Anda di halaman 1dari 66

BARRIER POLYMERS

Introduction
Barrier polymers can be broadly dened as macromolecules having the ability to
signicantly restrict the passage of gases, vapors, and liquids. Since all polymers
restrict the transport of penetrants to some degree and the barrier performance
of polymers to different penetrants depends on a variety of factors, it is difcult to
provide a concise, objective denition. In a practical sense, however, the denition
of a barrier polymer depends upon the end-use requirements, and a material that
provides sufcient barrier for a particular application can be considered to be a
barrier polymer for that purpose. In the present discussion, polymers that have
resistance to transport of gases, vapors, and liquids as one of their key attributes
will be considered to be barrier polymers.
Polymers have found wide acceptance as alternatives to traditional materials
such as glass, paper, and metals, in food, beverage, and other packaging indus-
tries. Akey characteristic of glass and metals as packaging materials is their total
barrier to transport of gases and vapors. While polymers can provide an attrac-
tive balance of properties such as exibility, toughness, light weight, formability,
and printability, they do allow the transport of gases and vapors to some extent.
Unfortunately, an inexpensive, recyclable polymeric material possessing high bar-
rier properties to every gas or vapor in addition to good mechanical, thermal, and
optical properties is not available. For this reason, the selection of a barrier poly-
mer for a particular application typically involves tradeoffs between permeation,
mechanical, and aesthetic properties as well as economic and recycling consider-
ations. Additionally, there is an ongoing interest in optimizing property sets of
barrier polymers to provide an efcient and economical method for packaging and
for extending the shelf life of packaged foods and beverages.
198
Encyclopedia of Polymer Science and Technology. Copyright John Wiley & Sons, Inc. All rights reserved.
Vol. 5 BARRIER POLYMERS 199
This article discusses various types of barrier polymers and structures, their
permeability characteristics and the underlying phenomena involved, measure-
ment techniques, ways to predict and improve barrier properties, and current as
well as potential future applications for barrier polymers.
Fundamentals of Permeation in Polymers
The permeability coefcient, or simply permeability, characterizes the steady-
state rate of mass transport of penetrant molecules through polymers. In a dense
polymer lm, the permeability P is dened as the molar ux of penetrant through
the polymer relative to a xed coordinate system N
A
normalized by the lm thick-
ness L and the difference between the upstream (p
2
) and downstream (p
1
) partial
pressures (1):
P=
N
A
L
( p
2
p
1
)
(1)
Accordingly, permeability has dimensions of quantity of penetrant (either
mass or moles) times thickness divided by area, time, and pressure. Several units
have been used to report permeability of gases and water vapor in the literature.
In the United States, a commonly used unit for permeability of gases in barrier
polymers is (cm
3
(STP)mil)/(100 in.
2
dayatm). Table 1 provides conversion factors
for several permeability units, including the SI unit (molm)/(m
2
sPa), which is
commonly preferred in technical encyclopedias.
The steady-state transport properties of water vapor in barrier polymers are
characterizedby water vapor transmissionrate (WVTR). The dimensions of WVTR
are quantity of water transmitted through a lm times thickness divided by area
and time, and a common unit for WVTR is (gmil)/(100 in.
2
day). Table 2 provides
conversion factors for some WVTR units, including the SI unit (molm)/(m
2
s).
WVTR can be converted to water vapor permeability by dividing by the water
partial pressure difference (which can be calculated from the specied relative
humidity and temperature).
Penetrant transport through polymers is described by the so-called solution-
diffusion model (1). According to this model, permeation through a at sheet or
lm occurs in three steps: penetrant dissolves into the upstream (ie, the high
partial pressure or high thermodynamic activity) side of the lm, diffuses through
the lm, and desorbs from the downstream (ie, the low partial pressure or low
thermodynamic activity) side of the lm. The rate-limiting step in this process
is diffusion through the lm. In one dimension, penetrant diffusion through a
polymer typically follows Ficks law:
N
A
= D
dC
dx
(2)
where D is the effective diffusion coefcient for the penetrant in the polymer and
dC/dx is the local concentration gradient of the penetrant.
When the downstream side penetrant partial pressure and concentration
are negligible relative to those on the upstream face of the lm, using Ficks law of
Table 1. Table of Common Gas Permeability Units with Conversion Factors
a
To obtain
Given Barrer
cccm
cm
2
scm Hg
cccm
cm
2
satm
cccm
cm
2
sPa
molcm
cm
2
scm Hg
molm
m
2
sPa
ccmil
100 in.
2
dayatm
cc20m
m
2
dayatm
Barrer 1 1.00 10
10
7.60 10
9
7.501 10
14
4.461 10
15
3.346 10
16
1.668 10
2
3.283 10
3
cccm
cm
2
scm Hg
1.00 10
10
1 76 7.501 10
4
4.461 10
5
3.346 10
6
1.668 10
12
3.283 10
13
cccm
cm
2
satm
1.316 10
8
1.316 10
2
1 9.869 10
6
5.87 10
7
4.403 10
8
2.195 10
10
4.32 10
11
cccm
cm
2
sPa
1.333 10
13
1.333 10
3
1.013 10
5
1 5.948 10
2
4.461 10
3
2.224 10
15
4.377 10
16
molcm
cm
2
scm Hg
2.241 10
14
2.241 10
4
1.703 10
6
16.81 1 7.501 10
2
3.738 10
16
7.359 10
17
molm
m
2
sPa
2.988 10
15
2.988 10
5
2.271 10
7
2.241 10
2
13.33 1 4.984 10
17
9.81 10
18
ccmil
100 in.
2
dayatm
5.996 10
3
5.996 10
13
4.557 10
11
4.497 10
16
2.675 10
17
2.007 10
18
1 19.68
cc20m
m
2
dayatm
3.046 10
4
3.046 10
14
2.315 10
12
2.285 10
17
1.359 10
18
1.019 10
19
5.08 10
2
1
a
Given permeability in the units shown in one element of the rst column, convert it to the units shown in one element of the rst row by multiplying the
original permeability by the factor at the intersection of the row and column of interest. For example, a value of 2 (cc20 m)/(m
2
dayatm) is equal to (2
0.0508) or 0.1 (ccmil)/(100 in.
2
dayatm). In this table and throughout the article cc (or cm
3
) has been used to denote cubic centimeters of gas as measured
at standard temperature and pressure (STP) conditions, which are 0

C and 1 atm.
2
0
0
Vol. 5 BARRIER POLYMERS 201
Table 2. Table of Common WVTR Units with Conversion Factors
a
To obtain
Given
molm
m
2
s
gmil
100 in.
2
day
gcm
m
2
day
molm
m
2
s
1 3.95 10
9
1.55 10
8
gmil
100 in.
2
day
2.53 10
10
1 3.94 10
2
gcm
m
2
day
6.45 10
9
25.4 1
a
Given a WVTR value in units shown in one element of the rst column,
convert it to the units shown in one element of the rst row by multiplying
the original WVTR value by the factor at the intersection of the row and
column of interest. For example, a value of 2 (gcm)/(m
2
day) is equal to
(2 25.4) or 50.8 (gmil)/(100 in.
2
day).
diffusion (eq. 2), permeability can be expressed as product of the effective diffusion
coefcient D and the solubility coefcient S, which is the ratio of the equilibrium
penetrant concentration in the polymer at the upstream side of the lm divided
by the penetrant partial pressure or activity in the contiguous phase (1):
P= DS (3)
According to equation 2, the diffusion coefcient is a kinetic term characterizing
the mass ux of penetrant through a polymer lm in response to a concentration
gradient (1). Diffusion coefcients have units of (length)
2
/time, and are often ex-
pressed in cm
2
/s. The solubility or partition coefcient is a thermodynamic factor
that links the equilibrium penetrant concentration in the polymer, C, with the
penetrant partial pressure contiguous to the polymer surface, p (1):
C=S p (4)
When the penetrants of interest are vapors, liquids, or solids, the partial pressure
is often replaced by penetrant activity. For an ideal gas, penetrant activity is equal
to the ratio of penetrant partial pressure to its saturation vapor pressure (1). For
nonideal systems thermodynamic models must be used to estimate penetrant
activity (2).
The diffusion process of penetrants in polymers can be broadly classied into
two categories: Fickian (which obeys the Ficks law of diffusion) and non-Fickian.
Penetrants in rubbery polymers and at low activities in glassy polymers typically
exhibit Fickian behavior (3). The signature of Fickian diffusion in a thin polymer
lm contacted on both faces with a constant partial pressure (or activity) of pen-
etrant is a weight increase due to penetrant absorption that is initially a linear
function of the square root of the contact time and then asymptotically approaches
a xed equilibriumvalue (3). For Fickian diffusion-controlled kinetics of penetrant
transport in a plane lm whose thickness (L) is much smaller than its width or
length, the time required to reach steady state in a permeation experiment, t
SS
,
is given by (3):
t
ss
=
L
2
D
(5)
202 BARRIER POLYMERS Vol. 5
The diffusion coefcients of large penetrants (eg, avor and aroma compounds) in
barrier polymers, which can be of the order of 10
14
cm
2
/s or less, coupled with a
typical lm thickness of 10 m result in extremely large time scales (over 3 years)
to reach steady-state transport. Hence, over the shelf life of packaged products
(typically of the order of several months), avor and aroma compounds may never
achieve equilibrium or steady-state conditions. In such cases, the steady-state
permeability does not provide sufcient information to predict package shelf life.
As discussed later, more detailed knowledge of penetrant solubility and diffusivity
is required to accurately predict migration of such compounds.
In glassy polymers, deviations from Fickian behavior can occur. These devia-
tions are generally believed to arise as a consequence of the nite rate of polymer
structure reorganization in response to penetrant-induced swelling during the
sorptiondiffusion process (4). An example of so-called non-Fickian diffusion be-
havior is the penetrant sorbing into the polymer in two stages, an initial Fickian-
like stage followed by a protracted, slowdrift toward the nal equilibriumsorption
value (5). In such cases, the time required to achieve steady-state transport may
be much higher than that predicted based on equation 5. This type of diffusion
behavior is often observed when organic vapors at relatively high activity sorb
into amorphous glassy polymers (6). For example, toluene vapors exhibit Fickian
diffusion behavior in amorphous poly(vinyl chloride) (PVC) lms at activities up
to 0.4 and show increasingly non-Fickian behavior at higher activities (7).
Mechanism of Penetrant Transport in Dense Polymers
The rate-limiting step for penetrant diffusion is the creation of transient gaps in
the polymer matrix via local scale polymer segmental dynamics involving several
polymer chains (8). Penetrant molecules vibrate inside local cavities inthe polymer
matrix at frequencies much higher than the frequency of polymer chain motion
required to open a gap of sufcient size to accommodate the penetrant. These steps
are shown schematically in Figure 1. In Figure 1a, a penetrant molecule is shown
dissolved in a polymer matrix. The penetrant vibrates inside a gap or molecular
scale cavity in the polymer matrix at very high frequency (ca 10
12
vibrations/s or
1 vibration/ps) (8). The polymer molecules do not occupy the entire volume of the
polymer sample. Because of packing inefciencies and polymer chain molecular
motion, some of the volume in the polymer matrix is empty or free and this
so-called free volume is redistributed continuously as a result of the random,
thermally stimulated molecular motion of the polymer segments (1).
In Figure 1b, local polymer segmental motion has opened a connecting chan-
nel between two free-volume elements in the polymer matrix and the penetrant
molecule can, as a result of its own Brownian motion, explore the entire corridor
between the initial free-volume element which it occupied and the second free-
volume element which is connected to it via the opening of a transient gap in
the polymer matrix. Eventually, local segmental motion of the polymer segments
closes the connection between the two free-volume elements and if the penetrant
happens to be away from its original position, as shown in Figure 1c, when the
gap in the polymer matrix is closed, the penetrant will be trapped in another free-
volume element in the polymer matrix and will have executed a diffusion step. The
Vol. 5 BARRIER POLYMERS 203
Local, penetrant-sized gap
in polymer due to chain
motion
Penetrant dissolved in
polymer matrix
(a)
Opening of transient,
penetrant-sized gap in
polymer matrix due to
cooperative segmental
motion
(b)
Gap closes behind penetrant,
which has executed a
diffusion step to a new region
of polymer matrix
(c)
Fig. 1. Schematic depicting mechanism of penetrant transport in polymers (8).
process shown in Figure 1 has been called the Red Sea mechanism of penetrant
transport in polymers (8).
Figure 2 shows a schematic of two polymer chains undergoing coordinated
local segmental motion as a result of random, thermally stimulated movements
of the polymer chains to open a gap between the polymer chains of sufcient
size to permit passage of a penetrant molecule from one free-volume element to
another (9). This schematic emphasizes that the polymer segmental motion is the
rate-controlling step in penetrant diffusion.
Figure 3 provides a very simplistic schematic of the molecular processes in-
volved in the local segmental motions of polymers that contribute to the formation
of transient gaps in the polymer matrix important for penetrant diffusion. This
gure shows what is believed to be a typical example of intramolecular coopera-
tive local segmental motion of the polymer backbone of a polyethylene chain. This
204 BARRIER POLYMERS Vol. 5
Prior to jump
Penetrant
Polymer chain
Activated state
After jump completed
l
d
l
d
Fig. 2. Schematic depicting polymer chain position before, during, and after a diffusion
step by a penetrant molecule (9).
so-called crankshaft motion requires the cooperation of several adjacent ethylene
units and can act to create gaps in the polymer matrix of sufcient size to accom-
modate small penetrant molecules. It should be emphasized that the detailed un-
derstanding of the molecular level motions in polymers that contribute to diffusion
is evolving rapidly as a result of more detailed atomistic simulation of penetrant
transport in polymers. As a result, more realistic descriptions of the important
molecular processes for penetrant diffusion in polymers should be expected in the
near future.
The motion of polymer segments to produce a gap of sufcient size to accom-
modate a penetrant molecule is much slower than the vibration of the penetrant
in a gap in the polymer matrix. For example, computer simulations of oxygen dif-
fusion in a polyimide reveal that the oxygen molecules execute a diffusion step
approximately once every nanosecond (1 ns = 1000 ps) (10). The rate of produc-
tion of gaps of sufcient size to accommodate penetrant molecules decreases with
increasing size of the penetrant. That is, there are fewer gaps produced per unit
time in a polymer matrix of sufcient size to accommodate methane (kinetic di-
ameter = 0.38 nm), for example, than there are for hydrogen (kinetic diameter =
0.289 nm).
Computer simulation results of diffusion in poly(2,6-dimethyl-1,4-phenylene
oxide) (PPO) have been reported (11). Figures 4 and 5 present the results for
the displacement of an oxygen molecule and a nitrogen molecule, respectively, as a
Vol. 5 BARRIER POLYMERS 205
1
2
3
4
5
6
7
1
2
3
4
5
Fig. 3. Crankshaft motionAn example of local segmental dynamics in polyolens. Each
ball represents, for example, a methylene (ie, CH
2
) unit in PE and the solid lines represent
the covalent chemical bonds between neighboring CH
2
groups. Such molecular motions are
believed to be important in providing pathways for diffusion of small molecules in polymers.
The crankshaft motion requires the simultaneous rotation of several contiguous methylene
units about bonds 1 and 7 or bonds 1 and 5 (9).
0 100 200 300 400 500 600 700 800 900 1000
O
2
Diffusion jumps
0
0.5
1.0
1.5
2.0
2.5
3.0
D
i
s
p
l
a
c
e
m
e
n
t
,

n
m
Time, ps
Fig. 4. Computer simulation of the displacement of an oxygen molecule in poly(2,6-
dimethyl-1,4-phenylene oxide) as a function of time (11).
206 BARRIER POLYMERS Vol. 5
0
0.5
1.0
1.5
2.0
2.5
3.0
0 100 200 300 400 500 600 700 800 900 1000
N
2
Diffusion
jump
Time, ps
D
i
s
p
l
a
c
e
m
e
n
t
,

n
m
Fig. 5. Computer simulation of the displacement of a nitrogen molecule in poly(2,6-
dimethyl-1,4-phenylene oxide) as a function of time (11).
function of time. There is about one diffusion jump for oxygen (kinetic diameter =
0.346 nm) every 300350 ps but only one diffusion step for nitrogen (kinetic di-
ameter = 0.364 nm) over the entire 1000 ps duration of the computer simulation
(11). The oxygen molecule spends most of its time rattling within a small cage (or
free-volume element) with average displacements of the order of 0.2 nm or so. The
diffusion jumps occurring approximately every 300350 ps involve displacements
of the oxygen atom of the order of 0.40.5 nm. As shown in Figure 5, the diffusion
jump length for nitrogen is longer (approximately 1 nm) and the jumps occur less
frequently. However, these results are obtained for very short periods of time ow-
ing to computational limitations, and very long simulations would be required to
generate precise estimates of jump lengths and jump frequencies. Because of the
extremely demanding computational resources required for such molecular-level
simulations, they are only now becoming possible for small molecule migration in
relatively permeable polymers. As yet, computers are not fast enough to provide
realistic simulations of phenomena such as migration of large avor molecules in
high barrier polymers.
Factors Affecting Permeability, Diffusivity, and Solubility
Free Volume. The dependence of penetrant transport properties on chain
packing in polymers is often described using correlations involving the fractional
free volume (FFV) of polymers. FFV is the fraction of volume in a polymer that is
available to assist in penetrant transport, and does not include volume occupied by
polymer molecules andvolume inthe polymer matrix that is otherwise unavailable
Vol. 5 BARRIER POLYMERS 207
for penetrant transport. It is often estimated using group contribution methods.
One popular method for estimating FFV is based on the following expression (1):
FFV=
V V
0
V
(6)
where V is the polymer specic molar volume and V
0
is the so-called occupied
volume that is not available to assist in penetrant transport. The occupied volume
is usually estimated by Bondis method as follows (12):
V
0
=1.3V
w
(7)
where V
w
is van der Waals volume of the molecule. A good estimate of V
w
can be
obtained from bond radii, van der Waals radii of constituent atoms, and geometric
factors. Bond radii are nearly constant from one molecule to another since the
same chemical bond will always have the same radius. The most complete list of
V
w
values is available in the compilation given in Reference 13. The dependence
of diffusion coefcients on FFV can be expressed as (14):
D= Aexp
_
B
FFV
_
(8)
where A and B are empirical constants. The higher the FFV, the larger the dif-
fusion coefcient. Figure 6 shows the effect of free volume on gas diffusion coef-
cients in a series of substituted polysulfones (15).
The dependence of solubility on FFV is usually weaker than that of diffusiv-
ity, especially in amorphous polymers (16). Therefore, permeability often follows
a similar dependence on free volume as penetrant diffusivity. Attempts have been
made to correlate FFV of polymers with gas permeability (17,18). As shown in
Figure 7 (17), a nearly linear correlation was found to exist between the logarithm
of oxygen permeability coefcients and the inverse of FFV in several families of
amorphous, glassy polymers and high barrier liquid crystalline polymers. Many
barrier polymers are glassy materials, since their use temperature is below their
glass-transition temperature. In glassy polymers, which are nonequilibrium ma-
terials, free volume can be altered to some extent by the processing history of
the sample (19). For instance, higher rates of cooling create higher free volume
in the glassy state, and vice versa. A more effective way to alter free volume is to
vary the chemical structure by, for example, adding or removing pendant groups
on the polymer backbone (19). The presence of polar groups with low specic vol-
umes can reduce the free volume (and hence, penetrant diffusion coefcients) by
facilitating more efcient packing of the polymer chains due to stronger interac-
tions between them (19). For example, barrier polymers such as polyacrylonitrile
(PAN) and poly(vinyl alcohol) (PVOH) have polar pendant groups, which lead to
strong energetic interactions between the polymer chains, efcient polymer chain
packing, low free volume, and, in turn, low permeability coefcients.
Free volume in polymers can be measured using probes such as elec-
trochromic, photochromic, and uoroscent probes, as well as xenon nuclear mag-
netic resonance, small-angle x-ray scattering, density measurements, andpositron
208 BARRIER POLYMERS Vol. 5
10
10
10
9
10
8
10
7
6 8 7 9 10
D
i
f
f
u
s
i
o
n

c
o
e
f
f
i
c
i
e
n
t
s
,

c
m
2
/
s
1/FFV
CO
2
CH
4
O
2
N
2
Fig. 6. Correlation of gas diffusion coefcients with inverse of polymer fractional free
volume in a series of substituted polysulfones (15).
annihilation lifetime spectroscopy (pals) (20). Each method has its strengths and
limitations, and a simple, direct measure of FFV is not available. The pals tech-
nique has, however, emerged inthe past several years as a valuable nondestructive
probe of free volume in polymers (20,21). PALS uses orthoPositronium (oPs) as a
probe of free volume in the polymer matrix. oPs resides in regions of reduced elec-
tron density, such as free-volume elements, that typically range in radius from
0.2 to 0.4 nm. This range of cavity radii compares well with nonbonded inter-
atomic distances in polymers and molecular radii of diffusing penetrants (20).
PALS permits an estimate of both the size and concentration of free-volume el-
ements in the polymer matrix. Transport properties of barrier polymers, copoly-
mers, and polymer blends have been well correlated with FFVas measured by pals
(2026).
Temperature. The temperature dependence of permeability and diffusiv-
ity are usually modeled using Arrhenius equations of the following forms (1):
P= P
0
exp
_
E
p
RT
_
(9)
D= D
0
exp
_
E
D
RT
_
(10)
Vol. 5 BARRIER POLYMERS 209
10
2
10
1
10
0
10
1
10
2
10
3
10
4
1/FFV
5 6 9 10 7 8
O
x
y
g
e
n

p
e
r
m
e
a
b
i
l
i
t
y
,

(
c
m
3


m
i
l
)
/
(
1
0
0

i
n
.
2


d
a
y


a
t
m
)
Fig. 7. Correlation of oxygen permeability with inverse of polymer fractional free volume
for several families of amorphous, glassy, and liquid crystalline polymers (17). Polystyrene
(35

