Вы находитесь на странице: 1из 6

Available online at www.sciencedirect.

com

Procedia Engineering 55 (2013) 408 413

6th International Conference on Cree


ep, Fatigue and Creep-Fatigue Interaction [CF-6
6]

Effect of Joining Processs on the Accumulation of Creep


Deformation and Cavvitation across the Weld joint of
316L(N
N) Stainless Steel
T. Sakthivel, M. Vasudevan, K. Laha, P. Parameswaran, K. S. Chandravathii,
M. D. Maathew, A. K. Bhaduri
Metallurgy and Materials Group, Indira Gaandhi Centre for Atomic Research, Kalpakkam, India 603102

Abstract
Detailed analysis of the progressive accumulation of creep strain and cavitation across the joints were carried out on
performing interrupted creep tests on joint specimenss fabricated by single-pass activated TIG (A-TIG) and multi-paass TIG
(MP-TIG) welding processes having grid marking insscribed on them before creep test. The localized creep deformation rate
in the weld metal was faster in the MP-TIG joint thann in the A-TIG joint. Creep cavitation was more pronounced in the
t MPTIG joint. The different deformation and cavitation responses of the joints have been attributed to -ferrite conteents and
higher strength disparity within the weld metal.
2013
Authors.
Published
by Elsevier
Open access
CC BY-NC-ND

2013The
The
Authors.
Published
by Ltd.
Elsevier
Ltd.under
Selection
and/orlicense.
peer-review under responsibility of the Indira
Selection and
peer-review
under responsibility
Gandhi
Centre
for Atomic
Research. of the Indira Gandhi Centre for Atomic Research.

Keywords: 316L(N) stainless steel; localized creep deformattion; cavitation; activated TIG; multi-pass TIG

1. Introduction
Austenitic stainless steel (SS) type 316L(N) is being used as a major structural material for Sodium Cooled
C
Fast Reactors (SFR). The selection of the mateerial is based on its good combination of strength and du
uctility;
and its high resistance to oxidation / corrosion and sensitization. Generally, fusion welding is widely used
u
to
fabricate the reactor vessel and piping due to their complexity and size involved. Cracking in the weld jo
oints is
the life limiting factor for welded components [ 1]. Tungsten Inert Gas (TIG) welding process with and without
w
filler metal is generally employed to join thiss class of material. The major disadvantages of TIG welding
w
process lie in the limited thickness of the materiial which can be welded in a single pass, poor tolerance to
o some
material composition and the low productivity. The penetration capability of the arc in TIG welding can be
significantly increased by application of a fluxx coating containing certain inorganic compounds on th
he joint
surface prior to welding [2 - 12]. It has been claimed that the activated-TIG (A-TIG) welding proceess can

Corresponding Author:
E-mail address: tsakthivel@igcar.gov.in

1877-7058 2013 The Authors. Published by Elsevier Ltd. Open access under CC BY-NC-ND license.
Selection and peer-review under responsibility of the Indira Gandhi Centre for Atomic Research.
doi:10.1016/j.proeng.2013.03.272

T. Sakthivel et al. / Procedia Engineering 55 (2013) 408 413

achieve a full penetration up to 12 mm in a single pass without using a bevel preparation or filler wire in
comparison with conventional TIG welding (Multi-pass TIG) process [10-12]. This has been attributed to the
constriction of the arc as well as reversal of marangoni flow in the molten weld pool [10]. In the present
investigation, localized creep deformation and cavitation behavior of 316L(N) austenitic stainless steel weld
joints fabricated both by single-pass activated TIG (A-TIG) and multi-pass TIG (MP-TIG) welding process
have been reported.
2. Experimental details
Austenitic stainless steel type 316L(N) plates of 6 mm thickness were butt welded by single-pass activated
TIG welding process after applying a thin coating of multi-component flux in the form of paste on surfaces of
the joint before welding. Conventional TIG welding process was used to join the plates of 6 mm thickness
using 1.6 mm diameter matching filler wire. Both the A-TIG and MP-TIG weld pads were X-ray radiographed
to ensure its soundness. Creep specimens with dimensions of 50 mm gauge length, 9 mm width, 4.5 mm
thickness were fabricated from the cross weld joints. Constant load creep tests in air were carried out on the
weld joints at 923 K and 180 MPa stress with interruption of 0.2, 0.5, 0.8, 1.0 time to rupture (tr). Temperature
was controlled to within 2 K along the gauge length of the specimen during the creep test. Square grid
markings were inscribed on both the weld joint specimens. Creep tests were carried out on these specimens to
monitor the variation of creep deformation across the joint during creep exposure by measuring the dimension
changes of the square grids. Vickers microhardness test, optical microscopy studies were carried out on the
joints.
3. Results and discussion
3.1. Variations of microstructure and hardness across the weld joints
Austenitic stainless steel type 316L(N) base metal had an equiaxed grain structure with grain size of around
75 m. In both the joints, weld metals consisted of columnar and equiaxed grain structures with -ferrite
(Fig.1). Distribution of both the columnar and equiaxed grains in the weld metal and orientation of the
columnar grain along with -ferrite with respect to welding direction were found to be appreciably different
with welding techniques adopted. Adjacent to the base metal, the columnar grains and -ferrite were oriented
nearly transverse to the welding direction in the A-TIG weld joint, whereas in the case of MP-TIG weld joint
columnar grains and -ferrite oriented more or less towards short transverse (across the plate thickness) to the
welding direction (Figs.1 and 2). These preferential alignments were due to the maximum thermal gradient
during solidification across the transverse and short transverse to the welding direction in the A-TIG and MPTIG weld joint respectively. Coarser columnar and equiaxed grain microstructures were observed in the A-TIG
joint than in the conventional MP-TIG joint (Fig.1), because of higher heat input during welding in the A-TIG
joint than that MP-TIG joint. Equiaxed grain size in the weld metal of A-TIG and MP-TIG weld joints were
about 230 m and 125 m respectively. Similarly, columnar grain sizes of both the weld joints were
comparable. In the A-TIG weld joint, equiaxed zone of width 850 m was found to grow parallel to the
welding direction at the central location of the weld metal and it was sandwiched between columnar grains.
Discontinuous formations of columnar and equiaxed structure were observed across the weld metal in the MPTIG joint (Fig.1).

