You are on page 1of 16

Original Article

Temperature evolution and material


removal mechanisms in nanosecondpulsed laser ablation of polycrystalline
diamond

Proc IMechE Part B:


J Engineering Manufacture
116
IMechE 2014
Reprints and permissions:
sagepub.co.uk/journalsPermissions.nav
DOI: 10.1177/0954405414535773
pib.sagepub.com

Qi Wu1, Jun Wang1, Chuanzhen Huang2 and Huaizhong Li3

Abstract
A study of the single-pulsed laser ablation process for a polycrystalline diamond is presented. A simulation of the laser
ablation process using a finite element model is carried out to understand the temperature evolution, material removal
process and mechanisms, as well as the other physical phenomena associated with this process, that is, carbon phase
transformation, liquid-phase ejection and vapour/plasma shielding effect. It is found that mass material removal can be
achieved through surface evaporation under a higher laser pulse energy. It is further found that diamond graphitization
under laser irradiation is responsible for heat losses due to the large heat accumulation in the graphitized diamond, while
cobalt melting suppresses the evaporation of cobalt phase because of the heat consumption for solidliquid transition.
Crater depth and surface formation are also investigated experimentally on the polycrystalline diamond using singlepulsed laser ablation. The predicted crater depths are in reasonably good agreement with the corresponding experimental results.

Keywords
Polycrystalline diamond, laser machining, temperature, material removal

Date received: 25 November 2013; accepted: 23 April 2014

Introduction
With highly attractive properties, such as extremely
high hardness, high fracture toughness and high degree
of chemical inertness, polycrystalline diamond (PCD) is
being widely used as tool materials for machining
difficult-to-machine materials such as glasses and ceramics as well as for drill bits in oil and gas exploration.1
However, the high hardness and brittleness of the PCD
materials make them difficult to machine with low
machining efficiency using the conventional grinding
process. Machining technologies using the electrical discharge machining (EDM) principle including wired
EDM, wire electro-discharge grinding (WEDG), electrical discharge milling and ultrasonic EDM are regarded
as the most desirable methods for shaping PCD because
of their capability to create complex shapes and less
dependent on the mechanical properties of the workpiece material.2,3 However, these EDM-related technologies are limited to processing electrically conductive
materials and are associated with tool wear and slow
material removal rate (MRR). Furthermore, they often

cause diamond grain detachments, residual stresses and


damages on the machined PCD surfaces.4
In recent years, studies have been undertaken to use
the laser machining technology for the high efficient
processing of brittle and hard materials such as PCD.
This includes the investigations to assess the feasibility
of applying short-pulsed lasers, ultra-short-pulsed lasers
and other novel laser-related machining techniques for
material processing.57 Precision and near damage-free
machining could be achieved by using lasers of ultra1

School of Mechanical and Manufacturing Engineering, The University of


New South Wales (UNSW), Sydney, NSW, Australia
2
Center for Advanced Jet Engineering Technology, School of Mechanical
Engineering, MOE Key Laboratory of High-efficiency and Clean
Mechanical Manufacture, Shandong University, Jinan, China
3
Griffith School of Engineering, Griffith University (Gold Coast Campus),
Southport, Queensland, Australia
Corresponding author:
Jun Wang, School of Mechanical and Manufacturing Engineering, The
University of New South Wales (UNSW), Sydney, NSW 2052, Australia.
Email: jun.wang@unsw.edu.au

Downloaded from pib.sagepub.com by guest on January 20, 2015

Proc IMechE Part B: J Engineering Manufacture

short pulses, but their capability at this stage of development makes them unsuitable for high MRR. Other
lasers like nanosecond-pulsed lasers are more desirable
for high efficient micromachining if the process-induced
damages can be properly controlled.
Some important studies on nanosecond-pulsed laser
machining have been carried out to investigate the
machining performance and material removal mechanisms in order to achieve a better control of the machining process.8,9 However, the ablation mechanisms
cannot be adequately understood to improve the
machining process through experimental studies since
the physical process involved cannot be observed or
measured experimentally. As such, various models have
been developed to simulate the lasermaterial interaction and understand the material removal mechanisms.
As the nanosecond-pulsed laser machining process
uses thermal energy for melting and vaporizing material, analytical models have been developed based on
heat conduction equations associated with energy conservation for predicting the temperature distribution
and material removal process for a particular target
material.10 However, the complicated model formulation and long computation time of the analytical
approaches make it difficult in dealing with non-linear
problems and complex boundary conditions.
By contrast, numerical techniques for solving thermal
equations can easily include the variation in material
parameters in the calculations. Furthermore, such calculations can be performed for almost any geometry of target material and for complex boundary conditions.
Numerical modelling of material processing with laser
pulses of nanosecond duration has been studied extensively for a better understanding of the material removal
process. These studies have been primarily focused on the
processing of metal, semiconductor and ceramic targets,
and focused on studying the process parameter selection,
material removal and surface formation, residual stress
and thermal damages aiming at process optimization.11
13
There has also been some attention given to investigating material removal mechanisms under various laser fluence regions, as well as the physical phenomena in the
ablation process in terms of the ionization and expansion
of vaporized plume, and energy and mass transfer
between the material and the vaporized matters.14,15
Despite these studies, little is known about the laser
PCD interaction in nanosecond-pulsed laser ablation.
An important aspect in the modelling of PCD machining by lasers is the capability of predicting the transient
diamondgraphite transition and different ablation
processes between the multiphase (diamond and cobalt
binder) of the PCD target which significantly affect the
absorption of laser energy and heat diffusion.
Furthermore, other phenomena take place in laser
ablation, such as the formation of vapour/plasma
plume, and recoil pressure induced by the evaporation
of materials, whose effects remain to be studied.
Likewise, the specific properties of PCD under the irradiation of laser pulses require a careful examination.

Nevertheless, some important studies have been


reported. Windholz and Molian16 modelled material
removal process and simulated the ablation threshold in
the excimer laser machining of a deposited diamond
layer through chemical vapour deposition (CVD), but
the other aspects such as carbon phase change and
plasma shielding effect have yet to be considered. The
femtosecond laser ablation process for PCD has also
been investigated using molecular dynamic (MD) simulation,17 although further studies are required to understand the various phenomena involved in the ablation
process. Moreover, models for predicting the graphitization of diamond under laser irradiation have been proposed.18 It should be noted that the above-mentioned
modelling efforts focus on only one or some of the physical phenomena occurred in the laser machining of PCD,
and an investigation is required to give an in-depth
understanding of the laserPCD interaction.
This study is attempted to understand the physical
phenomena associated with the laser ablation of PCD
through a numerical simulation and experimental investigation. A finite element (FE) model based on the heat
conduction equation is developed and used to simulate
the laser ablation process under different laser pulse
energies. The temperature-dependent thermophysical
properties and the thermal convection and radiation
boundary conditions are considered in the computation. PCD ablation experiment is performed using a
nanosecond-pulsed fibre laser. The ablation depths are
measured and ablation mechanisms are analysed under
different laser pulse energies. The computational model
is verified quantitatively by comparing the predicted
ablation depths by the model with the corresponding
experimental data. The laser heating and the associated
physical processes in the laser ablation process are then
analysed using the computational model.

