Hot Cracking2
Hot Cracking2
123
Prof. Dr.-Ing. Thomas Böllinghaus Prof. Dr. Carl E. Cross
Bundesanstalt für Bundesanstalt für
Materialforschung und–prüfung Materialforschung und–prüfung
Unter den Eichen 87 Unter den Eichen 87
12205 Berlin 12205 Berlin
Germany Germany
[email protected] [email protected]
c 2008 Springer-Verlag Berlin Heidelberg
This work is subject to copyright. All rights are reserved, whether the whole or part of the material is
concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting,
reproduction on microfilm or in any other way, and storage in data banks. Duplication of this publication
or parts thereof is permitted only under the provisions of the German Copyright Law of September 9,
1965, in its current version, and permission for use must always be obtained from Springer. Violations are
liable to prosecution under the German Copyright Law.
The use of general descriptive names, registered names, trademarks, etc. in this publication does not imply,
even in the absence of a specific statement, that such names are exempt from the relevant protective laws
and regulations and therefore free for general use.
9 8 7 6 5 4 3 2 1
springer.com
Preface
Abstract
One can see that, over the years, much more effort has been put on
the conditions required for hot tearing occurrence rather than on the
mechanisms of crack initiation and propagation. And when it comes to
the nucleation and propagation of hot tears, an educated guess
frequently replaces experimental proof.
6 L. Katgerman, D.G. Eskin
The existing hot tearing criteria as reviewed elsewhere [17, 50] can be
conditionally divided into the two categories: non-mechanical and me-
chanical. The former type of criteria deals with vulnerable temperature
range, phase diagram, and process parameters; and is represented by the
criteria of Clyne & Davies [6], Feurer [40], and Katgerman [36]. The latter
type of criteria involves critical stress [9, 27, 31, 32, 45], critical strain [9,
29, 51], or critical strain rate [23, 37, 38, 39, 49, 52].
Different casting processes impose specific requirements on the
application of hot tearing criteria. That is why some criteria are working
better for shape casting whereas others are more suitable for direct-chill
casting. There is no doubt that a good hot tearing criterion for DC casting
should correctly respond to the casting parameters, e.g. casting speed,
ramping rate, and alloy composition; and predict the vulnerable section of
a billet or an ingot, e.g. the centre of a round billet. Most of the existing
criteria have been tested for the composition sensitivity by calculating the
hot tearing susceptibility of several binary alloys with an attempt to
reproduce the so-called lambda-curve showing the maximum susceptibility
at a certain composition. And most of the existing criteria can do this
successfully. However, dynamic parameters such as casting speed and
strain rate are usually kept constant upon such testing. Therefore, the
compositional sensitivity of a hot tearing criterion does not assure its
successful application to a particular casting technology.
The basic phenomena that lead to hot cracking are well established and
understood, but a generic criterion that will predict hot cracking under
varying process conditions is still not available. Although the earlier
simple criteria based on the thermal history of the casting have been
extended and improved to include shrinkage and deformation, they are still
unable to give reliable predictions under all process conditions. Most of
the existing hot tearing criteria do not incorporate the nucleation and
propagation of a hot tear, focusing more on the macroscopic and
microscopic conditions that may result in rupture.
The ultimate hot cracking criterion needs to combine aspects of thermal
history, shrinkage and porosity formation, and constitutive behaviour in
combination with the evolution of the semi-solid microstructure. Current
research efforts are aimed, in particular, at the quantitative description of
structure evolution and its correlation to cracking.
Recently, several mechanical and non-mechanical hot tearing criteria
were evaluated by implementing them into a thermo-mechanical model of
In Search of the Prediction of Hot Cracking in Aluminium Alloys 9
DC casting [50]. The criteria show different results in predicting the hot
tearing susceptibility as shown in Table 2.
Table 2. Sensitivity of hot tearing criteria to the casting parameters and practice
upon direct-chill casting [50, 53]
Criterion Hot tearing More hot Ramping casting Correlation with
increases with tears in the speed during actual cracking
casting speed billet centre start-up of the observed in
casting reduces practice
hot tearing
Clyne & Davies [6] No No No N/A
Katgerman [36] Yes Yes No N/A
Feurer [40] Yes Yes No N/A
Novikov [9] No No No N/A
Magnin et al. [29] Yes No No No
Prokhorov [37] Yes Yes No No
Rappaz et al. [38] Yes Yes Yes No
Braccini et al. [23] Yes Yes No N/A
Suyitno [48] Yes Yes Yes Yes
The criteria of Clyne & Davies and Novikov give results that are
inconsistent with casting practice, not showing any sensitivity to the casting
speed and position within the billet volume. It is noteworthy to mention that
these criteria are very successful in predicting the compositional dependence
of hot tearing and are frequently used for shape casting. The criteria of
Feurer, Katgerman, Magnin et al., Prokhorov, Rappaz et al., and Braccini et
al. respond correctly to the casting parameters, demonstrating that the
increasing casting speed results in an increasing hot tearing susceptibility in
the centre of billet, which is in accordance with casting practice. However,
most of the tested criteria, except those by Rappaz et al. [38] and Suyitno
[48, 53], are not sensitive to the ramping of casting speed during the start-up
phase of casting (which is a usual practice to prevent hot cracking). And,
when confronted with casting practice, the criteria of Prokhorov, Magnin et
al., and Rappaz et al. predict the occurrence of hot cracks whereas no cracks
have been found in billets cast under given conditions. Only the criterion of
Suyitno adequately responds to all tested parameters, i.e. casting speed,
ramping rate, grain size, position in a billet, and casting practice. The
sensitivity of this criterion is, however, a function of correctly chosen values
of properties such as Young’s modulus of the mush, surface tension between
liquid and solid, and permeability of the mush. These parameters are
scarcely available and needed to be determined experimentally, while the
existing experimental techniques are not reliable.
10 L. Katgerman, D.G. Eskin
a b
Fractured
bridges
Fig. 3. Schematic illustration of possible hot crack initiators and some crack
propagation mechanisms (the initial sketch is adopted from [58])
Despite a large body of literature on hot tearing, only few efforts have
been spent on the mechanism of hot tear propagation. One can mention the
application of the Griffith criterion for brittle fracture [48, 53] and va-
cancy-diffusion-controlled growth [21]. The apparatus for the description
of crack propagation is well developed within fracture mechanics for
In Search of the Prediction of Hot Cracking in Aluminium Alloys 13
various situations, including those resembling hot cracking, i.e. liquid film
rupture, pore coalescence, high-temperature creep, and liquid-assisted
fracture [58]. It is clear that the propagation of the hot crack in the potential
presence of liquid (above or below the solidus) should involve the follow-
ing aspects: liquid feeding (involves permeability and, inevitably structure
evolution), pore coalescence, stress transfer by solid bridges, plastic de-
formation and creep of solid bridges in the absence of liquid, and brittle
fracture of solid bridges in the presence of liquid. It is also obvious that the
hot crack, like any other crack, can develop catastrophically (which is usu-
ally assumed), have sustained growth, or stop. Liquid feeding plays a dual
role. Firstly, adequate feeding of the shrinking material with liquid does
not eliminate the causes of hot tearing but rather “patches” the conse-
quences, which is reflected in the term “crack healing”. One can say that a
semi-solid alloy containing enough liquid at the last stage of solidification,
having a microstructure that enables adequate permeability of the mush,
and subjected to tensile stresses is a self-healing material. On the other
hand, the development of solid bridges between grains at high solid frac-
tions in the absence of any liquid would build up enough strength and duc-
tility to prevent any brittle rupture. Hence, the liquid feeding should be just
enough to supply some liquid to solid bridges that enables their liquid em-
brittlement, otherwise only the mechanisms of ductile fracture, e.g. high-
temperature creep and pore coalescence will be active [8]. The evidence of
plastic deformation during hot tearing has been observed in direct observa-
tions [59] and upon examination of fracture surfaces [9].
We can suggest the following approach to treating the nucleation and
propagation of hot cracks. Several distinct mechanisms are operational in
different temperature or compositional ranges or, in other words, at differ-
ent fractions of solid. Here we will consider only the decreasing temperature
as the factor affecting the solid fraction, though the composition is obvi-
ously the other factor that acts in a similar manner. It is important to note
that the microstructure, e.g. grain size and morphology, affects the critical
temperatures and fractions of solid. At relatively low fractions of solid
below the coherency temperature, the permeability of the mushy zone
enables adequate feeding of the solidification shrinkage and most of the
precipitating gas bubbles can float to the liquid part of the sample. In this
case, the tensile stresses caused by non-uniform thermal contraction of the
coherent dendrites may cause the formation of cavities and gaps that are
immediately filled with liquid, or “healed”. On further decreasing tempera-
ture, the tensile stress builds up to such an extend that the liquid film sepa-
rating grains ruptures and the formed gap cannot be filled with liquid due
to the already insufficient permeability of the mush and to the increasing
capillary pressures required to fill ever narrowing openings between grains.
14 L. Katgerman, D.G. Eskin
On further cooling, the bridging between solid grains replaces former entan-
glement and touching of grains, and the stress can now be transmitted over
larger distances through the rigid solid skeleton, hence the semi-solid body
acquires macroscopic strength. This critical temperature is called the
“rigidity temperature” and can be determined experimentally [9, 60]. Note
that the rigidity temperature strongly depends on the structure [60]. The
permeability at such fractions of solid is not enough to completely com-
pensate the solidification shrinkage, gas cannot escape from narrow pas-
sages and, as a result, porosity is formed at available interfaces, mainly on
grain boundaries. The non-uniform thermal stress causes rather significant
strains in the semi-solid material that can be or cannot be sustained by the
solid bridges. Even limited access of the liquid to the solid bridge will re-
sult in its brittle fracture by the mechanism of liquid-metal embrittlement.
This is partially reflected in the proposal of van Haaften et al. [61] to use
the fraction of grain boundaries covered with liquid rather than the fraction
of liquid in the constitutive equation for the mechanical behaviour or semi-
solid aluminium alloys. The same mechanism may act at subsolidus tem-
peratures when some amount of non-equilibrium liquid is present at grain
boundaries or other stress concentrators because of non-equilibrium char-
acter of solidification [8] or local remelting [57]. There is also a possibility
that the semi-solid material fails macroscopically in a brittle manner (be-
cause of film rupture and liquid-metal embrittlement) with ductile rupture
of some solid bridges on the microscopic level [59]. And finally, the frac-
tion of bridged grain boundaries becomes so overwhelmingly large and the
remaining liquid is so scattered in the solid network that the semi-solid
material behaves like completely solid and fails in a ductile manner by
ductile pore coalescence and high-temperature creep. The crack initiator in
all these cases could be represented by pore, liquid pool or film, interface
with an intermetallic particle, or non-metallic inclusion. The outline of
these mechanisms is given in Table 3. Figures 1 and 3 illustrate the corre-
lation between these mechanisms and the development of the structure
during solidification.
A criterion that can predict not the probability but the actual occurrence
and extent of hot tearing should be based on the application of multi-phase
mechanics and fracture mechanics to the failure of semi-solid materials,
which is today limited by the lack of knowledge about the actual nuclea-
tion and propagation mechanisms. The mechanisms outlined in Table 3 are
based on the common sense and interpretation of very few experimental
observations. What is needed is a thorough and systematic study of frac-
tures occurring in solidifying materials with the aim to single out the na-
ture and the critical dimensions of defects or structure features that can
cause the nucleation of hot cracks. It is also necessary, in our opinion, to
In Search of the Prediction of Hot Cracking in Aluminium Alloys 15
Concluding Remarks
The existing hot tearing criteria based on different principles have limited
applicability to commercial casting processes, e.g. to direct-chill casting,
due to their probabilistic character. The best of the available criteria can
successfully predict the probability of hot tearing in its dependence on
some casting parameters but fail to quantify the event, in other words can-
not forecast the actual occurrence of hot cracks in ingots and billets. There
are two main challenges in this endeavour. Firstly, we are lacking the
16 L. Katgerman, D.G. Eskin
Acknowledgements
References
1. H.F. Bishop, C.G. Ackerlind, and W.S. Pellini: Trans. Am. Foundrymen’s
Soc., 1952, vol. 60, pp. 818–833.
2. W.S. Pellini: Foundry, 1952, vol. 80, pp. 124–133, 192–199.
3. J.C. Borland: Brit. Weld. J., 1960, vol. 7, pp. 508–512.
4. S.A. Metz and M.C. Flemings: Trans. Am. Foundrymen’s Soc., 1970, vol. 78,
pp. 453–460.
5. U. Feurer: Giessereiforschung, 1976, vol. 28 (2), pp. 75–80.
6. T.W. Clyne and G.J. Davies: in Solidification and Casting of Metals, Metals
Society, London, United Kingdom, 1979, pp. 275–278.
7. B. Rogberg: Scand. J. Met., 1983, vol. 12, pp. 51–66.
8. J. Campbell: Castings, Butterworth-Heinemann, Oxford, United Kingdom, 1st
edn. 1991; 2nd edn., 2003.
9. I.I. Novikov: Goryachelomkost tsvetnykh metallov i splavov (Hot shortness of
non-ferrous metals and alloys), Nauka, Moscow, USSR, 1966.
10. G.K. Sigworth: Trans. Am. Foundrymen’s Soc., 1996, vol. 104, pp. 1053–1062.
11. J.A. Spittle and A.A. Cushway: Met. Technol., 1983, vol. 10, pp. 6–13.
12. L. Ohm and S. Engler: Giessereiforschung, 1990, vol. 42 (4), pp. 149–162.
13. M.L. Nedreberg: PhD Thesis, University of Oslo, Oslo, Norway, 1991.
14. J.A. Spittle, S.G.R. Brown, J.D. James, and R.W. Evans: in Proc 7th Int.
Symp. on Physical Simulation of Casting, Hot Rolling and Welding, National
Research Institute for Metals, Tsukuba, Japan, 1997, pp. 81–91.
In Search of the Prediction of Hot Cracking in Aluminium Alloys 17
42. M. M’Hamdi and A. Mo: in Light Metals 2002, W. Schneider, ed., The Minerals,
Metals, and Materials Society, Warrendale, USA, 2002, pp. 709–716.
43. M. Rappaz, P.-D. Grasso, V. Mathier, J.-M. Drezet, and A. Jacot: in Solidification
of Aluminum Alloys, M.G. Chu, D.A. Granger, and Q. Han, ed., The Minerals,
Metals, and Materials Society, Warrendale, USA, 2005, pp. 179–190.
44. K. Patterson: Giesserei, 1953, vol. 40 (12), pp. 597–605.
45. J.A. Williams and A.R.E. Singer: J. Inst. Met., 1968, vol. 96. pp. 5–12.
46. J.J. Gilman: in Proc of 2nd Symp. on Naval Structural Mechanics, E.H. Lee and
P.S. Symonds, ed., Pergamon, London, United Kingdom, 1960, pp. 43–99.
47. E. Orowan: in Fatigue and Fracture of Metals, W.M. Murray, ed., Massachu-
setts Institute of Technology, MIT Press, Cambridge, USA, 1950, p. 139.
48. Suyitno, W.H. Kool, and L. Katgerman: Mater. Sci. Forum, 2002, vol. 396–402,
pp. 179–184.
49. J.F. Grandfield, D.J. Cameron, and J.A. Taylor: in Light Metals 2001, J.L. Anjier,
ed., The Minerals, Metals, and Materials Society, Warrendale, USA, 2001, pp.
895–901.
50. Suyitno, W.H. Kool, and L. Katgerman: Metall. Mater. Trans. A, 2005, vol.
36A, pp. 1537–1546.
51. L. Zhao, Baoyin, N. Wang, V. Sahajwalla, and R.D. Pehlke: Int. J. Cast Metals
Res., 2000, vol. 13 (3), pp. 167–174.
52. J.-M. Drezet and M. Rappaz: in Light Metals 2001, J.L. Anjier, ed., The Minerals,
Metals, and Materials Society, Warrendale, USA, 2001, pp. 887–893.
53. Suyitno: PhD Thesis, Delft University of Technology, Delft, The Netherlands,
2005.
54. Suyitno, D.G. Eskin, V.I. Savran and L. Katgerman: Metall. Mater. Trans. A,
2004, vol. 35A, pp. 3551–3561.
55. Y. Ju. PhD Thesis, Norwegian University of Science and Technology,
Trondheim, Norway, 2004.
56. D. Eskin, Suyitno, and L. Katgerman: in Aluminium Cast House Technology
2005, Proc. 9th Australasian Conference, J. Taylor, I. Bainbridge, and
J. Granfield, ed., Collingwood: CSIRO Publishing, 2005, pp. 77–84.
57. A. Deschamps, S. Péron, Y. Bréchet, J.-C. Ehrström, and L. Poizat: Mater.
Sci. Technol., 2002, vol. 18, pp. 1085–1091.
58. M. Janssen, J. Zuidema, and R.J.H. Wanhill: Fracture Mechanics. Delft
University Press Blue Print, Delft, The Netherlands, 2002.
59. W.-M. van Haaften, W.H. Kool, and L. Katgerman. J. Mater. Eng. Perform.,
2002, vol. 11, pp. 537–543.
60. D.G. Eskin, Suyitno, J.F. Mooney, and L. Katgerman: Metall. Mater. Trans.
A, 2004, vol. 35A, pp. 1325–1335.
61. W.-M. van Haaften, W.H. Kool, and L. Katgerman: Mater. Sci. Eng. A, 2002,
vol. 336, pp. 1–6.
Application of the Rappaz-Drezet-Gremaud Hot
Tearing Criterion to Welding of Aluminium Alloys
Abstract
Introduction
∂g s §K ·
ρ + (1 + β ) g s (ε yy + εzz ) − div ¨ ( gradPl − ρl g ) ¸ = 0. (1)
∂t ©μ ¹
β = (ρs/ρƐ - 1) is the solidification shrinkage, ε yy and εzz are the two
components of the strain rate of the solid perpendicular to the thermal
gradient (see Fig. 1), K is the permeability of the mush, μ is the viscosity
and pƐ the local pressure in the interdendritic liquid, g being the gravity
vector. Note that ρs and ρƐ have been assumed constant, while gs is con-
stant in the directions perpendicular to the thermal gradient (directional
solidification). It is interesting to notice that although a deformation is
applied, it is the strain rate that appears in Eq. (1). This expression is
fairly general and can be interpreted as follows: solidification shrinkage
(1st term) and/or deformation of the solid (2nd term) have to be compen-
sated by liquid flow (3rd term) if pores or hot tears are to be avoided. In
the case a third phase (pores or hot tears) is considered, the right hand
term of Eq. (1) is simply replaced by - ∂gp/∂t, where gp is the fraction of
pores (or hot tears) [11].
Under steady directional solidification conditions at a velocity vT, the
maximum pressure drop, Δpmax, across the mushy zone and associated with
deformation and solidification shrinkage is given by:
L L
E g
Δpmax = (1 + β ) μ ³ dx + vT βμ ³ l dx (2)
0
K 0
K
where E is the cumulated strain rate E ( x) = g s (ε yy + εzz )dx . These two
³
integrals can be transformed into integrals over temperature, thus introduc-
ing a “competition” between strain rate and thermal gradient, G, for the
first contribution, and the standard vT/G ratio for the shrinkage term as
already derived by Niyama in his porosity criterion [12].
22 J.-M. Drezet, D. Allehaux
Fig. 1. Schematics of hot tearing formation and of the two-phase problem consid-
ered in the RDG approach [11]
The RDG criterion simply states that a hot tear forms if the local pres-
sure in the liquid, i.e., the metallostatic pressure minus the pressure drop,
falls below a given cavitation pressure. This criterion has nevertheless a
few limitations:
• using only the perpendicular component of the plastic strains is not
strictly valid, the longitudinal component also inducing some suction (or
expulsion) of the liquid;
• the lower bound of the integrals of Eq. (2) is ill-defined. As gs tends to-
wards unity, the permeability goes to zero and the calculation diverges.
In practical situations, this bound is set up to a value of gs at which the
solid is considered as coherent, i.e., the liquid remains only as liquid
pockets (no continuous liquid films).
• in relation to grain boundaries, the method does not consider any local-
ization of the strains and feeding.
Stating the local pressure in the liquid films must not fall below a cavita-
tion pressure if no hot tears should form is equivalent to state that the
strain rate must remain lower than a maximum value. In other words, the
mush can sustain some deformation but its rate should remain low enough
in order to permit liquid feeding. The deformation is indeed limited by the
ductility of the solid + liquid mixture. Ductility curves obtained by tensile
tests exhibit the typical shape of U when liquid and solid phases are pre-
sent as shown in Fig. 2 [13, 14]. Some ductility is present at temperatures
Application of RDG Hot Tearing Criterion to Welding of Al Alloys 23
Fig. 2. U ductility curve for the Al-Cu 4.5 wt. pct. alloy together with two ther-
momechanical paths (adapted from [14])
Finally, the RDG criterion, originally derived for aluminium alloys, has
been extended to steel by Drezet et al. [16] by taking into account the peri-
tectic reaction that transforms ferrite into austenite. Moreover, Rindler
et al. [17] have further worked out the criterion for steels by releasing the
assumption of a uniform strain rate acting over the mush.
Under the assumption that the strain rate applied to the mushy zone is
uniform ε = ε yy + εzz and that steady state is reached, the liquid pressure
drop through the mush is made out of two contributions, Δpsh due to the
24 J.-M. Drezet, D. Allehaux
solidification shrinkage and Δp mec due to deformation [10]. The two con-
tributions can be written as:
180μ ª (1 + β)Bε º
Δp sh +Δp mec = v βA +
2 « T »¼ (3)
Gλ 2 ¬ G
T
f s2 ³ f s dT
Tliq 2 Tliq Tcg
with A = ³
f sdT and B= ³ dT (4)
Tcg (1 − f s ) −
2 3
Tcg (1 f s)
λ2 , the secondary dendrite arm spacing, is the typical length of the micro-
structure used to define the permeability K. The two integrals A and B are
related to microporosity formation induced by the solidification shrinkage
and to hot tearing, respectively. The larger these quantities, the larger the
pressure drop and therefore the higher the risk to initiate a hot tear. Note
that the lower bound of the two integrals A and B are set to Tcg, the tem-
perature at which coalescence of the grains occurs.
Within the framework of the European Wel-Air project [18], the
RDG hot tearing criterion is used to analyse the influence of the alloy
composition per se and of the filler material on the hot tearing suscepti-
bility for new generation aircraft aluminium alloys. To do so, the solidi-
fication path of the alloy (plus the filler material) was obtained using
ProPHASE, a microsegregation program developed at Alcan Péchiney
[19]. As the solidification path should be relevant with the solidification
conditions undergone by the alloy during laser welding, a Scheil ap-
proach (no solid-state diffusion) is used owing to the very high cooling
rate experienced by the alloy. The grain coalescence temperature is not
easy to determine; the temperature corresponding to a solid fraction of
98% or the eutectic temperature if more than 2% eutectic has formed,
was adopted as it correctly predicted the hot cracking susceptibility for
binary alloys, the so-called Λ curve [1]. Figure 3 shows the solidification
paths computed with ProPHASE for five aluminium alloys. Their com-
position is detailed in Table 1.
Application of RDG Hot Tearing Criterion to Welding of Al Alloys 25
1
0.9
0.8
2098
0.7
Solid fraction (-)
6013
0.6
0.5 6056
0.4 2139
0.3 2022
0.2
0.1
0
490 510 530 550 570 590 610 630 650
Temperature (°C)
1
0.99
2098
0.98
6013
0.97
6056
Solid fraction (-)
0.96
2139
0.95
2022
0.94
0.93
0.92
0.91
0.9
490 500 510 520 530 540 550 560 570
Temperature (°C)
Fig. 4. Zoom at solid fractions higher than 90% of the solidification paths
1.E+05
6.E+04
4.E+04
2.E+04
0.E+00
2098 6013 6056 2139 2022
Alloy
Fig. 5. HCS of the WEL-AIR alloys computed with the RDG criterion
30000
25000
20000
HCS (Integral A)
15000
10000
5000
0
2098 2098 + 4047 15% 2098 + 4047 25% 2098 + 4047 33%
Filler content
this quantity is a function of the two integrals A and B but also of the local
thermal gradient G and solidification speed VT:
Δp max = Δpsh + Δp mec - ȡgh ≤ Δp cav
ε p ≤ ε pmax (G, VT , A, B, ǻpcav , Ȝ 2 ,...) (5)
1
HCS =
ε max
p
The computation of HCS for the AA6056 alloy (cf. Table 1) welded at a
speed of 1 m/min with a 3 kW laser is presented as an example. The steady
state temperature field in the butt joint is represented in Fig. 8. An eulerian
approach is used, i.e. the material is transported at the laser speed under
the fixed laser beam. The characteristics of the heat source can be found in
[20]. Using the local solidification conditions along the coalescence iso-
therm, HCS is computed with the help of CalcoSOFT3D [21]. The results
are shown in Fig. 9. As steady state is reached, HCS can be represented in
a plane perpendicular to the laser speed. A maximum appears close to the
bottom of the weld pool. Influence of the laser speed, power and heat
source parameters can be assessed with such an approach.
Finally, note that the present approach does not require any mechanical
computation, as the maximum strain rate the mush can sustain simply de-
pends on thermal quantities (cf. Eq. 5). However, only steady state can be
treated with the thermal criterion. The inclusion of a constitutive model for
the mushy alloy and the investigation of transients are presented in the
next section.
Fig. 8. Steady state temperature field during laser welding of the AA6056 alloy
Application of RDG Hot Tearing Criterion to Welding of Al Alloys 29
Thermo-Mechanical Approach
By including the mechanical behaviour of the mush, the strain rate tensor
undergone by the dendrites during solidification can be computed and
therefore the induced pressure drop; the larger this value, the higher the
risk to initiate a hot tear [22]. The main challenge is to establish a reliable
constitutive model for the mechanical behaviour of the solidifying mate-
rial. Although the mush should be treated as a compressible medium with
the help of internal variable models [23], a simple incompressible model is
adopted as a first step and for sake of simplicity. Indeed, many mechanical
tests in the solidification interval are required to determine the numerous
parameters describing the compressibility of the mushy alloy [23].
In this section, the mechanical strain rates undergone by the mushy zone
at the rear of the weld pool are assessed in a simple configuration, the butt
joint, in order to predict how process parameters can decrease the risk of
hot tearing in steady state as well as transient regimes such as run-in and
run-out. The results presented here deal again with the AA6056 alloy. Its
BTR is considered to be 510°C–550°C and corresponds to solid fractions
higher than 95% (cf. Fig. 4).
Thermal contraction arises as soon as the dendritic network is well de-
veloped and interconnected, that is at a solid fraction of 85% i.e. at a tem-
perature of 600°C (coherency temperature). The rheology of the alloy is
given by the classic viscoplastic Ludwik’s model and is detailed in [20].
30 J.-M. Drezet, D. Allehaux
The mesh used for the butt joint configuration is presented in Fig. 10.
Due to symmetry, only one half of the whole domain is meshed. The laser
is supposed to travel along the z-axis and mesh is refined in the regions
that undergo melting and solidification. The dimensions of the domain are
given in Fig. 10. The length of the domain (dimension in z) was deter-
mined so that the transient run-in and run-out regimes are well separated
by a steady state regime in the so called on-going zone.
The laser heat input is modelled by a volume heat generation within a
cylinder that is moving over the surface of the parts to be welded [20]. As
the mechanical field does not influence the thermal field (owing to the
absence of any air gap formation), the thermomechanical computation is
un-coupled. The thermal field is computed first and then used as a loading
for the mechanical calculation. As the two parts are tack welded prior to
laser beam welding, they are considered to be part of the same continuum
and therefore no contact elements are used. All plastic deformations are
reset to zero above 600°C, using the *anneal temperature feature in
Abaqus 6.5 [24].
25
y 15
x 3
z
Fig. 10. FE mesh for the butt joint configuration (dimensions in mm)
HTI1 =
Mean
(trε )
510°C ≤ T ≤ 550°C
(6)
HTI2 = Max
(trε )
510°C ≤ T ≤ 550°C
These values are saved during cooling only (i.e. during solidification)
and at each integration point using the user-subroutine UVARM of Abaqus
6.5. The larger this value, the higher the risk to initiate a hot crack.
The case presented here corresponds to a laser speed of 50 mm/s, i.e. 3
m/min, with a constant power all over the specimen length. The thermal
field when the laser reaches the location z = 18 mm is presented in Fig. 11.
The weld pool corresponding to temperatures higher than 650°C is repre-
sented in grey.
Fig. 11. Temperature distribution when the laser is located at z = 18 mm. The
liquid pool appears in grey
The distribution of the Von Mises equivalent plastic strain rate (positive
quantity) and the two hot tearing indexes, HTI1 and HTI2, is presented in
Fig. 12 as a function of the position, z, along the specimen slightly below
the top surface (y = 2.5 mm, x = 0). As expected, the three regimes of
welding are evidenced: in the run-in, the strain rate and the hot tearing in-
dexes exhibit high values; then they decrease and plateau in the on-going
zone. Finally, in the transient run-out regime, the three quantities increase
again. Both HTI1 and HTI2 present the same trends. One can notice that
the equivalent plastic strain rate cannot be considered as an indicator for
hot tearing since it is always positive by definition, whereas HTI1 and
32 J.-M. Drezet, D. Allehaux
HTI2 can have negative values (i.e. no risk of hot cracking). This is the
case when the strain rate reaches its maximum in the run-out whereas the
two HT indexes get negative. In the steady state regime, HTI1 is 0.05 /s
whereas HTI2 is close to 0.16 /s. HTI1 and HTI2 both exhibit a maximum
in the run-in and run-out regimes, which means that hot cracking is prone
to occur at those locations. This is in accordance with industrial observa-
tions, where no hot tears are observed in the steady state regime but appear
in the run-in and run-out. Finally, the HT indexes are larger in the run-in
than the run-out, which seems to be less prone to hot cracking.
3
run-in steady state run-out
2.5
HTI1
1.5 HTI2
0.5
0
0 5 10 15 20 25
-0.5
Specimen length, z (mm)
Fig. 12. Distribution of the equivalent plastic strain rate and HT indexes along
the specimen
1 γ gb − 2γ sl
ΔTb = (7)
Δs f δ
where γgb is the grain boundary energy, γsƐ is the solid-liquid interfacial en-
ergy, Δsf is the volumetric entropy of fusion and δ is the thickness of the
diffuse interfaces. For an alloy, coalescence is reached when a coalescence
line (or surface) parallel to the liquidus, but ΔTb below, is reached. Within
a grain, there is no grain boundary energy and interfaces become attractive
as soon as they get within interaction distance, i.e., distance δ. At “repul-
sive” grain boundaries, γgb > 2γsƐ, bridging is reached at some ΔTb > 0.
This concept of bridging undercooling has been tested by experiment
[27] where two crystals of a nickel-base superalloy are laser welded to-
gether under well defined conditions with increasing misorientations. At
small misorientation (typically less than 15º), no hot crack forms along the
weld centreline, whereas at larger values, a crack is initiated under the
same conditions, thus showing the influence of the grain boundary energy.
The simulations done by molecular dynamics or by phase field are interest-
ing but correspond to very small regions near the incoming interfaces,
typically a few tens of nanometers or micrometers, respectively. In practi-
cal situations, hot tears are indeed located at grain boundaries, but the con-
figuration of these boundaries associated with nucleation and growth of
grains is essential. For this reason, a simplified approach of coalescence
for a large population of equiaxed grains was undertaken first by Mathier
et al. [7] and then by Vernède et al. [8, 9]. In this granular approach, a
random set of nucleation centres with random orientations is first gener-
ated in a given volume. Considering that the grains nucleate at the same
time and that the temperature difference across each grain is small with re-
spect to the growth undercooling, the grain boundaries correspond to the
Voronoï tessellation of the nucleation centres (Fig. 13a), i.e., the grain
boundary between grains I and K is the median line. Assuming globular
grains, the smooth solid-liquid interfaces is first approximated by linear
segment in each triangle linking a nucleation centre (open circle) and two
vortices of the tessellation (open squares) (Fig. 13b).
34 J.-M. Drezet, D. Allehaux
Fig. 13. Schematics of the granular model used for the simulation of solidifica-
tion and feeding in a network of equiaxed globular grains: Voronoï tessellation
(a), microsegregation model (b), feeding KPL model (c)
of a few grains are formed but the liquid films remain continuous and in-
terconnected. In region (c) characterised by 0.97 < gs < 0.99, larger clusters
are visible, with a few isolated liquid films remaining inside. Finally, in
region (d) (0.99 < gs < 1), the solid network is continuous and liquid only
remains as isolated regions. As can be seen, this granular model is able to
predict the gradual transition from a continuous intergranular film network
to a continuous fully coherent solid. It should be emphasised that cluster
formation is directly induced by the stochastic nature of the nucleation
centre locations, a feature that has not been considered in past simulation
works related to hot tearing. Further analysis of the transition regions is
given in [11].
Fig. 14. Calculated mushy zone for an Al-1wt%Cu alloy cooled down at –1 K/s
in a gradient of 6000 K/m. Grains in solid contact are shaded with the same grey
level [11]
36 J.-M. Drezet, D. Allehaux
Conclusion
Hot tearing is a complex defect that involves many phenomena, in particu-
lar thermal and solidification aspects, stress-strain in an increasingly co-
herent solid, feeding in a gradually disappearing liquid film network. The
RDG criterion provided the first two-phase approach, which was further
improved using a more rigorous formalism and the complex rheology of
porous media. Nevertheless, these approaches are still based on averages
and do not consider any localization of strains and feeding at grain
boundaries. Granular models, while still limited to small portions of a so-
lidification process, have certainly the potential to answer some of these
questions, once mechanical aspects will be fully built in and the model will
be extended to 3 dimensions. The numerical simplicity of such approaches
makes it feasible from a CPU time point of view.
Acknowledgements
The European Commission and the whole consortium of the Wel-Air project,
Specific Targeted Research Project from the 6th framework, are acknowl-
edged for financial support. Isabelle Bordesoules from Alcan Péchiney CRV is
greatly thanked for providing the Prophase computations.
References
1. M. Rappaz, J.-M. Drezet, and M. Gremaud: Met. Mat. Trans. 30A (1999), 449.
2. J. Campbell: Castings (Elsevier, 2003).
3. D.G. Eskin, Suyitno and L. Katgerman: Progr. Mater. Sci. 49 (2004) 629.
4. M. M'Hamdi, A. Mo, and C.L. Martin: Met. Mat. Trans. 33A (2002) 2081.
5. M. M'Hamdi, H.G. Fjaer, A. Mo, D. Mortensen, and S. Benum in TMS 2004.
6. V. Mathier, J.-M. Drezet, and M. Rappaz: Two-Phase Modelling of Hot
Tearing in Aluminium Alloys using a Semi-Coupled Approach, Modelling
Simul. Sci. Eng. 15 (2007) 121–134.
7. V. Mathier, M. Rappaz, and A. Jacot: Mod. Simul. Mater. Sc. Engng 12
(2004) 479.
8. S. Vernède and M. Rappaz: Transition of the Mushy Zone from Continuous
Liquid Films to a Coherent Solid, Phil. Mag. 86 (2006) 3779–3794.
9. S. Vernède, P. Jarry, and M. Rappaz: A Granular Model of Mushy Zones-
Formation of a Coherent Solid and Localization of Feeding, Acta Materialia
54 (2006) 4023-4034.
Application of RDG Hot Tearing Criterion to Welding of Al Alloys 37
10. J.-M. Drezet and M. Rappaz: In the proceedings of the First EsaForm Conference
on Material Forming, Ecole des Mines de Paris, CEMEF, Sophia Antipolis,
France, (Mar. 1998) pp. 49–52.
11. M. Rappaz, J.-M. Drezet, V. Mathier, and S. Vernède: In Materials Science
Forum vols. 519–521 (July 2006) pp. 1665–1674, Trans Tech Publications,
Switzerland.
12. E. Niyama, T. Uchida, M. Morikawa, and S. Saito: AFS Int. Cast Metals J.
(Sept. 1982) 52.
13. I.I. Novikov and O.E. Grushko: In Materials Science and Technology, (Sept.
1995) vol. 11, p. 926.
14. B. Magnin et al.: In Materials Science Forum vols. 217–22 (1996) pp. 1209–1214,
Trans Tech Publications, Switzerland.
15. W. Rindler, E. Kozeschnik, and B. Buchmayr: Computer simulation of the
Brittle Temperature Range (BTR) for hot cracking in steels, in Steel Res.
2000, vol. 71 (11), pp. 460–465.
16. J.-M. Drezet, M. Gremaud, R. Graf, and M. Gaümann: A new hot tearing cri-
terion for steel, proceedings of the 4th European Continuous Casting Confer-
ence, IOM communications, Birmingham, UK, (Oct. 2002) pp. 755–763.
17. W. Rindler, E. Kozeschnik, N. Enzinger, and B. Buchmayr: A modified hot
tearing criterion for steels, in Math. Modelling of Welding Phenomena 6,
(2002) pp. 819–835.
18. EU- FP6 Project: WEL-AIR.
19. C. Sigli, R. Dif, R.B. Commet, and T. Warner: In Materials Science Forum
vol. 426–432, Issue 1, (2003) pp. 351–356.
20. J.-M. Drezet et al.: In Mathematical Modelling of Weld Phenomena 8, Ed.
H. Cerjak (2007) pp. 137–152.
21. Calcosoft3D user manual, distributed by Calcom ESI, http://www.calcom.ch.
22. J.-M. Drezet and M. Rappaz: “Prediction of hot tears in DC cast billets” in
Light Metals, Ed. J.-L. Anjier, TMS, (2001) New Orleans, pp. 887–893.
