MMP2003 1

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

MARCEL DEKKER, INC.

• 270 MADISON AVENUE • NEW YORK, NY 10016


©2003 Marcel Dekker, Inc. All rights reserved. This material may not be used or reproduced in any form without the express written permission of Marcel Dekker, Inc.

MATERIALS AND MANUFACTURING PROCESSES


Vol. 18, No. 1, pp. 1–22, 2003

Research and Progress in Laser Welding of Wrought


Aluminum Alloys. I. Laser Welding Processes

X. Cao,* W. Wallace, C. Poon, and J.-P. Immarigeon

Institute for Aerospace Research, National Research Council Canada, Ottawa,


Ontario, Canada

ABSTRACT

With the wide application of Al alloys in automotive, aerospace and other industries, laser
welding has become a critical joining technique for aluminum alloys. In this review, the
research and progress in laser welding of wrought Al alloys have been critically discussed
from different perspectives. The primary objective of this review is to understand the
influence of welding processes on joint quality and to build up the science base of laser
welding for the reliable production of Al alloy joints. Two main types of industrial lasers,
carbon dioxide (CO2) and neodymium-doped yttrium aluminum garnet (Nd:YAG), are
currently applied but special attention is paid to Nd:YAG laser welding of 5000 and 6000
series alloys in the keyhole (deep penetration) mode. In this part of the review, the main
laser welding processing parameters including the laser-, process-, and material-related
variables and their effects on weld quality are examined. In part II of this article in this
journal, the metallurgical microstructures and main defects encountered in laser welding
of Al alloys such as porosity, cracking, oxide inclusions, and loss of alloying elements are
discussed from the point of view of mechanism of their formation, main influencing
factors, and remedy measures. In part II, the main mechanical properties such as hardness,
tensile, and fatigue strength and formability are also discussed.

Key Words: Review; Laser welding; Laser beam welding; Aluminum alloys;
Wrought aluminum alloys; CO2 laser; Nd:YAG laser; Keyhole mode; Deep penetration
mode; Process parameters; Weld quality.

*Correspondence: X. Cao, Institute for Aerospace Research, National Research Council Canada,
Montreal Road, Ottawa, Ontario K1A 0R6, Canada; E-mail: xinjin.cao@nrc.ca.

DOI: 10.1081/AMP-120017586 1042-6914 (Print); 1532-2475 (Online)


Copyright q 2003 by Marcel Dekker, Inc. www.dekker.com
MARCEL DEKKER, INC. • 270 MADISON AVENUE • NEW YORK, NY 10016
©2003 Marcel Dekker, Inc. All rights reserved. This material may not be used or reproduced in any form without the express written permission of Marcel Dekker, Inc.

2 Cao et al.

1. INTRODUCTION

Aluminum and its alloys are widely used in automotive, aerospace, and other
industries because of their low density, high specific strength, good corrosion resistance,
good workability, high thermal and electrical conductivity, attractive appearance, and
intrinsic recyclability. Wrought Al alloys can be categorized into two main groups, namely
non– heat-treatable and heat-treatable. Non –heat-treatable Al alloys including 1000,
3000, 4000 containing only Si, and 5000 series can be strengthened by solid solution, cold
working, and grain refinement. The commercial heat-treatable Al alloys cover 2000 (Al –
Cu or Al –Cu – Mg), 6000 (Al –Mg –Si), and 7000 (Al – Zn –Mg) series alloys.[1,2] Though
arc, resistance, friction, electron beam, and laser welding processes have all been
industrially employed to weld Al alloys,[3] laser welding is one of the most promising
joining methods because it provides high productivity, high weld quality, low distortion,
manufacturing flexibility, and ease of automation.[4,5] For instance, it was recently
reported that for the first time in the aircraft industry laser welding has been used instead of
rivets for A318 and A380 fuselage shells[6 – 9]; therefore, considerable effort has been
made to develop a robust laser welding process for Al alloys.
A laser beam, as a controllable, clean, and concentrated high-intensity heat source,
can readily heat, melt, and evaporate almost all materials. Thus the laser beam has been
used as a welding heat source. Laser welding has many advantages, including

1. low and precise heat input, thus less thermal distortion and greater accuracy;
2. small heat-affected zone;
3. deep and narrow fusion zone with almost parallel fusion boundaries—the aspect
ratio (depth/width) of keyhole laser-welded seams is commonly around 4:1 but
can be as high as 10:1[10];
4. high productivity resulting from high welding speed and simple joint design and
preparation[11];
5. welding of a wide range of materials (steel, Al, Mg, Ti, and superalloys) and
dissimilar materials; and
6. good process flexibility and reliability.

Laser beams are not affected by magnetic fields. Laser welding can be performed with
or without filler material, in various environments (vacuum, air, pressurized chambers, or
controlled atmospheres) and in some locations that are normally inaccessible or accessible
only from one side. In addition to simple linear welds, the latest laser welding facilities are
capable of performing complex angular and curvilinear (nonlinear) welds.[12] Multibeam
techniques are easily realized in laser welding.
Laser welding, however, also has some disadvantages, including:

1. high equipment and operating costs;


2. stringent requirements for the clamping and fitting of workpieces;
3. requirement for accurate beam and joint alignment; and
4. safety requirement with respect to robotic manipulation and eye protection.[13,14]
MARCEL DEKKER, INC. • 270 MADISON AVENUE • NEW YORK, NY 10016
©2003 Marcel Dekker, Inc. All rights reserved. This material may not be used or reproduced in any form without the express written permission of Marcel Dekker, Inc.

Laser Welding of Al Alloys. I 3

Although direct diode lasers with a wavelength of 0.81 mm and delivered power up to
4 kW are now entering the market for aluminum welding,[15,16] two main types of
industrial lasers, carbon dioxide (CO2) and Nd:YAG, are currently applied in welding
operations. The CO2 laser has been used because of its greater power output, higher
efficiency, proven reliability, and safety.[17] With the development of high-output power,
improvement of laser beam quality, and the possibility of glass fiber delivery, the Nd:YAG
laser has entered the application fields dominated by the CO2 laser.[18] The shorter
wavelength Nd:YAG laser has higher absorptivity, thus less power (approximately one-
third less for 6082 alloy) is required with the Nd:YAG laser than with the CO2 laser at the
equivalent penetration depth and welding speed.[19] Compared to the CO2 laser,
the Nd:YAG laser has less severe plasma effects (absorption and defocusing) because the
absorption coefficient of the laser radiation in the plasma scales with the square of
the wavelength.[20] The shorter wavelength Nd:YAG laser beam allows light transmission
via fiber optic cables rather than the articulated mirror delivery system for CO2 laser.[21]
Thus the Nd:YAG laser beam is easier to manipulate and control. For instance, the laser
source may be remote from the welding stations up to a distance of 200 m.[13,22] Another
benefit of fiber optic delivery is that the light beam from a single laser source can be
delivered to a number of workstations.[23] By multifiber beam delivery, which can include
time or power options, multiple processing operations involving welding, cutting, and
surface treatment are possible simultaneously using one Nd:YAG laser source.[22] A
Nd:YAG beam will produce a slightly larger and more stable keyhole than a CO2
beam.[24,25] Thus the wider Nd:YAG beam profile provides the ability to span larger gaps
while maintaining welding quality and speed. Furthermore, the higher stability of the
Nd:YAG laser leads to a more uniform weld bead with less or no porosity.[24] Compared
with the CO2 laser, the Nd:YAG laser can operate at lower welding speeds because of the
reduced interaction among beam, plasma, and metal, and thus thicker metal can be welded.
The reduced plasma effect also permits the use of argon as a shielding gas, whereas helium
is essential for a CO2 laser. The integration of Nd:YAG lasers into conventional machine
tools can be realized more easily, as compared with CO2 lasers.[18] The fiber-guided
Nd:YAG laser has lower adjusting and maintenance requirements than the CO2 laser.[18]
In addition, Nd:YAG lasers are relatively small and compact, and they have no moving
parts, thus they are less prone to breakdown.[18] Therefore, Nd:YAG lasers delivered
through optical fibers with multikilowatt power will probably provide the most significant
revolution in welding technology since the introduction of arc welding.[26]
The effectiveness of laser welding depends greatly on the physical properties of the
material to be welded.[17] Table 1 compares the typical physical properties of pure Al and
iron. All Al alloys possess certain inherent characteristics, such as low absorptivity to laser
beam, tenacious oxide films, low boiling point elements, high thermal conductivity, high
coefficient of thermal expansion, relatively wide solidification temperature ranges, high
solidification shrinkage, a tendency to form low melting constituents, low viscosity, and
high solubility of hydrogen in liquid state.[1,2] Therefore, some defects such as lack of
penetration, excessive porosity and blowholes, liquation and solidification cracking, loss
of alloying elements, degradation in mechanical properties, and inconsistent welding
performance can be encountered in laser welding of Al alloys. The primary objective of
this review is to understand the influence of welding processes on joint quality and to build
up the science base of laser welding for the reliable production of Al alloy joints. In this
MARCEL DEKKER, INC. • 270 MADISON AVENUE • NEW YORK, NY 10016
©2003 Marcel Dekker, Inc. All rights reserved. This material may not be used or reproduced in any form without the express written permission of Marcel Dekker, Inc.

