Vous êtes sur la page 1sur 10

3.

1 Introduction
The idea of cutting with light has appealed to many from the first time they burnt paper on a
sunny day with the help of a magnifying glass. Cutting centimetre-thick steel (Figure 3.1) with a
laser beam is even more fascinating!

Laser cutting is today the most common industrial application of the laser; in Japan around 80% of
industrial lasers are used in this way. Apart from the fascination, which is rarely a driving force for
investment in hard industry, the reason is most probably that in cutting there is a direct process
substitution into an established market and the laser, in many cases, happens to be able to cut
faster and with a higher quality than the competing processes. The comparison with alternative
techniques is listed in Table 3.1

Figure 3.1 Metal cutting of 5- mm-thick stainless steel with a


CO2 slab laser of 2 kW. (Courtesy of AILU through http://www.designforlasermanufacture.com
and Rofin-Baasel UK)
The significant advantages of the laser are seen in the table but possibly these need some further
explanation. Thus, the advantages can be divided into two categories – cut quality and process
characteristics: Cut quality characteristics:

1. The cut can have a very narrow kerf width giving a substantial saving in material. (Kerf is the
width of the cut opening.)

2. The cut edges can be square and not rounded as occurs with most hot jet processes or other
thermal cutting techniques.

3. The cut edge can be smooth and clean. The cut is reckoned to be a finished cut, requiring no
further cleaning or treatment.

4. The cut edge is sufficiently clean that it can be directly rewelded.

5. There is no edge burr as with mechanical cutting techniques. Dross adhesion can usually be
avoided.

6. There is a very narrow HAZ, particularly on dross-free cuts. Usually there is a very thin
resolidified layer of micron dimensions. Thus, there is negligible distortion.

7. Blind cuts can be made in some materials, particularly those which volatilise, such as wood or
acrylic.

8. Cut depth is limited and depends on the laser power. The current range for highquality cuts with
2–5-kW laser power is 10–20 mm.

Process characteristics:

1. It is one of the faster cutting processes.

2. The workpiece does not need clamping, although it is usually advisable to do so to avoid the
workpiece shifting with the table acceleration and for locating when using a computer numerical
control (CNC) program.

3. There is no tool wear since the process is a noncontact cutting process, but the lens must be
kept clean.

4. Cuts can be made in any direction; but see Section 3.6.1.3 on polarisation.

5. The noise level is low.

6. The process can easily be automated with good prospects for adaptive control in the future.

7. Tool changes are mainly “soft”, that is, they are only programming changes. Thus, the process is
highly flexible.

8. Some materials can be stack-cut, but there may be a problem with welding between layers.
9. Nearly all materials can be cut. They can be friable, brittle, electric conductors or
nonconductors, hard or soft. Only highly reflective materials such as aluminium, copper and gold
can pose a problem, but with proper beam control these can be cut satisfactorily.

3.2 The Process – How It Is Done


The general arrangement for cutting with a laser is shown in Figure 3.2. The principle components
are the laser itself with some shutter control, beam guidance train, focus-

Figure 3.2 General arrangement for laser cutting: a using transmissive optics, and b using reflective
optics. CNC computer numerical control

In cutting, the laser evaporates a hole through the material and then the “hole” is traversed to
make a cut. The drilling and piercing processes are different from cutting in a line because they do
not have an open-sided hole. Since every cut must start with piercing, if it does not start at an
edge, our first topic here will be drilling and piercing.

3.3 Laser Drilling and Piercing


3.3.2 Drilling Process Variations

The importance of laser drilling as an industrial process has led to many variations on how to
achieve quick, high-quality holes with good repeatability. They include (see Figure 3.3):

• Single-shot drilling: One pulse makes and finishes the hole.

• Double-pulse drilling: The energy is divided between two pulses which follow each other in very
rapid succession to interact with the plasma more efficiently.

• Percussion drilling: Single or multiple shot with no movement ofthe workpiece or the beam.

