Sei sulla pagina 1di 13

Engineering Failure Analysis 11 (2004) 575–587

www.elsevier.com/locate/engfailanal

Effect of welding processes on toe cracking behaviour


of pressure vessel grade steel
V. Balasubramaniana,*, B. Guhab
a
Department of Manufacturing Engineering, Annamalai University, Annamalainagar 608 002, Tamil Nadu, India
b
Department of Metallurgical Engineering, Indian Institute of Technology Madras, Chennai 600 036, India

Received 8 September 2003; accepted 25 September 2003

Abstract
The effect of welding processes on fatigue crack growth behaviour of load carrying cruciform joints has been
analysed. Cruciform joints were fabricated from pressure vessel grade (ASTM 517 ‘F’ grade) steel using shielded metal
arc welding (SMAW) and flux cored arc welding (FCAW) processes. Fatigue crack growth experiments were carried
out in a mechanical resonance vertical pulsator (SCHENCK 200 kN capacity) with a frequency of 30 Hz under
constant amplitude loading ðR ¼ 0Þ. It was found that the toe crack growth rates were relatively lower in the joints
fabricated by SMAW process than the joints fabricated by FCAW process. The heat affected zone (HAZ) region of
SMAW joints contains a low carbon martensitic structure and exhibited better fatigue resistance compared to the
bainitic HAZ microstructure of FCAW joints. Relatively higher heat input involved in FCAW process resulted in the above
variation in HAZ microstructure and led to inferior fatigue performance of FCAW joints compared to SMAW joints.
# 2004 Elsevier Ltd. All rights reserved.
Keywords: Weld fatigue; Cruciform joint; Stress concentrations; Fatigue crack growth; Welding process

1. Introduction

Fillet welded cruciform joints are widely used in various structures including offshore and nuclear
installations. Linking the effects of weld defects and failure analysis of weldments suggests that fatigue
alone accounts for most of the disruptive failures and often precedes the onset of brittle failure [1]. The
fatigue resistance of weld metal and heat affected zone (HAZ) of various steels are better than or equal to
those of the base metal. However, problems arise when there is an abrupt change in section by excess weld
reinforcement, undercut, inclusion of slag or lack of penetration or fusion [2]. The fatigue crack growth
behaviour of welded joints depends on the material, loading and in particular, the geometric configurations
of the weld and plate [3].

* Corresponding author. Tel.: +91-4144-239734 (O), +91-4144-241147 (R); fax: +91-4144-238080.


E-mail address: visvabalu@yahoo.com (V. Balasubramanian).

1350-6307/$ - see front matter # 2004 Elsevier Ltd. All rights reserved.
doi:10.1016/j.engfailanal.2003.09.005
576 V. Balasubramanian, B. Guha / Engineering Failure Analysis 11 (2004) 575–587

Two types of cracking normally will cause failure of a fillet welded joint. They are: (i) root cracking and
(ii) toe cracking. Fatigue cracks initiate at a fillet weld toe when the fillet weld size is large enough and
initiate at a weld root when weld size is inadequate [4]. The toe region is highly susceptible to fatigue
loading due to the presence of weld defects, stress concentration due to toe geometry and adverse metal-
lurgical conditions such as tensile residual stresses and coarse HAZ microstructure. Moreover, the toe
region also experiences high dilution and thermal stresses depending upon the type of welding process and
procedures adopted during fabrication [5].
Automatic welding processes are favoured over manual processes for the fabrication of welded joints for
a number of reasons. Among these are increased productivity, lower costs, a better control of geometry
and access to weld in difficult environments. However, it has been observed that the service lives of these
automatic welds are usually shorter than that of their manual counterparts against fatigue loading [6]. A
twofold increase in fatigue lives was observed for manual welds as compared to automatic welds. This was
attributed mainly to a higher stress concentration arising from a poor weld profile caused by the automatic
processes such as submerged arc welding (SAW) [7].
Many investigators [8–11] have studied the effect of welding process and parameters on the fatigue
strength of fillet welded joints and they compared fatigue performances of fillet welds fabricated either by
manual or by fully automatic welding processes. However, in recent years, the flux cored arc welding
(FCAW) process is becoming more popular due to higher deposition rate and a better weld quality as
compared to the shielded metal arc welding (SMAW) process, at the same time exhibiting good weld metal
toughness similar to the SAW process [12]. As a consequence, it was decided to compare the fatigue per-
formances of cruciform joints fabricated by SMAW (manual) and FCAW (semi-automatic) processes to
study the suitability of any one of the processes for industrial applications.

