You are on page 1of 14

Reinforcement of Adhesion and Development of

Morphology at PolymerPolymer Interface via Reactive


Compatibilization: A Review

Genjie Jiang, Hong Wu, Shaoyun Guo


The State Key Laboratory of Polymer Materials Engineering, Polymer Research Institute of Sichuan University,
Chengdu 610065, China

Polymer blending is a common and effective way to


develop new materials with desirable physical and mechanical properties. Since most polymer pairs are immiscible, reactive compatibilization has been extensively studied to stabilize morphology of polymer
blends and improve their mechanical properties. In the
past several years, considerable interest has been
expressed in understanding the fundamental kinetics
and mechanisms of the interfacial reaction, investigating the reinforcement of the interfacial adhesion and
the development of morphological structure at polymerpolymer interface induced by the interfacial reaction. The present review focused on some theoretical
and experimental results that include the formation
and growth of copolymers at the interface, and also
the major factors such as reaction conditions, the concentration and bulk properties of the functionalized
polymer, the thermodynamic interaction between the
functionalized polymer and the matrix, which can inuence the interfacial adhesion and morphological development. POLYM. ENG. SCI., 50:22732286, 2010. 2010
Society of Plastics Engineers

INTRODUCTION
To prepare polymer materials with desirable properties,
blending two or more polymers have been used in many
aspects. The most important point in designing such materials is that most polymer pairs are thermodynamically
immiscible, which results in poor interfacial adhesion.
Commonly, one way to improve interfacial adhesion
between two immiscible polymers is the addition of a
third component (often a diblock copolymer), which can
connect immiscible polymer molecules through enhanced
entanglement on both side of the interface [110]. HowCorrespondence to: Shaoyun Guo; e-mail: nic7702@scu.edu.cn
Contract grant sponsor: National Natural Science Foundation of China;
contract grant number: 50703027; contract grant sponsor: Special Funds
for Major State Basic Research Projects of China; contract grant number:
2005CB623800.
DOI 10.1002/pen.21686
Published online in Wiley Online Library (wileyonlinelibrary.com).
C 2010 Society of Plastics Engineers
V

POLYMER ENGINEERING AND SCIENCE-2010

ever, it is almost impossible for all the added copolymer


molecules to reach the interface during melt blending.
Therefore, in most cases, a number of the copolymer molecules prefer to form micelles [1113] in the blends rather
than locate at the interface, consequently, the compatibilization efciency is low. Another way to improve the
interfacial adhesion is in situ reactive compatibilization
[1430]. When two polymer chains with reactive functional groups meet at the interface during melt blending,
they can react to form block or graft copolymers, which
can reduce interfacial tension and stabilize the morphology of the blend. The compatibilization efciency of this
strategy is relatively high [31] because the in situ formed
copolymers are pinned at the interface. Given the importance of reactive compatibilization in theoretical elds
and industrial applications, it is desirable to understand
the fundamental factors inuencing the formation of the
copolymers at the interface, the relationship between the
interfacial adhesion and the in situ formed copolymers.
Since Ide and Hasegawa [32] found that the mechanical
properties of nylon 6/polypropylene (PA6/PP) blends were
remarkably improved with increasing the concentration of
maleic anhydride added to the blend in 1974, there have
been many articles published to study the reactive compatibilization in different polymer blend systems theoretically [3338] and experimentally [3950] during melt
mixing, where there exists complex ow such as a combined shear/elongational ow. To exclude the effect of
complex ow on the reactive compatibilization, a planar
interface in layered sandwich [5155] was used as a
model for theoretical studies on the coupling reaction
mechanism of the functionalized polymers in the past decade. The correlative research was mainly concentrated on
the following two aspects. One is how the reaction
kinetics depends on functional group reactivity, the diffusion of the functionalized polymer chains, and the
increased amount of in situ formed copolymers at the
interface. The other is how the interfacial tension, morphology, and adhesion are inuenced by the in situ
formed copolymers. Macosko et al. [56] reviewed recent
research progress in their group on the major factors

REACTION KINETICS AND MECHANISMS

FIG. 1. Schematic of reactive compatibilization between end-functionalized polymer chains A and B to form copolymers at a at interface
between two immiscible polymers.

It is well known that the chemical reactions between


small molecules are usually under diffusion control due to
the reduced mobility of reactive sites. When both the reactive agent and the reaction media are highly entangled
polymeric chains, the reaction kinetics and the mechanism
become even more complicated. Theoretical analyses of
the polymerpolymer interfacial reaction kinetics have
been explored by the simple model system, which is
designed as the irreversible reaction between end-functionalized polymers at a planar interface.
In the past several years, there are two main theories
which were used to depict the reaction kinetics and the
mechanism of the coupling reaction at polymerpolymer
interface: one is diffusion-controlled (DC) models suggested by Fredrickson and coworkers [53, 63, 64], and the
other is reaction-controlled (RC) models established by
Oshaughnessy and coworkers [54, 55, 6567]. For simplicity, it was assumed in both models that the degree of
the polymerization (N) of two polymers is the same, a at
interface was formed between them under quiescent conditions, and reactive groups were located at the chain end
of two polymers. The increase of the number of copolymer chains per unit area (S) with the increasing reaction
time (t) can be expressed as dS/dt k(t)nAnB, where k(t)
is the reaction rate, nA and nB are the content of the reactive chains with function groups A and B, respectively.
Diffusion-Controlled Model

inuencing the interfacial reaction such as the inherent


reactivity of functional polymers [57, 58], thermodynamic
interaction between polymers [59], functional groups location along a chain [60], and the effect of processing ows
[61], however, some interesting experimental results
reported by other groups were not involved in this review.
Creton et al. [62] discussed recent theoretical development of the relationship between macroscopic fracture
toughness and interfacial stress transfer in immiscible
polymer blends, whose interfaces are compatibilized by
the addition of block or random copolymers.
In this article, we review recent theoretical and experimental developments in model polymer interfaces reinforced through in situ reactive compatibilization in the absence of ow. The simple schematic of the reaction is
given in Fig. 1. For simplicity, the functional groups are
all located at the end of polymer chains. When the model
interfaces are annealed under certain conditions, the coupling reaction will occur between two reactive functional
groups of polymer A and B once they meet each other at
the interface. As a result of that, the copolymers are
formed in the interfacial region. The article mainly highlights the interfacial reaction kinetics, the mechanisms for
the formation of the copolymers at the interface, as well
as the relationship between interfacial reaction, macroscopic fracture toughness, and microscopic interfacial
structure.
2274 POLYMER ENGINEERING AND SCIENCE-2010

For DC reaction, the formation of the copolymers at


the interface is controlled by the mass transfer of the
functionalized polymer chains. The reaction rate constant
(k) was described by [53]:
k  DRg 2Rg SV = ln N

(1)

where D, Rg, and N are the diffusion coefcient, radius of


gyration, and degree of polymerization of a reactive
chain, respectively. Sv is the ration of the interfacial area
to total volume fraction. The factor 2RgSV/ln N in Eq. 1
can be interpreted as the volume fraction of interfacial
region of an inhomogeneous system that is accessible to
AB coupling. Reduction of this accessible volume fraction by the factor of 1/ln N  1/ln t1 is associated with
the subdiffusive Rouse dynamics on the time scale that
local equilibrium is achieved inside the interface. It can
be obtained that k scales with molecular weight in the
present Rouse regime as k  1/ln N for unentangled
chains (NhNei). When the polymer chains exceed the
entanglement threshold (N Ne), the reputation time of
chain t  N3 according to tube model of Gennes, resultantly, k  1 (N ln N).
On the basis of the previous theoretical framework,
Fredrickson and Milner [63] proposed that there are three
time regimes to characterize the growth of the copolyDOI 10.1002/pen

mers. The initial growth rate of the copolymers is


described by:
ds=dt K0 r20

(2)

K0 40:7R4g =t lnt=t0

(3)

where s(t) is the number of copolymer chains per area of


interface that have been formed in the time interval t after
initiating the reaction. r0 is the concentration of the reactive
chains with function groups A and B. t0 is a reference time
whose denition depends on the state of entanglement.
In the early stages of the reaction (s  t  sq), when
the density of reactive chains in the interfacial region can
be considered to be the same as the value in the bulk,
coupling reaction is controlled by the quasilocal reaction
rate coefcient.
ri t  r0