23. O. Ludwig, C.-L. Martin, J.-M. Drezet, and M. Suéry in Met. Mat. Trans, vol
36A, (June 2005) pp. 1525–1535.
24. Abaqus user manual, http://www.abaqus.com.
25. C. Monroe and C. Beckerman: Development of a hot tear indicator for steel
castings, in Mat. Sc. and Eng. A, vol. 413–414 (2005), pp. 30–36.
26. M. Rappaz, A. Jacot, and W. J. Boettinger: Last-stage solidification of alloys:
theoretical model of dendrite-arm and grain coalescence, in Met. and Mat.
Trans. vol. 34A, (Mar. 2003) pp. 467–479.
27. N. Wang, S. Mokadem, M. Rappaz and W. Kurz: Acta Mater. 52 (2004) 3173.
Weld Solidification Cracking: Critical Conditions
for Crack Initiation and Growth
Abstract
Introduction
Thermo-Mechanical Analysis
Crack Initiation
Critical Strain
This topic has traditionally been approached assuming that the mushy
zone, and its associated liquid films at grain boundaries, can withstand a
limited strain before failure. This follows from early work of Pellini [1],
and has been further developed by Prokovorov [2] and Senda et al. [3] who
have established characteristic ductility curves for specific alloys. These
curves define the critical strain that can be tolerated over the solidification
range (brittle temperature range- BTR), bounded by liquidus and non-
equilibrium solidus temperatures. Examples of several different ductility
curves are given in Fig. 1 for different alloy systems. It is observed that
minimum ductilities vary between 0.1–0.5% for aluminum alloys [4],
2–5% for plain carbon steels [5], and 1–9% for stainless steels [5].
Weld Solidification Cracking 41
TL TS
BTR
B
Strain
A
εmin dε/dT
C
Temperature
(a) (b)
(c) (d)
Fig. 1. Ductility curve comparison showing (a) general schematic, (b) aluminum
alloys [4], (c) plain carbon steel alloys [5], and (d) austenitic stainless steel alloys [5]
dε dt dε ε
= ⋅ = (1)
dT dT dt T
42 C.E. Cross, N. Coniglio
(a) (b)
Fig. 2. Strain rate dependence of hot crack ductility for (a) aluminum alloys [4]
and (b) plain carbon steel alloys [9]
Fig. 3. Predicted strain along centerline of aluminum 2024 weld for two different
weld starting positions: at plate edge and 12.5 mm from plate edge [12]
terms of strain per temperature (dε/dT). Measured values for critical strain
rate are compared in Table 1, where it is observed that values range be-
tween 0.1 and 5.0%/s. These values can be taken as a direct comparison of
weldability, with low values representing poor weldability.
Table 1. Critical strain rate to initiate weld solidification cracking derived from [4]
Critical Strain
Test Aluminum Alloy
Rate (%/s)
2017 0.15
5083 0.47
Slow Bend Trans-Varestraint 2219 0.50
Test 5052 0.64
5154 0.70
1070 5.00
§ ρS ·
where E ( x) = ³ f εdx , ß = shrinkage factor ¨©
S ρ L − 1¸¹ , K = perme-
ability, v = growth velocity, fS = fraction solid, and μ = viscosity.
TL
low driving
force (Δp)
TC
increased
coherency
TS
Crack Growth
Only a limited amount of work has been devoted to understanding the fun-
damentals of crack growth in a liquid film. This curious lack of attention
may come from a presumption that conditions appropriate for crack initia-
tion must also result in crack growth. This assumption finds support in
many weldability tests (e.g. circular patch test), where it is observed that
once a crack is initiated, the crack will follow behind the weld pool until
the torch is extinguished. Another common problem is the failure to rec-
ognize crack initiation and growth as two distinguishable events, involving
two different mechanisms.
Two approaches for modeling liquid crack growth can be found: one
involving breaking of bonds [20]; and one involving the flow of liquid
[21]. Considering the relatively slow speed of crack propagation in
welding (i.e. at torch velocities around 4 mm/s), it may be difficult to
justify a mechanism based upon bond breakage. For the liquid flow
model, within a differential period of time, the transverse strain will be
46 C.E. Cross, N. Coniglio
Table 2. Critical strain rate to stop cracking as a function of the aluminium base
metal, filler, and welding speed, as measured by Matsuda et al. [6]
Critical
Welding Speed Aluminium Alloy
Test Deformation
(mm/min) (Base Metal/Filler)
Rate (mm/s)
240 1100/1070 0.15
400 1100/1070 0.25
VDR
600 1100/1070 0.40
Variable
800 1100/1070 0.50
Deformation Test
400 5052/1070 0.05
400 5083A/5183B 0.18
have been used over time, including real-time observation of scribe marks
[23] and moiré-fringe analysis of grid patterns [24].
Use has also been made of extensometers spanned across the weld, at-
tached either above or below the plate surface [25]. When applied to the
top surface, small diameter pins are attached on opposite sides of the joint,
and extension arms connect these pins to a remote extensometer (i.e. re-
moved from the torch path and possible exposure to heat). This allows
measurement of strain as the welding torch passes between the affixed
pins. When applied below the surface, the extensometer can be placed di-
rectly at the point of measurement, but this requires use of through-
thickness welds and a thin plate (i.e. plane strain) to provide any useful in-
formation. An example of one such measurement for an autogenous Al
6060 weld is given in Fig. 5a, where an extensometer (10.5 mm gage) has
been placed at the mid-span of a 100 mm long joint and the torch passes
over the extensometer at approximately 15 s. A compression cell is ob-
served ahead of the weld pool (negative strain), followed by a tensile cell
behind the weld pool, with a steadily increasing strain rate (approximately
0.5%/s in mushy zone).
(a) (b)
Fig. 5. (a) Output from extensometer placed across the root of an autogenous, full-
penetration GTA weld (10.5 mm gage) made on 4 mm aluminum 6060-T6 plate,
(b) first derivative of strain-time curve in (a) showing transverse strain rate
One technique has been developed to measure strain directly in the weld
mushy zone making use of small discontinuities found on the weld surface
[27]. The relative movement of these discontinuities is tracked using high
speed photography, typically providing an effective gage length of around
1 mm. Measurements made using this technique, referred to as measure-
ment by means of in-sito observation (MISO), are given in Table 3 for
values representing the minimum strain to initiate cracking (at high strain
rate) in a tensile test. When compared against corresponding values based
upon applied strain in a trans-varestraint test (TVT), it is observed that
MISO values are significantly higher. This discrepancy may be due in part
to a hinge effect, discussed below in the weldability section, and reflects a
difficulty to control applied strain in the TVT.
(a) (b)
Fig. 6. (a) Light optical DIC measurements made on bottom side of plate during
autogenous, full-penetration, bead-on-plate GTA welding of aluminum 6060 plate.
Note black paint speckle pattern sprayed onto plate prior to welding. (b) Strain
across mushy zone of weld pool showing a constant outward displacement with
distance from weld centerline
Weldability Testing
A broad variety of weldability tests have been developed over the years
that rely upon special conditions of joint configuration or applied strain to
generate cracks as outlined in Table 4. In most cases, weldability is deter-
mined based upon an evaluation of crack length measurements (i.e. crack
length parametrics). In other cases specifically involving application of
controlled strain rate, weldability can be related directly to a critical strain
rate (crack versus no-crack) for a given alloy and welding condition. As
argued above, this latter case holds the greatest promise for predicting
cracking because it relates to an actual cracking mechanism. Crack length
measurements, on the other hand, can only be used to rank the cracking
susceptibility of different alloys (i.e. provide a relative comparison), and
even this has come under increased scrutiny as discussed below.
weldability tests
intrinsic
circular patch test (CPT)
houldcroft test
extrinsic crack length
longitudinal varestraint test (LVT) parametrics
transverse varestraint test (TVT)
modified varestraint test (MVT)
sigma-jig test
Crack length measurements are most often reported in terms of total accu-
mulated crack length (TLC), as it is generally assumed that for fixed test-
ing conditions, an alloy with higher susceptibility must somehow result in
more extensive cracking. In longitudinal varestraint specimens, for exam-
ple, the TCL constitutes the sum of several smaller crack lengths, where
the number of cracks is bounded by the number of grain boundaries and
the length of each crack is bounded by size of the mushy zone. It is found
50 C.E. Cross, N. Coniglio
(b)
(a)
(c)
Fig. 8. (a) Schematic illustration showing hinge effect in TVT testing, (b) strain
gage placed on Al 6061 plate for TVT test with 0.5% strain die block, (c) strain
gage place on weld metal in pre-welded Al 6061 coupon, TVT tested with 0.5%
die block [28]
Weld Solidification Cracking 53
The following topics are believed to have major influences on weld solidi-
fication cracking, but are themselves not well understood and need further
clarification in order to expand our understanding of cracking behavior.
Most of the local strain rate measurements considered above, using exten-
someter and DIC methods, have been made in the vicinity of the mushy
zone, but not in the mushy zone itself and not across a single grain bound-
ary. In order to relate local measurements to actual strain conditions across
a single grain boundary, for ideal application of theory, it is important to
know how strain (and hence strain rate) is partitioned within the mushy
zone. For this purpose, it becomes convenient to treat the mushy zone as a
composite material consisting of parallel strips of liquid and solid phases,
with behavior proportioned to the relative amount of each phase:
ε = f S εS + f LεL (4)
Role of Impurities
In ferrous alloys, the impurities sulfur and phosphorous play a major role
in determining weldability and this has been fairly well documented. In
aluminum alloys, however, the role of impurities is less clear and this re-
mains a little studied subject [33]. Impurities oxygen and hydrogen, for ex-
ample, could conceivably affect weldability; where oxide films may serve
as crack initiation sites, and hydrogen may serve to reduce these oxides.
Hydrogen also results in pore formation that may help feed shrinkage. Dis-
solved hydrogen could also make it easier to initiate cracks through cavita-
tion. The impurity iron has been shown to affect both porosity and crack-
ing behavior in aluminum castings [34], but its affect in welding aluminum
remains largely unexplored.
Solidification Path
(a) (b)
Fig. 9. Solidification structure for Al 6060 (a) casting (9°C/s) and (b) gas tungsten
arc weld (54°C/s). Shown in (a) are dendrites with interdendritic constituents and
in (b) grain boundaries with finer constituents [35]
Weld Solidification Cracking 55
Crack Prediction
Future approaches for predicting the onset of cracking for a given indus-
trial application will involve applying material information to a cracking
mechanism with the aid of numerical simulation. Inputs to this simulation
will include pertinent welding information (heat input, travel speed), alloy
composition (base metal and filler dilution), and restraining conditions
(weld geometry, fixture design). The ability to accurately predict cracking
behavior will require a thorough understanding of the cracking mechanism
itself, something that does not yet exist.
Meanwhile, and until a reliable cracking model is developed, a more
pragmatic approach may prove useful. This could, for example, involve
development of composition-strain rate maps as demonstrated in Fig. 10
[8]. Based upon Controlled Tensile Weldability (CTW) testing, a boundary
has been established for the aluminum 6060/4043 alloy system that defines
how much 4043 filler metal is required to avoid cracking for a given local
strain rate. Using this map, a welding engineer can make informed deci-
sions regarding weld procedure development (including filler dilution)
based upon local strain rate measurements. Such measurements could con-
ceivably be made using a laser scanner across the mushy zone. While the
given data is for the Al 6060-4043 alloy system, a similar type of map
could be generated for any ferrous or non-ferrous alloy.
Fig. 10. Critical strain rate – ductility map showing conditions required for weld
solidification cracking in Al 6060 + 4043 filler using gas-tungsten arc process
(full penetration, 4 mm plate, 4 mm/s travel speed) [8]
56 C.E. Cross, N. Coniglio
Summary
References
29. Lippold JC (2005) Recent developments in weldability testing. In: Hot cracking
phenomena in welds, Springer, pp 271–290
30. Unpublished research (2006), BAM, Germany
31. Brooks, JA (1990) Weld microsegregation: modelling and segregation effects
on weld performance. In: Weldability of Materials, ASM, pp 41–47
32. Savage WF, Aronson AH (1966) Preferred orientation in the weld fusion
zone. Welding J 45: 85s–89s
33. Coniglio N, Cross CE (2007) Weld parameter and minor element effects on
solidification crack initiation in aluminium. In: this conference proceedings,
pp. 277–310
34. Lu L, Dahle AK (2005) Iron rich intermetallic phases and their role in casting
defect formation in hypoeutectic Al-Si alloys. Met Mat Trans 36A: 819–835
35. Coniglio N, Cross CE (2006) Characterization of solidification path for alu-
minium 6060 weld metal with variable 4043 filler addition. Welding in the
World 50: 14–23
36. Alexandrov BT, Lippold JC (2006) A new methodology for studying phase
transformations in high strength steel weld metal. In: Proc 7th int conf trends
in welding research, ASM Int, pp 975–980
Consideration of the Welding Process
as a Thermo-Physical Mechanism to Control
Cracking in Weldments
H. Herold, M. Streitenberger
Abstract
Introduction
Failure analysis and assessment concepts are made during the lifetime of
welded structures if cracks have formed and grown to macroscopic scale,
endangering the reliability of structural components. Cracks originate at
the weakest link of material behaviour in the microstructure, caused by
locally positioned thermo-mechanical overstress within the welded
structural component on which general thermo-mechanical stress is
imposed by welding.
On the example of solidification cracking, the paper follows the
philosophy: When we can describe and analyse the hot cracking
phenomena in both material behaviour and welded structural
components, concentrating on the crack – initiating thermo-mechanical
processes due to welding and on the state of art in hot-cracking
assessment, then we can develop more efficient engineering tools for
their prevention in stages of manufacture.
Weldment Cracking
The welding procedure supplies the thermal energy for local fusion with
the intensity and efficiency of the heat source (arc, beam). The material of
the assembled weld component absorbs the heat by its thermal capacity
and mass (volume), with heat losses due to heat transfer by resistance,
convection and emission. The heat dissipates thermodynamically into the
structures’ volume by heat conductivity, changing the physical properties
of the metal by thermal expansion. Quasi-steady temperature fields and
temperature gradients around the moving welding pool characterize the
changing local behaviour, which results in superposition of both the
microstructural material transformation and thermo-mechanical structural
kinetics, balancing the local thermal, mechanical and constitutional
yielding.
Thermal cycles are used to describe the constitutional and thermo-
dynamic reaction of the composition of the alloy on solidification and
HAZ. They also form the basis for calculation programs for temperature-
time-transition, correlation of microstructure and mechanical behaviour,
grain growth, and cooling time concepts in general. Quasi-steady
temperature fields describe the 2D or 3D thermo-physical and thermo-
dynamical heat transport around the weld pool in relation to the welded
component, which is used for the calculation of temperature-induced
mechanical reactions as distortion, displacement, stress and strain with the
help of Finite Element Analysis (ANSYS, ABACUS).
Thermo-Physical Mechanism to Control Cracking in Weldments 61
welding
material structure
ambient residual
weld metal atom
media stress
super-
position
1400 and 1000°C by reduced heat input per unit length and increased weld
speed all over the welded seam as well as by special designing and
assembling for welding [5, 6].
Welding reliability in structural transient stages has become more
important as refers to hot cracking prevention and hot cracking assessment.
Weldability Tests
tear- processing
zones Time,
weld procedure
associated load
In the past, Pellini [1] was the first to describe the effects of solidification
cracking by the hot spot in casting. The residual melt, enriched with
segregating elements, lowers the effective solidification temperature and
delays solidification time. The assessment quantifies the stress-concentration
over the length of the casting due to the linear expansion coefficient during
cooling down from liquid`s temperature, thus producing strain by
shrinkage. Campbell [19] describes the local conditions for solidification
crack formation in castings with five feeding mechanisms fundamentally,
by liquid feeding, mass feeding, interdendritic feeding (eutectic, solid
solution), solid feeding (elastic, plastic) and burst feeding. These
descriptions initiated two different directions for further investigation of
solidification mechanisms, giving priority either to the effects of thermo-
mechanical changes at solidification, or to the changing constitution during
solidification.
Assessment Concepts
Hot cracking occurs when as well the critical cross displacements during
the existence time of the BTR Δuy min are exceeded by the total cross
displacement Δuy BTR, due to component design and assembly for
welding, and/or the critical speed of cross displacement is exceeded by the
local speed of cross displacement [31, 32, 48].
The welding-process associated local speeds of cross displacement vq
are calculated by FEA with the help of the program ANSYS at discrete
points along the seam during one-side welding of the actual welding
components, taking into account the effects of design and assembly by
dimension of run-outs, tacks and magnetic clamping. The speed of cross
displacement vq in Fig. 3 will maximize systematically when the arc comes
near the plate’s end, and the far side of the weld pool passes from the plate
to the run-out due to the stiffness jump. It increases locally during fusion
of each tack, additionally releasing assembling stresses. These short and
rapid increases of cross displacement with regard to time and versus the
welded seam are the most important factors to initiate centerline
solidification cracking of longer butt weld seams. However, the estimation
of the critical vcr needs an empirical assessment diagram for determination
of the critical speed of cross displacement vcr in correlation with the
applied material and welding process.
The assessment diagram in Fig. 4 offers the complexity of hot
cracking by three sub-diagrams. The assessment diagram for centreline
solidification cracking can be seen on the right side. The multitude of
calculated curves of the weld component and process-associated speed of
cross displacements vq, measured with the help of the welding-process
associated hot-cracking test method, are plotted here as cooling time of
the weld metal versus the change of the measured and calculated cross
displacement ΔU (t) in [mm]. The critical speed of cross displacement vcr
is defined here as the threshold curve, separating the regions with
occurring hot cracking from crack-free regions, established by empirical
approximation. The sub-diagram in the centre shows the correlation with
the hot cracking theory by Prokhorov, setting up internal deformation
versus temperature. The left sub-diagram shows the thermal cycle of the
applied welding procedure.
Thermo-Physical Mechanism to Control Cracking in Weldments 69
test under the use of fish bone specimens. In their model, they adopted
(crack) opening displacement from elastic-plastic fracture mechanics, con-
sidering the crack propagation as the formation of new surfaces, without
taking into account the effects of temperature on the strength of the inter-
face element. The influence of BTR width and the opening displacement at
the crack were examined, from which followed that the opening displace-
ment must be adjusted together with the BTR width, so that the tempera-
ture-dependent interface element represents the solidification cracking sus-
ceptibility of the material.
Ploshikhin et al. [37] developed a thermo-mechanical and metallurgical
model for the solidification cracking mechanism in welds. These
solidification cracking criteria are based on the tensile strain accumulated
in the segregating inter-granular liquid film at the weld centreline by the
solidified dendrites from the completely solid region of the base material
until cracking appears. The local critical parameter δacc represents the
ability of the segregating liquid film to accumulate strains without tearing,
measured by the micro-structural morphology (by fraction of residual
liquid, primary and secondary dendrite arm spacing). Solidification
cracking will result when the maximum accumulated strain exceeds the
critical value.
Conclusion
References
Abstract
Introduction
Experimental Procedure
The investigations were carried out selecting Ni-base Alloy 602 CA (Ma-
terial No. 2.4633) as test material with composition given in Table 1 [6]. It
was available in solution-annealed condition. In preceding MVT- and
PVR-hot cracking tests, the metallurgical effect of N on hot cracking resis-
tance was already identified [7]. It was hence possible, by the use of two
Determination of Critical Strain Rate by Numerical Simulation 79
The numerical calculations were verified with the help of the MVT-
Test [8], in which the specimen loading rate was varied additionally [1, 2].
For quantification of the mechanical specimen loading, strain gauges
were applied to the specimens subjected to MVT-testing using various
die radii. The size of MVT-specimens was 100 mm × 40 mm × 10 mm.
The various welding parameters can be seen from Table 2. The tests P1 –
P4 were conducted in combination with the shielding gases Ar as well as
Ar + 1% N2.
P1 250 18 5 9.0
P2 182 17 3 10.3
P4 200 17 6 5.7
Numerical Approach
§ 1
½ ·
μ
1
¨ °ª ºω ° ¸
ª § τ
· º ¨ °« φ » ¸
«1 − ¨ z ¸ » ¨ °« § x ·
¨ ¸ »
°
¸
« ¨© Lx , SB ¸¹ » ¨B ¸ °
¨ °« » ° θ¸
T ( z , x) = − H SB ⋅ ¬ ¼ © SB ¹
⋅ ¨ ®«1 − » − Δ¾ + Δ ¸
(1 − Δ )θ + Δθ ¨ °« ª ν η
º
φ
» ° ¸ (1)
¨ °« « §¨ z ·¸ » » ° ¸
¨ °« «1 − ¨ L ¸ » » ° ¸
¨¨ °«¬ ¬ © x , SB ¹ ¼ »¼ ° ¸¸
©¯ ¿ ¹
Determination of Critical Strain Rate by Numerical Simulation 83
The geometry parameters BSB, Lx,SB and HSB were determined by meas-
urement of the weld end crater length, penetration depth and weld width.
The method of directly implementing the weld pool geometry in the nu-
merical simulations is particularly suitable with regard to the defined
specimen geometries and welding conditions in MVT-Tests. It was, for
example, possible to generate reproducible weld pool geometries by varia-
tion of test parameters such as cross-beam travel speed vtrav or total strain
εtot without the need of geometry and shape parameter matching [12, 13].
In order to determine the geometry and shape parameters, TIG-remelt
welds were produced on MVT-specimens using the parameters given in
Table 2.
84 M. Wolf et al.
Fig. 5. Weld end crater (left hand side) and transverse section (right hand side) of
a TIG-remelt weld of Alloy 602 CA; welding parameters P1; shielding gas:
Ar + 1% N2
A raster was projected with a grid spacing of 0.2 mm (Fig. 5). This
raster enabled light-microscopical analyses of the weld pool geometry pa-
rameters for the individual welding parameters P1–P4 (penetration depth
HSB, weld width BSB, total weld pool length Lx,SB = LV,SB + LR,SB). For
shape parameter determination, the coordinates of the weld end crater
functions and of the fusion line functions were measured in the transverse
section. Figures 6 and 7 show a comparison between the measured coordi-
nates and each matched curve by respective examples relating to the weld-
ing parameters P1 with Ar + 1% N2, for both the crater trailing edge and
the fusion line. The weld pool shape determined by this method was de-
pendent on the shielding gas.
Fig. 12. Strain distribution for P1 Fig. 13. Strain distribution for P4
with Ar, vtrav,crit = 0.05 mm/s with Ar, vtrav,crit = 0.1 mm/s
pool, because increasing welding speed (P4) involves a more tapered weld
pool trailing edge geometry [14, 15].
This example shows the advantages of direct implementation of the
weld pool geometry in the numerical simulation. Geometric influences of
the weld pool geometry on the local strain rates can thus be adequately
considered in the numerical calculation.
Further quantification of the relationship between the global specimen
loading rate vtrav and the local strain rate was performed by numerical
simulation of four different specimen loading rates between 0.05 mm/s and
0.4 mm/s for the parameter set P1, where 0.05 mm/s represents the critical
specimen loading rate vtrav,crit.
Fig. 15. Local strain versus speci- Fig. 16. Local strain rate versus speci-
men loading rate for P1 with Ar 4.8 men loading rate for P1 with Ar 4.8
Table 5. Local strain rate depending on the global strain rate for P1
vtrav global strain rate local strain rate εloc
εglo [1/s] εloc [1/s] εglo
0.50 0.020 0.065 3.3
0.25 0.010 0.031 3.1
0.10 0.004 0.013 3.3
0.05 0.002 0.005 2.5
• A new method could be developed which was used to reconstruct the weld
pool geometry from transverse sections as well as from the crater of the
weld end and to incorporate it as a three-dimensional analytical function
into numerical simulation. The comparisons between numerical simulation
and the measured temperature distributions show high correlation.
Determination of Critical Strain Rate by Numerical Simulation 91
• Validation was carried out using the MVT-hot cracking test in which the
specimen loading rate was varied additionally. The solidification crack-
critical limiting temperature could be calculated for Alloy 602 CA de-
pending on various welding parameters and on the investigated shield-
ing gases.
• In view of a transferability of hot cracking tests to components, critical
specimen loading rates were determined and the local crack-critical
strain rates in the immediate vicinity of a weld pool were calculated.
• It could be demonstrated that the local critical strains and strain rates
can be a crack criterion for a transferability.
References
10. Wei Y, Dong Z (2005) Simulation and predicting weld solidification cracks.
In: Hot Cracking Phenomena in Welds, Springer-Verlag Berlin Heidelberg,
ISBN: 3-540-22332-0, pp 185–222
11. Prokhorov NN, Prokhorov NN (1969) A general equation for the surface of
the solidification front during welding. Svar. Proiz. 8, pp 1–4
12. Wolf M (2006) Zur Phänomenologie der Heißrissbildung beim Schweißen
und Entwicklung aussagekräftiger Prüfverfahren. Dissertation, Verlag für
neue Wissenschaft GmbH, ISBN-10: 3-86509-599-2
13. Wolf M, Schobbert H, Böllinghaus T (2005) Influence of the weld pool geome-
try on solidification crack formation. In: Hot Cracking Phenomena in Welds,
Springer-Verlag Berlin Heidelberg, ISBN: 3-540-22332-0, pp 243–268
14. Katayama S (2000) Solidification phenomena of weld metals (1st report).
Characteristic solidification morphologies, microstructures and solidification
theory. Welding International 14 (12), S. 939–951, 2000
15. Hunziker O, Dye D, Reed RC (2000) On the formation of a centreline grain
boundary during fusion welding. Acta Materialia 48, pp 4191–4202
Part II
Solidification Cracking of Ferrous
and Nickel-Base Alloys
Classification and Mechanisms of Cracking
in Welding High-Alloy Steels and Nickel
Alloys in Brittle Temperature Ranges
b
Fig. 9. Variations in strength and ductility: (a) – high-alloy cast metal with austenitic
structure [5]; (b) – heat-resistant nickel alloy
Classification and Mechanisms of Cracking 103
Fig. 10. Flow diagram of the process of solidification of metal with stable-
austenitic structure in the presence of the concentration overcooling zone (a),
variations in character of dendrite growth in bi-crystalline metal (b) [1], and varia-
tions in strength and ductility of metal in solidification (c) [5]: Tz.d. – zero ductility
temperature; Tz.s. – zero strength temperature. Characteristics of structure of hot
crack surface [6]: D – dendritic; F – flat; D+F - transient
104 K.A. Yushchenko, V.S. Savchenko
Variations in the creep rate with the values of stresses at fixed values of
temperature for stable-austenitic steel AISI310 were evaluated to find
causes and determine the main mechanism of creep in the ductility dip
temperature range. Experimental points obtained as a result of experiments
are shown in Fig. 17.
Exponential dependence of variations in the creep rate for the selected
temperature of 650oC is proved by the straight line describing variations of
ε with σ in the selected coordinate system.
Empirical dependence of the steady-state creep rate upon the stresses
was plotted on the basis of the experimental results using the following
formula:
Q − mσ
ε = A ⋅ e RT
where:
ε is the variation in the deformation rate;
A is the constant;
Q is the creep activation energy, kJ/mole;
σ is the stress at a given temperature, MPa;
R is the gas constant, J⋅mole–1⋅K–1;
m is the exponent, m=f(T);
T is the temperature, K
Fig. 17. Dependence of the rate of steady-state creep of steel AISI310 upon
stresses at test temperature of 650oC. Critical strain rate, below which intergranu-
lar fracture is probable, is shown by the mark in straight line
Classification and Mechanisms of Cracking 109
As seen from the calculations, the creep rate can be described by the fol-
lowing equation:
46000 mσ
Fractography also showed the presence of the slip planes escaping to the
fracture surface (Fig. 19a), which proves existence of the intragranular
plastic deformation and the dislocation mechanism of creep in the tem-
perature range studied. It also can be concluded on the basis of the data ob-
tained that the decisive role in the fracture process is played by a tough
(liquid) interlayer, the traces of the deformation of which can be seen in
Fig. 19b.
Fig. 20. Hot ductility and sodium concentration on grain boundaries of Al-5%Mg
alloy as a function of sodium content [8]
112 K.A. Yushchenko, V.S. Savchenko
Fig. 21. Kinetics of variations in stress-strain state in welding of stainless steel (a)
and nickel alloy (b) at distance of 0.5 mm from fusion line [13]
Conclusions
References
11. Okrainets PH, Pishchak VK (1978) Stress dependence of creep rate for metals
with fcc lattice (in Russian). J Fizika Metallov I Metallovedeniye 46, Issue
3:597–602
12. Skleniþka V, Saxl I, öadek I (1983) The role of grain boundaries in deforma-
tion and fracture at high temperatures (in Czech). Kovove materaly 4, vol 21,
pp 364–382
13. Yushchenko KA, Makhnenko VI, Savchenko VS, Velikoivanenko EA,
Chervyakov NO (2006) Investigation Of Thermal-Deformation State Of
Welded Joints In Stable-AusteniticSteels And Nickel Alloys. IIW Doc.IX-
2224-06
Submerged Arc Welding – A Test
for Centreline Cracking
K. Håkansson
Abstract
Introduction
The present chapter describes briefly the work carried out at Kockums
Laboratory, to design a simple test to examine centreline cracking
behaviour during submerged arc welding of high strength steels. The
welding was done with addition of cold wire containing a high amount of
nickel. It discusses in more detail the results of the addition of cold wire
containing 98% nickel to the weld. The diameter of the arc wire was
changed on purpose to get different amounts of nickel in the weld metal.
An increase in the nickel concentration promotes austenitic solidification
and therefore should promote solidification cracking, the effect being
exaggerated when both the nickel and the sulphur concentrations are large
[1]. It is widely accepted that sulphur and phosphorus are harmful
impurities with respect to the solidification cracking of steel welds.
An excerpt from “The Fabricators and Erectors Guide to welded Steel
Construction” [2] states that centreline cracking is characterized by
separation in the centre of a given weld bead. If the weld bead happens to
be in the centre of the joint, which is normal for a single-pass weld,
116 K. Håkansson
centreline cracks will be in the centre of the joint. In the case of multi-pass
welds, where several runs per layer may be applied, a centreline crack may
not be in the geometric centre of the joint, although it will always be in the
centre of the bead.
Centreline cracking is the result of one of the following phenomena:
-segregation induced cracking
-bead shape induced cracking
-surface profile induced cracking
Unfortunately, all three phenomena reveal themselves in the same type of
crack, and it is often difficult to identify the cause. Moreover, experience
has shown that often two or even all three of the phenomena will interact
and contribute to the cracking problem. Understanding the fundamental
mechanism of each of these types of centre line cracks will help in
determining the corrective solutions [2].
Segregation induced cracking occurs when low melting point
constituents such as phosphorous, zinc, copper and sulphur compounds in
the admixture separate during the weld solidification process. Low melting
point components in the molten metal will be forced to the centre of the
joint during solidification, since they are the last to solidify and the weld
tends to separate as the solidified metal contracts away from the centre
region containing low melting point constituents.
In submerged arc welds, the cracking risk may be assessed by
calculating the Units of Crack Susceptibility (UCS) from the weld metal
chemical composition (weight %) [3, 4, 7]:
Experiment
mm and the thickness 35 mm was placed in the centre. On the other side of
the square plate, a round bar of 25 mm diameter was tack welded in the
centre. This bar was used to mount the test piece in a manipulator during
welding. Welding was made using a rectifier type constant voltage with
1400A as maximum output and both wires were fed by a synergic cold
wire kit (SCW), which feed the cold wire in synergy with the arc wire into
the weld pool where it melts. This means that the arc and the cold wire
ratio always remain constant after suitable wire diameter is selected. See
Fig. 1.
Fig. 1. The test set up where the manipulator drives the test piece during the
welding of the fillet weld by submerged arc welding
The chemistry of the weld and the deposition rate was controlled and
pre-selected [6].
Fig. 2. Adjustment of the arc wire and cold wire in the single bevel groove joint. It
is essential that the round test piece is aligned with the driving shaft of the
manipulator
Submerged Arc Welding 119
The arc wire for all welds was a lean alloyed wire with different
diameters. This wire gives about 0.5% manganese in the weld.
A basic agglomerated welding flux for welding high strength steels and
low temperature steels was used. As cold wire, a wire containing 98%
nickel and 2% titanium with a diameter of 0.8 mm was used. Figure 2
shows the alignment of both wires.
The theoretical chemical composition of the weld metal depends on the
difference in diameter of the arc wire and the cold wire. See Table 1.
The weld metal will have a different composition due to dilution from
the base material. In submerged arc welding this dilution is assumed to be
about 50–60%. The chemical composition of the base material, the arc
electrode and the cold wire is given in Table 2.
Welding Procedures
Welding was performed in flat position without any preheat and in one
single pass. The travel speed was set by the rotation performed by the
manipulator. The travel speed was aimed at 40–45 cm/min; the welding
current was aimed at 350 ampere and the voltage at around 28 V. These
figures give a heat input of 1.3–1.5 kJ/mm.
Results
Material from the weld metal was removed from each of the welded
samples by drilling holes in the weld metal. The chips were used for
chemical analysis of the weld metal. The chemical composition was
analysed by Inductively Coupled Plasma emission Spectroscopy ICP. The
result of the chemical analysis of the weld metal is shown in Table 4.
Submerged Arc Welding 121
Fig. 3. The test piece with the highest amount of nickel shows a solidification
crack almost all around the circumference
122 K. Håkansson
From each weld two macro test pieces were taken transverse to the weld. A
cut divided the test plate into two halves, and the macro tests were removed
from each exposed side. The tests were prepared metallographically and
etched in Nital. The tests were inspected using a stereo microscope at ten
times magnification and the results are shown in Table 5.
a b
c d
a b
Fig. 5. Micrographs of welds (a) HC1-1 and (b) HC4-2 showing root cracks
124 K. Håkansson
10 0
90
80
70
60
50
40
30
20
10
0
0 ,7 2,5 4 3,4 1 4,77
P e r c e nt a ge n i c k e l i n t he we l d me t a l
Microstructure
The hardness test was made according to EN 1043-1 with the load of 5 kg,
with the exception of that no hardness readings were made in the base
material [8]. Three indentations were made in the heat affected zone of the
square plate, in the weld metal and in the heat affected zone of the round
plate. The result of the hardness survey is shown in Table 6.
The hardness in the weld metal increases as the nickel content increases
and this may be due to changes in the microstructure in the weld metal to a
more martensitic microstructure. The hardness indicates that the weld
metal may have a tensile strength of about 1200 MPa for the two highest
nickel levels.
Submerged Arc Welding 125
The hardness of the weld metal with no addition of nickel is also rather
high, which can be due to the high amount of highly alloyed base material
mixed into the weld metal. The fillet weld itself normally experiences a
more rapid cooling, which increases the hardness.
Conclusions
References
1. Kazutoshi Ichikawa, Bhadeshia HKDH, and MacKay DJC (1996) Modell for
solidification cracking in low alloy steel weld metal. Science and Technology
of Welding and Joining Vol. 1 no. 1, pp 43–50
2. The James F. Lincoln Arc Welding Foundation. Weld Cracking. An Excerpt
from the Fabricators’ and Erectors Guide to Welded Steel Construction. pp 1–4
126 K. Håkansson
Abstract
Introduction
stresses are present. Further stresses with opposite algebraic sign develop
in other areas, due to the balance of forces [4-8, 12-18].
It is an undisputed fact that in front of the moving weld pool compres-
sive stresses are directly formed. The magnitude of this stress cell is
strongly affected by the welding parameters, particularly by the travel
speed [4-8]. In reaction to this, tensile stresses in the still cold base mate-
rial appear. The expansion in front of the weld pool continues directly at
its sides. The maximum of the longitudinal extensions is within this range.
At high welding speeds, this strain maximum can extend to the region be-
hind the weld pool. This has direct consequences on the solidified material
already cooling down there. Normally, contraction due to cooling creates a
tensile stress in this region, as with slow welding speeds. However, ac-
cording to Chihoski [17, 18], the expansion of the surrounding material
can thus superimpose the contraction within the cooling range -supported
by longitudinal expansions- resulting in a pressure cell. Thereafter, under
uninfluenced contraction, a tensile cell develops. The investigations of
Feng and Zacharia [4-8] confirm the presence of a pressure cell directly
behind the weld pool. Figure 1 illustrates this circumstance. Thus, solidifi-
cation cracking can be prevented.
The development of compressive and tensile cells around the weld pool
crucially depends on the welding parameters, namely on the heat input.
From this, it becomes clear that the place of a possible crack initiation can
be shifted relative to the weld pool. Whether the mechanisms of crack ini-
tiation are the same in different locations, needs to be clarified, particularly
since any thermal pre-treatment of the component should have an addi-
tional effect [12]. Up to now, the emphasis of most investigations has been
on the influence of welding parameters on hot crack formation. The num-
ber of investigations into the design specific influences is smaller.
It becomes clear that the influence of external load or geometry itself
during the welding procedure acts in a similar way, i.e. with movement of
compressive and tensile cells. First, a dependence of the stress/strain dis-
tributions of the momentary weld pool position on the specimen is to be
emphasized. Consequently, this distribution changes with distance and
proximity to boundary regions. Without the supporting effect of these ar-
eas, the material behind the weld pool seems clearly higher stressed and
more rapidly loaded under tensile stress, which should favour cracking
[4, 6].
130 A. Kromm, Th. Kannengießer
The application of an external, static load has similar effects, as used with
the Sigmajig – test [19]. Applied tensile stresses directly affect the level of
the locally evolving deformations. These can be increased or decreased [4,
5, 7, 8]. The influence of external restraint was examined recently also by
Kannengiesser [20]. Results show that higher restraint intensities can lead
to a decrease of the solidification cracking susceptibility.
