You are on page 1of 9

Available online at www.sciencedirect.

com
Physics Procedia 12 (2011) 232240
1875-3892 2011 Published by Elsevier Ltd.
doi:10.1016/j.phpro.2011.03.030
LiM 2011
Understanding of Humping Based on Conservation of Volume Flow
Peter Berger*, Helmut Hgel, Axel Hess, Rudolf Weber, Thomas Graf
Institut fr Strahlwerkzeuge (IFSW), Pfaffenwaldring 43, 70569 Stuttgart, Germany


Abstract
A generalised view on the causes leading to humping during high-speed seam welding is proposed. It is based on the fact that,
whenever the melt velocity in a welding bead exceeds considerably the welding speed, the weld groove cannot be completely
filled by melt at any location and any instant. As a consequence of solidification from the walls of the weld groove, the shape of
the free cross section for the melt flow is changed in such a way that the rest melt stream experiences an upwards deflection
which initiates the formation of a hump. Conservation of volume in an incompressible fluid requires that in the solidified seam
regions of smaller and larger cross sections compared to that of the groove must exist. The initiation of humping is discussed for
different situations in laser beam welding characterized by the aspect ratio welding depth over spot diameter. As this figure is
typical for the geometry of the melt pool/ weld groove, it appears as a meaningful parameter for correlating the onset of humping.

Keywords: Welding; Quality; Instabilities; Humping; Surface Structure
1. Introduction
The humping phenomenon is generally understood as the formation of drop like piles on the top of a weld bead
that alternates with the appearance of undercut between the individual humps. It occurs during seam welding at feed
rates above a certain value and is likewise observed in electron beam, arc and laser beam welding. As it imposes a
restriction to the attainable welding speed and, hence, limits the productivity of these joining technologies, it has
been intensively studied for more than 30 years. To explain the causes of humping several approaches were pursued
which are based on quite different conceptions. None of them, however, is capable to fully explain the phenomena
observed under a great variety of process parameters; moreover, some of them are questioned in recent literature.
Three facts, however, are generally accepted, namely that (1) the humps are finally formed at the end of the molten
region, (2) the onset of humping is related to high melt velocities within the weld pool, and (3) the high-velocity
melt flow is due to the ablation pressure at the keyhole front in e-beam as well as in laser welding and is produced
by electromagnetic forces in the discharge during arc welding.
The herein presented view of the causes giving rise to humping should be regarded as an extended and
generalized interpretation of experimental evidence so far assessed under similar, yet more special aspects; they are
to be found in [1], [2] and [3]. These approaches, and also the occasionally discussed one of a decaying jet, shall be
briefly outlined, before in sections 2 and 3 more recent experiments are presented and interpreted with regard to the
conservation of mass.

* Corresponding author. Tel.: +49-711-685-66843; Fax: +49-711-685-66841.
E-mail address: peter.berger@ifsw.uni-stuttgart.de.
Peter Berger et al. / Physics Procedia 12 (2011) 232240 233