4 Cao et al.

Table 1. Properties of pure aluminum and iron at their melting points.

Properties Unit Aluminum Iron Reference

Ionization energy eV 6 7.8 [27]


Specific heat J/kg/K 1080 795 [28]
Specific heat of fusion J/kg 4 £ 105 2.7 £ 105 [27]
Melting point K 933 1811 [29]
Boiling point K 2603 3343 [27]
Viscosity kg/m/s 0.0013 0.005 [29]
Surface tension N/m 0.84 1.86 [29]
Thermal conductivity W/m/K 94.03 38 [28]
Thermal diffusivity M2/s 3.65 £ 1025 6.80 £ 1026 [28]
Expansion coefficient 1/K 24 £ 1026 10 £ 1026 [27]
Density kg/m3 2385 7015 [28]
Elastic modulus N/m3 7.1 £ 1010 21 £ 1010 [27]

part of the review (part I), the main laser welding processing parameters, including the
laser-, process-, and material-related variables and their effects on weld quality, are
discussed. Details of the defects in welds, the reasons why they form and how they can be
minimized or avoided, and the main mechanical properties of welds are described in the
follow-on article (part II).

2. LASER WELDING PROCESSES OF WROUGHT Al ALLOYS

Laser welding generally has two modes:[30] conduction and deep penetration
(keyhole). In conduction welding, the surface of the material is heated to above its melting
point but below its vaporization temperature. Fusion occurs only by heat conduction
through the welding melt pool. A hemispherical weld bead in cross section with an aspect
ratio of 1.2 or less is formed in a similar manner to conventional fusion welding
processes.[31] Conduction welding is usually done using Nd:YAG rather than CO2 lasers
because of its shorter wavelength.[10] Conduction welds, however, are limited to materials
with thin thickness.[30] In contrast, keyhole laser welding needs a higher power density to
cause local vaporization. Thus a narrow and deeply penetrated vapor cavity with an aspect
ratio higher than 1.2[31] is formed by multiple internal reflection of the beam, as shown
schematically in Fig. 1.
The basic difference between the two modes is that the surface of the weld pool
remains unbroken during conduction welding but opens up to allow the laser beam to enter
the melt pool in keyhole welding.[33] The conduction mode causes less perturbation of the
weld pool and provides less tendency to entrap oxides and gases during welding. Keyhole
mode welding, however, results in better energy coupling, higher penetration, and high
speed necessary for economic justification.[33] Therefore, most applications of laser
welding are centered about the deep penetration process. The research into stable laser
welding has involved identifying and controlling the parameters influencing process
stability and reproducibility to reliably produce defect-free welds at high welding speeds.
MARCEL DEKKER, INC. • 270 MADISON AVENUE • NEW YORK, NY 10016
©2003 Marcel Dekker, Inc. All rights reserved. This material may not be used or reproduced in any form without the express written permission of Marcel Dekker, Inc.

Laser Welding of Al Alloys. I 5

Figure 1. Deep penetration (keyhole) mode laser welding.[32]

The following discussion focuses on the main processing parameters including the laser-,
process-, and material-related variables and their effects on the welding process and weld
quality, with attention being paid to the keyhole mode welding.

2.1. Laser-Related Variables

2.1.1. Wavelength

A short wavelength laser beam enhances the process stability and enlarges the
processing window in two respects: first, the threshold value for keyhole formation is
lower (Fig. 2) as a consequence of higher Fresnel absorptivity; second, the plasma effects

Figure 2. Intensity threshold for laser welding of Al – Mg – Si1.0 alloy from Ref.[34].
MARCEL DEKKER, INC. • 270 MADISON AVENUE • NEW YORK, NY 10016
©2003 Marcel Dekker, Inc. All rights reserved. This material may not be used or reproduced in any form without the express written permission of Marcel Dekker, Inc.

6 Cao et al.

(refraction and absorption) are less severe because the absorptive coefficient in the plasma
approximately scales with the square of the wavelength.[20,35] As a consequence, the
temperature and the viscosity of the plasma are lower, which allows an undisturbed vapor
or plasma flow out of the keyhole.[35] Thus a wavelength of 1.06 mm (Nd:YAG laser) in
comparison to 10.6 mm (CO2 laser) leads to a higher welding depth at equal line energy.[18]
At comparable focus diameter and given laser power, a higher velocity can be achieved
with Nd:YAG than with CO2 laser.

2.1.2. Power

High-power density at the workpiece is crucial to achieve keyhole welding and


control the formation of welds. Compared with steel, Al alloys have higher initial
reflectivity to laser beams and greater thermal conductivity, thus greater power densities
are needed. If the power density is too low, the coupling of laser energy to workpiece and
penetration may be lost.[22] Power density that is too high may cause spatter, undercut,
underfill, and “drop out.”[36] In practice, it is generally advisable not to exceed 107 W/cm2
to avoid heavy ejection of molten material.[10] The laser power should be set according to
the material and its thickness. As shown in Fig. 2, thicker material will require greater
beam intensity at a set welding speed because an increase in laser power yields a
proportional increase in penetration depth. For a set material thickness, higher laser power
will increase welding speed. Higher power has also been reported to enlarge the operating
window.[37]
Currently, CO2 lasers are available with power up to 60 kW,[38] whereas continuous
wave (CW) Nd:YAG lasers up to 4 kW at the workpiece surface are available.[15] Six to
ten kilowatt Nd:YAG lasers are being developed.[24,38] The welding of aluminum alloys
with 6 kW CW Nd:YAG laser leads to longer focal length, additional freedom in tracking
the seam because of the greater distance to the workpiece, increased tolerance with regard
to the process, and doubled welding speed at the same seam depth and width as compared
with a 3 kW laser.[39] Thus a higher power can enlarge process-operating window
(tolerance and parameter fields). However, more laser-induced plasma or plume can be
emanated in deep penetration welding with the increase of laser power.[40]

2.1.3. Spot Size

At the same output power, smaller spot size means higher power density but the welds
may become narrower than necessary or even not fully fused. Laser welding seams are
usually less than one-quarter the width of a tungsten-arc inert gas (TIG) weld for the same
material thickness.[41] With lap joints, a weld narrower than the thinner sheet thickness
cannot provide enough strength.[42 – 46] Joint fit-up and beam alignment are more critical
for a small spot size, thus the processing operation window is reduced with the smaller
spot size.[21] A small spot size may also lead to more loss of elements by vaporization,
causing undercut and underfill defects resulting from too high a power density.[36] Thus
small spot size cannot ensure good welding performance. For a Nd:YAG laser beam, the
spot diameter is dependent on the fiber diameter and the optical system of the output
MARCEL DEKKER, INC. • 270 MADISON AVENUE • NEW YORK, NY 10016
©2003 Marcel Dekker, Inc. All rights reserved. This material may not be used or reproduced in any form without the express written permission of Marcel Dekker, Inc.

