Вы находитесь на странице: 1из 9

Current Opinion in Colloid & Interface Science 17 (2012) 132–140

Contents lists available at SciVerse ScienceDirect

Current Opinion in Colloid & Interface Science


journal homepage: www.elsevier.com/locate/cocis

Free volume and mass transport in polymer nanocomposites


G. Choudalakis ⁎, A.D. Gotsis
Department of Sciences, Technical University of Crete, GR-73100, Hania, Greece

a r t i c l e i n f o a b s t r a c t

Article history: This review relates the free volume properties and the morphology to the mass transport coefficients of poly-
Received 14 December 2011 mer nanocomposites. Direct, utilising the method of Positron Annihilation Life-time Spectroscopy (PALS), or
Received in revised form 13 January 2012 indirect measurements of the free volume in the nanocomposites are discussed and the influence of spherical
Accepted 16 January 2012
or anisometric nanoparticles on its properties is examined.
Available online 26 January 2012
© 2012 Elsevier Ltd. All rights reserved.
Keywords:
Free volume
Positron annihilation
Gas transport
Diffusion
Solubility
Permeability
Nanocomposites

1. Introduction: polymer nanocomposites elements that carry the applied load in the polymer, while the particles
also deflect the evolving cracks. The stress transfer to the reinforcement
Polymer nanocomposites materials are hybrid systems that consist of phase results in significant tensile and toughening improvement. An-
a polymeric matrix and dispersed inorganic particles of nanometer scale. other major advantage of the nanocomposites is their diverse barrier
When the dispersed inorganic phase consists of spherical nanoparticles properties. Spherical nanoparticles can enhance the permeation of
with diameter of the order of 50 nm or less, e.g. fumed silica (FS), carbon gases through the PN, the rate depending on the size of the gas mole-
black (CB), TiO2, the materials are referred to as Polymer Nanocompo- cule. These materials are good candidates for gas separation applica-
sites (PN). If the dispersed particles originate from the family of phyllosi- tions. On the other hand, PCNs reduce the gas permeation rate
licates and have flat shape with high surface area, the composites are because the impermeable clay layers, with their large surface area,
called Polymer Clay Nanocomposites (PCN). Some common particles force a tortuous pathway for a permeant. These materials are ideal as
used for PCNs are montmorillonite (MMT), rectorite, hectorite, vermicu- barrier layers in packaging against the penetration of unwanted gases,
lite etc. e.g. O2, or to prevent gasses like CO2 or H2 to escape. Recent reviews of
The creation of a PN or PCN is not a simple issue. In powder form the the methods used for synthesis of PN and PCN and for their various
inorganic phase consists of aggregates or stacks of nanoparticles held to- properties can be found in the literature [1, e.g.].
gether by electrostatic forces. The formation of the nanocomposite
requires the deagglomeration of the primary nano-particles or nano- 2. Free volume
layers (Fig. 1). This can be done if the electrostatic forces between the
particles are weakened and the interactions with the organic polymer The term “free volume” (f.v.) in a polymer refers to the volume of the
are enhanced. To exfoliate the clay, therefore, modification of the sur- total mass that is not actually occupied by the molecules themselves. This
faces of the mineral particulates is needed. The objectives of the modifi- is the room that the parts of the chains have to move around. The concept
cation process are (i) the reduction of the electrostatic forces (swelling) of the free volume is not uniquely defined. Various definitions are often
and (ii) the compatibilisation of the filler with the matrix. The latter de- used, such as hole, configurational, fluctuation and excess free volume:
termines the degree of interaction between the two phases, which is cru-
cial for the final nanocomposite morphology. Hole free volume: even if the polymer chains were perfectly
The nanocomposites can have excellent mechanical properties. The aligned, there is free space (holes) between them.
nano-dimensions maximise the number of available reinforcing Configurational free volume: additional free space is created, due to
the insufficient chain packing.
⁎ Corresponding author. Fluctuation (dynamic) free volume: the polymer main- or side-
E-mail address: gchoudalakis@isc.tuc.gr (G. Choudalakis). chains are not fixed but can move (rotate, vibrate etc.) randomly,

1359-0294/$ – see front matter © 2012 Elsevier Ltd. All rights reserved.
doi:10.1016/j.cocis.2012.01.004
G. Choudalakis, A.D. Gotsis / Current Opinion in Colloid & Interface Science 17 (2012) 132–140 133

molecular chains and the resulting mobility change considerably


and abruptly.
High Tg polymers consist of inflexible chains and are limited in
their chain mobility. Low Tg polymers have, in general, more flexible
chains and present higher segmental mobility. The higher the chain
stiffness, the higher the Tg and the broader the distribution of f.v.
hole sizes, since the available variety of chain conformations is
restricted.
At T b Tg the segmental motion is frozen and thermal expansion of
free volume is almost impossible. In this case, the size, shape and
position of the f.v. holes remain constant (very long relaxation
times) and a definite, time-averaged Gaussian size distribution of
holes, F(Vh), is established. Assuming N spherical holes:
" #
2
N ðV −V Þ
Fig. 1. Delamination of the primary mineral particles into individual nano-layers. The F ðV h Þ ¼ pffiffiffiffiffiffi exp − h 2 0 : ð2:1Þ
layers are located on top of each other like the pages of a book. The Van der Waals σ 2π 2σ
gaps between the layers, are called galleries.
The width of this distribution is:
rffiffiffiffiffiffiffiffiffiffiffiffiffiffi
due to thermal activation. These motions generate transient gaps, V h RT g
σ¼ ; ð2:2Þ
which create extra free volume. B
Excess free volume: at temperatures below Tg, the polymer usually
has an actual volume greater than what it would have had, if it had with B the bulk modulus at Tg and Vh the average hole size [3 •].
Amorphous polymers in the rubbery state have high chain mobil-
been allowed to cool under equilibrium conditions. The difference
ity and short relaxation times. Motions of much longer segments of
between equilibrium free volume and the actual volume under
the backbone chain can take place. These motions are directly related
quenched conditions is the “excess free volume” (Fig. 2). to the increase in free volume caused by the thermal expansion of the
In time, the polymer molecules will try to rearrange themselves, to system.
approach the equilibrium volume of the material, thus eliminating  
the excess free volume. This time dependent slow decrease in volume FFV ¼ ðFFV ÞT g þ T−T g α T ð2:3Þ
is known as “physical ageing” and affects all material properties
which change drastically around Tg. with αT the thermal expansion coefficient for free volume. The WLF
The sum of hole and configurational f.v. is the “static” free volume. theory gives (FFV)Tg = 0.025, αT = 4.8 × 10 − 4 K − 1. The positions of
The excess f.v. may be static or dynamic, depending on the time scale the free volume holes change with time following the motions of
of the relaxation processes. The ratio of static to dynamic excess f.v. is the polymer chains. That is, the holes move freely inside the polymer,
thought to be greater than 1 and depends on temperature. as no energy change is required for their redistribution.
The static f.v., Vf = V − V0, with V the specific volume of the poly- At a given temperature the number of the holes remains constant,
mer and V0 the volume occupied by the atoms of the polymer chains while their size, shape and position vary with time. The number of
at 0 K, is often expressed in terms of the Van der Waals volume, Vw, holes increases with temperature, i.e. additional free volume sites
calculated using the group contribution method [2] as V0 = 1.3Vw. are created. The increase is induced by thermal activation (Brownian
The coefficient 1.3 is arbitrary and may sometimes result in faulty in- segmental motion) and not by an orientational effect. The size of the
terpretations. The fractional free volume is: FFV = Vf/V. free volume holes is sensitive to the applied stresses during mechan-
The size, shape and position of the polymer f.v. holes vary with ical deformation but the number of the holes is not affected [4].
time. Around the glass transition temperature, Tg, the polymer un- Crosslinking does not contribute to changing the free volume size in
dergoes a dramatic transformation, as the degrees of freedom of the the glassy state, since the macromolecular motions are frozen (vitri-
fied). In the rubbery state, crosslinking can decrease the free volume,
as it confines the mobility of chain segments. The effect is stronger
when starting with a looser, mobile polymer network (e.g. crosslinking
flexible linear chains). In rigid network structures (e.g. thermosetting
polymers) the influence of crosslinking is expected to be smaller.

