1 s2.0 S1350630722009050 Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Engineering Failure Analysis 144 (2023) 106938

Contents lists available at ScienceDirect

Engineering Failure Analysis


journal homepage: www.elsevier.com/locate/engfailanal

Failure analysis of glass fiber reinforced composite pipe for high


pressure sewage transport
Dandan Liao a, Biao Huang b, Jie Liu c, Xiwen Qian a, Fei Zhao c, Jun Wang a, *
a
School of Mechanical Engineering, Sichuan University, Chengdu 610065, China
b
College of Material Science and Engineering, Sichuan University, Chengdu, Sichuan 610065, China
c
Petro China Planning & Engineering Institute, Beijing 100083, China

A R T I C L E I N F O A B S T R A C T

Keywords: The research was conducted on failed GFRP pipes used in sour oilfield gathering and trans­
GFRP composite pipe portation pipelines, with the matrix material being an epoxy resin of an aromatic amine system
Epoxy resin that destructively cracked after 15 years of operation. To determine the cause of the failure, the
Ageing
failure area of the GFRP pipe was compared to the undamaged area, and the pipe’s micro-
Failure analysis
Sewage Transport
morphological structure, chemical composition, and mechanical properties were examined. The
results show that there are cavity defects in the GFRP pipe with insufficient air bubbles and resin
filling, that the overall resin content of the pipe is low, that the curing degree of the outer layer of
the damaged area (DA-OL) is severely insufficient, and that the fiber bonding strength is weak.
The degradation of the amine curing agent in the outer layer of the pipeline and the oxidative
decomposition of the resin were characterized using Fourier transform infrared spectroscopy and
X-ray electron spectroscopy, and an irreversible chemical degradation process occurred, reducing
the performance of the fiber–matrix interface. The data from nanoindentation and dynamic
thermomechanical characteristics confirmed the obvious reduction of nano-hardness and elastic
modulus in the outer layer of GFRP tubes. The failure of the pipeline is caused by the interaction
of the above factors.

1. Introduction

Fiber-reinforced composite materials use metal or polymer as the matrix material and natural or synthetic fibers as the reinforcing
phase to bear the main load. It achieves outstanding overall performance thanks to the synergistic effect of the two phases combined
with various preparatory technologies. Fiber reinforced composites typically have a high strength-to-weight ratio, great corrosion
resistance, outstanding durability, and wear resistance and tensile qualities, making them appropriate for a wide range of applications
[1–3]. As a reinforcing material, glass fiber has both economy and applicability with excellent mechanical properties and anti-
deterioration properties. In particular, wound-processed glass fiber reinforced polymer (GFRP) pipes are well suited for carrying
fluids that are corrosive to the majority of metals and alloys, effectively avoiding corrosion issues [4]. The petrochemical and oil and
gas industries are increasingly using GFRP pipes. It is frequently used in the transportation of H2S and CO2 containing oil and gas due to
its strong corrosion resistance. As a low-cost substitute for conventional steel pipe materials used in transportation and mining, it has
emerged as a significant method for reducing the corrosion of steel pipes [5–8].

* Corresponding author.
E-mail address: srwangjun@scu.edu.cn (J. Wang).

https://doi.org/10.1016/j.engfailanal.2022.106938
Received 13 September 2022; Received in revised form 1 November 2022; Accepted 12 November 2022
Available online 17 November 2022
1350-6307/© 2022 Published by Elsevier Ltd.
D. Liao et al. Engineering Failure Analysis 144 (2023) 106938

However, due to the intricacy of the pipe and the interfacial interaction of the fibers with the matrix, the production process, mode
of transportation, and environmental factors can readily affect the properties of fiber-reinforced composite pipes. The relevant
literature examines the relationship between various load modes and the best filament winding angle method, and discovered that the
winding is best at 55◦ for hoop pressure load, 75◦ for biaxial pressure load, and 85◦ for biaxial loads comprising axial compression
[9–12]. At the same time, insufficient fiber winding tension will result in weak fiber bonds, and excessive winding tension will break
the fibers. Samanci A et al. [13,14] conducted surface prefabrication of open cracks with different aspect ratios, and studied the fatigue
failure behavior of GRP pipes at stress levels of 50 %, 40 % and 30 % of the ultimate hoop stress intensity of three pipes. They found
that the three processes of fatigue failure of GRP pipes were in turn debonding and delamination resulting in whitening, fiber breakage
and leakage, and finally burst failure [15,16]. Gemi et al. [11,17–19] systematically studied the failure mechanism of GRP pipes
subjected to different shock loads under internal pressure. Delamination, radial and surface matrix crack formation were observed as
the main failure mechanisms on the outer surface of the hybrid pipe. Rafiee et al. [20,21] used the progressive damage modeling
technique of stiffness degradation to evaluate the fatigue failure of GRP pipes. The experimental data verified the applicability and
acceptable level of accuracy of the model in predicting the fatigue life of composite pipes. Residual strength over a span of 50 years of
continuous operation [22–24]. The impact of aging conditions, such as high temperature, water aging, high salt or acidic environment,
on the performance degradation of GRP pipes has been extensively investigated in the literature. The most important effect on shear
modulus is caused by high temperature, which will hasten the attenuation of the modulus and tensile strength of glass fiber reinforced
polymers [25]. The failure process manifests itself as fiber brittle fracture and partial loss of epoxy binder at low temperature (100–150
℃), and sheet split failure at high temperature (200–250 ℃) [26]. The water absorption and hydrothermal aging behavior of GFRP
pipes have been extensively studied in the literature. The relationship between mechanical property deterioration and failure mode has
been investigated in steam environments or high temperature water immersion conditions. It has been determined that damp heat and
water aging can cause fiber matrix debonding and interfacial cracking [27–32].
In the process of pipeline transportation, the failure process will be affected by many factors, such as the quality of the pipeline
itself, the forming process, the corrosive properties of the transport medium as well as the temperature and pressure [6,33]. The
deterioration of pipeline performance will further lead to leakage perforation or blasting rupture, and induce serious safety accidents.
Therefore, it is necessary to investigate the failure causes of engineering pipelines, and further analyze the factors affecting the service
life of pipelines, which can effectively prevent catastrophic accidents and reduce economic losses. In this study, we evaluated the
failure of a water delivery GFRP pipe used in Xinjiang Oilfield, and analyzed the degradation mechanisms present in the assembly to
find out the cause of its failure. At the same time, the factors affecting the service capacity of the pipeline are analyzed, which provides
a reference for the engineering application of GFRP pipes, and provides support for further ensuring the safe and stable operation of the
pipeline.