(4)

st  K0 r0 2 t

(5)

In the intermediate regime (sq  t  sr), a depletion


hole of reactants in the interfacial region is formed with
the progression of reaction,
ri t  r0 pt=tr 1=4
st  2=p1=2 s t=ts 1=2 2=p1=2 r0 D0 t1=2

(6)
(7)

Therefore, the reaction is controlled by center-of-mass


diffusion of the reactive chains to ll the depletion hole.
Finally, the reaction time corresponds to s sr, the
potential barrier arising from previously formed copolymers at the interface suppressed the reaction in the last regime, therefore, the diffusion of the functionalize chains
across the barrier plays a very important role in the further coupling reaction at the interface. The copolymer
grows slowly as
st  s ln1=2 N 1=2 t=ts lnN

(8)

Therefore, the predicted time dependence is a linear


growth of r(t) in the initial time regime, a t1/2 growth in
the intermediate regime, and a (ln t)1/2 growth in the last
regime, respectively. In most circumstances, only the
crossover from center-of-mass diffusion-limited behavior
(t1/2) to saturation behavior [(ln t)1/2] can be observed as
expected. Furthermore, Muller [68] investigated the rst
two regimes using Mote Carlo simulations and found that
the simulant reaction rate agreed with the theoretical prediction for the intermediate regime of the DC reaction,
however, the simulation indicated a higher reaction rate
than predicted by the theory.
According to the Fredricksons model, if the interfacial
coupling reaction was diffusion-controlled, the functionalDOI 10.1002/pen

ized chains located in the vicinity of the interface would


rst participate into the reaction. As the reaction going on,
more functionalized chains need to diffuse from the bulk to
the interface, then a concentration gradient of the functionalized chains between the interface region and the bulk
would be established. Clarke et al. [69, 70] monitored the
physical adsorption progress of deuterated polystyrene endfunctionalized with carboxylic acid (dPS-COOH) from PS
matrix with various molecular weight to a special oxide
layer of a silicon substrate, which has immobile reactive
sites. It was shown that depletion zones in the concentration
prole of dPS-COOH were up to four times longer than
necessary for the chains to diffuse from surface to interface,
indicating the progress could not be depicted simply in the
term of diffusion model. Norton et al. [71] investigated the
similar interface composed of dPS-COOH chains and functionalized epoxy network, and also found that the extent of
the reaction could be explained more than DC model.
However, Harton et al. [72, 73] provided direct evidence of depletion hole of the functionalized chains at an
interface by studying the coupling reaction between deuterated hydrozy terminated PS (dPS-OH) and methyl
methacrylate and methacrylic acide copolymer (PMMAMAA) at a relatively low temperature of 1208C. In their
experiment, dynamic secondary ion mass spectrometry
(DSIMS) was used to examine the dPS volume fraction
(udPS) at the interface. The real-space DSIMS depth proles of the concentration of dPS-OH at the interface from
0 to 100 h were shown in Fig. 2.
From Fig. 2ae, it could be observed that a depletion
hole formed and grew to a width with 40 nm apparently
from 0 to 24 h, which was a direct evidence of DC coupling reaction at the interface. When the reaction time
was up to 100 h, shown in Fig. 2f, the depletion hole
disappeared completely and the dPS-OH/PS layer approached diffusion equilibrium. On the basis of the previous experimental results, they concluded that the coupling
reaction mechanism at an immiscible polymerpolymer
interface was DC, at least in the early stage.
Therefore, the in situ chemical reaction at the interfaces
is controlled by DC mechanisms or not depending on the
diffusion step of the polymers put in contact and interactions between them. It has been experimentally veried
that the diffusion step of polymer chains at interface of
two polymers is clearly dependent on the structure of the
polymers [74, 75], the relaxation process of chains [76],
the molecular weight and distribution of polymers [77
81], and the composition [82, 83] and compatibility of the
interfaces [84] besides the experimental conditions such as
time [85], temperature [86, 87], and shearing ow [88,
89]. Bousmina et al. [90] also studied the effect of surface
topology and chemistry on the diffusion step by a novel
rheometry technique and theoretical analysis that allow
determination of diffusion coefcients of both polymer
polymer interfaces and bulk polymers. It was revealed that
the diffusion step is strongly dependent on the initial chain
ends distribution at the surface before contact.
POLYMER ENGINEERING AND SCIENCE-2010 2275

FIG. 2. Real-space DSIMS depth proles of the dPS volume fraction


(udPS), including both unreacted dPS-OH within the PS layer and
PMMA-g-dPS located at the interface, are shown as a function of sample
depth for (a) 0 h, (b) 3 h, (c) 6 h, (d) 12 h, (e) 24 h, and (f) 100 h at
1208C. (Reproduced from Ref. 73).

Reaction-Controlled Model
The RC mechanisms for interfacial reactive compatibilization were theoretically depicted in detail by Oshaughnessy and Sawhney [65, 66]. They considered that the
reaction depends on the reactivity of the functional groups
and the stage of the reaction. They predicted that the
reaction kinetics obeys mean eld theory in the reaction
between functional groups with weak reactivity (Q Q*):
kt Qa3 h

(9)

where Q, a, and h were the local functional group reactivity,


the group size, and the interfacial thickness, respectively.
When the functional groups are sufciently reactive
(Q Q*), k(t) would be controlled by the diffusion of the
reactive chains.
For unentangled melts (NhNe)
kt a4 =ta 1=lnR=h  1=ln N

(10)

and for entangled melts (N Ne)


kt a4 =ta 1=N=Ne lnN=Ne   1=N ln N
2276 POLYMER ENGINEERING AND SCIENCE-2010

(11)

where R was the coil size, ta was the relaxation time of a


single chain unit, and Ne was the critical degree of polymerization for entanglements.
Finally, the crowded copolymer chains formed at the
interfaces suppressed k(t) to exponentially small values.
Furthermore, Oshaughnessy and Vavylonis [66] systematically investigated four types of reaction kinetics,
through all time regimes, as a function of the reactivity
and bulk reactive group densities. Three distinct kinetic
sequences which gave rise to different types of reaction
kinetics as a function of reaction time were established.
Each of these three regions involved a different sequence
of reaction kinetics regimes with increasing reaction time.
At short time scale, simple mean eld (MF) kinetics was
applied, the second order rate constant was then independent of time. After this MF phase, there were two possibilities, a direct transition to rst-order kinetics may
occur. However, if the functional groups were very reactive, then a second-order diffusion regime would onset. In
this time regime, it was predicted that S  t/(ln t) for
unentangled chains and S  t/(ln t) during the two t1/4
regimes, whereas S  t1/2 for the t1/8 regime for
entangled chains. In long time regimes, rst-order DC
kinetics would be observed, time dependencies during
these regimes are S  t1/4 for unentangled chains,
whereas for entangled systems, in chronological order S
 t1/4, S  t1/8, and S  t1/4. When the accumulating copolymer ultimately saturated the interface, rst-order DC
kinetics with S  t1/2 were predicted.
Oyama et al. [91, 92] proposed a new reaction model
and kinetic equation for interfacial reaction at polymer
polymer interface by discussing the reaction kinetics and
mechanisms of interfacial reaction between amorphous
PA and polysulfon end-functionalized with phthalic anhydride (PAH), and triazine. They considered that the coupling reaction at polymerpolymer interface was the progression of the formation of a two-dimensional monolayer
at the interface between the polymers. The vacant sites
available for reaction at interface was dened as the parameter (S* 2 S), where S corresponds to the areal density of copolymers formed at time t and S* is a critical
value of the areal density of copolymers which can form
the monolayer at the interface. It was demonstrated that
the coupling reaction at polymerpolymer interface was
similar to the reaction at the gas/solid interface in surface
science, and the reaction followed pseudo-rst-order
kinetics in the parameter (S* 2 S) over the whole time
scale based on the assumption that the coupling reaction
would be terminated at S S*. Except for the systems in
their experiment, the reaction kinetics in published articles
for reactive polymer interfaces such as amorphous styrenemaleic anhydride copolymer/amine-terminated butadieneacrylonitrile copolymer (SMA/ATBA), SMA/PA11 [93],
and carboxyl end-functional polystyrene/precured epoxy
containing excess epoxide groups [94] were all analyzed by
using their model. The results showed that all the experimental data agreed quite well with presudo-rst-order
DOI 10.1002/pen

