Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Materials Science and Engineering A292 (2000) 74 82

www.elsevier.com/locate/msea

Microstructural features of dissimilar welds between 316LN


austenitic stainless steel and alloy 800
M. Sireesha a, V. Shankar b, Shaju K. Albert b, S. Sundaresan a,*
b

a
Department of Metallurgical Engineering, Indian Institute of Technology, Madras 600 036, India
Materials Technology Di6ision, Indira Gandhi Centre for Atomic Research, Kalpakkam 603 102, India

Received 3 February 2000; received in revised form 26 April 2000

Abstract
For joining type 316LN austenitic stainless steel to modified 9Cr 1Mo steel for power plant application, a trimetallic
configuration using an insert piece (such as alloy 800) of intermediate thermal coefficient of expansion (CTE) has been sometimes
suggested for bridging the wide gap in CTE between the two steels. Two joints are thus involved and this paper is concerned with
the weld between 316LN and alloy 800. These welds were produced using three types of filler materials: austenitic stainless steels
corresponding to 316, 16Cr8Ni2Mo, and the nickel-base Inconel 1821. The weld fusion zones and the interfaces with the base
materials were characterised in detail using light and transmission electron microscopy. The 316 and Inconel 182 weld metals
solidified dendritically, while the 1682 (16%Cr8%Ni 2%Mo) weld metal showed a predominantly cellular substructure. The
Inconel weld metal contained a large number of inclusions when deposited from flux-coated electrodes, but was relatively
inclusion-free under inert gas-shielded welding. Long-term elevated-temperature aging of the weld metals resulted in embrittling
sigma phase precipitation in the austenitic stainless steel weld metals, but the nickel-base welds showed no visible precipitation,
demonstrating their superior metallurgical stability for high-temperature service. 2000 Elsevier Science S.A. All rights reserved.
Keywords: Type 316LN; Austenitic stainless steel; Power plant application

1. Introduction
Dissimilar metal welds between ferritic steel and
austenitic steel tubing and piping are commonly employed in fossil fuel fired power plants and gas cooled
and liquid metal cooled fast breeder reactors. Such
transition joints are necessary because the austenitic
stainless steels with superior creep strength and oxidation resistance are required in the higher temperature
regions such as superheaters and reheaters, while creepresisting ferritic steels such as 2.25Cr 1Mo or 9Cr
1Mo steels are commercially more attractive for the
lower temperature primary boiler and heat exchanger
sections. The early use of austenitic stainless steel consumables for these joints led to several problems during
* Corresponding author. Tel.: +91-44-4458598; fax: + 91-442350509.
E-mail address: ssundar@acer.iitm.ernet.in (S. Sundaresan).
1
Inconel and Inco-weld are registered trademarks of the Inco
group of companies.

service: thermally induced cyclic stresses resulting from


the difference in thermal coefficients of expansion
(CTE) between the ferritic and austenitic steels, migration of carbon to the austenitic steel resulting in a
carbon-denuded region in the ferritic steel heat-affected
zone (HAZ) and preferential stress-oxidation at the
weld metal/ferritic steel interface leading to the formation of oxide notches [13]. In order to overcome these
problems nickel-base consumables were introduced in
the 1960s [24]. These have a CTE lying between those
of the ferritic and austenitic base materials; carbon
diffusion is also reduced on account of the decreased
carbon activity gradient between ferritic steel and
nickel-alloy weld metal and the low diffusivity of carbon in the nickel-base alloy [2]. While the introduction
of nickel-base filler materials led to a considerable
improvement in performance, service failures have been
reported even with these joints [4,5]. These failures have
been traced to the formation of creep voids associated
with alloy carbide precipitates in the ferritic steel adjacent to the weld fusion line [6].

