Sunteți pe pagina 1din 59

Materials Science and Engineering, R13 (1994) 265-324 Reports: A Review Journal

Environmental durability of glass-fiber composites


Carol L. Schutte
Polymer Composites Group, Po~mers Division, National Institute of Standards and Technology, Gaithersburg, MD 20899, USA

Received 10 March 1994; accepted 24 March 1994

Abstract

Durability of glass-fiber/polymer composites is dictated by the durability of the components:


glass fiber, matrix, and the interface. Environmental attack by moisture, for example, can degrade
the strength of the glass fiber; plasticize, swell, or microcrack the resin; and degrade the fiber/
matrix interface by either chemical or mechanical attack. The relative rates of these degradation
processes are a function of the chemistry of the resin, temperature, length of time of exposure,
degree of stress (whether cyclic or static), chemistry and morphology of coating of coupling agent
on the glass fiber, and type of glass fiber. Several examples illustrate how the chemistry and
morphology of the coatings of coupling agents that are on the glass fiber influence the strength
and durability of the interfacial region.

1. Introduction

The use of glass fiber/polymer composites offers many potential advantages with respect
to longer lifetimes, reduction of weight, and cost in industries such as transportation and
infrastructure. In order for polymer composites to be useful in their fullest capability both
in terms of structural and customer confidences, we need to integrate these materials further
into our economic infrastructure by having the capability to predict their performance
accurately and reliably throughout their lifetime. Understanding the mechanisms responsible
for the loss of mechanical properties of glass fiber/polymer composites is complex, however,
with each constituent having the capability of being the weak component in the structure.
Furthermore, synergy between these mechanisms of degradation has the potential to reduce
properties of the composite at a rate that is faster than is suggested by those rates of
degradation for the individual components.
Recent reports and workshops have addressed the issues of methodologies used for
prediction of lifetime behavior in load-bearing structures of continuous-fiber polymer com-
posites [1-4] and of polymers [5]. According to these reports, the state of the art in predictive
methodologies for life-time behavior of polymer composites lacks assimilation of the large
base of information available in the community that develops the use of these materials.
Furthermore, some phenomenological approaches accompanied by experimental work exist
for this purpose; however, it is most desirable that this methodology be based on a more
fundamental understanding of processes that are responsible for the mechanisms of failure
[1].
Having in mind further understanding of fundamental mechanisms responsible for
composite failure, the goal of this review is to evaluate our current level of understanding
of environmental durability of glass fiber composites and identify areas where further studies
of these issues would aid our capabilities of life-time predictions. The focus of this work,

0927-796X/94/26.00 © 1994 Elsevier Science S.A. All rights reserved


SSD! 0927-796X(94)00168-R
266 C.L. Schutte / Environmental durability of glass-fiber composites

therefore, addresses issues of durability for each component of the composite - fiber,
matrix and interface -- in order to identify their "Achilles' heel(s)". Also, because of the
significance of the interfacial region in governing the properties of composites, this work
emphasizes studies that focus on understanding the development of morphology and
characterizing mechanical performance at the interface. The loss of properties of the
components of the composite involves stress corrosion and time-dependent failure of the
glass fibers, diffusion of water into the polymer, subsequent swelling and hydrolysis of the
matrix, chemistry and morphology of the coupling agent on the surface of glass, and the
effect of that chemistry and morphology on durability, and mechanochemical degradation
at the interface.
This review focuses on studies of composites that contain glass fibers: however, where
studies using carbon fibers were relevant, as in single-fiber tests, this review includes those
results. Theoretically, many environmental elements can degrade polymer composites; how-
ever, the majority of studies addressed the problems of exposure to moisture. The report
that follows is an attempt to combine these individual studies in order to further broaden
the "picture" that we have of glass fiber/composite durability.

2. Effect of the environment on properties of glass fibers

Because glass fibers are the reinforcement for many commercial composites, the durability
of the fibers under environmental exposure will dictate the retention of many of their fiber-
dominated mechanical properties. This section covers a model of crack growth for bulk
glasses, time-dependent fracture of bulk glass and fibers, and stress corrosion of fibers with
and without resin. The mechanisms governing crack growth and time-dependent fracture
in bulk glasses are necessary in order to understand the fundamental behavior of glass
fibers in the polymer composite. Studies of the growth of cracks in bulk glasses identified
factors that control the rate in stress corrosion of glass. Although knowledge of these
controlling mechanisms is useful in explaining composite behavior in general terms, the use
of models based on these mechanisms alone is insufficient for accurate extrapolation of
the behavior in macroscopic composites. Specific information on the diffusion of water
through the resin to the fiber and the degree of stress of the fiber, for example, is necessary
for accurate prediction of stress corrosion of the fiber within the composite.

2.1. Crack-growth model in bulk glasses

The literature reports data from studies of crack growth in bulk glasses and ceramics
as an empirical relationship between the velocity of the growth of the crack, V, and the
stress intensity factor K~. Although other forms exist [6], the following empirical equation
is in common use, where the constant A and exponent n depend on the material and the
environment.
V=AKI n (1)
Under conditions of slow crack growth, where V and Kt are measurable experimentally,
three regions exist with different rate-limiting mechanisms [6,7]. Fig. 1 illustrates these
different regions. Changes in the mechanism that governs the growth of cracks result in
C.L. Schutte / Environmental durability of glass-fiber composites 267

K ic

J
//
-6
>

S t r e s s Intensity, KI

Fig. 1. Schematic illustration of Regions I, II and I I I (chemical-reaction rate-limited, diffusion-of-reactant rate-limited, and
dependent on stress at the crack tip, respectively) of a typical crack growth curve for glasses and ceramics [6] (by permission
of the publishers, American Geophysical Union).

distinct characteristics relating the crack velocity and the stress intensity factor. For crack
growth at very low velocities, Region I, the process is chemical reaction rate-limited; therefore,
traditional reaction rate theory applies. The slope [7] n of logVvs, log(K~) can range between
11 and 100. In this region, studies of the effects of stress corrosion of water in a variety
of environments show that molecules with an orbital containing a lone electron pair near
a site capable of donating a proton (such as water, ammonia, hydrazine and formamide)
significantly increase the rate of crack growth [8]. The rate of crack growth was faster for
glass immersed in water than in deuterated water. This isotope effect is also consistent
with the growth of the crack being chemical reaction rate-limiting under these conditions.
It is likely that water is the active species in mixed solutions of formamide and water [9].
Fig. 2 illustrates the increase in the rate of crack growth with increased content, activity,
of water in solutions containing mixtures of formamide and water. For these different
solutions, the slopes for the velocity-K! portion of the curve are similar both to each other
and to that for pure water. This result is consistent with a model whereby the content of
water determines the rate-controlling step in this regime, even in the trace amounts that
are present in the 100% formamide solution.
In Region II the rate of reaction of the glass with the species in solution is faster
than the rate of diffusion through solution for that species to the surface of the glass; thus
diffusion of the reactant(s) to the crack tip is the rate-controlling process [7]. The velocity
of crack growth has a slight dependence on KI and on the size of the crack, with the value
of n ranging between 2 and 6.
Region III has a steep dependence on KI, and the slope [7] is typically greater than
100. Rather than having the rate of crack growth depend on either the concentration of
water or the environments, Region III appears to be dependent on factors such as the
chain length of the alcohol in which the water is mixed [10]. In order to explain the
mechanism that controls the rate of growth of cracks in Region III, Wiederhorn et al.
268 C.L. Schutte / Environmental durability of glass-fiber composites

--2 I I I I I •
• formomide
0 9 5 ~ f o r m o r n i d e / 5 % woter
--3 • 90% formomlde/1096 woter • ~ 0
• 50% f o r m o m i d e / 5 0 ~ , woter

--4

~-~ -5 •T ¢-~
• • 0 ~0

E o .':'
v
>
--6
"o" Iff"-~"•
0
" --7 i• "o
• 0 • •
O
-8 0 • •

-9
IIII

-10 ~' , I I I I
0.2 0.3 0.4 0.5 0.6 0.7 0.8
K I (MPa.m 1/2)
Fig. 2. Illustration of the influence of the content of water in a mixed solvent on the rate o f crack growth for soda-lime
glass in formamide/water mixtures: 100/0, 95/5, 90/10, and 50/50 [9] (reprinted by permission of the American Ceramics
Society).

proposed a model [11] that relates the formation of charge on the strained and breaking
Si-O chemical bond to the magnitude of the volume of the activated species.
More recent studies under conditions of chemical reaction rate-limited crack growth
presented two possible mechanisms for attack by water on glass. It appears that the active
species is the molecule of water u n d e r acidic and neutral conditions, and under basic
conditions, the hydroxide ion. These proposed mechanisms were based on results showing
that the n value, slope, was similar for acidic and neutral conditions, but was lower in
magnitude for crack growth under basic conditions [12]. Furthermore, the rates of crack
growth under basic conditions m a d e from different alkali-metal hydroxides, NaOH, CsOH,
LiOH and KOH, were the same except for solutions m a d e with LiOH in concentrations
greater than 1 mol 1-1, where the rate was slower but had a greater slope than data
collected in 1 mol 1-a K O H solution. The logV vs. log(Ki) curve for the solution of 1 mol
1-1 LiOH was superimposable with that for neutral water. Because of its affinity with
hydroxide, Li ÷ could influence the activity of the hydroxide ion in solution. These kinetic
data suggested that the relatively high concentration of Li ÷ in solution had the potential
to associate with a large percentage of the hydroxide ions, compared to sodium ions. Because
of the extent of this association, the presence of the lithium ions possibly influenced a
change in the reacting species from the hydroxide ion to water, as in the solutions with a
more acidic pH.

2.Z Time-dependent fracture of glass


Models of time of failure of glass and ceramics based on principles of fracture mechanics
[13-16] are useful in predicting their lifetime [17-20]. For example, Evans and Wiederhorn
[13] proposed an analytical m o d e l for the time-dependent failure of ceramics involving the
C.L. Schutte / Environmental durability of glass-fiber composites 269

rate of propagation of cracks under a constant load. This approach assumes the presence
of pre-existing flaws in the material so that the initiation of flaws is not the rate-determining
process. They derived a relationship that determines the time required for a subcritical
flaw to grow to a crack with critical size. Using both the definition for the rate of crack
growth, da/dt=v, where a is the crack length, t is time, and v is velocity, and the relationship
K~= traYa1/2 for the stress intensity factor K~, where tra is the applied load, and Y is the
geometric factor, as well as assuming that failure is instantaneous once the final stress
intensity factor reaches the level of the critical intensity factor; and assuming that the
critical stress intensity factor raised to 2 - n is much less than the initial stress intensity
factor raised to the same power, Kic 2-" <<K~i2-n, since n is between 9 and 50 for ceramics;
then the following approximation determines the time r to failure:

r = 2 [(n - 2)A~r~Zy2K~c"-z] (2)


This approach requires "proof testing" at proof stress o'p under conditions where failure
is crack-propagation-controlled and the values for n and a are known, and Evans and
Wiederhorn presented time-to-failure analyses for soda-lime glass. Furthermore, Evans and
Wiederhorn combined the analysis for slow crack growth with probabilities of failure for
fast crack growth to determine the statistics of failure.
Ritter successfully applied these principles of fracture mechanics coupled with a Weibull
function to predict the failure of optical glass fibers [21] consisting of fused silica with
polymeric cladding. The agreement between prediction and experiments under dynamic
fatigue was consistent with slow crack growth of pre-existing flaws being the controlling
mechanism in failure of optical fibers. Subsequently, Phani [22] improved the fit for the
data of Ritter et al. on fracture strength, using a two-parameter Weibull model where
Ritter et al. had previously employed only a single parameter. Phani argued that the single-
p a r a m e t e r distribution was insufficient for the total distribution of strength.

2.3. Stress corrosion of glass fibers

D u n c a n studied the weakening of silica optical fibers under wet environments [23]. He
showed that both the degree of static strain and the duration of time under conditions of
static strain decreased the time to failure of the silica optical fiber. The experimental data
did not fit his empirical plot of breaking strain vs. time under strain, derived from V=AKI n.
Possibly the assumptions that Duncan made for these calculations were not totally valid
(n = 23) or the presence of the silicone coating influenced the test (a 50-~m coating on a
200-/.~m fiber). More recently, Rondinella and Matthewson [24] illustrated the importance
of proper preconditioning on coated optical fibers prior to testing because of consideration
of diffusion through the coating. Duncan's results showed that exposure to water decreased
the strength of the fiber, and exposure to deuterated water had a less severe effect on the
rate of crack growth than did non-deuterated water. This result was consistent with the
chemical reaction of water with the glass being the rate-determining step for failure of the
fiber.
E-glass (by weight %, 16% CaO, 14.5% A1203, 9.5% B203, 5% MgO, ~ 1% Na20 and
K20, and 55% SiO2 [25]) includes metal oxides that make it intrinsically alkaline and
270 C.L. Schutte / Environmental durability of glass-fiber composites

hygroscopic, thus facilitating stress corrosion of the fibers [26]. Schmitz and Metcalfe found
that an induction period existed in the stress corrosion of filaments of E-glass in the presence
of water. Once a surface flaw had formed, catastrophic failure of the filament occurred.
According to Bascom [25], this result implied that it was necessary for water to adsorb
onto the surface of the glass filament and form a film through which the transport of the
material that was responsible for corrosion occurred.
Both Barker [27] and Ishai [28] showed that rough surfaces of the glass fibers indicated
that water had attacked the glass. Samples that exhibited these rough surfaces also exhibited
a loss of strength from exposure to water. They did not determine whether this surface
roughening was the only cause of loss of strength for the sample. Scanning electron
micrographs showed the rough surfaces, and analyses by infrared spectrometry indicated
that material leached from the fibers into the resin after the sample's exposure in hot
water.
Agarwal and Broutman [29] described the general features of stress corrosion. First,
the strength of the fiber is independent of the time after application of a load under an
inert atmosphere and/or at low temperature. Second, the strength is independent of exposure
to active environments (water) before the application of the load. Third, the strength is
d e p e n d e n t on exposure to environments under an applied stress; the residual strength of
the fiber after its exposure to corrosive environments and an applied stress is less than
that for unexposed fibers, suggesting that the applied stress influenced the corrosion of the
glass. Fourth, the residual strength for fibers simultaneously exposed to a corrosive environment
and an applied stress decreases with increasing temperature.
Using a two-point bending test, France et al. [30] studied the durability of sodium
borosilicates with either barium- or germanium-oxide-modified cores to alter their index of
refraction. They modified the outside layers of the fibers with either AI, Mg or Zn to
improve the durability of the fiber. Modified dynamic bending fatigue studies at temperatures
between - 140 and + 100 °C at various relative humidities (RH), showed that the susceptibility
to stress corrosion was d e p e n d e n t on the relative content of water. The strengths of glass
fibers as a function of exposure in 20 °C distilled water were d e p e n d e n t on their composition.
The relative rates of degradation were: sodium borosilicate (high content of sodium) > sodium
borosilicate (low content of sodium) > sodium magnesium borosilicate ( ~ 1-2 mole %
Mg) > sodium aluminum borosilicate ( ~ 1-2 mole % A1) > sodium zinc aluminum borosilicate
( ~ 1-2 mole % each, AI and Z n ) > sodium magnesium zinc aluminum borosilicate ( ~ 1-2
mole % each, Mg, A1 and Zn).
Static fatigue studies both as a function of composition of the glass in 20 °C air and
after exposure to 20 °C distilled water showed that relative susceptibility towards static
fatigue was: sodium borosilicate>sodium aluminum borosilicate>sodium zinc aluminum
borosilicate > sodium zinc aluminum borosilicate. The relative ranking of times to failure
for fibers pretreated in distilled water followed the same trend as for the susceptibility
towards static fatigue; however, the times to failure were shorter for fibers immersed in
distilled water compared to those exposed to air.
Furthermore, the plots of percent applied strain vs. median time to failure for fibers
previously exposed to distilled water curved downward, and a straight line could not be
fitted to those data. This departure from the theory of stress corrosion motivated France
et al. to modify a theory on static fatigue to include consideration for both stress corrosion
and static fatigue (aging under zero stress), where for failures at early times the aging
under zero stress dominated, and, for failures at longer times, fatigue dominated. This
C.L. Schutte / Environmental durability of glass-fiber composites 271

effort by France et al. yielded the following equation for the time to failure tf,

[ 1+?)
t~ 2
\ ~ ( N - 2)

=BS°N-2°~-N (3)

where a and [3 are parameters relating residual strength of the fiber as a function of time
t to the initial inert strength S(t)= S0(1 + at)- ~, N is defined as the susceptibility to corrosion
under stress, So is the initial inert strength, presumably the inherent strength, of the glass
fiber, and (r is the applied constant stress. This modification fitted the data well.
Ritter [31] used the model of France et al. [30] that described modified static fatigue
to study the properties of four different optical fibers with a variety of coatings: amorphous
carbon, TiO2, a plastic coating and TiC. All of these coatings had polymeric coatings for
added protection. Studying fatigue involved exposure of the coated fibers in distilled water
at 25, 60 and 85 °C. The C and TiC coatings measurably increased the resistance to fatigue
and reduced the effect of temperature on strength of the optical fibers compared with the
TiO2 and plastic coatings. Moreover, up to 500 h, where the fiber with the carbon coating
showed no aging under stress and under exposure in distilled water at either 60 or 85 °C,
fibers with the other coatings experienced a reduction in strength due to zero-stress aging,
particularly at the higher temperature of 85 °C. Unfortunately, none of the fibers exhibited
optimal properties with respect to strength, fatigue and aging. Certainly, the need exists
for further research and development of glass fibers that have minimized these problems
in long-term behavior.
By creating flaws of different sizes, Dabbs and Lawn [32] examined the effect of the
size of the pre-existing flaw on the growth of cracks in glass fiber. They compared subthreshold-
with post-threshold-sized flaws and found significant differences in their behavior. A dis-
continuity exists between the inert strength in fibers containing subthreshold flaws, which
had a higher inert strength, vs. those containing post-threshold flaws. This result was
consistent with the growth of cracks being initiation-controlled for those flaws that were
below the critical size, and propagation-controlled for those above, implying that only in
the post-threshold domain can fracture mechanics define the criterion for failure. They also
speculated that the influence of residual stress also influenced the susceptibility of the fibers
towards fatigue under these two conditions; the fibers containing subthreshold flaws exhibited
a greater susceptibility towards fatigue than did those containing post-threshold flaws.
The significance of this approach lies in the ability to study the behavior of the strength
and fatigue of fibers as a function of size of the flaws. That the authors found such large
differences in the behavior of the glass fiber between those containing subcritical vs.
postcritical flaws highlights the limitations of our understanding of fracture mechanics of
glass failure. They pointed out that the data collected for fibers containing post-threshold
flaws would not adequately describe the behavior of fibers containing subthreshold flaws,
and further work in understanding the subthreshold domain is certainly needed.
Aveston et al. [33] compared the time to rupture under stress for fibers of E-glass
compared with that for a corrosion-resistant glass, Cemfil, after previous exposure to water,
under either no load or a static load. Wet E-glass fibers that were under static load prior
to testing had their strength decrease faster than fibers that had no prior load, regardless
of whether the fibers were uncoated or coated in polyester or epoxy. Because of their
corrosion-resistant nature, the Cemfil fibers retained their strength to a greater degree than
272 C.L. Schutte / Environmental durability of glass-fiber composites

did the E-glass fibers. Furthermore, from bulk samples of these glasses, Aveston et al.
derived a law which describes the rate of growth of cracks, utilizing Weibull statistics, and
a theory of stress corrosion for bundles of fibers that predicts the time to failure of these
glasses. The theory fits their experimental data well in the range of growth of cracks below
10 - 1 ° m s -1.

2.4. Stress corrosion of fibers in composites

Although stress corrosion of glass fibers can occur in composites, the isolation of this
mechanism from other processes that are responsible for damage is not always straightforward.
Jones et al. [34] showed for a glass/epoxy composite, made of 11 plies at 0°/90° orientation
with 28___2.5 wt.% E-glass and preconditioned under 65% R H at room temperature for
four months, that the tensile strength was slightly higher than for the unexposed samples,
its tensile modulus was 10% lower, and the effect on the fatigue behavior was insignificant.
The changes in properties were similar for samples of both orientations, 0°/90 ° and _ 45%
Jones et al. also boiled samples in water for three weeks and subsequently stored them in
water at room temperature, and these samples showed a reduction in mechanical properties.
The tensile strength decreased by 50% and 84% for 0o/90° and + 45 ° orientations, respectively.
The tensile modulus decreased by 15% for the samples at 0°/90° orientation. The fatigue
behavior of the previously boiled samples, illustrated in an S-log N curve (S is the stress
to failure, and N the n u m b e r of cycles), started at a level 50% below that of the dry sample
and crossed the curve for the dry sample at 104 to 105 cycles for 0°/90% and followed a
line slightly lower than for a dry sample for + 45 °. This damage from boiling water, illustrated
in the results of tensile tests of previously wet and subsequently dried samples, was not
reversible.
The authors concluded that the fatigue behavior of the boiled samples shows them to
be more resistant to fatigue at low stresses (200 MPa or less) than unexposed samples.
Their rationale involves the fact that, although the initial strength of the wet samples drops
to 50% of that of the dry sample, the slope of the plot does not drop as drastically as
that for the dry sample. The authors [35] suggest three possible reasons for this result:
first, water could plasticize and swell the resin, potentially increasing the strain to failure
of the resin and relaxing thermal strains developed in processing; second, the ability of the
interface to transfer stress to the fiber might have decreased, and reduced the concentration
of stress at broken fibers or resin microcracks, resulting in a possible increase of the size
of the critical crack nucleus; and, third, the distribution of strength of the fibers may change
as a result of stress corrosion, as previously illustrated by Mandell et al. [36].
The rate of the test of fatigue appears to have a significant influence on the residual
strength of glass-fiber composites. Because this was not the case for either carbon or Kevlar
fiber composites, Jones et al. attributed the loss of 75 MPa per decade of speed of testing
to damage to the strength of the glass fibers in the dry samples. Similarly, Metcalf and
Schmitz found a decrease in the strength of their glass fiber composite of 300 MPa per
decade of rate of application of load in the presence of water [37].
For their conditions, Jones et al. showed that previous loading of the sample did not
increase the uptake of moisture on exposure to boiling water. Loading samples prior to
testing u n d e r conditions of fatigue, for orientations at 0°/90 ° and at + 45 °, produced no
measurable differences in properties when loaded at 50% of the fracture stress, 0.4 GPa,
under dynamic fatigue. For samples that had been previously loaded at 80% of the fracture
C.L. Schutte / Environmental durability of glass-fiber composites 273

stress, the S-log N curves, for both orientations, were lower and converged at lower levels
of stress than those for previously untreated samples. The authors suggested that additional
water was present in these samples, owing to previous damage from prior loading. They
also observed fiber damage from the application of prior loading.
U n d e r different conditions, when prestressing at lower levels of load or boiling for
shorter times, the samples appeared to have a mechanism of damage that was controlled
by the fibers rather than the resin. For example, the results of Sims and Gladman [38,39]
and of Boiler [40] showed convergence, at low stress, of the S - l o g N curves for samples
that had been previously boiled in water with those of samples that had not. Their results
agreed with those of Mandell et al. [36], suggesting that damage to the fiber was the
dominant mechanism for loss of properties u p o n short-time exposure to water. Sims and
Gladman boiled their samples for only 64 h, compared to the treatment for 500 h employed
by Jones et al. Furthermore, Jones et al. saw an effect of orientation, 00/90 ° vs. ___ 45 °, in
his tests of fatigue, where Sims and Gladman did not. Jones et al. suggested that, based
on their data and those of Sims and Gladman, the slope of the S-log N curves decreased
with increasing damage to the composite, whether by boiling or preloading before the test
of fatigue [34].
Using Mandell's model -- where the slope B of the S-log N curve is plotted vs. tensile
strength -- one obtained a slope of 10(tensile strength/B) for samples of composites whose
mechanism of damage appeared to be due to loss in strength of the glass fiber. The data
of Sims and Gladman yielded slopes of 9.4 for glass/epoxy and 8.4 for glass/polyester
composites, when analyzed by the model of Mandell et al. The S-log N curves of Jones et
al. [34], however, are not linear. A poor fit of a straight line to these data resulted in a
slope of 13 (tensile strength/B) for the previously boiled sample at 00/90 ° and 8.7 for the
dry sample. Either the m o d e l of Mandell et al. is not applicable to these samples or to
their environmental treatment, or the scatter in the data was too great for further interpretation
of the mechanism of damage for the samples of Jones et al., or both.
Noble et al. [41] showed that, for E-glass/epoxy composites, stress corrosion occurred
on exposure to hydrochloric acid solutions. The level of stress, the concentration of acid
and the accessibility of the surface of the glass fibers to the solution influenced the time
to failure of the sample. The surface of the glass fibers at the location of fracture was
depleted in aluminum, calcium, sodium, and potassium, as determined by X-ray analysis.
Furthermore, the fracture surfaces exhibited debonding at the fiber/matrix interface. The
authors found that a mechanism of attack on the fiber was consistent with their results.