• Trepanning: Rotating the beam around the perimeter of the hole, a form of cutting.
• Helical trepanning: Starting near the middle of the hole and rotating around the perimeter,
gradually deepening the hole with each rotation spirally machining into the workpiece, sometimes
also with a change in the focal position following the hole base downward.

Using these methods holes can be made:

• Normal to the surface – the usual method

• At an angle to the surface – particularly relevant for turbine blades. No particular problem has
been found with making angled holes except that at oblique angles the hole depth would usually
have to be greater. At the start of drilling at an angle, the plume is emitted at right angles to the
surface, which is an advantage, but once the keyhole has formed, it aligns with the hole axis [2]

• As blind holes – for electronic vias.

These process variations will now be discussed in turn. The physical mechanism describing how
the beam penetrates a material is roughly the same for all of these processes. Percussion drilling is
chosen as the example to discuss the physical mechanisms.

3.3.3 Percussion and Single- or Double-shot Drilling


Laser percussion drilling techniques are the quickest methods for achieving a hole by comparison
with almost any other process, but they are not as precise as trepanning techniques. Percussion
drilling is the preferred method in many aerospace applications

for cooling holes owing to its speed, sufficient accuracy and repeatability. It requires pulses of
between 105 and 107 W cm−2 [3].

3.3.3.1 How Can Light Drill a Hole? The Physical Picture

In drilling, which relies on vaporisation, the focused beam first heats up the surface to boiling
point and so generates a keyhole. The keyhole causes a sudden increase in the absorptivity owing
to multiple reflections causing the hole to deepen quickly. Figure 3.4 [4] shows the calculated
change in drilling speed as the hole deepens. As it deepens, so vapour is generated and escapes.
This evaporation exerts a reaction force on the melt surface as the vapour accelerates away; also
the temperature gradient across the surface of the melt exerts forces through variations in surface
tension. Both of these forces drive the melt to the side of the forming hole. At the side the melt is
driven by the high pressure at the base of the hole and the drag forces of the escaping vapours to
flow up the walls and out as spray with the vapour [5]. When the laser pulse finishes, this flow will
cease and the melt will fall back as splatter around the lip of the hole or remain as a recast layer
within the hole. In general, the more melt there is, the poorer the quality of the resulting hole.

This is the usual method of drilling for pulsed lasers or in the cutting or drilling of materials which
sublime and do not melt, such as wood, carbon and some plastics. Rykalin et al. [6] showed that at
approximately 3 × 106 W cm−2 metals start to evaporate and at 108 W cm−2 a plasma is formed
that will stop or prevent further drilling by blocking the beam through absorption within the
electron cloud in the plasma. This will occur when the plasma frequency approaches the laser
frequency.

This simplified view of laser drilling contains the overall picture but added to this are the detailed
interactions of the incoming radiation within the keyhole and the fluid flow of the melt that is to
be ejected. The incoming radiation passes into the forming hole through the exiting hot gas and
dust, which will have a scattering effect on the beam. The beam will be absorbed by Fresnel
absorption on the walls of the keyhole as it is waveguided to the base of the hole or reflected out
of the hole, but it will also be absorbed by the electrons in the hot gas, which in turn will radiate to
the walls of the forming hole. If the power is sufficient (more than 108 W cm−2), a plasma whose

Figure 3.4 Time variation of keyhole depth. (After Noguchi et al. [4])

frequency is near that of the laser beam will form and effectively block the beam from reaching
the substrate. The pressure will still be there, which is the basis for the process of shock peening
(see Section 6.19).

The melt generated will flow up the walls of the hole owing to the high pressure at the base and
the drag forces from the exiting gases. The speed of ejection of the gases will in most cases be
supersonic and shock waves will form in the area of the hole. Added to this picture is the coaxial
assist gas blowing down onto the hole entrance. This will initially create a pressure countering the
exiting gases, but, if it is oxygen, it may react with the molten metal and generate more heat. On
breakthrough, the assist gas pressure will help to drive the vapour through the hole and help scour
the sides.