Hence it follows that potential cracking can either be increased or de-
creased by the design specific or mechanical boundary conditions. In [17,
18] this statement is also confirmed. It seems that the causes leading di-
rectly to cracking (local strains and strain rates, respectively) can alter with
the geometry of the material (e.g. round vs. rectangular specimens) [13,
14]. That would mean that crack-critical local values for the same material
vary with the geometry (restraint intensity). In the present study a contri-
bution is made to the question of how local strains and strain rates are in-
fluenced with respect to external load and how solidification cracking is
affected.
Influence of Local Weld Deformation on Solidification Cracking Susceptibility 131
Experimental
Test Material
The solidification cracking behaviour was examined both under free con-
traction and under varying external load. For this purpose, a newly devel-
oped, externally loaded hot cracking test (CTW-Test) was used for the first
time. In this test a defined tensile load can be applied during welding,
transverse to the welding direction [11]. The CTW-Test setup is shown in
Fig. 2.
132 A. Kromm, Th. Kannengießer
The specimen geometry and dimensions are shown in Fig. 3. The welding
parameters are summarized in Table 4. In order to avoid additional notch
effects and cracking at the weld start, special clamping adapters were used,
and the weld start and end were placed 5 mm away from the edge of the
sample resulting in a weld length of 45 mm.
Influence of Local Weld Deformation on Solidification Cracking Susceptibility 133
Strain Monitoring
In order to relate the weld pool dimensions and position to the strain meas-
urements, the weld pool movement was recorded optically. This was real-
ized using a high-speed camera. Furthermore, the records served to deter-
mine of the following parameters:
• Time of Solidification
• Time of crack initiation
• Location of crack initiation
• Characteristics of crack initiation
Before the investigations were started, the samples were provided with a
mark perpendicular to the welding direction, which was used as point of
reference in order to determine the weld pool dimensions. Furthermore, the
mark represents the place of the hot crack initiation determined in prelimi-
nary tests.
The high speed records helped to determine the dimensions of the weld
pool and to observe crack initiation and crack growth. The following Figs.
5–7 illustrate the formation and progress of a solidification crack at a con-
stant cross head speed of va,crit = 4 mm/min in the CTW-Test. As shown in
Fig. 5a, the crack arose at the solidification line. This is at the same point
of time as the weld pool begins to narrow.
Influence of Local Weld Deformation on Solidification Cracking Susceptibility 135
The hot crack grows perpendicular to the solidification line toward weld
pool centre. The growth direction corresponds to the solidification direc-
tion. This is represented in Fig. 5b. At the same time, a second hot crack
develops on the opposite flank of the travelling weld pool (Fig. 6a). Both
cracks grow toward the weld pool centre, until they unify at a certain point
in time at the trailing edge of the weld pool, as shown in Fig. 6b.
(a) (b)
Weld Pool
Crack initiation
Crack
Solidification Path
(a) (b)
Crack Unification
Opposite
Crack
Fig. 6. Initiation of opposite crack and crack unification on trailing edge of the
weld pool
(a) (b)
Centreline
Crack
Restraint Free
In the following diagrams, strain and strain rate in the near (Fig. 8) and
far field (Fig. 9) of the weld pool are represented. Solidification cracking
did not occur with these welding conditions. The weld pool position is
highlighted. In the curve of the near field, first a rise in strain rate with
approaching weld pool is observed due to the forward-moving heat. The
consequence is a slight strain of approx. 0.3%. With the arrival of the weld
pool, the strain rate decreases and changes its algebraic sign. The pin
movement towards the melt is now accelerated. With narrowing of the
Influence of Local Weld Deformation on Solidification Cracking Susceptibility 137
weld pool, this movement is slowed and goes finally towards zero during
cooling. The cooling-conditioned compression takes a value of approxi-
mately –4%.
Fig. 8. Strain and strain rate in near field of restraint free specimen
A qualitatively similar curve shape is obtained for the far field (Fig. 9).
However, the absolute values here are obviously smaller. In [15], similar
strain processes were measured and simulated when welding an aluminium
alloy.
Fig. 9. Strain and strain rate in far field of restraint free specimen
It remains to be stated that the material in the near field of the weld pool is
exposed to both positive and negative strains. These strains are mainly
thermally conditioned, since the material was welded under free contrac-
tion. In the near field, the amounts of the arising strains and strain rates are
138 A. Kromm, Th. Kannengießer
significantly higher than in the far field. The observed strain behaviour
agrees in principle to the discussed model conceptions of Zacharia [7, 8]
and Chihoski [17, 18].
In Figs. 10 and 11, strain and strain rate curves are represented when weld-
ing in the CTW-Test under external load at a cross head speed of
va = 3 mm/min for near field and far field, respectively. A hot crack was
not observed at this load. The results of measurement show that, in com-
parison to Figs. 8 and 9, an external load directly affects the weld pool
near strains and strain rates. In contrast to welding under free contraction,
near and far field curves show higher strain caused by the external load.
In addition, significantly different strain behaviour is observed as a
function of the weld distance, while the weld pool passes the measuring
point. Two strain maxima are found in the near field. Strains and strain
rates are smaller in the far field. It should be emphasized that the already
discussed typical strain behaviour in the weld pool vicinity, under free
contraction, was registered qualitatively here, although the thermally
caused strains were superimposed by an additional external load. Indeed,
additional strains acting on the solidifying weld metal are present, but the
strain rate does not seem to be high enough to initiate cracking.
Fig. 10. Strain and strain rate in near field applying an external load of va = 3 mm/min
Influence of Local Weld Deformation on Solidification Cracking Susceptibility 139
Fig. 11. Strain and strain rate in far field applying an external load of va = 3 mm/min
In Figs. 12 and 13, strain and strain rate are represented when welding in
the CTW-Test applying cross head speeds of va,crit = 4 mm/min and
va = 5 mm/min, respectively. These loads triggered solidification crack
formation in the form of a centreline crack as mentioned above. Passing of
the crack at the measuring point is marked in the Figures.
Fig. 12. Strain and strain rate in near field applying an external load of
va,crit = 4 mm/min (critical cross head speed to induce a centreline crack)
Note that the already mentioned typical strain behaviour in both cases was
observed here again. However, the absolute values rise at increased load.
The curves of strains and strain rates in the far field show lower values and
resemble qualitatively the far field measurements in Fig. 11.
140 A. Kromm, Th. Kannengießer
Fig. 13. Strain and strain rate in near field applying an external load of
va = 5 mm/min
Assuming that when a centreline crack follows directly the trailing edge of
the weld pool, the time when the centreline crack propagates through the
measuring range could be determined with the help of the high speed video
records as follows (see Fig. 3).
Pcrack − Pstart + lpool (1)
tcrack =
vs
(2)
lpool = (tedge − tarc) × vs
Fig. 14. Strain and strain rate for cracked and crack free welds
Although the results are subject to relatively large scatter, it becomes clear
that solidification crack propagation requires high critical strain rates. It is
obvious from the ranges A and B in Fig. 14 that quantification of the ex-
ternal restraint depending hot crack initiation should be possible based on
the local strain rates in the weld pool vicinity.
Fig. 15. Strains (indicated by arrows) caused by pure thermal load in vicinity of
the weld pool
Just behind the arc, when the temperature at the fusion line begins to
fall, the contraction by cooling starts in the direct vicinity of the weld pool.
The areas outside this region, however, are still heated up, i.e. they still
expand. With progressive cooling and contraction of adjacent areas, the
heated material here is more and more able to let its expansion proceed
inward towards the liquid area. The strain registered changes into a com-
pressive one, although heating (expansion) is still present. Due to progres-
sive cooling, the compressive strain will remain. Figure 15 clarifies this re-
lation. In the literature [17, 18], it is reported that the expansion of areas
still heated up can superimpose the contraction of already solidified areas
in such a manner that even compressive stresses develop there. According
to [4-8], this compressive cell can reach behind the weld pool and may
prevent cracking (see Fig. 1).
When an external load is applied, the deformation behaviour (Fig. 16)
changes. The described thermo-mechanics are superimposed by external
load. In the initial state not yet warmed up material strains, due to the load.
With heating of the material these strains become higher, since the rigidity
of the material decreases. Additionally, thermal expansion begins. Thermal
mechanics and load mechanics interact here in the same direction.
As previously mentioned, it is possible for the heated material to pro-
ceed with its expansion towards the weld pool. This thermo-mechanical in-
fluence interacts with the external load and at first prevails. With progres-
sive external load; its influence continues to increase until it finally
exceeds the thermal mechanics. Now again; a tensile strain is registered.
Thermo-mechanics and external load act against each other.
The presence of this tensile cell enables cracking in agreement with the
discussed model conceptions (see Fig. 1).
Influence of Local Weld Deformation on Solidification Cracking Susceptibility 143
Fig. 16. Strains (indicated by arrows) caused by interacting thermal and external
load in the vicinity of the weld pool
Conclusion
1. According to [1, 2], local strain rates in the vicinity of the weld pool can
significantly affect hot crack formation, and can be regarded as a suit-
able hot crack criterion.
2. By means of high speed video technology, the weld pool dimensions as
well as the location and point in time of crack initiation and growth
could be reconstructed. This permitted a direct correlation with the de-
termined weld pool near strains and strain rates.
3. The observed typical strain behaviour in the weld pool vicinity is in
principle in agreement with the model conceptions developed and dis-
cussed by Zacharia [7, 8] and Chihoski [17, 18]. An external load causes
an additional significant increase in the strains and strain rates. In the
near field, the magnitudes of the arising strains and strain rates are
clearly higher than in the far field.
4. Contrary to samples under free contraction, additional tensile strains
evolved in appropriate areas of the solidifying weld pool. Mechanical
and thermal loads superimpose and interact differently in the near and
far field.
5. The hot crack formation in the weld pool vicinity is directly affected by
the local strain rate, which seems suitable for the quantification of
144 A. Kromm, Th. Kannengießer
restraint specific hot crack susceptibility. It could be shown that for hot
crack propagation high critical strain rates are necessary.
6. By application of a constant external load, solidification cracks can be
initiated with narrowing weld pool on both sides of the solidification
line. In the present case, the cracks grow perpendicular to the solidifica-
tion line towards the weld centreline where they unify to become a cen-
treline crack.
Acknowledgements
The authors would like to thank A. Hannemann und M. Lammers for the
technical support during this work.
References
J.C. Lippold1, J.W. Sowards1, G.M. Murray1, B.T. Alexandrov1, A.J. Ramirez2
1
The Ohio State University, Columbus, Ohio, USA
2
Brazilian Synchrotron Light Laboratory, Campinas-SP, Brazil
Abstract
*
Inconel® is a registered trademark of Special Metals Company, a PCC company.
Hastelloy® is a registered trademark of Haynes International. In this paper, a
shortened version of these alloys will be used that does not include the trademark.
148 J.C. Lippold et al.
Introduction
t
ε= (1)
2R + t
where: İ = strain
t = sample thickness
R = radius of curvature of die block
Threshold Saturated
Strain Strain
Applied Strain
Fig. 2. Transverse Varestraint test results plotted as maximum crack distance
versus strain
MCD (2)
SCTR = ∗ CR
VW
Experimental Procedures
Materials
The Ni-base alloys tested in this investigation were in the form of filler
metals. The compositions of the filler metals, as provided by the
consumable manufacturers, are shown in Table 1.
Sample Preparation
Fig. 3. Sample preparation and testing sequence. Grooves in Alloy 600 plate were
filled in four layers with approximately 15 total weld passes. The weld
reinforcement was then removed and Varestraint testing conducted within the
deposited weld metal
Varestraint Testing
Characterization
Weld metal samples of 617, Hast X, and Haynes 230W were prepared for
metallographic analysis by electrolytic etching with 10% chromic acid at
2.5 V for 30–60 s. The 625 and Hast W samples were etched
electrolytically with 10% oxalic acid at 4 V for 30 s. Scanning electron
microscopy (SEM) and EDS analysis was performed on a Sirion
SEM/FEG and Phillips XL-30 ESEM/FEG at 15 kV.
Calphad Calculations
compositions of the actual weld deposits (Table 2) were used as input for
the calculations. The phases used for the calculations were liquid, δ, γ, γ´,
Laves, Ș, μ, σ, R-phase, MC, M6C, M7C3, and M23C6.
Alloy 625 L
L+γ
Temperature [°C]
γ + M6C
γ + M6C + ı
γ + MC
γ + M6C + ı + ȝ
γ + M6C + į + ȝ
γ + M23C6 + į γ + M23C6 + į + ȝ
γ + M23C6 +
γ + M23C6 + į + ȝ + γ’
į + γ’
Weight Percent Mo
Hastelloy X L + BCC
L+ı
L
L+γ + BCC
L+ı
L+ı
Temperature [°C]
+ M6C
γ
γ + M6C
γ + M6C + ı
γ + M23C6 γ + M6C + ı + γ’
γ + M6C + ı + ȝ
γ + M6C + ı + ȝ + γ’
Weight Percent Mo
Fig. 4. Calculated phase diagrams for Alloy 625 (top) and Hastelloy X (bottom)
systems as a function of Mo content. The vertical line represents the nominal Mo
content of each alloy
156 J.C. Lippold et al.
L
γ γ
Mole Fraction
Alloy 625
M23C6
ȝ M 6C
γ’ į
Temperature [oC]
L
γ
Mole Fraction
Hastelloy X
M23C6
ȝ
γ
γ’ σ M 6C
Temperature [oC]
L → L+ γ → L+ γ + M6C → L+ γ +
Inconel 617 1226 1386 160
M6C + δ → γ + M6C + δ
L → L+ γ → L+ γ + MC → L+ γ +
Inconel 625 1112 1355 243 MC + δ → L + γ + MC + δ +σ → γ +
MC + δ +σ
L → L+ γ → L+ γ + M6C → L+ γ +
Hastelloy W 1060 1385 325
M6C + μ → γ + M6C + μ
L → L+ γ → L+ γ + M6C → L+ γ +
Hastelloy X 1236 1396 160 MC + M6C + σ → γ + MC + M6C +
σ
L → L+ γ → L+ γ + M6C → L+ γ +
Haynes
1277 1402 125 M6C + M23C6 → L+ γ + M23C6 → L+ γ
230W
+ M6C + M23C6 → + γ + M6C + M23C6
158 J.C. Lippold et al.
The alloys examined in this study solidify as austenite (face centered cubic
crystal structure). Austenite solidification promotes strong segregation
resulting in local variations in composition in the weld metal. Due to this
segregation, many of these weld metals form a terminal eutectic
constituent (or constituents), as predicted by the Scheil-Gulliver
simulations. Figs. 6–8 show the solidification structure of the alloys
examined in this study. Included with the micrographs are composition
spectra (measured with SEM/XEDS) showing the composition of both the
austenite matrix and interdendritic eutectic constituent.
Fig. 6. Micrographs showing Alloy 617 (top) and Alloy 625 (bottom)
solidification structure and XEDS spectra of matrix and eutectic compositions
Weld Solidification Cracking of Ni-Base Alloys 159
Fig. 9. Back-scattered electron SEM image (1000X) and XEDS spectra of NbC
and matrix composition of Alloy 625
the travel speed. The measured values of DAS and SDAS (shown in Table
6) were comparable for each of the alloys indicating similar solidification
behavior.
Both the maximum crack distance (MCD) and total number of cracks
(TNC) were determined from the tranverse Varestraint test samples for the
five filler metals. The results from this analysis are shown in Figs. 10–12.
Except for Hastelloy W, cracking was observed at even the lowest applied
strain (0.5%). Based on these results, the 617 and Haynes 230W filler
metals exhibited the lowest cracking susceptibility. The MCD results for
all five filler metals are compiled in Fig. 13.
2.5 TNC
40
10
0.5
5
0 0
0 1 2 3 4 5 6 7 8
Applied strain (%)
2.5 TNC
40
Total number of cracks
35
2
30
1.5 25
20
1
15
10
0.5
5
0 0
0 1 2 3 4 5 6 7 8
Applied strain (%)
Fig. 10. Transverse Varestraint results for Alloys 617 and 625 filler metals
Weld Solidification Cracking of Ni-Base Alloys 163
2.5 TNC
40
1.5 25
20
1
15
10
0.5
5
0 0
0 1 2 3 4 5 6 7 8
Applied strain (%)
2.5 TNC
40
Total number of cracks
35
2
30
1.5 25
20
1
15
10
0.5
5
0 0
0 1 2 3 4 5 6 7 8
Applied strain (%)
Fig. 11. Transverse Varestraint results for Hastelloy W and Hastelloy X filler
metals
164 J.C. Lippold et al.
2.5 TNC
40
1.5 25
20
1
15
10
0.5
5
0 0
0 1 2 3 4 5 6 7 8
Applied strain (%)
Fig. 12. Transverse Varestraint results for Haynes 230W filler metal
3.5
IN617
IN625
3
Hastelly X
Hastelloy W
Maximum crack distance (mm)
1.5
0.5
0
0 1 2 3 4 5 6 7 8
Applied strain (%)
Fig. 13. Summary of maximum crack distance (MCD) versus strain for all the
filler metals tested
Weld Solidification Cracking of Ni-Base Alloys 165
250
Hast X
625
Hast W
200 210
190
Haynes 230
SCTR, °C
150
145
Type 304L, FN6
617
100
95
85
50
31
0
Fig. 14. Solidification cracking temperature range for the filler metals tested. Type
304L with Ferrite Number (FN) 6 is listed for comparison
and Scheil-Gulliver values. For Alloy 617 and Hastelloy W, the SCTR and
SS-DTA values are nearly equivalent, while the SCTR for Hastelloy X is
greater than both the simulated and measured solidification temperature
range. Since these materials are also susceptible to ductility dip cracking
(DDC), it is possible that some portion of the crack length measured in the
Varestraint test is actually DDC extension off the solidification crack. This
would result in an artificially high value of the SCTR. Additional analysis
is ongoing to rationalize these differences.
The SCTR determined using the Varestraint test is based on the MCD at
saturated strain. In the case of the weld metals tested here, the saturated
strain at which SCTR was determined was 5%. At such high applied
strains, the effect of crack backfilling by excess liquid at the end of
solidification can be reduced or eliminated. Under normal welding
conditions, where actual strains in the solidifying weld metal may be much
lower, backfilling may in fact be effective in reducing cracking
susceptibility.
Photomicrographs taken near the crack tips of Varestraint samples from
Alloy 625 and Haynes 230W are shown in Fig. 15. Note that in Alloy 625
considerable residual liquid (arrows in Fig. 15A) is observed near the
crack tips. This represents liquid that was along the solidification grain
boundaries at the end of solidification and is drawn into the tips of the
cracks by capillary attraction during testing at high strain levels. In
contrast, virtually no residual liquid is observed near the crack tips of the
Haynes 230W weld metal (Fig. 15B).
Weld Solidification Cracking of Ni-Base Alloys 167
This residual liquid effect is also observed when examining the fracture
surface of the solidification cracks in these weld metals, as shown in Fig. 16.
In both cases, the characteristic dendritic nature of solidification cracking
is evident. However, in the case of Haynes 230W, the dendritic fracture
morphology is much more distinct. This occurs since there is little residual
liquid at the end of solidification and the solidifying dendrites are covered
by only a thin layer of liquid. In contrast, the large amount of terminal
liquid during solidification of Alloy 625 acts to coat the dendritic structure
and make it less distinct.
A)
B)
Fig. 15. Microstructure near the crack tip of a solidification crack in (A) Alloy
625, and (B) Haynes 230W. Note difference in magnification
168 J.C. Lippold et al.
(A)
(B)
Fig. 16. Solidification crack surface of (A) Alloy 625, and (B) Haynes 230W.
Note difference in magnification
Weld Solidification Cracking of Ni-Base Alloys 169
Under lower applied strain conditions, the terminal liquid in the Alloy
625 weld metal would be available to heal any solidification cracks that
would form, improving the weld solidification cracking resistance relative
to what the SCTR ranking predicts. This may explain why Alloy 625 is
generally considered to have good resistance to weld solidification
cracking in actual practice, particularly at low to moderate restraint levels
where backfilling can be effective in healing cracks that form..
For Alloy 625, differences between the solidification cracking
susceptibility determined using the Varestraint test and actual fabrication
experience has been reported by other investigators [6, 7]. This suggests
that for weld metals that achieve resistance to solidification cracking
through a crack backfilling mechanism, the Varestraint test may not
accurately reflect their susceptibility to cracking. The high applied strains
used during this test force cracking to occur even in cases where
considerable terminal liquid is present. The use of a threshold strain for
cracking rather than MCD at saturated strain may be a more appropriate
measure of susceptibility for these weld metals.
Conclusions
Acknowledgements
References
1. Savage, W.F. and C.D. Lundin, The Varestraint Test, Welding Journal,
October 1965, pp. 433s–442s.
2. J.C. Lippold, 2005. Recent Developments in Weldability Testing, 1st
International Workshop, Hot Cracking Phenomena in Welds, Berlin, March
2004, publ. by Springer-Verlag, pp. 271–290.
3. Alexandrov B.T. and J.C. Lippold, Single Sensor Differential Thermal
Analysis of Phase Transformations and Structural Changes during Welding
and Postweld Heat Treatment, IIW Doc. IX-2199-06, 59th Annual Assembly
of IIW, Quebec City, Canada, August 2006, accepted for publishing in
Welding in the World.
4. Alexandrov B.T. and J.C. Lippold, Relationship Between the Solidification
Temperature Range and Weld Solidification Cracking Susceptibility of
Stainless Steels and Ni-base Alloys, IIW Doc. IX-2163-05, 58th Annual
Assembly of IIW, Prague, July 2005, www.iiw-iss.org.
5. Alexandrov B.T. and J.C. Lippold, In-Situ Weld Metal Continuous Cooling
Transformation Diagrams, Welding in the World, Vol. 50, No. 9/10, 2006, pp.
65–74 (Doc. IIW 1744-06, ex-doc. IX-2162-05).
6. Lingenfelter, A.C., Varestraint Testing of Nickel Alloys, Welding Journal,
September 1972, pp. 430s–436s.
7. Karlsson, L., E.L. Berquist, S. Rigdal, and N. Thalberg, Evaluating Hot
Cracking Susceptibility of Ni-base SAW Consumables for Welding of 9% Ni
Steel. Proceeding of 2nd International Workshop on Hot Cracking, Berlin,
March 2007.
Hot Cracking Susceptibility of Ni-Base Alloy
Dissimilar Metal Welds
Abstract
Introduction
The Ni-base alloy dissimilar metal welds are typically made using Alloy
182 and Alloy 82. Recently, Alloy 52 has been used both in new construc-
tions as well as in repair welding. The trend towards alloys with higher
contents of chromium is driven by the observed environment-assisted
cracking (EAC) in Alloy 182, and recently also in Alloy 82. One driving
force towards the more EAC resistant alloys is also the challenges and
172 H. Hänninen et al.
52 was more susceptible to DDC than Alloy 82, which also exhibited heat-
to-heat variation in susceptibility. Thus, because of varying observations in
various tests there is a major need to rank the Ni-base alloy weld metals
for their hot cracking susceptibility and to understand the mechanisms of
solidification cracking properly.
In this study the hot cracking tendency of six different dissimilar metal
weld joints was tested by Varestraint testing [6]. The Varestraint equip-
ment, Fig. 1, was built and the obtained hot cracking and weld microstruc-
tures were studied by optical and scanning electron microscopy (FEG-
SEM/EDS). The studied materials and welds are typical for nuclear power
plant and oil and gas industry applications. In Table 1 the chemical com-
positions of the base materials and filler metals used in the manufacturing
of the studied dissimilar metal welds are presented.
Linear drive
Programming unit
Video recorder
TIG torch
Power source
Automatic voltage
control (AVC)
Display
Pneumatic sylinder (18 kN)
Fig. 1. General view of Varestraint testing equipment built for hot cracking testing
174
Table 1. Chemical compositions of the base materials and filler metals used in the dissimilar metal welds (wt.-%)
C Si Mn P S Cr Mo Ni Nb Ti Fe Al Cu
H. Hänninen et al.
Base materials
42CrMo4 0,42 0,7 0,4 1,05 0,225
AISI 321 < 0,08 17-19 9-12 5*C-0,7
AISI 316 NG < 0,03 16,5-18,5 2,5-3 11-14
SA508 (Grade 2) 0,24 0,21 1,38 0,009 0,16 0,54 0,58 0,015 0,12
Inconel 600 0,07 0,19 0,21 0,007 0,001 16,3 72,86 Nb+Ta 0,10 0,28 9,44 0,229 0,13
Inconel 690 0,02 0,04 0,16 0,001 29,46 59,82 0,33 9,96 0,2 <0,01
Filler metals
E 347 (OK 61.81) 0,07 0,7 1,7 0,021 0,006 19,8 0,1 10 0,61 0,14
AISI 347 (OK Tigrod 16.11) 0,04 0,85 1,4 0,02 0,01 19 0,02 9,5 0,55 0,03
E 347 (OK Band 11.62) < 0,025 0,4 1,8 20 10 0,7
AISI 308L (OK Band 11.61) < 0,025 0,4 1,8 20 10
AISI 309L (OK Band 11.65) < 0,025 0,4 1,8 24 13
Inconel 182 (OK 92.26) 0,03 0,8 6,5 0,01 0,003 15,7 68 1,8 0,1 6,7 < 0,01
Inconel 82 (OK Tigrod 19.85) 0,039 0,03 2,98 0,001 0,004 19,94 72,6 Nb+Ta 2,47 0,34 1 0,01
Inconel 52 0,03 0,13 0,24 <0,001 <0,001 29,2 0,03 59,28 Nb+Ta <0,02 0,51 9,8 0,72 0,04
Inconel 52M 0,02 0,09 0,8 0,003 0,001 30,06 0,01 59,54 Nb+Ta 0,83 0,224 8,22 0,11 0,02
Inconel 52MS 0,014 0,12 0,68 0,004 0,0007 29,53 0,02 60,14 0,78 0,19 8,33 0,13 0,03
Inconel 152 0,048 0,41 3,48 0,003 0,003 28,74 0,01 55,2 Nb+Ta 1,54 0,09 10,39 0,06 <0,01
Inconel 152M 0,0293 0,3276 3,247 <0,001 0,0042 28,98 <0,005 57 Nb+Ta 1,533 0,039 8,75 <0,005 <0,005
Hot Cracking Susceptibility of Ni-Base Alloy Dissimilar Metal Welds 175
80 mm Inconel 182
42 CrMo4
Inconel 42CrMo4
Inconel
82 Inconel
E 347 182
82
E 347
E 347
E 347
Inconel 82
AISI 321
E 347 E 347
Fig. 2. Groove geometry, materials and cross-section of Neste Oil mock-up dis-
similar metal weld. Length of the specimen is 300 mm
50 °
8 mm
2,5
80 mm
50 °
8 mm
2,5
80 mm
AISI 309
AISI 308
16 °
8 mm
1,5
80 mm
AISI 309
AISI 308
4 4 45 °
8 mm
1,5
2,5
80 mm
Inconel
Inconel 690
SA 508 152
Inconel
52 Inconel
152
Figure 7 presents the test piece for Varestraint testing of pure weld met-
als and weld pass sequence of the test specimen welded with Alloy 82.
Base material was AISI 304 plate into which a 20 mm wide, 60 mm long
and 4 mm deep groove was machined and then welded with different filler
metals. The surface of the weld was ground to the level of the base metal
so that the surface quality was equal in all Varestraint tests. Filler metals
(supplier in brackets) used in the pure weld metal tests were: Alloy 182
(Esab OK 92.26), Alloy 82 (Esab Tigrod 19.85), and Alloys 152, 152M,
52, 52M, and 52MS (Special Metals). Table 2 presents the welding pa-
rameters used in the Varestraint testing.
20
Welding
filling with
with filler filler
material metal
8 AISI 304
4
60
300
60
Fig. 7. Test piece for Varestraint testing of pure weld metals and weld pass se-
quence of the test specimen welded with filler metal Alloy 82
178 H. Hänninen et al.
Experimental Results
Figure 8 (a) presents the surface (Alloy 182 weld, augmented strain 4%) of
Neste Oil mock-up with hot cracks after Varestraint test. Optical microscopy
of a hot crack (dark areas) in the cross-section of the weld (b) and scanning
6 mm
a) b)
c) d)
Fig. 8. (a) Surface (Alloy 182 weld, augmented strain 4%) of Neste Oil mock-up
with hot cracks after Varestraint test. (b) Optical micrograph of the hot cracks
(dark areas) in a cross-section of the weld (note the white phase ahead of the hot
crack tip). (c), (d) Scanning electron micrographs (SEM) of weld hot crack tips
showing details of the irregular shaped white phase
Hot Cracking Susceptibility of Ni-Base Alloy Dissimilar Metal Welds 179
electron microscopy (FEG-SEM) of weld hot crack tips (c) and (d) show
details of the irregular shaped white phase connected to hot cracks of this
alloy. The composition of the white phase areas was examined by EDS
analysis. Figure 9 presents X-ray element maps of the white phase area in a
dendrite boundary region and Fig. 10 shows a SEM image and EDS
analyses of the white phase showing marked Si, Nb, Mn and P enrichment,
which is compensated by depletion of Cr, Fe and Ni. Based on the high
magnification BSE-images and the EDS analyses the observed white phase
consists of two phases, white and grey, with varying Nb, Si, Mn and Ti
contents. Inside and along the dendrite/grain boundaries white Nb-rich
particles, Nb(C, N), are present.
Fig. 9. X-ray element maps of the white phase area in a dendrite boundary region.
Note marked Nb enrichment in the white phase and less enrichment in a wider
area along the dendrite boundary. Marked Si, Mn and P enrichment compensated
by depletion of Cr, Fe and Ni can also be seen
180 H. Hänninen et al.
Fig. 10. SEM image and EDS analyses of the white and grey phase showing Nb,
Si, Mn and P enrichment and Nb rich particles, Nb(C, N). The spot size in EDS
analysis has an interaction diameter about 1 μm, and therefore the surrounding
matrix affects also the result
The hot crack fracture surfaces were opened for fractography and phase
analysis with an FEG-SEM/EDS system. Figure 11 shows an opened fracture
surface of Alloy 182. With a low acceleration voltage of 7 kV the informa-
tion comes from close to the surface. The wavy morphology of the fracture
a) b)
Fig. 11. Fracture surface of Alloy 182 hot crack showing the white phase in the
valleys between dendrites
Hot Cracking Susceptibility of Ni-Base Alloy Dissimilar Metal Welds 181
surface can be clearly seen, which is an evidence of the molten phase along
the dendrite boundaries in which the separation has occurred.
Study of the fracture surfaces of the Neste Oil mock-up weld (Alloy
182) hot cracks revealed marked segregation of the same elements on frac-
ture surfaces as seen in the white phase in the cross-sections of the hot
cracks. On the fracture surface Nb-rich particles, Nb(C, N), and the lamel-
lar white phase are present intermixed in the final microstructure. The
white phase exhibits similar enrichment of Nb, Si, Mn, and P as observed
in the cross-sections of the weld metal.
Pure Alloy 82 weld metal hot cracking took place only at higher heat
input (Q = 9,4 kJ/cm, augmented strain 4%). The formed hot cracks were
markedly oxidized, probably because of the high heat input and also be-
cause of element segregation to the final melt. Figure 12 presents a thick ox-
ide layer on the fracture surface of Alloy 82 weld metal hot crack. Table 3
presents EDS analyses of three areas of the oxide layer on the hot crack
surface presented in Fig. 12. The oxide layer is enriched in Mn, Nb, Fe,
and Ti.
Fig. 12. Oxide layer on the hot crack fracture surface of Alloy 82 weld metal
Table 3. EDS analyses of three areas of the oxide layer on the hot crack surface
presented in Fig. 12 (wt.-%)
Element O Al Si Ti Cr Mn Fe Ni Nb
Area 1 23,4 0,02 0,18 1,2 8,6 13,4 11,7 30,1 11,6
Area 2 1,4 0,03 0,17 0,3 8,9 2,5 5,9 77,2 3,6
Area 3 1 0,04 0,15 0,3 10 2,2 5,2 79,2 1,9
Figure 13 (a) presents SEM image and (b) X-ray element maps of a
cross-section of hot cracks exhibiting a strong Nb segregation and less
marked Si, P and Ti segregation to the dendrite boundary at the hot crack
182 H. Hänninen et al.
tip. Figure 14 shows SEM image and X-ray element map of fracture surface
of Alloy 82 exhibiting distribution of Nb and Ti.
a)
O Si P
Ti Cr Mn
Fe Ni Nb
b)
Fig. 13. (a) SEM image and (b) X-ray element maps of a cross-section of hot
crack exhibiting strong Nb segregation and additional segregation of Si, P and Ti
to the dendrite boundary along and ahead of the hot crack
Hot Cracking Susceptibility of Ni-Base Alloy Dissimilar Metal Welds 183
a) b)
Fig. 14. (a) SEM image and (b) X-ray element map of fracture surface of Alloy 82
showing distribution of Nb and Ti
The chemical composition of the oxide layer on the hot crack fracture
surface of Alloy 82 shows marked increase of Nb, Mn and Fe contents,
which are the elements of a high tendency for oxidation and are available
on the fracture surface of a hot crack. Under the oxide layer the white
phase rich in Nb, Mn, Si and P as well as the Nb- and Ti-rich (Ti content
of Alloy 82 0,34 wt.-%) particles, Nb, Ti(C, N), are also clearly visible. In
the pure weld metal tests it was also observed that cracking took place
under the weld bead made in the Varestraint test by a liquation cracking
mechanism following the boundaries of lower weld passes (see Fig. 7).
Figure 15 presents SEM images of a cross-section of hot cracks in mock-
up 1 weld metal (Alloy 152 weld, augmented strain 4%) after Varestraint
test and Fig. 16 shows SEM images of fracture surface, where the presence
of the white phase in the valleys between the dendrites is clearly visible.
SEM image and X-ray element map of the hot crack fracture surface of
Alloy 152 exhibits high enrichment of Nb to the white phase (see Fig. 17).
a) b)
Fig. 15. SEM images of a cross-section of hot cracks in Alloy 152 weld metal
exhibit the connection of the white phase to hot cracks
184 H. Hänninen et al.
a) b)
Fig. 16. SEM images of hot crack fracture surface of Alloy 152 weld metal. Note
the brittle nature of the lamellar white phase
a) b)
Fig. 17. (a) SEM image and (b) X-ray element map of the fracture surface of Al-
loy 152 exhibit high enrichment of Nb to the white phase
Alloy 152 weld metal shows similar hot cracking tendency as compared
to Alloy 182. The Nb-rich white phase is connected to the hot cracks and
the phase can be easily found in the cross-sections of the hot cracks as well
as on the fracture surfaces (local coverage can be as high as 30%, see Fig.
16). The dendrite boundaries containing the white phase show in addition
to Nb a marked increase of Si, Mn and P contents both in the white phase,
but also in the wider zone along the dendrite boundaries. Simultaneously
depletion of Cr, Ni and Fe is observed in these zones. In addition to the
white phase both in the cross-section as well as on the fracture surfaces
plenty of small particles, Nb, Ti(C, N), are present.
Hot Cracking Susceptibility of Ni-Base Alloy Dissimilar Metal Welds 185
Filler metal Alloy 52 contains Ti (0,51 wt.-%), but does not contain Nb
as Alloys 182, 82 and 152 and, thus, hot cracks form along the dendrite
boundaries without the presence of the Nb-rich white phase. Figure 18 pre-
sents SEM images of mock-up 2 weld metal (Alloy 52, augmented strain
4%) hot cracks after Varestraint test. A lot of small particles (TiN) are pre-
sent along the dendrite boundaries. Figure 19 presents a SEM image and
X-ray element maps of the hot crack fracture surface.
Fig. 18. SEM images of mock-up 2 (Alloy 52, augmented strain 4%) weld hot
crack cross-sections after Varestraint test
Hot cracking of Alloy 52 does not show any indications of the Nb-
rich white phase observed in the earlier cases with Alloys 182, 82 and
152, since Alloy 52 does not contain Nb. Alloy 52 contains Ti and
therefore a lot of precipitation of TiN(C) was observed both in the
cross-sections along the dendrite boundaries as well as on the fracture
surfaces of the hot cracks, where the coverage of the hot crack fracture
surface by TiN(C) can be quite extensive. The dendritic large TiN(C)
phase particles are presumed to have formed in the melt prior to final
solidification of the weld. On the fracture surface also MnS phase
particles are observed as separate phases. Thus, in Alloy 52 the only
marked phase on the hot crack surfaces is TiN(C).
186 H. Hänninen et al.
a)
b)
Fig. 19. (a) SEM image and (b) X-ray element maps of fracture surface of Alloy
52 revealing the dendritic phase of TiN(C), and separate MnS particles are also
present
Total crack lengths (TCL) of Neste Oil mock-up, mock-up TV and mock-ups
1, 2 and 3 are shown in Fig. 20 (a) and maximum crack lengths (MCL) in
Fig. 20 (b), respectively, with augmented strains of 0%, 0.5%, 1%, 2%, 3%
and 4%. In the legend box markings are as follows: case, tension side and
filler metal. Table 4 presents summary of Varestraint tests (augmented strain
4%) made for pure Ni-base alloy weld metals (see, Table 1 and Fig. 7).