From the results in [1] it clearly becomes evident that under conditions leading to humping in e-beam welding of
thin metal sheets the picture of a melt groove that is completely filled with melt is not valid. Instead, a high-velocity
melt jet starts to solidify from the side-walls (here in the case of through-welds) of the groove thereby narrowing the
free flow channel. This causes the rest of the flowing melt to be deflected upwards and towards the center of the
bead where it finally piles up to humps. Such a mechanism is regarded as a deciding factor for the onset of humping
in gas tungsten arc (GTA) welding as well [4]. Quite similar arguments are formulated along these lines in [5] and
(6) (although both publications deal with gas metal arc (GMA) welding, where additional momentum is transferred
to the melt by the droplets from the wire, the basic physical phenomena are identical): A pinching of the thin metal
film moving along the wall of the weld groove and its rapid solidification is determined as a requirement for hump
formation. The elongated wall jet a perfect term used in [6] to describe the high-speed melt flow in the thin film
is sensitively exposed to rapid solidification which can choke off it from a growing hump. As a consequence, a
melt bridge remains for some time between the solidifying hump at the pool end and a new swelling occurring
closer to the arc position. Such a bridging was observed also in laser beam welding [2] where this fact was
interpreted as a periodic switching of melt pool length.
That surface tension effects should play an important role during the formation of humps suggests itself and shall
be underlined by two examples from e-beam [7] and GTA welding [4]. Most convincing evidence is given in [7]
where a higher oxygen content is demonstrated to drastically enhance the tendency to humping. Physically the same
effect is documented in [4] by observing that with lower sulfur content the welding speed becomes higher at which
humping begins to develop. Furthermore, surface tension is implicit related to the otherwise purely geometrical
issues of the Raleigh instability theory as it determines the growth time of a disturbance leading to the break-up of a
cylindrical fluid jet [8], [9] and [10]. It is interesting to note, how these authors thought about the characteristic
distance between the individual humps. In [8] it is stated that the calculated critical length l
c
after which a liquid jet
breaks up into droplets does not sufficiently agree with the experimentally observed pool length l
m
(the reported
values of l
c
are up to 30% larger than l
m
), whereas in [9] and in [10] l
c
= l
m
was taken as an a priori assumption.
Implicitly this equating indicates the obvious view of the authors that solidification at the pool end plays an
important role and must somehow be related to the time scale of hump formation. In fact, the theoretical treatments
in [9] and [10] using that ansatz end up by identifying geometrical features of the solidified bead (ratios of melt pool
width to length [9] and, additionally, seam height to width [10]) as characteristic figures to describe humping.
Despite this fact it remains questionable (see also [11]) whether that approach is applicable against the background
of experimental evidence (e.g. a free jet definitely does not exist, humps my start to grow close behind the
keyhole).
Other authors [2], [12] calculated the 2D-flow field around a keyhole and regarded the high melt velocity in the
center of the pool as responsible for the humps, being formed as a consequence of high stagnation pressures at the
rear end of the weld pool. The high velocity values reported in [2] are confirmed by more recent FEM calculations
[13]. They can exceed the feed rate by more than an order of magnitude and, during deceleration of the flow,
produce high stagnation pressures.
Yet another cause for humping is discussed in [3] and [12] where the flow field close to the keyhole is regarded
responsible. An upwards directed flow component in the pool behind the keyhole interacts with the flow around the
keyhole thereby leading to the formation of a hump. Possible causes for the underlying momentum in the melt may
be a momentum transfer via friction and direct impact of the vapor flow onto the rear keyhole wall. In welding with
disk and fiber lasers an additional component might arise from the piston effect of moving steps at the keyhole
front [14]: the melt film accelerated there along the direction of the laser beam axis is deflected at the bottom of the
weld groove below the keyhole tip and gives rise to an upwards flow.
2. Experimental and theoretical results substantiating the proposed conception
In addition to the investigations performed in the authors` labs [2], [13], [14], [15] and [16], reliable facts from
other publications have been evaluated, among them and in particular those reported in [10], [17], [19] and [20]. The
lasers used in the various experiments included all types that were and are employed in industrial applications,
namely CO
2
, Nd:YAG, disk and fiber lasers. The most common diagnostic method in all recent studies is high-speed
234 Peter Berger et al. / Physics Procedia 12 (2011) 232240
photography which allows to observe and distinguish in detail keyhole geometry, melt stream(s) and hump
formation (and, of course, spattering in its great variety).
The pictures in Figure1 show the appearance of different fluid dynamic states as the travel speed is increased at
otherwise constant parameters. In these experiments [15], [16] a single mode fiber laser with a BPP of
0.34 mm mrad was used. The focusing conditions were such that a (theoretical) spot diameter of 50 m was
established at the work-piece surface. The line energy P/v of 600 J/m was kept constant and adjusted in such a way
that the welding depth in the metal sheet (stainless steel 1.4310, thickness 0.5 mm) was close to 0.5 mm but no full
penetration did occur.
v = 20 m/min
short weld pool, tiny spatters,
round keyhole
stainless steel; s = 500 m; d
f
= 50 m; P/v= 600 J/m
v = 40 m/min
longer weld pool, standing melt wave,
elongated keyhole
v = 50 m/min
long waves, longer weld pool,
elongated keyhole
v = 60 m/min
periodic beaded weld seam,
elongated keyhole
v = 70 m/min
long region of underfill, hump formation
starting far behind region of energy coupling
v = 80 m/min
distance between humps becomes smaller
resize of window display

