You are on page 1of 24

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/275057712

Pipeline failures in corrosive environments – A


conceptual analysis of trends and effects

Article in Engineering Failure Analysis · July 2015


DOI: 10.1016/j.engfailanal.2015.03.004

CITATIONS READS

14 899

3 authors:

Chinedu Ossai B. Boswell


Curtin University Curtin University
19 PUBLICATIONS 50 CITATIONS 40 PUBLICATIONS 103 CITATIONS

SEE PROFILE SEE PROFILE

Ian J. Davies
Curtin University
189 PUBLICATIONS 1,122 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Synthesis bio-based epoxy/clay nanocomposite and characterisation View project

All content following this page was uploaded by Ian J. Davies on 05 May 2015.

The user has requested enhancement of the downloaded file.


Engineering Failure Analysis 53 (2015) 36–58

Contents lists available at ScienceDirect

Engineering Failure Analysis


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

Review

Pipeline failures in corrosive environments – A conceptual


analysis of trends and effects
Chinedu I. Ossai ⇑, Brian Boswell, Ian J. Davies
Department of Mechanical Engineering, Curtin University, GPO Box U1987, Perth, WA 6845, Australia

a r t i c l e i n f o a b s t r a c t

Article history: Pipeline corrosion is a major challenge facing many oil and gas industries today because of
Received 30 November 2014 the enormous downtime associated with corrosion related failures. Fatigue stress initiation
Received in revised form 10 February 2015 in pipelines has been attributed to corrosion defects whose growth is enhanced by cyclic
Accepted 17 March 2015
loading caused by the operating pressure of the transported fluids. This work reviews
Available online 2 April 2015
the concept of oil and gas transmission pipeline failures in corrosive environment by high-
lighting the corrosion mechanisms, dominant stress corrosion cracking trends, hydrogen
Keywords:
induced cracking and predominant models for burst pressure estimation. Fatigue stress
Corrosive environment
Stress corrosion cracking
failure trends of corroding pipelines were also explained whilst describing some pipeline
Pipeline integrity manufacturing processes that increases the susceptibility to fatigue stress failure.
Fatigue stress failure Optimization framework for pipeline integrity assurance against corrosion fatigue failures
Failure pressure was also shown to incorporate different steps that includes – strategic policy initiation,
policy implementation, information analysis and reviews and implementation actions.
Ó 2015 Elsevier Ltd. All rights reserved.

1. Introduction

Pipeline failure can be described as the bane of ruined cities due to the destructions that have followed their ruptures.
According to the Pipeline and Hazardous Materials Safety Administration (PHMSA) [1] of the US department of transport,
360 fatalities, 1368 injuries and 894 incidences relating to oil, gas and hazardous fluid pipeline failures occurred in US alone
between the years 1995 and 2014. Fig. 1 depicts some notable incidence of pipeline explosion and the ruin they caused to the
cities. The figure shows Kaohsiung city in South Taiwan on 1/8/2014 after a gas pipeline explosion which killed 25 people
and injured more than 259 others [2] and Ghislenghien city, Belgium on 30/7/2004 whose gas pipeline explosion caused
24 fatalities and injured more than 120 people [3].
Fatigue can be described as a structural damage of materials due to stress [4]. Whereas human and natural phenomena
can cause fatigue failure in pipelines, research has shown that corrosion plays a pivotal role in fatigue stress initiation [5–7]
whilst cyclic loading has contributed to the growth of this stress and eventual failure of the pipelines [8]. Assessment of cor-
rosion induced fatigue failures in pipelines can be done by using stress-life, strain-life, linear elastic fracture mechanics,
crack propagation and statistical based techniques [4,9–12]. Fatigue stress analysis is vital for ensuring that the level of
safety required for the operation of oil and gas transmission pipelines are maintained [13].
Fatigue stress behaviour of pipelines can be significantly affected by the operating environment, geometry and size of cor-
roded sections, pipe material properties, exposure time and time dependent corrosion propagation [14]. Hence, operators of

⇑ Corresponding author.
E-mail address: ossaic@gmail.com (C.I. Ossai).

http://dx.doi.org/10.1016/j.engfailanal.2015.03.004
1350-6307/Ó 2015 Elsevier Ltd. All rights reserved.
C.I. Ossai et al. / Engineering Failure Analysis 53 (2015) 36–58 37

Nomenclature

Acronyms
API American Petroleum Institute
ASME American Society of Mechanical Engineers
ASSY Average Shear Stress Yield
BSI British Standards Institute
CAPEX capital expenditure
CAPP Canadian Association of Petroleum Producers
CRAs Corrosion Resistance Alloys
ERW Electric Resistance Welding
FAD Failure Assessment Diagram
FEA Finite Element Analysis
HAZs Heat Affected Zones
IEAW Indirect Electric Arc Welding
MAOP Maximum Allowable Operating Pressure
MIC Microbiologically Induced Corrosion
OPEX operating expenditure
SAW Submerged Arc Welding
SINTAP structural integrity assessment procedure
SSC Stress Sulphide Cracking

Notations
C0, C1, C2, C3, C4 constants
d defect depth (m)
D nominal pipeline external diameter (m)
Di pipeline internal diameter (m)
D0 original average external diameter of pipeline (m)
E modulus of elasticity (MPa)
F inflation factor
G yield theory dependent constant
Kr normalized stress intensity factor
L longitudinal corrosion defect length (m)
Lr non-dimensional loading
Lmax
r maximum non-dimensional loading
M bulging factor
n strain hardening exponent
N second correction factor
Pf burst pressure (MPa)
Q correction factor (m1/2)
R radius of pipeline (m)
t pipe-wall thickness (m)
t0 original pipe-wall thickness (m)
y number of years
rY yield stress (MPa)
rU ultimate tensile strength (MPa)
ruts engineering ultimate tensile strength (MPa)
eref reference strain
rref reference stress (MPa)
l first correction factor

oil and gas industries and numerous experts are concertedly working to ensure that the right operating environments which
can reduce the risk of corrosion are created during exploration and production of oil and gas.
Corrosion problems have been estimated to cost approximately $2.2 trillion to the world economy (3.0% of the Gross
Domestic Product) with oil, gas and petrochemical industries accounting for a total of $1 trillion or 45% of this amount
[15,16]. With records showing that over 70% of the current oil and gas fields being developed around the world are highly
corrosive [17], this cost is bound to increase since it will cost more to manage facilities in corrosive environments.
Based on Bhaskaran et al. [18] estimated global direct cost of corrosion for the year 2004 and inflation factors for years
2005 to June 2014 [19], the global direct cost of corrosion for different countries were calculated and the summary shown in
38 C.I. Ossai et al. / Engineering Failure Analysis 53 (2015) 36–58

Fig. 1. Some catastrophic pipeline failures.

Table 1
Summary of the direct cost of corrosion.

Country Average inflation rate (2005– Inflation factor Direct cost of corrosion Direct cost of corrosion Percentage contribution
June 2014) [F = (1 + i)y] 2004 ($B) 2014 ($B) 2014 (%)
USA 2.33 1.26 303.76 382.44 26.11
Japan 0.14 1.01 59.02 59.84 4.08
Germany 1.62 1.17 49.26 57.86 3.95
UK 2.74 1.31 8.51 11.15 0.76
Australia 2.79 1.32 7.32 9.64 0.66
Canada 1.79 1.19 3.38 4.04 0.28
South 5.56 1.72 3.18 5.46 0.37
Africa
India 8.37 2.23 3.78 8.45 0.58
China 6.84 2.36 61.00a 144.23 9.85
Others 781.77 53.37
World 4 1.48 990 1464.88

En-Hou Han (as cited by Lieser and Xu [20]).


i: inflation rate; y: number of years.
a
Cost of corrosion in 2001.

Table 1. In order to estimate these direct costs of corrosion, It was assumed that they will vary with economic activities based
on inflation factor [F = (1 + i)y]. The average inflation rate (i) for the durations under review as obtained from Ref. [19] and the
number of years (y) between these periods were used to extrapolate the expected direct corrosion cost for the countries in
2014. This table shows that the world faces an indirect cost of over $1.4 trillion (approximately 2% of the world Gross
Domestic Product) annually in corrosion related problems. United States accounts for over 26% of the world direct cost of
corrosion whilst China accounts for about 10%. However, with the increasing infrastructural development and expansion
of Chinese economy, the total direct and indirect cost of corrosion in China is sure to increase at a higher rate over the coming
years [20].
C.I. Ossai et al. / Engineering Failure Analysis 53 (2015) 36–58 39

2. Pipeline corrosion mechanisms

Corrosion is one of the most predominant causes of pipeline failures in oil and gas production and accounts for between
one quarter to two third of the total downtime in the industry [21–23]. Every year, enormous amount of money is spent on
different forms of corrosion control measures in order to maintain the integrity of pipelines. Unfortunately, the difficulties
associated with getting appropriate designs, predictions, monitoring and mitigation strategies highlight the billions of dol-
lars lost in the economies around the world to corrosion related impairment of pipelines and other structures. Research has
shown that over 80% percent of failed pipelines are monitored in one form or another whilst between 20% and 65% of the
amount spent on corrosion problems could be saved if there was a good knowledge of corrosion inhibitors, protection
and control techniques [16,24]. Corrosion of pipelines can be attributed to numerous causes which are related to the physical
and chemical factors and includes both environmental conditions and the characteristics of the materials. Fig. 2 summarizes
the causes of pipeline corrosion.
The existence of H2S, CO2, organic and inorganic acids in a pipeline operating environment may cause corrosion defects of
different sorts [25–27]. To this end, Li et al. [28] investigated the effect of H2S concentration on X60 grade pipe corrosion in
an environment of coexistence of H2S/CO2. These authors concluded that H2S concentration below 0.05 mmol/l, CO2 corro-
sion rate is decreased and at H2S concentration of between 0.05 and 2 mmol/l corrosion rate of CO2 corrosion shows no sig-
nificant changes but increased at concentration above 2 mmol/l. This coexistence of H2S and CO2 can be attributed to the
formation of sulphide films on steel surfaces with crystal structure forms that includes – mackinawite, pyrrhotite, troilite,
pyrite and greigite [27,29,30]. Water–oil interface and water splashes or slugs contribute to the spread of corrosion on pipeli-
nes [31].
When fluid in multiphase flow transits from stratified flow to a stable water-in-oil flow regime, dispersion takes places.
This only happens if the oil phase turbulence is intense enough to cause the water phase break up into droplets [32].
This broken droplets migrate to the pipeline walls when it gets beyond a certain critical diameter causing or accelerating
corrosion [32]. Khaksarfard et al. [31] investigated the impact of inclination angles on the internal corrosion of multiphase
flow pipelines using computational fluid dynamics (CFD). They found out that, when flow is upwards on the pipeline, the
water-wet surface is more than when the flow is downwards. This implies that more surfaces are prone to corrosion with
an upward flow than downward flow hence the use of flexible pipes made of Corrosion Resistance Alloys (CRAs) for some
critical oil and gas operations such as well tubing, gas lift lines and subsea risers due to the high corrosion resistance ability
[33]. When a pipelines has no defect and is uniformly corroded, it may be relatively straightforward to predict the expected
corrosion wastage using linear and non-linear models [34–37]. However, the existence of defects such as dents, cracks, out-
of-roundness, gouges and buckles (see Fig. 3) around the corroded areas of the pipelines make the estimation more complex.
This is because the effects of these defects will contribute to more stress load on the pipeline and ultimately affect the
burst pressures at such corroded sections. One of such defects that is attracting attention to researchers is crack-in-corrosion
defect which can be described as a hybrid defect that has cracks which are coincident in a corroded area and has claimed
more than 10% of the pipe wall thickness [38]. To show the effects of these hybrid defects, Bedairi et al. [38] evaluated

Fig. 2. Causes of pipeline corrosion.