2. Experimental

A high strength quenched and tempered steel (ASTM 517 ‘F’ grade) of weldable quality in the form of
rolled plates of 8 mm thickness has been used as the base material throughout the investigation. This
material is widely used for welded constructions of all kinds such as pressure vessels, penstocks, bridges
and structures as well as transport vehicles, hoisting and earthmoving equipment utilized in different
climatic conditions [13]. The rolled plates were cut into the required sizes and profiles by oxy-fuel cutting
and grinding. The initial joint configuration was obtained by securing the long plate (300  100 mm) and
the stem plate (300  75 mm) in a cruciform position using tack welding. Subsequently, fillets were made
between the long plate and stem plate by laying weld metal using SMAW and FCAW processes with
strength matching consumables. All necessary care was taken to avoid joint distortion and the joints were
made without applying any clamping devices. The chemical composition and mechanical properties of base
metal and weld metals are presented in Table 1.
Full penetration cruciform joints (without lack of penetration) were fabricated to enable to study the
crack growth behaviour and subsequently to evaluate the fatigue life of the joints failing from toe region
alone. These joints were fabricated without leaving an unfused gap between each pair of fillets by having
sharp root faces obtained by prior machining. The fillet toe angle was maintained at 45 by controlling
electrode to work piece angle and arc length. Then the cruciform specimens were sliced from the joints to
the dimensions (as shown in Fig. 1). The welding conditions and process parameters used in the fabrication
of cruciform joints are given in Table 2.
The fatigue crack growth experiments were conducted using a vertical pulsar of 200 kN capacity (CARL
SCHENCK – Type: PVT/N) with a frequency of 30 Hz under constant amplitude loading ðR ¼ 0Þ. Before
loading, the specimen surface was polished using metallographic procedures and illuminated suitably to
enable the crack growth measurement. A traveling microscope (Make: WESWOS) was used to monitor the
Table 1

V. Balasubramanian, B. Guha / Engineering Failure Analysis 11 (2004) 575–587


Chemical composition of base metal and weld metals (Panel A), mechanical properties of base metal and weld metal (Panel B)

Elements C Si Mn P S Cr Mo Ni Cu Co V Fe

Panel A
Base metal 0.19 0.72 0.95 0.01 0.002 0.8 0.35 0.007 0.03 0.004 0.002 Balance
SMA weld metal 0.08 0.5 1.65 0.010 0.030 0.35 0.27 1.73 0.10 0.008 0.012 Balance
FCA weld metal 0.08 0.4 1.50 0.007 0.015 0.56 0.44 2.25 0.19 0.006 0.005 Balance

Yield Tensile Elongation Vickers Charpy


strength strength (%) hardness impact energy
(MPa) (MPa) (30 kg) at room temperature (J)

Panel B
Base metal 690 790 19 250 110
SMA weld metal 710 810 16.5 285 175
FCA weld metal 700 800 15 270 150

577
578 V. Balasubramanian, B. Guha / Engineering Failure Analysis 11 (2004) 575–587

Fig. 1. Cruciform test specimen.

