JMEPEG (2006) 15:23-31 ©ASM International
DOI: 10.1361/105994906X83385 1059-9495/$19.00
Laser Welding of 3 mm Thick
Laser-Cut AISI 304 Stainless Steel Sheet
Harish Kumar, P. Ganesh, Rakesh Kaul, B. Tirumala Rao, Pragya Tiwari, Ranjeet Brajpuriya, S.M. Chaudhari, and A.K. Nath
(Submitted November 11, 2004; in revised form August 14, 2005)
The objective of the present work was to study the laser weldability of laser-cut 3 mm thick AISI 304
austenitic stainless steel sheet (using oxygen as an assist gas). For minimizing heat input during laser
cutting, which is an important factor influencing the thickness of the oxide layer on the cut surface, laser
cutting was performed in pulsed mode. The results of the study demonstrated that although the laser
welding of laser-cut specimens did not result in the formation of weld defects, the resultant laser weldments
exhibited reduced ductility with respect to base metal and bead-on-plate laser weldments. Laser-cut and
laser-welded specimens also displayed higher notch sensitivity than the base metal. However, laser-cut and
laser-welded specimens still possessed enough ductility to pass guided bend tests.
Keywords austenitic stainless steel, CO2 laser, laser cutting, irradiated metal to a high-pressure gas jet coming out of the
laser welding cutting nozzle. For laser cutting, the focal point of the LB is set
very close to the nozzle opening. The mechanism of laser cut-
ting involves the rapid laser melting and blowing out of molten
metal with a high-pressure gas jet. Laser-cutting efficiency in
1. Introduction ferrous materials is enhanced by using oxygen as an assist gas.
The use of oxygen during laser cutting provides an additional
High-power CO2 lasers are finding extensive industrial ap- heat source in the form of an exothermic reaction involving the
plication in metal sheet cutting. Cutting with a laser beam (LB), oxidation of iron, and so the formed iron oxide, by virtue of its
a highly focused source of energy, offers a high degree of low adherence, is easily blown out of the cut front. In the laser
precision, a narrow kerf width, a low roughness of the cut cutting of mild steel sheets, 60 to 70% of the total energy
surfaces, and a minimum heat-affected zone. A laser, inte- requirement is met by the oxidation energy (Ref 2). The use of
grated with a computer-controlled workstation, facilitates the oxygen thus facilitates cutting at higher speeds. Ivarson et al.
easy cutting of complex shapes. Often, laser-cut sheets are (Ref 3, 4) reported that in the laser cutting of mild steel, about
required to be welded. Nielsen and Broden (Ref 1) studied the 50% of the ejected iron gets oxidized. FeO is the major con-
weldability of laser-cut austenitic stainless steel (ASS) sheets stituent (97.6 mol.%) of the ejected oxides, while Fe2O3 ac-
with gas tungsten arc welding. He reported that stainless steel counts for the rest (2.4 mol.%). The oxidation reaction in-
(SS) sheets, laser-cut with 100% oxygen, were not weldable. volved in the formation of FeO is as follows:
On the other hand, SS sheets, cut with a gas mixture of
60% helium and 40% oxygen, produced sound welds. But laser Fe + 1 Ⲑ 2 O2 → FeO ⌬H = −257.58 kJ Ⲑ mol at 2000 K
cutting with the gas mixture brought about a 25% reduction (Ref 5)
in the cutting speed (Ref 1). The work did not address me-
chanical property characterization of the resultant welds. The The laser cutting of ASS sheets is usually carried out with
present work was undertaken with the objective of studying either oxygen or an inert gas. Oxygen-assisted cutting is the
the laser weldability of laser-cut 3 mm thick AISI 304 ASS more common method. The use of oxygen in cutting SS sheet
sheets using oxygen as the cutting gas, and to characterize the involves the formation of refractory chromium oxide (melting
microstructural and mechanical properties of the resultant temperature 2453 K), which poses difficulty in its efficient
weldment. ejection from the cut front (Ref 6). The chemical reactions
taking place during the laser cutting of SS are more complex
than the simple Fe → FeO reaction involved in the case of mild