3. Mass transfer coefficients


id
Polymer Volume

qu
Li
rim
lib

Gas transport through a polymer membrane is a complex process


ui
Eq

m Glass that includes the sorption of gas molecules on the surface of the mem-
uilibriu
Non Eq
Excess Free brane; the diffusion through it; and, finally, the desorption of gas from
Volume the other surface of the membrane. The permeability coefficient, P
Glass (mol Pa − 1m− 1s− 1), of the gas molecules through the membrane is
ibrium
Equil the product of the diffusion coefficient, D (m2s− 1), and the sorption co-
efficient or solubility, S (mol m 3Pa− 1):

P ¼ D⋅S: ð3:1Þ
Tg
Temperature These three transport coefficients are discussed separately and
their connection with the polymer free volume is analysed in the fol-
Fig. 2. Definition of the excess free volume. lowing paragraphs.
134 G. Choudalakis, A.D. Gotsis / Current Opinion in Colloid & Interface Science 17 (2012) 132–140

3.1. Solubility coefficient

In the rubbery state of the polymer the solubility coefficient of the


gas is determined by Henry's law, C = kdp, with C the concentration of
the sorbed gas; kd Henry's constant (mol m 3 Pa − 1); and p the pres-
sure. In the glassy state we cannot neglect Langmuir's contribution
of the concentration of the sorbed gas:

CH b
C¼ p: ð3:2Þ
1 þ bp

The parameter CH characterises the amount of unrelaxed free vol-


ume in the glassy state (excess free volume):
!
V g −V l 22400
CH ¼ : ð3:3Þ
Vg Vp Fig. 3. Schematic representation of the diffusion process of a gas molecule through free
volume in a polymer.
Vg and Vl are the specific volumes in the glassy and the liquid (rubbery)
states and Vp is the molar volume of the sorbent gas (ml/mol). The pa-
rameter b characterises the tendency of a given penetrant to sorb into are much slower than those of polymer relaxations (Fickian diffu-
the excess unrelaxed volume in the non equilibrium matrix. Taking sion). When diffusion and relaxation rates are comparable the diffu-
into account both contributions, the solubility coefficient in the glassy sion is non-Fickian.
state becomes: A molecule can jump into a hole if the hole is larger than a critical
size υ*. The probability for such a process is proportional to
CH b exp(− γυ*/υf), where υf is the molar mean hole free volume and γ is
S ¼ kd þ : ð3:4Þ
1 þ bp an overlap factor (0 b γ b 1) indicating that the same hole may be
available to more than one molecules. The diffusion coefficient is [2]:
The relations between solubility and free volume are not always !
 
obvious, since the way penetrants dissolve in the polymer depends −E −γξυ
D ¼ D~0 exp exp ; ð3:5Þ
on the total pressure, the affinity between gas and polymer and the kT υf
temperature. For a specific polymer nanocomposite system, however,
the solubility coefficient will change only if the f.v. changes due to the where ξ = υs*/υ*; and υs* is the critical volume of the penetrant. The ac-
presence of the particles. Henry's constant depends on the facility of tivation energy, E*, for the diffusion of trapped molecules from one hole
forming holes in the space among chains [5] and cannot be associated to another is related to the molecular “jumping length”. The proportion-
with these changes. For f.v. holes that are all large enough to accom- ality coefficient D~0 is related to the size and shape of the permeant:
modate gas molecules, the solubility coefficient should reflect directly
their number, remaining unaffected by changes in their sizes. Other- RTσ 2

wise, kd has a tendency to increase with increasing average hole D~0 ¼ 1=2 : ð3:6Þ
M
size as well.
The connection between the solubility coefficient and the number of where σ is the Lennard–Jones size parameter and M is the molecular
free volume holes has been verified experimentally in acrylic resin/ weight of the permeant [2].
laponite nanocomposites [6•]. The total f.v. fraction, measured by Posi- The amount of free volume alone does not provide information
tron Annihilation Lifetime Spectroscopy, has an identical trend, as a about the possible connectivity of the holes. When f.v. elements are
function of laponite particles loading, with that of the solubility coeffi- added randomly to a system, the percolation threshold will be reached
cient, which is measured separately in a permeation cell. On the other at some point and a continuous pathway will exist for the permeating
hand, Langmuir's sorption is principally governed by the available free gas in the system. This is commonly identified with a pore-flow diffu-
volume. Consequently, the variation in solubility is attributed mostly sion process.
to CH, which is determined by the difference Vg − Vl (Eq. (3.3)). This, Diffusivity should be correlated to the accessible (for a given gas)
in turn, is influenced by the Tg. Therefore, higher Tg polymers generally free volume fraction, AFV. For glassy polymers the percolation thresh-
exhibit higher solubility coefficients. It can be stated that the FFV is re- old is in the range of 2–4% AFV. This is equivalent to 25–31% FFV [7 •].
lated to the solubility because it corresponds to the free space in the In the rubbery state the picture of fixed holes needs to be replaced,
polymer matrix where gas molecules can locate [5]. since the polymer chains are mobile and the f.v. holes show a dynamic
variation about their size, shape and position. The gas molecules diffuse
3.2. Diffusion coefficient within the fluctuating interstitial f.v. with much greater mobility than in
the glassy state. That is to say, the jumping step is larger in polymers
In contrast to the solubility, the diffusion is a kinetic process which with a stiff backbone chain than in flexible chains [2]. The factor
depends on the mobility of gas molecules in the polymer matrix. At exp(− E*/kT) can now be neglected, since less energy is required for dif-
temperatures below Tg, the polymer backbone is considered to be in fusional jumping, and Eq. (3.5) is often written in the following form:
a frozen state, segmental chain motions are drastically reduced, the
number of f.v. holes is fixed and no hole redistribution is likely. Gas D ¼ ART expð−B=FFV Þ; ð3:7Þ
transport is, therefore, assumed to take place via fixed (pre-existing)
holes. A gas molecule must “find its way” from hole to hole along where B is equal to γυs*/υ* and A is related to the size and shape of the
pathways involving only minor segmental rearrangements (Fig. 3). permeant and replaces the term σ2/M1/2 of Eq. (3.6). It seems that B de-
This means that the diffusivity depends largely on the number of pends not only on the type of gas but also on the polymer, since υ* is the
the holes with an appropriate size, able to accommodate the diffusing minimal free volume hole size required for the penetrant diffusion. The
gas molecule. This mechanism is expected when the diffusion rates size of the diffusing molecule has a smaller influence if the polymer is in
G. Choudalakis, A.D. Gotsis / Current Opinion in Colloid & Interface Science 17 (2012) 132–140 135