C), polycarbonates (35

C), polyesters (30

C), polyamides (25

C), liquid crys-


talline polymers (35

C).
where E
p
and E
D
are activation energies for permeation and diffusion, and P
0
and
D
0
are preexponential factors. The effect of temperature on solubility is usually
expressed by a vant Hoff relationship (1):
S=S
0
exp
_
H
S
RT
_
(11)
where S
0
is a preexponential factor and H
s
is the heat of sorption of penetrant in
the polymer. Since steady-state permeability is the product of diffusivity and solu-
bility, the activation energy of permeation can be dened as the sum of activation
energy of diffusion and the heat of sorption (1):
E
p
= E
d
+H
S
(12)
E
d
is always positive; H
s
is often positive for light gases (such as H
2
, O
2
, and N
2
),
but can be negative for larger, more soluble penetrants (such as C
3
H
8
and C
4
H
10
).
For polymers such as low density polyethylene (LDPE) and poly(vinyl chloride)
(PVC), E
p
is always positive (1). Therefore, permeability increases with increasing
temperature. To illustrate this behavior, Figure 8 shows the effect of temperature
on oxygen permeability of four widely used barrier polymers (27).
210 BARRIER POLYMERS Vol. 5
10
3
10
2
10
1
10
0
10
1
10
2
Temperature, C
PET
AN
PVDC
EVOH 27
0 10 20 30 40 50
O
x
y
g
e
n

p
e
r
m
e
a
b
i
l
i
t
y
,

(
c
m
3


m
i
l
)
/
(
1
0
0

i
n
.
2


d
a
y


a
t
m
)
Fig. 8. Effect of temperature on oxygen permeability at 75% RH. PET is poly(ethylene
terephthalate), AN is an acrylonitrilestyrene copolymer, PVDC is vinylidene chloride
vinyl chloride copolymer (coextrusion resin grade), and EVOH 27 is ethylene vinyl alcohol
copolymer containing 27 mol% ethylene (27).
At temperatures far from the transition temperatures (eg, glass-transition
temperature and melting point) the Arrhenius relationship (eq. 10) is obeyed,
and with a known activation energy of diffusion, E
D
, the diffusion coefcient of a
penetrant in a polymer can be estimated at any temperature. In cases where E
D
is not reported, it can be estimated using a known correlation (28,29) between D
0
and E
D
:
lnD
0
=a
E
D
RT
b (13)
where a and b are independent of penetrant type. The parameter a is indepen-
dent of polymer type and has a universal value of 0.64 (30). b has a value of
9.2 [ln(10
4
cm
2
/s)] for rubbery polymers (ie, polymers above their T
g
) and
11.5 [ln(10
5
cm
2
/s)] for glassy polymers (ie, polymers below their T
g
) (13).
Equation 13 is often referred to as a linear free-energy relation. Similar rela-
tions between D
0
and E
D
are observed for viscosity of organic liquids, molten salts,
and metals (31) and for rst-order chemical reaction kinetics (32), which are also
activated processes described by the Arrhenius equation. Additionally, common
barrier polymers such as poly(ethylene terephthalate) (PET) and polycarbonate
Vol. 5 BARRIER POLYMERS 211
10
2
10
1
10
0
10
1
E
D
/RT
D
0
,

c
m
2
/
s
5 10 15 20 25
Fig. 9. Correlation of D
0
with E
D
/RT for various penetrants in glassy polymers (33).
Poly(ethylene terephthalate), bisphenol-A polycarbonate, tetramethyl bisphenol-A
PC.
(PC) are known to follow this relation both above and below the glass-transition
temperature, as shown in Figure 9 (33). Combining equations 10 and 13 gives
D=exp
_
b
_
1a
_
E
D
RT
_
(14)
On the basis of a single value of the diffusion coefcient at one temperature, E
D
can be estimated fromequation 14. This equation can then be used to estimate the
diffusion coefcient at other temperatures, provided that the two temperatures
do not traverse thermal transitions (such as glass transition) and the polymer
morphology is otherwise unchanged.
Chemical Structure. The presence of polar groups onor inpolymer chains
often increases chain rigidity, which can increase glass-transition temperature
and improve mechanical properties, and increases packing density (1). Conven-
tional barrier polymers, such as PAN, have very low gas permeability as a result,
in part, of restricted chain mobility due to the presence of polar groups. Polymer
chain interactions can be quantied in terms of cohesive energy density (CED),
and CED has a strong inuence on penetrant diffusion. CED of a polymer is the
square of its solubility parameter and characterizes the strength of attraction (or
interactions) between the polymer chains. It can be estimated using group contri-
bution techniques (13). In a simple model of penetrant diffusion in polymers, the
activation energy for diffusion is directly proportional to the CED of a polymer
212 BARRIER POLYMERS Vol. 5
10
3
10
2
10
1
10
0
10
+1
10
+2
10
+3
0.2 0.4 0.6 0.8 1.0
CED, kJ/cm
3
PE
PS
PVA
PVC
PAN
PVOH
O
x
y
g
e
n

p
e
r
m
e
a
b
i
l
i
t
y
,

(
c
m
3


m
i
l
)
/
(
1
0
0

i
n
.
2


d
a
y


a
t
m
)
Fig. 10. Correlation of oxygen permeability (23

C, 0% RH) with CED for six barrier poly-


mers: PE, PS, PVA, PVC, PAN, PVOH (36,37).
(34). On the basis of this model and the linear free-energy relation (35), the loga-
rithm of penetrant diffusion coefcients should decrease linearly with increasing
CED. Solubility of relatively nonpolar penetrants usually has a weaker depen-
dence on CED than diffusivity, and hence, the logarithm of permeability should
decrease linearly with increasing CED. As shown in Figure 10, a nearly linear
correlation was found between oxygen permeability and CED of barrier polymers
with permeability values ranging over 5 orders of magnitude (36,37).
Penetrant diffusivity, and hence permeability, can also be decreased by
adding substituents to the polymer chain that reduce chain exibility. Bulky side
groups or rigid linkages such as aromatic groups decrease chain exibility and
hence reduce penetrant diffusion coefcients. Flexible linkages, such as ether or
methylene groups, produce the opposite effect (1). Several methods are used to
characterize chain exibility (1). The glass-transition temperature is a measure
of long range or bulk molecular motion, and high T
g
materials are usually rigid
and inexible. Sub-T
g
relaxations are another indication of molecular motion, and
correlations of O
2
and CO
2
permeability with sub-T
g
relaxations have been ob-
served within a family of amorphous polyesters and copolyesters (38). However,
the exact nature of molecular motions which control penetrant diffusion are com-
plex and unclear, and hence T
g
and sub-T
g
relaxations do not, in general, provide
predictive correlations with penetrant diffusivity (1).
Changes in molecular structure of the polymer often affect more than one
factor inuencing permeability and the net effect can be difcult to anticipate.
Vol. 5 BARRIER POLYMERS 213
For example, addition of bulky side groups can stiffen the polymer chains, which
is expected to reduce the diffusion coefcients. However, the same modication
could also decrease chain packing in the amorphous phase and reduce the level
of crystallinity in the polymer, which are expected to increase the diffusion coef-
cients (39). The net result of these competing effects can be difcult to predict a
priori.
Crystallinity. Increasing crystallinity ina polymer generally decreases gas
permeability (16) (see SEMICRYSTALLINE POLYMERS). Crystallinity inuences both
solubility and diffusion coefcients. For most polymers and penetrants of interest,
crystalline regions, which are much more dense and well ordered than amorphous
regions, preclude penetrant sorption, thereby reducing penetrant solubility. Ad-
ditionally, the presence of impermeable crystallites in a polymer matrix acts as
barriers to diffusion, increasing the path length for diffusion and, in some cases,
increasing chain rigidity, which also reduces diffusion coefcients (16). Although
crystallite size, shape, and orientation do not usually inuence solubility in poly-
mers signicantly, these factors canbe important inpenetrant diffusion. The effect
of crystallinity on penetrant diffusion can be expressed using the following model
(39):
D=
D
a

(15)
where D
a
is the penetrant diffusion coefcient in the amorphous polymer, is a
geometric impedance (ie, tortuosity) factor, and is a chain immobilization factor.
Impermeable crystalline regions force penetrants to follow a tortuous pathway
through the permeable amorphous regions. This effect is captured by the factor ,
which is the ratio of the average distance traveled by a penetrant molecule to the
thickness of the sample (39). can be a complex function of crystalline content
as well as crystallite size, shape, and orientation (16,40). Crystallites can also
restrict segmental mobility by acting as physical cross-links. This effect is taken
into account by the factor and is generally more pronounced in exible rubbery
polymers such as polyethylene (PE). In glassy polymers such as PET, the inherent
rigidity of the chain backbone imposes more impedance to chain mobility than the
crystallites and hence, is one (16) (see AMORPHOUS POLYMERS).
A two-phase model is often used to describe penetrant solubility in a
semicrystalline polymer (41):
S=S
a

a
(16)
where S
a
is solubility coefcient in the amorphous regions of the polymer and
a
is
amorphous phase volume fraction. This model assumes that the solubility of the
crystalline regions is zero, and that the presence of crystallites does not change
the amorphous phase solubility coefcient. For polymers used inbarrier packaging
applications, the assumption of zero solubility in the crystalline regions is gener-
ally accurate (42). The second assumption that the amorphous phase solubility is
independent of crystalline content is not necessarily obeyed, particularly in glassy
polymers, whose state of amorphous phase structural organization may be inu-
enced signicantly by common processing protocols (eg, orientation, stretching,
214 BARRIER POLYMERS Vol. 5
10
0
10
1
10
2
10
3
10
4
0 0.1 0.2 0.3 0.4 0.5
Amorphous phase volume fraction
CO
2
O
2
N
2
Polyethylene, T = 25C
G
a
s

p
e
r
m
e
a
b
i
l
i
t
y
,

(
c
m
3


m
i
l
)
/
(
1
0
0

i
n
.
2


d
a
y


a
t
m
)
Fig. 11. Effect of crystalline content on gas permeability in PE at 25

C and 0% RH (47).
annealing, and contact with crystallization-inducing agents) (4346). Neverthe-
less, the simplest and most commonly used model for the effect of crystallinity on
steady-state permeability is based on these assumptions and is expressed by the
following relationship:
P= (S
a

a)
_
D
a

_
(17)
Figure 11 shows the effect of crystallinity on gas permeability in PE at 25

C. Per-
meability decreases with increasing crystallinity primarily because of decreasing
diffusion coefcients. The effect of crystallinity is more pronounced on tortuosity
factor than the chain immobilization factor (47). In glassy polymers, the most
widely used approximations for and are =
a
1
and =1. Introducing these
values into the above equation yields P =P
a

a
2
, where P
a
is the amorphous phase
permeability.
In certain polymers, the simple assumptions of the two-phase model do not
hold. For example, poly(4-methyl-1-pentene) (PMP), which is a highly permeable
polymer, has a very low density crystal structure, and hence, penetrant molecules
can sorb into its crystalline phase (42). In PET, it has been observed that the
presence of crystalline regions increased the concentration of sorption sites in the
Vol. 5 BARRIER POLYMERS 215
amorphous regions remaining in the polymer after crystallization (41). Others
have recently examined this phenomenon in PET in more detail (48,49). It has
been reported (43) that exposing amorphous PET to a strongly sorbing penetrant
at high activity results in penetrant-induced crystallization. This process affects
the amorphous phase penetrant solubility of PET, resulting in a marked increase
in overall solubility with increasing crystallinity. For example, acetaldehyde sol-
ubility in PET increased by more than 300% as a result of penetrant-induced
crystallinity of about 36 wt% (43). This effect was attributed to the creation of
microvoids in the polymer as a by-product of penetrant-induced crystallization
(43). Evidence for microvoid formation in PET due to exposure to strongly sorbing
penetrants has been presented in studies investigating solvent treatments to im-
prove the dyeability of PET yarns (44), and in studies investigating the effect of
crystallizing liquids on the morphology of PET (45,46). For example, it has been
reported that dye uptake in PET bers exposed to dimethylformamide (DMF) was
four to ve times higher than in unexposed bers (44). The dye diffusion coefcient
was also more than 2 orders of magnitude higher in the DMF-exposed samples
than in the unexposed samples (50).
Chain Orientation. Stretching or drawing of polymer lms can improve
mechanical properties, and under certainconditions, barrier properties (see FILMS,
ORIENTATION). The degree of chain orientation achieved is dependent on the draw
ratio and other process conditions (16). Orientation is usually characterized by
birefringence and quantied by the Hermans orientation function f (51):
f =
1
2
(3 cos
2
1) (18)
where is the average angle between the polymer chain axis and the draw
direction.
Depending on the mode of deformation and the physical processes that oc-
cur during orientation, permeability may either increase or decrease with in-
creasing orientation (16). Impermeable polymer crystallites may become oriented
into plate-like structures during deformation, and this process generally de-
creases penetrant diffusivity by increasing tortuosity (16). In addition, drawing of
semicrystalline polymers can improve barrier properties through stress-induced
crystallization and orientation of the remaining amorphous phase. Hence, the re-
duction in permeability caused by orientation of crystallizable polymers can be
greater than that in noncrystallizable polymers (16). The dramatic effect of orien-
tation is supported by oxygen permeability data for PET in the literature. A 4
biaxial orientation (ie, draw ratio = 4 in each axis) decreased the permeability
of oxygen in PET by a factor of about 2 (52). For other systems, however, in-
creases in permeability upon biaxial orientation have also been reported. Avinyli-
dene chloride (VDC)vinyl chloride copolymer, for example, showed an increase
in oxygen permeability from 0.2 (cm
3
mil)/(100 in.
2
dayatm) to 0.3 (cm
3
mil)/
(100 in.
2
dayatm) upon 2.5 biaxial orientation (53). The permeability increase
was attributed to microvoid development during orientation of the polymer chains
after crystallinity was fully developed. Table 3 shows the effect of orientation on
oxygen permeability of semicrystalline and amorphous barrier polymers (53). In
semicrystalline VDC copolymer and nylon MXD-6 polymers, under the conditions
studied, the orientation process results in a slight increase in oxygen permeability
216 BARRIER POLYMERS Vol. 5
Table 3. Effect of Orientation on Oxygen Permeability Characteristics of
Semicrystalline and Amorphous Barrier Resins
a
Barrier polymer resins Oxygen permeability,
b
(cm
3
mil)/(100 in.
2
dayatm)
Semicrystalline resins
VDC copolymer
c
Compression molded lm 0.20 0.02
Extrusion cast lm 0.20 0.01
Biaxially oriented 2.5 0.30 0.01
Aromatic Nylon MXD-6
d
Extrusion cast 0.37 0.09
Biaxially oriented 2 0.39
Amorphous resins
Amorphous Nylon Selar PA 3426
e
Extrusion cast lm 1.40 0.31
Uniaxially oriented 2.5 1.14 0.07
Biaxially oriented 2.5 1.01 0.01
Polyacrylic-imide XHTA-50A
f
Extrusion cast 3.12 0.17
Uniaxially oriented 2 2.95 0.04
Uniaxially oriented 2.5 2.84
Biaxially oriented 2 2.76 0.03
a
Ref. 53.
b
At 23.5