Table 2
Welding conditions and process parameters

Parameter SMAW FCAW

Welding machine DC generator LINDE


Electrode/filler AWS E110 18-M AWS E100-T5K4
Electrode diameter (mm) 3.15 1.6
Arc voltage (V) 24 30
Welding current (A) 120 200
Welding speed (mm/s) 2.2 3.0
Heat input (kJ/mm) 1.3 2.0
Preheat temperature ( C) 150 150
Interpass temperature ( C) 250 250
Root run Using 2.5 mm dia electrode by SMAW process Using 2.5 mm dia electrode by SMAW process
Gas flow rate (l/min) – 1.2

crack growth with an accuracy of 0.01 mm [14]. After the specimen was gripped between the upper and
lower grips, the pulsating load was applied to the specimen. The crack initiation from the toe region for an
initial crack length of 0.3 mm and corresponding number of cycles was recorded as the crack initiation life.
The subsequent crack propagation at each 0.5 mm interval and corresponding number of cycles was
recorded until final failure of the specimen.

3. Fracture mechanics analysis

The fatigue crack growth experiments were conducted at four stress levels of 120, 160, 200 and 240 MPa.
All the specimens were tested under constant amplitude loading condition ðR ¼ 0Þ. The variations in toe
V. Balasubramanian, B. Guha / Engineering Failure Analysis 11 (2004) 575–587 579

crack length (a) and the corresponding number of cycles (N), obtained from fatigue crack growth
experiments, are plotted as shown in Fig. 2 for SMAW and FCAW joints. The fracture mechanics analysis
is based on the Paris equation [15] given below:

da=dN ¼ CðDKÞm ; ð1Þ

where da=dN is the crack growth rate, DK the stress intensity factor (SIF) range, and C and m are constants.

Fig. 2. Crack growth curves.


580 V. Balasubramanian, B. Guha / Engineering Failure Analysis 11 (2004) 575–587

Eq. (1) can be normalized and adopted for cruciform joints [16], in the form given below:

dða=Tp Þ CðDKÞm D m
¼ ; ð2Þ
dN D m Tp

where D is the nominal stress range, a the toe crack length, and Tp the plate thickness.
The SIF range ðDKÞ, for a crack emanating from the toe region is expressed as given below [17]:

Ms Mt Mk
DK ¼ D½pa1=2 ; ð3Þ
0

where Ms , Mt , 0 and Mk are correction factors (greater than unity) that correspond to (i) free surface of
the crack, (ii) finite plate thickness, (iii) crack front shape, (iv) stress concentration due to toe apex,
respectively.
The crack growth rate, da=dN for the propagation stage (between 106 and 103 mm/cycle), was
calculated at different intervals of crack length increment. The relationship between SIF range ðDKÞ and
the corresponding crack growth rate dðaÞ=dN on a log–log scale, in terms of the best fit line, is shown in
Fig. 3 for both the joints. Most of the data points correspond to the second stage (linear region) of the
sigmoidal relationship between da=dN and DK. The exponent m, which is the slope of the line on the log–
log plot is found to be 4.3 and 4.8 for SMAW and FCAW joints, respectively. The value of the constant C,
which is the intercept of the line on the log–log plot, is found to be 4.3  1011 and 2.4  1011 for the
SMAW and FCAW joints, respectively. From Fig. 3 it can be seen that the fatigue crack growth rate,
da=dN, is relatively less in SMAW joints than in FCAW joints. When the crack growth rate is around 103
mm/cycle, the curve tends to become parallel to Y-axis and the corresponding DK value is taken as critical
SIF range ðDKcr Þ. At lower values of DK (below 106 mm/cycle), the curve again becomes parallel to the
Y-axis indicating a threshold SIF range ðDKth Þ below which a crack may not propagate under fatigue loading
conditions. The values of DKcr and DKth for both the joints have been evaluated and presented in Table 3.

Fig. 3. Crack growth rate curves.