0921-5093/00/$ - see front matter 2000 Elsevier Science S.A. All rights reserved.
PII: S 0 9 2 1 - 5 0 9 3 ( 0 0 ) 0 0 9 6 9 - 2

M. Sireesha et al. / Materials Science and Engineering A292 (2000) 7482

75

Table 1
Chemical composition of materials used, wt.%
Base materials

Undiluted filler materials

Element

316LN

Alloy 800

316

16-8-2

Inconel 182

Inconel 82

C
Si
Mn
Ni
Cr
Mo
Ti
Nb
Fe

0.02
0.3
1.8
12.1
17.9
2.4

Balance

0.09
0.7
1.0
31.8
19.9

0.36

Balance

0.052
0.6
1.7
11.5
18.6
2.2
0.04
0.01
Balance

0.07

1.5
8.8
16.2
1.6

Balance

0.05
0.5
7
69
15

0.1
2
Balance

0.015
0.1
2.8
72.6
19.6

0.37
2.68
Balance

For prolonging the life of the transition joints another approach involving the use of a trimetallic
configuration has been suggested; this uses an insert
piece made of a material having a CTE intermediate
between those of the ferritic and austenitic steels. The
more gradual variation in CTE results in a lowering of
stresses during fluctuations in service temperature [7].
Among the materials which can be contemplated for
the insert piece, alloy 800 has been found to be the
most attractive on account of its excellent creep and
oxidation resistance, in addition to an appropriate CTE
[8].
In the present work, a trimetallic transition joint has
been investigated involving modified 9Cr 1Mo steel
(designated as P91/T91) and 316LN austenitic stainless
steel as the base materials and alloy 800 as the intermediate piece. Two joints are thus necessary: one between
P91 and alloy 800 and the other between alloy 800 and
316LN. The present paper is concerned with welds
produced for the alloy 800/316LN joint.
Several types of filler materials can be considered for
these joints including iron-base and nickel-base alloys.
King et al. [7] compared various austenitic filler materials corresponding to types 309, 312, 347 and 1682
(16%Cr8%Ni2%Mo). The main problem encountered in these welds was solidification cracking, to
which type 347 was the most susceptible and 1682
the least. While type 312 exhibited little hot cracking,
the deposit had a high delta ferrite content which was
not acceptable because of possible embrittling precipitations during service. Bhaduri et al. [9,10] made a comparative evaluation of 16 8 2 and Inconel 82/182
consumables for joining alloy 800 to type 304 stainless
steel. They found that Inconel 82/182 displayed a
higher tendency for microfissuring than 16 82 but
was superior in all-weld tensile testing over a range of
temperatures. However, transverse tensile tests showed
that the joints with 16 8 2 filler were only marginally
inferior. Additionally, 16 8 2 provided a more satisfactory transition in CTE. It was therefore concluded
that 1682 should be preferred for the joint.

In optimizing the choice of transition piece and welding consumables, factors other than thermal expansion
characteristics and weldability must also be considered.

Fig. 1. (A) Micrograph of 316LN base material. (B) Micrograph of


alloy 800 base material.

76

M. Sireesha et al. / Materials Science and Engineering A292 (2000) 7482

In particular, metallurgical compatibility and long term


stability must also be taken into account in view of the
elevated temperature service. These are primarily dependent on the microstructure in the weld fusion zones
and the interfaces with the two base metals. This paper
is concerned with the microstructural features in alloy
800/316LN weldments that control the behaviour during welding and during subsequent high-temperature
exposure.

2. Experimental work

Fig. 2. TEM micrograph showing cuboidal precipitate particle in


alloy 800 base material.

