You are on page 1of 20

Degradation of Polymer Matrix Composites☆

Rodney H Martin, Materials Engineering Research Laboratory Ltd., Wilbury Way, Hitchin, United Kingdom
Md Sayem H Bhuiyan, University of Malaya, Kuala Lumpur, Malaysia
r 2018 Elsevier Inc. All rights reserved.

1 Introduction 2
2 Ageing Mechanisms 4
2.1 Physical Ageing and Time-Dependent Effects 4
2.2 Hygrothermal Effects 5
2.3 Thermooxidative Degradation 6
2.4 Chemical Ageing 6
2.5 UV and Weathering Degradation 7
2.6 Mechanical Degradation 7
2.7 Fire Testing 8
2.8 Synergistic Effects 8
3 Accelerated Ageing 8
4 Ageing Associated With Supersonic Flight 9
5 Ageing in the Oil and Gas Industry 10
6 Ageing in the Chemical Processing Industry 13
6.1 ASTM Standard for Long-Term Chemical Resistance 16
6.2 The Arrhenius Relationship 16
6.3 Using a Semiempirical Corrosion Approach 17
7 Ageing in the Marine Industry 18
8 Conclusions 19
References 19
Relevant Website 20

Abbreviations FTIR Fourier transform infrared


CPI Chemical processing industry GC Gas chromatography
CRA Corrosion resistant alloy GRP Glass reinforced plastic
DCB Double cantilever beam MS Mass spectroscopy
DMA Dynamic mechanical analysis TGA Thermogravimetric analysis
DSC Differential scanning calorimeter TMA Thermomechanical analysis
ECR Corrosion resistant e-glass UV Ultraviolet
ESC Environmental stress cracking/corrosion Vf Fiber volume fraction
FRP Fiber reinforced plastic

Symbols Mt Total mass absorbed at time t (kg)


a Corrosion depth factor M1 Total mass at equilibrium (kg)
B Factor in the presence of deposits U Depth of corrosion (m)
c Chlorine dioxide concentration (g L1) R Universal gas constant (J K1 mol1)
D Diffusion coefficient (m2 s1) t Time (s)
Ea Activation energy (J mol1) Tg Glass transition temperature
E11 Young’s modulus in the fiber direction (GPa) t95 Time to reach 95% of original property (s)
E22 Young’s modulus transverse to the fiber direction (GPa) T Temperature (K)
G12 Shear modulus (GPa) W Rate of weight loss (kg s1)


Change History: November 2017. Md Sayem H Bhuiyan has added heading to Abstract, and changes the heading of Conclusions, added keywords and
updated Sections 1–3, 5 and 7.

Reference Module in Materials Science and Materials Engineering doi:10.1016/B978-0-12-803581-8.11222-6 1


2 Degradation of Polymer Matrix Composites

Glossary Pultrusion A manufacturing process that consists of


Extrusion A manufacturing process in which a softened pulling a fiber reinforcing material through a resin
blank of plastic material is forced through a shaped die to impregnation bath and a shaping die prior to final curing.
produce a continuous ribbon of the formed product. Thermoplastic A polymeric material (for example nylon)
Gelcoat A polymer material that is used to provide a high- which, upon curing, forms one- or two-dimensional
quality finish to the visible surface of a fiber reinforced molecular structures and is capable of being repeatedly
composite material. softened by increases in temperature and hardened by
Polymer resin A polymerized synthetic or chemically decreases in temperature.
modified natural material including thermoplastic materials Thermoset/thermosetting A polymeric material (e.g.,
such as polyvinyl, and thermosetting materials such as epoxy) which, upon curing, results in a three-dimensional
epoxies, that are used with other components such as fillers structure and decomposes rather than melting upon
to form plastics. heating.

1 Introduction

In the past, the primary reason for the use of fiber reinforced plastics (FRPs) is because of weight saving for their relative stiffness
and strength. As an example, a carbon fiber composite can be five times stronger than 1020 grade steel while having only one-fifth
the weight. Aluminum (6061 grade) is much nearer in weight to carbon fiber composite, but the composite can have twice the
modulus and up to seven times the strength. A comparison of some properties of metals and composites is shown in Fig. 1.
A composite, which is generically a material that is made up of two or more distinct (i.e., macroscopic, not microscopic)
materials, comes in different configurations but is essentially a plastic material within which there are embedded fibers or particles.
The plastic, or more correctly a polymer, is more often termed the matrix or sometimes the resin. The fibers or particles dispersed
within the polymer are known as the reinforcement. The reinforcement is generally orders of magnitude stiffer than the matrix,
thus resulting in a stiffened plastic. This stiffer reinforcement can either be randomly distributed short fibers or continuous fibers
(sometimes in mat form) that are laid in a particular direction within the matrix. The resulting material thus may have different
properties in different directions (anisotropic or orthotropic, depending on the lay-up). This characteristic is usually exploited to
optimize the design and results in the composite material being tailorable in that its mechanical properties can be formed to meet
specific directional loading.
For example, the PU is a widely used polymer due to it versatile nature. However, under harsh conditions PU alone fails to give
satisfactory thermal, mechanical and corrosion resistance performance. Thus, there is a need to induce such structural properties in
PUs through their blending and composites formations, which make them able to provide good services even under harsh
environmental conditions. It has been observed that the synergistic effect between CPs (filler) and insulating PUs (matrices)
enhances their proccessability, stability, solubility, thermal, mechanical, electrical and optical properties.1 Polymer matrix com-
posites (PMCs) have been manufactured to improve the effective thermal, mechanical and electrical properties of pure polymeric
materials. The benefits of PMCs, such as light weight, ease of process, high strength, durability, and multifunctionality, have been

Fig. 1 Comparison of specific strength of different metals and FRPs (www.merl-ltd.co.uk).


Degradation of Polymer Matrix Composites 3

clearly shown in aerospace applications. However, two main limitations are currently slowing down a more widespread exploi-
tation of such carbon based composites: a weak fiber-matrix adhesion, which typically leads to a progressive degradation of the
initial properties; poor out-of-plane properties, due to the anisotropic nature of nanotube or graphene nanofillers.2
Most engineering materials are essentially isotropic. That is, they have the same properties such as strength and modulus in any
direction. There may be ‘grain’ in some metals because of the manufacturing process, but it is only in critical applications that this
matters. Most machining or casting processes do not have to take directional differences into account. Because of the different fiber
orientation, mentioned above, most composites will have very different properties in different directions. For example, a carbon
FRP may be up to 100 times stronger under tension than it is in shear, and the stiffness may differ in the two directions by similar
ratios.
Several different terms have been used by different industries to describe a composite material, and these acronyms have
become jargons. More technical descriptions include:
None of these terms are incorrect, although some cover a wider range of material classes. FRP is a generic term and will be used
throughout this article when referring to composite materials generically.
Both thermosetting and thermoplastic polymers can be used for the matrix material, and the most common thermosetting
matrix materials include polyester, vinyl ester and epoxy, and bismaleimide for higher-temperature applications. Thermoplastic
matrix materials tend to have higher performance ability and include polymers such as polyimide, phenolic, polypropylene,
polyetheretherketone (PEEK), and polyphenylene sulfide (PPS).
Another primary benefit of FRPs is their durability in hostile environments, often termed corrosive environments, although in
reality, composite materials do not ‘corrode.’ While the matrix is primarily the binder that transfers the load to the reinforcement,
it is also very resilient to environmental attack. The type of resin needs to be carefully selected so that it can withstand the service
temperature (high or low) conditions, and it should not be significantly affected within its lifetime by fluids (gases or liquids) with
which the FRP will come into contact. Hence, FRPs are used in very corrosive applications such as for acid storage in the chemical
processing industry and longevity in marine environments such as in vessels and offshore oil and gas applications. These
applications are discussed in more detail in this article.
The effect of exposure to various environments such as heat, moisture, solvents, acids, ozone, hydrocarbons, loads, etc., and
more importantly a combination of these parameters, may degrade the material's key properties. For load bearing applications,
these properties may include stiffness and strength. Cracks may appear and accumulate ultimately leading to failure. For fluid
containment applications, cracks may coalesce leading to a leak path for this fluid even though the structure is globally still intact.
Despite these safety critical uses, the degradation of an FRP is still not a well understood phenomenon compared to the corrosion
of metals.
The reason that the long-term degradation of FRPs is not well understood is because it is a very complex process with many
interrelated factors and has led to much ongoing research into the topic. Degradation is often given the generic term ‘ageing’ that
can include the more benign physical ageing effects such as swelling from moisture absorption, that is largely reversible, to the
more serious chemical ageing where material properties are permanently (irreversibly) degraded. Additionally, synergistic effects to
ageing from mechanical loading, such as creep, need to be considered in isolation (or in addition) to that associated with the
environment and often change the rate of degradation. Environmental ageing of FRPs occurs from the surface or edge inwards and
requires time to penetrate into the material's centre, and the rate of ageing can be directional (like the mechanical properties);
further, it is generally dependent on the temperature and stresses. The complexity described above makes predicting ageing around
detailed geometries such as stress concentrations nontrivial.
Added to the issue of ageing of FRP materials is the variety of different manufacturing methods available to manufacture
components. The reader is pointed toward the works of Jones,3 Matthews et al.,4 and in ASM Handbook 5 for detailed information
on manufacturing. The methods available depend very much on the nature and quality of the component to be manufactured. For
the low-end marine industry, uncured resin is poured over chopped glass fibers laid in a mould to form what is often termed
‘fiberglass’. A step-up on this approach is known as filament winding where the fibers are wrapped around a net-shape mandrel,
and the resin poured on or the fibers are dipped in the resin before winding. Both methods are low-cost manufacturing techniques
and the resin is generally cured at room temperature. Alternative lower-cost manufacturing methods include injecting the resin
into a mould or extruding or pultruding the composite profile through a die to form the part desired. Additionally, resin-rich
layers, known as gel coats can be put on the outside of the parts for better resistance or appearance. To improve the quality and
performance of the resin, vacuum, pressure and temperature are used in a variety of combinations. A vacuum is employed by
placing the component in a bag and the vacuum is used to help to pull the resin through the fibers and, more importantly, to
remove air and other gases from the final composite part reducing porosity. Pressure is used on the outside of the vacuum bag to
help form and consolidate the part again, helping to reduce porosity. Temperature is used for stronger resins to assist in curing and
consolidation. The combination of these three manufacturing parameters requires an autoclave and produces a high-quality, high-
strength part and is still the manufacturing method of choice for the aerospace industry.
A further difficulty is introduced with the need to predict long-term ageing using short-term testing and analysis. For long-term
life prediction based on short-term data, representing the true service history (load, temperature, time, fluid, etc.) for long-term
structural life prediction is a vital step to validate any short-term-based predictive methodology. The coupon tests must reflect the
effect of ageing on the polymer (i.e., matrix-dominated properties) because the fibers may mask any property loss in the resin.
However, fibers can sometimes degrade in some environments quicker than the polymer and additionally the fiber–matrix
interface can be attacked and this too must be represented in any accelerated ageing methodology. Further, the chemical bond
4 Degradation of Polymer Matrix Composites