2. Material and methods

2.1. Background of the engineering

The failed FRP composite pipe came from a purified sewage pipeline at a Chinese oilfield in Xinjiang. The pipeline was completed
and put into service in September 2006, and it was shut down due to a burst failure in August 2021, after 15 years of operation. The

Fig.1. Macroscopic topography of the damaged pipeline: (a) and (b) failure scene; (c) morphology of the failed pipe section; (d) and (e) local
magnification of the damaged area.

2
D. Liao et al. Engineering Failure Analysis 144 (2023) 106938

pipe is made of multiple layers of continuous E-glass fibres cured by a ±45◦ spiral winding process with a bisphenol A epoxy resin
matrix. The pipe has an external diameter of 94.5 mm, a wall thickness of 9.5 mm and a single section length of 8.8 m. The pipe is
connected by means of a threaded connection, the average depth of burial is 1.5 m and the total length of the pipeline is 1.846 km. The
design working temperature of this high-pressure water injection pipeline is 60 ◦ C and the design pressure is 16 MPa. The experi­
mentally measured heat deflection temperature (HDT) value is 65.3 ◦ C and the glass transition temperature is approximately 120 ℃.
The actual working pressure is 15–15.8 MPa, and the actual water conveying temperature is 20–25 ℃. The conveyed medium is
strongly acidic and highly mineralised purified sewage with a pH value of 1.38 and a total salinity of 24482.8 mg/L. The type of water
sample conveyed is calcium chloride type. The service environment is a desert covered with dunes. Local gas phase reports indicate a
maximum temperature of 42 ◦ C and a minimum temperature of –32 ◦ C in the service environment. According to the preliminary
analysis of the site, the total length of this branch line is 1.846 km, the damage position is 630 m away from the source of the pipeline.
The buried depth of the damage position is 1.8 m, and the macroscopic appearance of the damage position is illustrated in Fig. 1. The
crack originated from a position 2.5 cm above the metal joint, the length of the notch crack is about 10 cm, and the maximum width is
close to 3.5 cm. The crack direction is parallel to the fiber winding direction, and the fiber delamination is obvious at the fracture
position.

2.2. Analytical method

Based on the objectives of the study, a comprehensive analysis of background information and operating conditions that may lead
to pipeline failures was carried out. A failure analysis strategy was developed, with non-destructive testing followed by destructive
testing. Meanwhile, the variation of the performance gradient in different areas of the failed pipe section needed to be considered in
both axial and radial directions. Several sections of the failed pipe were sampled and the total length of the pipe available for study was
36 cm, divided into three zones at equal intervals along the axial direction of the pipe and labelled in turn as the Damaged Area (DA),
Transition Area (TA) and Undamaged Area (UDA). The outer layer (OL), the middle layer (ML) and the inner layer (IL) were also
distinguished in the radial direction of the pipe to facilitate a disaggregated investigation of each part of the pipe, as shown in Fig. 2.
The appearance, physicochemical, thermal and mechanical properties of the failed pipes were analysed sequentially to characterise the
pattern of performance decay during failure and to explore the root causes of pipe failure.
In the experimental analysis, optical microscope and scanning electron microscope (ZEISS mini 300) were used to observe the
macroscopic and microscopic morphology of the fracture of the composite pipe, analyze the structural features and defects, and
calculate the winding angle by analyzing the images of different fiber layers. The burnout test was first carried out according to ASTM-
D2584 to determine the fiber volume fraction and porosity. Then, the extraction method recommended by the ISO 1172 standard is
adopted to determine the content of insoluble components through acetone solvent to evaluate whether it meets the requirements. The
composition and structure of the GFRP structural layer in the composite pipe were analyzed by Fourier transform infrared spectroscopy
(Nicolet 6700) and X-ray photoelectron spectroscopy, respectively. At the same time, thermogravimetric analyzer (TGA550) and
dynamic thermomechanical analysis equipment (DMA Q800) were used to verify the decay law of thermal stability of matrix phase
materials in service environment. Finally, a nano-indenter (Bruker Hysitron TI980) was used to analyze the change rule of the me­
chanical properties of GFRP pipes from the outside to the inside along the section direction, and the nano-hardness and elastic modulus
were compared.