Solution-annealed base plates of 316LN austenitic


stainless steel and alloy 800, 12 mm in thickness, were
welded together by manual metal arc welding using 316
austenitic stainless steel and Inconel 182 electrodes and
by gas tungsten-arc welding using 1682 and Inconel
82 filler wires. The chemical compositions of the base
materials and the undiluted filler materials are given in
Table 1. Transverse sections including weld metal, HAZ
and base metal regions of the weldment were examined
in the light microscope using standard metallographic
techniques. Samples from each weld metal were aged at
750C for 100 h in order to study possible embrittling
precipitation in accordance with the design requirements for qualifying welding consumables for the prototype fast-breeder reactor [11]. For these specimens
transmission electron microscopy was also used. After
mechanical thinning, the final electrochemical dissolution was accomplished in a twin-jet electropolishing
unit using a mixture consisting of 10% perchloric acid
and 90% methanol. A potential of 25 V and a current
of 80 mA were used and a temperature of 50C was
maintained using liquid nitrogen. The TEM investigations were conducted in a Philips scanning transmission
electron microscope.

3. Results and discussion

3.1. Base metal microstructures

Fig. 3. (A) Microstructure of 316 weld metal. (B) Isolated regions in


316 weld metal showing AF and FA modes of solidification.

The microstructure of the 316LN base metal (Fig.


1A) consists of equiaxed annealed austenite grains. In
addition, a small amount of delta ferrite in the form of
stringers and fine spherical inclusions which are probably Al2O3 were also present.
The base metal microstructure of alloy 800 is shown
in Fig. 1B. Alloy 800 contains 0.36% titanium and
0.09% carbon and hence has a tendency to form titanium carbide and titanium carbonitride during hightemperature exposure. The grain structure shows large
elongated deformed grains of austenite, some of which
are surrounded by fine grains formed by dynamic recrystallization during hot rolling. In addition, the mi-

M. Sireesha et al. / Materials Science and Engineering A292 (2000) 7482

crostructure of the alloy 800 base plate shows the


presence of three types of precipitates. The first are the
large precipitates with cuboidal morphology which are
seen in the austenite matrix and which, based on similarity to an earlier investigation on alloy 800 [12] may
be identified as nitrides or carbonitrides of titanium
formed during solidification from the melt. A TEM
micrograph of the same is shown in Fig. 2. Secondly, a
network of discontinuous precipitates, probably titanium carbides, are found along the unrecrystallised
grain boundaries, with size less than 1 mm. The third
type of precipitates are the ones that are located intragranularly and are associated with a precipitate-free
zone close to the grain boundaries. These precipitates
are of very fine size, less than 0.1 mm, which are also
expected to be titanium carbide. It has been observed
that titanium carbides and carbonitrides do not dissolve
easily during solution annealing even if high soaking
temperatures are employed [13].

77

3.2. Weld metal microstructures


The microstructure of the 316 weld metal is shown in
Fig. 3A. This reveals a nearly fully austenitic structure
with a dendritic morphology showing well-developed
side-branches. While the bulk of the 316 fusion zone
solidified as above, the region near the 316LN base
metal fusion boundary displayed a cellular structure.
Also, there were a few regions in the weld metal which
exhibited austeniticferritic (AF)/ferriticaustenitic
(FA) modes of solidification (Fig. 3B). Metallographic
examination revealed the ferrite content in these regions
to be 34% although magnetic measurement showed
only 0.5% on average. The structure of the regions that
solidified in the primary austenitic mode is coarse with
a limited amount of delta ferrite segregated in the
intercellular boundaries. However, in the regions that
solidified in the FA mode, there was a greater amount
of vermicular ferrite. The mode of solidification in

Fig. 4. (A) Microstructure of 1682 weld metal. (B) Region near root of 16 8 2 weld metal showing retained ferrite. (C) Region near 1682
weld metal/316 base metal interface showing retained ferrite.

78

M. Sireesha et al. / Materials Science and Engineering A292 (2000) 7482

Fig. 5. (A) Microstructure of Inconel 82 weld metal. (B) Microstructure of Inconel 182 weld metal.

welds of austenitic stainless steels that form a few


percent of delta ferrite has been found to be sensitive to
compositional as well as kinetic factors and the latter
may be more important in determining weld metal
microstructure. It has been shown that even autogenous
welds may exhibit both AF and FA modes in microscopically adjacent regions with no apparent differences
in composition [14,15].
The 1682 weld metal microstructure (Fig. 4A) is
predominantly cellular, although there are regions
which have solidified dendritically. The fully austenitic
cellular microstructure extended almost throughout the
fusion zone except for a few isolated regions near the
root (Fig. 4B) and near the weld metal 316LN interface (Fig. 4C) where some intercellular ferrite was
present because of the effects of dilution from the
316LN base metal.