409

410

T. Sakthivel et al. / Procedia Engineering 55 (2013) 408 413

Fig.1. Microstructures of the weld metal of (a)


( A-TIG and (b) MP-TIG weld joints in the as-weld condition.

Fig.2. Microstructures of weld metal of (aa) A-TIG and (b) MP-TIG weld joint in the as weld condition.

Fig.3. (a) Variations of microhardness across the


t as-welded A-TIG and MP-TIG and (b) grid marking inscribed
acrosss the weld joint specimen.

Higher content of -ferrite was observed in MP-TIG


M
joint (6.2 FN) than in the A-TIG joint (1.3 FN). This is
due to the higher content of ferrite stabilizing element chromium, lower content of austenite stabilizing element
nickel and higher cooling rate because of low heat input in the MP-TIG weld joint resulting in the stabilization
of more amount of -ferrite. The variations of haardness across the A-TIG and MP-TIG weld joints are shown in
Fig.3. Hardness variation across the joints is refflected in the microstructural variations. The MP-TIG welld joint

T. Sakthivel et al. / Procedia Engineering 55 (2013) 408 413

displayed higher hardness variations across it in comparison with the A-TIG weld joint due to the presence of
more inhomogeneous structure across the weld metal in the MP-TIG joint (Fig.1).
3.2. Creep deformation and creep cavitation damage of the joints
Higher creep rupture strength has been observed in the A-TIG weld joint than the MP-TIG weld joint
(Fig.4). The difference in rupture strength between the base metal and weld joint was found to increase with
creep exposure in the case of MP-TIG joint. Failure in both the weld joints occurred in the weld metal.
Variations in the accumulation of creep strain across the different constituents of the weld joints with creep
exposure were assessed by conducting interrupted creep tests on the joint specimens with grid marking (Fig.3
(b)) inscribed on them prior to creep testing and measuring the change in dimensions of the grids on preplanned
interruptions using optical microscope. Figure 5 illustrates the longitudinal and transverse creep strain
accumulations across the A-TIG and MP-TIG weld joint creep tested at 923 K and 190 MPa. In both the weld
joints, creep deformation was found to concentrate progressively in the weld metal and the equiaxed zone of
the weld metal was found to deform at higher rate than the columnar zone (Fig.5). This might be due to the
different dislocation sub-structures of the two micro-constituents of the weld metal. This is also reflected in the
hardness profile across the weld joints (Fig.3). Presence of more open bcc -ferrite increased the creep rate of
weld metal than the base metal in both the weld joint and higher content of -ferrite in the MP-TIG joint than
the A-TIG joint had increased the creep rate of weld metal in MP-TIG joint than that in the A-TIG joint. The interface could have also acted as effective diffusion path to increase creep deformation rate of weld metal
than the base metal; finer structures of MP-TIG weld metal than A-TIG weld metal might have also contributed
higher creep deformation in the weld metal of MP-TIG joint. The localized preferential deformation in weld
metal has a significant role in the premature failure of the joints under creep condition.

Fig.4. Variations of rupture life with applied stress at 923 K for base metal and weld joints.

411

412

T. Sakthivel et al. / Procedia Engineering 55 (2013) 408 413

Fig.5. Comparison of the accumulation of longitudinal and transverse creep strain across the
A-TIG and MP-TIG joint on creep exposure at 190 MPa and 923 K.