C, 65% RH.
c
Dow Chemical Companys vinylidene chloridevinyl chloride copolymer (experimental grade XU
32009.02).
d
Trademark of Mitsubishi Gas Chemical Co., Japan.
e
Trademark of E. I. du Pont de Nemours & Co., Wilmington, Del.
f
Trademark of Rohm & Haas Co., Philadelphia, Pa.
with chain orientation, whereas the reverse is true for amorphous Selar 3426 and
polyacrylic-imide barrier polymers.
Penetrant Concentration (or Partial Pressure). The inuence of pen-
etrant concentration on solubility, diffusivity, and, in turn, permeability varies,
depending on the penetrantpolymer system. Rubbery and glassy polymers typ-
ically show little or no concentration dependence for solubility, diffusivity, and
permeability of light gases such as H
2
, N
2
, and O
2
. Consistent with this notion,
Figure 12a shows essentially no inuence of pressure on H
2
permeability in PE
(19). Gases such as CO
2
, which are more soluble than light gases, typically have
a permeabilitypresssure response in glassy polymers similar to that shown in
Figure 12b. Permeability decreases monotonically with increasing pressure as
predicted by the dual-mode sorption model (54). The magnitude of the permeabil-
ity decrease depends upon the amount of so-called nonequilibrium excess volume
in the polymer, which can increase with increasing T
g
, the afnity of the pene-
trant for the nonequilibrium excess volume, and the mobility of the penetrant in
the nonequilibrium excess volume relative to its mobility in the equilibrium free
volume (1). The permeability of a rubbery polymer to an organic vapor often ex-
hibits the behavior shown in Figure 12c. The monotonic increase in permeability
is often due to increases in penetrant solubility with increasing pressure coupled
with increases in diffusivity with increasing pressure (1). The response shown in
Vol. 5 BARRIER POLYMERS 217
(b) CO
2
/PC
800
1200
1600
0 100 200 300
P
e
r
m
e
a
b
i
l
i
t
y
Pressure, psia
( d) Acetone/EC
0 4 8 12
P
e
r
m
e
a
b
i
l
i
t
y
Pressure, psia
310
5
410
5
510
5
610
5
1500
1600
1700
1800
0 100 200 300
P
e
r
m
e
a
b
i
l
i
t
y
Pressure, psia
(a) H
2
/PE
(c) C
3
H
8
/PE
4000
8000
12000
0 40 80 120
P
e
r
m
e
a
b
i
l
i
t
y
Pressure, psia
Fig. 12. Typical permeabilitypressure dependence in rubbery and glassy polymers. (a)
Hydrogen in PE at 30

C, (b) carbon dioxide in PC at 35

C, (c) propane in PE at 20

C,
(d) acetone in ethyl cellulose (EC) at 40

C (19). The permeability values have units of


(cm
3
mil)/(100 in.
2
dayatm).
Figure 12d is typical for strongly interacting penetrants (eg, organic vapors) in
glassy polymers at sufciently high penetrant partial pressures. It can be viewed
as a superposition of the behaviors in Figures 12b and 12c (1). The sharp increase
in permeability begins as the penetrant plasticizes the polymer. Plasticization
occurs when penetrant molecules dissolve in the polymer matrix at sufcient con-
centration to force polymer chain segment separation, thereby increasing the free
volume, and in turn, facilitating polymer segmental motion. This increase in seg-
mental mobility, which may be observed by the depression in T
g
, results in an
increase in penetrant diffusion coefcients and, in turn, permeability (1).
Humidity. The absorption of water can increase, decrease, or have no effect
on gas permeability of barrier polymers (27). Increasing the relative humidity
(RH) from 0 to 50% increases the oxygen permeability of cellophane (regenerated
cellulose) by an order of magnitude, and exposure to 90% RH removes it from
the class of high barriers by further increasing the permeability by more than an
218 BARRIER POLYMERS Vol. 5
10
3
10
2
10
1
10
0
10
1
10
2
0 20 40 60 80 100
Relative humidity, %
AmNY
BON
MXD-6
EVOH 44
EVOH 32
O
x
y
g
e
n

p
e
r
m
e
a
b
i
l
i
t
y
,

(
c
m
3


m
i
l
)
/
(
1
0
0

i
n
.
2


d
a
y


a
t
m
)
T = 20C
Fig. 13. Effect of relative humidity on oxygen permeability of hydrophilic barrier poly-
mers. AmNY is amorphous nylon (Selar), BON is biaxially oriented nylon-6, MXD-6 is
oriented poly(metaxylylenediamineadipic acid), EVOH 44 and 32 are ethylene vinyl alco-
hol copolymers containing 44 and 32 mol% ethylene (27).
order of magnitude (27). For packaging of foods that require protection against
oxygen ingress, cellophane is coated or laminated with water barriers such as
polyolens (27). Other hydrophilic barrier polymers, with the exception of certain
amorphous polyamides, also lose their barrier properties with increasing RH, as
shown in Figure 13. This is because water acts as a plasticizer and increases the
free volume of the polymer (55). However, at low to moderate RH, amorphous
polyamides and PET show slightly improved barrier properties with increasing
RH(27). This behavior has been explained as the water molecules not swelling the
polymer, but occupying some of the polymer free-volume sites instead, resulting
in reduction in permeability of other gases (56). VDC copolymers, acrylonitrile
copolymers, and polyolens show essentially no effect of RH on gas permeability
(57).
Techniques for Measuring Transport Properties
The determination of permeability, solubility, and diffusivity requires direct or
indirect measurement of mass transfer under controlled conditions. The perme-
ability of barrier polymers can be determined directly by measuring the pressure
change or other physical evidence of transfer or indirectly by using an indicator
of permeation, such as chemical reaction of the transferring gas with another
Vol. 5 BARRIER POLYMERS 219
substance. The preferred methods of measurement differ for different classes
of penetrants: light gases, water vapor, condensable vapors, and food avor and
aroma compounds.
There are two basic methods for measuring permeability: isostatic and quasi-
isostatic (55). Isostatic methods employ a continuous owon both sides of the poly-
mer lm to provide constant penetrant concentrations. Quasi-isostatic methods
use a continuous ow to maintain constant penetrant concentration only on the
upstream side and allow penetrant accumulation on the downstream side of the
lm. However, this accumulation is limited to a very low concentration, and hence
the penetrant partial pressure difference can be approximated as a constant (55).
Figure 14a typically shows the course of an isostatic permeability experiment for
a barrier polymer lm of uniform thickness exposed to constant penetrant par-
tial pressure p on the upstream side and constant removal of penetrant that has
permeated through the lm to the downstream side. Using a specied initial con-
dition (concentration in the lm uniformly equal to zero) and boundary conditions
(constant penetrant concentration C at the upstreamside and zero penetrant con-
centrationat the downstreamside), this situationcanbe describedby the following
mathematical expression (58):
q =
DC
L
_
t
L
2
6D
_

2LC

n=1
_
1
_
n
n
2
exp
_
Dn
2

2
t
L
2
_
(19)
where q is the total mass of penetrant permeating per unit lm area in time t, D
is the diffusion coefcient, L is lm thickness, and C is penetrant concentration at
the upstream side in equilibrium with the upstream penetrant partial pressure
p. When steady state is reached, t becomes large enough to make the exponential
term negligibly small, and the above equation reduces to
q =
DC
L
_
t
L
2
6D
_
(20)
A plot of q vs t yields a straight line whose slope is the steady-state penetrant ux
(N
A
= DC/L) and whose x-axis intercept is called the time lag (t
L
).
t
L
=
L
2
6D
(21)
The time lag can be related to the time required to achieve steady state (t
SS

2.7t
L
) (27). The diffusion coefcient can be calculated by rearranging the above
equation (58):
D=
L
2
6t
L
(22)
Permeability can be calculated from equation 1 and the steady-state ux value
(P = N
A
L/p = DC/p). It should be noted that for concentration-dependent
diffusion coefcients, t
L
will vary with the pressure difference across the
polymer lm, and as a result, this simple time lag analysis may yield signicant
220 BARRIER POLYMERS Vol. 5
t
L
Steady state
t
0
q
t
1/2
t
SS
1.0
0.5
0
(a)
(b)
N
A
/
(
N
A
)
S
S
Fig. 14. (a) Mass of permeating penetrant per unit lm area (q) as a function of time
(providing a measure of time lag t
L
). (b) Normalized penetrant ux (N
A
) as a function of
time (providing a measure of halftime t
1/2
). t
SS
is time required to achieve steady state.
errors in the diffusion coefcients estimated using equation 22 (58). More general
expressions for the time lag have been developed by including the concentration
dependence of D (59). Alternatively, a concentration-averaged diffusion coef-
cient can be obtained by plotting normalized penetrant ux (ie, ux at any time
t divided by the steady-state ux) as a function of time (Fig. 14b). The diffusion
coefcient can be estimated using the following relationship (60):
D=
L
2
7.2t
1/2
(23)
Vol. 5 BARRIER POLYMERS 221
Moist N
2
Moist O
2
Inside chamber
Outside chamber
Barrier film
RH probe
RH probe
To sensor
Fig. 15. A schematic of a permeation cell in the Ox-Tran (MOCON, Inc.) oxygen trans-
mission rate measurement system (62).
where t
1/2
is the halftime (ie, time required for the penetrant ux to reach half of
its steady-state value). Thus, permeability, diffusivity, and hence solubility (S =
P/D) can be determined from a single experiment. Given the ready availability of
computing power, it is nowpossible to use any of a variety of numerical techniques
to t the entire response and extract the desired parameters and, when applicable,
their concentration dependence.
Oxygen and Carbon Dioxide Permeation. The most widely used com-
mercial instrument for measuring oxygen transmission rates of at lms and
packages is the Ox-Tran (Modern Controls Inc., Minneapolis, Minn.), and mea-
surements are made inaccordance withASTMmethod D3985 (61). Inthis isostatic
coulometric method, at lm samples are clamped into a diffusion cell, which is
then purged of residual oxygen using an oxygen-free carrier gas such as N
2
. The
carrier gas is routed to the instrument sensor until a stable zero has been estab-
lished. Pure oxygen is then introduced into the outside chamber of the diffusion
cell (see Fig. 15) (62). Oxygen molecules diffusing through the lm to the inside
chamber are conveyed to the sensor by the carrier gas. The Ox-Tran system uses
a patented coulometric sensor (Coulox) to detect oxygen transmission through
both at lms and packages. This sensor provides parts-per-billion sensitivity to
oxygen even in the presence of water vapor. Digital pressure and ow controls
allow for RH control. Alternative instruments for measuring oxygen transmis-
sion rates include Oxygen Permeation Analyzers from Illinois Instruments Inc.
(Ingleside, Ill.).
Modern Controls, Inc. (MOCON) also makes instruments for measuring car-
bon dioxide permeation. Their Permatran-C line of instruments uses an infrared
detector to detect carbon dioxide that permeates through the test lm.
222 BARRIER POLYMERS Vol. 5
Water Vapor Permeation. WVTRs can either be measured by the tra-
ditional gravimetric cup method (63) or by newer electronic instruments. The
newer method (eg, ASTM method F1249 (64)) uses infrared detection to mea-
sure water vapor transmission through barrier lms. One of the most widely used
commercial WVTR systems is Permatran-W (Modern Controls Inc., Minneapolis,
Minn.). The newest model of this system (Permatran-W 3/31) uses a patented
modulated infrared sensor to detect water vapor transmission through at lms
and packages. It provides a sensitivity in the range of parts per million. Various
models are available with different temperature and RH capabilities.
Lyssy AG (Zollikon, Switzerland) also manufactures automatic water vapor
permeability testers designated as the L80 line of instruments. The L80-5000 is
the newest member and the fth generation in this series.
In the traditional method (ASTM method E96) (27), a sample cell containing
either a desiccant or distilled water is covered with the sample lm and placed in
a controlled atmosphere. Typical conditions for the desiccant method are 100

F
(37.8

C) and an external RHof 90%, although the standard also allows for temper-
atures between 70 and 90

F (21 and 32

C) at (50 2)% RH. The cell assembly is


weighed periodically until steady state is reached. WVTR can be calculated from
the steady-state rate of change in the weight of the cell.
Flavor and Aroma Compounds. Measurement of transport rates of a-
vor and aroma vapors in plastics is more complicated than that of either water
vapor or light gases. Elaborate equipment and sensitive analytical devices are re-
quired to obtain reliable results. Since the transport behavior of these compounds
is often strongly concentration dependent, measurements must be made in the
activity range in which the compounds are present in practice. Thus, some of the
major complexities include providing precisely mixed quantities of the condens-
able vapor in an inert carrier such as nitrogen or argon at very lowconcentrations,
typically a few parts per million, and assuring that the concentration is main-
tained. Temperature must also be carefully controlled to prevent condensation on
equipment surfaces (27).
As such, no single instrument has gained the widespread acceptance noted
above for instruments for O
2
, CO
2
, and water vapor, though several methods have
been used to measure transport properties of avor and aroma compounds in
barrier polymers. These include isostatic permeation techniques and gravimet-
ric techniques (65). The permeation techniques directly yield permeability and
diffusivity of avors in barrier polymers. Solubility can then be calculated indi-
rectly using the relation P = D S. The MAS 2000 Organic Vapor Permeation
Test System (MAS Technologies Inc., Zumbrota, Minn.) is an example of a sys-
tem that was commercially available for avor permeation measurements. Mass
spectrometry and ame ionization detection have also been successfully used as
vapor concentration detectors (65). In contrast, gravimetric techniques permit di-
rect and independent measurements of both solubility and diffusivity (66). It is
possible to measure sorption and desorption of organic avors in barrier polymers
using sensitive gravimetric sorption instruments such as the McBain spring bal-
ance and the Rubotherm magnetic suspension balance (6770). A schematic of
a McBain spring balance assembly is shown in Figure 16. The polymer sample
is suspended from a sensitive helical quartz spring inside the sorption chamber.
After introducing the penetrant, the spring position relative to a xed reference
Vol. 5 BARRIER POLYMERS 223
To
vacuum
Vent
Liquid nitrogen
cold trap
Penetrant
gas/vapor
inlet
To pressure read-out
and power supply
Transducer
(01000 torr)
12-L
gas/vapor
reservoir
CCD camera
Water in
Water out
Quartz spring
Reference
rod
Polymer
sample
Water-jacketed
sorption cell
Computer
Fig. 16. A schematic of a McBain spring balance apparatus for measuring sorption and
desorption of organic vapors in barrier polymers (66).
rod hanging inside the chamber is recorded using a CCDcamera. Fromthe kinetic
uptake data, solubility and diffusivity values are estimated (71). Acceptable lev-
els of agreement have been reported for solubility coefcients of ethyl acetate in
LDPE, linear LDPE (LLDPE), and ionomer lms obtained from gravimetric and
isostatic permeation techniques (65).
Techniques for Predicting Transport Properties
Modeling Transport Properties of Gases and Condensable Vapors
in Polymers. The permeation of low molecular weight gases such as O
2
, N
2
,
CO
2
as well as large avor and aroma compounds is an essential consideration
in the selection and schematic of food packages and containers. Predictive models
for permeation would minimize the number of experiments required in package
material selection and development. Perhaps, more importantly, they also provide
an insight into the underlying factors controlling permeation in barrier polymers.
The permachor method has been used to predict permeabilities of low molec-
ular weight penetrants in barrier polymers (72). Although originally developed
224 BARRIER POLYMERS Vol. 5
for O
2
, N
2
, and CO
2
, this method can be extended to other gases and vapors, pro-
vided there is no specic interaction between the penetrant and the polymer. It
has been successfully used for over 60 different polymers (72). In this method,
numerical values (ie, group contributions) are assigned to polymer segments. An
average numerical value can then be obtained for the polymer, which is referred
to as the permachor value of the polymer. A simple equation is used to relate gas
permeability P to polymer permachor value :
P= Ae
s
(24)
where A and s are temperature-dependent constants. This method also takes into
account the reduction in permeability caused by the orientation of crystalline
polymers using the following expressions:
P=
_
A

0
_
e
s
(25)

1.13

a
(26)
where
0
is tortuosity related to crystallite orientation and
a
is the amorphous
phase volume fraction. A good agreement has been reported between experimen-
tal and model predictions of O
2
, N
2
, and CO
2
permeability values in a variety
of polymers (72). This method has also been modied for predicting liquid per-
meation through polymers (63). The permachor method works well for polymers
and copolymers, but is not applicable to polymer blends (27). As with any group
contribution method, caution should be exercised when attempting to perform
predictions which are outside the data set used to generate the correlation.
Other methods for correlating gas permeability in barrier polymers with
polymer molecular structure have been developed using free-volume theory
(17,18,73). Groupcontributiontechniques canbe usedto estimate polymer free vol-
ume from densities and intrinsic volumes of various polymer components. In the
method proposed in Reference 18, polymer-specic free volume was used, which
was dened as (VV
0
)/M, where V is specic volume, V
0
is specic occupied vol-
ume, and Mis polymer molecular weight. V
0
can be calculated according to Bondis
method from van der Waals volumes of the various groups in the polymer struc-
ture (18). In this model, free volume was dened on a unit weight basis so that
various molecular structures could be compared on the same weight basis. The
model predicts a linear relationship between logarithm of gas permeability and
the reciprocal of polymer-specic free volume. Other improvements to this model
have been suggested (17,73). Gas permeability has often been correlated with FFV
(as dened in eq. 6) using the following relation (73):
P= Aexp
_
B
FFV
_
(27)
where A and B are constants for a particular gas. It has been observed that when
this model is limited to a specic family of polymers, eg, polyesters and polyamides
Vol. 5 BARRIER POLYMERS 225
(17), a reasonably good correlation can be obtained. However, when the correla-
tion is broadened to include a wider range of polymer types, there is consider-
able scatter in the data, particularly at low values of gas permeabilities. Even
though these free-volume-based models have some fundamental basis for corre-
lating transport properties, they have the following limitations (18,73): (1) the
assumption of solubility being independent of free volume and polymer structure
is clearly an approximation, (2) the concept of free volume cannot capture all
the factors affecting gas permeability (such as chain exibility and CED), and (3)
there may be errors in values of van der Waals volumes available in the literature.
Attempts have been made to rene these models by introducing more empirical
parameters and making them more predictive, and these efforts have resulted in
signicant improvements in the accuracy of the correlations (73).
Larger and more condensable penetrants, eg, avor and aroma compounds,
can have extremely low diffusion coefcients in common barrier polymers, result-
ing in extremely large time scales to achieve steady state. For example, d-limonene
has diffusion coefcients of the order of 10
14
cm
2
/s in PET (74). When coupled
with a typical lmthickness of 10 m, this leads to time scale to reach steady state
of more than 3 years. Hence, as noted before, over the shelf life of the packaged
product (typically of the order of several months), avor and aroma compounds
may never reach steady-state transport. Therefore, independent predictions of
their diffusion and solubility coefcients become necessary. Several methods have
been proposed for predicting solubility coefcients. A widely used method is based
on a thermodynamic approach that relates penetrant sorption to solubility param-
eters of the penetrant and the polymer (75). This dependence can be expressed as
follows:
S=S
0
exp[(H
vap
H
mix
)/RT ]
H
mix
=
1

2 (
1

2)
2
(28)
where S
0
is a constant for a particular polymer,
1
is partial molar volume of the
penetrant,
2
is volume fraction of polymer in the mixture,
1
and
2
are solubility
parameters of penetrant and polymer respectively, and H
mix
is enthalpy change
on mixing of penetrant molecules with polymer segments. The values of
1
,
1
, and

2
can be obtained from the literature (36). The enthalpy change on vaporization
of the penetrant (H
vap
) can be calculated from the penetrant boiling point by
using available correlations (2). Reasonably good agreements have been reported
between model-predicted and experimentally observed solubility coefcients of
several penetrants in vinylidene chloridevinyl chloride copolymers and LDPE at
85 and 30

