Вы находитесь на странице: 1из 11

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

net/publication/317312441

FATIGUE CRACK GROWTH RATE ANALYSIS IN AN API 5L X70 STEEL PIPE.

Conference Paper · September 2015

CITATIONS READS

0 433

2 authors:

Bruno Antonio Sorrija Marcelino Pereira do Nascimento


University of São Paulo São Paulo State University
5 PUBLICATIONS   6 CITATIONS    33 PUBLICATIONS   305 CITATIONS   

SEE PROFILE SEE PROFILE

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

Surface Treatments to Improve Structural Integrity of Aeronautic Components View project

Structural Integrity of Linepipes and Pipelines View project

All content following this page was uploaded by Bruno Antonio Sorrija on 02 June 2017.

The user has requested enhancement of the downloaded file.


IBP1239_15
FATIGUE CRACK GROWTH RATE ANALYSIS IN AN
API 5L X70 STEEL PIPE.
Bruno A. Sorrija , Marcelino P. Nascimento 2
1

Copyright 2015, Brazilian Petroleum, Gas and Biofuels Institute - IBP


This Technical Paper was prepared for presentation at the Rio Pipeline Conference & Exposition 2015, held between September, 22-
24, 2015, in Rio de Janeiro. This Technical Paper was selected for presentation by the Technical Committee of the event. The
material as it is presented, does not necessarily represent Brazilian Petroleum, Gas and Biofuels Institute’ opinion or that of its
Members or Representatives. Authors consent to the publication of this Technical Paper in the Rio Pipeline Conference &
Exposition 2015.

Abstract
In this work a study of the fatigue crack growth rates in specimens taken from 90o and 180o of a welded joint of an API
5L X70 steel pipe, which is used in the manufacture of oil and gas pipelines, has been made. This class of steel show
high strength and ductility values, and has been increasingly studied due the growing demand of oil and natural gas,
which in consequence, increases the needing of new pipelines to transport them. The material studied has been directly
taken from a tube provided by Tenaris Confab, according to the standards API 5L and ASTM E647, for the tensile tests
and fatigue crack growth respectively. The results of the tensile tests obtained are in accordance with the specifications
of the API 5L. The results of the fatigue crack growth rate tests showed significant differences between regions situated
90º and 180º of the welded joint region, which can be potentially related to the steel pipe manufacturing process.
Metallographic and microhardness analysis of the base material and welded joint complemented this work.

1. Introduction
Fatigue is defined by ASTM (2013a) as “the process of progressive localized permanent structural change
occurring in a material subjected to conditions which produce fluctuating stresses and strains at some point or points
and which may culminate in cracks or complete fracture after a sufficient number of fluctuations” and today it is
recognized as the most common form of failure of structural components (Callister, 2008; Abdullah et al., 2012). Due
cyclic loadings during transportation (transit fatigue) and during service life, pipelines used in the oil and gas
exploration are no exception, as they suffer from significant fatigue damage, which if not enough to lead to a fracture,
may severely affect the structural integrity of these entities.
Pipes may suffer fatigue damage before they enter in service due a mechanism called transit fatigue. Several
studies in the 60’s have been done by pipes manufacturers to investigate why so many pipes were being rejected in
hydrostatic testing prior to their assembling in the field (Godoy, 2008). Many different tests have been done, but only
after fatigue tests researchers could find cracks in the specimens that were very similar to the ones that had been found
in the hydrostatic test rejected pipes. In fact, during transportation pipes are subjected cyclic stresses, related to inertial
and gravitational forces (Godoy, 2008; Mohammed et al., 2009), which are responsible for the nucleation and growth of
fatigue cracks inside them, that compromise their structural integrity.
Taking in consideration offshore line pipes, they may also suffer from fatigue damage during their service life.
This happens because these structures are subject to cyclic loading originated from: cyclic pressure and thermal
expansion loads (Bai and Bai, 2005), waves movement induced loads and specially VIV – Vortex Induced Vibration
(Albuquerque, 2004; Riva, 2004; Bai and Bai, 2005; Lima, 2007; Valadão, 2011). Fluid flowing around a bluff-body
may lead to an unstable flowing region characterized by a Kármán Vortex Street (Bai and Bai, 2005; Gabbi, 2013). The
VIV is related to this phenomenon because line pipes like risers, flowlines and pipelines have a bluff-body profile,
therefore, when ocean currents flows around them, vortex shedding at a frequency f may happen, like shown in Figure
1.