FIG. 3. Schematic of two sample geometries used in Ref. 95.

kinetics. So, it was suggested that the coupling reaction process between polymers was probably reaction-controlled,
and the spatial restriction at the interface played an important role during the coupling reaction.
Schulze et al. [95, 96] measured the interfacial excess
(Z*/Rg) at three molecular weights of dPS end-functionalized with an amine group (dPS-NH2-22, 237, and 292)
and PMMA end-functionalized with anhydride (PMMAanh) interface using forward recoil spectrometry (FRES).
The measured interface (shown in Fig. 3a) was prepared
by oating lms containing 8.4 wt% dPS-NH2 and PS
onto PMMA-anh layers that had been spun from the solution onto Si wafer. In addition, a special designed trilayer
interface geometry shown in Fig. 3b was used to observe
independently the diffusion of dPS-NH2 chains through
the PS matrix and the reaction with PMMA-anh eventually. The results of the measured growth of Z*/Rg with
time for all three molecular weights could be described
with polymer chain reactivity, rather than the diffusion of
the functionalized chains. Compared with the concentration proles obtained from these two different designed
interface geometries indicated that the dPS-NH2 could diffuse through the PS layer many times before signicant
reaction could be observed under experimental condition.
These results proved that the coupling reaction rate at PSNH2/PMMA-anh interface was not limited by the diffusion of the functionalized polymer chains.
Later, Yin et al. [9799] also probed the interfacial
reaction between PMMA-anh and PS-NH2 with high and
low molecular weight by size exclusion chromatography
with a UV detector (SEC-UV). Reactive polymers were
labeled with anthracene to increase the sensitivity. It was
found that the conversion of PS-NH2 was faster when the
reactive chains were shorter. It likely showed that the
reaction kinetics was DC controlled because the diffusion
coefcient depends on the chain length. However, when
the dependence of the interfacial roughness on the molecular weight of the reactive chains was observed by AFM
and TEM, it was found that the interfacial modication
was slower than the diffusion of the reactive chains. The
same conclusion was obtained by them as well as Schulze
et al. that the interfacial reaction between PS-NH2 and
PMMA-anh was not controlled by diffusion but rather by
the reaction kinetics of the anhydride and amine functional groups.
DOI 10.1002/pen

The activation energy of the reaction is another clue,


which can be used to judge whether the coupling reaction
kinetics at polymer interface are controlled by diffusion
or reactivity. It is believed that the DC reactions usually
have low activation energy (\30 kJ/mol) because it is a
physical process, whereas the RC mechanisms have
higher activation energy attributed to a chemical process.
Oyama and Inoue [91] estimated the activation energy for
the reaction at SMA/ATBA interface was 120 kJ/mol,
indicating that the coupling reaction at this interface was
reaction-controlled. Kim et al. [100] investigated the
effect of surface functionalization on the interfacial adhesion between PP and PA6. The functional groups of PP
surface modied by low-energy ion-beam irradiation
method are concentrated on a very shallow surface layer
), which makes the diffusion of reactive
(less than 70 A
chains, can be neglected signicantly. The interfacial
reaction kinetics in this system can be considered as reaction-controlled. The calculated activation energy was 97.8
kJ/mol, which was also in favor of reaction-controlled
mechanisms. Kim et al. [101, 102] studied the effect of
molecular weight of dPS-NH2 and reaction temperature
on the reaction kinetics at PS/poly(2-vinylpyridine)
(P2VP) interface. It was experimentally showed that the
activation energy of the reaction was 156 6 7 kJ/mol,
and the value was not affected by the molecular weight
of dPS-NH2. Jiao [103] measured the activation energy of
the reaction between dPS-NH2 and PSMA interface, and
the value was 207 kJ/mol. For the reaction between the
same functional groups, the experimental values of the
activation energy were not the same, but all the experimental results suggested that the interfacial reaction
kinetics was controlled by the reactivity of the functional
groups.
Interfacial Adhesion (Gc)
When two immiscible polymer chains are connected
by the covalent bond of copolymers formed at the interface, the interfacial adhesion between two immiscible
phases will be enhanced. In the case of planar interfaces
enhanced via reactive compatibilization, the interfacial adhesion is characterized by the fracture toughness (Gc) of
that. It has been shown that an asymmetric double cantilever beam (ADCB) test is a reliable test for measuring Gc
of polymerpolymer interfaces [104106]. It involves
propagating a crack along the interface between two polymer beams using a wedge. The specic conguration of
the test is shown in Fig. 4. The wedge, typically a razor
blade, is inserted at the interface to induce a crack. The
elasticity of the two polymer beams enhances the stress at
the crack tip, causing the crack to be advanced some distance ahead of the wedge. Adhesion acts to suppress
crack propagation, counteracting the elasticity. Therefore,
the distance of the crack advances ahead of the wedge is
an indication of the adhesion. Generally, constant rate
mode is introduced, in the method, the razor blade of
POLYMER ENGINEERING AND SCIENCE-2010 2277

FIG. 4. The schematic drawing of the geometry used in the ADCB


test.

width D is driven at a constant rate of 3 lm/s, and the


crack length (a) at the interface is continuously monitored. Gc is then calculated according to the following
equation [107, 108]
"
#
3D2 E1 h1 3 E2 h2 3 E1 h1 3 C2 2 E2 h2 3 C1 2
Gc
8a4
E1 h1 3 C2 3 E2 h2 3 C1 3 2

(12)

where Ei and hi denote the Youngs modulus and the


thickness of material i, D is the thickness of the razor
blade, a is the crack length ahead of the blade, Ci is the
correctional factor for material i, which is expressed as
follows:
Ci 1 1:92hi =a 1:22hi =a
0:39hi =a=1 0:64hi =a

13

On the basis of the above measured method and other


analysis techniques, the effect of reaction conditions and
the bulk properties of polymers on the Gc of model reactive compatibilization interfaces were extensively studied,
including reaction time and temperature, the crystalline
structure near the interface, molecular weight, and architecture of the formed copolymers, as well as the relationship between Gc and the areal density of copolymers at
the interface (S).

hydride grafted PP was blended with PP for reactive compatibilization. As shown in Fig. 5, Gc increased with the
annealing time and then reached a level of saturation for
all annealing temperature. Many other researchers
obtained similar experimental results in other experimental systems [111113]. When the annealing temperature
was a litter higher than the melt temperature of PA6, i.e.,
2238C, the rate and the level of saturation of Gc increased
dramatically. Laurens et al. [114] also studied the effect
of annealing temperature on Gc of PP/PA6 by using functionalized PP (PPf) with different molecular weight as the
compatibilizer. It was found that annealing the samples
above the PA6 melting temperature increased strongly the
adhesion in samples containing high molecular weight
PPf, while such an increase was not observed in other
samples. Therefore, they believed that a sufcient molecular weight of functionalized polymer chains should be a
necessary condition to obtain a dramatic increase in interfacial adhesion at high reaction temperature.
However, some other researchers observed a maximum
in the saturation Gc at a certain reaction temperature when
they studied the similar experimental systems. Seo and
Ninh [115] found Gc of PP/PA6 compatibilized with
1 wt% maleic anhydride grafted PP showed a maximum
around 2208C and then decreased at higher annealing
temperature. Cho and Li [116] also observed an unexpected maximum in the saturation Gc of PP and amorphous PA6 interface at around 1708C, which was ascribed
to a different interfacial morphology around this temperature. Kim et al. [100] obtained a maximum saturation
value of Gc at 2008C when they studied the effect of surface functionalization by low-energy ion-beam irradiation
on the enhancement of interfacial adhesion between PP
and PA6. Since the PP molecules were all in melt state,

Effect of Reaction Time and Temperature on Gc


Theoretically, when the reaction temperature is higher
than the glass translation temperature of two polymers,
interfacial reaction can occur since the polymer chains
near the interface can diffuse each other. But very long
time is needed to form effective adhesion at low reaction
temperature. When the reaction temperature is above the
melt temperature of both polymers, Gc will increase rapidly within rst several minutes. To control the variation
in Gc over wide order of magnitude as a function of suitable reaction time, the reaction temperature is usually chosen between the melt temperatures of two polymers near
the interface.
Boucher et al. [109, 110] investigated Gc of PP/PA6
interface annealed at the temperature between 1858C and
2238C as a function of the annealing time, and 5 wt% an2278 POLYMER ENGINEERING AND SCIENCE-2010

FIG. 5. Gc versus annealing time for PP/PA6 interface at four annealing temperatures. (Reproduced from Ref. 109).