2. Background steel cutting. In this case, Fe, Cr, and Ni all oxidize simulta-
neously to form oxides according to the following reactions
The laser cutting of metal sheets is carried out by scanning (Ref 7):
the surface of the sheets with a focused LB while subjecting the
2Fe + 3 Ⲑ 2 O2 → Fe2O3 ⌬H = −826.72 kJ Ⲑ mol at 2000 K
Harish Kumar, P. Ganesh, Rakesh Kaul, B. Tirumala Rao, Pragya
Tiwari, and A.K. Nath, Industrial CO2 Laser Section, Synchrotron 2Cr + 3 Ⲑ 2 O2 → Cr2O3 ⌬H = −1163.67 kJ Ⲑ mol at 2000 K
Utilization Division, Centre for Advanced Technology, P.O.: CAT,
Indore, 452 013 India; and Ranjeet Brajpuriya and S.M. Chaudhari, Ni + 1 Ⲑ 2 O2 → NiO ⌬H = −248.23 kJ Ⲑ mol at 2000 K
University Grants Commission-Department of Atomic Energy Con-
sortium for Scientific Research, Indore, 452 017 India. Contact e-mail: It has been estimated that the respective concentrations of
[email protected]. Fe2O3, Cr2O3, and NiO in the ejected oxides are 78.4, 19.2, and
Journal of Materials Engineering and Performance Volume 15(1) February 2006—23
Table 1 Chemical composition of austenitic stainless
steel sheet (in wt.%)
C Cr Ni Mn Si Mo Ti S P
0.032 19.37 8.84 1.66 0.50 0.33 0.003 0.021 0.037
Table 2 Parameters of laser cutting and laser welding
Laser cutting
Average laser Frequency Duty cycle Cutting speed Gas Pressure
power
240 W 300 Hz 40% 5 mm/s Oxygen 0.5 MPa
Laser welding
Laser power Welding speed Shroud gas Fig. 1 Low-magnification view of the laser-cut surface of the SS
2.3 kW 12 mm/s Ar sheet: (right) top edge; (left) bottom edge
2.4 mol%, respectively. However, the oxides constituted about
35.5 mol% of the total amount of the ejected material (Ref 8).
In the applications demanding the welding of laser-cut
sheets, the weldability of the cut sheets is significantly influ-
enced by thickness and the chemistry of the oxide layer left on
the cut surface and by the roughness of the cut edge. An al-
ternative for avoiding the oxide layer on the laser-cut edges is
to use nitrogen as the assist gas. However, it results in a 30 to
60% reduction in the cutting speed (Ref 8, 9). Recent work
carried out in the authors’ laboratory demonstrated that the
laser cutting of 1 mm thick silicon steel sheets in pulse mode
(with oxygen as the cutting gas) yields improved quality of the
cut surface compared to that produced by cutting in continuous
wave (CW) mode (Ref 10).
3. Experimental
The current study was carried out on a 3 mm thick AISI 304
ASS sheet. Table 1 presents the chemical composition (in
wt.%) of the SS sheet. Laser cutting was performed with an
indigenously developed 2.5 kW transversely excited transverse
flow CO2 laser system (Ref 11). The laser can be operated in
CW mode as well as in pulsed mode. The laser yields a maxi-
mum output power of 750 W in TEM01* mode. TEM01* mode
was obtained by incorporating a suitable intracavity aperture.
The plane-polarized LB coming out of the laser cavity was
made circularly polarized with a phase-retarding mirror. The
circularly polarized LB was subsequently focused with a me-
niscus zinc selenide (ZnSe) lens of 100 mm focal length. The Fig. 2 (a) Parallel fibers near the top edge of the laser-cut surface. (b)
focusing lens was housed in a coaxial copper nozzle (exit di- The EDS spectrum of one of the fibers shown in (a)
ameter 1.5 mm) through which the assist gas (oxygen) was
blown. The focal point of the LB (typical spot diameter 200
m) was located at a distance of about 1 mm from the nozzle beam delivery system and a computer-controlled three-axis
tip. To control heat input during laser cutting, an important workstation. The raw LB, emanating out of the laser system,
parameter controlling surface roughness and thickness of the was folded with a 45° plane gold-coated copper mirror, and the
oxide layer on the cut surface (which, in turn, affects the weld- folded LB was subsequently focused with a 100 mm focal
ability of laser-cut sheets), the process was performed in pulsed length ZnSe lens. The focusing lens was mounted in a water-
mode. cooled copper nozzle. Autogenous laser welding was per-
The resultant laser-cut SS sheets were subsequently butt- formed by scanning the sheet interface with the focused LB.
