You are on page 1of 36

Accepted Manuscript

Epoxy Toughening via Formation of Polyisoprene Nanophases with Amphi-


philic Diblock Copolymer

Yixin Xiang, Sen Xu, Sixun Zheng

PII: S0014-3057(17)31644-0
DOI: https://doi.org/10.1016/j.eurpolymj.2017.11.032
Reference: EPJ 8166

To appear in: European Polymer Journal

Received Date: 16 September 2017


Revised Date: 30 October 2017
Accepted Date: 17 November 2017

Please cite this article as: Xiang, Y., Xu, S., Zheng, S., Epoxy Toughening via Formation of Polyisoprene
Nanophases with Amphiphilic Diblock Copolymer, European Polymer Journal (2017), doi: https://doi.org/10.1016/
j.eurpolymj.2017.11.032

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Epoxy Toughening via Formation of Polyisoprene Nanophases with Amphiphilic
Diblock Copolymer

Yixin Xiang1,2, Sen Xu1 and Sixun Zheng1*


1
College of Chemistry and Chemical Engineering and the State Key Laboratory of
Metal Matrix Composites, Shanghai Jiao Tong University, Shanghai 200240, P. R.
2
China; College of Biological and Chemical Engineering, Anhui Polytechnic
University, Wuhu 241000, P. R. China

*
To whom correspondence should be addressed. Email: szheng@sjtu.edu.cn

1
ABSTRACT

In this contribution, we reported the investigation of epoxy toughening via the


formation of nanostructures by the use of poly(-caprolactone)-block-polyisoprene
(PCL-b-PI) diblock copolymer. The novel diblock copolymer was synthesized via the
combination of ring-opening polymerization and reversible addition-fragmentation
chain transfer polymerizations. This diblock copolymer can be successfully used to
access the nanostructures of epoxy thermosets. Transmission electron microscopy
(TEM) and small angle X-ray scattering (SAXS) showed that the nanostructured
thermosets were successfully obtained. In the nanostructured thermosets, the spherical
or vesicular polyisoprene microdomains were dispersed in continuous epoxy matrix.
It is the formation of the PI nanophases that endowed the thermosets with
significantly improved fracture toughness.

(Keywords: polyisoprene; poly(-caprolactone); diblock copolymers; RAFT


polymerization; nanostructured thermosets; toughness improvement)

2
INTRODUCTION

Epoxy polymers are a class of important thermosets; they have been widely
used as adhesives, matrix of high performance materials and electronic encapsulation
materials [1-7]. However, epoxy thermosets are inherently brittle because of their high
crosslinking densities. Therefore, epoxy thermosets must be toughened for their
successful application. Traditionally, epoxy thermosets are toughened by introducing
elastomers or thermoplastics. In general, curing reactions are started from the
homogenous mixtures of epoxy precursors with these modifiers and phase-separated
morphologies were then formed with the occurrence of the curing reaction with
sufficiently conversion [8,9]. It is the formation of the heterogeneous morphologies
that endows the materials with improved toughness [10,11]. In general, the
heterogeneous morphologies were only generated on the scale of micrometer since
these modifiers are some homopolymers or random copolymers. Recently, it is
realized that microphase-separated morphologies (viz. nanostructures) can be formed
if amphiphilic block copolymers are used instead of the conventional modifiers. The
generation of the nanostructures can impart significantly improved fracture toughness
to epoxy thermosets [12,13]. It is recognized that the nanostructure can be formed via
self-assembly [14,15] or reaction-induced microphase separation (RIMPS) [16,17]
approach. In the former approach, block copolymers are self-organized into the
nanoobjects in the precursors of thermosets and the pre-formed microphase-separated
morphologies are then fixed via curing reaction. In RIMPS approach, the mixtures
composed of precursors of thermosets and block copolymers are fully homogenous,
i.e., no nanophases are generated before curing reaction. The nanophases would not
be formed until the curing reaction was performed at a sufficiently high conversion.
Since Bates et al first reported the preparation of the nanostructured epoxy thermosets
by using blocks copolymers [14,15], a variety of block copolymers have been
synthesized toward this end [18-49]. No matter which mechanism is involved, the
block copolymers used for control over the formation of nanophases in thermosets
must contain both thermoset-phobic and thermoset-philic blocks [13].

Diene polymers such as polyisoprene (PI) and polybutadiene (PB) are of great
interest owing to their low glass transition temperatures (Tg’s), degradability and the
latent functionalization with ease through their double-bond-rich structures. PI and/or
PB are the materials of ideal choice for the generation of nanophases in thermosets.
3
Nonetheless, this requires the synthesis of the block copolymers containing PI or/and
PB. In general, both polyisoprene and polybutadiene are synthesized via traditional
radical polymerization in the bulk or in emulsion and living anionic polymerization.
For traditional radical polymerization, it is quite difficult to control the molecular
weights and molecular weight distribution. In addition, is not easy to synthesize the
polymers with blocky architectures. For living anionic polymerization, it is long a
challenge to control the cost and stringent condition of polymerization. In the past
decades, living radical polymerization has emerged as a class of versatile techniques
to synthesize the polymers with controlled molecular weights and macromolecular
architectures; investigators have explored the living radical polymerizations of
various vinyl monomers through atomic transfer, nitroxide-mediated and reversible
addition-fragmentation chain transfer radical polymerization [50-53]. Of a great
number of the previous reports, only a few were concerned with isoprene and
butadiene monomers. Benoit et al. [54] first reported the synthesis of PI
homopolymers and block copolymers via the nitroxide-mediated radical
polymerization of isoprene. Jitchum and Germack et al [55,56] investigated the RAFT
polymerization of isoprene and found that a trithioester chain transfer agent can be
used to mediate the radical polymerization at elevated temperatures to afford well
defined PI homopolymers. The RAFT approach has been successfully applied to
synthesize several polyisoprene-containing block copolymers such as
polyisoprene-block-polystyrene, polyisoprene-block-poly(n-butyl acrylate) diblock
copolymers, poly(styrene-alt-maleic anhydride)-block-polyisoprene diblock
copolymers and the triblock copolymers containing polyisoprene, poly(t-butyl
acrylate) and polystyrene with ABC and ACB architectures [57-59]. Unfortunately,
these block copolymers cannot be used to prepare the nanostructured epoxy
thermosets since they did not contain epoxy-philic (or miscible) blocks.