______________________________
1
Mechanical Engineer and Postgraduate Student – Department of Materials and Technology - State
University of São Paulo – UNESP/FEG/DMT
2
Ph.D. – Department of Materials and Technology - State University of São Paulo –
UNESP/FEG/DMT
Rio Pipeline Conference & Exposition 2015

Figure 1. Ocean current flow around a pipe resulting into vortex shedding at a frequency f (Gabbi, 2013, modified).

When this vortex shedding happens at or near of a natural frequency of line pipe, undesired vibration of the structure
will happen (Ribeiro, 2002; Bai and Bai, 2005). These vibrations induced by the vortex shedding then will act as a
dynamic loading, leading to high frequency cyclic stresses. Especially when the ocean current flows speeds are high,
the phenomenon can result in severe fatigue damage, greatly reducing the pipes service life (Albuquerque, 2004; Bai
and Bai, 2005; Lima, 2007). VIV must be taken into consideration in the line pipe project, and while acceptable if
within certain limits, it may lead to some serious fatigue failure, being responsible, for example, for 14 failures between
1965 and 1976 on subsea pipelines in Alaska and for two failures in the Ping Hu Pipeline (East China Sea) in 2000
(Sollund and Velded, 2012; Sollund et al., 2014). It must be noticed though that VIV only may happen if there is fluid
flowing around the pipe, therefore, while risers are always susceptible to it, for pipelines and flowlines the phenomenon
will only happen if there are free spans (portions of pipe that are not in contact with the seabed) along the structure,
regions where there is a possibility of fluid flowing around the pipe (Valadão, 2011; Sollund and Velded, 2012; Sollund
et al., 2014).
Considering large diameter pipes (like the one that had been tested in this research), the “UOE” is the most
effective pipe manufacturing process, (Herynk et al., 2007; Kostryzhev et al., 2008; Ren et al., 2015). It consists
basically in three steps: first, cold forming of a steel plate first in an U-shape and after that, in a circular “O-shape”,
second, union of the remaining edges by welding (most commonly through SAW - Submerged Arc Welding process)
and finally, the third step is the a mechanical expansion of the already welded pipe, to guaranty its final form and
circularity (Herynk et al., 2007; Kostryzhev et al., 2008; Toscano, 2009; Ren et al., 2015).
Despite the high efficiency of the UOE process, the pipes obtained from it are far from having an isotropic
nature. First, the welding process alone introduces a discontinuity in the pipe, due inhomogeneous properties between
the base metal, the heat affected zone and the fusion zone (addition metal). Second, the cold forming can also lead to
some significant changes in the mechanical properties of the manufactured pipe. While there might be some hardening
due the cold working, some strength decrease may also happen due the Bauschinger effect (Kostryzhev et al., 2008;
Ogata, 2009; Toscano, 2009), which reduces specially the steel resistance to compressive loading (compressive yield
strength) (Toscano, 2009; Louis, 2010), a problematic issue for deep water pipelines because it turns them more
susceptible to collapse failure, especially when combined with undesirable shape imperfections like eccentricity and
ovality (Toscano, 2009; Louis, 2010). Finally cold work also leads to a wide distribution of residual stress along the
circumferential dimension of the pipe (Ceglias, 2012), which plays an important role when we consider fatigue
resistance of pipelines, because while residual compressive stresses are benefic and will restrain fatigue cracks growth,
tensile residual stresses have the opposite effect (Ceglias, 2012).
There are different ways to study fatigue damage in line pipes. Godoy (2008) for example developed a study of
transit fatigue using full scale specimens (steel pipe rings), and noticed that fatigue cracks may nucleate and grow
during pipe transportation, without occurrence of fracture. Other researchers, like Beltrão et al. (2011), focused their
work in the study of fatigue crack growth rates in the welded joint of steel pipes.
The aim of this paper is investigate how the fatigue growth rates of cracks vary along the circumferential
direction of the pipe, in regions that are not directly affected by the submerged arc welding process. For this, specimens
had been taken from angular positions situated 90º and 180o from the welded joint and the study of the fatigue cracks
growth rates has been done through the analysis of da/dN x ∆K curves, focusing on their Region II (refer to Figure 2)
where the cyclic growth of the fatigue crack is linear (on a log-log plot) and can be predicted by the Paris-Erdogan Law:


=  ∆ (1)


where da/dN is the fatigue crack growth rate, C and m are scaling constants and ∆K is the stress intensity factor range.
The Paris-Erdogan Law only considers the ∆K increasing for the fatigue crack growth rate evaluation, ignoring the
effects of the load ratio (Suresh, 1998; Branco et al., 1999; Dowling, 2007) and load frequency (Suresh, 1998; Dowling,
2007), which also may impact the values of C and m. Therefore, for a given material it is only adequate to evaluate
crack growth rates when the environmental conditions, temperature as well the load ratio and frequency (all variables
that may influence C and m) are fixed (Suresh, 1998). In the present study, specimens have been tested with the same
2
Rio Pipeline Conference & Exposition 2015

loading and environmental conditions, therefore differences


d in the C and m constants for the 90º and 180º positions may
indicate thatt the cold forming and mechanical expansion of the UOE process impact in the fatigue properties
pro of the
pipe, notably its fatigue crack growth rates.
rate In complement to the fatigue crack growth rate study, chemical analysis,
tensile tests of the pipe and an analysis of the welded joint through microhardness and metallography have been made.

Figure 2. Typical fatigue crack growth curve for metals showing its Regions I, II and III (Anderson, 1995).

2. Experimental Procedures
2.1. Steel Pipe Rings
In the present work a HRLA (High Resistance Low Alloy) microalloyed steel, used in the manufacturing of
pipes used for line pipe applications like risers, flowlines and pipelines, has been used. It was provided by its
manufacturer in the form of rings obtained
ined from an API 5L class, X70 grade, PSL 2 steel pipe. The steel sheets were
produced in Brazil by USIMINAS S/A through Thermo-Mechanical Control Process (TMCP) without accelerated
cooling. The pipe had an outer diameter of 700 mm, 15.9 mm thickness and was as manufactured in Brazil by Tenaris
Confab,, through the UOE process, being longitudinally welded by Submerged Arc Welding (SAW). The pipes have
been divided in four angular regions of interest: 0º (center of the welded joint, which has been used as the reference
re
point), 90º, 180º and 270º (which is equivalent to the 90º position),
position) as presented in Figure 3.After
.After this division, the pipes
have being initially sectioned by Oxy-Fuel
Fuel Cutting process and then cut into smaller pieces
pieces with a hydraulic band saw.
These smaller pieces had been machined then into the specimens that have ave been used in this research in the mechanical
tests, metallography and chemical analysis.

Figure 3. Steel pipe and its four regions of interest (not drawn to scale).

2.2 Chemical Analysis


To evaluate if the chemical composition of the steel pipe used in this work was in accordance to the API 5L
X70 PSL 2 requirements, an analysis by optical emission spectroscopy has been made using a Spectro Maxx 2007 arc
spark optical emission spectrometer. The procedures of the analysis were in accordance to the ASTM E1806: Standard
Practice for Sampling Steel and Iron for Determination of Chemical Composition,
Composition and the samples for the chemical
analysis were taken from
rom both 90º and 180º positions.

2.3 Tensile Tests


In order too evaluate the mechanical properties of the steel pipe, tensile tests
test have been performed. As required
by the API 5L: Specification for Line Pipe standard, transverse tensile specimens had been taken from the 180º18 position
of the pipe. In complement, specimens also had been taken from the 90º 9 (transverse and longitudinal direction) position

3
Rio Pipeline Conference & Exposition 2015

of the pipe, although, these tensile tests are not required by the referred API standard, and were used just for mechanical
properties comparison between the 180º and 90º positions. Dimensions of the specimens and testing conditions were all
in accordance to the API 5L and ASTM A370: Standard Test Methods and Definitions for Mechanical Testing of Steel
Products standards. The tensile tests were carried out in a 100 kN capacity, Instron 8801 servo hydraulic testing
machine, at room temperature with a low displacement rate of 1mm/min. Figure 4 shows the layout of the tensile test
specimens that have been used. There was no need of flattening for the longitudinal tensile test specimens, but it was
needed for the transverse tensile test specimens.

Figure 4. Tensile test round specimen layout, in accordance to the standards API 5L and ASTM E370.

2.4 Microhardness Test and Metallographic analysis


To better understand the characteristics of the welded joint, it has been characterized through microstructural
analysis and Vickers microhardness. For both analysis of the welded joint, specimens were manually polished using
abrasive papers with grit values ranging from 100 to 1200 mesh, after this a subsequent final mechanical polishing by
colloidal silica has been made. For the metallographic specimens after the polishing specimens were etched by 2% Nital
reagent during 3 to 5 seconds, with all the procedures being in accordance with the standard ASTM E3: Standard Guide
for Preparation of Metallographic Specimens. Concerning the Vickers microhardness specimen, after polishing it was
etched by 10% Nital reagent during 15 seconds, an attack which was sufficient to clearly reveal the boundaries of
welded joint. A total of 15 indentations have been made along the medium diameter line of the steel pipe, on its welded
joint, to measure the HV values of the BM (Base Metal), HAZ (Heat Affected Zone) and WM (Weld Metal). The
measurements have been made with a Buehler Wilson 401 MVD Hardness Tester, with a 500 g load and 10s
application interval.

