You are on page 1of 11

3-dimensional laser heating model including a moving

heat source consideration and phase change process


B. S. Yilbas
Abstract The use of a Fourier heating model in high in-
tensity laser material processing is limited due to the as-
sumptions made in the model. An electron-kinetic theory
may offer an alternative solution to the problem. Conse-
quently, in the present study an electron-kinetic theory
approach is introduced to model the 3-dimensional laser
heating process. The phase change and conduction effects
are encountered when driving the governing equations. To
simulate the moving heat source, a scanning velocity of the
laser beam is considered, in this case, the laser beam scans
the workpiece surface with a constant velocity. The gov-
erning heat transfer equation is in the form of integro-
differential equation, which does not yield the analytical
solution. Therefore, a numerical method employing an
explicit scheme is introduced to discretize the governing
equations. To validate the theoretical predictions, an ex-
periment is conducted to measure the surface tempera-
tures of the workpiece substrate during Nd YAG laser
heating process. It is found that the rapid increase in
temperature occurs in surface vicinity due to the succes-
sive electron-lattice site atom collisions. The depth of
melting zone increases as the heating progresses and the
temperature remains almost constant at the melting
temperature of the substrate in the surface vicinity. In
addition, the theoretical predictions agree well with the
experimental ndings.
1
Introduction
For many applications laser have been utilized as practical
and economical tool in industrial production [1]. The
processing quality depends on the energy-time behavior of
the laser emission and on the optical, thermal and me-
chanical specications of the workpiece material. This
quality is limited by formulation of a thermally affected
region in and around the laser-workpiece interaction zone.
Many metal-working problems have been investigated
and the corresponding interaction mechanisms have been
developed [2, 3, 4]. The heating and vaporization of solids
by intense laser beams have been examined by the use of a
Fourier heating model [5, 6, 7]. Diniz Neto and Lima [8]
numerically solved the 3-dimensional non-linear Fourier
heat conduction equation for a pulsed laser input. How-
ever, the study covered the conduction only process and
the phase change processes were left obscure. Blackwell [9]
investigated the phase change process initiated by a laser
beam with the use of a Fourier heating model. He indicated
that the maximum temperature lied below the surface. The
phase change processes were also studied by Duley [10] by
introducing a steady state evaporation. A comprehensive
work was carried out by Yilbas [11] to whom indicated that
the laser power intensity inuenced considerably the
resulting temperature proles. On the other hand, the
applicability of the Fourier heating model in laser material
processing is limited due to the assumptions made in the
theory. In the analysis of the Fourier theory of heating, the
heat ux through a given plane is considered as being a
function of the spatial temperature gradient at that plane.
This depends upon the assumption made that the tem-
perature gradient remains constant between two consecu-
tive closely spaced planes. The distance between these
planes is nite. However, the higher order terms (o
3
T=ox
3
),
which are neglected in Fourier theory become important at
high laser power intensity applications [12], which in turn
results in errors in the solution. Moreover, the absorption
depth of laser radiation is of the order of the mean free path
of electrons, which is comparable to the atomic spacing of
solids. Therefore, over the scale of distances required to
examine the problem the material can no longer be a
homogeneous continuum and the heat ux through given
plane is electron energy distribution dependent through
the material. Consequently, when formulating the problem,
a new model is needed to examine the problem on a
microscopic level.
The problems related to the applicability of the Fourier
theory in laser heating process have let to the development
of a theory of heat conduction based on a kinetic model of
the energy transfer [13]. The mechanism is essentially a
kinetic energy transfer process which occurs when elec-
trons and lattice site atoms with different energies interact.
When the electrons absorbs the energy from the incident
laser beam, the free electron gas in the substrate proceeds
to impart its excess energy to the lattice site atoms through
collisional process. However, the phase change process is
the complicated one and the energy exchange mechanism
occurs through electron-free molecule, free molecule-bond
molecule and free molecule-vacancy collisions [14]. The
Heat and Mass Transfer 33 (1998) 495505 Springer-Verlag 1998
495
Received on 4 August 1997
B. S. Yilbas
King Fahd University of Petroleum and Minerals
Mechanical Engineering Department
P.O. Box 1913,
Dhahran 31261, Saudi Arabia
The author acknowledge the support of King Fahd University
of Petroleum and Minerals, Dhahran, Saudi Arabia for this work.
free molecules from the interior of the material and at the
surface may escape if they have sufcient energy. However,
the thermal motion of the molecules can be taken into
account by abandoning the concept of thermal motion of
the molecules as being small vibrations about xed
equilibrium positions. Consequently, the irreversible
displacements, where molecules leave their lattice site and
become free, should be included.
In the light of the above arguments, the present study is
carried out to model the 3-dimensional laser heating
process using the electron-kinetic theory approach. The
conduction and convection affects in the heating process is
encountered and the resulting equations are derived. The
heat transfer equations so derived are in the form of
integro-differential equations, which do not yield analyti-
cal solutions. Therefore, a numerical method employing an
explicit scheme is introduced to discretize the governing
equations. In order to simulate the moving heat source
problem, a laser beam with constant scanning velocity is
considered, i.e. the laser beam scans the surface along y-
axis at a constant velocity. The experiment using an optical
system is conducted to measure the surface temperature
during the Nd YAG laser heating. This enables to validate
the theoretical predictions. However, the difculties are
encountered in the measurement after the attainment of
the surface evaporation, therefore, the laser power inten-
sity is reduced to obtain the melting process at the surface,
i.e. the high power intensity results in evaporation and
subsequent plasma formation at the surface.
2
Mathematical analysis of heating
The mathematical analysis of the heating process is con-
sidered according to the processes involved. These are
given under the appropriate sub-headings.
The electron movement and control volume is shown in
Fig. 1.
2.1
Heat conduction process
The kinetic theory of heat conduction model relies on the
energy exchange mechanism that occurs during the elec-
tron and lattice site atom collisions. To simplify the
physical process, some useful assumptions needs to be
made. These are: omission of thermionic emission, at-
tainment of space charge, a net ow of electrons in the
substrate by the presence of an electron source at innity
and a fraction of electron excess energy given to lattice site
atom during a single collision.
In the conduction heating, the energy transfer mecha-
nism takes place through successive collisions between the
electrons and lattice site atoms as well as phonon relax-
ation process. The electron excess energy, due to absorp-
tion of the laser beam, is conducted to the lattice site
atoms through successive electron lattice site atom colli-
sion in the surface vicinity. This results in an increase in
the amplitude of the lattice vibrations. As a result of this
process, the temperature of the lattice site increases. The
motion of electrons is 3-dimensional in the solid substrate,
therefore, three dimensional modeling of the heating
process is necessary.
The basic concepts of the mathematical model is given
here, but the details of which can be found in [13]. The
number of electrons having the average velocity V
x
en-
tering the control volume in x-axis across the area A
x
at
time dt is: N
x
V
x
A
x
dt where A
x
= dy dz. The probability of
the electrons moving in x-axis and making successive
collision is:
_