2.5. Summary

The failure of glass fibers under environmental exposure depends on the presence of
moisture, stress, time, compositions of the glass, and the presences of sub- or post-critical
flaws. General characteristics of the models of crack growth are applicable to the failure
of fibers in polymer resins; however, the time-dependent distribution of moisture in the
sample, particularly at the surface of the fiber, and the time-dependent distribution of
stress, associated with swelling stresses, will also influence the rate of crack growth. These
factors need to be included in models describing the time to failure of glass fibers. Although
the work that has been done thus far on models for crack growth, time-dependent failure,
stress corrosion, and the effect of composition of glass on susceptibility to failure has been
extremely useful in understanding the behavior of glass in composites, further understanding
274 C.L. Schutte / Environmental durability of glass-fiber composites

of the details of these p h e n o m e n a in the environment of the composite is necessary. For


many applications, E-glass is the fiber of choice because of cost/performance constraints:
however, revised requirements for use of polymer composites in new markets may d e m a n d
further optimization of glass-fiber performance (strength, fatigue, aging and stress corrosion)
and cost requirements through modified composites for long-term load-bearing structures.

3. Effect of the environment on properties of the matrix

The properties of the matrix are influenced by environmental exposure, and these
effects need to be considered when looking at composite durability. The examples cited
here address reversible (plasticization and swelling) and irreversible (hydrolysis and mi-
crocracking) changes in the resin properties. Furthermore, the mechanism of diffusion of
water, far from fitting a Fickian model, is not well understood. The ability to predict the
diffusion of water and its influence on the resin properties are necessary to predict long-
term behavior. Where distribution of water in polymers would yield more detailed information
of diffusion mechanisms in polymers, most studies of diffusion of water measure the uptake
of water only through weight gain. Marom's work showed the development of stress in the
samples as solvents diffused into the polymer, and this approach allows for more detailed
studies of the distribution of solvent as a function of time.
Water plasticizes epoxy resins, making t h e m more compliant [33,34,42,43]. Tsai looked
at the loss of strength over time as a function of environmental exposure, relative humidity
or heat, that would swell or expand the resins [44]. Ashbee and Wyatt [45] also observed
that swelling caused debonding of the fibers of E-glass if the generated swelling stress was
too great. Possibly, the swelling resulted in stresses that were greater than the effective
interracial strength; thus, failure at the interface occurred.
Some studies on the diffusion of water into resins assumed Fickian behavior; however,
there were many examples of non-Fickian based diffusion [46]. Ghorble and Valentin [47]
looked at Langmuir-type diffusion [48] to explain the penetration of water in polyester and
vinyl ester resins. The Langmuir model considered two phases for the water, bound and
unbound, with the equilibrium between these phases influencing the uptake of water, and
this approach adequately described the data for these systems. The authors also observed
plasticization, hydrolysis, debonding and formation of cracks in their samples, which should
also influence the diffusion of water in the material.
T h e structure of the resin has the potential to influence the uptake of water for that
material. For example, Carfagna et al. [49,50] found that the equilibrium content of water
scaled with the content of amine cross-linking agent in a series of diglycidylether bisphenol-
A epoxies cured with different contents of triethylenetetramine cross-linking agent, possibly
because of the increased hydrophilic nature of the network due to the presence of different
contents of the amine and hydroxyl functionalities. Furthermore, the thermal history of the
resin influenced the equilibrium content of water; an epoxy sample that had been pre-
equilibrated in water at 70 °C and cooled to 20 °C had a higher content of water than
one that had been equilibrated at 20 °C. This difference they attributed to microcavitation
[51], or irreversible damage, at the higher temperature. Possibly, differences in the content
of water between the two samples are a kinetic phenomenon; the diffusion of water is
faster at the higher temperature, and the sample exposed to water only at 20 °C may not
have been at its equilibrium content of water. Furthermore, these wet epoxies had Tg values
that were lower than the dry samples from plasticizing by the water. For samples that had
C.L. Schutte / Environmental durability of glass-fiber composites 275

low contents of cross-linking agent, the free volume theory [52] adequately described the
decrease in Tg values. However, for sample with higher contents of amine, an entropy model
[53] was more appropriate.
Water could not only swell and plasticize the resin, but also hydrolyze it. Johnson [54]
showed that the relative rate of degradation of properties in polyester resins was ortho
phthalic e s t e r > i s o phthalic ester>vinyl e s t e r > b i s p h e n o l - A ester. Presumably this trend
was due to the relative rates of hydrolysis of the polyesters.
Apicella et al. [55] studied the influence of the chemistry of polyester resins on the
retention of their mechanical properties after exposing polyester/glass fiber composites to
water at 25 °C and 90 °C. They found that the relative hydrolytic damage decreased as
isophthalic resins > bisphenol-B > bisphenol-A > vinyl ester, and suggested that the suscep-
tibility to hydrolytic attack increased with an increase in the n u m b e r of ester groups in
the polymer repeat unit. They did not, however, perform chemical analysis on their samples
to determine if hydrolysis did indeed occur.
As Illinger and Sprouse [56] found chemical reaction between water and the epoxy
group for a novolac/diglycidylether bisphenol-A system, so did Netravali et al. [57] for a
tetraglycidyl 4,4'-diaminodiphenylmethane/4,4'-diaminodiphenysulfone ( T G D D M / D D S )
epoxy. Whereas exposure of this resin to water at 20 °C resulted in recoverable changes,
exposure to water at 70 °C produced irreversible damage. Saturation at 20 °C caused
plasticizing of the resin, as observed by a lower wet Tg that was recovered after drying.
Exposure of the epoxy to water at 70 °C resulted in a lower Tg for contents of water up
to 4.2 wt.%, but no additional change in the Tg with an increased content of moisture.
The authors attributed this lack of further plasticization to evidence of microcracking in
the resin where water clusters accumulated, without the interactions with the resin to
decrease Tg. Further evidence for irreversible damage to the resin from exposure to water
at 70 °C included a lower Tg after drying, a decrease in the epoxide peak at 904 c m - 1
after exposure to water in the infrared spectrum, and a decrease in the curing energy
compared to that for a sample exposed to water at 20 °C from differential scanning
calorimetry.
Wimolkiatisak and Bell [58] showed that the tensile strength of an epoxy resin exposed
to water at either 50 or 100 °C decreased from 70 MPa to 58 MPa and 62 MPa, respectively.
In contrast, the dehydrated samples recovered their original tensile strengths; in fact the
one previously treated at 50 °C exhibited an increase in its tensile strength, to approximately
75 MPa, possibly through increased cross-linking from the thermal treatment.
M a r o m and co-workers studied extensively the effect of swelling stresses on diffusion
processes in resins and composites. They presented [60,61] a model for stress transfer to
explain deformation and failure for Case II diffusion [59] of an epoxy in methylene chloride
solvent, and poly(methlymethacrylate) (PMMA) in methanol [62], which they studied by
monitoring the uptake of solvent and the evolution of stress within samples. For both types
of systems where an advancing boundary of solvent occurred during the diffusion process,
their findings included: (a) use of the analogy of the mechanical constraints imposed at
the fiber/matrix interface for the transfer of stress at the interface between the unswollen,
stiff core (fiber) and the swollen and stressed outer resin that explained the different stages
of swelling that they observed during the diffusion process; (b) determination of the swelling
strains that d e p e n d e d on the size and relative dimensions of the unswoUen core; (c)
attainment of a critical core geometry that governed the relaxation and rearrangement of
strains in the sample (the content of solvent for which this occurred was d e p e n d e n t on
276 C.L. Schutte / Environmental durability of glass-fiber composites

temperature); (d) observation that the directional hygroelastic coefficients were dependent
on the dimensional proportions of the sample, and insensitive to both the temperature and
the surface-to-volume ratio; and (e) cracking of certain geometries experiencing swelling
stresses that could be predicted by the stress-transfer model.
In further work on styrene/alkyd resins exposed to water where the Fickian model
fitted the data, Cohn and Marom [63] showed that the coefficient of hygroelasticity was
dependent on the Young's modulus E of the resin, and not on its composition. In contrast,
they found a profound influence of the chemical structure on the hygroelastic properties
of an epoxy/methylenediamine cured system. These differences in the content of hardener
influenced the morphology of the epoxy network, the network with the low content of
hardener had a high free volume that allowed for high absorption of water without the
generation of stresses that would be associated with a network that was more highly cross-
linked. Furthermore, because formation of morphological phases [64] with different degrees
of cross-linking could occur, the morphological characteristics have the potential to influence
these hygroelastic properties as well.
Antoon and Koenig [65] showed that stress increased the rate of hydrolysis of ester
functional groups in an epoxy cross-linked by nadic methyl anhydride (NMA) in an aqueous
solution at high pH. By monitoring the frequencies unique to the ester group (1744 cm-1)
and the NMA (1860, 1080 and 918 cm -1) using Fourier transform infrared (FTIR)
spectroscopy, they found that hydrolysis of the ester and loss of NMA occurred initially,
while structural changes consistent with re-orienting of hydrogen bonding occurred later.
The application of stress to the resin increased the rate of loss of hydrolysis. Furthermore,
the presence of powdered E-glass increased the rate of hydrolysis of the ester group, either
because the glass is a concentrator of stress or its surface has a high pH.
In an effort to design conditions for accelerated testing that simulated environmental
exposure of low earth orbit and of salt water, Garton et al. [66] studied the oxidation of
polyethylene after the removal of antioxidants, under atmospheres of air, deionized water,
0.01 mol 1-1 NaC1, and 0.1 mol 1-1 NaCI. By measuring the concentration of carbonyl
functional groups with FTIR and UV spectroscopies, they determined that an insignificant
amount of oxidation occurred under air at 70-92 °C, whereas significant oxidation occurred
after 650 h in deionized water and in both salt solutions. By converting carboxylic acids
to acid fluorides (ketones would remain unreacted) and analyzing the carbonyl region of
the FTIR, they determined that a high percentage of the carbonyls were carboxylic acids
- a by-product of chain scission in polyethylene. Because the relative content of carboxylic
-

acid in the carbonyl species was unexpectedly high compared to the distribution of products
formed under non-aqueous conditions, they concluded that the accelerated tests cannot be
reliable without a basic understanding of the mechanisms responsible for the degradation.
Physical aging in polymers [67,68] can influence their behavior in composites. Sullivan
et al. [69] studied the creep behavior of several types of composite systems: thermosetting,
semicrystalline, amorphous and highly filled amorphous thermoplastic samples. Their results
showed that procedures employing time/aging-time and time/temperature superposition for
prediction of short-term creep were valid. However, because aging retarded the long-term
creep behavior, these models did not apply for long-term predictions.
3.1. Summary
The influence of environmental exposure can cause changes in the resin through a
diffusion of water into the sample, leading to a distribution of swelling stresses, hydrolysis
C.L. Schutte / Environmental durability of glass-fiber composites 277

and physical aging. Changes in the resin can be either reversible (plasticization and swelling)
or irreversible (too much swelling, hydrolysis and microcracking). The chemical composition
appears to influence both the solubility of water in the resin as well as its susceptibility
towards hydrolysis. The equilibrium content of water will determine the degree to which
swelling stresses occur. The larger the content of water, the greater can be the swelling
stresses that may cause failure of the polymer by hydrolyzing or microcracking it. Optimization
of resins for long-term environmental exposure would involve making them more hydrophobic
in order to decrease their equilibrium water content at saturation, less susceptible to
hydrolysis, less prone to changes in properties from physical aging, and more resistant to
microcracking from swelling stresses.

4. Effect of the environment on properties at the interface

In order for a polymer composite to have properties that are advantageous, the interface
between the fiber and polymer matrix must have good (although not necessarily maximum)
adhesion; in order for the composite to maintain its properties on exposure to the environment,
that interface must resist environmental degradation. This section focuses on studies of
environmental attack at the interface or interphase region, where this region is the weak
link and therefore responsible for the failure of the composite. To identify the mechanism
of attack at the interface, one must understand first the fundamental issues that govern
the chemistry [70], structure, morphology, and modes of failure at that interface. This
section addresses the factors that influence the chemistry and morphology of coupling agents
at and with the surface of the glass, interactions of the coupling agents with the resin, and
the role of surface treatments in maintaining properties after exposure to the environment.
Previously published reviews addressed mechanisms of reinforcement of glass-fiber
composites under wet conditions [71] and molecular microstructure of coupling agents [72].
Theoretically, coupling agents can react chemically with both the glass, through silicon
hydroxyl groups, and the resin, through an organic functional group that is compatible with
the chemistry of the resin. This view is overly simplistic, as experimental studies of coatings
of coupling agents reveal a complicated, multilayer structure. In general, deposition of
coupling agents from water results in three layers, or phases, on the surface of the glass:
a monolayer, a chemisorbed layer and a physisorbed layer. The role that these phases play
in enhancing the durability between the glass fiber and the resin is not totally understood;
however, some of the studies cited here suggest how the phases may influence the properties
of the interface and interphase.
Because of the coupling agents' ability to significantly influence adhesion at the interface
and the morphology of the interphasal region, and the interdependence between these
factors and the durability of the interface, this section presents an overview of the chemistry
and morphology of coupling agents. Evidence for the morphological model of the coating
of coupling agent includes radio-labeled tracer experiments, as well as infrared, X-ray
photoelectron, diffuse-reflectance and ion-scattering spectroscopies, coupled with secondary-
ion mass spectrometry. Furthermore, the chemistry of the coupling agent influences its
adsorption onto the substrate. Coatings with tailored morphology can be made from solutions
by both selective adsorption and desorption, with the pH of the aqueous solution and the
solvent used for non-aqueous deposition having significant influence on the structure. Also,
the thermal pretreatment of the coating influences the structure and degree of interpenetration
278 C.L. Schutte / Environmental durability of glass-fiber composites

in the interphase region and its susceptibility to environmental attack. Fowkes looked at
the interaction between the coupling agents and the resin as acid/base interactions at the
surface, and his experimental results cite examples where this interpretation explains the
results well. The morphology of the coating influences interfacial properties, and studies
that investigated these structure/property relationships on interfacial strength and durability
are included. These studies illustrate the usefulness of single-fiber tests, especially with
respect to environmental exposure.

4.1. Chemistry of coupling agents

4.1.1. Structure and reaction of coupling agents on surfaces of glass


The chemical structure of a coupling agent contains one end of the molecule that has
the potential to react with the glass Si-OH functional groups, typically a trifunctional silicon
with either chlorine or alkoxides, and the other end of the molecule can react chemically
with the resin. Most glasses contain silicon hydroxyl groups as well as additional metal
hydroxyl groups, but most characterization studies focused on reaction of coupling agents
with surface S i - O H groups. Both infrared [73] and Raman [74] spectroscopic studies provide
evidence that the silanol (SiOH) groups of the coupling agent vinyltriethoxysilane (VTES)
chemically reacted with the surface of small particles of silica. Ishida and Koenig [73]
assigned infrared absorption bands at 1170 and 1080 cm -1 to antisymmetric stretching of
the Si-O-Si bond between the coupling agent and silica. They quantified the degree of
reaction for the coupling agent by measuring the intensity of the SiOH group for vinyl-
triethoxysilane as a function of thermal history. Raman spectroscopy yielded evidence for
chemical bonding between the coupling agent and the surface of the fiber. They assigned
the weakly scattering band at 1085 cm -1 to Si-O-Si bonds formed between the glass and
hydrolyzed VTES.
A further study [75] examined the structure of T-methacryloxypropyltrimethoxysilane
(MPS) and VTES on the surfaces of fibers made of E-glass. Two findings of this study
were that the structure of the coupling agent in solution appeared to influence the structure
of the coating, and that the glass substrate influenced the degree of condensation of the
SiOH groups, presumably because condensation of the SiOH groups is catalyzed by base
and the surface of the glass has a higher pH than that of the solution. The amount of
deposited silane depended on the concentration of coupling agent in solution.
In a series of Raman spectra, the symmetric stretch for the vinyltrihydroxysilane at
678 cm-1 increased abruptly for samples coated from a solution at 1.5% relative to solutions
at lower concentrations. Increased aggregation of the coupling agent in solution influenced
the hydrogen bonding in the coupling agent; it is this change in the hydrogen bonding that
apparently influenced the signal at 678 cm -1, associated with hydrogen bonding, in the
Raman spectrum of the coating. The microstructure of the coupling agent deposited on
the surface therefore appeared to depend on its structure in solution.
For the MPS system, the SiOH stretching mode in the IR at 880 cm -~ shifted to a
lower frequency with increased density of cross-links. Ishida and Koenig attributed the
increase in the rate of condensation of the siloxanes on the surface of the glass, compared
to the rate without a substrate, to "ordering" of the siloxane on the surface of the glass
[75] and to the ability of the surface of the glass to catalyze condensation reactions due
to the high pH of the metal hydroxides in the glass [76]. The carbonyl group in the MPS,
while capable of hydrogen bonding with SiOH functionality (1720 and 1700 cm-1; the H-
C.L. Schutte / Environmental durability of glass-fiber composites 279

b o n d e d carbonyl is at lower frequency), did not show as much H-bonding on the glass
substrate as it did under conditions without a substrate. Therefore, the authors suggested
that MPS deposited in a head-to-head arrangement on the surface of the glass, where SiOH
functional groups were unavailable for hydrogen bonding with the carbonyl group.

4.1.2. Reaction of coupling agent with polymeric resin


The coupling agent has the potential to react chemically with the resin, and the following
works cite evidence for chemical reactions of this type. A study of the interaction of coupling
agents with resin [77] looked at the loss of radio-labeled resin from non-labeled coupling
agents on fibers of E-glass. Fibers, previously treated with either (methacryloxy)propyl-
trimethoxysilane (MPS) or vinyltrimethoxysilane, coated with methyl-14C-methacrylate/styrene
90/10 and benzoyl peroxide initiator, and subsequently polymerized at 90 °C for 18 h,
retained a coating of resin that was approximately 0.235 ~ m thick. This retention was for
samples containing a coating of coupling agent that was boiled in refluxing tetrahydrofuran
for 3 h and subsequently dried at room temperature. In contrast, labeled resin on previously
untreated fibers retained only a film of thickness 1.0 nm, as determined by the concentration
of radioactive material, after the same treatment. These results were consistent with the
double bond in the coupling agent reacting with either the styrene or the methyl methacrylate
monomer. This chemical bond between resin and coupling agent (chemically bound to the
glass-fiber substrate through the coupling agent) served to retain the resin through the
treatment in tetrahydrofuran.
Polymerization of methyl methacrylate in the presence of vinylsilane-treated fibers
[74] yielded poly(methyl methacrylate) that could not be extracted by methyl ethyl ketone
or methanol from the surface of the treated fibers. This result, plus the disappearance
in the R a m a n spectrum of the vinyl functionality of the silane coupling agent, served as
evidence that copolymerization between the coupling agent and methyl methacrylate
occurred.
Liao [78] studied the interfacial reaction between coupling agent and epoxy resin using
FTIR. The system contained an epoxy formed from the reaction of diglycidylether bisphenol-
A and nadic methylanhydride with N-methylaminopropyltriethoxysilane (MAPS)-coated fibers
of E-glass. Absorbance in the F T I R at 1640 c m - 1 indicated the presence of an amide
functional group that was formed, presumably, between the amine in the coupling agent
and a carbonyl species in the resin, the extent of which was greatest for a solution p H of
10, or the natural, unbuffered solution. Liao also presented evidence of hydrothermal attack
of the amide bond by demonstrating that the intensity of the amide absorption decreased
with increased exposure to boiling water. A decrease in the intensities of the nadic methyl
anhydride at 1860, 1080 and 916 cm -~ provided evidence of leaching of the anhydride
component.

4.2. Characterization of morphology

Coatings of coupling agent that have been deposited from aqueous solution have a
three-layer morphology, as shown in Fig. 3. Generally, these layers consist of the following
morphological components.
(1) A physisorbed phase forms at the outside layer of the coating. Immersion of the
sample in cold or warm water removes this layer. Because of its ease of removal, this layer
has the potential to mix with the resin during processing of the resin; thus the properties
280 C.L. Schutte / Environmental durability of glass-fiber composites

Idealized "Menolayer"
Idealized "Monolayer" plus
Idealized "Monolayer" plus Chemisorbed and
Chemisorbed phase Physisorbed phase

~--- "

/ /
Fig. 3. Illustration of the three-phase morphology of a coupling agent as deposited from an aqueous solution.

of the interphase can be significantly influenced by the presence and structure of the coating
of coupling agent.
(2) A chemisorbed layer forms below the physisorbed phase. This layer has a higher
degree of cross-link density than does the physisorbed phase. This difference in cross-link
density was evident from studies of selective desorption; the chemisorbed phase was stable
in cold or warm water, but it was extractable into boiling water.
(3) A monolayer, or monolayer-equivalent, phase existed at the surface of the glass.
It appeared that this phase was chemically bound to the surface of the glass, through
Sig-O-Sic linkages between the glass and the coupling agent.
The relative thicknesses of the chemisorbed and physisorbed phases were dependent
on the conditions of depositions. Because of the influence that the morphology of the
coating of coupling agent has on the mechanical properties of the interface, as well as on
its durability (discussed in later sections), this section covers the experimental evidence for
the morphological model of the coating. These results include the use of radio-labeled
traces and infrared, X-ray photoelectron and diffuse-reflectance spectroscopies, as well as
secondary-ion mass spectrometry and size-exclusion chromatography.

4.2.1. Characterization by radio-labeled tracers


F r o m the surface of Pyrex glass, Schrader [79-81] selectively desorbed ~4C-labeled 7-
aminopropyltriethoxysilane that was labeled at the carbon alpha to the amine -
(CH3CH20)aSiCH2CH2C*H2NH2. His results showed three fractions of coupling agent:
fraction 1 represented a maximum of 98% of the total content of silane that was physically
adsorbed to the surface of the fiber (as much as 270 monolayer equivalents) and was readily
removed by a rinse with cold water; fraction 2 was chemisorbed polymer (approximately
10 monolayer equivalents) and could be removed by an extraction in boiling water for 3
C.L. Schutte / Environmental durability of glass-fiber composites 281

to 4 h; and fraction 3 was a monolayer that could not be removed after prolonged exposure
to boiling water.

4.2.2. Characterization by infrared spectroscopy (IR)


Koenig and co-workers [82,83] characterized the products of polymerization of vinyl-
trimethoxysilane-treated fibers of E-glass by Raman spectroscopy. The results of monitoring
the absorbance in the spectrum at 1600 cm -1, corresponding to the C = C in the vinyl
moiety of the coupling agent, both before and after reaction with chlorine, suggested
multilayer coverage of the coupling agent. Because the absorbance line only decreased in
intensity and did not disappear, they determined that not all of the vinyl groups were
accessible to chlorine. If chlorine were available to the double bond, complete reaction
would occur, and the line at 1600 cm -1 should disappear, as it did for a fine powder of
pure vinyltrimethoxysilane. The presence of vinyl groups in layers below the "accessible
phase" was consistent with these results.
A further study [84] of VTMS-treated E-glass fibers after both exposure to water at
80 °C for 1464 h and heating to 150 °C for 2 h showed that the FTIR absorption attributed
to Si-OH functional groups at 920 cm -1 decreased and shifted to 856 cm -1. Apparently,
exposure to heat caused some condensation of Si-OH ~to Si-O-Si bonds. Also, a lower
content of Si-OH available for hydrogen bonding appeared to decrease the frequency of
the Si-OH mode that was hydrogen-bonded. This work serves as an example of how FTIR
can be used to characterize the degree of network formation, or morphology, of the coating
derived from a coupling agent.

