Case Studies in Engineering Failure Analysis: M. Ghalambaz, M. Abdollahi, A. Eslami, A. Bahrami

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Case Studies in Engineering Failure Analysis 9 (2017) 52–62

Contents lists available at ScienceDirect

Case Studies in Engineering Failure Analysis


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

Short communication

A case study on failure of AISI 347H stabilized stainless steel pipe


T
in a petrochemical plant

M. Ghalambaza, M. Abdollahib, A. Eslamib, , A. Bahramib
a
Jam Petrochemical Complex, Asalouyeh, Iran
b
Department of Materials Engineering, Isfahan University of Technology, Isfahan 84156-83111, Iran

AR TI CLE I NF O AB S T R A CT

Keywords: In this study failure of AISI 347 stabilized stainless steel pipe after 60,000 of working in a pet-
Failure rochemical plant was investigated. Result showed that the main cause of failure was thermal
Thermal fatigue stress fatigue. Fatigue cracks were formed at the outer surface of the investigated pipe, and were
Chromium rich phase grown towards the inner surface at the fusion line of welded area. The formation of chromium-
Stainless steel
rich phases together with thermal fatigue stresses were found to be main causes of failure.
Welding

1. Introduction

A petrochemical plant consists of many types of equipment, including pressure vessels, furnaces, drums, towers, and pipes. All
these components must have good corrosion resistance and mechanical properties. Materials, used in a petrochemical plant, must
have acceptable creep resistance, hot corrosion resistance, and resistance to fatigue.Also a reliable weld of good mechanical/cor-
rosion characteristics is a crucial requirement for many applications and components. Austenitic stainless steels which have great
high temperature properties are commonly used for parts that require high working temperatures [1,2]. Austenitic stainless steels
such as 347H, 321, 304H are vastly used in superheater and re-heater pipes that are exposed to high temperatures. Olefin furnaces are
good examples in which these alloys are commonly used. Among mentioned austenitic stainless steels, type 347H shows an excellent
high temperature performance, and is one of the most commonly used grades [3]. Niobium, used in this grade, significantly improves
the yield strength, mainly due to solid solution and precipitation hardening. Moreover, 347H austenitic stainless steel is a well-known
grade for its excellent resistance to intergranular corrosion and intergranular stress corrosion cracking (SCC) [4]. Some grades of
stainless steels (such as 347 and 321 types) are subjected to stabilization process. Stabilization is done, aiming at improving the creep
and intergranular corrosion resistance of the alloy. In these grades Ti and Nb will prevent chromium carbide precipitation by forming
carbides at 900–950 °C. Stabilization curve for titanium and niobium carbides is shown in Fig. 1 [5–8].
In austenitic stainless steels, presence of chromium and molybdenum elements increase the corrosion resistance. However, for-
mation of chromium carbide and intermetallic phases can make steel weak against corrosion. In this regard, Sigma phase formation is
a typical problem in welding of stainless steels and during working service at high temperatures (in the range of 450 °C to 850 °C).
Sigma phase can precipitate at elevated temperature environments, during casting, rolling, welding, forging, aging, and more im-
portantly during service. Precipitation of Sigma phase, is one of the main causes of deterioration of mechanical properties, corrosion
resistance, and weldability of stainless steels [9]. The Chromium content in the presence of delta ferrite greatly influences the
formation of Sigma phase. Also, Sigma phase can precipitate from austenite phase or austenite–austenite grain boundary with a slow
rate [7,10,11]. In stabilized grades, Sigma phase precipitation is faster than other grades of stainless steel grades (for example as is


Corresponding author.
E-mail address: m.eslami@cc.iut.ac.ir (A. Eslami).

http://dx.doi.org/10.1016/j.csefa.2017.07.001
Received 11 May 2017; Received in revised form 9 July 2017; Accepted 26 July 2017
Available online 01 August 2017
2213-2902/ © 2017 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/BY/4.0/).
M. Ghalambaz et al. Case Studies in Engineering Failure Analysis 9 (2017) 52–62

Fig. 1. Schematic of TTT stabilization curves in Ti stabilization heat treatment [7].

shown in Fig. 2, precipitating in alloy 347 is slightly faster than that in alloy 321) [12,13].
Sigma phase is brittle in nature and therefore it decreases the fracture toughness and ductility of stainless steels especially at low
temperatures (below 250 °C).
This study investigates the causes of failure in 347H austenitic stainless steel pipe, with attention being paid to the formation of
Sigma phase and its influence on the performance of the pipes. Some suggestions are given to decrease the risk of similar failures for
future applications.