Computational model
Computational domain and thermal analysis
Figure 1 shows a schematic view of the model setup. A
Gaussian-distributed laser beam with its spot centre at
the origin of a Cartesian coordinate system is considered. This gives an axisymmetric laser energy distribution on the target material. Thus, a two-dimensional
(2D) geometry along the cross section of the laser beam
and containing the laser beam axis is used for the computational model, and due to its axis-symmetry, a half
of the cross section is actually considered. Two major
compositions, cobalt as the bonding material and diamond grain, contained in the PCD material are of particular interest in understanding the laser ablation
mechanisms. The diamond grains have a large average
size of 25 mm in diameter, and the cobalt content is very
small (about 2%). The diamond grains and cobalt distribute randomly in the PCD from the fabrication process. It is apparent that a model to represent the actual
diamondcobalt arrangement insider the PCD will be

Downloaded from pib.sagepub.com by guest on January 20, 2015

Wu et al.

Figure 1. Schematic view of a laser beam profile and the ablation model at the cross section of the laser beam axis.

very complicated, if not impossible. To simplify the


model, a randomly chosen strip of 1.5 mm is embedded
in the FE model to represent the cobalt phase, as shown
in Figure 1, so that the model can facilitate the investigation of the ablation mechanisms for both the diamond and cobalt phases.
The PCD target is in the Z . 0 region with its surface fixed at Z = 0, while laser beam propagates in the
+Z-direction. Under the laser radiation, the temperature of the target increases and surface vaporization
may occur. As the target is being ablated, the target
surface recedes in the +Z-direction with a vaporization or material removal velocity of ve. In this coordinate system, the heat conduction equation in the PCD
can be given as19
rcp





T T
T
T
T
_

k
k

Qt
t
x
x
z
z

where r, cp and k are the mass density (kg/m3), specific


heat (J/kg K) and thermal conductivity (W/m K) of the
_
target material, respectively, and Qt
represents the
laser energy absorbed by a unit volume (W/m3) of the
target. In the ablation process, the heat losses associated with phase change, L (J/kg), are considered for the
latent heat of diamond graphitization Lg and evaporation Lvc, and the latent heat of cobalt melting Lm and
evaporation LvCo.

In this study, the laser beam has a spatial profile that


is rotationally symmetric about the axis of the laser
beam propagating into the target. The laser energy distribution in the radial direction can be approximately
formulated as the Gaussian function. Thus, in a coordinate system that is fixed with reference to the laser
beam, the laser energy deposited in the PCD can be
written as20




x2
Qx; z; t 1  Rf aI0 expaz exp  2 vt
wz
2

where Rf and a are the reflectivity and absorption coefficient of material, respectively, x is the radial distance
from the laser beam centre, wz is the laser beam radius
(mm) and v(t) is the temporal dependence of the laser
pulse (pulse shape).
In this study, the beam radius represents the full
width at which the intensity drops to 1/e2 of the maximum value, as shown in Figure 1. Using z as the distance between the focal plane and the working plane,
the laser beam radius wz at any working plane can be
expressed as21

wz w0

 2 !1=2
z
1
zR

Downloaded from pib.sagepub.com by guest on January 20, 2015

Proc IMechE Part B: J Engineering Manufacture

Figure 2. (a) Technical data of laser pulse shape in the time domain and (b) approximated uniform laser pulse shape in the time
domain.

where w0 is the beam radius at the focal plane and zR is


the Raleigh range.
Due to the complexity and high non-linearity of the
pulse shape v(t) in the time domain (Figure 2(a)), the
laser source is considered to have an equivalent uniform pulse by introducing a pulse shape coefficient kp
based on the technical date of the laser used in this
study, that is

kp vtdt 0:914
4
The total pulse energies under the actual and the
approximated condition in equation (4) are approximately equal. The pulse duration (42 ns) is defined as
the width where the intensity attenuates to ;0.82Ipeak
(Figure 2(a)). Thus, the time function of the laser intensity used in the model is given by (see Figure 2(b))

0:914; pulsing
vt
5
0;
others

Boundary conditions
The initial temperature of Ta = 300 K is applied to all
surfaces, where Ta is the ambient temperature. The top
surface in Figure 1 is exposed to the laser irradiation
and the corresponding convection and radiation. Thus,
the heat transfer balance across this surface can be
determined by19
ki



T
Iabs  ha T  Ta  ei sB T4  T4a
z

where Iabs is the laser intensity absorbed by the target


surface, ha is the convective heat transfer coefficient
(10 W/m2 K),22 ei is the emissivity of material for

radiation and sB is the StefanBoltzmanns constant


(5.67 3 1028 W/m2 K4).19
Since the depth of the laser-irradiated region is considerably smaller than the thickness of the PCD model,
the Neumann boundary condition for the bottom surface is applied where the boundary satisfies the adiabatic condition as denoted by
dT
0
dn

For all the other surfaces (considering the axissymmetry of the model shown in Figure 1), a cooling
condition considering free convection and thermal
radiation is used,19 that is




T T

ha T  T0 ei sB T4  T40
8
ki
x z

Surface evaporation. Material ablation by infrared lasers


in a nanosecond pulse duration takes place essentially
by a thermal process.23 Based on the laser intensities
applied in this study (1.526.07 GW/m2), normal vaporization is considered as the dominant material removal
process in laser ablation of PCD. Thus, the velocity of
material vaporized from the target surface, ve, can be
approximated by the HertzKnudsen equation,24 that is
r
1  b
mi
vei T
9
Ps Ts
ri
2pkB Ts
where b is the condensation coefficient and is 0.08 for
stationary vaporization;25 mi is the atomic mass of the
vaporized particle, that is, carbon and cobalt atoms; ri
is the mass density for carbon or cobalt; kB is the
Boltzmann constant (1.38 3 10223 J/K) and Ps(Ts) is
the equilibrium vapour pressure above the target

Downloaded from pib.sagepub.com by guest on January 20, 2015

Wu et al.