Figure 1: Appearance of keyhole and melt pool with increased travel speed (note that scale is changed in the two last pictures), see text.
At low values of travel speed, the classical picture of a thermally determined laser weld can be seen, i.e. a more
or less circular keyhole orifice and a typical elliptically elongated weld-pool geometry. Tiny spatters leave from the
rim of the keyhole. This indicates that a strong vapor flow exerts high shearing strains via friction on the keyhole
wall thereby tearing off small droplets. As the speed is increased the keyhole gets elongated under the direct impact
of the vapor on its rear wall, see also [17], [19]. At the same time, a melt wave appears behind the keyhole; it is
initiated by momentum which stems from vapor impact and from the piston effect of moving steps at the keyhole
front, see e.g. [14]. For speed values around 60 m/min and above, behind the keyhole there exists a region of
underfill or depression, where the volume flow V

, which was molten at the front of the pool, does by far not fill
up the weld groove; this phenomenon is seen also in the experiments described in [17], [19] and [20]. The fact that
the melt flow is concentrated mainly at the bottom and less at the side-walls is due to an obviously existing velocity
component along the front of the keyhole pointing towards the tip of the keyhole. Only recently, it has been directly
visualized and measured to be around 8 m/s for the conditions investigated in [20]. It is noted that in situations
showing this underfill the term keyhole should not be understood in the sense that a closed beam trap does exist,
but rather as the geometrical region where energy is coupled in. All together, the appearance of underfill is a clear
indication that the velocity of the melt stream by far exceeds the welding speed. Downstream of this region then the
melt starts to pile up and form humps. During their growth they are still faster than the welding speed. At the very
rear of the pool, where they begin to solidify, they then move with the travel speed, i.e. they are fixed to the frame of
Peter Berger et al. / Physics Procedia 12 (2011) 232240 235
the work piece. Although the typical aspect ratio, a = s/d
f
, welding depth s over spot diameter d
f
, is an order of
magnitude larger here as in [17], [19], the observed fluid dynamic phenomena are essentially the same.
Driving force for the high-speed melt jet downstream of the region of energy coupling is the high ablation
pressure there [19]. Results of 2D-calculations [13], presented in Figure 2, illustrate the fluid dynamic conditions.
The underlying assumption
of a cylindrical keyhole
does not impose restrictions
on the comparability of this
data with measurements
since the calculations reveal
that practically all energy is
deposited at the front. It can
be seen that the melt
velocities at the side walls
are about an order of
magnitude higher than the
welding speed. Also, the
theoretical seam widths
agree well with the meas-
ured ones of 110, 80, and
70 m corresponding to
weld speeds of 15, 50 and
100 m/min. Furthermore, it
appears reasonable to
assume that the flow com-
ponent along the front is of
the same order of magnitude as the velocity u
mx
on the sides of the keyhole presented in Figure 2. Thus, from both,
experiments and theoretical descriptions, the existence of a high-speed melt stream behind the region of energy
coupling can be regarded as given fact for process parameters yielding long weld pools.
Start and growth development of humps are studied from sequences like those presented in Figure 3 which were
obtained at 15000 fps. After a region of underfill where the melt velocity component in travel direction u
mx
is larger
than the travel speed v, the melt begins to rise up. This position l
s
is interpreted to be the start of hump development.
As the hump moves farther backwards, its velocity becomes lower. At the end of the pool l
m
, when solidification of
the hump commences, it moves with v. Under some conditions, however, the previously produced hump is still
liquid. This has led to assessments according to which the weld pool length fluctuates [2] respectively that there
exist more than one hump within l
m
[6], [16]. With increased v the distance between two solidified humps d
h

decreases slightly. This tendency qualitatively agrees with findings in [10] and [18] where a proportionality
v P d
h
/ ~ (1)
was deduced from experiments.