2. Background and working conditions

The heart of an Olefin plant is the Ethylene furnace, a plug-flow reactor in which the thermal cracking of hydrocarbons takes
place. Due to its slightly endothermic nature, pyrolysis requires a relatively large amount of heat to sustain the reaction at high
temperature. This heat is supplied by combustion of fuel in the space surrounding the coils, or firebox. The heat is transmitted
primarily by thermal radiation since the radiant coil operates at high temperature. The residual part of fired duty is recovered in the
furnace convection section. This heat is recovered from the flue gas as schematically shown in Fig. 3. In this regard boiling feed water
(BFW) is conveyed to the furnace convection section, and superheated by flue gas in the first High Pressure Steam Superheater
(HPSSH-I) and then in second High Pressure Steam Superheater (HPSSH-II) in series. The temperature leaving HPSSH-II is controlled
by injection of amine-free boiling feed water (BFW) at the VHP Steam De-superheater. The De-superheater is situated in the crossover
pipe between HPSSH-I and HPSSH-II. The superheated VHP steam from all furnaces is combined into a common header and sent the
grid at a temperature of 510 °C and a pressure of 106 bar.
The convection coil located in HPSSH-II consists of countercurrent parallel pipes in form of coils. There are two rows of sixteen
pipes in each coil. Schematic of a coil is shown in Fig. 4. The material of inlet manifold and its transition piece is made from A335
GR.P22, while the pipes and the outlet manifold material are made from stainless 347H.The transition piece of inlet manifold and
pipes were welded with ER-NiCr3filler using GTAW welding method. Stress relieve heat treatment was done on welded components
during construction time at 620 °C for 2 hours for this welding joint (failed welding joint area). The pipe section had worked for
around 60,000 hours h before failure.

Fig. 2. Precipitation of sigma in different grades of austenitic stainless steels [12,13].

53
M. Ghalambaz et al. Case Studies in Engineering Failure Analysis 9 (2017) 52–62

Fig. 3. Schematic of convection area of ethylene furnace in an olefin plant.

Fig. 4. Schematic of a coil with pipes: a) front view b) side view.

Fig. 3 shows Schematic of convection area of ethylene furnace in an olefin plant. In this Figure, FPH stands for feed preheater for
preheating hydrocarbon feed, ECO stands for Economizer for preheating BFW, HTC stands for High Temperature Coil for preheating
hydrocarbon feed + dilution steam, DSSH stands for Diluted steam super heat for preheating dilution steam, and HPSSH stands for
High pressure steam super heat for superheating VHP steam.

3. Analytical techniques

Initial examination of fracture surface of the failed pipe was done by Stereographic Microscope. The surface of the sample were
electrolytic etched in nitric acid solution at 1–2 V for 10 s and their microstructure was observed by optical microscopy
(OM).Specimens were etched in Groesbeck solution at above 75 °C for 4 h to investigate the formation of chromium carbide and

54
M. Ghalambaz et al. Case Studies in Engineering Failure Analysis 9 (2017) 52–62

Fig. 5. Microhardness distribution at different regions of the welded area.

sigma phases. To further evaluate the fracture surface, Scanning Electron Microscopy (SEM) was used. Secondary phases were
chemically analyzed using energy-dispersive X-ray spectroscopy (EDS). Vickers Micro-hardness test with weight 300 gr was per-
formed on different regions of the damaged pipe, in accordance with ISO 6507-1 standard. For phase characterization X-ray dif-
fraction (XRD) technique was used.

4. Results and discussion

4.1. Micro-hardness

Micro-hardness from weld metal, heat affected zone and base metal is shown in Fig. 5b. As can be seen from this Figure, micro-
hardness of HAZ and surrounding area has increased in comparison with the base metal. This could be due to chromium carbide and
secondary phases formation. The difference between the base metal and the weld metal micro-hardness could be due to the dis-
similarities of chemical composition and microstructure in these regions.
Also, as can be seen from Fig. 5b-Line 1, the micro-hardness has increased from the outer side of tube towards the inner side of the
tube in HAZ area of the failed pipe. With crack growth, there would be localized plastic deformation around the growing crack. This
affects the microhardness profile around the crack, i.e: the closer to the crack face; there would be relatively higher hardness.
[14–16]. This explains the sharp drop in microhardness profile along line 1, which could be due to relative distance to the crack face,
as schematically shown in Fig. 5a. Also, as can be seen from Fig. 5b, by moving from weld zone towards the base material (lines 2 and
3), micro-hardness has relatively increased. In this regard, stabilization precipitates (NbC in this case) should have been dissolved in
surrounding areas of the joint during welding process, facilitating formation of chromium carbides. Chromium carbides could grow in
working temperature (450–850 °C), resulting in a drop of hardness in this area. By moving away from the joint area, dissolution of
(NbC) precipitates is lowered, resulting in an increase in hardness. This has also been observed in other studies [5].