welded with an indigenously developed 10 kW transversely During the course of the laser welding, argon gas was passed
excited transverse flow CW CO2 laser (Ref 12, 13). The laser- through the nozzle that served the dual purpose of protecting
welding setup consisted of the laser system integrated with a the expensive ZnSe lens from possible particulate emissions
24—Volume 15(1) February 2006 Journal of Materials Engineering and Performance
Fig. 3 Cross section of the laser-cut SS sheet. OL, oxide layer; RL, resolidified layer
the right side of Fig. 1, most of the cut surface was largely
featureless. Near the top edge, the cut surface carried regular
pockets of parallel fibers, as shown in Fig. 2(a). The EDS
analysis identified the fibers as Cr-rich oxides, as shown in Fig.
2(b). A few gas porosities were also noticed near the bottom
edge of the laser-cut surface.
It is believed that parallel fibers represent chromium-rich
oxides that are formed on the elongated inclusions near the top
surface where heat input was too low to facilitate the efficient
ejection of refractory chromium oxide. The mechanism in-
volved during the initial part of laser cutting is the formation of
a depressed melt pool under the influence of gas jet. The cavity
so formed rapidly deepens and results in a keyhole. At this
stage, the mechanism of material removal involves material
evaporation, which proceeds until the LB moves past the re-
Fig. 4 The FZ microstructure of the LCW specimen. The inset shows gion under consideration. Due to the very high temperatures
the macroscopic view of the weldment. involved during the intermediary stage of laser cutting, chro-
mium oxide particles were not found in this region. In the final
stage of laser cutting, the molten metal is dragged out of the cut
from the substrate being welded as well as shielding the sub- front by the force of the assist gas (Ref 14).
strate being welded against oxidation. In addition to the laser
welding of laser-cut SS sheets, some bead-on-plate laser-
welded SS specimens were also prepared. Table 2 presents
4.3 Metallographic Examination
experimental parameters for laser cutting and laser welding.
Laser-cut, laser-welded, as well as base metal specimens The laser cutting of the ASS sheet produced a kerf width
were characterized by surface roughness measurement, x-ray of 350 to 400 m. Metallographic examination of the trans-
photoelectron spectroscopy (XPS), optical microscopy, micro- verse cross sections of the laser-cut specimens revealed a
hardness measurement, scanning electron microscopy (SEM), thin-melted and resolidified layer below the cut edge. The
energy-dispersive spectroscopy (EDS), tensile testing of layer was covered with a thin layer of oxide. The average
smooth and notched specimens, and guided bend test. In the thickness of the melted and resolidified layer registered an
subsequent part of the article, “laser-cut and laser-welded” increase toward the bottom edge of the SS sheet. The thickness
specimens are referred to as LCW, “bead-on-plate laser- of the melted and resolidified layer was 10 to 40 m near
welded” specimens are referred to as LW, and “base metal ” the top edge, while it increased to 50 to 80 m near the bottom
specimens are referred as BM. edge of the laser-cut sheet. Figure 3 presents a magnified
view of the transverse cross section of the laser-cut SS sheet
4. Results and Discussion specimen.
Both kinds of laser weldments (i.e., LCW and LW) exhib-
4.1 Surface Roughness Measurement ited full penetration. No defects were found in the weldments.
The surface roughness measurement of (the cut surface of) The fusion zone (FZ) of the two welds exhibited the primary
laser-cut SS sheet measured with a Surtronic 3+ (Taylor Hob- ferrite mode of solidification. Figure 4 presents the FZ micro-
son Ltd., UK) roughness tester revealed an Ra value of about structure of the LCW specimen. Figure 4 carries the macro-
5 m. scopic view of the laser weldment in the inset. A closer look at
the FZ microstructure revealed a largely uniform dispersion of
fine inclusions, as shown in Fig. 5(a). The EDS analysis of the
4.2 Scanning Electron Microscopy Examination of
inclusions confirmed the presence of C and O along with Fe,
Laser-Cut Surface
Cr, Ni, and Mn, as shown in Fig. 5(b). It is believed that the
Laser-cut surfaces were not found to be associated with inclusions were the result of the oxide layer present on the
well-defined characteristic “striations.” Figure 1 presents a laser-cut surfaces. Such inclusions were not found in the FZ of
low-magnification photomicrograph of the laser-cut surface. the LW specimen. Figure 6 presents the FZ microstructure of
Except for a small zone near the top edge, which is shown on the LW specimen.