C respectively (76).
For penetrants that interact with the polymer matrix primarily via disper-
sion (ie, van der Waals) forces, penetrant solubility scales with measures of pene-
trant condensability such as penetrant boiling point, critical temperature, or the
force constant in the LennardJones potential model (43). The following relation
between penetrant critical temperature and penetrant solubility has been derived
using a classical thermodynamics model (13,77):
lnS
a
= N+MT
c
(29)
226 BARRIER POLYMERS Vol. 5
0 100 200 300 400 500 600 700
10
5
10
4
10
3
10
2
10
1
10
0
10
1
10
2
Critical temperature T
c
, K
MIPK
MEK
CH
2
Cl
2
(CH
3
)
2
CO
CH
3
OH C
6
H
6 H
2
O
CH
3
COOC
2
H
5
C
2
H
4
O
n-C
5
H
12
n-C
4
H
10
i-C
4
H
10
i-C
5
H
12
CO
2
CH
4
O
2
N
2
He
Ar
C
6
H
5
CH
3
I
n
f
i
n
i
t
e

d
i
l
u
t
i
o
n

p
e
n
e
t
r
a
n
t

s
o
l
u
b
i
l
i
t
y
,

c
m
3

(
S
T
P
)
/
(
c
m
3


c
m

H
g
)
Fig. 17. Correlation of innite dilution penetrant solubility with penetrant critical tem-
perature in PET((71) and unpublished data). Nonpolar penetrants, polar and quadrupo-
lar penetrants, aromatic penetrants. MEK is methyl ethyl ketone and MIPK is methyl
isopropyl ketone. The slope M = 0.019 0.001 K
1
and intercept N = 9.6 0.4.
In this expression, N is a parameter that depends primarily on polymerpenetrant
interactions and polymer free volume. T
c
is the penetrant critical temperature,
which is widely tabulated for many penetrants of interest (2). M is constant and
has a value of approximately 0.016 K
1
for gas dissolution in liquids and in rub-
bery and glassy polymers (13). While N varies from polymer to polymer, average
values of 9.7 and 8.7 for rubbery and glassy polymers at 35

C, respectively,
when solubility is expressed in cm
3
(STP)/(cm
3
cm Hg) have been recommended
(13). Penetrants with strong dipole or quadrupole moments may be more soluble
in a polar polymer matrix, such as PET, than predicted based on equation 29 (71).
Although equation 29 is strictly valid for penetrant sorption in equilibrium ma-
trixes, suchas liquids or rubbery polymers, it also provides anexcellent description
of equilibrium solubility in glassy polymers (71). Figure 17 shows the correlation
of penetrant solubility in PET with penetrant critical temperature (71). Nonpolar
penetrants showexcellent agreement with the model presented in equation 29 and
polar or quadrupolar penetrants exhibit signicant scatter around the correlation
line.
Over wider ranges of critical temperature, it has been suggested that pene-
trant solubility coefcients may be better correlated with the square of reciprocal
reduced temperature (T
c
/T)
2
(78):
Vol. 5 BARRIER POLYMERS 227
ln S
a
=n+m
_
T
c
T
_
2
(30)
where T is the temperature of the experiment, and m and n are the slope and in-
tercept of the correlation line, respectively. This equation may also be derived from
fundamental thermodynamic considerations (43). Other semiempirical methods,
eg, UNIFAC group contribution model, have also been proposed for predicting
penetrant solubility in polymers (79,80).
Several predictive and correlative methods have been developed for diffusion
coefcients of penetrants in polymers. An empirical relationship has been devel-
oped for correlating diffusion coefcients with penetrant critical volume (81):
D
a
=

V

c
(31)
where D
a
is amorphous phase diffusion coefcient, V
c
is penetrant critical volume,
and , are adjustable constants. This equation has been proposed based on
analogy with correlations of diffusion coefcients with critical volume of small
molecules inliquids. For larger penetrants (eg, long-chainhydrocarbons), diffusion
steps may occur via motion of only part of the molecule, and critical volume is
not expected to capture the effective size of a penetrant unit participating in a
diffusion step (43). In such cases, diffusion coefcients would be less sensitive to
penetrant size than indicated in the above equation. Also, critical volume fails
to capture the effect of penetrant shape on diffusion coefcients (71). Figure 18
shows a plot of D vs. V
c
for PET at 25

C. Diffusion coefcients of penetrants (up to


molecular weights of 100 Da) in PVC, PS (polystyrene), and PMMA [poly(methyl
methacrylate)] have been correlated empirically with other measures of penetrant
size, such as molecular diameter (82).
A theoretical model based on polymer free volume, temperature, and pene-
trant size and shape has been developed (83). According to this model, diffusion
coefcients of large penetrants in amorphous rubbery polymers are given by
D=
V
2
f
6
_
eRT
M
_
1/2
_
1

1
Ac
2
1
+
1

2
Ac
2
2
+
1

3
Ac
2
3
_
exp
_
E
D
RT
_
(32)
where V
f
is average free volume per polymer chain segment, e is the base of
natural logarithm, M is penetrant molecular weight,
i
is length of penetrant
molecule along a given direction i (taken as the principal axis of inertia), Ac
i
is
effective penetrant molecular cross-sectional area perpendicular to the direction i,
and E
D
is the activation energy of diffusion. V
f
can be calculated fromthe following
equation (83):
V
f
=V [0.025+
f
(TT
g2
+kw
1
)] (33)
where V is the total volume per mole of the polymer repeat unit,
f
is tempera-
ture coefcient of free-volume expansion, T
g2
is glass-transition temperature of
the polymer, k is plasticizing efciency of the penetrant for the polymer, and w
1
is
228 BARRIER POLYMERS Vol. 5
10
16
10
14
10
12
10
10
10
8
10
6
0 100 200 300 400 500
Critical volume V
c
, cm
3
/mol
He
O
2
N
2
CO
2
H
2
O
Ar
CH
4
C
3
H
8
n-C
4
H
10
n-C
5
H
12
C
6
H
5
CH
3
CH
2
Cl
2
CH
2
Cl
2
C
6
H
6
MIPK
i-C
5
H
12
(CH
3
)
2
CO
MEK
I
n
f
i
n
i
t
e

d
i
l
u
t
i
o
n

p
e
n
e
t
r
a
n
t

d
i
f
f
u
s
i
v
i
t
y
,

c
m
2
/
s
Fig. 18. Effect of penetrant size on innite dilution, amorphous phase diffusion coef-
cients in PET at 25

C ((71) and unpublished data). Branched pentrants. MEK is methyl


ethyl ketone and MIPK is methyl isopropyl ketone. The best-t parameters of equation 31
are = (5.7 1.2) 10
8
(cm
2
/s) (cm
3
/mol)
9.1
and = 9.1 0.9.
weight fraction of the penetrant in the polymer. Equation 32 is strictly valid for
large penetrants satisfying the size criterion V
s
V
f
, where V
s
is effective steric
volume of the penetrant (83). The penetrant molecular shape dependence is rep-
resented by the three-termed expression within the parentheses of equation 32.
The equation can be applied to penetrants of a wide variety of molecular shapes
(83). This model has been tested using data for diffusion of plasticizers in PVC
(83).
Migration Modeling of Polymer Additives into Packaged Foods
and Beverages. One of the key applications of barrier polymers is food and
beverage packaging. Several low molecular weight components, eg, monomers
and oligomers, as well as additives such as lubricants, stabilizers, and plasticizers,
which are necessary for processing and stability, can be present in polymers used
for packaging. Hence, there exists a potential for permeation (or migration) of
these additives into the food or beverage with subsequent contamination (84
86). To ensure the safety of packaged food components, the U.S. Food and Drug
Administration (FDA) established a Threshold of Regulation approach which
sets upper limits on the additive concentrations in the food (84). Since traditional
migration testing methods are time consuming, expensive, and the analysis can
Vol. 5 BARRIER POLYMERS 229
be difcult (especially at low penetrant concentrations), the FDA has developed
models for predicting the additive concentration in the food simulant and rate of
transport of additives (84):
q =2C
0

_
D
p
t

(34)
where q is total mass of permeating species per unit surface area, C
0
is initial
additive concentration in the polymer, is polymer density, D
p
is additive diffu-
sion coefcient, and t is the package shelf life. This equation assumes that (84) (1)
permeation is diffusion-controlled and follows Ficks law, (2) no solubility-limited
partitioning occurs between the polymer and the food, and (3) other external phase
mass transfer resistances (eg, mixing, reaction with food) are negligible. The fol-
lowing empirical equation has been developed for predicting diffusion coefcients
(84):
D
p
= 10
4
exp
_
A
p
aMW b
_
1
T
__
(35)
where D
p
is additive diffusion coefcient (cm
2
/s), A
p
is constant for a particular
polymer, MW is additive molecular weight (g/mol), and a, b are correlation con-
stants with values of 0.01 mol/g and 10,450 K respectively. The A
p
values are 9
for LDPE, 3 for PET, and 5 for HDPE and PP (84). A semiempirical model for
predicting diffusion coefcients has also been developed (87):
lnD
P
= lnA+
_
MW
_
1/2

K
_
MW
_
1/3
T
(36)
where A, , and K are constants determined fromexperimental data. The diffusion
coefcients follow Arrhenius-type behavior and are taken to be independent of
penetrant concentration (84). These models do not explicitly account for the effect
of polymer crystallinity or orientation on additive diffusion coefcients.
The approach of the European Commission (EC), on the other hand, has been
to assign specic migration limits to different substances with adverse toxicologi-
cal properties (86). A Fickian diffusion-based model, called the Piringer migration
model, which uses the Migratest Lite program, has been used (85,86), and its
mathematical form is given below:
M
t
=C
0
L
_

1+
_
_
1

n=1
2
_
1+
_ _
1+ +
2
q
2
n
_
exp
_
D
p
tq
2
n
L
2
__
,
=
1
K
V
F
V
p
, tanq
n
= q
n
(37)
where L is lmthickness, K is partition coefcient, and V
F
and V
p
are the volumes
of food and polymer, respectively. The diffusion coefcient can be estimated from
equation 35.
The sorption and transport of avor and aroma compounds from the food
simulant into the packaging walls can affect the migration of additives from the
230 BARRIER POLYMERS Vol. 5
walls into the food simulant. Neither approach (FDA or EC) takes this effect
into account. In general, these migration models provide conservative estimates
(froma safety viewpoint) of additive concentrations in the food simulant and their
diffusion coefcients. However, there are certain cases when they can fail (84) (eg,
in cases where the avor and aroma compounds from the food simulant plasticize
the polymer, or if the additive reacts either with the polymer or with the food
simulant to produce a species that is not detectable), and attempts are being
made to improve them.
Chemical Structures and Properties of Barrier Polymers
Barrier polymers can be broadly classied as high barrier and moderate to low
barrier polymers, depending on the degree to which they restrict the passage of
gases such as O
2
or CO
2
and water vapor. The boundaries between these classica-
tions, while somewhat arbitrary, are based on the effect of the barrier properties of
the polymer on the shelf life of the packaged products. This section discusses var-
ious properties of different classes of barrier polymers. The selection of a barrier
polymer for a particular packaging application depends not only on its barrier
properties but also on other physical properties, and a comparison of physical,
mechanical, and optical properties of some commonly used barrier polymers is,
therefore, presented in Table 4 (88). Permeabilities of light gases (O
2
and CO
2
)
and water vapor are presented in Tables 5 and 6 respectively (27), and Table 7
presents permeability, diffusivity, and solubility of avor and aroma compounds in
various high and moderate barrier polymers (63). Permeability data of light gases
and water vapor in barrier polymers (9397) are more widely available than those
of avor and aroma compounds (98100). Proper caution should be exercised when
using or comparing data from different sources since, as discussed in the previous
sections, polymer permeability values depend on a wide variety of factors.
High Barrier Polymers. High barrier polymers are generally understood
to be those polymers which offer a high resistance to gas transmission. There
are no specic limits for the gas transmission rates, but this category comprises
polymers with gas permeabilities low enough to signicantly prolong the shelf life
of packaged products.
EthyleneVinyl Alcohol Copolymers. The general structure of ethylene
vinyl alcohol (EVOH) resins is as follows (101):
EVOH resins are random copolymers of ethylene and vinyl alcohol made
by the hydrolysis of ethylene vinyl acetate copolymers (101). The leading manu-
facturers are Kurray (EVALCA) and Nippon Gohsei (57). In commercial grades
of EVOH used in packaging, ethylene concentration ranges from 29 to 44 mol%
(EVAL Co., U.S.). No additives are required in their processing as the presence
of ethylene units renders the otherwise intractable vinyl alcohol melt process-
able in conventional molding and extrusion equipments. A high concentration of
Table 4. Physical, Mechanical, Optical, and Chemical Properties of Some Commonly Used Barrier Polymers
a
Property HDPE LDPE PP PET PVC PS PVDC Nylon-6 EVOH
Density, g/cm
3
0.9450.967 0.9150.925 0.90 1.4 1.221.36 1.05 1.61.7 1.14 1.141.19
Glass-transition 55 25 20 80 80 100 17 50
temperature,

C
Yield, Pa
1
, 1 mil 4.1 4.2 4.4 2.9 2.8 3.8 2.4 3.5 3.33.5
Tensile strength, MPa 17.241.4 10.334.5 137.8206.7 172.2227.4 27.555.1 55.182.7 55.1110.2 172.2255 8.211.7
Tensile modulus, 0.9 0.10.3 2.4 4.8 2.44.1 2.73.3 0.31 1.72 2.02.6
1% secant, GPa
Elongation at break, % 200600 200600 50275 70130 100400 230 50100 70120 120280
Tear strength, kN/m
b
17.587.5 175263 175350 17.552.5 52.5175 0.35 87.5140
Chemical resistance Inert Inert Inert Inert Inert Inert (except Inert, Inert
oils, greases) sorbs water
Haze, % 3 510 3 2 12 1 15 1.5 12
Light transmission, % 65 80 88 90 92 90 88 90
Heat-seal temperature 135155 120175 90150 135175 135170 120175 120150 120175 175200
range,

C
Service temperature 40 to 120 55 to 80 5 to 120 70 to 150 30 to 65 60 to 80 15 to 135 70 to 200 15 to 150
range,

F
a
Ref. 88.
b
To convert kN/m to lb/in., divide by 0.175.
2
3
1
232 BARRIER POLYMERS Vol. 5
Table 5. Oxygen and Carbon Dioxide Permeabilities of Various High and Moderate
Barrier Polymers
a
Gas permeability,
b
(cm
3
mil)/(100 in.
2
dayatm)
Barrier polymer Oxygen Carbon dioxide
EVOH, 27 mol% ethylene 0.03 (20

C) 0.04 (20

C, 65% RH)
P(VDCAN) barrier coating 0.04 0.1
Liquid crystalline polymer
c
0.06 (100% RH)
EVOH, 44 mol% ethylene 0.07 (20

C) 0.2 (20

C, 65% RH)
PVDC coextrusion resin 0.10 0.25
Nylon MXD-6 oriented lm 0.17
BLOX 4000 series PHAE resin
d
0.2 (80% RH)
Nitrile resin 0.65 1.6
Amorphous nylon 1.2 4.0 (30

C, 80% RH)
PEN polyester
e
1.2
Nylon-6, biaxially oriented 2.6 5.8 (dry)
PET, 25% crystalline (bottle wall) 4.8 24
PVC, rigid 5.0 20
Nylon-6, unoriented 6.6 10.2
Aclar 33C PCTFE lm
f
7.0 (0% RH) 16 (0% RH)
PP, biaxially oriented 150 548
HDPE, molded 185 580
MDPE, molded 250 1000
PS lm, oriented 365 900
LDPE, molded 498 2500
a
Ref. 27.
b
At 75% RH and 2325

C except as noted.
c
Vectra A950 LCP lm (Celanese AG) (89).
d
BLOX is a trademark of The Dow Chemical Co. (Midland, Mich.) (90).
e
Ref. 57.
f
Aclar is a trademark of Honeywell (Morristown, N.J.) (91).
ethylene is also recommended for thermoforming (102). At low to moderate RH,
EVOHcopolymers provide an excellent barrier to gases. The hydroxyl units (or OH
groups) contribute strongly to increasing chain cohesive energy density and im-
proving barrier properties, and so the greater the fraction of OH groups, the lower
the permeability (103). Figure 19 shows the effect of ethylene concentration on
oxygen and water vapor permeability of EVOH (27). This gure also presents per-
meation properties of poly(vinyl alcohol) (PVOH) [ie, fully hydrolyzed poly(vinyl
acetate)] and HDPE (which has no OH groups) for comparison with the EVOH
series (see VINYL ALCOHOL POLYMERS). As RH increases, the barrier properties of
EVOH copolymers decrease (cf Fig. 13). The OH groups are also responsible for
the hydrophilic nature of the polymer. Hence, the greater the percentage of vinyl
alcohol units, the greater the inuence of humidity on gas barrier properties of
EVOH copolymers (27). Conversely, high proportions of ethylene units improve
resistance to moisture (as shown in Fig. 13) and decrease the water vapor trans-
mission rate, as shown in Figure 19. Moisture sensitivity can also be decreased
somewhat by biaxial orientation (57). EVOH also offers very high barrier to a-
vor and aroma compounds and these barrier properties are not as sensitive to
Vol. 5 BARRIER POLYMERS 233
Table 6. Water Vapor Transmission Rates of Various High and Moderate Barrier
Polymers
a
Barrier polymer Water vapor transmission rate,
b
(gmil)/(100 in
2
day)
P(VDCAN) barrier coating 0.02
Aclar 33C PCTFE lm
c
0.025
PVDC coextrusion resin 0.09
Liquid crystalline polymer
d
0.15 (100% RH)
PP, biaxially oriented 0.25
HDPE, molded 0.3
MDPE, molded 0.7
PVC, rigid 0.9
LDPE, molded 1.0
Nylon MXD-6 oriented lm 1.2
EVOH, 44 mol% ethylene 1.4
PET, 25% crystalline (bottle wall) 1.8
Nitrile resin 4.0
EVOH, 27 mol% ethylene 5.7
PS lm, oriented 7.1
Nylon-6, biaxially oriented 10
Amorphous nylon 10
Nylon-6, unoriented 15
a
Ref. 27.
b
At 90% RH, 37.8

C.
c
Aclar is a trademark of Honeywell (Morristown, N.J.) (91).
d
Vectra A950 LCP lm (Celenese AG) (89).
moisture as its oxygen barrier (57). In the majority of commercial applications,
EVOH is used in a multilayer structure with moisture barrier and/or structural
layers on each side, a typical example being multilayer bottles with polypropylene
(PP) for ketchup (57). EVOH has also been used as a avor barrier on the inside of
PE-coated paperboard containers, a typical application being packaging of orange
juice, where it minimizes the loss of limonene from the juice into the PE layer
(57). EVOH is used as a barrier layer in many other rigid and exible packaging
applications.
Nitrile Polymers. The general structure of nitrile copolymers is shownbelow
(101):
CH
2
C
H
CN
CH
2
C
H
R
x y
where R = or COOCH
3
Application of PAN in packaging started in the 1960s. Despite processing
difculties, PAN was used because of its excellent barrier properties. A group
of copolymers were developed in the 1970s. Comonomers are methyl acrylate
and styrene at concentration up to 20 mol% (103). They are efcient oxygen
barriers and have high grease and oil resistance, strength, and stiffness. Lopac
(Monsanto Co.), Barex (Sohio), and Cycopac (Borg-Warner Chemicals) are three
234 BARRIER POLYMERS Vol. 5
Table 7. Permeability, Diffusivity, and Solubility Coefcients of Flavor and Aroma
Compounds in Various High and Moderate Barrier Polymers (63) at 25