The Inconel 82/182 weld metals, with nickel contents


of about 72 and 69%, respectively, are fully austenitic
because they do not undergo allotropic transformation.
The microstructures of both these weld metals are
similar (Fig. 5A and B) showing recrystallised features
with extensive grain boundary migration where the
recrystallised grain boundaries cut across the solidification substructures. The Inconel 182 weld metal deposited by manual metal arc welding shows, however, a
large number of fine inclusions.
The Inconel welds as well as the 316 welds with
higher solute contents have solidified dendritically,
whereas the 1682 weld with a lower solute content
has solidified in a cellular fashion. This observed difference can be rationalized on the basis of constitutional
supercooling which relates the changes in the mode of
solidification to the solute content, growth rate and
temperature gradient. For a given set of welding conditions, the solidification mode will change from cellular
to dendritic, as solute content of the liquid increases
[16,17].
The above microstructural features of the different
types of weld metals play an important role in deciding
their cracking susceptibility and mechanical properties.
For example, a quantitative evaluation of weld solidification cracking showed that the 316 weld metal exhibited a greater tendency to cracking than the 1682
weld metal [18]. This was in spite of the fact that the
316 weld metal showed a slightly larger amount of
localised ferrite. It is believed that the greater vulnerability to cracking of 316 weld metal is to be attributed
to its pronounced dendritic morphology. It is known
that dendritic structures are associated with a greater
degree of segregation and are more prone to cracking
[19].
In the case of the Inconel 82 and Inconel 182 weld
metals, mechanical property evaluation of the fusion
zones showed that the Inconel 82 welds displayed a
consistently higher tensile elongation and Charpy impact toughness when compared to the Inconel 182 weld
metal [18]. This is clearly to be attributed to the higher
inclusion content in the case of the Inconel 182 weld
metal which was produced using flux-coated electrodes
while Inconel 82 was deposited using gas tungsten-arc
welding.

3.3. Interfacial microstructures


The interface between the 316 weld metal and 316LN
base metal is shown in Fig. 6A. The microstructure
reveals ferrite stringers in some of the austenite grains
in the base metal close to the HAZ. These ferrite
stringers may be remnants from the high-temperature
primary processing of the base plate from the ingot. It
is well known that the homogenisation of these segre-

M. Sireesha et al. / Materials Science and Engineering A292 (2000) 7482

gates is very ineffective especially in austenitic alloys


[20]. The ferrite stringers have expanded and grown in
width close to the fusion boundary, presumably because more ferrite is usually retained on rapid cooling
after the formation of delta ferrite during the heating
cycle. Delta ferrite is also known to be retained in the
HAZ grain boundaries of welds having a composition
with a positive ferrite potential [21]. Such retained
ferrite is also visible. In addition to the delta ferrite
stringers extending from the base metal, the interfacial
region shows the presence of a prominent unmixed
zone with a much higher semicontinuous delta ferrite
content that has formed in a vermicular or intercellular
pattern. The unmixed zone exists as a laminar layer
where a small portion of the base metal has totally
melted and resolidified without undergoing filler metal

79

dilution. This zone is clearly seen in the higher magnification micrograph Fig. 6B.
The microstructure of the interface between 316 fusion zone-Alloy 800 heat-affected zone (Fig. 6C) shows
extensive grain boundary melting and liquation. The
partially melted zone on this side of the joint appears to
be much wider than on the 316LN side, as also the zone
of austenitic grain coarsening. As mentioned earlier, the
base metal microstructure of the alloy 800 consisted of
large elongated grains present as a result of incomplete
recrystallisation during hot deformation. In the HAZ,
these grains have recrystallised into a fine equiaxed
grain structure that is found close to the fusion
boundary. Even this recrystallised zone, however, has a
much coarser grain structure when compared with the
HAZ of the 316LN. Further, a well-defined partially