N
x
V
x
A
x
exp
[x s[
k
_ _
ds
k
dx
k
:
The negative bound of the integral is due to a mirror
image introduced at the free surface of the electrons [13].
The net transfer of energy during the electron-phonon
collision across the area A
x
in x-axis is:
DE
x;t
=
_

N
x
V
x
A
x
exp
[x s[
k
_ _
ds
k
dx
k
[E
s;t
E
x;t
[
where E
s;t
and E
x;t
are the electron and phonon energies in
x-axis respectively.
The transfer of energy across the area A
y
in y-axis
during the collision processes is:
DE
y;t
=
_

N
y
V
y
A
y
exp
[y g[
k
_ _
dg
k
dy
k
[E
g;t
E
y;t
[
where E
g;t
and E
y;t
are the electron and phonon energies in
y-axis respectively.
Similarly, the transfer of energy across the area A
z
in
z-axis during the collision processes is:
Fig. 1. 3-dimensional motion of electrons and molecules
496
DE
z;t
=
_

N
z
V
z
A
z
exp
[z f[
k
_ _
df
k
dz
k
[E
f;t
E
z;t
[
where E
f;t
and E
z;t
are the electron and phonon energies in
z-axis respectively.
Energy increase per unit volume in the material across
the area A
x
and along the increment dx during the time
interval dt is:
DE
x;t
A
x
dx dt
:
Similarly equation can be written for energy increase
per unit volume in the material across the areas A
y
and A
z
and along the increments dy and dz during the time
interval dt:
The energy absorbed by the substrate per unit volume
during time dt may be formulated considering the laser
beam intensity scanning the surface (moving in y-axis) with
a constant velocity u
s
(Fig. 1). In this case, the energy ab-
sorbed per unit volume by the substrate during time dt is:
DE[
abs
=
I
o

2p
_
a
exp
(y u
s
Dt)
2
z
2
a
2
_ _
d exp (dx)A
x
dx dt
where Dt is the time duration for a laser beam to scan the
surface under consideration.
Consequently, total energy gain per unit volume is:
DE
x;y;z;t
dx dy dz dt
=
DE
x;t
A
x
dx dt

DE
y;t
A
y
dy dt

DE
z;t
A
z
dz dt

DE[
abs
A
x
dx dt
knowing that
A
x
= dy dz :; A
y
= dx dz :; A
z
= dx dy
The total energy increase per unit volume is equal to
internal energy gain of the substrate, assuming thermal
properties are independent of temperature, i.e.:
DE
x;y;z
dx dy dz dt
= qC
p
o
ot
T(x; y; z; t)
Energy balance attains in the element during time incre-
ment dt, due to absorption and electron-phonon collision
processes, should be equal to the total energy increase in
the control volume. Therefore,
qC
p
o
ot
T(x; y; z; t)
=
I
o

