10.12.2012 Views

Laser cutting of metallic coated sheet steels - World Lasers, Inc.

Laser cutting of metallic coated sheet steels - World Lasers, Inc.

Laser cutting of metallic coated sheet steels - World Lasers, Inc.

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Abstract<br />

Journal <strong>of</strong> Materials Processing Technology 74 (1998) 234–242<br />

<strong>Laser</strong> <strong>cutting</strong> <strong>of</strong> <strong>metallic</strong> <strong>coated</strong> <strong>sheet</strong> <strong>steels</strong><br />

G.V.S. Prasad, E. Siores, W.C.K. Wong *<br />

School <strong>of</strong> Mechanical and Manufacturing Engineering, Queensland Uni�ersity <strong>of</strong> Technology, G.P.O. Box 2434, Brisbane,<br />

Qld. 4001, Australia<br />

Received 1 June 1996<br />

This paper discusses the laser-beam machining <strong>of</strong> <strong>metallic</strong> <strong>coated</strong> <strong>sheet</strong> <strong>steels</strong> such as ZINCALUME, ZINCANNEAL and<br />

GALVABOND <strong>of</strong> 1 mm thickness. These materials are essentially zinc and aluminium coatings <strong>of</strong> varying skin thickness on steel.<br />

The experimental work explores some methods for reducing thermal damage <strong>of</strong> the coatings vis-a-vis the parent/base metal which<br />

impose severe machining restrictions by virtue <strong>of</strong> their high reflectivity and thermal conductivity. A 500 W continuous-wave, 10.6<br />

�m CO 2 CNC laser centre was used to improve the cut quality in terms <strong>of</strong> good surface finish, reduced kerf width and dross. An<br />

analytical model was developed to establish the finite-element characteristic <strong>of</strong> the <strong>cutting</strong> process and it has been clarified that<br />

an efficient choice <strong>of</strong> the process parameters is a pre-requisite for minimum thermal damage <strong>of</strong> the coatings. The topographical<br />

characteristics <strong>of</strong> the uncut-through kerf and surface roughness are discussed. Some visualisational experiments were also<br />

performed for further understanding <strong>of</strong> the micro- and macro-mechanics <strong>of</strong> the <strong>cutting</strong> process. It is proven that the <strong>cutting</strong> speed<br />

is a function <strong>of</strong> the input power and that the laser processing <strong>of</strong> these materials is a commercially viable option. © 1998 Published<br />

by Elsevier Science S.A.<br />

Keywords: <strong>Laser</strong> centre; Metallic <strong>coated</strong> <strong>sheet</strong> <strong>steels</strong>; Process parameters<br />

1. Introduction<br />

<strong>Laser</strong>s are used in many industrial machining operations,<br />

especially for processing the ‘difficult-to-machine’<br />

materials. <strong>Laser</strong> machining has several advantages over<br />

conventional methods. First, it is a non-contact process<br />

that eliminates such effects as tool wear, machine vibration<br />

and mechanically induced thermal damage. Second,<br />

laser machining is a thermal process and materials<br />

with favourable thermal properties can be successfully<br />

processed regardless <strong>of</strong> their mechanical properties.<br />

Third, laser machining is a flexible process.<br />

Metallic <strong>coated</strong> <strong>sheet</strong> <strong>steels</strong> have been machined successfully<br />

using conventional equipment such as presses<br />

and guillotines for quite some time now but these<br />

methods have brought with them problems like low<br />

productivity and rapid tool wear although they too are<br />

chipless machining methods. This paper examines the<br />

application <strong>of</strong> a high energy laser beam as a potential<br />

tool for processing these materials.<br />

* Corresponding author. Fax: +61 7 38641529.<br />

0924-0136/99/$19.00 © 1998 Pubished by Elsevier Science S.A. All rights reserved.<br />

PII S0924-0136(97)00276-8<br />

The basic <strong>cutting</strong> mechanism is dependent upon the<br />

formation <strong>of</strong> the radiation trap giving rise to a molten<br />

pool at a localised spot that is then ejected through the<br />

root <strong>of</strong> the workpiece using a suitable assist gas jet.<br />

Previous studies on the laser–material interactions at<br />

the <strong>cutting</strong> zone for metals indicate that properties such<br />

as reflectivity and thermal conductivity dictate the efficiency<br />

<strong>of</strong> the <strong>cutting</strong> process, as most metals are highly<br />

reflective at the laser wave lengths. Due to this, the<br />

coupling <strong>of</strong> the beam and the workpiece is <strong>of</strong>ten inefficient<br />

and very low.<br />

However, the absorption coefficient <strong>of</strong> the material is<br />

a function <strong>of</strong> temperature, which changes during the<br />

transient phase <strong>of</strong> the process. The initial weak absorption<br />

at the surface <strong>of</strong> the workpiece begins to increase<br />

the workpiece temperature directly under the optical<br />

beam and thus decreases the reflectivity quite rapidly.<br />

Temperature and absorption increase until melting and<br />

evaporation temperatures are reached that permit a<br />

keyhole or radiation trap to form at the localised spot.<br />

The laser beam acts as an energetic line heat source<br />

within the material and initialises the <strong>cutting</strong> process.