C and 0% RH
Permeability, Diffusivity, Solubility,
Flavor/aroma compound 10
22
(kgcm)/(cm
2
sPa)
a
cm
2
/s kg/(cm
3
Pa)
Ethylene vinyl alcohol copolymer
b
Ethyl hexanoate 0.41 3.2 10
14
1.3 10
9
Ethyl 2-methylbutyrate 0.3 6.7 10
14
4.7 10
10
Hexanol 1.2 2.6 10
13
4.6 10
10
trans-2-Hexenal 110 6.4 10
13
1.8 10
8
d-Limonene 0.5 1.1 10
13
4.5 10
10
3-Octanone 0.2 1.0 10
14
2.0 10
9
Propyl butyrate 1.2 2.7 10
13
4.5 10
10
Vinylidene chloride copolymer
b
Ethyl hexanoate 570 8.0 10
14
7.1 10
7
Ethyl 2-methylbutyrate 3.2 1.9 10
13
1.7 10
9
Hexanol 40 5.2 10
13
7.7 10
9
trans-2-Hexenal 240 1.8 10
13
1.4 10
7
d-Limonene 32 3.3 10
13
9.7 10
9
3-Octanone 52 1.3 10
14
4.0 10
7
Propyl butyrate 42 4.4 10
14
9.4 10
8
Dipropyl disulde 270 2.6 10
14
1.0 10
6
Low density polyethylene
Ethyl hexanoate 4.1 10
6
5.2 10
9
7.8 10
8
Ethyl 2-methylbutyrate 4.9 10
5
2.4 10
9
2.3 10
8
Hexanol 9.7 10
5
4.6 10
9
2.3 10
8
trans-2-hexenal 8.1 10
5
d-Limonene 4.3 10
6
3-Octanone 6.8 10
6
5.6 10
9
1.2 10
7
Propyl butyrate 1.5 10
6
5.0 10
9
3.0 10
8
Dipropyl disulde 6.8 10
6
7.3 10
10
9.3 10
7
High density polyethylene
d-Limonene 3.5 10
6
1.7 10
9
2.5 10
7
Menthone 5.2 10
6
9.1 10
9
4.7 10
7
Methyl salicylate 1.1 10
7
8.7 10
10
1.6 10
6
Polypropylene
2-Butanone 8.5 10
3
2.1 10
11
4.0 10
8
Ethyl butyrate 9.5 10
3
1.8 10
11
5.3 10
8
Ethyl hexanoate 8.7 10
4
3.1 10
11
2.8 10
7
d-Limonene 1.6 10
4
7.4 10
12
2.1 10
7
Poly(ethylene terephthalate)
c
d-limonene 1.5 6.0 10
13
a
To convert the permeability values to (cm
3
mil)/(100 in.
2
dayatm), multiply the values provided in
the table by (4.98 10
22
)/MW, where MW is penetrant molecular weight (g/mol).
b
Values have been extrapolated from higher temperatures.
c
At 25

C; values are not expected to show any signicant variations with RH (92).
commercial nitrile copolymers that have been used in packaging applications.
Their compositions are shown in Table 8. Barex and Cycopac are rubber-modied
for improved mechanical properties, and the barrier properties of these copolymers
are relatively insensitive to moisture. Concerns about the possible migration of
Vol. 5 BARRIER POLYMERS 235
10
1
10
0
10
1
10
3
10
2
10
1
10
0
10
1
10
2
10
3
0 20 40 60 80 100
mol% Ethylene PVOH HDPE
EVOH
W
V
T
R
,

(
g


m
i
l
)
/
(
1
0
0

i
n
.
2


d
a
y
)
O
x
y
g
e
n

p
e
r
m
e
a
b
i
l
i
t
y
,

(
c
m
3


m
i
l
)
/
(
1
0
0

i
n
.
2


d
a
y


a
t
m
)
Fig. 19. Effect of ethylene content on oxygen permeability (2023

C, 65%RH) and WVTR


(40

C) of EVOH. The gure also shows data for PVOH and HDPE for comparison (27).
acrylonitrile monomer, a toxic compound (57), have limited the use of nitrile
polymers in food contact applications (see ACRYLONITRILE AND ACRYLONITRILE
POLYMERS).
Vinylidene Chloride Copolymers. Copolymers of VDC with vinyl chloride
and acrylonitrile were among the rst high barrier polymers to be widely used.
During their commercial appearance in the late 1930s, they had the lowest perme-
abilities among plastics to gases and water vapor (103). Poly(vinylidene chloride)
(PVDC) homopolymer is soluble only in hot dichlorobenzene (among common sol-
vents) and has a melting point only a fewdegrees belowits decomposition temper-
ature (103). These characteristics make it difcult to fabricate by melt-processing
Table 8. Compositions of Commercial, High Barrier Nitrile Copolymers
a
Polymer Manufacturer
b
Chemical composition
c
Lopac Monsanto Co. 70% acrylonitrile + 30% styrene
Barex Sohio 74% acrylonitrile + 26% methyl methacrylate
+ 10% butadiene graft rubber
Cycopac Borg-Warner 74% acrylonitrile + 26% styrene
Chemicals + 10% butadiene graft rubber
a
Ref. 63.
b
Manufacture are listed from Ref. 63. However, these copolymers are now manufactured by BP
Amoco Chemicals.
c
Data from FDA regulations for corresponding materials.
236 BARRIER POLYMERS Vol. 5
techniques. Copolymers were synthesized to overcome these drawbacks. Acrylates
were found to be among the most useful comonomers, along with vinyl chloride
and acrylonitrile (103). By adding comonomers, the melting point can be decreased
to a range of 140175

C (as compared to 198205

C for PVDC), thus making melt


processing feasible (104). These copolymers are semicrystalline and soluble in
only a limited range of solvents. The most notable attributes of VDC copolymers
are their chemical resistance and extremely low permeabilities to gases and wa-
ter vapor (cf Tables 5 and 6). The structure of the most widely used vinylidene
chloridevinyl chloride copolymer is shown below:
where x is 8590 mol% (103). VDC copolymers are commercially available under
a variety of trade names such as Saran (The Dow Chemical Co.), Daran (W. R.
Grace), Amsco Res (Union Oil), and Serfene (Rohm and Haas) within the United
States; Haloex (Neoresins), Diofan (Solvin), Ixan (Solvin), and Polyidene (Scott-
Bader) in Europe (see VINYLIDENE CHLORIDE POLYMERS). Copolymers are available
in the following forms (103): (1) lattices of approx. 10001500

A which can be
applied as coatings to paper and plastic lms to improve their barrier proper-
ties, (2) soluble resins for coating plastic lms (especially cellophane) to improve
their barrier properties, (3) melt-processable resins for extrusion, coextrusion, and
molding, and (4) clear, transparent lms for commercial packaging applications.
Small amounts of processing aids and heat stabilizers are added to extrusion and
molding resins.
Polyamides. The standard semicrystalline polyamides (PA) (nylon-6,
nylon-6,6, etc) used in packaging have medium gas barrier properties, and their
barrier properties are affected by humidity (101). However, specialty grades of
polyamides with higher gas barrier properties are available. Commercial grades
of Selar amorphous polyamides (AmPA) (E. I. du Pont de Nemours & Co., Wilm-
ington, Del.), for example, exhibit good O
2
barrier and reduced dependence of gas
barrier properties on RH. In fact, their gas barrier properties improve with in-
creasing RH (cf Fig. 13). At an RH of 80% or more, their O
2
barrier is similar to
that of PAN (102). Moreover, at 95100% RH, the O
2
barrier is equivalent to that
of EVOH at similar conditions, and substantially better than that of nylon-6. The
amorphous nature of Selar results in a much broader range of processing con-
ditions than those of semicrystalline nylon-6 (105). The mechanical and barrier
properties of AmPA can be improved by orientation (see POLYAMIDES, PLASTICS;
POLYAMIDES, AROMATIC).
Another high barrier polyamide is MXD-6 resin (Mitsubishi Gas and Chem-
ical Co., and Toyobo, Japan), which was developed in the 1970s. It is made from
the reaction of m-xylylenediamine and adipic acid (57):
Vol. 5 BARRIER POLYMERS 237
It provides improved clarity, mechanical, thermal, and barrier properties rel-
ative to standard nylons. It has better gas barrier properties than nylon-6 and PET
at all humidities, and is better than EVOH at 100% RH. The barrier properties of
MXD-6 are relatively unaffected by moisture up to an RH of 70% (57). Because of
its cost and the lack of a domestic source, MXD-6 has found limited applications in
the United States. It has a much wider market in Japan, appearing in commercial
applications such as nonpasteurized plastic beer bottles and carbonated soft drink
bottles (57).
Polyesters. The most widely used member of the polyester family for food
and beverage packaging applications is PET (see POLYESTERS, THERMOPLASTIC).
However, PET offers a moderate barrier to gases and water vapor. Poly(ethylene
naphthalate) (PEN) offers a much higher barrier to gases and water vapor than
PET and can be classied as a high barrier polyester.
PEN is a homopolymer of dimethyl-2,6-naphthalene dicarboxylate (NDC)
and ethylene glycol (106):
The rigid double-ring structure in the polymer backbone results in increased
mechanical strength, heat stability, and barrier properties as compared to PET.
As with PET, orientation produces a substantial reduction in gas permeability,
and the oxygen permeability of oriented PEN is a factor of 5 lower than that of
oriented PET (57). Teijin (Japan) and ICI (U.K.) have been the leaders in man-
ufacturing PEN lms. Teijin manufactures both resin and lms, the latter un-
der the trade name Teonex. Since 1990, DuPont and Teijin have had a world-
wide joint venture in polyester lms, including PEN. DuPont has also acquired
ICIs Melinex PET lm and Kaladex PEN lm operations. The main disadvan-
tages of PEN are its high cost, and unsettled sources of monomer technology
and supply (57). One of the leading manufacturers of NDC, BP Amoco Chemi-
cals (Decatur, Ala.) has developed a technology that uses o-xylene rather than
naphthalene as the feedstock, which could reduce the manufacturing cost of PEN
resin (57). Other newer NDC technologies from Kobe Steel (Japan) and Mobil
Chemicals could also lead to reduced cost for PEN. Another way of addressing
the price issue is by using blends or copolymers of PEN with PET. Gas and va-
por permeability has been found to decrease continually as PEN is added to PET
(107,108).
Liquid Crystalline Polymers. Liquid crystalline polymers, main-chain (qv)
(LCPs) offer excellent thermal and chemical resistance, and exhibit very high
barrier properties that are almost unmatched by existing barrier polymers
(16,109,110). They also offer adequate mechanical properties for certain packag-
ing applications (111). LCPs are very efciently packed, highly oriented, and often
semicrystalline materials. Commercially available LCPs such as Vectra (Celanese
AG), Zenite (E. I. du Pont de Nemours & Co.), and Xydar (Solvay Advanced
238 BARRIER POLYMERS Vol. 5
Polymers, Alpharetta, Ga.) are aromatic copolyesters that have a signicantly
higher degree of chain orientation than typical polyesters such as PET (111).
LCPs of this kind were rst introduced to the market in the 1980s.
Table 9 shows a comparison of properties of biaxially oriented LCP lms with
those of PET lms (112). Figure 20 compares barrier properties of LCPs with
other barrier polymers. LCPs offer the best combination of water vapor and O
2
barrier properties among all known classes of polymers. They also offer excellent
barrier to CO
2
, N
2
, and other gases and vapors (111). However, commercial ap-
plications of LCPs have been limited primarily because of their high cost, lack
of transparency, and processing characteristics (111). One key to unlocking the
potential of LCPs in barrier packaging is to be able to process thin uniform layers
in coextruded multilayer structures. Recently, three-layered structures (PETtie
layerLCP) have been coextruded (Superex Polymer, Inc., Waltham, Mass.) with
a total thickness of 2550 m and 1030% LCP layer thickness. This multilayer
lm is claimed to offer a high performancecost ratio (111).
Poly(hydroxy amino ethers). Poly(hydroxy amino ethers) (PHAE) are a
new family of high barrier epoxy-based thermoplastics. The general chemi-
cal structure of amide-containing PHAE, which are formed by the reaction of
Table 9. Comparison of Properties of Biaxially Oriented LCP and PET Films
a
Biaxially oriented Biaxially oriented
Property LCP lm
b
PET lm
c
Tensile strength, MPa
d
240 170
Tensile modulus, GPa
e
12.4 3.5
Oxygen permeability
f
, 0.05 4.8
(cm
3
mil)/(100in.
2
dayatm)
Water vapor permeability,
g
0.02 1.7
(gmil)/(100in.
2
dayatm)
Density, g/cm
3
1.4 1.4
Upper use temperature,

C Over 250 120
Tear resistance, kN/m
h
Initiation 595 35
Propagation 175525 953
a
Ref. 112.
b
Vectra (Celanese AG) isotropic LCP lm (orientation angle = 45

).
c
Mylar (DuPont) isotropic PET lm (orientation angle = 45

).
d
To convert MPa to psi, multiply by 145.
e
To convert GPa to psi, multiply by 145,000.
f
Permeability value at 25

C (and unspecied RH).


g
Permeability value at 25

C and 90% RH.


h
To convert kN/m to lb/in., divide by 0.175.
Vol. 5 BARRIER POLYMERS 239
10
2
10
1
10
0
10
1
10
2
Thermotropic
LCP
EVOH
Nylon-6
MXD-6
PAN
PET
PEN
PVDC
PCTFE
Biax PP
HDPE
10
2
10
3
10
2
10
1
10
0
10
1
10
3
W
a
t
e
r

v
a
p
o
r

p
e
r
m
e
a
b
i
l
i
t
y
,

(
g


m
i
l
)
/
(
1
0
0

i
n
.
2


d
a
y


a
t
m
)
Oxygen permeability, (cm
3
mil)/(100 in.
2
day atm)
Fig. 20. A comparison of oxygen and water vapor barrier properties of various high and
moderate barrier polymers at 23

C (111).
amide-containing bisphenols with aromatic diglycidyl ethers (113), is shown
below:
where R can be (CH
2
)
n
(n = 1, 2, 3. . .) or an aromatic group, and Ar designates
an aromatic moiety.
These polymers are amorphous, with T
g
values ranging from 90 to 133

C.
They incorporate both amide and hydroxyl moieties on the chain backbone. In
order to prevent cross-linking, the hydroxyl groups are generated during the poly-
merization of preferred amide-containing monomers (113). The presence of aro-
matic groups between the amide groups in the polymer backbone can result in
high T
g
and good barrier properties. It has also been observed that the presence
of m-phenylene units instead of p-phenylene units can reduce O
2
permeabilities
(as much as 3040%) by increasing the chain packing efciency of the polymer
(113). This phenomenon is, in fact, rather general among aromatic polymers. Of-
ten, meta-linked aromatic rings in polymer backbones have lower permeability
coefcients than their para-linked analogues (114,115). Lower O
2
permeabili-
ties can also be obtained when hydrogen bonding interactions in the polymer
backbone are increased either by reducing the number of nonpolar methylene
units or by increasing the population density of polar amide groups (113). Un-
like other hydroxyl-containing polymers such as EVOH, the barrier properties
240 BARRIER POLYMERS Vol. 5
of amide-containing PHAE improve with increasing RH (19,113). For example,
a decrease in O
2
permeability from 1.4 (cm
3
mil)/(100 in.
2
dayatm) at 5% RH to
0.8 (cm
3
mil)/(100 in.
2
dayatm) at 7580% RH has been reported. On the basis
of preliminary work involving density and pals studies, it has been postulated
that water molecules not only occupy free-volume elements in these materials but
also enhance interchain cohesion, thus inhibiting the transport of other nonpolar
gases (116).
A series of polymers from the PHAE family has recently been commercial-
ized by DowChemical Co. (Midland, Mich.) under the trade name BLOX Adhesive
and Barrier Resins (90,117). They offer high gas barrier properties, excellent ad-
hesion to a variety of substrates, high optical clarity, and good mechanical proper-
ties. For example, BLOX 4000 series resins exhibit an oxygen transmission rate
of 0.1 (cm
3
mil)/(100 in.
2
dayatm) at 23

C and 60% RH (90). The BLOX resins


have found some commercial applications in barrier packaging, starch-based foam
packaging, and powder coatings (117).
Polychlorotriuoroethylene. Polychlorotriuoroethylene (PCTFE) is a ex-
ible thermoplastic made from uorinatedchlorinated resins. It was rst discov-
ered in the 1950s and has been commercially produced since the 1960s (118). Its
chemical structure is as shown below:
The key characteristics of this polymer are its high optical clarity and ex-
cellent moisture barrier properties (118). In commercial Aclar resins (Honeywell,
Morristown, N.J.), the polymer is generally modied by copolymerization, result-
ing in a semicrystalline material with T
g
of about 45

C and melting point of about


190

C (104). It shows excellent thermal and chemical stability, and high water
vapor and O
2
barrier properties (120) (cf Tables 5 and 6). For example, Aclar lms
typically have water vapor transmission rates less than 0.04 (gmil)/(100 in
2
day)
(23

C) and oxygen transmission rates less than 14 (cm


3
mil)/(100 in
2
day) (25

C)
(119). PCTFE is most widely used for producing blister packs in pharmaceutical
applications (118). The moisture barrier properties can be tailored by varying the
PCTFE layer thickness in coextruded multilayer structures, thereby adjusting
the performance-to-cost ratio. An example of a commercial coextruded structure
is Aclar NT AE-1 (Honeywell, Morristown, N.J.), which is 33 m thick, and con-
tains about 8.4 m of PCTFE (118). PCTFE can also be laminated with PP, PAN,
PET, HDPE, LDPE, and PVC and then formed into blister packs. Other appli-
cations of the same technology include electronic component packaging where
lower moisture barrier may be acceptable, resulting in more favorable economics
(118).
Moderate Barrier Polymers. Polymers in this category include
polyesters, polyolens, poly(vinyl chloride), polystyrene, and certain semicrys-
talline polyamides. The polymers included in this section are typically more widely
available, have more manufacturers, and are less expensive than the specialty
barrier resins described in the previous section.
Polyesters. Polyesters represent a class of versatile barrier plastics. PET
is by far the most important member in this family from a commercial
Vol. 5 BARRIER POLYMERS 241
viewpoint. It is extensively used in the food and beverage packaging industries
and is especially known for its widespread use in bottles for carbonated beverages
(120) (see POLYESTERS, THERMOPLASTIC). The other members in this category are
poly(trimethylene terephthalate) (qv) (PTT) and Polylactide (PLA).
Poly(ethylene terephthalate). PET is a linear thermoplastic made fromethy-
lene glycol and terephthalic acid, or ethylene glycol and dimethyl terephthalate
(101). The structure of PET is shown below:
PET is used in many rigid food and beverage containers because of a good
balance of physical and mechanical properties, barrier properties, processibility
and formability, ecological and toxicological characteristics, and economics (121).
As a result, PET bottles have virtually replaced glass packages for carbonated
soft drinks in the United States. In the glassy state, it is strong, stiff, ductile,
and tough. It can be oriented by stretching during molding and extrusion, which
further increases its strength, stiffness, and barrier properties. One of the early
drawbacks of PET was its low heat distortion temperature of 60