DOI 10.1002/pen

the inuence of the PP crystallization on the Gc could be


eliminated. They argued that the appearance of a maximum at 2008C was attributed to the change of fracture
mechanisms at the interface.
Effect of Functionalized Chains Content on Gc
In general, 5 wt% functional polymer is enough for reactive compatibilization in immiscible polymer systems
prepared by melt blending, because new reactive interface
can be formed continually due to shear effect. But the reactive interface with model planar geometry is constant,
whether more functionalized polymer or not need to be
added for reactive compatibilization in this system. Cho
and Li [116] investigated the effect of maleic anhydride
grafted PP (mPP) content on the interfacial fracture
toughness of amorphous PA6/PP at the same bonding
temperature. It was shown that Gc could not be enhanced
when the amount of mPP was lower than 0.5 wt%, even
at a bonding time of 2 h. As the mPP content increased,
Gc increased obviously which was ascribed to the in situ
formation of copolymers at the interface. When the mPP
concentration reached 3 wt%, Gc became slightly leveled
off. Gc of the samples as a function of bonding time for 3
and 5 wt% mPP were nearly overlapped, indicating 3
wt% mPP were enough to compatibilize this system effectively.
Koriyama et al. [117] examined the interfacial thickness and adhesion of the reactive polysulfon (PSU) and
amorphous PA6 interfaces as a function of MAH group
content in PSU. It was observed that the interfacial thickness was almost independent of the MAH group content,
especially when the MAH group content was more than
0.56 wt%. According to their results, the interfacial adhesion increased with increasing interfacial thickness following the relationship, Gc  k2. So, they believed that
MAH group content of 0.2 wt% was sufcient to form a
thick interface, thus, efciency interfacial reinforcement.
Effect of Molecular Architecture of Functionalized
Polymer on Gc
When the miscibility between functionalized polymer
and the matrix is really poor, it will induce a lack of
entanglements or cocrystallization between the in situ
formed copolymers and the matrix. Therefore, it is not efcient to transfer stress across the interface during fracture test. On the contrary, when the functionalized polymer has a good compatibility with the matrix, the block
of the copolymers can entangle or cocrystallize with the
matrix chains well, leading to an effective reinforcement
of interface by developing a plastic deformation zones at
the craze tip. The molecular weight of the functionalized
polymer also inuences the entanglement between the
copolymers and the matrix near the interfaces. Laurens
[118] synthesized a series of succinic anhydride functionalized PP with different molecular architectures (isotaticDOI 10.1002/pen

PP, isotatic-PP with 5 wt% PE, syndiotactic-PP, metallocene-PP) and discussed the role of the architecture of PP
block of the in situ formed copolymers on the interfacial
fracture toughness enhancement for isotactic-PP/PA6
interface. It was found that the kinetics of fracture toughness enhancement of samples annealing at the same conditions was strongly inuenced by the miscibility of functionalized PP (PPf) with the matrix PP. When the PPf was
isotatic-PP, Gc of i-PP/PA6 interface increased rapidly
and then reached saturation value with annealing time due
to its good compatibility with the matrix. When the partially compatible PPf (PE-PPf) was incorporated, the
extent of the increase in Gc became less obvious and a
plateau was not always observed. In addition, the molecular weight of PPf also inuenced Gc, which was shown
that the maximum Gc was larger for the high molecular
weight PPf. They proposed that long copolymer chains
were more efcient than short copolymer ones to reinforce the interface, which was attributed to the fact that
long copolymer chains can link several lamellae in the
vicinity of the interface. Eastwood and Dadmun [119]
compared the ability of a series of styrene and methyl
methacrylate copolymers with varying architectures to
compatibilize PS/PMMA interface. It was demonstrated
that the ability of reinforcement of the interface decreased
in the order pentablock i triblock i diblock i heptablocks i
random copolymers, and the multiblock copolymers
except heptablocks provided good interfacial adhesion,
which was attributed to multiple interface crossings
between blocks of monomer and homopolymers. To form
adequate entanglements with homopolymer, a critical molecular weight was required for block length. This was
exactly the reason that why the heptablock copolymers
give the relatively weak interface.
Besides the aforementioned factors, The enhanced adhesion between immiscible polymer interfaces through in
situ reactive compatibilization depends on many other parameters, such as cooling rate [116], chains orientation
[120], and crystalline structure at the interface [121],
especially for semicrystalline polymer interfaces.
Relationship Between Gc and R
It is evident that the covalent bond connection at the
interface will build up with the increasing S, thus, Gc
increases as the progress of interfacial reaction. The
changes of Gc are strongly depended on the fracture
mechanisms at the interface. To understand which mechanisms are responsible for the interface reinforcement, it is
useful to examine the relationship between Gc and S. In
the case of glassy polymer interfaces [122, 123], two different relationships between Gc and S have been
observed: when the fracture mechanisms of the interface
are simple chain scission or chain pull-out without any
plastic deformation, Gc is found to vary as S linearly;
when crazing and plastics deformation are initiated at the
interface, then Gc is observed to be scaled with S2. For
POLYMER ENGINEERING AND SCIENCE-2010 2279

conditions but also on the molecular weight of functionalized polymers and the miscibility between the functionalized polymer and the matrix. The structure of the in situ
formed copolymers and the crystalline and orientation
behavior near the interfaces should also be considered.
Because of the complicated inuencing factors, a consistent and universal model has not been established to
describe the relationship between Gc and S for reactive
compatibilized semicrystalline polymer interface.
Interfacial Morphological Development by Reaction

FIG. 6. Gc of PP/PA6 interface versus S in a loglog plot at four


annealing temperatures. (Reproduced from Ref. 110).

semicrystalline polymer interfaces, the relationship


between Gc and S is more complicated because the crystalline structure at the interface has profound effects
[124127] on the plastic deformation properties of the
polymers. Boucher et al. [109] rst established a theoretical model to correlate Gc with S at semicrystalline PA6/
PP interfaces. By the combination of the results of ADCB
test and XPS analysis, they were surprised to obtain
that all date plotted on loglog scale in agreement Gc !
S2.03 6 0.18 scaling, which was similar to that observed
for glassy polymer interface in the crazing regime predicted by Brown [104].
But in further work by Boucher [110], the relationship
between Gc and S was not t to the above model when
PP/PA6 interface was annealed a litter above the melt
temperature of PA6, especially when high molecular
weight PPf was incorporated as compatibilizer. As shown
in Fig. 6, when the annealing temperature was below
2208C, it was likely that the relation between Gc and S
was not inuenced by the conditions for preparing the
samples, and the interfacial fracture mechanisms were
similar for all values of S. However, it was not right in
the case of the samples prepared at a temperature of
2238C. According to Bouchers further investigation, the
change in fracture mechanisms for sample with high molecular weight made at 2238C was related to the fact that
the b-form PP crystalline structure was developed at the
interface. It implied that the interfacial fracture mechanisms were not only inuenced by the copolymers formed
but also by the crystalline structure near the interfaces.
Lauren [114, 118] reported that the relationships between
Gc and S were not consistent with Bouchers model in
some experimental systems. He proposed that the relationship between Gc and S depends not only on annealing
2280 POLYMER ENGINEERING AND SCIENCE-2010