In this article, we reported the synthesis of poly(-caprolactone)-block-


polyisoprene (PCL-b-PI), a novel diblock copolymer, which can be used to access the
nanostructures in epoxy thermosets. The PCL-b-PI diblock copolymers is designed by
knowing that PCL is miscible with epoxy after and before curing reaction whereas
polyisoprene is immiscible with epoxy. It is expected that PCL-b-PI diblock
copolymer can self-assemble into nanophases in the precursors of epoxy, which can
be further fixed via curing reaction. Toward this end, we first synthesized a

4
monohydroxyl-terminated PCL; the monohydroxyl-terminated PCL was then allowed
reacting with 2-methyl-2-[(dodecylsulfanylthiocarbonyl)sulfanyl]propanoic acid
(MDPA) to afford a trithioester-terminated PCL. Second, the trithioester-terminated
PCL was used to mediate the polymerization of isoprene. To the best of our
knowledge, there has been no previous report on the synthesis of PCL-b-PI diblock
copolymers. The goal of this work is twofold: i) to explore the synthesis of PCL-b-PI
diblock copolymers and ii) to investigate the toughness improvement of the
thermosets via the formation of polyisoprene nanophases.

EXPERMENTAL

Materials
-Caprolactone (CL) was purchased from Acros Co. Shanghai, China; it was
dried over CaH2 for one week and distilled under reduced pressure. Benzyl alcohol,
stannous (II) octanoate [Sn(Oct)2], 4,4’-methylenebis (2-chloroaniline) (MOCA),
4-dimethylaminopyridine (DMAP) and 1-[3-(dimethyl
amino)propyl]-3-ethylcarbodiimide (EDC) were purchased from Shanghai Reagent
Co, China. Isoprene and di-tert-butyl peroxide were obtained from TCI Co, Shanghai,
China and it isoprene was passed through a basic alumina column and distilled. All
the solvents were purchased from Shanghai Reagent Co, China.
2-Methyl-2-[(dodecylsulfanylthiocarbonyl)sulfanyl]propanoic acid (MDPA) was
synthesized by following the method of literature [60]. Diglycidyl ether of bisphenol
A (DGEBA) which has a quoted epoxide equivalence of 185 g  mol-1 was purchased
from Shanghai Resin Co, China.

Synthesis of PCL-b-PI Diblock Copolymers

First, a monohydroxyl-terminated poly(-caprolactone) (denoted PCL-OH) was


synthesized as detailed elsewhere [34]. The PEO-OH had the molecular weight of
Mn=7,100 Da. The above PCL-OH was reacted with MDPA to obtain a
trithioester-terminated PCL. PCL-OH (15.300 g, 2.19 mmol with respect of hydroxyl
groups), MDPA (4.500 g, 12.36 mmol), EDC (2.373 g,12.36 mmol), DMAP (261 mg,
2.19 mmol) and dichloromethane (30 mL) were added to a flask. At room temperature,
5
this reaction was performed for 48 hours. After concentrated via rotary evaporation,
the reacted mixture was dropwise added into 300 mL of cold methanol to obtain the
precipitates. The excess MDPA was removed through three circles of
o
dissolution-precipitation procedure. After drying at 25 C in vacuo for 12 hours, the
macromolecular chain transfer agent (PCL-CTA) (14.250 g) was obtained with the
yield of 93.1%.

The above PCL-CTA was used for the synthesis of the PCL-b-PI diblock
copolymers. Typically, to a Schlenk flask, PCL-CTA (11.000 g, 1.6 mmol), isoprene
(44.000 g, 647.1 mmol), di-tert-butyl peroxide (80 mg, 0.55mmol) and 1,4-dioxane
(80 mL) were charged. The reactive mixture was purged with nitrogen at -40 oC for 30
min and then the flask was immersed into oil bath at 125 oC, at which the
polymerization was carried out for 20 hours. After polymerization, the solvent and the
unreacted isoprene were removed and then 5 mL of dichloromethane was added. The
as-obtained solution was dropwise added into 400 mL of cold methanol. After drying,
the product (22.000 g) was obtained with the conversion of isoprene of 25%. 1H NMR
(CDCl3, ppm): 4.06 [2H, OCO(CH2)4CH2], 2.30 [2H, OCOCH2(CH2)4], 1.65 [4H,
OCOCH2CH2CH2CH2CH2], 1.37 (2H, OCOCH2CH2CH2CH2CH2), 4.65 ~ 4.79 [2H,
-C(CH3)=CH2], 1.6 ~ 1.8, 2.0 ~ 2.2, methyl and methylene protons of PI), 4.85 ~ 5.05
(2H, -CH=CH2), 5.1 ~ 5.5 [2H, -CH=C(CH3)-], 5.6 ~ 5.9 (1H, -CH=CH2). GPC:
Mn=14,100 with Mw/Mn=1.26.

Preparation of Nanostructured Epoxy Thermosets

Desired amount of PCL62-b-PI103 was added to DGEBA with vigorous stirring


and then a stoichiometric amount of MOCA with respect of DGEBA was added. The
mixtures were vigorously stirred at 100 oC until the systems became homogenous and
transparent. The mixtures were cured at 150 oC for 3 hours plus180 oC for 2 hours.

Measurements and Techniques


A Varian Mercury Plus 400 MHz nuclear magnetic resonance (NMR)
spectrometer was used to obtain the spectra. Gel permeation chromatography (GPC)
was conducted on a Waters 1515 system equipped with Waters RH columns and a

6
Dwan Eos (Eyatt Technology) multi-angle laser scattering detector. Tetrahydrofuran
was used as the eluent at the rate of 1.0 mL/min; the molecular weights were
expressed relative to polystyrene standard. Differential scanning calorimetry (DSC)
was performed on a TA Instruments Q2000 differential scanning calorimeter. The
morphologies were observed by means of JEM 2010 high-resolution transmission
electron microscope at an accelerating voltage of 200 kV. The ultrathin sections (c.a.
70 nm in thickness) of the thermosets were stained with OsO4. Small-angle x-ray
scattering (SAXS) was conducted in the Shanghai Synchrotron Radiation Facility
(SSRF), Shanghai, China as detailed elsewhere [43]. Dynamic mechanical thermal
analysis (DMTA) was performed on Fracture
toughness was measured with the notched three-point bending tests as detailed
elsewhere [31].

