C-P10 Summary of Composite Creep v2

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

A Review of Creep in Carbon and Aramid Fiber Composites as

Applicable to SKA Designs for Radio Telescopes

Prepared for: National Research Council Canada, Dominion Radio Astrophysical Observatory
Date: November 2009
Authors: Dr. Bree Sharratt and Geoffrey Wood, Profile Composites Inc.

1.0 Relevance to the SKA Composite Radio Telescope Designs


Creep issues can develop in fiber-reinforced polymer composites when design requirements
and materials selection place either excessive loads on the fibers, such as aramid fibers, or on
resin matrices unsupported by reinforcement. Creep issues were studied relative to the
composite radio telescope designs proposed for SKA by National Research Council Canada –
Dominion Radio Astrophysical Observatory, for the 12-meter and 15-meter baseline
approaches. The analysis included review of the relevant background technical literature of
carbon fiber and aramid fiber reinforced polymer composites, and comparison of the findings in
the literature to the proposed designs.

Creep can become an issue under long-term loading, and is exacerbated by high temperatures.
With respect to the fiber materials, it is especially important in situations where loads are a
significant fraction of the ultimate fiber failure load in the case of aramid (e.g. Kevlar™) fiber, but
is generally not a factor in carbon fiber even at high fractions of ultimate strength or at high
temperatures. With respect to the polymer matrix, creep may present an issue when load is
applied in directions unsupported by fiber reinforcement and hence can affect potentially both
carbon fiber and aramid fiber composites. The two primary load directions that need to be
recognized as a possible source of creep concerns are shear loading on interfaces, and
through-thickness tension or compression loads perpendicular to the fiber reinforcement. The
analysis looks at the SKA radio telescope design stresses as a percentage of ultimate fiber
strength, at strains in the directions where creep could become an issue, and at the design and
construction approach for stresses in the polymer matrix that are not immediately obvious.

The designs for SKA radio telescopes operating in the desired frequency bands are heavily
dominated by stiffness requirements, and also by the need to have a surface stable enough at
all operating conditions to not degrade the signal quality. Designs were developed by NRC-
DRAO based on anticipated operating conditions including gravity loads under the various
operating orientations, at wind loads during operation and stowage, and at temperatures
expected in the dry, desert conditions of South Africa or Australia. For any composite material,
the material properties are developed as a function of the design and manufacturing methods,
and hence properties were estimated based on test data and known material properties from
similarly constructed composites. With a 12-meter baseline design, and a composite fabrication
approach that ensured the reflector dish surface tolerance of 0.5 mm RMS (as demonstrated in
the 10-m Mk2 CART telescope), finite element analysis was used by DRAO to determine the
local strains in the reflector dish and the support structural beams. SKA performance
requirements resulted in very high global and local structural stiffness. This was attained by
sandwich panel design of composite skins over a structural foam core, and thus keeping the
fibers at an extremely low strain relative to their ultimate strain. Similarly, the shear strains were
very low in the resin interfacial regions due to the overall thickness of the composite sandwich
structure. The long-term nature of creep behavior is such that the study looked at the typical
gravity and wind-based operating loads from the SKA design operating conditions, and not the
momentary loads associated with stowing in extreme wind cases. Comparison was made of the
operating stresses in the aramid composite with the ultimate strength of the materials, and it is
seen that the NRC-DRAO design is under 1% of the ultimate strength. Hence, in the case of
aramid fiber composites, a prediction of the median stress-rupture life of the telescope indicates
greater than 106 year lifespan based on creep alone.

The baseline designs of both the carbon and aramid composite components will use fibers
oriented in all loading directions. The current design philosophy of a quasi-isotropic structure
using stitched fabric materials minimizes any interlaminar shear on the resin system, and keeps
the resin strain well within creep parameters for the telescopes. This is enhanced by the
sandwich structure design which develops low total strain in both global and local regions.
Strain levels in the carbon fiber beams and support structures are of the order of 1.1% of
ultimate, so these too are stiffness dominated designs that exhibit no creep issues.