C (140

F), which
prevented it from being used in applications requiring lling at elevated tem-
peratures (121). However, careful heat-treatment increases the heat distortion
temperature of crystalline PET containers and recently commercialized processes
claim resistance to temperatures of up to 9092

C (194196

F) (121). It has mod-


erate barrier properties for light gases, but is a good barrier for avors and aroma
compounds. Its barrier properties can be improved by increasing crystallinity and
orientation (121). A 4 biaxial orientation of amorphous PET at temperatures
near 100

C produces signicant strain-induced crystallization and decreases the


permeability of O
2
and CO
2
by a factor of about 2 (71).
Oriented and heat-set PET lms have also found use in a broad range of
exible packaging applications because of their high strength, good barrier, high
clarity, heat resistance, and good metalizability. There are more than 50 specic
application areas for PET lms (122). Manufacturers have met the product re-
quirements of each of the diverse end-use markets by tailoring formulations and
process conditions. For example, in the food and beverage packaging industries,
many types of PET lms have been developed, including metallized PET lms
for packaging of coffee, wine, and meats, poly(vinylidene chloride)-coated PET
lms for meat and cheese packaging, and coextruded multilayer PET lms for
heat-sealable packaging (122). Commercial manufacturers of PET lms include
DuPont Teijin Films (Mylar, Melinex, Tetoron) and Mitsubishi Polyester Film,
LLC. (Diafoil, Hostaphan).
Poly(trimethylene terephthalate). PTT is made from the polycondensation
reaction of trimethylene glycol (also called 3G) with either terephthalic acid or
dimethyl terephthalate (57). Although this polymer was rst synthesized in 1941,
it was never commercialized because of lack of an economical source of 3G. In the
242 BARRIER POLYMERS Vol. 5
early 1990s, Shell Chemical Co. announced a catalyst breakthrough to make 3G
economically by hydroformylation of ethylene oxide (57).
PTT is an engineering resin and has been targeted mainly toward injection-
molding applications. However, there have beenclaims about its barrier properties
(57), and it is therefore included in this article for completeness.
Because of the presence of an odd number of methylene units in the chain
backbone, PTThas physical properties different fromPET(57). Its O
2
permeability
is about 6 (cm
3
mil)/(100 in.
2
dayatm) at 0% RH, and moisture absorption rate
is typically 0.03% after 24 h and 0.15% after 14 days (57). Typical properties of
PTT are as follows: it has a melting point of 228

C, a T
g
of 4565

C, and can be
produced with crystallinity values up to 45 wt%.
Polylactide. Polylactide (PLA) is a semicrystalline, linear thermoplastic
made from lactic acid, especially as derived from corn (maize). Polylactide (qv)
is new to the commercial market (123). Its applications are still being explored.
However, its barrier is adequate for some food packaging uses, especially for man-
agement of avor and aroma proles. It can be heat sealed and thermoformed.
It has a glass-transition temperature of 5565

C. The principal manufacturer of


PLA is Cargil Dow, LLC (NatureWorks).
Polyamides. This category of moderate barrier polymers includes nylon-6
and nylon-6,6. Nylon-6 is made by the polymerization of caprolactam, and nylon-
6,6 is made by the reaction of hexamethylenediamine and adipic acid (101). Their
chemical structures are shown below:
Ingeneral, nylons have good gas and aroma barrier properties, but poor mois-
ture barrier properties (57) (cf Tables 5 and 6). Absorbed water has a plasticizing
effect that leads to a reduction in tensile strength and increase in impact strength.
Uniaxial and biaxial orientations can improve their ex-crack resistance, mechan-
ical, and barrier properties (101). Biaxially oriented nylons offer better gas barrier
properties, softness, and puncture resistance, compared to oriented PET, which
offers better rigidity and moisture barrier properties (101). Nylons are less widely
used in the packaging industry than polyolens or PET, with the majority of the
applications being blow-molded bottles (coextrusion and blending of nylon-6 with
PE), for industrial and household chemical markets (57).
Polyolens. PEand PPare two of the most widely used polymers in the food
and beverage packaging industry. These polymers nd use as lms, moldings,
coatings, adhesives, and closures (124). They are available in a wide variety of
Vol. 5 BARRIER POLYMERS 243
Density High Low
Property
Low
High
Stiffness
Tensile strength
Chemical resistance
Gas and vapor transmission
Elongation at break
LTI
ESCR
Fig. 21. Typical effect of density on various properties of PE. LTI is low temperature
impact strength and ESCR is environmental stress-crack resistance (124).
types and grades. Although they have much higher permeabilities to gases than
many other barrier polymers, they are very good moisture barriers (cf Tables 5
and 6).
PE, whose structure is shown below, was one of the rst olenic polymers to
be used commercially in the packaging industry:
It is classied on the basis of density (see ETHYLENE POLYMERS, HDPE;
ETHYLENE POLYMERS, LDPE; ETHYLENE POLYMERS, LLDPE; ETHYLENE POLYMERS,
VLDPE). Figure 21 shows the effect of density on various properties of PE (124).
A branched structure for low density PE (LDPE) results from exceptionally high
temperature and pressure during its manufacture. It is tough, exible, can be
easily melt processed, and has good moisture barrier properties. It is a semicrys-
talline polymer with crystallinity typically in the range of 40%. Medium density
PE(MDPE) is stronger, stiffer, and has better barrier properties than LDPE. High
density PE (HDPE) is essentially unbranched and is the strongest and most rigid
polymer in this family. It offers barrier properties to moisture and gases that are
superior to those of LDPE and MDPE.
If unsaturated comonomers such as butene, hexene, or octene are added to
the HDPEpolymerization process in the presence of a stereospecic catalyst, it re-
sults in the formation of a linear polymer with short branch-like pendant groups
(104). Its density is in the same range as LDPE, but the degree of branching
is greatly reduced. This polymer is called linear LDPE (LLDPE) and its den-
sity depends on the amount of comonomer added. The larger the amount of the
comonomer, the lower the density of the copolymer (104). LLDPE combines the
244 BARRIER POLYMERS Vol. 5
clarity and excellent heat-seal properties of LDPE with the strength and tough-
ness of HDPE. It is often blended with LDPE in order to optimize the benet
obtained from both materials (104).
PP, whose chemical structure is shown below, can be made by the catalytic
polymerization of propylene at high temperature and pressure (124):
Isotactic PP (with all methyl groups on same side of the polymer
chain) is the commercially desired form for packaging applications (124) (see
PROPYLENE POLYMERS (PP)). The concentration of atactic PP (with irregular ar-
rangement of methyl groups) is kept low by suitable catalysts and polymerization
conditions (124). PPoffers high resistance to water vapor permeation and is widely
used in rigid as well as exible food packaging applications. PP lms can be ori-
ented, which improves their barrier properties, mechanical strength, and optical
properties. For example, oriented PP has about three times higher resistance to
water vapor transmission than unoriented PP (27). These properties can be varied
over a wide range by the choice of the manufacturing process.
Poly(vinyl chloride). Poly(vinyl chloride), also known as vinyl or PVC, is
made by lowpressure free-radical polymerizationof vinyl chloride at temperatures
in the 3871

C range (101). Its chemical structure is shown below:


PVC is a versatile polymer that can be formulated to meet the requirements
of many applications in packaging and other industries (88) (see VINYL CHLORIDE
POLYMERS). Its properties can be tuned over a very wide range by incorporating
comonomers, plasticizers, and other additives. When used as a rigid sheet or bot-
tle, little modication is required, but PVC requires the addition of plasticizers
to make it useful as a barrier lm for exible packaging (88). The plasticizers
increase chain exibility and reduce the processing temperature of PVC. For ex-
ample, addition of 40 vol% dioctyl phthalate plasticizer reduces the T
g
of PVC
from 100

C to about 5

C (125). The increase in chain exibility of plasticized PVC


also results in a reduction in its gas barrier properties primarily as a result of
higher diffusion coefcients in plasticized lms. Figure 22 presents a comparison
of diffusion coefcients of several penetrants in plasticized and unplasticized PVC
lms (126) (see PLASTICIZERS).
PVC has some drawbacks as a food-packaging material (88): Vinyl chloride
monomer is an animal carcinogen and causes liver cancer in humans. Thus the
amount of monomer in the nished polymer should be brought down to 1 ppm or
less. There is concern about the toxicity of plasticizers and other additives used
in PVC. Moreover, there have also been concerns that incineration of chlorine-
containing plastics can possibly lead to the formation of dioxin, a chlorinated
toxic molecule. Finally, the release of plasticizers over long periods of time can
lead to gradual embrittlement of PVC lms. However, none of these drawbacks
signicantly affect the functionality of PVC and it is used to make exible lms,
Vol. 5 BARRIER POLYMERS 245
10
18
10
16
10
14
10
12
10
10
10
8
10
6
1.2 1.0 0.2 0.4 0.6 0.8
Penetrant size, nm
D
i
f
f
u
s
i
o
n

c
o
e
f
f
i
c
i
e
n
t
s
,

c
m
2
/
s
Unplasticized
Plasticized
Fig. 22. Diffusion coefcients in unplasticized and plasticized PVC at 30

C as a function
of penetrant molecular diameter (126).
rigid sheets, and bottles for a variety of food (eg, fresh fruits, vegetables, and
poultry packaging in the United States) and nonfood packaging applications (104).
Polystyrene. Polystyrene (PS) is made by the peroxide-catalyzed bulk or
suspension polymerization of styrene (101) (see STYRENE POLYMERS). The poly-
merization reaction takes place at low pressure and temperature in the range of
250400

F, and the polymer chemical structure is given by


Polymer molecular weight, which affects the processing characteristics of
the resin, is often in the range of 40,000220,000; variations in molecular weight
can be obtained by changing the catalysts and polymerization conditions. PS is
an amorphous, clear, hard, brittle, low-strength material with poor impact re-
sistance (101). It has low to moderate moisture and gas barrier properties (88).
Copolymerization with butadiene or other rubbers increases its impact strength
and decreases stiffness (88). This copolymer is commonly referred to as high im-
pact polystyrene (HIPS). Comonomers such as -methyl styrene can improve the
heat resistance of PS by increasing the heat distortion temperature to 100

C or
246 BARRIER POLYMERS Vol. 5
higher (88). PS can be foamed by adding foaming agents such as pentane to the
reaction mixture during the suspension polymerization. This so-called expanded
polystyrene (EPS) is a very low density, yet highly rigid material that is used for
making egg cartons, and trays for meat, poultry, and other products (88). However,
EPS has poor gas barrier properties.
Improving Barrier Properties of Polymers
Barrier Structures. As mentioned in the previous sections, the principal
application of barrier polymers is in the food and beverage packaging industries.
Combining two or more polymers, or other materials, can achieve performance
advantages not available in any of the materials taken alone. In many cases, to
achieve better barrier performance, it is more efcient and economical to use a
thin layer of an expensive high barrier polymer (eg, EVOH) sandwiched between
layers of less expensive, moderate barrier, structural polymers (eg, PP, PET) than
to increase the monolayer thickness of the moderate barrier polymer. Multilayer
structures can be obtained by coextrusion, lamination, and coating. Barrier poly-
mers may also be combined to form miscible and immiscible blends.
Multilayer Structures. Steady-state barrier properties of multilayer lms
can be described by their permeability, which can be calculated by treating the
individual layers as resistances in series (63):
1
P

=
L
t
P
t
=
L
1
P
1
+
L
2
P
2
+ +
L
n
P
n
(38)
where P
t
, L
t
, and P

are permeability, thickness, and permeance of the compos-


ite structure, and P
1
, P
2
, . . . , P
n
and L
1
, L
2
, . . . , L
n
are the permeabilities and
thicknesses of the individual layers. The permeability P
t
(or permeance P

) can
be used to evaluate the performance of the composite structure for packaging
applications. Figure 23 illustrates the effect of barrier layer thickness on per-
meability of a hypothetical two-layer sheet, the permeability of the barrier layer
being 0.1 (cm
3
mil)/(100 in.
2
dayatm) and that of the nonbarrier layer being 100
(cm
3
mil)/(100 in.
2
dayatm). For moisture-sensitive polymers (eg, EVOH or ny-
lon), it is important to use the permeability that corresponds to the effective RH
that the layer will experience in the composite structure.
Coextrusion (qv) is one of the most cost-effective techniques for producing
multilayer barrier polymer lms (127). Although coextruded lms typically have
37 layers, as many as 11 layers can be extruded simultaneously (128). This
technique allows for the thinnest possible layers of the individual polymer resins
within the structure. Resins that do not bond well can be bonded together with
an adhesive tie layer (128). The materials must be compatible in terms of their
melt temperatures and viscosities in order to undergo simultaneous coextrusion.
Figure 24 presents a comparison of processing temperatures of various commonly
used barrier polymers (129). Figure 25 shows a schematic of a typical 9-layer co-
extruded structure. A coextrusion line will have several different extruders, each
extruder responsible for supplying individual polymer resins (127).
Vol. 5 BARRIER POLYMERS 247
10
1
10
0
10
1
10
2
0 0.2 0.4 0.6 0.8 1
Barrier layer thickness fraction
P
e
r
m
e
a
b
i
l
i
t
y
,

(
c
m
3


m
i
l
)
/
(
1
0
0

i
n
.
2


d
a
y


a
t
m
)
Fig. 23. Effect of thickness of barrier layer on permeability of a hypothetical two-layer
barrier composite structure.
100
150
200
250
300
350
MXD-6
Nylon-6
Nylon-6,6 PET PC
PE
PP
EVOH
P
r
o
c
e
s
s
i
n
g

t
e
m
p
e
r
a
t
u
r
e
,

C
Fig. 24. A comparison of processing temperatures of various commercially used barrier
polymers (129).
Coating and lamination are two additional processes for producing multi-
layer structures and are especially useful in applications where a nonpolymeric
material is part of the structure. Coatings can be melt extrusions of a polymer onto
a base lm or can be made by applying solutions or dispersions of polymers to the
248 BARRIER POLYMERS Vol. 5
Skin layer (1530%)
Adhesive layers (23% ea.)
Barrier layers
(515%)
Adhesive layers
Recycle layer (3050%)
Skin layer (1530%)
Fig. 25. An example of a typical symmetrical nine-layer coextruded barrier composite
structure (128).
base lm (128). Barrier polymers that resist water vapor and provide gas barrier
are often laminated to paper and paperboard. Cellophane, a exible transparent
gas barrier polymer, can also be made moisture-proof by coating or laminating
with other polymer lms (27). A commercial application of solvent coating with
barrier polymers is VDC resin dissolved in a polar solvent and coated onto cel-
lophane or PET (63). Water-based emulsion coatings of VDC are used for plastic
PET beer bottles, primarily in the United Kingdom. They can typically lower O
2
permeability of PET bottles by nearly 60%(57). Several companies produce PVDC
emulsions for coating lms such as nylon, PE, PP, and PET to improve their O
2
and moisture barrier properties (57). In the mid-1990s, PPG Industries, Inc. in-
troduced Bairocade external epoxy-amine organic coatings for improving barrier
properties of PET bottles (130,131). It is claimed that these coatings can reduce
O
2
permeability of PET bottles by a factor of 6 or more with the level of barrier
improvement depending on the thickness and formulation used, and that coated
bottles extend the shelf life of carbonated beverages by a factor of 3, and beer by
a factor of 20 relative to uncoated bottles (131).
Another way of improving barrier properties of polymers is by coating them
with thin inorganic layers (57,132,133). This can typically improve barrier prop-
erties by a factor of 100, whereas the thickness of the barrier coating applied is
less than 0.5% of the base lm (57). Laminations and coatings of aluminum foil or
aluminum oxide on barrier polymers can provide signicant improvements in gas
barrier properties (57). These metallized lms contain an extremely thin layer
of aluminum, which not only enhances the barrier properties of the base lm,
but also provides a shiny metallic appearance. Because the layer is very thin, it
does not appreciably affect the strength and exibility of the base lm (134). The
barrier properties of metallized lms can approach those of pure aluminum foil.
However, unlike aluminum foil, metallized lms are not subject to ex-cracking
and hence, are better at maintaining their barrier properties (134). Barrier poly-
mers that have been successfully metallized include PP, nylons, PET, and unplas-
ticized PVC. As an example, 99% decrease in oxygen permeability and 98.5% de-
crease in water vapor permeability have been reported for metallized PET lms as
compared to unmetallized PET lms (135). Metallized lms are widely used in the
Vol. 5 BARRIER POLYMERS 249
exible packaging industry, an example being potato chip packages consisting of
multilayer structures of oriented PP, PE, and metallized PP (136).
However, one of the main disadvantages of metallized structures is that they
may not be transparenta desirable feature for many packaging applications. In
part to address this shortcoming, several alternative high barrier coating tech-
nologies, typically producing amorphous carbon or glass-like layers, are being
investigated. These newer coatings have achieved limited market penetration to
date, with most of the commercial activity in Europe and Japan. It is still un-
clear what their eventual importance will be, but these technologies are being
actively pursued. Among the various glass and ceramic materials used as barrier
coatings, silicon dioxide (silica) has been the most widely used (57,133). These
barrier composite structures are clear, microwaveable, and recyclable. Electron-
beam treatment and plasma-enhanced chemical vapor deposition (PECVD) are
two of the most widely used techniques for depositing thin silica layers on barrier
polymers (57,137). In the electron-beam treatment, a high energy electron-beam
source is used to vaporize silica, which then precipitates onto the polymer lm,
forming a continuous coating as the lm passes through a vacuum chamber. The
resulting coatings are uniform and can be as thin as 0.04 m. PECVD has the
advantage of being able to control coating density and thickness, by changing
the process variables (137). Figure 26 presents the effect of coating thickness on
oxygen permeability of a PECVD silica-coated PET lm (137). Barrier properties
of silica-coated polymers can be superior to those of high barrier polymers such
10
7
10
6
10
5
10
4
10
3
10
2
10 100
Coating thickness, nm
O
x
y
g
e
n

t
r
a
n
s
m
i
s
s
i
o
n

r
a
t
e
,

c
m
3
/
(
1
0
0

i
n
.
2


d
a
y


a
t
m
)
Fig. 26. Effect of plasma-deposited silica coating thickness on steady-state oxygen trans-
mission rates (33

C, 0% RH) of PET lms that have an uncoated thickness of 13 m (137).