G.V.S. Prasad et al. / Journal <strong>of</strong> Materials Processing Technology 74 (1998) 234–242 235<br />

Thus, it is evident that for this keyhole-<strong>cutting</strong> process<br />

to be initiated, it is essential that the power density be<br />

high enough to overcome the reflection barrier. Once<br />

this is achieved, the process can be controlled using the<br />

melting and evaporation relationships.<br />

The goal in any laser machining process is to maximise<br />

the material removal rate whilst minimising the<br />

heat affected zone (HAZ). The objectives <strong>of</strong> this experimental<br />

study are: (i) to identify the parameters that<br />

have detrimental influence on the outcome <strong>of</strong> the <strong>cutting</strong><br />

process; (ii) to establish a relationship between<br />

traverse speed and input power; (iii) to examine the<br />

surface quality aspect through an analysis <strong>of</strong> the HAZ<br />

formed during the laser–material interaction by virtue<br />

<strong>of</strong> the oxidation <strong>of</strong> the coatings directly under the<br />

optical beam; and (iv) to analyse the trade-<strong>of</strong>f between<br />

the material removal rate and the HAZ.<br />

2. <strong>Laser</strong>–metal interactions<br />

Although the CO 2 laser <strong>cutting</strong> <strong>of</strong> metals has become<br />

a well established manufacturing process, the processing<br />

<strong>of</strong> <strong>metallic</strong> <strong>coated</strong> <strong>sheet</strong> <strong>steels</strong> is considered as<br />

‘difficult’. These materials are cut at lower speeds and<br />

at thinner maximum sections than for most other<br />

metals. The reasons behind this reduction in the <strong>cutting</strong><br />

efficiency as compared with, for example <strong>steels</strong>, can be<br />

accounted for by examining the physical properties <strong>of</strong><br />

these materials: (i) their reflectivity to the 10.6 �m CO 2<br />

laser radiation is very high, up to 99% at room temperature;<br />

and (ii) their thermal conductivity is approximately<br />

thrice that <strong>of</strong> other metals such as mild <strong>steels</strong>.<br />

The principles <strong>of</strong> laser processing <strong>of</strong> metals suggest<br />

that the <strong>cutting</strong> process depends upon the establishment<br />

<strong>of</strong> a localised area <strong>of</strong> melting and/or evaporation<br />

throughout the depth <strong>of</strong> the workpiece. The melt thus<br />

generated by the focused beam is removed from the cut<br />

zone by the incident gas jet, which is also chemically<br />

reactive with the melt. The chemical reaction most<br />

<strong>of</strong>ten employed is the exothermic oxidation <strong>of</strong> the<br />

<strong>metallic</strong> surface under the heat <strong>of</strong> the laser radiation.<br />

The melt is chemically degraded and the reaction forms<br />

a secondary heat input to the cut zone.<br />

The high reflectivity and thermal conductivity makes<br />

it very difficult to establish a localised molten zone. At<br />

ambient temperatures, all metals have high reflectivity<br />

(�99%) to the incident laser beam. The small proportion<br />

<strong>of</strong> the absorbed light has the effect <strong>of</strong> heating the<br />

area under the beam and the subsequent rise in temperature<br />

is accompanied by a reduction in reflectivity. This<br />

reduction in reflectivity results in further heating until a<br />

highly absorptive molten pool is established.<br />

The small percentage <strong>of</strong> heat that is absorbed by the<br />

material is converted into heat but is quickly dissipated<br />

across the surface <strong>of</strong> the <strong>sheet</strong> by virtue <strong>of</strong> its high<br />

thermal conductivity. As a result <strong>of</strong> the low thermal<br />

input and the rapid dissipation <strong>of</strong> the heat, a highly<br />

absorptive, localised hot spot is established less readily<br />

that in the case <strong>of</strong> other metals. Continuing the comparison<br />

with <strong>steels</strong>, the aluminium exothermic reaction:<br />

(4)Al+(3)O2=(2)AL2O3+1670 kJ/mol (1)<br />

1is far less effective as a heat source in the cut zone than<br />

the similar reaction employed when <strong>cutting</strong> <strong>steels</strong>:<br />

(4)Fe+(3)O2=(2)Fe2O3+822 kJ/mol (2)<br />

The aluminium oxidation reaction is capable <strong>of</strong> generating<br />

more energy than the iron reaction but the<br />

oxide generated forms an impermeable seal on the<br />

surface <strong>of</strong> the underlying aluminium and thereby suppresses<br />

any further reaction with oxygen. During laser<br />

<strong>cutting</strong>, this seal is continuously fractured due to the<br />

turbulent nature <strong>of</strong> the melt flow out <strong>of</strong> the cut zone.<br />

Consequent to this turbulence, the oxidation reaction<br />

can act as a substantial thermal input to the <strong>cutting</strong><br />

process although its contribution is not <strong>of</strong> the same<br />

order <strong>of</strong> magnitude as that <strong>of</strong> the oxidation <strong>of</strong> iron<br />

during the <strong>cutting</strong> <strong>of</strong> <strong>steels</strong>.<br />