The next area of potential concern then becomes feed-leg support structures, and the
transmission of loads from the data-collection system into the feed-legs, and thence through the
structure. This design is currently all carbon-fiber composite, selected for reasons of very high
dimensional stability and minimum cross-sectional areas, and is a tubular section with an outer
fairing. The loads in the feed-leg structure are transmitted directly to the reflector backside
support beams and not through the reflector surface. Although the stresses in the feed-legs will
be large compared to the ultimate strength, the fabrication with carbon fiber composites negates
any potential creep effects.

The one area of concern then becomes the interface between the hub and yoke assembly.
Although the hub is a carbon fiber composite, the nature of the load is such that creep issues
could develop in the resin matrix if not correctly accounted for. The hub would likely be
mounted to the yoke assembly through means of multiple bolts fastened to a bearing ring.
Typical designs would see relatively large loads on the structure taken up via through-thickness
stresses on the hub composite material. This approach places loads on the resin (through-
thickness properties will have matrix-dominated conditions) that could result in creep behavior in
operation. Cost limitations for both the reflector design as well as the bearing will very likely
result in a minimum diameter bearing used that can meet SKA accuracy and durability
requirements. The smaller the bearing diameter, the larger the concentrated stresses
transmitted across the hub-to-bearing interface. The SKA design will require specific attention
to this location and to materials and manufacturing designs to mitigate the creep issue both on
the resin (through-thickness) and bolt-hole boundaries. Several approaches are possible to
mitigate any potential shortfall in SKA composite telescope designs, and the ones
recommended for this approach to avoid over-building and excessive cost are as follows:
• Use of a three-dimensional (3D) stitched or braided preform of carbon fibers as the base
of the hub structure and interface with the yoke bearing ring.
• Use of a “trapped hub” where the composite hub nests into a recess in the bearing ring
and loads are transmitted over a large surface area rather than through discrete bolts.
Both of these approaches provide fiber reinforcement in all load directions, and can
accommodate loads seen during operating conditions where bolt compression may be
increased or decreased through orientation (gravity based) as well as wind loads. In each case,
the design requirements can be met by ensuring stresses are kept low and are oriented in the
fiber directions.

It is recommended that specific design, materials testing, and component testing be carried out
for a representative hub-to-bearing SKA design, and that design guidelines be developed for
this critical area. With an appropriate design the local temperature should not pose an issue for
this region but would need to be considered in the above recommended design guidelines
development. As will be seen, moisture can have an effect on the creep behavior, similar to that
of elevating the temperature of the composite. Hence, the hub-to-bearing interface design and
hub materials development should look to any potential localized moisture effects, and the
design should eliminate possibilities of trapped moisture in these regions.

The following sections expand on the literature review performed as background to determine
any potential areas of concern with creep of the SKA radio telescope designs, and identify the
areas in which creep can become an issue. It gives the background for creep in composite
constituent materials and references for work done in evaluating composite creep behavior.

2.0 Literature Review

Durability, i.e. good long-term performance, of composites is particularly important in the civil
infrastructure and marine sectors, as evidenced by research funding emanating from the Office
of Naval Research (ONR) (e.g. (Cain et al. 2006; Raghavan et al. 1997)), the US Federal
Highway Administration (e.g. (Karbhari et al. 2003)), the European Commission (e.g. (Canning
et al. 2007)), etc. Long-term performance of such structures and component parts is influenced
by the extent of and resistance to creep, which can result from long-term degradation of the
material and lead to failure at stress levels below the ultimate tensile strength (UTS). Creep
under sustained stress is of special concern in situations where stresses are applied in a matrix-
dominated direction, especially when the ambient temperature is elevated. Other factors that
can accelerate or enhance creep include the presence of moisture or other chemical species,
thermal cycling, UV exposure, and/or loading conditions that result in shear or compression
along the fiber direction (e.g. bending) (MIL-HDBK-17-1F 2002). Since simple application of a
UV-protective coating can alleviate the effects of UV exposure, this aspect is not addressed
here. The following provides an overview of long- and short-term creep of fiber-reinforced
polymer composites, including the aforementioned factors, commencing with a short
introduction to the creep response of the constituent materials.