Some investigation has shown that the copolymers


generated at the interface can form polymeric surfactants,
thus, decrease the interfacial tension between the incompatible polymer phases and result in a signicant diminution in the particle size of dispersed phase [128131].
Many studies have found that the nal morphology of
polymer blends with an in situ reactive compatibilizer
depends on the bulk properties of the components, blend
composition, the amount and the molecular weight of the
formed copolymer, viscosity ratio between different
phases, and the processing parameters during melt blend.
Among them, the shear force is a very important factor
because it can induce droplet breakup and bring continual
new reactive interfaces. For at polymer interface system,
discussed in this article, there is no shear effect during
reaction compatibilization. How does the morphology develop in this kind of experimental condition is another
interesting topic.
From the theoretical studies [63, 65] concerning the
kinetics of coupling reaction at a melt polymer interface,
a free energy barrier (l*/kBT) to the coupling reaction
will develop due to the entropy loss involved in stretching
the brush of grafted chains as the ratio of interface
excess of the grafting chain to its radius of gyration (Z*/
Rg) increases. The buildup of such a barrier during the
reaction will suppress the coupling reaction and ultimately
limit the Z*/Rg achievable within experimental time.
Oshaughnessy and Sawhney [65] asserted that the reaction rates will slow down to near zero before the reaction
can form sufcient copolymers at the melt polymer interface to effectively diminish the interfacial tension. However, Jiao et al. [132] demonstrated that it is possible to
make interfacial tension at polymer interface decrease signicantly by reactive compatibilization for short functionalized chains, and the decreases of the interfacial tension
is large enough to cause polymer interface instability,
which leads to interfacial roughness and eventually the
formation of microemulsions. A theoretical model correlated to 2Dc/c0 and Z*/Rg was developed to predict the
transition of the interface from at to corrugated, where
the interfacial tension becomes to zero.
On the basis of the self consistent mean eld calculations of Shull [133], the decrease in interfacial tension
(2Dc) due to the increased interfacial excess of copolymer chains can be written in the following equation:
DOI 10.1002/pen

 
Dg
r0 a
Z
p

a
kB T
R
6N
g

(14)

where r0 is the monomer number density, a is the statistical segment length or polymer, a is a function of Z*/Rg
tabulated by Shull for the limit where the unreactive matrix chains are much longer than the end-functional
chains.
For the high molecular weight polymers without graft
copolymers, the initial interfacial tension (c0) is given by
the Helfand and Tagami results [134],
g0
r0 a
kB T

r
w
6

(15)

where w is the FloryHuggins interaction parameter.


So, the ratio of the decrease in interfacial tension to
the initial interfacial tension can be expressed as
 
Dg
1
Z
p a
g0
wN Rg

(16)

if w and a are known, the model depicted by the equation


can be used to predict the Dg/g0 versus Z*/Rg for functionalized chains with different molecular weight. In
Jiaos work [132], the critical value of Z*/Rg (Z*c /Rg) for
vanishing interfacial tension of PS-NH2/PSMA were
investigated based on a self-consistent mean eld theory.
For reactive PS-NH2/PSMA interface, it was found that
Z*/Rg decreased with decreasing N of PS-NH2, and the
theoretical predication for two dPS-NH2 chains were Z*c /
Rg 1.5, 2.4 for N 55 and 270 chains, respectively.
On the other hand, the onset of the interface instability
occurred at Z*c /Rg of about 1.8 for N 55 and 2.5 for N
270 PS-NH2 chains by determining the root-meansquare (RMS) surface roughness, as shown in Fig. 7. The
experimental results of transition value of Z*c /Rg are in
qualitatively agreement with theoretical predication [135].
Some recent experimental results [98, 102, 136] also
proved that the decrease of interfacial tension due to reactive compatibilization can result in the interfacial roughening or interfacial emulsication at a planar polymer
interfaces in the absence of ow. The development of
morphology is closely related to the reaction conditions,
the concentration, and molecular weight of reactive functionalized polymer.
Lyu et al. [137, 138] examined the morphological development at layered reactive PS-NH2/PMMA-anh interface as a function of annealing time by AFM and TEM
observation. According to TEM results, the interface
between PS and PMMA was at before annealing, shown
in Fig. 8a. The interface became quite rough after annealing the sample for only 20 min, and some parts of the PS
domain appeared to have pinched off at the interface and
moved into PMMA phase, as shown in Fig. 8b. When the
annealing time increased to 1 h, the interfacial roughness
increased further and the magnitude of the width of the
DOI 10.1002/pen

FIG. 7. RMS roughness of the PS/PSMA interface after washing with


cyclohexane measured by SFM as a function of Zc /Rg. N 55 (&) and
N 270 (*). (Reproduced from Refs. 132 and 135).

roughening zone was about 0.5 lm shown in Fig. 8c. The


interfacial roughening during annealing was also observed
by AFM. It was observed that the magnitude of the width
of the roughening zone increased from 0 to 0.2 lm when
the sample was annealed for 1 h.
Zhang et al. [139] studied the relationship between the
interfacial roughening process and the extent of coupling
reaction at PS/PMMA interfaces. The coupling reaction
was controlled by varying the concentration of PS-NH2 in
an unreactive PS matrix, while maintaining the same pure
PMMA-anh layer. It was obtained that interfacial roughness was strongly inuenced by the concentration of PSNH2 in the PS layer. For 10 wt% PS-NH2 sample, the
RMS roughness increased to about 1.4 nm in the rst
5 min and then remained the same value within 120 min.
When the concentration of PS-NH2 was high, take 75
wt% for example, the RMS roughness increased to above
10 nm dramatically in about 10 min and then increased to
above 15 nm very slowly over 120 min. There was an abrupt increase in RMS roughness when the PS-NH2 concentration changed from 25 to 30 wt% within the same
annealing time of 1 h at 1758C, just shown in Fig. 9. To
explain the development of interfacial roughness, the
maximum interfacial coverage (S*) was dened. They got
a conclusion that when the concentrations of PS-NH2 was
low, S \ S*, the interfacial roughness increased with
reaction conversion but remained low. When the concentration was increased to 30 wt%, S [ S*, the interfacial
roughness increased dramatically, and interfacial emulsication phenomena happened, which was proved by TEM
observation of the interfacial morphology.
The length of the functionalized chains, i.e., the molecular weight also inuence the interfacial roughness [98, 102].
POLYMER ENGINEERING AND SCIENCE-2010 2281

FIG. 9. RMS roughness versus PS-NH2 concentrations after 1 h annealing of PS-NH2/PMMA-anh-pyr. (Reproduced from Ref. 139).

On the one hand, the short reactive chains can diffuse to


interface faster than long ones, on the other hand, when
long functionalized chains are introduced, the high molecular weight copolymers formed at the interface do not
leave the interface, and suppress the reaction rate to zero,
whereas, the copolymers formed by short chains can diffuse across the interface, which is favorable to the reaction progress and interfacial roughness.
When reactive compatibilized polymer blends are prepared in the presence of shear (or elongational) effect,
micelles are usually formed in the matrix phase depending
strongly on the amount of in situ formed copolymers as
well as the molecular structure of those [140143]. In the
Jiaos and Zhangs research, the interfacial roughness
both remain nearly constant after the transition point
though the extent of reaction increases. They attributed this
to the fact that the entire emulsied region (copolymers
coated droplets) formed after the transition point was
removed by the selective solvent during the preparation of
sample for AFM observation. Direct TEM cross-sectional
imaging of the interface supported their assumption. Kim
and coworkers [144, 145] rst reported the formation of
microemulsions at reactive interface of PS-mCOOH/
PMMA-GMA in the absence of shear when the samples
were annealed at 1808C for 17 h. A simple illustration
given in Fig. 10 was used to explain the interfacial morphological change with reaction stages and the formation

FIG. 8. Representative morphologies of PS-NH2/PMMA-ah interface


after static reaction of (a) 0, (b) 20, and (c) 60 min. (Reproduced from
Ref. 137).