Much of this was witnessed by French et al. [7] using filming at 42,000 frames per second in which
they identified three stages in the development of a hole: stage 1 – wide conical ejection of melt;
stage 2 – melt ejection from the hole in a ribbonlike form; stage 3 – when the hole is deeper and
the assist gas is oxygen, there is spark droplet ejection in all directions. They found that the assist
gas affected the velocity of the ejected material: for an increase in assist gas pressure from 1 to 6
bar the velocity of the melt ejection was reduced by 50 % from the observed value of 17–27 m
s−1.
3.3.3.2 Rate of Penetration.
3.3.3.2.1 Lumped Heat Capacity Model

The rate of penetration of the beam into the workpiece can be estimated from a lumped heat
capacity calculation, assuming the heat flow is one-dimensional and all of it is used in the
vaporisation process – that is, that the heat conduction is zero. This fairly gross assumption is not
ridiculous if the penetration rate is similar to or faster than the rate of conduction.

Thus, the volume removed per second per unit area equals the penetration velocity, V (m s−1).
The volume removed per seconds equals the power divided by the heat capacity of a unit volume,
and we have

where F0 is the absorbed power density (W m−2), ρ is the density of the solid (kg m−3), L is the
latent heat of fusion and vaporisation (J kg−1), Cp is the heat capacity of the solid (J kg−1 ○C−1), Tv
is the vaporisation temperature (○C) and T0 is the temperature of the material at the start (○C).

If we substitute values into this equation, we can derive the approximate maximum penetration
rate possible for different materials. Assuming we have a 2 kW laser focused to 0.2 mm beam
diameter, the power density will be

The rate of penetration of such a beam into various materials is shown in Table 3.2. These
penetration figures are of the same order as those found experimentally [8]. If the penetration
rate is around 1 m s−1, then the vapour velocity from a cylindrical hole would be defined by the
ratio of the densities of vapour and solid to be of the order of V ρv/ρs = 1,000 m s−1. At these
sonic speeds compression effects and variations in the hole shape will mean the actual velocity of
exit of the vapour is much less, but nevertheless sonic flow and shock waves will occur (there is
usually a distinct bang with each laser pulse). Such high-velocity flow will be capable of
considerable drag in eroding the walls of the forming hole.

Thus, in this form of drilling or piercing the material is removed partly as vapour and partly as
ejecta. Gagliano and Peak [9] estimated from their experiments that around 60% of the material
was removed as ejecta.This figure was partially confirmed in Voisey et al.’s [10] detailed analysis of
ejecta during drilling. They found using 0.5-ms pulses that the melt ejection fraction varied
between 35 and 60% for most metals except the really heavy metals, such as tungsten, for which
the fraction fell to nearer 10%. They also showed that the size of the ejected particles grew with a
reduction in the pulse energy or an increase in the length of the pulse; larger ejected particles
coming from long, lower-powered pulses. The average particle size for a 2.5 J, 0.5 ms pulse was
found to be around 10 μm. Using high-speed photography, they found the velocity of ejection
averaged approximately 8 m s−1 for a 1.4 J, 0.5 ms pulse and 13 m s−1 for a 2.4 J, 0.5-ms pulse – a
figure in agreement with that of Ng and Li [11]. There was, of course, a range of velocities,
presumably dependent on particle size and direction.

There are a number of side effects from this almost explosive evaporation. One is the recoil
pressure required to accelerate the vapour away. Bernoulli’s equation is able to give a rough
estimate of the value of this pressure for an exit velocity of 1,000 m s−1, even though it assumes
incompressible flow, as

One atmosphere is 105 N m−2. A pressure rise of this order (40 atm) will cause a rise in the
vaporisation temperature (given by the Clapeyron–Clausius equation dp dT = ΔH TΔV ). This
pressure causes stress in the surface, which is amplified by the thermal stresses generated in the
heated material. Together they represent quite a considerable stress. If this can be applied very
quickly, in a few nanoseconds (10−9 s), then the effect is similar to being hit, as in shot peening, as
noted earlier.

