You are on page 1of 45

4.

06 Laser Ablation
D Zhang, Huazhong University of Science and Technology, Wuhan, China
L Guan, Hebei University, Baoding, China
Ó 2014 Elsevier Ltd. All rights reserved.

4.06.1 Laser Ablation Technology: Introduction and Physical Fundamentals 125


4.06.1.1 Laser Ablation 125
4.06.1.2 History of Laser Ablation Technology 126
4.06.1.3 Long-Pulsed Laser Ablation 128
4.06.1.4 Femtosecond Laser Ablation 130
4.06.2 Applications of Laser Ablation in Material Processing I 130
4.06.2.1 Laser Drilling and Cutting 130
4.06.2.2 Laser Welding and Laser Modification of Physical Properties of Functional Materials 133
4.06.2.3 Laser Surface Modification 133
4.06.2.3.1 Surface Hardening 134
4.06.2.3.2 Laser Cladding 134
4.06.2.3.3 Alloying 134
4.06.2.4 Femtosecond Laser-Induced Surface Periodic Structure and Nanogratings 135
4.06.3 Applications of Laser Ablation in Material Processing II 137
4.06.3.1 Microimage of Femtosecond Laser Interaction with Transparent Material 137
4.06.3.2 Applications of Femtosecond Laser Ablation of Transparent Materials 139
4.06.3.3 Applications of Femtosecond Pulsed Laser in Nanoparticle Formation 139
4.06.4 Pulsed Laser Ablation and Pulsed Laser Deposition Technology 143
4.06.4.1 Physical Picture of Pulsed Laser Deposition 143
4.06.4.2 Introduction to Plasma Expansion in Pulsed Laser Deposition 144
4.06.4.3 Introduction to Film Growth in Pulsed Laser Deposition 146
4.06.5 Thermodynamics of Laser Ablation 147
4.06.5.1 Theoretical Framework of the Thermodynamics of Long-Pulsed Laser Ablation – The Basic Equation
and Plasma Shielding Effect 148
4.06.5.2 Theoretical Framework of the Thermodynamics of Long-Pulsed Laser Ablation – Dynamic Physical
Parameters and the Vaporization Effect 150
4.06.5.3 Main Theoretical Results of the Thermodynamics of Long-Pulsed Laser Ablation 152
4.06.5.4 Femtosecond Laser Ablation Models: Classic and Improved Two-Temperature Equations 156
4.06.5.5 Femtosecond Laser Ablation Models: Unified Two-Temperature Equations and Density of State Effect 160
References 165

4.06.1 Laser Ablation Technology: Introduction and Physical Fundamentals


4.06.1.1 Laser Ablation
Laser ablation is the thermal or nonthermal process of removing atoms from a solid by irradiating it with an intense continuous
wave (CW) or pulsed laser beam. As one of the most important techniques for material processing, laser ablation can be used for
drilling extremely small, deep holes through very hard materials such as metals or diamonds, producing thin films or nanoparticles,
preparing material surface in a micro- and nano-controlled fashion, and so on.
The CW laser beam refers to a continuous output once the laser system is powered on, while the pulsed laser refers to a short time
(e.g., milliseconds to femtoseconds) output, as illustrated in Figure 1. The peak output power of a pulsed laser beam is much higher
than that of a CW laser beam, giving the same average output power.
When a solid surface is irradiated by a CW laser beam or long-pulsed (e.g., nanoseconds pulsed) laser beam, the material is
heated by the absorbed laser energy. The thermal motion of some particles is accelerated. Once the absorbed energy exceeds the
sublimation energy, these particles evaporate or sublimate and become vaporized particles; that is, part of the target is ablated. The
laser ablation rate N_ is defined as the vaporized particles per unit area per second:
_ ¼ rd=sm
N [1]

where s is the duration time of the laser pulse, r is the density of the target, d is the thickness of the ablated material, and m is the
average mass of ablated atoms.
However, for a short-pulsed laser (e.g., femtosecond laser), the laser-material interaction time is very short. Thus, the heat energy
has no time to diffuse in lattice. The irradiated zone of the material quickly reaches vaporization temperature and the ablated

Comprehensive Materials Processing, Volume 4 http://dx.doi.org/10.1016/B978-0-08-096532-1.00406-4 125


126 Laser Ablation

Peak output power

Average output power Average

Output power

Output power
& Peak output power output power

Time Time
(a) Continuous wave (CW) (b) Pulsed laser beam
laser beam

Figure 1 Illustration of the peak output power and the average output power of a CW laser beam and pulsed laser beam.

Figure 2 The characteristics of different laser ablation processes. Reproduced from Yang, J. J. Laser Optoelect. Prog. 2004, 41, 44 (in Chinese).

particles evaporate from the surface. The duration time of the pulse is much less than the time taken by excited molecules, atoms,
and electrons to release heat energy by moving or rotating motion (i.e., electron–phonon relaxation time), which is too short for
linear absorption processes such as the single-photon process, electron–phonon interaction process, and thermal diffusion process
to happen. Consequently, nonlinear absorption of laser energy (e.g., multiple-photon process) does occur during short-pulsed laser
ablation. The short-pulse laser process is a nonthermal melting process. The characteristics of different laser ablation processes are
shown in Figure 2 (1).