The foregoing theoretical aspects were considered in<br />

the experimental investigations conducted on the specimens.<br />

The work was carried out using a Cincinnati<br />

CL-5 CNC <strong>Laser</strong> Centre with the combination <strong>of</strong> high<br />

power modulation and good mode, attaining extremely<br />

high energy densities, so that the problems <strong>of</strong> high<br />

reflectivity and thermal conductivity could be overcome.<br />

3. Experimental methodology<br />

These series <strong>of</strong> experiments were carried out using<br />

the Cincinnati CL-5 CNC <strong>Laser</strong> Centre at different<br />

power inputs with a view to optimise the cut quality.<br />

The machine produces a beam with a wave length in<br />

the range <strong>of</strong> 3×10 −7 –3×10 −3 �m. The beam was<br />

focused using a 127 mm focal length lens and a simple<br />

conical <strong>cutting</strong> nozzle that had a exit diameter <strong>of</strong> 1.7<br />

mm with the nozzle–workpiece stand<strong>of</strong>f distance being<br />

1 mm.<br />

The <strong>cutting</strong> head assembly <strong>of</strong> the machine is designed<br />

such that turning the materials follower changes the<br />

distance from the lens to the workpiece. Adjusting the<br />

material follower thus moves the beam focal point<br />

above or below the material surface. The material<br />

follower rotates (relative to the lens assembly) on<br />

threads that move it vertically 0.05 in. (1.27 mm) per<br />

revolution, over a total range <strong>of</strong> 0.4 in. or eight full<br />

turns.<br />

1 Enthalpies <strong>of</strong> formation at 293 K (for comparison only).


236<br />

G.V.S. Prasad et al. / Journal <strong>of</strong> Materials Processing Technology 74 (1998) 234–242<br />

Oxygen was employed as the assist gas which is also<br />

the primary assist gas specified for the machine, whilst<br />

up to eight different gases can be incorporated into the<br />

experimental design. The flow rates and the operating<br />

pressures <strong>of</strong> the assist gas generally depend on each<br />

specific application. Three process parameters were<br />

identified as important in the <strong>cutting</strong> <strong>of</strong> the specimens<br />

under study viz., input power, <strong>cutting</strong> velocity and the<br />

assist gas pressure.<br />

3.1. Input power<br />

The drastic fluctuation in the melting and evaporation<br />

temperatures <strong>of</strong> the coatings and the parent/base<br />

metal renders the input power one <strong>of</strong> the crucial factors<br />

in achieving optimum machining quality. The initial set<br />

<strong>of</strong> experiments conducted for varying power input in<br />

the range 450–700 W produced deteriorating quality <strong>of</strong><br />

cuts in all <strong>of</strong> the specimens. It was decided to modulate<br />

the power at 500 W and vary the <strong>cutting</strong> velocity and<br />

the assist gas pressure.<br />

3.2. Cutting �elocity<br />

Most laser systems are based on a low pressure<br />

<strong>cutting</strong> head that means that the maximum <strong>cutting</strong><br />

pressure is limited by the optical system in the <strong>cutting</strong><br />

head. Lenses normally made <strong>of</strong> GaAs or ZnSe are<br />

specified to withstand a maximum pressure <strong>of</strong> about 5<br />

bar. It is generally known that the <strong>cutting</strong> velocity<br />

increases with increasing gas pressure. There is a particular<br />

area in which high quality cuts appear. The maximum<br />

velocity is found at pressures <strong>of</strong> around 5 bar.<br />

Investigations indicate that there is a gap between the<br />

theoretically calculated and the experimentally obtainable<br />

<strong>cutting</strong> velocities, which indicate more scope for<br />

improvements.<br />

3.3. Assist gas pressure<br />

The oxygen pressure was increased in the range 5–20<br />

bar. At levels <strong>of</strong> 20 bar, the material begins to act like<br />

a mirror and there is no interaction between the material<br />

and the incident beam. This implies that for a<br />

specific laser power, there is a particular pressure range<br />

within which the material can be processed. The gas<br />

pressure variation was thus limited to a maximum <strong>of</strong> 14<br />

bar. The <strong>cutting</strong> rate as a function <strong>of</strong> input power was<br />

investigated.<br />

The cuts were evaluated in terms <strong>of</strong> fine, good,<br />

acceptable and poor quality. The qualities were optimised<br />

by optimising the focal point position <strong>of</strong> the<br />

focusing optics <strong>of</strong> the laser system.<br />

4. Results and discussion<br />

The basic philosophy behind the discussion is that<br />

the wasted energy, which does not contribute to the<br />

<strong>cutting</strong> process, should not be viewed as an aspect <strong>of</strong><br />

the laser–material interaction. Ignoring the reflected<br />

energy which is not, by definition, an input to the<br />

<strong>cutting</strong> process, the losses by conduction, convection<br />

and radiation can be treated solely as a function <strong>of</strong> the<br />

melt on its surroundings.<br />

A simple analytical model <strong>of</strong> the above can be interpreted<br />

with the help <strong>of</strong> the following energy balance<br />

equation:<br />

Input energy=(energy used in <strong>cutting</strong>)+<br />

(losses by conduction, convection and radiation)<br />

As an initial approximation, assume that the specific<br />

<strong>cutting</strong> energy used to remove a unit volume <strong>of</strong> cut<br />

material is independent <strong>of</strong> the material thickness. The<br />

energy used in <strong>cutting</strong> is, therefore, a function <strong>of</strong> this<br />

