0% found this document useful (0 votes)
10 views19 pages

Understanding The Effect of Galvanic Corrosion On The Sulfide Stress Corrosion Cracking of X80 - Inconel 625 Weld Overlay

This study investigates the failure of the monocable head (MH-22) adapter in a Veracruz well due to exposure to high levels of hydrogen sulfide (H2S) and carbon dioxide (CO2), leading to sulfide stress cracking (SSC) and material embrittlement. Comprehensive analysis, including chemical composition and mechanical testing, revealed that the material did not meet H2S resistance standards, exacerbated by excessive torque during tightening. Recommendations for preventing similar failures include using more suitable materials, such as nickel-based alloys, to enhance safety and reliability in harsh environments.

Uploaded by

Yesenia
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
10 views19 pages

Understanding The Effect of Galvanic Corrosion On The Sulfide Stress Corrosion Cracking of X80 - Inconel 625 Weld Overlay

This study investigates the failure of the monocable head (MH-22) adapter in a Veracruz well due to exposure to high levels of hydrogen sulfide (H2S) and carbon dioxide (CO2), leading to sulfide stress cracking (SSC) and material embrittlement. Comprehensive analysis, including chemical composition and mechanical testing, revealed that the material did not meet H2S resistance standards, exacerbated by excessive torque during tightening. Recommendations for preventing similar failures include using more suitable materials, such as nickel-based alloys, to enhance safety and reliability in harsh environments.

Uploaded by

Yesenia
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 19

Engineering Failure Analysis 171 (2025) 109421

Contents lists available at ScienceDirect

Engineering Failure Analysis


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

Failure analysis of the monocable head (MH-22) adapter exposed


to high H2S and CO2 environments
A. Albiter , Lucila Cruz-Castro , A. Contreras *
Instituto Mexicano del Petróleo, Eje Central Lázaro Cárdenas Norte 152, San Bartolo Atepehuacan, C.P. 07730 CDMX, Mexico

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

Keywords: This study aims to investigate the failure occurred in the main adapter of the monocable head
Failure analysis (MH-22) during operation of a well in Veracruz, Mexico. The cause of the failure was investigated
Sulfide stress cracking (SSC) through a comprehensive analysis, including information on the operating conditions of the well,
Hydrogen sulfide (H2S)
chemical composition, hardness testing, Scanning Electron Microscopy (SEM), and Energy-
Monocable head
Intergranular cracks
Dispersive Spectroscopy (EDS). The pressure and temperature conditions of the well were 4
MPa and 30 ◦ C at the wellhead, 13.65 MPa and 109 ◦ C at the bottom of the well. Fluid
composition revealed 6.05 % mol H2S and 4.8 % mol CO2 in the liquid phase, increasing to 9.89 %
mol H2S and 7.85 % mol CO2 in the gas phase. The results confirmed that the material suffered
sulfide stress cracking (SSC) due to the high H2S content. Excessive torque and fatigue when
tightening the MH-22 component, in addition to hardness of 39 HRC, contributing to SSC sus­
ceptibility of the material. The analysis of the physical conditions of the component, including
deformation, fracture type, and cracks, revealed embrittlement due to H2S exposure, which
caused the fracture under tensile loads. Based on the findings, the material did not meet the
required standards for H2S resistance, and the failure could have been prevented with the use of
more suitable materials, such as nickel-based alloys.

1. Introduction

Failure analysis in oil wells is crucial for identifying the causes of equipment malfunctions and structural failures in the oil and gas
industry [1–3]. The challenging environments of drilling, which involve high pressures, extreme temperatures, and corrosive gases like
hydrogen sulfide (H2S) and carbon dioxide (CO2), can lead to significant mechanical and structural issues [4,5]. Failure analysis helps
to maintain equipment integrity, avoid operational interruptions, and improve safety and efficiency in drilling activities. By utilizing
visual inspections, laboratory tests, and operational data, engineers can develop strategies for mitigation and offer recommendations
for equipment design and maintenance. Understanding failure mechanisms such as corrosion, stress cracking, and material embrit­
tlement is essential for extending the lifespan of critical components and reducing downtime [6].
Common failure mechanisms in oil and gas infrastructure are diverse and complex, with numerous case studies highlighting issues
that lead to leaks and ruptures. These studies provide essential insights into the root causes of failures, which is key to reducing risks
and ensuring the long-term reliability of downhole tools and pipelines. Metallurgical failure mechanisms, including manufacturing
defects, third-party damage, and corrosion, are particularly significant [7,8].
Li et al. [9] studied the failure of a 101.6 mm drill pipe and found that solids settling and sticking caused friction-induced heat,

* Corresponding author.
E-mail address: [email protected] (A. Contreras).

https://doi.org/10.1016/j.engfailanal.2025.109421
Received 9 December 2024; Received in revised form 29 January 2025; Accepted 13 February 2025
Available online 14 February 2025
1350-6307/© 2025 Elsevier Ltd. All rights are reserved, including those for text and data mining, AI training, and similar technologies.
A. Albiter et al. Engineering Failure Analysis 171 (2025) 109421

Nomenclature

AA Pipe for general service


AH-38 Adapter head-38
BB Pipe for general service (low CO2)
BL Bottom liner
BPD Barrels per day
C2H6 Ethane
C3H8 Propane
CC Pipe for general service (low CO2)
CCL Casing collar locator
CH4 Methane
CO2 Carbon dioxide
DD Sour service
EDS Energy-dispersive spectroscopy
EE Sour service (low CO2)
EUE External upset end
FF Sour service (medium CO2)
HE Hydrogen embrittlement
H2S Hydrogen sulfide
HH Sour service (high CO2)
HRC Hardness Rockwell C
HSC Hydrogen stress cracking
Inc. Inclination
KOP Kick-off point
MH-21 Monocable head-21
MH-22 Monocable head-22
MMCFD Millions of cubic feet per day
MVA Master valve assembly
N-80 Grade N-80 pipe (80 ksi)
Ni Nickel
OM Optical microscopy
ppCO2 Partial pressure of carbon dioxide
ppH2S Partial pressure of hydrogen sulfide
ppm Parts per million
PT Pressure–temperature
Pwf Bottomhole pressure
QFE Quick fit equipment
S Sulfur
SEM Scanning electron microscopy
SSC Sulfide stress cracking
SCC Stress corrosion cracking
TOC Top of cement
TP Tubing production
TRC Trade reference code
TRP Tubing retrievable packer
TVD True vertical depth
VTOP Premium thread connection used in production tubing for oil wells

ultimately leading to fracture. Saithala et al. [10] examined tubing hanger failures in high-pressure, high-temperature wells, attrib­
uting the failures to incorrect heat treatment and detrimental metallurgical phases. Shen et al. [11] and Shi et al. [12] also explored
fracture failures in drill pipes and casings, identifying factors such as heat treatment uniformity and large oxide metallurgical in­
clusions as contributing to failure. These studies emphasize the need for quality control and material optimization to improve the
integrity of drilling equipment.
Studies on corrosion in pipelines, like those by Mubarak et al. [13] and Wu et al. [14], highlight the importance of failure analysis
and advanced techniques for identifying corrosion mechanisms and implementing preventive measures. Liu et al. [15] conducted
corrosion fatigue crack propagation tests on S135 low carbon high-strength drill pipe steel under different strain rates in a simulated
H2S-containing drilling environment. The results showed that the fatigue crack propagation rate curve of S135 steel exhibited stress
corrosion and corrosion fatigue characteristics, with distinct zones: near threshold, platform, and rapid propagation. H2S significantly