3. Results and discussion

3.1. Microstructural characterizations

First, observe the deterioration degree of the macroscopic features of the receiving tube by visual inspection. It was concluded that
the general condition of the surface was acceptable. Fig. 3 shows that unevenly distributed resin spalling was observed in the

Fig. 2. Schematic diagram of pipeline zoning and sampling.

3
D. Liao et al. Engineering Failure Analysis 144 (2023) 106938

undamaged area of the pipe section, and the spalling occurred in the outermost resin-rich protective layer with a maximum peeling
length of 15 mm and a maximum depth of about 500 μm. Linear scour marks were found on the pipe surface, towards the direction of
GFRP pipe fiber entanglement. At the same time, pinholes with a diameter of about 1 mm were found on the outer surface, but no
defects such as resin dripping and inclusions were found, and the pipe size deformation was also acceptable. The peeling of the surface
resin did not cause damage and breakage of the glass fibers, and the structural layer still showed continuity, which would not affect the
reinforcing effect of the structural layer.
The microstructure of the GFRP tube at different positions was observed with a scanning electron microscope, and the overall
structural features are shown in Fig. 4. There are 11 layers of glass fiber winding structure in the fiberglass pipe. The thickness of the
single-layer wound glass fiber is 496 μm, the diameter of a single glass fiber is 10.9 μm, and the thickness of the resin bonding layer is
24.7 μm. The fibers are alternately wound symmetrically in two directions, and the fiber winding angle is ±45◦ . In the undamaged area
of the GFRP pipe, the fibers are evenly dispersed, the infiltration is complete, the glass fiber and the resin are well bonded, and no
defects such as interlayer cracks and bubbles are found. There are holes formed by the protruding fibers in the local position of the
cross-section direction, and no obvious resin-rich layer is found on the outer surface of the resin, which lacks the outer protective effect
of the resin layer.
In the cross-sectional morphology of the damaged area, it is found that there are obvious delamination cracks in the structural layer,
and there are pores and insufficient resin filling in local locations, resulting in cavities, accompanied by cracks at the fiber and resin
interface. The glass fiber layer showed obvious brittle fracture characteristics along the radial direction at the failure section position,
and the fibers were pulled out and formed holes at local positions. The fibers were still covered with resin after being pulled out,
indicating that the fibers were debonded from the resin, and the failure occurred in the interface area between the fibers and the resin.
Similarly, Perrier et al. [34] demonstrated that the failure of a single hemp/epoxy composite under water-aging conditions is due to a
significant degradation of the fiber–matrix interface properties, such as a decrease in modulus.

3.2. Resin structure and thermal stability

The resin content in GFRP pipes and their degree of curing are key factors that affect their performance in service. Excessive resin
will lead to a decrease in the bearing capacity of the composite pipe and insufficient pipe stiffness; insufficient resin content will lead to
weak bonding strength of reinforcing fibers and reduced pipe strength. The resin exhibits a branched structure when it is not cured.
After curing, it forms a cross-linked network structure to show excellent performance [35,36]. Therefore, the curing degree of the resin
is also an important factor affecting the performance of GFRP pipes. The resin content and curing degree of the sample tubes were
tested by calcination and acetone extraction, respectively. The experimental results (Fig. 5) show that the resin content of the pipeline
in the failed and non-failed areas is consistent, with a slight change, and the overall resin content of the composite pipe is about 18.8 %.
This result is significantly smaller than the allowable resin content range (30 ± 5 %) in the industry standard API Spec 15HR High-
pressure Fiberglass Line Pipe. It shows that the overall resin content of the composite pipe is too low to meet the requirements of
continuous operation.
In the test of resin curing degree, it was found that the content of insoluble components in the resin decreased to a certain extent in
the damaged area of the pipeline compared with the undamaged area, and the outer layer of the damaged area decreased seriously, the

Fig. 3. Macroscopic topography of the upper, middle and bottom of the undamaged area of the GFRP pipeline.

4
D. Liao et al. Engineering Failure Analysis 144 (2023) 106938

Fig. 4. Microstructure and Morphological Characteristics of the damaged area of the GFRP pipeline.

Fig. 5. Test results of resin content (a) and insoluble component content (b) in each part of GFRP pipe.

content of insoluble components in this position is only 53.57 %. It shows that the degree of curing at this position is insufficient and
the mechanical properties are relatively poor, which is a weak point that is prone to failure. The conclusion of the thermal decom­
position experiment also verifies that the thermal stability of the outer layer resin of the pipe decreases, the thermal decomposition
temperature drops from 364 ℃ to 344 ℃, as shown in Fig. 6. The possible reasons for the low insoluble content of the analyzed resin
are defects in the curing process during the preparation of the pipeline; or the decomposition of the curing agent during service,
resulting in a significant decrease in the degree of resin curing at this position.