2p
_
a
exp
(y u
s
t)
2
z
2
a
2
_ _
d exp(dx)
3f
k
2k
2
T(x; y; z; t)

fk
4k
3
_

0
exp
[x s[
k
_ _
T(s; y; z; t) ds
_

_
x
0
exp
[x s[
k
_ _
T(s; y; z; t) ds

_

x
exp
[x s[
k
_ _
T(s; y; z; t) ds
_

fk
2k
3
_

0
exp
[y g[
k
_ _
T(x; g; z; t) dg
_ _

fk
2k
3
_

0
exp
[z f[
k
_ _
T(x; y; f; t) df
_ _
(1)
where T(s; y; z; t), T(x; g; z; t) and T(x; y; f; t) are the
electron temperatures in x, y and z axes respectively.
However, the equation for the relation between the
energy of lattice site atoms and the electrons was devel-
oped using standard scattering rate formulas [15]. In this
case, the relation yields:
In x-axis:
T(s; y; z; t)
oT(s; y; z; t)
ot
=
3

h
pk
B
k(x
2
)[T(s; y; z; t) T(x; y; z; t)[
In y-axis:
T(x; g; z; t)
oT(x; g; z; t)
ot
=
3

h
pk
B
k(x
2
)[T(x; g; z; t) T(x; y; z; t)[
In z-axis:
T(x; y; f; t)
oT(x; y; f; t)
ot
=
3

h
pk
B
k(x
2
)[T(x; y; f; t) T(x; y; z; t)[ (2)
where x
2
is the second moment of the phonon spectrum,
k
B
is the Boltzmann's constant and k is the electron-
phonon coupling parameter.
To obtain a temperature eld for electrons and lattice
site atoms, Eqs. (1) and (2) must be solved simultaneously.
2.2
Phase change processes
In the case of phase change processes the energy exchange
may occur due to free molecular, free molecule-bound
molecule, electron-free molecule and free molecule-va-
cancy collisions. The bound molecules at the surface and
free molecules from the interior of the material may escape
through the surface if they have sufcient energy. In order
to consider the heat transfer process in detail, these effects
must be included. The thermal motion of the molecules
can be taken into account by including the irreversible
displacements of the molecules. In this case, a molecule
leaves its lattice site and becomes free, i.e. this irreversible
displacement in the surface vicinity allows molecules to
escape (evaporate) from the surface.
In most materials the effects of the solid-liquid phase
change are relatively small compared with the effects which
occur on evaporation. These facts, together with that of the
marginal change which occurs in specic heat on fusion
mean that the character of the thermal motion in liquids is
not fundamentally different from that in solids, where it is
mainly reduced to small molecular vibrations about xed
lattice sites. As the temperature is increased, these vibra-
tions are augmented by larger, irreversible movements
497
which shift the equilibrium site to a new position at a
distance of the order of intermolecular spacing. Finally, at a
high enough temperature, the whole lattice structure is
destroyed and the material is completely composed of free
molecules. The model of a material which has been pre-
sented is thus able to explain all these effects and therefore
is particularly suited as a starting point for the subsequent
analysis of the laser heating process where very high tem-
peratures and gradients co-exist. Before passing on to the
development of the energy transfer model it is worth noting
that there are two possible ways in which molecules may
leave the material. These are by direct evaporation of the
surface layers and by the escape of the free molecule gas
through the surface barrier. The latter process is governed
by the magnitude of the surface potential and it is useful to
consider the equilibrium situation when there is a saturated
vapor above the surface. In this case, the effect of the
surface barrier is to act outwards from the surface a small
distance. Molecules which move away from the surface
through this force eld are slowed down, whilst they are
accelerated if they move towards the surface. The overall
velocity distribution of molecules is Maxwellian. Since the
system is at equilibrium, there are equal number of mole-
cules condensing as there are evaporating. However, in
laser machining the process is non-equilibrium, therefore,
molecules which evaporate immediately escape. Since the
evaporation process is made up of contributions from the
surface and from the interior, the surface effect will be
unchanged apart from causing the surface to move, but
there will be a net outow of molecules from the interior of
the material. This may cause shock waves to propagate into
both the vapor and the material. The effects of shock waves
will be neglected but the total deciency of molecular ow
will be accounted for. The distribution of molecular
velocities in the surface vicinity will also become non-
equilibrium. However, it is assumed that this reaches an
equilibrium distribution.
Within the material, the molecules which escape from
the lattice must also have a minimum velocity. In the same
way, those molecules which escape form a Maxwellian
distribution. Since there is a minimum velocity which
correspond to a molecule being able to free, and there are
free molecules with velocities less than this, then they may
be easily trapped by vacancies to become high energy
bound molecules. Furthermore, these molecules may also
convert some their potential energy to kinetic energy and
therefore stable equilibrium distributions may be main-
tained in both the free and bound molecules. There is,
therefore, no mean temperature difference between free
and bound molecules in any section of the material.
In the analysis, the free molecular motion for evapo-
ration is considered as one-dimensional and it takes place
in the axis perpendicular to the surface, i.e. one-dimen-
sional motion only. However, the volumetric phase change
may be introduced as that the interior process is three-
dimensional, but it is modied at the free surface by the
fact that the evaporating molecules may escape in one-
dimension. The analysis of the phase change process is
given briey here, but the details can be found in [14].
The number of free molecules per unit volume
throughout the material at temperature / is
N
fm
= N
o
exp
U
1
k
B
/
_ _
and the number of vacancies per unit volume is:
N
h
= N
o
exp
U
2
k
B
/
_ _
where U is the activation energy and N
o
is the number of
lattice sites molecules per unit volume. It is also evident
that
N
bm
N
h
= N
o
where N
bm
is the number density of bound molecules.
The probability of the collision (P
h
) taking place be-
tween vacancy and the existing species (bound molecules,
free molecules and vacancies) may be expressed as [14]:
P
h
=
N
h
N
bm
N
fm
N
h
=
N
h
N
o
N
fm
=
exp
U
2
k
B
/
_ _
1 exp
U
1
k
B
/
_ _ :
The probability of the collision (P
fm
) taking place between
free molecule and the existing species (bound molecules,
free molecules and vacancies) may be expressed as [14]:
P
m
=
N
fm
N
bm
N
fm
N
h
=
exp
U
1
k
B
/
_ _
1 exp
U
1
k
B
/
_ _
Similarly, the probability of the collision (P
bm
) taking
place between a bound molecule and the existing species
(bound molecules, free molecules and vacancies) may be
written as [14]:
P
bm
=
1 exp
U
2
k
B
/
_ _
1 exp
U
1
k
B
/
_ _ :
The amount of energy transferred during free molecule-
vacancy collision is:
U(s) U(x) 3k
B
[/(s; y; z; t) /(x; y; z; t)[ (3)
where /(s; y; z; t) and /(x; y; z; t) are the temperatures of
the molecules at locations s and x respectively.
However, the recaptured molecule can give up an
amount of energy
U(s) 3k
B
/(s; y; z; t)
where /(s; y; z; t) is the lattice temperature at position s
(as shown in Fig. 1) and time t. Similarly, the escaping
molecule removes an amount of energy
U(x) 3k
B
/(x; y; z; t)
where /(x; y; z; t) is the lattice temperature at position x
(as shown in Fig. 1) and time t.
Therefore, the overall process may be considered as one
where free molecules created at ds collide in any section dx
and in doing so give up an amount of energy as indicated
by Eq. 3.
The amount of energy transferred during free molecule
bound molecule collision is:
498
1
2
g k
B
[/(s; y; z; t) /(x; y; z; t)[ :
Similarly, the amount of energy transferred during free
moleculefree molecule collision is:
1
2
h k
B
[/(s; y; z; t) /(x; y; z; t)[ :
When a collision takes place between molecules, it is
assumed that the proportions g and h of the excess kinetic
energy are transferred to bound and free molecules, re-
spectively. Moreover, constant proportions of the excess
energy (in this case g and h) are assumed to be transferred
during collision with bound and free molecules respec-
tively. However, the details of the energy transfer mecha-
nism and its formation is given in [14]. Consequently, the
total energy transfer due to evaporation is:
_

