You are on page 1of 15

Journal of Composite

Materials http://jcm.sagepub.com/

Langmuir-Type Model for Anomalous Moisture Diffusion In Composite Resins


Harris G. Carter and Kenneth G. Kibler
Journal of Composite Materials 1978 12: 118
DOI: 10.1177/002199837801200201

The online version of this article can be found at:


http://jcm.sagepub.com/content/12/2/118

Published by:

http://www.sagepublications.com

On behalf of:

American Society for Composites

Additional services and information for Journal of Composite Materials can be found at:

Email Alerts: http://jcm.sagepub.com/cgi/alerts

Subscriptions: http://jcm.sagepub.com/subscriptions

Reprints: http://www.sagepub.com/journalsReprints.nav

Permissions: http://www.sagepub.com/journalsPermissions.nav

Citations: http://jcm.sagepub.com/content/12/2/118.refs.html

>> Version of Record - Jan 1, 1978

What is This?

Downloaded from jcm.sagepub.com at Monash University on October 24, 2014


Langmuir-Type Model for
Anomalous Moisture Diffusion
In Composite Resins

HARRIS G. CARTER AND KENNETH G. KIBLER

Materials Research Laboratory


General Dynamics
Fort Worth Division
Fort Worth, Texas 76101

(Received January 16, 1978)

ABSTRACT
Considerable evidence has recently been presented for anomalous mois-
ture diffusion in epoxy matrix composites. In an earlier paper we noted
that there is no a priori reason why moisture should obey simple diffusion
theory rather than Kirkwood’s (linear) generalization of the Boltzmann
transport equation. With the intent of keeping the analytical aspects of the
moisture diffusion problem as simple as possible, we here present a slightly
generalized but linear model which involves sources and sinks of diffusing
water molecules. With respect to diffusive characteristics, our model is
related to the simplest form of neutron transport theory. With respect to
bound and unbound particles it is similar to the Langmuir theory of ad-
sorption isotherms, although we here assume bulk absorption in the resin
with no implication that surfaces are involved. An approximation to our
exact solution of the coupled linear differential equations is used to fit
mildly anomalous moisture uptake curves for 5208 resin exposed to several
rclative humidities for two years. The fact that the same parameters give
equally good fits to the data at all humidities suggests that the absorption
anomaly does not result from nonlinear (e.g. concentration-dependent)
effects.

SCOPE

J~EcAusE OF Knows humidity-induced effects on the physicals properties and


stress characteristics of epoxy-composite structures, the composite technology
community has recently shown some concern over long term moisture absorption
results which are anomalous in terms of simple (Fickian) diffusion theory. If the
anomalous moisture uptake is due to nonlinear elements in the governing equation
for water transport in composite resin, as evidently is assumed by most workers

118

Downloaded from jcm.sagepub.com at Monash University on October 24, 2014


presently considering the subject, then the prospects for making reliable predictions
of moisture distributions in structures exposed for years to variable service environ-
ments are somewhat clouded. ,

In the hope of keeping the analytical aspects of the moisture diffusion problem
as simple as possible, we here present a slightly generalized but linear model which
involves sources and sinks of diffusing water molecules. In regard to diffusive
characteristics, our model is related to the simplest form of neutron transport
theory. In regard to bound and unbound particles it is similar to the Langmuir
theory of adsorption isotherms, although we here assume bulk absorption in the
resin and do not imply that any surfaces are involved.
For the case of an initially dry one dimensional specimen exposed to a constant
moisture environment, we give exact solutions of the resulting pair of coupled
equations, which describe both the spatial distribution of moisture and total mois-
ture uptake as a function of time. We find that, under conditions which appear to
be satisfied by reasonably thin specimens, or by thicker specimens at higher temper-
atures, an approximation for total moisture content applies which is scarcely more
difficult to evaluate than that of simple diffusion theory. Agreement between this
approximation and available data on long term, mildly anomalous, moisture uptake
in epoxy at room temperature is encouraging, especially in the respect that the
same diffusion parameters apply equally well to all humidities considered.