4.2.3. Characterization by X-ray photoelectron spectroscopy (XPS)


More recent studies on the structure of y-APS using XPS revealed a variation of
structure depending on the position of the coupling agent in the coating [85]. This coating,
deposited from a 1.5 wt.% solution in deionized water, was on a flat plate of E-glass. At
the outer surface, the amino group was in the form of free amine, as indicated by a binding
energy of the N ls peak at 400.3 eV. Removal of this physisorbed phase occurred in water
at 50 °C for 24 h. The chemisorbed phase appeared to have a network containing hydrogen
bonding between the amine end-groups of the y-APS and silanol groups. The peak for
binding energy of N ls for this phase, consistent with a hydrogen-bonded amine, was 401.4
eV. The authors suggested an intermolecular rather than an intramolecular hydrogen-bonded
structure, possibly at the ends of the oligomeric coupling agents. Boiling the sample in
water at 100 °C resulted in removal of the chemisorbed phase. The remaining coating of
coupling agent appeared to have free amine, N ls binding energy 400.3 eV, and a high
density of cross-links. XPS revealed the presence of Ca and A1 in the coupling agent,
extracted from the E-glass, and these ions decreased in content after the extraction of the
physisorbed phase in water at 50 °C but increased after treatment in water at 100 °C. The
concentration of A1 increased after the boiling-water extraction, and Wang and Jones
suggested that aluminum was incorporated into the monolayer equivalent structure of the
coupling agent at the surface of the glass.

4.2. 4. Characterization by diffuse-reflectance infrared spectroscopy (DRIFT) and size-exclusion


chromatography (SEC)
Nishiyama et al. [86] studied the adsorption of y-methacryloxypropyltrimethoxysilane,
MPS, on colloidal silica. Specifically, they employed D R I F T and SEC to look at the
282 C.L. Schutte / Environmental durability of glass-fiber composites

chemisorbed and physisorbed fractions of MPS as a function of degree of surface coverage.


They studied the influence of concentration of silane in solution, of drying time, and of
pH of the solution on the morphological characteristics of the MPS derived coatings. In
general: (a) increasing the concentration of silane in solution increased the total and relative
amounts of physisorbed silane and increased the amount of chemisorbed silane, up to a
solution concentration of silane equal to 133 mole m -3 (33 mg ml-X); (b) increasing the
time of drying decreased the amount of physisorbed silane, increased the degree of
condensation between SiOH groups, and increased the molecular weight of the extractable
fraction; and (c) lowering the p H increased the molecular weight (800 to 4000 polystyrene
equivalent molecular weight range -- sizes of molecules equivalent to that for polystyrene,
i.e. between 800 and 4000 molecular weight as determined by SEC) of the physisorbed
silane oligomers and increased the relative amount of physisorbed silane.
Miller and Ishida [87,88] studied the intermolecular condensation reactions of MPS
on the surface of lead oxide as a function of surface coverage, using D R I F T . They chose
lead oxide as the substrate (surface area, 0.0005 m 2 kg -1) to avoid interferences from
substrate/coupling agent reactions in the IR; the Pb-O--Si absorption appeared at 960 cm-1,
while that for intermolecular condensation, Si-O--Si, appeared at 1130 and 1040 cm -1.
Discontinuities in the degree of intermolecular condensation of MPS as a function of degree
of surface coverage existed in a plot of the normalized content of intermolecular condensation
vs. degree of surface coverage. There was no detectable intermolecular reaction up to a
concentration of 0.0008 mol kg-1 (0.2 mg g-1) substrate. This content of intermolecular
reaction increased to 0.0020 tool kg -1 (0.5 mg g - l ) substrate, flattened out at 0.0040 mol
kg-1 (1.0 mg g-1) substrate, and increased thereafter. These data were consistent with the
multilayer/multiphase model describing the morphology of coupling agents.
The increase in the Si-O-Si bond formation between 0.0 and 0.0020 mol kg-1 (0 and
0.5 mg g-1) substrate was consistent with the formation of a monolayer on the surface.
The relative amount of intermolecular condensation between the molecules of MPS was
approximately 5%: the remainder were attributed to P b - O - S i substrate/coupling agent
linkages. The concentration of P b - O - S i functionality increased linearly from 0.0 to 0.004
mole m -z (0.0 to 1 mg m -z) coverage of the substrate. The change in adsorption behavior
- the flattening out of the curve of intermolecular condensation as a function of surface
-

coverage of coupling agent between 0.0020 and 0.0028 mol kg-1 (0.5 and 0.7 mg g-1) _
beyond the monolayer coverage initially produced a less "ordered" layer, followed by an
increase in the siloxane condensation, with a corresponding increase in the degree of order,
as the surface coverage increased. The influence of the lead oxide surface compared to
that of E-glass was not straightforward [89].

4. 2. 5. Characterization by secondary-ion mass spectrometry (SIMS) and ion-scattering spectroscopy


(ISS)
DiBenedetto and Scola [90] employed SIMS and ISS to study the structure of 3,-
aminopropyltriethoxysilane on S-glass, a glass formulated for high strength, containing (by
weight) 64.2% SiO2, 24.8% AI/O3, 0.21% Fe203, 0.01% CaO, 10.27% MgO, 0.27% Na20,
0.01% BzO3 and 0.2% BaO) [91]. They too found a morphology consisting of three phases.
Their results showed that the degree of condensation was greatest for that fraction nearest
the surface of the glass (monolayer equivalent) followed by the fraction at the air/coating
interface (physisorbed phase), with the region between these phases (chemisorbed phase)
having the lowest degree of Si-O-Si condensation. Unless the outer layer comes off by
C.L. Schutte / Environmental durability of glass-fiber composites 283

failure at the interlayer in solvent during preparation of the sample for analysis, these
results are not consistent with the easily removed physisorbed outer layer described by
Schrader. Specifically, analysis by SIMS showed that the concentrations of both nitrogen
and hydrogen were relatively constant with increasing distance from the air/silane interface.
At a depth of approximately 16 to 24 nm, the content of nitrogen increased and the simple
spectra from the SIMS analysis suggested the presence of low-molecular-weight oligomer.
Between a depth of 24 n m and the silane/glass interface, the contents of nitrogen and
hydrogen again were constant, but at levels that were three times higher than those at the
outer layer. This higher content of nitrogen at the silane/glass interface compared to that
at the air/silane interface is consistent with a higher concentration of coupling agent,
suggesting a tighter network at the surface of the glass. Therefore, this XPS result is
consistent with the Schrader model.

4.3. Adsorption and desorption characteristics

Further studies to elucidate the morphology of the coating of coupling agents involved
their adsorption and desorption characteristics. Factors that govern these properties are:
the chemistry of the coupling agent, p H of aqueous solution of deposition, the solvent used
for non-aqueous depositions, and the substrate. The studies characterizing the desorption
characteristics of the coating illustrate methodologies whereby tailoring of the interface
morphology is achievable. This general approach to tailoring surfaces allows for the study
of the influence or role of each layer on the strength and durability of that interface/
interphase, by building the layers one at a time and studying the behavior of the intermediate
structure. With the added insight that studies in surface chemistry coupled with mechanical
charactization can give, the selective deposition of the coating can then be optimized for
a given requirement for performance.

4.3.1. Influence of the chemistry on adsorption characteristics


The chemical structure of the coupling agent can have a profound effect on the adsorption
characteristics of the coupling agent. For example, Johannson et al. [77] used 14C-labeled
coupling agents -- y(amino)propyltriethoxysilane-y-14C (APS), y(methacryloxy)propyltri-
methoxysilane-y-14C (MPS), and 3-(2',3'-epoxy-propyl) propyltrimethoxysilane-3-]4C -- to
study their adsorption onto fibers of E-glass as a function of the concentration of coupling
agent in solution. Adsorption of these coupling agents appeared to be kinetically controlled,
with rapid adsorption occurring in the first 30 s, followed by a slow buildup of additional
silane thereafter. The adsorption behavior of the neutral coupling agents involved an increase
in the relative amount of coupling agent adsorbed with increasing concentration of silane.
In contrast, the relative adsorption of the amine-containing silane initially increased, but
then saturated the surface and remained constant beyond a concentration of 0.10 wt.%
silane. They attributed this difference in behavior between these coupling agents to the
ability of the amine functionality to adsorb onto the glass and, possibly, to influence its
conformation on that surface. They cited additional data on the strong interaction of amines
with glass from their study of adsorption of allylamine-l-~4C to E-glass compared to that
of propenyl-3-~4C-methacrylate, demonstrating two orders of magnitude greater adsorption
of the amine compared to that of the methacrylate.
284 C.L. Schutte / Environmental durability of glass-fiber composites

4.3.2. Effect of the initial coverage on selective desorption


A study [77] of the desorption of coupling agent in solution from fibers of E-glass
showed that, after exposure to boiling water, the rate of loss of coupling agent was independent
of the initial coverage. Also, the amount of coupling agent that remained was dependent
on the initial loading of silane. The increase in the relative retention of coupling agent
increased with the presence of a coating of resin on the pretreated fiber. Johannson found
94% retention of radio-labeled coupling agent, MPS, on a fiber that was subsequently coated
with unlabeled methyl methacrylate and styrene and boiled in water for 5 days.

4.3.3. Effect of p H on deposition


The pH of the solution influenced the amount of aminopropylsilane, APS, deposited
on a substrate, increasing from pH 2 to 10.6 and then decreasing thereafter; the pH did
not, however, influence the nature of deposition of methacroxypropyltriethoxysilane, MPS,
or vinyltriethoxysilane. Apparently, according to Koenig and co-workers, the interaction of
the amine and the silicon hydroxyl influenced not only the conformation of the coupling
agent in solution, but also its conformation on the surface of E-glass [92], possibly through
hydrogen bonding. Carbon dioxide reacted with 54 mole% of the APS (the remaining
material being hydrogen-bonded amine and therefore inaccessible to the carbon dioxide),
to form an amine bicarbonate salt at a pH > 5.6; for pH < 5.6, the primary amine exists as
an amine hydrochloride. The degree of reaction was dependent on pH and on concentration
of coupling agent, with the greatest uptake of coupling agent being at the unbuffered pH
of 10.6 [93]. Upon heating for 2 min at 100 °C, the samples lost CO2. This trend of salt
formation with the primary amine also existed for diamine, N-2-aminoethyl-3-aminopro-
pyltrimethoxysilane, and triamine, trimethoxysilylpropyldiethylenetriamine, coupling agents.
Differences in salt formation and/or hydrogen bonding potentially could influence the Si-O-Si
condensation of the APS coupling agent on the surface of the glass [94].

4.3. 4. Effect o f solvent on deposition


An early investigation, using ellipsometry, by Tutas et al. [95] of the thickness of
coatings of coupling agents revealed that the solution used for depositing the coating agent
significantly influenced the thickness of the coating. Deposition of vinyl tris(2-methoxy-
ethoxy)silane from a 1% solution in methy ethyl ketone onto surfaces of flat glass plates
resulted in a coating whose thickness was only 0.5-1.0 nm, with an index of refraction of
1.52, indicating a denser coating compared to that deposited from water, whose index of
refraction was 1.50. Coatings made from 1 wt.% aqueous solutions were significantly thicker
than single monolayers and more representative of the three-phase morphology obtained
from aqueous solutions: 14--18 nm for vinyl tris(2-methoxyethoxy)silane and 5-18 nm for
3,-aminopropyltriethoxysilane.
The solvent used for deposition of multiamino coupling agents also influenced their
structure. FTIR analyses [96] showed that deposition of N-3-aminoethyl-3-aminopropyltri-
methoxysilane (AAPS) and of trimethoxysilylpropyldiethylenetriamine (TPDT) in toluene
resulted in "completely" condensed coatings of coupling agent. In contrast, deposition from
dilute water produced incomplete cure, as indicated by the absorption at 940 cm-x, attributed
to the Si-O stretching mode of the uncured silanol. If the condensation reaction for the
SiOH groups is reversible, then the presence of water should "push" this equilibrium to
the reactants, SiOH, rather than to the condensed Si-O--Si groups that produce more water.
The absorbance of the Si-OH stretch decreased in intensity after the sample was heated,
C.L. Schutte / Environmental durability of glass-fiber composites 285

indicating condensation of the SiOH groups. The amount of coupling agent that deposited
on the surface of glass depended on its concentration in solution and on the coupling agent
itself: the surface coverage for AAPS was greater than that for TPDT at a given concentration.
The relative loss of AAPS from previously coated samples of silica powder (Cab-O-Sil) in
boiling water was greater for coupling agent whose degree of cure was less.
Early studies by Bascom [97] looked at the structure of silane coupling agents
(vinyl-, y-chloropropyl-, p-chlorophenylethyl-, and y-aminopropyltrialkoxysilane) deposited
from polar and non-polar solvents (cyclohexane, methyl ethyl ketone, 2-propanol, distilled
water, and a mixture of 2-propanol:water 1:1 v/v) by depositing them by dip-coating onto
surfaces of Pyrex, stainless steel, germanium and a mixture of the thallium salts of chloride
and bromide. After initial removal of the physisorbed fraction of the coupling agent by a
wash with acetone, measurement of the equilibrium advancing and receding contact angles
of water, formamide, thiodiglycol, methylene iodide, a-chloronapthalene, and bicyclohexyl
showed that, except for measurements with methylene iodide, a large hysteresis existed for
these films. This hysteresis indicated the presence of roughness of the surface or wetting
and swelling of the film. Although the substrate on which the coating lay had no significant
influence on these measurements, the time of contact of the surface with the solvent did,
possibly because the solvent swells the coating. Therefore, the coating and solvent were
pre-equilibrated for 20 h prior to the measurements of the contact angle.
The spectra of the coating on germanium or on the thallium salt, measured by Attenuated
Total Internal Reflectance Infrared Spectroscopy, showed the presence of the characteristic
functional groups in the coatings both before and after rinsing. Coatings deposited from
cyclohexane yielded spectra that were more intense than those deposited from the other
solvents. Instead of the formation of coatings of vinyltriethoxysilane and chloropropyltri-
methoxysilane on surfaces of stainless steel, scanning electron microscopic analyses revealed
spherical particulate material. The thicknesses of the films derived from y-chloropropyl-
trimethoxysilane, rinsed with acetone, and measured ellipsometrically, were in the range
of 2.5 to 4 nm for samples deposited from methyl ethyl ketone, 2-propanol, and the 2-
propanol:water mixture. The thickness of the film deposited from cyclohexane was significantly
thicker at 274.8 nm, even after the rinse with acetone. The solvent used for deposition
therefore had significant influence on the thickness of the chemisorbed fraction of the
coating.
The solvent used for deposition of the coupling agent influenced the orientation of
the coupling agent [98]; generally, coatings of y-aminopropylsilane that were deposited from
polar, non-aqueous solvents had accessible amine functional groups that were instantaneously
neutralized with HC1, while the amine groups of samples made from aqueous solution
required 2-3 min for complete neutralization.

4.3.5. Influence of the substrate


In order to better understand the role of both the substrate and the coupling agent
in the dissolution of polymer from the surfaces of substrates, Belton and Joshi [99] studied
the effect of the surface-treated substrate on the dissolution of poly(amic acid-imide) in
aqueous N a O H and KOH solutions. Their goal was to determine the efficacy of the polymer/
substrate interface to resist solutions useful for developing and processing photoresists for
applications in electronics. They found that, for the relative rate of loss of poly(amic acid-
imide), the ranking of substrates treated with y-APS, slowest to fastest, was as
A1 < SixNr < Si< SiOz. Substrates not treated with coupling agent showed no difference in
286 C.L. Schutte / Environmental durability of glass-fiber composites

their relative rate of loss of polymer, a rate that was significantly higher than for systems
whose substrates were previously treated with APS. Furthermore, the rate of loss of polymer
decreased with increasing content of APS.
The sample preparation in these experiments required thermal treatment at 140 °C
for 45 min; this treatment allowed silicon hydroxyl groups in the coupling agent to condense
to Si-O--Si, with the release of water, and increase in cross-link density. The relative amount
of coupling agent that was extractable in ambient water decreased after being heated to
140 °C for 45 min compared to a sample that had no prior thermal treatment. This result
was consistent with the silane condensing further after the thermal treatment. Chemical
evidence for this transformation would help to confirm the hypothesis. This trend in rate
of loss of polymeric coating from the substrate, from slowest to fastest, A1 < SixNy < Si < SIO2,
illustrated the relative adhesion of the substrate to the coupling agent. These researchers
also observed an effect from the thickness -- and therefore the distribution of phases --
of the coupling agent. The results of this work illustrated the added degree of complexity
with regard to the durability of the interface. The substrate can have an influence on the
durability; however, the exact details of the mechanism responsible for this influence need
further study.

4.4. Acid~base interactions

Fowkes and co-workers theorized that acid/base interactions were responsible for the
adhesion between substrate and polymer [100-103]. O n e study involved the adhesion of
polymers to glass whose surface was high in p H because of the content of sodium in the
glass. Although they could not loosen acidic polymer, chlorinated poly(vinylchloride), CPVC,
from the surface of the glass in water, they could remove the basic polymer poly(methyl
methacrylate), PMMA, u n d e r the same conditions. W h e n the glass was previously washed
with HCI to acidify the surface, the trend reversed: CPVC could be removed and P M M A
could not [104]. Fowkes extended his work to the modification of surfaces of glass to
promote adhesion to polymers. Measurement of the surface acidity/basicity for E-glass fibers
involved measuring the molar heats of adsorption of phenol using flow microcalorimetry.
These results showed that the surface of E-glass was basic.
Moreover, titration of the surface with pyridine showed that the surface contained
three times as many basic sites as acidic ones. These basic sites were three times as strong,
as measured by heats of adsorption of phenol, as the acidic sites measured by adsorption
of pyridine. Using Drago's approach to estimating heats of mixing of acids and bases [105],
Fowkes et al. [106] estimated the electrostatic contribution of the basic sites, EB, for
the heat of mixing and that for the covalent contribution, CB, to be 0.4 and 0.8 (kJ
mol-1)1/2, respectively. Likewise, they estimate the acidic components for electrostatic and
covalent contributions to be EA=0.3 and CA=0.4 (kJ mol-1) 1/2. Treatment of E-glass with
coupling agents containing aminosilane and methacryloxysilane resulted in surfaces that
contained a lesser n u m b e r of accessible basic sites but at twice the strength compared to
those sites on untreated E-glass.
Fowkes suggested that there was incomplete coverage by the coupling agents on the
glass, resulting in a smaller number of stronger, accessible sites. Because the observed
strength of the amino groups in the silane was not greater than that for the accessible
sites on the glass, Fowkes suggested that the amine reacted with carbon dioxide to form
less basic a m m o n i u m carbonate salts at the surface. Characterization of the surface of the
C.L. Schutte / Environmental durability of glass-fiber composites 287

glass with the aminosiloxane by angle-dependent X-ray photoelectron spectroscopy showed


both a broader peak in the carbon ls region, 285 eV, and the presence of a shoulder at
higher energy than the ls peak for nitrogen, approximately 400 and 402 eV. These results
were consistent with the presence of an a m m o n i u m carbonate species in the coating. Results
of Fowkes and co-workers illustrate the importance of acid/base interaction between the
surface of the glass and the resin. The optimization of these types of interactions for
improved adhesion, particularly in the presence of water, may be difficult, however, as the
water can potentially displace components in an acid/base pair.

4.5. The influence of the coupling agent on the properties of the interfacial region

The chemistry and morphology of the coating derived from the coupling agent can
significantly influence its strength and durability. Studies of durability as a function of both
degree of siloxane condensation and of order, as influenced by chemical structure, as well
as the n u m b e r of siloxane attachments in the chemical structure of the coupling agent are
addressed. Studies of the effect of morphology of the interracial region on its mechanical
properties and durability included their mechanical characterization through peel tests and
morphological characterization in terms of interdiffusion of the coupling agent with the
resin as a function of cure temperature of the coating of coupling agent.

4.5.1. Effect of chemistry on the strength and the durability of the interface
Work by Ishida and Koenig using F T I R added insight into the hydrolytic stability of
coupling agents based on their chemical nature as deposited from aqueous solutions. Ishida
and Koenig [107,108] looked at silane coupling agents on E-glass fibers using F T I R by
following the loss of y-methylacryloxypropyltrimethoxysilane (MPS), vinyltrimethoxysilane
(VTMS), and cyclohexyltrimethoxysilane (CHTS) in water at 80 °C. The desorption of MPS
followed an exponential decay and leveled off at a coverage of approximately 20 molecules
nm 2. VTMS differed by having no significant loss of coupling agent until approximately 600
h, after which the concentration decreased linearly to a surface coverage of 100 molecules
n m z. CHTS appeared relatively stable, with no significant loss up through 5500 h exposure
to water at 80 °C. This higher degree of stability [109] for CHTS was consistent with the
high degree of order that CHTS can achieve upon adsorption onto the surface of the glass.
The authors attributed differences in the relative rates of loss of the coupling agents
MPS and VTMS to their relative degree of Si-O-Si condensation. After a sample of MPS
was heated to 150 °C for 1 h, the magnitude of the absorbance for the SiOH group decreased
to 32_+2% of its initial intensity, while that for VTMS decreased to 12_+2%. The lower
degree of condensation of the MPS after thermal treatment appeared to correlate with a
higher concentration of cyclic oligomers in the solution of MPS compared to that of VTMS.
Furthermore, the chemical nature of the surface treatments changed: the MPS coatings
exhibited a shift in the absorbance of the carbonyl group, from 1720 to 1740 cm -1, with
increased time in water at 80 °C as the a,/3 unsaturated carbonyl transformed to a saturated
(polymerized) carbonyl moiety. There was also a corresponding decrease in intensity for
the C - - C unsaturation at 1638 cm -1. T h e absorption at 870 cm -x, attributed to the presence
of SiOH groups, increased with time in water at 80 °C for the VTMS coating, indicating
hydrolysis of the Si-O-Si functional groups. Surprisingly there was little relative change in
the absorption for the terminal C = C unsaturation at 1412 cm -1, indicating that the vinyl
groups in VTMS were more stable toward polymerization than those of MPS in water at
288 C.L. Schutte / Environmental durability of glass-fiber composites

80 °C. Also, further evidence for hydrolysis of Si-O-Si groups included an increase in the
absorbance for the SiOH functional groups, at 900-920 cm-1, with a corresponding decrease
in the Si-O--Si absorbance at 1150-1000 cm -1. In general, these trends were consistent
with the three-phase morphology found by Johannson [77] and Schrader [79-81].
The chemical structure of the coupling agent had significant influence on the hydrolytic
stability of a flat plate of silica glass/poly(2,2-bis-((4-methacryloxy-2-hydroxy-propyl)-phenyl)-
propane), referred to as poly(BIS-GMA), after immersion in boiling water. Nishiyama et
al. [110] coated three different coupling agents onto a fiat silica plate. The coupling agents
contained one, two, and three silicon units respectively; they were: y-methacryloxypropyl-
triethoxysilane (y-MPS), 1,4-bis(2-(triethoxysilyl-propyl-carbamoyloxy)-l-methacryloxyprop-
oxy)-butane (BSMB), and 1-(tri-(2-(triethoxysilyl-propyl-carbamoyloxy)-l-methacryloxyprop-
oxy-methyl))-propane (TSMP).
Samples consisting of coatings of coupling agent deposited on silica glass and adhered
to poly(BIS-GMA) initially showed cohesive failure after immersion in boiling water. Interfacial
failure occurred in the sample coated with 3,-MPS and TSMP after immersion for 60 h,
and after 80 h for that coated with BSMB. The tensile bonding strength was higher for
the samples treated with BSMB, the bifunctional silane. Furthermore, the time of onset
for interracial failure was consistently later for the sample made with BSMB in a wider
range of concentrations of solutions for the initial deposition compared to that for either
the 3,-MPS or the TSMP. These results suggest that the chemical structure of the coupling
agent significantly influenced the deposition behavior from solution, the structure, the
adhesive strength, the hydrolytic stability of the coating, and the transition between cohesive
(in the coating) and interracial (at the coating/glass or coating/polymer interface) failure.