surface at a temperature Ts, which can be described by


the ClausiusClapeyron relation26



Lvi mi 1
1
Ps Ts Pb exp

10
Tbi Ts
kB
In equation (10), Pb and Tb are the reference boiling
pressure (105 Pa) and equilibrium boiling temperature,
respectively, and the other symbols are as defined in
Notation.
Prediction of diamond graphitization. In the nanosecondpulsed laser ablation process, PCD target is under fast
laser irradiation during which the atomic transformation of diamond to graphite occurs once the individual
carbon atoms overcome the potential barrier by
absorbing laser photons. Meanwhile, the temperature
rise in the lattice system stimulates the thermal graphitization of diamond grains.27 As introduced by
Strekalov et al.,28 a graphitization probability PG can
be employed for estimating the graphitization rate of
diamond in the laser field and is given by
p
3
p 2pmc =2 aD 3
PG
dE0 2
3
h2p
h 3



p 
3T
ea  hn
hn 3 1
3t
ea  
3 exp 
4 e a  
T
hn

laser machining of PCD, when temperature reaches the


melting point of cobalt, it is possible that the cobalt
phase is removed through melting and liquid-phase
ejection by the recoil pressure. Therefore, the ejection
velocity of liquid cobalt from the molten layer vl can be
expressed as33
s
3:23104 Ts Tm
2pr
1 pr  2:26 
wz
vl
t
13

rCo rCo
wz
where Tm is the cobalt melting temperature and pr is
the recoil pressure generated due to material evaporation upon the melted layer, which can be expressed as a
function of the incident laser intensity I, as well as the
cobalt thermal conductivity kCo, atomic mass mCo,
latent heat of vaporization LvCo and surface temperature Ts,34 as
2 q 3
kTs
1:69 4
mCo LvCo
5
14
pr Iz p
LvCo 1 2:2 mkCoLTs
Co vCo
Accordingly, the machined depth of cobalt due to
liquid-phase ejection hl can be estimated by
t

hl vl dt

15

11

where mc is the atomic mass of carbon (2 3 10226 kg),


aD is the lattice constant of diamond structure
(1.78 3 10210 m),28 d denotes the dipole moment
(7.69 3 10231 K m) produced during the graphitization
process,29 T represents the target temperature, ea
denotes the activation energy required for diamond
graphitization (5.4 eV),30 h is Planck constant
(4.135 3 10215 eV s),31 hn represents the photon energy
absorbed by carbon atom (1.12 eV) and E0 represents
the electric field induced by the electromagnetic wave
produced in the laser system, which can be expressed
by laser intensity I0, velocity of light c (3 3 1028 m/s)
and magnetism permeability in the vacuum m0
(4p 3 1027 N/A2),32 that is
p
E0 2I0 m0 c
12

Vapour/plasma shielding effect. The ablation of solids by


nanosecond laser pulses is usually associated with the
formation of vapour plume when evaporation is the
dominant ablation mechanism.35 When the laser intensity is higher than 109 W/cm2, the plasma plume may
form above the target surface as a result of the ionization of vaporized matters, so that plasma behaves similarly to a thick optical medium and part of the laser
radiation is shielded from reaching the target surface,
as illustrated in Figure 3. After the onset of material
ablation, the particle density and plume temperature
are affected by the rate of material evaporation and

Therefore, based on the target temperature and processing time, the transition probability of the carbon
phase from the diamond state (sp3-bonded) to the graphite state (sp2-bonded) can be estimated using equations (11) and (12).
Cobalt melting and liquid-phase ejection. In the process of
laser machining of metals, material removal takes place
primarily in the liquid and vapour phases. The evolving
vapour applies a recoil pressure on the surface which
ejects the molten material out of the ablated area and
creates a machined crater. In the nanosecond-pulsed

Figure 3. Schematic of vapour/plasma shielding effect.

Downloaded from pib.sagepub.com by guest on January 20, 2015

Proc IMechE Part B: J Engineering Manufacture

laser energy absorbed by the plume.36 Thus, the timedependent optical thickness of vapour/plasma plume
Lt can be expressed by the transient ablation depth
Dz and the absorbed energy Ea as37
Lt aDz bEa

16

where Dz is the ablation depth with time, and Ea is a


self-consistent quantity and is given by38,39
t

Ea t I0 t0 1  expLt0 dt0

17

The laser intensity reaching the target after passing


through the plume can be calculated based on the
LambertBeer law19
Iabs t It expLt

18

Equations (16)(18) for describing the dynamics of


laser radiation absorption by the plasma plume are
employed to estimate the attenuation of laser energy in
this study.

FE analysis
The thermal and ablation models described above are
solved using the FE code in ANSYS 12.1. Some trail
studies were carried out first to study the effect of
model dimension so that a compromise could be
reached between the model accuracy and computational time (both are expected to increase with the
model dimension within certain range). It has been
found that when x . 14 mm in the model, the temperature developed to a near-ambient condition under
all laser pulse energies and laser spot sizes considered
in the study. Thus, 15 mm is used in the X-direction for
the model. Likewise, to approximate the ablation depth
and the isolated temperature field in the Z-direction,
the depth of the FE model should be larger than the
maximum ablation depth and make sure that the bottom surface temperature is in the ambient condition. It
has been found that 8 mm in the Z-direction is sufficient to meet this need. As a result, the FE model
shown as a half of the 2D domain in Figure 1 is defined
at 15 3 8 mm in the X- and Z-axes, respectively.
A mesh independence study and a convergence test
were performed. It has been found that the maximum
temperature increased by only about 2% when the
mesh density (model dimension/element dimension)
was increased from 160 to 800, as shown in Figure 4.
Considering the solution accuracy and computation
time, the mesh density of 160 is used, which corresponds to the mesh size of 0.1 and 0.05 mm in the
X- and Z-axes, respectively. Thus, plane elements with
uniform mesh size of 0.1x 3 0.05z mm are used in the
FE model, so that the model has a total of 24,000 elements and 24,311 nodes.
In the computation, the time-dependent problem is
solved sequentially, by using the output of nth step as

Figure 4. Calculated maximum surface temperature versus


mesh density.

the initial conditions for the (n + 1)th step. In each


step of the simulation, computations were performed to
calculate the properties of the element and the laser
incident position based on the calculated temperature
and element coordinates. For the diamond phase, graphitization is assumed to take place if the graphitization
probability PG . 1 at the end of a particular step. For
the cobalt phase, melting is assumed to occur once the
element temperature reaches the melting point, and the
liquid-phase ejection depth is then calculated according
to equations (13)(15). The material removal via surface evaporation for the carbon and cobalt elements is
estimated according to the HertzKnudsen equation.
After one step of computation, the new surface (if a
material removal took place) is considered and the corresponding new elements are now heated by the laser
beam. The element temperature from the previous step
is applied as the initial temperature for this step. The
boundary conditions are to be updated after each step.
The computation repeats until the desired simulation
time is reached. The parameters used in the simulation
are given in Table 1, while the thermophysical and optical properties of the diamond, graphite and cobalt used
in the simulation are given in Tables 24.