the rubbery state than on a rigid glassy membrane, where the hole size Here Ii is the relative intensity and τi is the lifetime of the i-th Debye
distribution is frozen. Glassy polymers seem to be more suitable for gas process. In polymers the lifetime spectrum (Eq. (4.1)) is usually re-
separation membranes. solved into three terms. The first term (I1, τ1) has a relaxation time (life-
While correlations between lnD and 1/FFV of the type of Eq. (3.7) time) τ1 ∼ 0.125 ns and is related to the fast annihilation of the p-Ps and
have been verified for various glassy polymers the values of parame- the free positrons. The second process (I2, τ2) with a relaxation time
ters A and B show a significant variance. A linear fit of data for the dif- τ2 ∼ 0.4 ns corresponds to the direct annihilation of the free positrons.
fusion of CO2 in four polymers (PS, PMMA, PC, PP) gives A = 4⋅ 10 − 6 The third process (I3, τ3) is the “pick-off” annihilation process of the o-
cm 2/s and B = 0.97 [2]. On the other hand, values of A = 3.9⋅ 10 − 7 Ps and has a relaxation time τ3 in the range of 1 to 5 ns. In some rare
cm 2/s, B = 0.297 were reported for diffusion of CO2 in polycarbonates cases of high free volume polymers there is a fourth significant lifetime
[8], and A = 7.4⋅ 10 − 5 cm 2/s, B = 0.1336 in polyurethanes [9]. This component, (I4, τ4), with τ4 greater than 10 ns [11 etc.] Lifetime, τ4, is
variance indicates that the parameters depend also on the polymer ascribed to the bimodal distribution of possibly non spherical free vol-
and not only on the type of gas. ume elements.
There is also the case where the free volume in the material consists To o-Ps lifetime is related to the size, R, of the free volume hole
of a network of large holes connected by channel-like holes. Diffusing where it is trapped (Tao-Eldrup model):
molecules may now move with little friction within the network.
Large variations in the size or in the number of larger holes have small " #
2πR −1
1 R sin RþΔR
effect on the diffusion through the free volume network, whereas a τ3 ¼ 1− þ ; ð4:2Þ
2 R þ ΔR 2π
small decrease in the size of the smaller holes may destroy the connec-
tivity of the channel network. In this case the mobility of diffusing mol-
ecules is strongly hindered [10]. This means that a broader distribution where ΔR ∼ 0.1656 nm is the electron layer thickness around the hole,
of free volume holes can result in higher values of the gas diffusion co- usually assumed to be uniform. This model is sufficient for characteris-
efficient than a narrow distribution with the same average size. Both the ing small voids with cavity diameter smaller than the thermal de Broglie
amount of the free volume and its distribution over the hole sizes are wavelength of Ps (∼ 6 nm). There is also a low threshold for detecting
important factors that influence the diffusion coefficient. f.v. holes by PALS, equal to the minimum for o-Ps localisation Rc:

!1=2
4. Direct measurement of the free volume using PALS π ℏ2
Rc ¼ : ð4:3Þ
2 2mV~ 0
Positron Annihilation Lifetime Spectroscopy (PALS) is an accept-
able method for the direct measurement of polymer free volume.
The positrons are used for bulk measurements as they can penetrate ℏ is Planck's constant, m the electron mass and V~ 0 the potential barrier
a few tenths of a mm inside the sample. When a positron meets an (∼ 1.4 eV). Typically, Rc ≈ 2 Å.
electron, they either annihilate generating photons or they form a The intensity I3 is related to the formation probability of o-Ps in the
temporary pair, a positronium, Ps. There are two possible Ps atoms: f.v. holes and is connected to their number. The fractional free volume
p-Ps corresponding to the antiparallel spin configuration of the two in polymers is FFV∼ VhI3, where Vh = (4/3)πR3, in Å3, is the volume of
particles, and o-Ps corresponding to the parallel. Ps formation can be the average spherical hole. The total fractional free volume cannot be
achieved only in low electron density regions such as in voids or in estimated, since I3 originates only from the holes, where the o-Ps can
the free volume of the polymer (Fig. 4) and its annihilation is accom- be formed, which are not the total free volume. Thus, a proportionality
panied by the emission of 511 keV γ-photons. If no such areas are parameter, C, is used:
present, then the positrons will be annihilated directly.
To estimate the lifetimes, the spectrum of the emissions due to FFV ¼ CV h I3 ; ð4:4Þ
positron annihilations is resolved into N Debye decaying processes:
where C depends on the polymer and can be deduced by comparing the
  values of FFV from PALS measurements with values of FFV from P–V–T
X
N
Ii t
F ðt Þ ¼ exp − : ð4:1Þ data. Indicative values are 0.325 nm− 3 for poly(vinyl acetate) [12]
i¼1
τi τi and 1.8 nm− 3 for amine-cured epoxy polymers [13].
When a fourth mode (I4, τ4) is found, the size can be estimated by
an extended Tao-Eldrup model [14]:

Thermalization  "  #−1


R−Ra b 1
τ4 ¼ 0:0626 1− þ ; ð4:5Þ
R þ ΔR 140

511 keV
e where Ra = 8 Å and b = 0.55 are fitted parameters.
In molecular solids, a decrease in I3 can be due either to a decrease in
the number of o-Ps annihilation sites or to a decrease in the probability
e
+
of Ps formation. Negatively charged polar groups, M− (nitro, halogen
and hydroxyl), can suppress Ps formation considerably by arresting
the positrons. This competing mechanism, which reduces the o-Ps for-
+
e mation and results in a decrease of I3, is often observed upon addition
Diffusion
of salts. The latter increase the dipole moment of the polymer molecules
511 keV
enhancing localisation of the negative charges. This decrease in I3 is ap-
proximately exponential with anion concentration, approaching a satu-
ration value [15].
Besides its effect on I3, the chemistry of the material also affects τ3. A
Trapping and Annihilation stronger polarity effectively quenches the Ps atoms and, thus, shortens
the Ps lifetimes. The size of the cavities may remain constant, but the
Fig. 4. The process from the time of positron implantation to the time of Ps annihilation. ortho-Ps lifetimes can be quenched by various oxidisers, including
136 G. Choudalakis, A.D. Gotsis / Current Opinion in Colloid & Interface Science 17 (2012) 132–140