CONCLUSIONS AND SIGNIFICANCE


At least part of the recently available data on anomalous moisture absorption in
epoxy materials can be explained quantitatively by assuming that absorbed mois-
ture consists of mobile and bound phases. Molecules of the mobile phase diffuse
with a concentration-and stress-independent diffusion coefficient D’Y and are
absorbed (become bound) with a probability per unit time 7 at certain sites whose
nature is unspecified. Molecules are emitted from the bound phase, thereby becom-
ing mobile, with a probability per unit time (3. The diffusion process therefore is
described by the same modification of the simple diffusion equation which has
long been used to describe neutron flux in absorbing and multiplying media. The
local weight fraction of moisture approaches an equilibrium value, moo, when the
number of mobile molecules per unit volume, n, and the number of bound mole-
cules per unit volume, N, approach values such that &dquo;’Ill (W, as in Langmuir =

adsorption theory.
When both y and j3 are small compared to the parameter which determines the
rate of saturation of a one dimensional specimen of thickness 2 6 in simple dif-
fusion theory, namely,

119

Downloaded from jcm.sagepub.com at Monash University on October 24, 2014


then the total moisture uptake in an initially dry one-dimensional specimen, after
exposure time t, is given approximately by .

where

is thesame sum which appears in ordinary simple diffusion theory. This approxima-
tion gives good fits to mildly anomalous moisture uptake curves for 5208 resin
exposed for two years at room temperature with constant relative humidities rang.
ing from 45% to 100%.
The fact that the fits for different relative humidities are based on the same
values ofy, {3, and D’Y’ and that the humidity dependence taken for m_ is similar to
those assumed in previous simple diffusion analyses, strongly suggests that the
absorption anomaly does not result primarily from non-linear moisture-concentra-
tion effects. If these results are corroborated by analyses of data on specimens of
different thicknesses, or exposed to different temperatures, it would be reasonable
to adapt the Langmuir-type linear model proposed here to the long term prediction
of moisture distributions in service environments.

BACKGROUND ON ANOMALOUS MOISTURE-ABSORPTION EFFECTS


Much evidence for what was originally designed by Alfrey, et. al. [1 ] as anoma-
lous moisture diffusion was presented in connection with epoxy composite materi-
als at the September 1977 ASTM D-30 Committee Symposium in Dayton, Ohio

[2,3,4]. Aside fiom aging effects associated with thermal spikes, which demon-
strably produce irreversible changes in the diffusion characteristics of epoxy materi-
als [5, 6, 7] , two types of anomalies can be distinguished. The first of these is
manifested in the curve which describes total moisture gain in completely dry
specimens upon exposure to a constant relative humidity and temperature. Under
prolonged exposures, the moisture gain, plotted against square root of exposure
time, tends to rise above the &dquo;equilibrium&dquo; plateau predicted by fitting van
Amerongen’s solution of the simple diffusion equation [8] to the data at shorter
exposures. In some cases there also appears to be evidence for an ultimate secon-
dary &dquo;equilibrium&dquo; level. --

120

Downloaded from jcm.sagepub.com at Monash University on October 24, 2014


The second type of anomaly relates to a history-dependence of saturation level,
or equilibrium moisture content (mOo). For a given final humidity and temperature
it is found that the apparent equilibrium moisture content depends on the speci-
men’s previous history of exposure, even when no temperature changes are involved
and there is no reason to believe that irreversible changes have occurred in the resin.
Specifically, at a given final humidity, a polymeric material appears to have a
greater capacity for holding water if it was saturated initially at a higher humidity
than if it was initially dry or saturated at a lower humidity. In regard to composite
materials this effect seems to have been noted quite recently for the first time by
Shirrell [2] , although its existence has been known for some time in the case of
..

organic fibers [9] .