between the fiber and resin itself (often know as the interface region) can be degraded. All of these degradation methods are
important to understand. Loss of fiber properties greatly reduces the strength and stiffness of the FRP laminate because that is their
main purpose. Similarly, a break down of the bond between the fiber and the resin leads to the fibers not providing stiffness and
strength through transfer of shear between the two. Reduction of properties of the resin is also important because it provides the
load transfer between fibers.
It is this shortage of long-term data or of an accelerated ageing methodology to determine the residual properties of a FRP
component at any time in its life that still hinders the uptake of FRPs in some applications. However, over the decades when
composites have been used, much has been learned, and this article addresses some of this work detailing the nature of ageing of
composites and the use of composites in certain industry sectors where corrosion or degradation is an issue. This article will
describe how it is the degradation or ageing of a composite that needs to be understood when these materials are used in hostile
environments. This article is not intended to be a general introduction to FRPs; there are many publications available for this
purpose3–5 or informative internet-based sites such as The Virtual Composite Material Consultant.4

2 Ageing Mechanisms

The ageing of an FRP comprises an ongoing development of damage that is dependent on the type of polymer matrix, the fiber
type and architecture, the fiber volume fraction and the operating conditions of temperature, relative humidity, oxidative attack,
solvent infusion, internal moisture concentration, air pressure, mechanical loads, and the time of exposure. These degradation
factors are often related and varying. For example, a FRP pipe used to transport hydrocarbons may experience hot oil field fluids
internally, ultraviolet (UV) radiation from sunlight, and moisture externally, external loads from supports and other loading. The
external temperature may vary between  40 and þ 401C depending on the location while the internal temperature may be that of
the oil at temperatures up to þ 1001C. However, a fuselage or wing FRP structure may experience temperature cycles of  55 to
þ 401C between flight and ground with ice, cleaning solutions, moisture, jet fuel, ozone, UV and other environmental conditions,
making the determination of residual life a complex procedure.
Numerous studies have focused on the environmental degradations of Polymer Matrix Composites (PMCs). The synergistic
aging of a group of commercially available Glass Reinforced Polymer (GRPs) under combined UV radiation, moisture, tem-
perature and time have been studied by Tianyi Lu et al.6 The UV components of sunlight which reach the ground are in the range of
280e400 nm. The energy of ground reaching UV photons is comparable to the dissociation energies of polymer covalent bonds
resulting in a loss of surface gloss, surface discoloration, chalking, flaking of surface resin, pitting, microcracking, and a severe loss
of resin in GRPs. The damaging effect of water or moisture on polymer composites, on the other hand, is not as harsh as
degradation just by UV radiation even at elevated temperatures. However, moisture diffusion into polymer matrix/fiber interfaces
can damage the interfaces by microcracking especially at elevated temperatures. In addition, hydrolysis of chemical bonds may
lead to permanent chemical degradation and moisture induced swelling of polymers and their composites. The combined action
of both UV and water on polymers and PMCs as a function of time and temperature could be even more severe than the individual
effects6. Another type of degradation called fatigue degradation of polymer composite is observed during cyclic loading. The
process of fatigue degradation of polymeric composites consists of three stages: at the first stage initial damage accumulation in a
form of microckracks of a polymeric matrix can be observed; further, at the second stage, monotonically growing damage
accumulation in the matrix as well as at the fiber/matrix interface occurs; and finally, at the third stage, macroscopic cracks
development in the matrix, and immediately after that, fiber breakage occurs which leads to failure of the structure.7
Therefore, ideally these conditions must be represented in any simulation of ageing for predictive purposes to accurately
reproduce the degradation mechanisms. In reality, only the key conditions considered to be the more potentially damaging may
be investigated. In any test program, it is important to use relevant material or structural properties as an indicator of the state of
ageing of the material. This may be the mass change, the glass transition temperature Tg, the crack density, the modulus, strength or
a combination of these properties. Therefore, a complete understanding of all the degradation processes should be known or an
assumption made on what is the primary degradation mechanism. The section below describes some of the key degradation
mechanisms that individually or combined form the ageing of a composite material.

2.1 Physical Ageing and Time-Dependent Effects


Polymers, both in their rubbery (above Tg) and in their glassy state (below Tg), are characterized by viscoelasticity, a time-
dependent mechanical behavior. The characteristic times of this constitutive behavior are much longer in the glassy state and
decrease at temperatures near and above Tg. Typical evidences of viscoelasticity are creep and stress relaxation. Creep is a time-
dependent deformation that usually has three stages when under a constant load, rapid primary creep, a steady-state secondary
creep rate and a rapid acceleration in the creep rate leading to rupture. Stress relaxation is a related phenomenon where stresses are
relieved when under a constant deformation. In composite materials, the stress transfer between the viscoelastic matrix and the
elastic reinforcement, or between layers with different fiber directions (laminae) needs to be considered for long-term ageing
characterization. For long-term property estimation, especially when considering FRP properties that are more dominated by the
resin than the fiber, long-term viscoelastic behavior can be accelerated by adopting the time–temperature superposition
Degradation of Polymer Matrix Composites 5

principle.8,9 This principle is based on the premise that the behavior for long time periods at low temperatures corresponds to that
for short time periods at higher temperatures. This is because in a polymer, the polymer chains are more mobile and easier to
arrange at higher temperature resulting in a higher compliance or lower modulus. The time–temperature superposition principle is
evident from the compliance curves in creep experiments (or modulus curves for stress-relaxation experiments) at different
temperatures that are related to one another by a simple temperature-dependent shift on the log time scale.
The fundamentals on physical ageing and the dependency of mechanical properties on time and temperature become more
pronounced when testing close to the Tg. The Tg itself may well be affected by moisture; lowering Tg, and service temperatures may
be close to wet Tg requiring aspects of physical ageing to be a vital part of a durability analysis.
With physical ageing, polymers below their Tg exhibit a time-dependent rearrangement of their structure. The mutual inter-
action between viscoelasticity and physical ageing imposes the introduction of time–ageing and time–temperature shift factors.
Other time-shifting effects may also exist based on the applied stress, moisture condition and other environmental factors. These
time-dependent properties and physical ageing has an affect on the material's overall strength and stiffness. Thus, ideally an
accelerated ageing analysis needs to include an aspect of physical ageing such that the strength/failure criteria includes time-
dependent coefficients and a nonlinear viscoelastic–viscoplastic damage analysis.
For accelerated physical ageing tests, the relationships are obtained from short-term tests using the time-dependent experi-
mental data described above. It is important that no chemical ageing (discussed further below) is included. The typical properties
measured depend on the application, but most testing presented in the literature has been on off-axis plies because they give a
good representation of matrix-dominated properties.9 However, the literature often falls short in producing such time-dependent
relationships for more useful design properties such as static properties of FRP laminates with or without holes.