paramagnetic ions and chlorides. Oxygen is particularly efficient in pro- Changes of the crystallinity of the matrix: the incorporation of nano-
moting the conversion of o-Ps to p-Ps [10]: particles often changes the crystallinity of the matrix polymer. The
 particles perturb the polymer chain packing and can reduce the
o−Ps þ O2 → ½Ps−O2  → p−Ps þ O2 ; ð4:6Þ
crystallinity in the polymer, alleviating, thus, some of the decrease
At high temperatures, well above Tg, the o-Ps lifetime may reflect the in chain mobility that accompanies crystallisation. Inevitably, this
formation of so-called o-Ps bubbles. When the surface tension is small enhances the free volume. On the other hand, the nanofillers may
enough, the repulsive forces between the Ps and the surrounding poly- also act as nucleating agents, increasing the degree of crystallinity
mer chains are sufficient to form a “bubble” around the o-Ps atom. This and reducing the free volume. The net result will depend on the
bubble formation is only possible if the segmental relaxation times are crystallisation characteristics of the matrix and its interaction with
in the order of, or shorter than the o-Ps lifetime [16]. Obviously, when nucleation agents.
bubbles are present, the o-Ps lifetime does not reflect the actual size of Changes of the cross-linking density of the matrix: Adding nano-
a f.v. hole. particles in a thermosetting polymer can have two opposite re-
sults. The particles may serve as crosslinkers, which increases
5. Effects of nanoparticles on polymer free volume the cross-link density and decreases the polymer mobility and
the free volume. Or they may hinder the cross-linking reaction
The incorporation of inorganic nano-fillers into the polymer matrix,
and enhance the free volume. The net effect depends on the chem-
will inevitably change the morphological features of the matrix and,
consequently, the free volume properties. The following effects are istry of the crosslinking reaction, the possible catalytic action of
expected in the nano-composite: the nano-particles and the interfacial properties of the system.

Interfacial regions: because of their large surface area the inorganic All these effects on the polymer free volume coexist in any nano-
composite system. Which of them become dominant depends pri-
particles create interfacial regions in the organic polymer. These
marily on the degree of interaction between the two components,
regions are numerous and have finite thickness because of the
the volume fraction of the filler and the geometrical features of the
weak interaction/miscibility between the inorganic reinforcement particles.
and the organic polymer and the need for compatibilisation. The
interfacial layers help the stress transfer between the nanoparti- 6. Free volume measurements in polymer nanocomposites
cles and the matrix, but contribute to the enhancement of the
overall free volume. Since clays create more interfacial area than 6.1. Spherical nano-particles
spherical particles, the interfacial effect on free volume is expected
to be stronger in PCN than in PN. The changes in free volume that nanoparticles induce in the ma-
Interstitial cavities in the filler agglomerates: if exfoliation in a PCN is trix polymer have been verified experimentally in several reports.
not complete, stacks of nano-layers (intercalated morphology) However, because of the complexity of the studied systems, some un-
exist inside the polymer matrix. The interlayer spacing of these certainties remain on what actually is measured.
The incorporation of particles has small effect on the polymer free
stacks varies from 2 to 5 nm. If polymer chain penetration in the gal-
volume. In poly(dimethyl siloxane)/polysilicate nanocomposites, 30%
leries is incomplete, the volume of the galleries contributes to the
nanoparticles content reduces the hole radius from 4.35 Å to 4.31 Å
overall free volume of the composite. This extra f.v. is in layers rather
[18], while the intensity I3 remains almost constant. Crosslinking has
than spherical holes. However, experimental results of Na-MMT no effect on the f.v. properties. All this casts some doubt on the suitabil-
powder showed that the FFV was only 1.2% [17]. There is some un- ity of PALS to reveal differences in dynamic free volume, which is in
certainty in these results, since the interlayer distances in this clay principle affected by temperature and chain mobility. The o-Ps may
are very small: (i) Ps formation may have been hindered by the neg- “see” these sites as “occupied” if the frequency of the molecular motion
ative charges on the clay particle surfaces; and (ii) the penetration of through the sites is faster than the annihilation rate.
the positrons may be limited by the presence of the cations in the The longer lifetime, τ4, that has been observed in some polymers,
galleries. In organomodified clays the FFV is expected to be much is ascribed to large, possibly interconnected, free volume elements. In
greater. poly(1-trimethylsilyl-1-propyne) the average radius of these cavities
was about 5.5 Å and increased to 11.5 Å with the addition of fumed
In the case of spherical nanoparticles, the weak interaction can
silica (FS) nanoparticles [11]. The enhancement of the longer o-Ps life-
lead to the formation of filler aggregates enclosing interstitial vol-
time (τ4) was a consequence of o-Ps annihilating in the interstitial
ume inaccessible to polymer segments. These enclosed small voids cavities of the FS agglomerates. If the interstitial voids in the FS pow-
can be in the order of 10 Å [11] and create extra free volume. der are filled with polymer chains, then τ4 is not present.
Chain segmental motion immobilisation: the randomly distributed In general, the o-Ps intensity, I3, decreases with particles volume
filler particles may restrict the chain segmental motion and reduce fraction due to the interfacial layers between the two components.
the chain mobility and the amount of free volume. Such effects are In HDPE/CaCO3 nanocomposites τ3 remains practically constant with
common to both PN and PCN. Due to their high surface area, nano- particles content, but I3 decreases [19 •]. The number of interfacial
clays are expected to be more effective in immobilisation because layers between CaCO3 and HDPE increases with adding more CaCO3
they act as bridges over polymer chains, limiting their capability to and the probability for Ps formation decreases, due to the increasing
fraction of positrons annihilating directly in the interfacial layers:
change their conformation.
The reduction in o-Ps intensity is accompanied by the enhancement
Insufficient chain packing: the filler particles may confine the poly-
in I2. Under these circumstances the interpretation of I3 as being pro-
mer chains and disrupt their packing. This effect can enhance the
portional to the number of f.v. holes may be inaccurate.
free volume due to this inefficient packing or due to the increase Highly electronegative atoms or radicals on the particle surfaces
of the distances between polymer segments. may also induce a reduction in o-Ps intensity, which does not always
Changes of the free volume hole size distribution: the particles may reflect changes in the free volume. Decrease in I3 was observed in PP/
alter the f.v. hole sizes distribution in the system [6 •]. CB and LDPE/CB nanocomposites [20]. This reduction was ascribed to
G. Choudalakis, A.D. Gotsis / Current Opinion in Colloid & Interface Science 17 (2012) 132–140 137