2282 POLYMER ENGINEERING AND SCIENCE-2010

FIG. 10. Schematic describing variations of interfacial morphology development for (PS-mCOOH)/(PMMA-GMA) bilayer at 1808C. (Reproduced from Ref. 145).
DOI 10.1002/pen

progress of microemulsions in the studied system. Three


distinct stages were considered: in Stage I, the functionalized chains reacted each other mainly near the interface and
the interface was not roughened. When the in situ formed
copolymers were covered at least a single layer at the interface (Stage II), the functionalized chains diffuse into the
brush like copolymer layer, the interface became corrugated. At the beginning of the Stage III, the reactant chains
diffuse through the brush like copolymer layer, the interface became more corrugated with the process of reaction.
When t [ tmicroemulsion, the copolymers began to pinch off
and then PS chains were encapsulated by PMMA chains,
which nally became microemulsions. They veried the
hypothetic process by TEM observation [144].
They suggested that the microemulsions structure in
the planar interface depended on the reaction time and the
molecular weight of the functionalized chains. Longer
reaction time and lower molecular weight more likely to
form microemulsions structure. Since the above experiment was performed in plate rheometer, a very small oscillatory shear force was applied on reactive bilayer sample during reaction, it might be argued that whether the
microemulsion formation was induced by this very small
shear effect. According to the experimental results of
Kim, the morphological development near the interface of
the (PS-mCOOH)/(PMMA-GMA) bilayer sample annealing in the presence of oscillatory shearing used in their
rheological measurements was essentially the same as that
in the absence of oscillatory shearing. Microemulsions
were also observed in the bilayer sample annealing at
1708C under quiescent conditions.
In a word, the copolymers formed at the planar interface is expected to decrease the interfacial tension of this
polymer interface. Instability and roughening of the interface is observed when the interfacial tension decreases to
a critical value. The extent of interfacial roughness
depends on the concentration and molecular weight of
functionalized chains and reaction conditions. There exists
a critical values for Z*/Rg and the concentration of the
functionalized chains, where the interfacial tension vanishes and the interfacial morphology changes dramatically
from at to corrugated one.
CONCLUSIONS
This article has intended to provide a summary of the
recent developments in the eld of interfacial reactive
compatibilization for simple model systems, which are
assembled by bilayer or trilayer polymer lms or sheets
with well-dened interfaces. Such systems enable one to
determine the kinetics parameters and mechanisms of the
reaction, to study the effect of the in situ formed copolymers on the enhanced interfacial adhesion, to observe the
interfacial morphological development induced by reactive compatibilization, etc. Since the interfacial reactive
compatibilization is performed at a circumstance excluding the complex effect of shear and elongational ows, it
DOI 10.1002/pen

is possible to correlate the progress of interfacial reactive


compatibilization directly with some factors such as processing conditions, the bulk properties of polymer matrix,
and functionalized chains. The following achievements
related to different aspects are worth mentioning.
Though the question related to the reaction kinetics at
the interfaces are controlled by the reactivity or diffusion
of functionalized chains is still in discussion theoretically
and experimentally, the reactivity and diffusion of the
functional chains at the interfaces play an important role
in the reaction kinetics.
The measured fracture toughness of reinforced interfaces is crucially dependent on the details of annealing conditions, the content and molecular weight of the functionalized chains, and the miscibility between the formed
copolymers and matrix. The reinforcing mechanisms can
be depicted by the Browns model in some systems. However, in the case of some semicrystalline polymer interfaces prepared under some conditions, the relationship
between Gc and S is deviated from the theoretical model,
which is attributed to the morphology and the crystallinity
near the enhanced interface.
With the progress of interfacial reaction, the interfacial
tension at the interfaces will decrease to negative value at
a certain critical grafting density, which is related to the
content and molecular weight of the functionalized chains,
thus, leads to the transition from at interface to the corrugated one. Possibility of the formation of the microemulsions at reactive interface in the absence of ow is
found in some cases.
Important theoretical and experimental developments
in reactive compatibilization have been achieved by
studying the coupling reaction at simple polymer interface, but some important problems related to that are still
worth paying more attention. More accurate theoretical
model should be established to describe the kinetics and
mechanisms of coupling reaction at polymerpolymer
interfaces, and to predict the interfacial adhesion due to
the formation of copolymers; Reliable analysis techniques
are necessary to be developed to characterize the crystalline structure and morphology, and to understand the reinforced effect between semicrystalline polymer interfaces
clearly. Given the importance of shear effect in commercial reactive processing, another important aspect is the
problem concerning the effect of shear force applied on
the reactive polymer systems. It was reported that Kim
[145] found that high strain and high frequencies inhibit
the interdiffusion process. Other studies [146, 147]
showed that large strains may generate high density of
chain ends, which in turn enhance the interdiffusion process. Until now, the links between the shear ow and the
interdiffusion process or interfacial reaction are still a
controversial point. The interfacial morphology in coextruded multilayer is well dened by number of layers and
layer thickness. The large number of layers offers large
amount of specic reactive area, and the shear effect can
be controlled by the rate of extrusion, which makes it a
POLYMER ENGINEERING AND SCIENCE-2010 2283

suitable model system to study interfacial reaction [148,


149]. According to one of our recent work [150], as the
layer number of PE/PE-MAH/PA6 microlayer samples
increased, the areal density of the formation of copolymers at the interface increased obviously because of the
stronger and more durable shearing and elongational
forces [151] on the melts owing through the laminatingmultiplying elements. This caused that the interfacial adhesion between PA6 and HDPE layers was improved.
Further deep study was in progress. The ultimate aim of
the work is a creation of the theoretical and experimental
results that would enable one to control the reactive processing as a powerful way to produce polymer composites
with desire properties.
REFERENCES
1. L.A. Utracki, Polym. Eng. Sci., 22, 1166 (1982).
2. S.B. Chun and C.D. Han, Macromolecules, 33, 3409 (2000).
3. R. Fayt, R. Jerome, and Ph. Teyssie, J. Polym. Sci. Polym.
Phys. Ed., 7, 775 (1989).
4. M.F. Daz, S.E. Barbosa, and N.J. Capiati, Polym. Eng.
Sci., 46, 329 (2006).
5. C. Auschra, R. Stadler, and I.G. Voigt-Martin, Polymer,
34, 2081 (1993).
6. C. Harrats, R. Fayt, and R. Jerome, Polymer, 43, 5347
(2002).
7. C.A. Fleischer, A.R. Marales, and J.T. Koberstein, Macromolecules, 27, 379 (1994).
8. H.R. Brown, K. Char, V.R. Deline, and P.F. Green, Macromolecules, 26, 4155 (1993).
9. C. Scott and C.W. Macosko, J. Polym. Sci. Part B: Polym.
Phys., 32, 205 (1994).
10. H. Ruckdaschel, J.K.W. Sandler, V. Altstadt, H. Schmalz,
V. Abetz, and A.H.E. Muller, Polymer, 48, 2700 (2007).
11. S. Xu, H. Zhao, T. Tang, L. Dong, and B. Huang, Polymer,
40, 1537 (1999).
12. S.H. Anastasiadis, I. Gancarz, and J.T. Koberstein, Macromolecules, 22, 1449 (1989).
13. C. Harrats, K. Dedecker, G. Groeninckx, and R. Jerome,
Macromol. Symp., 198, 183 (2003).
14. N.C. Liu and W.E. Baker, Adv. Polym. Technol., 11, 249
(1992).
15. S. Thomas and G. Groeninckx, Polymer, 40, 5799 (1999).
16. Y.J. Sun, G.H. Hu, M. Lambla, and H.K. Kotlar, Polymer,
37, 4119 (1996).
17. H. Cartier and G.H. Hu, Polymer, 42, 8807 (2001).
18. P.C. Lee, W.F. Kuo, and F.C. Chang, Polymer, 35, 5641
(1994).
19. Y. Son and R.A. Weiss, Polym. Eng. Sci., 42, 1322 (2002).
20. C. Jiang, S. Filippi, and P. Magagnini, Polymer, 44, 2411
(2003).
21. C. Harrats, T. Omonov, G. Groeninckx, and P. Moldenaers,
Polymer, 45, 8115 (2004).
22. S.M. Hong, S.S. Hwang, Y. Seo, I.J. Chung, and K.U.
Kim, Polym. Eng. Sci., 37, 646 (1997).