Experiment
The experiment was conducted to ablate a PCD of
0.7 mm thickness sintered on a tungsten carbide (WC)
substrate. The PCD contains 92% diamond and 2%
cobalt with the Knoop hardness of 50008000 kg/mm2.
A Manlight nanosecond-pulsed Ytterbium laser operating at 1080 nm wavelength and 42 ns pulse duration
was used. The laser was randomly polarized due to the
birefringence of the optical fibre as supplied by the
manufacturer. The laser beam was expanded by an 8X
beam expander before being delivered to a lens of
50 mm focal length where it was focused onto the PCD
surface, as shown in Figure 5. As for the simulation,
four levels of laser pulse energy (E = 0.2, 0.4, 0.6 and

Downloaded from pib.sagepub.com by guest on January 20, 2015

Wu et al.

Table 1. Parameters used in the finite element simulation.


Process parameters

Levels and values

Laser pulse energy (E) (mJ)


Laser beam diameter (D) (mm)
Model size (mm)
Element size (mm)
Time step (dt) (ns)
Pulse duration (t) (ns)
Ambient temperature (Ta) (K)
Coefficients of plasma plume a, b

0.2
0.4
20
15x 3 8z
0.1x 3 0.05z
0.5
42
300
a = 2.5 3 105, b = 5.7 3 1025 39

0.6

0.8

Table 2. Thermal and optical properties of diamond grains.40


Properties

Value

Density (rd) (kg/m3)


Thermal conductivity (kd) (W/m K)
Specific heat at constant pressure (cpd) (J/kg K)
Graphitization temperature (Tg) (K)
Latent heat of graphitization (Lg) (J/kg)
Thermal expansion (1026/K)
Emissivity (ed) (300 K)

4000
540
516 (300 K), 2058 (1800 K), 2192 (3000 K)
1050
1.58 3 105
3.8
0.81

Table 3. Thermal and optical properties of graphite.4145


Properties

Value

Density (rg) (kg/m3)


Thermal conductivity (kg) (W/m K)

2200
6 3 104/T when T \ 3000 K
20 when T . 3000 K
1727 + 0.333T 2 3.106 3 105/T when T 4 1500 K
2019.4 when T . 1500 K
4473
5.927 3 107
1.992 3 10226
0.212.83 3 1025 (T 2 300) when T \ 7000 K
0.02 when T . 7000 K
0.82
1.75 3 107

Specific heat at constant pressure (cpg) (J/kg K)


Boiling temperature (Tvc) at ambient pressure (K)
Latent heat of sublimation (Lvc) (J/kg)
Atomic mass of carbon (mc)
Reflectivity (Rfg)
Emissivity (eg) (300 K)
Absorption coefficient (ag) (m21)

Table 4. Thermal and optical properties of cobalt.4449


Properties

Value

Density of cobalt (rCo) (kg/m3)

20.432T + 8994.3 (solid)


7760 2 1.09 (T 2 TmCo) (liquid)
20.0367T + 94.2 (solid)
21.5 + 0.0082T (liquid)
0.22T + 380.06 (solid)
685.9 (liquid)
12.2
1768
3200
2.59 3 105
6.38 3 106
9.787 3 10226
0.67
0.28 (solid); 0.36 (liquid)
6.65 3 107

Thermal conductivity (kCo) (300 K) (W/m K)


Specific heat of cobalt at constant pressure (cpCo) (J/kg K)
Thermal coefficient of linear expansion (1026/K)
Melting temperature (TmCo) (K)
Boiling temperature (TbCo) (K)
Latent heat of melting (LmCo) (J/kg)
Latent heat of vaporization (LvCo) (J/kg)
Atomic mass of cobalt (mCo)
Reflectivity (RfCo)
Emissivity (eCo) (300 K)
Absorption coefficient (aCo) (m21)

Downloaded from pib.sagepub.com by guest on January 20, 2015

Proc IMechE Part B: J Engineering Manufacture

0.8 mJ) were investigated with six levels of the number


of laser pulses (N = 1, 2, 4, 8, 16 and 32). The repetition rate and laser beam diameter at the target surface
were keep constant at 20 kHz and 20 mm, respectively,
where the 20 mm beam size was achieved by adjusting
the relative position of the target surface with respect to
the laser focal position within the Raleigh range. A full
factorial design scheme was used to construct the ablation tests, resulting in 24 test runs each of which was to
generate a crater on the specimen. Each test was
repeated for three times. The maximum crater depth for
each test was measured using a laser three-dimensional
(3D) microscope (Keyence VK-X200) with a 0.5-nm
vertical resolution, and the average was taken as the
final reading. The ablation rate was then obtained from
dividing the crater depth by the number of laser pulses
applied.

Model verification
Figure 6 shows a comparison between the modelpredicted and experimental ablation depth per pulse
under the corresponding conditions. It can be seen
from the figure that without considering the vapour/
plasma shielding, the predicted ablation rate increases
almost linearly with the increase in pulse energy. The
predicted results are much higher than the experimental
results, particularly when the pulse energy is greater
than about 0.4 mJ. By contrast, the calculated results
tend to approach saturation with the increase in pulse
energy when considering the shielding effect and show
a reasonable agreement with the corresponding experimental data. It is thus apparent that the model considering vapour/plasma shielding effect can be regarded as
a good description of the PCD ablation process with
nanosecond-pulsed laser radiation. Considering the
inherent uncertainty of thermophysical properties of
the PCD and the laser output instabilities, the agreement between the model-calculated and experimental

ablation depths is reasonable when the vapour/plasma


shielding effect is considered.
The differences of the predicted ablation depths with
and without considering shielding effect are likely to be
due to the absorption of laser energy by the plume
formed during the pulse. In the laser ablation process,
the target surface temperature can be raised to and
maintained at a relatively high value due to the small
absorption depth of the target material. Hence, based
on the HertzKnudsen model,24 the vaporization flux
will be large, generating a relatively high mass density
in the small region right above the target surface in the
vapour/plasma plume. The high mass density may yield
relatively high laser energy absorption and plasma
radiation intensity. It is possible that a significant part
of the incoming energy is transferred to the plasma and
only part of it is dissipated in the target. Obviously, a
higher laser pulse energy produces a higher surface
temperature and vaporization flux, and hence, a larger
radiation intensity zone right above the target surface,
enhancing the shielding effect in the laserplasma interaction. Therefore, without considering the vapour/
plasma shielding effect, the model significantly overestimated the ablation depths under the higher laser energy
region.