Because of its relevance to moisture-induced changes in material properties and
stressdistributions in structures subject to prolonged moisture exposure, the first of
the above cited anomalies has recently attracted considerable interest. Although
energetic efforts are being made by some to account for this anomaly in terms of
moisture-concentration dependent, or swelling-stress dependent diffusion coeffi-
cients (both of which lead to a non-linear Fickian diffusion equation), it would
appear that other possibilities should be explored. In a recent paper describing a
quick test of the spatial moisture distribution predicted by the simple diffusion
theory [10], we noted that there is no a priori reason why moisture should obey
simple (Fickian) diffusion theory rather than Kirkwood’s (linear) generalization of
the Boltzmann transport equation [11] . In the latter case, for example, the dura-
tion of interactions between water molecules and molecular groups of the resin
could be long compared to times in which the distribution of water molecules
changes appreciably.
Our data [10] on the spatial distribution of moisture in epoxy and polyimide
composites exposed for only two or three days were quite consistent with simple
diffusion theory. However, the onset of anomalies in long term gross moisture
uptake observed by another group at our facility at about the same time led us to
postulate [12] a more general relationship of the previously mentioned Kirkwood
type. This description amounts to simple diffusion with sources and sinks of dif-
fusing water molecules; that is, we assume that absorbed water is divided into
mobile and strongly-bound types. With regard to transitions between bound and
unbound molecular states, our model is quite analogous to the Langmuir theory of
adsorption on surfaces. We have obtained an exact solution of the corresponding
equations for the usual experimental case of an initially dry specimen exposed to a
uniform moisture environment on both sides.
Our decision to offer our results for publication at this time is prompted in part
by the fact that a verbal description of a similar if not identical model was given by
Gurtin at the October 1977 Mechanics of Composites Review in Dayton, Ohio
[13]. We strongly wish to advance a description of moisture diffusion which does
not involve non-linear differential equations. We also feel that data from exposures

121

Downloaded from jcm.sagepub.com at Monash University on October 24, 2014


started two years ago now provide the experimental authority for a theory that
conflicts with some admittedly well-reasoned arguments for nonlinear Fickian dif-
fusion. Finally, our treatment appears to offer a natural explanation for the re-
cently re-discovered history dependence of apparent equilibrium moisture content,
although a discussion of this effect must be deferred.
MODEL FORMULATION

We assume that at a given time and place in a resinous material there are n
mobile H2 0 molecules per unit volume which diffuse with a diffusion coefficient
D’Y and become bound at a rate per unit volume 7r:. At the same time and place
there are N bound molecules per unit volume which become mobile at a rate per
unit volume 0 N. As in the Langmuir theory of adsorption ~14] , the number-densi-
ties of mobile and bound molecules at equilibrium, ry and N_ respectively, depend
upon the relative humidity H and satisfy the relation

The long-term solubility of water, expressed as percent by weight of the dry resin is

where M,, is the molecular weight of water,NA is Avogardro’s number and p, is the
density of the dry resin. Conversely, from Eqs. (1) and (2) the equilibrium number
densities nee (H) and N, (ll) can be expressed in terms of empirically-determined
solubilities as a function of humidity.
We assume that the diffusion of mobile molecules conforms to simple diffusion
theory augmented by sources and sinks, as in the theory of neutron chain reactions
[15] Thus, for the one dimensional case, the number densities of position z and
time t satisfy the coupled pair of equations

These relations represent a case of Kirkwood’s generalization of the


special
Boltzmann transport equation [11]. In the usual experimental case of an initially
dry slab of thickness 28 exposed to a constant moisture environment on both sides
at t =
0, the appropriate boundary and initial conditions are I

122

Downloaded from jcm.sagepub.com at Monash University on October 24, 2014


and (4)

Method of Solution
We have obtained an exact solution of the above problem by means of Laplace
transforms. Equations (3) yield for n {x,p), the Laplace transform of n (x,t) with
respect to time, the ordinary differential equation

where

From Equations (4) we obtain the corresponding boundary conditions

From Equations (5) and (7) we obtain for the Laplace transform of the mobile-
molecule number density.

The corresponding solution for bound molecules is

The inversion of Equation (9) is formally expressed as

where C is an arbitrary positive number.