2.1 Viscoelastic creep of constituent materials

Creep is a time-dependent permanent strain that evolves under a constant stress. Generally
measured as time-dependent strain, with reference to Fig. 1, this viscoelastic process is
characterized by an initial elastic deformation (often left off creep curves), then a region of
primary or transient creep during which the creep rate decreases, then a region of secondary or
steady-state creep where the creep rate is relatively constant, and finally a third region of
tertiary creep or creep rupture where the creep rate accelerates until failure occurs (Wolff 1989).

Figure 1: A general schematic of strain or creep compliance as a function of test time on a linear plot.

Carbon fibers are generally considered to be "creep-resistant" and exhibit little to no chemical-
induced strength degradation (Karbhari et al. 2003). While the nature of the viscoelastic
response, i.e. linear versus non-linear, of aramid fibers has been under dispute (Burgoyne and
Alwis 2008), the fibers are susceptible to creep (Burgoyne and Alwis 2008). Burgoyne and Alwis
note that the source of the dispute between linear and non-linear viscoelastic behavior may be
attributable to the stiffening of fibers at ~50% of their short-term strength as observed by
Chambers (Chambers 1986) and therefore performed a comprehensive study of creep and
relaxation of 1000-filament Kevlar 49 yarns over a range of loads. The authors confirmed that
creep is non-linear for loads below 40% of the average fiber break load and linear with little
change in the creep compliance for loads greater than 40% of the average fiber break load (Fig.
2). They further observed that the greatest creep compliance is observed at the lowest levels of
load for the longest test time. For example, at the lowest load and longest time, the creep strain
can be estimated at 0.14% after 800 hours, which is significantly lower than measured creep
compliances for aramid/epoxy composites as discussed below, but could have bearing on any
comparisons to carbon/polymer composties. Insofar as steady-state behavior is concerned,
Maksimov and Plume observed that the most significant creep occurs in aramid fibers in the first
few months of testing and that, while after 5 years creep had slowed considerably, the fibers did
not demonstrate a true equilibrium creep strain (Fig. 3) (Maksimov and Plume 2001).
Figure 2: Creep compliance as a function of time and load as a % of average fiber break load (ABL) = 445 N (2.64
GPa) (Burgoyne and Alwis 2008).

Figure 3: Creep curves for aramid fibers at (1) 10%, (2) 30%, and (3) 50% of UTS = 3.3 GPa (Maksimov and Plume
2001).

Epoxy and vinylester resins show extensive viscoelastic behavior over a range of temperatures
(Sperling 2001). Raghavan et al reported creep test results for an epoxy resin (3501-6) at
temperatures ranging from 22°C to 160°C, which showed the characteristic increase in
compliance with time and temperature. The effect of temperature was attributed to decreased
resistance to motion within the polymer network, i.e. softening of the polymer, with temperatures
climbing toward the Tg. The authors reported strains on the order of 0.3% recorded after 104
sec (2.8 hr) under a 15 MPa applied stress (Raghavan et al. 1997). Maksimov and Plume
reported ~1.25% strain under a 13.6 MPa applied stress after 104 sec for a different epoxy resin
(EDT-10) (Maksimov and Plume 2001). The variation in these results indicates the dramatic
variability that can be observed within this family of resins. It is generally accepted that the
creep response of carbon/polymer composites is dominated by the resin matrix properties, with
the fibre reinforcements moderating the extent of creep. Factors affecting the creep response of
resins, such as temperature and moisture levels and the extent of cure, will therefore play a role
in the creep of a composite laminate (Karbhari et al. 2003).