Results and discussion


Crater formation
Figure 7(a) shows the top view of a typical crater
machined in the test, while Figure 7(b) depicts the
cross-sectional profile in a plane containing the laser
beam axis. The crater surfaces formed in single-pulse
ablation under different pulse energies are shown in
Figure 8, where the micrographs show the ablated area.
It can be noticed that under the pulse energy of 0.2 mJ,
microstructure changes occurred in the area irradiated
by the laser beam. This is because the laser energy was
not sufficient to perform a deep cut, but heated the target and caused phase transition.

Figure 5. Schematic of the laser modulation system.


PCD: polycrystalline diamond.

Downloaded from pib.sagepub.com by guest on January 20, 2015

Wu et al.

Figure 6. Comparison of measured (the scatter bars show


standard deviation) and simulated ablation depth under different
pulse energies (laser pulse frequency = 20 kHz).

Figure 7. Micrographs of the crater formed in single-pulse


ablation when E = 0.8 mJ, D = 20 mm and t = 42 ns: (a) crater
top view and (b) crater cross-sectional profile.

Recast was observed on the crater bottom surface


under the pulse energy of 0.4 and 0.6 mJ, as shown in
Figure 8(b) and (c), where the sign of liquid flow can be

seen. The recast has long tails under the pulse energy of
0.4 mJ, while the liquid appears to be scattered as the
laser pulse energy increased to 0.6 mJ. Since the recoil
pressure increases almost linearly with the incident laser
intensity, as shown in equation (14), the higher recoil
pressure under the higher laser pulse energy accelerates
the ejection of melted material and its collision with the
crater boundary so that part of the melt is broken into
small droplets to fly away. After the irradiation of the
laser beam, cooling and subsequent solidification took
place to form such surface features.
Further increasing the pulse energy to 0.8 mJ
resulted in new morphology features, where the bottom
surface of the crater has an ablated appearance as can
be seen in Figure 8(d), while no obvious melted phase
was found. It appears that thermal ablation in terms of
the normal vaporization dominates the material
removal under higher power densities.

Laser heating and ablation process


Figure 9(a)(f) shows some simulated temperature profiles on the PCD sample under a single laser pulse irradiation at different time durations from 1 to 50,000 ns,
using a 42-ns (full width at half maximum (FWHM))
laser beam with the pulse energy of 0.6 mJ and laser
beam diameter of 20 mm.
Under the condition, the surface temperature reaches
the boiling point of diamond and cobalt after 1 ns of
the laser irradiation, as shown in Figure 9(a), so that
the ablation of diamond and cobalt takes place at this
moment. The temperature gradient is gradual from target surface to inside the bulk. Since cobalt has a lower
thermal diffusivity than that of diamond, the absorbed
heat is conducted into a thin upper surface layer of the
cobalt phase within such a short time and leads to a
rapid rise in the surface temperature of cobalt, while the
temperature in the deeper layers remains unchanged.
By contrast, the thermal diffusivity of diamond is much
higher; the heat is thus conducted to the deeper layer of
diamond so that the surface temperature rises at a lower
rate, but the thickness of the heated layer is larger than
in cobalt, so is the ablated thickness of the diamond.
With further laser heating, the ablation of diamond
continues at 10 ns of the laser irradiation, as shown in
Figure 9(b). Meanwhile, the melting of cobalt takes
place by this moment and the melted phase of cobalt
requires further heating for evaporation.
As the irradiation time increases, the ablation of diamond continues to be faster than the evaporation of
cobalt, resulting in a depression in the diamond matrix,
as shown in Figure 9(c). Meanwhile, less heat is diffused
to the deeper layer of cobalt because of cobalt surface
melting and evaporation which takes energy from the
target, and cobalt temperature near the cobalt/diamond
boundary rises more slowly than the surrounding diamond. Consequently, heat is conducted from diamond
to cobalt which decreases the diamond temperature in
the vicinity of cobalt/diamond boundary. This

Downloaded from pib.sagepub.com by guest on January 20, 2015

10

Proc IMechE Part B: J Engineering Manufacture

Figure 8. Micrographs of surface morphology under different laser pulse energies (graphs show the central part of single-pulsed
laser-ablated area; D = 20 mm and t = 42 ns): (a) E= 0.2 mJ, (b) E = 0.4 mJ, (c) E = 0.6 mJ and (d) E = 0.8 mJ.

temperature decrease reduced the diamond evaporation


rate in this region, and thus steps are formed in the diamond matrix in the cobalt/diamond boundary, as
shown in Figure 9(b) and (c).
After 40 ns of the laser radiation, the depressions in
the diamond phase continue to increase and the cobalt
phase stands above that of diamond (Figure 9(d)).
Because of the attenuation of laser radiation that incidents on the target surface with vapour/plasma plume
shielding, the surface temperature starts to decrease
and the material ablation slows down. This is evident
from the reduced increase rate of the ablated layer
thickness from 20 to 40 ns compared to that from 10 to
20 ns.
At 50 ns (Figure 9(e)), the temperature has decreased
because the laser pulse period is 42 ns and there have
been already 8 ns in which no laser energy is applied, so
that the target temperature decreases because of heat
conduction and surface cooling. Finally, the temperature decreases to the ambient condition at 50,000 ns
(Figure 9(f)).

Carbon phase transition


Figure 10(a) and (b) shows, respectively, the variations
in the graphite layer thickness and propagation velocity
with time under different laser pulse energies. As heat
diffuses into the target and the associated material is

removed, the graphite layer thickness increases with


time, as shown in Figure 10(a). It is noted that under
all levels of pulse energy, the graphite layer thickness
increases initially, and then reaches saturation after a
certain period of time depending on the laser pulse
energy applied. The time needed to reach the saturation
decreases with an increase in the laser pulse energy,
while the graphite layer thickness at the saturation
stage increases with pulse energy. For instance, the graphitization depth increases over almost the entire laser
pulse-ON time under the laser pulse energy of 0.2 mJ,
while it reaches the saturation before 20 ns with the
0.8 mJ pulse energy. This is because when the incident
energy density is below or just slightly exceeds the ablation threshold, heat is quickly conducted into the PCD
and results in thermal graphitization. As soon as the
rate of energy consumption (by heat diffusion and for
material ablation) and energy absorption from the laser
beam becomes equal, the phase transition in the target
is sustained.
It can be seen from Figure 10(b) that the propagation velocity of graphitediamond interface towards
the solid diamond bulk reaches the maximum value at
the very start of the laser pulse under all levels of pulse
energy, and then decreases abruptly before decreasing
gradually. In general, the thickness of the graphitized
layer is determined by the optical absorption
p depth
(a21) and thermal diffusion length ( 2at ) of the

Downloaded from pib.sagepub.com by guest on January 20, 2015

Wu et al.