40 C.I. Ossai et al. / Engineering Failure Analysis 53 (2015) 36–58

Fig. 3. Some defects that enhance fatigue stress failure of corroded pipelines.

cracking, corrosion and crack-in-corrosion defects of pipelines using Finite Element Analysis (FEA) based on elastic–plastic
fracture mechanics. These authors experimentally validated this work which was on a 508 mm diameter and 5.7 mm thick
pipe. Their conclusion after analysing 200 mm long defect with different percentages of the corroded depths and cracked
wall thicknesses is that the corrosion defect was most conservatively predicted with 3.2% variation existing between the
experimental result and FEA data. This was followed by crack only defect and crack-in-corrosion defects which had 12.4%
and 17.4% variations between the FEA and experimental results respectively. This finding shows that the accuracy of predic-
tion of crack-in-corrosion defect may need to be improved in order to effectively manage the integrity of ageing pipelines.
It is also evident from different researches that flow pattern influences pipeline corrosion [39,40] hence Biomorgi et al.
[41] practically measured the effects of variation of operating conditions on the corrosion of oil and gas pipelines. After
4 months of observation, these researchers concluded that in addition to the influence of flow pattern, sand, iron carbonate
and sulphide scales predominately caused under deposits in the studied 102 mm and 154 mm diameter pipelines. They fur-
ther stated that slug patterned flow contributed to more corrosion than bubbles on the pipelines whereas the pitting corro-
sion rates was also influenced by the internal diameter of the pipelines. Similarly, Papavinasam et al. [42] ascertained that
subcutaneous substances such as sand contributed to the increased rate of pitting corrosion in oil and gas pipelines. The
authors showed that the pitting corrosion rates varied from position to position due to flow type and velocities whilst
Mazumder et al. [43] attributed this variation to particle-to-particle, particle-to-fluid and particle-to-wall interactions at
such positions.
Manufacturing processes have also been found to contribute to the susceptibility of pipelines to corrosion attack as was
shown after testing the impact of hydrostatic pressure on Fe–20Cr material by Zhang et al. [44]. They concluded that
increased hydrostatic pressures used for testing of pipes resulted in loss of corrosion resistance of the materials. This was
evident in increased pit generation and growth rates and enlargement of metastable pits into cavities.

2.1. Sweet corrosion

When oil and gas flow through pipelines, the wall of the pipeline can be wet with oil or water depending on the fluid that
is in the continuous phase. Oil-wet-surface occurs if water is trapped in the oil which is in continuous phase and in direct
contact with the wall of the pipeline (water-in-oil phase) whereas water-wet-surface occurs when oil is trapped in the water
which is in direct contact with the pipeline wall and also in continuous phase (oil-in-water phase) [45]. Oil-wet-surface may
C.I. Ossai et al. / Engineering Failure Analysis 53 (2015) 36–58 41

not readily cause pipeline corrosion since oil may not form an electrolyte for redox reaction. Water-wet-surface is caused by
oil-in-water phase flow, wet gas and condensation in gas transmission pipelines are potential sources of water [46,47] which
may result in CO2 absorption [26,48], cathodic and anodic reactions [7] and oxide films formation [26]. The acidification of
CO2 gas via reaction with water can increase corrosion rate of pipelines [48] as the system tends to balance the chemical
equilibrium by creating more anodic and cathodic reactions. The electrochemical reaction at the water-wet surface for
CO2 induced corrosion is shown below:

– Absorption of CO2:

CO2 ðgÞ þ H2 OðlÞ $ H2 CO3 ðaqÞ ð1Þ

H2 CO3 $ Hþ þ HCO3 ð2Þ

HCO3 $ Hþ þ CO2
3 ð3Þ
– Cathodic reaction may occur either by hydrogen reduction or carbonate reduction [45]:
2Hþ þ 2e ! H2 ð4Þ

1
H2 CO3 þ e ! HCO3 þ H2 ð5Þ
2

1
HCO3 ! CO2
3 þ H2 ð6Þ
2
– The anodic reaction follows an oxidation process for iron:
Fe ! Fe2þ þ 2e ð7Þ
– Oxide films formation can follow any of the following steps:
Fe2þ þ CO2
3 ! FeCO3 ð8Þ

Fe2þ þ 2HCO3 ! FeðHCO3 Þ2 ð9Þ

FeðHCO3 Þ2 ! FeCO3 þ CO2 þ H2 O ð10Þ


Numerous experts have commented on different aspects of sweet corrosion of pipelines using empirical and experimental
approaches. Nesic [26] reviewed the physico-chemical modelling of CO2 corrosion in carbon steel and stated that the corro-
sion mechanism centres on the anodic dissolution of iron and the cathodic evolution of hydrogen. This involves a direct elec-
trochemical reactions of reduction of carbonic acid and water. The iron carbonate produced during the electrochemical
reaction of CO2 corrosion forms a protective Layer which prevents more corrosion from taking place by covering the carbon
steel surface from the corrosive species [6]. However, erosion and turbulence in pipelines mechanically remove them giving
room for more corrosion [39,49] or else dissolution of the scale due to chemical reactions necessitated by microorganisms
[50] or operating condition of the flowing oil and gas [47].

2.2. Sour corrosion

This is a sulphide ion induced corrosion in which hydrogen reacts with iron to form iron sulphide that may act as a pro-
tective scale depending on the concentration of the hydrogen sulphide and the prevalent environmental condition. The gen-
eral equation for this reaction is shown in Eq. (11) below [51].

H2 S þ Fe þ H2 O ! FeSx þ 2Hþ þ H2 O ð11Þ


Sour corrosion has resulted in fatigue failure of pipelines by causing hydrogen embrittlement [52], pitting corrosion [42],
lamination [53] and Stress Oriented Hydrogen Induced Cracking (SOHIC) which could be a manifestation of Sulphide
Stress Corrosion (SSC) [46,54]. Elemental sulphur have been reported by some researchers as being responsible for the local-
ized pitting corrosion attack on carbon steel material however, Song et al. [55] suggested that other substances such as SO2
4 ,
SO2 2 +
3 , S2O3 and H may have contributed to the localized attach in a sour environment. The probable mechanism of macki-
nawite film formation in sour environment according to Popoola et al. [51] is shown in Fig. 4.

2.3. Microbiological Induced Corrosion (MIC)

This type of corrosion is induced by microorganisms in the pipelines. These microorganisms initiate different chemical
activities that could include – production of organic and inorganic acids [56], oxidation and reduction of some elements
[57]. The activities of these microorganisms on pipelines are found to be most predominate at pH of 4–9 and temperature
42 C.I. Ossai et al. / Engineering Failure Analysis 53 (2015) 36–58

Fig. 4. Probable mechanism for mackinawite formation in sour corrosion environment.

of 10–50 °C [57]. In general, microorganisms produce biofilms which form a good site for both aerobic and anaerobic activi-
ties that enhance the corrosion of carbon steel materials. This is done by physical deposition of toxic substances, production
of corrosive acidic bye-products and depolarization of the corrosion cell through increased utilization of hydrogen, oxygen or
iron compound in the environment [56]. These secreted metabolites can enhance pitting corrosion [58,59], de-alloying of
metals [60,61], galvanic corrosion [47,58], stress corrosion cracking and hydrogen induced cracking [62,63]. Research has
shown that at a linear velocity of 3.5 m/s, the biofilm formed by Sulphur Reducing Bacteria (SRB) could not adhere to the
wall of a pipelines [64] hence Sulphur Reducing Bacteria (SRB) may not be actively enhancing internal pipeline corrosion
for fluids under turbulent flow regime.
Experimental observation has shown that microorganisms that induce corrosion in steel materials are more active at
lower concentration of the corrosive environment than at higher concentrations (see Fig. 5) [57]. It is evident from this figure
that at 1% NaCl concentration, the activities of Iron Oxidizing Bacteria (IOB) were higher as the microorganism attacked the
carbon steel material by more oxidation than it did at 6% NaCl concentration.

Fig. 5. Epifluorescence micrographs of Iron Oxidizing Bacteria (IOB) inoculated medium containing (a) 1% NaCl and (b) 6% NaCl, adapted from
Chandrasatheesh et al. [57].
C.I. Ossai et al. / Engineering Failure Analysis 53 (2015) 36–58 43

Table 2
Summary of bacteria associated with the corrosion of pipelines.

Bacteria Mode of corrosion attack Remarks


Sulphur Reducing Respire sulphate and produce sulphide while oxidizing diverse Increase of anodic reaction to establish a chemical
Bacteria (SRB) electron donors [66] equilibrium results in more corrosion of the material
Acid Producing Produce acidic compounds such as acetic and sulphuric acids that Can be introduced by injecting water from seas and rivers
Bacteria (APB) may chemically attack the protective surface in pipelines into oil and gas fields [50]
Iron Reducing Attack iron compounds by reducing Fe3+ to Fe2+ [57] Promotes corrosion by dissolving the protective film
Bacteria (IRB) layers on material surfaces
Iron Oxidizing Oxidizes Fe2+ to Fe3+ by depositing metabolites [58] Can produce low pH that enhances Fe3+ dissolution
Bacteria (IOB)
Manganese Deposits MnO2 that favours the growth of sulphur reducing bacteria Sulphur reducing bacteria simulates depolarization
oxidizing [56] which brings about increase in the corrosion process
bacteria

Microorganisms such sulphur reducing bacteria can be introduced into the pipelines through formation water injection
[50,64] and can enhance external corrosion of underground buried pipelines (with disbonded coating) from the soil [65].
Sulphur reducing bacteria can also exist in high and low temperatures, deep in subsurface reservoirs and can be introduced
into oilfield during drilling operation [66]. Table 2 summaries different kinds of bacteria that may cause corrosion of
pipelines.