3.3. Elemental and composition analysis

In order to further find traces of degradation of GFRP tubes and possible traces of foreign contaminants, infrared spectroscopy and
XPS analysis were performed on the damaged area near the crack on the inner surface of the sample and the undamaged area away
from the crack for reference. The infrared spectrum is based on the characteristics of the absorption peak generated by the stretching
vibration of the group to identify the group. In this paper, the change of the content of curing agent in epoxy resin was studied by
infrared spectroscopy. After normalization, the group concentration was determined by peak strength.
The representative infrared absorption spectrum of the obtained epoxy resin is shown in Fig. 7. It is reported that the N–H peak at
wave number 3436 cm− 1 is related to amine curing agents [37,38]. According to the experimental results, it can be found that the

5
D. Liao et al. Engineering Failure Analysis 144 (2023) 106938

Fig. 6. Thermal decomposition test of GFRP pipe.

concentration of N–H bond (3436 cm− 1) in the inner layer (IL) of the pipeline is the largest and the concentration of the outer layer
(OL) is the smallest in the damaged area. Combined with the analysis results of insoluble content, it shows that the content of the curing
agent in the outer layer of the pipeline decreases most obviously and the aging is most serious. In the undamaged area, the change of
peak position at this position is not obvious, indicating that the content of corresponding bonds has not changed significantly, and the
resin loss is not obvious enough. The changes in the intensity of the absorption peaks of the C– – O bond at the wavenumber of 1726
cm− 1 and the C–N bond at 1236 cm− 1 are opposite, indicating that the concentration of functional groups containing two bonds in the
outer layer of the failure position increases. The concentration of C–
– O bond and C–N bond increases from the inner layer to the outer
layer. In essence, the epoxy resin in the matrix is oxidized to aldehyde-based compounds during the aging process. Under the action of
oxygen in the service environment, GFRP The outer surface of the tube is more susceptible to oxidative degradation than the inner wall.
In the process of oxidative degradation, part of the group connected to the hydroxyl group in the epoxy resin is oxidized to an aldehyde
group (–CHO), and the methylene group connected to N in the curing agent is oxidized to C– – O [39], forming C– – O in the amide, while
the molecular backbone is broken and low molecular weight polymers are produced. The infrared test results in the undamaged area
(Fig. 7.b) are also compared at the inner wall, the middle layer and the outer surface of the pipeline. The results show that the intensity
of the absorption peaks of corresponding chemical bonds at the relevant positions varies weakly, indicating that the concentration of
the bond content changed little.
The pipeline samples were analyzed by XPS spectrum, and the peak spectrum was calibrated and fitted according to the relevant
research literature [39–41], and the results are shown in Figs. 8 and 9. The C elements all come from the resin matrix, and the Si
elements all come from glass fibers. Both epoxy resin and glass fibers provide the source of O elements. The results show that the
oxygen content increases in the outer layer of the GFRP tube, and the degree of oxidation intensifies. The maximum oxygen con­
centration in the outer layer is 27.12 %. The decrease of the C content of the outer layer also proves the decrease of the resin content of
the outer layer. The C1s peaks of the samples at each position are fitted to obtain the concentration changes corresponding to each
chemical bond, as shown in Table 1. The thermal degradation process of the cured epoxy resin will produce peracid and perester
functional groups [37]. The formation of O-C– – O bonds also leads to an increase in the concentration of C– – O bonds [39], and the

4000 3500 3000 2500 2000 1500 1000 500

Fig. 7. Infrared absorption spectrum results of different positions of GFRP tube.

6
D. Liao et al. Engineering Failure Analysis 144 (2023) 106938

Fig. 8. Wide Scan XPS Spectra (a) and Elemental Content (b) each group of samples.

Fig. 9. XPS fine spectrum of C 1 s of each group of samples.

Table 1
Functional group compositions of each group of samples.
Types Functional group content (%)
C–H C–C C–O C–
–O O-C–
–O

DA-IL 20.13 47.99 14.28 9.97 7.62


DA-ML 14.59 46.48 19.14 12.68 7.11
DA-OL 10.96 48.03 20.42 13.25 7.34
UDA-IL 21.6 46.65 14.91 9.5 7.34
UDA-ML 6.61 51.29 23.30 11.48 7.33
UDA-OL 10.67 50.93 19.20 11.96 7.23

7
D. Liao et al. Engineering Failure Analysis 144 (2023) 106938

results show that the oxidative degradation process is more pronounced in the outer layer of the damaged zone.

3.4. Dynamic mechanical analysis

Nanoindentation and DMA testing are widely used methods to measure the micromechanical properties of thin films and micro­
scale materials [42], and similar studies have shown that these two methods can be well applied to analyze the interfacial properties of
fiber-reinforced composites [34,43]. The relationship between the dynamic modulus and mechanical loss of FRP pipe under vibration
load with temperature was analyzed by dynamic thermomechanical performance analysis method, and the stability of GFRP pipe
under high temperature dynamic load was judged. The storage and loss moduli of undamaged (UDA) and damaged (DA) specimens
versus temperature at 1 Hz were measured by DMA. The experiment also evaluated the change of glass transition temperature (Tg) in
each region. The experimental results in Fig. 10 show that the undamaged region in the initial stage of loading has a relatively large
storage modulus value of about 1.83 × 1010 Pa at the inner layer position, and the modulus decreases by 17.5 % at the outer layer
position. In the damaged area, the modulus value in the direction from the inner layer to the outer layer decreases from 1.69 × 1010 Pa
to 1.38 × 1010 Pa, the modulus retention rate in the failure area is 81 %, which corresponds to a decrease of 10 % compared with the
UDA specimen. The performance attenuation of the pipeline along the radial direction is greater than the axial performance atten­
uation from the non-failure position to the failure position, and the outer surface of the failure location is the weak point of the
mechanical properties. The glass transition temperature exhibits the same regular decay. This study showed a significant reduction in
the glass transition temperature outside the damaged area, and the glass transition temperature dropped to 110.86 ℃, which is lower
than the range allowed for normal use in the standard API Spec 15HR. This is attributed to the sharp drop in the cured resin content of
the outer layer of the damaged zone.
Fig. 11 shows the three-dimensional morphology of the indentation and the stress and displacement curves during loading and
unloading of the DA/UDA samples during the nanoindentation experiment, respectively. The experiment was carried out in high load
mode, the maximum load was 0.3 N, and 5 points were uniformly taken from the outside to the inside of the pipe section. The
maximum indentation depth (hmax), Young’s modulus (Er) and nano-hardness (H) were used as evaluation indicators and summarized
in Table 2. We can find that the hmax of the damaged area is 746 μm, and the hmax in the undamaged area is only 698.46 μm. Compared
with the undamaged area, the elastic modulus declines more obviously, from 5.56 GPa to 4.62 GPa, the attenuation ratio is 16.9 %, the

