Sie sind auf Seite 1von 5

X.

Zhao
Y. C. Shin1
e-mail: shin@purdue.edu
Center for Laser-Based Manufacturing,
School of Mechanical Engineering,
Purdue University,
West Lafayette, IN 47907

Ablation Dynamics of Silicon


by Femtosecond Laser and the
Role of Early Plasma
In this paper, the femtosecond laser ablation of silicon is investigated by a twodimensional hydrodynamic model. The ablation depth of the silicon wafer ablated in air
at different laser intensities is calculated, and the corresponding experimental measurements are carried out for validation. Two different ablation regimes have been identified
by varying the laser fluence. While two-photon absorption dominates in the low fluence
regime (<2 J/cm2), electron heat diffusion is a major energy transport mechanism at
higher laser fluences (>2 J/cm2). The ablation efficiency first increases with the laser fluence, and reaches the peak value at the laser fluence around 8 J/cm2. It starts to drop
when the laser fluence further increases, because of the early plasma absorption of the
laser beam energy. [DOI: 10.1115/1.4025805]

Introduction

Ultrashort pulse lasers have been proven to be powerful tools


for precise micromachining applications [1,2]. The major advantages of femtosecond laser ablation include reduced heat affected
zones by rapid pulse energy deposition, clean and precise machining, and capability of fabricating a wide range of materials.
Because of its advantages and attractive applications, a lot of studies have been dedicated to femtosecond laser ablation of various
materials, such as metals [3,4], semiconductors [5,6], and transparent materials [7,8]. However, many aspects of the femtosecond
ablation remain unclear, due to the involved complicated multiphysical phenomena, such as laser energy absorption, electron excitation, electron heat diffusion, electron-lattice energy exchange,
surface electron emission, ionization, plasma generation and
expansion, and etc.
In this work, a two dimensional computational model is presented for theoretically predicting the femtosecond laser ablation
process of silicon in air. The model [9] comprises a twodimensional (2D) axisymmetric hydrodynamic model and a twotemperature model (TTM), supplemented by a quotidian equation
of state (QEOS) model to calculate the thermal and electrical
properties of materials at different conditions. Parametric studies
are carried out to predict the ablation depth and ablation efficiency
by varying the laser fluence. The simulation results are validated
against the experimental measurement under the same condition.
The laser interaction with the early plasma and its effect on the
ablation efficiency at high fluence are also investigated.



@Te 1 @
@Te
@ 2 Te
ke r

2  GTe  Ti S
Ce
r @r
@t
@r
@z
Ci

2.1 Energy Transfer Inside the Target. For ultrashort lasermaterial interactions, the electron and lattice temperatures are different before the thermodynamic equilibrium is reached, and
hence the energy transfer inside the target can be described by the
well-known two-temperature model [10,11]

2.2 Laser Absorption. The laser absorption term S can be


calculated as [14]
S aI bI 2 Hne I  Eg

@ne
@t

(3)

where a and b are single- and two-photon absorption coefficient,


H is the free carrier absorption coefficient, Eg is the band gap of
the semiconductor, ne is the electron number density, I is the laser
intensity, which has a Gaussian profile for both spatial and temporal distributions.
2.3 Electron-Hole Dynamics Inside Target. The electron
dynamics in silicon are described by [12,13] (hole dynamics are
similar)
@ne 1
 r  J~ Ge  Re
e
@t

(4)

where J~ is the electric current density including both drift and diffusion terms
~ eDe rne
J~ ene lE

(5)

In above equation, l and De are the electron mobility and diffu~ is the electric field generated because
sivity, respectively, and E
of breaking of local neutrality inside the target. In Eq. (4), Ge and
Re represent the electron generation and loss term, respectively.
The term Ge is given by [1214]
Ge

1
Corresponding author.
Manuscript received April 29, 2013; final manuscript received October 18, 2013;
published online November 18, 2013. Editor: Y. Lawrence Yao.

(2)

where Ce ; Ci ; Te ; Ti are the volumetric heat capacities and temperatures of electrons and lattice, respectively, ke is the electron thermal conductivity, G is the electron-lattice coupling term, all the
values of which are obtained from QEOS model, and S is the laser
absorption term.

Theoretical Model

The schematic diagram for the model setup is shown in Fig. 1.


The silicon target is located in z < 0 region, and the region where
z > 0 is ambient gas (air). The silicon-air interface is located at
z 0, and the symmetric axis is at r 0.