3. Pipeline fatigue stress failure mechanisms

3.1. Stress corrosion cracking

The interaction of corrosive environment and cyclic stress in materials results in stress corrosion cracking [8,67]. In oil
and gas transmission pipelines, both high pH and low pH stress corrosion cracking have been reported by numerous authors
[62,68,69]. Stress corrosion cracking has the ability of increasing the risk of failure of oil and gas transmission pipelines and
results in approximately 15–20% gas pipeline failures depending on the ages of the pipelines [5,70]. It occurs by different
mechanisms that include hydrogen embrittlement, adsorption-induced cleavage, atomic surface mobility, film rupture,
stress-accelerated dissolution, film-induced cleavage, tunnel pitting and tearing and localized surface plasticity [71].
Whereas low pH stress corrosion cracking occurs in the presence of dilute solutions such as carbonates and bicarbonates
having a pH of approximately 6, high pH stress corrosion cracking occurs at a pH of approximately 9 [72] and the rate of
crack growth is influenced by temperature [73]. Fig. 6 shows a leaking jet fuel transmission pipeline subjected to stress cor-
rosion cracking.
Chemicals such as chloride ions pumped into oil and gas reservoirs during water injection [66,74] for improved oil and
gas recovery, pass through the pipelines with the extracted oil and gas causing different forms of localized corrosion such as
pitting [25,66]. These corroded areas of pipelines have become sites for stress corrosion cracking as cyclic stress induced on
the pipelines from the transported fluids systematically initiates cracking over time. Similarly, undercutting caused by tur-
bulence at 6 o’clock position in pipelines (see Fig. 7) results in localized pitting corrosion which leads to stress corrosion
cracking and pipeline leakage [40].
The point of fatigue stress failure due to stress corrosion cracking on the pipelines has been shown to have the chemistry
of oxides changing with distance moving away from the oxide, from metal-rich to oxygen-rich [75,76] (see Fig. 8). Again,
when pipelines are increasingly exposed to a corrosive environment under repeated stress load caused by the operating

Jet fuel leaking due to stress corrosion cracking

Fig. 6. Leaking jet fuel transmission pipeline subjected to stress corrosion cracking.
44 C.I. Ossai et al. / Engineering Failure Analysis 53 (2015) 36–58

Fig. 7. Schematic of turbulence induced localized corrosion that can result in stress corrosion cracking, adapted from Ilman and Kusmono [40].

Fig. 8. Corrosion products of localized corrosion region of X52 pipeline; (a) corrosion product; (b–d) EDX-spectra taken from regions marked A–C
respectively in (a), adapted from Ilman and Kusmono [40].
C.I. Ossai et al. / Engineering Failure Analysis 53 (2015) 36–58 45

Extended crack by fague


(a) (b)
Fig. 9. (a) Cracked profile of extended Stress corrosion cracking (etched in 5% Nital) hoop plan view, adapted from Gamboa et al. [68]. (b) Micrograph
showing change of crack morphology (etched in 5% Nital) hoop plan view, adapted from Gamboa et al. [68].

pressure, existing cracks on them grow as new ones are initiated [77]. Gamboa et al. [68] also showed that dormant Stress
Corrosion Cracking (SCC) in carbon steel pipelines will grow by fatigue during the service life by experimentally applying
cyclic pressures on X65 grade of pipe. This fatigue extended stress corrosion cracking on X65 grade material subjected to
fatigue loading is shown in Fig. 9a and b.
After analysing pipeline failure due to chloride ingression, Kumar et al. [78] recommended the following mitigation
actions – use of low chloride insulation, coating of pipeline prior to insulation, substituting chloride containing glass wool
with calcium silicate based insulation and water proofing insulation. The transgranular fracture and secondary cracks along
the grain boundaries of the stainless steel due to chloride stress corrosion cracking is shown in Fig. 10.
The impact of stress corrosion cracking on Heat Affected Zones (HAZs) of pipelines were also studied by some researchers
such as Contreras et al. [69] who tested the susceptibility of X52 and X70 welded grades of pipeline to stress corrosion crack-
ing by using strain rate test in H2S saturated environment. They concluded that a ductile type failure of the materials hap-
pened outside corrosive environment and brittle type failure happened in corrosive environment which is evidence that the
mechanical properties of the material are reduced in the corrosive environment [79]. The authors further stated that the
permeability of hydrogen on X52 weld is higher than X70 weld. This suggests that heat affected zones of X52 grade pipe
may be more susceptible to stress corrosion cracking in corrosive environment than those of X70. X70 pipeline has also been

Fig. 10. Transgranular fracture and secondary cracks along the grain boundaries of chloride stress corrosion cracking, adapted from Kumar et al. [78].
46 C.I. Ossai et al. / Engineering Failure Analysis 53 (2015) 36–58

Table 3
Summary of some factors responsible for stress corrosion cracking in fatigue failed pipelines.

S/ Causes Examples Remarks


N
1 Disbondment of 254 mm X52 gas pipeline failed due to action of sulphur Due to carbonate/bicarbonate reaction with the carbon
coating reducing bacteria and stress corrosion cracking [82] steel material of the pipeline
X60 and X42 gas pipeline failed due to the combination of The failure of X60 occurred after 40 years whilst X42
electrochemical reaction and stress on the disbonded occurred after 6 months due to the electrochemical
portion of the pipeline [8,83] reaction in the disbonded area
2 Manufacturing defect Error on sand blasted X46 grade gas transmission pipeline Due to the reduction of the wall thickness and formation
caused cavity due to erosion mechanism [84] of a stress raiser resulted in the failure of the pipeline
during hydrostatic testing
Faulty trimming of a longitudinal step on a pipeline surface This 457.2 mm pipeline failed after 25 years of service
[85]
Faulty procedure of manufacturing a 762 mm  457.2 mm The fault was discovered during hydrostatic testing
butt welded tee-joint [86]
3 Electrochemical Both electrochemical reaction and erosion contributed to The failure occurred after 27 years
reaction and Flow the failure of X52 gas transmission pipeline [40]
assisted corrosion Erosion mechanism resulted in the rupture of the A leaking gas caused the sand blasting of a buried pipeline
609.6 mm gas pipeline [87] which ruptured after substantial Abrasive reduction of
the wall thickness
Turbulent gas stream caused pitting corrosion of X52 gas The turbulence resulted in under deposit at the 6 o’clock
transmission pipeline [39] position and eventually localized corrosion
4 Ingression of X46 pipeline absorbed hydrogen which subsequently lead The action of H2S assisted the mechanism in the failure of
corrosive species to hydrogen Induced cracking of the material [75] this X46 grade transmission pipeline
Chlorine ingression from the wool used for thermal This stainless steel pipeline supplied hydrogen to a hydro
insulation of the stainless pipeline caused localized pitting cracking reactor of a petrochemical industry
corrosion [78]

shown not to form stable oxides on the surface in near neutral pH solution [80]. Kentish [81] also used slow strain rate test-
ing method to determine the effect of surface roughness created by grit blasting on the resistance of X70 grade pipe to stress
corrosion cracking. This author concluded that the time for failure ratios decreased with increasing roughness which is an
indication of reduced resistance to stress corrosion cracking. This implies that even essential manufacturing techniques such
as grit blasting which creates adhesive surface for coating of pipelines can potentially increase the risk of stress corrosion
cracking. Some of the causes of stress corrosion cracking on pipelines are summarized in Table 3.

3.2. Hydrogen induced cracking

Structural integrity evaluation is vital for risk minimization in the operation of oil and gas pipelines hence the reason why
many authors investigated hydrogen permeability of different carbon steel and Corrosion Resistance Alloys (CRAs)
[52,71,88]. Hydrogen can permeate into carbon steel materials exposed to aqueous H2S environment and deposit on defec-
tive locations of the metallic matrix as inclusions which can affect the fractural behaviour of the material [71]. This can hap-
pen when the material is exposed under load externally to a hydrogen containing environment or internally by hydrogen
located within the bulk of the material [71]. Hydrogen embrittlement which is a mechanism by which carbon steel materials
especially high strength materials become brittle and fracture when exposed to an environment where hydrogen is gener-
ated [89] can occur by high-pressure bubble formation [71]. It can also occur by reduction in surface energy (adsorption
mechanism) [90], reduction in lattice cohesive force (decohesion mechanism) [91], hydrogen interaction with dislocation
[88] and hydride formation [88,90]. This could result in the reduction of physical and chemical properties of the material
especially around the heat affected zones [79] and can cause fatigue failure in extreme cases. Natividad et al. [92] showed
that the microstructures of pipeline materials are distorted at the heat affected zones resulting from welding operation used
for joining pipes during pipeline installation. These authors also experimentally tested the susceptibility of X65 pipeline
weldment produced by three different processes – Indirect Electric Arc Welding (IEAW), Submerged Arc Welding (SAW)
and Metal Inert Gas (MIG) to sulphide stress cracking and discovered that indirect electric arc weldment has a better
Stress Sulphide Cracking (SSC) resistance than others. The authors believed that the microstructural modification of the
welded bead from ferrite in a needle-like form to a fine grain microstructure was responsible for the reduced susceptibility
of indirect electric arc weldment at 25 °C to sulphide stress cracking. The weldments have been previously tested using Slow
Strain Rate Tests (SSRT) and electrochemical measurements at NACE solution saturated with H2S at 25 °C, 37 °C and 50 °C.
Since Electric Resistance Welding (ERW) and Submerged Arc Welding (SAW) which are predominantly used in the oil and
gas industry for pipeline installation result in a wide area of Heat Affected Zones (HAZs) and create heterogeneous
microstructures that increases the risk of cracking [93,94], development of narrow HAZs with indirect electric arc welding
may help to reduce sulphide stress cracking of pipelines [92].
C.I. Ossai et al. / Engineering Failure Analysis 53 (2015) 36–58 47

Fig. 11. Formation of step-wise crack in a material by pressure mechanism, adapted from Gonzalez and Morales [96].

Other researchers such as Azevedo [75] investigated the failure of X46 pipeline and linked the failure to hydrogen blisters
in the metallic matrix whereas Beidokhti et al. [52] experimentally tested the impact of manganese content (1.4% and 2.0%)
of X70 pipe on hydrogen induced cracking and sulphide stress cracking of submerged arc welded test pieces. They found out
that the weldment with 2.0% manganese content was more susceptible due to the higher hardness of the weldment. Capelle
et al. [88] tested the hydrogen absorption ability of X52 pipe by estimating the burst pressure of externally notched pipe
under the pressure of hydrogen and methane gas with the conclusion that at a low pressure of 2 MPa, hydrogen could pene-
trate the near surface layer of the material. They also stated that the penetration rate of hydrogen increases with time and
applied stress caused by the natural gas. Additionally, Park et al. [95] experimented on X65 pipe and showed that degener-
ated pearlite, acicular ferrite, bainite–ferrite and martensite–austenite microstructures have the ability to trap and diffuse
hydrogen with acicular ferrite having the highest absorption efficiency. Woodtli and Rolf [91] on their part suggested that
hydrogen gas contact with steel surface at room temperature may not pose a problem considering the very high dissociation
constant (429.5 kJ/mol) and very low diffusion coefficient (1.5  108 cm2/s). However, cracks can be formed on the steel due
to the impact of mechanical stress concentration and hydrogen dissolution. From the foregoing discussion, it implies that
most pipelines transporting oil and gas are vulnerable to hydrogen ingression given the low dissociation constant and stress
pressure necessary for diffusion hence, the need for high quality manufactured materials free of inclusions and other
microstructural defects that will act as grounds for hydrogen related defects initiation.
Lamination which is multiple stepwise cracks on the wall of pipelines affect the residual strength of materials [96]. In
sour gas pipelines, hydrogen absorption may cause lamination on different sections of pipelines. The growth of lamination
may result in hydrogen induced cracking (HIC) when increment of internal pressures in the lamination lead to interaction of
stress fields of adjacent crack tips (see Fig. 11) [97]. Similarly, the experiment on X70 grade steel pipe at different H2S con-
centrations, stress ratios and load frequencies indicated that fatigue crack growth rate increased with stress ratio, decreased
with load frequency but showed no linear relationship with H2S concentration [98]. Furthermore, at 775 mV potential X80
grade pipe showed no apparent reduction in area but at 1000 mV to 1200 mV [99], the area varied between 38.96% and
21.25% showing that, at reduced potentials, the susceptibility of the material to hydrogen induced cracking is higher [99].
Because hydrogen diffusion into the microstructure of pipelines has resulted in an increased susceptibility of this asset to
fatigue stress failures as the yield strength increased, the strength of corrosion resistance alloy for service in sour environ-
ment was limited to 690 MPa or 22 HRC [100].
The measure of the appropriate level of hydrogen necessary for a pipeline material requires a guideline which could
include those shown in Table 4.