RESULTS AND DISCUSSION

Synthesis of PCL-b-PI Diblock Copolymers

First, PCL-OH was synthesized via ROP of -caprolactone with benzyl


alcohol as the initiator and then the as-obtained PCL-OH was reacted with MDPA,
yielding a trithioester-capped PCL. The latter was used to mediate the radical
polymerization of isoprene to obtain PCL-b-PI diblock copolymers as depicted in
Scheme 1. By controlling the molar ratios of PCL-CTA to isoprene, the diblock
copolymers with various lengths of PI blocks were obtained. Shown in Figure 1 are
the 1H NMR spectra of PCL-OH, PCL-CTA and PCL62-b-PI103. For PCL-OH, the
signals of resonance at 1.37, 1.64, 2.31 and 4.06 ppm are attributable to the protons of
methylene groups of PCL chain as indicated. The minor peak at 3.65 ppm is
assignable to the methylene protons of the terminal hydroxymethyl group. According
to the ratio of the integral intensity of this peak to that of other methylene protons, the
molecular weight of the PCL-OH was estimated to be Mn=7,100 Da. With the
occurrence of the esterification of PCL-OH with MDPA, notably, the signal at 3.65
ppm was fully shifted to 3.71 ppm. Concurrently, a minor peak appeared at 1.27 ppm,
assignable to the methyl protons from MDPA. The 1H NMR spectroscopy shows that
the terminal groups of PCL-OHs were fully capped with trithioester moieties, i.e., the
PCL-CTA was successfully obtained. Also representatively shown in Figure 1 is the

7
1
H NMR spectrum of PCL62-b-PI103. Several new signals at 0.96, 2.06, 4.5~6.5 ppm
were displayed; they are assignable to the proton resonances of methyl and methane
groups of PI block. It is noteworthy that 1,2-, 3,4- and 1,4-additions could be involved
with the polyaddition of this diene monomers. The 1,2-addition can be evidenced by
the signals of resonance at 5.60 ~ 5.90 and 4.85 ~ 5.05 ppm, respectively. The former
is attributable to the protons of methine groups whereas latter to those of methylene
with 1,2 addition. The broad peak in the range of 5.10 ~ 5.50 ppm resulted from the
methine proton in the -CH=C(CH3)- structural units whereas those at 4.60 ~ 4.85 ppm
are assignable to the protons of methylene group in the structural unit of
-C(CH3)=CH2 with 3,4 addition. In terms of the ratio of integral intensity of the
resonance of these protons, the percentages of 1,2, 3,4 and 1,4 addition structural units
were estimated to be 10, 10 and 80%, respectively. The molecular weights of the
diblock copolymers were measured by means of GPC. The GPC curves are presented
in Figure 2 and the data are listed in Table 1. In all the cases, unimodal distribution of
molecular weights was exhibited. Notably, the peaks increasingly shifted to the left
sides with increasing the lengths of PI blocks. In all the cases, the values of
polydispersity index were in the range of 1.22 ~ 1.28. The narrow distribution of
molecular weights suggests that the radical polymerization of isoprene mediated with
PCL-CTA could be in a living/controlled manner. The results of 1H NMR and GPC
showed that the PCL-b-PI diblock copolymers were successfully synthesized.

Figure 3 shows the DSC curves of PCL-OH and the PCL-b-PI diblock
copolymers. For PCL-OH, an endothermic peak was exhibited at 58 oC, assignable to
the melting transition of this crystalline polymer. Notably, the glass transition was
feeble, suggesting that the crystallinity of the PCL-OH was quite high. For each
PCL-b-PI diblock copolymer, one glass transition and one melting transition were
displayed at -60 and 58 oC, respectively. The former is ascribed to PI block whereas
the latter to PCL block. Notably, the glass transition temperatures (Tg’s) remained
unchanged, irrespective of the composition of the diblock copolymers. The melting
points (Tm’s) of the PCL blocks were slightly lower than that of PCL-OH. With the
enthalpy values of the melting transitions, the crystallinity of PCL (X %) was
estimated according to the following equation:

8
H f
X%  100% (1)
H of  f PCL

where Hf is the measured enthalpy of melting peak and Hfo is that of perfectly
crystalline PCL and has been reported to be Hfo = 16.2 KJ  mol-1 [57]; fPCL is the
mass fraction of PCL block in the diblock copolymer. The plot of crystallinity (X%) of
PCL as a function of the mass fraction of PCL in the diblock copolymers was also
shown in Figure 3 (See the inset). It is noted that the crystallinity remained almost
invariant, irrespective of the copolymer composition. The fact that the Tg’s of PI
blocks, the Tm’s and crystallinity (X%) of PCL blocks were remained invariant and
irrespective of the compositions of the diblock copolymers indicates that both of the
blocks were less affected with each other, i.e., the PCL-b-PI diblock copolymers were
microphase-separated. Also shown in Figure 3 are the cooling scans of the PCL-OH
and PCL-b-PI diblock copolymers. For PCL-OH, an exothermic transition occurred at
30 oC, which is attributable to the crystallization of PCL in the cooling scan. The
similar crystallization peaks were also detected for the diblock copolymers, assignable
to the PCL blocks. The temperatures of crystallization (Tc’s) were slightly lower than
that of PCL-OH. While the lengths of the PI blocks were sufficiently long, the peaks
of crystallization became very weak, e.g. for PCL62-b-PI147; in the meantime, there
appeared a minor peak at a lower temperature (i.e., at -42 oC). For PCL62-b-PI515, no
crystallization was even exhibited. This observation can be explained on the basis of
the restriction of PI microdomains on the crystallization of PCL blocks. It is realized
that the crystallization of crystallizable blocks in bulks is quite dependent on
microdomain connectivity, mechanical states and Tg’s of the continuous matrices
[58-62]. While the mass fraction of PI block was much higher than that of PCL block,
the PCL blocks formed the isolated microdomains which were dispersed in
continuous PI matrix. In this case, the crystallization of PCL microdomains was in a
restricted manner although the Tg’s of PI was much lower than the Tm’s of PCL owing
to their too big inter-microdomain connectivity. As a result, the crystallization must be
driven by a much larger supper cooling (i.e., Tm-Tc).