2.2 Hygrothermal Effects


Hygrothermal effects are related to the uptake of water molecules from a humid environment or water in direct contact with the
FRP when the component is submerged in water (or other solution). The water is absorbed by the polymer matrix and results in
plasticization of the matrix leading to hydrolytic degradation or hydrolysis which lowers the molecular weight and hence
mechanical properties of the FRP. The rate of hydrolysis may be accelerated by the application of load or more significantly with
elevated temperature and high- or low-pH conditions. In contact with the polymer, low molecular weight liquids are absorbed by
and diffuse into the polymer until saturation (equilibrium) is reached. The rate of diffusion and equilibrium are dependent on the
chemical structure and morphology, such as the degree of crystallinity of the polymer. For an FRP, these properties are also related
to the type of fibers, the volume fraction and the nature of the fiber–matrix interface.
The relationship between time and moisture mass uptake can classically be characterized using simple approaches such as Fick's
law, but generally for FRPs, diffusion is more complex with competing factors affecting the rate of diffusion.10 With time, moisture
diffuses into the FRP through the surfaces and there is a moisture-content concentration gradient through the thickness. Indeed, for
some FRPs there may be a different concentration through the edges than through the surfaces. If the moisture content or
temperature is cycled, this will lead to local dryness at the surface plies compared to the centre of the laminate. This moisture
distribution can lead to significant residual stresses within the laminate that can in turn cause outer ply delamination of blisters.
Short-term hygrothermal effects on FRPs may extend beyond plasticization of the matrix and lead to mechanical damage such
as the creation of microvoids, matrix cracking and blistering. These microcracks, in turn, provide fast diffusion paths (for both
liquid water and water vapor) and then alter the moisture-absorption characteristics of the laminate.
Long-term exposures may not cause mechanical damage, described above, but the plasticization may well change matrix-
dominated mechanical properties such as interlaminar and intralaminar shear modulus and the transverse tensile strength and
modulus. Properties like compression modulus and strength are greatly affected by the shear properties of the matrix; hence, under
hot and wet conditions compression properties may be severely degraded. The high temperature and humidity exposure of FRPs
effectively plasticizes the matrix, more so closer to the Tg of the material which reduces when the material is saturated.
Post-cure reactions, hydrolysis and leaching of products within the FRP coupled with the decrease of Tg when moisture is
absorbed, make modeling the effects of moisture complex and warrent further investigation. Such modeling can supply important
information to establish the effects of moisture on time–temperature shift factors, by scaling the temperature axis on the basis of Tg
change. As a result, Fick’s law needs to be extended to model three-dimensional diffusion in anisotropic materials. For FRPs, the
value of the diffusion coefficient, D, for the neat polymer needs to be modified to account for the tortuous path created by the
presence of the fibers (their size, distribution, aspect ratio, orientation, etc.). Hence, D becomes an anisotropic property that is
dependent on fiber direction. In FRP laminates with several layers of different orientation, the different diffusivity in each layer
must be considered for the overall laminate. In ideal Fickian behavior, moisture absorption increases linearly with the square root
of time for values of Mt/M1 r 0.5 (where Mt is the total amount of moisture absorbed at time t and M1 is the total amount
absorbed at equilibrium). Time–thickness scaling allows the long-term prediction of absorption in thick laminates from thin-
sample measurement, when Fick's law applies. Diffusion can be accelerated by increasing the temperature and follows the
Arrhenius relationship Eq. (1)

  
Ea
D ¼ D0 exp ð1Þ
RT
6 Degradation of Polymer Matrix Composites

where Ea is the activation energy of the diffusion rate, R is the universal gas constant, and T is the absolute temperature of the
exposure in Kelvin.
As discussed above, Fickian law may not always be demonstrated in composites where, following reaching equilibrium, there
can be a slow long-term increase in mass uptake or a decrease in mass or both. The decrease may arise from products leaching from
the FRP over time and the increase from a two-stage absorption process where the first is Fickian and the second a relaxation-based
phenomenon.11

2.3 Thermooxidative Degradation


While the application of temperature and moisture can have negative effects on the composite, exposure to temperature in an air
environment can cause thermooxidative ageing. This ageing mechanism arises from the thermal instability of the polymer and
leads to oxidative attack that comprises chain scission, cross-linking, and thermooxidative reactions. Thermal degradation of FRPs,
in an inert atmosphere, is exclusively a thermolysis phenomenon, while in air it is dominated by oxidation. Thermolysis is the
result of breaking of covalent bonds in the polymer network and, in general, thermal stability of high-performance polymers in an
inert atmosphere is very good.9,11–14 Oxidation is usually characterized in terms of weight loss generally from the surfaces of the
laminate. This weight loss is predominantly from the matrix and will penetrate into the laminate at higher temperatures and long
exposures. Eventually, microcracks will form leading to a larger reduction in the mechanical properties of the FRP. These weight
losses are generally permanent (irreversible) and even a small weight change can be an indication of the onset of damage at the
surface. For oxidation to occur at the centre of the laminate, the diffusion process must be understood, including the diffusion of
oxygen and reactions products throughout the polymer matrix in a similar manner to liquid diffusion discussed above. The pattern
of oxidation reactions and subsequent chemical changes depends on the exposure temperature, the resin system, and the oxygen
content. Simply increasing the temperature to accelerate the characterization of oxidation is nontrivial. The changes in mechanical
properties postexposure are a function of sample thickness and surface protection as well as the ageing temperature and the
material’s Tg. Therefore, the examination of the thermooxidative stability of an FRP must include not only the degradation of
mechanical properties, but also the failure mechanisms to ensure that they are consistent when comparing two laminates aged at
different temperatures in an attempt to accelerate thermooxidative ageing. The temperature and time are identified as the key
process parameters for controlling thermal degradation in both the natural fiber and the bio-based matrix material.15 Degradation
of mechanical properties is one of the weaknesses of polymer matrix composites (PMCs) during thermal cycling. This weakness
limits applications of this light material with this limited strength in aerospace industry. During the past decades, different studies
are done to identify thermal cycling effects on the PMCs specimens. The observations showed that hole diameter changes failure
modes of composite specimens; while thermal cycling changes only failure regions without significant change in failure mode.
Moreover, thermal cycling decreases fiber-matrix deboning area and increases number of fiber splitting, consequently. Due to
surface oxidation and surface matrix degradation, fiber splitting after thermal cycling is larger than unexposed specimens. Hole in
the specimen caused significant stress concentration and mechanical property instability in comparison with thermal cycling
effects.16

2.4 Chemical Ageing


Chemical ageing follows principles very similar to oxidative ageing but occurs when the environmental fluid is an acid, alkali
solvent, or many other fluids to which an FRP may be exposed. These chemicals actually attack the resin, the fiber, or the
fiber–matrix interface (or any combination of these three). When the FRP is subjected to exposure to, for example, an acid
solution, depending on the concentration of the solution, competing ageing mechanisms may occur. The liquid will first be
absorbed in a hygrothermal fashion as described above. In time (dependent on the strength and nature of the chemical involved
and the compatibility of the polymer), the material will chemically react with the FRP. This results in an increase in cross-linking
density that can severely affect the mechanical properties by densification and increasing the Tg. Also, products may be leached out
of the FRP in an irreversible manner. This will result in weight loss and a potential decrease in properties. With further time, actual
physical material loss can occur, resulting in a more dramatic change in properties.
In order to perform studies of long-term and accelerated chemical ageing, usually a large test program is required with long
periods of time and a selection of parameters (temperature, chemicals, loads). A variety of specimen types is also required to assess
the degradation of mechanical properties parameters such as specimen type, lay-up, thickness, material, etc. As with other ageing
types, chemical ageing is a diffusion-related phenomenon. Therefore, the use of thick specimens will delay observable changes in
mechanical properties although significant surface damage may have occurred. Different lay-ups will affect the interlaminar
stresses, which will affect diffusion rates.17 Additionally, edge effects can play a role and this is often where mechanical failure can
initiate. As a result, chemical ageing exposures are often conducted on panels of FRP and mechanical specimens cut from these
panels once the ageing is complete. Another approach is to coat the free edges of the test specimens with a resilient coating to
prevent unrepresentative edge damage.
An additional approach to characterizing chemical ageing to better understand the change in the polymer is to conduct thermal
analysis on the FRP or the matrix. Of all the properties, Tg is the most common one measured using thermal analysis because it is a
good indication of physical changes. Methods and approaches to measure Tg vary and all are reasonably well established as
Degradation of Polymer Matrix Composites 7

standards. These include dynamic mechanical analysis (DMA) that may be performed in torsion or flexure depending on the test
apparatus; differential scanning calorimetry (DSC); or thermomechanical analysis (TMA). These methods all raise the temperature
of the FRP and look for inflection points of peaks in the property being measured. The DMA tends to be more sensitive than the
DSC for materials such as phenolics. Thermogravimetric analysis (TGA) is used to measure the rate of change of weight versus time
and temperature in a controlled gas environment. TGA is highly dependent on the available surface of the specimens placed into
the test chamber and hence the size of the pieces of materials that are used can effect measurements. It is sometimes useful to
identify how the polymer has changed during ageing and other equipments such as Fourier transform infrared (FTIR) spectroscopy
and gas chromatography (GC) as well as mass spectroscopy (MS) are used to evaluate the degradation products. These techniques
can also be used for thermooxidative ageing.