2
A. Albiter et al. Engineering Failure Analysis 171 (2025) 109421

accelerated the corrosion rate, making it approximately ten times faster than in air.
Yu et al. [1] focused on frequent drill pipe failures, identifying overload ductile fractures caused by corrosion pits on the inner wall.
Their study suggested that addressing material performance, stress, and environmental factors could help prevent similar failures.
Jun Jing et al. [16] examine the impact of pressure, temperature, and fluid composition on the failure of offshore oil and gas well
strings due to corrosion. Their analysis highlights how increased temperature reduces the yield strength of pipe materials, while
elevated pressure and temperature can cause changes in pipe wall thickness due to thermal expansion and pressure effects. These
changes subsequently influence the residual collapsing, internal pressure, and tensile strength of the pipe. Furthermore, higher gas
production and increased oil–water content generally enhance the pipe’s residual strength. Zeng et al. [5] examined gas well failures in
the Northwest Oilfield, attributing failures to water-accumulation corrosion. They recommended using P110ss steel and nickel-based
alloys to prevent such issues.
Albiter et al. [17] examined a brittle failure caused by excessive torque and fatigue, emphasizing the need for proactive mainte­
nance to detect and mitigate failure risks. In a second study, Albiter et al. [18] analyzed coiled tubing failure due to external corrosion,
highlighting material degradation, corrosion pits, and mechanical damage as key contributors. Albiter [19] analyzed the behavior of
X52 steel under severe corrosion conditions induced by H2S and CO2, showing that while the steel did not exhibit cracking under
certain conditions, subsurface cracks were detected in environments with high concentrations of H2S and CO2 at 150 ◦ C, providing
insights on using carbon steel pipelines in extreme oil and gas conditions.
This study analyzed the failure of the MH-22 main adapter exposed to high H2S and CO2 concentrations, resulting in hydrogen
sulfide stress cracking (SSC) and material embrittlement. By combining visual inspection, operational data review, and laboratory
analysis, it identifies the root causes of the rupture. Through the results of this failure analysis, it is visualized solutions to improve the
safety and reliability of downhole equipment in similar environments, focusing on prevention and material selection.

2. Experimental procedure

The analysis process for this case of failure involved several sequential steps. First, a visual inspection of the MH-22 component was
performed, followed by an analysis of the fluid composition in the well. Subsequently, chemical analysis of the MH-22 component was
conducted using a portable energy-dispersive X-ray fluorescence analyzer to determine the grade, type, compliance with standards,

Fig. 1. Flowchart showing the process used to perform the root cause failure analysis.

3
A. Albiter et al. Engineering Failure Analysis 171 (2025) 109421

and any deviations contributing to the failure. A hardness test was conducted using a hardness testing device. The fracture surface
examination was performed using optical microscopy (OM) and Scanning Electron Microscopy (SEM) coupled with Energy-dispersive
Spectroscopy (EDS). Finally, an analysis of the pre-selection of the material for systems with CO2 + H2S was conducted using a
methodology based on ANSI/NACE MR 0175 [20] and API Spec 6A [21] standards, considering the operating conditions.
Fig. 1 presents a flowchart of the process used to conduct the root cause failure analysis. This investigation focused on failure
incidents during well operations, utilizing a combination of failure analysis, chemical and mechanical tests, as well as observations
through OM, SEM, and EDS. Additionally, material preselection was performed based on normative analysis.

2.1. Mechanical conditions of the well

The mechanical conditions of the well are shown in Fig. 2. It includes the wellhead Master Valve Assembly (MVA) of 13–5/8″, 3–1/
16″ and 2–1/16″ at 10 m, which controls the fluid flow from the surface; the surface casing of 30″ at 40 m, which provides initial
stability; the intermediate casing of 20″ at 700 m, which prevents aquifer contamination; the kick-off point (KOP) at 1,750 m with an
inclination of 18.27◦ ; the production casing of 13–3/8″ at 2,100 m and 9–5/8″ at 3,204 m, which supports production and protects
productive zones; the top of cement (TOC) at 2,604 m and 3,478 m, which secures the casing; additional bottom liners (BL) of 9–5/8″ at
1,942 m and 7″ at 3,062 m; production tubing of various sizes and materials (3–1/2″ and 4–1/2″ in TRC-95 and P-110) to transport
fluids; a packed tubing retrievable packer (TRJ 10 K Packed) at 3359 m to seal and ensure the well’s integrity; and final tubing
production of 2-7/8″ N-80 External upset end (EUE). The drilling fluid was filtered water with density of 1.00 g/cm3 to maintain
pressure and stabilize the well walls. This general schematic of the well design highlights the depths of the casings and tubing and
critical points such as the TOC and packers, which are crucial for the well’s integrity and efficient production. VTOP is a premium
thread connection used in production tubing for oil wells.

2.2. Well-fluid composition

The gas composition was analyzed using an Agilent 7890B gas chromatograph following the ASTM D1945 standard [22]. The fluid
composition is presented in Table 1. In the liquid, they are 6.05 mol% H2S and 4.8 mol% CO2. In the gas phase, these values increase to
9.89 mol% H2S and 7.85 mol% CO2. According to data provided on the date of the failure, the well was producing around 8 bpd of
water, which represents 28 % of production. And an oil production of 20 BPD and 0.1 MMCFD of gas.

Fig. 2. Mechanical condition of the well indicating the zone where quick fit equipment (QFE) with the MH-22 and AH-38 adapter was working.

4
A. Albiter et al. Engineering Failure Analysis 171 (2025) 109421

2.3. Quick fit equipment MH-22 adapter

The quick fit equipment (QFE) is a weight bar designed to assist in lowering calibration tools, shooting strings, and completion logs
from diameters of 1-3/8″ to 2-1/2″. QFE tools are composed of a robust frame with an insulated conductor capable of carrying the
electric current signal to the devices connected below them, and they are compatible with a string perforating system capable of
absorbing the impact of the shot. They are used in cased wells that contain liquid or air.
One of the adapters to the QFE tool is the MH-22. The MH-22 main adapter connects to the drilling equipment and serves as the
initial coupling point, and the AH-38 adapter attaches to the main adapter and the drilling or sampling tube, ensuring a firm
connection. The principal components are the pinion, which facilitates the mechanical connection between the adapters; the threads
and pin, which ensure that the adapters remain firmly attached during operation; and the drilling tube, which contains the extracted
core sample. The quick coupling function allows for rapid connection and disconnection between the different components of the
drilling system, increasing operational efficiency. The transfer of movement and torque ensures that the energy and motion generated
in the drilling equipment are efficiently transferred to the sampling tube. Fig. 3 shows a schematic representation of the MH-22 and
AH-38 components, as well as a cross-sectional view of the internal parts of the component.