Laser Welding of Al Alloys. I 7

housing used for collimating and focusing the beam.[47] Increasing the fiber diameter leads
to larger spot diameters. Migliore[10] recommended that the spot size should be about 30%
of the butt-weld width for keyhole welding. The typical spot sizes in industrial laser
welding usually range from about 0.1 to 1.0 mm, with 0.3 mm being a common size. For
the overlapping welds the laser spot size is usually set to about the desired weld width.[10]

2.1.4. Focal Length

The shorter the focal length (lower F number), the smaller the beam waist diameter
and the depth of focus, and the larger the convergence angle.[36] A shorter focal length or
depth of focus makes the setting of the focus position more critical because of less
tolerance to the variation in the position of the workpiece. A shorter focal length also
generates spatter, thermal, and vapor damage to the cover-glass slide used to protect the
focusing optic. Too short a focal length was also reported to result in an oxidized root
weld.[48] The larger beam convergence angle can reduce the laser beam’s ability to access
certain joints and narrow gaps.[36] Increasing the F number by applying optics with larger
focal lengths enlarges the working distance. Therefore, the welding process becomes less
sensitive to variations in the working distance.[20] Other advantages include the reduced
contamination and damage of the sensitive lenses, cheaper guidance system, and increased
access. However, a longer lens requires significantly greater power levels to induce
keyhole formation.[17] Typically, focal lengths of 100 to 200 mm are used, although it is
usually possible to obtain lenses of focal lengths from 50 to 300 mm.[13,22] The choice of a
lens will additionally depend on the access requirements for a particular application.[22]

2.1.5. Depth of Focus

The depth of focus (DOF) indicates the allowable distance the part can deviate from
the point of the best focus for acceptable welding quality. Within the DOF, the beam
diameter has not increased beyond 5% of the minimum waist diameter and the power
density will decrease no more than 10%.[36] The DOF decreases with the square of the
focal length. With a shorter DOF, weld depth will decrease and the vertical tolerances of
workpiece become tightened. The DOF is also dependant on both power and welding
speed and usually decreases with increasing speed.[42]

2.1.6. Focal Plane Position

The focal plane position is the placement of the focal spot or minimum waist diameter
to the workpiece surface. Straying outside the DOF in relation to the workpiece is unwise
because small deviations cause large variations in beam diameter, especially with low
F numbers, where the beam divergence angle is large.[36] The focus plane should be set
where the maximum penetration depth or best process tolerances are produced.
Focusing the laser beam above the workpiece (positive defocusing) produces more
plasma that defocuses the beam and reduces the irradiance on the surface, but the keyhole
MARCEL DEKKER, INC. • 270 MADISON AVENUE • NEW YORK, NY 10016
©2003 Marcel Dekker, Inc. All rights reserved. This material may not be used or reproduced in any form without the express written permission of Marcel Dekker, Inc.

8 Cao et al.

is inherently more stable at positive defocusing.[49] The placement of the laser beam focus
on or below the surface of the workpiece can optimize the laser beam coupling to the
material and increase the irradiance inside the weld pool.[29,33] The threshold power
density for keyhole formation is lower at negative defocusing than at positive
defocusing.[49] Therefore, the weld pool size at negative defocusing is larger than that for
the same extent of positive defocusing at a given power and spot size. By slightly focusing
the beam below the surface of the material, welding speed could be increased at the same
penetration depth than the beam focused at the material surface.[50]
Jones et al.[51] recommended that the tolerances of focus position (mismatch in beam
direction) for butt joints are ^ 1 mm and ^ 0.5 mm for CO2 and Nd:YAG lasers,
respectively, for 2 mm thick aluminum alloys at a focal length of 150 mm. When filler
metal was used, it was reported that the deviation of the focal or the wire position can be
tolerated up to about ^ 0.4 mm at a focal length of 100 mm for 3 kW Nd:YAG laser.[46]

2.1.7. Beam Alignment (Mismatch of the Beam Across Seams)

Unless movable clamping and beam guiding systems are used in production, a
mismatch of the beam across the seam is unavoidable. The acceptable tolerant zones
clearly depend on the expected application. For a butt joint, the aim is to ensure a
homogeneous seam with sufficient depth, whereas in the case of overlap joints, the
welding width of the joint plane has to be at least the amount of the thinner joining partner
for best strength.[45,46] Tolerances of butt beam and joint misalignment of less than 0.3 and
0.5 mm have been recommended for 2 mm thick aluminum alloys welded using CO2 and
Nd:YAG lasers, respectively.[51] Schinzel et al.[45] claimed that the beam mismatch across
the seam could be tolerated from 0 to 0.6 mm for a gapless butt joint of Al– Mg – Si0.5
alloy with 2 mm thickness. For overlap joints, however, the mismatch of the beam across
the seam is not critical, which is a great advantage in comparison with butt joints.[45]
During laser welding of components with different thicknesses or metallurgical
properties, it is often advantageous to offset the beam profile to one side of the seam. For
example, when butt-welding thicker to thinner material from the step side, it has been
suggested that the center of the beam focus be offset to allow more laser power for thicker
material.[33] This smoothens the weld and ensures full penetration on both sides of the
weld without the possibility of burning through the thinner component. The laser welding
process is very sensitive to this offset. Depending on the thickness of workpiece, an offset
of 0.1 to 0.2 mm to the thicker material is recommended.[48,52]

2.1.8. Operational Mode

Both pulsed wave (PW) and CW lasers are used for welding of aluminum alloys. It
has been reported that a pulsing laser beam reduces energy coupling by approximately
25%.[53] The keyhole has to reform at the start of each pulse during pulsed laser welding.
The metal surface becomes flat after the withdrawal of the pulsed beam and internal
reflection does not occur until the keyhole is reformed and deepened. Thus the energy
transfer is poor for welding. For a first approximation, CW laser welding is used for
MARCEL DEKKER, INC. • 270 MADISON AVENUE • NEW YORK, NY 10016
©2003 Marcel Dekker, Inc. All rights reserved. This material may not be used or reproduced in any form without the express written permission of Marcel Dekker, Inc.

Laser Welding of Al Alloys. I 9

high-speed welding, whereas PW lasers are used for precision welding.[10] The high-
power CW Nd:YAG lasers will assume a dominant role in future applications compared
with PW lasers because CW lasers involve lower investment and operating expenses and
excellent industrial suitability as compared with PW lasers because of simpler main supply
technology and longer lamp service life.[54] Stable Al alloy welds free of porosity, cavity,
or hot cracking and with good bead appearance have been obtained using CW Nd:YAG
laser.[19,55] With PW lasers, bead appearance is not as good, and underfill,[55,56] porosity,
cavity, and crack[1,2,19,56] may occur. Because of the vaporization of magnesium during
PW laser welding of 5000 alloys, spatter becomes intense, thus the bead surface becomes
rough and underfill occurs.[55] For the same average power, PW Nd:YAG lasers provide
deeper penetration than CW Nd:YAG lasers.[56,57] At a constant peak power, the pulse
energy influences the penetration depth and the seam width. At constant pulse energy, the
peak power mainly influences the penetration depth.[47] When certain pulse parameters are
required, the pulse frequency influences the welding speed[47] or bead continuity.[22] It has
been reported that higher pulse frequency produced narrower welds.[51] Thus the
achievement of a deeper penetration and optimization of the pulse waveform to prevent
spattering, porosity, and solidification cracks would be further investigated. It has also
been reported that pulse options of the lasers might expand the variety of materials to be
processed.[54]

2.1.9. Beam (Spatial) Mode

A transverse electromagnetic mode (TEM00) beam with the intensity profile of a


Gaussian distribution can be focused to the smallest spot size and propagate with the
lowest divergence angle from any given waist diameter.[42] Thus the TEM00 beam is
favored in cutting and drilling applications. In welding, however, the TEM00 mode
produces deep and narrow welds. Thus a higher order beam may be desirable in
welding.[42] High-order or multimode output beams cannot be focused to as small a spot
diameter as can the TEM00 beam. The larger focal spots from high-order modes provide
leeway in beam placement tolerance with respect to the weld seam. For example, in butt-
welding, a higher order mode allows wider gap tolerances.[42] The power density is also
more evenly distributed. Smaller changes in workpiece position within this focal depth do
not affect welding performance. High-order modes are less sensitive to resonator
distribution and put less thermal load on the laser optics. These properties increase the
reliability of the laser even though high-order modes diverge more rapidly.[42] Thus high-
order or multimode welding is a more stable and consistent process than the TEM00 mode.
A CO2 laser beam typically has a Gaussian multimode shape with a central peak and
gradual drop intensity in the radial direction.[24] At the output of the delivery fiber of
Nd:YAG lasers, the laser beam usually has a typical top-hat profile.[24]

2.1.10. Multibeam Technique

The term multibeam or multifocus technique is used whether the spots are produced
by combining several beams or by splitting a single beam into several foci. The original
MARCEL DEKKER, INC. • 270 MADISON AVENUE • NEW YORK, NY 10016
©2003 Marcel Dekker, Inc. All rights reserved. This material may not be used or reproduced in any form without the express written permission of Marcel Dekker, Inc.