@Ti
GTe  Ti
@t

(1)

1  RaI 1  R2 bI 2
na

dne
hv
2hv
na ni

(6)

where d is the impact ionization coefficient, hv is the photon


energy, R is the surface reflectivity, and na , ni are atom and ion
number density, respectively.

Journal of Manufacturing Science and Engineering


C 2013 by ASME
Copyright V

DECEMBER 2013, Vol. 135 / 061015-1

Downloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 12/20/2014 Terms of Use: http://asme.org/terms

vei is the electron (ion) momentum,


where !
p ei mei nei~
Pei nei kB Tei is the electron (ion) pressure, and
eei 1:5nei kB Tei 0:5mei nei v2ei is the total energy of electron (ion) with mei as the electron (ion) mass, kB as the Boltzmann constant, and ~
vei as the electron (ion) velocity. Tei is the
electron (ion) temperature. Here, ejected electrons are treated as
ideal gases in thermal equilibrium. Sei is the source
! term for electron (ion) generation in the bulk of the plasma, f ei is the force
term consisting of electric forces, collision forces and ponderomotive force, ~
qT is the thermal flux of electron (ion) by conduction,
and Q is the energy term generated by collision, impact ionization,
and recombination.

Fig. 1 Schematic diagram of the model setup

The term Re denotes the three photon Auger recombination process given by [15]
Re cn3e

(7)

with c the Auger recombination coefficient.


There will be surface electron emission at the silicon surface,
and the total emission rate is the summation of thermal emission
rate Jth and the n-photon photoelectric emission rate Jn [16]


U
2
(8)
Jth ATes exp 
kB Tes


 e n
nht  U
2
Jn a n
(9)
AI n 1  Rn Tes
F
ht
kB Tes
where Tes is the surface electron temperature, A is a universal constant equal to 120 A/cm2K2, U is the work function of the material, h is the Planck constant, t is the photon frequency, F is the
Fowler function, and an is the n-photon photoelectric emission
coefficient. The photoelectric emission takes place when the total
photon energy absorbed by surface electrons is higher than the
work function to overcome the surface energy barrier and escape
from the solid. For a wavelength of 800 nm in this study, the photon energy ht is 1.55 eV, and the work function U for silicon is
4.6 eV [17]. Therefore, three photons are necessary to overcome
the work function and three-photon photoelectric emission is
expected to be the dominant mechanism [12]. The three-photon
emission coefficient is a3 1030 cm2 s=C3 [16].

2.5 Numerical Solution. The simulation starts at t 0, when


the temperature of both the target material and the ambient gas
stays at room temperature (Te Ti 300K), the velocities of
electrons, and ions are all set to zero (~
vei 0), the total pressure
P is set to 1 atm. The TTM and the electron dynamics inside target
are first solved in the region where z < 0 by finite volume method
[18]. At the siliconair interface, the hot electron emission rates
are calculated based on the obtained target surface temperature,
and are implemented as the boundary conditions for Eqs.
(10)(12) for the early plasma. The hydrodynamic equations are
solved by convex essentially nonoscillatory high order schemes
[19] in the region where z > 0 to simulate the early plasma behavior. Free boundary conditions are adopted at the siliconair interface, and the plasma and air are assumed to be able to move
freely. The simulation domain is chosen to be large enough such
that the boundaries of the domain are not disturbed at all. Therefore, fixed boundary conditions are applied at the upper, lower
and right boundaries of the simulation domain. At the left boundary (r 0), the values at points left of the r 0 axis are equal to
those right of the axis, because of the symmetric properties of the
simulation.

Experimental Setup

Experiments were carried out to measure ablation depth of silicon at different laser intensities. The experimental setup is shown
in Fig. 2. A regeneratively amplified Ti:sapphire laser (Spitfire,
Spectra Physics) was used, which has the characteristics of 100 fs
pulse duration, the central wavelength of 800 nm and the repetition rate of 1 kHz. The pulse energy could be adjusted from 5 lJ
to 1 mJ by a half waveplate and a polarizer. A 10 microscope
objective was used to focus the laser beam onto the target for ablation. A commercial grade silicon wafer was used as the target, and
it was irradiated by a single pulse for every spot. The ablation
depth and crater shape were measured by an optical surface profilometer (ADE MicroXAM).

2.4 Plasma Dynamics. The electron and ion dynamics inside


the plasma during the laser ablation can be simulated by hydrodynamic equations [9]
@nei
vei Sei
r  nei~
@t
@!
p ei
!
vei rPei nei f ei
r  !
p ei~
@eei@t
vei r  Pei~
vei  r  ~
qT Q
r  eei~
@t

061015-2 / Vol. 135, DECEMBER 2013

(10)
(11)
(12)
Fig. 2

Experimental setup

Transactions of the ASME

Downloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 12/20/2014 Terms of Use: http://asme.org/terms

Fig. 3 Spatial distribution of the lattice temperature at 10 ps.