3. Results and discussion

3.1. Background of failure

During pressure and temperature recording operations in the well, when the quick fit equipment tool was brought to the surface of
the well, a failure was detected in the MH-22 component. The QFE was working up to a depth of 3800 m. Real-time pressure and
temperature (P-T) shown in Table 2 presents the data collected during the operation (one data block at the start of the well and another
data block at the bottom of the well). The P-T data indicates that the data recording was always continuous as seen in Table 2. This is
because the logging cable never broke and maintained its continuity in the pressure and temperature records until the tool was pulled
out of the well.
A head pressure of 4.0 MPa, a bottom pressure of 13.65 MPa, and temperatures of 30.4 ◦ C at the head and 109 ◦ C at the bottom were
registered. These values were used as reference points in the analysis of the MH-22 component behavior and in the preselection
material analysis. In addition to the high concentrations of H2S and CO2 (9.89 mol% H2S and 7.85 mol% CO2 respectively).
After the QFE was recovered to the surface of the well, a bar gauge with CCL and a 2.300″ paraffin cutter calibrator were used. At
1700 m, a signal attenuation was detected, prompting a reduction in speed and checks of operating parameters, but with no
improvement. The tool was stopped 100 m before reaching the surface, and roles were assigned to the staff for receive the tool. The bar
gauge was retrieved, and pressures were reduced to 0 MPa. H2S and CO2 levels were monitored, and lubricators were disconnected.

3.2. Visual inspection

Fig. 4 shows recovered components for inspection. Damage to the pinion of the MH-22 adapter was found, leading to its
replacement with MH-21 (Fig. 4b). As shown in Fig. 4c, the AH-38 and MH-22 components were connected by a monocable, and a cut
was made to inspect the fracture surface and separate the two pieces. As can be observed in Fig. 4c, some scratches are visible on the
surface of MH-22 component; these scratches are generated during the operations of tightening and loosening the MH-22 and AH-38
components with the Stilson keys.
A cut was made to the fluid-air connector of the MH-22 component, as shown in Fig. 5, using a Milwaukee® M18-FUEL angle
grinder with a Milwaukee® H-11266 thin cutting disc, taking care not to damage the fractured surfaces of the MH-22 components.
According to the scratches and the type of fracture observed in the components, components suffer excessive torque and fatigue
when tightening the MH-22 and AH-38 adapter.
It was also determined that a section of the MH-22 thread is internally in the AH-38 component, so it is also necessary to carefully
extract the fractured surface. For this purpose, machining was performed using a conventional TOS 40C lathe, wearing down the AH-
38 component to extract the piece of the MH-22 that was internally in the AH-38, taking care not to damage the MH-22 component.
After obtaining both fractured pieces of MH-22 (Fig. 6), an attempt was made to join the fractured pieces to analyze their orien­
tation and fracture behavior. The fracture surface needed to be cleaned for observation under OM and SEM.

Table 1
Chemical composition of well fluid.
Component Liquid Gas

Wt.% Mol % Wt.% Mol %

CO2 1.76 4.80 12.37 7.85


H2S 1.72 6.05 12.08 9.89
N2 0.75 3.23 5.29 5.28
CH4 4.74 35.47 33.33 58.01
C2H6 0.79 3.18 5.59 5.19
C3H8 1.04 2.85 6.85 4.34

5
A. Albiter et al. Engineering Failure Analysis 171 (2025) 109421

Fig. 3. MH-22 and AH-38 components adapted for quick-fit equipment, a) general view, b) cross-sectional view.

Table 2
Real-time pressure and temperature (P-T) data collected during the operation.
TVD (m) Depth (m) Pressure (psi) Pressure (MPa) Temperature (◦ F) Temperature (◦ C) Pressure gradient (MPa/m)

0 0 576.93 3.98 96.41 35.78 0


100 100 582.78 4.02 107.89 42.16 0.0004
300 300 593.03 4.09 123.48 50.82 0.0004
600 600 607.98 4.19 138.97 59.43 0.0003
900 900 623.15 4.30 157.52 69.72 0.0004
1200 1200 637.21 4.39 168.07 75.59 0.0003
1360 1360 645.04 4.45 170.44 76.91 0.0003
1410 1410 648.32 4.47 172.03 77.79 0.0005
1460 1460 672.84 4.64 172.11 77.84 0.0034
1500 1500 695.43 4.79 173.54 78.63 0.0039
…. …. …. ​ …. ….. ​
3463.63 3540 1659.8 11.44 220.22 104.57 0.0000
3492.16 3570 1695.48 11.69 221.37 105.2 0.0086
3511.17 3590 1721.97 11.87 222.46 105.81 0.0096
3530.18 3610 1748.43 12.06 223.32 106.29 0.0096
3549.19 3630 1775.18 12.24 224.1 106.72 0.0097
3568.22 3650 1801.79 12.42 223.99 106.66 0.0096
3577.74 3660 1815.02 12.51 224.24 106.8 0.0096
3615.8 3700 1868.65 12.88 225.38 107.43 0.0097
3696.89 3785 1986.18 13.69 227.71 108.73 0.0100

*TVD: True vertical depth.

3.3. Material identification of the MH-22 component

To identify the type of MH-22 material, a portable energy-dispersive X-ray fluorescence (XRF) analyzer (TITAN Model LE 600,
Bruker) equipped with a 2 W rhodium anode was used. This device has a library of 900 material types, allowing it to determine and
compare the detected material with those already established in its database according to ASTM, ASME, DIN, and ISO standards. For
identification, it was necessary to remove surface contamination and provide a larger analysis area for using the TITAN.
The identified material was a Ni-Hard 1 alloy. Ni-Hard is a family of white cast irons alloyed with nickel and chromium, providing
high hardness and outstanding resistance to abrasion [23]. Table 3 shows the chemical composition analysis of the MH-22 specimen,
highlighting the main elements and their concentrations compared to the ASTM A532 standard [24] for Ni-Hard 1 alloy. This table
presents percentages of key elements such as iron (Fe), nickel (Ni), and chromium (Cr), which are the main components of this alloy
known for its high wear and abrasion resistance. These results confirm that the material of the MH-22 specimen corresponds to Ni-Hard
1, making it suitable for high-abrasion environments.
High-chromium alloys, with significant Cr-rich M3C2 carbides, exhibit high hardness but increased brittleness. In contrast, Ni-Hard
1, with minimal chromium, lacks these carbides, resulting in lower hardness but enhanced toughness and ductility. While Ni-Hard 1
offers better impact resistance, it sacrifices the extreme hardness seen in chromium-rich alloys [25].

6
A. Albiter et al. Engineering Failure Analysis 171 (2025) 109421

Fig. 4. Failed component, a) visual inspection of AH-38 and MH-22, b) MH-22 component connecting the AH-38 adapter, c) image of the MH-22
damaged component.

Fig. 5. Image of the cut of the monocable to separate the components and analyze the fracture surface of the MH-22 component.

Ni-Hard 1 has a microstructure consisting of carbides and a martensitic-austenitic-bainitic or predominantly martensitic matrix.
The microstructure of Ni-Hard 1 consists of primary dendrites and the carbide eutectic; the eutectic is a mixture of M3C carbide plates
and matrix [26]. After solidification the austenite is transformed to martensite, some bainite, retained austenite, and secondary
carbides. The final matrix will depend on chemical composition, rate of cooling after casting, and eventual heat treatment.
The hardness of casting with a given carbon content is directly related to the martensite content of its matrix. The abrasion
resistance of iron and steel correlates reasonably well with their carbon content and microstructure. Ni-Hard 1 is especially suited for
bimetallic castings because Ni-Hard attains its abrasion resistance in the as-cast condition or after only a low-temperature heat

7
A. Albiter et al. Engineering Failure Analysis 171 (2025) 109421

Fig. 6. Visual inspection of both fracture surfaces and joining of surfaces was performed for analysis.