N
sx
2l
2
V
sx
h k
B
P
fm
exp
[x s[
l
_ _
[/(s; t) /(x; t)[ ds

N
sx
2l
2
V
sx
g k
B
P
bm
exp
[x s[
l
_ _
[/(s; t) /(x; t)[ ds

N
sx
l
2
V
sx
P
h
exp
[x s[
l
_ _
U(s) U(x) 3k
B
[/(s; t) /(x; t)[ [ [ ds : (4)
In laser heating process, in general, the material is
heated beyond the evaporation temperature. Therefore,
when formulating the problem, the conduction contribu-
tion should be added to convection part. Consequently, the
total energy equation becomes:
qC
p
o
ot
T(x; y; z; t)
=
I
o

2p
_
a
exp
(y u
s
t)
2
z
2
a
2
_ _
d exp(dx)
3f
k
2k
2
T(x; y; z; t)

fk
4k
3
_

0
exp
[x s[
k
_ _
T(s; y; z; t) ds
_

_
x
0
exp
[x s[
k
_ _
T(s; y; z; t) ds

_

x
exp
[x s[
k
_ _
T(s; y; z; t) ds
_

fk
2k
3
_

0
exp
[y g[
k
_ _
T(x; g; z; t) dg
_ _

fk
2k
3
_

0
exp
[z f[
k
_ _
T(x; y; f; t) df
_ _

N
sx
2l
2
V
sx
h k
B
P
fm
exp
[x s[
l
_ _
[/(s; t) /(x; t)[ ds

N
sx
2l
2
V
sx
g k
B
P
bm
exp
[x s[
l
_ _
[/(s; t) /(x; t)[ ds

N
sx
l
2
V
sx
P
h
exp
[x s[
l
_ _
U(s) U(x) 3k
B
[/(s; t) /(x; t)[ [ [ ds :
(5)
However, when solving the conduction part, Eqs. (1)
and (2) has been taken into account and they should be
solved simultaneously to obtain electron and lattice site
atom temperatures. To determine the evaporating front
velocity (V
s
), the following equation can be used [16]:
V
s
=