the existence on the particle surfaces of radicals associated with the the clay loading exceeds 19 wt.%. At these loadings the samples are
aromatic structures of the carbon sheets and with functional groups rather microcomposites with large particle aggregates. Since the
containing oxygen. A significant reduction on both lifetime and inten- clay on its own shows an o-Ps lifetime in the order of 2.17 ns and
sity of o-Ps was observed in PMMA/TiO2 [15], presumably caused by the neat polyamide shows a lifetime of 1.67 ns, it is expected that ad-
the − OH groups on the surface of the titania particles. An in- dition of clay above 10 wt.% will lead to an increase in τ3, simply by
crease of τ3 has been reported for polystyrene/SiO2 [21] and in applying the rule of mixtures.
poly-(ethylene-alt-propylene)/SiO2 [22 •] nanocomposites. In all There are some reports that clays may enhance significantly the
these cases the o-Ps intensity remains unchanged or decreases slightly. polymer free volume also for clay loadings lower than 10%. In Ref.
It seems that silica particles cause an enhancement on free volume in [29] the mean volume of the holes increased by 26% when 5 wt.% of
any matrix they are dispersed. This is due either to chain packing dis- MMT was added in PVA-PES hydrogels. This behaviour is attributed
ruption or to the interstitial voids in the silica agglomerates. to the f.v. in the interfacial layers between the organic and inorganic
components. As in the case of epoxy/rectorite, a reduction in I3 was
6.2. Clay nano-particles accompanied by an enhancement in I2. Both the o-Ps lifetime and
the intensity increased in cyanate/ bentonite systems [30]. The en-
The influence of the clay on the polymer free volume is different hancement of f.v. was about 4.5% for the sample containing 2.5 wt.%
from that of the isometric nanoparticles. Most experimental results of particles. This increasing trend is ascribed to the insufficient
show a reduction of the amount of the f.v. as a function of clay load- chain packing that the particles cause in the polymer.
ing. This reduction concerns mainly the o-Ps intensity, I3, and not Summarising, the free volume in the polymer matrix remains
the lifetime, τ3, which remains mostly unaffected: While the mean mostly unaffected by the addition of clay at low volume fractions or
hole size remains constant, the number of the f.v. holes seems to de- even decreases slightly. This small reduction is seen mainly as a de-
crease. This is a clear evidence that the clays inhibit the polymer chain crease of the number of the f.v. holes, while their average size remains
motions resulting in a reduction of the dynamic f.v. relative to the the same. Exceptions to this involve systems with large interfacial re-
neat polymer. However, if the interaction between the two phases gions, where extra f.v. may develop. Further, at high clay volume frac-
is very weak, then the contribution of interfacial layers on f.v. can tions exfoliation/deagglomeration is incomplete. The free volume,
overwhelm the immobilisation effect resulting in an increased free then, is enhanced by the presence of interstitial empty space inside
volume. the aggregates, which is inaccessible to the polymer.
In elastomers, such as styrene-butadiene rubber/MMT [17], SBR/
Rectorite [23] etc., a decrease in FFV is observed as a function of clay 7. The permeability of polymer nanocomposites
loading, mainly due to the o-Ps intensity reduction. The hole radius
in these elastomers is about 3 Å and does not vary with the addition In polymer nanocomposite systems there is an additional feature
of filler. The reduction of the FFV is stronger (27%) in nanocomposites that strongly affects the gas diffusion process. The essentially imper-
of SBR with 40 phr rectorite, while the other materials present a slight meable nanoparticles act as obstacles to the gas molecules, forcing
reduction of the order of 5 − 6% for filler contents of 2.5–10 wt.%. them to follow longer and complicated routes to diffuse through the
Similar trends for I3 have been observed for rectorite in epoxy [24] material. This extension/complication of the path is called tortuosity
and ethylene-co-(vinyl acetate) [25] nanocomposites. The o-Ps life- and is schematically represented in Fig. 5. The tortuosity factor is de-
time is unaffected or slightly reduced, due to the increasing fraction fined as τ ¼ ‘′=‘, where ‘′ is the diffusion path length when particles
of positrons that annihilate in the interfacial layers between silicate are present and ‘ is the length in the neat polymer. The effect of tor-
platelets and polymer matrix. This reduction is accompanied by a tuosity on the diffusion process depends on the geometrical features
small increase in I2. In EVA/rectorite systems the 1% reduction in of the dispersed nanoparticles, their arrangement in space, their ori-
FFV is due to the chain immobilisation effect. Adding MMT to an entation compared to the diffusion direction and their volume frac-
epoxy/polyurethane (EP/PU) blend has no significant effect on FFV tion [31]. It is obvious that nanoclays will be more efficient in
for clay loadings in the area 0–5 wt.% [26]; for laponite particles reducing diffusion, since their surface area is much greater than the
added to an acrylic resin [6 •]; and in HDPE/PA6 blend/clay nanocom- spherical nanoparticles. In the case of nanocomposites, therefore,
posites [27]. Eq. (3.6) should be modified to account for the tortuosity effect.
These results indicate that the incorporation of clay in the polymer
matrix does not alter the free volume properties significantly, espe- A′
D¼ expð−B=FFV Þ: ð7:1Þ
cially at relatively low clay volume fractions. Higher clay loadings τ
have been found to have a stronger effect on f.v. The exfoliation of
the particles and the successful formation of the nanocomposite be- The diffusion coefficient in a polymer nanocomposite depends
come particularly difficult in these cases. In polyamide-6/MMT an en- both on the fractional free volume and the tortuosity factor. These
hancement in o-Ps lifetime has been reported [28], especially when two parameters can be either in accordance or in competition. The

Fig. 5. Schematic representation of the diffusion process of a gas molecule through a polymer/clay nanocomposite.
138 G. Choudalakis, A.D. Gotsis / Current Opinion in Colloid & Interface Science 17 (2012) 132–140