2.5 UV and Weathering Degradation


Similar to oxidative ageing, weathering affects the surface of the FRP laminate. Generally, UV ageing affects the appearance of the
laminate (particularly color loss) and this degradation in appearance can be considered failure (aesthetically). Often, the change in
color has a minor effect on structural performance. Weathering is a combination of ultraviolet radiation, temperature from the sun,
oxygen and any ambient moisture such as humidity or precipitation.18,19 With the use of coatings and additives, the structural
performance of FRPs can be maintained and the weathering effects tend to be surface related and appear as:

• fading and darkening;


• yellowing;
• blooming; and
• loss of gloss and chalking.
Yellowing, fading, and darkening are generally due to chemical degradation in the polymer. Blooming can be caused by an
additive coming to the surface, whereas loss of gloss is generally caused by erosion of the surface layer. While the loss of surface
quality may be considered benign, it can lead to leaks and weakening.
The severity of the ageing due to weathering is strictly dependent on the nature of the climate or the geographical region, for
example, tropical, desert, arctic, etc. Predicting the weathering performance of FRPs is carried out both in laboratories using
artificial weathering and in long-term experimental field ageing. The latter takes several years and becomes very specific to the
material and location selected for the study. The specimens are directly exposed to outdoor conditions at a fixed angle relative to
the horizontal and in a fixed direction. To accelerate weathering, the UV radiation is concentrated onto the test specimen using
special mirrors. Also, locations with high levels of sunshine, for example, Florida and Arizona are used to compare with longer-
duration exposures in regions with less sunshine hours.
There are a larger number of laboratory-based ageing tests along with actual case studies where specimens are exposed to UV
radiation from a variety of UV light sources. There are currently three key laboratory methods used for artificial weathering.19 The
carbon arc lamp uses two strong emission bands that peak at 358 and 386 nm which are much more intense than natural sunlight.
These have a lower effect than solar radiation on materials that absorb only short-wavelength UV radiation. The xenon arc lamp
gives a broad spectrum of light that matches the solar spectrum quite closely. Filters are used to reduce the short-wavelength UV
light from these lamps that are not present in sunlight. Fluorescent lamps have special phosphors selected to emit UV light at a
particular waveband. Exposure is carried out under controlled conditions of temperature and moisture. With the varied nature of
weather, conducting accelerated tests to simulate specific weather exposure for predictive purposes is difficult. As a result, accel-
erated methods tend to focus on the worst-case weather scenarios and any correlations must be qualitative.
Environmental degradation due to moisture, heat and ultraviolet (UV) radiation is of practical concern, particularly to aero-
space structures. Carbon fiber epoxy-matrix composites degrade under hygrothermal aging (i.e., aging in the presence of heat and
moisture), resulting in a reduction of the mechanical properties. The epoxy absorbs moisture, thus increasing in volume and
increasing the contact electrical resistivity of the inter laminar interface of the laminate. Compared to epoxy-matrix composites,
thermoplastic-matrix composites have less problem with moisture, but they tend to degrade under UV accelerated weathering
conditions, with synergistic effects of UV, moisture and heat. Among the thermoplastics, polyether-ketone-ketone (PEKK) is
attractive for its high values of the glass transition temperature (Tg), strength, stiffness and fracture toughness, its low moisture
absorption, and its good environmental resistance. Nevertheless, UV accelerated weathering condition decreases the storage
modulus of a carbon fiber PEKK-matrix composite from 40 to 10 GPa and decreases the Tg from 147 to 1051C, thereby causing
decreases in the modulus and the allowable service temperature.20

2.6 Mechanical Degradation


This article focuses more on the ageing of FRPs under environmental loads, and, strictly speaking, mechanical degradation is not
part of ageing. Nevertheless, it is often a consequence of ageing. Mechanical damage processes such as matrix cracking, delami-
nation, plastic strain, interfacial failure are generally irreversible. In an FRP, some of these damage modes may be seen to be benign
or subcritical; however, the damage may accumulate or lead to damage elsewhere by transferring load ultimately leading to failure.
The most common mechanical degradation, especially when operating at high temperature, is the formation of cracks in the
8 Degradation of Polymer Matrix Composites

matrix either within the ply or transverse to the ply. Fatigue, environmental loading, and residual stresses can all promote the onset
and accumulation of these cracks. The laminate strength, stiffness, and thermal properties as well as failure modes can be affected
by transverse matrix cracking which can also promote higher uptake of moisture deeper in the laminate. The prediction of these
failure modes are often parts of a long-term durability assessment, especially under fatigue loads. The reader is directed towards
Jones,1 Mathews and Rawlings,4 and ASM Handbook,5 for a description of predicting mechanical failure in FRP laminates.

2.7 Fire Testing


While the above sections discuss various environmental forms of ageing, FRPs are often, perhaps surprisingly, used in high-risk fire
applications. These include the aerospace industry in the event of on-board fires and crashes, marine vessels with very high
requirements for resistance to fire, railways, where the fire resistance for underground trains is perhaps the most stringent, and
offshore oil and gas requirements. Many of these fire-resistance requirements arise from unfortunate incidents. Glass and carbon
fibers do not burn but the resin may well set alight or give off noxious fumes or both. Different resins have different fire resistance
to jet fires (where a flame is concentrated on one spot) and pool fires. The combination of low thermal conductivity, resin
endotherm effects, structural integrity and coatings mean FRPs can have excellent fire-resistance performance but this does need to
be strictly controlled using established small-scale tests to evaluate the combustibility and ignitability. Models based on the energy
flux involved in decomposition have been used to predict fire performance of materials.21 However, no further discussion on fire
testing is given in this article because it goes beyond the scope of less severe forms of ageing and degradation.

2.8 Synergistic Effects


Each of the sections above described individual possible ageing mechanisms that may occur to an FRP laminate. However, an FRP
component is a combination of fibers, the fiber–matrix interface, and the polymer matrix that may be subject to long-term ageing
of a combination of the above ageing scenarios. Hence, the mechanical properties and fitness for purpose may change through a
combination of thermal oxidation, physical ageing, hygrothermal effects, UV exposure, and chemical attack. Therefore, macro-
scopic changes need to account for the synergies between these ageing mechanisms and any accelerated ageing methodology
should really account for these combined factors – a nontrivial task!

3 Accelerated Ageing

As discussed above, one of the biggest advantages of composite materials is their long life in harsh environments. Ironically, it is
this long life and the difficulties in being able to reliably predict the long life that has hindered the uptake of composites. The
development of a robust accelerated ageing methodology is a key technology development for FRPs. An accelerated-ageing
methodology should accelerate the action of degradation factors without altering the underlying molecular mechanisms of ageing.
The analogy that a chicken does not hatch quicker from the egg by boiling the egg applies here. As discussed, the ageing process
may cause both reversible and irreversible changes and the accelerated ageing methodology must simulate these processes. The
design of an accelerated methodology relies on the knowledge of material properties and their relationship with environmental
conditions (temperature, pressure, loads, relative humidity, UV, etc.) and upon the fundamental ageing or damage mechanisms as
described in the previous section.
The purpose of accelerated ageing methodology is simply to speed up the accumulation of damage or deformation, potentially
leading to failure. This is achieved by establishing relationships between time and various parameters, such as temperature, fluid
concentration, load, etc. that can be related back to the in-service conditions. Accelerated testing also aims to determine the
material microstructure and damage at the end-of-life or ideally at any time during the life of the component. For most structural
applications, the durability and mechanical fatigue are the main degradation processes in which frequency and load are the main
acceleration factors used. For FRPs in corrosion applications, exposure temperature is the primary means of accelerating ageing, but
other parameters include increasing the concentration of a degradative chemical, applying pressure, or a load. Most accelerated
ageing experiments consider a single accelerating factor because it is difficult to test multiple accelerated conditions and under-
stand the synergistic effects as discussed above. The preferred approach is to incrementally subject samples to accelerated con-
ditions. For the approach to be valid, a fundamental understanding of material response and the reversible and irreversible
degradation processes is required and needs to be incorporated into the modeling.
An accelerated test methodology must follow the following steps22:

(a) identification of the material, including the polymeric matrix and the fiber (including type, volume fraction, orientation,
sizing, etc.);
(b) identification of the primary ageing mechanism for the environment, for example, thermooxidation, hydrolysis, UV ageing,
etching,…;
(c) identification of the physical properties, most important to the application, to be measured in the accelerated ageing test (e.g.,
mass loss, stress, strength, toughness, leak resistance, color, etc.);
(d) selection of the environmental acceleration of degradation factors such as temperature, time, concentration, etc.;
Degradation of Polymer Matrix Composites 9

(e) carry out the accelerated tests under conditions given in (c) and (d);
(f) use theoretical modeling (e.g., Arrhenius relationship, time–temperature superposition, curve-fitting) of the degradation
factors; and
(g) validating some of the modeling by performing long-term tests that are representative of in-service conditions.