Hot Cracking Susceptibility of Ni-Base Alloy Dissimilar Metal Welds 187
50
Neste Oil, surface, 182
45 Mock-up TV, root, 182
35 Mock-up 2, root, 52
25
20
15
10
0
0 1 2 3 4 5 6 7 8
Augmented strain [%]
a)
8
Neste Oil, surface,182
7,5
7 Mock-up TV, root,182
6,5 Mock-up 1, root, 152
Maximum crack length [mm]
6
Mock-up 2, root, 52
5,5
Mock-up 3, root, 152
5
4,5
4
3,5
3
2,5
2
1,5
1
0,5
0
0 1 2 3 4 5 6 7
Augmented strain [%]
b)
Fig. 20. Total solidification crack lengths (TCL) (a) and maximum crack lengths
(MCL) (b) as a function of augmented strain of Neste Oil mock-up, mock-up TV
and mock-up 1, 2 and 3 samples in Varestraint testing
188 H. Hänninen et al.
Table 4. Summary of Varestraint tests made for pure Ni-base alloy weld metals
(augmented strain 4%, heat input 4,7 kJ/cm, except for Alloy 82)
Inconel 182 (Q=4,7 kJ/cm) TCL = 19,3 mm / 2nd test -> cracks
Inconel 82 (Q=4,7 kJ/cm) no cracks / 2nd test -> no cracks
Inconel 82 (Q=9,4 kJ/cm) TCL = 32,9 mm
Inconel 152 (Q=4,7 kJ/cm) TCL = 19,7 mm / 2nd test -> cracks
Inconel 152M (Q=4,7 kJ/cm) TCL = 8,1 mm
Inconel 52 (Q=4,7 kJ/cm) no cracks / 2nd test -> no cracks
Inconel 52M (Q=4,7 kJ/cm) TCL = 3,1 mm / 2nd test -> no cracks
Inconel 52MS (Q=4,7 kJ/cm) no cracks / 2nd test -> no cracks
Discussion
Nb segregation and the amount of Laves phase [11]. Thus, for controlling
the Laves phase in the Nb-bearing Ni-base weld metals lowest possible
heat input with high cooling rates have to be used. Additionally, control of
dilution by Fe, Si, and C is very important, since these elements lower the
liquidus and solidus temperatures, increase the solidification temperature
range and favour eutectic-type solidifications.
Conclusions
The following conclusions can be drawn based on the results of this study:
− Dissimilar metal weld joints of Ni-base weld metals show in their mi-
crostructure marked segregation of Nb, Si, P and Mn to the last liquid
to solidify wetting the dendrite boundaries. In the Nb-bearing alloys
eutectic Laves and Nb(C, N) phases form along the dendrite bounda-
ries and in Ti-alloyed Alloy 52 TiN(C) phase forms, respectively.
− Hot cracking occurred along the dendrite boundaries and the suscepti-
bility to hot cracking in the weld mock-up samples was observed to
follow the order: Alloy 152 > Alloy 52 > Alloy 182 > Alloy 82.
− In pure weld metal hot cracking tests the hot cracking susceptibility
was observed to follow the order: Alloy 182 Alloy 152 > Alloy 52
Alloy 82.
− The differences are probably related to the dilution effects because Fe,
Si, and C enhance the eutectic phases and expand the solidification
temperature range.
References
Abstract
Introduction
The cast pin tear test was originally developed by Hull [1] and used in the
1970’s as a method for evaluating the susceptibility of alloys to hot
cracking during welding and casting. It consisted of levitation melting and
casting of small charges of material in copper molds to produce conical
cast pins with varying geometry. The charge mass in this test was constant
(19 g). The test sensitivity was controlled by the pin length and geometry,
and by the mold diameter. The pins were examined for circumferential
cracks and the total crack length was plotted versus the mold number.
Based on the percent of circumferential cracking and mold number, the
CPTT generated arbitrary ranking of the susceptibility to hot cracking.
Compared to the other hot-cracking tests, it required only several hundred
grams of material to develop a cracking susceptibility curve and was
suitable for alloy development purposes.
The CPTT of Hull was modified by the Welding and Joining
Metallurgy Group at The Ohio State University [2]. The costly and
complex levitation melting technique was replaced by an arc melting
technique. A standard gas tungsten arc welding (GTAW) torch and power
supply are used to melt a small charge of material on a water-cooled
copper hearth under an inert (argon) atmosphere. The hearth was equipped
with a simple delivery system that transferred the molten charge into a
copper mold positioned below the hearth. Although a successful system
was initially developed using this approach, problems with complete mold
filling and reproducibility of the cracking results were encountered.
Because of the problems with mold filling, a second-generation cast
pin tear test that is more robust and capable of ranking the weldability of
Ni-base superalloys has been recently developed at the Ohio State
University [3]. The reproducibility and reliability of test results was
improved by further development of the CPTT apparatus and by
optimizing the melting and mold filling procedures. The test sensitivity
was improved by introducing smaller increments in cast pin length. The
capacity of testing various alloys was expanded by using mold materials
with different thermal conductivity and by extending the range of mold
lengths.
Weld Solidification Cracking in Ni-Base Superalloys Using CPTT 195
Parameter Value
Mold Material Cu; Cu-Be
Shielding Gas Ar
Gas Flow Rate, ml/s 70–80 (9–10 cf h–1 )
Arc Current, A 250
Arc Length, mm 15 (0.6 in)
Arc Time, s 5–7
Over-Pressure, MPa 0.007–0.014 (1–2 psi)
Cast Pin Diameter, mm 9.525 (0.375 in)
Cast Pin Length, mm 12.7–50.8 (0.5–2 in)
Length Increments, mm 3.175 (0.125 in.)
The hearth design is presented in Fig. 2. The hearth opening limits the
heat extraction from the charge during its melting, thus allowing for
sufficient superheating of the latter and facilitating its quick transfer into
the mold. The mold cavity is designed to produce a cylindrical cast pin
with restraining head and foot, Fig. 3. The CPTT test utilizes molds and
cast pins with constant diameters and varying lengths. The charge mass
corresponds to the cast pin length, Table 2.
Head
Foot
Table 2. Range of mold (cast pin) lengths and correspondent pin/charge volume
and mass (for pure Ni samples)
The rate of heat extraction from the molten metal is controlled by the
thermal conductivity of mold material. In this study, two mold materials
were used: high purity copper (C10100) and a Cu-Be-Co alloy (C17000)
with thermal conductivities of 391 W/m-K and 118 W/m-K respectively.
The lower rate of heat extraction of the Cu-Be-Co molds made it possible
to test some Ni-base alloys that had been previously problematic with
respect to filling of copper C10100 molds.
Longitudinal tensile stress is generated in the cast pin as it solidifies
and cools down to room temperature. The stress level increases with the
pin length. At the critical stress level (pin length) for particular alloy, hot
cracking occurs at the pin surface usually just below the pin head, as
indicated on Fig. 4. With further increase in the cast pin length, the total
surface crack length increases to 100% circumferential cracking and
eventually results in complete pin separation, Fig. 5.
Fig. 4. Cast pin samples of alloy 625. Fig. 5. Complete separation in 1.25 in.
Arrows show the typical location of long pin of alloy 800H
solidification cracks
The cast pins are examined for cracks using a binocular microscope at
magnifications from 10x to 70x. During this examination the cast pins are
rotated around their longitudinal axes by a specialized device. The
projected crack length in a plane that is perpendicular to the pin axis is
measured in degrees. The percentage of cracking is calculated as follows:
LT
% Cracking = ×100 , %,
360
where LT is the total length in degrees of all cracks measured on the pin
surface.
Weld Solidification Cracking in Ni-Base Superalloys Using CPTT 199
Maximum circumferential
70% cracking response curve
60%
50%
0% to 100%
40% cracking range
10%
1
2 2 3 3 1
0%
0.5 0.625 0.75 0.875 1 1.125 1.25 1.375 1.5 1.625
Pin Length (in)
These alloys had previously been the subject of an IIW round robin
study of externally restrained weld hot cracking tests [4, 5]. The study was
conducted by ten participating organizations and included the Varestraint,
Trans-Varestraint, PVR, MVT, MIS-1, and LTP-1-6 tests [4]. Each
participating laboratory applied its own testing procedures.
That study produced a total of twelve separate rankings of the hot
cracking susceptibility of the tested alloys. Most of the rankings, generated
by the different hot cracking tests and by different laboratories using same
tests, were conflicting. This poor correlation was related to the different
aspects of hot cracking evaluated by the different tests, and to the different
testing procedures used with the same hot cracking tests [4, 5]. A summary
Weld Solidification Cracking in Ni-Base Superalloys Using CPTT 201
ranking of the tested alloys that is based on the results from that study is
presented in Table 4.
The CPTT results for these alloys, in terms of maximum pin length for
no cracking and response curve for circumferential cracking, are
summarized in Fig. 7. The CPTT ranked alloy 800H as the most
susceptible, followed by alloys 926, AC66 and 825. These results are in
fairly good correlation with the general results of the IIW round robin test.
A very good correlation was found between the CPTT and PVR test
rankings. This is not surprising since both the CPTT and the PVR test
restrain the test samples in the direction of solidification.
Table 4. Hot cracking susceptibility ranking of the IIW round robin study alloys
100%
% Circumferential cracking
90%
80%
70%
Alloy 800H Alloy 926 Alloy 825
60%
50%
Alloy AC66
40% .
30%
AC66 Max
20% 800H Max
825 Max
10% 926 Max
0%
0.5 0.625 0.75 0.875 1 1.125 1.25 1.375 1.5
Pin length, in.
a) b)
Fig. 8. Alloy 926 fracture surface tested with Varestraint (a) and CPTT (b)
Ni 76 53 58 57 60 59 57
Ta – – – – – 3.80 6.35
Nb – 5.10 – – – 0.10 –
Hf – – – – – 1.55 1.50
Cu 0.15 0.15 – – – – –
100
90
80
70 R'80
R'125
60
R'142 R'125 R'77 718 R'80 Waspalloy 600 R'142
50 R'77
40 600
718
30
Waspalloy
20
10
0
0.50 0.75 1.00 1.25 1.50 1.75 2.00
10 15 20 25 30 35 40 Pin
45length, in50 mm
Pin Length, in (mm)
Solidification Microstructure
c) Central area with finer equiaxed d) Isolated central area with large
dendrites in cast pin of alloy R’142 equiaxed dendrites in cast pin of alloy
R’77
Table 7. Primary dendritic arm spacing in the tested Ni-base alloys in the
conditions of CPTT (average of 30 measurements)
Shrinkage Porosity
Some shrinkage porosity was found along the cast pin center line in the
alloys forming large equiaxed dendrites in the last region of the pin to
solidify. Alloys René 125 and René 77 (Fig. 11d) had more significant
center-line shrinkage porosity, followed by alloy 718 (Fig. 11e) and René
80 with some moderate porosity. The shrinkage porosity was typically
found in the longer pins and decreased to an insignificant level with
decreasing the pin length. In general, the shrinkage prosity was not related
to solidification cracking but instead increased the percent of invalid
samples at the higher mold lengths.
Alloys 600, Waspaloy, and René 142 exhibited very little dispersed
shrinkage porosity that was not related to solidification cracking. The
dispersed shrinkage porosity typically forms in the cast pin central area, it
is evenly distributed, and the size of the separate pores is smaller then the
dendrite arm spacing. Therefore, this dispersed porosity is not expected to
affect the general stress distribution in the cast pin, and the processes of
crack nucleation and propagation, which occur in the pin surface area. This
kind of porosity was considered acceptable and not affecting the test
results.
From the total of 27 sectioned cast pin samples, only three had large
shrinkage pores that caused cracking. The test results from these samples
were not included in the results. Most of the tested samples had either
insignificant or no shrinkage porosity and, in general, porosity did not
affect the test results.
208 B.T. Alexandrov, J.C. Lippold, N.E. Nissley
Fracture Morphology
20 um 40 um
a) René 142 at 0.625 in. b) René 142 at 0.625 in.
20 um 20 um
c) René 125 at 0.75 in. d) René 77 at 1.25 in.
20 um 20 um
e) Inconel 718 at 1.75 in. f) René 80 at 1.5 in.
20 um 20 um
g) Inconel 600 at 2 in. h) Waspalloy at 2 in.
Fig. 13. Solidification cracking fracture morphology in cast pins of the tested
Ni-base superalloys
Weld Solidification Cracking in Ni-Base Superalloys Using CPTT 211
Discussion
The second-generation cast pin tear test has been developed in an effort to
create a robust weldability test that is able to rank the solidification
cracking susceptibility of both “difficult-to-weld” alloys and “standard”
alloys. The cast pin apparatus and testing procedure were designed to
provide controllable and reproducible testing conditions, and to produce
cast pins of good integrity. The cases of unacceptable shrinkage porosity
in difficult-to-cast alloys were reduced to less than 10% of the tested
samples. As a result, the reproducibility and reliability of the CPTT results
were significantly improved.
The solidification conditions in the CPTT closely simulate the
solidification conditions of low to medium heat input welds, such as those
made by the GTAW process. These include the morphology of the
solidification microstructure and the sequence of solidification, starting
with a fine-grained “chill zone” at the mold wall, followed by cellular and
columnar dendritic growth, and finishing with equiaxed dendritic growth
at the center of the pin. The primary dendritic arm spacing was between 13
and 30 μm. Even higher weld heat input conditions can be simulated by
using lower cooling capacity (lower conductivity) molds.
Metallographic and fractographic examination of cast pins have
confirmed that the CPTT also simulates the weld solidification cracking in
actual welds relative to the nature of crack nucleation and propagation,
crack morphology, and crack healing. For example, cracking was primarily
along solidification grain boundaries, just as it is in actual welds in Ni-base
alloys and stainless steels.
The reliability of the CPTT results has been confirmed by comparison
to six externally restrained hot cracking tests. Very good correlation was
found between the weldability rankings of four high alloy stainless steels
generated by CPTT and by PVR test.
The solidification cracking susceptibility ranking of the seven tested
Ni-base superalloys is in good correlation with the field weldability
experience with these alloys. This indicates that the improved, second
generation CPTT technique is a reliable tool for evaluating the weldability
of materials with wide variations in solidification cracking susceptibility,
particularly those with high to moderate susceptibility.
The solidification cracking behavior of the Ni-base superalloys is a
function of their complicated alloying systems and of the solidification
conditions provided by the CPTT. The combination of chemical
composition and cooling rate determine the solidification morphology,
212 B.T. Alexandrov, J.C. Lippold, N.E. Nissley
tendency for formation of eutectic phases, the properties of the latter and
their capability to filling interdendritic shrinkage, and the general liquid-
solid and solid-state shrinkage behavior.
A series of experimental and modeling investigations will be
conducted in the future, aimed at quantifying the parameters that control
the solidification cracking phenomenon under the conditions of the CPTT.
The mold and cast pin design facilitates the application of relatively simple
finite element modeling to study the processes of heat transfer,
solidification, and development of stresses during solidification. The
technique of single sensor differential thermal analysis – SS-DTA [6] will
be utilized in conjunction with the CPTT to determine the solidification
temperature ranges and the possible liquid-state precipitation and eutectic
reactions. Such combined experimental- modeling approaches will allow
the solidification morphology and solidification cracking behavior to be
related to quantifiable parameters as cooling rate, solidification
temperature range, liquid- and solid-state volume shrinkage, and stresses
associated with solidification.
Conclusions
References
1. Hull F. C., Cast-Pin Tear Test for Susceptibility to Hot Cracking, Welding
Journal, Vol. 38 (4), 1959, pp. 176s–181s.
2. Ryan D. P., Development of Modified Cast-Pin Tear Test to Evaluate the
Solidification Cracking Susceptibility of Fully Austenitic Materials, MS
Thesis, The Ohio State University, 2003.
3. Alexandrov B., Nissley N., Norton S., and Lippold J., Development of a
Weldability Test for High Performance Base and Filler Materials, Report No.
MR0606, Edison Welding Institute, July 2006.
4. Wilken K., Investigation to compare Hot Cracking Tests – Externally Loaded
Specimen, IIW Doc. IX-1945-99 (II-C-168-99).
5. Finton T., Lippold J., and Bowers R., Comparison of Weld Hot Cracking
Tests, Summary of an IIW Round Robin Study, IIW Doc. IX-H-459-99 (II-C-
175-99).
6. Alexandrov B. T. and Lippold J. C., Relationship Between the Solidification
Temperature Range and Weld Solidification Cracking Susceptibility of
Stainless Steels and Ni-base Alloys, IIW Doc. IX-2163-05.
SAW Cold Wire Technology – Economic
Alternative for Joining Hot Crack Sensitive
Nickel-Base Alloys
Abstract
Introduction
Testing Equipment
Conventional SA equipment has been used for the weld examinations. The
equipment had been modified with an additional cold wire feeding unit,
which works without applying any additional electrical energy to the cold
wire. In order to adjust the feed speeds for electrode and cold wire
independently of each other, two separate feeding units have been used.
This allowed varying the feeding quantity of the cold wire without
changing the cold wire diameter which resulted in a higher flexibility of
the process.
SAW Cold Wire Technology 217
Stickout
Workpiece
surface
The following tests show the influence which cold wire addition exerts on
the welded joint. In particular, the weld geometry, the structure, the
chemical composition and the connection between cold wire addition and
hot-cracking susceptibility have been considered.
Weld Geometry
The first tests dealt with the determination of the influence which cold
wire exerts on the weld geometry. For that, welds with different quantities
of cold wire were carried out. The parameter “melting ratio” has been
introduced for these tests. It specifies the ratio between the mass of the
cold wire which has been added during welding and the mass of the
electrode. With the increasing quantity of cold wire, weld reinforcement
and weld width were also increasing, while the penetration depth was
reduced (Fig. 3). The smaller penetration area resulted in a lower degree of
dilution with the base metal (Fig. 4).
Caused by cold wire addition, the weld shows a tendency to reinforce-
ment of the weld beads. In order to compensate the reinforcement, the
SAW Cold Wire Technology 219
14
12
10
Weld height
8 Weld width
[mm]
6
4
2
0 Penetration depth
0% 20% 40% 60% 80% 100%
deposition ratio [%]
Fig. 3. Influence of cold wire addition on welds reinforcement, weld width and
penetration depth
60%
50%
dilution [%]
40%
30%
20%
10%
0%
0% 20% 40% 60% 80% 100%
deposition rate [%]
Fig. 4. Influence of cold wire addition on the dilution with the base metal
220 U. Reisgen et al.
The cooling speed during crystallisation has great influence on the for-
mation of hot cracks. Cold wire addition results in faster cooling of the
molten metal; the crystallisation processes in the molten pool are greatly
influenced.
These tests were performed in order to determine a potential influence
on the temperature gradient in the molten pool in a solid state, provided
that the deposition ratio had been varied.
For that, thermocouples have been dipped into the molten pool, right
behind the cold wire through the slag which was by the time still liquid.
The thermocouples allowed the reproducible measurement of the
temperature gradient only within a temperature range of below 1000°C. It
has been observed that the solidification interval for alloy 617 is between
1380°C and 1330°C. Because of the homogenisation of the temperature
fields in the component it is not possible to establish an influence of the
cold wire addition on the temperature profile below 1000°C. For further
examinations it would be advisable to select a measuring method which
allows the measurement of temperature ranges above the melting
temperature of the filler materials.
The macro-sections (Fig. 5) do show differences with regard to the weld
geometry. It is observed that – with the increase of the cold wire addition -
the penetration area is getting smaller and the reinforcement is increasing
(Fig. 6). This causes the shortening of the fusion line, the heat dissipation
Fig. 5. Micro-sections of surface welds with increasing welding voltage and a cold
wire addition of 50%. Used wire electrode: UTP UP 6170 Co Mod., weld flux:
UTP UP FX 6170 Co, base material: austenitic CrNi-steel, welding speed:
50 cm/min
SAW Cold Wire Technology 221
into the base metal is reduced. Particularly in SA welding, the heat dissipa-
tion into the base material is of special importance and the alteration of the
fusion line may have large effects on the thermal balance of the molten
pool and thus also on the solidification conditions.
Figure 6 shows that, up to the increase of the cold wire addition to 50%,
also an increase of the weld cross-section takes place. This means that an
increased quantity of metal is molten without any variation of the heat
input. The heat in the molten pool is more effectively used through the
cold wire addition than this is the case with standard SA welding methods.
A conclusion is that the average temperature of the molten pool which has
been produced with the SA cold wire method is lower than that of the
molten pool produced with SA standard methods. This also brings about a
changed curve of the molten pool isotherms and thus a lower temperature
gradient during a SA cold wire process which again affects the
crystallisation processes and the welded structure.
If the deposition rate is increased further, the excessive heat of the
molten pool is used for melting the increasing quantities of cold wire ad-
dition. The heat of the molten pool is no longer sufficient for the melting
of equal quantities of base material. The penetration area is reduced and
the weld reinforcement area is, at the same time, increased.
75
60 Reinforcement area
Weld area
area {mm²]
45
Penetration area
30
15
0
0 25 50 75 100
deposition ratio [%]
Fig. 6. Influence of the deposition ratio on the weld area, the penetration area and
the reinforcement area
222 U. Reisgen et al.
The fusion line is shortened. The heat transport from the liquid molten
pool into the solid base material is carried out via the fusion line. That im-
plies that: the longer the fusion line, the higher the excessive heat of the
molten pool (deposition ratio – 0%), and, the shorter the fusion line, the
colder the molten pool (deposition ratio – 100%). The increase of the cold
wire quantities is accompanied by the displacement of the weld cross-
sectional area in the direction of the workpiece surface.
If the deposition ratio is too high (above 100%), strong fluctuations of
the weld geometry occur. This is characterised by changes of the weld
width and the weld height and also by non-parallel weld flanks. The reason
for this is that the cooling of the molten pool has been too strong. These
fluctuations show that the welding process is no longer running stable and
that it will be collapsing if the quantity of the cold wire addition is further
increased.
Structure
The examinations of the structure have been carried out by means of light
and scanning electron microscopy. The macrographs of the surfacing
beads showed mainly columnar crystals in the weld metal
As from a 200x magnification and higher, a finer distribution of the
structure in the individual crystals in the vicinity of the fusion line is
observed. With increasing distance to the fusion line, a coarsening of the
crystal structure is observed, (Fig. 7). The coarsening is a result from the
alteration of the cooling conditions during the solidification. It results from
the decrease of the temperature gradient with increasing distance to the
fusion line.
Fig. 7. Fusion line between two weld beads in multiple layer welding, using the
SA Cold wire method, SEM micrographs
SAW Cold Wire Technology 223
Fig. 8. Grain boundaries in the bottom and the top weld bead in multiple-layer
welding using the SA cold wire method. SEM micrographs
224 U. Reisgen et al.
Fig. 9. Inheritance of the grain boundary from the base material to the weld metal,
SEM-micrographs
Fig. 10. Widening of the grain boundaries (formation of micro-cracks) in the weld
metal, SEM micrographs
SAW Cold Wire Technology 225
SA cold wire
Um= 35,7 V, Im= 355 A
SA standard
Um= 35,7 V, Im= 353 A
Fig. 11. Macro-sections with detailed photographs of a weld, Cold wire deposition
ratio: 50% (top) , without cold wire addition (bottom), welding speed: 50 cm/min
226 U. Reisgen et al.
(a) (b)
Fig. 12. Centre region of the weld bead in the SA cold wire method (a), contacting
columnar crystals in the centre region of the weld bead in SA standard weld-
ing (b), SEM micrographs
SAW Cold Wire Technology 227
(a)
(b)
(c)
Fig. 13. SA standard method, lower region, transverse section (a), SA cold wire
method, vs = 50 cm/min, lower region, transverse section (b), SA cold wire
method, vs = 70 cm/min, lower region, transverse section (c)
228 U. Reisgen et al.
flank
flank
middle
middle
Fig. 14. SA standard method (left) and SA cold wire method (right)
Chemical Composition
225
200
175
HAZ WM Centre standard
150
0 2 4 6 8 10
Distance [mm]
275
250
Flank cold wire
Hardness [HV1]
225
200
Fig. 15. Comparison of the hardness traverse of a weld made with the SA standard
method and of a weld made with the SA cold wire method and a deposition ratio
of 50% in weld centre and flank (welding speed 50 cm/min)
For the welding tests, a modified wire of the type UTP UP 6170 Co Mod.
has been used. The chemical composition of this wire is specified in Table 1.
For testing the influence of the cold wire addition on the Al content in the
pure weld metal, welds on the austenitic CrNi steel were carried out. The
specimens for the multiple-layer welding tests were built up in 5 layers.
The chemical composition of the material used in these welding tests was
each determined from the top layer and is shown in Table 2.
Table 1. Chemical composition of the welding wire and the base material
C Cr Mo Fe Al Co Ti Ni
[%] [%] [%] [%] [%] [%] [%] [%]
Wire analysis
0,05 21,77 8,38 0,24 1,23 10,44 0,31 57,00
acc. to ISF
Base material
analysis acc. to 0,05 18,47 0,27 <69,6 0,01 0,14 0,01 8,51
ISF
Four welding tests with a cold wire addition of 100% of the electrode
mass have been carried out and the welding parameters voltage and current
have been varied. The table shows that the Al burn-off remains constant
with approximately 1%. The fifth test was carried out without additional
cold wire feeding. The chemical analysis shows a clear decrease of the Al
content down to 0.77%.
In comparison to the conventional SA welding, the tests with cold wire
addition show a clear increase of the Al content in the weld metal. The
reason is that through the trailing arrangement of cold wire addition the
melting of the cold wire takes place at distinctly lower temperatures and is
thus significantly reducing the aluminium burn-off.
SAW Cold Wire Technology 231
Joint Welding
The wire diameters 1.2 mm, 1.6 mm and 2 mm were examined. For all
welds, electrodes and cold wire with equal diameters have been applied.
During the tests it has been established that with the increase of the
current, the melting potential of the molten pool was also increased.
The application of the electrode diameter of 1,2 mm resulted in a slight
increase of heat input only. This was the reason why only a restricted,
additional quantity of cold wire was fed to the process. Also, very precise
positioning of the cold wire was necessary. Therefore, the process stability
was not always ensured. With the application of the electrode diameters
1.6 mm and 2.0 mm, a very high process stability was achieved.
Figure 16 shows an example of a joint on alloy 617, plate thickness: 16
mm, and an electrode diameter of 2 mm. For the weld preparation, a V-
type weld with a preparation angle of 60° had been chosen. Both plates
were firmly clamped during welding. The root pass was welded manually
using the TIG method. The 2 filler passes and 2 final weld passes were
welded with the SA method, with a cold wire addition of 47% (Table 3).
The macro-section in the picture shows good overlap and reliable fusion to
the weld flanks. With those final weld passes, a good weld interface with
the base material was obtained. A distinct increase of the deposition rate
was achieved.
It has been observed that, also during welding in the groove, most
crystals of each individual bead are continuing uninterruptedly from the
fusion line to the weld surface. This facilitates the transport of the low-
melting phases in the slag which again results in a positive influence on the
hot cracking resistance [5].
The bead reinforcement and the increased deposition rate which are
typical of the SA cold wire method have also been observed during
welding tests made in the groove.
The slightly reinforced beads in the groove lead to the reduction of the
tensile stress peaks which may result in longitudinal solidification cracks
in the centre region of the weld bead [5].
Figure 17a shows a horizontal projection of the joint weld surface where
a dye penetration test had been carried out. Surface cracks have not been
observed. The results from the radiography examinations which are
depicted in Fig. 17b do not show any slag inclusions, cracks or other
discontinuities in the joint weld.
232 U. Reisgen et al.
16mm
root (TIG)
Fig. 16. Transverse section of a joint weld (Alloy 617) with a TIG welded root
pass
(a) (b)
Fig. 17. Weld surface (a) and radiograph (b) of the weld
The next step was the examination of the cold wire influence on the
angular distortion. For that purpose, joint welds with and without
additional cold wire feeding have been carried out where one plate had
been fixed and the other plate was allowed to shrink without restrictions
(Fig. 18).
Through the cold wire addition which was caused by the higher
deposition rate, the number of necessary weld beads was reduced.
Moreover, the angular distortion was clearly reduced if welding was
performed with cold wire addition
For a comparison of the supporting effect of the individual passes in
welding, the welding speed in welding with cold wire addition had been
increased in order to enable welding with the same weld build-up as in
conventional SA welding. In this case also a clear decrease of the angular
SAW Cold Wire Technology 233
distortion was the result (Fig. 18). The exact influence of heat input and
weld build-up is still subject to more detailed examinations.
110mm
Clamping force
16mm
30 16
25 26,9 mm
15,3° 12
20 22 mm
Distortion [mm]
11,9°
Distortion [°]
19,4 mm
15 10,4° 8
5 layers 7 layers 7 layers
10
SA cold wire SA Standard SA cold wire 4
A=50% A=0% A=50%
5 Um= 35,5 V Um= 35,7 V Um= 35,8 V
Im= 371 A Im= 359 A Im= 358 A
vS=50cm/min vS=50cm/min vS=70cm/min
0 0
Fig. 18. Comparison of the distortion values using the SA standard method and
the SA cold wire method
234 U. Reisgen et al.
For a closer examination of the hot cracking tendency, MVT tests were
made on SA welds. This is a sensitive and material-oriented testing
method where the specimen is loaded externally. In the MVT test, the
specimen is bent around a die with a defined radius and a defined speed.
During the bending; the SA weld bead is re-molten by a TIG arc.
The three regions which are depicted in Fig. 19 (weldable, restrictedly
weldable and not weldable) have been determined empirically through a
large number of tests at the BAM, Berlin and have been stored in an expert
data base. These three regions serve orientation purposes only since
material-specific discrepancies may occur.
The diagram shows the results from the longitudinally welded SA
specimens with the dimensions 100 mm × 40 mm × 10 mm. For the SA
welding tests, constant welding current and constant welding voltage have
been applied. The energy-per-unit length had been realised through the
variation of the welding speed, its length is 50 cm/min and 70 cm/min,
respectively.
30,0
Sektorgrade 1
27,5
Sektorgrade 2
25,0 KD 50-50 L2
22,5 St L
Total crack length [mm]
20,0 KD 50-70 L
KD 50-50 L1 Weldable
17,5
with restrictions
15,0
not
12,5 weldable
10,0
7,5
5,0
2,5 weldable
0,0
0 1 2 3 4 5
surface bending strain [%]
Fig. 19. Total crack length after application of the MVT test
SAW Cold Wire Technology 235
The MVT tests have been performed at the BAM, Berlin. The
Varestraint mode has been applied. All specimens have been tested with a
travelling speed of 2.0 mm/s, 4% - strain and energy per unit length of
7.5 kJ/cm. The entire length of the hot cracks which have been produced
during the test will be evaluated.
It is, however, important to mention that all results from the SA cold
wire welded specimens are all within a close range and are forming a
group which is close to the restrictedly weldable area. The diagram shows
that the results do not show a clear difference with regard to the changed
energy-per-unit length caused by the increase of the welding speed from
50 to 70 cm/min.
As already mentioned before, the addition of the cold wire is the
dominating factor with regard to the deposition of the micro-alloying
elements into the weld. This had been confirmed through the chemical
analysis of the SA weld metal by means of MVT tests where a similar
arrangement with regard to the Al content has been determined (Table 4).
The low Al content of the weld made with the conventional SA method is
correlating with a somewhat lower hot cracking susceptibility.
Industrial Implementation
Finally, a pipe weld had been carried out at the company Essener
Hochdruck Rohrleitungsbau, using the SA cold wire method.
For this, two tube sections made of the material Alloy 617 with a wall
thickness of 78 mm have been welded. A common circumferential weld
preparation with a weld angle of approximately 24° has been chosen. The
root pass was welded using the TIG method. The groove was then filled by
manual arc welding up to a width of approximately 20 mm. The remaining
passes were welded with the SA method and additional cold wire feeding.
The applied weld current was 290–310 A and the voltage was
approximately 32–33 V. The welding speed was 45 cm/min. The cold wire
was added with a ratio of 50% of the electrode feed speed.
Conclusion
The tests which have been carried out at the ISF demonstrated that the
addition of cold wire resulted in a distinct reduction of distortion,
compared with the conventional SA method.
The practicability of the process modification was demonstrated through
the industrial implementation of the test results. It has been possible to
realise the increase of the deposition rate by approx. 50%, referring to the
deposition rate of the conventional SA welding methods.
References
Abstract
Introduction
Experimental Procedure
The base material used in this work is AA6056, the chemical composition of
which is given in Table 1. The material in the shape of 1.6 mm thick sheets
was butt welded in the T4 state (namely after a solution treatment, water
quench and several months of natural ageing). Welding was carried out
along the rolling direction, on 400 mm × 150 mm samples, after degreasing.
Butt joints were produced with a Nd:YAG laser operating at 3 kW as shown
in Fig. 1 [8]. The focal point of the laser beam was set at the surface of the
sheets.
Sheets
Welding direction to weld
Results
This first part is dedicated to the study of the number of hot tears on a
given length of weld (40 cm) as a function of the process parameters.
Figure 2 shows the influence of the welding speed on the occurrence of hot
tearing for different filler wire speeds. The number of hot tears increases
when the welding speed increases whatever the filler wire speed. These
results are in agreement with those obtained by Cicala [11] on the same
alloy but for different welding and filler wire speeds. This increase of the
sensitivity to hot tearing can be due either to a higher instability of the
weld pool (leading to the closure of the keyhole) or/and to the higher
solidification rate when the welding speed increases.
The influence of the solidification rate on hot tearing has been widely
studied over the past. It has been shown in particular that the presence of
a second beam to decrease the cooling rate decreases the number of hot
tears [7]. Early in the 70’s, some models, predicting the sensitivity to hot
tearing, considered that decreasing the solidification rate lead to less
defects [12, 13]. To study the influence of the cooling rate different
inserts have been used between the sheet to weld and the welding table
(Fig. 3). These experiments confirm that the lower the cooling rate, the
lower the number of hot tears. This could partly explain the higher
number of hot tears observed in Fig. 2 when the welding speed was
increased to 6 m.min–1.
Hot Tearing During Laser Butt Welding of 6xxx Aluminium Alloys 245
Figure 2 also shows that the quantity of filler wire has an influence on
the occurrence of hot tearing: for higher filler wire speeds, the number of
hot tears drops whatever the welding speed. The filler wire speed changes
the chemical composition of the weld pool and thus of the weld nugget. In
our case, the silicon concentration will increase as the quantity of filler
wire increases. The silicon concentration has been calculated using optical
observation and dilution consideration (for more details see [14]). Figure 4
shows the number of hot tears as a function of the silicon composition. The
number of hot tears is high for low silicon content and then drops when the
composition in silicon increases. This is in accordance with the literature
[15]: for most of the alloying elements (Mg, Si, Cu), the hot tearing
sensitivity during casting decreases with increasing alloying element
concentration. This study shows that in the case of laser welding, the Si
composition acts in the same way. Further experiments are required to
verify and explain the minimum of hot tears observed for 2% of Si.
60
Nbr. of hot tears for a 40cm weld
50 fv=0
fv=1m/min
fv=2m/min
40
fv=2.5m/min
fv=3m/min
30
20
10
0
3,5 4 4,5 5 5,5 6 6,5
welding speed (m/min)
Fig. 2. Influence of the welding speed on the number of hot tears in a 40 cm long
weld for various filler wire speeds (fv=filler velocity)
246 D. Fabrègue et al.
100
Nbr. of hot tears for a 40cm weld
80
60
40
20
0
- 0 + ++
Thermal contact between sheet and support
Fig. 3. Influence of the thermal contact between the sheets to weld and the
welding table on the number of hot tears. The welding speed is 6 m/min and the
filler wire speed is 1.5 m/min. The sign – corresponds to cork insert whereas the
signs + and ++ correspond to an insert of a single sheet of brass and the sandwich
configuration respectively
60
Nbr. of hot tears for a 40cm weld
50
40
30
20
10
0
0,5 1 1,5 2 2,5 3 3,5
Silicon content of the weld nugget
Fig. 4. Influence of the silicon content on the number of hot tears in a 40 cm long
weld for a welding speed of 6 m/min
60
Nbr. of hot tears for a 40cm weld
mechanical
50 vacuum
40
30
20
10
0
0 0,5 1 1,5 2 2,5 3 3,5
Filler wire speed (m/min)
Fig. 5. Influence of the fastening device on the number of hot tears for various
filler wire speeds. Welding speed is equal to 6 m/min
Figure 6 shows the EBSD cartography of the fusion zone perpendicular to the
welding direction. A hot tear is observed in black. On this figure the grain
248 D. Fabrègue et al.
Fig. 6. EBSD map of the fusion zone perpendicular to the welding direction.
Black zones correspond to the hot tear
Hot Tearing During Laser Butt Welding of 6xxx Aluminium Alloys 249
Hot tearing takes place at the very end of the solidification process [17],
usually for solid fractions higher than 0.9. During laser welding,
solidification occurs first at the fusion line because the main phenomenon
responsible for cooling is conduction in the solid [19] (heat extraction is
mainly assures by this means). Figure 8 shows the micrograph in the
rolling plane of a laser weld containing hot tears. From this figure, a hot
tear can be seen. The crack exhibits a quite planar shape when it goes
between the columnar grains and a round shape around the equiaxed grains
of the centre of the fusion zone. The hot tear results from the strains
applied on the mushy zone. These stresses come from firstly the
solidification shrinkage which is quite high in aluminium alloys (except
for the 4xxx series) and secondly from thermal contraction of the solid
skeleton. Moreover external loads (for example coming from the
fastening) can apply on the mush zone.
250 D. Fabrègue et al.
Equiaxed
grain
Columnar
grain
Base material
Welding Fusion line
direction
more accurately the centre zone of the hot tears. The deviation around the
equiaxed grains of the centre, as guessed from the 2D micrograph (Fig. 8),
is clearly visible here. The planar surface of the borders becomes more
round as the crack goes around the grains.
Outer
surface
Hot tear
Fusion
zone
Base
material
Welding
direction
Fig. 9. X-ray picture showing a hot tear. The rings are artefact
252 D. Fabrègue et al.
Welding
direction
Fig. 10. 3D X-ray tomography of the fusion zone containing two hot tears. The
grey surface represents the outer surface of the fusion line
Hot Tearing During Laser Butt Welding of 6xxx Aluminium Alloys 253
Welding
direction
Fig. 11. 3D X-ray tomography of the fusion zone seen from upward containing
two hot tears. The grey surface represents the outer surface of the weld bead
Conclusions
Acknowledgements
The authors want to thank L. Salvo and J.-J. Blandin for their help during
the ESRF experiments. The authors are grateful to the French Ministry of
Industry for funding in the framework of the ASA (Allègement des
Structures dans l’Aéronautique) RNMP project, carried out in
collaboration with EADS and ALCAN, which are also thanked for
providing the materials. L. Kirschner of EADS-CCR and J.C. Ehrström of
ALCAN CRV are acknowledged for fruitful discussions.