Table 3
Chemical composition of MH-22 specimen and comparison with Ni-hard 1 specification range
(ASTM A532).
Element Range Ni-Hard 1 Sample (%)
ASTM A532

Fe 91–97 92.73
Ni 3.3–5 3.94
Cr 1.4–4 1.83
Mn 2.0 max 0.34
C 2.8–3.6 2.6
Si 0.8 max 0.32
Mo 1.5 max 0.26
S 0.15 max 0.06
Cu − 0.26
Al − 0.6

treatment. The cooling rate should be as slow as possible, especially in the critical range of martensite formation.
Ni-Hard 1 was originally intended to be put in service in the as-cast condition. However, almost all castings are heat-treated. The
most used are: 1) Single-stage tempering (225–275 ◦ C) for stress relief and improved toughness; 2) Double-heat treatment at 450 ◦ C
followed by 275 ◦ C to improved repeated impact fatigue resistance. The result is that more austenite will be transformed to martensite.
3) Heat treatment at 750–850 ◦ C followed by slow cooling, to improve hardness and abrasion resistance.

3.4. Material identification of the AH-38 component

Similarly, to identify the material of the AH-38 component, a portable energy-dispersive X-ray fluorescence analyzer (TITAN Model
LE 600, Bruker brand) was used. Several analyses were conducted, and it was determined that the material of the AH-38 component
was Monel K500 compared to the AMS-4676 [27], as shown in Table 4. Monel K-500 is a nickel-copper alloy with small amounts of
aluminum and titanium and is known for its high mechanical strength, good predisposition to weldability, excellent corrosion
resistance, and toughness over a wide temperature range [28]. It is used in chemical and petrochemical industries in heat exchangers,
pressure vessels, pipes, and chemical processing equipment [29]. Monel K-500 is used for service applications in sour gas environments
and has good resistance to stress corrosion cracking (SCC) [30].

8
A. Albiter et al. Engineering Failure Analysis 171 (2025) 109421

3.5. Hardness measurements on MH-22 component

The hardness test was conducted in one section of the MH-22 component. The measurements were taken with an AFFRI Digitronic
Mod. 102 hardness tester, as is shown in Fig. 7. Ten hardness readings were performed on the cross section and five measurements on
the surface of component using the Rockwell-C (HRC) scale, and an average hardness was obtained. The results showed an average
hardness of 39 HRC (110 HRB). According to the results obtained, the hardness values are homogeneous throughout the piece,
including on the surface of the component. However, materials exposed to sour environments with high H2S concentrations should not
exceed a hardness of 22 HRC to avoid SSC. The MH-22 component with an average hardness of 39 HRC is highly susceptible to SSC,
which was verified by the presence of internal cracks in the material (ANSI/NACE MR0175-2021/ISO 15156) [20].
The hardness of Ni-Hard 1 is a function of its carbide (M3C) content and matrix structure [25]. Hardness is usually regarded as an
indication of the abrasion resistance of a material. Abrasion resistance increases with hardness. However, it must be recognized that
hardness does not improve abrasion resistance by itself but is meaningful only as an indicator of the microstructure of a specific alloy.
The abrasion resistance of a material depends on the system where abrasion takes place. The abrasion resistance of iron and steel
correlates reasonably well with their carbon content and microstructure. Generally, abrasion resistance will increase with carbon
content. The highest abrasion resistance is achieved by materials with martensitic matrix structures. Any tempering will reduce
abrasion resistance and may occur during work-place usage due to frictional heating.
The practical meaning of fracture resistance for an abrasion-resistant white iron like Ni Hard-1, is that it does not break or crack
during service.
The most significant property of Ni-Hard castings is their hardness. In this work, hardness is the only property that was determined.
The determination tensile strength was not possible for the small component provided. For this kind of alloy (white cast irons),
obtaining specimens for tension testing is difficult and expensive. The machining of test pieces with this type of alloy is difficult by the
hard and brittle material. However, the values of tensile strength reported range from 280 to 450 MPa [26].

3.6. Optical microscopy (OM)

The microstructure of MH-22 was analyzed using a ZEISS AxioVert 40 MAT Microscope at magnifications of 100X. Cubic samples
were extracted from the cross-sectional and longitudinal areas adjacent to the failure zone. These samples were prepared through
standard grinding and mirror polishing processes. To analyze the microstructure of the MH-22 component, a section was cut and
roughened using silicon carbide sandpaper, followed by mirror polishing with diamond paste. Chemical etching with 2 % Nital was
then applied to reveal the microstructure. Fig. 8 shows the microstructure of the MH-22 component observed through optical mi­
croscopy . This sample was observed with OM before chemical etching (Fig. 8a) and after chemical etching (Fig. 8b).
In Fig. 8a, without chemical etching, the phases of the microstructure are not clearly visible due to the mirror-polished surface, only
the presence of some precipitates of different sizes can be seen. In Fig. 8b, after etching with 2 % Nital, the chemical etching highlights
the different nodular precipitates in the material. Ni-hard 1 steel cools rapidly after solidification, generating a martensitic structure in
its matrix (Fig. 8b). Martensite is a hard and strong phase that provides the mechanical strength of the material. In addition to
martensite, Ni-hard 1 contains a large amount of iron and nickel carbides such as Fe3C, Ni3Cr, and FeNi3. These carbides (usually Fe3C)
are dispersed precipitates within the martensite matrix. Under certain cooling and composition conditions, a small fraction of
austenitic phase may exist in the microstructure of Ni-hard 1, although martensite generally predominates. The austenitic phase may
contribute to toughness and resistance to deformation at higher temperatures.

3.7. Scanning Electron microscopy (SEM)

3.7.1. EDS analysis


SEM images were obtained using a Jeol Model JSM-IT Scanning Electron Microscope. Fig. 9a shows an EDS analysis indicating that
precipitates are primarily composed of Fe, Ni, C, Al and Cr. In this type of alloys, Cr3C2 and Fe3C carbides are commonly found, which
are used to improve the wear resistance and hardness of the alloys. Some intermetallic compounds such as Ni3Cr and FeNi3 can also
form in these alloys. According to the elemental analysis of the precipitates, it is possible that they are precipitates of Fe3C and FeNi3.

Table 4
Chemical composition for AH38 component (Monel K500).
Element Range (%) Sample (%)
AMS-4676

Ni 63–70 63.84
Cu 18.09–37.63 29.65
Al 2.30–3.15 3.96
Fe 2.0 max 1.12
Mn 1.5 max 0.67
Ti 0.35–0.85 0.51
S 0.01 max 0.02
Si 0.50 max 0.01
C 0.25 max ​

9
A. Albiter et al. Engineering Failure Analysis 171 (2025) 109421

Fig. 7. a) image of the hardness measurements performed in the mh-22 component, b) results of the hardness measurements.

Fig. 8. Image obtained by OM of the MH-22 component polished to a mirror finish: a) without chemical etching and b) with chemical etching
(2% Nital).

Fig. 9b shows the presence of nodular precipitates typical of Ni-Hard 1 steel, confirming that the microstructure corresponds to this
type of steel. Fig. 9c shows the elemental chemical mapping of a nodular precipitate. Ni-Hard 1 is a white cast iron alloyed with nickel
(Ni) and chromium (Cr), featuring a structure composed of a hard carbide matrix that becomes martensitic due to the nickel content.