Fig. 10. Dynamic thermomechanical analysis results of GFRP pipes.

8
D. Liao et al. Engineering Failure Analysis 144 (2023) 106938

Fig. 11. Nanoindentation morphologies and loading–unloading curves: (a) (b)- damaged area; (c) (d)- undamaged area.

average nano-hardness also decreases slightly. Similarly, some researchers [34] analyzed the microscopic properties of hydrothermally
treated hemp/epoxy composites through nanoindentation and DMA testing, and found that water causes a decrease in nano­
indentation performance, with the most pronounced reduction in modulus occurring in the interfacial region. Passilly et al. [43] using
the same analytical method for thermally oxidatively aged MTR6 epoxy resin, it was found that the Young’s modulus of the core of the
sample also decreased significantly with increasing temperature.

3.5. The failure mechanism analysis

There are a certain number of manufacturing defects inside the pipe, such as internal pores, insufficient resin filling, and debonding
of the resin fiber interface (Fig. 12.a), which can form weak areas inside the pipe that are easily damaged. During the long-term service
of the pipeline, the media and oxides in the environment induce the oxidation and degradation of the resin matrix, which further leads
to the decrease of the resin content of the GFRP pipe and the decrease of the resin curing degree. This process will lead to the decrease
of the bonding strength between the resin matrix and the fiber and the decline of the two-phase interface properties, and the overall
strength of the resin will decrease, which will lead to the decrease of fracture toughness and brittle fracture. As the pipe continues to
age over time in service, mechanical property such as hardness and modulus also demonstrate a decline in mechanical properties at this
location. It has been reported that multi-stage load can accelerate the creep behavior of GFRP tubes under internal pressure and
transverse load, and shorten the failure time rapidly [44–46]. The hoop stress generated by the internal pressure of the medium, the
thermal stress caused by the change of ambient temperature and the residual stress caused by the assembly of this GFRP pipeline are
coupled with each other and accelerate the explosion failure of the pipeline from inside to outside.
The specific performance of the failure process is that in the early stage of failure, a crack source is generated at the weak part of the

9
D. Liao et al. Engineering Failure Analysis 144 (2023) 106938

Table 2
Summary of nanoindentation experimental data.
Types Pmax (μN) hmax (nm) Er (GPa) H (GPa)

DA 2996.74 746.00 4.62 0.21


UDA 2996.62 694.84 5.56 0.24

Fig. 12. Schematic diagram of GFRP pipe body structure and failure defects: (a) before failure; (b) after failure.

outer surface of the pipeline (defect position) due to the impact of external load or the aging of the pipeline, and the phenomenon of
whitening is observed. The macroscopic appearance is the formation of flaky whitish areas parallel to the fiber direction, which is a
sign of debonding and delamination. This is essentially due to the multi-layer bonded structure of the GFRP pipe, which generates
radial expansion under the action of hoop stress, and the interlayer structure is subjected to shear stress, and the direction of force is
parallel to the direction of fiber winding. The hoop stress generated by the internal pressure of the pipe accelerates the expansion of the
crack source in the pipe to form a large-scale crack (as shown in Fig. 12.b), which leads to the cracking of the matrix and the separation
of the fiber matrix, causing delamination and debonding, and the growth rate increases rapidly and then slows down [47]. Pinholes
and leaks occur in the second stage of failure, with pinholes appearing due to delamination of the whitened area. Under the pressure of
the fluid, the micropores extend outward from the inner surface, and the leakage begins when reaching the outer surface to form
through cracks. The leakage starts in the pinhole region in the form of tiny droplets, and after many cycles, a strong leakage occurs. In
the final stage of failure, the fiber adhesion decreases after the pipeline resin is debonded, and large-scale fiber brittle fracture occurs
after loading, accompanied by the pulling and pulling out of fibers to form holes.