Self-assembly in Epoxy Thermosets

In this work, one of the diblock copolymers (viz. PCL62-b-PI103) was


9
incorporated into epoxy to obtain the nanostructured thermosets. Depending on the
content of PCL62-b-PI103 diblock copolymer, all the epoxy thermosets were
transparent or translucent (See Figure 4). The transparency of the samples suggests
that if any the microdomains dispersed in the continuous epoxy matrices had the sizes
smaller than (or close to) the wavelength of visible lights. The morphologies of the
thermosets were further investigated by means of transmission electron microscopy
(TEM). Figure 5 shows TEM micrographs of the samples. In order to increase the
contrast in electron density, the ultrathin sections with the thickness of 70 ~100 nm
were stained with OsO4. It is proposed that PI blocks would be preferentially stained
through their unsaturated C=C bonds whereas epoxy remained less affected.
Therefore, the dark microdomains are attributable to PI block whereas the light to the
matrix of epoxy. It is seen that all the thermosets containing PCL62-b-PI103 were
heterogeneous on the scale of nanometer and the microphase-separated morphologies
were quite dependent on the concentration of the diblock copolymer in the thermosets.
While the content of PCL62-b-PI103 was 10 wt%, the spherical PI microdomains were
formed with the diameter of 20 nm (See Figure 5A). With increasing the content of
PCL62-b-PI103, the size of the spherical PI microdomains slightly increased and there
appeared some vesicular microdomains appeared. It is seen that the number and size
of the vesicular microdomains increased with the content of PCL62-b-PI103 (See
Figure 5B through 5C). For the thermosets containing 40 wt% of PCL62-b-PI103, some
giant vesicular nanoobjects were formed with the size as high as 150 nm in diameter.

The formation of the nanostructures was further confirmed by means of SAXS.


As shown in Figure 6, the strong scattering peaks were displayed in the SAXS curves,
indicating that these thermosets were indeed microphase-separated. The primary
scattering peaks gradually shifted to the higher q positions with increasing the content
of PCL62-b-PI103, indicating that the long periods of the nanostructures were decreased.
The formation mechanism of the nanostructures was readily distinguished by SAXS.
Representatively shown in Figure 7 are the SAXS curves of the mixture of DGEBA,
MOCA and 30 wt% of PCL62-b-PI103. At room temperature, a ramp and broad
scattering peak was exhibited with q = 0.24 nm-1 (See Curve A in Figure 6). The
scattering peak is indicative of the inhomogeneous morphology, i.e., there were the
nanophases in the ternary mixture, i.e., the diblock copolymer was self-assembled into
the microdomains. The SAXS result was further confirmed by means of atomic force

10
microscopy (AFM) and the AFM images of the mixture are shown in Figure 8. The
left and right are topological and phase shift images, respectively. The topological
image shows that the specimen was quite flat and thus the phase shift image can be
used to indicate the microphase behavior. According to the difference in
viscoelasticity among the components (i.e., DGEBA + MOCA, PCL and PI), it is
judged that the dark region is attributable to the PI microdomains. It is seen that the PI
spherical and worm-like microdomains were dispersed into the continuous matrix.
The ternary mixture was rapidly heated to the curing temperature (i.e., 150 oC),
notably, the scattering peak was still preserved; the position of the scattering peak
remained unchanged but became sharper (See Curve B in Figure 7). This
phenomenon suggests that at the curing temperature the self-assembly nanostructure
still existed in the ternary mixture, i.e., the curing reaction would be started in the
presence of the self-assembly nanostructure. After curing, the thermoset still displayed
the scattering peak (See Curve C in Figure 7). Compared to the mixture at 150 oC,
notably, the scattering peak of the thermoset was observed to shift to q = 0.21 nm-1, a
lower q value. This observation was associated with the extraction of the precursors of
epoxy from the swollen PI microdomains with the occurrence of the curing reaction.
The SAXS results indicate that in the thermosets the PI nanophases were formed
through a self-assembly approach.

The results of TEM and SAXS showed that in the nanostructured thermosets,
PI microdomains were demixed out of epoxy matrix in the forms of spherical or
vesicular microdomains whereas PCL block could remain intimately mixed with
epoxy. The microphase separation in the thermosetting blends of epoxy with
PCL-b-PIp diblock copolymer can be confirmed by DMTA. Figure 9 shows the
DMTA curves of plain epoxy and the nanostructured thermosets. For plain epoxy, the
 transition was exhibited at 156 oC; this peak is responsible for the glass-rubber
transition of the crosslinked polymer. In addition,  transition was discernable at -56
and 70 oC, respectively. The former is attributable to the motion of hydroxyether
structural units of amine-crosslinked epoxy whereas the latter to diphenyl groups of
bisphenol-A type epoxy [63-65]. Notably, Tg’s of the matrix gradually decreased with
increasing the content of PCL62-b-PI103 diblock copolymer. The decreased Tg’s is
attributed to the plasticization of the miscible PCL blocks on epoxy matrices. For each
nanostructured thermoset, there was a second major transition at c.a. -54 oC. The
11
intensity of the second major transitions increased with increasing the content of
PCL62-b-PI103 diblock copolymer. The second major transition is attributable to the PI
microdomains. Both separated glass transitions revealed that the epoxy thermosets
containing PCL62-b-PI103 diblock copolymer was indeed microphase-separated as
shown by TEM and SAXS. Owing to the inclusion of the elastomer (i.e. PI), the
nanostructured thermosets at low temperature had the moduli much lower than that of
plain epoxy (See Figure 9). It is the inclusion of elastomeric microdomains that
imparts the improved fracture toughness to the thermosetting polymer.