3.7.2. Fracture surface analysis


A cleaning process was necessary to observe the fractured pieces of the MH-22 component under SEM, as the fractures were covered
with hydrocarbons, giving them a completely black appearance. This chemical cleaning involved immersing the piece in acetic acid for
approximately 10 min in an ultrasonic bath. Since this cleaning was insufficient to obtain a clean fracture surface, inhibited acid was
applied to the surface for 5–10 s. Subsequently, the sample was washed with acetone and air-dried. Immediately after, the piece was
subjected to SEM observation.
Fig. 10 shows general images of the complete fracture surface of MH-22 component after the cleaning process. SEM images of the
fracture surface of MH-22 component are shown in Figs. 11 and 12. These images mainly reveal the presence of internal cracks on the
material surface caused by the effect of H2S.
Fig. 11 shows micrographs with different internal cracks or defects in the MH-22 component. Fig. 11a shows internal cracks that
follow the grain boundaries within the material. These cracks are indicative of intergranular fracture, where separation occurs along
the grain boundaries rather than through the grains themselves. Fig. 11b shows a micrograph that also depicts intergranular internal
cracks. However, these cracks are attributed to material decohesion, where the cohesive forces within the material are compromised,
leading to separation along the grain boundaries. Fig. 11c focuses on an area where the material has become embrittled. Embrittlement
is a process that makes a material more brittle and prone to cracking under stress and presence of H2S. The embrittled area shows

10
A. Albiter et al. Engineering Failure Analysis 171 (2025) 109421

Fig. 9. Compositional analysis of the nodular precipitates in the MH-22 component: a) EDS analysis and b) SEM image of nodular precipitates, c)
elemental chemical mapping of a nodular precipitate.

numerous cracks and likely a more brittle fracture surface. Fig. 11d displays both microcracks and larger internal macrocracks within
the material. Macrocracks are large cracks that can significantly affect structural integrity, while microcracks are smaller and may
develop into macrocracks over time. Arrows point to both microcracks and larger internal macrocracks.
These micrographs illustrate different types of cracking phenomena within a material, which could be due to various factors, such
as stress, embrittlement, and material decohesion. Understanding these microstructural features is crucial for assessing material
integrity and performance in different applications.

11
A. Albiter et al. Engineering Failure Analysis 171 (2025) 109421

Fig. 10. General view of the fracture surface of the MH-22 component a) before cleaning and b) after cleaning.

Fig. 12 shows micrographs illustrating the stages of crack formation and propagation in a material. In Fig. 12a, internal macro­
cracks are observed, indicating significant internal damage, as well as a region characterized by a brittle fracture, where the material
has fractured with little or no plastic deformation. In Fig. 12b, the crack initiation zone is shown, with arrows indicating the devel­
opment of initial cracks and the visible brittle zone, demonstrating that the material in this region fractures in a brittle manner, and the
arrows indicate the direction of crack propagation. In Fig. 12c, the focus is on crack propagation through the material, with arrows
indicating the crack paths, highlighting the advancement from the initial zones deeper into the material. Fig. 12d shows the final crack
arrest zone, where crack propagation has halted, indicating the end of the propagation process. These images together represent the
sequence of crack development, from initiation through propagation to final arrest in a brittle material.

3.8. Material preselection (normative analysis)

Using a methodology for material preselection for systems with CO2 + H2S, based on the ANSI/NACE MR0175/ISO 15156-2, Part 3
[20] and API Spec 6A [21] standards, the materials were selected according to the operating conditions. Table 5 presents the infor­
mation required to apply the methodology. Based on the results, recommendations were made for materials that are resistant to
corrosion, aligning with the material’s performance in these specific conditions.
ANSI/NACE MR0175/ISO15156 [20] provides general guidelines for estimating the pH of the water phase under various condi­
tions, as shown in Fig. 13. This estimated pH can be used if specific calculations are not provided or if “in situ” measurement techniques
are not reliably available.
Where:
ppH2S (kPa) = (% mol H2S*10e-2) (working pressure)
ppCO2 (kPa) = (% mol CO2*10e-2) (working pressure)
Obtaining a partial pressure of 154 and 193 psi for CO2 and H2S respectively. Using partial pressure as a measure to predict
corrosion the following relationships have been found:

• ppCO2 < 7 psi → no corrosion occurs


• ppCO2 7–30 psi → moderate corrosion
• ppCO2 > 30 psi → severe corrosion

According to results from ppCO2 severe corrosion may occur. However, corrosion depends on many factors such as metallurgy,
temperature, water content, pressure, among others. In this case, significant corrosion in the MH-22 component was not observed.
It should be noted that the MH-22 component is occasionally used for well logging (P-T) and is not exposed for long periods of time
to a well environment. According to the visual inspection and observations with SEM and OM, there is no significant corrosion in the
component, thus CO2 corrosion was not the cause of the MH-22 failure.
In ANSI/NACE MR0175/ISO15156 standard [20], the pH is calculated at 20 ◦ C and 100 ◦ C due to the influence that temperature
has on the corrosion of materials in the presence of H2S environments, such as in the oil and gas industry. The main reasons for
calculating the pH at these two temperatures are:
Influence of temperature on corrosion: Corrosion induced by H2S is highly dependent on temperature. At higher temperatures, the
corrosion rate and the severity of mechanisms such as sulfidation and SSC can increase. Measuring pH at 100 ◦ C allows for the
evaluation of how the conditions of the corrosive environment change at higher temperatures, which may represent the actual con­
ditions of the reservoir or process.
pH variation with temperature: The pH of an aqueous solution can change significantly with temperature. At 20 ◦ C, a standard

12
A. Albiter et al. Engineering Failure Analysis 171 (2025) 109421

Fig. 11. SEM images of the fracture surface of MH-22 component showing a-b) internal cracks with intergranular appearance, c) areas of material
embrittlement, d) larger internal cracks.

reference close to ambient temperature is obtained, while at 100 ◦ C, the pH tends to decrease due to increased water ionization and
changes in the physicochemical properties of the solution. These two values provide a reference range to assess the behavior of
materials under different operating conditions.
Evaluating pH at both temperatures is important to meet the material selection criteria for H2S corrosion resistance, following the
guidelines of the standard. This ensures that the selected materials can withstand both storage conditions (at ambient temperature) and
operating conditions (at elevated temperatures), minimizing the risk of corrosion failure in equipment exposed to H2S environments.
In Fig. 14, the severity of the environment for carbon and low-alloy steels was determined, the limits for pH values of 3.2 and 3.4 are
observed, with an upper and lower limit of 20 %, and a ppH2S (kPa) value of 1064 kPa (154 psi). According to the operating conditions
of the MH-22 carbon steel component, it falls outside the limits established by the NACE MR 0175 standard [20] and exhibits a high
susceptibility to SSC [31], as illustrated in Fig. 15. Additionally, an analysis was conducted to classify the service type and medium
corrosivity according to API Spec 6A [21]. SCC in pipelines occurs due to a combination of tensile stresses and corrosive environments.
SCC can be influenced by residual and externally applied stress, as well as the type of surface corrosion (sweet or sour). The micro­
structure of pipeline steel, which includes phases such as austenite, pearlite, and ferrite, significantly affects SCC susceptibility [32].
In Region 0 (ppH2S < 0.3 kPa or 0.05 psi) special precautions are generally not needed. However, steels susceptible to SSC and
Hydrogen Stress Cracking (HSC) can crack, and their physical and metallurgical properties influence their resistance to SSC and HSC.
High-strength steel can suffer from HSC in aqueous environments without H2S, special attention must be given to their composition

13
A. Albiter et al. Engineering Failure Analysis 171 (2025) 109421

Fig. 12. SEM images of the fracture surface of MH-22 component showing a) brittle zone and cracks, b) and c) crack initiation and propagation
zone, and d) final crack arrest zone.