11

Figure 9. Simulated temporal variations in temperature distribution in the PCD under the pulse energy of 0.6 mJ: (a) 1 ns,
(b) 10 ns, (c) 20 ns, (d) 40 ns, (e) 50 ns and (f) 50 ms (the strips within the dashed lines represent cobalt).

Figure 10. Simulated temporal variations in (a) graphite layer thickness and (b) graphitediamond interface propagation velocity in
PCD under different laser pulse energies (considering plasma shielding effect).

material. For nanosecond-pulsed laser irradiation, the


optical absorption depth of carbon is much smaller
than thermal diffusion distance which defines the

temperature profile in the PCD. Therefore, the initial


graphitized layer thickness is noticeably larger than the
optical absorption depth, and thus the propagation

Downloaded from pib.sagepub.com by guest on January 20, 2015

12

Proc IMechE Part B: J Engineering Manufacture

Figure 11. Simulated temperature profiles in PCD along the symmetry axis under different laser pulse energies (considering plasma
shielding effect).

velocity reaches a high value. As time progresses, the


graphite front propagation velocity depends on the
heat diffusion and the associated material ablation into
the target. The existence of graphite layer enhances the
accumulation of heat in the upper layer of the target
due to the higher heat capacity and lower thermal conductivity of graphite under higher temperature, and
thus suppresses heat diffusion into the deeper layer of
the target.

Transient temperature field


To investigate the effect of laser pulse energy on the
temperature field in the laser ablation process, the temporal evolution of the temperature profile in the PCD
along the laser beam axis is given in Figure 11(a) and
(b), respectively, for the laser pulse energy of 0.4 and
0.6 mJ, noting that the horizontal axis represents the
distance from the surface into the target. It can be seen
that under both conditions, the temperature decreases
with an increase in the distance from the target surface.
This is attributed to the fact that the absorbed laser
energy decreases exponentially with an increase in the
depth below the target surface, according to the
LambertBeer law.19
It can be noticed from Figure 11(a) that there is a
small turning from segment A to segment B on every
curve, which suggests that the temperature decreases
relatively slowly in the target upper layer as compared
to the deeper layer. This is because in the upper layer
(segment A) where graphitization has occurred, thermal

conductivity of graphite is lower than the diamond in


the substrate, which results in a slow heat diffusion
from the irradiated region into the target. Furthermore,
the higher heat capacity of graphite allows more heat to
be accumulated in the irradiated region. As a result, the
material in the upper layer gains a higher temperature,
and hence yields a higher surface evaporation rate.
Material ablation takes heat away from the surface and
a smaller thermal gradient results in a slower decrease
in temperature in the upper layer regions.
By contrast, since the diamond bulk covered by a
graphite layer is beyond the absorption depth, the contribution of the absorbed laser radiation to its temperature rise is not significant. Temperature change along
the depth direction is mainly driven by the temperature
gradient. The high rate of energy gain from the irradiated area increases the surface temperature, and
hence, the temperature gradient next to the surface
becomes high, which enhances the heat conduction into
the bulk. Thus, the temperature decrease is sharp near
the graphitediamond interface region (segment B). As
the distance from the target surface further increases
towards the bulk material, the temperature decrease
becomes gradual, as shown in Figure 11(a) (segment
C), due to the balance between the energy gained from
the laser irradiation and the energy transported to the
solid bulk through heat diffusion.
Comparing to the condition at 0.4 mJ pulse energy
shown in Figure 11(a), the temperatures under the pulse
energy of 0.6 mJ are higher at 10 and 20 ns, while there
is a sharp decay of the maximum temperature at 30 and

Downloaded from pib.sagepub.com by guest on January 20, 2015

Wu et al.

13

Figure 12. Simulated temporal variation in (a) ablation depth and (b) evaporation velocity under different laser pulse energies.
Magnified view of (c) ablation depth from 0 to 5 ns and (d) evaporation velocity from 0 to 5 ns (considering plasma shielding effect).

40 ns, as shown in Figure 11(b). This is attributed to


the higher evaporation rate at the early stage of the
laser irradiation under the 0.6 mJ pulse energy, which
enhances plume shielding effect, and thus a higher
attenuation coefficient to the laser beam. After 30 ns of
laser pulse, the plume that composed of evaporated
matters shielded most of the laser radiation from the
target, and thus temperature decreases.

Surface evaporation
The temporal variations in ablation depth and evaporation velocity under different laser pulse energies are
shown in Figure 12(a) and (b), respectively. For a
clearer presentation for the earlier stage of ablation, the
change in ablation depth and ablation velocity from
the start to 5 ns is shown in Figure 12(c) and (d). It can
be seen from Figure 12(c) that material ablation initiates immediately after the laser beam is applied and a

lower laser energy (e.g. 0.2 mJ) takes a longer time to


cause material removal than a higher pulse energy. A
maximum depth appears to exist for the laser pulse
energy applied and an increase in the pulse energy
increases the maximum ablation depth (Figure 12(a)).
It is noticed from Figure 12(b) that the ablation
velocity (or the velocity of material evaporation)
increases rapidly at the start of laser irradiation (segment A), and then rises gradually at segment B until it
reaches the maximum value, and finally decreases at
segment C. Ablation starts when the laser fluence
exceeds the threshold for ablation. Although the maximum energy absorbed from the irradiated area takes
place at the surface of the target, the rate of evaporation is limited because of the high value of the latent
heat of evaporation, so that the increase in evaporation
(or ablation) velocity becomes slow over a period from
segment A to segment B shown in Figure 12(b). During
the target material evaporation, a vapour/plasma

Downloaded from pib.sagepub.com by guest on January 20, 2015

14

Proc IMechE Part B: J Engineering Manufacture

Figure 13. Simulated temporal variation in cobalt melting


depth (considering shielding effect).

plume is formed above the target surface which blocks


a significant part of the incoming laser energy from
being dissipated into the target, and the laser pulse
energy absorbed by the target thus decreases, so does
the ablation velocity as shown in segment C in Figure
12(b). It is further noticed from Figure 12(b) that the
evaporation velocity is higher under a higher than a
lower laser pulse energy, as may be expected.
It can be seen from Figure 12(b) that the decrease in
ablation rate starts earlier with a higher than a lower
pulse energy. This may be due to the more significant
vapour/plasma shielding effect under a higher laser
energy where mass material evaporation takes place.50
During the laser irradiation, a higher laser pulse energy
produces a higher target surface temperature and a
higher vaporization flux, and thus a larger vapour/
plasma intensity zone right above the target surface. As
a result, some laser energy cannot reach the target surface, but is absorbed to heat the plasma. Meanwhile,
the evaporation takes some heat away from the target
surface, so that the overall effect causes the surface
temperature and the evaporation rate to decrease. A
higher pulse energy results in a quicker temperature
increase in the target and requires a shorter time to
cause material evaporation, and hence an earlier start
of decrease in the ablation rate.