4. Conclusions

(1) The macro-morphological analysis of the GFRP pipe shows that there are scouring marks on the surface of the pipe along the
fiber winding direction, and the resin on the outer surface is partially peeled off, but no fiber breakage and damage caused by
the peeling of the surface resin are found. Micro-morphological analysis found that there were defects such as pores and
insufficient resin filling inside, the interface between resin and glass fiber in the damaged area was debonded, fiber delami­
nation and tearing were obvious, and the fracture surface showed brittle fracture characteristics.
(2) The overall combined resin content of the pipe was 18.8 %, with the lowest percentage of insoluble resin in the outer layer of the
damaged area being 53.57 %, indicating a low resin content and insufficient cure for continued use.
(3) Infrared spectroscopy experiments show that the N–H bond of the amine curing agent in the damaged area of the pipeline has a
reduced functional group concentration in the outer layer. It is speculated that the curing agent is decomposed in the local
position of the pipeline, and the content decreases from the inside to the outside. XPS spectra verified that the concentration of
C–– O bond functional groups increased from the inner layer to the outer layer, indicating that the hydroxyl groups of epoxy
resins were oxidized to aldehyde-based compounds during the aging process, and their content increased with the aggravation
of aging.
(4) The analysis of the dynamic mechanical properties of the composite pipeline also verifies that the outer position of the pipeline
damage zone is the weak zone of the mechanical properties. The elastic modulus of the damaged area decreased significantly,
from 5.56 GPa of UDA to 4.62 GPa of DA. The inner layer position in the undamaged area has the largest storage modulus of
1.83 × 1010Pa, the modulus retention rate in the failure area is 81 %, and the attenuation range of the modulus from the non-
failure position to the failure position is about 10 %.

10
D. Liao et al. Engineering Failure Analysis 144 (2023) 106938

5. Measures to failure prevention

According to the analysis of this accident, the following preventive measures should be taken for pipeline failure under this working
environment.

(1) The acceptance of pipes shall be carried out in strict accordance with the quality acceptance specification of glass fiber pipes.
The key performance indicators of the pipeline, such as resin content of the structural layer, glass transition temperature, resin
curing degree and hydrostatic pressure of the pipeline, shall be checked. It is also suggested that the manufacturer should
conduct inspection during the pipeline preparation to control the product quality.
(2) It is suggested that the non-metallic pipe joints in service in this area should be inspected and evaluated in time. The damage and
hidden danger of the pipeline shall be checked in time, and the pipeline with significantly reduced performance shall be
repaired, replaced and repaired in time to avoid leakage and failure accidents.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Data availability

The authors are unable or have chosen not to specify which data has been used.

Acknowledgements

The work was supported by the Tarim Oilfield of China National Petroleum Corporation and the China Petroleum Planning
Institute. The authors also thank Sichuan University for providing sophisticated instrumentation for this research investigation.

References

[1] R. Hsissou, R. Seghiri, Z. Benzekri, M. Hilali, M. Rafik, A. Elharfi, Polymer composite materials: a comprehensive review, Compos. Struct. 262 (2021), https://
doi.org/10.1016/j.compstruct.2021.113640.
[2] D. Rajak, D. Pagar, P. Menezes, E. Linul, Fiber-reinforced polymer composites: manufacturing, properties, and applications, Polymers (Basel). 11 (10) (2019),
https://doi.org/10.3390/polym11101667.
[3] R. Rafiee, On the mechanical performance of glass-fibre-reinforced thermosetting-resin pipes: a review, Compos. Struct. 143 (2016) 151–164, https://doi.org/
10.1016/j.compstruct.2016.02.037.