250 BARRIER POLYMERS Vol. 5
10
3
10
2
10
1
10
0
10
1
10
2
10
3
102
1.3
1.0
<0.003
OPP
SiO
x
ORMOCER
OPP OPP
SiO
x
OPP
ORMOCER
O
x
y
g
e
n

t
r
a
n
s
m
i
s
s
i
o
n

r
a
t
e
,

c
m
3
/
(
1
0
0

i
n
.
2


d
a
y


a
t
m
)
Fig. 27. Inuence of ORMOCER coatings on oxygen transmission rates (OTRs) of SiO
x
-
coated oriented polypropylene (OPP) (OTR measured at 23

C and 70% RH) (138).


as PVDC and EVOH. Moreover, they are not inuenced by moisture and temper-
ature (57). However, because of poor adhesion and mechanical properties, for all
practical applications silica-coated lms have to be laminated.
Recently, inorganicorganic hybrid polymers have been developed, which can
be used as laminating agents and in conjunction with silica to enhance barrier
properties of polymers (138,139). An example is ORMOCER or organically modi-
ed ceramic (Fraunhofer Gesellschaft, Germany). Such materials can be used as
coatings as well as high barrier laminating agents in multilayer structures. The
use of ORMOCER as a top layer on silica-coated PP can signicantly improve its
oxygen barrier properties, as shown in Figure 27. The composite structure also
offers good water vapor barrier properties. Similar improvements in barrier prop-
erties of PET have been reported (138). Figure 28 shows a comparison of O
2
and
water vapor barrier properties of different barrier polymer composite structures
(140).
Another potential candidate as an effective barrier coating is diamond-like
carbon (DLC) (132). DLC refers to a group of amorphous, hard, and chemically
inert materials consisting of carbon bonded partially as diamond (sp
3
) and par-
tially as graphite (sp
2
) and containing 040% hydrogen atoms. These coatings are
transparent, exible, extremely impermeable, biocompatible, and adhere well to
a wide range of polymers.
Among the most recent commercial applications of barrier polymeric con-
tainers coated with thin inorganic layers is plastic beer packaging. Glaskin (Tetra
Vol. 5 BARRIER POLYMERS 251
10
2
10
1
10
0
10
1
10
2
10
3
BOPP (met)/PE
EVOH composites
PA/PE composites
PET/Met./PE
PET (met)/PE
Multilayer bottles
PETPEN (met)/PE
ORMOCERs
with inorganic layers
Al composites
10
2
10
1
10
0
10
1
10
3
10
4
O
x
y
g
e
n

t
r
a
n
s
m
i
s
s
i
o
n

r
a
t
e
,

c
m
3
/
(
1
0
0

i
n
.
2


d
a
y


a
t
m
)
Water vapor transmission rate, g/(100 in.
2
day)
Fig. 28. A comparison of oxygen and water vapor transmission rates of various bar-
rier polymer composite structures (140). BOPP (met) is biaxially oriented, metallized
polypropylene; PET (met.) is metallized poly(ethylene terephthalate); PEN (met) is metal-
lized poly(ethylene naphthalate); Al composites are aluminum composites.
Pak, Geneva, Switzerland) is a proprietary technology that utilizes a vacuum de-
position process for coating clear, extremely thin layers of silicon oxide on the
inside of blown plastic bottles (141). It offers excellent barrier properties for O
2
,
CO
2
, and avor compounds, and it is completely recyclable. For beer packaging, a
shelf life of 6 months has been claimed using this technology. BESTPET (Barrier
Enhanced Silica Coated PET) is another silica-based coating technology developed
by the Coca Cola Co. (Atlanta, Ga.) in conjunction with Applied Films (Longmont,
Colo.) and Krones (Nevtraubling, Germany). The exterior surface of PET bottles is
coated with a thin silica layer and improvements in barrier properties of PET by a
factor of at least 2 have been claimed, which results in a shelf life of over 6 months
for packaged beer (142). A carbon-based high-barrier bottle coating technology
is ACTIS (Amorphous Carbon Treatment on Internal Surface technology) (Sidel,
France) (143). In this process, the internal surface of plastic bottles is coated with
a 0.15 m thick layer of highly hydrogenated amorphous carbon obtained from a
food-safe gas (eg, acetylene) in its plasma state. This technology has been claimed
to improve O
2
barrier properties of PET bottles by a factor of 30, and CO
2
barrier
properties by a factor of 7. Another carbon-based DLC coating technology (Kirin
Brewery, Japan) claims to improve the barrier properties of PET for O
2
, CO
2
, and
H
2
O by factors of 20, 7, and 8 respectively (142).
Miscible and Immiscible Blends. Polymer blending offers an alternative,
simple, unique, and economical approach for improving barrier properties for
several applications (see POLYMER BLENDS). In general, the goal is to add small
252 BARRIER POLYMERS Vol. 5
amounts of a high barrier polymer (generally more expensive) to a selected ma-
trix polymer (generally low cost) (40). High cost polymers such as PEN and LCPs
can be blended with lower cost polymers such as PET to achieve a balance of
barrier properties and cost. A common objective of blending is to attenuate the de-
ciencies while maintaining as much as possible the desirable properties of each
component. Occasionally, synergistic effects result in blend properties better than
those of the individual components (40). Reactive blending is also possible, for
example, with the mixing of different polyester polymers (transesterication), or
polyester and nylon (polyesteramide formation) (40).
In general, polymer blends can be broadly classied as homogenous or misci-
ble blends and multiphase or immiscible blends. The permeability coefcient P of
miscible blends as well as copolymers often follows an empirical semilogarithmic
additivity rule (40):
lnP=
1
lnP
1
+
2
lnP
2
(39)
where
i
is volume fraction of the ith component and P
i
is the components per-
meability coefcient. This simple additivity rule is generally obeyed only if there
are no interactions between the components. Deviations from this rule can ei-
ther be positive or negative depending on the nature and magnitude of inter-
actions (40). For example, a miscible blend of styreneacrylonitrile copolymers
(SAN) containing 9% acrylonitrile (AN) and tetramethyl bisphenol A polycar-
bonate (TMPC) shows negative deviations from the linear additivity rule, indi-
cating strong polymerpolymer interactions. On the other hand, blends of SAN
containing 13.5 and 28% AN, and PMMA show positive deviation from the linear
additivity rule (40).
When immiscible polymers are blended, or when inorganic ller is added to
a polymer matrix, it results in the formation of a dispersion of one component in
a continuous matrix of the other. Figure 29 shows several schematic examples of
suchsystems (144). Immiscible polymer blends are far more commonthanmiscible
blends (40). Barrier properties of animmiscible blenddependonthe permeabilities
of the individual components, their volume fractions, phase continuity, and the
aspect ratio of the dispersed (or discontinuous) phase (40). The aspect ratio L/W
refers to the shape of the particles in the dispersed phase. Spheres and cubes have
an aspect ratio of 1, whereas platelets and rods have higher aspect ratios.
The presence of an impermeable dispersed phase lowers the permeability
by increasing tortuosity. The Maxwell model can be used to calculate the perme-
ability of a polymer blend, P, with impermeable spherical particles dispersed in a
continuous phase (40):
P=
P
m
_
1
d
_
1+
d
/2
(40)
where P
m
is permeability of the continuous polymer phase and
d
is volume
fraction of the dispersed phase. Different models for permeation in a heteroge-
neous medium wherein the dispersed phase is impermeable and represented by
Vol. 5 BARRIER POLYMERS 253
(a) (b)
(c) (d)
Fig. 29. Examples of heterogeneous, immiscible blends. (a) Random spheres in a dis-
persed phase, (b) aggregated spheres in a dispersed phase, (c) oriented platelets in a dis-
persed phase, and (d) oriented rods in a dispersed phase (144).
different geometrical shapes have been reviewed (144). Several modications to
the Maxwell model have been made to describe permeation behavior in platelets
and ellipsoid-shaped particles dispersed in a more permeable continuous phase
(40). For example, oxygen barrier properties of blends of oriented PET and EVOH
have been successfully modeled with one such modication (Fricke model) (145).
Robeson extended the Maxwell model by applying it to blends in which both the
polymers contribute to the continuous phase (40). This model considers the prac-
tical implications of attempts to increase the barrier performance of a moderate
barrier polymer by adding small amounts of a high barrier polymer to it. Accord-
ing to this model, a multilayer structure provides the highest barrier followed by
a blend in which the high barrier polymer is the continuous phase (40).
Figure 30 shows the improvement in O
2
barrier properties of PEEVOH
blends as a function of EVOH volume fraction. Blends of PET with MXD-6 nylon
254 BARRIER POLYMERS Vol. 5
10
2
10
1
10
0
10
1
10
2
10
3
0 0.1 0.2 0.3 0.4 0.5
EVOH volume fraction
a
b
O
x
y
g
e
n

p
e
r
m
e
a
b
i
l
i
t
y
,

(
c
m
3


m
i
l
)
/
(
1
0
0

i
n
.
2


d
a
y


a
t
m
)
Fig. 30. Effect of EVOH content on oxygen permeability of PEEVOH blends that exhibit
(a) discontinuous morphology and (b) co-continuous lamellar morphology (40).
(Mitsubishi Gas Chemical Co.) have shown signicant reduction in O
2
and CO
2
permeability coefcients relative to those of PETand are being investigated for use
in plastic containers (40). PETEVOH blends have been promoted (Kuraray Co.)
for potential applications in the beverage industry (57). The transport properties
of PETLCP blends have also been studied, and Table 10 shows the reductions
in permeabilities of O
2
, N
2
, and CO
2
that have been reported for blown lms
containing 2, 10, and 30 wt% LCP (146).
Polymer Nanocomposites. Polymer nanocomposites are immiscible blends
made by adding nanometer-size particles to barrier polymers. Since nanometer-
size grains, bers, and plates have dramatically increased surface area compared
Table 10. Gas Permeabilities
a
of LCPPET blends
b
%LCP
c
in PET CO
2
permeability O
2
permeability N
2
permeability
0 114.5 21.3 5.1
2 78.1 14.9 2.6
10 65.8 12.0 3.1
30 34.5 5.5 2.0
a
(cm
3
mil)/(100 in.
2
dayatm).
b
Ref. 40.
c
Rodrun LC3000 LCP.
Vol. 5 BARRIER POLYMERS 255
to conventional-size materials, the chemistry of nanosized materials is different
from other conventional materials. Polymers lled with nanometer-size particles
have signicantly different properties than those lled with conventional inor-
ganic materials (147). Properties of nanocomposites such as high tensile strength
can be achieved by using higher conventional ller loading, but other proper-
ties such as improved clarity cannot be duplicated by lled resins at any loading
(148).
Polymer nanocomposites were developed in the late 1980s, and were rst
commercialized by Toyota, which used nanocomposite parts in one of its car mod-
els for several years (57). Initial developments focussed on the use of nylon resins
and very ne smectite clay particles, with surface area of about 750 m
2
/g, as
llers (57). Ube Industries developed its rst nylon nanocomposite in 1989 for an
automotive timing belt cover. They have also developed other nylon nanocompos-
ites called nylon clay hybrids (57). More recently, novel nanocomposite nylon-6
resins developed by Honeywell Engineered Applications and Solutions (Morris-
town, N.J.) have been claimed to improve O
2
and CO
2
barrier properties by a factor
of 34 (149). It has also been claimed that these nanocomposites double the heat
resistance of nylon-6 and improve other mechanical properties by 3050% (149).
Others such as Nanocor, Inc. (150) are also actively developing nanocomposites for
enhanced barrier performance. Commercial products are available for PP and u-
oropolymers. Some of the other potential candidate polymers for nanocomposites
include polyesters, PS, and ethylene vinyl acetate copolymers (57). Several poly-
mer nanocomposites, including amorphous nylon and EVOH matrixes, intended
for high barrier applications are in various stages of development.
Oxygen-Scavenging Systems. Another strategy for improving the bar-
rier properties of polymers is the introduction of reactive groups in the polymer.
These groups can reduce the transmission of penetrants such as oxygen and wa-
ter vapor and the term active barrier is often used to describe this approach
to distinguish it from passive barrier packages that rely on reduced permeabil-
ity to decrease gas transmission. By using an oxygen scavenger, which absorbs
the residual oxygen after packaging, quality changes of oxygen-sensitive foods
can be minimized (151153). Although oxygen-scavenging technology is rapidly
evolving and encompasses a wide variety of chemistries, the majority of current
commercial oxygen-scavenger packages probably still employ sachets that remove
oxygen from the headspace via iron oxidation. Ageless (Mitsubishi Gas Chemi-
cal Co., Japan) oxygen absorbers are the most commonly used sachets that are
placed inside the food package (57). If the initial oxygen concentration and oxy-
gen permeability of the packaging polymer is known, then an oxygen scavenger
can be chosen with a higher capacity than the theoretically needed capacity, and
near total absence of oxygen can be maintained during the expected shelf life of
the product (151). Other iron-based oxygen-scavenger sachets are ATCO (Standa
Ind., France), Freshilizer (Toppan Printing Co., Japan), Vitalon (Toagosei Chem.
Ind., Japan), Freshpax (Multisorb Technologies Inc., U.S.), and Sanso-cut (Finetec
Co., Japan) (151).
An alternative to sachets is to incorporate the oxygen scavenger into the
barrier polymer structure itself (151). An example of this strategy is Oxbar
(Crown Cork and Seal, Philadelphia, Pa.), which involves cobalt-catalyzed
oxidation of nylon MXD-6 polymer used in multilayer PET bottles for packaging of
256 BARRIER POLYMERS Vol. 5
0
0.05
0.1
0.15
0.2
0 100 200 300
Time, days
No scavenger
(a) 50 ppm Co
(b) 200 ppm Co
B
o
t
t
l
e

w
a
l
l

o
x
y
g
e
n

p
e
r
m
e
a
n
c
e
,

c
m
3
/
(
1
0
0

i
n
.
2


d
a
y


a
t
m
)
Fig. 31. Reduction in oxygen transmission rates due to oxygen scavenging in blends of
PETand nylon MXD-6 (157). The measurements were made at 23