3.3.3.2.2 Melt Fraction During Drilling

The quality of the hole or cut is determined to a large extent by the quantity of melt which may
build up and cause debris on the surface or erosion marks on the wall. Thus, it is interesting to
calculate how quickly the boiling point is reached and how much of the laser pulse energy is used
in evaporation or melting.

For one-dimensional heat flow with constant energy input it can be shown (see Chapter 5) that the
surface temperature at any time, t, after the start of irradiation is

given by

where α is the thermal diffusivity (K/ρCp) (m2 s −1) and K is the thermal conductivity (W m−1 K−1).

Therefore, the time required to reach the boiling point on the surface, tv, is
The estimated time for a 2-kW laser beam to cause vaporisation is shown in Table 3.2. The thermal
gradient at that time would have penetrated [assuming a Fourier number (x2/α t) = 1)] to around 2
μm for iron and hence it can be seen that the HAZ is expected to be small in this case. The
previous calculation, based upon ignoring the heat conduction, is thus not so wide of the mark.

The energy and peak power of a pulse are also critical in determining how much of the energy will
be used to evaporate as opposed to melt. If the pulse is sufficiently short, the thermal penetration
distance is limited, being proportional to √(α tpulse). It is thus possible to do a heat balance on this
small volume and estimate how much of the energy is available for evaporating material.

Such an energy balance on the volume [ √∝ tpulse ( πd2/4 )] give

where A is a constant and m′ is a multiple of the sensible heat required to evaporate a unit mass of
material.

From this equation we get

which can be simplified as

where A′ and B are constants dependent on the material being drilled. If m′ > 1, then there is
sufficient energy in the pulse to evaporate all the heat-affected volume. Using the values for iron
and aluminium (see Table 3.2), we find that:

Figure 3.5 The pulse energy and duration required to evaporate all the material in the
heataffected zone, assuming no thermal conduction. The lines represent just sufficient energy for
evaporation
Figure 3.5 shows how the pulse energy varies with pulse duration to supply just sufficient energy
to evaporate the affected volume. In this simple calculation remember that the depth of the
evaporated volume is equal to √(α tpulse). So the shorter the pulse, the smaller the thickness of
the treated zone. Hence, a larger spot means a thinner depth of penetration is required if the
whole volume is to be evaporated with the energy supplied. If the pulse energy increases, then the
depth of penetration must increase to maintain the equation for evaporating all the affected
material. This is counterintuitive owing to the condition that all the material has to be evaporated.
A more intuitive equation would result if we had considered pulse power instead of energy. Since
the power is proportional to 1/ √t if all the material were to be evaporated. Shorter pulse times
(hence more intense pulses) by a factor of (58/76) = 0.76 (the ratio of A′ and B for iron and
aluminium) are required if melting aluminium is to be avoided.

The figure shows clearly that longer pulses of low energy generate more melt than short sharp
pulses and that pulses longer than 0.3 ms even with a very fine 50-μm spot size will generate
significant melt that has to be removed. A nanosecond pulse is a good pulse length to use to give
minimum melt. Many papers have shown that shorter pulses leave less spatter and recast
material.

3.3.3.2.3 One-dimensional Heat Flow Model

The simple view of the drilling mechanism just discussed can be greatly elaborated. In the paper
by Chan and Mazumder [12], they assumed that the laser radiation was absorbed by the surface of
the liquid melt at the bottom of the hole. They then modelled the heat flow from the solid to the
liquid–solid interface and through the liquid melt layer to the liquid–vapour interface. At the
liquid–vapour interface the power is absorbed and was modelled for a moving boundary by the
Stefan energy balance. The vapour is driven off with considerable force owing to the rate of
evaporation and thermal expansion to near sonic speeds within the distance of a molecular free
path or so. In this layer, known as a Knudsen layer, there will be a discontinuity in the
temperature,

Figure 3.6 Vaporisation, liquid expulsion and total


removal ratesfor aluminium versus laser beam power. (After Chan and Mazumder [12])
Figure3.7 Vaporisation, liquid expulsion and total removal rates versus laser power for superalloy.
(After Chan and Mazumder [12])

Vous aimerez peut-être aussi