4. Assessment of burst pressure of corroded pipelines

Safe operation of pipelines can be achieved if proper assessment of the fitness-for-service is carried out from design,
manufacturing, installation and operation [75,102]. Since this assessment is vital for downtime minimization and high pro-
ductivity in oil and gas industry, numerous standards have been developed by different regulatory agencies such as
American Petroleum Institute (API), American Society of Mechanical Engineers (ASME), British Standards Institute (BSI),
48 C.I. Ossai et al. / Engineering Failure Analysis 53 (2015) 36–58

Table 4
Laboratory methods for estimating materials susceptibility to hydrogen.

Type of hydrogen effect Standard Specific tests


Sulphide Stress Cracking (SSC) MRO175/ISO 15156-2a NACE TMO177
Stress-Oriented Hydrogen Induced Cracking (SOHIC) MRO175/ISO 15156-3b EFC publication 16
Soft zone cracking MRO175/ISO 15156-2a NACE TMO177
Step-wise cracking MRO175/ISO 15156-2a NACE TMO284
Hydrogen induced cracking MRO175/ISO 15156-2a NACE TMO285

Adapted from Papavinasam [101].


a
Carbon steel.
b
Corrosion Resistance Alloys (CRAs).

Standards Australia and Joint Industry Project (JIP). Many researchers have also proposed assessment methods for corrosion
defects [103–105] in order to establish the burst pressure at such corroded points by using empirical and numerical meth-
ods. These methods considered the longitudinal corrosion defect length, depth of the corroded section and the internal
operating pressure of the pipelines [106,107] in assessing the burst pressure. The predominant techniques for assessment
of burst pressure of corroded pipelines include – ASME B31G [107], modified ASME B31G (RSTRENG 0.85) [103], SHELL
92 [108], DNV-RP-F101 [106], PCORRC [109] and safe-SwRI [110].
ASME B31G and modified ASME B31G codes evaluate the remaining strength of corroded pipelines based on burst pres-
sure experiments which considered corrosion defects to be either rectangular or parabolic shaped [111] according to Fig. 12.
The codes assumed that the burst pressure of corroded pipelines are not impacted by the toughness of the material [111] but,
experimental analysis of aged pipelines showed that low toughness may have contributed to a variation of 14–24.5%
between the experimentally measured and predicted burst pressures using industry standards model such as ASME B31G
and ASME B31G modified [112].
The burst pressure (Pf) can be calculated by accounting for yield stress (rY), external pipeline diameter (D), defect depth
(d), wall thickness (t), longitudinal corrosion defect length (L) and bulging factor (M) using the equations shown below [107]:
ASME B31G:
 
2ð1:1rY Þt 1  ð2=3ðd=tÞÞ
Pf ¼ ð12Þ
D 1  ðð2=3Þððd=tÞ=MÞÞ
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
M¼ 1 þ 0:8ðL=DÞ2 ðD=tÞ ð13Þ
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
For 0:8ðL=DÞ2 ðD=tÞ  4, the defect is parabolic.
For rectangular defect,
2ð1:1rY Þt
Pf ¼ ½1  ðd=tÞ; M¼1 ð14Þ
D

L L

Parabolic shape assumed for Rectangular shape assumed for


short defects long defects

Fig. 12. Simple illustration of pipeline corrosion defect based on ASME B31G code.
C.I. Ossai et al. / Engineering Failure Analysis 53 (2015) 36–58 49

whilst,
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
0:8ðL=DÞ2 ðD=tÞ  4:

ASME B31G modified:


 
2ð1:1rY þ 69Þt 1  ð0:85ðd=tÞÞ
Pf ¼ ð15Þ
D 1  ð0:85ððd=tÞ=MÞÞ
8 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
<
1 þ 0:627ðL=DÞ2 ðD=tÞ  0:003375ðL=DÞ4 ðD=tÞ2 ; for ðL=DÞ2 ðD=tÞ 6 50
M¼ ð16Þ
:
3:3 þ 0:032ðL=DÞ2 ðD=tÞ; for ðL=DÞ2 ðD=tÞ > 50
DNV RP-F101 defect assessment model provides a codified formulation which considers pressure, bending, area and depth of
defects in calculating the burst pressure [106]. In the model, the failure pressure cannot exceed the maximum allowable
operating pressure the system is design to withstand but if that happens, the pipeline need to be repaired or replaced
[111]. The burst pressure can be calculated using Eq. (17) below:

Pf ¼ 2ðru Þt =ðD  tÞ½ð1  ðd=tÞÞ=ð1  ðd=tÞQ Þ ð17Þ


where the correction factor (Q) is given below:
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffi 2
Q¼ 1 þ 0:31ð1= DÞ ð18Þ
Choi et al. [113] used finite element analysis of corroded pipelines to propose a limit load function for calculating failure
pressure using the relationships shown below:
8 h pffiffiffiffiffi pffiffiffiffiffi 2 i pffiffiffiffiffi
>
< 0:9 2ðrDu Þt C 0 þ C 1 ðL= RtÞ þ C 2 ðL= RtÞ ; for ðL= Rt Þ < 6
i
Pf ¼ h pffiffiffiffiffi i pffiffiffiffiffi ð19Þ
>
: 2ðru Þt C 3 þ C 4 ðL= Rt Þ ; for ðL= RtÞ P 6
D i

where ru is the ultimate tensile strength, Di is the internal diameter of the pipeline, R is the radius of the pipeline and C0, C1,
C2, C3 and C4 are given by Eq. (20) below:
8
2
>
> C0
>
>
¼ 0:06ðd=tÞ  0:1035ðd=tÞ þ 1
>
> 2
>
< C1 ¼ 0:6913ðd=tÞ þ 0:4548ðd=tÞ  0:1447
2 ð20Þ
> C2 ¼ 0:1163ðd=tÞ  0:1053ðd=tÞ þ 0:0292
>
>
>
> C ¼ 0:9847ðd=tÞ þ 1:1101
>
> 3
:
C4 ¼ 0:0071ðd=tÞ  0:0126
Structural integrity assessment procedure (SINTAP) accounts for plastic collapse, brittle failure and the safety factors in burst
pressure analysis using Failure Assessment Diagram (FAD) shown in Fig. 13. Failure assessment diagram takes cognizance of
normalized stress intensity factor (Kr) and normal stress [108] in the estimation of defects on pipelines and have been used
by different researchers [114–117]. Adib-Ramezani et al. [111] used structured Integrity Assessment Procedure (SINTAP) to

Fig. 13. Failure assessment diagram used in structural integrity assessment procedure (SINTAP) failure analysis, adapted from Adib-Ramezani et al. [111].
50 C.I. Ossai et al. / Engineering Failure Analysis 53 (2015) 36–58

analyse X52 pipes subjected to corrosion and concluded that the burst pressure predicted are within the upper and lower
limits of those predicted by ASME B31G, modified ASME B31G and DNV RP-101. Meliani et al. [118] used a modified notch
failure assessment diagram to get a safety factor greater than 2 which made the result conservative since it was within struc-
tural integrity assessment procedure (SINTAP) safety factor acceptance limit. These authors initially used ASME B31G, modi-
fied ASME B31G, DNV F 101 and Choi’s method to calculate the burst pressure of the material in order to estimate the
normalized stress intensity and stress intensity factors. Adib et al. [117] used finite element analysis to investigate defects
of X52 grade gas transmission pipeline and obtained a safety factor of 3.4 which is higher than the minimum expected limit
of 2 prescribed in by the standard however, the security factor was found to be 1.7 which is less than 2 stipulated by the
model. Although many researches based on failure assessment diagram did not consider the constraining effects of supports
on pipelines, Wang and Reinhardt [119] considered the impacts of such supports on the stress intensity factor and load limits
of a through-wall projected crack on a tube. They concluded that the maximum allowable percentage of degraded area can
increase by a margin of 20–30% in comparison to assessment that did not take into consideration the restraining effects of
supports. This implies that numerous structural integrity analysis of pipelines that did not consider the effects of supports on
the structures whilst concluding on their reliabilities may have been significantly flawed.
The interpolating function (f(Lr)) can be computed using the equations below [120]:
8h  ið1=2Þ h i
>
> 1 þ L 2
=2 0:3 þ 0:7e lL6r
; for 0 6 Lr 6 1
>
> r
< pffiffiffiffiffiffiffiffi
f ðLr Þ ¼ 2=3½0:3 þ 0:7e Lr  l ðN1=2NÞ
; for 0 6 Lr 6 Lmax r
ð21Þ
>
>h ið1=2Þ
>
>
: ðEe =r Þ þ ð1=2ÞðL2 =ðEe =r ÞÞ ; for 0 6 Lr 6 Lmax
ref ref r ref ref r

where,
 
E
l ¼ min 0:001 ; 0:6 ð22Þ
rY
 
rY
N ¼ 0:3 1  ð23Þ
ru
8
< 1 þ ð150=rY Þ2:5
Lmax
r ¼ 1 hrY þru i ð24Þ
:
2 ru

rref ¼ Lr rY ð25Þ

where Lr is the non-dimensional loading, Lmax


is the maximum value of non-dimensional loading, E is the modulus of elas-
r
ticity, l is the first correction factor, N is the second correction factor, eref is the reference strain and rref is the reference
stress.
Failure pressure of corroded pipelines can also be determined by using von Mises, Tresca and Average Shear Stress Yield
(ASSY) theories which assumes that failure of pipelines follow a strain hardening models which follows pure power law
[121,122]. The failure pressure can be calculated based on Eq. (26) [121].
 nþ1  
G 4t 0
Pf ¼ ruts ð26Þ
2 D0
where t0 is the original wall thickness, D0 is the original average external diameter, ruts is the engineering ultimate tensile
pffiffiffi pffiffiffi
strength, G is the yield theory dependent constant and has a value of 1, ð2= 3Þ and ðð1=2Þ þ ð1= 3ÞÞ for Tresca, von Mises
and ASSY theories respectively, n is the strain hardening exponent and ranges from 0 to 0.3 for most carbon steel pipelines
[121]. Ma et al. [122] used von Mises strength failure criteria to predict the burst pressure of low, mild and high strength
pffiffiffiffiffiffi
corroded pipelines using Finite Element Analysis (FEA). The authors considered 195 cases of (L= Dt ) ranging from (0.15
to 13.69) and (d/t) ranging from (0.2 to 0.6) and established that failure pressure of corroded pipelines decrease with the
pffiffiffiffiffiffi pffiffiffiffiffiffi
increase of L= Dt and. However, at smaller values of L= Dt and d/t, the failure pressures only decreased slightly. To further
validate this work, the authors used 79 groups of pipeline burst test experimental data of low, mild and high strength grade
steels obtained from literature to compare failure pressures of the experimental results with those obtained from industry
standard models such as – ASME B31G [107], DNV F101 [106], PCOORC [109], Zhu [123], Choi’s [113] models and their pre-
diction model shown in Eq. (27). The result indicated that the prediction model could estimate the failure pressures of low
strength, mild strength and high strength steels to an error margins of 13.99%, 8.03% and 4.95% respectively. Hence, indicat-
ing that the predicted model was more conservative in calculating the failure pressure of high strength materials than low
and mild strength ones (see Fig. 14).
C.I. Ossai et al. / Engineering Failure Analysis 53 (2015) 36–58 51

Fig. 14. predicted and actual failure pressure of low, mild and high strength steel using Ma et al. [122] corroded pipeline failure model.