specific <strong>cutting</strong> energy multiplied by the volume <strong>of</strong> the<br />

material removed during the <strong>cutting</strong>. The losses by<br />

conduction, convection and radiation are a function <strong>of</strong><br />

the temperature <strong>of</strong> the <strong>cutting</strong> front and its surface area<br />

in contact with its surroundings. Under these conditions,<br />

the energy balance equation can be written as<br />

follows.If a laser power P can cut a line L in time t,<br />

then:<br />

(P−b)t(x/100)=E cutldk+tBdk/2(A+B+C) (3)<br />

where b is the laser power transmitted to the cut zone;<br />

x is the absorptivity <strong>of</strong> the cut zone; E cut is the specific<br />

energy needed to melt and remove a unit volume <strong>of</strong><br />

material from the cut zone; d is the material thickness<br />

and A, B and C are the conductive, radiative and<br />

convective loss functions.<br />

Theory suggests that during <strong>cutting</strong>, it is <strong>of</strong>ten the<br />

case that the trailing edge <strong>of</strong> the cut front does not<br />

extend to the full diameter <strong>of</strong> the incident beam. A<br />

proportion <strong>of</strong> the light therefore passes straight through<br />

the kerf without interacting with the cut front. Consequently,<br />

the absorptivity will partially be much higher<br />

than the theoretical values at ambient room temperature.<br />

This is because the cut zone has its absorptivity<br />

increased as a result <strong>of</strong> the high temperature, the presence<br />

<strong>of</strong> oxides, the shallow angle <strong>of</strong> incidence <strong>of</strong> the<br />

laser beam, the roughness and the absorptive layer <strong>of</strong><br />

vapour.<br />

The specific <strong>cutting</strong> energy can be assumed constant<br />

for any given material as all cuts appear similar and<br />

thus can be thought <strong>of</strong> as being produced by similar<br />

mechanisms. Based on this assumption, the average<br />

<strong>cutting</strong> temperature can also be assumed to remain<br />

constant for a given material. Considering the losses


G.V.S. Prasad et al. / Journal <strong>of</strong> Materials Processing Technology 74 (1998) 234–242 237<br />

indicated in the above equation, the conductive losses<br />

per unit area <strong>of</strong> <strong>cutting</strong> front can again be assumed to<br />

be constant for a given material, so that while the<br />

conductive heat loss is generally determined by the<br />

temperature <strong>of</strong> the heat sink and the heat source, this<br />

factor will not interfere with the central idea <strong>of</strong> this<br />

discussion.<br />

The foregoing will also conveniently imply that the<br />

convective and radiative heat losses per unit area can be<br />

approximated to be proportional to the surface area <strong>of</strong><br />

the front. It follows that in the equation, the energy<br />

used in <strong>cutting</strong> is independent <strong>of</strong> the <strong>cutting</strong> time. The<br />

losses will then be proportional to the <strong>cutting</strong> time, so<br />

Fig. 1. Graphs <strong>of</strong> input power (W) vs. <strong>cutting</strong> rate (mm min 1 ): (a)<br />

GALVABOND; (b) ZINCANNEAL; (c) ZINCALUME.<br />

that the proportion <strong>of</strong> the useful and the wasted energy<br />

will change if the <strong>cutting</strong> speed is changed in order to<br />

cut materials <strong>of</strong> different thickness.<br />

4.1. Effect <strong>of</strong> material thickness on <strong>cutting</strong> speeds<br />

In the equation, suppose that d is halved at the same<br />

laser power input:<br />

(P−b)(t/2)(x/100)=E cutlk(d/2)+tBdk/4(A+B+C)<br />

(4)<br />

For the sake <strong>of</strong> comparison, let everything be doubled<br />

in Eq. (4):<br />

(P−b)t(x/100)=E cutldk+(t/2)Bdk(A+B+C) (5)<br />

It is clear that the imbalance in the equation with<br />

respect to Eq. (3) is that the losses have been halved.<br />

Thus, the equation can be balanced by simply manipulating<br />

t.<br />

The foregoing clearly implies that there is a specific<br />

limit to the material thickness beyond which the <strong>cutting</strong><br />

mechanism breaks down and cannot be re-established<br />

at any <strong>cutting</strong> speed. The reason for this is the relative<br />

increase in thermal losses from the cut zone as the<br />

<strong>cutting</strong> speed is decreased. In the case <strong>of</strong> the machining<br />

<strong>of</strong> the materials under study, the thermal conductivity<br />

<strong>of</strong> the coatings dictates the <strong>cutting</strong> speed although the<br />

material thickness is not the primary concern.<br />

The effects <strong>of</strong> the rapid oxidation reactions have to<br />

be considered in determining the optimum selection <strong>of</strong><br />

the input power and the proportional increase in the<br />

<strong>cutting</strong> speeds. In this case, it follows that with an<br />

increase in the material thickness, there is a proportional<br />

increase in the energy wasted, but this is more<br />

gradual, due mainly to the dissipation <strong>of</strong> the energy<br />

across the material surface by virtue <strong>of</strong> high thermal<br />

conductivity. In other words, there is less concentration<br />

<strong>of</strong> energy absorbed in any particular region across the<br />

material surface.<br />

4.2. Impact <strong>of</strong> the incident beam on the surface <strong>of</strong> the<br />

material<br />

Fig. 1 shows graphs plotted for input power versus<br />

<strong>cutting</strong> rates. It can be inferred that GALVABOND<br />

specimens are cut faster than the ZINCALUME and<br />

ZINCANNEAL specimens. This is due to the aluminium<br />

coating being highly absorptive at the laser<br />

wavelength <strong>of</strong> 10.6 �m. The oxides thus formed are<br />

firmly bonded to the substrate and are not vaporised by<br />

the incident beam, owing to their high melting and<br />

boiling points. This highly absorptive and refractory<br />

surface replaces the original aluminium surface and so<br />

the problems associated with reflection are minimised.


