results.
J. Mater. Res., 2015
http://journals.cambridge.org
Downloaded: 11 Mar 2015
IP address: 200.145.17.118
T. Brocks et al.: Efficiency comparison of hyperbranched polymers as toughening agents for a one-part epoxy resin
In addition to the higher KIC values, blends prepared with
5% HPBA did not show significant changes in thermal and
thermomechanical properties nor viscosity behavior.
By selection of an adequate amount of modifier, a reduction in particle aggregation and regular phase separation
can be obtained, thereby contributing to a good epoxy and
HBP interfacial interaction through chain entanglement. As
a consequence, stress transfer from the matrix to HBP has
been improved leading to higher toughness levels.
Fibril formation is further evidence of better entanglement and efficiency of HPBA as a toughness agent.
ACKNOWLEDGMENTS
The authors would like to thank São Paulo Research
Foundation (FAPESP - 2012/13431-7, 2011/11311-1) for
financial support, the National Research Council of Italy
(CNR-Pozzuoli-NA) and University of Naples “Federico
II” for the equipment use license and scientific support.
They would also like to thank Cytec Industries and
Perstorp for kindly providing Cytec 890 RTM and
HPOH, respectively.
REFERENCES
1. C. Wu and W. Xu: Atomistic simulation study of absorbed water
influence on structure and properties of crosslinked epoxy resin.
Polym 48, 5440 (2007).
2. F. Meyer, G. Sanz, A. Eceiza, I. Mondragon, and J. Mijovic:
The effect of stoichiometry and thermal history during cure on
structure and properties of epoxy networks. Polym 36, 1407
(1995).
3. M. Flores, X.F. Francos, F. Ferrando, X. Ramis, and A. Serra:
Efficient impact resistance improvement of epoxy/anhydride
thermosets by adding hyperbranched polyesters partially modified
with undecenoyl chains. Polym 53, 5232 (2012).
4. G. Xu, W. Shi, M. Gong, F. Yu, and J. Feng: Curing behaviour
and toughening performance of epoxy resins containing
hyperbranched polyester. Polym. Adv. Technol. 15, 639
(2004).
5. X. Zhang, W. Xu, X. Xia, Z. Zhang, and R. Yu: Toughening of
cycloaliphatic epoxy resin by nanosize silicon dioxide. Mater.
Lett. 60, 3319 (2006).
6. H. Yahyaie, M. Ebrahimi, H.V. Tahami, and E.R. Mafi:
Toughening mechanisms of rubber modified thin film epoxy
resins. Prog. Org. Coat. 76, 286 (2013).
7. J.D. Liu, H-J. Sue, Z.J. Thompson, F.S. Bates, M. Dettloff,
G. Jacob, N. Verghese, and H. Pham: Strain rate effect on
toughening of nano-sized PEP–PEO block copolymer modified
epoxy. Acta Mater. 57, 2691 (2009).
8. D. Ratna and G.P. Simon: Thermal and mechanical properties
of a hydroxyl-functional dendritic hyperbranched polymer and
trifunctional epoxy resin blends. Polym. Eng. Sci. 41, 1815
(2001).
9. Q. Guo, J. Huang, L. Ge, and Z. Feng: Phase-separation in
anhydride-cured epoxy-resin containing phenolphthalein poly
(ether ether ketone). Eur. Polym. J. 28, 405 (1992).
10. A. Saxena, B. Francis, V.L. Rao, and K.N. Ninan: Epoxy-tert-butyl
poly(cyanoarylene ether) blends: Phase morphology, fracture
toughness, and mechanical properties. J. Appl. Polym. Sci. 100,
3536 (2006).
11. P. Bussi and H. Ishida: Composition of the continuous phase in
partially miscible blends of epoxy resin and epoxidized rubber by
dynamic mechanical analysis. Polym 35, 956 (1994).
12. M. Rico, J. López, B. Montero, and R. Bellas: Phase separation
and morphology development in a thermoplastic modified
toughened epoxy. Eur. Polym. J. 48, 1660 (2012).
13. R. Mezzenga, L. Boogh, and J-A.E. Manson: A review of
dendritic hyperbranched polymer as modifiers in epoxy composite.
Compos. Sci. Technol. 61, 787 (2001).
14. G. Cicala, A. Recca, and C. Restuccia: Influence of hydroxyl
functionalized hyperbranched polymers on the thermomechanical
and morphological properties of epoxy resins. Polym. Eng. Sci.
45, 225 (2005).
15. L. Boogh, B. Pettersson, and J-A.E. Manson: Dendritic hyperbranched polymers as tougheners for epoxy resins. Polym
40, 2249 (1999).