Cain et al [2006] studied E-glass/vinyl ester resins by way of symmetric [0/+45/90/-45/0]s and
[±45]2s laminates. Isothermal creep characterization of the laminates showed that the presence
of an uncured oligomer phase within the vitrified resin resulted in greater short-term creep
response as compared to the same panel tested for creep after resting at room temperature for
an extended period, during which the uncured phase was thought to increase the degree of cure
of matrix resin. More highly post-cured specimens were not observed to advance in cure
significantly with aging at room temperature.

2.2 Aramid composites

Long-term creep of unidirectional arabid/epoxy composites with a 50% fiber volume fraction was
examined by Maksimov and Plume at 20°C (Maksimov and Plume 2001). Creep responses in
the fiber direction were obtained at 50-60% UTS and showed that the aramid/epoxy composite
exhibits considerable creep as compared to an idential layup of glass/epoxy, with the total strain
after 50,000 hours exceeding the instantaneous viscoelastic strain by 220%. The inclusion of
some volume fraction of glass fibers, which exhibit negligible creep on their own, resulted in
decreased creep levels over time (Fig. 4).

Figure 4: Creep curves for unidirectional aramid/glass composites with volume fractions of aramid (a) and glass (g)
fibers: (1) 50% a, 0% g; (2) 40% a, 10% g; (3) 29% a, 21% g; (4) 18% a, 32% g; (5) 10% a, 40% g; and (6) 0% a,
50% g (Maksimov and Plume 2001).

The viscoelastic behavior of aramid composites has also been studied under flexure (Fischer
and Roman 1986). Fischer and Roman examined flexural creep of unidirectional aramid/epoxy
with a fiber volume fraction of 46%. The authors reported compressive creep strains on the
order of 2 - 4% for applied compressive stresses between 475 MPa and 563 MPa, with failure
occurring due to excessive strain on the tensile side.

2.3 Carbon composites

Creep in carbon composites has been shown to follow the profile provided in Fig. 1 (Fischer and
Roman 1986; Goertzen and Kessler 2006; Hu 2007; Miyano et al. 2005; Sperling 2001; Tuttle
and Brinson 1986). While long-term creep testing of carbon/epoxy composites has been
accomplished, efforts to effectively predict long-term creep behavior have been focused on the
use of momentary, or short-term, creep studies combined with measured aging effects for the
prediction of the long-term creep response.

As mentioned in Sec. 3.0 it is generally accepted that the viscoelastic behavior of laminates is
influenced primarily by the resin matrix and conditions that affect creep thereof. Nevertheless,
as indicated by Tuttle and Brinson, the extent of the viscoelastic deformation is also dependent
on layup and levels of stress in relation to the UTS of the laminate (Tuttle and Brinson 1986). In
their article published in 1986, the authors reported the creep compliance of unidirectional [0]8
and [90]8 carbon/epoxy laminates as measured over 69.4 days at 149°C, ~30°C below the
laminates' glass transition temperature. The results showed negligible creep for the fiber-
dominated [0]8 laminates and a typical creep response in the matrix-dominated [90]8 laminates.

Similar results were obtained by Raghavan and Meshii who expanded the data set to include
[10], [30], and [60] 8-ply carbon/epoxy laminates tested at temperatures ranging from 22°C to
160°C over 24 hrs. Results showed increasing creep compliance with stress, temperature, and
time characteristic of a non-linear viscoelastic response, but with creep compliances well below
that of the neat epoxy resin. Further, it was found that creep under in-plane shear loading, is
higher than that under tensile loading but that the resolved in-plane shear is the same, as in the
case of the [30]8 and [60]8 laminates, greater creep will evolve in the laminate with the greater
resolved transverse stress (Raghavan and Meshii 1997).