Fracture Toughness of Nanostructured Thermosets

Plain epoxy and the nanostructured thermosets were subjected to three-point


bending tests, with which the critical stress field intensity factors (denoted KIC)
together with critical fracture energy (GIC) were measured. Shown in Figure 10 are the
plots of KIC and GIC as functions of the content of PCL62-b-PI103 diblock copolymer.
For plain epoxy, the values of KIC and GIC were measured to be 0.52 MN  m-3/2 and
0.19 kJ  m-2, respectively. Upon adding PCL62-b-PI103 diblock copolymer to epoxy,
the KIC and GIC values were significantly enhanced. In the composition range
investigated, the KIC and GIC values of the nanostructured thermosets were much
higher than that of plain epoxy. The nanostructured thermosets containing 20 wt% of
PCL62-b-PI103 possessed the values of KIC and GIC to be 1.43 MN  m-3/2 and 0.95 KJ
 m-2, respectively; the values were almost three times as that of plain epoxy. The
values of KIC and GIC indicate that epoxy thermosets were significantly toughened
with the inclusion of PI microdomains. It is proposed that the enhanced fracture
toughness resulted from the following factors: i) the formation of elastomeric
nanophases and ii) the strong adhesion between the dispersed PI microdomains and
epoxy matrix. In this work, the microdomains of PI were formed via the self-assembly
of the PCL62-b-PI103 diblock copolymer. Compared to traditional macroscopic
phase-separated morphologies, the microphase-separated morphologies provided a
great amount of interface between thermosetting matric and the modifiers, which is
critical to the interfacial interactions in terms of several proposed mechanisms of
toughness improvement. Due to the connectivity of the copolymer blocks, there
existed the strong chemical linkage between epoxy matrix and the dispersed PI

12
microdomains. The latter is also important for the toughness enhancement of
thermosets with fine phase-separated morphologies.

CONCLUSION

In this work, we successfully PCL-b-PI diblock copolymers via sequential


living polymerization approach. First, a PCL-OH was synthesized via ROP approach
of -caprolactone; the PCL-OH was allowed reacting with MDPA to afford a
trithioester-terminated PCL (PCL-CTA). Second, PCL-CTA was used to mediate the
radical polymerization of isoprene. By controlling the molar ratios of PCL-CTA to
isoprene, a series of PCL-b-PI diblock copolymers were obtained. DSC showed that
the diblock copolymers were microphase-separated. One of these PCL-b-PI diblock
copolymers was used to modulate the formation of nanostructures in epoxy
thermosets. The results of TEM and SAXS showed that the nanostructured thermosets
were successfully obtained. The incorporation of the diblock copolymers into epoxy
formed the PI microdomains with the spherical and vesicular morphologies whereas
the PCL blocks remained mixed with epoxy matrix. It is the formation of the
nanostructured that endowed the thermosetting polymer with improved fracture
toughness.

ACKNOWLEDGEMENT
The financial supports from National Natural Science Foundation of China
(Grant No. 51133003, 21274091 and 21774078) were acknowledged. The authors
thank the Shanghai Synchrotron Radiation Facility for the support under the projects
of No. 13SRBL16B14042.

REFERENCES

1. C. B. Bucknall, I. K. Partridg, Phase separation in epoxy resins containing


polyethersulphone, Polymer 24 (1983) 639-644.

2. R. A. Pearson, A.F. Yee, Toughening mechanisms in elastomer-modified epoxies,


J. Mater. Chem. 24 (1989) 2571-2580.

13
3. H. S. Y. Hsich, Morphology and properties control on rubber‐epoxy alloy
systems, Polym. Eng. Sci. 30 (1990) 493-510.

4. J. L. Hedrick, I. Yilgor, M. Jurek, J.C. Hedrick, G. L. Wilkes, J. E. McGrath,


Chemical modification of matrix resin networks with engineering thermoplastics:
1. Synthesis, morphology, physical behavior and toughening mechanisms of
poly(arylene ether sulphone) modified epoxy networks, Polymer 32 (1991)
2020-2032.

5. C. B. Bucknall, A.H. Gilbert, Toughening tetrafunctional epoxy resins using


polyetherimide, Polymer 30 (1989) 213-217.

6. R. J. J. Williams, B. A. Rozenberg, J.-P. Pascault, Reaction-induced phase


separation in modified thermosetting polymers, Adv. Polym. Sci. 128 (1997)
97-153;

7. S. Zheng, J. Wang, Q. Guo, J. Wei, J. Li, Miscibility, morphology and fracture


toughness of epoxy resin/poly (styrene-co-acrylonitrile) blends, Polymer 37(1996)
4667-4673.

8. J.-P. Pascault, R. J. J. Williams, In Polymer Blends, D. R. Paul, C. B. Bucknall


Eds. New York: Wiley; 2000. p. 379-415.

9. L. Ruiz-Pérez, G. J. Royston, J. A. Fairclough, A. J. Ryan, Toughening by


nanostructure, Polymer 49 (2008) 4475-4588

10. S. Zheng, In Epoxy Polymers: New Materials and Innovations, J.-P. Pascault, R. J.
J. Williams, Eds. Weinheim: Wiley-VCH; 2010. p.79-108

11. R. Bagheri, B. T. Marouf, R. A. Pearson, Rubber-Toughened Epoxies: A Critical


Review, Polym. Rev. 49 (2009) 201-225

12. A. Gilbert, C. Bucknall, Epoxy resin toughened with thermoplastic. Vol 45, Wiley
Online Library, Amsterdam, 1991, pp 289-298

13. M. A. Hillmyer, P. M. Lipic, D. A. Hajduk, K. Almdal, F. S. Bates, Self-assembly


and polymerization of epoxy resin-amphiphilic block copolymer nanocomposites,
J. Am. Chem. Soc. 119 (1997) 2749-2750

14. P. M. Lipic, F. S. Bates, M. A. Hillmyer, Nanostructured thermosets from

14
self-assembled amphiphilic block copolymer/epoxy resin mixtures, J. Am. Chem.
Soc. 120 (1998) 8963-8970

15. F. Meng, S. Zheng, W. Zhang, H. Li, Q. Liang, Nanostructured thermosetting


blends of epoxy resin and amphiphilic poly(ε-caprolactone)-block-
polybutadiene-block-poly(ε-caprolactone) triblock copolymer, Macromolecules
39 (2006) 711-719

16. F. Meng, S. Zheng, H. Li, Q. Liang, T. Liu, Formation of ordered nanostructures


in epoxy thermosets: a mechanism of reaction-induced microphase separation,
Macromolecules 39 (2006) 5072-5080.