0 1
    
4 t0 d pffiffiffi Þð1dÞ
ð0:4174L
0:1151
Pf ¼ @ pffiffiffi nþ1 A ðruts Þ 1  1  0:7501e Dt t
ð27Þ
D0 t
ð 3Þ n

Basel et al. [103] also used von Mises burst pressure prediction method to estimate the failure pressure of corroded area of a
high grade carbon steel pipe (X100) by using finite element analysis to evaluate plastic flow in the areas with 25% and 75%
wall thickness losses. The conclusion of the work shows that the pressure containment capability of the pipe was not sig-
nificantly affected at those wall thickness losses. Furthermore, Xu and Cheng [105] also used finite element analysis to evalu-
ate corroded X65, X80 and X100 pipes and stated that the burst pressures decreased with increase in corrosion depths as
expected whereas the defect geometry on the pipe was dependent on the local strain/stress distribution of the environment.
Additionally, increase in longitudinal defect length on pipelines resulted in decrease of burst pressures due to the elastic load
carrying capacity decrease of the material [7,104].
The work of other researchers that addressed failure pressure and reliability estimation of pipelines includes Wang and
Zarghamee [14] who used non-linear finite element analysis to analytically investigate the variability of pipeline strength at
different corroding environments. They also determined a reliability model for the material strength and predicted a frame-
work for fitness-for-service testing. Hasan et al. [124] used Monte Carlo simulation and first order second moment method
for analysing the burst pressure of internally corroded pipelines and had results which compared reasonably with industry
standard failure assessment methods. On the other hand, Teixeira et al. [125] used first-order reliability method, 3D non-lin-
ear finite element analysis and experimental validation to predict the burst pressure of corroded pipelines. Markov decision
process has also been used by other researchers [126,127] to determine the pitting corrosion growth of pipelines in a bid to
estimate the remaining useful life.

5. Pipeline integrity optimization framework

The increased need for energy and the attendant pressure on oil and gas pipelines entails that the integrity of this asset
should be maintained to minimize non-productive time [128], increase the efficiency of oil and gas transportation and pro-
tect the environment. Kishawy and Gabbar [129] summarized pipeline integrity management to involve – identification of
process for understanding pipeline failure mode, assessment plan for the failure modes, integrity analysis procedure for fail-
ure and consequences, setting criteria for repair action and information analysis, continuous process of assessment to keep
maintaining integrity, mitigation and preventative actions to forestall downtime, effectiveness measure and review tech-
niques. Integrity management involves deciding the type of material to use in laying a pipeline at the design stage [130].
This is vital in risk management and cost optimal operation seeing that pipeline failure accounts for more than 50% of
the total downtime in oil and gas industry [22]. If Carbon steel material is used for a pipeline, there may be high operational
expenditure (OPEX) involved in running the pipeline due to the high corrosion susceptibility of the material whilst the use of
Corrosion Resistance Alloys (CRAs) may bring about high capital expenditure (CAPEX) as the initial cost of investment on the
material is very high but it has high corrosion resistance qualities [130]. It is therefore imperative that decision makers
should strike a balance between these two costs (OPEX and CAPEX) in material selection in order to maintain the appropriate
reliability of the pipeline and reduce risk whilst optimizing profit through minimal lifecycle cost of the pipeline. Fig. 15
52 C.I. Ossai et al. / Engineering Failure Analysis 53 (2015) 36–58

Fig. 15. Variation of pipeline lifecycle cost, operating expenditure (OPEX) and capital expenditure (CAPEX) with pipeline integrity.

shows a trade-off between operating expenditure and capital expenditure in determining the appropriate material for pipe-
line design.
This figure indicates that very low capital expenditure on pipeline material may result in very high operating expenditure
which may include- cost of chemical injection, inspection, maintenance and repair and environmental damage penalties
should pollution occur whilst high capital expenditure may bring about minimal operating cost since the material is
expected to be highly resistant to corrosion and very low risk. Both extremes of the costs may not be to the advantage of
any company since there is very high risk in adopting low capital expenditure and high operating expenditure whereas there
may be problem of profitability in adopting high capital expenditure and low operating cost. This is why the balance
between these two costs is vital for the management of pipeline integrity and enhancing profitability of an organization
since risk and cost will be optimized.
Different frameworks for corroded pipelines integrity management and fitness-for-service testing have been established
by different regulatory agencies and they includes API 579-1/ASME FFS-1 [131], Canadian Association of Petroleum
Producers (CAPP) 5-M approach [21], BS 7910 [132] and ASME B31.8S [133]. Fig. 16 shows a typical pipeline integrity
optimization framework that may be used for managing pipeline against corrosion induced fatigue failures.
According to this figure, to strategically manage pipeline deterioration and fatigue failures may entail addressing health,
safety and environmental risks [13,134], threats and barriers for risk escalation [135], risk based inspection activities [22],
proactive and reactive actions for integrity control [22,129] and assurance activities for performance standards [102].
Corrosion risk assessment which aid in identification, removal and mitigation of corrosion through monitoring and inspec-
tion programs [7,136] is also vital for mapping the expected risk level of the pipeline under different operating environ-
ments. Probabilistic approach have been applied to risk level assessment of corroded pipelines [137] in order to ensure
that the pipeline is fit-for-purpose [138]. This can be achieved if total quality management is adopted at design, manufacture,
installation and operation [129] by testing, monitoring and inspection to stipulated standards. The techniques for control,
monitoring and inspection of pipeline against corrosion related defects are shown in Fig. 17.
Data used for pipeline integrity management have been collated by in-line-inspections, risk based inspection and process
monitoring. Unfortunately, imperfection in the collected data has been identified to be caused by variability and uncertainty
[4] and poor knowledge transfer through inadequate communication [15]. Singh and Markeset [4] attributed variability to
lack of uniformity of diameter, corroded pit, corroded length, pipe wall thickness whilst uncertainty could be caused by
equipment calibration and mal-functionality, personnel skill, and variability in operating characteristics of pipeline across
different zones of the pipeline.
Since changes in operating conditions may potentially influence the integrity of pipelines, process control of key operat-
ing parameters such as pH, water chemistry, pressure, CO2 and H2S partial pressures, oxygen, carbonates, bicarbonates and
bacteria [7,24,139] is vital for monitoring the level of corrosion on pipelines. Comparing the measured values of these
parameters against established key performance indicators will aid in predicting the integrity of the pipeline at any point
in time. Other corrective and proactive actions that may be useful in pipeline integrity management includes procedural
changes to chemicals applications [140], inspection strategies [141] and in some extreme cases repair or replacement of por-
tions of the pipelines [142]. Although the use of corrosion inhibitors have been attributed to internal corrosion reduction in
pipelines, the right concentration of the inhibitor is needed to achieve this. Martínez et al. [53] showed that 50 ppm of amine
type inhibitor gave the best corrosion inhibition in X52 pipeline whereas 100 ppm increased the corrosion rate at static con-
dition. The concern about environmental pollution in use of corrosion inhibitors led to the testing of lanthanum-4 hydroxyl
cinnamate compound for their corrosion inhibitive ability [143]. The work showed that this compound inhibited both the
C.I. Ossai et al. / Engineering Failure Analysis 53 (2015) 36–58 53

Fig. 16. Pipeline Integrity assurance cycle.

Fig. 17. Summary of pipeline integrity management techniques.


54 C.I. Ossai et al. / Engineering Failure Analysis 53 (2015) 36–58

Table 5
Summary of pipeline leakage detection techniques.

S/N Methodology Principle Remarks


1 Mass balance technique Uses law of mass transfer Identifies leakage by balancing the difference
between input volume and output volume
2 System rate of change Uses change in pressure or flow rate After accounting for frictional losses, decreases in
pressure and flow rate may be attributed to
leakage.
3 Pressure point analysis Uses statistical analysis of pressure and flow rate May require a number of transient runs to detect
leakage of various magnitude from pipelines
4 Wave alert system Uses wavelength variation Detects leakage through bidirectional arrival of
negative pressure wavelength from ruptured
pipeline
5 Supervisory control and data Uses system data monitoring Can pick-up information about blocked pipes at
acquisition (SCADA) metering points and valve sites
6 Fibre optics Uses fibre optics sensors such as – clustered sensor Uses the change in temperature, mechanical stress,
system, discrete sensor system, continuous surface coating or material absorption of the fibre
distributed sensor systems optics material
7 Chemical detection Uses chemical sensors such as- discrete sensors, Chemical probes and sensors can detected the
distributed chemical sensors, distributed fibre optics accurate point of pipeline leakage
sensors
8 Acoustic method Uses turbulent velocity fluctuation Fluid leakage may result in the change of turbulent
characteristics of such a pipeline

anodic and cathodic reactions at 0.01 M NaCl solution exposure of carbon steel after using linear polarisation resistance and
cyclic potentiodynamic polarisation measurements to check the rate of corrosion of the material. Other environmental
friendly inhibitors that are in predominant use in controlling pipeline corrosion in the oil and gas industry includes imidazo-
line and nitrogen-based inhibitors [144].
Because monitoring is essential for pipeline integrity monitoring, Huang and Ji [145] experimentally tested the use of Cu–
Zn galvanic sensor to monitor internal corrosion of pipeline in seawater environment by measuring the difference between
dissolved oxygen entrant and exit. After comparing the corrosion rates monitored with this sensor and corrosion coupon, the
authors affirmed that the sensors can be used to reasonably predict internal corrosion of pipelines despite the need for reli-
able sensor design and data-base accumulation for the corrosive environment. Although repair of a pipeline by repeated
welding operation has been shown not to affect the integrity of X52 pipelines [146], Fazzini and Otegui [142] after hydro-
statically testing several rectangular, elliptical and other geometry type welded sections of X52 gas transmission pipeline at
200% of Maximum Allowable Operating Pressure (MAOP) concluded that repaired patches that pose most risks to pipeline
integrity are those with poor quality of welds, placed with the pipeline at less than half the MAOP, rectangular and roughly
twice longer than the width and repairs that have large dimensions with the corrosion defect more than 40% of the nominal
wall thickness.
To effective manage pipelines against fatigue induced failures, leakage may be monitored with some of the techniques
listed in Table 5 [147].