Laser pulse duration: 100 fs, wavelength: 800 nm, laser fluence:
2.5 J/cm2.

Results and Discussions

It is revealed that at the fluence above 0.5 J/cm2, the dominant


ablation mechanisms for ultrashort laser ablation might be critical
point phase separation (CPPS) [20], fragmentation [21], and vaporization [21]. However, since fragmentation and vaporization
occur in the region near the target surface, the total ablation depth
will be determined by CPPS [21]. When CPPS occurs, the ablation depth is determined by the location where the local temperature exceeds the separation temperature (Tsep ), which is the
minimum temperature required for the material to be ablated, as
described by [10]
Tsep Tc

 2=3
q0
qc

 
I
L l ln H
Ith

(15)

where d is the optical penetration depth, l is the electronic heat


M
H
and Ith
are the threshold of ablation for
penetration depth, and Ith

(13)

where Tc , qc , and q0 are the critical temperature (7925 K for silicon [22]), critical density (0.76 g/cm3 for silicon [22]), and normal
density (2.33 g/cm3 for silicon [22]), respectively. Therefore, the
separation temperature of silicon is calculated to be 16,724 K.
Figure 3 shows the spatial distribution of the lattice temperature
inside the target at 10 ps during the ablation process. The laser fluence is 2.5 J/cm2, with the wavelength of 800 nm, and the pulse
duration of 100 fs. With this result, the crater geometry and the
ablation depth can be determined by the locations where the maximum local temperature exceeds the separation temperature.
Figures 4(a) and 4(b) present the predicted crater geometry at the
fluence of 2.5 J/cm2 and 20 J/cm2, respectively. The experimental
measurements by the optical surface profilometer are also shown
for comparison. It can be seen that the simulation results predict
the crater shape very well compared with the experimental results,
except for some fluctuations on the crater wall and the ridges at
the edge.
Figure 5 represents the calculation of the ablation depth from
low to high laser fluence, validated by the experimental measurement. The ablation depth shows logarithmic dependence on the
laser pulse fluence, and two distinct regimes have been experimentally observed for both metals [1,4] and semiconductors
[23,24]. The dependence of the ablation rate could be expressed
based on the experimental data by two different logarithmic
functions [4,23]
 
I
(14)
L d ln M
Ith
Journal of Manufacturing Science and Engineering

Fig. 4 Ablated crater geometry at the fluence of (a) 2.5 J/cm2


and (b) 20 J/cm2. Dashed line: experimental measurements;
solid line: simulation results.

Fig. 5 Ablation depth versus laser fluence. (a) From low to


high fluence, (b) zoom in low fluence range. Laser pulse duration: 100 fs, wavelength: 800 nm.

DECEMBER 2013, Vol. 135 / 061015-3

Downloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 12/20/2014 Terms of Use: http://asme.org/terms

laser energy. The absorption of the air breakdown increases fast


as the laser fluence increases. Compared with the air breakdown
absorption, the absorption of the silicon plasma (plasma caused by
surface hot electron emission) is the minor issue, and can be
neglected in the laser fluence range covered in this study. The air
breakdown absorption is the dominant mechanism for laser-early
plasma interaction as illustrated in Fig. 7.

Fig. 6 Ablation efficiency versus laser fluence. Laser pulse duration: 100 fs, wavelength: 800 nm.

low and high laser fluence ranges, respectively. For silicon,


d 145 nm, l 322 nm, IthM 0:458J=cm2 , and IthH 0:637 J=cm2
[23]. The fitting curves of Eqs. (14) and (15) are compared with
the experimental data in Fig. 5, and a good agreement is shown.
In the case of low-fluence ablation (<2 J/cm2 for silicon), the
number density of hot electrons is so low that energy transfer
occurs only within the area characterized by the skin depth. At
higher fluences (>2 J/cm2 for silicon), the contribution of carrier
conduction becomes important and the heat-affected region is
defined by the electron-driven heat penetration depth. The simulation results are in good agreement with the experimental data. The
two ablation regimes are accurately predicted by the simulation.
The ablation efficiency dependence on the laser fluence predicted by the model is shown in Fig. 6, which is defined as the
ablation volume divided by pulse energy. The experimental measurement is provided for validation, and the simulation results
agree with the experimental data very well. It can be seen that the
ablation efficiency first increases fast with the laser fluence. It
reaches the peak value at the laser fluence around 8 J/cm2. After
that, the ablation efficiency drops quickly when the laser fluence
further increases. At the laser fluence of 10 J/cm2, the ablation efficiency drops by 50%, compared with the peak value at 8 J/cm2.
This result indicates that the optimum laser fluence for silicon
ablation in air is about 8 J/cm2. The decrease of the ablation efficiency in the high fluence regime (> 8 J/cm2) is attributed to the
strong interaction of the laser pulse with the laser-induced plasma
[6,23]. When the laser fluence is high enough, laser-induced air
breakdown will occur, and the strong hot electron emission from
the target surface will generate strong early plasma [25]. The early
plasma will significantly absorb the incident laser beam energy,
and reduce the laser energy deposited into the target material. Figure 7 summarizes the energy loss of the incident laser pulse due to
the early plasma absorption. The air breakdown starts occurring at
the laser fluence of 3 J/cm2, and begins absorbing the incident