k
B
/(0; 0; 0; t)
2pm
_
exp
U
o
k
B
/(0; 0; 0; t)
_ _
: (6)
It should be noted that Eq. 5 is valid for temperatures
greater than evaporation temperature. Hence, by knowing
the surface temperature, the value of V
s
can be obtained.
3
Numerical solution to governing equations
and laser pulse property
The energy equation governing heating processes is in the
form of integro-differential equation, which does not yield
analytical solution. Therefore, a numerical method using
an explicit scheme is introduced to solve the equations.
The general procedure of the discretization process is
given in the appendix. In the case of conduction limited
heating, the stability criteria should satisfy:
3fkDt
2k
2
qC
p
_ 1
The calculated time increment meeting the stability crite-
rion is of the order of 10
9
s and this value is used in the
computation. The stability requirements for conduction
and phase change processes should meet:
3fkDt
2k
2
qC
p

P
fm
h k Dt
2l
2
qC
p
_

P
bm
g k Dt
2l
2
qC
p

2P
h
k Dt
2l
2
qC
p
_
_ 1
The time increment satisfying the stability criterion is of
the order of 10
10
s and it is this time increment which is
being used in the computation.
3.1
Pulse property
The laser power intensity distribution is shown in Fig. 2.
1=e points of the power intensity distribution correspond
to a diameter of 260 lm. The temporal variation of the
heating pulse is also considered resembling the actual laser
output pulse (Fig. 2). The energy contents of the actual
pulse was measured and then adapted for the theoretical
prole ensuring that both pulse proles have the same
energy contents.
4
Experimental
The experimental setup is shown in Fig. 3. An Nd-YAG
laser delivering output energy of 715 J in 1:48 ls pulses
499
was used to irradiate the workpiece surface. A care was
taken not to exceed the limiting value of the output power
intensity, which results in surface evaporation of the
substrate. An energy-power meter was used to monitor the
laser output energy and power. 51 mm nominal focal
length lens was used to focus the laser beam. To measure
the surface temperature an optical method was employed.
The principle of the optical temperature measurement is
based on the emissive power ratios emitted from the
heated spot knowing that there is a unique ratio of
monochromatic emissive power intensities corresponding
to two spectral frequencies. Consequently, after obtaining
the emissive power ratios, the corresponding temperature
can be calculated, i.e. recording the temporal variation of
this emissive power ratios, the surface temperature prole
with time can be determined. The details of the measure-
ment is not given here due to lengthy arguments, but re-
ferred to [17]. The photodetector output are coupled into a
computer where the data processing was carried out.
A stainless steel thick sheet is used as workpiece.
The spectral analysis of the light emitted from the
heated spot was carried out using a grating spectrometer.
Two fast response photodetectors were used to detect the
light emitted from the irradiated spot. The detectors were
spaced 2 mm apart at the exit slid and the spacing of
photodetectors corresponded to 164 A spectral width. To
integrate the monochromatic emissive power intensity
seen by the photodetectors, a 5.62 A spectral band width
was considered, since the diameter of the photodetector
corresponded to 5.62 A spectral width. The spectrometer
was positioned at 45