References
Abstract
Introduction
Ten years ago laser beam welding was establish in the commercial aircraft
industry (Fig. 1). At the present time skin and stiffener in form of stringers
are welded in the lower shells of some sections at Airbus types A318,
A340 and A380 (Fig. 2). Against the riveted structures the welded ones
have a completely different fracture mechanical behavior. At riveted joints
the stringer is able to take over loads when at the same time the skin col-
lapses. By contrast skin and stringer fail simultaneously at a welded struc-
ture. The different behavior bases on the differential and integral design.
New aircraft aluminum alloys are designed after criteria like weight, re-
sidual strengths, damage tolerance or corrosion but not under considera-
tions concerning weldability. New materials like AA2098 or 7xxx-alloys
contain elements like Li or Zn. Both are characterized by a high vapor
pressure. This circumstance provokes instabilities of capillary during the
welding. Experiments have shown that these base alloys have a lower so-
lidification cracking resistance compared to the applied alloys in fuselage
like AA6013 or AA6056 [1].
Solidification cracks result from transverse shrinkage. They extend
some millimeters and could be visible on the weld seam surface (Fig. 3) or
limit on the inner weld seam. Hilbinger argues that the reason for the ap-
pearance of hot cracks lies in the magnitude of its dimension [2]. That is
that microscopic cracks based on process instabilities and macroscopic
ones have a global cause like process design and position of weld seam.
Solidification cracks belong to the second group.
The Integral Approach enables the combination of the structural behav-
ior under welding design criterions with an optimization of weldability.
The objective of this approach is to improve the performance of the
welded structure and the transition to the upper shell. There are higher ten-
sile stresses in different to the lower one. A structural compressive stress is
somehow preferred for reduced susceptibility to hot cracking because
compressive stresses work against the thermal expansion due to the local
heat input and heat conduction in the structure.
The Integral Approach 259
Fig. 1. Joining of stringer and skin by both sides simultaneous laser beam
welding
Fig. 2. skin and stringer are welded in the lower shell of the section no. 16 by
A318 (gray-coloured)
Fig. 4. Modification of the stringer with the Integral Approach by designed two
slots, each at the stringer end. It is only shown one end to simplify matters
The secondary parts, that are located at outer stringer regions lead to a
decrease of the discontinuous between skin and stringer (eccentricity) and
therefore to a higher stiffness at the ends of the outer primary parts. In this
way the bending stresses that result by the loads during the flight phase are
reduced locally. The notch stresses that depend on the bending stresses de-
crease as well.
From the technical welding point of view the tendency that the weld
seam cracks is reduced by the so called clamping effect. After welding of
the inner primary and the secondary parts both areas which surround the
weld seams shrink. The still not-welded outer primary part is pulled down.
In these areas residual compressive stresses are accumulated. With the
welding of the outer primary part the compressive stresses are induced as
well, but both compressive stresses have an opposite direction to each
other. Therefore the material expansion is impeded by the heat input, re-
sulting in a lower volume shrinkage due to the cooling and the resulted
tensile stresses on the mushy zone are lower as well.
This work is focussed on the structural behaviour during the welding
process, less the material behaviour. The next sections prove the influence
of slot on the structural behavior as well as on the hot cracking behavior.
The Integral Approach 261
The fuselage consists of several sections which each other are joined with
stringer couplings. Each coupling is composed of one short stringer and a
circumferential butt strap. At all Airbus types couplings are located below
a frame (Fig. 5). The cabin pressure and the self-weight of the fuselage
lead to tensile stresses in the structure.
Fig. 5. The stringer coupling for the connection of several fuselage sections
Stringers with and without slots are compared in order to evaluate the in-
fluence of slots on the stress concentration. In order to reduce the stress
concentration at the stringer ends the current skin-stringer-joining is re-
vised by a socket below the stringer (Fig. 7 left).
The socket leads to a reduction of discontinuity between skin and
stringer. The reference parameters are dSTR = 15, lFH=1.4, lFW=6, lGD=0,
lGW=0. These ones also apply for the structure with slot. Figure 7 right
shows the structural model for a stringer with slot. The slot relevant
reference parameters in mm are dStr =15, lSLOT =15, hSLOT =15, lS =30.
The characteristic lines in Fig. 8 show that the maximum stress concen-
trations in the primary weld seam at stringers with slots are 20% lower
than at stringers without slots. But the stresses in the outer regions of the
secondary weld seams are increased about 11%. This fact is based on the
about ls longer lever of the stringer coupling.
The Integral Approach 263
Fig. 7. Structural models for the calculation of stress concentration. (left) for the
revised current skin-stringer-design without slot. (right) for a stringer with slot
2,0
P
1,49 without SLOT
1,5
S
1,34
P
ıx / ıVM
1,07
1,0
0,5
0,0
0 1 2 3 4
a [mm]
Fig. 8. The comparison of the stress concentrations for stringers with and without
slot at the start-end-positions in primary part (P) and secondary part (S) (Fig. 4).
The geometrical parameter for the stress concentration at the stringer-edge a is
shown in Fig. 6
The edge distance (dSTR) has not any significant influence on the maxi-
mum stresses in the bearing primary weld seam. But stresses in the secon-
dary weld seam (S) grow up about 65% at an edge distance dSTR (see
Fig. 7) by 15 mm compared to 5 mm. The longer lever is the reason for the
increasing of stresses again.
The secondary part length (lS) and the slot length (lSLOT) have not any in-
fluence on the stress concentration at the primary (p) and secondary (S)
weld seam. A modifying of the slot height (hSLOT) has a significant influ-
ence on the stress concentration in P and S. In case of reduction from 15 to
10 mm the stress notch factor in P decreases roughly 11% but the stress
264 H. Gruss et al.
sections shrink and pull down the still not-welded outer primary part (3) to
skin. A higher hot cracking resistance in P is achieved (Fig. 9).
The residual tensile stresses (1 and 2) obtain its maximum within the
weld seams and decrease with increasing stringer height. Thereby it is im-
portant that both residual tensile stress areas are joined by a so called
“Bridge”. This bridge ensures that after the first two welding steps the
outer primary zone is pulled down to the stringer whereby there are oc-
cured residual compressive stresses in this one (Fig. 10).
At the third welding step the base metal is heated due to the thermal
stress and it tries to expand. The residual compressive stresses around P
and the surrounding cold areas constrain the volume expansion. Depending
on the residual compressive stresses in the outer primary part it results
lower tensile stresses within the BTR. At this principle it is spoken about a
clamping effect, because the third welding section is enclosed by two ten-
sile stress areas via a bridge.
50
0
ı [MPa]
ıD1,P
-50
1. cooling step
-100 ıD2,P
-150
2. cooling step
-200
0 30 60 90 120
Time [s]
Fig. 10. The resulted stresses (y-direction) at the start-end-position of the primary
part (P) after the first two welding steps. The characteristic compressive stresses
lines refer to the reference model
Determinations
There are used two Nd:YAG-Lasers with a focus diameter by 0.3 mm. The
heat source’s calibration for the simulation bases on a laser power by 2.3
kW, on a welding velocity by 4 m/s. The focus point is fL=0. Shielding gas
is He with a flow rate by 30 l/min (Fig. 11).The used base metal is
AA6056 and the filler metal is AlSi7 (Table 1), 82% dilution rate. The
stringer thickness is 2 mm and the thickness of skin is 3 mm. Before the
welding process the stringer is tacked on the skin.
Alloy Al Si Fe Cu Mn Mg Cr Zn
Fig. 11. Comparison of the cross section and the simulated geometry of welding
seam
Influencing Parameters
Fig. 12. Influencing parameters controlling the hot cracking behaviour at the
position P (start-end-position in the primary part)
The slot width (lSLOT) and height (hSLOT) are fixed on the minimal allow-
able dimension by 7 mm. The secondary part length is defined by 25 mm.
The parameter e is determined by 10 mm (Table 2). The displacements at
start-end-position of the primary part (P) are the reference value (R) for the
following investigations (Fig. 13).
8
2 3
[ yBTR(x) / yBTR(x P, Ref) -1 ] x 100%
1
6
2 R
0 P
-2
-4
-6
0 10 20 30 40 50 60
x [mm]
The slot length is enlarged from 7 to 20 mm. The modification has no in-
fluence on the displacements of the inner primary and the secondary weld
seams. At the beginning of the third weld step (P) the maximum strains
grow up about 50% in comparison with the reference model.
The enlarged slot height from 7 to 20 mm has an effect on the strain be-
havior at the secondary and the outer primary weld seams. The rising of
strains about 90% in the outer primary weld seam results by the reduction
of the web width above the slot (Fig. 14). That leads to a decrease of the
mechanical section modulus und therefore to a reduction of residual com-
pressive stresses in the outer primary stringer part, which counter the heat
expansion (Fig. 15).
8
[ yBTR(x) / yBTR(xP, Ref) -1 ] x 100%
2 3 1
6
4
1
2
0,9
0 P
-2
-4
TIS fest
-6
0 10 20 30 40 50 60
x [mm]
Fig. 14. Effect of slot height on displacements at P. The enlarged slot height from
7 to 20 mm causes higher displacements about 90%
The Integral Approach 271
50
0
ı [MPa]
-50
-200
0 30 60 90 120
Time [s]
Fig. 15. Effect of slot height on residual stresses at the start-end-position of the
primary part (P). The enlarged slot height leads to a lower stringer stiffness and
therewith to a lower level of residual compressive stresses in P.
50
0
ı [MPa]
-50
-100
ls = 25mm (Ref) -145
-150
ls = 40mm -154
-200
0 30 60 90 120
Time [s]
Fig. 16. Effect of the parameter lS on the residual stresses at the start-end-position
of the primary part (P). The residual compressive stresses in P increase insuffi-
ciently, in spite of a extended secondary part length
If the outer primary part is not tacked (d=10 mm), then this part is not able
to develop residual compressive stresses in P. The resulting low stress
level in the outer primary part (P) leads to higher displacements about
430%.
The shortening of the distance between the inner slot edge and the
start/end of the inner primary part from 10 to 2 mm leads to a partial melt-
ing of the inner primary weld seam during the welding of the outer primary
part. The tensile stresses which are assigned to the inner primary weld
seam are released. Due to this the resulted tensile stress in the inner pri-
mary part and the compressive stress in the outer primary part at the P-
position are decreased.
Welding Direction: A1 / A2
Welding Direction: B1 / B2
Welding Direction: C1 / C2
The change of the welding direction at the inner primary part results to
higher strains about 200%. Furthermore the end crater cracks are expected
on the start-end-positions at the primary and secondary parts (P and S).
274 H. Gruss et al.
The designed slots in the advanced stringer geometry reduce the stress
concentrations and increase the hot crack resistance at the outer primary
part (P) assumedly. In order to the conditions in the start-end-positions P
and S during the flight phase and the welding, there are following conclu-
sions for the structural design and the welding procedure:
1. The width and height of slots (lSlot, hSlot) should be chosen as small as
possible in order to reduce the stress concentrations and to increase the
superposition between both residual tensile stresses areas, which are
generated by several welding steps of the inner primary and the secon-
dary parts.
2. The main factor, limiting the residual comressive stresses at the start-
end-position of the primary part (P) is not the length of the secondary
part (ls), but rather the heat input.
3. A short edge distance (dSTR) – the distance between the edges of the
socket and the start-end-position of the secondary part (S) – reduces the
stress concentration at S.
4. In order to avoid end crater cracks the ends of weld seams should be not
located at any stringer edges but rather in regions within the stringer
with lower stress levels during the flight phase.
A complex welding order results by the existence of two slots (Fig. 17).
The welding direction has to be changed at the several welding steps. The
stringer must be tacked on the skin before welding procedure can start.
The composition of 4 welding steps and their welding directions result
by the welding analysis for the whole stringer geometry:
− Step 1: The first welding step starts in the middle of the stringer (inner pri-
mary part) and ends at the distance e in front of the slot at the stringer end.
− Step 2: The welding direction of the second step is diametrical to the
first one and starts also in the middle of the inner primary stringer part.
This welding steps stops at the distance e in front of the slot at the
stringer start.
− Step 3: The welding direction changes again. Two weld seams are pro-
duced by this step. The first weld seam spans along the secondary part
and starting on the start –end-position SStart. The second one completes the
outer primary part and starts on the slot at the stringer start PStart.
− Step 4: The welding of the secondary slot needs the same welding step
sequence than step 3, but starting at Send and Pend in opposite direction.
The Integral Approach 275
Fig. 17. Recommendation for the order of the several welding steps and their
direction
References
Abstract
Earlier work has established that a critical amount of 4043 filler is required
to avoid solidification cracking when arc-welding 6060 aluminium, de-
pending upon local strain conditions. For example, when the mushy zone
behind the weld pool experiences a tensile strain from combined thermal
and shrinkage stresses, the possibility exists for crack initiation. For a
greater rate of strain, it has been determined that a greater 4043 dilution
(i.e. higher weld metal silicon content) is required to avoid crack initiation.
Making use of the Controlled Tensile Weldability (CTW) test and local
strain extensometer measurements, a boundary has been established be-
tween crack and no-crack conditions for different local strain rates and
filler dilutions, holding all other welding parameters constant. Using this
established boundary as a line of reference, additional parameters have
now been examined and their influence on cracking has been character-
ized. These parameter influences have included studies of weld travel
speed, weld pool contaminants (Fe, O, and H), and grain refiner additions
(TiAl3 + Boron). Each parameter has been independently varied and its ef-
fect on cracking susceptibility quantified in terms of a critical strain rate
required to initiate cracking for a given 4043 filler dilution.
Introduction
(e.g. grain size, iron content), welding parameters (torch travel speed), and
external contamination (hydrogen and oxygen contents).
Background
Each of the factors examined in this study could conceivably have an ef-
fect on weld solidification cracking susceptibility. Details regarding these
possible interactions are discussed below.
Role of Silicon
Table 1. Measured chemical analysis for aluminium alloy 6060 and 4043 mix-
tures (wet chemical analysis): (a) 6060-T4, (b) 6060-T4 + 10%4043, (c) 6060-T4
+ 20%4043, and (d) 4043
Composition (wt.%)
Si Mg Fe Mn Cu Cr Ni Zn Ti Ga Zr
(a) 0.42 0.59 0.19 0.02 0.01 0.004 0.004 0.009 0.02 0.01 0.001
(b) 0.90 0.53 0.19 0.02 0.04 0.00 0.00 0.02 0.04 0.01 0.00
(c) 1.39 0.47 0.19 0.02 0.07 0.00 0.00 0.03 0.06 0.01 0.00
(d) 5.25 0.001 0.21 0.02 0.3* --- --- 0.1* 0.2* --- ---
*handbook values
--- no value specified
Fig. 1. Quasi-binary line superimposed on ring casting data of Jennings et al. [20]
showing solidification cracking susceptibility for Al-Mg-Si ternary alloy system
Role of Iron
Iron, even when present at low impurity levels, plays an important role in
determining solidification microstructure as noted above. Although nor-
mally found at around 0.20 wt.% (0.30 wt.% max.) in most wrought alu-
minium alloys, iron has a 0.80 wt.% max. limit in the 4043 filler alloy
[23]. It is normally considered an undesirable element due to its resulting
in reduced mechanical properties and corrosion resistance, where both
problems are tied largely to the formation of coarse, iron-bearing intermet-
allic phases that are cathodic relative to the aluminium matrix [24].
During solidification of castings, high temperature iron bearing inter-
metallic phases are believed to block interdendritic channels, leading to
poor feeding of shrinkage and porosity formation [25-27]. Singer et al.
measured the cracking susceptibility of high purity Al-Fe-Si alloys by
means of restrained welds [28]. Plotting the mean crack length as a func-
tion of iron and silicon content (Fig. 2), they observed no cracking for the
condition Fe/Si>1. Cracking susceptibility increased when increasing sili-
con content or decreasing iron content.
Lu and Dahle [25] studied the influence of iron content on the solidifica-
tion path and cracking susceptibility of cast Al-7Si-0.4Mg alloys. Cast
samples were quenched at around 590°C to observe the microstructure. In-
creasing the iron content from 0.3 to 0.7 wt.% favours the binary β-phase
(forming at a temperature over 600°C) over the ternary β-phase (forming
at temperatures lower than 580°C). Moreover, an increase in iron content
resulted in an increase in porosity, but decreased cracking susceptibility. It
is believed that, because these coarse binary β needles form at high tem-
perature, they block the liquid feeding and promote porosity [25] and serve
282 N. Coniglio, C.E. Cross
Fig. 2. Mean crack length (in inches) on restrained welds of Al-Fe-Si ternary al-
loys as a function of iron and silicon content [28]
Grain Refinement
Grain size directly affects the strain distribution over the grain boundaries
in the weld mushy zone. Refining the grain structure increases the number
of grain boundaries, thereby reducing the strain seen by each boundary.
Reduced cracking susceptibility of aluminium alloy 7108 has been related
to grain refinement, adding grain refiners scandium and titanium-boron,
i.e. TiBor [29]. Using the circular patch test, cracking was avoided when
adding a minimum of 0.25 wt.%Sc or (0.02 wt.%Ti + 0.004 wt.%B),
Weld Parameter and Minor Element Effects 283
Travel Speed
Strain and stress fields around the weld pool have been experimentally ob-
served and simulated [10, 11, 16, 30-34]. These fields are dependent on
weld processing parameters as well as restraining conditions. The influ-
ence of torch travel speed on solidification cracking during welding has
been extensively studied, but is not well understood [16, 30-34]. Some
studies have observed an improvement in weldability when increasing the
torch travel speed [30-32]. These measurements were made for the GTA
process and travel speeds between 2.5 and 13 mm/s. Chihoski [30, 31] ex-
perimentally observed compressive and tensile cells around a moving weld
pool, and postulated that no cracking will occur if the mushy zone is under
compression. At high travel speeds, a compressive cell is located at the
mushy zone, thereby avoiding cracking. With decreasing travel speed, this
compressive cell diminishes and is replaced with a tensile cell that is
highly susceptible to cracking. However, some simulations [33] and ex-
perimental observations [34] show an increase in cracking susceptibility
with increasing travel speed. But in these situations, the welding speeds are
high; between 50 and 100 mm/s for laser welding [34] and between 16 and
25 mm/s for GTA welding [33].
Weld travel speed also influences the weld pool shape, which is known
to have a strong effect on weldability [35]. Increasing weld travel speed
modifies the weld pool shape from round to teardrop, which favours stray
grain formation and long continuous centreline grain boundaries. The ori-
entation of these long grain boundaries are perpendicular to the thermal
and mechanical strains, leading to a high cracking susceptibility. This
counters Chihoski´s observations where high torch travel speed should im-
prove weldability. Therefore, both weld pool shape and local strain cells
must be considered when determining solidification cracking susceptibility.
hence tend not to separate if mixed. Entrapped oxide films may influence
the cracking susceptibility, providing sites for crack nucleation due to de-
cohesion [36]. Such oxide films may also impair the fluidity and feeding
ability of the alloy [24], or affect the solidification path, serving as sites for
phase nucleation [37]. Above 750°C, the γ-Al2O3 will be present, which
acts as the preferred nuclei for β-Al5FeSi needles. The α-Al2O3 oxide sup-
presses nucleation of the β-phase, which as noted before is the phase be-
lieved to promote porosity.
Hydrogen contamination can arise from moisture or hydro-carbons (e.g.
machine oil). Hydrogen in molten aluminium often results in porosity, be-
cause of the sharp drop in solubility when going from liquid to solid. Dur-
ing solidification of pure aluminium, the solubility of hydrogen drops from
500 to 50 cubic millimetre per 100 grams of aluminium with 1 atm of hy-
drogen over the coupon [24]. The role of porosity on cracking is not well
understood, and it could conceivably have either positive or negative ef-
fects. Dissolved hydrogen should make it easier to cavitate, which is one
of the possible mechanisms for crack formation [36, 38, 39]. But pores
also serve to feed shrinkage, reducing the interdendritic volume that needs
to be fed, and hence lowering the interdendritic pressure drop [40].
more vertical this boundary (i.e. higher slope), the less influence filler dilu-
tion has on weldability. Accordingly, an alloy is considered more weldable
if less 4043 is needed to avoid cracking.
Fig. 3. Critical strain rate – dilution map for alloy 6060-T4, characterizing crack-
ing susceptibility in terms of local strain rate and 4043 filler dilution using CTW
test [5]
Experimental Approach
(a) (b)
Fig. 4. (a) Overview of CTW test machine and (b) dimensions of test coupon
Welding Parameters
Welding was performed using the gas-tungsten arc, cold-wire feed process
(GTAW-CWF). The base welding parameters are listed in Table 3. The arc
voltage was kept constant using an arc voltage control system, maintaining
a 2 mm arc gap corresponding to a 17.8 V arc voltage. The heat input was
chosen to obtain a bead-on-plate, full penetration weld. Filler wire speed
and tensile cross-head speed were experimental variables. Prior to the
CTW test, the oxide layer on the test coupon was chemically removed
(etch E1, Table 4), followed by degreasing with acetone.
Test Sequence
strain rate measurements were made at the observed crack initiation site to
correlate these measurements with those measured at weld mid-length. The
gage length of the extensometer was 10.5 mm which, by definition, influ-
ences local strain measurements.
The strain and strain rate values were taken at the coherency tempera-
ture, which corresponds to the point during solidification where the secon-
dary arms of two adjacent dendrites are beginning to be in contact. Ac-
cording to some theories [13, 18, 41–43], the coherency temperature
corresponds to the critical temperature where interdendritic liquid feeding
ability (healing) is sharply reduced due to the continuity of the solid phase.
The calculated strain rate value was also taken at the coherency tempera-
ture, at the critical stage where cracking begins. The position of coherency
relative to the torch along the centreline was determined by temperature
and weld pool shape measurements.
The arc was suddenly extinguished at the end of each test, freezing in
the weld pool shape, the boundary of which corresponds to the liquidus of
the weld metal. The position of the liquidus relative to the torch along the
weld centreline was measured. Temperature measurements were then per-
formed to determine the position of coherency point relative to the liquidus
along the weld centreline.
Temperature Measurement
The cooling curve for the weld pool was monitored during solidification
for all experimental conditions to determine the position of the coherency
point relative to the liquidus. A hole, 0.6 mm diameter and 1 mm deep,
was drilled from the bottom of the weld coupon, 35 mm from the start of
the weld. A 0.5 mm outer diameter, sheathed, electrically grounded,
nickel/chrome-nickel thermocouple was pre-placed inside the hole before
welding. The thermocouple output was recorded at a 200 Hz frequency
during solidification using CATMAN 4.5 computer software for data acqui-
sition and a Spider 8 analog-to-digital converter.
Material
Extrusions
Extruded aluminum 6060-T6 bar was used as the base material, and was
compared against 6060-T4 bar material studied previously [5]. Both mate-
rials came from different heats and had slightly different compositions.
Weld Parameter and Minor Element Effects 289
Composition of the 4043 filler and both 6060 heats were measured with
wet chemical analysis and spectrometry (Table 6). 6060-T4 and 6060-T6
extrusions had a different hardness (respectively 40 and 83 HV0.5) and
varied, in particular, in alloying elements Mn, Cu, and Zn. The size of the
weld coupons were 120 mm × 40 mm × 4 mm (Fig. 4b).
Table 6. Measured wet chemical analysis and emission spectrometry for alumi-
num alloys 6060, 4043, and controlled mixtures of 6060 with master alloys Al-
10Fe and Al-5Ti-B (wt.%) : (a) 6060-T4, (b) 6060-T6, (c) 4043, (d) insert
6060+Tibor (6060+15%(Al-5Ti)), (e) insert 6060+Fe (6060+15%(Al-10Fe)),
(f) weld pool 6060+insert (6060+Fe), and (g) weld pool 6060+ insert (6060+Tibor)
Composition (wt.%)
Si Mg Fe Mn Cu Cr Ni Zn Ti Ga Zr
(a) 0.42 0.59 0.19 0.02 0.01 0.004 0.004 0.009 0.02 0.01 0.001
(b) 0.51 0.51 0.21 0.04 0.03 0.003 0.003 0.04 0.02 0.01 0.001
(c) 5.25 0.001 0.21 0.02 0.3* --- --- 0.1* 0.2* --- ---
(d) 0.43 0.43 0.18 0.03 0.02 0.003 0.003 0.04 0.77 0.01 0.001
(e) 0.43 0.43 1.68 0.03 0.02 0.003 0.003 0.04 0.02 0.01 0.001
(f) 0.50 0.50 0.43 0.04 0.03 0.003 0.003 0.04 0.02 0.01 0.001
(g) 0.50 0.50 0.21 0.04 0.03 0.003 0.003 0.04 0.14 0.01 0.001
*handbook
--- value not specified
Inserts
The objective was to vary weld metal composition using inserts having a
controlled composition (Table 6). Inserts were machined from cast ingots
made from controlled mixtures of 6060-T6 with master alloys (e.g. Al-5Ti-
B or Al-10Fe). These inserts were then pre-welded into 6060-T6 CTW
coupons in preparation for CTW wedability testing.
Controlled mixtures of 6060-T6, 6060+15%(Al-5Ti-B), and
6060+15%(Al-10Fe) were cast. The composition of the resulting ingots,
when using master alloys, differs from 6060 respectively for titanium (0.8
wt.% instead of 0.02 wt.%) and iron (1.7 wt.% instead of 0.2 wt.%). The
casting mold is shown in Fig. 5, with inside dimensions 140 × 10 × 30 mm,
and mold thickness 10 mm. All mold materials with direct exposure to
molten aluminum were pre-coated with boron-nitride spray. Each cast heat
weighed approximately 130 grams and was melted in a graphite crucible
placed inside an electric furnace held at 800°C. Oxide dross was skimmed
just prior to casting, and the melt was rigorously stirred in order to ensure
thorough mixing with the master alloy. In the case of Al-5Ti-B, 6060-T6
was thoroughly melted before adding the TiBor master alloy, in order to
290 N. Coniglio, C.E. Cross
reduce time exposure of TiAl3 particles in the melt, and thus minimize
particle dissolution and promote efficient grain refinement. Five inserts
(2 mm × 2 mm × 140 mm) were machined from each cast ingot. For pur-
poses of providing an experimental control, inserts of cast 6060 material
were also prepared, left untreated with any master alloy.
(a) (b)
Fig. 5. Casting mold (a) photograph and (b) schematic, used to obtain ingots for
weld insert production
Shielding Gas
The crack- no crack boundary in a critical strain rate – dilution map was
determined for each experimental condition by measuring cracking suscep-
tibility at two levels of dilution: 0% and 15–20%. 4043 filler speed was
chosen to reach a dilution between 15% and 20%. Dilution is an important
concept in determining weld pool composition, approximated using Eq. (1):
B+C
filler dilution = × 100 (%) , (1)
A+B+C
where A is the melted area of 6060 base metal, and B+C is the difference
between the total area of the weld metal and A (Fig. 7). Calculated dilu-
tions are summarized in the Appendix.
For both low and high dilution levels, the transverse cross-head speed
was incremented in steps of 0.017 mm/s, noting the value where cracking
292 N. Coniglio, C.E. Cross
Fig. 7. Illustration for filler dilution calculation from weld metal cross-section
Strain measurements were made at the point where the mushy zone (co-
herency point) passes between the extensometer. The weld pool tempera-
ture was recorded during solidification to know where the mushy zone and
coherency was located relative to the torch position. Thermal analyses per-
formed previously measured the liquidus at 660°C and coherency at 624°C
for the 6060 alloy [5]. For all experimental conditions, it has been found
that the weld temperature drops during solidification from 660°C to 624°C
in approximately 0.2 s. Thus, knowing the travel speed, the distance from
the fusion line to the coherency point can be calculated.
The weld pool shape and resulting grain structure was observed and char-
acterized from the top surface. The application of the etch E2 (Table 4) on
the top surface of the weld revealed the grain structure and weld pool shape,
obtained by extinguishing suddenly the arc. Figure 8 illustrates the measured
distances to characterize the weld pool shape. For all the welding conditions,
the dimensional characteristics of the weld pool were: top width 8 mm, and
distance behind electrode 6 mm, with a deviation of ± 1 mm. The measure-
ments of the weld pool shape are summarized in the Appendix.
Weld Parameter and Minor Element Effects 293
(a) (b)
Fig. 8. Weld pool measurements : (a) top width (A) and (b) distance behind elec-
trode (B) and in front of electrode (C)
(a) (b)
Fig. 9. Weld cross sections of 6060-T6 coupons made with (a) 0% and (b) 17%
4043 dilution
294 N. Coniglio, C.E. Cross
(a) (b)
Fig. 10. (a) CTW test sequence superposed on the measured strain and (b) calcu-
lated strain rate, for the test conditions 6060-T4 with 9% 4043 dilution and 0.067
mm/s cross-head speed
Aluminium 6060 extrusions from two different heats and tempers were
evaluated with the CTW test using the same base welding parameters
(Table 3). A critical strain rate – dilution map comparing 6060-T4 and
6060-T6 is shown in Fig. 11 as discussed in the background. 6060-T6 has
been studied at 0 and 17% 4043 dilution, whereas data for 6060-T4 exists
from previous work (Fig. 3) [5] where CTW tests were run over a broad
range of dilutions (0%, 5%, 9%, 11%, 14%, and 16% 4043). A large dif-
ference in critical dilution is observed at high strain rates, where higher
4043 dilution is required to avoid cracking in 6060-T6. For example, at a
local strain rate of 0%/s, 7% 4043 dilution is required to avoid cracking in
6060-T4, while 17% dilution is required for 6060-T6. It is not known at
this point whether this difference in weldability is due to differences in
Weld Parameter and Minor Element Effects 295
Fig. 11. Critical strain rate-dilution map: (a) 6060-T4, previous data from [5] and
(b) 6060-T6, with data points shown for 0 and 17% 4043 dilution
CTW tests were made on aluminium 6060-T6 using the base parameters
(Table 3), but varying the current and weld travel speed. In addition, the
welding current was varied so as to maintain a constant weld pool size.
Weld travel speed was run at 2, 4, and 6 mm/s, and the current was varied
respectively from 80A to 145A. Data for 4 mm/s was obtained from a pre-
vious test (Fig. 11). Dilutions studied were 0%, and either 18% or 16%
4043 respectively for 2 and 6 mm/s weld travel speeds. The crack – no
crack boundaries for these conditions are compared against the boundary
found using the base welding parameters (Table 3), i.e. 4 mm/s weld travel
speed and 110A, as shown in Figs. 12 and 13. Increasing weld travel speed
should improve weldability following the relationship of Chihosky [30,
31], and this appears to be the case, but only at high strain rates. Specifi-
cally, at 18% 4043 dilution, solidification cracking occurs for a +0.12%/s
local strain rate at 2 and 4 mm/s weld travel speed, while no crack occurs
at +0.25%/s strain rate for a 6 mm/s travel speed. On the other hand, 6060-
T6 is less weldable at 0% 4043 dilution with increasing weld travel speed.
296 N. Coniglio, C.E. Cross
This is likely due to a change in weld metal grain structure, where stray
centreline grains were observed to occur only at 4 and 6 mm/s weld travel
speed.
Fig. 12. Critical strain rate – dilution map comparing crack-no crack boundaries at
weld travel speeds of (a) 4 mm/s and (b) 2 mm/s when welding 6060-T6. Data
points are shown for 2 mm/s
Fig. 13. Critical strain rate – dilution map comparing crack no crack boundaries at
weld travel speeds of (a) 4 mm/s and (b) 6 mm/s when welding 6060-T6. Data
points are shown for 6 mm/s
Weld Parameter and Minor Element Effects 297
Fig. 14. Critical strain rate – dilution map comparing crack-no crack boundary for
(a) 6060-T6 and (b) 6060-T6 with 6060 insert. Data points are shown for 6060 in-
serts
298 N. Coniglio, C.E. Cross
Fig. 15. Critical strain rate – dilution map comparing crack-no crack boundary for
6060-T6 with (a) 6060 insert and (b) 6060+1.7%Fe insert. Data points are shown
for 6060+1.7%Fe insert
CTW tests were performed with inserts high in TiBor content machined
from a 6060+0.8%Ti+0.16%B ingot (Table 6). A few scattered macro-
pores were observed in the weld cross-sections. The influence of TiBor on
cracking susceptibility was first observed in cast ingots, where 6060 ingots
cracked at the base of the pouring spout, and 6060+0.8%Ti+0.16%B ingots
Weld Parameter and Minor Element Effects 299
did not crack. The welding parameters were the same as those used for
6060 inserts. The 4043 dilutions studied were 0% and 15%. The weld pool
composition with the Ti rich insert differed by its higher titanium content
(from 0.02 to 0.14 wt.%Ti) as given in Table 6. Crack-no crack boundaries
are compared for both insert compositions in Fig. 16. The grain refiner ad-
dition significantly improved weldability, particularly at low 4043 dilu-
tions. This is not surprising considering the potent effect of refinement re-
ported for aluminium weldability improvement [29]. Moreover, it seems
that TiBor improves weldability with more efficiency than 4043, the
boundary being completely moved to higher strain rates for all dilutions.
Finally, the boundary is almost vertical; revealing that 4043 dilution has
little influence on the weldability of grain refined 6060 welds. In essence,
the grain refinement resulted from TiBor exceeds any metallurgical effects
resulting from 4043 filler dilution.
4043 filler is known to reduce the mechanical properties of 6xxx base
metal, and so the improved weldability comes at a price. This suggests that
it may prove feasible to develop a filler that is high in grain refiner, while
low in Si. It is interesting to note that Ti content in 4043 filler tends to be
very high (0.2 wt.%Ti) compared to the initial Ti content in 6060-T6 (0.02
wt.%Ti). Increasing 4043 filler dilution increases the Ti content in the
weld metal, promoting grain refinement [5]. Thus, improved grain refine-
ment should be achieved with increasing 4043 dilution, which may explain
why the crack-no crack boundaries tend to be closer at high 4043 dilutions
(Fig. 16). Future studies are required to determine the improvement in
weldability when increasing the only Si content of the weld without in-
creasing the Ti content.
CTW tests were performed using an oxygen containing shielding gas (flow
rate: 0.33 L/s He + 0.03 L/s Ar + 1%O2). This required a decrease in the
welding current to 100A to keep a constant weld pool size. 4043 dilutions
of 0% and 18% were studied. High quantities of oxides were observed at
the weld surface. Crack-no crack boundaries are compared between nor-
mal and oxygen-containing gas in Fig. 17, where use of oxygen is found to
slightly improve weldability. At 18% 4043 dilution and a local strain rate
of +0.09%/s, a crack should occur when welding with helium, but does not
initiate with oxygen added to the helium. This trend goes counter to the be-
lief that oxygen forms oxides films that can nucleate cracks [24]. However,
300 N. Coniglio, C.E. Cross
it may just be that a heavy oxide is formed at the pool surface and does not
get mixed into the weld pool.
Fig. 16. Critical strain rate – dilution map comparing crack-no crack boundary for
6060-T6 with (a) 6060 insert and (b) 6060+0.8%Ti insert. Data points are shown
for 6060+0.8%Ti insert
Fig. 17. Critical strain rate - dilution map comparing crack-no crack boundary for
6060-T6 welded with (a) 0.33 L/s He and (b) 0.33 L/s He + 0.03 L/s (Ar+1%O2).
Data points are shown for He+(Ar+1%O2) shielding gas
Weld Parameter and Minor Element Effects 301
One of the aims of the present work was to use the CTW test to evaluate
how small variations in alloy composition may affect weldability. In pur-
suing this aim, the sensitivity and limitations of this new weldability test
have also been established. Some of the inherent problems associated with
this test are discussed below.
Position of Extensometer
In executing the CTW test sequence (Table 5), the extensometer used to
measure local weld strain was positioned at weld mid-length. However,
ideally the best place to measure strain is at the crack initiation site, which
was found to always occur at the point where the mushy zone sits when
strain is first applied, 25 mm after the weld start. In order to determine
how strain at these two locations may differ (i.e. 25 mm versus 50 mm),
four tests were performed with the extensometer positioned 25 mm from
the weld. These values were then compared against the standard mid-
length measurements, as shown in Table 8. These values show good
agreement (less than 0.05%/s difference), except for one condition at high
cross-head speed where there was a large difference (0.29%/s). This sug-
gests that the standard sequence should be modified for future tests, to re-
locate the extensometer to the 25 mm position.
302 N. Coniglio, C.E. Cross
Fig. 18. Un-etched weld cross sections of 6060-T6 coupon made with He+Ar+H2
and 0% 4043 filler dilution revealing interdendritic pores (arrows)
Fig. 19. Critical strain rate – dilution map comparing crack-no crack boundary for
6060-T6 welded with (a) 0.33 L/s He and (b) 0.33 L/s He + 0.27 L/s (Ar+2%H2).
Data points are shown for He+( Ar+2%H2) shielding gas
Weld Parameter and Minor Element Effects 303
As can be seen in any of the strain rate-dilution maps, cracks that occur at
low 4043 dilutions involves a negative critical strain rate (i.e. inward
movement of material to feed shrinkage). At first hand this appears counter
intuitive, but may actually reflect upon the material’s very poor weldabil-
ity. Even with the inward movement of base material, the solidification
shrinkage is apparently not sufficiently compensated to avoid cracking.