2284 POLYMER ENGINEERING AND SCIENCE-2010

23. C.C. Su and C.K. Shih, Polym. Eng. Sci., 46, 337 (2006).
24. H.J. Kim, K.J. Lee, Y. Seo, S. Kwak, and S.K. Koh, Macromolecules, 34, 2546 (2001).
25. H.K. Jeon and J.K. Kim, Macromolecules, 31, 9273
(1998).
26. J.S. Oh, A.I. Isayev, and M.A. Rogunova, Polymer, 44,
2337 (2003).
27. W. Feng and A.I. Isayev, Polymer, 45, 1207 (2004).
28. A.I. Isayev and C.K. Hong, Polym. Eng. Sci., 43, 91 (2003).
29. J. Li, S.Y. Guo, R. Slezak, and B. Hausnerova, Macromol.
Chem. Phys., 206, 2429 (2005).
30. G.S. Chen, S.Y. Guo, and H.L. Li, J. Appl. Polym. Sci.,
86, 23 (2002).
31. H.K. Jeon, J. Zhang, and C.W. Macosko, Polymer, 46,
12422 (2005).
32. F. Ide and A. Hasegawa, J. Appl. Polym. Sci., 18, 963
(1974).
33. C. Komalan, K.E. George, S. Jacob, and S. Thomas,
Polym. Adv. Technol., 19, 351 (2008).
34. K. Dedecker, G. Groeninckx, and T. Inoue, Polymer, 39,
5001 (1998).
35. C. Sailer and U.A. Handge, Macromolecules, 40, 2019
(2007).
36. C.H. Yeung and A. Kimberly, Macromolecules, 36, 229
(2003).
37. E. Riemann, H.J. Cantow, and C. Friedrich, Macromolecules, 30, 5476 (1997).
38. D. Shi, Z. Ke, J. Yang, Y. Gao, J. Wu, and J. Yin, Macromolecules, 35, 8005 (2002).
39. S.S. Dagli, M. Xanthos, and J.A. Biesenberger, Polym.
Eng. Sci., 34, 1720 (1994).
40. P. Guegan, C.W. Macosko, T. Ishizone, A. Hirao, and S.
Nakahama, Macromolecules, 27, 4993 (1994).
41. J.J. Benkoski, P. Flores, and E.J. Kramer, Macromolecules,
36, 3289 (2003).
42. Y. Lee and K. Char, Macromolecules, 31, 7091 (1998).
43. B.L. Sauvarop, S. Jareerat, A. Taweechai, and M. Wiriya,
Polym. Eng. Sci., 39, 312 (1999).
44. E.A. Kang, J.H. Kim, C.K. Kim, S.Y. Oh, and H.W. Rhee,
Polym. Eng. Sci., 40, 2374 (2000).
45. S. Jose, B. Francis, S. Thomas, and J. Karger-Kocsis, Polymer, 47, 3874 (2006).
46. H.J. Zhou, M.Z. Rong, M.Q. Zhang, W.H. Ruan, and K.
Friedrich, Polym. Eng. Sci., 47, 499 (2007).
47. C. Sailer and U.A. Handge, Macromolecules, 41, 4258
(2008).
48. C. Sailer and U.A. Handge, Macromol. Symp., 254, 217
(2007).
49. N.C. Beck Tan, S.K. Tai, and R.M. Briber, Polymer, 37,
3509 (1996).
50. H. Asthana and K. Jayaraman, Macromolecules, 32, 3412
(1999).
51. M. Bousmina, H. Qiu, M. Grmela, and J.E. Klemberg-Sapieha,
Macromolecules, 31, 8273 (1998).
52. H. Qiu and M. Bousmina, Macromolecules, 33, 6588
(2000).

DOI 10.1002/pen

53. G.H. Fredrickson, Phys. Rev. Lett., 76, 3440 (1996).


54. B. Oshaughnessy and U. Sawhney, Phys. Rev. Lett., 76,
3444 (1996).
55. B. Oshaughnessy and D. Vavylonis, Europhys. Lett., 45,
638 (1999).
56. C.W. Macosko, H.K. Jeon, and T.R. Hoye, Prog. Polym.
Sci., 30, 939 (2005).
57. N.C. Liu, H.Q. Xie, and W.E. Baker, Polymer, 34, 4680
(1993).
58. C.A. Orr, J.J. Cernohous, P. Guegan, A. Hirao, H.K. Jeon,
and C.W. Macosko, Polymer, 42, 8171 (2001).
59. T.D. Jones, J.S. Schulze, C.W. Macosko, and T.P. Lodge,
Macromolecules, 36, 7212 (2003).
60. H.K. Jeon, C.W. Macosko, B. Moon, T.R. Hoye, and Z.
Yin, Macromolecules, 37, 2563 (2004).
61. P.J. Cole, PolymerPolymer Adhesion in Melt-processed
Multilayered Structures, Ph.D. Thesis, University of
Minnesota (2000).
62. C. Creton, E.J. Kramer, H.R. Brown, and C.Y. Hui, Adv.
Polym. Sci., 156, 53 (2001).
63. G.H. Fredrickson and S.T. Milner, Macromolecules, 29,
7386 (1996).
64. G.H. Fredrickson and L. Leibler, Macromolecules, 29,
2674 (1996).
65. B. Oshaughnessy and U. Sawhney, Macromolecules, 29,
7230 (1996).
66. B. Oshaughnessy and D. Vavylonis, Macromolecules, 32,
1785 (1999).
67. B. Oshaughnessy and D. Vavylonis, Phys. Rev. Lett., 84,
3193 (2000).
68. M. Muller, Macromolecules, 30, 6353 (1997).
69. C.J. Clarke, R.A.L. Jones, and A.S. Clough, Polymer, 37,
3813 (1996).
70. C.J. Clarke, R.A.L. Jones, J.L. Edwards, A.S. Clough, and
J. Penfold, Polymer, 35, 4065 (1994).
71. L.J. Norton, V. Smigolova, M.U. Pralle, A. Hubenko, K.H.
Dai, E.J. Kramer, S. Hahn, C. Berglund, and B. DeKoven,
Macromolecules, 28, 1999 (1995).
72. S.E. Harton, F.A. Stevie, R.J. Spontak, T. Koga, M.H.
Rafailovich, J.C. Sokolov, and H. Ade, Polymer, 46, 10173
(2005).
73. S.E. Harton, F.A. Stevie, and H. Ade, Macromolecules, 38,
3543 (2005).
74. J.E. Martin and D. Adolf, Macromolecules, 22, 4309 (1989).
75. K.R. Shull, E.J. Kramer, G. Hadziioan, M. Antonietti, and
H. Sillescu, Polym. Prepr. (Am. Chem. Soc. Div. Polym.
Chem.), 30, 65 (1989).
76. E. Jabbari and N.A. Peppas, Macromolecules, 28, 6229 (1995).
77. R.J. Composto, E.J. Kramer, and D.M. White, Macromolecules, 25, 4167 (1992).
78. W.W. Graessley, Adv. Polym. Sci., 47, 67 (1982).
79. E. Jabbari and N.A. Peppas, J. Mater. Sci., 29, 3969
(1994).
80. H. Qiu and M. Bousmina, CJChE, 80, 1206 (2002).
81. R. Levesque, S. Vaudreuil, M. Bousmina, and S. Kaliaguine, J. Polym. Eng., 25, 501 (2005).