V. Balasubramanian, B. Guha / Engineering Failure Analysis 11 (2004) 575–587 581

Table 3
Fatigue crack growth parameters
p p
Joint m C DKth (MPa m) DKcr (MPa m) De (MPa)
11
SMAW 4.3 4:3  10 9.0 50 60
FCAW 4.8 2:4  1011 8.0 40 45

Normally, in steels, the threshold values are obtained at a crack growth rate of 108 mm/cycle. Because of
the specimen configuration and loading conditions, crack propagation rates in the region of 108 mm/cycle
could not be obtained and hence the value obtained in this analysis cannot be taken as the design value.
The fatigue crack growth (fracture mechanics) parameters of the two joints are compared in Table 3.
The S–N behaviour (fatigue life) of the two joints is depicted in Fig. 4. The endurance stress range ðDe Þ
corresponding to the failure life of 107 cycles for SMAW and FCAW weldments are 60 and 45 MPa,
respectively. The crack initiation life is also evaluated experimentally using ‘crack initiation criteria’. The
criterion is based on the number of cycles required for growing a toe crack to 0.3 mm length at the earlier
growth stage under a particular stress level. Otegui et al. [18] adopted a similar criterion to analyse the
fatigue crack growth behaviour of fillet welded joints. From Fig. 5, it is evident that welding processes
significantly influence the crack initiation life of the joints failing from the toe region.

4. Metallurgical analysis

The microstructures of the HAZ region of the SMAW and FCAW cruciform joints, from where the toe
crack has initiated and propagated has been analysed using an optical microscope and are depicted in
Figs. 6 and 7, respectively. Figs. 6(a) and 7(a) display the optical micrograph of WM–HAZ interface region

Fig. 4. S–N curves.


582 V. Balasubramanian, B. Guha / Engineering Failure Analysis 11 (2004) 575–587

Fig. 5. Relationship between DKi and Ni.

Fig. 6. Microstructure of the SMAW joint along toe crack path. (a) WM–HAZ interface, (b) CGHAZ, (c) FGHAZ.
V. Balasubramanian, B. Guha / Engineering Failure Analysis 11 (2004) 575–587 583

Fig. 7. Microstructure of the FCAW joint along toe crack path. (a) WM–HAZ interface, (b) CGHAZ, (c) FGHAZ.

of SMAW and FCAW cruciform joints, respectively, at lower magnification. The microstructure of the
coarse grain heat affected zone (CGHAZ) of the SMAW joints was composed of low carbon martensite
with some bainitic regions. The martensitic structure in Fig. 6(b) consists of elongated packets, which
formed in bundles within prior austenite grains. However, the CGHAZ of the FCAW joints contained
entirely coarse bainite. The bainitic structure, shown in Fig. 7(b) consists of parallel bainitic ferrite laths
separated by elongated martensite–austenite (MA) regions. The fine grain heat affected zone (FGHAZ)
microstructure of SMAW and FCAW joints are depicted in Figs. 6(c) and 7(c), respectively. From the
photomicrographs, it can be seen that there is a small difference in the grain size of the FGHAZ, i.e., the
FCAW joints contain coarse grains than SMAW joints in FGHAZ region.
The fracture surface appearance corresponding to crack initiation, crack propagation and the final
failure regions of SMAW and FCAW joints, as observed under the scanning electron microscope (SEM), is
shown in Figs. 8 and 9, respectively. Here, the fatigue crack initiation (FCI) region is corresponding to 0.3
mm from the toe region; fatigue crack propagation (FCP) region is referred to 0.3–3.5 mm; final failure
(FF) region is 4 mm away from the crack initiation region. It is evident from the fractographs that the
fracture facet size corresponded to the dimensions of the individual martensite packets. The finer the
packet size, the higher the resistance to fatigue crack growth. The large facet size of the crack initiation
region related to the bainitic packet size. The most important factor is that fatigue cracks can easily
propagate across a bainite packet with little resistance encountered at low angle boundaries within a
packet. However, not much difference has been observed in the final failure regions of both the joints, since
unstable crack growth occurred in these regions.
584 V. Balasubramanian, B. Guha / Engineering Failure Analysis 11 (2004) 575–587

Fig. 8. SEM fractograph of SMAW joint viewed at different regions. (a) FCI region, (b) FCP region, (c) FF region.