Fig.6. Creep cavity initiation in the weld joint of


(a) A-TIG and (b) MP-TIG, creep tested at 923 K and 180 MPa.

Fig.7. Creep cavity associated with -phase in the weld joint of


MP-TIG, creep tested at 923 K and 180 MPa.

T. Sakthivel et al. / Procedia Engineering 55 (2013) 408 413

Detailed metallographic investigation on the interrupted creep specimens carried out at 923 K and 180 MPa
revealed that in the 1.0 tr specimen fracture took place at the interface between the columnar and equiaxed grain
of the weld metal in the A-TIG and MP-TIG weld joint (Fig.6). Higher creep cavitation occurred in the weld
metal of MP-TIG joint than in the A-TIG joint (Fig.6). Interrupted creep test revealed that the creep cavitation
behaviour was not appreciable in the creep specimen up to 0.8 tr. Weld joint specimens interrupted at 0.2, 0.5
time to rupture were not shown any indication of creep cavitation behaviour. Area fraction of creep cavitation
was about 2.30 % and 0.15 % in MP-TIG and A-TIG joints respectively, creep ruptured (1.0 tr)at 180 MPa and
923 K. Nucleation of creep cavity was found to be associated with brittle -phase (Fig.7). Greater alignment of
the columnar grains and -ferrite along the applied stress direction in the case of A-TIG joint than in the MPTIG joint might be also the reason for lower creep cavitation in the A-TIG joint. Adopting A-TIG joining
process is expected to minimize the failure in welded 316L(N) components under creep conditions.
4. Conclusions
Joining of 316L(N) austenitic stainless steel adopting single-pass Activated TIG (A-TIG) welding process
increased the creep rupture life of the steel weld joint.
Progressive localization of creep deformation and cavitation in the weld metal led to the premature failure of
the joints in the weld metal. Accumulation of creep deformation at higher rate was observed in the weld
metal of the MP-TIG joint than that in A-TIG joint had been attributed to the higher amount of -ferrite and
finer microstructural features.
Alignment of columnar grains with -ferrite along applied stress direction and less strength disparity
between columnar and equiaxed structure of weld metal in A-TIG joint than in MP-TIG joint had been
attributed to initiate less creep cavitation.

References
[1]

M.P.Mishra, H.U.Borgstedt, M.D.Mathew, S.L.Mannan, P.Rodriguez, A comparative study of creep rupture behaviour of
modified 316L(N) base metal and 316L(N)/16-8-2 weldment in air and liquid sodium environments, Int. Journal of Pressure
Vessels & Piping, 72(1997)111-118.
[2] M.Vasudevan, A.K.Bhaduri, and Baldev Raj, Effect of Activating Fluxes on the Weld Bead Geometry, Microstructure and
Mechanical Properties of Austenitic Stainless Steels, International Welding Congress, Mumbai, 16 19 February, 2005.
[3] C.R.Heiple and J.R.Roper, Mechanism for minor element effect on GTA fusion zone geometry, Welding Journal, 61(4), p.97s102s, 1982.
[4]
K.C.Mills, and B.J.Keene, (1990), Factors affecting variable penetration, International Material Reviews, 35(4), p. 185-216.
[5]
P.C.J.Anderson, and R.Wiktorowicz, A-TIG welding The effect of the shielding gas, TWI Bulletin, July/August, p. 76-77.,
1995.
[6] P.C.J.Anderson and R.Wiktorowicz, Improving productivity with A-TIG welding, Welding and Metal Fabrication, 64(3), p.108109, 1996.
[7] W.Lucas and D.S.Howse, Activating flux increasing the performance and productivity of the TIG and plasma processes,
Welding and Metal Fabrication, 64 (1), p.11-17, 1996.
[8] T.Paskell, C.Lundin and H.Castner, (1997), GTAW flux increases weld joint penetration, Welding Journal, 76 (4), p.57-62.
[9] M.Tanaka, T.Shimizu, H.Terasaki, M.Ushio, F.Koshi-ishi and C.L.Yang, Effects of activating flux on arc phenomena in gas
tungsten arc welding, Science and Technology of Welding and Joining, 5(6), p. 397-402, 2000.
[10] M.Vasudevan, Computational and Experimental Studies on Arc Welded Austenitic Stainless Steels, PhD Thesis, 2007, Indian
Institute of Technology, Madras, India.
[11] M.Vasudevan, A.K.Bhaduri and Baldev Raj, et al., A Penetration Enhancing Flux Formulation for Tungsten Inert Gas (TIG)
Welding of Austenitic Stainless Steel and Its Application US Publication number US 2010-0068559-A1 dated18/03/2010.
[12] K.A.Yushchenko, et al, A-TIG welding of carbon-manganese and stainless steels, Intl. Conf. on Welding Technology, Paton
Institute, paper 2 Cambridge, 13-14, October, 1993.

413

Вам также может понравиться