For accelerated ageing, the synergistic effects of time, pressure, and conditions on FRP degradation should be established. Free
space exists between molecular chains to a greater or lesser degree leading to a balance between stiffness and flexibility. However,
this same free space can mean that polymers can absorb fluids to which they are exposed, especially those with similar solubility
parameters. Such absorption can physically weaken the polymer to provide an effect of ageing. In addition, the fluid might
chemically attack the polymer to provide an additional effect. The kinetics of these two ageing effects is governed by: (i) diffusion,
and (ii) chemical kinetics, both of which are governed by Arrhenius relationships with regard to the influence of temperature. This
characterization is commonly based on weight loss and, where possible, mechanical property reduction measurements as a
function of time and temperature. The rate of weight loss (W) versus time can be modeled by using an Arrhenius relationship,
Eq. (2).11
  
Ea
W ¼ A exp ð2Þ
RT
where W represents the ratio between the weight loss and the starting weight of the sample and A is an empirical factor. However,
because the thermooxidative degradation in an FRP laminate occurs at the surface, the above equation does not necessarily relate
to the same reduction in mechanical properties that are dominated by the bulk properties, and not just the surface properties. A
good example of this is an interlaminar fracture test such as the double cantilever beam (DCB)16 where the fracture data are
measured from a delamination in the centre of the specimen. During an ageing test, much of this delamination is not exposed to
the environment.
There are few published works on experimental work using a change in a specific mechanical property of relevance, say
modulus or strength in Eq. (3), to allow modeling of the global effect on a component from local ageing degradation at the
surface. Chemical kinetics, classically involving concentrations of reactants and products, can employ the fact that for cross-linked
polymers, concentration of cross-links is approximately proportional to the modulus or stiffness. Hence, measurements of changes
in modulus from ageing can be plotted logarithmically against linear time (for first-order reactions) at each temperature. From a
series of such ageing plots at different temperatures, times to attain the same degree of modulus change can be used to develop the
Arrhenius plot, Eq. (3).
l Ea
ln ¼A ð3Þ
t95 RT
where t95 is the time for a property to reach 95% of its original value (although other values can be used), T is the absolute
temperature, R is the universal gas constant (8.314 J deg1 mol1), A is a constant, and Ea the activation energy. This expression
holds well where there is only one degradation mechanism taking place and this is seldom the situation in the chemical processing
environments.
The mechanics and chemistry of degradation can be different at higher-temperature exposures to that at lower-temperature
exposures. Mass-loss curves can change significantly with temperature because of different mechanism occurring, including
removal of volatiles, additional cross-linking, and structural rearrangements of the matrix. Testing at temperatures too close to the
Tg gives degradation rates that are nonlinear, making estimates of useful remaining life invalid.
One acceleration method for thermooxidative stability is to increase the temperature and the pressure of the oxygen,23 thus
accelerating the rate of degradation due to oxidation. However, other attempts to accelerate ageing in acids by increasing the acid
concentration were less successful, with only small increases in the ageing rate being achieved.24
While the above sections have described the current technology and gaps in predictive methodologies, there are FRP com-
ponents in-service with long duration lives. The approaches used to give some estimation on long-term life include single-point
data (properties after a specific set of ageing conditions) and short-term ageing extrapolation. The following section describes some
applications where ageing has been considered.

4 Ageing Associated With Supersonic Flight

Many of the earliest applications of carbon FRPs were in military aircraft where the main ageing mechanisms were hygrothermal.
As discussed above, this reduced the modulus of the matrix lowering some mechanical properties. As a result, most aircraft with
FRP components use hot wet properties as part of the design allowables. Several aerospace programmes have been concerned with
commercial supersonic aircraft to replace the Concorde. Long-term thermooxidative ageing became a key technology development
for these aircraft. The first major study on ageing of FRPs for supersonic aircraft was performed by Kerr and Haskins.25 The
materials investigated included carbon- and boron-epoxy and carbon polyimide that were exposed to elevated temperatures for up
to 50,000 h. The effect of altitude was also included representing the reduced oxygen levels at high altitudes. The laminates used
were typically [01/7451]s and temperatures ranging from 122 to 1771C representing temperatures up to Mach 2.4 flight. The tests
performed included tensile, compression, and shear and also tests to study the effects of moisture and creep after the exposures in
10 Degradation of Polymer Matrix Composites

an attempt to identify the thermooxidative ageing mechanisms. These tests were chosen as those typically used in an aircraft design
process as opposed to those required to identify the effects of ageing.
For some materials, edge cracking and severe property degradation was identified at temperatures of 1771C after 5000 h of
ageing with a 0.1 MPa pressure. This degradation was not noted after 25,000 h at the same temperature but at 0.014 MPa.
Degradation was observed for epoxy specimens at times under 5000 h when aged at 1771C, which was close to the Tg of materials.
Degradation was a severe crumbling of the matrix at the specimen surface. Ageing damage was far less at 1221C after many more
hours exposure. This demonstrates that raising the temperature to accelerate ageing may not be a valid approach when the
exposure temperature reaches Tg. The polyimide-based specimens had much better thermooxidative stability. At 2321C, after
50,000 h the tensile properties decreased, although no macroscopic damage to the resin could be observed, although clearly some
degradation had occurred from the weight-loss results. In this instance, the selection of a matrix-based mechanical test during the
ageing program would perhaps identify matrix degradation after shorter exposure times.
During NASA’s High-Speed Research (HSR) program in the 1990s, studies were conducted that focused mainly on high
temperature polymers for a new generation of High-Speed Civil Transport (HSCT) aircraft with a speed target between Mach 2.0
and 2.4 with a design requirement of 60,000–120,000 h. The work in this program focused on new materials' development for
high temperature use and included carbon bismaleimides and polyimides.
One of the primary difficulties in characterizing the effects of ageing on changes in structural properties is the correlation of
‘ageing’-related data, such as weight loss, to changes in material mechanical properties such as strength, stiffness, and toughness.
An investigation in the change of lamina properties (E11, E22, and G12), Tg and weight loss for this program can be found
elsewhere.25 An IM7/8320 carbon thermoplastic and an IM7/5260 carbon bismaleimide were aged in air-circulating ovens at 125
and 1751C for 5000 h (representing 10% of the life of proposed life for HSCT). In this time frame, the materials showed little signs
of ageing at 1251C. The IM7/5260 had a 2% weight loss after 5000 h at 1751C. This corresponded to a 10% decrease in E22 and a
2% decrease in G12, but no significant change in Tg or E11.
A test program to evaluate the synergistic effects of stress, temperature, moisture, time, radiation, and oxygen level on the
properties of bismaleimide composites was conducted as part of the HSR program.13,14,23,26–28 Isothermal ageing as well as
thermal cycling, and creep tests were conducted at temperatures up to 2501C. As for the Kerr and Haskins work, exposure
temperatures close to the materials Tg led to very rapid ageing and anomalous effects that made it difficult to develop an
accelerated ageing approach.
Martin29 has investigated the delamination onset under thermal mechanical fatigue with modeling used to predict delami-
nation onset. To verify the methodology, isothermal statics tests were conducted at temperatures of 125 and 1751C (representing
Mach 2 and Mach 2.4 flight, respectively) on quasi-isotropic laminates fabricated from carbon bismaleimide (IM7/5260). Raising
the temperature had the effect of decreasing the value of strain energy release rate (G) at the edge by reducing residual thermal
stresses. However, the materials fracture toughness was also lower at higher temperatures. The isothermal ageing results of these
materials discussed by Kerr et al.25 would need to be incorporated as additional effects of ageing. These would need to be included
in addition to the change in properties related to temperature and fatigue.
For aerospace applications, open-hole tension and compression of laminates are important properties for damage tolerance
designs. As discussed above, thermooxidative ageing attacks the surface and an investigation was conducted by Morgan et al.27 to
determine the influence on isothermal ageing of laminates with a hole drilled before and after ageing and edge cut before and after
ageing. Quasi-isotropic laminates fabricated from IM7/5260, IM7/8320, and IM7/K3B a carbon polyimide added to the HSR
program were aged at temperatures of 125 and 1751C (additional exposures at 2001C were carried out on the IM7/K3B) for
periods up to 5000 h. For IM7/5260 and IM7/8320, significant damage was seen on the surfaces and this degraded the stress to
cause delamination at the edges but did not affect ultimate strength of the laminate, because of the presence of 01 fibers in the
loading direction. Removal of these damaged edge by cutting them off brought the edge delamination stress back to the unda-
maged value demonstrating that thermooxidative ageing begins at the surface. For the K3B laminate, no edge delamination was
observed before lamination failure reflecting the higher toughness of this material and hence the better resistance to thermo-
oxidative ageing. Little difference was identified whether the holes were drilled before or after ageing. While there would be some
damage to the matrix within the holes after ageing, the failure is still dominated by the local stress concentration on the 01 fibers at
the hole.

5 Ageing in the Oil and Gas Industry

FRPs are key candidate materials for replacing carbon steel in the Oil and Gas Industry because of their good corrosion resistance
and light weight. However, this industry is very risk averse and the uptake of these materials has been slow and is directly related to
the operational risks for that component. The shortage of long-term performance data of FRPs in oil and gas applications does not
aid their uptake. However, the potential improvements in component and system performance using FRPs will ultimately reduce
capital and maintenance requirements, and this fact has enabled some usage of composites in the industry.30,31
The environments in the oil and gas industry can be very harsh. The fluids with which FRPs might come into contact include
those illustrated in Fig. 2, along with some of the potential ageing effects. The composition of crude oil varies around the world
from well to well. It mainly comprises aliphatic alkanes (hydrocarbons) such as hexane, octane, decane, etc. Aromatic hydro-
carbons may also be present and these are known to swell epoxy matrices. The presence of corrosive media such as hydrogen
Degradation of Polymer Matrix Composites 11

Fig. 2 Possible fluids used in the oil and gas industry and their effects on polymers.