Data from previous work has shown similar behavior, where cracking may
occur even with inward movement of base material [5].
Crack Boundary
The nature of the crack-no crack boundary for the 6060/4043 alloy system
is such that variations in critical dilution of 0–20% corresponds to a maxi-
mum difference in critical strain rate, expressed in terms of cross-head
speed, of roughly 0.1 mm/s. In terms of test sensitivity, it would be prefer-
able that this number be much larger. As it is, this means that test incre-
ments approaching 0.01 mm/s are needed to accurately identify the loca-
tion of a boundary (i.e. one tenth of the full range). In this work,
increments of 0.017 mm/s were used, which only roughly determined the
boundary for each experimental condition. Thus, the use of finer incre-
ments should be considered for future tests on this alloy. Other alloy sys-
tems will likely exhibit their own unique sensitivity to strain rate.
Interacting Factors
the weld metal, but increases the size and shape of the weld bead. Thus, it
becomes difficult to separate the metallurgical effects from the mechani-
cal. Likewise, variations in travel speed or changes in shielding gas will al-
ter the heat input and, although corrected with adjustments to current, there
may be subtle changes to pool shape that can affect grain structure. Thus, it
becomes difficult to know if the effect of travel speed on weldability is due
to a change in local strain rate or grain structure. The application of model-
ling may help to answer these questions in the future.
Conclusion
The newly developed CTW test, evaluated in terms of critical strain rate –
dilution maps, has been successfully used to quantify the weldability of the
aluminium 6060/4043 alloy system. In particular, the effect of minor alloy
additions and travel speed on solidification cracking susceptibility has
been studied by observing how these changes shift the crack-no crack
boundary in a strain rate-dilution map. Small additions of oxygen or iron
were found to have little or no effect on weldability. Improvements in
weldability were experienced with the addition of hydrogen or a grain re-
finer, or with the use of faster travel speeds.
Acknowledgement
The authors are grateful to BAM for internal funding of this project, and
specifically wish to thank Th. Böllinghaus and Th. Kannengiesser for their
insightful support of this research effort. In addition, technical support at
BAM from A. Hannemann, P. Friedersdorf, K. Scheideck, R. Breu, M.
Lammers, and M.Richter was greatly appreciated. The authors are also
grateful to Metallurg Aluminum for donation of master alloys.
References
3. Mousavi MG, Cross CE, Grong Ø, Hval M (1997) Controlling Weld Metal
Dilution for Optimized Weld Performance in Aluminum. Sci. Tech. Weld.
Joining 2:275–278
4. Liptak JA, Baysinger FR (1968) Welding Dissimilar Aluminum Alloys.
Welding Journal 47:173s–180s
5. Coniglio N, Cross CE, Michael Th, Lammers M (2008) Defining a Critical
Weld Dilution to Avoid Solidification Cracking in Aluminum. Welding Jour-
nal (in review)
6. Pellini WS (1952) Strain Theory for Hot Tearing. Foundry 80:125–199
7. Prokhorov NN (1956) The Problem of the Strength of Metals While Solidify-
ing During Welding. Svarochnoe Proizvodstvo 6:5–11
8. Prokhorov NN, Bochai MP (1958) Mechanical Properties of Aluminum Al-
loys in the Crystallization Temperature Range during Welding. Svarochnoe
Proizvodstvo 2:1–6
9. Prokhorov NN, Gavrilyuk MN (1971) Strain Behaviour of Metals during So-
lidification after Welding. Welding Production 13(6):8–13
10. Zacharia T (1994) Dynamic Stresses in Weld Metal Hot Cracking. Welding
Journal 73(7): 164s–172s
11. Feng Z (1994) A Computational Analysis of Thermal and Mechanical Conditions
for Weld Metal Solidification Cracking. Welding in the World 33:340–347
12. Kannengiesser T, McInerney T, Florian W, Böllinghaus T, Cross CE (2002)
The influence of local weld deformation on hot cracking susceptibility. In:
Mathematical modelling of weld phenomena 6, Maney, pp 803–817
13. Feurer U (1977) Influence of Alloy Composition and Solidification Condi-
tions on Dendrite Arm Spacing, Feeding, and Hot Tear Properties of Alumi-
num Alloys. In: Proc. international symposium engineering alloys, Delft,
131–145
14. Senda T, Matsuda F, Takano G, Watanabe K, Kobayashi T, Matsuzaka T
(1971) Experimental Investigations on Solidification Crack Susceptibility for
Weld Metals with Trans–Varestraint Test. Trans. of JWRI 2(2):141–162
15. Arata Y, Matsuda F, Nakata K, Shinozaki K (1977) Solidification Crack Sus-
ceptibility of Aluminum Alloy Weld Metals (Report II) – Effect of Straining
Rate on Cracking Threshold in Weld Metal during Solidification. Trans. of
JWRI 6(1):91–104
16. Matsuda F, Nakagawa H, Nakata K, Okada H (1979) The VDR Cracking Test
for Solidification Crack Susceptibility on Weld Metals and its Application to
Aluminum Alloys. Trans. of JWRI 8:85–95
17. Tamura H, Kato N, Ochiai S, Katagiri Y (1977) Cracking Study of Aluminum
Alloys by the Variable Tensile Strain Hot Cracking Test. Trans. JWS 8:16–22
18. Rappaz M, Drezet JM, Gremaud M (1999) A New Hot–Tearing Criterion.
Met. Mat. Trans. 30A:449–455
19. Rindler W, Kozeschnik E, Enzinger N, Buchmayr B (2002) A Modified Hot
Tearing Criterion for Steels. In: H Cerjak (ed) Mathematical Modelling of
Weld Phenomena 6, Woodhead Publishing Limited, pp 819–835
306 N. Coniglio, C.E. Cross
20. Jennings PH, Singer ARE, Pumphrey WI (1948) Hot-Shortness of some High-
Purity Alloys in the Systems Al-Cu-Si and Al-Mg-Si. J. Inst. Metals 74:227–248
21. Mondolfo LF (1976) Aluminum Alloys–Structure & Properties, Butterworths,
London
22. Coniglio N, Cross CE (2006) Characterization of Solidification Path for Alu-
minum 6060 Weld Metal with Variable 4043 Filler Dilution. IIW-1755–06,
Welding in the World 50(11/12):14–23
23. The Aluminum Association (2006) International Alloy Designations and
Chemical Composition Limits for Wrought Aluminum and Wrought Alumi-
num Alloys
24. Van Horn KR (1967) Aluminum Properties, Physical Metallurgy, and Phase
Diagrams, vol.1. American Society of Metals, Ohio
25. Lu L, Dahle AK (2005) Iron-rich Intermetallic Phases and their Role in Cast-
ing Defect Formation in Hypoeutectic Al-Si Alloys. Met. Mat. Trans.
36A:819–835
26. Wang L, Makhlouf M, Apelian D (1995) Aluminum Die Casting Alloys : Al-
loy Composition, Microstructure, and Properties-Performance Relationships.
International Materials Reviews 40(6):221–238
27. Mbuya TO, Odera BO, Ng’ang’a SP (2003) Influence of Iron on Castability
and Properties of Aluminum Silicon Alloys: Literature Review. International
Journal of Cast Metals Research 16(5):451–465
28. Singer ARE, Jennings PH (1947) Hot-Shortness of some Aluminum-Iron-
Silicon Alloys of High Purity. J. Inst. of Metals 73:273–284
29. Mousavi MG, Cross CE, Grong Ø (1999) Effect of Scandium and Titanium-
Boron on Grain Refinement and Hot Cracking of Aluminum Alloy 7108. Sci-
ence and Technology of Welding and Joining 4(6):381–388
30. Chihoski RA (1972) The Character of Stress Field Around a Weld Arc Mov-
ing on Aluminum Sheet. Welding Journal 51(1):9s–18s
31. Chihoski RA (1979) Expansion and Stress Around Aluminum Weld Puddles.
Welding Journal 58(9):263s–276s
32. Hunziker O, Dye D, Roberts SM, Reed RC (2005) A Coupled Approach for
the Prediction of Solidification Cracking during the Welding of Superalloys.
In: Böllinghaus Th, Herold H (eds) Hot Cracking Phenomena in Welds,
Springer-Verlag Berlin Heidelberg, pp 299–319
33. Shibahara M, Serizawa H, Murakawa H (2005) Finite Element Method for
Hot Cracking Analysis using Temperature Dependent Interface Element. In:
Böllinghaus Th, Herold H (eds) Hot Cracking Phenomena in Welds, Springer-
Verlag Berlin Heidelberg, pp 253–267
34. Cicala E, Duffet G, Andrzejewski H, Grevey D, Ignat S (2005) Hot Cracking
in Al-Mg-Si alloy laser welding – operating parameters and their effects. Ma-
terials Science and Engineering 395A:1–9
35. Savage WF, Aronson AH (1966) Preferred orientation in the weld fusion
zone. Welding Journal 45:85s–89s
36. Campbell J (1991) Castings, Butterworth-Heinemann, Oxford, Boston
37. Westengen H (1982) Formation of Intermetallic Compounds During DC Cast-
ing of a Commercial Purity Al-Fe-Si Alloy. Z. Metallkde. 73:360–368
Weld Parameter and Minor Element Effects 307
38. Cross CE, Olson DL, Edwards GR (1993) The Role of Porosity in Initiating
Weld Metal Hot Cracks. In: Zacharia Th (ed) Int. Conf. Proc. On Modeling
and Control of Joining Processes, pp 549–557
39. Talbot DEJ (2004) The Effects of Hydrogen in Aluminum and Its Alloys,
Maney Publishing, London
40. M´Hamdi M, Mo A (2005) On Modelling the Interplay between Microporos-
ity Formation and Hot Tearing in Aluminium Direct-Chill Casting. Materials
Science and Engineering A:105–108
41. Arata Y, Matsuda F, Nakata K, Shinozaki K (1977) Solidification Crack Sus-
ceptibility of Aluminum Alloy Weld Metals (Report III) - Effect of Straining
Rate on Crack Length in Weld Metal. Trans. of JWRI 6(2):47–52
42. Borland JC (1960) Generalized Theory of Super Solidus Cracking in Welds
(and Castings). British Welding Journal: 508–512
43. Pumphrey WI, Jennings PH (1948) A Consideration of the Nature of Brittle-
ness and Temperature Above the Solidus in Castings and Welds in Aluminum
Alloys. J. Inst. Metals 75:235–256
Appendix
The following tables reveal the different measurements realized for all the
studied experimental conditions. They explicitly relates the calculated dilu-
tion to the 4043 filler feeding speed, the weld pool shape measurements
according to Fig. 8, and dimensional cross sectional characteristics of the
weld pools. Each table correspond to one experimental condition at low
and high 4043 filler dilutions.
Table A.1. 4043 filler dilution and dimensional characteristics for CTW welds of
alloy 6060-T4 [5]. Distances (A), (B), (C) are according to Fig. 8
Distance Distance in front
Filler Wire 4043
Top Width behind of electrode
Base metal Speed Dilution
(A) (mm) electrode (B) (C) (mm)
(mm/s) (%)
(mm)
6060-T4 0 0 7.5 6.1 4.7
6060-T4 41.7 16 8.1 5.9 4.8
Cross
Bead Root Over-Bead Over-Bead
Section
Thickness Width Width Curvature
Area
(mm) (mm) (mm) (mm–1)
(mm²)
25.1 4.2 5.5 7.5 –0.068
31.7 5.0 6.5 8.1 +0.011
308 N. Coniglio, C.E. Cross
Table A.2. 4043 filler dilution and dimensional characteristics for CTW welds of
alloy 6060-T6. Distances (A), (B), (C) are according to Fig. 8
Distance Distance in
Filler Wire 4043
Top Width behind front of
Base metal Speed Dilution
(A) (mm) electrode electrode
(mm/s) (%)
(B) (mm) (C) (mm)
6060-T6 0 0 7.5 6.3 5.6
6060-T6 41.7 17 7.8 5.9 5.3
Cross Sec- Bead Root Over-Bead Over-Bead
tion Area Thickness Width Width Curvature
(mm²) (mm) (mm) (mm) (mm–1)
26.1 4.8 5.8 7.8 –0.066
31.5 5.0 6.9 8.1 +0.009
Table A.3. 4043 filler dilution and dimensional characteristics for CTW welds of
alloy 6060-T6 with 6060 insert. Distances (A), (B), (C) are according to Fig. 8
Distance Distance in
Filler Wire 4043
Top Width behind front of
Base metal Speed Dilution
(A) (mm) electrode electrode
(mm/s) (%)
(B) (mm) (C) (mm)
6060-T6 0 0 8.0 5.6 6.1
6060-T6 41.7 17 7.9 6.0 5.0
Cross Bead Root Over-Bead
Over-Bead
Section Thickness Width Curvature
Width (mm)
Area (mm²) (mm) (mm) (mm–1)
24.2 4.3 5.7 8.1 –0.056
28.7 5.1 6.0 8.0 +0.025
Table A.4. 4043 filler dilution and dimensional characteristics for CTW welds of
alloy 6060-T6 with 6060+1.7%Fe insert. Distances (A), (B), (C) are according to
Fig. 8
Distance Distance in
Filler Wire 4043
Top Width behind front of
Base metal Speed Dilution
(A) (mm) electrode (B) electrode
(mm/s) (%)
(mm) (C) (mm)
6060-T6 0 0 8.3 6.0 6.4
6060-T6 41.7 20 7.5 5.1 5.5
Cross
Bead Root Over-Bead
Section Over-Bead
Thickness Width Curvature
Area Width (mm)
(mm) (mm) (mm–1)
(mm²)
26.7 4.3 5.7 8.5 –0.053
29.6 5.3 6.5 7.6 +0.025
Weld Parameter and Minor Element Effects 309
Table A.5. 4043 filler dilution and dimensional characteristics for CTW welds of
alloy 6060-T6 with 6060+0.8%Ti insert. Distances (A), (B), (C) are according to
Fig. 8
Distance Distance in
4043
Filler Wire Top Width behind front of
Base metal Dilution
Speed (mm/s) (A) (mm) electrode electrode
(%)
(B) (mm) (C) (mm)
6060-T6 0 0 8.9 6.2 6.3
6060-T6 41.7 15 9.0 6.0 5.6
Cross Bead Root Over-Bead
Over-Bead
Section Area Thickness Width Curvature
Width (mm)
(mm²) (mm) (mm) (mm–1)
25.2 4.3 4.5 8.4 –0,025
28.7 5.1 5.1 7.9 +0.015
Table A.6. 4043 filler dilution and dimensional characteristics for CTW welds of
alloy 6060-T6 welded at a torch travel speed of 2 mm/s. Distances (A), (B), (C)
are according to Fig. 8
Distance Distance in
Filler Wire 4043
Top Width behind front of
Base metal Speed Dilution
(A) (mm) electrode electrode
(mm/s) (%)
(B) (mm) (C) (mm)
6060-T6 0 0 8.2 6.6 7.2
6060-T6 20.8 18 8.3 6.3 6.7
Cross Bead Root Over-Bead Over-Bead
Section Thickness Width Width Curvature
Area (mm²) (mm) (mm) (mm) (mm–1)
25.7 4.0 6.3 8.3 –0.074
28.5 5.2 6.1 8.1 +0.000
Table A.7. 4043 filler dilution and dimensional characteristics for CTW welds of
alloy 6060-T6 welded at a torch travel speed of 6 mm/s. Distances (A), (B), (C)
are according to Fig. 8
Distance Distance in
Filler Wire 4043 Top Width behind front of
Base metal
Speed (mm/s) Dilution (%) (A) (mm) electrode electrode
(B) (mm) (C) (mm)
6060-T6 0 0 7.3 5.4 5.6
6060-T6 62.5 16 7.2 5.4 4.3
Cross Bead Over-Bead Over-Bead
Root Width
Section Thickness Width Curvature
(mm)
Area (mm²) (mm) (mm) (mm–1)
25.1 4.2 6.5 7.6 –0.107
30.1 5.1 6.7 7.9 –0.015
310 N. Coniglio, C.E. Cross
Table A.8. 4043 filler dilution and dimensional characteristics for CTW welds of
alloy 6060-T6 welded with He+(Ar+1%O2) shielding gas. Distances (A), (B), (C)
are according to Fig. 8
Distance Distance in
Filler Wire
4043 Top Width behind front of
Base metal Speed
Dilution (%) (A) (mm) electrode electrode
(mm/s)
(B) (mm) (C) (mm)
6060-T6 0 0 8.2 5.8 6.4
6060-T6 41.7 18 7.6 6.0 6.4
Cross Bead Over-Bead
Root Width Over-Bead
Section Area Thickness Curvature
(mm) Width (mm)
(mm²) (mm) (mm–1)
27.6 4.2 7.1 8.4 –0.108
29.1 5.1 6.4 7.6 +0.000
Table A.9. 4043 filler dilution and dimensional characteristics for CTW welds of
alloy 6060-T6 welded with He+(Ar+2%H2) shielding gas. Distances (A), (B), (C)
are according to Fig. 8
Distance Distance in
Filler Wire
4043 Top Width behind front of
Base metal Speed
Dilution (%) (A) (mm) electrode electrode
(mm/s)
(B) (mm) (C) (mm)
6060-T6 0 0 8.7 5.9 6.2
6060-T6 41.7 18 8.4 5.8 6.0
Cross Bead Over-Bead
Root Width Over-Bead
Section Area Thickness Curvature
(mm) Width (mm)
(mm²) (mm) (mm–1)
26.4 4.3 6.7 8.8 –0.084
31.4 5.1 6.4 8.5 +0.015
Using Simulation for Investigations of Hot
Cracking Phenomena in Resistance Spot Welding
of 6xxx Aluminum Alloys (AA6016 and AA6181)
Abstract
This study presents how Numerical Simulation, by using the Finite Ele-
ment Method (FEM), can help to avoid hot cracking phenomena for resis-
tance spot welding of 6xxx AlMgSi alloys. It is well known that aluminum
spot welding is a difficult task especially for 6xxx alloys. Small silicon
contents in these alloys increase the cracking sensitivity dramatically in
comparison with 5xxx alloys. In order to reduce tedious experimental work
and to understand such problems SORPAS®, a commercial available Finite
Element Method software for resistance welding, was used. Virtual simu-
lation experiments for the innovative resistance spot welding process
DELTA SPOT™ led to optimized welding parameters in order to get the
desired nugget size without surface cracks.
Introduction
Using the innovative resistance spot welding process DELTA SPOT™ [1]
one can weld a wide range of metal combinations like three-sheet alumi-
num, three-sheet steel and dissimilar metals like aluminum-steel. All this is
made possible by using a coated process tape in combination with a resis-
tance spot welding gun – Figs. 1 and 2.
312 A. Eder et al.
The tape is posed between the electrodes and the sheets to be joined.
The resistance caused by the tape generates an additional heat input into
the work piece. It not only allows dissimilar material welding but also
eliminates work piece-coating and contaminant effects. It leads to ideal
contact, splash-free welds and high electrode lifetime. After every weld the
process tape moves to its next position. The contact surface therefore al-
ways remains clean. Additionally the fine coating of the tape optimizes the
electrical contact to the work piece, especially for welding aluminum
where it prevents spatter formation. Each welding spot is reproducible –
there is always the same quality as at the start.
DELTA SPOT™ was designed with the automotive and supplier indus-
tries very much in mind. One goal is to solve difficult welding tasks aris-
ing from the use of new materials in the automotive industry. Therefore
welding aluminum, especially high strength aluminum alloys, is an impor-
tant topic. For resistance spot welding of aluminum the process tape has
another important effect. It can be used to control the heat input into the
workpiece. The additional heat due to the resistance of the process tape
compensates the negative effect of the high conductivity of aluminum. The
tape leads to high process efficiency and reliability.
This study concentrates on the problems arising by resistance spot weld-
ing of aluminum and its alloys – especially the problem of hot cracking.
The difference between welding with and without process tape is demon-
strated for AlMg3 (AA 5754). As an example for difficult to weld 6xxx
aluminum alloys, AlMg0.4Si alloys are investigated. For these alloys it is
demonstrated how simulation can be used to find an optimal process curve
for resistance spot welding in order to avoid hot cracking.
Using Simulation for Investigations of Hot Cracking Phenomena 313
These phenomena do not appear for welding AlMg3 with a process tape
(in this case PT1407). The process parameters where chosen as in the con-
ventional spot welding example, except the current which could even be
reduced. In the macro-section there are just some small eutectics visible.
Compare Fig. 5. The nugget size is larger and it has a cylindrical form
which leads to a higher tensile strength – Fig. 6 [6]. The cylindrical form
of the weld nugget is caused by the additional heat from the process tapes.
This changes the cooling behavior and the stress distribution in the sheets
which helps reduce cracking sensitivity. It also reduces the gap adjacent to
the weld nugget.
As mentioned above for spot welding with a process tape less current is
necessary due to the additional heat from the tape. It makes a weld more
economically. Additionally the electrodes stay clean due to the protection
of the tape. Therefore a high quality of the sheet surface appearance can be
reached.
Spot welding of the high strength aluminum alloys is much more sophisti-
cated to solve. Test welds of 6xxx (AA6016) alloys with the process tape
PT 1200 were made. The welding parameters were 3 kN with one pulse of
approx. 17 kA for two 1 mm sheets. The standard electrode R70 was used.
Even with the aid of a process tape cracks appeared at the surface and in
the nugget – see Fig. 7. Due to the large solidification interval of AA6016
and the critical concentration of the alloying elements a high risk of crack-
ing is present.
In order to understand and to solve this problem the temperature profile
during cooling of the spot weld was needed. Therefore a series of Finite
Element simulations with the commercial available resistance welding pro-
gram SORPAS® were made.
Using Simulation for Investigations of Hot Cracking Phenomena 317
The simulation of the resistance spot welding process involves strong in-
teractions between the work pieces, electrodes, interfaces, process tape,
coatings and machine settings. Figure 8 [7, 8] shows the main components
which have to be considered for resistance spot welding with a process
tape.
Fig. 9. FEM mesh for resistance spot welding with a process tape
Using Simulation for Investigations of Hot Cracking Phenomena 319
A typical FEM mesh for solving a resistance spot welding process with an
additional tape is shown in Fig. 9. An adaptation of the standard program
settings was necessary to be able to include the thin tape and coatings.
1 ∂ § ∂φ · ∂ § ∂φ · (1)
¨ rσ r ¸ + ¨σ z ¸ = 0,
r ∂r © ∂r ¹ ∂z © ∂z ¹
where σ is the electrical conductivity of the materials.
The temperature distribution T is a result of the transient heat transfer
differential equation with internal heat source Q
1 ∂§ ∂T · ∂ § ∂T · ∂T (2)
¨ rλr ¸ + ¨ λz ¸+Q = ρ C ,
r ∂r © ∂r ¹ ∂z © ∂z ¹ ∂t
where λ is the thermal conductivity, ρ is the mass density and C is the
heat capacity. All material data is given in a temperature dependent way,
which allows phase transformation calculations.
The deformation, the stress, strain distribution and the contact areas at
interfaces are calculated by the variation approach for the functional of the
potential energy
∂π (3)
π = ³ σ εdV − ³ FvdS , =0
V S
∂v
The first term on the right hand side is the potential energy of the bulk
deformation. The second term respects the boundary conditions due to ex-
ternal load or velocity and friction etc. Wanheim and Bay’s friction theory
for real contact areas [9] is used for the contact resistivity:
§σ ·§ ρ1 + ρ 2 ·
ρ contact = 3¨¨ s _ soft ¸¸¨ + ρ calibration ¸ ,
© σn ¹© 2 ¹ (4)
320 A. Eder et al.
where σ s _ soft is the flow stress of the softer metal, σ n is the contact
normal pressure at the interface. The subscripts 1 and 2 indicate the two
base metals in contact. To take account for surface contaminants such as
oil, water, dirt and oxides etc. a term called ρ calibration is introduced in the
model. The quantitative influence of the surface contaminants on the con-
tact resistance is usually not known. Therefore the parameter ρ calibration in
Eq. (4) can be used for calibration. In order to complete the mathematical
model, appropriate boundary and initial conditions are used for the electri-
cal potential, temperature field and mechanical model.
Solver Strategy
FEM techniques are applied for Eqs. (1–4) to obtain a numerical discretiza-
tion. Figure 10 shows the solver strategy, designed to obtain efficient spot
welding simulations [7].
Using Simulation for Investigations of Hot Cracking Phenomena 321
During resistance spot welding the main process parameters, welding cur-
rent and electrode force, can be varied in order to get high quality welds.
At the beginning of a spot weld the force increases in a linear manner
which leads to a good electrical contact between the work pieces. After
this preload the current pulse starts. Depending on the material there can
be used one or several current pulses. For example it can be useful to set a
preheating pulse to avoid critical cooling rates [10].
The process engineer has to choose the right process curve for the mate-
rials to be welded. Due to the variety of process curves this can be a diffi-
cult and time consuming task. SORPAS® allows any force and current
form as well as material combinations which makes it suitable for process
optimization. Several simulations can be run successively in a batch proc-
ess for parametric studies. The optimization loop applied for solving diffi-
cult spot welding problems is shown in Fig. 11.
Usually the simulation starts on the basis of a first practical trial. Ac-
cording to a spot weld macro-section resulting from a first guess of a
current-force pulse the simulation model is calibrated. The calibration usu-
ally is done via the contact resistance model – compare Eq. (4). Then vir-
tual parameter studies by variation of the shape of the current-force pulse
have to be done. This results in a series of work piece temperature profiles.
These temperature histories in combination with metallurgical knowledge
322 A. Eder et al.
The goal for resistance spot welding the aluminum alloys AA 6016 and
AA 6181 was to minimize hot cracking as well as to get the desired nugget
size. A virtual parameter study for different types of welding process
curves (current and force) was done in SORPAS®. Exemplary are some re-
sulting temperature histories close to the work piece surface on the sym-
metry axis for AA 6016 as shown in Fig. 12.
Fig. 12. Simulation of temperature history close to the work piece surface
Using Simulation for Investigations of Hot Cracking Phenomena 323
Conclusion
References
Abstract
A study of Ni-base weld metal hot cracking resistance was made compar-
ing results from Modified Varestraint-Transvarestraint Testing (MVT),
metallographic evaluation and mechanical testing. Submerged arc weld
metals were produced with Alloy 625, Alloy 59 and Alloy C-276 filler
wires at two levels of heat input. Welds were subjected to MVT testing,
three-point and wrap-around bend testing as well as metallographic inspec-
tion of weld cross-sections. All test methods generally ranked the lower
heat input welds as more crack resistant than those produced with a higher
heat input. However, MVT testing ranked alloy types in a different order
than mechanical testing and metallographic evaluation. Alloy 59 type weld
metals were judged most crack resistant with MVT testing and C-276
welds were ranked as slightly better than those produced with Alloy 625
wires. Mechanical testing and metallographic evaluation, on the other
hand, ranked the Alloy 625 type weld metals as most crack resistant and
Alloy C-276 weld metals as least resistant. Discrepancies between results
of the different test methods and practical experience can most likely be at-
tributed to a number of factors including metallurgical aspects and test
procedures. It is concluded that test methods simulating actual application
welding conditions as closely as possible are most likely to accurately pre-
dict hot cracking susceptibility of real weldments. An assessment of the
hot cracking susceptibility of welding consumables should therefore pref-
erably be based on results from a combination of test methods.
Introduction
Large vessels and tanks for transportation and storage of liquefied natural
gas (LNG) at cryogenic temperatures are since many years commonly
constructed from 9% Ni steel. One of the most important steps in the
330 L. Karlsson et al.
Experimental
Three types of Ni-base SAW weld metals were subjected to MVT testing,
metallographic evaluation and bend testing.
Weld Metal
SAW weld metals were produced with three types of 1.6 mm diameter Ni-
base filler wires (Alloy 625, Alloy 59 and Alloy C-276). Welding was
done with a basic flux at two levels of heat input for each filler wire type.
Typical all-weld metal composition and properties for the three SAW
wire/flux combinations are presented in Tables 1 and 2.
Metallographic Evaluation
Bend Testing
The presence of cracks was also evaluated using three-point bend and wrap
around bend testing. Three-point bend testing was done with 10 mm thick,
transverse side bend specimens and longitudinal face bend test specimens.
The diameter of the former was in both cases 40 mm and the bending an-
gle was 180°.
Wrap around testing was done using 5 mm thick, transverse side bend
test specimens. The diameter of the former was 24 mm and the bending
angle was 180°. All testing was done in accordance with SS-EN 910.
MVT Testing
In MVT testing a GTA torch is used to remelt the machined surface of the
weld metal to be tested. As the arc passes a predefined point the specimen
is bent at a constant rate around a die. The specimen surface is thereby
subjected to a strain of well-defined level and rate. As a result of the
334 L. Karlsson et al.
stress/strain distribution during bending the hot cracks form almost exclu-
sively at or near the specimen surface. On completion of the test the
specimen surface is visually examined for cracks at a magnification of
X25. The number and total length of all visible cracks are determined and
plotted as a function of the bending strain [5].
Specimens (10×40×100 mm) for MVT testing were machined 2 mm
below the weld top surface resulting in a test surface located largely in the
top layer of beads. Welding was done under Ar and with a nominal energy
input of approximately 0.75 kJ/mm. Bending was done transverse to the
welding direction at a bending rate of 1.8 mm/s with bending radii selected
to produce strains of 1%, 2% and 4%. One specimen was tested per weld
and strain level.
Results
Metallographic Evaluation
Bend Tests
Results from bend testing are summarised in Table 4. The two types of
side bend tests and the longitudinal bend test were in complete agreement.
Welds 625-L and 625-H passed without remarks except in side bend wrap
around testing where one small crack was detected at the edge of the
specimen. Only a limited number of small cracks shorter than 0.5 mm
(here called micro-cracks) were seen in 59-L and 59-H specimens (Fig. 3).
All 276-L and 276-H specimens contained numerous micro-cracks and in
most cases also larger cracks (Fig. 4). One 276-H 3-point side bend
specimen even broke at a bending angle of 120°.
Fig. 4. Large cracks and numerous micro-cracks on strained surface of 276-L bend
test specimens
(a) 3 point side-bend specimen
(b) Longitudinal face-bend specimen
MVT Tests
Results from MVT testing is presented as tables giving total cracks lengths
(Table 5) and total number of cracks (Table 6) for all strain rates, as a
standard crack length versus strain diagram (Fig. 5) and as a diagram
presenting number of cracks against strain (Fig. 6).
Evaluating Hot Cracking Susceptibility of Ni-Base SAW Consumables 339
Table 5. Results of MVT tests presented as combined crack lengths for all strains
Weld Combined crack length for 1%, 2% and 4% strain (mm)
GTA WM Weld metal GTA WM + “Distant Total
HAZ Weld metal cracks”
HAZ
59-L 10.2 3.8 14.0 5.3 19.3
59-H 6.8 4.1 0.9 2.1 13.0
625-L 16.8 3.1 19.9 1.6 21.5
625-H 17.6 2.6 20.2 2.2 22.4
276-L 12.1 5.6 17.7 4.0 21.7
276-H* 15.3 6.8 22.1 10.3 32.4
* only tested at 2% and 4% strain
Table 6. Results of MVT test presented as the total number of cracks for all
strains
Weld Number of cracks
GTA WM Weld metal GTA WM + “Distant Total
HAZ Weld metal cracks”
HAZ
59-L 21 10 31 18 59
59-H 16 12 28 7 35
625-L 29 14 43 10 53
625-H 38 19 57 4 61
276-L 22 14 36 17 53
276-H* 20 18 38 37 75
*only tested at 2% and 4% strain
Newly formed cracks, at a distance from the GTA bead fusion boundary
(ductility dip cracks), could not easily be distinguished from hot cracks
already present in the weld metals before MVT testing. The “distant”
cracks are therefore included in the tables but excluded from the data used
to construct diagrams.
Using combined crack lengths or total number of cracks can produce
various rankings of hot cracking sensitivity depending on whether cracks
in weld metal, HAZ or distant cracks are considered. Looking only at weld
metal cracks the 59 type weld metals were least crack sensitive and the 625
welds most prone to cracking. Ranking in HAZ depends on whether num-
ber of cracks or crack length is used as criteria. However, including or ex-
cluding distant cracks has the most dramatic effect on ranking. The C-276
type weld metals were clearly most crack prone at a distance from the
GTA bead whereas the 625 were least crack sensitive in this region.
340 L. Karlsson et al.
Including all cracks thereby shifted the ranking suggesting the 625 or 59 type
weld metals to be most resistant and the C-276 type more likely to crack.
15,0
Varestraint specimen:
Varestraint specimen:
HAZ
HAZ 59-L
14,0 l l1 1
WM
WM
13,0 HAZ
59-H
HAZ
i =1
li , SG +
m
i =1
li ,WEZ
625-L
n m
+
l ges
tot =
tot
¦ liWM,SGWM ++ ¦ lHAZHAZ
i ,WEZ
11,0 i =1 i =1
625-H
10,0
276-L
Total crack length [mm]
9,0
276-H
8,0
7,0
3
6,0
2
5,0
1
4,0
3,0
2,0 Sector:
625 weld metals most prone to hot cracking. According to the classification
normally used when interpreting MVT diagrams the 625 type weld metals
would be considered restricted weldable and the 59 and C-276 types weld-
able or restricted weldable depending on strain rate and heat input [5].
40,0
59-L
35,0
59-H
625-L
30,0
625-H
25,0
276-L
No. of cracks
276-H
20,0
15,0
10,0
5,0
0,0
0 1 2 3 4 5
intermediate strain were the largest number of cracks were found in the
276-H weld metal.
A limited metallographic study of cross-sections of MVT specimens
revealed that most but far from all cracks were surface breaking. Although
statistics is limited there appeared to be a tendency that relatively fewer of
the hot cracks formed in C-276 type welds penetrated the top surface (Fig. 7).
This was also the only alloy type where quite a few rather large cracks
were seen at a distance from the GTA bead fusion boundary.
Fig. 7. Cross section of a C-276 type MVT specimen showing non-surface break-
ing cracks and a crack at some distance from the GTA re-melted bead. The broken
line outlines the MVT GTA weld bead
Discussion
Weld metals were ranked with respect to hot cracking susceptibility for
each test (Table 7). As noted earlier, evaluation based on metallography
and bend testing was in very good agreement ranking the 625 weld metals
Evaluating Hot Cracking Susceptibility of Ni-Base SAW Consumables 343
as best and C-276 weld metals as worst. All methods also ranked the lower
heat input welds as more crack resistant. As noted earlier these results
were in poor agreement with MVT test results.
Table 7. Comparison of hot cracking tendency ranking based on results from met-
allographic evaluation, bend testing and MVT testing. One is most cracking
resistant and six most prone to cracking
Test method Criteria Weld
59 59 625 625 276 276
L H L H L H
Solidification cracks 4 1 2 2 6 5
Liquation & ductility dip 3 5 1 2 4 6
Metallography
cracks
Total # of cracks 3 4 1 2 5 5
Side bend,
3 3 1 1 5 6
3-point
# and size (semi-
Side bend,
quantitative) of cracks 3 4 1 2 5 5
wrap around
Face bend 3 4 1 1 6 5
Av. ranking of metallographic and bend tests 3.2 3.5 1.2 1.7 5.2 5.3
# of cracks in GTA WM 3 1 5 6 4 2
# of cracks in GTA HAZ 1 2 3 6 3 5
# of cracks in GTA WM &
2 1 5 6 3 4
HAZ
# of “distant” cracks 5 2 3 1 4 6
Total # of cracks 4 1 2 5 2 6
MVT:
all strain rates Crack length in GTA WM 3 1 5 6 4 2
Crack length in GTA HAZ 1 2 3 6 3 5
Crack length in GTA WM
2 1 5 6 3 4
& HAZ
Crack length “distant”
5 2 3 1 4 6
cracks
Total crack length 4 1 2 5 2 6
# of cracks at 1% strain 1 1 4 3 5 n.d
MVT: # of cracks at 2% strain 1 2 4 5 3 6
Cracks in # of cracks at 4% strain 3 1 5 6 4 2
GTA WM Crack length at 1% strain 1 1 4 3 5 n.d
and HAZ Crack length at 2% strain 2 1 6 5 3 4
Crack length at 4% strain 4 1 3 5 6 2
Av. ranking MVT tests 2.6 1.3 3.9 4.7 3.6 4.2
n.d. = no data
344 L. Karlsson et al.
Evaluation of MVT test results did show more scatter in the ranking
compared to metallography of cross-sections or bend testing. However,
there was fairly good agreement between different criteria in that the 59
type weld metals were most crack resistant. On an average the C-276
welds were ranked as slightly better than the 625 type weld metals
although the difference was small and ranking varied significantly with
criteria. With the exception of the 59 welds, also MVT testing ranked
lower heat input specimens as less crack susceptible than those produced
with higher heat input.
It is notable than none of the criteria used to evaluate the MVT
specimens produced a similar ranking as metallography of cross-sections
or bend testing.
The MVT ranking of the alloy types is in good agreement with previous
results from MVT testing predicting Alloy 59 base material to be most
crack resistant and 625 as most likely to crack [6]. Weld metals of 625 and
C-276 type were ranked in the same order as parent material and higher
heat input during testing produced generally more and longer cracks but
did not change the ranking order. However, a more recent comparison
reported Alloy 59 parent material to be significantly more susceptible to
cracking than Alloy 625 at higher strains although the difference was very
small at 1% or 2% strain [7]. Obviously care should be taken when
comparing tests performed on different heats of the same material.
Alloy 59 and Alloy 625 type weld metals have good reputation for
resistance to hot cracking in practical applications although problems are
sometimes encountered [3]. Opinions about Alloy C-276 type filler metals
vary but also these are successfully used in numerous applications. Com-
paring practical experience to test results it appears that there is agreement
on 59 type weld metals being relatively crack resistant. On the other hand,
metallography, bend testing and practical experience suggests a reasonably
good hot cracking resistance whereas MVT ranks 625 type weld metals as
the least resistant of the three alloys tested.