p
0
i
n

b
a
r
u
m
x
i
n

m
/
s
b
i
n

m

o
i
n

m
v in m/min v in m/min
d
f
= 25 m
d
f
= 25 m
d
f
= 25 m
d
f
= 25 m
d
f
= 50 m
d
f
= 50 m
d
f
= 50 m
d
f
= 50 m
Figure 2: Calculated fluid dynamic properties around a keyhole. The diagrammes show the total pressure
p
0
at the keyhole front, the melt velocity component u
mx
in welding direction at both sides of the keyhole,
the pool width b there and the melt film thickness o at the front.
236 Peter Berger et al. / Physics Procedia 12 (2011) 232240
The evaluation of pictures like those in Figure 3 yields
a dependence of the characteristic lengths on welding
speed at otherwise constant parameters as presented in
Figure 4. The data shows a steadily increasing growth of
both quantities with speed. This is explained as a result of
the rapid increasing u
mx
with v (see Figure 2). The high-
velocity, high-temperature melt stream leads to an
intensified energy transport to the rear so that
solidification sets in later (farther downstream). The
reproducible discontinuity at around 60 m/min might be
understood as a consequence of a changing keyhole/ melt
flow interaction: Up to this speed value an elongated
keyhole exists the rear wall of which is hit by vapor flow
from its front. At somewhat increased speed the high-
velocity melt flow around the front has a larger
momentum so that the side wall streams merge only at a
distance far away from the front; a direct impact of a
vapor jet on the melt does no longer occur. Hence, this
jump would characterize the change from a keyhole
dominated regime to a fluid dynamic dominated one.
0
500
1000
1500
2000
2500
3000
3500
0 10 20 30 40 50 60 70 80 90 100
l
s
,
l
m
i
n

m
v in m/min
P/v = 600 J/m
d
f
= 50 m
stainless steel
l
m
l
s

Figure 4: Weld-pool length l
m
and distance ls where flow appears to become deflected and humps start to grow.
3. Conception based on conservation of mass/volume flow
The approach presented in the following tries to understand humping in a most general way as a consequence of
mass, respectively volume conservation over the complete time period from melting during energy coupling and
acceleration of the molten material there until all the melt`s complete solidification. Accordingly, when somewhere
along the weld pool at some time instant the melt velocity considerably exceeds the feed rate, the weld groove
there cannot be filled completely with melt. As a consequence from the conservation of mass (volume appears a
more appropriate term as it accounts also for thermal expansion effects), at some other location material has to be
accumulated. In the interaction of fluid dynamics, solidification of the melt (mainly) from the sidewalls of the
groove and surface tension effects humps are formed. So, in a completely solidified seam there must exist regions
with smaller and larger cross sections as compared to that of the groove. This idea, which explicitly takes into
account the conservation of mass/volume, is graphically represented in Figure 5; it altogether follows largely Arata`s
[1] interpretation of phenomena observed during e-beam welding of metal sheets.
l
s
start of hump built up
l
m
length of molten pool
d
h
distance between humps
0.010318 s
0.010445 s
0.010572 s
l
m
l
s
d
h
P/v = 600 J/m
d
f
= 50 m
stainless steel

Figure 3: Development of humps at v = 60 m/min and definition of
characteristic dimensions, see text.
Peter Berger et al. / Physics Procedia 12 (2011) 232240 237
A
x
A(x) A
groove
=
d
h
v V/
-
}
=
h
d
groove
h
A dx x A
d
) (
1
for = const:
Figure 5: Sketch illustrating the consequence of mass/volume conservation on bead cross section. V