2.5 Fatigue Crack Growth Test

The material for the fatigue crack growth rate specimens had been taken from the medium diameter of the pipe
ring samples, in the angular positions 90º and 180º (using the same angular references of the previous sections) from the
welded joint, and they were machined into Compact-type (CT) specimens. There was no need of flattening of the steel
pipe samples before the machining of the CT specimens. Figure 5 (a) shows the CT specimens layout, while Figure 5
(b) represents from which part of the samples the specimens have been taken from.

Figure 5. (a) Compact-type (CT) specimen layout and (b) Scheme showing from which points of the samples
the CT specimens have been taken from.
4
Rio Pipeline Conference & Exposition 2015

After machining, the specimens were subjected to fatigue precracking (until a 1.5 mm length fatigue crack),
which has been carried out in a 100 kN capacity, Instron 8801 servo hydraulic testing machine in the dependences of
the Fatigue and Aeronautical Materials Research Group, in the Department of Materials and Technology of the São
Paulo State University, Campus of Guaratinguetá (FEG/UNESP). After precracking, the fatigue crack growth tests have
been carried out, in a 250 kN capacity MTS 810 servo hydraulic testing machine, in the Department of Materials of the
Engineering School of Lorena (EEL), University of São Paulo (USP). The fatigue precracking parameters are shown in
Table 1 while the fatigue crack growth test parameters and the number of tested specimens are presented in Table 2.

Table 1. Fatigue precracking parameters.

Max. Load Load Ratio Frequency Wave form


8 kN 0.1 10 Hz Sinusoidal
Fatigue Precrack length: 1.5 mm

Table 2. Fatigue crack growth test parameters and number of tested specimens.

Max. Load Load Ratio Frequency Wave form


9 kN 0.1 10 Hz Sinusoidal
90º Position 180º Position
Number of specimens:
3 3

All test parameters and specimens were in accordance with the standard ASTM E647: Standard Test Method
for Measurement of Fatigue Crack Growth Rates.

3. Results and Discussion

3.1 Chemical Analysis


The required chemical composition of the X70 grade steel, as defined by the API 5L standard, as well the
chemical composition determined by optical emission spectroscopy of the X70 steel that have been used in the present
work, are shown in Table 3. The values of the chemical composition of the Experimental 5L X70 presented in Table 3
are an average value of the samples taken from the 90° and 180° positions, and are within the limits required by the
referred standard.

Table 3. Chemical composition of the X70 grade, PSL 2 steel pipe with wall thickness ≤ 25 mm, as required by API and
average measured values.

Chemical element mass fraction (Max %) CE (Max.)


Steel Grade C Si Mn P S V+Nb+Ti Pcm
API 5L Standard X70 0.12 0.45 1.70 0.025 0.015 0.15 0.25
Experimental 5L X70 0.10 0.16 1.34 0.008 0.0013 0.1173 0.1984

3.2 Tensile Tests


The results of the tensile tests are shown in Table 4. Considering the 90º position, the results obtained indicate
that the values of both yield and ultimate strength are slightly higher for longitudinal direction of the pipe when it is
compared to the transverse position. This may be related to the hot rolling process of the steel plates that are used in the
manufacturing of the pipes, which can lead to some degree of anisotropy and explain the higher values of strength along
the longitudinal direction of the plates and consequently of the pipes. For the transverse specimen taken from the 180º
position the obtained results of yield strength (API 5L req.: 435 < σY < 635 MPa) and ultimate strength (API 5L req.:
570 < σU < 760 MPa) as well the ratio σY/σU (API 5L req.: ratio < 0.93) are all in accordance with the API 5L standard
requirements for a X70 grade PSL 2 steel pipe. Comparing both positions, the values of mechanical strength are higher
for the 180º when they are compared to the ones observed in the 90º position. Comparing both positions only for the
transverse specimens, the yield strength and the ultimate strength are respectively 10.7% and 8.8% higher for the 180º
position. The results obtained suggest that the UOE process effectively impact on the mechanical properties of the steel
plate along the circumferential direction of the pipe.