final increase or decrease of the diffusion coefficient upon adding was reported in PE/HDPE-g-MA20/MMT nanocomposite [40], where
nanoparticles depends on which of the two becomes dominant. the diffusion coefficient was reduced by 44% when 18.42 wt.% of
The curve of lnD vs. 1/FFV of some PVA-PA6/MMT nanocomposites MMT was added, while the solubility increased by 36% and the per-
[32] is linear in the range of 1/FFV b 15. That is, relations of the type of meability decreased by 24%. The poor compatibility between the
Eq. (7.1) are valid in this case, only when the fractional free volume is polymer and the inorganic particles creates interfacial regions,
higher than 6.7%. For smaller FFV the curve is not linear, showing a which enhance the f.v. This was the case in polyurethane/vermiculite
saturation with decreasing FFV. Similar results were obtained in nanocomposites [41 •], where the permeability of CO2 and N2, was re-
SBR/rectorite nanocomposites, where Eq. (7.1) holds mainly for duced only by 20 and 30%, although the high aspect ratio particles
FFV > 3.33% [23]. This suggests that when the FFV is relatively high, were expected to be more effective barriers.
it overcomes the influence of the tortuosity and the diffusion coeffi- The overall transport properties of the nanocomposites are affected
cient is determined by the free volume variations. by the presence of interfacial regions. In their multi-scale hierarchical
The incorporation of nanoparticles in a polymeric matrix can numerical model, Manke et al. [42•] defined a polymer-nanoparticle in-
cause redistribution of the free volume hole sizes. Even when the teraction strength parameter to quantify this effect. For high values of
total FFV is identical in two polymers, differences in their hole size this parameter, they showed that the polymer density next to the inter-
distribution may be adequate to induce different diffusivities, both face increases, indicating that the enhancement of the barrier properties
in glassy and rubbery polymers. This has been measured in acrylic are due also to the decrease of the free volume there. For large gas mol-
resin/laponite nanocomposites, where the particles do not influence ecules, however, Pryamitsyn et al. [43•] showed that the polymer-
the FFV but, nevertheless, the diffusion coefficient increases as a func- particle interactions are less important than the tortuosity effect. The
tion of laponite loading [6 •]. This enhancement is ascribed to changes mass transport properties, therefore, can be determined by the segmen-
in the hole size distribution. tal dynamics of the matrix.
The permeability coefficient is the product of the solubility and While the total amount of free volume affects the permeability of
diffusion coefficients. For polymer nanocomposites: a material as described above, the distribution of the f.v. hole sizes
may also influence the permeation process. This distribution can
A′ S strongly affect the diffusion process because the very small holes,
P¼ expð−B=FFV Þ: ð7:2Þ where the gas molecules cannot enter, do not contribute significant-
τ
ly to the diffusion. The solubility, which is connected to the overall
Therefore, P depends on the diffusion coefficient, the solubility f.v., remains unaffected by the hole size distribution. An example is
and the tortuosity. As described previously, the first two depend on the PMMA/MMT system [44]: While the solubility of CO2 in the
the free volume in the system. However, the permeability coefficient PMMA/MMT is approximately the same as that for pure PMMA, the
seems to be less sensitive to the FFV than each of these two coeffi- diffusion coefficient is much greater. This has been attributed to
cients is. I.e., the tortuosity of the system seems to play a more impor- changes in the f.v. size distribution towards a relatively greater num-
tant role in determining the value of P in nanocomposites. ber of larger holes.
Assuming an ideal periodical arrangement of nano-layers, the tor- In polymer nanocomposites filled with spherical nanoparticles the
tuosity factor is: gas permeability has an opposite trend compared to PCN. This type of
fillers often enhances the free volume and is expected to reduce the
αφw barrier properties of the matrix. The gas permeability is 3–4 times
τ ¼1þ ; ð7:3Þ
2 higher in 1,2-polybutadiene/TiO2 nanocomposites than in the unfilled
polymer. This is due mainly to the increase in gas solubility. Diffusion
where α, = L/W (Fig. 5). For example, the reduction of the permeabil- initially decreases with increasing particle loading, due to the in-
ity due to the tortuosity, that 3 vol.% spherical nanoparticles induce, is crease in tortuosity, and then it increases at higher loadings, as the
in the order of 15%. For the same vol. fraction of clay nanoparticles space in the interstitial cavities in the agglomerates becomes more
and for α = 100 the permeability decreases by 60%, all other parame- prominent [45].
ters remaining the same. These interstitial cavities seem to have a significant positive effect
In polymer clay nanocomposites, where the tortuosity dominates on the permeability of gases in the nanocomposites. For more than
the fractional free volume variations, the percentage reduction on 10 wt.% silica nanoparticles in PTMSP, this effect dominates over the
gas permeability is 20% – 90%. In polyimide/MMT nanocomposites a negative effect of the reduced diffusivity presented by Eq. (7.3) with
reduction of 83% in oxygen permeabilityis observed for only 1 wt.% α = 1. For lower filler content, the higher tortuosity seems to be the
MMT [33 •], while for polyamide 6/MMT a reduction in the order of dominant factor. The CO2 permeability in PTMSP containing 40 vol.%
77% was reported [34] for 2 vol.% of MMT. A 76% reduction of the He MgO particles is 4.5 times higher than that of the unfilled polymer.
permeability was measured for 8 wt.% MMT in polyurethane/MMT Up to 75% of this increase is due to the increase in the CO2 diffusion
nanocomposites [35 •], while the presence of 5 wt.% MMT in polyacry- coefficient [46]. In Poly(4-methyl-2-pentene)/FS the increased per-
late decreased the permeability by up to 85% [36]. The addition of meability is ascribed entirely to the diffusion coefficient, since
4 wt.% of MMT in Nylon-6 reduced the oxygen permeability by 50% fumed silica has little impact on gas solubility [47]. This is evidence
[37] and in Epoxy/MMT the reduction is of the order of 86% for for the effect of the free volume hole size distribution on gas perme-
7 wt.% MMT [38]. An appreciable reduction in gas permeability com- ability. Broader distribution results in higher gas diffusion coefficient.
pared to the unfilled polymer film was reported in PP/MMT nano- Enhanced gas permeability is found in PMP/FS [48] and in ethylene
composites [39], where both the diffusivity and the solubility were vinyl acetate/silica [49] nanocomposites a.o.
reduced by the presence of the filler. A reduction in gas permeability due to the incorporation of spher-
There are also cases, where the permeability is affected apprecia- ical nanoparticles has also been reported. In PVC/SiO2 nanocompo-
bly by free volume changes. The nitrogen permeability of SBR con- sites the oxygen and water permeability rates decreased with 3 wt.%
taining 40 phr rectorite decreased by 69% compared to the pure of filler [50]. As it was mentioned above, for low particle loading the
rubber [23]. This reduction, however, cannot be attributed solely to tortuosity becomes the dominant factor that determines the perme-
the tortuosity, since the fractional f.v. of the nanocomposite was ation properties, even though this effect is much weaker for nano-
also reduced by 27%. The two factors that affect f.v. and permeation spheres than for clay nanolayers. Similar results are observed in
properties may be competitive and the gas permeability can in polyurethane/silica systems, where CO2 permeability is reduced by
some cases show a small decrease or remain constant. Such a case about 35% for samples containing 12 vol.% nanoparticles [51].
G. Choudalakis, A.D. Gotsis / Current Opinion in Colloid & Interface Science 17 (2012) 132–140 139