5. Discussion

From the experimental results presented in the preceding sections, it can be understood that the welding
processes are having significant influence on the toe cracking behaviour and fatigue life of the cruciform
joints. It has been found from the investigation that the SMAW cruciform joints (without LOP) are
enduring more cycles as compared to their FCAW counterparts as can be revealed from Figs. 2 and 4.
Moreover, crack initiation is delayed in SMAW joints and subsequently the crack initiation life is longer
as compared to FCAW joints (Fig. 5). Further, the crack growth rate is slower in SMAW joints when
compared to FCAW counterparts, in the crack propagation stage as can be clearly seen from Fig. 3
The superior fatigue performance of SMAW joints could be due to the difference in HAZ microstructure
at the toe region (Figs. 6 and 7) and is caused by the variations in heat input involved in the fabrication of
the joints. The CGHAZ microstructure of the SMAW joints contains low carbon martensite and is pro-
duced by low heat input involved in the process. But, the higher heat input involved in FCAW process
caused the formation of a bainitic structure in the CGHAZ region. Normally, the low carbon martensitic
V. Balasubramanian, B. Guha / Engineering Failure Analysis 11 (2004) 575–587 585

Fig. 9. SEM fractograph of FCAW joint viewed at different regions. (a) FCI region, (b) FCP region, (c) FF region.

microstructure will have better notch impact toughness than the bainitic microstructure and this may be
attributed to the delayed initiation of fatigue cracks from the toe region of the SMAW joints. The crack
growth rate in the propagation region of the SMAW and FCAW joints also exhibits a small difference and
this could be due to the minor difference in FGHAZ microstructure. Similar observations were made by
other investigators also [16,19]. Gooch and Ginn [20] also observed that the HAZ toughness of 12% Cr
martensitic–ferritic steels was dependent mainly on the peak ferrite grain size produced by SAW and hence
they recommended low arc energy for optimum HAZ toughness.
Chapetti and Otegui [6] reported that the lower fatigue growth rates for manual welds are the result of
higher degree of irregularity of the crack path, which in turn is related to the irregularity or waviness of the
weld toe. Greater toe waviness induces a larger mismatch between the planes of small surface cracks pro-
ducing effects of load shedding and delaying coalescence. Because of the more uniform notch severity at
the toe, fatigue cracks in automatic welds have much higher surface growth rates and coalesce sooner into
low aspect ratios than do cracks in manual welds [21]. FCAW is a semi-automatic process and therefore,
the resulting toe waviness is lower compared to SMAW, which is a manual process. This may also be one
of the reasons for the delayed crack initiation in SMAW joints.
586 V. Balasubramanian, B. Guha / Engineering Failure Analysis 11 (2004) 575–587

6. Conclusions

The toe cracking behaviour of SMAW and FCAW load carrying cruciform joints has been analysed in
detail. From this investigation, the following conclusions have been derived:

(i) The fatigue crack growth behaviour and subsequently the fatigue life of full penetration, load
carrying cruciform joints is found to be influenced by the welding processes utilised to fabricate the
joints.
(ii) The experimental evidence showed that the microstructural features of the HAZ region play an
important role in toe cracking behaviour and influenced the fatigue life of the joints. The HAZ
region of SMAW joints contains a low carbon martensitic structure and exhibited better fatigue
resistance compared to the bainitic HAZ microstructure of FCAW joints.
(iii) Higher crack initiation life and lower fatigue crack growth rates of SMAW joints are probably due
to the differences in metallurgical properties of the weldments caused by the process parameters
employed in two processes. The relatively high heat input involved in the FCAW process resulted in
inferior fatigue performance of these joints compared to SMAW joints.