123

Downloaded from jcm.sagepub.com at Monash University on October 24, 2014


Once the inversion for bound molecules has been accomplished, a corresponding
expression for mobile molecules is easily obtained from the second of Equations
(3). Evaluation of Equation (10) is achieved by a contour integration in the plane
defined by the complex transform variable p, which, in terms of real variables r and
q, can be represented as

From Equations (6) and (9) it can be shown that 1V (z, p) is analytic and single
valued everywhere in the complex plane except at the points [r --(7~-~3), q =0], =

[r -13, q =0] , r = 0, q =0] , and at two infinite sets of points [r =p,, q =0]
=

and [r = Pp q = 0] where .

and

where I and j are positive odd integers, and

N (z, t) is the sum of the residues of Ñ {z,p)ept at the singularities defined above.
Exact Expressions
Evaluation of the residues at the points designated above yields the following
expression for the spatial distribution of bound molecules:

where

124

Downloaded from jcm.sagepub.com at Monash University on October 24, 2014


with&dquo; given by Equation (14). Then from the second of Equations (3), the spatial
distribution of mobile molecules is found to be

The validity of these solutions can be demonstrated rather easily by direct substitu-
tion into Equations (3) and (4).
Integration of the above results over slab thickness gives the following expression
for m,, the percentage by weight of moisture uptake after a time t:

where the long term solubility m- is related to the equilibrium particle densities by
Equation (2)..
,

Useful Approximations
A convenient approximation to Equation (18) applies when 2 y and 2 P are both
small compared to K. It can be valid at all times t provided the error is referred to
the exact result for mt rather than the difference between m= and tnt,This approx-
imation for the total moisture uptake in an initially dry slab is

125

Downloaded from jcm.sagepub.com at Monash University on October 24, 2014


The terms in brackets are the same as those which determine mt in simple dif-
fusion theory, to which Equation (19) reduces for y = 0. When the exposure time
is small enough that gt is less than about 0.7, Equation (19) yields the approxima-
tion

If, on the other hand, the exposure time is long enough that rct is large compared to
unity, the following approximation holds

Equation (14) shows that the accuracy of the above approximation increases
with increasing temperature and decreasing specimen thickness. For sufficiently
thin specimens and/or sufficiently high temperatures Equation (19) seems to be
consistent with a sort of quasi-equilibrium moisture content which appears at times
on the order of, say, t = 5/x and whose value is

Under exposure the moisture level would show a gradual rise towards
prolonged
.m_. Our data exposure at room temperature do not show double plateaus of
on
this type, but there is some indication of them in Shirrell’s data at higher tempera-
ture [2J .
APPLICATION OF THE MODEL

. We have applied the approximate expression given by Equation (19) to data on


moisture uptake in initially dry 5208 resin which have been taken for the last two
years at this facility by J. D. Reynolds and J. E. Halkias. Thus far, we have
considered only room temperature (21°C) exposures at humidities of 45.5%,
75.5%, 97.5% and 104% (water immersion). We represent the observed percentage

126

Downloaded from jcm.sagepub.com at Monash University on October 24, 2014


weight gain, mt, by the average of the values for three specimens at each humidity.
In evaluations involving D and K we have used the average thicknesses for the
specimens exposed at each humidity, namely 25 0.084 cm (0.033 in), 0.086 cm =

(0.034 in), 0.097 cm (0.038 in) and 0.086 cm (0.034 in) respectively. The lateral
dimensions of the specimens are approximately 2.54 cm (1.0 in) X 1.27 (0.5 in).
Although larger lateral sizes would have been desirable for a test of the one-dimen-
sional version of our model, it can be shown that three-dimensional effects do no
more than cause a slight change in the values of D’Y inferred from the data.
At short exposure times the observed values of mt, as usual, rise linearly on a
plot against (t)Y1. In one case, that of 97.5% humidity, no good linear fit could be
made to pass exactly through the origin (t)~’1 0, mi 0. On the basis of past
= =

experience, we assumed that a little moisture was distributed almost uniformly


through the specimen at t = 0, and, on this basis, made the initial uptake curve pass
through the origin by adding 0.30% to all values of mt for this humidity. The initial
slopes for the four humidities cited above were then assigned the following values
respectively: d mtld(t)1h 0.5543, 1.0737, 1.3759 and 1.6170% day - *.
=

At large values of t all four sets of experimental points showed behavior which
cannot be explained by simple diffusion theory. Although the anomalies are mild in
the respect that no positive second derivatives appear in mt, a distinct upward drift
can be distinguished after 100 days, at which time a good simple diffusion theory
fit to points at shorter exposures would predict essentially no further weight in-
crease.
Our assumed values of (3, ’Y and ZL for room temperature are based on the data
for 100% relative humidity alone. We drew a smooth curve through the experi-
mental points, then calculated j3 and y from two relations derivable from Equation
(21), namely