Cobalt melting and ejection


The change in the cobalt melting depth with time is
shown in Figure 13 under different laser pulse energies.
At the early stage of laser irradiation, the heat absorbed
from the irradiated area is conducted to the solid bulk
and causes the cobalt to melt. At the interface of the
liquid and solid cobalt, the rate of heat diffusion from
the liquid to solid phase is high, thus increasing the

temperature in the solid region next to liquid.


Moreover, the melting of cobalt increases the energy
gain from the irradiated area due to the large latent heat
of melting of cobalt and enhances the solidliquid
transition.
It is noticed that the melting depth (melting layer
thickness) reaches a maximum value after a period of
time depending on the pulse energy applied where a
higher pulse energy requires a shorter time to arrive at
the maximum because of more energy input. With further laser irradiation, part of the heat absorbed from
the laser is used for material evaporation which
together with the vapour/plasma shielding effect
results in a reduced amount of energy transferred into
the solid bulk, and hence a reduced melting depth. For
the high pulse energies applied (such as 0.8 mJ), the
removal of cobalt phase is mainly through evaporation
because of more energy input, while the ejection of
molten cobalt through a recoil pressure from the melting process dominants the cobalt removal process
under relatively low laser pulse energies.
After the laser pulse of 42 ns, there is no further
energy input so that the temperature decreases and resolidification of cobalt takes place. As a result, the melting depth decreases rapidly, as shown in Figure 13. It is
noticed that the reduction in melting depth is smaller
under a higher than a lower laser pulse energy. This is
possibly due to the higher latent heat contained in the
liquid to build evaporation with a higher pulse energy
that helps to maintain the liquid phase for some time
after the laser irradiation. The difference between the
maximum melting depth during the pulse-ON period
and the melting depth during the pulse-OFF period
shown in Figure 13 should represent the amount of resolidified cobalt on the ablated surfaces after the irradiation of one laser pulse.

Conclusion
An FE model for single-pulsed laser ablation of PCD
has been developed, considering the diamond and
cobalt phases in the PCD. The model has been verified
by comparing the model-calculated ablation depth with
the corresponding experimental data which showed a
good agreement when the vapour/plasma shielding
effect was considered. From the experimental and
simulation studies, three material removal mechanisms
have been identified, namely, the graphitization of diamond for all the energy levels considered, surface evaporation under higher laser pulse energies and ejection
of molten cobalt under lower pulse energies.
The removal of cobalt has been found to initialize
earlier than that of diamond under a laser irradiation.
However, as the melting of cobalt consumes a large
amount of the absorbed laser energy, the MRR for
cobalt is smaller than diamond. The graphitization of
diamond results in heat accumulation in the surface
region which leads to a high surface temperature, and

Downloaded from pib.sagepub.com by guest on January 20, 2015

Wu et al.

15

thus a high evaporation rate. It has been noted that


while increasing laser pulse energy raises the ablation
rate, it enhances the plasma shielding effect which
decreases the laser intensity reaching the target, and
consequently reduces the efficiency of energy use.
Declaration of conflicting interests
The authors declare that there are no conflicts of
interest.
Funding
Q.W. would like to thank the China Scholarship
Council for providing a scholarship for her PhD study.
References
1. Asmussen J. Diamond films handbook. New York: Marcel Dekker, 2002.
2. Sheu DY and Cheng CC. Micro three-dimensional cavities tools fabrication on PCD by m-EDM scanning process. Mater Manuf Processes 2012; 28: 4247.
3. Lauwers B, Kruth JP and Brans K. Development of
technology and strategies for the machining of ceramic
components by sinking and milling EDM. CIRP Ann:
Manuf Techn 2007; 56: 225228.
4. Ho KH and Newman ST. State of the art electrical discharge machining (EDM). Int J Mach Tool Manu 2003;
43: 12871300.
5. Wu Q and Wang J. An Experimental investigation of the
laser milling process for polycrystalline diamonds. J
Mater Process Tech 2011; 291294: 810815.
6. Perry MD, Stuart BC, Banks PS, et al. Ultrashort-pulse
laser machining of dielectric materials. J Appl Phys 1999;
85: 68036810.
7. Tangwarodomnukun V, Wang J, Huang CZ, et al. An
investigation of hybrid laser-waterjet ablation of silicon
PCDs. Int J Mach Tool Manu 2012; 56: 3949.
8. Li ZQ, Wang J and Wu Q. Ultrashort pulsed laser
micromachining of polycrystalline diamond. Adv Mater
Res 2012; 497: 220224.
9. Kononenko TV, Meier M, Komlenok MS, et al. Microstructuring of diamond bulk by IR femtosecond laser
pulses. Appl Phys A: Mater 2008; 90: 645651.
10. Araya G and Gutierrez G. Analytical solution for a transient, three-dimensional temperature distribution due to a
moving laser beam. Int J Heat Mass Tran 2006; 49: 4124
4131.
11. Samant AN and Dahotre NB. Computational predictions
in single-dimensional laser machining of alumina. Int J
Mach Tool Manu 2008; 48: 13451353.
12. Dobrev T, Dimov SS and Thomas AJ. Laser milling:
modelling crater and surface formation. Proc IMechE,
Part C: J Mechanical Engineering Science 2006; 220:
16851696.
13. Yan YZ, Li L, Sezer K, et al. Experimental and theoretical investigation of fibre laser crack-free cutting of thicksection alumina. Int J Mach Tool Manu 2011; 51: 859
870.
14. Gragossian A, Tavassoli SH and Shokri B. Laser ablation
of aluminum from normal evaporation to phase explosion. J Appl Phys 2009; 105: 10330411033047.