4.2. Macro- fractographic observation

As can be seen from the Digital Camera Image shown in Fig. 6 the fracture surface of the failed pipe consists of three sections,
which are crack origin zone, the crack growth zone, and the fracture zone. As can be seen from this Figure, there are two crack origin
regions on fracture surface. These represent regions with surface defects and stresses concentration. The crack origins are located on
the outer surface of the failed pipe. Cracks have grown towards the inner surface up to pipe failure. Also in the crack growth zone

55
M. Ghalambaz et al. Case Studies in Engineering Failure Analysis 9 (2017) 52–62

Fig. 6. Fracture surface of the failed pipe.

beach marks can be clearly seen. Beach marks are indication of fatigue failure mechanism [17]. This signature pattern results from
the crack propagation that takes place during every cycle above the threshold loading. These concentric cracks continue to propagate
until the cross-sectional area is reduced to the point where failure takes place due to the overload [17,18]. Fig. 7a is a digital camera
image showing surface defects after the liquid penetration test, and Fig. 7b is a digital camera image of the failed tube, indicating that
the crack has initiated is in the heat affected zone (HAZ) area of joint and propagated circumferential around the pipe.

4.3. Micro-fractographic observations

Optical microscopic image from the tip of the crack is shown in Fig. 8. As can be seen from this Figure the crack has grown
transgranular with a sharp tip.
Fracture surface of the failed pipe is shown in Fig. 9. As can be seen from this Figure the surface of failed specimen has no dimples
and it is perfectly smooth. The fracture surface can be categorized as brittle fracture. Brittle fracture in stainless steels is usually due to
the existence of brittle phases such as carbides, Sigma and Chi phases [5]. Fig. 10 shows the SEM image of Beach Marks in failure
surface of pipe.
Figs. 11 and 12 show Striations and Ratchet Marks in fracture surface respectively. These marks once again confirm the fatigue
fracture mechanism in the failed pipe [17,19].

4.4. EDS analysis

Chromium rich phase (Sigma phase) were observed at fracture surface (see Fig. 13). The sigma phase is a brittle phase which can
form during the heat treatment and working service at temperature range 450–850 °C. Formation of Sigma phase is normally as-
sociated with early unexpected failures [10,11,20]. High Cr content in this phase is an indication of chromium rich phase formation
(Fig. 14).

4.5. XRD analysis

Some studies have shown that the high temperature water vapor reduces the fatigue life of stainless steels. Anodic dissolution or
hydrogen friability at the crack tip could affect crack growth and propagation in high temperature water vapor. Also, crack initiation
in high temperature water vapor environment is more pronounced than that in high temperature air environment [21]. XRD analysis
from corrosion product on the failed pipe is shown in Fig. 15.
As shown in Fig. 15, phase analysis of surface products on the failed pipe suggests the presence of austenite and Fe3O4 phases
together with C, H, O-containing components. Chromium oxides (Cr2O3) in XRD analysis results were not detected. This could be due
to the consumption of Cr, during the welding process, in the heat affected zone. Formation of chromium oxide layer can cause high

56
M. Ghalambaz et al. Case Studies in Engineering Failure Analysis 9 (2017) 52–62

Fig. 7. a) Digital Camera image after liquid penetration test, showing surface defects on the outer side of tube after liquid penetration test, b) Digital Camera image of
the failed tube, indicating that the crack has initiated is in the heat affected zone (HAZ) area of joint and propagated circumferential around the pipe.

corrosion resistivity in stainless steels [22–24]. Formation of Fe3O4 and C, H, O components on the fracture surface is an indication of
corrosion. Fe3O4 is formed when oxygen activity is low. High temperature water vapor consists of low oxygen. Therefore, formation
of Fe3O4 is more probable (the high temperature water vapor crosses the pipe during the operation) [24]. Investigation on 310S
austenitic stainless steel in supercritical water at 500 °C showed formation of Fe3O4 at high temperature is more probable. After
formation of chromium oxide thin layer (1–2 nm scale), by iron (Fe) diffusion through the chromium layer from inner side to outer
surface, magnetite should have formed [22]. Fig. 16 shows corrosion products in the form of precipitates on the failed surface of the
pipe.