We then obtained m_ for 100% humidity using Equation (21) and, finally, ob-
tained D’Y from the short-term exposure data at 100% RH using Equations (20) and
(14). Having prescribed Dy, y and 0 as well as the above-cited initial slopes, we are
not at liberty to choose the longer-term solubilities m_ for the remaining three
humidities; these are dictated by Equation (20) in the form

127

Downloaded from jcm.sagepub.com at Monash University on October 24, 2014


The resulting values of the long-term solubilities moo (H) depart somewhat from a
Iineai function of the relative humidity H, i.e., long-term solubility does not pre-
cisely obey Henry’s law. The amount of curvature is similar to that cited by
McKague in describing apparent saturation of 5208 resin after shorter exposures
[16] .
Our parameters for 5208 resin at room temperature are summarized as follows:

y =
0.0003488 day-I, 13 = 0.002058 day-’
Dy =
9.090 X 10-1 cm2 /day (1.409 X 10-5 in2 /day) .

moo(H 45.5%) = =
2.525%
m_ (H = 75.5%) =
5.040%
m_ (H 97.5%)
= =
7.218%
m_ (H = 100%) =
7.590%
Since the values of K given by Equation (14) and the average thicknesses 26 are
approximately 0.1 day&dquo;’-’, the conditions for the validity of Equation (19) are well
satisfied.
Moisture uptake predictions, obtained by substitutig the above parameters into
Equations (14) and (19), are compared with the experimental data in Figure 1. The
four calculated curves would be the same when divided by moo (H) except for the
fact that the average thicknesses 25 were slightly different. It is seen that the
agreement between the approximate version of our model and the experimental
data is satisfactory for all the humidities considered. Further tests of the approxi-
mation can be based on the data of Shirrell [2]. His room temperature results for
T300/5208 composite specimens show the same type of gradual rise after pro-
longed exposures that is visible in Figure 1. At higher temperatures the data for his
thinnest specimens show a positive second derivative, which is also consistent with
the approximation of Equation (19). We intend to test our exact solution, Equation
(18), on the data of Whitney and Browning [3] who have subjected thick speci-
mens of 3501-5 resin to prolonged exposures at room temperature and higher.
We will not speculate here on the nature of the absorption sites at which we
assume that water molecules become bound. We can only note that the absorption

probability y is remarkably small at room temperature and that data at higher


temperature suggest a rather weak temperature dependence of both y and 13. This
latter behavior would be consistent with sites whose inaccessibility involved a large
&dquo;steric hindrance&dquo; rather than an activation energy.
Because of the small values off, the spatial distribution of mobile molecules can
become essentially uniform across the dimensions of a thin specimen long before
the specimen is saturated in regard to bound molecules. In such cases the approach
to ultimate saturation proceeds at the same rate everywhere in the specimen. A
further peculiarity of the Langmuir model implications relates to rate of approach
to saturation as inferred from measurements of weight gain. According to simple
diffusion theory, an initially dry 5208 resin specimen of thickness 0.089 cm

128

Downloaded from jcm.sagepub.com at Monash University on October 24, 2014


Figure 1. Moisture uptake in 5208 resin at 24°C (75°FJ. Solid curves from
Equation (19), using same values of parameters for each humidity.

(0.035&dquo;,) for example, would reach 99.0% of its final saturation level about 39.0
days after its exposure to moisture. On the other hand, according to Equation (21)
with the values of 7 and (3 cited above, the same degree of saturation would require
3.6 years.

NOMENCLATURE

D 7 (cm2fday) =
diffusion coefficient for mobile H20 molecules
H(%) =
relative humidity at a given temperature
mt(%) If = total weight-percent moisture uptake in slab after time
t

ml (%) =
total moisture uptake in slab at apparent saturation
(short term)
m (%) =
true moisture uptake at saturation (long term)
MW (gmfmole) =
molecular weight of water
n(moleculesfcm3) =
number density of mobile Hz 0 molecules .