4.5.2. Effect of morphology, degree of interdiffusion, on the strength and durability


Both the type of coupling agent and its concentration in the solution of deposition
influence the properties of the interracial region. Furthermore, both the degree of network
formation within the coating of coupling agent and the degree of interdiffusion of the
coupling agent with the resin influence the mechanical performance and durability of both
fiber/matrix interfaces and flat plate/resin adhesive joints. The ability to predict the optimum
curing temperature of these types of coatings, prior to application of the polymer, appears
to depend on the specific coupling agent, and therefore no general recommendations are
possible at this time. The examples cited in this section, however, serve to illustrate how
important the understanding of the relationship of chemical structure, degree of network
formation, degrde of cure prior to processing with the resin, degree of interdiffusion with
the resin, and the interdependence of these issues is for the coupling agent. Once the
fundamental structure/property relationships are known for these types of coatings, the task
of tailoring or engineering the interracial region for both optimal strength and durability
characteristics can be addressed.
Suzuki et al. [111] showed that the Mode I fracture toughness, as measured by double
cantilever beam, was significantly different for a sample prepared using a methacrylate
functional coupling agent, ~/-methacryloxypropyltrimethoxysilane (MPS), coated on E-glass
fabric and tested in a vinyl ester resin, compared to one containing an epoxide functional
group, y-glycidoxypropyltrimethoxysilane (GPS). They coated fibers from solutions of MPS
in water of pH 4 at concentrations of 0.01, 0.4, and 1.0 wt.%, and from a solution of GPS
at 0.4 wt.%. Samples that had been coated with the methacrylate functional coupling agent
from solutions of 0.4 and 1.0 wt.% had the strongest tensile (373, 371 MPa for 0.4 and
C.L. Schutte / Environmental durability of glass-fiber composites 289

1.0 wt.%, respectively), flexural (447, 460 MPa), and shear (39.4, 47.8 MPa) properties.
Monitoring by FTIR [112] the functional groups that are unique to the coupling agent
MPS (1720 cm -1, C = O in the coupling agent) and the resin (1430-1462 cm -1, methylene
from the vinyl ester resin), for the sample coated from the 1% solution of coupling agent,
revealed that the coupling agent migrated into the resin. Furthermore, the FTIR evidence
for this system suggested that the curing of the resin was influenced by the presence of
the coupling agent and the glass fibers; the methylene units, indicative of uncured resin,
were at a concentration that was measurable within 200 /~m of the surface of the fibers.
The presence of the coupling agent in the interphase therefore appeared to have a significant
influence on the development of the interphase.
An increase in the temperature of thermal pretreatment of the coupling agent increased
the cross-link density of the network [113,114]. This formation of the 3-D structure of the
coupling agent increased the strength of its adhesive bond, provided that the coupling agent
was not completely cross-linked initially. The rationale for this observation was that a
minimum amount of cross-linking was necessary for the coupling agent to have an initial
degree of strength. If the sample cross-linked completely, however, no intermixing of the
coupling agent and resin could occur. It appeared that this intermixing, or possible formation
of an interpenetrating network, was necessary for optimum adhesion.
Sung and co-workers looked at the effect of thickness of coupling agent, and of thermal
history (for the coupling agent alone and the coupling agent in the presence of resin) on
peel strength of A1203/y-APS/polyethylene [113] and AlzO3/y-APS/poly(caprolactam) [114].
For the polyethylene system, a maximum peel strength resulted when the substrate was
coated with a 2.0 wt.% solution of APS, yielding a coating with an approximate thickness
of 100 nm. Furthermore, when the coating of APS experienced no thermal treatment prior
to being coated with unoxidized polyethylene, the strength of the joint was greatest. Samples
whose coatings of coupling agents were not previously heated failed in a cohesive mode,
while those that experienced a thermal treatment prior to the application of polyethylene
experienced a mixed mode of failure. Sung and coworkers observed interdiffusion of the
silane with polyethylene after the samples were heated. The degree of interdiffusion between
these components was greatest when the APS coating was not previously heated. This result
suggested that the degree of interdiffusion significantly influenced the strength of the joint.
For the poly(caprolactam) study, the maximum peel strength existed in samples whose
coating of coupling agent was previously dried at 25 °C for 5 days and 55 °C for 1 day,
compared with those dried at 55 °C for 5 days [114]. These results, coupled with those
from XPS, suggested that interdiffusion of the polymer and coupling agent created a strong
bond.
Chaudhury et al. [115] also looked at the influence of the degree of thermal pretreatment
on the adhesive strength of N-(2-aminoethyl-3-aminopropyl trimethoxy silane) with poly(vinyl
chloride). They dried a series of coatings of coupling agent between temperatures of 25
to 175 °C, for 15 min, adhered the layer of coupling agent to poly(vinyl chloride), and
measured the adhesive strength using a peel test. For their system, the optimum temperature
of thermal pretreatment was 50 °C. Apparently, at temperatures below 50 °C the network
of the coupling agent was not formed adequately, and above 50 °C the network that formed
did not allow for sufficient interdiffusion with poly(vinyl chloride). Exposure to water involved
immersion of the samples in distilled and deionized water at room temperature for three
days. The wet adhesive strength decreased for samples that had been thermally pretreated
between 25 and 75 °C, and, above 75 °C, was not measurably different from the low value
290 C.L. Schutte / Environmental durability of glass-fiber composites

for dry samples. The peel force as a function of drying temperature also had a maximum
at 50 °C.
Chaudhury and co-workers looked at the depth profile of silicon (from the coupling
agent) into poly(vinyl chloride) using XPS. They sputtered away the surface of the sample
and analyzed the composition at different depths, up to approximately 600 nm. The composition
of silicon for the sample whose coupling agent was cured at 25 °C had a higher concentration
at deeper levels into the sample than did a sample that was cross-linked at 175 °C. This
result was consistent with the hypothesis that interdiffusion occurred and allowed for the
formation of a morphology suggesting an interpenetrating network.
Apparently, the optimum cure conditions for the coupling agent prior to application
of the polymer were system-dependent; the result of Chaudhury et al. [115], optimum
thermal pretreatment at 50 °C, was different from those of Sung et al. [113], who found
optimum conditions of adhesion after drying their coupling agent at room temperature only.
The durability of the sapphire/coupling agent/polymer system appeared to be dependent
on the interdiffusion and cross-link density of the coupling agent/polymer interphase. At
a lower cross-link density, Chaudhury et al. hypothesized that the coupling agent was more
hydrophilic and more susceptible to attack by moisture. Looking at the development of the
morphology of the coupling agent/polymer interracial region in this manner has the advantage
of relating these structure-property relationships to the chemistry and degree of chemical
reaction, Si-O-Si network formation, in the coupling agent. This type of approach should
allow for the optimization of future tailored interracial regions with respect to interracial
strength and durability.
Earlier, Schrader and Block [81] showed that, for samples whose surfaces were treated
with dilute solutions of previously hydrolyzed coupling agent, the life of the joint increased
from about 550 to 1100 h with increased solution concentration of APS. The resulting range
of coverage of these dilute monolayers of APS was 1 to 6 monomers per nm 2, most likely
within the monolayer-equivalent phase. These data on the lifetime of the joint were consistent
with the reaction between the glass/resin interphase of the sample and water being coupling-
agent concentration rate-limited, at least within this range of content of coupling agent.
Their findings of a correlation between an increased life of the joint for samples with the
largest lateral concentration of silane coupling agent at the surface would support a model
where the development of the monolayer-equivalent phase would increase the lifetime of
the interracial region.
For samples that were prepared by selective desorption of the coupling agent from
the surface of the glass, initially the life of the joint increased from approximately 50 to
1050 h with an increase of content of APS from 0 to 24 monomers nm -2. Beyond this
concentration, however, there was no further influence of the concentration of APS on the
time to failure of the joint. Schrader [81] commented that these results suggested mass
diffusion of water contributing towards the rate-limiting step for the hydrolysis. They did
not take into account the influences of the chemical and morphological nature of the phases
present after the preferential desorption of coupling agent. Ease of removal of the first
fraction was consistent with the removed fraction being the physisorbed fraction. Selective
desorption under these conditions would leave the monolayer-equivalent and chemisorbed
phase on the sample. Differences in behavior towards durability between these phases could
account for the initial improvement as the monolayer formed; however, as the chemisorbed
phase developed in the thicker layers, differences in the network formation may have
changed the governing factors with respect to attack by water. Based on their experimental
C.L. Schutte / Environmental durability of glass-fiber composites 291

protocol, the chemisorbed coupling agent that Schrader and Block referred to is most likely
the monolayer and chemisorbed phases as they are defined in the present paper.
Schrader and Block [81] studied the wet-strength adhesion of a sample of Pyrex glass/
epoxy/Pyrex glass that had "tailored" surface treatments of APS coupling agent, 14C-labeled
3,-aminopropyltriethoxysilane that was labeled at the carbon alpha to the amine,
(CH3CH20)3SiCH2CH2C*H2NH2, applied from aqueous solutions. Their results showed that
the presence or absence of the first fraction or physisorbed phase did not have significant
influence on the wet strength of the joint under a load of 50 lbs. If there was an excessive
amount of the physisorbed phase, however, this presence accelerated the rate of loss of
bond strength. They found that the larger the proportions of chemisorbed and "monolayer"
components relative to the physisorbed phase, the greater was the life of the joint. Furthermore,
they found that the debonding involved cohesive failure within the coupling agent: radioactivity
existed on the surfaces of both the glass and the resin.
Schrader and Block [81] also showed that the application of dilute solutions of coupling
agent, APS, in benzene formed the equivalent of two monolayers of silane and resulted
in joint lives that were longer than those derived from aqueous solutions with subsequent
selective desorption of coupling agents. The solution used for deposition of the coupling
agent significantly influenced both the morphology of the coupling agent and the performance
of that interface.
Graf et al. [116] adsorbed MPS onto the surface of E-glass fibers that they subsequently
incorporated into a bisphenol-A-fumarate polyester resin. To examine the influence of
removing the physisorbed layer from the coating vs. the presence of all three fractions,
they desorbed that fraction from selected fibers by washing them with tetrahydrofuran
(THF) and incorporated these fibers into polyester composite samples as well. The flexural
strength of the composites increased significantly after they were treated with coupling
agent. There was a slight maximum in the flexural strength of dry samples at a coverage
of MPS of 0.4 wt.%. Samples that were previously washed with T H F to remove the
physisorbed fraction of MPS had flexural strengths that were slightly higher than those
with unwashed fibers for surface coverages above 0.4 wt.%. For samples that were treated
in water at 80 °C for 48 h, the wet strength was greater than that for dry samples. It was
unclear if the resin, the fibers, or the interface governed the properties of the composite
under these wet conditions.

4.6. The use of single-fiber tests to characterize the fiber~matrix interracial environmental
durability

Single-fiber tests [117,118] have been useful as model systems for studying the fiber
and the fiber/matrix interface. Recent work by Baillie and Buxton [119] presented several
of the experimental issues that must be considered when using single-fiber tests for studies
of durability. This section summarizes results of studies of durability of interfaces using
these tests. Although this review focuses on durability of glass fibers, where studies employing
carbon fibers yield important and useful results that are relevant when considering glass-
fiber systems, they are included. This section presents both the single-fiber fragmentation
and the single-fiber pull-out tests.

4.6.1. Single-fiber fragmentation test


The single-fiber fragmentation test involves applying a tensile load to a dog-bone-
shaped specimen of solid resin containing a single fiber. As Fig. 4 illustrates, this applied
292 C.L. Schutte / Environmental durability of glass-fiber composites

. . . . . . . . . . . . . .

Fig. 4. Illustration of the fragmentation of the fiber during the single-fiber fragmentation test. Each dog-bone, top to bottom,
represents a sample under an increasing tensile stress (by permission of the publishers, Butterworth Heinemann Ltd.).

stress causes the fiber to fragment when the transfer of stress to the fiber through the
interface reached the breaking stress of the fiber. As the level of applied stress increases,
the fragmentation continues until the applied stress reached a certain level, mechanical
saturation, where the fiber fragments are so small that the interface can not transfer sufficient
stress to fracture the fiber further.
The single-fiber fragmentation test is one of many experiments designed to determine
the interracial strength between a fiber and resin [120--123]. Because of difficulties in
interpreting experimental results, measuring this interfacial strength unambiguously in bulk
composites is difficult. The focus of the single-fiber fragmentation test on a simple geometry,
that of a single fiber in a dog-bone specimen, has the potential to make the study of the
interracial properties more straightforward than in a bulk composite. Fig. 5 presents a
typical experimental setup for the single-fiber fragmentation test and Fig. 4 illustrates the
behavior of the sample undergoing testing. Kelly and Tyson [124] were the first to use this
approach with an analysis relating the shear strength of the interracial region to the strength
of the fiber and the average critical length of the fiber fragments, after achieving mechanical
saturation for tungsten and molybdenum wires in copper matrices. For a system exhibiting
plastic deformation of the resin and elastic deformation of the fiber, they related the effective
interracial shear strength or transferability ~- to the strength of the fiber, trf, and the average
critical length/diameter ratio of the fiber fragment, LJD:

L~ 1 trt
D 2 ~- (4)

Recently DiBenedetto [125] reviewed the study of thermomechanical stability of interfaces


using this Embedded Single-Fiber Test. He emphasized that, in order to quantify the value
of interracial shear transferability, or strength, from this type of test, the mode or modes
of failure must be known [126-128]. Rather than analyzing his data by the traditional
Kelly-Tyson approach [124], DiBenedetto and coworkers modified Piggott's [129] approach
of measuring the energy required to decouple the fiber from the interface using a fracture
criterion. Although the Lc of the fiber at mechanical saturation increased between temperatures
of testing between 25 and 175 °C, and this increase in Lc was also a function of surface
treatment, for surface-treated glass fibers in poly(caprolactam) [130] and in epoxy [131],
the energy of debonding decreased only slightly. The systems exhibited strong adhesion
between the fiber and resin, and according to DiBenedetto these results implied that the
mode of failure was plane-strain brittle fracture at the interface. For a moderately strong
C.L. Schutte / Environmental durability of glass-fiber composites 293

l
9

1 I\ 5

1 TV Camera 5 Microscope
2 TV Monitor 6 Single Figer Specimen
3 Loading Device 7 Linear Vm-iabre Differential Transformer
4 XY Translation Stage 8 Data Acquisition System
9 Displacement Gage

Fig. 5. Illustration of equipment used for the single-fiber fragmentation test (by permission of the publishers, Butterworth
H e i n e m a n n Ltd.).

interface determined by the fracture energy criterion, such as in a glass fiber/high density
polyethylene system [130], they observed slippage of the fiber.
Wagner and Lustinger [132] also used an energy-balanced approach to look at the
durability of interfaces of glass fiber/epoxy samples using the single-fiber fragmentation test
on exposure to water at 95 °C for up to 336 h. They quantified the interfacial energy by
equating the energy to create new surfaces in the fiber (fracture) and in the interface
(debonding) to that from both the strain-energy release in the region of the fiber that
became debonded and from the length of transfer of stress after the fracture of the fiber
occurred. This approach resulted in the following relationship:
2yi, = A + (B/Ld) (5)
where
A = (rtrf*2)/4E, (6)
and

(7)

where 2yif is the interfacial energy, Ld the length of the debond, r the radius of the fiber,
of* the stress at which the fiber breaks and interfacial debonding occurs, Ef the Young's
modulus of the fiber, ~ the length of transfer of stress, and ~/f the surface energy of the
294 C.L. Schutte / Environmental durability of glass-fiber composites

fiber. Fig. 6 illustrates the decrease in both the strength of the fiber, o-, and the interfacial
energy as a function of time of exposure to water at 95 °C, as determined by this approach.
The use of data generated by the single-fiber fragmentation test for their epoxy system,
acquired at stresses below those required to achieve mechanical saturation, illustrated the
potential of this approach, particularly where mechanical saturation of the sample was
neither reproducible nor achievable. Where brittle resins do not allow for the achievement
of mechanical saturation in this test, the ability to use the data prior to mechanical saturation
in this analysis significantly increases the number of material systems that potentially could
be characterized for durability studies. The debonded length increased as a function of
time of exposure to water at 95 °C, and this reduction in the interracial area results in an
apparent decrease in the interracial surface energy. Quantifying the interfacial surface energy
in this way allowed for further understanding the transition in the mode of fracture [133],
from the formation of cracks in the matrix of a well-bonded system to one where interfacial
debonding occurred from degradation by exposure to moisture.
Schadler and co-workers also measured ~ at low levels of strain from the lengths of
the stress-transfer zones in the photoelastic patterns, termed the lambda technique [134],
to study cyclic fatigue in single-fiber composites [135,136]. Their model included the exchange
of strain energy between a failed fiber and the matrix as well as visco-elastic relaxation of
the interphase. In order to probe the effect of cyclic loading, they studied single-fiber
specimens of carbon fiber in polycarbonate in tension under both axial and transverse
loading. In axial loading no damage occurred at the strains below the strain to failure of
the fiber. At strains high enough for breakage of the fiber, the fatigue behavior under cyclic
tension appeared to be dominated by the matrix. In cyclic loading of the single-fiber sample
with transverse orientation, they found that the interracial properties were dependent on
(a) the rate of strain. At the slower strains, under 0.005% strain s -I, the damage to the
interface was the greatest.

600 2.0 f T 1

0 Sigma (GPa)
5OO
• Interfacial Energy ( J / m 2)
&-- 1.5
E
4O0 0
3
8_
o ©
c
LIJ
300 0
1.0
E
5 C~
5 b~
13
2OO
"6 0.5
1 O0
• ©
©
o 0.0 t ± J"O--O--
0 IO0 2O0 300 400

Hours
Fig. 6. Plot of the calculated strength of the fiber, ~r, and interfacial energy as a function of time of exposure (h) to water
at 95 *C [132] (by permission of the publishers, Butterworth Heinemann Ltd.).
C.L. Schutte / Environmental durability of glass-fiber composites 295

(b) the amplitude of the strain. For a maximum strain amplitude of 2.0%, the sample
showed minimal damage, while at 2.70% a loss of ability to transfer load to the fiber
occurred (failure of the interphase occurred at a finite distance from the fiber/matrix
interface, suggesting that this was the location where the radial stress was at a maximum).
(c) the process of unloading and/or cycling of the strain. Tests in creep showed that, even
up to 2.8% strain, the interphase remained intact. The observed damage under conditions
of fatigue was not a function of the surface treatment of the fiber -- hence the interface.
DiBenedetto et al. [137] studied thermal and hydrothermal stability in model single-
fiber composites of treated fibers of E-glass in an epoxy matrix. Conditions for the environmental
exposure consisted of submerging the samples in water at 25, 50, or 75 °C during the
fragmentation test or boiling the samples for 2 to 24 h prior to performing the fragmentation
test. They used several coupling agents for coating the fibers of E-glass, and concluded
from measurement of transmission of shear (a function that is inversely d e p e n d e n t on Lc/
D) that 3-(phenyamino)propyltrimethoxy silane (PAPS), and 3-(4-hydroxy-3-methoxy-
phenyl)propyltrimethoxysilane (NMPS) exhibited stability at high temperatures, and 3-
(methylamino)propyltrimethoxysilane (MAPS), 3-(aminoethylamino)propyltrimethoxysilane
(AAPS) and bis(3-trimethoxysilylpropyl)amine (BTA) showed improved stability in moist
environments. The improvement in transmission of shear as a function of coating decreased
in the following series: NMPS > PAPS > BTA > MAPS > AAPS > APS > no coating.
In contrast to the results for the m e a s u r e m e n t of transmission of shear, both the
transmission of force to the fiber and the m i n i m u m energy release, energy of debonding
per unit area, in the range of those for epoxy materials, were insensitive to the types of
coating for the fiber. DiBenedetto et al. concluded that these results were consistent with
failure occurring in the matrix near the interface. Possibly the time of exposure was too
short to observe failure in the interracial region instead of in the resin. Their results showed
that exposure of the single-fiber composite to boiling water resulted in: (a) an increase in
the critical length of the fibers for all conditions of coating after a 2-h exposure, with fibers
coated with B T A and AAPS being the most resistant, (b) recovery of the initial properties,
after drying, of the uncoated samples and of those coated with BTA and MAPS, but
permanent, non-recoverable damage after drying of the sample for those with coatings of
APS and NMPS, (c) slippage of the fiber during the fiber fragmentation test in all undried
samples after being boiled for 24 hr, and (d) complete recovery for the uncoated sample
and the sample with a coating of MAPS after being boiled for 24 h with subsequent drying,
but irreversible damage (% reduction of the shear strength transmission) for samples coated
with BTA, 11%; PAPS, 57%; NMPS, 19%; APS, 15%; and AAPS, 34%. These results of
DiBenedetto et al. illustrate that, although the choice of coupling agent can influence the
transmission of shear at the interface, the region of failure under these conditions of
exposure was not affected, because the loci of failure appeared to be within the matrix.
Because it was unclear under what conditions these samples reach saturation with moisture,
possibly exposure of samples to water for longer periods of time may degrade the interfacial
region and change the region of failure to the interfacial region instead of the epoxy.
Gomez and Kilgour [138] illustrated that thermal exposure influenced the properties
at the interface by measuring the critical length and interfacial shear strength for an E-
glass/epoxy system at 25, 50 and 75 °C. The critical lengths increased with the increase in
temperature for fibers coated with 3-aminopropyltriethoxysilane ( A P S ) > bis-(3-trimethoxy-
silylpropyl)amine (BTA)>11-aminoundecyltrimethoxysilane (11-AUTMS). The critical
lengths, combined with the measurement of the strength of the glass fibers, yielded the
296 C.L. Schutte / Environmental durability of glass-fiber composites

interracial shear stresses; these decreased with increasing t e m p e r a t u r e for all three types
of samples. T h e relative values of the interracial strength decreased as B T A > l l -
A U T M S > APS. T h e different coupling agents studied also influenced the strength of the
glass, possibly through differences in protection from abrasion.
Schutte and coworkers [139,140] used the single-fiber fragmentation test to study
durability of fibers o f E-glass and of interfaces in an epoxy matrix on exposure to both 65
and 75 °C distilled water. T h e y found evidence that both the strength of the fiber and the
interfacial strength d e g r a d e d on exposure to water as the content of water increased in
the sample. Because the Lc increased as a function of time of exposure, it a p p e a r e d that
the relative rate of degradation of the interface was greater than that of the strength of
the fiber. Fig. 7 illustrates this t r e n d in increasing critical length with time of exposure to
hydrothermal conditions while the Lc remains the same for those that were thermally treated
only. Evidence that the fiber d e g r a d e d included the detection of ions unique to the E-glass
in the distilled water, using Inductively Coupled Plasma Mass Spectrometry. Furthermore,
the fiber f r a g m e n t e d at a lower applied strain with increased time in the w a r m water. This
result implied that the level of stress at which the fiber fragmented (applied stress plus
swelling stress) for the hydrothermally treated fibers was lower -- the apparent strength
o f the fiber was less -- t h a n for the dry sample. This work therefore showed that, for glass
fibers in resin, the degradation of both the fiber and the interfacial region needs to be
considered.
Drzal et al. [141] showed that reversible changes to the tensile modulus and Tg occur
to the high~Tg epoxy resin derived from diglycidyl ether bisphenol-A/metaphenylene diamine
( M P D A ) 100/14.5 w/w (Tg, 166 °C), w h e n the sample did not go above its plasticized Tg.
W h e n a low-Tg epoxy ( e p o x y / M P D A 100/7.5 w/w; Tg, 62 °C) is immersed in water at a
t e m p e r a t u r e above its Tg, however, the dry tensile properties were non-recoverable. Drzal
et al. described the interphase as a region that is m o r e brittle and has a lower Tg than
the bulk matrix [142], and they translated these effects on the resins into changes in the

70

60
~ -~drothermal
ormal

50
o
o 8
a
4O o
o
o 0
n o o o°
3O
0 o o
o o
20 8 •
o8 •

10

0
0 2000 4.000 6000
Time of Exposure (hrs)
Fig. 7. Plot of the averageLJD as a functionof time of exposurefor samplesof E-glassin epoxy[139]: @, thermalconditions
(75 °C ambient air); O, hydrothermalconditions (75 *C distilled water). (By permission of the publishers, Butterworth
Heinemann Ltd.)
C.L. Schutte / Environmental durability of glass-fiber composites 297

behavior of the interface/interphase on exposure to water. They hypothesized that plasticization


of the resin increased the critical lengths of carbon fibers as determined in the single-fiber
fragmentation test after exposure to water at 100% R H at 20, 75 and 125 °C. This hypothesis
implies that yield of the matrix at the interfacial region governs the interfacial shear strength.
Furthermore, the critical lengths were longer, the interfacial shear strength smaller, for the
carbon fiber that was "treated" and "finished" with an uncured epoxy compared to the
carbon fiber that was "treated" but did not contain an epoxy finish. Drzal and co-workers
indicated that the finish of uncured epoxy on the fiber behaved similarly to epoxy with a
lower content of curing agent; they explained that this interphase should have a lower Tg
and exhibits an irreversible change after being plasticized by hot water at a lower temperature
compared to an epoxy with a higher Tg that experiences reversible effects. In general, for
fibers containing the coating of uncured epoxy, the critical lengths were shorter but did
not recover their original dry critical lengths after dehydradation compared with those with
no epoxy finish coating. Apparently, the presence of the uncured epoxy phase influenced
both the adhesion of the fiber to the matrix and the behavior after exposure to water.
Here, Drzal et al. illustrated the influence of "finish" coatings on the properties of the
matrix and, ultimately, this interaction. The interaction between the coupling agent's chemistry
and morphology on the properties of the matrix in the proposed interphasal region is also
an area, as the work of Drzal and co-workers illustrates, that needs further work.
Drzal [143] used the fragmentation test to examine the effect of properties of the
matrix and residual stresses induced by processing on the maximum shear strength at the
fiber-matrix interface. He found that the interfacial shear strength decreased with decreasing
stiffness of the matrix. He made this conclusion based on results of measurements using
carbon fibers in resins whose compliance was varied by changing the molecular weight of
the diamine added to cure either the diepoxy diglycidyl ether of bisphenol-A, DGEBA, or
to the tetraepoxytetraglycidylmethylenedianiline, T G M D A . He concluded that such effects
influenced the ability of the resin to transfer stress into the fiber. These changes are due
either to a lower interracial strength or a lower yield of the resin.
There are few examples of comparisons between results of tests on single-fiber and
macroscopic composites. One study involved work by Dwight et al. [144], who treated
surfaces of E-glass (a) with coupling agents directly prior to incorporation into either epoxy
or polyester resins, (b) with coupling agents previously incorporated into resin, (c) with a
silicone to inhibit bonding between resin and fiber, and (d) with no coupling agent. The
coupling agents that they used were an aminopropylsilane for epoxy and a methacryloxy-
silane for polyester resins. They measured interfacial shear strength from the single-fiber
fragmentation test [143,144], the short-beam shear strength from a three-point bend, and
examined the residual plastic zone, by measuring the area of the region exhibiting birefringence,
using optical microscopy under polarized light. For the epoxy system, compared to the
untreated E-glass fiber, the pretreatment of the fiber with the aminopropylsilane resulted
in a significant increase in the interfacial shear strength, a minor improvement in the short-
beam shear strength, and an increase in the residual plastic zone. Dwight et al. attributed
the mode of failure for the three-point test in the epoxy system as a combination of shear
and microbuckling for the sample that had pretreated fibers, and as "pure bending" for
those samples with untreated fibers.
For the single-fiber samples containing the brittle polyester, the single-fiber fragmentation
test had to be performed at elevated temperatures where the resin was more ductile. The
authors saw no significant improvement in the interfacial shear strength between treated
298 C.L. Schutte /Environmental durability of glass-fiber composites

and untreated fibers. In contrast to the single-fiber fragmentation test results, they reported
a significant increase in both the short-beam shear strength and the residual plastic zone
for samples that had pretreated glass fibers compared to those whose fibers were untreated.
It was unclear whether testing of the polyester samples under three-point bending occurred
at the elevated temperature as well. The magnitude of the differences in the effect of
interracial treatments may change as a function of temperature, and this explanation may
be the reason for these anomalous results. According to Dwight et al., the mode of failure
for the polyester composite containing pretreated fibers was pure shear, and that for samples
containing untreated fibers involved internal slippage of the fibers. Possibly these observations
imply that matrix failure occurred for the short-beam shear test with previously treated
fibers and interracial failure for those with untreated fibers. He also reported that samples
with silicone-coated fibers experienced pure internal slippage, possibly interracial or interphasal
failure, for composites of both resins under three-point bending.