Such short-term, or momentary, creep studies have also been used to examine the effects of
physical aging on laminate creep response (Hu 2007; Sullivan 1990; Sullivan et al. 1993). Such
studies are generally conducted at elevated temperature with a constant stress applied for a
short duration, e.g. 105 sec, to record the momentary creep response followed by relaxation and
dwell at the nominal aging temperature (Ta < Tg). Momentary creep compliance curves for a
symmetric carbon/epoxy laminate (Hu 2007) showed that aging results in lower compliance
(Fig. 5). Similar effects have been found by decreasing the test temperature (Goertzen and
Kessler 2006; Sullivan 1990). For example, Goertzen and Kessler found that, for laminates
made of plain woven carbon fiber cloth and epoxy resin and tested in the warp direction (UTS =
645 MPa), tensile creep compliance curves generated at temperatures ranging from 30°C to
75°C showed progressively lower creep compliance with decreasing temperature (Fig. 6)
(Goertzen and Kessler 2006). The authors were able to apply time-temperature superposition
to describe a complete creep compliance curve over several decades of time.

(a) [(±45)3]s (b) [90/(+45/-45)3]s

Figure 5: Creep compliance curves for carbon/epoxy multi-directional laminates show that aging results in a
depression of creep compliance (Hu 2007).

Figure 6: Creep compliance curves measured at temperatures between 30°C and 75°C show that increasing
temperature softens the composite, resulting in elevated compliance. Also shown is the master curve onto which the
65°C data has been shifted using TTS (Goertzen and Kessler 2006).

Creep of carbon/polymer composites has also been approached from the perspective of creep
rupture strength, a measure of the time failure under constant stress. Miyano et al proposed an
accelerated testing methodology for long-term durability of plain woven carbon fiber/vinylester
composites based on the time-temperature superposition principle (Miyano et al. 2005). The
authors showed that the flexural creep strength at temperatures greater than 25°C could be
shifted to represent the flexural creep strength at 25°C after longer times (Fig. 7).

Figure 7: Left: experimental flexural creep strengths at 25°C, 80°C, and 100°C. Right: flexural creep strengths are
shown shifted to a nominal temperature of 25°C (Miyano et al. 2005).

2.4 Effects of temperature and moisture

Creep in carbon fibre reinforced polymer composites has primarily been addressed at
temperatures between 20°C and ~Tg-30°C, where Tg is the glass transition temperature of the
laminate in question, (e.g. (Goertzen and Kessler 2006; Hu 2007; Raghavan and Meshii 1994))
and there exists little data on the effects of temperature on the creep response of
aramid/thermoset composites. In general, the creep compliance of the composite laminate,
irrespective of layup, that demonstrates susceptibility to creep will exhibit greater compliance at
elevated temperature. The effect of temperature is most often addressed using time-
temperature superposition (TTS), also known as the time-temperature superposition principle
(TTSP), wherein the creep compliance curves are shifted onto a continuous master creep curve.
This method has been used extensively for unidirectional composite laminates over the
temperature ranges indicated above (e.g. (Raghavan and Meshii 1994; Tuttle and Brinson
1986)). Use of this method is either done by inspection (e.g. (Tuttle and Brinson 1986)) or by
implementing constant activation energy assumptions wherein the shift factor is determined
from an Arrhenius relationship such as the following (Goertzen and Kessler 2006):
ΔH ⎛ 1 1 ⎞
log(aT ) = ⎜⎜ − ⎟⎟loge (1)
R ⎝ T Tref ⎠

where aT is the shift factor, ΔH is the activation energy calculated from tan delta curves, R is the
universal constant, T is the temperature of the data curve to be shifted, and Tref is the
temperature of the reference curve to which the data is being shifted. This method was
successfully used by Goertzen and Kessler (Goertzen and Kessler 2006) for temperatures
between 30°C and 75°C. For a more extensive review, the reader is referred to (Scott et al.
1995).