15. Gusarov AV and Smurov I. Thermal model of nanosecond pulsed laser ablation: Analysis of energy and mass
transfer. J Appl Phys 2005; 97: 307320.
16. Windholz R and Molian PA. Nanosecond pulsed excimer
laser machining of chemically vapour-deposited diamond
and graphite - Part II - Analysis and modelling. J Mater
Sci 1998; 33: 523528.
17. Li ZQ, Wang J and Wu Q. Molecular dynamics simulation of the ablation process in ultrashort pulsed laser
machining of polycrystalline diamond. Adv Mater Res
2012; 500: 351356.
18. Strekalov VN. Graphitization of diamond stimulated by
electron-hole recombination. Appl Phys A: Mater 2005;
80: 10611066.
19. Cxengel YA. Heat and mass transfer: a practical approach.
3rd ed. New York: McGraw-Hill, 2007.
20. Jervis TR. Excimer-laser surface processing of materials.
In: LEOS93 conference proceedings: lasers and electrooptics society annual meeting, San Jose, CA, 1518
November 1993, pp.772773. New York: IEEE.
21. Kogelnik H and Li T. Laser beams and resonators. Pr
Inst Electr Elect 1966; 54: 15501567.
22. Tate A. A theory for the deceleration of long rods after
impact. J Mech Phys Solids 1967; 15: 387399.
23. Bauerle D. Laser processing and chemistry. 4th ed. Berlin/
Heidelberg: Springer, 2011.
24. Kelly R and Miotello A. On the role of thermal processes
in sputtering and composition changes due to ions or
laser pulses. Nucl Instrum Meth B 1998; 141: 4960.
25. Martynyuk MM. Phase explosion of a metastable fluid.
Combustion Explosion and Shock Waves. Combust
Explo Shock+ 1977; 13: 178191.
26. Miotello A and Kelly R. Critical-assessment of thermal
models for laser sputtering at high fluences. Appl Phys
Lett 1995; 67: 35353537.
27. Bundy FP, Bassett WA, Weathers MS, et al. The pressure-temperature phase and transformation diagram for
carbon; updated through 1994. Carbon 1996; 34: 141153.
28. Strekalov VN, Konov VI, Kononenko VV, et al. Early
stages of laser graphitization of diamond. Appl Phys A:
Mater 2003; 76: 603607.
29. Lin JF, Lin JW and Wei PJ. Thermal analysis for graphitization and ablation depths of diamond films. Diam
Relat Mater 2006; 15: 19.
30. Kittel C. Introduction to solid state physics. 8th ed. Hoboken, NJ: Wiley, 2005.
31. Omne`s R. Understanding quantum mechanics. Princeton,
NJ: Princeton University Press, 1999.
32. Yariv A. Optical electronics in modern communications.
5th ed. New York: Oxford University Press, 1997.
33. Semak V and Matsunawa A. The role of recoil pressure
in energy balance during laser materials processing. J
Phys D: Appl Phys 1997; 30: 25412552.
34. Harimkar SP, Samant AN and Dahotre NB. Temporally
evolved recoil pressure driven melt infiltration during
laser surface modifications of porous alumina ceramic. J
Appl Phys 2007; 101: 05491110549118.
35. Granse G and Vollmar S. Results of modeling laserinduced plasma in pulsed laser deposition. IEEE T
Plasma Sci 1996; 24: 4748.
36. Krokhin ON. Physics of high energy density. New York:
Academic Press, 1971, pp.278305.
37. Bulgakov AV and Bulgakova NM. Thermal model of
pulsed laser ablation under the conditions of formation

Downloaded from pib.sagepub.com by guest on January 20, 2015

16

38.

39.

40.
41.
42.

43.

44.
45.
46.
47.
48.
49.

50.

Proc IMechE Part B: J Engineering Manufacture


and heating of a radiation-absorbing plasma. J Quantum
Electron 1999; 29: 433437.
Tokarev VN, Lunney JG, Marine W, et al. Analytical
thermal-model of ultraviolet-laser ablation with singlephoton absorption in the plume. J Appl Phys 1995; 78:
12411246.
Bulgakova NM, Bulgakov AV and Babich LP. Energy
balance of pulsed laser ablation: thermal model revised.
Appl Phys A: Mater 2004; 79: 13231326.
Zaitsev AM. Optical properties of diamond: a data handbook. Berlin/Heidelberg: Springer-Verlag, 2001.
Bundy FP. Melting point of graphite at high pressure heat of fusion. Science 1962; 137: 10551057.
Chupka WA and Inghram MG. Direct determination of
the heat of sublimation of carbon with the mass spectrometer. J Phys Chem 1955; 59: 100104.
Malvezzi AM, Bloembergen N and Huang CY. Timeresolved picosecond optical measurements of laser-excited
graphite. Phys Rev Lett 1986; 57, 146149.
Ho CY, Powell RW and Liley PE. Thermal conductivity
of the elements. J Phys Chem Ref Data 1972; 1: 279421.
Grigoriev IS and Meilikhov EZ. Handbook of physical
quantities. Moscow: CRC Press, 1995.
Iida T and Guthrie RIL. The physical properties of liquid
metals. New York: Oxford University Press, 1988.
Wahlin HB and Knop HW. The spectral emissivity of
iron and cobalt. Phys Rev 1948; 74, 687689.
Brandes EA and Brook GB. Smithells metals reference
book. 7th ed. Oxford: Butterworth-Heinemann, 1998.
Makino T, Kawasaki H and Kunitomo T. Study of
the radiative properties of heat resisting metals and
alloys -1. optical constants and emissivities of nickel,
cobalt and chromium. Bull JSME 1982; 25: 804811.
Willmott PR and Huber JR. Pulsed laser vaporization
and deposition. Rev Mod Phys 2000; 72: 315328.

Appendix 1
Notation
a, b
aD
c
cp
d
D
E
Ea
E0
ha
hl
I
Iabs

coefficient of plasma plume


lattice constant of diamond structure
light velocity
heat capacity at constant pressure
dipole moment
laser beam diameter
laser pulse energy
laser energy absorbed by plasma plume
electric field
heat convection coefficient
liquid ejection depth
laser intensity
laser intensity absorbed by target

zR
Dz

maximum laser intensity at the beam


centre
thermal conductivity
Boltzmanns constant
laser pulse shape coefficient
latent heat of vaporization
atomic mass
initial atom density
surface pressure in normal condition
graphitization probability
recoil pressure
surface pressure
laser heat source
reflectivity
time
temperature
ambient temperature
boiling temperature
melting temperature
surface temperature
surface evaporation rate
melted layer ejection velocity
laser beam radius at focal plane
laser beam radius at z distance from the
focal plane
Cartesian system coordinate
Cartesian system coordinate
physical thickness of the plasma plume (as
shown in Figure 3)
Raleigh length
ablation depth

a
ap
b
e
ea
L
l
m0
r
sB
t
v
v0
h
hn

absorption coefficient (of target)


absorption coefficient of plasma plume
correction factor
emissivity
activation energy
optical thickness of plasma plume
light wavelength
magnetism permeability
mass density
StefanBoltzmanns constant
pulse width
laser pulse shape function
radiation frequency
Planck constant
photon energy

I0
k
kB
kp
Lv
m
N0
Pb
PG
pr
Ps
Q
Rf
t
T
Ta
Tb
Tm
Ts
ve
vl
w0
wz
X
Z
Z(t)

Downloaded from pib.sagepub.com by guest on January 20, 2015

You might also like