238<br />

G.V.S. Prasad et al. / Journal <strong>of</strong> Materials Processing Technology 74 (1998) 234–242<br />

Fig. 2. Comparison <strong>of</strong> kerf widths: (a) GALVABOND; (b) ZINCALUME; (c) ZINCANNEAL.<br />

4.3. Influence <strong>of</strong> the assist gas on <strong>cutting</strong><br />

The photographs in Fig. 4 show the effect <strong>of</strong> the<br />

<strong>cutting</strong> gas on the surface <strong>of</strong> the various specimens.<br />

There is greater/pronounced surface disintegration in<br />

the case <strong>of</strong> GALVABOND specimens as compared<br />

with the others. This is indicated by the distinct oxide<br />

formation along the length <strong>of</strong> the cut. The cut edge<br />

surfaces represent the extreme edge <strong>of</strong> the molten cut<br />

zone which is then left behind by the <strong>cutting</strong> process. In<br />

this region, the melt is in contact with the parent/base<br />

metal, where the temperature is not greatly in excess <strong>of</strong><br />

the melting point <strong>of</strong> the base metal.<br />

This low temperature melt has a higher surface tension<br />

than the much hotter top surface <strong>of</strong> the coating at<br />

the centre <strong>of</strong> the laser–material interaction zone. This<br />

surface tension gradient acts to draw the molten material<br />

towards the sides <strong>of</strong> the cut whilst it is at the same<br />

time being propelled vertically downwards by the impinging<br />

gas jet. In this way, the molten materials can<br />

accumulate on the bottom <strong>of</strong> the cut edge and will<br />

thereafter solidify as dross. Further, the melt zone is<br />

covered with the oxide coating, which tends to increase<br />

the overall melt surface tension. Also, experience from<br />

brazing has shown that a hotter surface tends to attract<br />

the melt more efficiently hence the hotter, the lower the<br />

surface tension.<br />

High surface tension forces tend to restrict the surface<br />

geometry <strong>of</strong> fluid to large radii. This reduction in<br />

the melt surface tension might have been expected to<br />

accelerate the <strong>cutting</strong> process. The gas moving through<br />

the cut zone acts primarily as a mechanical propellant<br />

<strong>of</strong> the liquid metal out <strong>of</strong> the cut zone. Chemically, it<br />

can possibly also act as a source <strong>of</strong> energy if oxygen is<br />

input to the <strong>cutting</strong> process but it must be borne in<br />

mind that the gas also serves to refrigerate the <strong>cutting</strong><br />

zone by forced convective cooling. Using this interpretation,<br />

it can be predicted that the higher specific heat<br />

and thermal conductivity <strong>of</strong> some other assist gas such<br />

as nitrogen or helium should render it more effective as<br />

a means <strong>of</strong> cooling the laser melt zone.<br />

4.4. Kerf width analysis<br />

The kerf width <strong>of</strong> all the cuts carried out for this<br />

experimental programme varied only slightly about an


G.V.S. Prasad et al. / Journal <strong>of</strong> Materials Processing Technology 74 (1998) 234–242 239<br />

Fig. 3. Microphotographs <strong>of</strong> the transverse edge: (a) GALVABOND; (b) ZINCANNEAL; (c) ZINCALUME.<br />

average value <strong>of</strong> 250 �m with the ZINCANNEAL and<br />

ZINCALUME specimens occupying the range 220–250<br />

�m and the GALVABOND specimens in the 250–270<br />

�m range (see Fig. 2). This is in close conformity with<br />

the studies undertaken by other researchers for similar<br />

metals and most <strong>of</strong> the results reported thus far exhibit<br />

this tendency towards uniformity <strong>of</strong> the kerf width.<br />

This almost independence <strong>of</strong> the kerf width with respect<br />

to <strong>cutting</strong> speed, material thickness or the type <strong>of</strong> assist<br />

gas used, is reminiscent <strong>of</strong> mechanical <strong>cutting</strong> methods<br />