17. J. Mijovic, M. Shen, J. W. Sy, I. Mondragon, Dynamics and morphology in


nanostructured thermoset network/block copolymer blends during network
formation, Macromolecules 33 (2000) 5235-5244.

18. R. B. Grubbs, J. M. Dean, M. E. Broz, F. S. Bates, Reactive block copolymers for


modification of thermosetting epoxy, Macromolecules 33 (2000) 9522-9534.

19. R. B. Grubbs, J. M. Dean, F. S. Bates, Methacrylic block copolymers through


metal-mediated living free radical polymerization for modification of
thermosetting epoxy, Macromolecules 34 (2001) 8593-8595.

20. Q. Guo, R. Thomann, W. Gronski, Phase behavior, crystallization, and


hierarchical nanostructures in self-organized thermoset blends of epoxy resin and
amphiphilic poly(ethylene oxide)-block-poly(propylene oxide)-block-
poly(ethylene oxide) triblock copolymers, Macromolecules 35 (2002) 3133-3144

21. S. Ritzenthaler, F. Court, E. Girard-Reydet, L. Leibler, J.-P. Pascault, ABC


triblock copolymers/epoxy-diamine blends. 1. Keys to achieve nanostructured
thermosets, Macromolecules 35 (2002) 6245-6254.

22. S. Ritzenthaler, F. Court, E. Girard-Reydet, L. Leibler, J.-P. Pascault, ABC


triblock copolymers/epoxy-diamine blends. 2. Parameters controlling the
morphologies and properties, Macromolecules 36 (2003) 118-126.

23. V. Rebizant, V. Abetz, T. Tournihac, F. Court, L. Leibler, Reactive tetrablock


copolymers containing glycidyl methacrylate. Synthesis and morphology control
in epoxy-amine networks, Macromolecules 36 (2003) 9889-9896

15
24. V. Rebizant, A. S. Venet, F. Tournillhac, E. Girard-Reydet, C. Navarro, J.-P.
Pascault, L. Leibler, Chemistry and mechanical properties of epoxy-based
thermosets reinforced by reactive and nonreactive SBMX block copolymers,
Macromolecules 37 (2004) 8017-8027.

25. J. M. Dean, N. E. Verghese, H. Q. Pham, F. S. Bates, Nanostructure toughened


epoxy resins, Macromolecules 36 (2003) 9267-9270.

26. I. A. Zucchi, M. J. Galante, R. J. J. Williams, Comparison of morphologies and


mechanical properties of crosslinked epoxies modified by polystyrene and
poly(methyl methacrylate) or by the corresponding block copolymer
polystyrene-b-poly(methyl methacrylate), Polymer 46 (2005) 2603-2607.

27. Y. S. Thio, J. Wu, F. S. Bates, Epoxy toughening using low molecular weight
poly(hexylene oxide)-poly(ethylene oxide) diblock copolymers, Macromolecules
39 (2006) 7187-7189.

28. E. Serrano, A. Tercjak, G. Kortaberria, J.A. Pomposo, D. Mecerreyes, N. E.


Zafeiropoulos, M. Stamm, I. Mondragon, Nanostructured thermosetting systems
by modification with epoxidized styrene-butadiene star block copolymers. Effect
of epoxidation degree, Macromolecules 39 (2006) 2254-2261.

29. C. Ocando, E. Serrano, A. Tercjak, C. Pena, G. Kortaberria, C. Calberg, B.


Grignard, R. Jerome, P.M. Carrasco, D. Mecerreyes, I. Mondragon, Structure and
properties of a semifluorinated diblock copolymer modified epoxy blend
Macromolecules 40 (2007) 4068-4074.

30. S. Maiez-Tribut, J.-P. Pascault, E. R. Soule, J. Borrajo, R. J. J. Williams,


Nanostructured epoxies based on the self-assembly of block copolymers: a new
miscible block that can be tailored to different epoxy formulations,
Macromolecules 40 (2007) 1268-1273.

31. W. Gong, K. Zeng, L. Wang, S. Zheng, Poly(hydroxyether of bisphenol


A)-block-polydimethylsiloxane alternating block copolymer and its
nanostructured blends with epoxy resin, Polymer 49 (2007) 3318-3326.

32. F. Yi, S. Zheng, T. Liu, Nanostructures and surface hydrophobicity of


self-assembled thermosets involving epoxy resin and poly(2, 2, 2-trifluoroethyl
acrylate)-block-poly(ethylene oxide) amphiphilic diblock copolymer, J. Phys.
16
Chem. B 113 (2009) 1857-1868.

33. C. Sinturel, M. Vayer, R. Erre, H. Amenitsch, Nanostructured polymers obtained


from polyethylene-block-poly(ethylene oxide) block copolymer in unsaturated
polyester, Macromolecules 40 (2007) 2532-2535

34. Z. Xu, S. Zheng, Reaction-induced microphase separation in epoxy thermosets


containing poly(-caprolactone)-block-poly(n-butyl acrylate) diblock copolymer.
Macromolecules 40 (2007) 2548-2558.

35. F. Meng, Z. Xu, S. Zheng, Microphase separation in thermosetting blends of


epoxy resin and poly(ε-caprolactone)-block-polystyrene block copolymers,
Macromolecules 41 (2008) 1411-1420;

36. W. Fan, L. Wang, S. Zheng, Nanostructures in thermosetting blends of epoxy


resin with polydimethylsiloxane-block-poly(ε-caprolactone)-block-polystyrene
ABC triblock copolymer, Macromolecules 42 (2008) 327-336.