4.6.2. Single-fiber pull-out test


The single-fiber pull-out test [117] measures the force required to extract a known
length of fiber from the resin. Fig. 8 illustrates a typical experimental setup for the microdrop
pull-out test. The force F required to pull out the fiber is a function of the radius r and
embedded length l of the fiber and the effective interracial shear strength r:
F = 2rrdr (8)

6
2

t
3 t
I
, Iv
A,I

1 LOAD SENSING DEVICE


2 MICROVISE
3 FIBER WITH DROPLETS
4 TRANSLATION STAGE
5 MOTORIZED ACTUATOR
6 ALUMINUM TAB
7 DATA ACQUISITION SYSTEM
Fig. 8. Illustration of equipment used for the single-fiber pull-off test.
C.L. Schutte / Environmental durability of glass-fiber composites 299

E m a d i p o u r et al. [145] looked at how the morphology influences both the interfacial
strength and the durability of that strength after immersion of the samples in water. They
showed that the use of N-2-aminoethyl-3-aminopropyltrimethoxysilane (AAPS) resulted in
the highest interracial shear strength compared to the use of 3-aminopropyltriethoxysilane
(APS) or N-methylaminopropyltriethoxysilane (MAPS) when coated on a single fiber of E-
glass e m b e d d e d in an epoxy matrix (1:1 w/w epoxy: nadic methylanhydride and accelerated
with N,N-benzyl dimethylamine). Samples whose fibers were coated from a solution of 1
wt.% coupling agent had interracial shear strengths of 189, 217, 221, and 2 7 7 x 1 0 -4 kg
m -2 for no coating, MAPS, APS and AAPS, respectively. Increasing the content of coupling
agent on the fiber by coating from a 5% solution, instead of a 1% solution, reduced the
shear strength at the interface for all types of coatings. These values were 125, 157, and
168 × 10 -4 kg m -2 for coatings of MAPS, APS and AAPS. These lower values in interracial
properties reflected the dramatic effect that changes in the concentration and the three-
phase morphology of the coupling agent can have on the properties and durability of the
interface. Possibly, the weaker c o m p o n e n t in the coating, the physisorbed phase, was thicker
in the coatings made from the higher concentration of coupling agent, and, because of its
weakness, it resulted in complete pull-out of the fiber with no improvement of adhesion
over that for an uncoated fiber.
Later experiments comparing the measured interfacial shear strength with the con-
centration of the AAPS solution showed a maximum value at 0.5% with an interracial shear
strength of 3 2 2 x 10 - 4 kg m -2 [146]. After exposure to water at 95 °C, embedded fibers
of E-glass with a coating from a 1% solution of AAPS lost strength as a function of time
of immersion, up to 168 h as determined by fiber pull-out [145,146]. For coatings deposited
from solutions of coupling agent of 1.0, 2.0, 5.0, and 10.0 wt.%, apparent removal of the
physisorbed and chemisorbed phases by immersion of the previously coated fibers in water
at 95 °C for 4 h resulted in interracial shear strengths that were considerably higher (all
approximately 3 3 0 x 1 0 - 4 kg m -2) than that of the original coating [145,146]. These results
are significant because they demonstrate the relationship between the interracial strengths
and the treatment of the coatings by selective desorption; the measured interracial shear
strength for those coatings that had been selectively desorbed, presumably by removing the
physisorbed and chemisorbed phases, were significantly stronger. The mechanical properties
of the interfacial region are improved either by the presence only of the highly cross-linked
fraction, the monolayer-equivalent and possibly some of the chemisorbed phase, or by the
lack of the presence of the physisorbed phase that could potentially interdiffuse with the
resin and modify the properties of the interphasal region, or both.
Piggott [147] evaluated the m o d e of failure of the fiber/resin interface for both thermosets
and thermoplastics. Both types of resins resulted in brittle failure at the interface in a
single-fiber pull-out test. Piggott suggested tailoring of these interfaces to have a modulus
that was intermediate between the resin and the fiber to improve the properties of the
composite that are governed by interfacial properties. Although he did not address the
change in the properties of the interface/interphase after exposure to environmental agents,
these concepts are useful in addressing the issues of durability.
Latour et al. [148] studied the mechanisms of fracture of the fiber/matrix interface for
composites consisting of either carbon or polyaramide fibers in a polyarylsulfone (PAS)
resin. They were interested in these systems as medical implants; therefore, their study
involved the fracture behavior of the interfaces of these composites and their resistance
to environmental agents. Specifically, they investigated the durability of the interface not
300 C.L. Schutte / Environmental durability of glass-fiber composites

only under fatigue loading, but also in a saline solution that approximated the electrolytes
in New Zealand White rabbits. As a model system, they chose the microdroplet pull-out
test, an evaluation of which was presented by Haaksma and Cehelnik [149]. By using the
same experimental setup for determining the interracial shear strength, they fatigued the
microdrop by applying a sinusoidal tensile-tensile load at a frequency of 1 Hz and a fatigue
ratio of R = 0 . 1 at four levels of stress. The ultimate and fatigue strength at the interface
for the carbon fiber/PAS was greater than that for the polyaramide fiber/PAS under both
dry (15% RH) and wet conditions (saline solution at 37 °C for 24 h). Exposure to the
saline solution, without mechanical fatigue, significantly reduced the force required to pull
out the fiber from the microdrop for samples containing both types of fiber. They concluded
that the mechanism of failure included interracial adhesive failure and cohesive failure in
the matrix, and they related the increase in cohesive failure with an increasing measured
interracial shear stress. Exposure to both moisture and conditions of fatigue reduce the
force required to pull out the fiber from the microdrop, and Latour et al. have associated
this with interracial damage.
In a further study [150], Latour and Black concluded that, for the geometry of the
microdrop test, the decrease in the required force to pull out the fiber occurred within
the first 10 min of immersion. Previous calculations [151] indicated that the sample achieved
an equilibrium content of water by this time. The presence of water potentially can serve
both to swell the resin, thus reducing residual curing stresses, and to attack the interracial
region chemically. After an initial decrease in the force required to pull out the fiber by
10 min of exposure to water, this force for samples containing fibers of Kevlar did not
change significantly. In contrast, those samples with carbon fibers continued to experience
a decrease in the pull-out force with increased time of exposure to various aqueous solutions.
According to Latour and Black, examination of the fracture surfaces by scanning electron
microscopy to determine the modes of failure revealed both apparent adhesive failure at
the carbon fiber/resin interface and a mixed mode of apparent adhesive and cohesive failure
for the Kevlar/resin interface. Further characterization of the fracture surface by a com-
plementary technique such as XPS, possibly showing no presence of resin on the fiber,
would be further evidence supporting the conclusion that the failure was adhesive.
More recent studies by Meyer et al. [152] investigated the interracial bonding between
unsized carbon fiber and both polyetheretherketone (PEEK) and polysulfone (PSF) as a
function of hydrothermal exposure to physiologic saline at 37, 65, and 95 °C for up to 5000
h. Thermal exposure involved the same thermal and temporal conditions in a dry oven.
Thermal exposure at all three temperatures did not result in measurable changes in the
pull-out force for the carbon fiber and either the P E E K or the PSF resin. In contrast,
hydrothermal exposure resulted in a decrease of this force. As the temperature or time of
exposure increased, so did the extent of decrease of the pull-out force. Further work focused
on understanding the effect of diffusion of water into the drop of resin is needed.
Gaur and Miller [153] studied the durability of E-glass fiber/epoxy interfaces. Whereas
exposure to 100 °C heat for 40 h resulted in only a 4% reduction in pull-out force, immersion
in water at 88 °C for 30 min resulted in a reduction of 57%. Fig. 9 illustrates the reduction
in the calculated interfacial shear strength, assuming interracial failure, with exposure to
hot water. Thereafter, exposure to hot water resulted in a pull-out force that was 42% of
the value of the unexposed sample. Samples that had been exposed to hot water and
subsequently dried showed only partial recovery of the resistance to pull-out of the fiber;
75% of that for the unexposed sample. Not only did the water degrade the wet properties
C.L. Schutte / Environmental durability of glass-fiber composites 301

40 I I I I

I J Contr°l J
Hydrothermally aged

30

u
c
20
O-

h
7,
N
z. \
\
\
\
\
10 \ \
\
\
\
\
\
"I
\
\
\
\
Nl nn
0
0 2O 3O 40 50 60 70

Tou (MPa)
Fig. 9. Distribution of shear strength for E-glass/epoxy before and after 72 h of exposure to water at 88 °C, using the single-
fiber puU-out test [153] (by permission of the publishers, Society of Plastics Engineers).

potentially by plasticizing the resin, leaching out components of the resin, and potentially
attacking the interface, but it also resulted in irreversible loss in resistance to fiber pull-
out that was interpreted as degradation of the interfacial properties. The geometry of the
microdrop test allowed for the rapid achievement of an equilibrium in saturation of water,
compared to that for the single-fiber fragmentation test, and this achievement of equilibrium
is advantageous. At a long time of exposure, however, the glass fiber has the potential to
degrade in strength where breakage of the fiber occurs over pull-out of the fiber; thus the
test may no longer be performed.

4. 7. Summary

The structure of the coating of a coupling agent depends upon experimental conditions
such as the chemistry of the coupling agent, the concentration of the solution, the solvent,
the pH if an aqueous solution was used, the substrate, the pretreatment of the coating by
solvent to desorb fractions of the coating, and the degree of cure prior to the application
of the polymer resin. The examples cited here serve to illustrate how the chemistry and
the morphology of the coating of coupling agent can have a profound influence on the
mechanical properties and durability of the interracial region. Although model coatings do
not contain all of the ingredients of complicated commercial coatings, they do allow for
the study of the role of each component in determining their contribution to strength and
durability of the interracial region. Further understanding of the role of each component
should allow for the design of more complicated model systems that are representative of
commercial systems. Determination of these chemical and morphological effects will allow,
hopefully, for the ability to tailor an interracial region for optimal mechanical and durable
performance.
302 C.L. Schutte / Environmental durability of glass-fiber composites

The general trends for results of both the single-fiber pull-out test and the fragmentation
test are consistent with a decrease in effective interfacial strength as the time of exposure
to water increases. These tests are excellent techniques for evaluating the relative durability
between different types of surface treatments. When interpreting mechanisms of failure,
however, care is necessary; in order to conclude that adhesive failure of an interface
occurred, for example, appropriate surface analysis must accompany the results of mechanical
tests. Also, the role of water in potentially reducing the residual stresses and imposing
swelling stresses in the interracial region on both types of test geometries must be better
understood through, perhaps, appropriate mechanical modeling coupled with experimental
data. In order to determine the ability to translate data from model systems to macroscopic
composites, results of tests from these model systems must accompany those from tests on
macroscopic composites.

5. Studies of changes in macroscopic mechanical properties of composites

5.1. Uptake of solvent

Although the mechanisms of diffusion of water into polymers are still not completely
understood, the additions of fibers in the resin have the potential to complicate the process
even further. The mechanism(s) of uptake of solvent in composites may be different from
those in the pure polymer. Many researchers have speculated that water can diffuse along
the fiber/matrix interface; however, this speculation is not easy to prove. Most studies
involving the uptake of moisture report the total weight change of the sample, and not
the distribution of water in the sample.
Mehta et al. [154] showed that biaxial orientation of cellulose acetate resulted in a
higher coefficient of equilibrium sorption for water compared to that for a sample that was
compression-molded. Furthermore, glass-fiber composites of these films showed an unexpected
increase in content of water as the content of glass fiber increased. Although these results
are contrary to expectations based on the solubility of water in resin for composite alone,
they serve to illustrate how potential incomplete wetting of the fiber, and the presence of
the fibers, may influence the conditions of processing of the resin, orientation, and resulting
voids and/or density of the resin that could control the solubility and diffusion of water in
the composite.
One study on the uptake of moisture in E-glass fiber/polyester composites showed that,
in water at 23 °C, the ratio of the apparent equilibrium content in distilled water compared
to that in salt water was 6.5, 8.4 and 2.3 for polyester sheet molding compounds (SMCs)
of approximately 25, 30 and 65 wt.% fiber respectively. Although the equilibrium moisture
content of water was slightly higher for some of the samples at 50 °C, the ratio for that
in distilled vs. salt water was not significantly different from those values at 23 °C. Because
each type of sample contained a different polyester resin and the SMCs containing 25 and
30 wt.% fiber also contained calcium carbonate, conclusions based on the uptake of water
as a function of content of glass fiber are not possible. The authors believed that this
increase in the equilibrium content of water in distilled water was due to the initial presence
of ions in the resin [155] -- where, because of osmotic pressure, a preferred state for the
system involves distilled water diluting the ions in the sample to a greater degree than
would salt water.
C.L. Schutte / Environmental durability of glass-fiber composites 303

Water can, potentially, cause debonding at the interface between the fiber and resin,
not only through chemical attack and reaction, but also through mechanochemical effects
such as osmotic pressure. Also, the resin can degrade chemically from attack by water.
Bascom emphasized the influence of both stress and water in accelerating failure in glass-
fiber composites [156]. The possible synergy between exposures to moisture and cyclic stress
that may be responsible for the acceleration of damage to composites is not well understood,
but the relation between stress and corrosion of glass fibers and their strength is well
documented.
Some studies of the influence of stress on diffusion of water showed consistent trends.
For example, Pollard et al. [157] studied the diffusion of water into E-glass/isophthalic
polyester resin composites. Because previous studies [158] suggested that pressure accelerated
the degradation of the mechanical properties, they looked at the effect of pressure on the
diffusion of water in their system. They also looked at the diffusion of water into epoxy
under stress [159]. Diffusion of sea water at atmospheric pressures initially increased with
increasing temperature, between ambient, 40 and 60 °C. At 80 °C the composite's properties
deteriorated. Longer-term exposure of the composites under these conditions revealed
damage to the composites at both 40 and 60 °C. They pointed out that increasing the
temperature did more than increase the rate of diffusion of water (by causing the possible
hydrolysis of the resin); therefore, accelerated tests of attack by moisture performed at
elevated temperatures may not accurately reveal mechanisms that occur at lower temperatures.
U n d e r atmospheric pressure and at 15, 35 and 70 MPa, the diffusion of deionized water
and of sea water, as measured by weight gain, did not differ significantly in this composite.
They found that the coefficient of diffusion was not a function of pressure, from analyses
of the diffusion data using the model of Shen and Springer [160], based on Fick's second
law, and assuming a concentration-independent diffusion coefficient, correcting for effects
of diffusion at the edges of the sample, assuming isotropic behavior, and neglecting a factor
for diffusion u n d e r stress. They also looked at their data based on a model that assumes
surface-adsorption-limited diffusion. Because their data did not fit this surface-dominated
model, they could not explain the lack of effect of pressure on the behavior of diffusion
in their samples.
Avena and Bunsell [161] showed that the uptake of water was d e p e n d e n t on the fiber/
matrix interface in samples containing 60 vol.% E-glass in an epoxy resin. They compared
the uptake of water at different pressures (5, 10, 20 and 29 MPa) for fibers that were
treated and untreated, and found a slight decrease in the equilibrium content of water
with increasing pressure. They saw no difference in the coefficient of diffusivity as a function
of pressure with the composite containing the sized fibers, and these results agree with
those of Pollard et al. [157]. In contrast, the diffusivity of water in the composites decreased
as the pressure increased for the samples containing unsized fiber. Furthermore, the uptake
of water was greater for the sample containing unsized fibers. Apparently the surface of
the fibers, and possibly the adhesion of the fiber/matrix interface and presence of voids,
influenced the solubility of water under these conditions. The decrease in the stress at
failure in a test using a three-point bend correlated with the increase in content of moisture;
therefore, the samples containing unsized fibers failed at a lower level of stress than did
samples containing treated fibers.
Marom and co-workers [162-165] established a theoretical model describing diffusion
of water in composites under stress. They relate the change in the free volume of the
sample, as described in the Doolittle equation, to the applied stress. Assuming Fickian
304 C.L. Schutte / Environmental durability of glass-fiber composites

diffusion, and using the laminate theory applying elastic properties of the fiber and resin,
but not allowing for damage-induced enhancement of diffusion such as through microcracks
or capillary flow along fiber/matrix interfaces, they showed excellent agreement between
their theoretical diffusion model and experimental measurements for samples of unidirectional
carbon-fiber/epoxy composites under stress at a variety of orientations and contents of fibers.
The equilibrium content of moisture decreased with increased applied stress. Furthermore,
their model showed that the diffusion constant under an applied stress was also influenced
by the orientation of the fibers, a 90 ° orientation being more strongly influenced than one
at 0 °. This model is the most comprehensive to date, and serves as an excellent basis for
prediction of the effect of stress on diffusion into composite samples.
Weitsman [166] considered the viscoelastic properties of polymers to describe the
deviation from Fickian behavior observed in many polymers and polymer composites. The
equations that he derived for the chemical potential and the strain revealed that the time-
d e p e n d e n t strain in the sample was expressible using reduced time that d e p e n d e d on the
content of moisture, and the viscoelastic behavior in the tensile properties d e p e n d e d on
the time d e p e n d e n c e of the chemical potential. This model indicated that the sample
required time to reach its equilibrium value for chemical potential; the uptake of water
will also require time to reach equilibrium.
Kasturiarachchi and Pritchard [167] found that the diffusivity of water into glass/epoxy
composites was not influenced by bending of the samples, as determined by macroscopic
measurements of weight gain under 95% R H at ambient temeperature, 45 and 80 °C.
Perhaps they m a d e this observation because they could not isolate the portion of the sample
under compression from that under tension. They observed that no sample reached an
equilibrium value for content of water and that at 80 °C deviation from Fickian behavior
occurred. They postulated that water reacted with the curing agent dicyandiamide. The
face of the sample under compression exhibited buckling of the fiber and rupture at the
surface. Samples that were under high stresses failed by delamination under hydrothermal
exposure at 45 °C.
Springer et al. [168] looked at the uptake of solvent in polyester-E-glass composites
(containing 25, 28 and 65 wt.% glass) under liquids and vapors. The liquids were distilled
water, NaCl-saturated water, No. 2 diesel fuel, jet A fuel, synthetic aviation lubricant and
Indoline HO 30, while the vapors were air at 40, 60 and 100 RH. They measured the
increase in weight after exposure. The apparent diffusivity, calculated using the solution
for a semi-infinite solid with Fick's second law [160] and the macroscopic weight gain of
the composite, increased with increasing temperature. The relative rank of the apparent
diffusivities and maximum content of solvent d e p e n d e d on the solvent, the composition of
the composite, the relative humidity, and the temperature. In a further publication [169],
Springer et al. investigated the mechanical properties of samples under exposure to these
conditions for both polyester and vinyl ester composites. Although a number of factors
such as the solvents, the temperature and the type of composite governed the relative loss
of properties of the composites, their exposure to the solvents such as saturated salt water,
Indoline and anti-freeze influenced the strength and modulus of the polyester composites
to the greatest extent. These properties were not recoverable after drying the samples.
In an attempt to study the distribution of water in wet samples of composites, Marzi
et al. [170] used dynamic-mechanical analyses (DMA) of epoxy resins that had been
previously exposed to water. They claimed a correlation between the distribution of water
in the composite -- at the interface, in the resin and in microvoids -- and structure of
C.L. Schutte / Environmental durability of glass-fiber composites 305

the resin. For liquid water at 0 °C, two samples of resins with no reinforcement, one with
the diglycidyl ether of bisphenol-A and bis(4-amino-3-methyl-cyclohexyl)methane, PACM,
D G E B A / P A C M and the other with triethylenetetramine, D G E B A / T E T A , reached a low
level of water content of 1.5 and 8.0%, respectively. The sample that contained D G E B A /
T E T A had a significant increase in the dielectric loss function (tan6) with an increase in
the content of water up to an equilibrium value of 8.0 wt.%. They attributed this increase
in both the equilibrium value of the content of water and the tan6 to a higher content of
microvoids in the D G E B A / T E T A resin. The normalized rate of uptake of water in glass-
fabric-reinforced D G E B A / T E T A compared to the unreinforced resin at 70 °C was initially
slower, but increased as the content of water increased at longer times; the equilibrium
contents of water for these samples were 10 and 8 wt.%, respectively. Furthermore, an
additional relaxation appeared in the D M A spectra of the reinforced composite; the intensity
and temperature of the relaxation both increased with increasing content of water. At the
intermediate temperature of 21 °C, the additional relaxation for the D G E B A / T E T A system
occurred only for samples whose glass fibers were previously treated with hydrophilic coatings.
The sample of D G E B A f F E T A showed an increase both in the equilibrium content of
water and the tan6 for reinforced composites at both 21 and 70 °C compared with the
unreinforced resin at 70 °C. These results were consistent with the authors' interpretation
of increased content of water at the interface. For the D G E B A / P A C M system, the unreinforced
resin had an equilibrium content of water at 0 °C of 1.5 wt.% of the resin. The reinforced
composite, however, reached the normalized content of water of 1.8 wt.%, and its tan6
also increased in magnitude.
Exposure to boiling water influenced the dynamic mechanical properties of epoxy
composites containing glass fibers and glass beads [171]. The beta-relaxation at - 5 0 °C
corresponded to the presence of the glyceryl unit, formed from the reaction of the diglycidyl
ether of 2,2-bis(4'-hydroxyphenyl-)-propane with 1,3-diaminopropane. The height and position
of the beta-relaxation appeared to correlate with the concentration of the glycidyl functional
groups [172], mechanical fatigue [173], thermal degradation [174] and content of water
[175]. As the content of water increased in the composite, the beta-transition increased in
both magnitude and transition temperature [176]. Samples with no fillers showed no change
in the beta-relaxation after exposure to boiling water. The composites, however, exhibited
a new beta-transition near - 3 0 °C that was more pronounced in those samples containing
fibers of glass or aramide than for those with carbon fibers. Moreover, the intensity of'the
transition and the transition temperature increased with time of exposure to boiling water.
The rate of these changes for composites containing glass fibers was greater for the sample
containing fibers whose silane coating was previously washed off, compared with the fibers
containing the silane treatment. Williams [171] suggested that one explanation for the
appearance of the loss peak could be changes at the interface of the fiber and matrix.