16. F. Samperi, S. Battiato, C. Puglisi, A. Scamporrino, V. Ambrogi,
L. Ascione, and C. Carfagna: Combined techniques for the
characterization of linear-hyperbranched hybrid poly(butylene
adipate) copolymers. J. Polym. Sci. Part A: Polym. Chem.
49, 3615 (2011).
17. D. Campbell, R.A. Pethrick, and J.R. White: Polymer Characterization: Physical Techniques, 2nd ed. (Stanley Thornes Ltd, Delta
Place, United Kingdom, 2000), pp. 72–75.
18. X.F. Francos, W.D. Cook, J.M. Salla, A. Serra, and X. Ramis:
Crosslinking of mixtures of diglycidylether of bisphenol-A with
1,6-dioxaspiro[4.4] nonan-2,7-dione initiated by tertiary amines:
III. Effect of hydroxyl groups on network formation. Polym. Int.
58, 1401 (2009).
19. D. Foix, M.T. Rodriguez, F. Ferrando, X. Ramis, and A. Serra:
Combined use of sepiolite and a hyperbranched polyester in the
modification of epoxy/anhydride coatings: A study of the
curing process and the final properties. Prog. Org. Coat. 75,
364 (2012).
20. D. Foix, Y. Yu, A. Serra, X. Ramis, and J.M. Salla: Study on
the chemical modification of epoxy/anhydride thermosets using
a hydroxyl terminated hyperbranched polymer. Eur. Polym. J. 45,
1454 (2009).
21. V.P. Carvalho, Jr., C.P. Ferraz, and B.S.L. Neto: Tailored
norbornene-based copolymer with systematic variation of norbornadiene as a crosslinker obtained via ROMP with alternative
amine Ru catalysts. Eur. Polym. J. 48, 341 (2012).
22. D. Foix, X. Ramis, A. Serra, and M. Sangermano: UV generation
of a multifunctional hyperbranched thermal crosslinker to cure
epoxy resins. Polym 52, 3269 (2011).
23. P. Supaphol and N. Krutphun: Effect of small amount of poly
(ethylene naphthalate) on isothermal crystallization and spherulitic
morphology of poly(trimethylene terephthalate). Polym. Test. 26,
985 (2007).
24. D. Santiago, X.F. Francos, X. Ramis, J.M. Salla, and
M. Sangermano: Comparative curing kinetics and thermal–
mechanical properties of DGEBA thermosets cured with a
hyperbranched poly(ethyleneimine) and an aliphatic triamine.
Thermochim. Acta 526, 9 (2011).
25. W. Brostow, R. Chiu, I.M. Kalogeras, and A.V. Dova: Prediction
of glass transition temperatures: Binary blends and copolymers.
Mater. Lett. 62, 3152 (2008).
26. R.J. Varley, J.H. Hodgkin, and G.P. Simon: Toughening of
a trifunctional epoxy system: Part VI. Structure property relationships of the thermoplastic toughened system. Polym 42, 3847
(2001).
27. M. Morell, X. Ramis, F. Ferrando, and A. Serra: Effect of polymer
topology on the curing process and mechanical characteristics
of epoxy thermosets modified with linear or multiarm star
poly( e-caprolactone). Polym 52, 4694–4702 (2011).
J. Mater. Res., 2015
http://journals.cambridge.org
Downloaded: 11 Mar 2015
9
IP address: 200.145.17.118
T. Brocks et al.: Efficiency comparison of hyperbranched polymers as toughening agents for a one-part epoxy resin
28. P.G. Parzuchowiski, M. Kizlinska, and G. Rokicki: New hyperbranched polyether containing cyclic carbonate groups as a toughening agent for epoxy resin. Polym 48, 1857 (2007).
29. D.M. Dhevi, S.N. Jaisankar, and M. Pathak: Effect of new
hyperbranched polyester of varying generations on toughening
of epoxy resin through interpenetrating polymer networks using
urethane linkages. Eur. Polym. J. 49, 3561 (2013).
30. J-F. Fu, L-Y. Shi, S. Yuan, Q-D. Zhong, D-S. Zhang, Y. Chen,
and J. Wu: Morphology, toughness mechanism, and thermal
10
properties of hyperbranched epoxy modified diglycidyl ether of
bisphenol A (DGEBA) interpenetrating polymer networks. Polym.
Adv. Technol. 19, 1597 (2008).
31. R. Bagueri and R.A. Pearson: Role of particle cavitation in rubbertoughened epoxies: 1. Microvoid toughening. Polym 37, 4529
(1996).
32. P. Dittanet and R.A. Pearson: Effect of bimodal particle size
distributions on the toughening mechanisms in silica nanoparticle
filled epoxy resin. Polym 54, 1832 (2013).
J. Mater. Res., 2015
http://journals.cambridge.org
Downloaded: 11 Mar 2015
IP address: 200.145.17.118