to the incident laser beam and the


light emitted from the heated spot was guided onto the
spectrometer inlet slit by two F6.3 lenses. The spectrometer
slit width was selected as twice the image diameter of the
irradiated spot as seen at the inlet slit, which was of the
order of 1 mm. The error estimated from the present ex-
periment based on the 95% condence level is about 5%.
5
Discussions
The results obtained from the 3-dimensional electron-
kinetic theory model is discussed according with the
Fig. 2. Laser power intensity distribution used in
the computation and actual laser output pulse
Fig. 3. Experimental set up
500
consideration of a moving heat source. However, the time
taken for surface to reach the melting points is consider-
ably short because of the high power intensity employed in
the present study, which is of the order of 4 10
10
W/m
2
.
In addition, the limitations encountered in the computa-
tional time and memory storage, the heating time in
computation was limited to only couple of ls. However,
the distance taken by the laser beam during this time is a
fraction of a lm, since the laser scanning velocity is set as
0.3 m/s. Therefore, the inuence of the laser scanning
velocity on the resulting temperature proles is negligible.
Nevertheless, in order to account for the complete de-
scription of the physical process, the scanning motion of
the laser beam should be included in the analysis.
Figure 4 shows the 3-dimensional temperature proles
corresponding to two different heating times while Fig. 5
shows the temperature distribution along x-axis obtained
at different y-axis locations and for different heating times.
In general, the melting point moves towards the surface of
the substance as the y-axis location increases. This because
of the spatial distribution of the laser power intensity in y-
axis. Therefore, the depth of melt isotherm associated with
y = 90 lm is less than that correspond to y = 0 lm. In this
case, laser power intensity available at y = 0 lm is high,
which is equal to the peak power intensity. As the heating
progresses the depth of liquid zone extends further inside
the solid substrate. The energy absorbed by the substrate
due to laser radiation is dissipated through free molecular,
free molecule-bound molecule, and electron-lattice site
atom collisions. The rate of liquid phase formation in-
creases in the surface vicinity than internal energy gain in
this region. Consequently, the temperature at melting oc-
curs as the heating progresses in this region. However, the
internal energy increase dominates the conduction loses in
early heating times. In this case, the energy absorbed by
the electrons is transferred to the lattice site atoms through
successive collisions, which in turn results in increased
lattice site energy. As the distance from the irradiated spot
center increases, the electrons loss their excess energy due
to i) excess energy of the electrons are dissipated in the
surface vicinity because of the previous successive colli-
sions and ii) the laser power available below the surface
reduces because of the exponential decay of the absorbed
power, i.e. less laser energy is available in this region.
Consequently, the energy conducted cannot be substanti-
ated resulting in rapid temperature decay in the surface
Fig. 4. a 3-dimensional view of temperature proles for
t = 0:00613 ls; b 3-dimensional view of temperature proles for
t = 1 ls
Fig. 5. a Temperature proles along x-axis for y = 0; b temper-
ature proles along x-axis for y = 20 lm; c temperature proles
along x-axis for y = 90 lm
501
vicinity. However, the rate of temperature decay changes
as the y-axis locations increase. This is also evident from
Fig. 6, in which dT=dx variation along x-axis is shown.
This may occur because of the laser power intensity dis-
tribution along y-axis as indicated earlier. However at the
surface (x = 0), the temperature gradient reaches to
minimum, which may indicate that the attainment of
equilibrium occurs at this point, i.e. the energy absorbed
by the substrate balances the energy losses due to phase
change, conduction and internal energy gain of the sub-
strate.
Figure 7 shows the temperature proles along y-axis
locations obtained at different x-axis locations and for
different heating times. As the heating progresses, the
depth of the liquid layer extends to 80 lm along y-axis and
a depth of 5 10
8
m for 1 ls heating time. This may
indicate that the electron-lattice site atom collision is the
sole mechanism in the surface vicinity. The rate of tem-
perature decay along y-axis is considerably less than that
occurs along x-axis. This is because of the laser energy
absorbed in the surface vicinity extends to some distance
in y-axis due to the spatial distribution of the intensity in
this region. In this case, electrons gain extra energy from
the laser radiation and dissipates this extra energy through
collisional process in this region. Therefore, the lattice site
atoms at some depth gain energy through this collisional
process, which in turn results in high lattice site temper-
ature, i.e. the small temperature gradient is attained along
the y-axis inside the solid substrate.
Figure 8 shows the temporal variation of the surface
temperature proles associated with different y-axis loca-
tions. The temperature increases rapidly at y = 0 where
the symmetry axis of the laser power intensity distribu-
tions is located. However, as the distance from this loca-
tion increases the rise of temperature proles to reach the
melting temperature reduces. Therefore, the inuence of
laser power intensity distribution is evident. The rapid rise
of temperature may indicate that the high rate of energy
absorption results in increased rate of collisions. In this
case, the excess energy of the electrons in the surface vi-