Journal of Materials Engineering and Performance Volume 15(1) February 2006—25
Fig. 5 (a) The uniformly dispersed fine inclusions (marked with arrows) in the FZ of the LCW specimen. (b) The EDS spectrum of one of the
fine inclusions in the FZ of the LCW specimen
LW and LCW) specimens. The measurements revealed that in
comparison to the largely uniform microhardness profile across
the LW weldment, the LCW specimen exhibited a relatively
higher microhardness (213–271 VHN) in the FZ with respect to
the base metal (204–230 VHN). Figure 7 presents microhard-
ness profiles across the cross sections of laser weldments.
4.5 X-Ray Photoelectron Spectroscopy
The XPS measurements were performed on an as-laser-cut
surface and the FZ of the LCW specimen using workstation of
photo-electron spectroscopy beam line installed at INDUS-1, a
synchrotron radiation source at the Center for Advanced Tech-
nology, Indore, India. The sample for the FZ of the LCW
specimen was extracted by taking out a thin slice of material
along the longitudinal cross section of the weld. The specimens
Fig. 6 The FZ microstructure of the LW specimen for XPS were prepared by the side-by-side placement of four
numbers of the concerned pieces to present a surface area of 10
× 10 mm. All of the measurements were made by using MgK␣
radiation having 1253.6 eV photon energy. At the pass energy
4.4 Microhardness Measurement
of 50 eV, the measured resolution was better than 0.8 eV. The
Microhardness measurements (load 0.981 N) were per- binding energies (BEs) of the observed peaks were determined
formed on the transverse cross section of laser-welded (i.e., using an internal reference (defined as C-1s) at 284.7 eV and
26—Volume 15(1) February 2006 Journal of Materials Engineering and Performance
Fig. 7 Microhardness profile across the cross section of laser-welded
specimens
Fig. 9 (a) Detailed XPS spectrum of O-1s signals of the laser-cut SS
surface. (b) Detailed XPS spectrum of O-1s signals of the FZ of the
LCW specimen
proportion of Fe2O3 and Cr2O3 was found to be 47:53 (atomic).
Fig. 8 The XPS survey scans of the laser-cut surface and FZ of the
LCW specimen This value matches quite well with the 48:52 (atomic) figure
reported by Powell and Menzies (Ref 7). The observed O-1s
spectrum for the FZ of the LCW specimen was quite similar to
external references as Au-4f7/2 and Au-4f5/2 at 83.8 eV and
that of the laser-cut surface, but with a marginally different
87.5 eV, respectively. Before taking test measurements, the
Fe2O3-to-Cr2O3 ratio of 41:59 (atomic). The similar natures of
specimens were sputtered with Ar+ ions for 10 min to remove
the O-1s spectra of the laser-cut surface and the FZ of the LCW
surface contamination. The survey scans of both of the speci-
specimen indicate that the nature of the oxides remained un-
mens are shown in Fig. 8. The survey scan exhibited intense
changed during the laser welding of laser-cut sheets. On the
peaks of C-1s (BE 285.2 eV) and O-1s (BE 530.7 eV). In
basis of the results of the metallographic examination and the
addition, smaller peaks of Fe 2p and Cr 2p were also detected
XPS, it is evident that during the course of laser welding the
in the survey scan. The narrow scans of O-1s and C-1s peaks
surface oxides were significantly refined and dispersed in the FZ.
were also separately recorded, and are shown in Fig. 9 and 10,
Figure 10 presents C-1s spectra for the two specimens ana-
respectively. Figure 9(a) and (b) present the detailed XPS spec-
lyzed. The peak position in the two spectra matched well with
tra of O-1s for the laser-cut specimen as well as the FZ of the
that of C-1s, indicating that the C was largely present in the
LCW specimen. The observed peaks were broad and were
elemental form.
deconvoluted by using standard XPSpeak41 software (Ref. 15).