Fig. 6. (A) Interface between 316 weld metal and 316LN base metal showing unmixed zone (arrowed). (B) Higher magnification micrograph of
316 weld metal and 316LN base metal interface revealing unmixed zone (arrowed) with higher ferrite content. (C) Microstructure of interface
between 316 weld metal and alloy 800 base metal showing unmixed zone (arrowed).

80

M. Sireesha et al. / Materials Science and Engineering A292 (2000) 7482

Fig. 7. (A) Microstructure at interface between Inconel 182 weld


metal and 316LN base metal showing unmixed zone (arrowed). (B)
Microstructure at interface between Inconel 82 weld metal and Alloy
800 base metal showing unmixed zone (arrowed).

melted zone is visible on this side of the joint that is


much wider than in the 316LN HAZ. This partially
melted zone is characterised by grain boundary melting
and thickening. The original grain boundaries in many
cases have moved to locations a few microns away from
the previous sites. The tendency of grain boundaries to
melt in alloy 800 HAZ is well known and is attributed
to the enrichment of titanium at these boundaries
[22,23]. Titanium at these boundaries not only lowers
the melting point constitutionally but also forms lowmelting carbide-austenite eutectics during solidification
[22].
The interface of the Inconel 182 weld metal with
316LN base metal (Fig. 7A) as well as with alloy 800
(Fig. 7B) shows the presence of an unmixed zone.
However, the unmixed zone on the Inconel 182-316LN

base metal interface is much wider and exhibits a


greater degree of dendritic morphology than that on the
alloy 800 side. This may be attributed to the greater
compositional difference between the Inconel weld
metal (nickel-based) and the 316LN base metal (ironrich). Welds between widely dissimilar combinations
are known to exhibit much wider unmixed or partially
mixed zones where the microstructure and chemical
composition are quite different from that of the surrounding weld metal [24]. These large unmixed zones
tend to form near the interface with the base metal
having a significantly higher melting point. In the
present case, the melting point of the 316LN base metal
is higher than that of the Inconel weld metal, and the
convection currents are not able to promote adequate
fluid flow and mixing. On the other hand, being significantly highly alloyed, the Alloy 800 has a melting point
as well as composition closer to that of the Inconel
weld metal. In this case a wide unmixed zone cannot
form as convection in the weld puddle would ensure
that only a thin laminar layer would remain unmixed.
The occurrence of grain boundaries in the weld metal
of dissimilar welds just adjacent to the fusion boundary
and running parallel to it has been discussed recently
and designated as type II boundaries. These are different from type I boundaries usually observed in homogeneous (similar base and filler metals) welds. In the
latter type of welds epitaxial growth causes grain
boundaries from the base metal substrate to run continuously across the fusion boundary in a direction
roughly perpendicular to it.
The occurrence of type II boundaries was originally
attributed to a transition in primary solidification behaviour (say from ferritic to austenitic) due to the
compositional gradient normal to the fusion boundary
[25]. More recent work [26] has shown that the type II

Fig. 8. Microstructure at region between Inconel 182 weld metal and


T91 base metal in joint between Alloy 800 and T91, showing type II
grain boundary.