N(molecules/cm3 ) =
number density of bound H70 molecules
noo,N&dquo;&dquo;, (molecules/cm3 ) =
number density of mobile, bound H20 molecules at
long-term saturation

129

Downloaded from jcm.sagepub.com at Monash University on October 24, 2014


NA (molecules/mole) =
Avogadro’s number
p(days -1 ) =
Laplace transform variable
t(days) =
exposure time
z(cm) =
position in slab relative to center-plane ,

5(cm) = half-thickness of specimen .

7(days-1 ) =
probability per unit time that a mobile H20 molecule
becomes bound
0(days-’) =
probability per unit time that a bound H2 O molecule
becomes mobile
x(days’1) =
arZD7/(28)z =
characteristic diffusion constant
Pr(gmfcm3) =
dry resin density

REFERENCES
1. T. Alfrey, E. F. Gurnee and W. G. Lloyd, "Diffusion in Glassy Polymers," J. Polymer
Science: Part C, (1966), p. 249.
2. C. D. Shirrell, "Diffusion of Water Vapor in Graphite/Epoxy Composites" Paper presented
at the ASTM Symposium on Environmental Effects on Advanced Composite Materials held
in Dayton, Ohio on September 29-30, 1977.
3. T. M. Whitney and C.E. Browning, "Some Anomalies Associated with Moisture Diffusion in
Epoxy Matrix Composite Materials," Paper presented at the ASTM Symposium on Environ-
mental Effects on Advanced Composite Materials held in Dayton, Ohio on September
29-30, 1977.
4. H. T. Hann and R. Y. Kim, "Swelling of Composite Laminates," Paper presented at the
ASTM Symposium on Environmental Effects on Advanced Composite Materials held in
Dayton, Ohio on September 29-30, 1977.
5. R. Delasi and J. B. Whiteside, "Effect of Moisture on Epoxy Resins and Composites," Paper
presented at the ASTM Symposium on Environmental Effects on Advanced Composite
Materials held in Dayton, Ohio on Sepgember 29-30, 1977.
6. C. E. Browning, "The Mechanisms of Elevated Temperature Property Losses in High Per-
formance Structural Epoxy Resin Matrix Materials After Exposures to High Humidity
Environments," AFML-TR-76-153, Air Force Materials Laboratory Wright-Patterson Air
Force Base, Ohio, March 1977.
7. E. L. McKague, J. E. Halkias and J. D. Reynolds, "Moisture in Composites: The Effect of
Supersonic Service on Diffusion," J. Composite Materials, Vol. 9 (1975), p. 2.
8. G. J. van Amerongen, "Diffusion in Elastomers," Rubber Chemistry and Technology, Vol.
37 (1964), p. 1065.
9. R. Meredith in Rheology, F. R. Eirich, Ed., Vol. 2, Chapter 7, Academic Press Inc., New
York, 1958.
10. H. G. Carter and K. G. Kibler, "Rapid Moisture-Characterization of Composites and Possi-
ble Screening Applications," J. Composite Materials, Vol. 10 (1976), p. 355.
11. J. O. Hirschfelder, C. F. Curtiss, R. B. Bird, and E. L. Spotz, "The Kinetic Theory of
Gases," Thermodynamics and Physics of Matter, F. D. Rossini, Ed., Princeton University
Press (1955), p. 192.
12. H. G. Carter, "Moisture Absorption in Epoxy-Type Structures," Paper presented at the
AIAA North Texas Mini-Symposium at the University of Texas at Arlington, February 21,
1976.

130

Downloaded from jcm.sagepub.com at Monash University on October 24, 2014


13. M. E. Gurtin, "Continuum Theory of Fracture," Mechanics of Composites Review, 25-27
October 1977, Dayton, Ohio, p. 85.
14. S. Glasstone, Textbook of Physical Chemistry, 2nd Ed., Chapters XIII and XIV, D. Van
Nostrand Co., Inc., New York, 1946.
15. A. M. Weinberg and E. P. Wigner, The Physical Theory of Neutron Chain Reactors, Chapter
VIII, The University of Chicago Press, 1958.
16. E. L. McKague, J. D. Reynolds and T. E. Halkias, "Moisture Absorption, Swelling and Glass
Transition Relationships for Epoxy Resin," J. Applied Polymer Science, in press, 1977.

131

Downloaded from jcm.sagepub.com at Monash University on October 24, 2014

You might also like