sulfide or even carbon dioxide in water (carbonic acid) can chemically attack the entire composite, especially if a GRP is under load
promoting environmental stress cracking/corrosion (ESC). In addition, most oil and gas operations are conducted at high tem-
peratures (up to 2001C) and at high pressures, helping to accelerate ageing of the material.
The notable uses of FRPs in this industry are in pipeworks, gratings, and blast and fire protection. Apart from hundreds of
kilometers of GRP pipe used for transporting hydrocarbon and water lines in the Middle East, FRPs are also being increasingly used
to line steel pipes where the FRP provides corrosion protection for the pipe bore by acting as a barrier to the passage of transported
fluids. The composite-lined pipe solution is in direct competition with conventional corrosion-resistant alloy (CRA) systems for
downhole and flowline/pipeline applications. Statoil Hydro pioneered the use of downhole tubing with composite materials in
the North Sea. A Duoline 20 liner was examined after five years of service at up to 1101C and was found to be in excellent
condition.32 The technology has generally been confined to water injection service. A consortium project has been run by Oil States
Industries UK to help understand the chemical compatibility of various types of FRP exposed to different oil field fluids, with the
aim of obtaining 20-year life.33 These liners see a variety of hostile media including hydrocarbons at temperatures up to 1601C,
and short exposures to methanol, HCl/HF, and other fluids for cleaning operations. As part of the materials screening tests
conducted at MERL, Hitchin, UK, single-point data was generated for specific conditions such as several days at 1601C in
hydrocarbon and brine mix, 8 h of HCL exposure, etc. Several postexposure tests were conducted to ascertain surface degradation,
weight change and mechanical property change of the liners. A picture of a glass-epoxy liner exposed to HCl for 8 h shows
complete degradation (Fig. 3).
GRP piping onshore is primarily filament-wound glass-epoxy and it has been successfully installed in 100 km runs in the
Middle East for transport of hydrocarbons and other fluids. One of the shortcomings for these pipes is that the corrosion science of
metals is much more developed and understood than the ageing and degradation of these materials. International standards exist
for the design and installation of such pipework where short-term tests are used to account for such degradation and sufficient
safety factors employed. DNV-OS-C501 standard discusses the various effects on mechanical properties under the influence of
temperature, moisture and chemicals, and suggests that “… the degradation rates shall be obtained for the actual materials in
question…” but gives no method for quantifying these effects in the long term. The ISO 14692 Part 2 Annex D provides
information on defining partial factors A1/A2 to account for temperature and chemical resistance, respectively. These factors are
used to give knockdowns on properties that were measured at standard temperatures. While values for determining A1 are well
established, A2 can be generated for water but is less well identified for other chemicals. The standard makes recommendations of
12 Degradation of Polymer Matrix Composites

Fig. 3 Section of glass-epoxy liner submitted to an HCl exposure. Courtesy of MERL.

Fig. 4 A schematic of GRP degradation when transporting a hostile fluid.

other standards (such as ASTM C 581) that can be used for deriving these factors for other chemicals, although these standards
were not specifically written for determining A2 (see Section 6). While these approaches have allowed GRP piping to be specified,
it is generally accepted that an improved approach of quantifying such degradation is needed.
Ageing from contact of the transport fluid in these GRP pipes occurs from the surface inwards and requires time to penetrate
into the material’s centre, as discussed above. The ageing or degradation of an FRP pipe has many stages and it is up to the user to
determine at which phase failure is said to have occurred, Fig. 4. Almost immediately at Stage 1, the fluid will diffuse into the
polymer causing immediate physical changes. This is a complex phenomenon for an FRP material in contact with mixed fluids and
is discussed further below. In Stage 2, the fluids would have diffused further into the FRP and the inner surface might have begun
to chemically age. In Stage 3, the ageing is such that it has caused mechanical damage on the inner surface. Fluid diffusion and
chemical ageing continue in Stage 3, but with the presence of cracks the fluid can now penetrate quicker. In Stage 4 diffusion,
chemical ageing and mechanical damage has continued, but now the mechanical damage on the inner surface is sufficient for
material to be removed. Stage 4 also shows cracks on the outside of the pipe from other damages such as erosion or UV
degradation. Stage 5 shows complete damage through the pipe wall and subsequent leaking.
Degradation of Polymer Matrix Composites 13

Fig. 5 A glass-phenolic grating on the Mars Platform in the Gulf of Mexico, Courtesy of Strongwell.

In most pipe designs, ‘failure’ should be considered as somewhere between Stages 1 and 3 and the influence on strength of the
system evaluated to allow for sufficient safety factors (e.g., A2) to be applied.
In order to predict the life of an FRP laminate the following information needs to be known:

• The aged condition of the material and the profile through the thickness. This can be determined by knowing the rate of ageing
(physical and chemical) for a given set of conditions (fluid type, temperature profile, time, etc.).
• The key material properties at a given amount of ageing. This depends on what is classed as the residual life of the pipe
(see next bullet). The mechanical properties that are used to determine this residual life are those that have to be
measured.
• The residual life of the structure with given damage and degradation profile; this is dependent on what is considered as failure
for example, leak, global change in stiffness, % loss of material, inner cracking.

The technology for generating the information above does not exist at the time of publication.
The advantages of FRP gratings used offshore are obvious in the improved resistance to corrosion in a sea water
environment and their durability and light weight, compare to metallic options, Fig. 5. The products are generally pultruded glass
reinforced, offering a cost-effective manufacturing method. The use of phenolic resins allows the parts to meet fire-resistance
requirements offshore. An important driver for GRP gratings has been the ease of installation compared to steel. The ageing
of these materials follows the same aspects as marine composites discussed in Section 7 and includes hydrolysis and UV ageing.
The fire resistance of these phenolic gratings is also a key reason behind their selection, but fire properties goes beyond the scope of
this article.
Composites are also being increasingly used to protect subsea structures such as wellheads, and flowlines, protecting them from
impact events such as dropped objects and commercial fishing operations. This is an accepted area of use where the perceived risks
of adoption are outweighed by the benefits in performance and corrosion resistance. Another example of the use of composites in
hostile environments is a ProTek(tm) Jet Fire and Blast protection enclosure manufactured from composite panels by Solent
Composite Systems Ltd., Fig. 6. This enclosure was installed at the LNG plant in the Arctic town of Hammerfest in northern
Norway to protect emergency shutdown equipment. The composite structure resists an explosion pressure equivalent to 7 tonnes
per square meter and protects against 90 min exposure to the erosive and heating effects of a hydrocarbon jet fire with a flame
temperature of 11501C resulting from a high pressure gas leak. During this period the equipment temperature does not exceed
651C. The ageing aspects of long-term water exposure, biofouling, and hydrolysis are covered in Section 7.

6 Ageing in the Chemical Processing Industry

In the chemical processing industry (CPI), the environments for equipment such as reactor vessels, storage tanks, scrubbing towers,
stacks, piping, valves, etc. may be extremely harsh (Fig. 7). In many instances, corrosion-resistant alloys including highly alloyed
stainless steel, titanium, and nickel-based alloys have to be used and even these can corrode in these environments. One solution
adopted in this industry is the use of FRPs in manufacturing this equipment.
14 Degradation of Polymer Matrix Composites

Fig. 6 Composite fire-blast protection enclosure. Reproduced with permission from Solvent Composite Systems Ltd.

Fig. 7 FRP scrubbing tower. Courtesy of Dow Chemicals.

FRP materials are used to a large extent in plants which manufacture chlorine, chlorate and concentrated acids (e.g., sulfric,
hydrochloric, hydrofluoric, nitric) as well as metal chloride solutions (e.g., NaCl, FeCl3, AlCl3, MgCl2, NiCl2).34 FRPs are now also
used in desulfrization plants (flue-gas ducting, scrubbers, etc.) which also have applications in oil and gas production. In many
Degradation of Polymer Matrix Composites 15

Fig. 8 Failure in alkaline aqueous solution at 701C. Courtesy of G. Bergman.

instances, a thermoplastic liner is used in the metallic and FRP pipes to act as a corrosion or permeation barrier, or both. The
thermoplastic lining may be PA11, PVC, or PP; or for more aggressive service, PVDF. Because of the environmentally hostile nature
of some of the chemicals being transported or stored, failure is unacceptable. However, failures do occur (Fig. 8).35 These failures
may not only be very costly, but also present a health and safety risk to the workers at the plant, local residents and environment.
The consequences and liability of equipment failure, even minor leaks, are becoming increasingly severe and have resulted in a
very strict regulatory climate. Although FRP is often used to solve problems of corrosion on various metallic materials, these
materials can still be affected by these fluids.36
The types of resins and fibers depend very much on the application being considered. Thermosetting resins are suitable for large
moldings and for filament wound pipes. The most important resin systems used for corrosion protection in the CPI include
polyesters, epoxies, bismaleimides, phenolics, and vinyl esters.34,37 Epoxies have certain advantages, in that they have lower cure
shrinkage, adhere better to steel substrates, and offer better chemical resistance than polyesters towards hot alkalis. Vinyl esters are
favored because they offer a compromise between epoxies and polyesters, offering good corrosion resistance with moderate
relative cost. E-glass fiber is used as the reinforcement for the main structural part with a more chemical resistant fiber such as ECR-
glass used in the layer in contact with the corrosive fluid. A protective resin-rich layer is often used on the surface adjacent to
the acid.
Prediction methods for FRPs in the CPI must account for chemical ageing and include diffusion, hydrolysis and ultimately
cracking, blisters and other damage. Eventually, this degradation may continue and lead to actual material loss from the inside of
the vessels (Fig. 9).31
Degradation in FRP composites may be defined as one of any of the following in the exposed surface:
The published work related to chemical resistance of FRP materials in different environments has focused on immersion
testing, rather than the more realistic single-sided exposure. Much of the published data quote only single-point data, for example,
“The mass of material X increases by 10% in methanol after 10 days at 601C.” This information may be misleading because the
overall trend may involve a mixture of competing mechanisms as discussed in previous sections. For hybrid materials (e.g., the use
of a corrosion-resistant veil), the prediction method must account for the time for the ageing to reach each layer.
Further, FRPs may be subject to step changes in property changes leading to failure such as resin embrittlement. This failure is
often termed environmental stress cracking (ESC) and also applies to the glass fibers when undergoing strains in excess of 2% and
exposed to even dilute acids.38 In any long-term evaluation, it is necessary to use diagnostic equipment to ensure that the cause for
the change in properties of the FRP is understood.
Estimates of remaining life of FRP equipment in the CPI often rely on periodic visual inspection, intuition, and experience of
the inspector. The inspector searches out defects such as blistering, delamination, or signs of leaking. Often a bright light is shone
on the opposite side of the laminate to help reveal flaws. The inspector often also checks the condition of the laminate surface with
a Barcol indentation tester. Drastic reduction in Barcol readings since the previous inspection indicates degradation. On the basis
of these findings, the inspector will judge whether the equipment is still safe to operate and when the next inspection should
occur.36
16 Degradation of Polymer Matrix Composites

Fig. 9 Uniform corrosion in an FRP pipe after 14 years of service. The white line shows the original thickness. Courtesy of G. Bergman.