and it can be postulated that for a particular combination<br />

<strong>of</strong> laser–lens–metal, the focused laser assumed an<br />

effective width which is not necessarily changed by<br />

altering the process parameters. The metal itself determines<br />

this width as a result <strong>of</strong> its high thermal conductivity,<br />

which effectively cools all the material not<br />

directly irradiated by the beam and thereby prevents<br />

lateral expansion <strong>of</strong> the kerf width.<br />

4.5. Influence <strong>of</strong> assist gas pressure on the <strong>cutting</strong><br />

At pressure <strong>of</strong> around 5 bar, good quality cuts were<br />

obtained for ZINCALUME nad ZINCANNEAL, (see<br />

Fig. 3) with insignificant burr and heat-affected zone<br />

(HAZ) while GALVABOND exhibited a little wider<br />

HAZ. The <strong>cutting</strong> velocity increased with increasing gas<br />

pressure by about 60% as the gas pressure moved from<br />

5 to 20 bar. Repeat experiments for GALVABOND<br />

revealed that the parameter area in which good quality<br />

cuts are obtained narrowed compared to low gas pressure<br />

parameters. At low <strong>cutting</strong> velocities, the high O 2<br />

pressure resulted in a strong burring effect that was<br />

uncontrollable and produced a wide irregular kerf.<br />

It can be inferred that this is due to the high pressure<br />

<strong>of</strong> pure oxygen that reacts with zinc and steel, forming<br />

zinc and chromium oxides along the edges <strong>of</strong> the cuts.<br />

End results indicated that at higher O 2 pressures, the<br />

<strong>cutting</strong> velocities increased in the range <strong>of</strong> 40–70%,<br />

proving machining pr<strong>of</strong>itable for these specimens using<br />

the high energy laser.<br />

4.6. Metallographic in�estigation<br />

The micrographs in Fig. 5 show the front views <strong>of</strong><br />

the specimens. While the HAZ is very narrow in ZIN-<br />

CALUME and ZINCANNEAL specimens, it is more<br />

pronounced in the case <strong>of</strong> GALVABOND. This is also<br />

true for the case <strong>of</strong> oxide deposition across the length


240<br />

G.V.S. Prasad et al. / Journal <strong>of</strong> Materials Processing Technology 74 (1998) 234–242<br />

<strong>of</strong> cut. The figure covers the area towards the bottom <strong>of</strong><br />

the cut edge, and clearly shows the adhesive dross and<br />

the porous nature <strong>of</strong> the resolidified molten zone. The<br />

angularity <strong>of</strong> most <strong>of</strong> the pores implies that they are the<br />

result <strong>of</strong> the entrapment <strong>of</strong> the gas from the top to the<br />

bottom <strong>of</strong> the cut. Also, a kind <strong>of</strong> folding mechanism is<br />

revealed in the lines parallel to the cut edge. During the<br />

fluid flow from the central hot spot to the side <strong>of</strong> the<br />

cut zone, it could be possible that the two adjacent,<br />

oxidised surfaces can come into contact and become<br />

trapped as a linear inclusion <strong>of</strong> the type shown in the<br />

figure.<br />

Fig. 4. Microphotographs <strong>of</strong> the top surface: (a) ZINCALUME; (b)<br />

ZINCANNEAL; (c) GALVABOND.<br />

4.7. Control <strong>of</strong> dross<br />

While the deposition <strong>of</strong> dross on the lower edge <strong>of</strong><br />

the cut is an undesirable effect, it is rather easy to<br />

remove mechanically in the case <strong>of</strong> these specimens by<br />

either scraping or abrasion. To minimise the dross,<br />

there are numerous techniques available. One option<br />

could be the use <strong>of</strong> a pulsed laser beam rather than a<br />

CW mode. This is because the peak energy <strong>of</strong> a pulsed<br />

laser beam is much higher than the CW output but the<br />

average output is generally lower.<br />

The high peak powers <strong>of</strong> each pulse should act to<br />

rapidly melt and vaporise these <strong>metallic</strong> coatings. A cut<br />

can be carried out in this way with minimum surplus<br />

melting by conduction effects. This reduction in the<br />

surplus melting could further inhibit the generation <strong>of</strong><br />

dross although the <strong>cutting</strong> speeds tend to be lower than<br />

for the higher power CW mode.<br />

5. Conclusions<br />

The <strong>metallic</strong> <strong>coated</strong> <strong>sheet</strong> <strong>steels</strong> under study, i.e.<br />

ZINCALUME, ZINCANNEAL and GALVABOND,<br />

can be cut at commercially acceptable rates in the<br />

observed thickness range <strong>of</strong> 0.5–1.0 mm at high laser<br />

powers. While the <strong>cutting</strong> speed is same in the case <strong>of</strong><br />

ZINCALUME and ZINCANNEAL, it is slightly<br />

higher (about 20%) in the case <strong>of</strong> GALVABOND. The<br />

input power, <strong>cutting</strong> velocity and the assist gas pressure<br />

dictate the quality <strong>of</strong> cuts obtainable in the machining<br />

<strong>of</strong> these materials.<br />

Oxygen is quite effective as an assist gas for the<br />

<strong>cutting</strong> process as far as the <strong>cutting</strong> speeds are concerned.<br />

However, difficulties associated with localised<br />

overheating, particularly in the case <strong>of</strong> GALVABOND<br />

specimens, may be encountered if detailed work is<br />

required. In the laser <strong>cutting</strong> <strong>of</strong> GALVABOND specimens,<br />

the oxidised edges can be totally eliminated by<br />

using some other assist gas such as nitrogen or helium,<br />

which should render totally oxidised-free cut edges.<br />

If two different laser powers are compared, it is<br />

probable that the higher power will have an inferior<br />

mode quality which will not focus to as small spot as<br />

the lower power. This larger focal spot will produce a<br />

wider cut, thus rendering the process less efficient, as<br />

more material will have to be removed to generate the<br />

cut.<br />

The fluid dynamics <strong>of</strong> the cut zone play a very<br />

important role in determining the material removal<br />

rate. <strong>Inc</strong>rease in power input changes the inclination<br />

and the geometry <strong>of</strong> the cut front, which will in turn,<br />

induce changes in the material removal rates. Thus,<br />

above a limiting <strong>cutting</strong> speed, the viscosity <strong>of</strong> the melt<br />

may become the rate determining factor.