M. Sireesha et al. / Materials Science and Engineering A292 (2000) 7482

81

Fig. 9. (A) Microstructure of 316 weld metal after aging for 100 h at 750C showing sigma phase precipitation (arrowed). (B) Microstructure of
16 8 2 weld metal after aging for 100 h at 750C showing sigma phase precipitation (arrowed). (C) Transmission electron micrograph of 316 weld
metal showing sigma phase (arrowed). (D) Microstructure of Inconel 182 weld metal after aging showing absence of precipitation.

boundary is a result of allotropic transformation in the


base metal that occurs on cooling and produces grain
boundaries of the g/g type at the fusion boundary in
dissimilar metal (fcc/bcc) welds. It has been shown that
type II boundaries can form in such dissimilar metal
welds only when there is a FA phase boundary at
elevated temperature in the base metal.
In the current work all the micrographs taken from
the interface regions show that only type I boundaries
are present. This is, of course, to be expected because,
although dissimilar welds are involved, there is no
allotropic transformation on cooling in the case of
either of the two base metals. It is of interest to record
that, in another phase of the program concerning a
joint between alloy 800 and modified 9Cr 1Mo steel
using Inconel 182 filler material, type II boundaries

were indeed observed on the ferritic steel side. This is


shown in Fig. 8.

3.4. Aged microstructures


As mentioned earlier, the weld metals were subjected
to an aging treatment at 750C for 100 h on account of
a specification for qualifying welding consumables for
fast breeder reactor components. The microstructures
of the aged 316, 1682 and Inconel 182 weld metals
are shown in Fig. 9AC. Typically, both the stainless
steel fusion zones exhibit a discontinuous network of
the sigma phase and isolated fine particles apparently
consisting of carbides. A greater amount of sigma
phase formation is noticeable in Fig. 9A corresponding
to the 316 weld metal than in Fig. 9B corresponding to

82

M. Sireesha et al. / Materials Science and Engineering A292 (2000) 7482

16 82. It is likely that the sigma phase has formed


predominantly in regions of ferrite noticed in the aswelded microstructures. Fig. 9C is a transmission electron micrograph of the 316 weld metal which shows
roughly spheroidal particles of the sigma phase that
seem to have formed preferentially in a region previously occupied by the ferrite. The fact that more sigma
is noticed in 316 than in 16 8 2 may be attributed to
the higher chromium and molybdenum contents in the
former.
On the other hand, the micrograph of the Inconel
182 weld metal (Fig. 9D) does not reveal any kind of
precipitation after the ageing treatment. TEM examination also showed that no precipitation could be noticed.
The behaviour of the Inconel 82 weld metal was quite
similar to that of Inconel 182. In an earlier investigation on the behaviour of nickel-based weld metal during heat treatment in the 600 900C temperature range
[27], embrittling precipitation was found to occur at
700C, but only after ageing for 10 000 h. Precipitation
and toughness loss were less pronounced at 600C, and
even less in the range 800 900C. It is therefore not
surprising that in the current work the two nickel-based
weld metals did not exhibit any observable precipitation on ageing for 100 h at 750C. Clearly, these
observations demonstrate the superior thermal stability
of nickel-base alloy weld metals in relation to the
stainless steel weld metals. Indeed, measurements of
Charpy toughness of the aged specimens [18] showed
that the two stainless steel weld metals suffered an
appreciable reduction in toughness consequent to the
aging treatment. In the case of the nickel-base weld
metals, on the other hand, the lowering of toughness
after ageing was of a much smaller magnitude. It is
worth mentioning, however, that the toughness values
of even the stainless steel weld metals in the aged
condition exceeded the minimum prescribed under the
relevant specification.

4. Summary
Weld joints between 316LN austenitic stainless steel
and alloy 800 can be made with either stainless steel or
nickel-alloy filler material. The 16 8 2 weld metal with
the lowest solute content exhibits a cellular substructure, which was shown in a related study [18] to be
associated with a high resistance to solidification cracking. The Inconel and type 316 weld fusion zones, on the
other hand, solidify with more pronounced dendritic
branching and are more susceptible to hot cracking.
The Inconel weld metal, when deposited from fluxcoated electrodes, reveals a higher inclusion content
than under gas-shielded welding; this results in lower
ductility and toughness in the former case. High-tem.

perature ageing promotes sigma-phase formation in the


austenitic steel weld metal, while the nickel-base fusion
zones show high metallurgical stability.