The spectra were resolved into their components after subtract-
4.6 Tensile Test
ing the Shirley background (Ref 16). The O-1s spectrum for the
laser-cut surface was resolved in two components: Cr2O3 and Laser-welded tensile test specimens were fabricated as per
Fe2O3. On the basis of the relative area under the peaks, the the Boiler and Pressure Vessel Code (section VIII) (Ref 17).
Journal of Materials Engineering and Performance Volume 15(1) February 2006—27
The test specimen carried laser welds in the center of its gage specimens are indicative of lower ductility of these specimens
length. In addition to the laser-welded specimens (LCW and with respect to BM.
LW), the tensile tests were also performed on BM specimens of
the same dimensions. All of the specimens exhibited largely
similar tensile and yield strength values, although the LCW
4.7 Notch Tensile Test
specimens exhibited lower ductility than the BM and LW Notch tensile tests were performed with the objective of
specimens. All laser-welded specimens suffered failure in the comparing the notch sensitivity of the FZ of LCW specimens
FZ. The results of tensile tests are summarized in Table 3. with respect to BM. Figure 12 presents the drawing of the test
The SEM fractographic examination of the failed specimens specimen. Notch tensile testing of LCW and BM specimens
revealed dimpled surfaces reflecting the ductile nature of the magnified the difference between their ductility. Careful ob-
failure. The average size of the dimples was significantly servation of the specimens during the course of testing revealed
coarser in the BM specimen than that in laser-welded speci- that in BM specimens gross plastic deformation was introduced
mens. Figure 11 compares the fractographic features of BM (causing crack tip blunting) at an earlier stage than that in
and LCW specimens at the same magnification. Ductile failure the LCW specimens. In these specimens, the crack extended
of metals/alloys involves the formation of microvoids around with two zones of plastic deformation (appearing in the form
inclusions, their growth, and finally their coalescence to bring of relief on the surface of the specimens under testing), inclined
about fracture, giving rise to the dimpled nature of the fractured at about 45° with respect to the direction of crack extension.
surface. Microvoids are formed as a result of particle-matrix On the other hand, in LCW specimens the inclination of
decohesion or the cracking of second-phase particles. The pro- the plastically deformed zones (with respect to the direction
cess of microvoid growth involves considerable localized plas- of crack extension), which developed ahead of the crack
tic deformation and requires the expenditure of large amounts tip, was much lower. The BM specimens exhibited greater
of energy (Ref 18). Therefore, finer dimples in laser-welded notch tensile strength and ductility than those of the LCW
Table 3 Results of tensile testing
Yield Tensile Elongation
strength, strength, (gauge length = Failure
Specimen MPa MPa 15 mm), % location
Unnotched
BM 346 656 75 …
305 631 68 …
304 630 69 …
LCW 353 631 56.5 Weld
356 637 43.5 Weld
335 619 49.5 Weld
355 635 55 Weld
LW 345 634 59 Weld
325 620 62 Weld
349 621 65.6 Weld
Notch tensile Reduction Failure
Specimen strength, MPa in area, % location NSR
Notched
BM 564 62 … 0.86-0.90
562 56 …
Fig. 10 Detailed XPS spectra of C-1s signals of the laser-cut SS LCW 462 29 Weld 0.71-0.75
surface and the FZ of the LCW specimen 454 32 Weld
Fig. 11 Comparison of dimpled fracture surfaces of tensile tested (a) BM specimen and (b) LCW specimen
28—Volume 15(1) February 2006 Journal of Materials Engineering and Performance
Fig. 12 Drawing of a notch tensile test specimen
specimens. The LCW specimens displayed a lower notch
strength ratio (NSR; 0.71–0.75) than that of BM specimens
(0.86–0.9). The NSR is defined as the ratio of the notch tensile
strength to the tensile strength of the unnotched specimen
(Ref 19). The results of notch tensile tests are summarized in
Table 3.
The SEM fractographic examination of notch-tensile tested
LCW and BM specimens revealed sharp differences in their
surface morphologies near the notch root. In the BM specimen,
a narrow “smooth zone” was noticed at the roots of the two
notches, as shown in Fig. 13(a). The smooth zones were in-
clined with respect to the rest of the fractured surface and were
associated with elongated dimples (Fig. 13b), indicating the
influence of shear stresses near the notch root (Ref 20). The
rest of the fractured surface exhibited coarse equiaxed dimples.