Fig. 7 Plasma absorption of the incident laser beam energy.


Laser pulse duration: 100 fs, wavelength: 800 nm.

061015-4 / Vol. 135, DECEMBER 2013

Conclusions

In this study, the femtosecond laser ablation of silicon in air


was investigated by a 2D axisymmetric hydrodynamic model. The
model was combined with TTM, hydrodynamic and QEOS models, and covers the complex physics during laser-silicon interaction. It is revealed that two distinct logarithmic ablation regimes
are formed in low and high laser fluence ranges. At low laser fluence (<2 J/cm2), the ablation depth is determined by the optical
penetration depth, while electron heat diffusion becomes the dominant one at high laser fluence (>2 J/cm2). The ablation efficiency
first increases with the laser fluence, and then decreases. The maximum value of ablation efficiency appears at the laser fluence of
8 J/cm2, which is proved to be the most efficient ablation fluence
for silicon in air. The ablation efficiency drops by 50% when the
laser fluence increases to 10 J/cm2. The drop of the ablation efficiency at the laser fluence above 8 J/cm2 is attributed to the laser
interaction with the early plasma. Air breakdown is proved to be
the dominant absorption mechanism for early plasma absorption.
The early plasma starts to absorb the incident laser energy at the
laser fluence of 3 J/cm2, and the absorption rate increases with the
laser fluence, thus causing the eventual decrease of ablation
efficiency.

Acknowledgment
The authors wish to gratefully acknowledge the financial support provided for this study by the National Science Foundation
(Grant No: 0853890-CBET and CMMI-1030786).

References
[1] Momma, C., Nolte, S., Chichkov, B. N., Alvensleben, F. V., and T
unnermann,
A., 1997, Precise Laser Ablation With Ultrashort Pulses, Appl. Surf. Sci.,
109/110, pp. 1519.
[2] Furusawa, K., Takahashi, K., Kumagai, H., Midorikawa, K., and Obara, M.,
1999, Ablation Characteristics of Au, Ag, and Cu Metals Using a Femtosecond
Ti:Sapphire Laser, Appl. Phys. A, 69, pp. S359S366.
[3] Oh, B., Kim, D., Kim, J., and Lee, J., 2007, Femtosecond Laser Ablation of
Metals and Crater Formation by Phase Explosion in the High Fluence Regime,
J. Phys.: Conf. Ser., 59, pp. 567570.
[4] Nolte, S., Momma, C., Jacobs, H., and T
unnermann, A., 1997, Ablation
of Metals by Ultrashort Laser Pulses, J. Opt. Soc. Am. B, 14(10), pp.
27162722.
[5] Ahn, S., Hwang, D. J., Park, H. K., and Grigoropoulos, C. P., 2012,
Femtosecond Laser Drilling of Crystalline and Multicrystalline Silicon for
Advanced Solar Cell Fabrication, Appl. Phys. A, 108, pp. 113120.
[6] Lee, S., Yang, D., and Nikumb, S., 2008, Femtosecond Laser Micromilling of
Si Wafers, Appl. Surf. Sci., 254, pp. 29963005.
[7] Bellouard, Y., Said, A., Dugan, M., and Bado, P., 2004, Fabrication of High
Aspect Ratio, Micro-Fluidic Channels and Tunnels Using Femtosecond Laser
Pulses and Chemical Etching, Opt. Express, 12, pp. 21202129.
[8] Zhao, X., and Shin, Y. C., 2011, Femtosecond Laser Drilling of High-Aspect
Ratio Microchannels in Glass, Appl. Phys. A, 104, pp. 713719.
[9] Zhao, X., and Shin, Y. C., 2012, A Two-Dimensional Comprehensive Hydrodynamic Model for Femtosecond Laser Pulse Interaction With Metals, J. Phys.
D: Appl. Phys., 45(10), p. 105201.
[10] Wu, B., and Shin, Y. C., 2007, A Simple Model for High Fluence Ultra-Short
Pulsed Laser Metal Ablation, Appl. Surf. Sci., 253, pp. 40794084.
[11] Qiu, T. Q., and Tien, C. L., 1993, Heat Transfer Mechanisms During ShortPulse Laser Heating of Metals, ASME J. Heat Transfer, 115, pp. 835841.
[12] Mao, S. S., Mao, X., Greif, R., and Russo, R. E., 1998, Simulation of Infrared
Picosecond Laser-Induced Electron Emission From Semiconductors, Appl.
Surf. Sci., 127, pp. 206211.
[13] Bulgakova, N. M., Stoianm, R., Rosenfeld, A., Hertel, I. V., and Campbell,
E. E. B., 2004, Electronic Transport and Consequences for Material Removal in Ultrafast Pulsed Laser Ablation of Materials, Phys. Rev. B, 69,
p. 054102.
[14] Bulgakova, N. M., Stoian, R., Rosenfeld, A., Hertel, I. V., Marine, W., and
Campbell, E. E. B., 2005, A General Continuum Approach to Describe Fast