5
Rio Pipeline Conference & Exposition 2015

Table 4. Tensile tests results.

σY – Yield Strength σU – Ultimate Ductility


Specimen σY/σU
[MPa] Strength [MPa] [%]
Transversal Position 90º
Average Value 499.35 555.58 30 0.90
Standard Deviation 3.01 7.78 1 0.01
Longitudinal Position 90º
Average Value 507.79 568.39 31 0.89
Standard Deviation 7.93 9.54 2 0.01
Transverse Position 180º
Average Value 552.75 604.62 29 0.91
Standard Deviation 7.23 13.22 4 0.01

3.3 Metallographic Analysis


Figure 6 presents the base-material, HAZ and weld metal microstructures for the API 5L X70 grade PSL 2 steel pipe.

Figure 6. Welded joint microestructure as observed through optical microscopy and Nital 2% etching: (a) Base Metal
(BM), (b) Heat Affected Zone (HAZ), notably the coarse grains region, (c) interface between the HAZ and the Weld
Metal (WM) and (d) WM (500x).

According to Figure 6 (a), it can be noticed that the BM is essentially composed by a ferritic/pearlitic microstructure as
expected for this HSLA microalloyed steel, with the presence of Primary Polygonal Ferrite (αF) and Pearlite (P)
Moving right, to the center of the welded joint, the HAZ and its coarse grains sub region is presented (b), with larger
grains of PF (Primary Ferrite), when compared to the BM, as well the presence of Intragranular Primary Ferrite PF(I)
and Grain Boundary Primary Ferrite PF(G). At (c) there is a transition between the HAZ coarse grains sub region and
the WM, with the presence of aligned and Ferrite with Second Phase Aligned and Not Aligned, FS(A) and FS(NA)
respectively. Finally in (d), the WM fusion zone is shown, with the presence of PF(I), FS(NA) and specially Acicular
Ferrite (AF), which has a fine microstructure and is responsible for the high values of strength and toughness of this low
alloy microalloyed WM.

3.4 Vickers Microhardness Test


The results of the Vickers microhardness test, along the medium diameter of the steel ring on the welded joint
regions of interest (Base Metal, Heat Affected Zone and Weld Metal) are presented in Figure 7 through a microhardness
profile. In complement, Table 5 presents the average values of the Vickers microhardness for each region of the welded
joint. According to the microhardness test, the BM has the lowest values of hardness when compared to the HAZ and
WM, which is a partially unexpected result. While the highest value of hardness of the WM was expected and can be
related to the presence of Acicular Ferrite in its microstructure (which is responsible for the higher value of mechanical
strength), that is not the case for the HAZ. This region is a portion of the BM that is subjected to changes in its
microestructure due the welding process heat input and is composed by sub regions. Some sub regions inside the HAZ
6
Rio Pipeline Conference & Exposition 2015

have poor mechanical properties, like the Coarse Grains and the Intercritical sub regions, therefore, it was expected that,
in average, the microhardness value of the HAZ would be lower than the one measured in the BM. Although, this result
may be related to the reverse cold deformation of the UOE process, which can lead to some strength decrease in the BM
due the occurrence of the Bauschinger Effect (Kostryzhev et al, 2007).

Figure 7. Vickers microhardness profile of the welded joint, measured along the medium diameter of the steel pipe.

Table 5. Results of Vickers microhardness along the welded joint.

Vickers Microhardness (HV 0.5)


Region Average Value Standard Deviation
BM 169.5 7.1
HAZ 190.2 9.9
WM 224.0 5.8

3.5 Fatigue Crack Growth Tests


Some sources had compiled experimental values of the empirical constants C and m for various types of steels,
but these results are not absolute. For example, the standard BS7910, suggests mean values of m equal to 2.88 for
assessing Fatigue Crack Growth Rate (FCGR) in ferritic, austenitic and duplex ferritic-austenitic steels (with yield
strength ≤ 700 MPa), both for load ratios of R < 0.5 and R ≥ 0.5, while Barsom and Rolfe (1999) suggests m values of
3.00 and 3.25 for ferritic-perlitic and austenitic steels respectively (for R≈0). According to Suresh (1998) and Branco et
al. (1999) changes in the test conditions (related both to load and environment) can lead to changes in the values of
these constants, and this seems to be confirmed by the results obtained by Beltrão et al. (2011). In their work FCGR
tests have been carried out in three positions of an API 5L X70 steel pipe welded joint: WM, HAZ and BM, with values
of R equal to 0.1 and 0.5. According to their results, it could be noticed a significant difference between the values of C
and m for the tests carried out with R = 0.1 and R = 0.5. Therefore, taking this is into consideration, it is only fruitful to
compare Region II results of FCGR tests, using the Paris-Erdogan Law, when the conditions of two or more different
tests are strictly the same, which is the case for this research. The results of the FCGR tests are presented in Figure 8
and Table 6. From the results of the FCGR tests, it can be noticed that there is a significant difference in the values of C
and m (Equation 1 coefficients) when the results that have been obtained from the tests of the 90º and 180º pipe
positions are compared.