is the volume being molten per unit of time


and A
groove
is the cross section produced thereby. Volume losses due to spatter and vaporisation as well as volume increase by thermal expansion
are neglected in the discussion.
In laser beam welding (what is the main concern of this contribution) humping is observed within a wide field of
experimental conditions which, at a first glance, can be characterized by the spot diameter. In the case of large
values, say d
f
about 0.4 mm and larger (figures which are common for most CO
2
and Nd:YAG lasers), a situation is
found which shows close resemblance to that in arc welding. In such cases humping occurs in relatively shallow
grooves with ratios of weld depth to width of about one. Welding with disk and fiber lasers allowing spot diameters
as low as several tens of m also leads to humping with, however, very narrow weld seams and aspect ratios of the
order of 10 and more; a situation similar to e-beam welding of thick sections. Because of sometimes confusing
discussions without clearly addressing the particular situation, here the two cases shallow or deep and narrow
beads will be discussed on the basis of s/d
f
which appears to be the most appropriate figure for a general
description.
In Figure 6 an attempt is undertaken to illustrate the time and
space dependent development of humps. On top of the figure is
a sketch of a melt pool which is fixed in space. The positions of
the different humps A, B, C, D at various time steps can be
followed along the straight parallel lines v = dx/dt, the slope of
which corresponds to 1/v. Indicated are different flow regimes
distinguished by relating the longitudinal flow component u
mx
to
v. The region of very high melt flow velocity, u
mx
>> v, ends
where, due to solidification effects, the flow is deflected
upwards and melt is piled up to form a hump. Since the position
l
s
as well as the geometry there (i.e. the cross section of the melt
flow already solidified) depend on the spot diameter at
otherwise constant process parameters the particular features
will be accounted for in Figs. 7 and 8. In the x-t diagram the
high velocity of a fluid element travelling from the front towards
l
s
is indicated by the smallest slope; this region corresponds to
the appearance of underfill. After the hump has begun to grow,
it moves (still growing) towards the end of the pool where it
comes to rest relative to the work piece, i.e. u
mx
= v. The
stagnation pressure of the oncoming flow and surface tension
finally form the hump A(t
o
). A still liquid hump B(t
o
) could be
assessed as the existence of two humps within the melt pool; the
length of which in the notation of Figure 6 would then have a value of l
m
= l
s
+ d
h
. The distance d
h
is established in a
balance of fluid dynamic and thermal effects including cooling and solidification; its value cannot be predicted by
the quantitative argumentations developed herein (on the other hand, an equating of d
h
with l
m
is neither satisfying
from the point of gaining an understanding of physical causes).
x
t
u
m
x


v
u
m
x
>
v
u
m
x
~
v
l
m
d
h
A(t
0
+At) B(t
0
+At)
A(t
0
+2At)
A(t
0
) B(t
0
)
B(t
0
At) C(t
0
At) D(t
0
At)
C(t
0
) D(t
0
)
C(t
0
+At)
B(t
0
+2At)
A B C D
l
s
v
u
m
t
0
t
0
+At
t
0
+2At
t
0
At
t = t
0

Figure 6: Time dependent development of humps depicted
in a x-t diagramme, see text.
238 Peter Berger et al. / Physics Procedia 12 (2011) 232240
l
m
l
s
d
h
a b c d
e e
cross section a
cross section e cross section d
cross section b
cross section c
u
mx
v u
mx
v u
mx
~ v
bottomjet or
walljet region
deflection region /
start of hump formation
stagnation
region
melt stream
solidified wall jet
weld groove
b
s

Figure 7: Hump formation in a typical shallow and long pool geometry caracterized by s/d
f
1.

Here, in addition to melt
deflection by rapid
solidification from side
walls, u
mx
stemming from
vapor friction in keyhole
(and moving steps at its
front) initiates humping
u
mx
> v
l
m
l
s
d
h
s
s
b
melt stream
solidified melt
weld groove
cross section a cross section c
cross section d
d d
a
cross section b
a
b c