Transactions of the ASME

Downloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 12/20/2014 Terms of Use: http://asme.org/terms

[15]

[16]

[17]

[18]

[19]

Electronic Transport in Pulsed Laser Irradiated Materials: The Problem of Coulomb Explosion, Appl. Phys. A, 81, pp. 345356.
Van Driel, H. M., 1987, Kinetics of High-Density Plasmas Generated in Si
by 1.06- and 0.53-lm Picosecond Laser Pulses, Phys. Rev. B, 35, pp.
81668176.
Bechtel, J. H., Smith, W. L., and Bloembergen, N., 1977, Two-Photon Photoemission From Metals Induced by Picosecond Laser Pulses, Phys. Rev. B, 15,
pp. 45574563.
Sebenne, C., Bolmont, D., Guichar, G., and Balkanski. M., 1975, Surface
States From Photoemission Threshold Measurements on a Clean, Cleaved, Si
(111) Surface, Phys. Rev. B, 12, pp. 32803285.
Eymard, R., Gallouet, T. R., and Herbin, R., 2000, The Finite Volume Method, Handbook of Numerical Analysis, P. G. Ciarlet and J. L. Lions, eds.,
North-Holland, Amsterdam, The Netherlands, Vol. 7, pp. 7131020.
Liu, X., and Osher, S., 1998, Convex ENO High Order Multi-Dimensional
Schemes Without Field by Field Decomposition or Staggered Grids, J. Comput. Phys., 142, pp. 304330.

Journal of Manufacturing Science and Engineering

[20] Vidal, F., Johnston, T. W., Laville, S., Barthelemy, O., Chaker, M., Le Drogoff,
B., Margot, J., and Sabsabi, M., 2001, Critical-Point Phase Separation in Laser
Ablation of Conductors, Phys. Rev. Lett., 86, pp. 25732576.
[21] Perez, D., and Lewis, L. J., 2003, Molecular-Dynamics Study of Ablation of
Solids Under Femtosecond Laser Pulses, Phys. Rev. B, 67, p. 184102.
[22] Lorazo, P., Lewis, L. J., and Meunier, M., 2006, Thermodynamic Pathways to
Melting, Ablation, and Solidification in Absorbing Solids Under Pulsed Laser
Irradiation, Phys. Rev. B, 73, p. 134108.
[23] Hwang, D. J., Grigoropoulos, C. P., and Choi, T. Y., 2006, Efficiency of Silicon Micromachining by Femtosecond Laser Pulses in Ambient Air, J. Appl.
Phys., 99, p. 083101.
[24] Besner, S., Degorce, J. Y., Kabashin, A. V., and Meunier M., 2005, Influence
of Ambient Medium on Femtosecond Laser Processing of Silicon, Appl. Surf.
Sci., 247, pp. 163168.
[25] Hu, W., Shin, Y. C., and King, G. B., 2011, Early-State Plasma Dynamics
With Air Ionization During Ultrashort Laser Ablation of Metal, Phys. Plasmas,
18, p. 093302.

DECEMBER 2013, Vol. 135 / 061015-5

Downloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 12/20/2014 Terms of Use: http://asme.org/terms

Das könnte Ihnen auch gefallen