Table 6. Results of the Fatigue Crack Grow Rate Test, C and m coefficients according to Equation 1.

C [mm.Cycle-1/
Specimen (MPa√m)m] m
90° - 1 2.05E-07 2.11
90° - 2 2.53E-07 1.98
90° - 3 1.11E-07 2.19
Average Value 1.90E-07 2.09
180° - 1 5.75E-08 2.38
180° - 2 1.56E-08 2.65
180° - 3 8.82E-08 2.25
Average Value 5.38E-08 2.43

7
Rio Pipeline Conference & Exposition 2015

Figure 8. (a) Crack Length [mm]


mm] versus Number of Cycles es [N] for all tested specimens and (b) Results of the
Fatigue Crack Growth test for all specimens in Region II (refer to Figure 1).

As presented, the
he values of C are one order of magnitude higher for the 90º position while the values of m are,
in average, 16.3% higher for the 180º position.posit Being the testing and environmental conditions as well the
microstructure (both regions belong to the BM, thus they are expected to have the same ferrite-pearlite
ferrite microstructure)
the same for the 90º and 180º specimens, it would be expected, in average, the same values of C and m for both
positions. Considering Region II, the crack growth rate is not a strong function
function of the microstructure or monotonic flow
properties (Anderson, 1995; Suresh, 1998),
1998), there might be other factors, apart from test conditions, that also
al affect the
values of C and m,, like the fact for example that the values obtained in the tensile tests show different mechanical
strength values for the 90º and 180º position. Another point worth of discussing is that the he increase of any of these
coefficients increase the growth rate of a fatigue crack and because of this, while one region is more critical to the value
of C the other is more critical to the value of m. Nevertheless, due the power law nature of Equation 1, small changes on
the value of m have a big impact on the growth rate, rate becoming significant as the values of ∆K increase. Figure 9
comparess the crack grow rates for both positions taking in consideration a ∆K K range varying from 40 (around the
inferior limit of the Region II for 180º position) to 65 MPa.m0.5.

Figure 9.. Comparison between the average values of fatigue crack growth rate within a ∆K range contained
within the Region II for both 90º and 180º angular positions.

It can be noticed that from values of ∆K higher than 40 MPa√m the crack growth rate in the 180º will be
always higher than the one observed in the 90º position, despite its
its one order of magnitude higher value of C. Although,
it must be noticed that for the 90º position, the Region II (which presents the stable crack growth) is achieved at lower
values of ∆K,K, therefore while a fatigue crack will grow faster in the 180º
180º position, it will start growing early in the 90º
9
position.
It must be noticed also that the only expected differences between the samples taken from the 90º and 180º
positions are the strain levels that these are subjected
subjecte during the UOE cold forming and possibly the presence of
different values of residual stresses resultant from the manufacturing process. These factors might be responsible for the
8
Rio Pipeline Conference & Exposition 2015

differences in the values of C and m. Considering only the residual stresses, in fact, the influence of their presence at the
edge of a crack may be very significant, and can even allow fatigue crack growth during a pure compressive loading, if
these residual stresses are sufficient to promote the opening of the cracks surface (Branco, 1999).