Cand 50%RH. (a) 4 wt%


MXD-6 and 50 ppm cobalt (as metal); (b) 4 wt% MXD-6 and 200 ppm cobalt (as metal). For
comparison, the dotted line indicates oxygen permeance in the absence of cobalt catalyst
(no scavenging).
beverages. Another example of a polymer-based absorber is Amosorb (BP Amoco
Chemicals), which can reportedly be incorporated into various rigid and exi-
ble packaging structures (151). This is a rapidly evolving eld and several other
companies have also introduced or announced oxygen-scavenging resins. Among
those active in this area are EVAL Co., U.S., Kuraray Co., Japan and Darex
Co., U.S., with DarEval (154); Owens-Illinois/Continental PET Technologies with
CPTX-312; Honeywell International, Inc. with scavenging nylon resin (155); Cry-
ovac with OS 1000; and Chevron Phillips Chemical Co. with their OSP scavengers
(156). While the speed and capacity of oxygen-scavenging lms are considerably
lower than iron-based oxygen-scavenger sachets (151), oxygen ingress into the
package can be reduced to very low levels with appropriately designed scavenger
structures. Figure 31 presents an example of the reduction in oxygen transmis-
sion rate due to oxygen scavenging in blends of PET and nylon MXD-6 (157).
Scavengers can also be incorporated into the liner of bottle closures where they
can signicantly reduce the ingress of oxygen into the package through the closure
liner (158).
Vol. 5 BARRIER POLYMERS 257
A recent study (159) has shown that incorporation of inorganic llers as well
as reactive groups in barrier polymer lms can signicantly improve the shelf life
of packages. According to that theory, the presence of immobile reactive groups
dramatically increases the time lag (ie, the time required to achieve steady-state
permeability), but do not affect steady-state transport of penetrants across barrier
polymer lms. For example, in LDPE and PVDC lms containing 10% akes of
mica or clay and linolenic acid as the oxygen-scavenging species, the time lag for
permeation of O
2
dramatically increased by about 3 orders of magnitude to 40 h
and 3 years, respectively (159).
Conclusions
Barrier polymers are widely used in food, beverage, and other packaging indus-
tries. Some of the their advantages over traditional packaging materials such as
glass, paper, and metals are exibility, light weight, toughness, versatility, and
printability. The selection of a polymer for a particular packaging application de-
pends on its barrier as well as other physical properties. A comparison of these
properties and current as well as potential future applications for different types
of barrier polymers has been presented. Additionally, important factors govern-
ing permeability (eg, penetrant size, polymer chemical structure, temperature,
humidity) as well as ways to measure and predict permeability have also been
discussed.
Unlike glass and metals, no polymer offers an innite gas barrier. Despite
this limitation, monolayer polymer structures in many instances satisfy the bar-
rier requirements for a package. Inother situations, combinations of different poly-
mers, or polymers with inorganic materials, in the form of multilayer structures
or blends, can provide cost-effective barrier for the intended shelf life of packaged
products. As a result, plastic packaging is ubiquitous. Market pressures, however,
drive the need for continual improvements in packaging materials. Hence, there is
an ongoing interest in improving the barrier properties of polymers used in pack-
aging, and the search for improved barrier polymers and structures is ongoing.
Inorganic materials such as silicon and aluminum oxides and clays can be used
to signicantly enhance gas barrier and other mechanical properties of polymers.
One area of recent activity is polymer nanocomposites, which involves disper-
sion of nanoscale barrier particles in a polymer matrix. A considerable amount of
research is also being focussed on techniques for developing thin layers of inor-
ganic coatings on barrier polymer lms and containers. In addition, work in the
eld of oxygen-scavenging technologies (so-called active packaging systems) has
resulted in several developments. While advances in these newer barrier technolo-
gies have recently opened up several commercial applications, with plastic beer
packaging being perhaps the most publicized, it is much too early to accurately
predict their eventual success in the market place. Constructions with established
passive barrier polymers currently dominate the barrier packaging market, and
these materials are expected to be commercially important for many years to
come.
258 BARRIER POLYMERS Vol. 5
BIBLIOGRAPHY
Barrier Polymers in EPST 1st ed., Vol. 1, pp. 6595, by S. P. Nemphos, M. Salame, and
S. Steingiser, Monsanto Co.; Barrier Polymers in EPSE 2nd ed., Vol. 4 pp. 176192, by
W. A. Combellick, Goodyear Tire & Rubber Co.
1. K. Ghosal and B. D. Freeman, Polym. Adv. Technol. 5, 673 (1994).
2. R. C. Reid, J. M. Prausnitz, and B. E. Poling, The Properties of Gases and Liquids,
McGraw Hill, New York, 1988, p. 230.
3. J. Crank, The Mathematics of Diffusion, 2nd ed., Clarendon Press, Oxford, 1975,
p. 414.
4. C. E. Rogers, in J. Comyn, ed., Polymer Permeability, Elsevier Applied Science,
London, 1988, p. 11.
5. C. C. McDowell, B. D. Freeman, and G. W. McNeely, Polymer 40, 3487 (1999).
6. A. H. Windle, in Ref. 4, p. 75.
7. A. R. Berens, J. Appl. Polym. Sci. 37, 901 (1989).
8. F. Muller-Plathe, Acta Polym. 45, 259 (1994).
9. W. J. Koros and M. W. Hellums, in J. I. Kroschwitz, ed., Encyclopedia of Polymer Sci-
ence and Engineering, John Wiley & Sons, Inc., New York, 1990, Supplement Volume,
p. 724.
10. D. Hofman and co-workers, Polymer 37, 4773 (1996).
11. J. R. Fried, J. Membr. Sci. 149, 115 (1998).
12. A. Bondi, J. Phys. Chem. 68, 441 (1964).
13. D. W. vanKrevelen, Properties of Polymers: Their CorrelationwithChemical Structure;
Their Numerical Estimation and Prediction from Additive Group Contributions, 3rd
ed., Elsevier, Amsterdam, 1990, p. 875.
14. M. H. Cohen and D. Turnbull, J. Chem. Phys. 31, 1164 (1959).
15. K. Ghosal, R. T. Chern, and B. D. Freeman, J. Polym. Sci., Polym. Phys. Ed. 33, 657
(1995).
16. D. H. Weinkauf and D. R. Paul, in W. J. Koros, ed., Barrier Polymers and Structures
(ACS Symposium Series, Vol. 423), American Chemical Society, Washington, D.C.,
1990, p. 60.
17. D. H. Weinkauf and D. R. Paul, J. Polym. Sci., Polym. Phys. Ed. 30, 837
(1992).
18. W. M. Lee, Polym. Eng. Sci. 20, 65 (1980).
19. A. Singh and W. J. Koros, in Permeation Processes in Barriers and Membranes: Differ-
ences and Similarities, Polymers, Laminations &Coatings Conference, San Francisco,
Calif., 1998, p. 361.
20. A. J. Hill and co-workers, Eur. Polym. J. 32, 843 (1996).
21. B. D. Freeman and A. J. Hill, in M. R. Tant and A. J. Hill, eds., Structure and Properties
of Glassy Polymers, Vol. 710, American Chemical Society, Washington, D.C., 1997,
p. 306.
22. E. A. McGonigle and co-workers, Polymer 42, 2413 (2001).
23. V. P. Shantarovich and co-workers, Macromolecules 33, 7453 (2000).
24. J. E. Robertson, T. C. Ward, and A. J. Hill, Polymer 41, 6251 (2000).
25. G. Dlubek and co-workers, J. Phys.: Condens. Matter 10, 10443 (1998).
26. H. Yang and Y. C. Jean, Mater. Sci. Forum 255257, 40 (1997).
27. W. E. Brown, in W. E. Brown, ed., Plastics in Food Packaging, Marcel Dekker, Inc.,
New York, 1992, p. 292.
28. R. M. Barrer, Trans. Faraday Soc. 38, 322 (1942).
29. G. J. van Amerongen, J. Appl. Phys. 17, 972 (1946).
30. R. M. Barrer and G. Skirrow, J. Polym. Sci. 3, 549 (1948).
31. R. M. Barrer, Trans. Faraday Soc. 39, 48 (1943).
Vol. 5 BARRIER POLYMERS 259
32. J. E. Lefer, J. Org. Chem. 20, 1202 (1955).
33. W. J. Koros, in Ref. 16, p. 1.
34. P. Meares, J. Amer. Chem. Soc. 76, 3415 (1954).
35. T. K. Kwei and W. Arnheim, J. Chem. Phys. 37, 1900 (1962).
36. J. Brandrup, E. H. Immergut, and E. A. Grulke, Polymer Handbook, 4th ed., Wiley-
Interscience, New York, 1999, p. VI198.
37. M. Salame, in K. M. Finlayson, ed., Plastic Film Technology, Vol. 1, Technomic Pub-
lishing Co., Inc., Lancaster, Pa., 1989, p. 132.
38. R. R. Light and R. W. Seymour, Polym. Eng. Sci. 22, 857 (1982).
39. A. S. Michaels, W. R. Vieth, and J. A. Barrie, J. Appl. Phys. 34(1), 13 (1963).
40. P. M. Subramanian and I. G. Plotzker, in D. R. Paul and C. B. Bucknall, eds., Polymer
Blends, Vol. 2, John Wiley & Sons, Inc., New York, 2000, p. 359.
41. A. S. Michaels, W. R. Vieth, and J. A. Barrie, J. Appl. Phys. 34(1), 1 (1963).
42. A. C. Puleo, D. R. Paul, and P. K. Wong, Polymer 30, 1357 (1989).
43. G. E. Serad and co-workers, Polymer 42, 6929 (2001).
44. H. D. Weigmann and co-workers, Text. Res. J. 46, 574 (1976).
45. H. Jameel, J. Waldman, and L. Rebenfeld, J. Appl. Polym. Sci. 26, 1795 (1981).
46. H. Jameel, H. D. Noether, and L. Rebenfeld, J. Appl. Polym. Sci. 27, 773
(1982).
47. A. S. Michaels and R. B. J. Parker, J. Polym. Sci. 61, 53 (1959).
48. N. Qureshi and co-workers, J. Polym. Sci., Polym. Phys. Ed. 38, 1679 (2000).
49. D. J. Sekelik and co-workers, J. Polym. Sci., Polym. Phys. Ed. 37, 847 (1999).
50. H. D. Weigmann and co-workers, Text. Res. J. 47, 745 (1977).
51. P. H. Hermans and P. Platzek, Kolloid Z. 88, 68 (1939).
52. M. R. Tant and co-workers, Polym. Mater. Sci. Eng. 81, 374 (1999).
53. R. Shastri, H. C. Roehrs, C. N. Brown, and S. E. Dollinger, in Ref. 16,
p. 239.
54. D. R. Paul, Ber. Bunsenges. Phys. Chem. 83, 294 (1979).
55. R. J. Hernandez, S. E. M. Selke, and J. D. Culter, Plastics Packaging: Properties, Pro-
cessing, Applications, andRegulations, Hanser Gardner Publications, Inc., Cincinnati,
Ohio, 2000, p. 313.
56. R. J. Hernandez, J. Food Eng. 22, 495 (1994).
57. G. Strupinsky and A. L. Brody, in A Twenty-Year Retrospective on Plastics: Oxygen
Barrier Packaging Materials, Polymers, Laminations & Coatings Conference, San
Francisco, Calif., 1998, p. 119.
58. W. R. Vieth, Diffusion in and through Polymers: Principles and Applications, Oxford
University Press, New York, 1991, p. 20.
59. H. L. Frisch, J. Phys. Chem. 62, 401 (1958).
60. K. D. Ziegel, H. K. Frensdorff, and D. E. Blair, J. Polym. Sci. 7, 809 (1969).
61. Annual Book of ASTM Standards, Standard D3985, Vol. 15.09, American Society for
Testing and Materials, West Conshocken, Pa., 1994, p. 542.
62. Ox-TranOxygen Transmission Rate Measurement System, MOCON, Inc., Minneapo-
lis, Minn., 1998.
63. P. DeLassus, in J. I. Kroschwitz, ed., Encyclopedia of Polymer Science and Technology,
Vol. 3, John Wiley & Sons, Inc., New York, 1992, p. 931.
64. Annual Book of ASTM Standards, Standard F1249, Vol. 15.09, American Society for
Testing and Materials, West Conshocken, Pa., 1994, p. 1051.
65. C. D. Barr, J. R. Giacin, and R. J. Hernandez, Packag. Technol. Sci. 13, 157 (2000).
66. C. C. McDowell, D. T. Coker, and B. D. Freeman, Rev. Sci. Instr. 69, 2510
(1998).
67. P. Hernandez-Munoz, R. Gavara, and R. J. Hernandez, J. Membr. Sci. 154, 195
(1999).
260 BARRIER POLYMERS Vol. 5
68. M. Moaddeb and W. J. Koros, J. Appl. Polym. Sci. 57, 687 (1995).
69. B. I. Chaudhary and A. I. Johns, J. Cell. Plast. 34, 312 (1998).
70. O. Pfannschmidt and W. Michaeli, in Determination of the Solubility and Diffusivity
of Gases in Polymers by using a High Pressure Magnet-Suspension-Balance, An-
nual Technical Conference of the Society of Plastics (ANTEC), Atlanta, Ga., 1998,
p. 1918.
71. S. N. Dhoot and co-workers, J. Polym. Sci., Polym. Phys. Ed. 39, 1160 (2001).
72. M. Salame, Polym. Eng. Sci. 26, 1543 (1986).
73. J. Y. Park and D. R. Paul, J. Membr. Sci. 125, 23 (1997).
74. R. Franz, Packag. Technol. Sci. 6, 91 (1993).
75. S. C. Fayoux, A. Seuvre, and A. J. Voilley, Packag. Technol. Sci. 10, 145
(1997).
76. G. Strandburg, P. T. DeLassus, and B. A. Howell, in S. J. Risch and J. H. Hotchkiss,
eds., Food and Packaging Interactions II, Vol. 473, American Chemical Society,
Washington, D.C., 1991, p. 133.
77. R. M. Barrer and G. J. Skirrow, J. Appl. Polym. Sci. 3, 564 (1948).
78. S. A. Stern, J. T. Mullhaupt, and P. J. Gareis, AIChE J. 15, 64 (1969).
79. S. Li and J. S. Paik, Trans. ASAE 39, 1013 (1996).
80. J. S. Paik and M. A. Tigani, J. Agric. Food Chem. 41, 806 (1993).
81. B. D. FreemanandI. Pinnau, inB. D. FreemanandI. Pinnau, eds., Polymeric Materials
for Gas Separations: Chemistry and Materials Science, American Chemical Society,
Washington, D.C., 1999, p. 1.
82. R. C. Mason and co-workers, Tappi J. 75(6), 163 (1992).
83. C. S. Coughlin, K. A. Mauritz, and R. F. Storey, Macromolecules 24, 1526
(1991).
84. T. H. Begley, Food Addit. Contam. 14, 545 (1997).
85. O. Piringer and co-workers, J. Agric. Food Chem. 46, 1532 (1998).
86. A. OBrien, A. Goodson, and I. Cooper, Food Addit. Contam. 16, 367 (1999).
87. W. Limm and H. C. Hollield, Food Addit. Contam. 13, 949 (1996).
88. W. A. Jenkins and J. P. Harrington, in W. A. Jenkins and J. P. Harrington, eds.,
Packaging Foods with Plastics, Technomic Publishing Co., Lancaster, Pa., 1991,
p. 46.
89. Permeability and Other Film Properties of Plastics and Elastomers, Plastics Design
Library, Norwich, N.Y., 1995, p. 143.
90. BLOX
TM
Adhesion and Barrier Resins: Product Information, The DowChemical Com-
pany, May 2001.
91. A. B. Robertson and K. R. Habermann, in M. Bakker, ed., The Wiley Ency-
clopedia of Packaging Technology, John Wiley & Sons, Inc., New York, 1986,
p. 311.
92. J. Quezada-Gallo, F. Debeaufort, and A. Voilley, in S. J. Risch, ed., Food Packag-
ing: Testing Methods and Applications (ACS Symposium Series, Vol. 753), American
Chemical Society, Washington, D.C., 1999, p. 130.
93. Ref. 89, p. 706.
94. V. T. Stannett, Polym. Eng. Sci. 18, 1129 (1978).
95. R. J. Ashley, in J. Comyn, ed., Polymer Permeability, Elsevier Applied Science Pub-
lishers Ltd., Essex, England, 1985, Chapt. 7.
96. M. Salame, in Ref. 91, p. 48.
97. S. Pauly, in J. Brandrup and E. H. Immergut, eds., Polymer Handbook, 3rd ed., Wiley-
Interscience, New York, 1989, p. IV435.
98. P. T. DeLassus, in J. I. Gray, B. R. Harte, and J. Miltz, eds., Food
Product-Package Compatibility, Technomic Publishing Co., Lancaster, Pa., 1987,
p. 229.
Vol. 5 BARRIER POLYMERS 261
99. J. Landois-Garza and J. H. Hotchkiss, in J. H. Hotchkiss, ed., Food and Packag-
ing Interactions (ACS Symposium Series, Vol. 365), American Chemical Society,
Washington, D.C., 1988, p. 42.
100. M. G. R. Zobel, in Aroma Barrier Properties of Coextruded Films, Proceedings of 4th
Annual International Conference on Coextrusion Markets and Technology, Schotland
Business Research, Princeton, N.J., 1984, p. 75.
101. G. L. Robertson, in H. A. Hughes, ed., Food Packaging: Principles and Practice, Vol.
6, Marcel Dekker, Inc., New York, 1993, p. 9.
102. R. Catala and R. Gavara, Food Sci. Tech. Int. 2, 281 (1996).
103. Ref. 27, p. 132.
104. Ref. 55, p. 89.
105. Selar PA Processing Guide, E. I. du Pont de Nemours & Co., Inc., Wilmington, Del.,
2001.
106. P. M. Morse, Chem. Eng. News 8 (Nov. 10, 1997).
107. T. M. McGee and A. S. Jones, in The Effect of Processing Parameters on Physical
Properties of PET/PENBlends for Bottle Applications, 11thAnnual HighPerformance
Blow Molding Conference: Technical Innovations in Blow Molding, Cleveland, Ohio,
1995, p. 89.
108. C. C. McDowell and co-workers, J. Polym. Sci., Polym. Phys. Ed. 36, 2981
(1998).
109. G. R. Cantrell and co-workers, in C. Carfagna, ed., Liquid Crystalline Polymers, Perg-
amon Press, Oxford, 1994, p. 233.
110. N. R. Miranda and co-workers, J. Membr. Sci. 94, 67 (1994).
111. R. Lusignea, Tappi J. 80, 205 (1997).
112. R. Lusignea, in Flexible Multilayer Packaging with Oriented LCP Barrier Layer,
Polymers, Laminations & Coatings Conference, San Francisco, Calif., 1998,
p. 889.
113. D. J. Brennan and co-workers, Macromolecules 29, 3707 (1996).
114. K. Ghosal and co-workers, Macromolecules 29, 4360 (1996).
115. M. R. Pixton and D. R. Paul, in D. R. Paul and Y. P. Yampolskii, eds., Polymeric Gas
Separation Membranes, CRC Press: Boca Raton, Fla., 1994, p. 83.
116. C. H. Silvis, Trends Polym. Sci. 5(3), 75 (1997).
117. T. Glass, H. Pham, and M. Winkler, in New Thermoplastic Adhesive and Barrier
Resins, ANTEC 2000: Plastics, The Magical Solution, Orlando, Fla., 2000.
118. W. Erickson, in New Coextruded PCTFE-Based Moisture Barrier Films for High Per-
formance Packaging, Polymers, Laminations & Coatings Conference, San Francisco,
Calif., 1998, p. 355.
119. Aclar Fluoropolymer Films: Product Information, Honeywell International,
2001.
120. P. T. DeLassus, D. L. Clarke, and T. Cosse, Mod. Plast. 86 (Jan. 1983).
121. Ref. 27, p. 125.
122. E. Werner, S. Janocha, M. Hopper, and K. Mackenzie, in J. I. Kroschwitz, ed., Ency-
clopedia of Polymer Science and Technology, Vol. 12, Wiley-Interscience, New York,
1992, p. 193.
123. N. Whiteman, P. T. DeLassus, and J. Gunderson, in TAPPI Proceedings of PLC Con-
ference, 2001.
124. Ref. 27, p. 105.
125. J. L. Throne, in Ref. 91, p. 529.
126. A. R. Berens, Makromol. Chem., Macromol. Symp. 29, 95 (1989).
127. J. H. Levy, in Nylon 6 Barrier Coextrusions: A Cost-effective Packaging Route,
Polymers, Laminations & Coatings Conference, San Francisco, Calif., 1998,
p. 163.
262 BARRIER POLYMERS Vol. 5
128. Ref. 88, p. 11.
129. MXD-6 Nylon Polyamide Property Brochure, Mitsubishi Gas Chemical Co., Inc.,
2001.
130. D. A. Abramowicz and P. J. Heyes, in Comparing the Performance of Barrier-Coated
versus Multilayer PET Containers for Oxygen Sensitive Products, Nova-Pack Amer-
icas 2000, Orlando, Fla., 2000.
131. T. Hart, in New Effective Barrier Coating for Plastic Containers, Bev-Pak Americas
97, Directions 21, Inc., 1997.
132. D. S. Finch and co-workers, Packag. Technol. Sci. 9, 73 (1996).
133. N. Inagaki, in S. J. Clarson, J. J. Fitzgerald, M. J. Owen, and S. D. Smith, eds., Silicones
and Silicone-Modied Materials, ACS SymposiumSeries, Vol. 729, Oxford University
Press, Washington, D.C., 2000, p. 544.
134. Ref. 55, p. 217.
135. Ref. 101, p. 111.
136. Ref. 88, p. 195.
137. A. S. Da Silva Sobrinho, G. Czeremuszkin, M. Latreche, and M. R. Wertheimer, in W.
W. Lee, R. dAgostino, and M. R. Wertheimer, eds., Plasma Deposition and Treatment
of Polymers, Materials Research Society Symposium Proceedings, Vol. 544, Materials
Research Society, Boston, 1998, p. 245.
138. K. H. Haas and co-workers, Surf. Coat. Tech. 111, 72 (1999).
139. M. Hoffmann and co-workers, J. Sol Gel Sci. Tech. 1/2, 141 (1998).
140. H. Langowski, Kunstoffe 90(6), 88 (2000).
141. Packaging World 2 (May 2000).
142. I. Bucklow and P. Butler, Materials World 14 (Aug. 2000).
143. Packaging World 2 (June 1999).
144. R. M. Barrer, in J. Crank and G. S. Park, eds., Diffusion in Polymers, Academic Press,
New York, 1968, p. 165.
145. K. M. Kit, J. M. Schultz, and R. Gohil, Polym. Eng. Sci. 35, 680 (1995).
146. O. Motta and co-workers, Polymer 37, 2373 (1996).
147. E. Petrovicova and co-workers, J. Appl. Polym. Sci. 78, 2272 (2000).
148. Polymer Nanocomposites, Business Communications Co., Inc., Apr. 2000.
149. Packaging World (Nov. 6, 2000).
150. K. Kamena, in Emerging Nanocomposite Technologies for Barrier and Thermal
Improvements in PET Containers, Nova-Pack Americas 99, Orlando, Fla., 1999,
p. 1.
151. L. Vermeiren and co-workers, Trends Food Sci. Tech. 10, 77 (1999).
152. W. Scholl, in Active Barrier Packaging: A Technology for the 21st Cen-
tury, Polymers, Laminations & Coatings Conference, Atlanta, Ga., 1999,
p. 671.
153. M. L. Rooney, in M. L. Rooney, ed., Active Food Packaging, 1st ed., Blackie Academic
& Professional, Glasgow, U.K., 1995, p. 74.
154. S. Lambert, in New EVOH-Based Oxygen Scavenging Resins for Use in
Co-Injected PET Bottles, Nova-Pack Europe 2000, Neuss, Germany, 2000,
p. 163.
155. E. P. Socci, M. K. Akkapeddi, and D. C. Worley, in New High Bar-
rier, Oxygen Scavenging Polyamides for Packaging Applications, ANTEC,
2001.
156. B. D. Rodgers, in New High Capacity Oxygen Scavenging Polymer For
Use in Flexible Packaging, Flex-Pak Europe 2000, Amsterdam, Nov. 2000,
p. 61.
157. U. S. Pat. 5,639,815 (1997), M. A. Cochran and co-workers.
Vol. 5 BIODEGRADABLE POLYMERS, MEDICAL APPLICATIONS 263
158. F. N. Teumac, B. A. Ross, and M. R. Rassouli, MBAA Tech. Q. 27, 122
(1990).
159. C. Yang, E. Nuxoll, and E. Cussler, AIChE J. 47, 295 (2001).
SUSHIL N. DHOOT
University of Texas at Austin
BENNY D. FREEMAN
University of Texas at Austin
MARK E. STEWART
Eastman Chemical Company

Anda mungkin juga menyukai