On the other hand, in LCW specimens, no such smooth zones
were noticed at the roots of either of the two notches. Figure 14
presents a low-magnification SEM fractograph of the LCW
specimen showing the absence of the smooth zone near the
notch root. The whole fractured surface of the specimen dis- Fig. 13 (a) Low-magnification view of the fracture surface of notch
played finer equiaxed dimples, signifying a lower degree of tensile-tested BM specimen near the notch root. Note a thin smooth
plastic deformation (preceding failure) with respect to the BM zone near the notch root. (b) Elongated dimples on the smooth zone
specimen. A metallographic examination of the top surface of shown in (a). S, smooth zone; N, notch root
the fractured BM specimen exhibited a fine “V-notch” region
at the root of the machined notch, which corresponded to the
smooth zone noticed on the fractured surface, as shown in Fig.
13(a). The material around the V-notch region exhibited sig-
natures of plastic deformation, as shown in Fig. 15. The de-
velopment of a fine V-notch region during the initial process of
failure signifies fracture along planes at 45° to the stress-axis.
During the subsequent period of the failure, the specimen frac-
tured along the planes normal to the direction of applied stress.
In contrast to the BM specimens, no such features were noticed
in the LCW specimens.
Notch tensile tests demonstrated that, in contrast to LCW
specimens, BM specimens developed plane stress conditions at
the notch-root, which facilitated deformation on the planes of
maximum shear stress (at 45° to the stress-axis), causing blunt-
ing of the notch root (Ref 20). The argument is supported by
the development of fine V-notch region (with signs of plastic
deformation around it) at the root of the notch. The relatively Fig. 14 Low-magnification view of the fracture surface of notch
lower value of the NSR of LCW specimens (with respect to tensile-tested LCW specimen near the notch root (N)
that of BM specimens) also indicates that due to the limited
ability of the FZ of the LCW specimens to plastically deform
and blunt the crack tip, stress concentration effects were more with respect to BM is attributed to the presence of finely dis-
dominant and brought about the failure of these specimens at persed oxide inclusions in the FZ of LCW specimens, which
lower applied stress. This means higher notch sensitivity of the brought about an increase in the hardness of the material (al-
FZ of LCW with respect to BM (Ref 20). The higher notch though the rise is not very large) and suppressed plastic defor-
sensitivity and lower ductility of the FZ of the LCW specimens mation at the crack tip.
Journal of Materials Engineering and Performance Volume 15(1) February 2006—29
Fig. 15 Top surface of notch tensile-tested BM specimen showing a fine V-notch (encircled) at the root of the machined notch. The inset shows
signs of plastic deformation around the V-notch region.
laser-cut surface in the FZ of LCW weldment. These welds
also carried higher notch sensitivity than the base metal. How-
ever, laser-cut and laser-welded specimens were still ductile
enough to pass the guided bend test.
Acknowledgments
The authors thank Dr. D.M. Phase and Mr. Vinay Ahire of
the University Grants Commission-Department of Atomic En-
ergy Consortium for Scientific Research, Indore, for EDS
analysis. They also wish to thank Mr. Ram Nihal Ram, Mr.
C.H. Prem Singh, and Mr. Anil Adbol for their assistance at
Fig. 16 Transverse cross section of guided bend-tested LCW speci- various stages of this work.