10 Cao et al.

idea behind this approach, namely using two or more individual focal spots at the
workpiece instead of just one, is to gain improvements in the process performance.[20] In
general, the multifocus technique can increase power level. Thus higher penetration or
process velocity can be obtained.[19,20,58,59] Compared to the single-beam process, the
multibeam technique can improve process robustness and welding quality.[19,20,58,59] The
multifocus technique yields larger gap widths.[20] It has been reported that dual-beam
welding allows gaps up to 0.3 mm with minimal concavity.[24] Common ways to enlarge
seam width are to reduce travel speed or defocus the beam. Both measures work at the
expense of process efficiency. Defocusing the laser beam reduces the laser power density
and, thereby, the weld penetration depth, unless extra power is available to compensate.
The multibeam technique produces a less sensitive geometry by widening and ensuring an
unhindered escape of the evaporated material[59] and the flotation of gas and inclusions[24]
because the weld pool remains molten for an extended period of time. As a result,
blowholes and cavities of irregular shape and large size can be drastically reduced or even
avoided by this measure.[59] Multifocus technique can also increase flexibility for adapting
power density distribution to geometrical requirements of the joints to control the heat
input.[20,24,59]
The multifocus spots can be diverted either longitudinally in the direction of the
laser beams or transversely in the horizontal plane.[48] In the longitudinal arrangement,
the distance between the beams can be changed by adjusting the focal length. When
diverted transversely in the same horizontal plane, the points can be oriented in any
direction to the welding direction and with the opportunity to balance the energy
distribution between the points. Typical configurations in the horizontal plane includes
parallel (cross) or tandem (in-line) arrangements.[19] Compared to the single-beam
process, the multifocus technique provides additional controlling parameters such as the
number of foci, spot distance, power distribution, orientation angle to welding direction,
and arrangement.[59] Small distances between the foci lead to deep and narrow weld
seams, whereas large spot distances result in shallow and broad seam geometries. An
in-line arrangement of foci leads to high penetration depths resulting from the preheating
effect of the first beam. In contrast, the cross-arrangement leads to broad seams.[59]
Ga-Orza,[19] however, arrived at a different conclusion, namely that there is no great
difference in penetration for parallel and tandem arrangements when all other parameters
are optimum and the bead is much narrower in the parallel configuration.

2.1.11. Beam Spinning and Weaving

Beam spinning or weaving has been shown to improve tolerances to joint gaps.[41]
The Welding Institute examined laser beam spinning in a bid to improve the tolerance to
beam and joint alignment and gap at the joint faces. Beam spinning maintains the
maximum power density, but reduces the welding speed by at least 20%.[41]
Beam spinning or weaving equipment has been successfully developed and
applied to steel sheets.[51] The rotation for Nd:YAG laser has even combined with the
dual-beam technique to further optimize the energy input and improve process
robustness.[15,24]
MARCEL DEKKER, INC. • 270 MADISON AVENUE • NEW YORK, NY 10016
©2003 Marcel Dekker, Inc. All rights reserved. This material may not be used or reproduced in any form without the express written permission of Marcel Dekker, Inc.

Laser Welding of Al Alloys. I 11

2.2. Process-Related Variables

2.2.1. Weld Speed

Welding speed is a function of material, laser power, and penetration depth. Too low a
speed leads to excessive melting, loss of material, and weld perforation.[60] Penetration is
inversely proportional to the speed for a given mode, focal spot size, and power, thus an
increase in welding speed at constant laser power results in a corresponding decrease in
penetration depth (even insufficient penetration) and weld volume because of the
reduction in the heat quantity applied to the component.[61] The welding speed for a given
thickness or depth of penetration rises with power, as shown in Fig. 3. Higher welding
speeds reduce alloying element evaporation and produce fine microstructures, causing an
increase in tensile strength of the weld.[64] The demands for quality control and process
tolerance increase with welding speed. The range of welding speeds that produce
acceptable welds obviously varies with laser type, material properties, and thickness.
Laser welding of tailored blanks for automotive applications is typically carried out at 6 to
15 m/min to maintain profitability.[33] Although such high speeds are not possible in
heavier gauge welding, the tendency is always to maintain welding quality at high welding
speed.

2.2.2. Shield Gas

Even though Al alloys can be laser welded without shielding,[48,65] shield gases are
usually used to protect the face and root of the weld metal from oxidation, to control
convection in the melt pool, to control the plasma plume, and to protect the optic lenses
from weld spatter and fumes.[43,44] Blanketing of shielding gas requires a total
displacement of air (and thus oxygen) away from the weld surface to prevent gross
oxidation during welding.[66] Plasma or plume formation via the release of vaporized
material from the keyhole must also be suppressed, especially in CO2 laser welding,
because it can cause instabilities of the vapor-filled keyhole, reduce weld penetration,
produce coarse porosity, and reduce life of coverglass slides.[40] A poorly shielded weld

Figure 3. Effect of continuous-wave Nd:YAG laser power on welding speed for 5083 and 6082
alloys (Data from Refs.[21,40,62,63]).
MARCEL DEKKER, INC. • 270 MADISON AVENUE • NEW YORK, NY 10016
©2003 Marcel Dekker, Inc. All rights reserved. This material may not be used or reproduced in any form without the express written permission of Marcel Dekker, Inc.

12 Cao et al.

will exhibit porosity, undercut, and bead roughness.[15] Gas shielding of the weld pool is
more important when filler wire is used, especially where a gap is present, because the
weld pool is larger and longer than the autogenous laser welding. The relatively long
trailing edge of the weld is particularly vulnerable, and a trailing gas shield is often
necessary. Close control of the underbead shielding is also required to achieve a smooth
underbead without material drop-through for 5000 alloys.[63]
Pure helium or a helium – argon mixture is often preferred for CO2 lasers.[67] Helium
has higher ionization potential; therefore, it resists plasma ignition and offers greater weld
penetration and smoother weld surface, but it is more expensive than argon.[68] Also
helium is lighter than air, thus it tends to rise away from the weld. Argon usually provides
better shield for Nd:YAG lasers.[67] Argon is both cheaper than helium and heavier than
air, but it is readily ionized by the metal plasma and, thus absorbs considerable laser
energy. The percentage of absorption rises with increasing power levels. At lower power
levels, energy absorption beneficially slows solidification rate and yields smoother beads
than when produced in helium. The best solution may be a compromise (i.e., a mixture of
the two gases, helium and argon).[17]
Naeem[21] reported that welds carried out with nitrogen shielding gas for producing
butt and overlap welds gave the best results compared with helium and argon, with respect
to metallurgy and radiography. Welds produced with argon as the shielding gas exhibited
root and centerline porosity; helium showed finely distributed porosity. Nitrogen resulted
in lowest porosity and greater weld penetration depths in comparison with argon and
helium.[50] CW or PW Nd:YAG laser welding performed in oxygen shielding gas shows
that a deeply penetrated weld could be obtained but the bead had a rough surface and large
porosity.[56]
Gas flow is dictated by the shield geometry, type of nozzle, nozzle configuration,
nozzle diameter, pressure, and mass flow rate.[66] Process gas should be present in the
welding interaction region. Gas may be delivered coaxially with the wire or from a
separate nozzle.[69] Much care should be taken to direct the gas flow toward the metal
vapor coming out of the keyhole.[70]

2.2.3. Filler Metal

The addition of filler material lowers the sensitivity to joint gaps.[52] Typically, a gap
up to 0.5 mm can be bridged when using filler material compared with a maximum 0.3 mm
gap without filler wires.[48] Dawes[36] recommended that the joint gap should be slightly
wider than the diameter of the filler wire whenever possible because adding filler wire to
joint gaps narrower than the wire diameter is difficult without special equipment and
precise control. Addition of filler wire can eliminate underfill (concave bead surface) and
undercut from the top and bottom bead, reduce porosity and hot cracking, and improve the
mechanical properties.[21,48] Joints welded with filler material display a smoother
transition and thus reduce the sensitivity to fatigue failure.[21,48] Filler wire can also
stabilize the plasma and reduce the risk of explosions in the weld.[48] Moreover, filler
metals can modify the composition of the fusion zone and compensate for the loss of
volatile alloying elements. Thus filler material provides a means of controlling the
metallurgy of the weld bead and ensures weld quality. Most strain-hardenable alloys can
MARCEL DEKKER, INC. • 270 MADISON AVENUE • NEW YORK, NY 10016
©2003 Marcel Dekker, Inc. All rights reserved. This material may not be used or reproduced in any form without the express written permission of Marcel Dekker, Inc.