G.V.S. Prasad et al. / Journal <strong>of</strong> Materials Processing Technology 74 (1998) 234–242 241<br />

Fig. 5. Microphotographs <strong>of</strong> the front view: (a) GALVABOND; (b) ZINCALUME; (c) ZINCANNEAL.<br />

As a result <strong>of</strong> the reduction in thermal losses to the<br />

workpiece when <strong>cutting</strong> at higher speeds, the thermal<br />

gradients around the <strong>cutting</strong> zone become more severe<br />

as the material along the cut line is not preheated by<br />

the moving cut front and therefore requires more energy<br />

to become melted and ejected.<br />

Slag-free cuts are obtainable in the <strong>cutting</strong> <strong>of</strong> ZIN-<br />

CALUME and ZINCANNEAL but in the case <strong>of</strong><br />

GALVABOND, the oxides formed are concentrated in<br />

the slag. Owing to the high thermal conductivity and<br />

melting point, the slag solidifies before it leaves the<br />

kerf.<br />

It is observed that the slag is partly pressed into the<br />

melt zone in the cut kerf, which could cause problems<br />

when these specimens are subjected to further processing.<br />

Adherent dross is generally formed at the lower edge<br />

<strong>of</strong> the cut as a result <strong>of</strong> high surface tension forces and<br />

surface tension gradients within the melt. While dross<br />

can be easily removed, it can also be minimised by<br />

pulsed laser <strong>cutting</strong>, which is <strong>of</strong> course, at the expense<br />

<strong>of</strong> the <strong>cutting</strong> speed or probably by the use <strong>of</strong> a dross<br />

jet which directs all the dross onto the waste-material<br />

side <strong>of</strong> the cut.<br />

5.1. Implication <strong>of</strong> the arguments<br />

It is evident from the foregoing results and discussion<br />

that the analytical model thus developed exhibits a<br />

finite-element character, when applied to the materials<br />

under consideration. This is due to the presence <strong>of</strong><br />

different materials <strong>of</strong> various thicknesses and their<br />

sandwiching influence. The simplified model precludes<br />

the possibility <strong>of</strong> studying the laser–material interactions<br />

at the various interfaces, where the real time<br />

interactions continue to remain drastic and intricate, by<br />

virtue <strong>of</strong> different expansion rates <strong>of</strong> the metals, viz.<br />

zinc, aluminium and steel, as the beam power travels<br />

down the sandwich, thereby giving rise to different<br />

temperature gradients at the interfaces.<br />

Although it is intended to amplify the model to<br />

investigate the laser <strong>cutting</strong> <strong>of</strong> <strong>metallic</strong> <strong>coated</strong> <strong>sheet</strong>


242<br />

G.V.S. Prasad et al. / Journal <strong>of</strong> Materials Processing Technology 74 (1998) 234–242<br />

<strong>steels</strong> in future work, the situation is rather complex,<br />

considering the composition <strong>of</strong> these specimens. The<br />

model developed so far thus provide scope for further<br />

modifications in the light <strong>of</strong> the above-mentioned intricacies<br />

and can be modified to accommodate the interactions<br />

at the interfaces <strong>of</strong> the sandwich <strong>of</strong> the specimens<br />

in terms <strong>of</strong> the temperature gradients and the different<br />

coefficients <strong>of</strong> linear expansions <strong>of</strong> the various metals.<br />

Acknowledgements<br />

The authors wish to express their thanks and<br />

appreciation to BHP Steels Ltd., Australia, for<br />

providing the specimens and the Queensland<br />

Manufacturing Institute, Brisbane, Australia, for the<br />

Cincinatti CL-5 CNC laser centre to carry out the<br />

experiments.<br />

Bibliography<br />

[1] Y. Arata, I. Miyamoto, Metal heating by laser beam, Weld.<br />

Constr. 57 (8) (1978) 239.<br />

[2] Y. Arata, I. Miyamoto, Some fundamental properties <strong>of</strong> highpower<br />

laser beam as a heat source (report 2)—CO 2 laser absorption<br />

characteristics <strong>of</strong> metals, Trans. Jpn. Weld. Soc. 3 (1) (1972)<br />

152–162.<br />

[3] F.N. Birkett et al., Prevention <strong>of</strong> dross attachment during laser<br />

<strong>cutting</strong>, in: F. Kimmit (Ed.), Proc. 2nd Int. Conf. on <strong>Laser</strong>s in<br />

Manufacturing, 26–28 March 1985, Birmingham, UK, IFS,<br />

Bedford, UK, 1985, pp. 63–66.<br />

[4] G. Buness, S. Weber, New <strong>sheet</strong> metal <strong>cutting</strong> using CO 2 lasers,<br />