men
References
4.8 Guided Bend Test 1. S.E. Nielsen and G. Broden, Improved Weldability of Stainless Steel
Cut by Laser, Power Beam Technology, J.D. Russell, Ed., September
The LCW specimens were also subjected to a transverse- 10-12, 1986 (Brighton, U.K.), The Welding Institute, Proc. Int. Conf.,
face guided bend test conducted as per ASTM E 190-92 1987, p 256-267
(Ref 21). Three of these specimens were slowly bent to 180° 2. P.A. Molian, Laser Cutting of Thick Metallic Solids–A Reactive Gas
in such a way that the top weld face became the convex Flow Approach, Laser Advanced Materials Processing (LAMP)
(Osaka, Japan), High Temperature Society of Japan and Japan Laser
surface. A macroscopic examination of the convex surface of Processing Society, Proc. Int. Conf., May 21-23, 1987, p 245-250
the bent specimens revealed a smooth crack-free surface. A 3. A. Ivarson, J. Powell, and C. Magnusson, The Role of Oxidation in
metallographic examination of the transverse cross section of Laser Cutting Stainless Steel and Mild Steel, J. Laser Appl., 1991, 3
the bent specimens did not reveal any defects. Figure 16 (3), p 41-45
presents a magnified view of the cross section of the bent LCW 4. A. Ivarson, J. Powell, and C. Magnusson, Laser Cutting of Steels:
Analysis of the Particles Ejected during Laser Cutting, Weld. World,
specimen. 1992, 30 (5/6), p 116 -125
5. W. Bolton, Newnes Engineering Materials Pocket Book, Heinemann-
Newnes, 1990
5. Conclusions 6. W.M. Steen, Laser Material Processing, 2nd ed., Springer Verlag,
1998, p 115-116
The results of the study demonstrated that laser welding can 7. J. Powell and I. Menzies, Metallurgical Implications of Laser Cutting
Stainless Steel: Power Beam Technology, J.D. Russell, Ed., September
produce sound butt welds between 3 mm thick AISI 304 SS 10-12, 1986 (Brighton, U.K.), The Welding Institute, Proc. Int. Conf.
sheets that have been laser-cut using oxygen as an assist gas. 1987, p 269-284
For the improved weldability of laser-cut sheets, laser cutting 8. J. Powell, CO2 Laser Cutting, Springer-Verlag, 1993, p 57-65, 211-
was performed in pulsed mode with controlled heat input. The 215
welds of the resultant laser-cut sheets possessed largely similar 9. J. Powell, Laser Cutting of Metals: LIA Handbook of Laser Materials
Processing, 1st ed., J.F. Ready, Ed., Magnolia Publishing Inc., 2001,
ultimate tensile strength values as that of the base metal. The p 443-444
relatively lower ductility of the LCW specimen with respect to 10. B.T. Rao, R.K. Sinha, and A.K. Nath, Optimization of Si-steel Cutting
the BM and LW specimens is attributed to finely dispersed with CO2 Laser, Metals Mater. Proc., 2002, 14 (2), p 145-154
oxide inclusions, arising out of the oxide layer present on the 11. A.K. Nath, L. Abhinandan, and P. Choudhary, Characteristics of a
30—Volume 15(1) February 2006 Journal of Materials Engineering and Performance
Pulse-sustained dc-Excited Transverse-Flow cw CO2 Laser of 1.5-kW 16. D.A. Shirley, High-resolution X-Ray Photoemission Spectrum of the
Output Power, Opt. Eng., 1994, 33 (6), p 1889-1893 Valence Bands of Gold, Phys. Rev. B: Condens. Matter, 1972, 5, p
12. C.P. Paul, H. Kumar, T. Reghu, P. Bhargava, and A.K. Nath, Enhance- 4709-4714
ment of Output Power in a 5 kW Transverse Flow CW CO2 Laser, 17. Boiler and Pressure Vessel Code, Section IX: Welding and Brazing
DAE-BRNS National Laser Symposium, Allied Publishers Ltd., De- Qualification, ASME, 1986, p QW-151, QW-462
cember 2001, p 51-52 18. J.R. Davis, Ed., ASM Handbook, 2nd ed. ASM International, 1998, p
13. J. Khare, R. Sridhar, C.P. Paul, T. Reghu, and A.K. Nath, Operational 1214
Characteristics and Power Scaling of a Transverse Flow Transversely 19. G.E. Dieter, Mechanical Metallurgy, McGraw-Hill, 1988, p 314-316
Excited CW CO2 Laser, Pramana, 2003, 60 (1), p 99-107 20. R.W. Hertzberg, Deformation and Fracture Mechanics of Engineering
14. A. Bharti and R. Sivakumar, The Mechanism of Material Removal in Materials, 4th ed., John Wiley & Sons, Inc., 1996, p 295-297, 342-345
Laser Fusion Cutting, Lasers Eng., 1996, 5 (2), p 87-105 21. “Standard Method for Guided Bend Test for Ductility of Welds, ” E
15. XPS analysis data fitting shareware program; www.icg.nsrrc.org.tw 190-92, Annual Book of ASTM Standards, ASTM, 1992
Journal of Materials Engineering and Performance Volume 15(1) February 2006—31