[4] F.G. Alabtah, E. Mahdi, F.F. Eliyan, The use of fiber reinforced polymeric composites in pipelines: a review, Compos. Struct. 276 (2021), 114595, https://doi.
org/10.1016/j.compstruct.2021.114595.
[5] T. Jiadong, L. Mingchang, Z. Wanrong, Application and development of fiber reinforced plastic pipe in oil & gas field, Petrol. Tubul. Goods Instrum. 3 (2017)
1–4, https://doi.org/10.19459/j.cnki.61-1500/te.2017.05.001.
[6] Q.i. Guoquan, Y. Hongxia, Q.i. Dongtao, W. Bin, L.i. Houbu, Analysis of cracks in polyvinylidene fluoride lined reinforced thermoplastic pipe used in acidic gas
fields, Eng. Fail. Anal. 99 (2019) 26–33, https://doi.org/10.1016/j.engfailanal.2019.01.079.
[7] J.M.L. Reis, F.D.F. Martins, H.S. Da Costa Mattos, Influence of ageing in the failure pressure of a GFRP pipe used in oil industry, Eng. Fail. Anal. 71 (2017)
120–130, https://doi.org/10.1016/j.engfailanal.2016.06.013.
[8] Y.X. Zhou, B. Chen, J.Y. Li, Application and evaluation of non-metallic pipeline in lamadian oilfield, Adv. Mater. Res. 694 (2013) 521–525, https://doi.org/
10.4028/www.scientific.net/AMR.694-697.521.
[9] A.F. Hamed, Y.A. Khalid, S.M. Sapuan, M.M. Hamdan, T.S. Younis, B.B. Sahari, Effects of winding angles on the strength of filament wound composite tubes
subjected to different loading modes, Polym. Polym. Compos. 15 (3) (2007) 199–206, https://doi.org/10.1177/096739110701500304.
[10] H. Arikan, Failure analysis of (±55◦ )3 filament wound composite pipes with an inclined surface crack under static internal pressure, Compos. Struct. 92 (2010)
182–187, https://doi.org/10.1016/j.compstruct.2009.07.027.
[11] L. Gemi, N. Tarakçioğlu, A. Akdemir, Ö.S. Şahin, Progressive fatigue failure behavior of glass/epoxy (±75)2 filament-wound pipes under pure internal pressure,
Mater. Design. 30 (10) (2009) 4293–4298, https://doi.org/10.1016/j.matdes.2009.04.025.
[12] P. Krishnan, M.S. Abdul Majid, M. Afendi, A.G. Gibson, H.F.A. Marzuki, Effects of winding angle on the behaviour of glass/epoxy pipes under multiaxial cyclic
loading, Mater. Design. 88 (2015) 196–206, https://doi.org/10.1016/j.matdes.2015.08.153.
[13] A. Samanci, N. Tarakçioğlu, A. Akdemir, Fatigue failure analysis of surface-cracked (±45◦ )3 filament-wound GRP pipes under internal pressure, J. Compos.
Mater. 46 (2012) 1041–1050, https://doi.org/10.1177/0021998311414945.
[14] A. Samanci, A. Avci, N. Tarakcioglu, Ö.S. Şahin, Fatigue crack growth of filament wound GRP pipes with a surface crack under cyclic internal pressure, J. Mater.
Sci. 43 (16) (2008) 5569–5573, https://doi.org/10.1007/s10853-008-2820-x.
[15] N. Tarakcioglu, L. Gemi, H. Yapici, Fatigue failure behavior of glass/epoxy±55◦ filament wound pipes under internal pressure, Compos. Sci. Technol. 65 (2005)
703–708, https://doi.org/10.1016/j.compscitech.2004.10.002.
[16] N. Tarakcioglu, A. Samanci, H. Arikan, A. Akdemir, The fatigue behavior of (±55◦ )3 filament wound GRP pipes with a surface crack under internal pressure,
Compos. Struct. 80 (2) (2007) 207–211, https://doi.org/10.1016/j.compstruct.2006.05.015.
[17] L. Gemi, M. Kara, A. Avci, Low velocity impact response of prestressed functionally graded hybrid pipes, Compos. B Eng. 106 (2016) 154–163, https://doi.org/
10.1016/j.compositesb.2016.09.025.
[18] L. Gemi, M.A. Köroğlu, A. Ashour, Experimental study on compressive behavior and failure analysis of composite concrete confined by glass/epoxy ±55◦
filament wound pipes, Compos. Struct. 187 (2018) 157–168, https://doi.org/10.1016/j.compstruct.2017.12.049.
[19] L. Gemi, Investigation of the effect of stacking sequence on low velocity impact response and damage formation in hybrid composite pipes under internal
pressure. A comparative study, Compos. Part B: Eng. 153 (2018) 217–232, https://doi.org/10.1016/j.compositesb.2018.07.056.
[20] R. Rafiee, Experimental and theoretical investigations on the failure of filament wound GRP pipes, Compos. B Eng. 45 (2013) 257–267, https://doi.org/
10.1016/j.compositesb.2012.04.009.
[21] R. Rafiee, Stochastic fatigue analysis of glass fiber reinforced polymer pipes, Compos. Struct. 167 (2017) 96–102, https://doi.org/10.1016/j.
compstruct.2017.01.068.