Laser Welding of Al Alloys. I 13

be laser welded autogenously but filler wire can be used to improve welding process and
mechanical properties. For heat-treatable alloys, however, filler metal is generally used to
adjust the weld bead composition beyond the crack-sensitive range to avoid hot
cracking.[58]
There are no commercially available filler wires developed specifically for the laser
welding of Al alloys. Filler wire compositions are designed according to the materials to
be welded and the welding technique. The laser weld process is quite different from
conventional arc welding techniques. Therefore, it cannot be assumed that all metal-arc
inert gas welding (MIG) or TIG wires are ideal for laser welding. More wire and parent
material research must be completed before well-established filler wire data can be
presented for laser welding. However, useful guidance on the selection of appropriate filler
metal can be obtained from the existing standards covering MIG[71] or TIG[72] welding of
Al alloys, based on maximum weld metal strength, corrosion resistance, or freedom from
cracking.[5] It was suggested that both 4000 and high-Mg 5000 filler metal are appropriate
for 6000 alloys because Si or Mg reduces hot shortness. A 5000 filler wire is recommended
for 5000 metal. A 4000 wire should not be used for 5000 alloys with Mg higher than 2.5%,
because the weld joints may become brittle as a result of the formation of excess
magnesium silicide (Mg2Si). For joints of dissimilar alloys such as 5000 to 6000 alloys,
5000 wire should be used if needed to reduce hot cracking.[15]
Filler material can be applied to the weld region in several ways: via a continuous wire
feed system in the form of a preplaced wire or as a thin strip nipped between the joint
faces. Other techniques such as powder addition have also been researched.[36] The
successful application of feeding filler wire depends on the wire feed delivery angle,
position in relation to the laser beam and weld zone, and feed rate. When filler metal is fed
into the weld pool, its angle has little effect, but when fed into the beam, its angle plays a
major role in determining the energy absorbed by the weld.[69] The angle at which the wire
is fed relative to the workpiece normally lies between 10 degrees and 60 degrees but the
best results were reported to be 45 degrees.[36,69,73,74] During the CO2 welding process, the
filler is partially melted directly by the laser beam, partially by the laser-induced plasma,
and partially by the molten pool, depending on the wire position.[61] The gap between the
filler wire tip and the weld pool is perhaps the most important setting for the smooth
transfer of feed wire into the weld pool, and it depends on welding speed, focal diameter,
and the feeding direction of the wire.[61] The wire must not intersect the laser beam above
the weld pool to avoid excessive spatter, tooling damage, and reduced penetration.[36,74]
The wire has to touch the surface at the interaction zone. Deviations can be tolerated up to
^ 0.4 mm.[58] Filler wire may be fed directly into the laser beam above the workpiece into
the leading or the trailing edge of the weld pool. Such wire-feeding systems designed for
laser welding are now commercially available,[69] but the normal practice is to feed into
the leading edge of the weld pool.[36] An adaptive wire feed control system based on
measurement of air gap volume using a vision system has been developed.[69] Too high a
wire feed rate results in a reduced underbead or loss of full penetration and heavy topbead.
Lack of penetration may even be met by insufficient melting of wire because of excessive
feed rate. It was reported that the undercut of top bead and weld metal drop-through may
be caused by the opposite settings of these parameters because the weld characteristic
could be changed from lack of penetration to drop-through, merely by control of the wire
feed rate.[21]
MARCEL DEKKER, INC. • 270 MADISON AVENUE • NEW YORK, NY 10016
©2003 Marcel Dekker, Inc. All rights reserved. This material may not be used or reproduced in any form without the express written permission of Marcel Dekker, Inc.

14 Cao et al.

A two-wire technique has also been developed for laser welding of aluminum
alloys.[46] Compared with the cold-wire feed, “hot-wire feeding” (i.e., the filler is
preheated to a high temperature by resistance heating) is claimed to give better process
reliability and welding speed.[75]

2.3. Workpiece-Related Variables

2.3.1. Composition

Katayama[76] studied the melting of various Al alloys by laser beam processing


and rated their melting from easy to difficult: 2090 (Al – Cu2.7) , 5456
(Al – Mg5 – Mn1), 5083 (Al – Mg4.5 – Mn) and 5182 (Al – Mg4.5 – Mn) , 7075
(Al – Zn5.5 – Mg2.5 – Cu1), 7N01 (Al – Zn4.5 – Mg1.5) , 5052 (Al – Mg2.5) , 2024
(Al – Cu4 – Mg1), 6061 (Al – Mg1 – Si – Cu) and 6N01 (Al – Mg0.6 – Si) , 2219
(Al – Cu6 –Mn), and 3003 (Al – Mn1 – Cu) , 1100 (Al99.0 –Cu) and 1050 (Al99.5).
The results indicated that the alloys with higher volatile elements such as magnesium,
zinc, and lithium are more easily melted by the laser beam.[28] Because of their high
vapor pressures and lower vaporization temperatures, these volatile alloying elements
help to establish and stabilize keyholes when they vaporize.[77] Thus the threshold
power density to achieve satisfactory coupling between a laser beam and Al alloys is
reduced with the increase of volatile elements.[28] The threshold beam irradiances for
5000 alloys are reported to be 0.8– 3 £ 106 W/cm2 and for the 6000 series 1–
5 £ 106 W/cm2 for CO2 lasers.[29] In the case of Nd:YAG beam, the threshold
irradiances for 5000 and 6000 series aluminum alloys reduces to , 1 £ 106 W/cm2 and
, 2 £ 106 W/cm2, respectively.[29] Martukanitz and Smith[78] also reported that a
threshold power density of 1.0– 1.2 £ 106 W/cm2 was required for CO2 lasers, whereas
full penetration with CW Nd:YAG lasers was achieved at a power density of
7.5 £ 105 W/cm2 for Al alloys. The ranges of these values or differences depend on
different beam sizes, welding speed, and alloy composition. The 5000 alloys with
higher contents of Mg give rise to a more stable keyhole at lower power densities
than 6000 alloys containing lower levels of Mg. Thus, at a given power density and
penetration depth, 5000 alloys appeared to weld at faster speeds than 6000 materials;
at given speed and penetration depth, lower power densities were required; and at
a given power density and welding speed, deeper penetration was obtained with 5000
alloys, as shown in Fig. 3.

2.3.2. Thickness

As shown in Fig. 3, thicker aluminum plates require a higher power or lower welding
speed. The main restriction to the achievement of thicker section welds is the plasma
above the keyhole at high power for CO2 lasers[79] or power available for Nd:YAG lasers.
For thicker workpieces, multipass welding techniques should be explored. Little research
work has been reported on the potential of this approach.
MARCEL DEKKER, INC. • 270 MADISON AVENUE • NEW YORK, NY 10016
©2003 Marcel Dekker, Inc. All rights reserved. This material may not be used or reproduced in any form without the express written permission of Marcel Dekker, Inc.

Laser Welding of Al Alloys. I 15

2.3.3. Surface Condition

The surface condition of the material has a great influence on the energy absorption of
incident laser light and on the threshold power density for keyhole welding. Typical
energy absorptivity of laser light in pure aluminum and its alloys are shown in Table 2.
The data show that Al has very low absorptivity of laser beams in the as-received
condition. The scattered values may be due to the differences in surface oxide thickness,
roughness, and wavelength of laser beams.[28,80] The polished surface has lower
absorptivity, indicating higher reflectivity of the smooth surface. The energy absorption in
the sand-blasted surface was twice that of the as-received surface, indicating that rough
surfaces absorb more energy than do smooth ones. Huntington and Eagar[80] attributed the
higher absorptivity in the sand-blasted condition to glass beads embedded in the surface
and the increased absorption of anodized samples to the decreased free electron
concentration at the surface for CO2 laser beams. However, anodization had little
influence on the energy absorption of Nd:YAG laser beams compared with the as-received
condition. Ceramics such as Al2O3, MgO, and SiO2 were thought to be transparent in three
ranges of the electromagnetic spectra: ultraviolet (0.2 – 0.4 mm), infrared (0.7 –3.0 mm),
and radar (. 103 mm).[81,82] Clearly the Nd:YAG wavelength of 1.06 mm lies in the
infrared transparent region, which is the reason there are similar absorptive values for
Nd:YAG laser beam in the as-received and anodized specimens.[81] It is this transparency
that makes the delivery by optical fibers possible for Nd:YAG lasers but not for CO2
lasers. All the above results were obtained when a surface was irradiated by a laser beam
without any keyhole formation. Under such conditions the absorption of laser beam energy
by Al and its alloys is very low. However, the formation of a keyhole greatly increases the
absorption of laser energy as a result of the Fresnel effect.
Aluminum usually has continuous oxide layers on the workpiece surface because of
its high chemical affinity for oxygen. The presence of oxide layers can increase the
absorptivity for CO2 but not for Nd:YAG beam, as shown in Table 2. Some ways to
increase the absorptivity for CO2 laser beams are to modify the surface of Al alloys by
mechanical or chemical roughening, by deposition of various absorptive paint-on
coatings, or by anodizing and dyeing the aluminum surface. These methods have been
tried with varying degrees of success. For instance, some absorptive (low-reflective)
coatings such as graphite have been developed but they are not recommended because they
may cause porosity[5] or cracking.[51] A high-power Nd:YAG beam can easily penetrate

Table 2. Energy absorptivity of laser light for pure aluminum and its alloys (%).