5.2. Effect of solvent on mechanical properties

Morii and co-workers [177,178] studied the change in weight of randomly oriented
glass-fiber-reinforced bisphenol unsaturated polyester resin on exposure to water at 60 and
80 °C for up to 3000 h. The profile of the uptake of weight (water) contained three phases.
The authors attributed the initial phase to the resin absorbing water, the second to absorption
of water at the fiber/matrix interface accompanying that in the resin, and the third to
306 C.L. Schutte / Environmental durability of glass-fiber composites

dissolution (loss of weight) of material from both the matrix and interface accompanying
the absorption processes.
Further work by Morii et al. [179] involved exposure of samples of y-methacryloxy-
propyltrimethoxysilane-treated glass woven cloth in a vinyl ester matrix to water at 40, 60
and 80 °C for periods of time up to 1000 h. Their study investigated the effect of exposure
to water on both the tensile strength and the initial fracture stress, by optically monitoring
the damage to tensile specimens as they were stressed. In general, the tensile strength
decreased both as the temperature of the water increased and as a function of time of
exposure. These results appeared to be d e p e n d e n t on the rate of diffusion of water into
the sample and the influence of that content of water through plasticization of the resin
and/or interphase.
According to Morii et al. [179], changes in brightness of the sample indicated either
debonding or cracking of the matrix, and this change was associated with the stress where
initial fracture of the sample occurred. In contrast to the influence of temperature of the
water on the tensile strength of the samples, the initial stress at fracture, although it
decreased for all conditions of exposure within 100 h, was independent of temperature.
Morii et al. associated the observed damage with degradation of the fiber/matrix interfaces
leading to early debonding for fibers that are transverse to the direction of stress.
Farrar and Ashbee [180] concluded that osmotic pressure at an E-glass/epoxy interface
caused a "pressure pocket" due to the presence of alkaline and alkaline earth ions, Ca +
and K +, from E-glass and impurities from unreacted hardener. They related the presence
of these "pressure pockets" to loss of transfer of load between the fiber and the matrix.
Chateauminois et al. [181] studied the effects of hygrothermal aging using a three-
point bend test of unidirectional glass fiber/epoxy composites made by filament winding.
Hygrothermal exposure consisted of immersion of samples in distilled water at either 30,
50, 70 or 90 °C for 100 days. Choice of these conditions reflected the achievement of
saturation of content of moisture for these samples. Although the Young's modulus of the
samples did not change after exposure to water, their strength and strain at failure did,
as measured in three-point bend at a constant rate of strain. The loss in properties was
reversible on redrying for samples exposed to 30 and 50 °C. The m o d e of failure observed
in the untreated (dry) samples and in those exposed to hygrothermal conditions at 30 and
50 °C showed failure through microbuckling of the glass fibers on the side of the sample
under compression. For samples exposed to water at 70 and 90 °C, the m o d e of failure
involved cracking of the matrix, breakage of the fiber, and debonding of the interface on
the side of the sample that was in tension.
Static fatigue studies of samples exposed to the same conditions showed changes in
the mode of failure under three-point bend, from compressive to tensile, that were observed
in the test under monotonic loading. For the samples exposed to water at 30 and 50 °C,
the reduction of the lifetime of the samples was independent of temperature, and the
microbuckling that signified failure in the sample occurred to a greater extent than it did
for the dry samples. Whereas gradual propagation of cracks occurred in the tensile side
of the samples exposed to 70 °C, catastrophic failure occurred in those at 90 °C. The
propagation of the cracks was associated with breakage of the fibers at strains lower than
the failure strain for the fibers. This observation was consistent with the fibers and their
interfaces decreasing in strength from exposure to the warm water. Apparently, for this
case the attack on the strength of the fiber and its interface with the resin were the dominant
mechanism of environmental attack at 70 and 90 °C.
C.L. Schutte / Environmental durability of glass-fiber composites 307

Stoudt et al. [182] studied the effect of hydrothermal exposure of distilled water on
E-glass fiber/epoxy composites at both atmospheric pressure (0.1 MPa) and high pressure,
5.9 MPa. They found no measurable difference in either the uptake of water or the changes
in mechanical properties as a function of pressure. They compared the ultimate strength
and yield stress for samples of 00/90° fabric at 0 °, 90 ° and +45 ° both at saturation and
after being saturated and redried. After exposure to distilled water for 70 days at 20 °C
the saturated samples showed a slight decrease, 3-17%, in the ultimate strength, and a
measurable decrease in the yield stress, 17-38% of the original dry values. This decrease
in the yield stress was recoverable after the samples had been redried after saturation, and
suggests that plasticization of the resin by water may be responsible for this effect on the
yield stress. In contrast to the recoverable values for the yield stress, the ultimate strength
decreased further for the redried samples, by 9-37% of the original values. This result
suggests that exposure to the water resulted in p e r m a n e n t damage to the glass fibers, the
resin (through possible microcracking) or the interfacial region. Although the SEM images
of the fracture surfaces of the original dry, wet, and wet and redried samples show no
measurable change, the decrease in the ultimate strength for the +45 ° samples suggests
degradation of the matrix, the interface, or both.
Ishai did extensive research on the effect of environmental exposure on glass-fiber
reinforced composites [183-186]. He identified processes that caused reversible and irreversible
damage to the composites on exposure to water. Reversible processes included plasticization
and the stresses associated with swelling of a matrix. If the matrix experienced too great
a degree of swelling, however, the matrix cracked, and this resulted in irreversible damage
associated with loss of weight of the matrix and/or fiber from attack by water [184]. For
example, composites of E-glass/epoxy lost weight in water at 80 °C, from leaching of the
fiber as determined by infrared analysis that identified both components of the glass and
coupling agents, causing non-recoverable damage and loss of properties of the composite
compared to samples that were immersed in water at 22 °C for the same length of time.
A decrease in longitudinal strength due to loss of strength of the fibers, and a loss in
transverse strength due to degradation of the interracial strength, accompanied by the
leaching of the coupling agent, illustrated the loss in properties that Ishai observed.
Furthermore, he observed that the leaching of the fibers left voids that could take up more
water on further cycling. Stressing the composite prior to immersion could also affect the
retention of strength of the composite. Stretching the immersed composite up to 25% of
its ultimate strain resulted in an increase in the tensile strength (apparently the dominant
effect was plasticizing of the resin); strains greater than 25% but less than 40% of the
ultimate strain resulted in a decrease in the tensile strength (possibly both interfacial and
fiber degradation occurred under these conditions); and at 40% of its ultimate strain the
sample failed during environmental exposure, presumably from the accumulation of damage
from both loading and immersion. These results illustrate how the history of the sample
- its state of stress, for example -- influenced the degree of damage in the sample, the
-

dominant mechanism, and possibly its reversibility, after exposure to water.


Jones et al. [187] found that damage by stress corrosion of unidirectional polyester/
glass-fiber composites in sulfuric acid (1 M) appeared to involve degradation at the fiber/
matrix interface. Failure occurred in the region of the polyester sample that was exposed
to the solution of acid, as did a composite containing isophthalic polyester [188]. In contrast,
exposure of glass/epoxy composites resulted in failure of the sample at a region that was
not immersed in the acid solution [188,189]. Failure in an exposed region occurred in the
308 C.L. Schutte / Environmental durability of glass-fiber composites

epoxy system, however, when hydrochloric acid was used. Because the mode of failure
appeared to be d e p e n d e n t on the type of resin and acid, there did not appear to be a
universal mechanism responsible for failure of composites under these conditions. Despite
the differences in their modes of failure, samples of polyester resins appeared to protect
the fibers better than those made from epoxy [190].
Lieblein [191] took samples of glass fiber/polyester composites from submarines that
had been in contact with seawater during service for 10 to 19 years, and found an average
loss in mechanical properties, static and fatigue strength, of only 10 to 15%. Lieblein did
not state whether the samples of composites experienced stress or conditions of fatigue
while in use as part of the submarine.
Phillips [192] addressed the ability to reliably predict long-term properties of glass-
reinforced polyesters under exposure to aqueous environments by studying their stress-
rupture behavior. In general, three regimes of stress rupture that he found were: (a)
i n d e p e n d e n t of environment, (b) d e p e n d e n t on both environment and on level of stress,
and (c) independent of stress. He found that the type of polyester resin and the temperature
had a statistically significant influence on the time to failure. Also the environmental exposure
appeared to have insignificant influence on the failure at high stresses. Significant variability
in the data suggests that the assumption of a unique mechanism of failure was incorrect.
Whereas exposure to the environmental conditions influenced the time to failure, at lower
stresses the d e p e n d e n c e of time to failure on temperature is too great to be the result of
thermal activation of only a single mechanism.
Ellis and F o u n d [193] showed that the fatigue life of a polyester/chopped-glass mat
composite decreased with exposure to water. The resin debonded from the fiber during
immersion in water at 51 °C for 40 days, in the region where the uptake of water was
close to its equilibrium value. In contrast to the trend in fatigue life, the residual strength
increased with time of exposure to water. Because the authors saw a smaller number of
cracks in the matrix for the wet samples compared to those in the dry, they hypothesized
that the water m a d e the resin more extensible and possibly reduced residual stresses in
the composite.
Ishai and A r n o n [194] compared the mechanical properties, before and after exposure
to water, of two epoxy/glass composites containing 50 vol.% unidirectional glass fibers: one
for use below 100 °C (a flexible epoxy containing an elastomer), and one for use up to
180 °C (an epoxy, Novolac). The strength and Young's modulus of the composite containing
the epoxy designed for use at lower temperatures decreased after exposure to water. The
opposite trend existed for the high-temperature sample. Based on the results that the + 45 °
samples regained their original strength after exposure with subsequent drying, while the
90 ° and 0/90 ° specimens experienced a significant reduction in the original properties, they
attributed the mechanism of degradation to that of the fibers.
The orientation of E-glass fibers in an epoxy resin significantly influenced the reduction
in ultimate tensile strength (UTS) and Young's modulus after immersion of sample composites
containing 65 vol.% fibers in water at 70 °C for 20 days. Specifically, the results of Gopalan
et al. [195] showed a decrease in the UTS of 33.3, 8.0, 13.7 and 20.4% for resin, unidirectional,
bidirectional and chopped-strand mat, respectively. Likewise, for these samples, the Young's
modulus decreased 45.2, 5.0, 14.6 and 24.0%. Apparently, the orientation of the fibers in
these composites reflects the degree to which the interface and matrix influence the measured
properties compared to those that are dominated by the fiber. Analyses of samples by SEM
revealed loss of adhesion at the fiber/matrix interface and evidence of leaching of the resin.
C.L. Schutte / Environmental durability of glass-fiber composites 309

For unidirectional tensile testing at 0 ° orientation, the fibers dominate the property. Possibly,
the fiber also degrades in the water; however, the larger drop in properties for samples
t h a t are biaxial and random, as in the chopped-strand mat, illustrate how the interracial
properties, or lack thereof, dominate the strength.
If one mechanism dominates the durability of a composite for its lifetime, then accurate
predictions on the loss of properties should be possible. Unfortunately, without extensive
testing, one cannot be sure that a second mechanism responsible for loss of strength would
dominate at longer times of exposure, for example. If the mechanism responsible for loss
in properties does change, then predictions based on the initial one will be inaccurate. On
the loss in properties of E-glass chopped-strand mat laminate in a polyester resin, one
study concluded the existence of a dominant process or processes that allowed for treatment
by t i m e - t e m p e r a t u r e superposition. In this work, Phani and Bose [196] reported the loss
in flexural strength after exposure to distilled water at 25, 42, 50, 60, 80, 90 and 100 °C,
as determined in a three-point bend test for E-glass in resin made from isophthalic polyester
cured with methy ethyl ketone peroxide and cobalt napthanate. Potentially, the water could
attack (a) the fiber, by degrading its strength, (b) the matrix, by plasticizing, swelling,
microcracking and hydrolyzing it, and (c) the interface of the fiber and resin, by hydrolytic
attack and build-up of osmotic pressure. The authors assumed that plasticization of the
polyester and debonding of the interface were the predominant factors controlling the loss
of strength. They found that, for the periods of time for which they exposed their samples
(up to 480 h), the loss in strength was controlled by processes whose rate followed the
Arrhenius equation. Furthermore, they successfully applied the t i m e - t e m p e r a t u r e super-
position principle to the degradation of the strength. This study served as a example of
the use of a predictive model without specific knowledge of fundamental controlling
mechanisms of degradation.
Work by Bledzki et al. [197] illustrates the interaction between two competing mechanisms
of degradation of a composite of E-glass/ortho phthalic polyester cross-linked with styrene.
They examined the lifetime of this composite by both dynamic stress relaxation and fatigue-
crack propagation under environmental conditions of air and dilute sulfuric acid. For samples
that failed at early times of exposure and high initial stresses, the mechanism of damage
appeared to be dominated by mechanical failure, but evidence of chemical attack of the
glass fibers by the acid was present in these samples as well. At longer times of exposure
to the corrosive environment, however, chemical attack of the fiber dominated the damage
and the failure of the composite. Aqueous acids can leach alkalies and alkaline-earth metals
as well as Ca, Mg, AI and B from the E-glass. The authors examined single fibers in sulfuric
acid solutions and found a core/shell appearance and formation of surface spiral and axial
cracks accompanying the leaching process. The formation of the shell was consistent with
leaching of ions from the outside of the fiber. Furthermore, they showed that a decrease
in the strength of the glass fiber accompanied this leaching; the tensile strength decreased
to a value that was one-third of that for an unexposed fiber after only 100 h of exposure
to acid. The appearance of the fibers in the exposed composite resembled the single fibers
that they immersed in acid previously.
The authors [197] cited the core/shell appearance of the fibers and the surface cracks
as evidence for the chemical attack of the fiber with a loss in fiber strength. They reasoned
that for early times of exposure, where mechanical failure of the composites occurred, the
acid solution did not have enough time to attack and weaken the glass fibers sufficiently
to cause failure by this mechanism. For later times of exposure, however, the acid had
310 C.L. Schutte / Environmental durability of glass-fiber composites

degraded the strength of the fibers enough so that they were the weak c o m p o n e n t in the
system. Because the mechanism of degradation switched from one dominated by mechanical
failure to one governed by chemical degradation, any long-term predictions based on the
mechanism that governed failure at early times -- mechanical failure - would have been
inaccurate in predicting failures at long times.
A n o t h e r example of interaction of competing mechanisms is the work of Pritchard and
Speake [198], who studied the effect of temperature on the time to failure in water for
stress-rupture of E-glass/iso phthalic polyester laminate with 48 wt.% glass. Because the
time to failure was shortest at 45 °C compared to 60 or 30 °C, they reasoned that the
effect of temperature increased both the diffusion of water in the resin and its fracture
toughness. The shorter time to failure at 45 °C occurred, they explained, because the
diffusion of water was faster than for the 30 °C sample. At 60 °C, the fracture toughness
of the resin increased and this compensated for some of the damage from the water,
allowing for the longer time to failure at 60 °C compared to 45 °C. Estimation of the time
of failure at different temperatures requires knowledge of the effect of temperature on
both the rate of diffusion of water and the fracture toughness.
Marshall et al. [199] found that the cracks in the gel coat of chopped-strand mat of
E-glass in both vinyl ester and neopentyl glycol isophthalics served as stress concentrators
and assisted in diffusing the acidic solutions into the samples to the glass fibers. They
studied these samples under conditions of creep in flexure and found four stages of damage:
(a) resin creep and minor damage of the sample occurred; (b) diffusion of the solution
proceeded, and this served to plasticize the resin and make it more compliant; (c) the gel
coat cracked; and (d) stress corrosion of the fibers occurred. The environmental conditions
influence the time to failure; for example, at a critical strain, the time to failure at 20 °C
for a sample was over 103 h for air, 102 h for water, and instantaneous for an acidic solution.
Collins [200] found for an isophthalic-maleic polyester/E-glass filament-wound pipe
under stress while exposed to several acidic and basic solutions, that: (a) if the p H of the
solution was between 3 and 10 or the strain was less than 0.3%, stress-corrosion cracking
was not observed; (b) cracking of samples through strain corrosion was associated with
exposure of the samples to solutions of strong acids; (c) some cracking of the sample
was associated with exposure to high pH, although most likely the samples experienced
softening and blistering; (d) the concentration of the hydrogen ion in solution appeared
to correlate with the tendency for strain corrosion; (e) an increase in the temperature
of the acidic solution appeared to increase its corrosive properties; (f) prevention of the
strain corrosion accompanied the application of a gel coat; and (g) no correlation existed
between resistance to chemical attack and resistance to strain-corrosion cracking for
these samples.
The work of both Marshall et al. [199] and Collins [200] illustrates the importance of
the role of the gel coat in delaying or avoiding cracks through which the solvent can diffuse,
penetrate and degrade the fibers below. Hogg [201] found that the greater the fracture
toughness K~c of the resin, the longer the time to failure of the pipe in 0.6 mol I-~ HC1
solution at 20 °C. They found the following trend for time to failure from longest to shortest
(with K~c values in MN m-3/2): a flexibilized isophthalic polyester (0.77)> an unflexibilized
isophthalic polyester (0.62)> a teraphthalic polyester (not available)>vinyl ester (0.75)> an
epoxy modified bisphenol (0.49)> a bisphenol resin (0.45)> a brittle polyester (0.46). Post-
curing of the resin decreased the time to failure, presumably because the further cross-
linking of the resin decreased its toughness. Furthermore, he showed that samples containing
C.L. Schutte / Environmental durability of glass-fiber composites 311

a corrosion-resistant E-glass fiber had longer times to failure compared to samples containing
E-glass.
Hull and Hogg [202] compared the durability of hoop-wound and _ 55°-wound pipes
of E-glass terephthalic polyester on exposure to air and to 0.65 mol 1-1 HCI solution while
being stressed. In general, they found that the mechanism of stress and strain corrosion
was the same for pipes wound under both conditions. The time to failure of these samples
appeared to correlate with the stress parallel to the fiber. The transverse cracking that
occurred in the _ 55°-wound pipes had a significant influence on the time to failure, because
these cracks allowed for the solvent to diffuse into the sample and corrode the fibers faster
than if the solvent diffused into an uncracked pipe.
Juska [203,204] studied the effect of exposure of continuous glass-filled unidirectional
thermosets and thermoplastics to water at 50 °C for up to 9 months. This work involved
tests of 0 ° alignment of fibers in compression, 90 ° alignment of fibers in flexure, and short-
beam shear on continuous fiber composites of thermoplastic (polyphenylenesulfide, po-
lyetheretherketone, polyetherketoneketone, and a liquid-crystalline polymer) and thermosets
(epoxy and a polyester). The glass fiber/thermoset maintained its mechanical properties
better than the glass/thermoplastic systems. The latter showed evidence of degradation of
fiber/matrix debonding after exposure to water. Juska attributes this to the less-well-developed
status of coupling agents for use with thermoplastics, combined with the problems in
processing thermoplastic composites at such high temperatures. The results of the work on
thermoplastic composites showed no general trends with regard to their loss in properties
beyond the general result that the composites containing thermosets maintained their
properties better than did those containing thermoplastics.
Yeager et al. [205] studied the influence of automotive fluids - unleaded gasoline
with and without methanol or ethanol, fuel C, ethylene glycol/water, motor oil, transmission
fluid and brake fluid -- on the tensile strengths of stressed and unstressed injection-moldable
thermoplastic composites. All composites lost strength under the condition of strain, 0.25%,
at high temperature, 82 °C. Yeager et al. did not discuss possible mechanisms responsible
for their deterioration in properties.
Haarsma [206] offered a general test for studying the effect of the environment on
composites. He demonstrated that his test could be used to study the effects of 3' irradiation
and exposure to water, humidity, and crude oil by exposing segments of short cylinders of
composites that were cut from previously pulltruded rods containing continuous glass fibers
in resin. Therefore, these samples expose the interface of the fiber and matrix to the
environmental conditions at the ends of these cylinders. Haarsma used a compression test
that he modified from the ASTM Method C-496. This test puts the samples under compression
along their diameter. The measurement of transverse tensile strength was controlled by the
interfacial bond strength. Haarsma demonstrated that this test yielded results that were
reproducible, sensitive to changes in the properties of the samples, and appropriate for a
routine test. According to Haarsma, the geometry of the sample, particularly the faces
exposing the interface of the resin and matrix, allowed for the evaluation of the properties
on exposure to environmental agents.
Houston and Johnson [207] tested the tensile and compressive properties of liquid-
molded composites, made of vinyl ester, urethane and phenolic matrices, both before and
after exposure to elevated temperatures with and without high contents of moisture. The
most damaging conditions of exposure consisted of both high humidity and temperature,
after which the authors noted damage to interfaces between the glass fibers and resin. The
312 C.L. Schutte / Environmental durability of glass-fiber composites

tensile and compressive moduli decreased slightly for all three types of resins, and the
tensile and compressive strength decreased significantly for the vinyl esters, phenolic
(approximately 50% of the original values for both the vinyl esters and phenolic) and
urethane (between 70 and 90% of the original values). These decreases in tensile and
compressive strength depended on the conditions of exposure.
Johnson [208] compared the effects of weathering, immersion in water, and exposure
to chemicals on a hand lay-up glass-reinforced composite, GRP, (E-glass/polyester containing
chopped-strand mat, 20-35 wt.% glass). He also examined sheet molding compounds (SMCs)
containing E-glass/orthophthalic or bisphenol-A polyester/mineral filler (25-30 wt.% glass/
25-40 wt.% resin/35-50 wt.% filler). The filler consisted of clay, calcium carbonate or calcium
magnesium carbonate. Samples of the SMCs experienced loss of weight with subsequent
erosion of glass fibers after 1 to 2 years for orthophthalic polyesters, and after a few months
for those containing bisphenol-A. The exposure was in the Alps in Germany or on the
Dutch coast for 10 years. Initially, these samples experienced a reduction in flexural strength
for the first 2 to 3 years. The flexural strength leveled off at an average value of 80%, and
a range between 60 and 100%, of the initial strength. In contrast, GRPs with a gel coat,
a polymer-rich outer layer designed to protect the fibers, exhibited good gloss after 5 years
of exposure. After 2.5 years of exposure, these samples showed no loss in tensile properties,
a 20% decrease in flexural modulus, and a loss in strength of up to 10%.
After being immersed in water, the SMC samples had a content of water that was
significantly higher (increasing for approximately 1 year, thereafter leveling off at 5 wt.%
of the composite) compared to samples of GRPs containing no mineral filler, with an
equilibrium value of 1 wt.% of the composite for water. Exposure of the SMCs to deionized
water for 1 year resulted in a flexural modulus and strength of 87% and 69% of the original
values, respectively. Samples of GRPs retained 97% and 93% of their original flexural
modulus and strength under the same conditions.
The retention of flexural strength for SMC samples after their exposure to 5% sulfuric
acid for 2 years was: 38% for those containing isophthalic resin, 58% for bisphenol-A, 62%
for vinyl ester, and 82% for isophthalic neopentyl glycol. Samples of SMCs containing clay
and isophthalic neopentyl glycol retained a flexural strength of 72, 66 and 97%, and flexural
modulus of 86, 77 and 97% on exposure for 1 year to pH 4, pH 10 and saturated sodium
chloride, respectively. For these samples, exposure to saturated aqueous sodium chloride
had little effect on the properties of the SMC samples, while exposure to deionized water
degraded the properties. However, samples of GRPs containing bisphenol-A polyester did
not lose properties after being immersed in deionized water at 99 °C for 1 year. Similarly,
GRPs containing laminates of isophthalic and bisphenol-A polyesters retained 91% and
100% of their strength after being exposed to orthophosphoric acid at 50 °C for 4 months.
In general, the properties of the SMC samples degraded faster than those of the GRPs.
Possibly this difference in behavior is due to less protection of the fibers for the SMC
samples.