4. Conclusions
Pipes that are used into the assemblage line pipes are frequently subjected to the cyclic loading, both during
their transportation and service life, which can lead to fatigue damage and severely affect their structural integrity of the
pipe. Apart from this, there must be taken into consideration that pipes manufactured through the UOE process are far
from being isotropic, which can lead to some differences in the values of strength along their circumferential direction.
Based on the results obtained, the following conclusion may be drawn: i.) The tensile tests results indicate that the UOE
process cold forming may lead to a circumferential mechanical strength anisotropy of the steel pipe; ii.) The analysis of
the welded joint through metallography has confirmed a ferritic-pearlitc microstructure, as expected for a HSLA
microalloyed steel. The average values of microhardness in the welded joint have shown that WM has the highest
values of hardness (which can be related to the presence of AF). Microhardness values in the HAZ were higher than the
ones observed in the BM. This can be potentially related to the reverse cold deformation during the UOE cold forming
of the steel pipes, which can lead to a strength decrease due the occurrence of the Bauschinger Effect; iii.) The fatigue
crack growth rate tests, analyzed with the use of the Paris-Erdogan Paris law, have shown that there are significant
differences between the coefficients C and m when comparing samples taken from a 90º and 180º angular positions of
the steel pipe. While a fatigue crack will start to grow at lower values of ∆K for the 90º position, an analysis of the
growth rates in a ∆K range within the Region II for both cases, shows that the fatigue crack growth rate will be always
higher for the 180º position in these conditions, despite the fact that this region apparently presents higher values of
mechanical strength (according to the tensile test results); iv.) While most literature agrees that Region II of the crack
growth rate is affected specially by testing and environmental conditions (given a certain material), the obtained results
shows that there are other factors that also may be important. The authors propose that these factors (which are present
in the material even after the machining of the specimens) are related to the pipe anisotropy along its circumferential
direction and may be related to the presence of residual stress or/and strength increase (due cold work hardening) or
decrease (due Bauschinger Effect), despite the fact that literature usually considers that Region II is not high sensitive to
the mechanical strength.

5. Acknowledgements
The authors are grateful to Professor Carlos Antonio Reis Pereira Baptista and the Mechanical Tests Technician
Francisco de Paiva Reis from the Department of Materials of the Engineering School of Lorena (EEL), University of
São Paulo (USP), for the Fatigue Crack Growth Rate tests, and are grateful also to Tenaris Confab for providing the
steel pipes rings that had been used in this study.

6. References

ABDULLAH, S., ARIFFIN, A. K., BEDEN, S. M. Fatigue crack growth rate of API X70 pipelines under spectrum
loading. International Journal of Pressure Vessels and Piping 96-97, p. 9-12, 2012.
ALBUQUERQUE, M. C. S.; Comportamento à fadiga de juntas soldadas de tubulações marítimas tratadas pela técnica
TIG Dressing. Ph.D. Thesis in Process Engineering (in Portuguese), Federal University of Campina Grande, UFGC,
Brazil, 2004.
AMERICAN PETROLEUM INSTITUTE. API 5L: Specification for Line Pipe. 44th ed, Washington D.C: API
Publishing Services, October, 2008.
AMERICAN SOCIETY FOR TESTING AND MATERIALS. ASTM E18123: Standard Terminology Relating to
Fatigue and Fracture Testing. ASTM International, 2013a.
AMERICAN SOCIETY FOR TESTING AND MATERIALS. ASTM E647: Standard Test Method for Measurement of
Fatigue Crack Growth Rates. ASTM International, 2013b.
AMERICAN SOCIETY FOR TESTING AND MATERIALS. ASTM E370: Standard Test Methods and Definitions for
Mechanical Testing of Steel Products. ASTM International. 2010.
AMERICAN SOCIETY FOR TESTING AND MATERIALS. ASTM E3: Standard Guide for Preparation of
Metallographic Specimens. ASTM International, 2011.
AMERICAN SOCIETY FOR TESTING AND MATERIALS. ASTM E1806: Standard Practice for Sampling Steel and
Iron for Determination of Chemical Composition. ASTM International, 2009.
9
Rio Pipeline Conference & Exposition 2015