CO2 laser (200 J, 2 s pulse)[80] Nd:YAG laser (36 kJ)[81]

Surface condition 5456 Al alloy 99.999 Al 6063 Al alloy


As received 5 – 12 7 14
Electropolished 4 5 —
Sandblasted 22 20 30
Anodised 27 22 13
Anodised (polished) — — 10
MARCEL DEKKER, INC. • 270 MADISON AVENUE • NEW YORK, NY 10016
©2003 Marcel Dekker, Inc. All rights reserved. This material may not be used or reproduced in any form without the express written permission of Marcel Dekker, Inc.

16 Cao et al.

the surface oxides and produce consistent welds of Al alloys in the as-received condition,
so it is usually unnecessary to remove the oxide layer before laser welding.[15,73] However,
the porous oxide films absorb moisture, especially over extended periods in high humidity
and fluctuating temperature environments, which will cause porosity in the welds.[15]
Therefore, some traditional techniques such as mechanical methods (scraping or shaving),
chemical etching (10% NaOH), fluxing, and employing concentrated heat sources are
recommended to remove such oxides.[5,50,54]
In addition, in some cases workpiece surface may contain some lubricants, such as
oils, or may be specially treated. Little is known about the effects of these surface
conditions on the absorptivity of laser beams and the welding performance of Al alloys.

2.3.4. Joint Design and Fit-Up

Various types of joints can be used for laser welding.[33] The butt and lap joints,
however, are most widely used for laser welding of Al alloys. For the butt joint, part fit-up
must be good enough to maintain the alignment between the beam and the joint to avoid
bead concavity caused by air gaps. For lap joints, care must be taken to ensure the
cleanliness of the contacting surfaces. Much tighter control of fit-up is required to achieve
consistent weld quality in a high-speed production environment because of the much more
highly focused energy source.
The general process tolerances for high power Nd:YAG laser welding of steel are
recommended as follows[22]: focus position; ^ 1 mm; joint gap (butt/lap), 10% of material
thickness; beam/joint misalignment, 0.5 mm; joint mismatch, 50% of sheet thickness. It
has been suggested that the tolerances for Al welding are less than those of steel.[22] Some
other researchers recommend that the gap width for laser buttwelds should be 7% to 8% of
thinner material thickness,[83] or no greater than 15%.[10] Typical joint fit-up tolerance is
10% of material thickness.[17] Jones et al.[51] recommend that the tolerance of butt-joint
gaps be less than 0.3 and 0.5 mm for 2 mm thick Al alloys using CO2 and Nd:YAG lasers,
respectively.
Surface mismatches between butt-welded sheets is also of importance for laser
welding of tailored blanks. Jones et al.[51] recommended that the tolerance of surface
mismatch may be up to 50% of sheet thickness for CO2 and Nd:YAG lasers. For lap joints,
the presence of a gap between the sheets can cause problems with the control of weld
appearance. For 1 mm Al – Mg – Si1.0 alloy, a gap of 0.2 mm may be sufficient to cause
weld undercut and ideally, joint gaps should be kept to a minimum.[51]

2.3.5. Fixturing

Small laser beam diameter leaves little tolerance for the variation of workpiece
positions. Excessive heat input during Al welding can cause considerable distortion,
particularly in thin sheets. Compared to traditional welding, the laser process produces a
very narrow weld bead, with minimum thermal distortion. Aluminum components,
however, should be carefully fixtured before welding to assure the repeatable accuracy of
welding joints in a production environment.
MARCEL DEKKER, INC. • 270 MADISON AVENUE • NEW YORK, NY 10016
©2003 Marcel Dekker, Inc. All rights reserved. This material may not be used or reproduced in any form without the express written permission of Marcel Dekker, Inc.

Laser Welding of Al Alloys. I 17

3. CONCLUDING REMARKS

With the wider application of wrought Al alloys in automotive, aerospace, and other
industries, laser welding will become a critical joining technique for Al alloys. Great
progress has been made in the development of high-power and high-quality CO2 lasers.
However, high-power CW Nd:YAG lasers only have a relatively short history. Because
the research on the development of new equipment drives toward demonstrating the
maximum capabilities of a process as a prelude to satisfy the needs of industry, many
welding fundamentals have not been researched and, therefore, are not fully understood.
Laser welding of Al alloys is far from a mature and reliable process. Industry will certainly
expect a predictable, repeatable, consistent, and reliable laser welding technique before the
process is widely accepted. Wider welding operating windows are also welcome for wider
applications of laser welding processes. Much work should be done to define parameter-
operating windows for the different alloys of interest to produce defect-free laser-welded
joints. Process specifications for laser welding of Al alloys should be built up to assure the
reliable production of Al joints.
Different Al alloys and different thicknesses or shapes, applicable to sheets,
extrusions, castings, and hydroformed parts, will probably be laser welded in the future.
With the development of high-power lasers, laser welding of heavy structures could be one
of the future applications. To weld the thicker material or to improve process quality, laser
welding process combined with MIG, TIG, and plasma arc processes or even
combinations of similar or different lasers (e.g. multibeam techniques) will possibly be
used in industries. The easy manipulation and control of Nd:YAG lasers through optical
fiber delivery provides welding opportunities for complex geometries in which overhead,
vertical, horizontal, or all-position welding may be carried out. Efficient and effective
process control and monitoring are required for such applications, even though they are
still in their infancy. Difficulties arise because there is no electrical signal, as in the case
with arc and resistance welding, which may be used to monitor the process or provide a
feedback signal to the power supply for control purposes. Greater use may therefore be
needed for vision systems that are able to monitor key aspects of the process. Without a
doubt, these new fields in applications of laser welding processes will provide many
challenging tasks.

ACKNOWLEDGMENTS

One author (XC) would like to thank the Natural Science and Engineering Research
Council of Canada for the award of a visiting fellowship at the Aerospace Manufacturing
Technology Center, Institute for Aerospace Research, National Research Council Canada,
Canada.

REFERENCES

1. Cam, G.; Kocak, M. Progress in joining of advanced materials. Int. Mater. Rev.
1998, 43 (1), 1 –44.
MARCEL DEKKER, INC. • 270 MADISON AVENUE • NEW YORK, NY 10016
©2003 Marcel Dekker, Inc. All rights reserved. This material may not be used or reproduced in any form without the express written permission of Marcel Dekker, Inc.