5.3. Environmental effects on growth of cracks

O'Grady et al. [209] measured the influence of the environment on the growth of
cracks in a single layer of glass-fiber (chopped-strand mat)-reinforced polyester as a function
of exposure to water at 20, 50 and 70 °C, to both aqueous HCI (pH 1.5) and aqueous
N a O H (pH 13.0) at 20 °C. The samples were not pre-equilibrated in the solutions prior
C.L. Schutte / Environmental durability of glass-fiber composites 313

to the test. The critical stress intensity factors, Kc (MN m-3/2), w e r e : 5.65 in air at 20 °C,
5.95 for water at 20 °C, 4.68 for water at 50 °C, 3.65 for water at 70 °C, 5.33 for p H 13.0
at 20 °C, and 5.04 for p H 1.5 at 20 °C. The presence of water increased Kc at room
temperature. They explained that the larger zone of damage for the wet sample at 20 °C
compared to that for the dry sample may be due to plasticization of the resin by water.
Furthermore, increased water temperatures appeared to decrease the toughness of the
samples.
In contrast to the results of O'Grady et al. [209], Owen and Cann [210] found that
the presence of water, at room temperature and for prolonged periods of time of exposure,
decreased the value of Kc by 10% compared to that for a dry laminate of polyester reinforced
with chopped-strand mat or fabric of woven roving. Possibly, the time of exposure to water
allowed for the laminate used in this study to reach the equilibrium content of water, where
those of the single layer studied by O'Grady et al. did not, since they were tested without
pre-equilibration. Furthermore, at the short time of exposure experienced by the samples
of O'Grady et al., the increase in Kc may be due to plasticization of the resin. After samples
achieve their equilibrium content of water, however, degradation of the interface and of
the strength of the fiber could have occurred, and this explanation is consistent with the
decrease in K~ that Owen and Cann observed.
Sekine [211] found degradation at the glass fiber/resin interface and dissolution of A1 +,
Ca +, K ÷ and Na + from the fibers of E-glass to be associated with the loss of fracture
toughness in random chopped glass fiber, 60 wt.%, in a resin of unsaturated polyester after
immersion in distilled water at 25 °C for up to 1 year. With increased time of exposure
to water, the maximum load to failure and the fracture toughness decreased linearly.
Aveston et al. studied the long-term strength of glass under wet conditions [212]. They
measured the decrease in the strength of strands of E-glass fibers under conditions with
and without stress. Studies of stress rupture indicated the trend in the decreasing rate of
loss of strength for: strands of E-glass immersed in distilled water > unimpregnated fibers
exposed to atmospheric moisture > polyester-impregnated strands previously immersed in
distilled water > polyester-impregnated fibers. Both polyester- and epoxy-coated samples lost
approximately five-sixths of their original strength, with the sample containing epoxy exhibiting
an induction period, presumably because of the slower rate of diffusion of water in epoxy
compared to polyester. U n d e r wet conditions combined with no stress, samples lost a minimal
a m o u n t of their strength, even after exposure to water for the same periods where similar
samples had lost approximately 80% of their strength under conditions of stress. The loss
in strength of E-glass in water fitted a model based on slow crack growth for glass. The
strength of this type of sample can therefore be predicted under these conditions. Slow
cyclic stress (0.1 Hz) resulted in a decrease in time to failure for both the wet and dry
samples that was similar to that under static load, with the rate of loss of strength for the
wet samples being faster than for the dry samples.

5.4. Effects of fatigue coupled with environmental exposure

Romans et al. [212] showed that cyclic loading was significantly more damaging than
static stress on S-glass/epoxy samples. The composition of S-glass is optimized for strength;
it is 64.2 wt.% SIO2, 24.8 wt.% A1203, 0.21 wt.% Fe203, 0.01 wt.% CaO, 10.27 wt.% MgO,
0.27 wt.% Na20, 0.01 wt.% B203 [213]. These results were consistent with the findings of
Schadler and co-workers [136] who used the single-fiber fragmentation test on a carbon
314 C.L. Schutte / Environmental durability of glass-fiber composites

fiber/polycarbonate system. Likewise, Mandell [214] showed that, under moist conditions,
cyclic loading was significantly more damaging to polymer composites than static loading.
He looked at the propagation of cracks in E-glass/epoxy laminates in dry and wet samples
under conditions of static and dynamic fatigue. The growth of cracks under static conditions
of constant applied stress was faster for dry samples than for those that were wet. In
contrast, the rate of growth of cracks was faster for wet samples under cyclic loads. Previously,
Mandell et al. showed the same trends for polyester composites [215]. U n d e r constant load,
the presence of water corresponded to an increased amount of delamination at the crack
tip compared with that in dry samples. This increased zone of damage reduced the amount
of stress at the crack tip and reduced the rate of growth of the crack. U n d e r cyclic loading,
however, large zones of delamination existed for both dry and wet samples. The rate of
growth of cracks for wet samples was faster, presumably because the presence of water
weakened the material at the crack tip.
In later work, Mandell et al. [216] showed that, for fourteen different systems, cyclic
fatigue of composites in tension was dominated by the fatigue of the E-glass fiber. The
fibers studied by them, or by others, were in strands, unidirectional laminates, sheet-molded
compounds and injection-molded short-fiber composites. The decreases in strength, as
determined by the S - l o g N curves, were 10% per decade regardless of resin (polyester,
epoxy or rubber-toughened epoxy) or m e t h o d of fabrication, as shown in Fig. 10. In composites
made from either woven fabrics or woven rovings, the S-log N curve has a steeper slope
than 10% per decade between the UTS and 30% of the UTS; this appeared to be a result
of a more severe mechanism of damage, involving delamination. At maximum stresses below
30% of the UTS for these types of samples, the slope approached the more typical rate
of 10% per decade. Possibly, the mechanism shifted to that observed in samples without
woven fibers, whose dominant mechanism for fatigue was controlled by fiber fracture. The
ability to predict the long-term behavior of this type of sample requires further understanding
of the mechanisms involved as well as the conditions under which they are active.
Mandell et al. [223] studied cyclic fatigue of E-glass, surface-compressed glass, and
glass ceramic single fibers in tension to further study the mechanism(s) responsible for
damage of glass fibers in fatigue. Fig. 11 shows their results, illustrating that the loss in
strength for the single fibers was less than that for strands of fibers: 3--4% per decade
compared with 10% per decade. These results suggested that damage in a composite under
tensile, cyclic fatigue is due to fiber-fiber damage (abrasion). The single fiber degraded
from stress corrosion at a lower rate than did strands because the fibers did not experience
fiber-fiber contact. If contact with other fibers caused the damage responsible for the
degradation of glass fibers in composites, then fibers that are more resistant to surface
damage should degrade at a slower rate than the rate for E-glass fibers.
Surface-compressed macrofibers typically are strengthened by ion-exchange process
whereby the small, diffusible cations near and at the surface of the fiber can exchange with
heavier and bulkier cations. Because of the presence of the larger ions at and near the
surface, the surface of the glass was in compression, and this process strengthened the
glass. This strengthened surface should reduce the growth of surface scratches or flaws and
experience less susceptibility to fatigue. Mandell and coworkers realized these expectations
with both surface-compressed glass and glass/ceramic macrofibers. Single macrofibers of
both types showed no susceptibility to fatigue under the conditions of their experiment,
up to six decades. The glass fibers showed some sensitivity to brittle matrices, however.
U n d e r cyclic and static conditions, cracks propagated through the matrix and the fiber.
C.L. Schutte / Environmental durability of glass-fiber composites 315

1000 I I I I

~" z t
o_ 800
v
t,9
t--
i 600

400

u>, //~3 13

0)
8
.E 200 5~¢o tz

11
0 t t L I
0 20 40 60 80 100

Slope of S - N Curve, 13, (MPo/decode)


Fig. 10. Diagram illustrating that the fundamental fatigue behavior is dependent on that of the strand of glass fibers, regardless
of the resin. Plot of ultimate tensile strength, single-cycle strength, vs. slope B of S - N curve for non-woven glass-fiber
composites in tensile-tensile fatigue for R = 0 to 0.1 [216] (by permission of the publisher, ASTM).
1. Unidirectional glass/epoxy [217], Vf=0.5
2. Unidirectional glass/epoxy [217], Vf=0.33
3. Unidirectional glass/epoxy [217], Vf=0.16
4. 0°/90° glass/epoxy [218]
5. 30--40 wt.% glass in poly(hexamethylene adipamide) [219], injection-molded
6. 30--40 wt.% glass in polycarbonate [219], injection-molded
7. 30--40 wt.% glass in polyphenylenesulfide [219], injection-molded
8. 30-40 wt.% glass in poly(amide-imide) [219], injection-molded
9. Chopped-strand mat/polyester [220]
10. Sheet molding compound (SMC) of rubber-modified epoxy
11. SMC, rubber-modified epoxy
12. SMC, R50
13. (0~/+45°/90 °) glass/epoxy [221]
14. Chopped-strand mat/polyester [222]

Composites of the macrofibers in epoxy and in polysulfone showed some fatigue under
cyclic loading; their loss in strength was significantly less than that of E-glass, however.

5.5. Summary

Studies of the effect of environmental exposure on polymer composites potentially


includes the attack of each of the components covered in previous sections. Furthermore,
a combination of these processes may result in a rate of degradation of the composite that
is different from that predicted by the isolated mechanisms. This potential synergy may
exacerbate the loss in performance of the composite. For example, as the swelling stress
associated with the uptake of moisture progresses, so may the rate of stress corrosion of
the glass fibers through an increase in both the content of water and the stress on the
fiber. Tests on macroscopic properties of composites, although they are more difficult to
interpret with regard to interfacial, fiber and matrix effects, are necessary in determining
316 C.L. Schutte / Environmental durability of glass-fiber composites

1.2 I I I I I I

(-

1.0 Vy
~ ~-3%/decade

k~ 0.8 ~ o
5 /'h~o vv •
lO%/decade" []

>
0
0
0.6 . 4

E
E 0.2 • single fiber
"7<
0 [] u n i m p r e g n a t e d s t r a n d
© epoxy-impregnated strand
0.0 i t t t t
0 I 2 3 4 5

Log (cyclesto foilure)


Fig. 11. Plot of the normalized data of S - N fatigue of fibers of E-glass (diameter, 25 p.m): single fiber, strands of 30 fibers
unimpregnated, and strands of 30 fibers impregnated with epoxy [223] (by permission of the publishers, Society of Plastics
Engineers).

how the degradative processes interact. T h e competitions between degradative mechanisms


can make predictions difficult. For example, both mechanical degradation and chemical
corrosion of fibers occurred in the system that Bledzki et ai. studied [197]. Also, the influence
of temperature on both the rate of diffusion of water and the fracture toughness of the
resin that Pritchard and Speake [198] presented illustrates the complicated interactions that
exist for these types of systems. Understanding the detailed mechanisms and how they
interact is essential for accurate predictions of composite performance under conditions of
actual use.

6. Conclusions and future opportunities

The previous sections address the effect of the environment, particularly water, on
each component of the composite: the glass fiber, the resin and the interracial region. These
sections illustrated that many factors contribute to the failure of composites on exposure
to the environment. The dominant m o d e of failure can be due to failure of the glass fibers,
matrix, interface, or a combination thereof. It is necessary, therefore, to understand the
influence that environmental exposure will have on each of the components of the composite.
Understanding the "Achilles' heel(s)" of the materials used for these composites should
allow for the development and improvement of more durable structures. For example, the
chemical composition of the glass will determine its relative durability. Although it is unlikely
to avoid completely this type of damage, by judicious choice of the type of glass one should
be able to decrease the fiber's relative susceptibility.
A d d e d complications exist from swelling stresses that can crack and fail the resin as
well as contribute to the stress corrosion of the glass fiber. The stress on the fiber and
C.L. Schutte / Environmental durability of glass-fiber composites 317

the content of water at the surface of that fiber increase as water diffuses into the sample;
both of these factors will accelerate the corrosion of the glass fiber. Although the influence
of water and stress on the interracial region is not completely understood, results of model
studies suggest that water, swelling stresses and conditions of fatigue will have a significant
influence on the strength of the interracial region. Furthermore, water can not only stress
the resin through swelling, but also potentially hydrolyze the functional groups in the resin
as well. The hydrophilicity of the resin will be one of the determining factors for the
equilibrium content of water, to a first-order approximation. The content of water in the
composite will influence the swelling stresses of the resin and glass fiber. Because these
stresses influence both the potential microcracking in the resin and the stress corrosion of
the fiber, the choice of a more hydrophobic resin could reduce the rate of these associated
processes and the degree to which they contribute to the degradation of the sample.
The choice of coupling agent, the m e t h o d of application, and the morphology of its
coating influence the properties of the interface. The studies characterizing the desorption
characteristics of the coating illustrate methodologies whereby tailoring of the interfacial
morphology is achievable. This general approach to tailoring surfaces allows for the study
of the influence or role of each layer on the strength and durability of that interface/
interphase, by building the morphology and studying the behavior of the intermediate
structure. With the added insight that studies in surface chemistry coupled with mechanical
charactization can give, the selective deposition of the coating can then be optimized for
a given performance requirement. Once the fundamental structure/property relationships
are known for this area, then the task of tailoring or engineering the interfacial region for
both optimal strength and durability characteristics can be addressed.
One major unresolved issue of concern is the difficulty that exists in choosing conditions
for the accelerated tests, and very little attention has been focused on this issue. If harsher
conditions are employed in order to hasten the degradation of samples, the mechanisms
that are dominant under the harsher conditions may not be those controlling the degradation
processes under the conditions of use.
Shortcomings in this area create the opportunities to address the following issues:
1. Fundamental materials limitations: a continuation of studies that address further under-
standing of fundamental mechanisms governing the loss of properties of polymer composites
u p o n exposure to the environment. Identification of the materials limitations, the "Achilles'
heels", of the components should allow for their improvement in terms of mechanical
performance and durability.
2. Lack of understanding of the synergy between contributing mechanisms: for example, the
loss in properties in the resin may affect the properties at an interface that is also
undergoing a disruption of adhesion, and these changes determine how the properties
of the composite degrade.
3. Choice of conditions for accelerated tests: difficulty in assessing the translatability of
accelerated testing in the laboratory to actual conditions of exposure. For example, it
may be that increasing the temperature for environmental exposure does more than just
increase the rate of diffusion of water into composites: the relative rates of competing
mechanisms could change. Estimations of long-term performance based on the dominant
mechanism at high temperature may not accurately predict the true performance when
a different mechanism controls the durability of a composite at low temperature.
4. Development and implementation of predictive models that are not only based on accurate
data on mechanisms of failure but also "simple" enough to address the criteria for choice
318 C.L. Schutte / Environmental durability of glass-fiber composites

by design engineers. Although many of the examples cited here illustrate a focus on the
understanding of the mechanisms that limit the performance of these components and
of how their loss in properties influences the behavior of the remaining components,
reliable predictive models for durability are still needed. Collaborative efforts between
the efforts in research with those in composite development would be of great help to
address these ends. Accurate models for prediction of lifetime performance, therefore,
would be based on the loss in properties based on the most aggressive mechanisms of
failure and their interaction.

Acknowledgements

I am grateful to Dr. Martin Chiang, Polymers Division, NIST, and P.J. Herrera-Franco
for the figures describing the experimental setup for the single-fiber tests.

References

[1] Life Prediction Methodologies for Composite Materials, Committee on Life Prediction Methodology for Composite Materials,
National Research Council, National Materials Advisory Board, National Academy Press, Washington, DC, Oct. 1991.
[2] C.A. Baillie, D.A. Dransfield and Y.W. Mai, Improving the delamination resistance of CFRP by stitching, a review,
Composites Sci. Technol., in press.
[3] L. Babic, C. Dunn and P.J. Hogg, Plastics Rubber Proc. Appl., 12 (1989) 199--207.
[4] C.M.R. Dunn, H. Battjes, E. English, E. Fouch, P.J. Hogg, R.J. Morgan and S. Turner, in A.H. Cardon and G. Verches
(eds.), Durability of Po~,mer Based Composite Systems for Structural Applications, Elsevier, Amsterdam, 1991, pp. 129-148.
[5] C. Arnold-McKenna and G.B. McKenna, J. Res. Natl. lnst. Standards Technol., 98 (4) (1993) 523-533.
[6] S.W. Freiman, J. Geophys. Res., 89 (1984) 4072-4076.
[7] B. Lawn, in Fracture of Brittle Solids, Cambridge University Press, 2nd edn., 1993, Chapter 5.
[8] T.A. Michalske and S.W. Freiman, J. Am. Ceram. Soc., 66 (1983) 284-288.
[9] S.W. Freiman, G.S. White and E.R. Fuller Jr., J. Am. Ceram. Soc., 68 (3) (1985) 108--112.
[10] S.W. Freiman, J. Am. Ceram. Soc., 57 (1974) 350-353.
[11] S.M. Wiederhorn, S.W. Freiman, E.R. Fuller Jr. and C.J. Simmons, J. Mat. ScL, 17 (1982) 3460-3478.
[12] G.S. White, S.W. Freiman, S.M. Wiederhorn and T.D. Coyle, J. Am. Ceram. Soc., 70 (1987) 891-895.
[13] A.G. Evans and S.M. Weiderhorn, Intl. J. Fracture, 10 (3) (1974) 379-392.
[14] S.M. Weiderhorn, in R.C. Brandt, D.P.H. Hasselman and F.F. Lange (eds.), Fracture Mechanics o f Ceramics, Vol. 2,
Plenum Press, New York, 1974, p. 613.
[15] J.E. Ritter Jr., in R.C. Brandt, D.P.H. Hasselman and F.F. Lange (eds.), Fracture Mechanics of Ceramics, Vol. 4, Plenum
Press, New York, 1978, p. 667.
[16] J.E. Ritter Jr. and J.A. Meisel, J. Am. Ceram. Soc., 59 (1976) 478.
[17] S.M. Weiderhorn, A.G. Evans and D.E. Roberts, in R.C. Brandt, D.P.H. Hasselman and F.F. Lange (eds.), Fracture
Mechanics of Ceramics, Vol. 2, Plenum, New York, 1974, p. 829.
[18] S.M. Weiderhorn, A.G. Evans, E.R. Fuller and H. Johnson, J. Am. Ceram. Soc, 57 (1974) 319.
[19] A.G. Evans, S.M. Weiderhorn, M. Linzer and E.R. Fuller, Am. Ceramics Soc. Bull., 54 (1975) 576.
[20] J.E. Ritter Jr. and S.A. Wulf, Am. Ceram. Soc. Bull., 57 (1978) 186.
[21] J.E. Ritter Jr., J.M. Sullivan Jr. and K. Jakus, J. Appl. Phys., 49 (9) (1978) 4779.
[22] K.K. Phani, J. Mat. Sci. Let., 6(12) (1987) 1389-1391.
[23] W.J. Duncan, J. Am. Ceramics Soc., 69 (1986) C-132-C-133.
[24] V.V. Rondinella and M.J. Matthewson, Z Am. Ceram. Soc., 76 (1993) 139-144.
[25] W.D. Bascom, The surface chemistry of moisture-induced composite failure, in E.P. Plueddemann (ed.), Interfaces in
Po~,mer Matrix Composites, Academic Press, New York, 1974, pp. 79-108.
[26] G.K. Schmitz and A.G. Metcalfe, Ind. Eng. Chem. Prod. Res. Dev., 5 (1966) 1-8.
[27] A.J. Barker and T.R. Bott, Trans. Instn. Chem. Engrs., 47 (1969) T212-T221.
[28] O. Ishai, Potyrn. Eng. ScL, 15 (1975) 491-499.
[29] B.D. Agarwal and L.J. Broutman, Analysis and Performance of Fiber Composites, SPE Monograph, Wiley, New York,
1980, p. 269.
[30] P.W. France, W.J. Duncan, D.J. Smith and K.T. Beales, J. Mat. Sci., 18 (1983) 785-792.
C.L. Schutte / Environmental durability of glass-fiber composites 319

[31] J.E. Ritter, T.H. Service and IC Jakus, J.. Arr~ Ceram. Soc., 71 (11) (1988) 988-992.
[32] T.P. Dabbs and B.R. Lawn, J. Am. Ceram. Soc., 68 (11) (1985) 563-569.
[33] J. Aveston, A. Kelly and J.M. Sillwood, in A.R. Bunsell, C. Bathias, M. Martrencher and G. Verchery (eds.), Advances
in Composite Materials: Proc. 3rd Intl. Conf. on Composite Materials, 1980, Vol. 1, Pergamon Press, Oxford, UK, 1980,
pp. 556-568.
[34] C.J. Jones, R.F. Dickson, T. Adam, H. Reiter and B. Harris, Proc. R. Soc. Lond. A, 396 (1984) 315-338.
[35] C.J. Jones, R.F. Dickson, T. Adam, H. Reiter and B. Harris, Composites, 14 (1983) 288- 293.
[36] J. Mandell, D. Huang and F.J. McGarry, Proc. 35th Ann. Tech. Conf. of the Reinforced Plastics~Composites Institute of
SPI, New Orleans, LA, 1980, paper 19-A, Society of Plastics Industry, Washington, DC, 1980.
[37] A.G. Metcalfe and G.R. Schmitz, Glass Technol., 13 (1972) 5-16.
[38] G.D. Sims and D.G. Gladman, Plast. Rubber Mater. AppL, 1 (1980) 122-128.
[39] G.D. Sims and D.G. Gladman, Plast. Rubber Mater. Appl., 1 (1978) 41-48.
[40] K.H. Boiler, Mod. Plast., 41 (June 1964) 145-150,188.
[41] B. Noble, S.J. Harris and M.J. Owen, J. Mat. Sci., 18 (1983) 1244-1254.
[42] H. Lilholt and R. Talreja (eds.), Fatigue and Creep of Composites Materials, Riso National Laboratories, Riso, Norway,
1981.
[43] T.S. Ellis and F.E. Karasz, Polymers, 25 (1984) 664-669.
[44] S.W. Tsai, in F.W. Wendt, H. Liebowits and N. Perrone (eds.), Mechanics of Composite Materials, Pergamon Press,
Oxford, UK, 1970, p. 749.
[45] K.H.G. Ashbee and R.C. Wyatt, Proc. R. Soc. Lond., 312A (1969) 553-564.
[46] T.C. Wong and L.J. Broutman, Polym. Eng. Sci., 25 (9) (1985) 521-534.
[47] I. Ghorbel and D. Valentin, Polym. Composites, 14 (4) (1993) 324-334.
[48] H.G. Carter and K.G. Kibler, J. Compos. Mat., 12 (1978) 118-131.
[49] C. Carfagna, A. Apicella and L. Nicolais, J. Appl. Poly. Sci., 27 (1982) 105-112.
[50] A. Apicella, L. Nicolais and C. Carfagna, in J.C. Seferis and L. Nicolais (eds.), The Role of the Polymeric Matrix in the
Processing and Structural Properties of Composite Materials, Plenum Press, New York, 1981, pp. 215-229.
[51] A. Apicella and L. Nicolais, Ind. Eng. Sci, 20 (1981) 138.
[52] M.L. Williams, R.F. Landel and J.D. Ferry, Z Am. Cherr~ Soc., 77 (1955) 3701-3707.
[53] H.G. Carter and K.G. Kibler, J. Compos. Mat., 11 (1977) 265-275.
[54] A.F. Johnson, Engineering Design Properties of GRP, British Plastics Federation, London, UK, 1979.
[55] A. Apicella, C. Migliaresi, L. Nicolais, L. Iaccarino and S. Roccotelli, Composites, 14 (1983) 387-392.
[56] J. Illinger and J.F. Sprouse, Am. Chem. Soc. Org. Coat. Plast. Chem., 175th ACS Meeting, 38 (1978) 497-502.
[57] A.N. Netravali, R.E. Fumes, R.D. Gilbert and J.D. Memory, Z Appl. Polym. Sci., 30 (1985) 1573-1578.
[58] A.S. Wimolkiatisak and J.P. Bell, Polym. Composites, 10 (1989) 162-172.
[59] T. Alfrey Jr., E.F. Gurnee and W.G. Lloyd, J. Polym. Sci., Part C, 12 (1966) 249-261.
[60] D. Cohn and G. Marom, in J.C. Seferis and L. Nicolais (eds.), The Role of the Polymeric Matrix in the Processing and
Structural Properties of Composite Materials, Plenum Press, New York, 1981, pp. 245-259.
[61] D. Cohn and G. Marom, Z Appl. Polym. Sci., 28 (1983) 1981-1992.
[62] D. Cohn and G. Marom, Polynt Eng. Sci., 22 (1982) 870-877.
[63] D. Cohn and G. Marom, Polymer, 24 (1983) 223-228.
[64] Y. Diamant, G. Marom and L.J. Broutman, Z Appl. Polym. Sci, 26 (1981) 3015-3025.
[65] M.K. A n t o o n and J.L. Koenig, J. Polytr~ Phys., 19 (1981) 197-212.
[66] A. Garton, J.L. Henry, P.D. McLean and W.T.K. Stevenson, Polym. Eng. Sci., 29 (1989) 23-28.
[67] C. G'Sell and G.B. McKenna, Polymer, 33 (10) (1992) 2103-2114.
[68] G.B. McKenna, Polym. Prepr., Am. Chem. Soc., Div. Polym. Chem., 33 (1) (1992) 364-365.
[69] J.L. Sullivan, E.J. Blais and D. Houston, Composite Sci. Tech., 47 (1993) 389-403.
[70] E.P. Plueddemann, Silane Coupling Agents, Plenum Press, New York, 2nd edn., 1991.
[71] H. Ishida and J.L. Koenig, Polym. Eng. Sci., 18 (1978) 128-145.
[72] H. Ishida, Polym. Composites, 5 (1984) 101-123.
[731 H. Ishida and J.L. Koenig, J. Colloid Interface Sci., 64 (1978) 555-564.
[74] J.L. Koenig and P.T.K. Shih, J. Colloid Interface Sci., 36 (1971) 247-253.
[75] H. Ishida and J. Koenig, J. Colloid Interface Sci., 64 (1978) 565-576.
[76] H. Ishida and J.L. Koenig., J. Polym. Sci., Polytr~ Phys. Edn., 17 (1979) 1807-1813.
[77] O.K. Johannson, F.O. Stark, G.E. Vogel and R.M. Fleischman, Z Compos. Mat., 1 (1967) 278-292.
[78] Y.T. Liao, Polym. Composites, 10 (6) (1989) 424-428.
[79] M.E. Schrader, Radioisotope studies of coupling agents at the interface, in E.P. Plueddemann (ed.), Interfaces in Polymer
Matrix Composites, Academic Press, New York, 1974, pp. 109-129.
[80] M.E. Schrader, I. Lerner and F.J. D'Oria, Mod. Plastics, 45 (1967) 195.
[81] M.E. Schrader and A. Block, J. Pob~'n. Sci. C, 34 (1971) 281-291.
[82] J.L. Koenig, P.T.K. Shih and P. Lagally, Mater. Sci. Eng., 20 (1975) 127- 135.
[83] H. Ishida and J.L. Koenig, Appl. Spectroscopy, 32 (1978) 462-479.
320 C.L. Schutte / Environmental durability of glass-fiber composites