cinity is extracted substantially and remaining excess en-
ergy of the electrons may not be sufcient to increase the
energy of the lattice site atoms at some depth below the
surface. Hence, as the electron excess energy increases the
rate of energy extracted from the electron increases, in this
case, rapid rise of lattice site temperature occurs in the
surface vicinity. As the temperature reaches the melting
temperature, it remains almost constant at this tempera-
ture. This may indicate that the energy absorbed, due to
Fig. 6. Temperature gradient (dT=dx) along x-axis for
t = 0:00613 ls
Fig. 7. a Temperature proles along y-axis for 0:0613 ls heating
time; b temperature proles along y-axis for 0:141 ls heating
time; c temperature proles along y-axis for 1 ls heating time
Fig. 8. Temporal variation of surface temperature obtained from
theory and experiment
502
incident laser radiation, in the surface vicinity balances the
energy consumed for the liquid phase formation and
conduction losses, i.e. the internal energy gain of the
substrate remains the same. However, when comparing the
theoretical predictions with the experimental ndings,
both results are found to be in a good agreement. The
small discrepancies between both results may be attributed
to the experimental errors and the assumptions made in
the analysis.
6
Conclusions
Electron-kinetic theory approach introduces a new heating
model which is appropriate to the laser workpiece inter-
action process. However, the heating process takes place
due to free molecular, free molecule-bound molecule, free
molecule-vacancy, electron-free molecule and electron-
lattice site atom collisions. Therefore, the rate of collision
and the distribution of these species are the determining
factors in the heat transfer mechanism. In the initial phase
of the laser heating process, the energy gain by the elec-
trons is transferred to lattice site atoms at a high rate in the
surface vicinity of the substrate. This increases the internal
energy gain of the substrate rapidly in this region. How-
ever, as the distance from the surface increases, the excess
energy of the electrons reduces due to the previous suc-
cessive collisions and the energy absorbed because of the
incident laser radiation also reduces in this region, since
the absorbed energy decays exponentially with increasing
depth. Consequently, energy transferred by the electrons
to the lattice side atoms during the successive collisions
reduces in this region.
The inuence of the laser power intensity distribution,
across the heated spot, on the resulting temperature pro-
les is more pronounced in the surface vicinity, since the
temperature gradient along y-axis becomes considerably
lower than that occurs along x-axis. However, this inu-
ence reduces as the distance from the surface towards the
solid bulk increases.
As the heating progresses, the melting temperature
occurs in the surface vicinity and the temperature attains
almost constant at some depth in the substrate. In this
case, internal energy increase is minimal and the energy
absorbed due the incident laser radiation is consumed
through phase change process, i.e. the number of mole-
cules in the liquid phase increases.
References
1. Yilbas BS (1987) Study of affecting parameters in laser hole
drilling of sheet metals. Trans of ASME, J of Eng Materials and
Technology 109: 282287
2. Majumder J; Steen WM (1980) Heat transfer model for cw
laser material processing. J Appl Phys 51(2): 941947
3. Yilbas BS; Sami M (1997) Heat transfer analysis of a semi-
innite solid heated by a laser beam. Heat and Mass Transfer
32: 242253
4. Zhang HJ (1990) Non-quasi-steady analysis of heat conduc-
tion from a moving heat source. ASME, J Heat Transfer 112:
777779
5. Yilbas BS (1997) Analytical solution for time unsteady laser
pulse heating of semi-innite solid. Int J Mechanical Sciences
39(6): 671682
6. Schultz D; Becker D; Franke J; Kemmerling R; Herziger G
(1993) Heat conduction losses in laser cutting of metals.
J Phys D: Appl Phys 26: 13571363
7. Yilbas BS (1995) Study into a numerical solution for a pulsed
CO
2
laser heating process. Numerical Heat Transfer, An In-
ternational Journal of Computation and Methodology: Part
A Applications, 28(4): 487502
8. Diniz N; Lima C (1994) Non-linear three-dimensional tem-
perature proles in pulsed laser heated solids. J Phys D: Appl
Phys 27: 17951804
9. Blackwell FJ (1990) Temperature prole in semi-innite body
with exponential source and convective boundary conditions.
ASME J Heat Transfer 112: 567571
10. Duley WW (1983) Laser processing and analysis of materials.
1st Ed., Plenum Press, New York
11. Yilbas BS (1997) Laser heating process and experimental
validation. Int. J Heat and Mass Transfer 40(5): 11311143
12. Yilbas BS (1988) The validity of Fourier theory of radiation
heating of metals. Res Mechanica 24: 377382
13. Yilbas BS (1986) Heating of metals at a free surface by laser
radiation an electron kinetic theory approach. Int J Engng Sci
24(8): 13251334
14. Yilbas BS; Sami M (1995) Laser heating mechanisms includ-
ing evaporation process semiclassical and kinetic theory
approaches. Jpn J Appl Phys 34: 63916400
15. Dresselhaus SB; Kazeroonian A; Moodera J; Face D; Cheng T;
Ippen E; Dresselhaus G(1990) Femtosecond room temperature
measurement of the electron-phonon coupling constant k in
metallic superconductors. Phys Rev Letters 64(18): 21722175
16. Yilbas BS; Sahin AZ; Davies R (1995) Laser heating mecha-
nism including evaporation process initiating the laser drill-
ing. Int J Mach Tools and Manufact 35(7): 10471062
17. Yilbas BS (1986) Optical methods for measurement of the
temporal variation of temperature at the surface of a metal
heated by a laser beam. J of Optics 15(4): 108116
Appendix
Explicit formulation
The energy equation resulting from the kinetic theory
analysis is,
qC
p
d
dt
T(x; y; z; t)
=
I
o