DOI 10.1002/pen

82. J. Kanetakis and G. Fytas, Macromolecules, 22, 3452


(1989).
83. E. Kim, E.J. Kramer, J.O. Osby, and D.J. Walsh, J. Polym.
Sci. Part B: Polym. Phys., 33, 467 (1995).
84. S.S. Voyutskii, J. Adhes., 3, 69 (1971).
85. P.F. Green, Macromolecules, 24, 3373 (1991).
86. C.R. Bartels, B. Crist, and W.W. Graessley, Macromolecules, 17, 2702 (1984).
87. S.J. Mumby, B.A. Smith, E.T. Samulski, L.P. Yu, and
M.A. Winnik, Polymer, 27, 1826 (1986).
88. E. Benoit, J.F. Koening, and R. Muller, Macromol. Symp.,
158, 155 (2000).
89. H. Qiu and M. Bousmina, J. Rheol., 43, 551 (1999).
90. M. Bousmina, H. Qiu, M. Grmela, and J.E. Klemberg-Sapieha,
Macromolecules, 31, 8273 (1998).
91. H.T. Oyama and T. Inoue, Macromolecules, 34, 3331 (2001).
92. H.T. Oyama, T. Ougizawa, T. Inoue, M. Weber, and K.
Tamaru, Macromolecules, 34, 7017 (2001).
93. C. Scott and C.W. Macosko, J. Polym. Sci. Polym. Phys.
Ed., 32, 205 (1994).
94. E.J. Kramer, Isr. J. Chem., 35, 49 (1995).
95. J.S. Schulze, J.J. Cernohous, A. Hirao, T.P. Lodge, and
C.W. Macosko, Macromolecules, 33, 1191 (2000).
96. J.S. Schulze, B. Moon, T.P. Lodge, and C.W. Macosko,
Macromolecules, 34, 200 (2001).
97. Z. Yin, C. Koulic, C. Pagnoulle, and R. Jerome, Macromolecules, 34, 5132 (2001).
98. Z. Yin, C. Koulic, C. Pagnoulle, and R. Jerome, Langmuir,
19, 453 (2003).
99. Z. Yin, C. Koulic, C. Pagnoulle, and R. Jerome, Macromol.
Symp., 198, 197 (2003).
100. H.J. Kim, K.J. Lee, and Y. Seo, Macromolecules, 35, 1267
(2002).
101. B.J. Kim, G.H. Fredrickson, and E.J. Kramer, Macromolecules, 40, 3686 (2007).
102. B.J. Kim, H. Kang, K. Char, K. Katsov, G.H. Fredrickson,
and E.J. Kramer, Macromolecules, 38, 6106 (2005).
103. J. Jiao, Reaction Kinetics and Morphological Changes at
Polymer Interfaces and Surfaces, Ph.D. Thesis, Cornell University (1997).
104. H.R. Brown, Macromolecules, 24, 2752 (1991).
105. J.W. Smith, F.J. Kramer, F. Xiao, C.Y. Hui, W. Reichert,
and H.R. Brown, J. Mater. Sci., 28, 4234 (1993).
106. B. Bernard, H.R. Brown, C.J. Hawker, A.J. Kellock, and
T.P. Russell, Macromolecules, 32, 6254 (1999).
107. C.F. Creton, E.J. Kramer, C.Y. Hui, and H.R. Brown, Macromolecules, 25, 3075 (1992).
108. H.R. Brown, J. Mater. Sci., 25, 2791 (1990).
109. E. Boucher, J.P. Folkers, H. Hervet, L. Leger, and C. Creton, Macromolecules, 29, 774 (1996).
110. E. Boucher, J.P. Folkers, C. Creton, H. Hervet, and L.
Leger, Macromolecules, 30, 2102 (1997).
111. Y. Seo, T.H. Ninh, S.M. Hong, S. Kim, T.J. Kang, H.
Kim, and J. Kim, Langmuir, 22, 3062 (2006).
112. J.E. Bidaux, G.D. Smith, N. Bernet, J.A.E. Manson, and J.
Hilborn, Polymer, 37, 1129 (1996).

POLYMER ENGINEERING AND SCIENCE-2010 2285

113. Y. Seo, J. Lee, T.J. Kang, H.J. Choi, and J. Kim, Macromolecules, 40, 5953 (2007).
114. C. Laurens, R. Ober, C. Creton, and L. Leger, Macromolecules, 37, 6806 (2004).
115. Y. Seo and T.H. Ninh, Polymer, 45, 8573 (2004).
116. K. Cho and F. Li, Macromolecules, 31, 7495 (1998).
117. H. Koriyamaa, H.T. Oyamaa, T. Ougizawaa, T. Inouea, M.
Weberb, and E. Kochb, Polymer, 40, 6381 (1999).
118. C. Laurens, C. Creton, and L. Leger, Macromolecules, 37,
6814 (2004).
119. E.A. Eastwood and M.D. Dadmun, Macromolecules, 35,
5069 (2002).
120. C. Laurens, R. Ober, C. Creton, and L. Leger, Macromolecules, 34, 2932 (2001).
121. C.J.G. Plummer, H.H. Kausch, C. Creton, F. Kalb, and L.
Leger, Macromolecules, 31, 6164 (1998).
122. K. Char, H.R. Brown, and V.R. Deline, Macromolecules,
26, 4164 (1993).
123. C. Creton, H.R. Brown, and V.R. Deline, Macromolecules,
27, 1774 (1994).
124. K.K. Jozsef, Polym. Eng. Sci., 36, 203 (1996).
125. C.J.G. Plummer, P. Menu, N. Cudre-Mauroux, and H.H.
Kausch, J. Appl. Polym. Sci., 55, 489 (1995).
126. C.J.G. Plummer and H.H. Kausch, Macromol. Chem. Phys.,
197, 2047 (1996).
127. X. Zhang and G. Shi, Polymer, 35, 5067 (1994).
128. K.R. Shull, A.J. Kellock, V.R. Deline, and S.A. MacDonald, J. Chem. Phys., 97, 2095 (1992).
129. L. Leibler, Makromol. Chem. Macromol. Symp., 16, 1
(1988).
130. J. Noolandi and K.M. Hong, Macromolecules, 15, 482
(1982).
131. J. Noolandi and K.M. Hong, Macromolecules, 17, 1531
(1984).
132. J. Jiao, E.J. Kramera, S. Vosb, M. Mollerc, and C. Koning,
Polymer, 40, 3585 (1999).

2286 POLYMER ENGINEERING AND SCIENCE-2010

133. K.R. Shull, J. Chem. Phys., 94, 5723 (1991).


134. E. Helfand and Y. Tagami, Polym. Lett., 9, 741 (1971).
135. J. Jiao, E.J. Kramera, S. Vosb, M. Mollerc, and C. Koning,
Macromolecules, 32, 6261 (1999).
136. J.K. Kim, W.Y. Jeong, J.M. Son, and H.K. Jeon, Macromolecules, 33, 9161 (2000).
137. S.P. Lyu, J.J. Cernohous, F.S. Bates, and C.W. Macosko,
Macromolecules, 32, 106 (1999).
138. S.P. Lyu, F.S. Bates, and C.W. Macosko, AIChE J., 46,
229 (2000).
139. J. Zhang, T.P. Lodge, and C.W. Macosko, Macromolecules, 38, 6586 (2005).
140. C. Auschra, R. Stadler, and I.G. Voigt-Martin, Polymer,
34, 2081 (1993).
141. C.W. Macosko, P. Guegan, A.K. Khandpur, A. Nakayama,
P. Marechal, and T. Inoue, Macromolecules, 29, 5590
(1994).
142. S.H. Anastasiadis, I. Gancarz, and J.T. Koberstein, Macromolecules, 22, 1449 (1987).
143. P. Cigana and B.D. Favis, Polymer, 39, 3373 (1998).
144. H.Y. Kim, U. Jeong, and J.K. Kim, Macromolecules, 36,
1594 (2003).
145. H.Y. Kim, D.H. Lee, and J.K. Kim, Polymer, 47, 5108
(2006).
146. W. Yu, M. Bousmina, M. Grmela, and C.X. Zhou, J.
Rheol., 46, 1401 (2002).
147. H. Qiu, M. Bousmina, and J.M. Dealy, Rheol. Acta, 41, 87
(2002).
148. J.S. Schulze, Reaction Kinetics of End-functionalized Chains
at PolymerPolymer Interfaces, Ph.D. Thesis, University of
Minnesota (2001).
149. J. Zhang, Ph.D. Thesis, University of Minnesota (2007).
150. Q. Du, G.J. Jiang, J. Li, and S.Y. Guo, Polym. Eng. Sci.,
50 (2010).
151. J.B. Shen, J. Li, and S.Y. Guo, Polym. Adv. Technol.,
in press.

DOI 10.1002/pen

You might also like