Figure 8: Hump formation in a typical deep, slender and short pool geometry characterized by s/d
f
10.
The situation in shallow beads s/d
f
1 is discussed in more detail along the various sections through the bead as
sketched in Figure 7. Here, the region of underfill might extend up to several d
f
as reported in [10]. The melt flow is
concentrated towards the bottom of the groove without wetting the cold side walls up to the work-piece`s surface
(section a); upon final solidification this effect gives rise to undercut in the region between two humps (section d).
Solidifying melt from the wall and the rear causes a narrowing and change of the free cross section so that at some
position l
s
(section e) the local high-speed volume flow is deflected upwards. Viewing many high-speed films, a
figure of thumb for l
s
typically lies between 0.5 l
m
and l
m
(the very end of the pool). Thus, the hypothesis presented
Peter Berger et al. / Physics Procedia 12 (2011) 232240 239
g y ( )
in [2] that the stagnation of a high-velocity melt jet at the pool end causes humping would hold for the case of
l
s
= l
m
.
The two arrows drawn at the hump in its stagnation position (section c) shall indicate that its size might still
become larger due to the decelerating oncoming flow as well as by surface tension effects pulling material from the
melt bridge between the actual and the preceding hump.
A somewhat different flow field will be established in geometries that can be characterized by s/d
f
= 10,
nevertheless, with the same basic mechanisms involved. In very slender and deep weld pools as indicated in Figure
8 the cooling from the side walls is very efficient (note that the cooling rate from a geometrically defined body
depends on the ratio of its surface area to volume). This implies that, assuming b ~ d
f
, the cooling rate increases as d
f

is made smaller. As a consequence, the melt pool (at least the part below the surface and in the depth) is much
shorter in such cases in comparison to its extension
for s/d
f
= 1. The, hence, wedge-like solidification
front at the rear causes a deflection of the x-
component of melt flow to occur far closer to the
keyhole. Thus, the piling up of melt is observed to
appear at distances rather near the keyhole [18],
[23] defined by the location behind the keyhole at
which the weld pool more or less abruptly changes
in depth. This is quite the situation illustrated at the
high welding speeds in Figure 4.
Thermal expansion [24] can account for an
increase of volume flow, which might be a reason
that underfill is not so pronounced in these deep and
slender welds. In addition to the deflection
mechanism, other effects too will contribute to an
upwards flow of the melt under these conditions. In
particular, the momentum transferred to the keyhole
wall by friction from the vapor flow might be of major importance. This can be concluded from many high-speed
films and other evidence, examples of the latter are given in Figure 9. The cross section at the left indicates how
tracer material (sheet metal without Cr, 50 m thick, 50 m below surface) is transported from lower regions of the
pool towards its top and, hence, is found in the humps. On the right is demonstrated how by taking into account
vapor friction in theoretical calculations this mechanism changes the melt pool shape (in these calculations, not
allowing for geometrical changes of the surface, the upwards flow u
mz
is just deflected close to the (here fixed)
surface giving rise to a longer melt pool there. Altogether, from these discussions a picture evolves that in essence
is quite similar to that given in [3].
4. Conclusions
The way in which the melt flow evolves as function of time and space in a solidifying melt pool is the central
aspect that was pursued in the conception presented herein. It is postulated that humps start to develop in a region
where the bulk melt flow experiences an upwards deflection which is primarily caused by a change/narrowing of the
bead`s cross section due to solidification starting from the wall of the weld groove. As this is a geometrical effect,
process parameters yielding the corresponding pool/groove shapes should play a deciding role for the onset of
humping. In fact, characterizing typical geometries by the aspect ratio, the description given along these lines can
explain the observed phenomena without being in conflict with other approaches, particularly those in [2] and [3].
Although the conception formulated herein takes into account the various phenomena observed in the
experiments and is supported by basic physical correlations, it is far away from representing a model allowing
quantitative statements. To establish such descriptions it is necessary to obtain more insight into e.g. the
solidification under particular conditions, the mechanisms giving rise to the upwards flow close behind a keyhole,
the effects of thermal volume expansion (of interest mainly for welding with extreme small spot diameters) and not
least, the role of surface tension and other material properties.