6. Conclusions

Safe operation of oil and gas transmission pipelines requires mitigation of fatigue stress caused by corrosion defects in
operating environments. These corrosion defects are caused by the electrochemical reaction of water and iron content of
the pipeline material in the presence of CO2 (sweet corrosion), H2S (sour corrosion) and/or microorganisms (microbiologi-
cally induced corrosion) such as bacteria. The corrosion reaction produces protective scales such as FeCO3, CaCO3 and
FexSx which can temporarily prevent corrosion however, hydrodynamic force in the fluid flow removes the scales exposing
the pipeline surfaces to more corrosion. The microorganisms introduced into the pipeline via injection water pumping, dril-
ling operations or those surrounding buried pipelines secrete metabolites whose activities accelerate anodic/cathodic reac-
tions which accelerates corrosion rate.
Localized corrosion sites on pipelines are points of initiation of stress corrosion cracking due to the cyclic loads induced
on the pipelines by the operating pressures. Stress corrosion cracking occurs in both low and high pH environments and
potentially caused fatigue failures of 20–25% depending on the age of the pipeline. Stress corrosion cracking have been found
to be majorly caused by disbonded coatings, manufacturing defects, electrochemical reaction, flow assisted corrosion and
ingression of corrosive species such as hydrogen into carbon steel materials.
Pipelines exposed to external and internal cyclic stress in sour environment have hydrogen diffused to the carbon steel
surface of the pipeline. This ingression which can happen at a pressure as low as 2 MPa can potentially cause distortion of the
microstructure of the carbon steel at defective metallic matrix in the microstructure hence the need for improved quality of
manufactured pipes. The Heat Affected Zones (HAZs) on pipelines create heterogeneous microstructures which give room for
sulphide stress cracking due to their increased susceptibility to hydrogen ingression as a result of modified metallic matrix.
C.I. Ossai et al. / Engineering Failure Analysis 53 (2015) 36–58 55

Because indirect electric arc welding produces narrow heat affected zone, it is preferable to Submerged Arc Welding (SAW)
and Electric Resistance Welding (ERW) that are predominantly used for pipeline installation in oil and gas industry. Carbon
steel materials are more susceptible to stress corrosion cracking at reduced potentials and fatigue failure on Corrosion
Resistance Alloys (CRAs) in service in sour environment can be reduced if the strength is not more than 690 MPa or 22 HRC.
Since pipeline corrosion may not be eliminated in its entirety in oil and gas transmission pipelines, it is imperative that
optimal performance of pipelines be achieved based on integrity assurance cycle. Hence, core activities that includes strate-
gic policy initiation, policy implementation, information analysis and reviews and implementation actions are vital for fati-
gue stress minimization. This integrity programme applies corrosion controls, monitoring and inspections strategies to
ensure that the fatigue failure is minimized by carrying out fitness-for-service testing at design, manufacture, installation
and operation. Furthermore, there should be enhanced manufacturing processes for pipes in order to reduce defects such
as cracks, dents, buckles, bulges and out-of-roundness which increase the stress load on corroded pipelines. Again, when
pipelines are produced to high quality – free from inclusions and other microstructural defects, the risk of hydrogen induced
cracking is reduced and the chances of fatigue stress failures minimized.
Finally, it is recommended that more research work need to be done on:

 Stress corrosion cracking measurement techniques of pipelines since the present techniques have not adequately pro-
vided enough help for in-situ prediction of stress corrosion cracking.
 The effect of manufacturing processes such as welding and hydrostatic testing on the susceptibility of pipelines to corro-
sion, stress corrosion cracking and hydrogen induced cracking need to be investigated more especially with respect to a
safe testing pressure that may not increase the risk of fatigue stress on pipelines.
 There is need for more experimental analysis on the correlation of pipeline burst pressure and toughness of ageing
pipelines.

References

[1] US DOT. Pipeline and Hazardous Materials Safety Administration <https://hip.phmsa.dot.gov/analyticsSOAP/saw.dll?Portalpages> [access 07.02.15].
[2] http://www.bbc.com/news/world-asia-28594693 [assessed 07.02.15].
[3] http://www.iab-atex.nl/publicaties/database/Ghislenghien%20Dossier.pdf [assessed 07.02.15].
[4] Singh M, Markeset T. Simultaneous handling of variability and uncertainty in probabilistic and possibilistic failure analysis of corroded pipes.
International Journal of System Assurance Engineering and Management 2014;5:43–54.
[5] Pilkey AK, Lamberf SB, Plumtree A. Stress corrosion cracking of X-60 line pipe steel in a carbonate-bicarbonate solution. Corros Sci 1995;51(2):91–6.
[6] Fekete G, Varga L. The effect of the width to length ratios of corrosion defects on the burst pressures of transmission pipelines. Eng Fail Anal
2012;21:21–30.
[7] Ossai CI. Advances in asset management techniques: an overview of corrosion mechanisms and mitigation strategies for oil and gas pipelines. ISRN
Corrosion, vol. 2012. http://dx.doi.org/10.5402/2012/570143, Article ID 570143.
[8] Meresht ES, Farahani TS, Neshati J. Failure analysis of stress corrosion cracking occurred in a gas transmission steel pipeline. Eng Fail Anal
2011;18:963–70.
[9] Hong HP. Inspection and maintenance planning of pipeline under external corrosion considering generation of new defects. Struct Saf
1999;21:203–22.
[10] Sinha SK, McKim RA. Probabilistic based integrated pipeline management system. Tunn Undergr Space Technol 2007;22:543–52.
[11] Pandey MD. Probabilistic models for condition assessment of oil and gas pipelines. NDT&E Int 1998;31(5):349–58.
[12] Breton T, Sanchez-Gheno JC, Alamillac JL, Alvarez-Ramirez J. Identification of failure type in corroded pipelines: a Bayesian probabilistic approach. J
Hazard Mater 2010;179:628–34.
[13] Cumber PS. Predicting outflow from high pressure vessels. Process Saf Environ Prot 2001;79:13–22.
[14] Wang N, Zarghamee M. Evaluating fitness-for-service of corroded metal pipelines: structural reliability bases. J Pipeline Syst Eng Pract
2014;5:04013012.
[15] Elliott P. Corrosion-when you see the signs, which way do you go. Mater Perform NACE Int 2013;52(6):30–4.
[16] Hays GF. Now is the time; 2010 <http://corrosion.org/wp-content/themes/twentyten/images/nowisthetime.pdf> [assessed 30.06.14].
[17] Miller R. Why pipes matter: the importance of clad pipes in the oil and gas industry; 2013 <http://breakingenergy.com/2013/03/15/why-pipes-
matter-the-importance-of-clad-pipe-in-the-oil-and-gas/> [assessed 01.07.14].
[18] Bhaskaran R, Palaniswamy N, Rengaswamy NS, Jayachandran M. Global cost of corrosion—a historical review. ASM Handbook, vol. 13B: Corrosion:
Materials. In: Cramer SD, Covino BS, Jr, editors; 2005, p. 621–8. http://dx.doi.org/10.1361/asmhba0003968.
[19] Yardeni E, Johnson D, Quintana M. Global economic briefing: global inflation; 2014 <http://www.yardeni.com/pub/infltrglb_bb.pdf> [assessed
01.07.14].
[20] Lieser MJ, Xu J. Composites and the future of society: preventing a legacy of costly corrosion with modern materials; 2010 <http://www.
ocvreinforcements.com/pdf/library/CSB_Corrosion_White_Paper_Sept_10_10_English.pdf> [assessed 01.07.14].
[21] CAPP. Best management practices: mitigation of internal corrosion in oil effluent pipeline systems, 2009. Calgary Canada; 2009 <http://www.capp.ca/
getdoc.aspx?DocId=155641&DT=PDF> [accessed 19.06.14].
[22] Review of corrosion management for offshore oil and gas processing. HSE offshore technology report 2001/044; 2001 <http://www.hse.gov.uk/
research/otopdf/2001/oto01044.pdf> [accessed 22.06.14].
[23] Alberta Energy regulator (AER). Report 2013 – B: pipeline performance in Alberta, 1990–2012; 2013 <http://www.aer.ca/documents/reports/R2013-
B.pdf> [accessed 17.09.13].
[24] Brown GK. Appendix – internal corrosion monitoring techniques for pipeline systems. In: Singh Ramesh, editor. Pipeline integrity
handbook. Boston: Gulf Professional Publishing; 2014. p. 255–92.
[25] Fang H, Brown B, Nescaronicacute S. Effects of sodium chloride concentration on mild steel corrosion in slightly sour environments. Corrosion
2011;67(1).
[26] Nesic S. Key issues related to modelling of internal corrosion of oil and gas pipelines – a review. Corros Sci 2007;49:4308–38.
[27] Singer M, Brown B, Camacho A, Nesic S. Combined effect of carbon dioxide hydrogen sulfide and acetic acid on bottom-of-the-line corrosion.
Corrosion 2011;67(1).
56 C.I. Ossai et al. / Engineering Failure Analysis 53 (2015) 36–58