A number of factors contribute most likely to discrepancies between
results of the different test methods and practical experience. These
include metallurgical aspects such as solidification, segregation and
precipitation behaviour of the alloys, tendencies to form non-surface
breaking cracks and differences in heat input between actual welding
procedures and MVT testing.
Evaluating Hot Cracking Susceptibility of Ni-Base SAW Consumables 345
Both Alloy 625 and C-276 are known to have rather complex
solidification behaviour [8, 9, 10]. Significant segregation occurs and
solidification is completed by precipitation of topologically close-packed
P-phase in Alloy C-276 whereas a eutectic-type reaction between the
austenitic matrix and various Nb-rich phases such as NbC and Laves phase
occurs in Alloy 625 [3, 11]. However, Alloy 59 is expected to solidify
largely as single-phase austenite with less segregation of alloying elements
and thereby less extension of the solidification range. Alloys 625 and C-276
are also significantly more prone to precipitation during heat treatment as
often seen in reheated regions of weldments [12].
A larger solidification temperature range of is contributing to a greater
risk of solidification cracking as is often observed when comparing 625
and 59 type weldments [3]. Back-filling can on the other hand contribute
to healing of solidification cracks thereby making the effect on cracking
tendency of real weldments smaller. This could partly explain the larger
cracking tendency in MVT testing of 625 welds compared to
metallography and bend testing. Strains are larger and solidification more
rapid during MVT testing, than in actual SAW applications, due to the
lower heat input. The strain rate will also be different during MVT testing
as compared to in actual welding. Another consequence of differences in
heat input is that slower cooling and larger weld pools lead to more
pronounced segregation. The segregated regions will be those where
liquation cracking occurs most easily due to a lower melting temperature.
The importance of heat input was confirmed by test results for all weld
metals, with one exception, ranking lower heat input welds as more crack
resistant (Table 7).
Apart from effects of cracking mechanisms the relatively greater
number of non-surface penetrating cracks in Alloy C-276 welds will
clearly confuse results in MVT testing. The reason is presently not
understood since the fraction of liquation cracks, identified from
metallographic cross-sections (Table 3 and Fig. 2), was equal or higher in
C-276 weld metals than for the other two types.
In conclusion there are indications that the MVT test weld pool size
and cooling rate provokes relatively more solidification cracking in
Alloy 625 compared to in submerged arc welding and thereby predicts
a higher cracking sensitivity than observed in practice. Quite contrary
significant subsurface cracking in C-276 weld metals renders MVT test
results misleading in the opposite direction. Although being a technique
capable of producing useful data from a small number of tests it is clear
that the MVT method should be complemented with other tests
simulating actual application welding conditions to permit a fair
assessment of welding consumables.
346 L. Karlsson et al.
Conclusions
Acknowledgement
References
Abstract
Fully austenitic materials, such as nickel base alloys and several stainless
steels, can be prone to hot cracking in the heat affected zone during
welding. Such cracks, due to their small size commonly denoted as micro-
cracks, were found in laser hybrid welds of a high nitrogen stainless steel
(UNS S 34565 / German material no. 1.4565).
Mechanical testing, such as tensile and Charpy V-notch tests as well as
cyclic loading, has been carried out on hot crack afflicted laser plasma
hybrid welds. The test results show that the ductility of the material is
decreased, whereas the strength is not influenced at all. SEM micrographs
of the fracture surfaces revealed that micro-cracks in the HAZ were
associated with failure.
For a thorough evaluation of the cracking mechanism, the orientation of
the micro-cracks alongside the fusion line of the hybrid welds is of great
importance. Thus, computer tomography using micro focus X-ray has been
carried out. Using this new approach, the orientation of the cracks towards
the fusion boundary was ascertained. It turned out that the critical strain
emerged at specific sites of the hybrid weld. The critical strain was
oriented tangentially to the fusion boundary and perpendicularly to the
welding direction. X-ray cone beam computer tomography turned out as a
very useful tool for the investigation of the crack distribution and hence,
for getting qualitative information about the critical stress/strain
distribution in real welds.
Introduction
Welding Experiments
Laser plasma hybrid welds of high nitrogen stainless steel plates of 6.6 mm
plate thickness were produced for the subsequent examinations. The
composition of the base metal UNS S 34565 (German material no. 1.4565)
and the filler powder are summarised in Table 1.
Hybrid welding has been carried out using a diode pumped Nd:YAG
laser. A scheme of the arrangement of laser beam and plasma torch can be
seen in Fig. 1. The parameters for the welds are summarised in Table 2.
S P N Nb others
UNS S 34565
<.001 0.016 0.466 0.010 Al: 0.024
6.6 mm plate
W: 0.34
filler powder 0.004 0.003 0.340 –
Cu: 1.01
Fig. 1. Arrangement of laser beam and plasma torch in the hybrid welding process
Assessment of HAZ Hot Cracking in a High Nitrogen Stainless Steel 353
Microscopic Examination
The cross sections of the welds are shown in Figs. 2a and 3a. The shape of
the hybrid weld is typically divided into two parts. The broader upper part
results from the plasma arc, whereas the lower part is tapered by the deep
penetration of the laser. The particular shape of the weld pool is influenced
by the applied welding parameters, such as arc current, focal point position
and plasma gas flow. In hybrid welding, these parameters can be varied
within a comparatively wide range, thus, a variety of weld pool geometries
can be produced.
Fig. 2. (a) Cross section of the laser plasma hybrid weld used for the tensile test
and cyclic tensile loading, (b) hot cracks in the HAZ of the base metal
354 K. Stelling et al.
Fig. 3. (a) Cross section of the laser plasma hybrid weld used for the Charpy test,
(b) hot cracks in the HAZ of the base metal
Mechanical Testing
Tensile Test
Fig. 4. Specimen geometry for the tensile test, loading direction was perpendicular
to the welding direction
Figure 5 shows the average values for the base metal specimens and the
weld metal specimens. The measured 0.2%-proof stress ranged from 400
to 450 MPa and the tensile strength ranged from 830 to 860 MPa,
respectively. There was no significant difference in the mechanical
strength between base metal and welded specimen.
In contrast, the ductility, represented by the reduction of area,
appreciably decreased from 65% in the base metal specimens to 40% in
both welded specimens. It has to be pointed out that the absolute value for
the ductility is still high. However, the welded specimens failed along the
fusion line and that would actually be a disqualifying criterion for quality
assurance.
356 K. Stelling et al.
Figure 6 shows the fractured weld metal tensile test specimen. From
Fig. 6b which shows the base metal side of the fracture, it is obvious that
the fracture surface has two different appearances. The upper part of the
fracture surface shows shiny sites which may indicate brittle fracture,
whereas the lower part shows ductile fracture. From the SEM micrographs
in Fig. 7, the different surfaces, i.e. the flat appearance of the shiny fracture
areas which might indicate portions of brittle fracture (a) and the dimple
structure of the ductile fracture (b), become apparent.
Fig. 5. Results of the tensile test, weld metal specimens fractured along the fusion
line
Fig. 6. (a) Fractured tensile specimen, (b) fracture surface of (a), base metal side
Assessment of HAZ Hot Cracking in a High Nitrogen Stainless Steel 357
Fig. 7. SEM micrographs of the surface in Fig. 9b, (a) fracture surface within the
upper part, (b) ductile (dimple) fracture in the lower part
Charpy Test
Fig. 8. (a) V-notch specimen according to DIN EN 10045 and (b) estimated
fracture line during testing at –120°C
Fig. 10. Survey of the fracture surface of the V-notch specimen which cracked
partially along the fusion line, on the left: weld metal side, on the right: base metal
side
and thus, the surface appearance of the cracks obviously depends on the
distance from the fusion zone.
Fig. 11. SEM pictures, section sites are marked in Fig. 10, weld metal side
Fig. 12. SEM pictures, section site is marked in Fig. 10, base metal side
Assessment of HAZ Hot Cracking in a High Nitrogen Stainless Steel 361
The above described hot cracks are surrounded by an area which has a
flat, transcrystalline appearance without pronounced dimple structure. This
surface structure supposedly results from a complex stress state which was
induced by the micro-cracks during the impact test.
Therefore, at least four different surface structures can be distinguished:
The first three of the different surface types are indicated in section no. 4,
Fig. 13. The fourth one was shown in detail of section no. 2, Fig. 11.
Fig. 13. SEM micrograph of section no. 4, demonstration of the different types of
fracture surfaces, section site is marked in Fig. 10, base metal side
The SEM examination revealed that the reduced ductility of the weld at
lower temperatures can be attributed to the micro-cracks along the fusion
line. Although ductile dimple fracture represents the major part of the
fracture surface, the micro-cracks were obviously involved in failure and
governed the fracture path during testing, especially at low temperatures.
362 K. Stelling et al.
The major part of the fracture surface was at a right angle to the
direction of the tensile stress which indicates fatigue failure. In Fig. 14b
and Fig. 15, the portion of fatigue fracture and of final catastrophic failure
characterised by strong plastic deformation can clearly be distinguished.
An examination with the naked eye revealed that cracking initiated on the
top side of the specimen surface close to the fusion line. The macroscopic
fracture propagation direction is indicated by directional markings on the
fracture surface.
Assessment of HAZ Hot Cracking in a High Nitrogen Stainless Steel 363
Fig. 14. Macrographs of the failed test specimen, (a) side view, position of the
weld is distinguished by means of a characteristic surface structure of the
specimen after extensive plastic deformation, (b) top view
Fig. 15. Directional markings on the fracture surface of the failed specimen
indicate the crack initiation and the propagation direction of the fracture.
observed on the fracture surface, i.e. they are oriented according to the
crystallographic planes of the underlying metal.
Resuming the results of the cyclic tensile loading tests it can be stated
that under extreme loading conditions, which have been applied to the
welded specimens, micro-hot cracks can lead to fatigue failure if they are
opened towards the surface.
Fig. 16. SEM micrographs showing the initiation site of the fracture at different
magnification, (a) survey, (b) surface of the intergranular micro-crack, (c) smooth
surface structure revealing slight slip line formation
Fig. 17. (a) transition area of the micro-crack and the transcrystalline fatigue
fracture surface, (b) transcrystalline fatigue fracture surface
of the cracks relative to the fusion line which would indicate the direction
of the critical strain during the welding process did not become clear either
from the SEM examination or from the two dimensional micrographs.
For that reason, a weld section was examined using three-dimensional
computer tomography. This technique is known to be a very useful tool for
revealing the distribution of imperfections inside a component.
The principle of three-dimensional or cone beam tomography (CT) is
shown in Fig. 18. The tomograph was developed at BAM in the
department for non-destructive testing.
During testing, the sample is turned in the cone beam of a micro focus
X-ray tube and manifold slices are measured using an area detector. With
the usage of magnification techniques, a spatial resolution of less than
10 μm could be reached depending on the object size.
In order to get highest resolution and quality of the CT images, the
specimen should have an almost cylindrical geometry and a low geometric
measure in the radiation direction.
For this reason, a small piece has been cut off from the weld. Then the
specimen was oriented such that the lowest specimen thickness was in
radiation direction, i.e. the turning axis of the specimen was perpendicular
to the welding direction as well as to the plate normal. For the analysis of
crack distribution, the cracks which can be identified in the tomography
due to their lower attenuation were manually marked in every single plane
of the CT in order to create a three-dimensional picture of the crack-
distribution.
366 K. Stelling et al.
3. The ratio of the fusion boundary area and the underlying base
metal volume is highest in the inwardly curved regions of the
weld. If shrinkage starts during cooling tangential strains are
imposed on these regions.
However, inwardly curved regions are found in the lower part of the
weld, i.e. the laser part, too. Nevertheless, hot cracks in these regions are
rarely found. Therefore, it must be a combination of increased local
shrinkage strains and adverse temperature fields, i.e. adverse metallurgical
conditions, causing the typical crack distribution.
Fig. 20. (a) Crack distribution along the weld flank in the upper weld part and
(b) cracks visualized by computer CT with the viewing direction perpendicular to
the fusion boundary; (c) schematic illustration of the direction of critical strain
towards the fusion boundary
368 K. Stelling et al.
Conclusions
1. The high nitrogen stainless steel UNS S 34565 is susceptible to hot
cracking in the heat affected zone of the base metal during laser plasma
hybrid welding.
2. Cracking typically occurred in the upper part of the weld produced by
the trailing plasma process during hybrid welding.
3. Transverse tensile testing and Charpy V-notch testing of welded
specimens indicate a decrease in ductility. But the strength of the
material is obviously not affected.
4. SEM surface examination of fractured specimens revealed that hot
cracks were involved in failure. The intergranular hot cracks show
different surface structures. Some grains exhibited an almost dendritic
surface, whereas others revealed a more even but still very smooth
surface. Thus, it is concluded that cracking occurred at different stages
of liquation and/or grain boundary ductility depending on the distance
from the fusion line.
5. By means of X-ray computer tomography it was found out that the
crack planes were located perpendicularly to the fusion boundary and
parallel to the welding direction. It could be shown in this contribution
that computer tomography is a very useful tool for the evaluation of
critical regions for hot cracking inside the material.
References
1. Brooks JA, Thompson AW, Williams JC (1980) Weld Cracking of Austenitic
Stainless Steels. In: R. Kossowsky (ed) Physical Metallurgy of Metal Joining,
Warrendale, Pennsylvania, pp 117–136
2. Berns H, Gavriljuk VG (1999) High Nitrogen Steels. Springer, Berlin
Heidelberg New York
3. Lin W, Lippold JC, Baeslack WA III (1993) An Evaluation of Heat-Affected
Zone Liquation Cracking Susceptibility, Part I: Development of a Method for
Quantification. Welding Research Supplement, April:135s–153s
4. Savage WF, Lundin CD (1965) The Varestraint Test. Welding Journal
34(10):433s–442s
5. Nissley NE, Collings MG, Guaytima G, Lippold JC (2002) Development of
the Strain-to-Fracture Test for Evaluating Ductility-Dip Cracking in
Austenitic Stainless Steels and Ni-Base Alloys. Welding in the World:32–40
6. Lundin CD, Lee CH, Menon R, Osorio V (1988) Weldability Evaluation of
Modified 316 and 347 Austenitic Stainless Steels: Part I – Preliminary
Results. Welding Research Supplement February:35s–46s
Assessment of HAZ Hot Cracking in a High Nitrogen Stainless Steel 369
1
Advanced Materials Analysis, Enschede, The Netherlands
2
Budapest University of Technology and Economics, Budapest, Hungary
Abstract
Hot cracks may appear in metal alloys on heating or on cooling when the
tensile strains and related stresses, caused by thermal expansion or
shrinkage and usually enhanced by various restraints, cannot be
compensated by a local plastic deformation of an alloy. In general, these
metal alloys, which have large thermal expansion coefficients, are
susceptible to hot cracking during casting, welding or e.g. re-heating for
hot rolling. The metallurgical quality of an alloy affects its susceptibility to
hot cracking, in particular the chemical and microstructural
inhomogeneities influence this susceptibility. Additional plastic strains,
applied in hot forming, usually extend the cracks leading to damage. In
this work an attempt is made to describe the micro- and macro-
mechanisms of the cracking as well as microstructural and mechanical
factors assisting the damage, based on observations of continuously cast
ingots in which internal hot cracks and cavities were formed during initial
steps of hot rolling. Characteristic of these ingots was chemical
segregation resulting in differences of hot ductility in different parts of the
ingots. Thus during plastic deformation, due to interaction between the
microstructurally and mechanically different regions, cracks and cavities
did nucleate and grow. Discussed here are physical data necessary to
adequately describe the behaviour of material during deformation and
cracking, as well as physical simulation methods to gain these data. The
gained data and identification of the damage processes are then used for
computer modelling with an aim to determine critical conditions of
controlling the application process in order to avoid the cracking and to
assure the manufacture of sound products. The micro-scale conditions
characteristic of e.g. welds and the macro-scale situations typical for
rolling of steel billets are discussed.
372 S.T. Mandziej, G. Krallics
Introduction
During hot rolling after continuous casting of steel various cracks may
occur resulting from the size and shape of the product as well as from
temperature, strain and strain rate of the applied processing. Thus during
rolling of 410 mm diameter continuously cast ingots into square 200 mm
cross section billets, periodic large internal cavities (bursts) were observed.
They often appeared in a medium-carbon Cr-Mo construction steel,
however never in any plain carbon steel of similar carbon content,
although the casting and rolling conditions were identical.
Fig. 1. Drawing stress (Vxf) as a function of reduction (r%), with indicated burst-
sensitive zone [1]
Crack Appearance in Hot Rolled Billets 373
Relative Reduction, ∈
Fig. 3. Schematic of rolling with marked area where the pulling force is created
due to friction [3]
Fig. 4. Cross-section of 200 mm billet with marked locations from which samples
were taken for hot tensile testing
causing rupture. Intragranular voids form, grow, and coalesce into internal
cavities, which result in a fracture with a finite reduction in area.
Fig. 7. Three stages of ductile fracture: (a) Void initiation at hard particles (b) Void
growth (c) Void linking
The material under consideration here in its initial form is a round ingot of
410 mm diameter made of 34CrMo4 grade steel, which has been
continuously cast in vertical direction with travel speed of 0.5 m/min at a
casting temperature 1390°C. Cutting off a 6 m long piece from the con-
cast ingot appeared after cooling to about 1050°C and transferring to re-
heating/soaking oven at inlet temperature of about 900°C. The exit
378 S.T. Mandziej, G. Krallics
temperature from this furnace to the rolling track was 1280–1290°C, and
rolling with the rate of about 5/s at the temperature of 1240°C into 200 mm
square billet. After this procedure, internal burst cavities were found by
NDT, and typical for these cavities was their periodicity in the length
direction; they were repeating after each 1.2–1.5 m.
To check for the susceptibility to liquation cracking/grain boundary
sensitization, strain-induced crack opening (SICO) tests were run on 10 mm
diameter samples after re-heating to nil strength temperature (NST). The
SICO test was developed by Dynamic Systems Inc. in USA [5], for
determining formability limits at the situations when high thermal
gradients occur. It comprises heating a rod-like sample mounted in “cold”
Gleeble’s jaws, so that in the middle of the sample’s span the uniformly
heated zone is formed. This central hot zone is then axially compressed to
form a bulge. The test has been dedicated to Gleeble physical simulators,
in which the use of high-rate electric resistance heating may cause a non
controlled overheat of anvil-work piece interface in a conventional
uniaxial compression test. The SICO procedure was in the first instance
applied here for determining the melting and solidification parameters. For
this 25 mm diameter bars were used and from them the melting (solidus)
temperature was found to be ~1390°C, the nil strength temperature
~1350°C and the nil ductility temperature ~1300°C. Then samples for the
10 mm SICO tests were taken from different locations on the cross section
of the 200 mm square billet:
Sample A - taken ~30 mm from the central burst cavity,
Sample B - taken ~15 mm from the central burst cavity,
Sample C - taken exactly at the central burst cavity.
These locations have been marked on the cross-section of the billet shown
in Fig. 4. The results obtained from SICO tests are given in Table 1.
Fig. 9. SICO sample A showing a small Fig. 10. SICO sample C2 showing a
crack near to perimeter of the bulge chain of small cracks along perimeter
Following the above SICO results, additional samples for hot tensile
testing were taken from the as continuously cast 410 mm diameter ingot,
from the locations adequate to these which showed different hot crack
sensitivity in the SICO tests. These samples were marked as given in
Table 2.
In the hot tensile tests the nil strength temperature (NST) of 1345°C was
taken as the maximum of the heating cycle and from it the samples were
cooled down to the pulling temperatures. Two samples were taken from
the center of the ingot, location (1) where a small shrinkage cavity as well
as visible porosity appeared with dendritic crystals inside. The mechanical
results from these samples are of course irrelevant nevertheless the
samples were useful for observation of fracture mode.
380 S.T. Mandziej, G. Krallics
The results of the hot tensile testing on the sound samples showed that
at the temperature of 1240°C, i.e. equivalent to the application conditions,
the material from location (3) deformed slightly slower in the crosswise
direction than this from location (2), while their yield strengths were
comparable. As the procedure of checking for susceptibility to liquation
cracking comprises heating-up a sample with fixed Gleeble’s jaws that due
to thermal expansion causes slight bulging and increase of sample’s
diameter in the central uniformly heated zone, it appeared on one sample
from the location (3), that it begun already cracking on heating in a brittle
manner, failing during subsequent pulling at very small extension (Hl) and
with “negative” reduction in area. This odd behaviour was caused by a
liquid phase, which already on heating under the compression due to
thermal expansion spread along the grain boundaries as a semi-continuous
film. Except for this, all other samples tested at 1200°C showed high
ductility, i.e. the RA of more than 85%. The case of brittle failure and then
the scatter of results indicate the influence of segregation and/or porosity
on the results of the hot tensile testing.
Metallographic Examination
In the first instance the central burst cavities were examined. As the
internal defects were found in the ingots, one might suspect that the
formation of the burst cavity might be related to the porosity in the middle
of the ingot, however it is not. The walls of the burst cavity appeared to be
Crack Appearance in Hot Rolled Billets 381
entirely of shear fracture type, Fig. 11, with numerous fibre-like “pillars”
breaking by tensile necking, Fig. 12.
Fig. 11. Shear-type separation surfaces Fig. 12. Fibre-like pillar with defined
dominating inside the burst cavity necking portion in the burst cavity
Fig. 13. Dendritic crystals in the cavity Fig. 14. Outer surface of the tensile
of hot tensile sample 1 sample 1 showing shear-slip lines
382 S.T. Mandziej, G. Krallics
Fig. 15. Shear-slip lines on the surface Fig. 16. A cross section through a
of the burst cavity in the 200 mm fibre-pillar protruding from wall of the
square billet burst cavity; nital etched - DF
Fig. 17. Coarse prior austenite grains Fig. 18. Finer prior austenite grains at
in the fibre-like pillar; nital etched – about 10 mm below the pillar’s root;
dark field image nital etched - DF
On the shear-slip type walls of the burst cavity evidences can be found
that during their formation hard particles existed which assisted the
deformation and separation process of the fibres and the walls, as well as
that some fibres had different ductility than the others. In the subsequent
Fig. 19 this situation is shown. After a slight electro-polishing and etching
of such shear-slip surface, the particles, which left the sharp shear traces on
the surfaces, can be revealed, in the size and amount adequate to explain
the appearance of “scratches” on the burst cavity walls, Fig. 20.
Crack Appearance in Hot Rolled Billets 383
Fig. 19. Shear-slip type of the burst Fig. 20. Fragment of polished and
cavity wall with sharp “scratches” etched surface of the burst cavity
caused by solid hard inclusion particles showing microstructure with inclusion
Fig. 23. A chain of AlN particles Fig. 24. Porosity near to dendrite
tracing primary dendrite boundary in boundary in as-cast microstructure of
the crack-sensitive zone (3) of the the ingot in zone (3); nital etched
ingot; nital etched
Discussion
then coalescence of these voids, not always can be applied without reserve.
The same remark may refer to the deformability maps.
Nevertheless, the cavitation during metal forming is still an existing
important phenomenon, which often causes premature failures of the
material during processing. It appears in a large number of steels, which
generate microscopic voids (or cavities) when subjected to large strains
under tensile loading. The formation of microscopic cavities often occurs
along the grain boundaries of primary i.e. as-cast structure during the
initial operations of high-temperature deformation. It is believed that the
cavitation may lead to premature failure at levels of deformation far less
than those at which the flow localization controlled failure would occur
[7]. For a given material, the extent of cavitation depends on the specific
deformation conditions like the amount of crosswise strain, strain rate and
temperature, as well as on the surface condition, which may affect the
friction between tool (rolls) and workpiece (billet) and this way the level
of tensile strains and dynamic stresses. An important requirement for
cavitation during flow under hot working is the presence of tensile stress.
On the other hand, under conditions of homogeneous compression,
cavitation is not observed. In fact, it is believed that the cavities, which
might be produced at the tensile flow situation, can be removed during
subsequent compressive flow. Due to this last belief, the metal-formers
often do not care too much about the fine porosities and cavities, which
may nucleate during the forming process, unless the coalescence of these
cavities leads to total failure of the product.
In the situation of billets containing the large central burst cavities, two
major factors should be highlighted. The first is the tensile stress/strain
situation that may occur on the entry of the ingot to the rolling stand and
which is assisted by the friction between the rolls and the oxidized surface of
the ingot. The oxide scale on the investigated Cr-Mo steel does not easily
strip-off as well as it limits the possibility of sliding between billet and rolls,
thus contributing to high tensile forces. The second factor is the segregation
of inclusions in particular this of AlN. For the purpose of de-oxidation and
grain refinement the investigated steel contained almost 300 ppm of Al and
100 ppm of N. Concentration of these and other inclusion particles in certain
regions of the ingot structure resulted in differences of hot deformability, in
particular in capacity of material to flow with different rates at the same
applied external stress. This phenomenon is consistent with the formation of
the large central burst cavities having shear-type walls.
In this chapter the whole complexity of the problem was neither
addressed nor solved. Our aim was mainly to describe the phenomena,
which assist the formation of the cracks – the central burst cavities. It
appeared that the differences in local ductility under dynamic deformation
386 S.T. Mandziej, G. Krallics
conditions could create the situation, which gets out of control. The hot
deformation tests: SICO and hot tensile testing, gave results showing the
differences in local material properties, while metallography revealed the
effects: micro-structure and cracks, as well as indicated the distribution of
particles which might assist the problem.
In result, measures were taken at steelworks to eliminate the problem, in
particular by reducing the amount of AlN in the cast as well as introducing
stirring to reduce segregation in the ingots.
Conclusions
References
Abstract
Introduction
The Ni-base superalloy, alloy 600 (Inconel 600), has been commonly used
in the reactor vessel, steam generator and pressuriser of nuclear power
plants (pressurised-water nuclear reactors). However, serious damage such
as the primary water stress corrosion cracking (PWSCC) has been found in
alloy 600 components [1]. Another Ni-base superalloy, alloy 690 (Inconel
690), is the material of choice for pressurised-water nuclear reactor
components because of its superior PWSCC resistance compared to alloy
600. It is planned that alloy 690 will replace the existing alloy 600 in ageing
nuclear power plants [2, 3]. In particular, the cladding of alloy 690 onto the
existing alloy 600 components and welds is one of the most practical
countermeasures for the preventive maintenance and/or repair of PWSCC
[4]. Recently, the laser cladding technique, especially when performed
underwater, has been developed and applied in practice for overlay welding
and seal welding of reactor components [5]. This technique possesses the
advantages of low heat input and reliable welding, compact welding
machine dimensions for welding in narrow spaces and a wide choice of
weld materials.
On the other hand, it has been pointed out that alloy 690 is highly
susceptible to hot cracking under heavy-restraint conditions such as the
welding of thick components [6]. Microcracking (microfissuring) in the
reheated weld metal during multipass welding can be of particular concern.
According to previous studies [7, 8, 9, 10, 11], microcracking in alloy 690
welds has been characterised as ductility-dip cracking. The authors have
reported that the primary cause of microcracking (ductility-dip cracking) in
alloy 690 was the reduction of hot ductility during cooling in welding
attributable to the grain boundary segregation of impurity elements such as
P and S [12, 13]. Furthermore, we have also proposed an effective method
for preventing microcracks in alloy 690 welds using La-containing alloy
690 filler metals, and confirmed that microcracks in multipass gas tungsten
arc (GTA) welds of alloy 690 could be completely prevented using the filler
metal with 0.01 mass% La added [14].
The objectives of the present study are to evaluate the microcracking
susceptibility in the multipass laser overlay weld metal of alloy 690, and to
Effect of La Additions on Micro-Cracking 391
clarify the effect of La addition to the filler metal with the objective of
improving microcracking susceptibility of the weld metal. We have
conducted microstructural analysis of the weld metal to clarify the
mechanism by which La addition improves microcracking behaviour.
Furthermore, in order to comprehend the occurrence of ductility-dip crack
during laser overlay welding, we have made theoretical approaches,
employing numerical analysis, to model the behaviours such as the grain
boundary segregation of P, S and the development of thermal strain in laser
overlay welds.
Materials
The base metal used in this study to simulate alloy 600 weld metal was alloy
132 (Inconel 132, DNiCrFe-1J) overlay weld metal (shielded metal arc
welding: SMAW, 4 layers), surface-ground to approx. 10 mm thick, on a 70
mm thick carbon steel, with the dimensions shown in Fig. 1. Five kinds of
filler metals of alloy 690/alloy 52 (Inconel 52, ERNiCrFe-7) in which the La
content varied between 0 and 0.07 mass% were used. The commercial filler
metal NF0 (La-free) was used as a standard of comparison for
microcracking susceptibility. The chemical compositions of the base metal
and filler metals are shown in Tables 1 and 2, respectively. The diameters of
the filler metals are 1.2 mm.
15
Carbon steel
60mm
Table 1. Chemical composition of base metal alloy 132 (DNiCrFe-1J) used (mass%)
C Si Mn Ni Cr Cu Fe P* S* Nb+Ta
DNiCrFe-1J
(Alloy 132) 0.04 0.21 2.80 Bal. 14.43 0.02 8.88 40* 30* 1.86
* : ppm
Experimental Procedures
Figure 2 illustrates the set-ups for multipass laser overlay welding (3 layers,
10-9-8 passes) of alloy 690 on alloy 132 was carried out in air and under
water using a YAG laser, and the conditions are summarised in Table 3. The
Optical fiber
Specimen
Jig
Controller
Numerical control
welding conditions were varied within the following ranges: laser power of
1.0–1.2 kW, laser travelling speed of 5.0–11.7 mm/s, wire feeding speed of
10.0–18.3 mm/s and shield gas flow rates (Ar) of 20 L/min for in-air
welding and 60 L/min for underwater welding. The microcracking
susceptibility was evaluated as the total crack length in the cross-section of
the overlay weld metal. The weld metal was sectioned and prepared for
optical microscope and scanning electron microscope (SEM) observation by
grinding and then polishing with #400–#1500 grit emery paper. The final
polishing was performed with 1 μm alumina paste. Electrolytic etching was
performed with a 10% oxalic acid ethanol solution at 1.5–3.0 V for 25–30 s.
The element distribution in the weld metal was analysed by EPMA.
Characterisation of Microcracks
The cross-sectional views of the laser overlay weld metal with varying La
content, using Condition A in Table 3, are shown in Fig. 3. In the case of
using the La-free filler metal (NF0), a few microcracks occur in the weld
passes of the second and third weld layers where they were reheated by the
subsequent weld beads (designated as “under-bead cracking”). Several large
microcracks can be seen in the weld metal (NF4) containing the high
proportion of La (0.07 mass%), mainly in the final weld layer. However, no
microcracks can be observed in the weld metal in the intermediate case
using 0.01 mass% La added to the filler metal (NF1). Figure 4 shows the
microstructures of microcracks in the weld metals of NF0 and NF4. A
microcrack in the La-free weld metal (NF0) occurs at a small distance from
the fusion line and is propagated at the austenitic grain boundary (G.B.).
From the fact that the crack surface indicates smooth and intergranular
fracture without any trace of melting, the majority of microcracks in the
La-free weld metal can be characterised as being ductility-dip cracks. On
the other hand, a microcrack in the weld metal containing 0.07 mass% La
(NF4) originates at the fusion line and is propagated at the solidification
394 K. Nishimoto et al.
1mm
Crack
1mm
1mm
Crack
Filler metal : NF4 (0.07mass%La)
Fusion line
Filler metal : NF0
High temp.
(La-free)
Low temp.
20μm 100μm 5μm
G.B.
High temp.
Filler metal : NF4
(0.07mass%La)
Fusion line
Low temp.
12
Total clack length (mm)
Condition A (underwater)
10
Condition B (underwater)
Condition C (underwater)
8
Condition D (underwater)
Condition E (underwater)
6
Condition C (in-air)
4
2
(0)(0)(0) (0)(0)(0)(0)(0)(0) (0)(0)(0)(0)(0)(0) (0)(0)(0)(0) (0)(0)
0
NF0 NF1 NF2 NF3 NF4
(La-free) (0.01mass%La) (0.03mass%La) (0.04mass%La) (0.07mass%La)
Filler metal
Fig. 5. Effects of La-addition and welding conditions on microcracking susceptibility in
laser overlay weld metal
2μm
La P
S O
2μm
La P
S Ni
conducted under the condition that the overlay weld metal was subjected to
welding thermal cycles generated by two successive overlay weld passes.
As described above, ductility-dip cracks in the alloy 690 overlay weld metal
would propagate preferentially along austenitic grain boundaries which
were coincident with solidification boundaries (dendrite boundaries) [12].
As a result of this conclusion, the theoretical model for microsegregation
used in the present simulation considered solidification segregation as
well as grain boundary segregation. The calculation procedures for
microsegregation are similar to those used previously for multipass GTA
welding [13, 16, 17]. Specially, segregation is assumed to happen during the
solidification stage of welding; subsequent segregation during the following
cooling/reheating stage occurs under the initial condition of non-uniform
distribution of segregating species formed during solidification segregation.
In order to simplify the computation procedure, segregation behaviour of P
and S to a grain boundary was calculated for pseudo-binary systems of
(Ni-30Cr-10Fe)-P and (Ni-30Cr-10Fe)-S using the finite differential
method (FDM) scheme. Furthermore, the co-segregation effect of P with S
was not considered in the present simulation. The model analysed here uses
one-dimensional diffusion in a rectangular triangle assuming that the
morphology of a dendrite cell is basically a hexagonal prism. Further details
of the theoretical models used in the present study can be found in Refs. [13,
16, 17]. The conditions used in the calculations are summarised in Table 4,
i.e., the initial P and S contents in the weld metal are 50 ppm and 16 ppm,
respectively, and the cooling and reheating rates were varied between four
P 0.0050
Initial concentration
(mass%) S 0.0016
GTA welding 30(20) / 50
Laser overlay 60(40) / 100
Cooling rate / welding-1
reheating rate Laser overlay
(K/s) 150(100) / 250
welding-2
Laser overlay 300(200) / 500
welding-3
( ) : Cooling rate between liquidus and solidus
400 K. Nishimoto et al.
levels in order to simulate laser overlay welding and GTA welding. Material
constants used for the numerical simulation are also given in Ref. [13].
The computer simulation was carried out with the peak temperature in both
the second and third cycles being 1500 K. Figure 8 shows the relation
between the elapsed time in welding thermal cycles and the calculated P and
S concentrations at a grain boundary for cooling/reheating rates of 150/250
K/s. In the solidification process, P and S concentrations in the rest liquid
phase increase with the progress of solidification, resulting in remarkably
high amount of P and S bring segregated to a grain boundary at the
completion of solidification. However, P and S concentrations are
drastically reduced to their equilibrium concentrations when the weld metal
cools only by approx. 200 K. After that, they increase again with further
cooling (drop in temperature) from approx. 1400 K to 700 K, and they
saturate at the constant level below 700 K. In the reheating process, P and S
concentrations at a grain boundary keep constant temporarily or increase
slightly with a rise in temperature. However, they decline rapidly at
temperatures over approx. 1100 K. They once more increase again during
the cooling stage in the reheating process, and then change cyclically
thereafter in synch with the thermal cycle. The effects of different cooling
1600
2.5 P
1400
Temperature (K)
2.0 1200
1000
1.5
800
1.0 600
400
0.5
200
0 0
0 5 10 15 20 25 30 35
Time (s)
Calculated S concentration at
0.35 0.0405
0.34 0.0400
0.33 0.0395
0.32 0.0390
0.31 0.0385
0.30 0.0380
GTAW LOW-1 LOW-2 LOW-3
C.R. : 30K/s C.R. : 60K/s C.R. : 150K/s C.R. : 300K/s
H.R. : 50K/s H.R. : 100K/s H.R. : 250K/s H.R. : 500K/s
Model Details
The high temperature stress and strain in multipass welds were analysed by
two-dimensional thermo elasto-plastic FEM (calculation code: Quick
Therm). The model used for analysis of multipass laser overlay welds (3
layers, 6 passes) on alloy 690 is illustrated in Fig. 10. The deposited metal of
each weld pass modelled in the present analysis has the rectangular shape
with 1873 K (estimated temperature of the molten metal) dimensioned as 2
mm wide and 0.5 mm thick as illustrated in figures; the penetration of each
weld pass is consequently determined by the heat conduction from this
rectangular heat source. The position used for the analysis of stress/strain in
the multipass weld metal is also indicated in the bottom figure; it is the
reheated location in the first weld pass adjacent to the fusion boundaries of
the subsequent weld passes. The temperature dependencies of the various
material constants of the alloy 690 employed in the FEM analysis are
summarised in Fig. 11. The solidification shrinkage of the molten metal was
introduced as a volume change of 3%, and the melting temperature of alloy
690 was set at 1660 K. The interpass temperature used in multipass welding
used for the FEM analysis is 473 K.
x Alloy 690
10mm
2mm
0.5mm
Pass-4 Pass-5
Fig. 10. Analysed models for stress and strain in multipass laser overlay welds
Effect of La Additions on Micro-Cracking 403
8
Specific heat (×10-3J/g•K)
7
Physical constants
5
Poisson's ratio (×10-1)
4
3
Thermal conductivity
2 (×10J/m•s•K)
The calculated results of the plastic strain at the analysed position in the first
weld bead perpendicular to the weld line/crack during laser overlay welding
404 K. Nishimoto et al.
2273
Temperature (K)
1773
1273
773
273
20
Stress, σx (MPa)
10
-10
-20
0.02
0.01
Total strain
0
-0.01
-0.02
-0.03
-0.04
-0.05
0 50 100 150 200 250 300 350 400
Elapsed time (s)
Fig. 12. Change in temperature, crack opening stress Vx and total strain during
multipass laser overlay welding
is shown in Fig. 13. In the present study, the plastic strain was expressed as
the increment in plastic strain [18] from the moment of solidification
(approx. 1660 K) of the first weld pass. A solid line indicates the increment
in plastic strain under the condition that the crack opening stress Vx is tensile
(corresponding nearly to the cooling stage) and a dashed line indicates that
the crack opening stress Vx is compressive (corresponding nearly to the
reheating stage). The plastic strain increases monotonically with cooling in
the first weld pass. However, in subsequent welds, once in the reheating
stage, it decreases, and then increases again in the cooling stage. It follows
that the plastic strain in the weld metal accumulates with increasing number
of weld passes. The relationship between the plastic strain-temperature
curve and the DTR, as a function of the La content in the weld metal are also
shown in Fig. 13. In the present study, the DTRs of the laser overlay weld
Effect of La Additions on Micro-Cracking 405
BTR
0.7
Base metal : Alloy 690
0.6 Filler metal : NF0,NF1
DTR Laser overlay welding
Plastic strain (%)
(0.01mass%La)
0.5
DTR
(0.003mass%La)
0.4
DTR
0.3 (La-free) ng
Reheati ass
ng in in 4th p
Cooli ass
0.2 4th p
Cracking
s
in d pas
0.1 ling
Coo pass ating
in 2n 3rd pass
2nd Rehe 1 s tp a s s
Coolin
g in Crack-free
0
1673 1473 1273
Temperature (K)
Fig. 13. Relation between plastic strain-temperature curve and ductility-dip
temperature ranges (DTRs) in multipass laser overlay welding
Conclusions
affect grain boundary segregation of P and S for the range of conditions used
here.