2p
_
a
exp
(y u
s
t)
2
z
2
a
2
_ _
d exp(dx)
3f
k
2k
2
T(x; y; z; t)

fk
4k
3
_

0
exp
[x s[
k
_ _
T(s; y; z; t) ds
_

_
x
0
exp
[x s[
k
_ _
T(s; y; z; t) ds

_

x
exp
[x s[
k
_ _
T(s; y; z; t) ds
_

fk
2k
3
_

0
exp
[y g[
k
_ _
T(s; g; z; t) dg
_ _

fk
2k
3
_

0
exp
[z f[
k
_ _
T(s; y; f; t) df
_ _
: (7)
Using the forward differences for the rst derivatives,
the energy equation may be written as
503
T(x; y; z; t dt)
= I
o
d
Dt
qC
p
exp
(y u
s
t)
2
z
2
a
2
_ _
exp(dx)

fk
4k
3
Dt
qC
p
_

0
exp
[x s[
k
_ _
T(s; y; z; t) ds
_

_
x
0
exp
[x s[
k
_ _
T(s; y; z; t) ds

_

x
exp
[x s[
k
_ _
T(s; y; z; t) ds
_

fk
2k
3
Dt
qC
p
_

0
exp
[y g[
k
_ _
T(s; g; z; t) dg
_ _

fk
2k
3
Dt
qC
p
_

0
exp
[z f[
k
_ _
T(s; y; f; t) df
_ _
T(x; y; z; t) 1
3f a dt
2k
2
_ _
(8)
This is the explicit nite difference representation of the
energy equation.
Evaluation of the integrals
The integrals involved in the energy equation are com-
puted using the trapezoidal method. According to this
method, a function f is computed from an initial point 1 to
a nal point n using the values of f at equally spaced
discrete points. Mathematically,
Integral =
h
2
f
1
2f
2
2f
3
. . . f
n
[ [
where f
1
; f
2
; . . . ; f
n
are the values of the function at points
1; 2; . . . ; n and h is the spatial increment between succes-
sive points.
Hence if the integrals are denoted by
G
1
=
_

0
exp
[x s[
k
_ _
T(s; y; z; t) ds
G
2
=
_
x
0
exp
[x s[
k
_ _
T(s; y; z; t) ds
G
3
=
_

x
exp
[x s[
k
_ _
T(s; y; z; t) ds
G
4
=
_

0
exp
[y g[
k
_ _
T(x; g; z; t) dg
G
5
=
_

0
exp
[z f[
k
_ _
T(x; y; f; t) df
then the trapezoidal approximation to one of these inte-
grals, G
1
, can be written as
G
1
=
Dx
2
exp
[x s
o
[
k
_ _
T(s
o
; y; z; t)
_
2 exp
[x s
1
[
k
_ _
T(s
1
; y; z; t)
_
exp
[x s
n1
[
k
_ _
T(s
n1
; y; z; t)
_
exp
[x s
n
[
k
_ _
T(s
n
; y; z; t)
_
; (9)
where
s
1
= Dx
s
2
= 2Dx
.
.
.
.
.
.
s
n
= nDx:
The expressions for G
2
to G
5
also involve the exponents,
so in order to generalize, the exponential terms are dened
as
A
7
(j; m) =

n1
j=1

n1
m=1
exp
[jDx mDx[
k
_ _
:
A
8
(j; m) =

n1
j=1

n1
m=1
exp
[jDx mDx[
k
_ _
:
A
81
(j; m) =

n1
j=1

n1
m=1
exp
[jDy mDy[
k
_ _
:
and
A
82
(j; m) =

n1
j=1

n1
m=1
exp
[jDz mDz[
k
_ _
:
Therefore, the integrals yield in a nal form:
G
1
=
Dx
2
_
A
8
(j; 1) T(1; y; z; t)
2

n
r=2
A
8
(j; r) T(r; y; z; t)
A
8
(j; n 1) T(n 1; y; z; t)
_
G
2
=
Dx
2
_
A
7
(j; 1) T(1; y; z; t)
2

j1
r=1
A
7
(j; r) T(r; y; z; t)
A
7
(j; j) T(j; y; z; t)
_
G
3
=
Dx
2
_
A
7
(j; j) T(j; y; z; t)
2

n
r=j1
A
7
(j; r) T(r; y; z; t)
A
7
(j; n 1) T(n 1; y; z; t)
_
G
4
=
Dy
2
_
A
81
(jy; 1) T(x; 1; z; t)
504
2

n
r=2
A
81
(jy; r) T(r; y; z; t)
A
81
(jy; n 1) T(x; n 1; z; t)
_
G
5
=
Dz
2
_
A
82
(jz; 1) T(x; y; 1; t)
2

n
r=2
A
82
(jz; r) T(r; y; z; t)
A
82
(jz; n 1) T(x; y; n 1; t)
_
:
As can be observed, the explicit representation gives the
future temperature at (x; y; z) in terms of the current
temperatures at (x; y; z) and its surrounding nodes. Hence,
knowing only the initial temperatures for all the nodes, the
individual nodal temperatures for the next time step can
be calculated. However, the explicit formulation faces the
problem of stability, and careful choice of the time in-
crement is required. The upper limit of the allowable time
increment is given by
3fKDt
2k
2
qC
p
_ 1 (10)
so that the second term on the right-hand side of the en-
ergy equation remains positive. If we let
A =
I
o
dDt
q C
p
exp
(y u
s
t)
2
z
2
a
2
_ _
B = 1
3fKDt
2k
2
qC
p
_ _
C =
f KDt
4k
3
qC
p
and G
1
to G
5
are as dened above, then equation (8) can
be written as

T(x; t Dt) = A exp(d x) B



T(x; t)
C (G
1
G
2
G
3
2G
4
2G
5
) (11)
and from there we obtain the stability criterion as
B _ 0 :
A computer program has been developed to solve the
energy equation.
505

You might also like