4.6. Failure mechanism

The failure mechanism has been summarized in Fig. 17. As shown in this Figure formation of Sigma phase during welding is the
main cause of failure. In more details, as shown in previous sections the pipe has failed from a crack initiated from the outer surface in
the heat affected zone of the welded pipe, toward the inner surface. Formation of Sigma phase could cause an increased local
hardness and make the pipe more brittle. Also, depletion of chromium by Sigma phase could form local areas, which are not resistant
to corrosion due to a drop in chromium content. It is well established that Sigma phase could form in the temperature range of
450–850 °C, which should have occurred during welding/service condition in the bonded area. Also, as shown in SEM images,
Figs. 10–12, fatigue cracks were observed on the fracture surface. Thermal Fatigue could occur during service conditions, where there
is change in temperature that could generate thermals stresses. The fatigue cracks have initiated from pipe surface (local stress risers
at the surface) and propagated in heat affected zone of the welded joint. This is the region which could have lower toughness, due to
possible formation of Sigma phase, which could accelerate the tube failure (although Sigma phase has not initiate the cracks, but

57
M. Ghalambaz et al. Case Studies in Engineering Failure Analysis 9 (2017) 52–62

Fig. 8. Optical microscopic image, showing transgranular surface crack.

Fig. 9. Fracture surface of 347H austenitic stainless steel.

Fig. 10. SEM image of beach marks in failure surface of 347H austenitic stainless steel.

58
M. Ghalambaz et al. Case Studies in Engineering Failure Analysis 9 (2017) 52–62

Fig. 11. SEM image of striations in fracture surface of 347H austenitic stainless steel.

Fig. 12. SEM image of ratchet marks in crack origin.

Fig. 13. Optical microscopic image of sigma phase precipitated in 347 stabilized stainless steel.

formation of this brittle phase should have facilitated the tube failure [5]).
For future failure prevention of the pipes the following practices are recommended:

• Controlled welding process: welding should be done in controlled condition with high cooling range and controlled chemical
composition.
• Using more resistance pipe material less sensitive to sigma phase formation [12,13].
• Heat treatment of welded region: Heat treatment of welded region above 1050 °C to dissolve Sigma phase, and holding at 950 °C,
in order to form Nb rich carbides. By formation of Nb riched carbides, the C content would be below the required limit for
formation of brittle phases such as the Sigma phase during the cooling process.

59
M. Ghalambaz et al. Case Studies in Engineering Failure Analysis 9 (2017) 52–62

Fig. 14. EDS analysis from sigma phase in AISI 347H austenitic stainless steel after 6 month working, a) SEM image of precipitation, b and c) precipitation analysis.

Fig. 15. X–Ray diffraction pattern of pipe fracture surface.

5. Conclusions

1. Pipe failure investigated in this study initiated from the outer pipe surface with brittle fracture surface appearance,
2. Crack growth and propagation had occurred by a thermal fatigue mechanism, within the HAZ
3. Sigma phase was observed in HAZ of the failed tube. This should have facilitated tube failure. Sigma phase could form at
450–850 °C temperature range during the heat treatment after the welding and also during service condition, reducing the
toughness of HAZ, and facilitating crack growth process.
4. Controlled welding process, using more resistance pipe material, or heat-treatment of welded region could be some remedies for

60
M. Ghalambaz et al. Case Studies in Engineering Failure Analysis 9 (2017) 52–62

Fig. 16. SEM image of corrosion product on fracture surface.

Fig. 17. Summary of failure mechanism: a) welding process; b) heat treatment process; c) Sigma phase formation; d) Crack initiation and growth.

future failure prevention.

Acknowledgments

The authors would like to thank Isfahan University of Technology and Jam Petrochemical Complex for their support.

References

[1] Hajiannia I, Shamanian M, Kasiri M. Microstructure and mechanical properties of AISI 347 stainless steel/A335 low alloy steel dissimilar joint produced by gas
tungsten arc welding. Mater Des 2013;50:566–73.
[2] Özyürek D. An effect of weld current and weld atmosphere on the resistance spot weldability of 304L austenitic stainless steel. Mater Des 2008;29(3):597–603.
[3] Lee HS, Jung JS, Kim D-S, Yoo K-B. Failure analysis on welded joints of 347H austenitic boiler tubes. Eng Fail Anal 2015;57:413–22.
[4] Lima A, Nascimento A, Abreu H, de Lima-Neto P. Sensitization evaluation of the austenitic stainless steel AISI 304L, 316L, 321 and 347. J Mater Sci