Acknowledgements

The authors are grateful to the Department of Manufacturing Engineering, Annamalai University for
the support rendered and making available the Metal Joining Laboratory to prepare the test specimens.
The authors are also thankful to the Department of Metallurgical Engineering, Indian Institute of Tech-
nology, Madras for giving permission to conduct fatigue crack growth test using Mechanical Testing
Laboratory. The authors are indebted to M/s. Bharat Heavy Electricals Limited (BHEL), Ranipet, Tamil
Nadu for the material (ASTM 517 ‘F’ grade steel) supplied to carryout investigation.

References

[1] Tsai CL. Fitness for service design of fillet welded joints. J Struct Eng 1986;112(8):1761–80.
[2] Blodgett OW. Designing weldments for fatigue loading. Welding J 1992;7:39–42.
[3] Branco CM, Ferreira JAM, Randon JC. Fatigue of fillet welded joints. Theo Appl Frac Mech 1985;3:13–22.
[4] Maddox SJ. Recent advances in the fatigue assessment of weld imperfections. Welding J 1992:39–44.
[5] Veereman Y, Bailon JP, Masaunove. Fatigue life prediction of welded joints – a reassessment. Fatigue Fract Eng Mater Struct
1987;10:17–36.
[6] Chapetti MD, Otegui JL. Importance of toe irregularity for fatigue resistance of automatic welds. Int J Fatigue 1995;17:531–8.
[7] Gurney TR. Fatigue of welded structures. 2nd ed. Cambridge: Cambridge University Press; 1979.
[8] Lassen T. Effect of welding processes on fatigue crack growth. Welding J 1990;69:75s–81s.
[9] Overbeke JL, Alting AJ, Jonkers PAM. Influence of weld parameters on fatigue strength under CA and VA loading of welded
joints in HSLA steels. IIW Doc. 1991; No. IIW-XIII-1404-91.
[10] Rading GO. Effect of welding on the fatigue crack growth rate in a structural steel. Welding J 1993;71:307s–11s.
[11] Bartosiewicz L, Krause AR, Sengupta A, Putatunda SK. A new model for fatigue threshold. Eng Fract Mech 1993;45:463–77.
[12] Parmar RS. Welding processes and technology. 2nd ed. New Delhi: Khanna Publishers; 1999.
[13] Crooker TW, Lange EA. Low cycle fatigue crack propagation in A201B, A3028 and A517F pressure vessel steels. Welding J
1971;7:322s–8s.
[14] Balasubramanian V, Guha B. Effect of L=Tp ratio on fatigue life prediction of SMAW cruciform joints of ASTM 517 ‘F’ grade
steels. Int J Pres Vess Piping 1998;74:907–18.
[15] Paris PC, Erdogan F. A critical analysis of crack propagation laws. J Basic Eng 1963;85:528–38.
[16] Guha B. Fatigue life estimate of C–Mn steel welded cruciform joints. Theo Appl Fract Mech 1994;21:121–9.
[17] Maddox SJ. An analysis of fatigue cracks in filler welded joints. Int J Fract 1975;11:221–43.
V. Balasubramanian, B. Guha / Engineering Failure Analysis 11 (2004) 575–587 587

[18] Otegui JL, Burns DJ, Kerr HW, Mohaupt UH. Growth coalescence of fatigue cracks at weld toes in steel. Int J Pres Ves Piping
1991;48:129–45.
[19] Krishnadev MR, McGrath JT, Bowker JT, Dionne S. HAZ microstructure and toughness of precipitation strengthened HSLA
steels. In: Proceedings of International Symposium on Welding Metallurgy of Structural Steels, Denver, USA, February 1987.
p. 145–62.
[20] Gooch TG, Ginn BJ. Heat affected zone toughness of SMA welded 12% Cr martensitic–ferritic steels. Welding J 1990;69:431s–
40s.
[21] Otegui JL, Mohauput UH, Burns DJ. Effect of weld process on early growth of fatigue on steel T-joints. Int J Fatigue 1991;13:
45–58.

Potrebbero piacerti anche