Summarising the effect of nanoparticles on the mass transport [4] Xie L, Gidley D, Hristov H, Yee A. Evolution of nanometer voids in polycarbonate
under mechanical stress and thermal expansion using spectroscopy. J Polym Sci,
properties of the nanocomposites, one could state that when spheri- Part B: Polym Phys 1995;33:77–84.
cal nanoplatelets are added to a polymer, then the gas permeability [5] Garcia A, Iriarte M, Uriarte C, Etxeberria A. Study of the relationship between transport
will most likely increase, since the tortuosity that the spheres impose properties and free volume based in polyamide blends. J Membr Sci 2006;284:173–9.
[6] Choudalakis G, Gotsis A, Schut H, Picken S. The free volume in acrylic
is small, while the dominant factor is the increase of the free volume •
resin/laponite nanocomposite coatings. Eur Polym J 2011;47:264–72.PALS mea-
at the interfaces and the interstitial regions within the agglomerates. surements of free volume in polymer/clay nanocomposties and analysis of the influ-
On the other hand, when exfoliated nano-clay is added to the poly- ence of its hole size distribution on gas transport.
[7] Thornton A, Nairn K, Hill A, Hill J. New relation between diffusion and free
mer, then the permeability is expected to be reduced significantly. •
volume: predicting gas diffusion. J Membr Sci 2009;338:29–37.An empirically re-
The resulting tortuosity of the diffusion path is the major cause for lation between gas diffusion and free volume derived using a large database of
this enhancement. The free volume may increase somewhat, due to 105 polymers with additional high free volume polymers.
[8] Jean Y, Yuan J, Liu J, Deng Q, Yang H. Correlations between gas permeation and
the presence of the expansive interfacial regions, but this is a minor
free-volume hole properties probed by positron annihilation spectroscopy. J
effect and, in most cases, cannot overrule that of the tortuosity. Of Polym Sci, Part B: Polym Phys 1995;33:2365–71.
course, this is the case when the clay particles are well exfoliated [9] Wang Z, Wang B, Yang Y, Hu C. Correlations between gas permeation and free-
and properly oriented. At lower degrees of exfoliation the situation volume hole properties of pu membranes. Eur Polym J 2003;39:2345–9.
[10] Consolati G, Genco I, Pegoraro M, Zanderighi L. Positron annihilation lifetime in
can point towards that of the isometric nano-fillers. poly[1-(trimethyl-silyl)propine]: free volume determination and time depen-
dence of permeability. J Polym Sci, Part B: Polym Phys 1996;34:357–67.
[11] Winberg P, DeSitter K, Dotremont C, Mullens S, Vankelecom I, Maurer F. Free vol-
8. Conclusions ume and interstitial mesopores in silica filled poly(1-trimethylsilyl-1-propyne)
nanocomposites. Macromolecules 2005;38:3776–82.
The free volume characteristics of polymers are altered signifi- [12] Kobayashi Y, Zheng W, Meyer E, McGervey J, Jamieson A, Simha R. Free volume
and physical aging of poly(vinyl acetate) studied by positron annihilation. Macro-
cantly by the incorporation of nanoparticles. The changes have been molecules 1989;22:2302–6.
verified experimentally by Positron Annihilation Lifetime Spectrosco- [13] Wang Y, Nakanishi H, Jean Y. Positron annihilation in amine-cured epoxy polymers-
py for various systems and types of fillers. The creation of interfacial pressure dependence. J Polym Sci, Part B: Polym Phys 1990;28:1431–41.
[14] Ito K, Nakahishi H, Ujihira Y. Extension of the equation for the annihilation life-
layers between the organic and inorganic components, the immobili- time of ortho-positronium at a cavity larger than 1 nm in radius. J Phys Chem B
sation of chain segments, the formation of interstitial cavities, and the 1999;103:4555–8.
changes of hole size distribution are the main effects nanoparticles [15]

Zhang J, Yang M, Maurer F. Effect of TiO2 formation on the free volume properties
of electrospun PMMA nanohybrids. Macromolecules 2011;44:5711–21.The influ-
have on the polymer, that can alter the free volume properties of
ence of the hydroxyl groups present in titania due to the strong inhibition effect
the material. that they cause on the o-Ps formation.
The transport properties of polymer nanocomposites are deter- [16] Winberg P, Eldrup M, Maurer F. Nanoscopic properties of silica filled polydi-
methylsiloxane by means of positron annihilation lifetime spectroscopy. Polymer
mined by the free volume changes and the tortuosity of the diffusion
2004;45:8253–64.
path that the presence of the particles causes. These factors can follow [17] Wang Y, Wu Y, Zhang H, Wang B, Wang Z. Free volume of montmorillonite/styrene-
the same trend or be competitive. The final effect on the transport co- butadiene rubber nanocomposites estimated by positron annihilation lifetime spec-
efficients depends on which of them dominates. In polymer clay troscopy. Macromol Rapid Commun 2004;25:1973–8.
[18] Dreiss C, Cosgrove T, Benton N, Kilburn D, Alam M, Schmidt R, et al. Effect of cross-
nanocomposites the tortuosity effect overcomes the free volume var- linking on the mobility of PDMS filled with polysilicate nanoparticles: positron
iations and finally determines the transport coefficients. In the case of lifetime, rheology and NMR relaxation studies. Polymer 2007;48:4419–28.
spherical nanoparticles the free volume properties dominate over the [19] Zhang M, Fang P, Zhang S, Wang B, Wang S. Study of structural characteristics of
HDPE/CaCO3 nanocomosites by positrons. Radiat Phys Chem 2003;68:565–7.
tortuosity, controlling, thus, the gas permeation rates. [20] Debowska M, Girulska JR, Jezierski A, Pasternak A, Pozniak R. Positron annihila-
An important role for the determination of the transport properties tion in carbon black-polymer composites. Radiat Phys Chem 2000;58:575–9.
of polymer nanocomposites seems to be played also by the free volume [21] Chen H, Jean Y, Lee L, Yang J, Huang J. Positron annihilation study in inorganic-
polymer nano-composites. Phys Status Solidi C 2009;6:2397–400.
hole size distribution. The total amount of the free volume itself does [22] Harms S, Ratzke K, Faupel F, Schneider G, Willner L, Richter D. Free volume of in-

not provide enough information about the free volume network terphases in model nanocomposites studied by positron annihilation lifetime
which determines the diffusion process. The higher the relative number spectroscopy. Macromolecules 2010;43:10505–11.The enhancement of the o-Ps
lifetime as function of filler loading is ascribed to out-diffusion of the positrons
of smaller holes, the lower the mobility of gas molecules. Therefore, the
from the nanoparticle surface. While the findings are interesting this interpreta-
changes in the free volume hole size distribution can strongly influence tion is under debate.
the diffusion coefficient, even if the total free volume of the system re- [23] Wang Z, Wang B, Qi N, Zhang H, Zhang L. Influence of fillers on free volume and
gas barrier properties in styrene-butadiene rubber studied by positrons. Polymer
mains unaffected. They may control, thus, the permeability of the sys-
2005;46:719–24.
tem, since they have little effect on the solubility. [24] Wang S, Liu L, Fang P, Chen Z, Wang H, Zhang S. Microstructure of polymer-clay
The search of high/low gas permeability nanocomposite materials nanocomposites studied by positrons. Radiat Phys Chem 2007;76:106–11.
requires more than information on the chemical compatibility between [25] Fang P, Chen Z, Zhang S, Wang S, Wang L, Feng J. Microstructure and thermal
properties of ethylene-(vinyl acetate) copolymer/rectorite nanocomposites.
the components. While the later may be adequate for the improvement Polym Int 2006;55:312–8.
of their mechanical performance, their transport properties are more [26] Jia Q, Zheng M, Zhu Y, Li J, Xu C. Effects of organophilic montmorillonite on hydrogen
complicated and require the study of the physicochemical conse- bonding, free volume and glass transition temperature of epoxy resin/polyurethane
interpenetrating polymer networks. Eur Polym J 2007;43:35–42.
quences that the nanoparticles cause on the polymer matrix, where [27] Fang Z, Xu Y, Tong L. Free-volume hole properties of two kinds thermoplastic
they are dispersed. nanocomposites based on polymer blends probed by positron annihilation life-
time spectroscopy. J Appl Polym Sci 2006;102:2463–9.
• [28] Winberg P, Eldrup M, Pedersen N, Van Es M, Maurer F. Free volume sizes in inter-
References calated polyamide 6/clay nanocomposites. Polymer 2005;46:8239–49.
[29] Paranchos C, Soares B, Machado J, Windmoller D, Pessan L. Microstructure and
[1] Robeson L, Paul D. Polymer nanotechnology: nanocomposites. Polymer 2008;49: free volume evaluation of poly(vinyl alcohol) nanocomposite hydrogels. Eur
3187–204. Polym J 2007;43:4882–90.
[2] Thran A, Faupel G. Correlation between fractional free volume and diffusivity of [30] Yu D, Wang B, Feng Y, Fang Z. Investigation of free volume, interfacial, and toughening
gas molecules in glassy polymers. J Polym Sci, Part B: Polym Phys 1999;37: behavior for cyanate ester/bentonite nanocomposites by positron annihilation. J Appl
3344–58. Polym Sci 2006;102:1509–15.
[3]