Table 5
Base information for the preselection of materials for CO2 + H2S systems.
Operating conditions Wellhead temperature 30 ◦ C
Bottomhole temperature 109 ◦ C
Wellhead pressure 4 MPa
Bottomhole pressure 13.65 MPa
Working pressure 13.7 MPa
Production data Gas production 0.1 MMCFD
Oil production 20 BPD
Water cut 28.5 %
Gas content H2S content 9.89 % mol
CO2 content 7.85 % mol
Component details Outer diameter 88.9 mm

Inner diameter 63 mm
Length 101.6 mm
Inclination angle 0 degrees

14
A. Albiter et al. Engineering Failure Analysis 171 (2025) 109421

Fig. 13. Determination of the pH of the hydrocarbon in contact with the MH-22 component.

Fig. 14. Determination of SSC susceptibility considering the pH and ppH2S.

and processing. In Regions 1, 2, and 3 (ppH2S ≥ 0.3 kPa or 0.05 psi), materials should be selected according to the appendices of ANSI/
NACE MR 0175 / ISO 15156 [20]. If no alternatives are available, carbon and low-alloy steels must be tested and qualified according to
NACE MR0175 / ISO 15156–1 [20]. Considering the operating conditions, the carbon steel material, and the specifications in API SPEC
6A [21], it is classified as a sour service (HH). However, it is necessary to determine the susceptibility to sulfide stress cracking (SSC)
due to the H2S content. According to this analysis, the material required for any component must be suitable for sour service and be a
special alloy resistant to H2S and CO2.
Using pre-selection methodology, the following recommendations are made for candidate materials: corrosion-resistant alloys
(CRAs) or carbon steels that comply with ANSI/NACE MR0175/ISO 15156-2 [20].
Based on this evaluation, the material for the MH-22 component should have been selected from the following options:

• A nickel-based alloy, known for its high corrosion resistance in H2S environments [33].
• Material like Sanicro 28, an alloy that combines excellent corrosion resistance with suitable mechanical properties for such
conditions.
• Duplex alloys like 2205 or 2507, which offer outstanding corrosion resistance to acids and corrosive gases, as well as high me­
chanical strength, making them ideal for severe H2S and CO2 conditions (Fig. 16).

15
A. Albiter et al. Engineering Failure Analysis 171 (2025) 109421

Fig. 15. Classification of service type according to API Spec 6A [21]. AA: Pipe for general service, BB: Pipe for general service (low CO2), CC: Pipe
for general service (high CO2), DD: Sour service, EE: Sour service (low CO2), FF: Sour service (medium CO2), HH: Sour service (high CO2).

Each of these options has been recommended by material suppliers to ensure the optimal performance of the component in sour gas
environments.

3.9. Failure mechanism analysis

To assess the influence of CO2 and H2S on a Ni-Hard steel component within a production pipeline, it is important to consider the
corrosion processes and specific effects of these gases. This analysis involves an approach that integrates thermodynamic, kinetic and
metallurgical properties.
H2S and CO2 gases in combination with water are the main causes of corrosion in the oil industry. CO2 corrosion can manifest itself

Fig. 16. Selection of materials according to H2S and CO2 content.

16
A. Albiter et al. Engineering Failure Analysis 171 (2025) 109421

as general wear or localized attack in the form of pitting. The severity of CO2 corrosion increases with temperature and CO2 partial
pressure.
Carbon dioxide dissolves in the presence of water, forming carbonic acid, which is corrosive to carbon steel. The reactions that take
place are:
CO2(g) + H2 O(l) →CO2(dissolved) (1)

CO2(dissolved) + H2 O(l) →H2 CO3 (2)

The main corrosion process can be summarized by three cathodic reactions and one anodic reaction:
2H2 CO3 + 2e− →H2 + 2H2 CO−3 (3)

2H2 CO3− + 2e− →H2 + 2CO−3 2 (4)

2H+ CO3 + 2e− →H2 (5)

Fe→Fe+2 + 2e− (6)

According to results from ppCO2 severe corrosion may occur. However, no significant corrosion in the MH-22 component was
observed. Which can be attributed that the component is occasionally used for well logging (P-T) and is not exposed for long periods of
time to a well environment. Therefore, CO2 corrosion was not the cause of the MH-22 failure.
Sulfide stress cracking is a phenomenon that occurs in metals, particularly steels, when exposed to environments with high con­
centrations of H2S and high pressure and temperature conditions.
H2S can penetrate and react with metals, generating atomic hydrogen (H). This hydrogen accumulates in high stress areas of the
material, which can lead to the formation of holes or intergranular cracking. The accumulated hydrogen reduces the ductility of the
metal, which facilitates the formation of cracks and their propagation [34–36]. At higher concentrations of H2S, the risk of cracking
increases. High temperatures and pressures also accelerate corrosion reactions and promote the diffusion of hydrogen into the metal.
High strength steels, with hardness greater than 22 HRC, are more susceptible to this type of damage.
Since the hydrogen evolution and permeation in steels is a concern regarding the SSC process, the interaction of hydrogen with
metallurgical defects such as inclusions plays a fundamental role in crack initiation and propagation under high H2S contents. Many of
these inclusions suffer corrosion and dissolution, leaving microvoids in the steel. Once hydrogen enters the steel it will be prone to be
trapped in these microvoids deforming the structure which results in cracks formation [34].
The trapping sites of hydrogen in a martensite steel are considered to be one of the key factors affecting the hydrogen embrittlement
(HE) of these steels. Safyari et al. [37] said that crack propagation in the hydrogen environment mainly depends on hydrogen ab­
sorption and distribution. On the other hand, Jack et al. [38] Investigated the hydrogen diffusion and cracking susceptibility of
pipeline steels after various heat treatments. They found distinct differences in micro structural characteristics, which significantly
affected their cracking behavior. In similar way, Eskinja et al. [39] study the influence of Mo carbides on the susceptibility of a
martensitic steel to HE. The results revealed that two-stage tempered steel exhibited superior resistance to HE, as a result of reduced
dislocation density and higher quantity of Mo2C carbides.
H2S can react directly with iron and other metals to form metal sulfides, such as iron sulfide (FeS). This process not only weakens
the material but can also generate products that promote corrosion. The typical reaction is:
Fe(s) + H2 S→FeS(s) + H2(g) (7)

and when water is present the reaction is:

Fe + H2 S + H2 O ↔ FeSH− + H3 O+ (8)

Depending on the pH, the following reactions can occur:

Fe + HS− →FeS + H+ + 2e− (9)

FeS + HS− →FeS2 + H+ + 2e− (10)

In this study, the combination of material hardness and high H2S content was explored through visual inspection, hardness mea­
surements, fracture surface observation, and microscopic analysis. The results revealed that MH-22 component suffered SSC. The
combination of material hardness and high H2S content made the MH-22 component highly susceptible to SSC, which was confirmed
by the embrittlement and presence of internal cracks in the material. The analysis revealed internal cracks with an intergranular
appearance, confirming embrittlement. These cracks, along with brittle fracture zones, were observed throughout the fracture surface,
supporting the H2S-induced cracking. In addition, the scratches and the type of fracture observed in the components indicate that
components suffer excessive torque and fatigue when tightening and loosening the MH-22 and AH-38 components with the Stilson
keys, this generates fatigue and internal stresses that promote SSC. Once the material became brittle, any impact during well oper­
ations was enough to cause it to fail.