ANDERSON, T .L. Fracture Mechanics – Fundamentals and Applications. 2nd ed., CRC Press, 1995.
BAI, Y., BAI, Q. Subsea pipelines and risers, 1st ed., Elsevier, 2005.
BARSOM, J. M., S. T. ROLFE. Fracture and Fatigue Control in Structures, 3rd ed., ASTM
International, 1999.
BELTRÃO, M. A. N., CASTRODEZA, E. M., BASTIAN, F. L. Fatigue crack propagation in API5L X-70 pipeline
steel welded joints under constant and variable amplitudes. Fatigue & Fracture of Engineering Materials &
Structures, V. 34, Issue 5, p. 321-328, 2011.
BRANCO, M. C., FERNANDES, A. A., CASTRO, P. M. S. Fadiga de Estruturas Soldadas (in portuguese). 2nd ed.,
Fundação Calouste Gulbenkian, 1999.
BRITISH STANDARD. BS 7910: Guide to methods for assessing the acceptability of flaws in metallic structures.BSI,
2013.
CALLISTER, W. D. Materials Science and Engineering: an Introduction (in Portuguese), 7th ed., LTC, 2008.
CEGLIAS, R. B. Análise de tensão residual em um tubo de aço API 5L X70. Master of Material Science Thesis (in
Portuguese), Military Engineering Institute, IME, Brazil, 110p., 2012.
DOWLING, N. E. Mechanical Behavior of Materials – Engineering Methods for Deformation Fracture and Fatigue. 3rd
ed., Pearson/Prentice Hall, 2007.
GABBI, R. Modelagem matemática do escoamento turbulento em canal axissimétrico com bluff body. Master’s Thesis
in Mathematical Modeling (in Portuguese), Regional University of Northwestern Rio Grande do Sul,
UNIJUÍ/DCEEng, 92 p., 2013.
GODOY, J. M. Study of transit fatigue crack in welded steel pipes. Master’s Thesis in Mechanical Engineering (in
Portuguese), State University of São Paulo, UNESP/FEG, Brazil, 132 p., 2008.
HERYNK, M. D., KYRIAKIDES, S., ONOUFIROU, A., YUN, H. D. Effects of the UOE/UOC pipe manufacturing
processes on pipe collapse pressure. International Journal of Mechanical Sciences 49, p. 533-553, 2007.
KOSTRYZHEV, A. G., STRANGWOOD, M., DAVIS, C. L. Influence of microalloying precipitates on Bauschinger
effect during UOE forming of line pipe steels. Materials Technology, v. 22, n. 3, p. 166-172, 2008.
LIMA, A. J. Analysis of slender offshore structures due to vortex induced vibrations Master of Science Thesis (in
Portuguese), Federal University of Rio de Janeiro, UFRJ/COPPE, 120 p., 2007.
LOUIS, J. G. Collapse of thick walled UOE pipes corrosive resistant cladding. Master of Science Thesis, University of
Oslo, UiO/Faculty of Mathematics and Natural Sciences, 92p., 2010.
MARQUES, P. V., MODENESI, P. J., BRACARENSE, A. Q. Soldagem: Fundamentos e Tecnologia (in Portuguese).
3rd ed., UFMG Press,2013.
MOHAMMED, A. A., RAO, S., LOBLEY, G. R. Pipeline failure by transitt Fatigue. Journal of Failure Analysis and
Prevention, Volume 9, Issue 1, p. 35-38, 2009.
OGATA, P. H. Caracterização do aço para tubo API 5L-X65 em diferentes regiões da chapa como laminada e após a
austenitização e resfriamento sob diversas taxas de resfriamento. Master’s Thesis in Engineering (in Portuguese),
University of São Paulo, USP/POLI, Brazil, 105 p., 2009.
REN, Q., ZOU, T., LI, D., TANG, D., PENG, Y., Numerical study on the X80 UOE pipe forming process. Journal of
Materials Processing Technology 215, p. 264-267, 2015.
RIBEIRO, P. A. R. Desprendimento de vórtices e controle em esteira de cilindros por simulação numérica direta.
Master’s Thesis in Water Supply and Environmental Sanitation (in Portuguese), University of Rio Grande do Sul,
UFRGS/IPH, Brazil, 94p., 2002.
RIVA, I. R. Análise de fadiga de estruturas metálicas com ênfase em offshore. Abstract of Undergraduate Project as a
partial fulfillment of the requirements for the degree of Civil Engineer (in Portuguese), Federal University of Rio de
Janeiro, UFRJ/COPPE, Brazil, 160 p., 2004.
SOLLUND, H. A., VEDELD, K., A semi-analytical model for free vibrations of free spanning offshore pipelines.
Research Report in Mechanics, n. 2, University of Oslo, UiO, 65p., 2012.
SOLLUND, H. A., VEDELD, K., HELLESLAND, J. Dynamic response of multi-span offshore pipelines. Marine
Structures 39, p. 174-197, 2014.
SURESH, S. Fatigue of Materials. 2nd ed., Cambridge University Press, 1998.
TOSCANO, R. G. Collapse and post-collapse behavior of steel pipes under external pressure and bending. Application
to deep water pipelines. Ph.D Thesis, University of Buenos Aires, UBA, 214p., 2009.
VALADÃO, A. M. F. Fatigue analysis of topside offshore structures – A case study. Abstract of Undergraduate Project
as a partial fulfillment of the requirements for the degree of Engineer (in Portuguese), Federal University of Rio de
Janeiro, UFRJ/POLI, Brazil, 73 p., 2011.

10

View publication stats

Вам также может понравиться