61
M. Ghalambaz et al. Case Studies in Engineering Failure Analysis 9 (2017) 52–62

2005;40(1):139–44.
[5] Guan K, Xu X, Xu H, Wang Z. Effect of aging at 700 °C on precipitation and toughness of AISI 321 and AISI 347 austenitic stainless steel welds. Nucl Eng Des
2005;235(23):2485–94.
[6] Jahromi SJ, Javadpour S, Gheisari K. Failure analysis of welded joints in a power plant exhaust flue. Eng Fail Anal 2006;13(4):527–36.
[7] Moura V, Kina AY, Tavares SSM, Lima L, Mainier FB. Influence of stabilization heat treatments on microstructure, hardness and intergranular corrosion re-
sistance of the AISI 321 stainless steel. J Mater Sci 2008;43(2):536–40.
[8] Thorvaldsson T, Dunlop G. Grain boundary Cr-depleted zones in Ti and Nb stabilized austenitic stainless steels. J Mater Sci 1983;18(3):793–803.
[9] Hsieh CC, Wu W. Overview of intermetallic sigma (σ) phase precipitation in stainless steels. ISRN Metall 2012;2012.
[10] Martins M, Casteletti LC. Sigma phase morphologies in cast and aged super duplex stainless steel. Mater Charact 2009;60(8):792–5.
[11] Schwind M, Källqvist J, Nilsson JO, Ågren J, Andrén HO. σ-phase precipitation in stabilized austenitic stainless steels. Acta Mater 2000;48(10):2473–81.
[12] Minami Y, Kimura H, Ihara Y. Microstructural changes in austenitic stainless steels during long-term aging. Mater Sci Technol 1986;2(8):795–806.
[13] Sourmail T. Precipitation in creep resistant austenitic stainless steels. Mater Sci Technol 2001;17(1):1–14.
[14] Grabulov A, Petrov R, Zandbergen HW. EBSD investigation of the crack initiation and TEM/FIB analysesof the microstructural changes around the cracks formed
under rolling contact fatigue (RCF). Int J Fatigue 2010;32:576–83.
[15] Hu Z, Lu W, Thouless MD, Barber JR. Effect of plastic deformation on the evolution of wear and local stressfields in fretting. Int J Solids Struct 2016;82:1–8.
[16] Kida K, Ishida M, Furuse M, Mizobe K, Santos EC. Effect of plastic deformation on magnetic fields around fatigue crack tipsof carbon tool steel (JIS, SKS93). Int J
Fatigue 2016;88:156–65.
[17] Pantazopoulos G, Papaefthymiou S. Failure and fracture analysis of austenitic stainless steel marine propeller shaft. J Fail Anal Prev 2015:1–6.
[18] Wu X, Shi J, Ding N, Guo W, Xu N, Zang Q, et al. Fracture failure of 304 stainless steel connectors on the isolating switches. J Fail Anal Prev 2015;15(3):364–9.
[19] Maruschak P, Panin SV, Stachowicz F, Danyliuk I, Vlasov IV, Bishchak R. Structural levels of fatigue failure and damage estimation in 17Mn1Si steel on the basis
of a multilevel approach of physical mesomechanics. Acta Mech 2016;227(1):151–7.
[20] Yousefi M, Farghadin MH, Farzadi A. Investigate the causes of cracks in welded 310 stainless steel used in the flare tip. Eng Fail Anal 2015;53:138–47.
[21] Wu H, Yang B, Wang S, Zhang M. Effect of oxidation behavior on the corrosion fatigue crack initiation and propagation of 316LN austenitic stainless steel in high
temperature water. Mater Sci Eng: A 2015;633:176–83.
[22] Wu H, Yang B, Wang S, Zhang M. Effect of oxidation behavior on the corrosion fatigue crack initiation and propagation of 316LN austenitic stainless steel in high
temperature water. Mater Sci Eng: A 2015;633:176–83.
[23] Behnamian Y, Mostafaei A, Kohandehghan A, Amirkhiz BS, Serate D, Zheng W, et al. Characterization of oxide scales grown on alloy 310S stainless steel after
long term exposure to supercritical water at 500 °C. Mater Charact 2016;120:273–84.
[24] Hooshyar H, Jonsson T, Hall J, Svensson JE, Johansson L, Liske J. The effect of H2 and H2O on the oxidation of 304L-Stainless steel at 600 °C: general behaviour
(part I). Oxid Met 2016;85(3–4):321–42.

62

You might also like