17
A. Albiter et al. Engineering Failure Analysis 171 (2025) 109421

For wells with high H2S content, it is crucial to select materials resistant to sour environments [5]. More suitable material, such as
nickel alloys (2205, 2507, 625, 718, Sanicro 28, etc.), should have been chosen to resist the high H2S concentrations. Nickel alloys are
known for their high corrosion resistance in H2S environments [31]. It is crucial to continuously monitor the pressure, temperature,
and fluid composition in oil and gas wells, as these factors directly impact the integrity of system components and the risk of failures
due to corrosion [4,16,40].
SSC is a serious phenomenon that can affect the integrity of metallic materials in industrial environments. These processes can be
controlled by proper material selection, monitoring operating conditions and the use of protective technologies.

4. Conclusions

The investigation into the MH-22 component failure identified key findings and necessary measures to mitigate future failures
under operating conditions:

1. Based on the operational conditions, the failure analysis of the MH-22 component confirmed that the rupture was primarily caused
by sulfide stress cracking (SSC), attributed to the combination of unsuitable material selection, high hardness, and elevated H2S
content in the well environment.
2. Hardness tests revealed an average hardness of 39 HRC, significantly exceeding the recommended maximum of 22 HRC for
components exposed to sour gas environments. This hardness increased susceptibility to embrittlement and cracking, corroborated
by the internal intergranular cracks observed on the fracture surface, confirming an SSC failure mechanism.
3. The type of fracture and scratches observed in the components indicate that components suffer excessive torque and fatigue when
tightening and loosening the MH-22 and AH-38 components; this generates fatigue and internal stresses that promote SSC. Once the
material became brittle, any impact during well operations caused the failure.
4. The material selection for the MH-22 component was inadequate for the operational conditions. The use of H2S-resistant alloys,
such as nickel-based alloys, would have provided the necessary mechanical and chemical stability, prevented SSC and extended the
component’s service life.
5. Replacing the current material with H2S-resistant alloys, such as nickel-based alloys (e.g., Inconel 718), is essential to improve
resistance to sour environments and significantly reduce SSC susceptibility. To prevent such failures in the future, material se­
lection processes should be enhanced, focusing on alloys specifically designed for sour gas conditions.
6. Preventative actions should focus on ensuring proper material hardness, using corrosion-resistant alloys, and applying protective
coatings. Use special alloys and sulfide-resistant materials, such as stainless steels, nickel alloys, and materials with protective
coatings, such as anticorrosive paints or surface heat treatments such as nitriding or cementation. Temperature and pressure
control to maintain operating conditions that reduce the risk of cracking, avoiding prolonged exposure to high concentrations of
H2S.

Funding
No funding was received to assist with the preparation of this manuscript.

CRediT authorship contribution statement

A. Albiter: Writing – review & editing, Writing – original draft, Visualization, Validation, Methodology, Investigation. Lucila Cruz-
Castro: Writing – review & editing, Writing – original draft, Methodology, Investigation. A. Contreras: Investigation, Methodology,
Validation, Visualization, Writing – original draft, Writing – review & editing.

Declaration of competing interest

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

Acknowledgments

The authors sincerely acknowledge Instituto Mexicano del Petróleo for support in the experimental work.

Data availability

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

References

[1] Z. Yu, D. Zeng, S. Hu, X. Zhou, W. Lu, J. Luo, Y. Fan, K. Meng, The failure patterns and analysis process of drill pipes in oil and gas well: a case study of fracture
S135 drill pipe, Eng. Fail. Anal. 138 (2022) 106171, https://doi.org/10.1016/j.engfailanal.2022.106171.