18 Cao et al.

2. Cam, G.; Kocak, M. Progress in joining of advanced materials—part 2: joining of


metal matrix composites and joining of other advanced materials. Sci. Technol.
Weld. Joining 1998, 3 (4), 159 – 175.
3. Lanza, M.; Lauro, A.; Scanavino, S. Fabrication and weldability in structures. AL
Alumin. Alloys 2001, 13 (135), 80– 86.
4. Kutsuna, M.; Yan, Q.U. Study on porosity formation in laser welds of aluminum
alloys (report 2)—mechanism of porosity formation by hydrogen and magnesium.
J. Light Met. Weld. Constr. 1998, 36 (11), 1– 17.
5. Ion, J.C. Laser beam welding of wrought aluminum alloys. Sci. Technol. Weld.
Joining 2000, 5 (5), 265 – 276.
6. Hinrichsen, J. The latest development of airbus A380. AL Alumin. Alloys 2001, 13
(136), 171 –178.
7. Wustefeld, F. Successful shot peen forming tests for airbus A380 fuselage shells.
Met. Finishing News 2001, 2 (8), 12.
8. Schmidt, H.; Voto, C.; Hansson, J. Tango metallic fuselage barrel validation of
advanced technologies. ICAF’2001: 21st Symposium of the International
Committee on Aeronautical Fatigue, Toulouse, France, June 27 –29, 2001.
9. Zink, W. Welding fuselage shells. Ind. Laser Solutions Manuf. 2001, 16 (4), 7– 10.
10. Migliore, L. The principles of laser welding. Ind. Laser Rev. 1998, 13 (7),
17– 21.
11. Rudlaff, T.; Heesen, N. Laser welding in automobile production. European Symp. on
Assessment of Power Beam Welds, GKSS Research Center, Geestharch, Germany,
Feb 4– 5, 1999.
12. Kochan, A. Laser welding adapts to non-linear tailored blanks. Assembly Automat.
2001, 21 (1), 48– 50.
13. Verhaeghe, G.; Riches, S.T.; Wiemer, K.; Ion, J.C. High power Nd:YAG laser
processing. ICAWT’97: Int. Conf. on Advances in Welding Technology, Columbus,
Ohio, Sept 17 –19, 1997; 197 – 210.
14. Kocak, M.; Dos Santos, J.F. Laser welding technology: trends in industrial
applications and quality assessment. Rev. De La Soud. 1999, 55 (4), 26– 35.
15. Clarke, J.A.; Christy, B. Aluminum tailored-welded blanks for automotive
applications. Aluminum 2001—Proc. TMS 2001 Aluminum Automotive and
Joining Sessions, Annual Meeting in New Orleans, LA, Feb 12 – 14, 2001; Das, S.K.,
Kaufman, J.G., Lienert, T.J., Eds.; 117– 128.
16. Herfurth, H.J.; Cryderman, M.; Clarke, J.A. Diode laser welding of aluminum. Ind.
Laser Solutions Manuf. 2001, 16 (4), 29– 31.
17. Sprow, E.E. The laser-welding spectrum: what it has to offer you. Tooling Prod.
1988, 54 (3), 56– 63.
18. Hack, R.; Dausinger, F.; Hugel, H. Cutting and welding applications of high power
Nd:YAG lasers with high beam quality. Proc. SPIE: ICALEO 1994, The
International Society for Optical Engineering: Orlando, Florida, 17– 20 Oct., 1994;
Vol. 2500, 210 –219.
19. Ga-Orza, J.A. Welding of aluminum alloys with high-power Nd:YAG lasers. Rev. de
Met. 1998, 34 (2), 227– 231.
20. Hugel, H. Solid state lasers—development trends and application potentials. 7th
NOLAMP Proc./NORDIC Conf. 1999, 1, 1 –12.
MARCEL DEKKER, INC. • 270 MADISON AVENUE • NEW YORK, NY 10016
©2003 Marcel Dekker, Inc. All rights reserved. This material may not be used or reproduced in any form without the express written permission of Marcel Dekker, Inc.

Laser Welding of Al Alloys. I 19

21. Naeem, M. Welding of aluminum alloys with continuous wave Nd:YAG Laser up to
5 kW. ICALEO 1999: Laser Materials Processing, San Diego, CA, Nov 15 –18,
1999; Vol. 87 (2), 206– 215.
22. Riches, S.T.; Ion, J.C. Guide to high average power Nd:YAG laser processing with
fiber optic beam delivery for metals. Automotive Ind. Core Res. TWI Publ. 2000,
1 – 39.
23. Rooks, B. The laser solution discovers its problems. Automotive Manuf. Solutions
2000, 1 (1), 52 –59.
24. Mueller, R.; Gu, H.; Ferguson, N. Nd:YAG laser welding for automotive
manufacturing applications. IBEC 1999: Int. Body Engineering Conf. and
Exposition, Detroit, MI, Sept 28 –30, 1999.
25. Das, S. Aluminum tailor welded blanks. Adv. Mater. Processes 2000, 157 (3),
41 – 42.
26. Russell, D. Welding with high power Nd:YAG lasers. Weld. Met. Fabrication (UK)
2000, 68 (3), 17 –20.
27. Trumpf, Welding of Aluminum and Aluminum Alloys, April 1993.
28. Zhao, H.; White, D.R.; DebRoy, T. Current issues and problems in laser welding of
automotive aluminum alloys. Int. Mater. Rev. 1999, 44 (6), 238– 266.
29. Leong, K.H.; Sabo, K.R.; Sanders, P.G.; Spawr, W.J. Laser welding of aluminum
alloys. Proc. SPIE 1997, 2993, 37 –44.
30. Tam, S.C.; Williams, R.; Yang, L.J.; Jana, S.; Lim, L.E.N.; Lau, M.W.S. A review of
the laser processing of aircraft components. J. Mater. Processing Technol. 1990, 23,
177 – 194.
31. Richter, K.; Bostanjoglo, G.; Dommaschk, R.; Mayrhofer, R.; Pathe, D.; Weber, H.
Comparative study on aluminum and steel welding with CW and repetitively
Q-Switched Nd:YAG lasers. Proc. SPIE, Nov 30– Dec 2, The International Society
for Optical Engineering: Cleveland, Ohio, 1995; Vol. 2789, 12 –20.
32. Schubert, E.; Klassen, M.; Zerner, I.; Walz, C.; Sepold, G. Light-weight structures
produced by laser beam joining for future applications in automobile and aerospace
industry. J. Mater. Processing Technol. 2001, 115, 2– 8.
33. Duley, W.W. Laser Welding; John Wiley & Sons Inc.: Toronto, 1999.
34. Behler, K.; Berkmanns, J.; Ehrhardt, A.; Frohn, W. Laser beam welding of low
weight materials and structures. Mat. Des. 1997, 18 (4/6), 261 – 267.
35. Hugel, H.; Beck, M.; Rapp, J.; Dausinger, F. Laser welding of aluminum. XI Int.
Symp. on Gas Flow and Chemical Lasers and High-Power Laser Conf., 1997; Vol.
3092, 516 –521.
36. Dawes, C. Laser Welding: A Practical Guide; Abington Publishing: Cambridge,
1992.
37. Steen, W.M. Laser Material Processing, 2nd Ed.; Springer-Verlag London Limited:
London, 1998.
38. Seto, N.; Katayama, S.; Matsunawa, A. Porosity formation mechanism and
suppression procedure in laser welding of aluminum alloys. Quart. J. Jpn Weld. Soc.
2000, 18 (2), 243 –255.
39. Alder, H.; Bloehs, W. Perspectives of welding with a 6 kW Nd:YAG laser for the
automotive industry. ICALEO’98: Laser Materials Processing, Orlando, Florida,
Nov 16– 19, 1998; Vol. 85 (2), 18– 24.
MARCEL DEKKER, INC. • 270 MADISON AVENUE • NEW YORK, NY 10016
©2003 Marcel Dekker, Inc. All rights reserved. This material may not be used or reproduced in any form without the express written permission of Marcel Dekker, Inc.