ZIS Mitt. 19 (1) (1977) 120–129.<br />

[5] J.S. Ekersley, <strong>Laser</strong> <strong>cutting</strong> <strong>of</strong> <strong>sheet</strong> steel with the revolutionary<br />

turbolase T1000 CO 2 laser, in: ICAEO 1982, Proc. Mater.<br />

Process. Symp. 31 (1982) pp. 20–23, September, 1982, Boston,<br />

MA, <strong>Laser</strong> Institute <strong>of</strong> America, Toledo, OH, pp. 131–134.<br />

[6] S.G. Gornyi et al, et al., An investigation <strong>of</strong> the process <strong>of</strong><br />

gas-laser metal <strong>cutting</strong>, Weld. Eng. 3 (1992) 31–32.<br />

[7] V.G. Gregson, <strong>Laser</strong> metal <strong>cutting</strong>, Mach. Tool Blue Book 76<br />

(9) (1981) 72–75.<br />

.<br />

.<br />

[8] A.G. Grigoryants, Basics <strong>of</strong> <strong>Laser</strong> Material Processing, Mir<br />

Publishers, Moscow, 1994.<br />

[9] S. Hoshinouchi, M. Kobayashi, et al., Quality improvement in<br />

CO 2 laser <strong>cutting</strong> <strong>of</strong> metals, in: Colloq. on Thermal Cutting and<br />

Flame Processes, 7 September, 1982, Ljubljana, Yugoslavia,<br />

International Institute for Welding, London, UK, 1982.<br />

[10] J.N. Kamulu, W.M. Steen, The importance <strong>of</strong> gas flow parameters<br />

in laser <strong>cutting</strong>, in: Metall. Soc. AMIE, Paper No. 81-38,<br />

Warrendale, PA.<br />

[11] E.P. Kudrayavstev, A.V. Tikhomirov, Gas laser <strong>cutting</strong> <strong>of</strong> materials,<br />

in: Colloq. on Thermal Cutting and Flame Processes, 7<br />

September, 1982, Ljubljana, Yugoslavia, International Institute<br />

for Welding, London, UK, 1982.<br />

[12] C.S. Lee, A. Goel, H. Osada, Parametric studies <strong>of</strong> pulsed laser<br />

<strong>cutting</strong> <strong>of</strong> thin metal plates, in: J. Majumdar (Ed.), ICALEO<br />

1984, Proc. Mater. Process. Symp. 44, 12–15 November 1984,<br />

MA, <strong>Laser</strong> Institute <strong>of</strong> America, Toledo, OH, 1985, pp. 86–93.<br />

[13] M. Lepore, et al., A quantative theory for the role <strong>of</strong> oxygen in<br />

the laser <strong>cutting</strong> process (Centro <strong>Laser</strong>, Italy), in: E.A. Metzbower<br />

(Ed.), ICALEO ’83, Proc. Mater. Process. Symp. 38,<br />

November 1983) Los Angeles, CA, <strong>Laser</strong> Institute <strong>of</strong> America,<br />

Toledo, OH, 1984, pp. 160–165.<br />

[14] G. Manassero, et al., <strong>Laser</strong> beam <strong>cutting</strong> <strong>of</strong> <strong>metallic</strong> and structural<br />

components, Commission <strong>of</strong> the European Communities,<br />

Luxembourg, 1985, p. 27.<br />

[15] J. Mazumdar, W.M. Steen, Heat transfer model for CW laser<br />

material processing, J. Appl. Phys. 51 (2) (1980) 941–947.<br />

[16] M. Mitani, et al., <strong>Laser</strong> <strong>cutting</strong> <strong>of</strong> aluminium thin film with no<br />

damage to under-layers, Ann. CIRP 28 (1) (1979) 113–115.<br />

[17] J. Powell, et al., CO 2 laser <strong>cutting</strong> <strong>of</strong> aluminium alloys, in: Proc.<br />

5th Int. Conf. on <strong>Laser</strong>s in Manufacturing, September 1988,<br />

IFS, pp. 15–24.<br />

[18] J. Powell, The influence <strong>of</strong> material thickness on the efficiency <strong>of</strong><br />

laser <strong>cutting</strong>, in: Proc. 6th Int. Conf. on <strong>Laser</strong>s in Manufacturing,<br />

September 1988, IFS, pp. 135–153.<br />

[19] J. Powell, Metallurgical implications <strong>of</strong> laser <strong>cutting</strong> <strong>of</strong> stainless<br />

<strong>steels</strong>, in: Proc. Int. Conf. on Power Beam Technology (10–12<br />

September 1986), Brighton, UK, Welding Institute.<br />

[20] R. Rothe, et al., <strong>Laser</strong> <strong>cutting</strong> <strong>of</strong> <strong>sheet</strong> metal ranging in thickness<br />

from 3 to 30 mm, in: Schweissen und Schneiden ’82, 29 September–1<br />

October, 1982, Berlin, Germany, pp. 40–48.<br />

[21] W.M. Steen, J.N. Kamalu, <strong>Laser</strong> Cutting, Imperial College <strong>of</strong><br />

Science and Technology, London, UK, 1985, p. 111.<br />

[22] A.V. Tikhomirov, et al., Technology and equipment for the gas<br />

laser <strong>cutting</strong> <strong>of</strong> thin <strong>sheet</strong> steel, Weld. Prod. 26 (11) (1979)<br />

11–13.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!