Arnold J. A free-volume hole-filling model for the solubility of liquid molecules in [31] Choudalakis G, Gotsis A. Permeability of polymer/clay nanocomposites: a review.
glassy polymers 1: Model derivation. Eur Polym J 2010;46:1131–40.The model is Eur Polym J 2009;45:967–84.
based on the principal of successive filling of pre-existing holes in a polymer using [32] Cui L, Yeh JT, Wang K, Tsai FC, Fu Q. Relation of free volume and barrier properties
a Gaussian distribution of hole sizes. The contributing terms include elastic expan- in the miscible blends of poly(vinyl alcohol) and nylon 6-clay nanocomposites
sion of small holes and bonding between liquid and polymer in larger holes. film. J Membr Sci 2009;327:226–33.
[33] Min U, Kim J, Chang J. Transparent polyimide nanocomposite films: thermo-
• •
of special interest optical properties, morphology, and gas permeability. Polym Eng Sci 2011;51:
140 G. Choudalakis, A.D. Gotsis / Current Opinion in Colloid & Interface Science 17 (2012) 132–140

2143–150.A remarkable reduction of the oxygen permeability by addition of only model takes the interfacial layer effect into account by incorporating an interaction
1wt % of nanoparticles. strength parameter.
[34] Fasihi M, Abolghasemi M. Oxygen barrier and mechanical properties of [43]

Pryamitsyn V, Hanson B, Ganesan V. Coarse-grained simulations of penetrant trans-
masterbatch-based PA-6/nanoclay composite films. J Appl Polym Sci in press; port in polymer nanocomposites. Macromolecules 2011;44:9839–51.A simulation
doi:10.1002/app.35647. study that examines the influence of the interfacial layers and the polymer matrix
[35]

Maji P, Das N, Bhowmick A. Preparation and properties of polyurethane nanocom- dynamics on the transport properties of polymer nanocomposites.
posites of novel architecture as advanced barrier materials. Polymer 2010;51: [44] Manninen A, Naguib H, Nawaby A, Day M. CO2 sorption and diffusion in poly-
1100–10.The effect of crosslinking density on the helium barrier properties of methyl methacrylate-clay nanocomposites. Polym Eng Sci 2005;45:904–14.
polyurethane has been demonstrated. [45] Matteucci S, Kusuma V, Swinnea S, Freeman B. Gas permeability, solubility and
[36] Herrera-Alonso J, Sedlakova Z, Marand E. Gas transport properties of polyacrylate/clay diffusivity in 1,2-polybutadiene containing brookite nanoparticles. Polymer
nanocomposites prepared via emulsion polymerization. J Membr Sci 2010;363:48–56. 2008;49:757–73.
[37] Sadeghi F, Fereydoon M, Ajji A. Rheological, mechanical and barrier properties of [46] Matteucci S, Kusuma V, Swinnea S, Freeman B. Gas transport properties of MgO
multilayer nylon/clay nanocomposite film. Adv Polym Technol in press; doi:10. filled poly(1-trimethylsilyl-1-propyne) nanocomposites. Polymer 2008;49:
1002/adv.20270. 1659–75.
[38] Dai C, Li P, Yeh J. Comparative studies for the effect of intercalating agent on the [47] Merkel T, Freeman B, Spontak R, He Z, Pinnau I, Meakin P, et al. Sorption, trans-
physical properties of epoxy resin-clay based nanocomposite materials. Eur port, and structural evidence for enhanced free volume in poly(4-methyl-2-
Polym J 2008;44:2439–47. pentyne)/fumed silica nanocomposite membranes. Chem Mater 2003;15:
[39] Villaluenga J, Khayet M, Lopez-Manchado M, Valentin J, Seoane B, Mengual J. Gas 109–23.
transport properties of polypropylene/clay composite membrane. Eur Polym J [48] Shao L, Samseth J, Hagg M. Crosslinking and stabilization of nanoparticle filled
2007;43:1132–43. pmp nanocomposite membranes for gas separation. J Membr Sci 2009;326:
[40] Jacquelot E, Espuche E, Gerard J, Duchet J, Mazabraud P. Morphology and gas bar- 285–92.
rier properties of polyethylene-based nanocomposites. J Polym Sci, Part B: Polym [49] Sadeghi M, Khanbabaei G, Dehaghani A, Sadeghi M, Aravand M, Akbarzade M,
Phys 2006;44:431–40. et al. Gas permeation properties of ethylene vinyl acetate-silica nanocomposite
[41]

Qian Y, Lindsay C, Macosko C, Stein A. Synthesis and properties of vermiculite- membranes. J Membr Sci 2008;322:423–8.
reinforced polyurethane nanocomposites. Appl Mater Interfaces 2011;3: [50] Zhu A, Cai A, Zhang J, Jia H, Wang J. PMMA-grafted-silica/PVC nanocomposites:
3709–17.While the tensile modulus and tensile strength showed a remarkable mechanical performance and barrier properties. J Appl Polym Sci 2008;108:
enhancement, the improvement of the gas barrier properties was weaker, due 2189–96.
to the free volume in the interfacial regions. [51] Sadeghi M, Semsarzadeh M, Barikani M, Chenar M. Gas separation properties of
[42]

Xiao J, Huang Y, Manke C. Computational design of polymer nanocomposite coatings: polyether-based polyurethane-silica nanocomposite membrane. J Membr Sci
a multiscale hierarchical approach for barrier property prediction. Ind Eng Chem Res 2011;376:188–95.
2010;49:7718–27.A numerical multiscale modeling and simulation method is devel-
oped in order to predict the relative permeability in a polymer nanocomposite. The

Вам также может понравиться