[28] Li D, Zhang L, Yang J, Lu M, Ding J, Liu M. Effect of H2S concentration on the corrosion behavior of pipeline steel under the coexistence of H2S and CO2.
Int J Miner Metall Mater 2014;21:388–94.
[29] Yin ZF, Zhao WZ, Bai ZQ, Feng YR, Zhou WJ. Corrosion behaviour of SM 80SS tube steel in stimulant solution containing H2S and CO2. Electrochim Acta
2008;53(10):3690.
[30] Davoodi A, Pakshir M, Babaiee M, Ebrahimi GR. A comparative H2S corrosion study of 304L and 316L stainless steels in acidic media. Corros Sci
2011;53(1):399.
[31] Khaksarfard R, Paraschivoiu M, Zhu Z, Tajallipour N, Teevens PJ. CFD based analysis of multiphase flows in bends of large diameter pipelines.
CORROSION/2013; 2013.
[32] Nesic S, Cai J, Lee KJ. A multiphase flow and internal corrosion prediction model for mild steel pipelines. Corrosion/2005. Paper No 05556; 2005.
[33] Guo B, Song S, Ghalambor A. Lin TR. Chapter 10 – introduction to flexible pipelines. Offshore pipelines. 2nd ed. Boston: Gulf Professional Publishing;
2014. p. 125–32. Offshore Pipelines, 2nd ed. ISBN 9780123979490. http://dx.doi.org/10.1016/B978-0-12-397949-0.00010-8.
[34] Velazquez JC, Valor A, Caleyo F, Venegas V, Espina-Hernandez JH, Hallen JM, et al. Pitting corrosion models improve integrity management reliability.
Oil Gas J 2009;107(28):56–62. <www.scopus.com> [accessed 28.04.14].
[35] Caleyo F, González JL, Hallen JM. A study on the reliability assessment methodology for pipelines with active corrosion defects. Int J Press Vessels Pip
2002;79(1):77–86.
[36] Ahammed M. Probabilistic estimation of remaining life of a pipeline in the presence of active corrosion defects. Int J Press Vessels Pip
1998;75(4):321–9.
[37] Brazan FAV, Beck AT. Stochastic process corrosion growth models for pipeline reliability. Corros Sci 2013;74:50–8.
[38] Bedairi B, Cronin D, Hosseini A, Plumtree A. Failure prediction for crack-in-corrosion defects in natural gas transmission pipelines. Int J Press Vessels
Pip 2012;96–97:90–9.
[39] Hernandez-Rodriguez MAL, Martínez-Delgado D, González R, Pérez Unzueta A, Mercado-Solís RD, Rodríguez J. Corrosive wear failure analysis in a
natural gas pipeline. Wear 2007;263:567–71.
[40] Ilman MN, Kusmono. Analysis of internal corrosion in subsea oil pipeline. Case Stud Eng Fail Anal 2014;2:1–8.
[41] Biomorgi J, Hernández S, Marín J, Rodriguez E, Lara M, Viloria A. Internal corrosion studies in hydrocarbons production pipelines located at
venezuelan northeastern. Chem Eng Res Des 2012;90:1159–67.
[42] Papavinasam S, Doiron A, Revie RW. Model to predict internal pitting corrosion of oil and gas pipelines. Corrosion 2010;66(3):E1–E11.
[43] Mazumder QH, Shirazi SA, McLaury B. Experimental investigation of the location of maximum erosive wear damage in elbows. J Pressure Vessel
Technol 2008;130. 011303-03.
[44] Zhang T, Yang Y, Shao Y, Meng G, Wang F. A stochastic analysis of the effect of hydrostatic pressure on the pit corrosion of Fe–20Cr alloy. Electrochim
Acta 2009;54:3915–22.
[45] Papavinasam S, Doiron A, Panneerselvam T, Revie RW. Effect of hydrocarbons on the internal corrosion of oil and gas pipelines. Corrosion
2007;63(7):704–12. <http://search.proquest.com/docview/223119315> [accessed 12.12.13].
[46] Park NG, Morello L, Wong JE, Maksoud SA. The effect of oxygenated methanol on corrosion of carbon steel in sour wet gas environments. CORROSION/
2007. Paper no. 07663 NACE; 2007.
[47] Wang H, Jepson PW, Cai JY, Gopal M. Effect of bubbles on mass transfer in multiphase flow. In: NACE int. conference series, corrosion/2000, paper no.
00050; 2000.
[48] Cole IS, Corrigan P, Sim S, Birbilis N. Corrosion of pipelines used for CO2 transport in CCS: Is it a real problem? Int J Greenhouse Gas Control
2011;5:749–56.
[49] Choi H, Al-Ajwad H. Flow dependency of sweet and sour corrosion of L-80 carbon steel downhole production tubing in Khuff gas wells. Saudi Aramco
journal of technology, Spring; 2007.
[50] Ferris FG, Jack TR, Bramhill BJ. Corrosion products associated with attached bacteria at an oil field water injection plant. Can J Microbiol
1992;38:1320–4.
[51] Popoola LT, Grema AS, Latinwo GK, Gutti B, Balogun AS. Corrosion problems during oil and gas production and its mitigation. Int J Ind Chem
2013;4:35.
[52] Beidokhti B, Dolati A, Koukabi AH. Effects of alloying elements and microstructure on the susceptibility of the welded HSLA steel to hydrogen-induced
cracking and sulfide stress cracking. Mater Sci Eng A 2009;507:167–73.
[53] Martínez D, Gonzalez R, Montemayor K, Juarez-Hernandez A, Fajardo G, Hernandez-Rodriguez MAL. Amine type inhibitor effect on corrosion–erosion
wear in oil gas pipes. Wear 2009;267:255–8.
[54] Kapusta S. Managing corrosion in sour gas system: testing, design, implementation and field experience. CORROSION/2008, paper no. 08641. NACE;
2008.
[55] Song Y, Palencsár A, Svenningsen G, Kvarekvål J, Hemmingsen T. Effect of O2 and temperature on sour corrosion. Corrosion 2012;68(7):662–71.
[56] Muthukumar N, Rajasekar A, Ponmariappan S, Mohanan S, Maruthamuthu S, Muralidharan S, et al. Microbiologically influenced corrosion in
petroleum product pipelines – a review. Indian J Exp Biol 2003;41:1012–22.
[57] Chandrasatheesh C, Jayapriya J, George RP, Kamachi Mudali U. Detection and analysis of microbiologically influenced corrosion of 316 L stainless steel
with electrochemical noise technique. Eng Fail Anal 2014;42:133–42.
[58] Chen Y, Howdyshell R, Howdyshell S, Ju L. Characterizing pitting corrosion caused by a long-term starving sulfate-reducing bacterium surviving on
carbon steel and effects of surface roughness. Corrosion 2014;70(8):767–80.
[59] Melchers RE. Extreme value statistics and long-term marine pitting corrosion of steel. Probabilist Eng Mech 2008;23(4):482–8. http://dx.doi.org/
10.1016/j.probengmech.2007.09.003. ISSN 0266-8920.
[60] Sana NO, Nazırb H, Donmezc G. Microbiologically influenced corrosion failure analysis of nickel–copper alloy coatings by Aeromonas salmonicida and
Delftia acidovorans bacterium isolated from pipe system. Eng Fail Anal 2012;25:63–70.
[61] Machuca LL, Murray L, Gubner R, Bailey SI. Evaluation of the effects of seawater ingress into 316L lined pipes on corrosion performance. Mater Corros
2014;65:8–17.
[62] Raman RKS, Javaherdashti R, Panter C, Pereloma EV. Hydrogen embrittlement of a low carbon steel during slow strain testing in chloride solutions
containing sulphate reducing bacteria. Mater Sci Technol 2005;21(9):1094–8.
[63] Javaherdashti R. Impact of sulphate-reducing bacteria on the performance of engineering materials. Appl Microbiol Biotechnol 2011;91(6):507–1517.
<http://search.proquest.com>.
[64] Wen J, Gu T, Nesic S. Investigation of the effects of fluid flow on SRB biofilm. In: NACE Int. conference series, corrosion/2007, paper no. 07516; 2007.
[65] Maruthamuthu S, Muthukumar N, Natesan M, Palaniswamy N. Role of air microbes on atmospheric corrosion. Corros Sci 2008;94(3):359–63.
[66] Gieg LM, Jack TR, Foght JM. Biological souring and mitigation in oil reservoirs. Appl Microbiol Biotechnol 2011;92(2):263–82. http://dx.doi.org/
10.1007/s00253-011-3542-6.
[67] Lee H, Choi J, Lee B, Kim T. Failure analysis of stress corrosion cracking in aircraft bolts. Eng Fail Anal 2007;14:209–17.
[68] Gamboa E, Linton V, Law M. Fatigue of stress corrosion cracks in X65 pipeline steels. Int J Fatigue 2008;30:850–60.
[69] Contreras A, Albiter A, Salazar M, Pérez R. Slow strain rate corrosion and fracture characteristics of X-52 and X-70 pipeline steels. Mater Sci Eng, A
2005;407:45–52.
[70] Ahmed TM, Lambert SB, Sutherby R, Plumtree A. Cyclic crack growth rates of X-60 pipeline steel in a neutral dilute solution. Corrosion (NACE)
1997;53(7):581–90.
[71] Eliaz N, Shachar A, Tal B, Eliezer D. Characteristics of hydrogen embrittlement. Stress Corros Crack Tempered Martensite Embrittlement High-
Strength Steels Eng Fail Anal 2002;9:167–84.
C.I. Ossai et al. / Engineering Failure Analysis 53 (2015) 36–58 57