11
D. Liao et al. Engineering Failure Analysis 144 (2023) 106938

[22] R. Rafiee, F. Elasmi, Theoretical modeling of fatigue phenomenon in composite pipes, Compos. Struct. 161 (2017) 256–263, https://doi.org/10.1016/j.
compstruct.2016.11.054.
[23] R. Rafiee, M.R. Habibagahi, On the stiffness prediction of GFRP pipes subjected to transverse loading, Ksce J. Civ. Eng. 22 (2018) 4564–4572, https://doi.org/
10.1007/s12205-018-2003-5.
[24] R. Rafiee, B. Mazhari, Evaluating long-term performance of Glass Fiber Reinforced Plastic pipes subjected to internal pressure, Constr. Build. Mater. 122 (2016)
694–701, https://doi.org/10.1016/j.conbuildmat.2016.06.103.
[25] A. Manalo, G. Maranan, S. Sharma, W. Karunasena, Y.u. Bai, Temperature-sensitive mechanical properties of GFRP composites in longitudinal and transverse
directions: a comparative study, Compos. Struct. 173 (2017) 255–267, https://doi.org/10.1016/j.compstruct.2017.04.040.
[26] R.A. Hawileh, A. Abu-Obeidah, J.A. Abdalla, A. Al-Tamimi, Temperature effect on the mechanical properties of carbon, glass and carbon–glass FRP laminates,
Constr. Build. Mater. 75 (2015) 342–348, https://doi.org/10.1016/j.conbuildmat.2014.11.020.
[27] K. Imielińska, L. Guillaumat, The effect of water immersion ageing on low-velocity impact behaviour of woven aramid–glass fibre/epoxy composites, Compos.
Sci. Technol. 64 (2004) 2271–2278, https://doi.org/10.1016/j.compscitech.2004.03.002.
[28] M.E. Deniz, R. Karakuzu, Seawater effect on impact behavior of glass–epoxy composite pipes, Compos. B Eng. 43 (2012) 1130–1138, https://doi.org/10.1016/j.
compositesb.2011.11.006.
[29] A. Stocchi, A. Pellicano, J.P. Rossi, C. Bernal, P. Montemartini, Physical and water aging of glass fiber-reinforced plastic pipes, Compos. Interface. 13 (8–9)
(2006) 685–697, https://doi.org/10.1163/156855406779366831.
[30] X. Jiang, H. Kolstein, F.S.K. Bijlaard, Moisture diffusion and hygrothermal aging in pultruded fibre reinforced polymer composites of bridge decks, Mater.
Design. 37 (2012) 304–312, https://doi.org/10.1016/j.matdes.2012.01.017.
[31] M. Assarar, D. Scida, A. El Mahi, C. Poilâne, R. Ayad, Influence of water ageing on mechanical properties and damage events of two reinforced composite
materials: Flax-fibres and glass-fibres, Mater. Des. 32 (2) (2011) 788–795, https://doi.org/10.1016/j.matdes.2010.07.024.
[32] F. Ellyin, R. Maser, Environmental effects on the mechanical properties of glass-fiber epoxy composite tubular specimens, Compos. Sci. Technol. 64 (2004)
1863–1874, https://doi.org/10.1016/j.compscitech.2004.01.017.
[33] H. Li, M. Yan, D. Qi, N. Ding, X. Cai, S. Zhang, Q.i. Li, X. Zhang, J. Deng, Failure analysis of steel wire reinforced thermoplastics composite pipe, Eng. Fail. Anal.
20 (2012) 88–96, https://doi.org/10.1016/j.engfailanal.2011.11.002.
[34] A. Perrier, E. Le Bourhis, F. Touchard, L. Chocinski-Arnault, Effect of water ageing on nanoindentation response of single hemp yarn/epoxy composites, Compos.
A Appl. Sci. Manuf. 84 (2016) 216–223, https://doi.org/10.1016/j.compositesa.2016.01.022.
[35] F. Xie, L.N Huang, X.B Wang, B. Li, J.H Ma, Research on depolymerisation and regeneration of epoxy resin cured with anhydrides, Polym. Polym. Compos. 19
(4–5) (2011) 265–270, https://doi.org/10.1177/0967391111019004-503.
[36] P. Yang, Q. Zhou, X.-X. Yuan, J.M.N. van Kasteren, Y.-Z. Wang, Highly efficient solvolysis of epoxy resin using poly (ethylene glycol)/NaOH systems, Polym
Degrad Stabil. 97 (7) (2012) 1101–1106.
[37] V.Y. Ignatenko, S.O. Ilyin, A.V. Kostyuk, G.N. Bondarenko, S.V. Antonov, Acceleration of epoxy resin curing by using a combination of aliphatic and aromatic
amines, Polym. Bull. 77 (3) (2020) 1519–1540, https://doi.org/10.1007/s00289-019-02815-x.
[38] S. Ahmad, A.P. Gupta, E. Sharmin, M. Alam, S.K. Pandey, Synthesis, characterization and development of high performance siloxane-modified epoxy paints,
Prog. Org. Coat. 54 (3) (2005) 248–255, https://doi.org/10.1016/j.porgcoat.2005.06.013.
[39] S. Hong, The thermal-oxidative degradation of an epoxy adhesive on metal substrates: XPS and RAIR analyses, Polym. Degrad. Stabil. 48 (1995) 211–218,
https://doi.org/10.1016/0141-3910(95)00042-K.
[40] T. Duguet, C. Bessaguet, M. Aufray, J. Esvan, C. Charvillat, C. Vahlas, C. Lacaze-Dufaure, Toward a computational and experimental model of a poly-epoxy
surface, Appl. Surf. Sci. 324 (2015) 605–611, https://doi.org/10.1016/j.apsusc.2014.10.096.
[41] N. Li, X. Y. LI, L. Liu, Acta Materiae Compositae SinicaSurface modification of carbon fiber(CF) deposited graphene oxide(GO) by electrophorestic deposition
and interfacial properties of GO-CF/epoxy composites, Acta Mater. Compos. Sinica, 37 (2020) 1571–1580, doi: 10.13801/j.cnki.fhclxb.20191120.001.
[42] J. Lee, M. Zhang, D. Bhattacharyya, Micromechanical behavior of self-healing epoxy and hardener-loaded microcapsules by nanoindentation, Mater. Lett. 76
(2012) 62–65, https://doi.org/10.1016/j.matlet.2012.02.052.
[43] B. Passilly, R. Delannoy, Characterization of the ageing of an epoxy resin using high temperature nanoindentation, Matér. Tech. 107 (2019) 206, https://doi.
org/10.1051/mattech/2019004.
[44] R. Rafiee, A. Ghorbanhosseini, Analyzing the long-term creep behavior of composite pipes: Developing an alternative scenario of short-term multi-stage loading
test, Compos. Struct. 254 (2020), https://doi.org/10.1016/j.compstruct.2020.112868.
[45] R. Rafiee, A. Ghorbanhosseini, Developing a micro-macromechanical approach for evaluating long-term creep in composite cylinders, Thin-Wall. Struct. 151
(2020), https://doi.org/10.1016/j.tws.2020.106714.
[46] R. Rafiee, A. Ghorbanhosseini, Experimental and theoretical investigations of creep on a composite pipe under compressive transverse loading, Fibers Polym. 22
(2021) 222–230, https://doi.org/10.1007/s12221-021-0265-x.
[47] C. Kaynak, O. Mat, Uniaxial fatigue behavior of filament-wound glass-fiber/epoxy composite tubes, Compos. Sci. Technol. 61 (2001) 1833–1840, https://doi.
org/10.1016/S0266-3538(01)00084-7.

12

You might also like