(6) The occurrence of ductility-dip cracking in the multipass alloy 690
laser overlay weld metal was predicted by simulation using a mechanical
approach. The plastic strain-temperature curve in the La-free weld metal did
not cross the DTR in the first weld pass, however, it intersected slightly the
DTR during the cooling portion of subsequent welds. On the other hand, for
the weld metals containing La, none of the plastic strain-temperature curves
of any weld passes crossed the DTR. The laser overlay welding test revealed
that ductility-dip cracking would occur in the overlay weld metal when the
plastic strain-temperature curve intersected the DTR.
References
1. B.P. Miglin and G.J. Theus: “Stress Corrosion Cracking of Alloy 600 and 690
in All Volatile Treated Water at Elevated Temperature”, EPRI Report
NP-5761M, Electric Power Research Institute, Polo Alto, USA, (1988).
2. S.D. Strauss: “Inconel 690 is Alloy of Choice for Steam-Generator Tubing”,
Power, 140-2 (1996), p. 29–30.
3. J. Miyaguchi, K. Tabuchi and T. Yamamoto: “Towards Maintenance Service
Supporting Secure Nuclear Energy”, Mitsubishi Juko Giho, 40-2 (2003),
p. 92–95 (in Japanese).
4. Y. Nakamura, K. Dambayashi, H. Sakamoto, Y. Fujiya and Y. Tsukamoto:
“Field Application of the Cladding of PWR Reactor Vessel Outlet Nozzle”,
Mitsubishi Juko Giho, 43-1 (2006), p. 4–5 (in Japanese).
5. Y. Kanazawa and M. Tamura: “Underwater YAG Laser Welding Technique”,
Toshiba Review, 60-10 (2005), p. 36–39 (in Japanese).
6. W. Wu and C.H. Tsai: “Hot Cracking Susceptibility of Fillers 52 and 82 in
Alloy 690 Welding”, Metall. Mater. Trans., 30A-2, (1999), p. 417–426.
7. V.R. Davé, M.J. Cola, M. Kumar, A.J. Schwartz and G.N.A. Hussen: “Grain
Boundary Character in Alloy 690 and Ductility-Dip Cracking Susceptibility”,
W.J., 83-1, (2004), p. 1s–5s.
8. N.E. Nissley and J.C. Lippold: “Development of the Strain-to-Fracture Test”,
W.J., 82-12, (2003), p. 355s–364s.
9. M.G. Collins and J.C. Lippold: “An Investigation of Ductility Dip Cracking in
Nickel-Based Filler Materials - Part I”, W.J., 82-10, (2003), p. 288s–295s.
10. M.G. Collins, A.J. Ramirez and J.C. Lippold: “An Investigation of Ductility
Dip Cracking in Nickel-Based Filler Materials - Part II”, W.J., 82-12, (2003),
p. 348s–354s.
11. M.G. Collins, A.J. Ramirez and J.C. Lippold: “An Investigation of Ductility
Dip Cracking in Nickel-Based Filler Materials - Part III”, W.J., 83-2, (2004),
p. 39s–49s.
408 K. Nishimoto et al.
Abstract
Introduction
DDC is unclear, although large grain size and long, straight grain
“migrated” grain boundaries in the weld metal have been shown to
increase susceptibility. The effect of filler metal composition and the
related nature of grain boundary precipitation phenomena are key elements
in the understanding of DDC, as recently described by Ramirez and
Lippold [1, 2]. The strain-to-fracture (STF) test has proved to be an
efficient and effective method for quantifying susceptibility to DDC [3]
and has been used here to compare a number of high-Cr, Ni-base filler
metals. This chapters is a continuation of the research that has been
reported previously at IIW annual assemblies [4, 5]. Readers are referred
to those chapter for the results of preliminary studies of the DDC
susceptibility of FM52 and FM52M. This chapter provides additional STF
data on other high-Cr, Ni-base filler metals and compares this data to the
FM52 and FM52M results reported previously.
Filler metals conforming to AWS A5.14, ERNiCrFe-7 (FM52, Sanicro
68HP) and ERNiCrFe-7A (FM52M) are frequently used in nuclear power
plant applications, often for welding Ni-base Alloy 690. These filler metals
have good resistance to stress-corrosion cracking and are used as a
replacement for lower-Cr, FM82 in environments where stress corrosion
cracking is a concern, such as in high-purity, primary water in nuclear
power plants. DDC has been observed when using these filler metals,
particularly in multipass welds made under medium to high restraint
conditions.
Experimental Procedures
Weld deposits of eight (8) different high-Cr filler metals were tested using
the strain-to-fracture (STF) test. This test was developed at The Ohio State
University for the evaluation of elevated temperature grain boundary
cracking in austenitic materials. Details of the test development and
procedures can be found in Ref. [3]. The compositions of the filler metals
tested in this investigation are provided in Table 1.
The FM52 and Sanicro 68HP filler metals conform to AWS
ERNiCrFe-7. The three FM52M filler metals were produced to conform to
ERNiCrFe-7A with Nb in the range from 0.5–1.0 wt%. Two Ni-30Cr
variants containing elemental variations and Nb and Mo (FM52X-D and
FM52X-H), and Sanicro 69HP with nominally 1.8 wt% Nb were also
tested.
Ductility-Dip Cracking in High Chromium, Ni-Base Filler Metals 411
The testing procedure used for the eight filler metals is essentially
identical to that used previously for FM52 and FM52M [5]. These
represent the FM52 and FM52M-A filler metals in Table 1. STF samples
were machined from multipass welds. An autogenous, gas tungsten arc
spot weld was made in the gage section of these samples under controlled
conditions to produce a radial pattern of grain boundaries in the spot weld.
Full temperature-strain plots were produced for the FM52, 52M, and
52X-H filler metals, and the Sanicro 68HP and 69HP filler metals. Filler
metals described as B, C, and D were only tested at 950qC. This
temperature was selected since it is normally the temperature that
represents the minimum threshold strain for cracking across the entire
DDC range and where the highest degree of cracking (number of cracks)
occurs as strain is increased in the range from 0–10%.
Temperature-Strain Curves
STF test results for FM52 and FM52M that have been published
previously [5] are shown in Fig. 1. Each symbol represents a single test
with an “X” for a strain-temperature condition where cracking was not
observed and a circular symbol for samples where cracking occurred. The
number adjacent to each point represents the number of cracks that were
observed on the surface of the sample under a binocular microscope using
the procedure described previously. A curve separates samples that
cracked from those that did not and represents the minimum strain for
cracking as a function of temperature. The threshold strain (minimum
strain for cracking in the ductility-dip temperature range) was similar in
both materials, approximately 2% at 950°C. The temperature range where
the threshold strain occurred (900–1000qC) is consistent with other STF
testing, and the threshold strain is similar to results on the same heat of
IN52 previously reported by Collins et al. [6]. As a reference, the STF
curve previously developed by Collins et al. for Filler Metal 82 is
included.
Results of STF tests for Sanicro 68HP and 69HP are provided in Fig. 2.
These filler metals exhibit a slightly higher threshold strain (2.5%) than
FM52 and 52M and have less severe cracking at higher strains at 950qC.
No increase in the minimum strain for cracking could be identified at the
high temperature end of the DDC range (>1150qC), possibly due to the
overlap of the DDC and liquation cracking range.
STF results for the experimental 52X-H filler metal are presented in
Fig. 3. The threshold strain for cracking for this filler metal is considerably
higher than for the filler metals shown in Figs. 1 and 2. In the temperature
range from 750 to 950qC, 52X-H is extremely resistant to DDC as
indicated by both a high threshold strain for cracking and the small number
of cracks that occur above threshold.
In Fig. 4, The STF results for 52X-H are compared to STF data pre-
viously generated for Type 304L and Filler Metal 82 weld metals [6, 7, 8].
This shows that the DDC resistance of 52X-H is superior to FM-82 over
Ductility-Dip Cracking in High Chromium, Ni-Base Filler Metals 413
Development of a full STF curve for a single filler metal requires 30–40
samples and may not be appropriate for preliminary screening of filler
metal compositions. In this investigation, only filler metals 52, 52M-A,
52X-H, and Sanicro 68HP and 69HP were tested over the full DDC
temperature range. For filler metals 52M-B, 52M-C, and 52X-D, testing
was only conducted at 950qC. This temperature was selected because the
minimum in DDC behavior usually occurs at, or near, 950qC and that
cracking response at this temperature correlates well with actual
fabrication behavior.
A comparison of the eight filler metals at 950qC is provided in Fig. 5.
The horizontal bar in the column above each filler metal represents the
threshold strain for cracking for that material while the numbers represent
the number of cracks that were present in the sample as a function of
strain. All the FM52 and FM52M weld metals exhibit a threshold strain in
the range from 1–2%, while the Sanicro weld metals are at approximately
2.5%. There is a considerable difference in cracking susceptibility, based
on number of cracks at applied strains above 2%, for these filler metals.
As reported previously [5], FM52 and FM52M-A exhibit extensive
cracking above the threshold strain. FM52M-C exhibits similar behavior,
but FM52M-B does not show extensive cracking until the strain exceeds
6%. The Sanicro 68HP and 69HP weld metals do not show extensive
cracking until strains reach almost 8%.
The experimental weld metals containing Nb levels above 1 wt% and
4 wt% Mo exhibit both the highest threshold strains and the best resistance
to cracking above the threshold. FM52X-D has a threshold strain of 3%,
while that for FM52X-H is approximately 6%. It should be noted that all
the other filler metals showed extensive cracking above 6% strain, but the
experimental filler metals were very resistant to DDC at these strain levels.
414
16%
14%
12%
FM-82
>100 >100
10%
J.C. Lippold, N.E. Nissley
>100 >100
8%
25 5
Strain (%)
>50 >50
6%
9 2
19 14 17 4
2 1 2
4% 3
FM-52M 7 5
2%
FM-52
0%
600 700 800 900 1000 1100 1200 1300
Temperature (°C)
Fig. 1. STF results for FM52 and FM52M. Numbers to the left of the open symbols are the number of cracks for the FM52 weld
metal, while the numbers to the right are for FM52M. A curve for FM-82 is included for comparison
16.0
14.0
12.0
FM-82
10.0
69 HP 35 31
8.0
>50 >25
>25 >25
6.0 15 4
Calculated Strain %
7 5 >50
8 7 5 4 2
4.0 9 8
5 2 1 1 23 2 2 1
3 10 4
1 1
2.0
68 HP
0.0
600 700 800 900 1000 1100 1200 1300
Temperature (°C)
Fig. 2. STF results for Sanicro 68HP and 69HP filler metals. Numbers to the left of the open symbols are the number of cracks for
Ductility-Dip Cracking in High Chromium, Ni-Base Filler Metals
the 68HP weld metal, while the numbers to the right are for 69HP
415
416 J.C. Lippold, N.E. Nissley
18.00
14.00
(14)
12.00
Strain (%)
(2)
10.00
(0.5)
(0) (0.5)
4.00 (0.5)
(0)
2.00
0.00
600 700 800 900 1000 1100 1200
Temperature (°C)
Fig. 3. STF results for Filler Metal 52X-H. No data was collected above 1050qC,
so a hypothetical extension of the curve is indicated by the dashed line
18.00
16.00 (0)
14.00
52X-H (14)
12.00
Type 304 SS, FN6
Strain (%)
(2)
10.00
(0.5)
(0) (0.5)
4.00 (0.5)
(0)
2.00
0.00
600 700 800 900 1000 1100 1200
Temperature (°C)
Fig. 4. Comparison of STF results for Filler Metal 52X-H with those of Type
304L (Ferrite Number 6) and Filler Metal 82 weld metal
Ductility-Dip Cracking in High Chromium, Ni-Base Filler Metals 417
14
Strain-to-Fracture Test 14
12 Results at 950qC
10
Applied strain (%)
35 20
8 >100 >100
11
6 32 2
0 15 4
1 7 5
19 14 3
4
42 2 9 8
2 0
7 5
2 1 24 0 0 0
0 0
0 1
0
IN52 IN52M-A IN52M-B IN52M-C IN52X-D IN52X-H San-68HP San-69HP
Fig. 5. Strain-to-fracture test results for high-Cr, Ni-base weld metals tested at
950qC. Horizontal bars represent the threshold strain for cracking. The numbers
represent the individual cracks counted on the sample surface
Fig. 6. Weld metal microstructure in a FM52 STF sample tested at 950qC and
8.5% strain. Ductility dip cracks are evident along the weld metal migrated grain
boundaries
M23C6
M(C,N)
Fig. 10. Electron backscattered diffraction (EBSD) patterns showing the nature of
the weld metal migrated grain boundaries in FM52X-H (left) and FM52M-C
(right)
Ductility-Dip Cracking in High Chromium, Ni-Base Filler Metals 421
The behavior of the high-Cr, Ni-base filler metals again shows the
importance of the nature and morphology of precipitates on resistance to
DDC. Nb additions above the level prescribed in ERNiCrFe-7A (0.5–1.0
wt%) lead to the formation of a higher fraction of Nb-rich M(C,N) at the
end of weld solidification. Because these precipitates are present at the end
of solidification they are very effective in pinning the migrated grain
boundaries that form from the original solidification grain boundaries. The
fraction of these solidification- related precipitates is apparently quite
important, since when the fraction is low, such as with FM52M-A and
FM52M-C, grain boundary pinning is minimal and DDC susceptibility is
high. Increasing Nb content above 1 wt% has a profound effect on DDC
resistance as evidenced by the behavior of FM52X-D, FM52X-H, and
Sancro 69HP reported here and previous data for FM82 [6]. This behavior
is consistent with the model proposed by Ramirez and Lippold [2] and
shown in Fig. 11.
MC Macroscopic GB Locking
Microscopic
GB Locking
M23C6
Fig. 12. Schematic showing the combination of grain boundary tortuosity that
provides macroscopic grain boundary locking (resistance to sliding) and
microscopic locking by the formation of grain boundary precipitates. This
combination appears to provide the best resistance to ductility dip cracking
Conclusions
Acknowledgements
The authors wish to thank the Special Metals Welding Products Company
for providing the FM52, FM52M, and FM52X welded samples, and
Sandvik AB for providing the Sanicro 68HP and 69HP consumables. Our
thanks to Kenny Izor, an undergraduate research assistant at OSU, for
performing some of the STF testing. We also wish to thank Mr. Sam Kiser
of Specials Metals and Mr. Claes-Ove Pettersson from Sandvik Materials
Technology, R&D for useful discussions and insight.
References
1. A.J. Ramirez and J.C. Lippold, 2004. High Temperature Behavior of Ni-base
Weld Metal, Part I, Ductility and microstructural characterization, Materials
Science and Engineering A, 380, p. 259–271.
2. A.J Ramirez and J.C. Lippold, 2004. High Temperature Behavior of Ni-base
Weld Metal. Part II, Insight into the mechanism for ductility dip cracking,
Materials Science and Engineering A, 380, p. 245–258.
3. N.E. Nissley and J.C. Lippold, 2003. Development of the strain-to-fracture
test for evaluating ductility-dip cracking in austenitic alloys, Welding Journal,
82(12):355s–364s.
4. J.C. Lippold, N.E. Nissley, and A.J. Ramirez, 2005. Recent advances in the
understanding of ductility-dip cracking, IIW-IX-2161-05, Prague, July 11–14,
2005.
5. J.C. Lippold and N.E. Nissley, 2006. Further investigations of ductility-dip
cracking in high chromium, Ni-base Filler Metals, IIW-IIC-326-06, Quebec
City, August 27–31.
Ductility-Dip Cracking in High Chromium, Ni-Base Filler Metals 425
Abstract
Introduction
plagues different alloy systems that present a FCC structure in these inter-
mediate temperatures, as Ni-alloys, Cu-alloys, steels and stainless steels
and has been the object of research for quite a long time [1, 2, 3, 4, 5].
Among the factors that have been reported to have influence on the DDC
phenomenon are: alloy chemical composition (with special influence of
impurities and interstitials), segregation to grain boundaries, precipitation
behavior, grain boundary migration and pinning, grain boundary orienta-
tion relative to the applied strain and recrystallization [6-12]. Depending
on the system, some of these factors become more important and eventu-
ally control the cracking mechanism. Therefore, as a result of this appar-
ently changing mechanism the technical and scientific literature on DDC
may become quite confusing and controversy involving these different
controlling factors has appeared. For example, the dominant factor on the
DDC of Ni-base alloys with moderate to high levels of impurities has
been proposed to be impurity segregation to the gain boundaries [11]. On
the other hand, the DDC of very “clean alloys’, as most commercial Ni-
base alloys, can not be directly linked to intergranular segregation of im-
purities [10]. However, S segregation has also been observed in these
“clean” alloys [13].
Based on microstructural analysis up to the nanoscale and the exhaus-
tive study of DDC behavior of as-welded Ni-base filler metals AWS A5.14
ERNiCrFe-7 (FM-52) and ERNiCr-3 (FM-82), it was verified that the dif-
ference on DDC resistance between these two alloys is due to the large
amount of primary intergranular precipitates found in ERNiCr-3
(NiCr20Mn3Nb), which is a near-matching filler metal for alloy 600 [8-
10]. These primary precipitates in as-welded ERNiCr-3 are (NbTi)(CN)
cabonitrides and their effect on DDC resistance is based on the capability
of such eutectic-like precipitates to preclude gain boundary (GB) migration
at the very end of solidification. The GB pinning results in tortuous GBs
that would make GB sliding more difficult and strain concentration at tri-
ple points less likely to happen, resulting in less severe void formation,
which improves DDC resistance [6, 8-10]. Figure 1 presents schematically
the proposed effect of GB tortuosity on GB sliding, strain concentration at
triple points and intergranular precipitates, and void formation.
In order to verify the Nb rich carbonitrides role on the DDC resistance
of ERNiCr-3, welds were prepared using the DDC susceptible filler metal
ERNiCrFe-7 with additions of Nb and Ti to induce a larger amount of pri-
mary precipitation. The remarkable DDC resistance improvement of these
modified ERNiCrFe-7 welds was verified using the strain-to-fracture test
Ductility-Dip Cracking Resistance Improvement of ERNiCrFe-7 429
a.
b.
c.
Actual weld metal samples have also been characterized and compared
with the simulations. Among the obtained results can be highlighted the
kinetic phase diagrams where the microstructure stability during cooling
after welding is related to Ti- and Nb-content. Atomic partitioning between
the austenitic matrix and the precipitates was also evaluated.
a. b.
c.
Fig. 2. Comparison of Strain-to-fracture results on unmodified (circles) and modi-
fied (inverted triangles) ERNiCrFe-7 alloy with small (a) Ti, (b) Nb, and (c) Nb+Ti
additions [15]
432 A.J. Ramirez, C.M. Garzón
Materials
The nominal chemical composition of the studied alloy AWS A5.14 ER-
NiCrFe-7 used as a master alloy is presented in Table 1. The base alloy
was in the form of welding filler metal wire of 1.2 mm diameter.
Samples were obtained by arc melting using inert gas atmosphere the
base alloy with additions of high purity Ti and Nb powder. Melted buttons
were cold forged at 25°C using a uniaxial hydraulic press to obtain flat-
like shapes of proximately 15 mm diameter and 3 mm thick. As-welded
microstructure with a radial distribution of columnar grains was obtained
by autogenously spot welding on top of forged samples. 5 mm diameter
spot welds were performed with manual GTAW using 115 A (ampere) for
3–5 s. Table 2 shows the targeted and measured chemical composition of
the melted samples, measured using energy-dispersive X-ray spectroscopy
(XEDS) associated with the scanning electron microscope (SEM).
Table 1. Chemical composition of welding filler metal wire AWS A5.14 ER-
NiCrFe-7 used as master alloy (wt.%)
Fe Cr Al Ti Mn Si C N S P Nb+Ta Ni
10.1 29.1 0.71 0.51 0.25 0.13 0.027* 0.015* <0.001 0.003 <0.01 Bal.
Notes: * N-content was between 0.008 and 0.0170 and carbon content was
between 0.025 and 0.029 wt.%.
ERNiCrFe-7 0.52 -*
0.74/0.95/1.45/2.7
0.7Ti/1Ti/1.5Ti/3Ti -*
respectively
0.6/1.2/1.6/2.9
0.5Nb/1Nb/1.5Nb/3Nb 0.52
respectively
1Nb-1Ti 0.95 1.1
*
Notes: : Measured content lower than the SEM-XEDS detection limit for the
used procedure.
For a confidence interval of 67%: Ti-content, between ± 0.05 and ± 0.15,
and Nb-content, between ± 0.1 and ± 0.2, increasing standard deviations
corresponding to increasing alloying contents in the table.
Simulation Methodology
state atomic diffusion Equation 1 [20]. This equation was solved with
Dictra® software [21-22], which uses the finite-difference method and
assesses thermodynamic and kinetic parameters from specific databases.
The commercially available databases Ni-data® and Mob2® were used for
the calculations performed. The migration rate for each phase interface
was computed by assuming that interface migration is controlled by the
mass balance obtained from fluxes of diffusing elements across the interface
(Eqs. 2 and 3).
Fig. 3. Schematic representation of the simulated system. (a) OM top image of the
welding spot; (b) SEM micrograph showing alloying micro-segregation at inter-
dendritic regions; (c) scheme of the one dimensional system used for numerical
simulations and its relevant dimensions
Ductility-Dip Cracking Resistance Improvement of ERNiCrFe-7 435
n
∂μ i
J K = −¦ L'ki (1)
i =1 ∂z
v L 2 / γ (c kL 2 − c kγ ) = J kL 2 − J kγ , k = 1,2, … , n (3)
Fig. 8. Computed average alloying content at the 0.5 μm tick outmost layer of the
austenite dendrite as a function of Ti and Nb additions
Phase Sample Ti Nb Cr Ni Fe Al
EDS NC EDS NC EDS NC EDS NC EDS NC EDS NC
MA/1Ti /3Ti 90.0 99.8 - - - - 4.5 0.2 4.5 <0.001 1.0 <0.001 NM <10-8
MX 1Nb 25 14.8 65 85 5.0 0.2 3.5 <0.01 1.5 <0.001 NM <10-8
1Nb-1Ti 45* 38 42* 62 6 0.2 5 <0.01 1.5 <0.001 NM <10-8
Notes: EDS: Experimentally measured by TEM-XEDS on extracted carbon rep-
lica;
NC: Numerical calculation;
MA: Master Alloy ERNiCrFe-7;
NM: Non measurable value;
* : MX particles with average Nb-content varying between 20 and 70
wt.% and Ti-content varying between 20 and 70 wt.% were identified.
Fig. 10. (a) Computed MX fraction as a function of Ti and Nb additions; (b) com-
puted and SEM-measured MX fraction as a function of Ti or Nb additions
the increase in volumetric fraction was higher than the increase in average
size. This is due to solute segregation at interdendritic regions that simul-
taneously increases driving force for MX precipitation and the number of
cell-triple intersections, the density of nucleation sites and the concomitant
number of particles per unit area being consequently increased. For modi-
fied alloys with high alloying contents (3Nb or 3Ti wt-%), the experimen-
tally verified intensification of MX precipitation was significantly higher
than the one numerically predicted, which most probably arises from errors
associated with neglecting nucleation events in the numerical algorithm as
well as with simplifying the analyzed 3D micro-segregation phenomena to
a 1D system. However, an actually wider solubility range of either carbon
or nitrogen inside the MX phase with regard to the one described by the
Ni-data® database can also be a factor affecting the relationship predicted
between alloying additions and primary carbonitride precipitation, i.e. the
actual carbon and/or nitrogen content of MX precipitates can be lower than
the one numerically predicted.
It is a reasonable approximation to relate the MX precipitates displaying
irregular morphologies with eutectic particles precipitated at lower tem-
peratures, where diffusion distances are shorter due to limited diffusivity
of substitutional atomic species in austenite. On the other hand, it is a good
approximation to relate the MX particles displaying regular, almost po-
lygonal or equiaxial morphologies with pro-eutectic particles precipitated
at high temperatures, where minimization of interfacial energy prevails
over minimization of diffusion distances. Thus, it can be inferred that Nb
and Ti additions induce an increase in MX particles with irregular mor-
phologies due to MX particles in modified alloys having more eutectic
character and they are precipitated at lower temperatures when compared
to MX particles in the unmodified alloy. Irregular primary precipitates are
expected to produce less cracking resistance microstructures due to the fa-
cilitated nucleation of cavities at such intergranular particles when com-
pared to smooth round-like precipitates. Therefore, Nb and/or Ti additions
should be optimized in order to have the right particle temperature forma-
tion, the right number and distribution of particles within the macrostruc-
ture and finally the right particle morphology. From the point of view of
morphology, the lower the addition of these elements would produce
needed round-like primary MX precipitates.
446 A.J. Ramirez, C.M. Garzón
Fig. 12. Computed growth kinetics of primary MX. The start and end of MX pre-
cipitation are shown as a function of Nb and Ti content of the alloy and tempera-
ture. Notice the reverse order of alloying content axis.
Ductility-Dip Cracking Resistance Improvement of ERNiCrFe-7 447
Fig. 13. Computed growth kinetics of primary MX. The 0%, 20%, 80% and 100%
transformed fractions are shown as a function of temperature, and Nb and Ti con-
tent of the alloy. Notice the reverse order of alloying content axis
As an example, Figs. 14 and 15 show some of the results from the TEM
analyses of M23C6 and γ’ precipitates observed in carbon replicas extracted
from the master alloy ERNiCrFe-7 and the sample 1Nb-1Ti. There are
shown in these figures the overall precipitate morphology, enlarged detail
with lattice resolution image, SAD patterns of the region shown, and
XEDS spectra colleted focusing an 5 nm diameter electron beam at the
centre of the correspondingly enlarged second-phase particles.
There were not detected significant changes in M23C6 or γ’ morphology
as a consequence of alloying additions. M23C6 carbides precipitated in
448 A.J. Ramirez, C.M. Garzón
Fig. 14. M23C6 precipitates observed in carbon replicas from master alloy ER-
NiCrFe-7. XEDS spectra of precipitate colleted using 5 nm diameter electron
beam at the center of precipitate shown on the lattice resolution image detail; SAD
pattern
Ductility-Dip Cracking Resistance Improvement of ERNiCrFe-7 449
or non alloying additions are made. These two behaviors can qualitatively
be explained by the average Ti- or Nb- (constituents of γ’ and γ”) and C-
content (constituent of M23C6) segregated to the inter-dendritic region (see
Fig. 8). The higher the alloying segregation, the higher the driving force
for secondary precipitation. Although, simultaneous presence of M23C6 and
γ’was numerically predicted, it was not observed. However, it is expected
that a microstructure containing γ’ precipitation may hinder the observa-
tion and identification of small fractions of M23C6 carbides. Nevertheless,
when the numerical and experimental results are analyzed together it is
found that M23C6 precipitation can be inhibited by the faster γ’ stability in-
crease compared to the M23C6 stability rise (samples 1Nb, and 1Ti1Nb).
This is probably due to the competition between these two phases for the
nucleation sites, with M23C6 nucleation being overrun by γ’ nucleation.
The prevalence of γ´ nucleation is due to the lower volumetric atomic dif-
fusion required and the lower energy interface formed when compared
with M23C6 nucleation. It is not clear to the date if intergranular M23C6 are
damaging for the DDC resistance. However, is believed that when these
precipitates appear in the microstructure the ductility dip cracks have al-
ready had the chance to form.
XEDS spectra (Figs. 14 and 15) showed the occurrence of minimal solu-
bility of Ni, Fe, Al, Nb and Ti in M23C6 and Fe and Cr in γ’. In a similar
way as presented above for XEDS analysis of MX particles, it is expected
that Ni and Fe content of M23C6 particles and Fe and Cr content of γ’
Ductility-Dip Cracking Resistance Improvement of ERNiCrFe-7 451
particles are lower than revealed by the XEDS analyses, which are de-
graded by spurious X-ray signal coming form regions away from the ana-
lyzed particles. Minor changes of chemical composition inside second-
phase precipitates were predicted by kinetic simulations, and therefore
these results are not shown. These numerical results show that atomic ele-
ment stoichiometry in second-phase precipitates can be described as Cr26C6
and Ni3(Ti,Nb,Al). XEDS analyses and numerical simulations of average
chemical composition of M23C6 and γ’ revealed that Ti additions did not
caused significant composition changes of M23C6 or γ’. In contrast, in Nb-
modified alloys, the here so-called γ’ particles displayed substitution of Ti
and Al for Nb. Therefore, the new formed phase may be γ”.
Summary
Based on the exposed results and previously reported data on the weldabil-
ity and performance of alloy ERNiCrFe-7, it is possible to establish appro-
priate composition ranges for the modification of such alloy with Nb and
Ti additions. Thus, the kinetic phase diagrams presented indicate that alloy
ERNiCrFe-7 modified with Ti-content up to ~1.0 wt-% and Nb-content be-
tween ~0.5 and ~1.0 wt-% will present optimized as-welded microstruc-
ture with regard to DDC resistance and in some degree to corrosion and
SCC resistances because of (i) M23C6 precipitation can be slightly inhibited
with regard unmodified ERNiCrFe-7, (ii) although some γ´ precipitation is
induced, still low driving force for γ’ precipitation is established, and (iii)
profuse MX precipitation above the ductility-dip temperature range (i.e.
~0.2–0.25 vol.% of MX above ~1150 °C) takes place. Intense precipitation
of primary carbonitrides above the ductility-dip temperature range and mi-
nor precipitation of secondary chromium carbides during the whole weld-
ing cooling cycle are desired because [1-5] primary carbonitride precipita-
tion results on increased DDC resistance due to the formed tortuous grain
boundaries, whereas chromium rich carbide precipitation at low tempera-
ture induces chromium depletion of the matrix phase near grain boundaries
and, consequently, may reduce corrosion resistance as well as SCC resis-
tance of the alloy.
It was observed that Ti or Nb additions are useful to induce increasing
precipitation of primary carbonitrides and simultaneously to inhibit secon-
dary precipitation of chromium rich carbides. This is due to alloying addi-
tions increasing the fraction of C partitioned in the (Ti,Nb)C phase with
regard to C partitioned in the γ phase. However, these Ti or Nb additions
lead to strong atomic segregation near the interdendritic regions and, as a
452 A.J. Ramirez, C.M. Garzón
Acknowledgements
References
10. Ramirez AJ, Lippold JC (2004) High Temperature Behavior of Ni-base Weld
Metal – Part II. Insight Into the Mechanism for Ductility Dip Cracking. Mater
Sci. Eng. A 380: 245–258.
11. Nishimoto K, Saida K, Okauchi H (2006) Microcracking in Multipass Weld
Metal of Alloy 690 Part 1 – Microcraking Susceptibility in Reheated Weld
Metal. Sci. Tech. Weld. Joining 11(4): 455–461.
12. Nishimoto K, Saida K, Okauchi H, Ohta K (2006) Microcracking in Multipass
Weld Metal of Alloy 690 Part 1 – Microcraking Mechanism in Reheated
Weld Metal. Sci. Tech. Weld. Joining 11(4): 462–470.
13. Capobianco TE, Hanson ME (2005) Auger Spectroscopy Results from Ductil-
ity Dip Cracks Opened under Ultra-High Vacuum. In: Proc. 7th Int. Conf. on
‘Trends in Welding Research’, Pine Mountain, GA, USA, May 2005, ASM
International: 767–772.
14. Nissley NE, Lippold JC (2003) Development of the Strain-to-Fracture Test.
Weld. J. 82(12): 355s–364s.
15. Ramirez AJ, Sowards JW, Lippold JC (2006) Improving the Ductility-Dip
Cracking Resistance of Ni-Base Alloys. J. Mat. Proc. Tech. 179: 212–218.
16. Lippold JC, Nissley NE (2006) Further Investigations of Ductility-Dip Crack-
ing in High Chromium, Ni-base Filler Metals. Document IIW-IIC-326-06:
1–10.
17. Nishimoto K, Saida K, Okauchi H, Otha K (2006) Microcracking in Multipass
Weld Metal of Alloy 690 Part 1 – Prevention of Microcraking in Reheated
Weld Metal by Addition of La to Filler Metal. Sci. Tech. Weld. Joining 11(4):
471–479.
18. Garzon CM, Ramirez AJ (2007) Second Phase Precipitation in As-Welded
Microstructures of ERNiCrFe-7 Ni Alloy Modified with Ti and Nb Additions.
Submitted to Mater. Sci. Eng. A.
19. Garzon CM, Ramirez AJ (2007) Kinetic phase diagrams for assessing second-
phase precipitation during welding a Ni-30Cr-10Fe alloy modified with Ti
and Nb additions. Submitted to Scripta Materialia.
20. Andersson JO, Ågren J (1992) Models for numerical treatment of. multicom-
ponent diffusion in simple phases. J. Appl. Phys. 72: 1350–1355.
21. Borgenstam A, Engström A, Hoglund L, Ågren J (2000) J. Phase Equilibria
21: 269–280
22. Andersson JO, Helander T, Höglund L, Shi P, Sundman B (2002) Thermo-
Calc & DICTRA, Computational Tools for Materials Science. Calphad 26:
273–312
23. Christensen AN (1978) Acta Chem. Scand. Ser. A 32(1): 89–90.
24. Huang X, Zhang Y, Liu Y, Hu Z (1997) Effect of Small Amount of Nitroten
on Carbide Characteristics in unidirectional Ni-Base Superalloy, Metall.
Mater. Trans. A 28A(10): 2143–2147.
25. Craven AJ, He K, Garvie LAJ, Baker TN (2000) Complex Heterogeneous
Precipitation in Titanium–Niobium Microalloyed Al-killed HSLA Steels—1.
(Ti,Nb)(C,N) Particles, Acta Mater. 48: 3857–3868.
Index
for critical strain rate determination, Prokhorov, N.N., 5, 6, 9, 40, 67, 68,
77–91 70, 82, 128, 278
experimental procedure, 78–80 Propagation, in Al alloys hot cracking,
local strain, 89, 90 5, 10, 13
MVT test, 80, 81 ProPHASE program, 24, 25
numerical approach, 79, 80 Pshennikov, A., 264
strain rates, 87–90 Pumphrey, W.I., 5, 42
temperature field, 82–87
test materials, 78, 79 Ramirez, A.J., 410, 419, 421, 422
weld pool geometry, 82–87 Rappaz, M., 5, 6, 9, 19, 30, 32, 44,
welding parameters, 78, 79 67, 128
Rappaz-Drezet-Gremaud hot tearing
Overlay weld metal criterion, see RDG
microcracking behaviour in, RDG (Rappaz Drezet Gremaud)
393–395 criteria, 44, 45
microcracking mechanism in, in Al alloys welding, 19–36
395–398 alloy hot tearing susceptibility,
improvement mechanism, 23–27
396–398 coalescence, 20
microcracks, characterization, deformation, 20–22
393–394 filler content influence, 27
microstructural analysis with added granular model, 32–35
La, 395, 396 macroscopic approach, 20
microporosity, 20
Patterson, W., 5 microscopic approach, 20
Pellini, W.S., 5, 40 percolation of grains, 20
Pellini, W.S.F., 66 solidification paths, 25–27
Ploshikhin, V., 43, 72 thermal hot tearing criterion,
Porosity formation, and hot tearing in 27–29
Al alloys, 6 thermo-mechanical approach,
Prediction, crack 29–32
weld solidification cracking, 55 coherency point, 19
alloy composition, 55 limitations, 22
composition-strain rate maps, 55 Resistance spot welding of 6xxx Al
numerical simulation, 55 alloys, 311–324
pertinent welding information, 55 See also under Simulation
restraining conditions, 55 investigations
Preheating, to avoid hot cracking, 313 Rindler, W., 23, 67
Primary water stress corrosion Robino, C.V., 51
cracking (PWSCC), 390
Process tape Savage, W.F., 148
resistance spot welding simulation Saveiko, V.N., 5
with a tape, 317–322 SAW cold wire technology, 115–125,
welding with and without, difference 215–236, 329, 330
between, 312 chemical composition of weld
spot welding of Al, 314–316 metal, 120, 121
464 Index