One can assume that the TTS principle holds for sub-zero temperatures, however, there is no
direct data to support this. Studies that have examined the effects of subzero temperatures are
generally related to civil infrastructure usage of composite laminates for repair and use as
reinforcement for concrete structures (Dutte 2002; Karbhari et al. 2003). While Dutte has argued
that freeze-thaw and sub-zero temperatures within a range of +30°C to -20°C will not
significantly affect the mechanical properties of composite laminates (Dutte 2002), Karbhari et al
showed that exposing an E-glass/vinylester composite cured under ambient conditions to
freezing at -10°C, as well as exposing the laminate to a freeze-thaw cycles, induced a reduction
in the glass transition temperature of the laminate as measured by DMTA (Karbhari et al. 2002).
The effect on creep, therefore, might be speculated to follow from the influence of these
conditions on the mechanical properties and the glass transition temperature, Tg. Since the Tg
is intimately related to the viscoelastic properties of a laminate, a reversal of the general effect
of temperature, i.e. decreasing creep compliance with decreasing temperature, could result.
Such an effect, or any effect, however, remains to be documented.

The adsorption of moisture into the matrix resin has also been shown to affect the mechanical
properties of laminates (Karbhari et al. 2002; Mohan and Adams 1985; Woo 1994). Woo
examined the effect of temperature and moisture on creep of [±45]2s carbon/toughened epoxy
prepreg and found that the presence of moisture, induced by soaking specimens in water at
60°C for 14 days prior to testing, lowered the Tg of the laminate and resulted in non-linear creep
compliance curves even at 26°C but with exaggerated effects at higher temperatures. Due to
the increased non-linearity, the creep compliance curves of "wet" laminates could not be shifted
onto a master curve using standard TTS methods. It was proposed that the increased moisture
content meant that 'wet' laminates tested at, for example, 26°C performed as would a 'dry'
laminate tested at elevated temperature, 60°C in this case. No mathematical expression was
provided to describe this equivalency. Mohan and Adams also investigated the effects of
temperature and humidity, testing the tensile and compressive creep responses of
carbon/epoxy prepreg specimens in relative humidity of 57%, 75%, and 100% RH at 25°C and
in standard ambient RH at temperatures between 25°C and 125°C. The authors reported
similar effects to those discussed above.

3.0 Summary and Conclusions of the Literature Review

The creep response of carbon/polymer and aramid/polymer laminates has been reviewed.
Overall, both carbon and aramid fiber composites can be susceptible to creep, primarily when
stress is applied in a matrix-dominated direction and especially so when in-plane shear stresses
are present. Creep is seen as a function of the ultimate strengths and is generally seen as
significantly reducing in severity under 10% of ultimate. In general the following conclusions
can be drawn:
1) Aramid-based composites show elevated creep levels due to the viscoelastic nature of the
aramid fibers but including elastic fibers can moderate the extent of creep.
2) The general mechanism of creep within the matrix does not change with the addition of
reinforcing fibers but strain levels are moderated.
3) Long-term creep of multi-directional laminates can be approximated from short-term creep
compliance curves for unidirectional laminates tested over a range of temperatures below Tg.
4) The time-temperature superposition principle is applicable when the creep response exhibits
a relatively high degree of linearity.
5) The effect of sub-zero temperatures on the creep of laminates has not been extensively
reported and the effect of such temperatures would benefit from further investigation, especially
if a significant portion of the operating life of the composite is predicted to be at low
temperatures.
6) The adsorption of moisture into the laminate results in non-linear creep response with a
depressed Tg; as such, creep curves can be approximated by those of a dry laminate at an
elevated temperature corresponding to the difference in Tg.