6.1 ASTM Standard for Long-Term Chemical Resistance


Materials are generally approved for long-term usage in the CPI using ASTM C 581 Standard Practice for Determining Chemical
Resistance of Thermosetting Resins Used in Glass-Fiber-Reinforced Structures Intended for Liquid Service. This standard requires immersion
of a material in a fluid at a single temperature. Various properties (Barcol hardness, flexural modulus and strength, and Tg) are
determined at intervals normally within one year. If the properties do not decrease by a certain amount, the material may be
approved for long-term usage. However, this method does not offer an approach to allow an extrapolation to longer-term usage at
varied of temperatures and is therefore little more than a screening test.
To demonstrate the use of ASTM C 581, an FRP laminate that may be used for the linings of tanks and vessels used to store
acids was tested.36 Plaques of materials were immersed in concentrated HCl or H2SO4 for 12 months. Periodically, Barcol
hardness, mass change, and flexural properties were measured on specimens cut from these plaques. The mass change is shown in
Fig. 10, illustrating that the material shows weight loss early in the exposures and the rate of weight loss begins to reduce as the test
progresses. The change in Barcol hardness, shown in Fig. 11, shows that there is an initial increase in both exposures, indicating
that some form of local hardening has occurred. After 6 months’ exposure, the hardness begins to decrease, leading to 10%–15%
from the starting values after 12 months. This indicates that there is more than one ageing mechanism occurring and the resulting
change in properties cannot be taken from only the end data point.

6.2 The Arrhenius Relationship


Using the different test parameters used in ASTM C 581 in the previous section, the feasibility of using an Arrhenius relationship
was presented elsewhere.36 Key to this approach being successful is that the elevated temperature exposures do not cause a change
in the ageing mechanisms observed at lower temperatures.
A Derakane 411 vinyl ester resin was tested with two plies of 45g E-Glass chopped-strand mat, with a 0.25-mm C-glass veil. The
materials were exposed to 37% HCl at temperatures of 40, 60, and 801C. Before and after the exposures, thickness and weight in
air and water were measured to determine any changes. Following the ageing and the weighing, suitable specimens were cut from
the plaques. Three point-bend tests were conducted and showed that the modulus initially increased before decreasing, whereas at
the higher temperatures, the modulus and strength continually decreased, showing a change in mechanism at the different
temperatures. The work demonstrated that there is an overall trend from test to test of a property change with exposure time, the
rate of which increases with time. It is this general concept that lays foundation for developing accelerated ageing approaches using
the Arrhenius equation if a single mechanism is present. The 5% and 10% change of flexural strength were plotted against the
reciprocal of the absolute exposure temperature and a straight line fitted. The shorter-term, higher-temperature exposure tests were
able to predict the properties at lower-temperature exposures. The energy of activation for this change is in the region of 80 kJ
mol1, which is indicative of chemical ageing. However, an Arrhenius curve cannot be drawn for the modulus because of the
initial increase in modulus in the early exposure times, this invalidating this method unless the different mechanisms are
separated.
Degradation of Polymer Matrix Composites 17

Fig. 10 Mass change of an FRP in concentrated acid.

Fig. 11 Percentage change in Barcol hardness of an FRP in concentrated acid.

6.3 Using a Semiempirical Corrosion Approach


Another approach to the long-term ageing of FRPs is to relate the effects of ageing to those of corrosion in metals, because on the
macrolevel the results are very similar. Many types of corrosion found in metals can also be found in FRP, such as uniform
corrosion (material loss), localized corrosion (pitting), selective corrosion, stress corrosion, corrosion fatigue, erosion corrosion,
and layer corrosion (delamination). CFRPs are conductive and galvanic corrosion effects may arise when being coupled to a metal
component.
On the basis of the results of corrosion analyses of samples taken from FRP equipment used for different chlorine dioxide
environments in various pulp mill applications over a period of 20 years, a semiempirical relationship for uniform corrosion
behavior of FRP in chlorine dioxide environments has been established,39
 
Ea
F ¼ Bt a cA exp ð4Þ
RT
18 Degradation of Polymer Matrix Composites

where F is the depth of corrosion (mm), B is a special factor in case of protective deposits on the surface (usually B¼ 0 or 1), t is
the time in service (years), a is a factor which depends on the thickness and degree of degradation of the corroded surface layer
(usually a is between 0.5 and 1), c is the concentration of chlorine dioxide (g l1), A is a material constant which depends on the
type of resin, the degree of curing, and the laminate structure, Ea is the activation energy of the rate-controlling step of the
corrosion process (kJ mol1), R is the general gas constant (8.3 J mol1 K), T is the temperature (K).
This expression is not proven for other applications and environments, although the premise should hold.

7 Ageing in the Marine Industry

One of the more established uses of FRPs with long and chopped fiber resins in marine vessels from small dinghies to racing
yachts. A review of marine composites is given elsewhere.40,41 The FRP materials are mainly E-glass reinforced polyester and epoxy,
but with the high-performance vessels carbon fiber composites and a range of sandwich materials are now used. Almost always, a
protective outer coating or gelcoat is used to help prevent diffusion and to give a better appearance. There is much experience on
the ageing of these materials in this environment. However, similar to the studies in other industries, research studies on ageing for
this industry have focused on short-term ageing tests under severe conditions that may have little relationship with in-service use.
Leisure craft or pleasure boats, defined as vessels built for recreational purposes, were one of the first applications of FRPs. The
average lifetime of a FRP pleasure boat is 30 years, demonstrating the anticipated long use life of these materials. For commercial
vessels including fast ferries, lifeboats, fishing boats, and some military vessels have been made from a conventional GRP material
since the 1960s. Some vessels such as mine countermeasure vessels are in excess of 40 m length and the nonmagnetic and
conducting properties of the GRP add another advantage to the use of these materials. The low weight of FRP materials make them
attractive for submarine use in internal and external structures such as sonar domes.
Woven or chopped glass strand mats in a polyester resin is the typical FRP use in marine applications and much work on ageing
of these materials in a sea water environment has been conducted.42 Higher-end materials, including stitched multiaxial cloth,
carbon epoxy are used for the higher end applications where ageing may be less of a concern such as in racing yachts. The
document also provides a database of material properties to be used if test data are not available. Typical hull laminate thicknesses
vary from 5 mm for pleasure craft up to 150 mm for hulls that must survive underwater explosions. The difference in thickness has
an affect on the time to reach equilibrium and hence begin hydrolytic ageing. Manufacturing techniques have been developed over
the past 10 years to utilize resin infusion to draw the resin through the dry fibers. This has resulted in the ability to have a higher
fiber to resin volume fraction giving thinner laminates with the same strength as the previously used wet lay-up manufacturing
methods with lower volume fractions. Sandwich panels comprising a core (foam or balsawood) surrounded by thin layers or skins
of FRP are a preferred manufacturing method for large hulls. However, the thin skins, their bonding to the core and degradation of
the core material present further concerns for ageing. As mentioned above, a gelcoat is almost always used and this serves to
protect the laminate.
There are a number of different damage modes for marine composites and the gelcoats, including osmosis, blisters, pinholes,
wrinkling, debonding/delamination, crazing, etc., can all be considered a form of ageing. Blisters are a particular ageing damage
mode in marine environments, where the osmotic pressure is developed between the gelcoat and the FRP. This is built up from
products such as glycols leached from the base FRP and can cause local delamination or blisters and is particularly apparent with a
gelcoat that is more permeable than the base laminate. The temperature and the composition of the seawater, which varies around
the world, are important aspects in consider ageing of marine FRPs as well as the loading from wave load, slamming load,
hydrostatic pressure, etc. Antifoul treatments are used on many boats to prevent biofouling or marine growth. These are applied to
the gelcoat and may in themselves age the coating.
Conventional marine ageing of a marine FRP follows the same degradation mechanisms as discussed in Section 2.2. This is
physical ageing causing plasticization and swelling (potentially leading to interfacial debonding and delaminations) and chemical
ageing such as hydrolysis of the matrix and, in saltwater environments, the fiber and the fiber–matrix interface. Many of the ageing
studies are conducted at higher temperatures to reduce the ageing times. A number of case studies are presented in Searle and
Summerscales,42 where an attempt to better understand the changes due to ageing of properties that can be used for design were
studied. In one study, weight change was measured along with shear strength and modulus for three GRPs (polyester, vinyl ester,
and epoxy, all with the same glass reinforcement).
Fiber reinforced polymer (FRP) composites have attracted great attention in marine applications, such as yacht marina, offshore
oil and gas platforms, rapid island buildings, and so on. They are suitable materials for structural applications where high
performance and long service life are required. FRP composites used in marine applications should retain their mechanical
properties. However, FR composites are susceptible to heat, moisture, NaCl molecules, algae, and ultraviolet (UV) radiation when
they are exposed to the marine environment. The major issue of FRP composites used in the marine environment is that the
interface between the fiber and the matrix would be debonded, thus the thermo-mechanical properties of composites are
affected.43
A work carried out by Searle et al.42 demonstrated that the resin and fiber interface within an FRP absorbs water, not just the
resin. In the same work, FRP materials were exposed to an accelerated exposure (elevated temperature), and the same panels were
exposed to the sea for up to a year. The results suggest that without due attention to the various ageing conditions, the accelerated
test will overpredict changes in properties.
Degradation of Polymer Matrix Composites 19