[72] Fazzini PG, Otegui JL. Experimental determination of stress corrosion crack rates and service lives in a buried ERW pipeline. Int J Press Vessels Pip
2007;84:739–48.
[73] Jiang X-C, Staehle RW. Effects of stress and temperature on stress corrosion cracking of austenitic stainless steels in concentrated magnesium chloride
solutions. Corrosion 1997;53(6):448–66.
[74] Alizadeh AH, Khishvand M, Ioannidis MA, Piri M. Multi-scale experimental study of carbonated water injection: an effective process for mobilization
and recovery of trapped oil. Fuel 2014;132:219–35.
[75] Azevedo CRF. Failure analysis of a crude oil pipeline. Eng Fail Anal 2007;14:978–94.
[76] Waanders FB, Vorster SW. Corrosion products formed on mild steel samples submerged in various aqueous solution. Hyperfine Interact
2002;139:239–44.
[77] Cerny I, Linhart V. An evaluation of the resistance of pipeline steels to initiation and early growth of stress corrosion cracks. Eng Fract Mech
2004;71:913–21.
[78] Kumar MS, Sujata M, Venkataswamy MA, Bhaumik SK. Failure analysis of a stainless steel pipeline. Eng Fail Anal 2008;15:497–504.
[79] Zhang C, Zhang Q, Chen W, Liu D, Ou T, Huang X, et al. Assessment of corrosion and mechanical properties degradation of pipeline steel after long
term service in high sour gas-field environments. In: International Petroleum Technology Conference (IPTC); 2013.
[80] Niu L, Cheng YF. Corrosion behaviour of X-70 pipe steel in near-neutral pH solution. Appl Surf Sci 2007;253:8626–31.
[81] Kentish P. Stress corrosion cracking of gas pipelines – effect of surface roughness orientations and flattening. Corros Sci 2007;49:2521–33.
[82] Sh Abedi S, Abdolmaleki A, Adibi N. Failure analysis of SCC and SRB induced cracking of a transmission oil products pipeline. Eng Fail Anal
2007;14:250–61.
[83] Majid ZA, Mohsin R, Yaacob Z, Hassan Z. Failure analysis of natural gas pipes. Eng Fail Anal 2010;17:818–37.
[84] Azevedo CRF, Sinatora A. Failure analysis of a gas pipeline. Eng Fail Anal 2004;11:387–400.
[85] Hasan F, Iqbal J, Ahmed F. Stress corrosion failure of high-pressure gas pipeline. Eng Fail Anal 2007;14:801–9.
[86] Ahmed F, Ali L, Iqbal J, Hasan F. Failure of pipe joints during hydrostatic testing. Eng Fail Anal 2008;15:766–73.
[87] Hasan F, Iqbal J. Consequential rupture of gas pipeline. Eng Fail Anal 2006;13:127–35.
[88] Capelle J, Gilgert J, Dmytrakh I, Pluvinage G. Sensitivity of pipelines with steel API X52 to hydrogen embrittlement. Int J Hydrogen Energy
2008;33:7630–41.
[89] Gang L, Kaxiras E. Hydrogen embrittlement of aluminum: the crucial role of vacancies. Phys Rev Lett 2005;94(15):1–5.
[90] Lynch S. Hydrogen embrittlement phenomena and mechanisms. Corros Rev 2012;30(3–4):105–23.
[91] Woodtli J, Rolf K. Damage due to hydrogen embrittlement and stress corrosion cracking. Eng Fail Anal 2000;7(6):427–50.
[92] Natividad C, Salazar M, Espinosa-Medina MA, Pérez R. A comparative study of the SSC resistance of a novel welding process IEA with SAW and MIG.
Mater Charact 2007;58:786–93.
[93] Omweg GM, Frankel GS, Bruce WA, Ramirez JE, Koch G. Performance of welded high-strength low-alloy steels in sour environments. Corrosion
2003;59(7):640–53.
[94] Cayard MS, Kane RD. Large-scale wet hydrogen sulphide cracking performance. evaluation of metallurgical, mechanical, and welding variables.
Corrosion 1997;53:227–30.
[95] Park GT, Koh SU, Jung HG, Kim KY. Effect of microstructure on the hydrogen trapping efficiency and hydrogen induced cracking of linepipe steel.
Corros Sci 2008;50:1865–71.
[96] Gonzalez JL, Morales A. Analysis of laminations in X52 steel pipes by nonlinear by finite element. J Pressure Vessel Technol 2008;130. 021706-06.
[97] Zacaria MYB, Davies TT. Formation and analysis of stack cracks in pipeline steel. J Mater Sci 1993;28:322–8.
[98] Wang J, Li X. A phenomenological model for fatigue crack growth rate of X70 pipeline steel in H2S corrosive environment. J Pressure Vessel Technol
2014;136. 041703-03.
[99] Liang P, Li X, Du C, Chen X. Stress corrosion cracking of X80 pipeline steel in simulated alkaline soil solution. Mater Des 2009;30:1712–7.
[100] Sedriks AJ. Stress corrosion cracking test. Methods 1989;1:1.
[101] Papavinasam S. Chapter 6 – modelling – internal corrosion’. In: Papavinasam Sankara, editor. Corrosion control in the oil and gas
industry. Boston: Gulf Professional Publishing; 2014. p. 301–60.
[102] Cosham A, Hopkins P, Macdonald KA. Best practice for the assessment of defects in pipelines – corrosion. Eng Fail Anal 2007;14:1245–65.
[103] Besel M, Zimmermann S, Kalwa C, Koppe T, Liessem A. Corrosion assessment method validation for high-grade line pipe. In: Proceedings of the 8th
international pipeline conference IPC2010, September 27–October 1, 2010, Calgary, Alberta, Canada, paper no-IPC2010-31664; 2010.
[104] Chiodo MSG, Ruggieri C. Failure assessments of corroded pipelines with axial defects using stress-based criteria: numerical studies and verification
analyses. Int J Press Vessels Pip 2009;86:164–76.
[105] Xu LY, Cheng YF. Reliability and failure pressure prediction of various grades of pipeline steel in the presence of corrosion defects and pre-strain. Int J
Press Vessels Pip 2012;89:75–84.
[106] Det Norske Veritas (DNV). Recommended Practice – RP-F101 Corroded Pipelines; 2010.
[107] ASME. Manual for determining the remaining strength of corroded pipelines. American Society of Mechanical Engineers. B31G, New York; 2009.
[108] Klever FJ, Steward G. New developments in burst strength prediction for locally corroded pipes. Shell Int Research; 1995.
[109] Leis N, Stephens DR. In: Proceedings of the 7th international offshore and polar engineering conference, Part I, Honolulu, USA, ASME International,
New York; 1997. p. 624–41.
[110] Wang W, Smith MQ, Popelar CH, Maple JA. A new rupture prediction model for corroded pipelines under combined loading. In: Proceedings of second
international pipeline conference (IPC 1998), Calgary, Alberta, Canada, American Society of Mechanical Engineers, 8–11 June 1998.
[111] Adib-Ramezani H, Jeong J, Pluvinage G. Structural Integrity evaluation of X52 gas pipes subjected to external corrosion defects using the SINTAP
procedure. Int J Press Vessels Pip 2006;83:420–32.
[112] Owen R, Chauhan V, Pipet D, Morgan G. Assessment of corrosion damage in pipelines with low toughness. In: 14th Biennial joint technical meeting on
pipeline research, Berlin; 2003.
[113] Choi JB, Goo BK, Kim JC, Kim YJ, Kim WS. Development of limit load solutions for corroded gas pipelines. Int J Press Vessels Pip 2003;80(2):121–8.
[114] Cravero S, Ruggieri C. Structural integrity analysis of axially cracked pipelines using conventional and constraint-modified failure assessment
diagrams. Int J Press Vessels Pip 2006;83:607–17.
[115] Qian X, Li Y, Ou Z. Ductile tearing assessment of high-strength steel X-joints under in-plane bending. Eng Fail Anal 2013;28:176–91.
[116] YongBo S, Tjhen LS. Parametric equation of stress intensity factor for tubular K-joint under balanced axial loads. Int J Fatigue 2005;27:666–79.
[117] Adib R, Schmitt C, Pluvinage G. Application of volumetric method to the assessment of damage induced by action of foreign object on gas pipes.
Strength Mater 2006;38:409–16.
[118] Meliani M Hadj, Matvienko YG, Pluvinage G. Corrosion defect assessment on pipes using limit analysis and notch fracture mechanics. Eng Fail Anal
2011;18:271–83.
[119] Wang X, Reinhardt W. On the assessment of through-wall circumferential cracks in steam generator tubes with tube supports. J Pressure Vessel
Technol 2003;125:85–90.
[120] Structural integrity assessment procedure (SINTAP). Final report E-U project BE95-1462. Brite Euram Programme Brussels; 1999 <http://www.
eurofitnet.org/sintap_Procedure_version_1a.pdf>.
[121] Zhu X, Leis BN. Theoretical and numerical predictions of burst pressure of pipelines. J Pressure Vessel Technol 2007;129:644–52.
[122] Ma B, Shuai J, Liu D, Xu K. Assessment on failure pressure of high strength pipeline with corrosion defects. Eng Fail Anal 2013;32:209–19.
[123] Zhu X, Leis BN. Average shear stress yield criterion and its application to plastic collapse analysis of pipelines. Int J Press Vessels Pip 2006;83:663–71.
[124] Hasan S, Khan F, Kenny S. Probability assessment of burst limit state due to internal corrosion. Int J Press Vessels Pip 2012;89:48–58.
58 C.I. Ossai et al. / Engineering Failure Analysis 53 (2015) 36–58

[125] Teixeira AP, Guedes Soares C, Netto TA, Estefen SF. Reliability of pipelines with corrosion defects. Int J Press Vessels Pip 2008;85:228–37.
[126] Caleyo F, Velázquez JC, Valor A, Hallen JM. Markov chain modelling of pitting corrosion in underground pipelines. Corros Sci 2009;51:2197–207.
[127] Bolanos-Rodriguez E, González Islas JC, Vega-Cano GY, Flores-García E. Modeling based on markov chains for the evolution of pitting corrosion in
buried pipelines carrying gas. Pap Presented ECS Trans 2011;35(17):37–42. <www.scopus.com> [accessed 28.04.14].
[128] Nicholson R, Feblowitz J, Madden C, Bigliani R. The role of predictive analytics in asset optimization for the oil and gas industry. IDC Energy Insight
2010. <http://tessella.com/documents/role predictive-analytics-asset-optimization-oil-gas-industry>.
[129] Kishawy HA, Gabbar HA. Review of pipeline integrity management practices. Int J Press Vessels Pip 2010;87:373–80.
[130] Papavinasam S. Chapter 7 – mitigation – internal corrosion. In: Papavinasam Sankara, editor. Corrosion control in the oil and gas
industry. Boston: Gulf Professional Publishing; 2014. p. 361–424.
[131] American Petroleum Institute (API). Fitness-for-service. API recommended practice 579, first ed.; 2000.
[132] British Standards Institution (BSI). Guide on methods for assessing the acceptability of flaws in fusion welded structures BS 7910. London (UK); 2005.
[133] ASME B31.8S. Managing system integrity of gas pipelines. New York (USA): American Society of Mechanical Engineers; 2004.
[134] Han ZY, Weng WG. An integrated quantitative risk analysis method for natural gas pipeline network. J Loss Prev Process Ind 2010;23:428–36.
[135] Jo Y, Crowl DA. Individual risk analysis of high-pressure natural gas pipelines. J Loss Prev Process Ind 2008;21:589–95.
[136] Venkatesh CD, Farinha PA. Corrosion Risk Assessment (CRA) in the oil and gas industry – an overview and its holistic approach; 2006 <http://
members.iinet.net.au/~extrin/ConferencePapers/CAP04Paper06.pdf> [accessed 18.10.13].
[137] Sahraoui Y, Khelif R, Chateauneuf A. Maintenance planning under imperfect inspections of corroded pipelines. Int J Press Vessels Pip 2013;104:76–82.
[138] Race JM. 4.42 – Management of corrosion of onshore pipelines. In: Tony JA Richardson, editor(s)-in-chief, Shreir’s corrosion. Oxford: Elsevier; 2010.
p. 3270–306. ISBN 9780444527875. http://dx.doi.org/10.1016/B978-044452787-5.00169-4.
[139] Ossai CI. Predictive modelling of wellhead corrosion due to operating conditions: a field data approach. ISRN Corrosion, vol. 2012, Article ID 237025;
2012. http://dx.doi.org/10.5402/2012/237025.
[140] Meng W, Ogea P, King M. Changes of operating procedures and chemical application of a mature deepwater tie-back – aspen field case study. Offshore
Technol Conf Proce 2011;2:881–92.
[141] Geary W, Hobbs J. Catastrophic failure of a carbon steel storage tank due to internal corrosion. Case Stud Eng Fail Anal 2013;1:257–64.
[142] Fazzini PG, Otegui JL. Influence of old rectangular repair patches on the burst pressure of a gas pipeline. Int J Press Vessels Pip 2006;83:27–34.
[143] Blin F, Koutsoukos P, Klepetsianis P, Forsyth M. The corrosion inhibition mechanism of new rare earth cinnamate compounds — electrochemical
studies. Electrochim Acta 2007;52:6212–20.
[144] Hu X, Barker R, Neville A, Gnanavelu A. Case study on erosion–corrosion degradation of pipework located on an offshore oil and gas facility. Wear
2011;271:1295–301.
[145] Huang Y, Ji D. Experimental study on seawater-pipeline internal corrosion monitoring system. Sens Actuat B: Chem 2008;135:375–80.
[146] Vega OE, Hallen JM, Villagomez A, Contreras A. Effect of multiple repairs in girth welds of pipelines on the mechanical properties. Mater Charact
2008;59:1498–507.
[147] Nayak C. Fault detection in fluid flowing pipes using acoustic method. Int J Appl Eng Res 2014;9(1):23–8.

View publication stats

You might also like