20 Cao et al.

40. Naeem, M.; Riches, S.; Russell, D. Material processing with a high average power
and high beam quality continuous wave Nd:YAG laser. ICALEO’98: Laser
Materials Processing, Orlando, Florida, 1998; Vol. 85 (2), 8 –19.
41. Dunkerton, S.B.; Dawes, C.J. The application of diffusion bonding and laser welding
in the fabrication of aerospace structures. AGARD CP (NATO) 1986, CP-398,
3.1– 3.12.
42. Migliore, L. Laser Materials Processing; Marcel Dekker Inc.: New York, 1996.
43. Dausinger, F.; Rapp, J.; Hohenberger, B.; Hugel, H. Laser beam welding of
aluminum alloys: state of the art and recent developments. Proc. Int. Body
Engineering Conf. IBEC’97: Advanced Technologies and Processes, Stuttgart,
Germany, Sep 30 –Oct 2, 1997; Vol. 33, 38– 46.
44. Dausinger, F. Laser welding of aluminum alloys: from fundamental investigation to
industrial application. In High-Power Lasers in Manufacturing, Proc. SPIE; Chen,
X., Fujioka, T., Mataunawa, A., Eds.; 2000; Vol. 3888, 367 –379.
45. Schinzel, C.; Hohenberger, B.; Dausinger, F.; Hugel, H. Laser welding of aluminum
car bodies—from research to production. Proc. Conf. ICALEO’98: Laser Materials
Processing, Orlando, Florida, 1998; Vol. 85 (2), 56– 65.
46. Schinzel, C.; Hohenberger, B.; Dausinger, F.; Hugel, H. Laser welding of aluminum
with filler wire. 7th NOLAMP Proc./7th NORDIC Conf.: Laser Processing of
Materials, Lappeenranta, Finland, Aug 23 –25, 1999; Vol. 84, 430 –440.
47. Tonshoff, H.; Schumacher, J. Welding of aluminum alloys using kW Nd:YAG laser
with beam guidance via optical fibers. Proc. 5th European Conf. ECLAT’94: Laser
Treatment of Materials, Bremen, Sept 26– 27, 1994; Vol. 163, 188 –196.
48. Larsson, L.; Palmquist, N.; Larsson, K. High quality aluminum welding—a key
factor in future car body production. Svetsaren 2000, 54 (2), 17 –24.
49. Pastor, M.; Zhao, H.; Martukanitz, R.P.; DebRoy, T. Porosity, underfill and
magnesium loss during continuous wave Nd:YAG laser welding of thin plates of
aluminum alloys 5182 and 5754. Weld. J. 1999, 207S – 216S.
50. Sekhar, N.C.; Hilton, P.A.; Scott, G. Laser beam welding of aluminum stiffened
panels. Proc. from Joining of Advanced and Specialty Materials, ASM International,
St. Louis, MO, Oct 9 – 11, 2000; 129 –135.
51. Jones, I.A.; Riches, S.T.; Yoon, J.W.; Wallach, E.R. Laser welding of aluminum
alloys. TWI J. 1998, 7 (2), 421 –481.
52. Torsten, C.; Niclas, P. Nd:YAG laser welding of aluminum blanks, utilized
technical aspects on weldability and formability. 7th NOLAMP Proc./NORDIC
Conf.: Laser Processing of Materials, Lappeenranta, Finland, Aug 23 –25, 1999;
Vol. 2, 441 –450.
53. Weston, J. Surface related features of laser welding of aluminum alloys. ISIJ Int.
2000, 40, S6 – S9.
54. Emmelmann, C. Suitability of high-power solid state lasers for welding applications
in particular aluminum. Laser Application in Manufacturing Seminar, Welding
Technology Institute of Australia, May, 1995.
55. Kanazawa, H. Welding performance of high power YAG lasers with aluminum
alloys. J. Light Met. Weld. Constr. 1997, 35 (7), 10 –15.
56. Katayama, S.; Hamada, S.; Matsunawa, A. YAG laser welding of aluminum alloys.
Proc. 6th JWS Int. Symp.: Role of Weld. Sci. Technol. 21st century, 1996; 249– 254.
MARCEL DEKKER, INC. • 270 MADISON AVENUE • NEW YORK, NY 10016
©2003 Marcel Dekker, Inc. All rights reserved. This material may not be used or reproduced in any form without the express written permission of Marcel Dekker, Inc.

Laser Welding of Al Alloys. I 21

57. Kocak, M.; Cam, G.; Kim, Y.J.; Dos Santos, J.F. Mechanical and fracture properties
of laser beam welded joints. Proc. 5th Int. Conf.: Trends in Welding Research, Pine
Mountain, Georgia, June 1– 5, 1999; 805 –815.
58. Weeter, L. Technological advances in aluminum laser welding. Pract. Weld. Today
1998, 2 (1), 56 –57.
59. Hohenberger, B.; Chang, C.L.; Schinzel, C.; Dausinger, F.; Hugel, H. Laser welding
with Nd:YAG multi-beam technique. ICALEO 1999: Laser Materials Processing
conf., Laser Institute of America, San Diego, CA, Nov 15 –18, 1999; Vol. 87 (2),
167 – 176.
60. Mazumder, J. Laser welding. In Laser Materials Processing; Bass, M., Ed.; North-
Holland Publishing Company: New York, 1983; Vol. 3, 115 –200.
61. Neye, G.; Heider, P. Laser beam welding of modern Al-alloy for the aircraft
industry. Proc. Conf. ECLAT’94/Dusseldorf: Deutscher Verband fur Schweis-
stechnik, 1994; 108 –117.
62. Dausinger, F.; Rapp, J.; Beck, M.; Faisst, F.; Hack, R.; Hugel, H. Welding of
aluminum: a challenging opportunity for laser welding. J. Laser Appl. 1996, 8 (6),
285 – 290.
63. Naeem, M.; Riches, S.T. Laser welding of aluminum alloy sheets for the automotive
industry. Materials for Lean Weight Vehicles (UK), Nov 24– 25, 1997; 185 – 192.
64. Weston, J.; Wallach, R. Mechanical properties of laser welds in aluminum alloys.
Proc. INALCO’98/7th Int. Conf.: Joints in Aluminum, Cambridge, April 15 –17,
1998; 332– 341.
65. Irving, B. The auto industry gears up for aluminum. Weld. J. 2000, 79 (11), 63 –68.
66. Blake, A.; Mazumder, J. Control of magnesium loss during laser welding of Al-5083
using a plasma suppression technique. J. Eng. Ind. 1985, 107 (3), 275– 280.
67. Clarke, J.A. Laser welding aluminum alloys. Ind. Laser solutions 2000, 15 (6),
17 – 22.
68. Yoshikawa, M.; Kurosawa, T.; Nakata, K.; Kimura, S.; Aoki, S. YAG laser welding
of aluminum alloys. J. Light Met. Weld. Constr. 1994, 32 (9), 15 –23.
69. Ion, J.C.; Jokinen, T.; Salminen, A.; Kjanpaa, V. Laser beam welding using filler
wire. Ind. Laser Solutions Manuf. 2001, 16 (2), 16 –18.
70. Hugel, H.; Berger, P.; Dausinger, F. Modelling of laser treatment process—a
versatile development tool. Proc. Conf. ICALEO’98: Laser Materials Processing,
Orlando, Florida, 1998; Vol. 85 (2), 141– 150.
71. British Standard. MIG Welding—Part 1: Specification for MIG Welding of
Aluminum and Aluminum alloys, BS 3571, 1985.
72. British Standard. TIG Welding—Part 1: Specification for TIG Welding of
Aluminum, Magnesium and Their Alloys, BS 3019, 1984.
73. Naeem, M.; Jessett, R. Aluminum tailored blank welding with and without wire feed,
using high power continuous wave Nd:YAG laser. IBEC’98: Proc. Int. Body
Engineering Conf. & Exposition: Body Manufacturing Assembly and Advanced
Manufacturing, Detroit, Sept 29, 1998; 247 –255.
74. Naeem, M.; Jessett, R. Welding aluminum tailored blanks with Nd:YAG laser for
automotive applications. Pract. Weld. Today 1999, 3 (1), 24 –28.
75. Aichele, G. New development in welding technology for light metals, part II.
Aluminum 2001, 77 (7/8), 575 –584.
MARCEL DEKKER, INC. • 270 MADISON AVENUE • NEW YORK, NY 10016
©2003 Marcel Dekker, Inc. All rights reserved. This material may not be used or reproduced in any form without the express written permission of Marcel Dekker, Inc.

22 Cao et al.

76. Katayama, S. Laser welding phenomena and melting of aluminum alloys. J. Light
Met. Weld. Constr. 1996, 34 (4), 23 –30.
77. Boisselier, D.; Lenoir, R. Weldability and manufacturing of YAG laser welds on
aluminum alloys. European Symp. on Assessment of Power Beam Welds, GKSS
Research Center, Geesthacht, Germany, Feb 4– 5, 1999.
78. Martukanitz, R.P.; Smith, D.J. Laser beam welding of aluminum alloys. Proc. 6th
Int. Conf. on Aluminum Weldments, Cleveland, AWS, 1995; 309 – 323.
79. Riches, S. Laser welding gathers pace. Met. Working (Australia) 1999, 14 (2),
42– 44.
80. Huntington, C.A.; Eagar, T.W. Laser welding of aluminum and aluminum alloys.
Weld. J. 1982, 62 (4), 105S – 107S.
81. Forsman, T.; Kaplan, A.F.H.; Powell, J.; Magnusson, C. Nd:YAG laser welding of
aluminum: factors affecting absorptivity. Lasers Eng. 1999, 8, 295 – 309.
82. Richerson, D. Modern Ceramic Engineering; Dekker: New York, 1982.
83. Blundell, N.; Biffin, J.; Johnson, T.; Page, C. High speed, augmented laser welding
of thin sheet metals. 5th Int. Conf.: Trends in Welding Research, ASM International,
Pine Mountain, GA, USA, 1998; 483 –487.

You might also like