For realistic accelerated ageing of marine FRPs, several aspects need to be considered. Virtually, all laminates, as stated above,
will have a gelcoat. While not an impermeable layer, it is a resin-rich layer and needs to be considered in ageing analysis. The work
in Searle and Summerscales42 demonstrated that the composition of the water can greatly affect the ageing of the FRP. Seawater
will diffuse into polymers slower than distilled water, and ageing exposure using distilled water may be overly conservative. In
addition, during ageing, products may leach out of the FRP into the water. These products will themselves change the composition
of the water more significantly if small vessels and hence small volumes of water are used. As for other applications, the use of
temperature as an ageing accelerator must be used with caution. The increase in temperature can serve to promote further cross-
linking in partially cured polyesters and use of ageing temperatures near or above the Tg can cause degradation mechanisms that
will never occur at lower temperatures. For true comparisons for sea environments, it is the inclusion of biofilms that will develop,
followed by biofouling, resulting in a complete covering. This needs to be considered when conducting weight-gain/loss mea-
surements and the influence of the living matter on the properties of the composites needs to be better understood.

8 Conclusions

This article has described some of the more common forms of ageing of FRPs. Ageing is a summation word for degradation and
can include physical changes of the matrix such as swelling or irreversible chemical changes such as hydrolysis or oxidation. The
extent of the degradation is very much dependent on the exposure the FRP material will experience including the fluid in contact,
time, pressure temperature, load applied, if any, etc. Some attempts have been to develop predictive modeling capability for the
long-term durability assessment of FRP components, but this still remains an area for further research. As the use of FRP materials
increases and the demands for extended performance grow, so does the need and importance to develop predictive capability for
assigning the fitness for the purpose of an FRP structure at any time during its service.

References

1. Khatoon, H., Ahmad, S., 2017. A review on conducting polymer reinforced polyurethane composites. Journal of Industrial and Engineering Chemistry 53 (Suppl. C), 1–22.
2. Bigdeli, M.B., Fasano, M., 2017. Thermal transmittance in graphene based networks for polymer matrix composites. International Journal of Thermal Sciences 117 (Suppl.
C), 98–105.
3. Jones, R.M., 1975. Mechanics of Composite Materials. Washington, DC: Scripta Book Company.
4. Matthews, F.L., Rawlings, R.D., 2002. Composite Materials: engineering and Science. Cambridge, England: Woodhead Publishing.
5. ASM Handbook, 2001. Volume 21 Composite. Ohio: ASM International.
6. Lu, T., Solis-Ramos, E., Yi, Y., Kumosa, M., 2016. Synergistic environmental degradation of glass reinforced polymer composites. Polymer Degradation and Stability 131,
1–8.
7. Katunin, A., Wronkowicz, A., Bilewicz, M., Wachla, D., 2017. Criticality of self-heating in degradation processes of polymeric composites subjected to cyclic loading: a
multiphysical approach. Archives of Civil and Mechanical Engineering 17 (4), 806–815.
8. Struik, L.C., 1978. Physical Ageing in Amorphous Polymers and Other Materials. New York: Elsevier Scientific.
9. Gates, T., Martin, R., 2008. Ageing of Composites. Cambridge, England: Woodhead Publishing (Chapter 1).
10. Jones, F.R., Pritchard, G., 1999. Reinforced Plastics Durability. Cambridge, England: Woodhead Publishing (Chapter 3).
11. Mensitieri, G., Iannone, M., Martin, R., 2008. Ageing of Composites. Cambridge, England: Woodhead Publishing (Chapter 9).
12. Harris, C.E., Gates, T.S., 1993. High Temperature and Environmental Effects on Polymeric Composites. American Society for Testing and Materials Pennsylvania ASTM
STP 1174.
13. Martin, R.H., Siochi, E.J., Gates, T.S., 1992. Proceedings of the American Society for Composites Seventh Technical Conference, October 13–15, University Park
Pennsylvania, pp. 207–217.
14. Tsotsis, T., Martin, R., 2008. Ageing of Composites. Cambridge, England: Woodhead Publishing (Chapter 5).
15. Khanlou, H.M., Woodfield, P., Summerscales, J., Hall, W., 2017. Consolidation process boundaries of the degradation of mechanical properties in compression moulding
of natural-fibre bio-polymer composites. Polymer Degradation and Stability 138 (Suppl. C), 115–125.
16. Ghasemi, A.R., Moradi, M., 2017. Effect of thermal cycling and open-hole size on mechanical properties of polymer matrix composites. Polymer Testing 59, 20–28.
17. Weitsman, Y., 1987. Journal of the Mechanics and Physics of Solids 35 (1), 73–93.
18. Layton, J., Pritchard, G., 1999. Reinforced Plastics Durability. Cambridge, England: Woodhead Publishing (Chapter 6).
19. Halliwell, S., Martin, R., 2008. Ageing of Composites. Cambridge, England: Woodhead Publishing (Chapter 15).
20. Chung, D.D.L., 2017. Processing-structure-property relationships of continuous carbon fiber polymer-matrix composites. Materials Science and Engineering: R: Reports 113
(Suppl. C), 1–29.
21. Dodds, N., Gibson, G., 1999. Composite Materials for Offshore Applications. American Bureau of Shipping. pp. 77–92.
22. Gates, T.S., 2003. On the use of accelerated test methods for characterisation of advanced composite materials, NASA TP-2003-212407 May.
23. Tsotsis, T.K., Lee, S.M., 1998. Composites Science and Technology 58, 355–368.
24. Hogg, P.J., Hull, D., Harris, B., 1983. In: Harris, B. (Ed.), Developments in GRP Technology 1. UK: Applied Science London, pp. 37–90.
25. Kerr, J.R., Haskins, J.F., 1987. Time temperature stress capabilities of composite materials for advanced supersonic technology applications, NASA CR-178272 May.
26. Martin, R.H., Siochi, E.J., Gates, T.S., 1992. 7th Technical Conference on Composite Materials. The Pennsylvania University, American Society for Composites.
27. Morgan, R.J., Shin, E., Dunn, C., Fouch, E., Jurek, B., Jurek, A., 1994. In: Proceedings of the 39th International SAMPE Symposium 1564–1575 April.
28. Martin, R.H., 1994. In: AIAA/ASME/ASCE/AHS/ACE Proceedings of the 35th Structures, Structural Dynamics and Materials Conference, Hilton Head, SC, April 18–20.
29. Harris, C.E., Gates, T.S., 1993. In: High Temperature and Environmental Effects on Polymeric Composites 1993. American Society for Testing and Materials, Pennsylvania.
30. Martin, R.H., 2007. A Technology Gap Review of Composites in the UK Oil and Gas Industry, www.merl-ltd.co.uk.
31. Frost, S., Martin, R., 2008. Ageing of Composites. Cambridge, England: Woodhead Publishing (Chapter 14).
32. Melve, B., Nice, P., 2003. In: Statoil, MERL Oilfield Engineering with Polymers Conference, London.
33. http://medlicott.uk.com/bus/osiuk_jip.htm.
34. Kelly, P., Pritchard, G., 1999. Reinforced Plastics Durability. Cambridge, England: Woodhead Publishing (Chapter 9).
20 Degradation of Polymer Matrix Composites

35. Bergman, G., Corrosion 2004 NACE Paper 04611.


36. Martin, R.H., Martin, R., 2008. Ageing of Composites. Cambridge, England: Woodhead Publishing (Chapter 17).
37. Pritchard, G., 1989. Advanced Composites. Elsevier Applied Science. pp. 163–196.
38. French, M.A., Pritchard, G., 1993. Composites Science and Technology 47, 257–263.
39. Bergman, G., 2007. Journal of Hazardous Materials 142 (3), 695–704.
40. Davies, P., Choqueuse, D., Martin, R., 2008. Ageing of Composites. Cambridge, England: Woodhead Publishing (Chapter 12).
41. Choqueuse, D., Davies, P., Martin, R., 2008. Ageing of Composites. Cambridge, England: Woodhead Publishing (Chapter 18).
42. Searle, T.J., Summerscales, J., Pritchard, G., 1999. Reinforced Plastics Durability. Cambridge, England: Woodhead Publishing (Chapter 7).
43. Fang, Y., Wang, K., Hui, D., et al., 2017. Monitoring of seawater immersion degradation in glass fibre reinforced polymer composites using quantum dots. Composites
Part B: Engineering 112 (Suppl. C), 93–102.

Relevant Website
http://urlm.co/www.vircon-composites.com
URL Metrics.

You might also like