4.0 References
Burgoyne, C. J., and Alwis, K. G. N. C. (2008). "Viso-elasticity of aramid fibres." Journal of
Materials Science, 43, 7091-7101.
Cain, J. J., Post, N. L., Lesko, J. J., Case, S. W., Lin, Y.-N., Riffle, J. S., and Hess, P. E. (2006).
"Post-curing effects on marine VARTM FRP composite material properties for test and
implementation." Journal of Engineering Materials and Technology, 128, 34-40.
Canning, L., Hodgson, J., Karuna, R., Luke, S., and Brown, P. (2007). "Progress of advanced
composites for civil infrastructure." Proceedings of the Institution of Civil Engineers,
Structures & Buildings, 160(S86), 307-315.
Chambers, J. J. (1986). "Parallel-lay aramid ropes for use as tendons in prestressed concrete.,"
University of London, London, UK.
Dutte, P. K. (2002). "Durability issues of composites in cold regions." Polymer Composites II:
Applications of Composites in Infrastructure Renewal and Public Development, R. C.
Creese and G. Hota, eds., CRC Press LLC, Boca Raton, FL, 125-136.
Fischer, S., and Roman, I. (1986). "A viscoelastic-plastic analysis of an aramic composite beam
in flexure: rate of testing effects, stress relaxation and creep." Materials Science and
Engineering, 84, 97-104.
Goertzen, W. K., and Kessler, M. R. (2006). "Creep behavior of carbon fiber/epoxy matrix
composites." Materials Science and Engineering A, 421, 217-225.
Hu, H. W. (2007). "Physical aging in long term creep of polymeric composite laminates." Journal
of Mechanics, 23(3), 245-252.
Karbhari, V. M., Chin, J. W., Hunston, D., Benmokrane, B., Juska, T., Morgan, R., Lesko, J. J.,
Sorathia, U., and Reynaud, U. (2003). "Durability gap analysis for fiber-reinforced
polymer composites in civil infrastructure." Journal of Composites for Construction, 7(3),
238-247.
Karbhari, V. M., Rivera, J., and Zhang, J. (2002). "Low-temperature hygrothermal degradation of
ambient cured E-glass/vinylester composites." Journal of Applied Polymer Science, 86,
2255-2260.
Maksimov, R. D., and Plume, E. (2001). "Long-term creep of hybrid aramid/glass-fiber-
reinforced plastics." Mechanics of Composite Materials, 37(4), 271-280.
MIL-HDBK-17-1F. (2002). "Composite materials handbook. Volume 1. Polymer matrix
composites guidelines for characterization of structural materials." D. o. Defense, ed.
Miyano, Y., Nakada, M., and Sekine, N. (2005). "Accelerated testing for long-term durability of
FRP laminates for marine use." Journal of Composite Materials, 39(1), 5-20.
Mohan, R., and Adams, D. F. (1985). "Nonlinear creep-recovery response of a polymer matrix
and its composites." Experimental Mechanics, 25(3), 262-271.
Raghavan, J., and Meshii, M. (1994). "Activation theory for creep of matrix resin and carbon
fibre-reinforced polymer composite." Journal of Materials Science, 29, 5078-5084.
Raghavan, J., and Meshii, M. (1997). "Creep of polymer composites." Composites Science and
Technology, 57, 1673-1688.
Raghavan, J., Meshii, M., and Feng, D. (1997). "Influence of reinforcing continuous carbon
fibers on the viscoelastic properties of the epoxy matrix." Polymer Composites, 18(1),
55-63.
Scott, D. W., Lai, J. S., and Zureick, A.-H. (1995). "Creep behavior of fiber-reinforced polymeric
composites: A review of the technical literature." Journal of Reinforced Plastics and
Composites, 14, 588-617.
Sperling, L. H. (2001). Introduction to physical polymer science, John Wiley & Sons, Inc., New
York, NY.
Sullivan, J. L. (1990). "Creep and physical aging of composites." Composites Science and
Technology, 39, 207-232.
Sullivan, J. L., Blais, E. J., and Houston, D. (1993). "Physical aging in the creep behavior of
thermosetting and thermoplastic composites." Composites Science and Technology, 46,
389-403.
Tuttle, M. E., and Brinson, H. F. (1986). "Prediction of the long-term creep compliance of
general composite laminates." Experimental Mechanics, 26(1), 89-102.
Wolff, E. G. (1989). "Creep." Reference Book for Composites Technology, S. M. Lee, ed.,
Technomic Publishing Company, Inc., Lancaster, PA, 111-142.
Woo, E. M. (1994). "Moisture-temperature equivalency in creep analysis of a heterogeneous-
matrix carbon fibre/epoxy composite." Composites, 25(6), 425-430.

You might also like