18
A. Albiter et al. Engineering Failure Analysis 171 (2025) 109421

[2] C. Mingjie, H. Mingmin, T. Leichuan, M. Dan, X. Jinchao, Failure analysis of large-size drilling tools in the oil and gas industry, J. Energy Res. Technol. 146
(2024), https://doi.org/10.1115/1.4065250.
[3] A.I. Mohammed, B. Oyeneyin, B. Atchison, J. Njuguna, Casing structural integrity and failure modes in a range of well types – a review, J. Nat. Gas Sci. Eng. 68
(2019) 102898, https://doi.org/10.1016/j.jngse.2019.05.011.
[4] Z. Chen, M. Xu, T. Hu, G. Xue, F. Chen, H. Zhao, H. Zhou, Y. Lei, K. Zhu, Effects of H2S content on the corrosion behavior of gas storage reservoir injection and
production pipeline steel in CO2-H2S environment, Mater. Today Commun. 41 (2024) 110364, https://doi.org/10.1016/j.mtcomm.2024.110364.
[5] D. Zeng, B. Dong, F. Zeng, Z. Yu, W. Zeng, Y. Guo, Z. Peng, Y. Tao, Analysis of corrosion failure and materials selection for CO2–H2S gas well, J. Nat. Gas Sci.
Eng. 86 (2021) 103734, https://doi.org/10.1016/j.jngse.2020.103734.
[6] E.D.W. Syah Putri, T. Triyono, A.R. Prabowo, Estimating failure mechanism of steel specimens using stress corrosion-cracking (SCC) testing methods: state and
development, Proc. Struc. Integ. 41 (2022) 266–273, https://doi.org/10.1016/j.prostr.2022.05.031.
[7] M.T. Albdiry, M.F. Almensory, Failure analysis of drill string in petroleum industry: a review, Eng. Fail. Anal. 65 (2016) 74–85, https://doi.org/10.1016/j.
engfailanal.2016.03.014.
[8] B. James, A. Hudgins, Chapter 1 - Failure analysis of oil and gas transmission pipelines, in: A.S.H. Makhlouf, M. Aliofkhazraei (Eds.), Handbook of Materials
Failure analysis with case studies from the oil and gas industry, Butterworth-Heinemann, 2016: pp. 1–38. 10.1016/B978-0-08-100117-2.00001-7.
[9] L. Li, Z. Lian, C. Zhou, Failure analysis of drill pipe during working process in a deep well: a case study, Processes 10 (2022), https://doi.org/10.3390/
pr10091765.
[10] J.R. Saithala, A. Kharusi, M. Suryanarayana, N. Behlani, T. Nabhani, Implications of failure of alloy 718 (UNS N07718) tubing hanger in sour well, Eng. Fail.
Anal. 120 (2021) 105060, https://doi.org/10.1016/j.engfailanal.2020.105060.
[11] D. Shen, K. Tong, Z. Fan, W. Yang, Q. Liu, H. Li, J. Zhao, S. Cong, Failure analysis of S135 drill pipe body fracture in a well, Eng. Fail. Anal. 145 (2023) 106998,
https://doi.org/10.1016/j.engfailanal.2022.106998.
[12] L. Shi, K. Luo, J. Wang, D. Feng, H. Zhang, X. Cao, Failure analysis of an offshore drilling casing under harsh working conditions, Eng. Fail. Anal. 120 (2021)
105018, https://doi.org/10.1016/j.engfailanal.2020.105018.
[13] G. Mubarak, M. Elkhodbia, I. Gadala, A. AlFantazi, I. Barsoum, Failure analysis, corrosion rate prediction, and integrity assessment of J55 downhole tubing in
ultra-deep gas and condensate well, Eng. Fail. Anal. 151 (2023) 107381, https://doi.org/10.1016/j.engfailanal.2023.107381.
[14] W. Wu, M. Sun, P. Chai, J. Yang, L. Song, Z. Liu, Failure analysis of a high strength Cr-containing tube used in an oil well, Eng. Fail. Anal. 150 (2023) 107335,
https://doi.org/10.1016/j.engfailanal.2023.107335.
[15] M. Liu, S. Luo, Y. Shen, X. Lin, Corrosion fatigue crack propagation behavior of S135 high–strength drill pipe steel in H2S environment, Eng. Fail. Anal. 97
(2019) 493–505, https://doi.org/10.1016/j.engfailanal.2019.01.026.
[16] J. Jing, Y. Tian, X. Zhu, Y. Zhou, C. Shi, Q. Lei, Safety evaluation of offshore oil and gas well string based on corrosion rate prediction, J. Loss Prev. Process Ind.
91 (2024) 105385, https://doi.org/10.1016/j.jlp.2024.105385.
[17] A. Albiter, L. Cruz-Castro, A. Contreras, Failure analysis of a bottom hole motor attached to a coiled tubing, Forces Mech. 14 (2024) 100250, https://doi.org/
10.1016/j.finmec.2023.100250.
[18] A. Albiter, L. Cruz-Castro, A. Contreras, Failure analysis of CT-90 coiled tubing used in oil and gas industry in Mexico, J. Fail. Anal. Prev. (2024), https://doi.
org/10.1007/s11668-024-01918-6.
[19] A. Albiter, Sulfide stress cracking assessment of carbon steel welding with high content of H2S and CO2 at high temperature: a case study, Engineering 12 (2020)
863–885, https://doi.org/10.4236/eng.2020.1212061.
[20] ANSI/NACE MR0175-2021/ISO 15156:2020 (2020). Petroleum and natural gas Industries-materials for use In H2S-Containing environments in oil and gas
production. Part 1,2 and 3.
[21] API 6A, (2010). Specification ISO 10423:2009 (Modified), Petroleum and natural gas industries– Drilling and production equipment–Wellhead and christmas
tree equipment 6A.
[22] ASTM D1945-14 (2019). Standard test method for analysis of natural gas by gas chromatography, American Society for Testing and Materials. American Society
for Testing and Materials. ASTM International.
[23] O. Muratal, R. Yamanoğlu, C. Duran, Y. Gönülalan, Y. Akyıldız, F. Koç, B. Barutçuoğlu, Production of Ni-hard alloy powders by gas atomization, Int. J. 3D Print.
Technol. Dig. Ind. 8 (2024) 124–129. 10.46519/ij3dptdi.1402760.
[24] ASTM A532/A532M-10 (2019). Standard specification for abrasion-resistant cast irons. American Society for Testing and Materials. ASTM International.
[25] K. D. Ajoy, A.N. Sudhakar, R. Markandeya, B. Srinivasa Rao, K. P Ajoy., D. Kaushik. Effect of alloying elements on the microstructure and mechanical properties
of high chromium white cast iron and Ni-Hard iron, Mater. Today Proc. 61 (2022) 1006–1014. 10.1016/j.matpr.2021.10.284.
[26] John Fischer, Properties and applications of Ni Hard alloys. A guide to the use of nickel containing alloys No 11017, Second Edition, 2021.
[27] SAE International, AMS-4676 Nickel-Copper alloy, corrosion-resistant, bars and forgings 66.5Ni - 3.0Al - 0.62Ti - 28Cu hot-finished, precipitation hardenable,
1964.
[28] A.G. Kostryzhev, O.O. Marenych, Z. Pan, H. Li, S. van Duin, Strengthening mechanisms in Monel K500 alloyed with Al and Ti, J. Mater. Sci. 58 (2023)
4150–4164. 10.1007/s10853-023-08248-2.
[29] Z. Chen, W. Fan, S. Gao, Y. Qi, S.Y. Kandukuri, K. Zhou, Effect of grain structure on the mechanical properties of a Monel alloy fabricated by laser-based directed
energy deposition, J. Mater. Sci. Technol. 164 (2023) 129–139, https://doi.org/10.1016/j.jmst.2023.04.044.
[30] Q.Y. Wang, X. Luo, X. Zhang, T. Liu, H. Zheng, L. Dong, Y. Xi, S. Bai, Stress corrosion cracking behavior and mechanism of aging treated Monel K500 alloy in
flowing seawater, J. Mater. Sci. 58 (2023) 1–19, https://doi.org/10.1007/s10853-023-08404-8.
[31] C.M. Hussain, C. Verma, J. Aslam, R. Aslam, S. Zehra, 9 - Stress corrosion cracking, in: C.M. Hussain, C. Verma, J. Aslam, R. Aslam, S. Zehra (Eds.), Handbook of
Corrosion Engineering, Elsevier, 2023: pp. 101–110. 10.1016/B978-0-323-95185-2.00009-5.
[32] M. Nnoka, T. Alaso Jack, J. Szpunar, Effects of different microstructural parameters on the corrosion and cracking resistance of pipeline steels: a review, Eng.
Fail. Anal. 159 (2024) 108065, https://doi.org/10.1016/j.engfailanal.2024.108065.
[33] N. Rathi, P. Kumar, S. Kumar, A. Gupta, Non-conventional machining of nickel-based superalloys: a review, Mater. Today Proc. (2023), https://doi.org/
10.1016/j.matpr.2023.02.176.
[34] B. Gu, J. Luo, X. Mao, Hydrogen-facilitated anodic dissolution-type stress corrosion cracking of pipeline steels in near-neutral pH solution, Corrosion 55 (1999)
96–106, https://doi.org/10.5006/1.3283971. 10.1016/j.engfailanal.2024.108562.
[35] Z.Y. Liu, X.G. Li, C.W. Du, L. Lu, Y.R. Zhang, Y.F. Cheng, Effect of inclusions on initiation of stress corrosion cracks in X70 pipeline steel in an acidic soil
environment, Corros. Sci. 51 (2009) 895–900, https://doi.org/10.1016/j.corsci.2009.01.007.
[36] W. Chen, F. King, E. Vokes, Characteristics of near-neutral-pH stress corrosion cracks in an X-65 pipeline, Corrosion 58 (2002) 267–275, https://doi.org/
10.5006/1.3279878.
[37] M. Safyari, S. Bhosale, M. Moshtaghi, Capacity of hydrogen traps affects H-assisted crack initiation and propagation mechanisms in martensitic steels, Eng. Fail.
Anal. 163 (2024) 108560, https://doi.org/10.1016/j.engfailanal.2024.108560.
[38] T.A. Jack, M. Nnoka, J. Zhang, J. Szpunar, Investigating the hydrogen diffusion and cracking susceptibility of pipeline steels after various heat treatments, Eng.
Fail. Anal. 161 (2024) 108311, https://doi.org/10.1016/j.engfailanal.2024.108311.
[39] M. Eskinja, G. Winter, H. Schnideritsch, J. Klarner, V. Razumovskiy, M. Moshtaghi, G. Mori, Influence of Mo carbides and two-stage tempering methodology on
the susceptibility of medium carbon martensitic steel to hydrogen embrittlement, Eng. Fail. Anal. 163 (2024) 108562.
[40] K. Liao, J. Leng, Y.F. Cheng, T. He, G. He, S. Zhao, X. Liu, Q. Huang, Effect of H2S concentrations on corrosion failure of L245NS steel in CO2-O2-H2S system, Pro.
Saf. Env. Prot. 168 (2022) 224–238, https://doi.org/10.1016/j.psep.2022.09.038.

19

You might also like