(a)
(b)
Figure 9: Experimental (left,see text) and theoretical proof of an upwards
directed (against the direction of laser beam) flow component for
d
f
= 50 m, P = 500 W, v = 30 m/min: pool shape (a) without, (b) with
taking into account vapor friction in the keyhole.
240 Peter Berger et al. / Physics Procedia 12 (2011) 232240
References
[1] Arata, Y.; Nabegata, E.: Tandem electron beam welding. Trans. JWRI, 7 (1978), 101.
[2] Beck, M. et al.: Aspects of keyhole/melt interaction in high speed laser welding. SPIE Vol. 1397 (1990), 769.
[3] Beyer, E.: Schweien mit Laser. Springer, 1995.
[4] Mendez, P.F.: Penetration and defect formation in high-current arc welding. Welding J. 82 (Oct. 2003), 296-s.
[5] Cho, M.H., Farson, D.F.: Understanding bead hump formation using a numerical simulation. Metall.&Mat. Transactions 38B (2007), 305.
[6] Nguyen, T. C. et al.: The humping phenomenon during high speed gas metal arc welding. Sci. and Techn. of Welding and Joining 10 (2005),
447.
[7] Sievers, E.-R.: Schmelzbadinstabilitten beim Elektronenstrahlschweissen von Grobblechen. Schweissen und Schneiden 58 (2006), 288.
[8] Albright, Ch. E.; Chiang, S.: High speed laser welding discontinuities. J. of Laser Appl.1 (1988), 18.
[9] Gratzke, U. et al.: Theoretical approach to the humping phenomenon in welding processes. J. Phys. D: Appl. Phys. 25 (1992), 640.
[10] Thomy, C. et al.: The occurance of humping in welding with highest beam qualities. Key Eng. Materials 344 (2007), 731.
[11] Kumar, A.; Debroy, T.: Toward a unified model to prevent humping defects in gas tungsten arc welding. Welding Journal 85, (Dec. 2006),
292-s.
[12] Pirch, N. et al.: Die Humping Instabilitt beim Schweien mit Laserstrahlung. Proc. 10th Int. Congress Laser 1991 (1992), 552.
[13] Schuster, R.: Melt bath shapes and keyhole vapour pressures. 21st Meeting of Mathematical Modelling of Material Processing with Lasers.
Igls/Innsbruck, 2008.
[14] Berger, P. et al.: Zur Bedeutung von gleitenden Stufen an der Kapillarfront beim Schweien und Schneiden mit Laserstrahlen. Accepted for
publication in Schneiden und Schweien, 18. Nov. 2010.
[15] Hess, A., Dausinger, F.: Humping mechanisms during high-speed welding with brilliant lasers. 3rd Pacific Int. Conf. on Appl. of Lasers and
in Optics, 2008, paper 3 on proceedings CD.
[16] Collective of authors: 2nd Intermediate Report on Project SCHARP, FGSW 2007, IFSW University of Stuttgart.
[17] Fabbro, R. et al.: Experimental study of the humping process during Nd:YAG laser welding. Proc. 4th Int. WLT Conference Lasers in
Manufacturing (2007), 277.
[18] Neumann, St. et al.: Influence of material and process parameters on the humping effect. Proc. ICALEO 2008, 345.
[19] Fabbro, R., Slimani, S.: Melt pool dynamics durig deep penetration cw Nd:YAG laser welding. Proc. 4th Int. WLT Conf. Lasers in
Manufacturing (2007), 259.
[20] Erikson, I. et al.: High speed camera investigations of the keyhole front. Proc. CD of 13th Int. Workshop on Process Diagnostics ond Control
in Laser Beam Welding and Cutting, 2011 (IFSW).
[21] Hgel, H., Graf, Th.: Laser in der Fertigung. Vieweg und Teubner (2009).
[22] Radaj, D.: Schweissprozesssimulation Grundlagen und Anwendungen. DVS Vol. 141 (1999).
[23] Kawahito, Y. et al.: Elucidation of high-power fibre laser welding phenomena of stainless steel and effect of factors on weld geometry. J.
Phys. D: Appl. Phys. 40 (2007), 5854.
[24] Sudnik, W. et al.: Numerical simulation of weld pool geometry in laser beam welding. J. Phys. D: Appl. Phys. 33 (2000), 662.

You might also like