Acknowledgements
The authors thank the Board of Research in Nuclear
Sciences, Department of Atomic Energy, for the financial support provided for the work.

References
[1] C.D. Lundin, Weld. J. 61 (2) (1982) 58s.
[2] R. Viswanathan, Proc. AWS/EPRI Conf. Joining Dissimilar
Metals, Pittsburg, PA (1982), pp. 7.
[3] K.G.K. Murti, S. Sundaresan, Weld. J. 64 (12) (1985) 329s.
[4] A.T. Price, CEGB experience with small diameter dissimilar
metal welds in coal fired boilers, in Proc. AWS/EPRI Conf.
Joining Dissimilar Metals, Pittsburg, PA, 1982, pp. 4879.
[5] P.E. Haas, Results of industry wide survey on dissimilar metal
weld performance, in Proc. AWS/EPRI Conf. Joining Dissimilar
Metals, Pittsburgh, PA, 1982, pp. 37 47.
[6] R.D. Nicholson, Metals Technol. 11 (3) (1984) 115.
[7] J.F. King, M.D. Sullivan, G.M. Slaughter, Weld. J. 56 (11)
(1977) 354s.
[8] W.B. Jones, R.M. Allen, Metall. Trans. A 13A (1982) 637.
[9] K. Bhaduri, I. Gowrisankar, V. Seetharaman, S. Venkadesan, P.
Rodriguez, Mater. Sci. Technol. 4 (11) (1988) 1020.
[10] A.K. Bhaduri, S. Venkadesan, P. Rodriguez, Intl. J. Pres. Ves.
Pip. 58 (1994) 251.
[11] PFBR/32040/SP/1002/R-0-Prototype Fast Breeder Reactor specification for the qualification of the welding consumables as
proposed by IGCAR, Kalpakkam, India.
[12] Lippold, Weld. J. 62 (1) (1983) 1s.
[13] P.G. Stone, J. Orr, J.C. Guest, in: S.F. Pugh (Ed.), Proceedings
of a BNES Conference, AERE Harwell, Oxfordshire, 1974, p.
15.
[14] H.K.D.H. Bhadeshia, S.A. David, J.M. Vitek, Mater. Sci. Technol. 7 (1) (1991) 50.
[15] N. Suutala, Metall. Trans. A 14A (2) (1983) 191.
[16] W.F. Savage, E.F. Nippes, J.S. Erickson, Weld. J. 55 (1976)
213s.
[17] G.J. Davies, J.G. Garland, Int. Mater. Rev. 20 (1975) 83.
[18] M. Sireesha, S.K. Albert, V. Shankar, S. Sundaresan, J. Nuclear
Mater. 279 (2000) 65.
[19] W.F. Savage, C.D. Lundin, A.H. Aronson, Weld. J. 44 (4)
(1965) 175s.
[20] E. Folkhard, Welding Metallurgy of Stainless Steels, Springer
Verlag, New York, 1988.
[21] W.A. Baeslack, J.C. Lippold, W.F. Savage, Weld. J. 58 (6)
(1979) 168s.
[22] J.C. Lippold, Weld. J. 62 (1) (1983) 1s.
[23] K. Saito, M. Aoki, H. Kondo, Trans. ISIJ 27 (1987) 580.
[24] S.K. Albert, T.P.S. Gill, A.K. Tyagi, S.L. Mannan, S.D. Kulkarni, P. Rodriguez, Weld. J. 76 (3) (1997) 135s.
[25] S. Duvall, W.A. Owczarski, Weld. J. 47 (3) (1968) 115s.
[26] T.W. Nelson, J.C. Lippold, M.J. Mills, Sci. Technol. Weld. Join.
3 (5) (1998) 249.
[27] J.O. Nilsson, B. Lundquist, M. Lonnberg, Weld. J. 73 (1) (1994)
45.

You might also like