[84] H. Ishida and J.L. Koenig, Z Polym. Sci., Poem. Phys. Edn., 18 (1980) 233-237.
[85] D. Wang and F.R. Jones, Z Mat. Sci., 28 (1993) 2481-2488.
[86] N. Nishiyama, R. Shick and H. Ishida, Z Colloid Interface Sci., 143 (1991) 146-156.
[87] J.D. Miller and H. Ishida, Langmuir, 2 (1986) 127-131.
[88] J. Miller and H. Ishida, 3". Chem. Phys., 86 (1987) 1593-1600.
[89] H. Ishida and J.D. Miller, Macromolecules, 17 (1984) 1659-1666.
[90] A.T. DiBenedetto and D.A. Scola, Z Colloid Interface Sci., 64 (1978) 480-500.
[91] G. Lubin, Handbook of Composites, Van Nostrand Reinhold, New York, 1982, p. 139.
[92] S. Naviroj, S.R. Ctdler, J.L. Koenig and H. Ishida, J. Colloid Interface ScL, 96 (1983) 308-317.
[93] S.R. Culler, S. Naviroj, H. Ishida and J.L. Koenig, Z Colloid Interface ScL, 96 (1983) 69-79.
[94] S. Naviroj, J.L. Koenig and H. Ishida, Z MacromoL Sci.-Phys., B22 (1983) 291-304.
[95] D.J. Tutas, R. Stromberg and E. Passaglia, SPE Trans., 4 (1964) 256-262.
[96] C. Chiang and J.L. Koenig, J. Colloid Interface Sci., 83 (1981) 361-370.
[97] W.D. Bascom, Macromolecules, 5 (1972) 792-798.
[98] G. Wong, J. Adhes., 4 (1972) 171-179.
[99] D.J. Belton, A. Joshi, in H. Ishida and G. Kumar (eds.), Molecular Characterization of Composite Interfaces, Plenum
Press, NY, 1983, pp. 187-204.
[100] D.W. Dwight, F.M. Fowkes, D.A. Cole, M.J. Kulp, PJ. Sabat, L. Salvati and T.C. Huang, J. Adhes. ScL TechnoL, 4
(1990) 619-632.
[101] F.M. Fowkes, D.O. Tischler, J.A. Wolfe, L.A. Lannigan, C.M. Ademu-John and M.J. Halliwell, Z Poem. Sci., Po~ym.
Chem. Edn., 22 (1984) 547-566.
[102] F.M. Fowkes, J. Adhes., 4 (1972) 155-159.
[103] F.M. Fowkes, M.B. Kaczinski and D.W. Dwight, Langmuir, 7 (1991) 2464-2470
[104] F.M. Fowkes, Z Adhes. ScL Technol., 1 (1986) 1.
[105] R.S. Drago, G.C. Vogel and T.E. Needham, Z Am. Chem. Soc., 93 (1971) 6014.
[106] F.M. Fowkes, D.W. Dwight, D.A. Cole and T.C. Huang, J. Non-Cryst. Solids, 120 (1990) 47-60.
[107] H. Ishida and J.L. Koenig, Z Po~yr~ Sci., Polym. Phys. Edn., 18 (1980) 1931-1943.
[108] H. Ishida, S. Naviroj and J.L. Koenig, in K.T. Mittal (ed.), PhysiochemicalAspects of Po~mer Surfaces, Vol. 1, Plenum
Press, New York, 1983, pp. 91-104.
[109] H. Ishida and J . L Koenig, Z Polyn~ Sci., Polym. Phys. Edn., 17 (1979) 615- 626.
[110] N. Nishiyama, T. Ishizaki, K. Horie, M. Tomari and M. Someya, J. Biomed. Mat. Res., 25 (1991) 213-221.
[111] Y. Suzuki, Z. Maekawa, H. Hamada, A. Yokoyama, T. Sugihara and M. Hojo, J. Mat. Sci., 28 (1993) 1725-1732.
[112] N. Ikuta, Y. Suzuki, Z. Maekawa and J. Hamada, Polymers, 34 (1993) 2445-2446.
[113] N.H. Sung, A. Kaul, I. Chin and C.P. Sung, Po~ym. Eng. Sci., 22 (1982) 637-644.
[114] A. Kaul and N.H. Sung, Po~,n. Eng. Sci., 25 (1985) 1171-1178.
[115] M.K. Chaudhury, T.M. Gentle and E.P. Plueddemann, Z Adhes. Sci., 1 (1987) 29-38.
[116] R.T. Graf, J.L. Koenig and H. Ishida, J. Adhes., 16 (1983) 97-114.
[117] M. Narkis and E.J.H. Chen, Polym. Composites, 9 (1988) 245-251.
[118] L.T. Drzal and PJ. Herrera-Franco, Engineered Materials Handbook, VoL 3: Adhesives and Sealants, ASM International,
Materials Park, OH, 1990, pp. 391-405.
[119] C.A. Baillie and A. Buxton, Composites (special issue dedicated to the papers presented at Interracial Phenomena in
Composite Materials '93, 13-15 Sept. 1993, Cambridge, UK), in press.
[120] M. Narkis, J.H. Chen and R.B. Pipes, Polym. Composites, 9 (4) (August 1988) 245-251.
[121] P.J. Herrera-Franco and LT. Drzai, Composites, 23 (1) (1992) 2-27.
[122] W.G. McDonough, P.J. Herrera-Franco, W.L. Wu, LT. Drzal and D.L Hunston, 23rd Intl. SAMPE Tech. Conf., Oct
21-24, Society for the Advancement of Materials and Process Engineering, Covina, CA, pp. 247-258.
[123] W.A. Fraser, F.H. Ancker, A.T. DiBenedetto and B. Elbirli, Po~ym. Composites, 4(4) (1983) 238-248.
[124] A. Kelly and W.R. Tyson, Z Mech. Phys. Solids, 13 (1965) 329-350.
[125] A.T. DiBenedetto, Composites Sci. Technol., 42 (1991) 103-123.
[126] M.R. Piggott, Composites ScL Technol., 30 (1987) 295-306
[127] M.R. Piggott and S.R. Dai, Composites ScL Technol., 31 (1988) 15-24.
[128] J. Mullin, J.M. Berry and A. Gatti, Z Compos. Mat., 2 (1968) 82-103.
[129] M.R. Piggott, Composites ScL Technol., 30 (1987) 295-306.
[130] W.A. Fraser, F.H. Ancker, A.T. DiBenedetto and B. Elbirli, Po~ym. Composites, 4 (1983) 238-248.
[131] A.T. DiBenedetto and P.J. Lex, Po~,m. Eng. Sci., 29 (8) (1989) 543-555.
[132] H.D. Wagner and A. Lustinger, Composites (special issue dedicated to the papers presented at Interfacial Phenomena
in Composite Materials '93, 13-15 Sept. 1993, Cambridge, LIK), 25 (7) (1994) 613-616.
[133] J. Mullin, J.M. Berry and A. Gatti, Z Compos. Mat., 2 (1) (1968) 82-103.
[134] J.C. Figneroa, L.S. Schadler and C. Laird, in C. Pantano and E.J.H. Chen (eds.), Interfaces in Composites, MRS Symp.
Proc., 170, Materials Research Society, Pittsburgh, PA, 1990, pp. 65-70.
[135] L.S. Schadler, C. Laird and J.C. Figueroa, Z Mat. Sci., 27 (1992) 4035-4044.
C.L. Schutte / Environmental durability of glass-fiber composites 321

[136] L.S..Schadler, J.C. Figueroa and C. Laird, in C. Pantano and E.J.H. Chen (eds.), Interfaces in Composites, MRS Symp.
Proc., 170, Materials Research Society, Pittsburgh, PA, 1990, pp. 345-350.
[137] A.T. Dibenedetto, J.A. Gomez, C. Schilling, F. Osterholtz, G. Haddad, in C. Pantano and E.J.H. Chen (eds.), Interfaces
in Composites, MRS Syrup. Proc., 170, Materials Research Society, Pittsburgh, PA, 1990, pp. 297-302.
[138] J.A. Gomez and J.A. Kilgour, 48th Ann. Conf., Composite Institute, The Society of the Plastics Industry, Inc., Feb. 1993,
Cincinnati, OH, Washington, DC, 1993.
[139] C. Schutte, W. McDonough, M. Shioya, M. McAuliffe and M. Greenwood, Composites (special issue dedicated to the
papers presented at Interfacial Phenomena in Composite Materials "93, 13-15 Sept. 1993, Cambridge, UK), 25 (7) (1994)
617-624.
[140] C. Schutte, W. McDonough, M. Shioya and D. Hunston, in R.L. Opila, F.J. Boerio and A.W. Czanderna (eds.),
Polymer~Inorganic Interfaces, MRS Proc., 304, Materials Research Society, Pittsburgh, PA, 1993, pp. 43-48.
[141] L. Drzal, M. Rich and M. Koenig, J. Adhes., 18 (1985) 49-72.
[142] L.T. Drzal, M.J. Rich and P.F. Lloyd, J. Adhes., 16 (1983) 133-152.
[143] L.T. Drzal, in C. Pantano and E.J.H. Chen (eds.), Interfaces in Composites, MRS Syrup. Proc., 170, Materials Research
Society, Pittsburgh, PA, 1990, pp. 275-283.
[144] D.W. Dwight, P.J. Sabat and H.F. Brinson, 4th Tech. Conf. on Composite Matls., October 3-6, 1989, Virginia Polytechnic
Institute, Blacksburg, VA.
[145] H. Emadipour, P. Chiang and J.L. Koenig, Res. Mechanica, 5 (1982) 165-176.
[146] J. Koenig and H. Emadipour, Polym. Composites, 6 (1985) 142-150.
[147] M.R. Piggott, in C. Pantano and E.J.H. Chen (eds.), Interfaces in Composites, MRS Symp. Proc., 170, Materials Research
Society, Pittsburgh, PA, 1990, pp. 265-273.
[148] R.A. Latour Jr., A. Robert, J. Black and B. Miller, Surf. lnterf AnaL, 17 (1991) 477-484.
[149] R.A. Haaksma and M.J. Cehelnik, in C. Pantano and E.J.H. Chen (eds.), Interfaces in Composites, MRS Syrup. Proc.,
170, Materials Research Society, Pittsburgh, PA, 1990, pp. 71-76.
[150] R.A. Latour Jr. and J. Black, J. Biomed. Mat. Res., 26 (1992) 596-606.
[151] R.A. Latour Jr., In vitro simulation of in vivo dynamic micromechanical failure of structural composite biomaterials,
Doctoral Dissertation, University of Pennsylvania, Philadelphia, PA, Univ. Microfilms Intl. DA9015125, Ann Arbor, MI,
Dec. 1989.
[152] M.R. Meyer, R.A. Latour Jr. and H.D. Shutte, Proc. Composites Materials Mechanics and Processing (8th Technical
Conference, Am. Soc. for Composites, Oct. 8-21, 1993, Cleveland, Ohio, 1993), Technomic Publishing Co. Inc., Lancaster,
PA, 1993, pp. 20-29.
[153] U. Gaur and B. Miller, Polym. Composites, 11 (4) (1990) 217-222.
[154] B.S. Mehta, A.T. DiBenedetto and J.L. Kardos, J. Appl. Polym. Sci., 21 (1977) 3111-3127.
[155] A.C. Loos, G.S. Springer, B.A. Sanders and R.W. Tung, in G.S. Springer (ed.), Environmental Effects of Composite
Materials, Technomic Publishing Co. Inc., Westport, CT, 1981, pp. 51-63.
[156] W.D. Bascom, in E.P. Plueddernann (ed.), Interfaces in Polymer Matrix Composites, Vol. 6, Academic Press, New York,
1974, pp. 79-108.
[157] A. Pollard, R. Baggott, G.H. Wostenholm and B. Yates, J. Mat. Sci., 24 (1989) 1665-1669.
[158] G. Dukes, Proc. Int. Conf. on Carbon Fibres, their Composites and Applications, The Plastics Institute, London, Feb.
2-4, 1971, paper 26.
[159] A.A. Fahmy and J.C. Hurt, Polym. Composites, 1 (1980) 77.
[160] C.H. Shen and G.S. Springer, J. Compos. Mat., 10 (1976) 2-20.
[161] A. Avena and A.R. Bunsell, Composites, 19 (1988) 355-357.
[162] S. Neumann and G. Marom, J. Compos. Mat., 21 (1987) 68-80.
[163] S. Neumann and G. Marom, Z Mat. Sci., 21 (1986) 26-30.
[164] G. Marom and L.J. Broutman, J. Adhes., 12 (1981) 153-164.
[165] S. Neumann and G. Marom, Polym. Composites, 6 (1) (1985) 9-12.
[166] Y. Weitsman, Moisture in composites: sorption and desorption, in K.L. Reifsnider (ed.), Fatigue of Composite Materials,
Elsevier, Amsterdam, Netherlands, 1991, pp. 385-429.
[167] K.A. Kasturiarachchi and G. Pritchard, Composites, 14 (1983) 244-250.
[168] A.C. Loos and G.S. Springer, in G.S. Springer (ed.), Environmental Effects on Composite Materials, Technomic Publ.
Co. Inc., Westport, CT, 1981, pp. 51-62.
[169] G.S. Springer, B.A. Sanders and R.W. Tung, in G.S. Springer (ed.), Environmental Effects on Composite Materials,
Technomic Publ. Co. Inc., Westport, C-'T, 1981, pp. 126-144.
[170] T. Marzi, U. Schroder, M. Heb and R. Kosfeld, in C. Pantano and E.J.H. Chen (eds.), Interfaces in Composites, MRS
Syrup. Proc., 170, Materials Research Society, Pittsburgh, PA, 1990, pp. 123-127.
[171] J.G. Williams, J. Mat. Sci., 17 (1982) 1427-1433.
[172] J.G. Williams, J. Appl. Polym. Sci., 23 (1979) 3433-3444.
[173] M. Schrager, J. Polym. Sci., 8 (1970) 1999.
[174] J.C. Patterson-Jones and D.A. Smith, Jr. Appl. Polym. Sci., 12 (1968) 1601.
[175] G.A. Pogany, Polymer, 11 (1970) 66.
322 C.L. Schutte / Environmental durability of glass-fiber composites

[176] M.P. Ebdon, O. Delatycki and J.G. Williams, J. Pobvn. Sci., Polynt Phys. Edn., 12 (1974) 1555.
[177] T. Morii, H. Hamada, Z. Maekawa, T. Tanimoto, T. Hirano and K. Kiyosume, Composite Struct., 25 (1993)
95-100.
[178] T. Morii, T. Tanimoto, H. Hamada, Z. Maekawa, T. Hirano and K. Kiyosume, Composites Sci. Technol., 49 (1993)
209-216.
[179] T. Morii, T. Tanimoto, Z. Maekawa, H. Hamada, K. Kiyoshumi and H. Ichihashi, Po~mer Potym. Composites, 1 (2)
(1993) 113-120.
[180] N.R. Farrar and K.H. Ashbee, Z Phys. D: Appl. Phys., 11 (1978) 1009-1015.
[181] A. Chateauminois, B. Chabert, J.P. Soulier and L. Vincent, Composites, 24(7) (1993) 547-555.
[182] M. Stoudt, E. Escalante and R.E. Ricker, Ceramic Trans., 19 (1991) 993-1000.
[183] O. Ishai, Poem. Eng. ScL, 15 (1975) 486--490.
[184] O. Ishai, Po(ym. Eng. Sci., 15 (1975) 491-499.
[185] O. Ishai and A. Mazor, Rheol. Acta, 13 (1974) 889-902.
[186] O. Ishai, in The effect of environmental-loading history on mechanical behavior of fiber-reinforced polymeric composites,
TDM 74-08, 1974 (Technical Research and Development Foundation Ltd., Technion -- Department of Mechanics,
Israel Institute of Technology).
[187] F.R. Jones, J.W. Rock and A.R. Wheatley, Jr. Mat. Sci. Lett., 2 (1983) 519-521.
[188] F.R. Jones, J.W. Rock, A.R. Wheatley and J.E. Bailey, Reinforced Plastics Congress, Brighton, UK, 1982, British Plastics
Federation, London, UK, paper 32.
[189] F.R. Jones, J.W. Rock and J.E. Bailey, J. Mat. ScL, 18 (1983) 1059-1071.
[190] F.R. Jones, J.W. Rock and A.R. Wheatley, Composites, 14 (1983) 262-269.
[191] S. Lieblein, Survey of long term durability of fiberglass-reinforced plastic structures, Contract Report C-39549oD, January,
1981, (NASA).
[192] M.G. Phillips, Composites, 14 (1983) 270-275.
[193] B. Ellis and M.S. Found, Composites, 14 (1983) 237-243.
[194] O. Ishai and U. Amon, in J.R. Vinson (ed.),Advanced Composite Materials -- EnvironmentalEffects, STP 658, American
Society for Testing and Materials, Philadelphia, PA, 1977.
[195] R. Gopalan, B.R. Somashekar and B. Dattaguru, Polym. Degrad. Stability, 24 (1989) 361-371.
[196] K.K. Phani and N.R. Bose, Composite Sci. Technol., 29 (1987) 79-87.
[197] A. Bledski, R. Spaude and G.W. Ehrenstein, Composite ScL TechnoL, 23 (1985) 263-285.
[198] G. Pritchard and S.D. Speake, Composites, 19 (1) (1988) 29-36.
[199] G.P. Marshall, M. Kisbenyi, D. Harrison and R. Pinzelli, 37th Ann. Conf., January 11-15, 1982, Reinforced Plastics/
Composites Institute, The Society of the Plastics Industry, Inc., paper 9-D.
[200] H.H. Collins, Plast. Rubber: Mat. Appl., 3 (1) (1978) 6-10.
[201] P.J. Hogg, Composites, 14 (3) (1983) 254-261.
[202] D. Hull and P.J. Hogg, in A.R. Bunsell, C. Bathias, A. Martrenchar and G. Verchery (ed.), Advances in Composite
Materials: Proc. 3rd Int. Conf. on Composite Materials, Paris, 1980, Vol. 1, Pergamon Press, Oxford, UK, 1980, pp.
543-555.
[203] T. Juska, Effect of water immersion on fiber/matrix adhesion, CarderockDiv-SME-92/38, January 1993 (Ship Materials
Engineering Department Research and Development Report, US Navy, Carderock Division, Naval Surface Warfare
Center, Bethesda, MD).
[204] T. Juska, Effect of water immersion on fiber/matrix adhesion in thermoplastic composites, CarderockDiv-SSM-93/02,
March 1993 (Survivability, Structures and Materials Directorate Technical Report, US Navy, Carderock Division, Naval
Surface Warfare Center, Bethesda, MD).
[205] C.M. Yeager, C.A. Carreno and D.W. Rogge, Automotive Polym. Design (August 1991) 2-7.
[206] J. Haarsma, Po~,ra. Eng. Sci., 19 (1979) 350-352.
[207] D. Houston and C.F. Johnson, Advanced Composite Materials: New Developments and Applications, Proc. 7th Ann. ASM/
ESD Advanced Composites Conf., 30 Sept.-3 Oct., 1991, Detroit, MI, ASM International, Materials Park, OH, 1991,
pp. 249-257.
[208] A.F. Johnson, Composites, 17 (1986) 233-239.
[209] D. O'Grady, M. Buggy and C. Birkinshaw, Key Eng. Matls., 32 (1989) 17-24.
[210] M.J. Owen and R.J. Cann, Jr. Mat. Sci., 14 (1979) 1982-1996.
[211] H. Sekine, K. Shimomura, N. Hamana, JSME Intl. Z, 31 (1988) 619-626.
[212] J.B. Romans, A.G. Sands and J.E. Cowling, Fatigue behavior of glass-filament wound composites in water, NRL Rep.
7246, 1971 (Naval Res. Lab., Washington, DC).
[213] C.E. Knox, in G. Lubin (ed.), Handbook of Composites, Van Nostrand Reinhold, New York, 1982, p. 139.
[214] J.F. Mandell, Polym. Eng. Sci., 19 (1979) 353-358.
[215] J.F. Mandell, F.J. MeGarry, W.D. Barton and R.P. Demchik, Proc. 31st Conf. SPI Reinforced Plastics~Composites lnst.,
1976, Paper 21-B.
[216] J.F. Mandell, D.D. Huang and F.J. McGarry, Composites TechnoL Rev., 3 (1981) 96-102.
[217] C.K.H. Dharan, J. Mat. Sci., 10 (10) (1975) 1665-1670.
C.L. Schutte / Environmental durability of glass-fiber composites 323

[218] J.F. Mandell and U. Meier, in Fatigue of Composite Materials, STP 569, American Society for Testing and Materials,
Philadelphia, PA, 1975, pp. 28--52.
[219] J.F. Mandell, D.D. Huang and F.J. McGarry, in Proc. 35th Ann. Conf. of RP/C Institute, Society of the Plastics Industry,
New York, Society of the Plastics Industry Inc., Washington, DC, 1980, paper 20D.
[220] J.F. Mandell, in Composite Reliability, STP 580, American Society for Testing and Materials, Philadelphia, PA, 1975,
pp. 515-527.
[221] H.T. Hahn and R.Y. Kim, J. Compos. Mat., 10 (1976) 156-180.
[222] R.J. Howe and M.J. Owen, in Proc. 8th Int. Reinforced Plastics Conf., Brighton, UK, 1972, British Plastics Federation,
London, UK, paper 21.
[223] J.F. Mandell, F.J. McGarry, A.J.-Y. Hsieh and C.G. Li, Polym. Composites, 6 (1985) 168-174.

S-ar putea să vă placă și