37. C. Ocando, A. Tercjak, M. D. Martín, J. A. Ramos, M. Campo, I. Mondragon,


Morphology development in thermosetting mixtures through the variation on
chemical functionalization degree of poly(styrene-b-butadiene) diblock
copolymer modifiers. Thermomechanical properties, Macromolecules 42 (2009)
6215-6224.

38. W. Fan, L. Wang, S. Zheng, Double reaction-induced microphase separation in


epoxy resin containing polystyrene-block-poly(ε-caprolactone)-block-
poly(n-butyl acrylate) ABC triblock copolymer, Macromolecules 43 (2010)
10600-10611.

39. E. Serrano, M. Larranaga, P. M. Remiro, I. Mondragon, P. M. Carrasco, J.A.


Pomposo, D. Mecerreyes, Synthesis and characterization of epoxidized
styrene-butadiene block copolymers as templates for nanostructured thermoset,
Macromol. Chem. & Phys. 205 (2004) 987-996.

40. E. Serrano, M. D. Martin, A. Tercjak, J. A. Pomposo, D. Mecerreyes, I.


Mondragon, Nanostructured thermosetting systems from epoxidized styrene
butadiene block copolymers, Macromol. Rapid. Commun. 26 (2005) 982-985

41. S. Wu, Q. Guo, S. Peng, N. Hameed, M. Kraska, B. Stühn, Y. W. Mai,


17
Toughening epoxy thermosets with block ionomer complexes: a
nanostructure–mechanical property correlation, Macromolecules 45 (2012)
3829-3840.

42. R. Yu, S. Zheng, Morphological transition from spherical to lamellar nanophases


in epoxy thermosets containing poly(ethylene
oxide)-block-poly(ε-caprolactone)-block-polystyrene triblock copolymer by
hardeners, Macromolecules 44 (2011) 8546-8557.

43. R. Yu, S. Zheng, X. Li, J. Wang, Reaction-induced microphase separation in


epoxy thermosets containing block copolymers composed of polystyrene and
poly(ε-caprolactone): Influence of copolymer architectures on formation of
nanophases, Macromolecules 45 (2012) 9155-9168.

44. H. Garate, I. Mondragon, N. B. D’Accorso, S. Goyanes, Exploring microphase


separation behavior of epoxidized poly (styrene-b-isoprene-b-styrene) block
copolymer inside thin epoxy coatings, Macromolecules 46 (2013) 2182-2187.

45. C. Zhang, L. Li, S. Zheng, Formation and confined crystallization of polyethylene


nanophases in epoxy thermosets, Macromolecules. 46 (2013) 2740-2753.

46. C. Ocando, R. Fernández, A. Tercjak, I. Mondragon, A. Eceiza, Nanostructured


thermoplastic elastomers based on SBS triblock copolymer stiffening with low
contents of epoxy system. Morphological behavior and mechanical
properties, Macromolecules 46 (2013) 3444-3451.

47. H. Cong, S. Xu, S. Zheng, Synthesis and Microphase Separation Behavior of


Random, Mixed Cylindrical Brush Copolymers Bearing Polystyrene and
Poly(ε-caprolactone) Side Chains, Chinese J. Polym. Sci., DOI:
10.1007/s10118-017-2001-y.

48. J. McLeary, F. Calitz, J. McKenzie, M. Tonge, R. Sanderson, B. Klumperman,


Beyond inhibition: a 1H NMR investigation of the early kinetics of
RAFT-mediated polymerization with the same initiating and leaving
groups, Macromolecules 37 (2004) 2383-2394.

49. T. Pintauer, K. Matyjaszewski, Atom transfer radical addition and polymerization


reactions catalyzed by ppm amounts of copper complexes, Chem. Soc. Rev. 37
(2008) 1087-1097.
18
50. C. J. Hawker, A.W. Bosman, E. Harth, New polymer synthesis by nitroxide
mediated living radical polymerizations, Chem. Rev., 101 (2001) 3661-3688.

51. G. Moad, E. Rizzardo, S. H. Thang, Radical addition-fragmentation chemistry in


polymer synthesis, Polymer 49 (2008) 1079-1131;

52. A. H. Müller, K. Matyjaszewski, Controlled and Living Polymerizations:


Methods and Materials. Weinheim,Germany: Wiley-VCH, 2009.

53. D. Benoit, E. Harth, P. Fox, R. M. Waymouth, C. J. Hawker, Accurate structural


control and block formation in the living polymerization of 1, 3-dienes by
nitroxide-mediated procedures, Macromolecules 33 (2000) 363-370.

54. D. S. Germack, K. L. Wooley, Isoprene polymerization via reversible addition


fragmentation chain transfer polymerization, J. Polym. Sci. Part A: Polym. Chem.
45 (2007) 4100-4108.

55. D. S. Germack, K. L. Wooley, RAFT-based synthesis and characterization of


ABC versus ACB triblock copolymers containing tert-butyl acrylate, isoprene,
and styrene blocks, Macromol. Chem. & Phys. 208 (2007) 2481-2491.

56. V. Jitchum, S. Perrier, Living radical polymerization of isoprene via the RAFT
process, Macromolecules 40 (2007) 1408-1412.

57. J. T. Lai, D. Filla, R. Shea, Functional polymers from novel carboxyl-terminated


trithiocarbonates as highly efficient RAFT agents, Macromolecules 35 (2002)
6754-6756.

58. L. H. Sperling. Introduction to physical polymer science, 4 th Edn. Oxford, Eng.


Oxford University Press, 2006.

59. C. De Rosa, C. Park, E. L. Thomas, B. Lotz, Microdomain patterns from


directional eutectic solidification and epitaxy, Nature 405 (2000) 433.

60. L. Zhu, B. Mimnaugh, Q. Ge, R. P. Quirk, S. Z. D. Cheng, E. L. Thomas, B. Lotz,


B. S. Hsiao, F. Yeh, L. Liu, Hard and soft confinement effects on polymer
crystallization in microphase separated cylinder-forming PEO-b-PS/PS blends,
Polymer 42 (2001) 9121-9131.

61. Y. L. Loo, R. A. Register, A. J. Ryan, Modes of crystallization in block copolymer

19
microdomains: breakout, templated, and confined, Macromolecules 35 (2002)
2365-2374.

62. R. V. Castillo, A. J. Mueller, M. C. Lin, H. L. Chen, U. S. Jeng, M. A. Hillmyer,


Confined crystallization and morphology of melt segregated PLLA-b-PE and
PLDA-b-PE diblock copolymers, Macromolecules 41 (2008) 6154-6164.

63. D. E. Kline, Dynamic mechanical properties of polymerized epoxy resins, J.


Polym. Sci. Part A: Polym. Chem. 47 (1960) 237-249.

64. L. M. Robeson, The effect of antiplasticization on secondary loss transitions and


permeability of polymers, Polym. Eng. & Sci. 9 (1969) 277-281.

65. A. A. Jones, Molecular level model for motion and relaxation in glassy
polycarbonate, Macromolecules 18 (1985) 902-906.

20
SCHEMES

Scheme 1 Synthesis of PCL-b-PI diblock copolymers

21
Table 1 Molecular weights of PCL and PCL-b-PI diblock copolymers via RAFT
polymerization

Samples LPCL(Da) LPI(Da) Mn(Da) Mw/Mn

PCL62 7,100 -- 7,100 1.20

PCL62-b-PI30 7,100 2,000 9,100 1.22

PCL62-b-PI66 7,100 4,500 10,600 1.23

PCL62-b-PI103 7,100 7,000 14,100 1.26

PCL62-b-PI147 7,100 10,000 17,100 1.25

PCL62-b-PI515 7,100 35,000 42,100 1.28

22
FIGURE CAPTIONS
1
Figure 1. H NMR spectra of PCL, PCL-CTA and PCL62-b-PI103 diblock
copolymer;

Figure 2. GPC curves of PCL-CTA and PCL-b-PI diblock copolymers;

Figure 3. DSC curves of plain PCL and PCL-b-PI diblock copolymers; inset: Plot
of PCL crystallinity as a function of the content of PCL for plain PCL
and PCL-b-PI diblock copolymer. The heating and cooling rates were 20
and 10 oC/min, respectively.

Figure 4. Photos of the epoxy thermosets containing: A) 0, B) 10, C) 20, D) 30


and E) 40 wt% of PCL62-b-PI103 diblock copolymer.

Figure 5. TEM images of the epoxy thermosets containing: A) 10, B) 20, C) 30


and (D) 40 wt % of PCL62-b-PI103 diblock copolymer;

Figure 6. SAXS profiles of the epoxy thermosets containing PCL62-b-PI103 diblock


copolymer;

Figure 7. SAXS profiles of the mixture of DGEBA, MOCA and 30 wt%


PCL62-b-PI103 diblock copolymer at: A) 25 oC, B) 150 oC and C) the
cured thermoset, respectively;

Figure 8. AFM images of the mixture of the epoxy precursor (i.e., DGEBA +
MOCA) with PCL62-b-PI103 (30 wt%). The left and right are the
topological and phase shift images, respectively.

Figure 9. DMTA curves of control epoxy and the nanostructured thermosets


containing PCL62-b-PI103 diblock copolymer. The contents of the
copolymer are 10, 20, 30 and 40 wt%, respectively;

Figure 10. Fracture toughness of control epoxy and the nanostructured thermosets
containing PCL62-b-PI103 diblock copolymer.

23
j l
O m k S
b c e e O n i
C12H25
a O S S
d f g m x y p z
O
g
n n

d e +o
TMS
CDCl3 p f

a +b c +n
i
k j l g'

O
e
m S
b c e e i

d
O C12H25
a O d f g(h) m
S S f
O g
a +b
c g' m +i

O d
b c e e OH e f
a O
d f
g
g( h) m

a +b h
c

8 7 6 5 4 3 2 1 0
Chemical shift (ppm)

Figure 1

24
PCL62
PCL62-b -PI30
PCL62-b -PI66
PCL62-b -PI103
PCL62-b -PI147
PCL62-b -PI515

10 12 14 16 18 20
Retention time (min)

Figure 2

25
100

80

Crystallinity (%)
60

40

20

0
0 25 50 75 100
PCL (wt%)
PCL62

PCL62-b -PI30
Endo

PCL62-b -PI66

PCL62-b -PI103

PCL62-b -PI147

PCL62-b -PI515

-40 0 40 80 120
o
Temperature ( C)

Figure 3

26
A B C D E

Figure 4

27
A B

500nm 500nm

C D

500nm 500nm

Figure 5

28
PCL62-b -PI103
(wt%)
Intensity (a .u .)

40

30

20

10

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4

q (nm-1)

Figure 6

29
PCL62-b -PI103
(wt%)
Intensity (a .u .)

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4


-1
q (nm )

Figure 7

30
Figure 8

31
104 1.5

103 1.2
Storage modulus (MPa)

PCL62-b -PI103
(wt%)
0
102
10 0.9
20

Tan 
30
40
101 0.6

100 T
g
(PI) 0.3

10-1 0.0
-100 -50 0 50 100 150 200

Temperature ( oC)

Figure 9

32
2.0 1.4

K
1c
G 1.2
1c
1.6

1.0

1.2
0.8
(MN/m3/2)

(kJ/m2)
0.6
1c

0.8

1c
K

G
0.4

0.4
0.2

0.0 0.0
0 10 20 30 40
PCL-b -PI (wt%)

Figure 10

33
GRAPHICAL ABSTRACT

Epoxy Toughening via Formation of Polyisoprene Nanophases with Amphiphilic


Diblock Copolymer

Yixin Xiang, Sen Xu and Sixun Zheng*

500 nm

34
HIGHLIGHTS
 PCL-b-PI diblocks were synthesized via ring-opening and RAFT polymerizations
 The diblock copolymer is capable of self-assembling in epoxy thermosets
 The nanostructured epoxy containing PCL-b-PI displayed improved fracture
toughness

35

You might also like