4.06.1.2 History of Laser Ablation Technology


The thermal effect of laser ablation is utilized in thin-film preparation, laser welding, laser surface treatment and modification, laser
cladding, drilling, cutting, and so on. In 1960, the first working laser, a ruby laser, was realized by Theodore Maiman at Hughes
Research Laboratories (2). In 1965, it was discovered that a laser could be used for diamond drilling using laser ablation mecha-
nisms (3). In 1967, British scientists started research on cutting titanium metal with a laser-assisted oxygen jet, which was later
widely used in aerospace projects. Due to its longer wavelength, CO2 laser was only used for nonmetal cutting (3). In the 1960s and
1970s, laser equipment consisted mainly of CW or long-pulsed lasers such as CO2 and Nd:YAG, which are typical high-power laser
systems. Figure 3 presents the main process of laser evaporation by thermal effect. Once the incident laser energy density exceeds the
threshold, the thermal effect leads to evaporation.
In the middle 1960s, it was found that in the process of laser beam irradiation, electrons, ions, and neutral atoms removed from
the solid surface form a glowing plasma near the surface with the temperature of 103–104 K. Then the ablated material is condensed
on the substrate, and the film is finally obtained. This is the initial principle of laser deposition (4,5). In 1965, Smith and Turner
deposited optical thin films by a ruby laser for the first time (6). This is one of the earliest pulsed laser deposition (PLD) techniques
for thin-film preparation. Limited to low laser energy density at this stage, the high quality of the deposited films was not achieved.
The long wavelength of the laser beam led to a deep melted liquid layer, where sputtering easily occurred. The sputtering produces
many droplets in the deposition process and strongly affects the quality of films.
In the 1980s to 1990s, with the development of lasers with high-energy density and short wavelength, the sputtering effect was
gradually reduced in film deposition. The PLD technique became one of the most successful film fabrication techniques (7,8). PLD
is a physical vapor deposition technique. Figure 4 gives a typical setup of PLD equipment. Under laser irradiation, the target is
Laser Ablation 127

Figure 3 The process of laser ablation by thermal effect.

Figure 4 A typical experimental setup of pulsed laser deposition technique.

ablated and vaporized to form plasma. The plasma plume expands in vacuum or a background gas, and finally arrives at a substrate
surface to deposit a film.
In the 1970s, since lasers with sufficient power and energy density to melt metals were realized, laser welding technology was
developed (9). The Avco Everett Research Laboratory brought the world a laser surface modification machine for heat treatment of
metals in 1973 (10). The laser beams produced in this era had a relatively long wavelength and duration time. The interaction
between the laser and the target is mainly the linear absorption of energy via electron photon resonance, resulting in melting,
vaporization, and removal of material. Therefore, the ablation essentially depends on the optical and thermal properties of the
material. A heat effect zone is unavoidable around the processing zone, which causes heat stress and other defects. In the 1980s and
1990s, various laser techniques were developed for many applications. As an example, laser cladding and PLD technology were fully
developed and made breakthrough progress in this period.
As laser power density is enhanced and pulse duration is shortened, a nonthermal effect was observed in laser ablation for
material processing. In the 1980s, ultraviolet excimer lasers with nanosecond or picosecond pulse width could accomplish many
precise processing tasks, such as drilling, etching, and surface heat treatment, on nonmetallic materials such as polymers and
ceramics (11,12). Each photon of an excimer laser has a large energy, which is enough to break the bonds of molecules or atoms to
generate plasma near the target surface. This is actually a photochemistry reaction rather than a thermal melting process, as shown in
Figure 5. Due to its short wavelength, the laser beam can easily focus on a tiny area, which greatly improves the processing precision.
In 1976, the dye laser for the first time was mode locked to produce pulses with durations as short as 0.3 ps (13). In 1981, Fork
et al. reported a colliding pulse mode-locking technique to produce a continuous and stable train of pulses shorter than 0.1 ps (14).
Asaki et al. utilized a self mode-locked technique to obtain an 11 fs pulse laser in 1993 (15). In 2000, the emergence of photonic
crystal fibers indicated a new stage of femtosecond laser development. The femtosecond laser has two prominent characteristics: (1)
the pulse width can be as short as a few femtoseconds, which is 10 trillion times smaller than the resolution limit of the human eye,
and (2) the transient power is extremely high (16). 200 TW peak power has been achieved experimentally by chirped pulse
amplification (17). The focused peak intensity of a femtosecond laser can run up to 1021 W cm2, or even higher (16,18).
The emergence of femtosecond laser technology further expanded the application field of nonthermal laser ablation. As
a nonthermal and ‘cold’ treatment technique, the femtosecond laser started a new era of laser material processing, showing many
advantages such as superprecision, high spatial resolution, versatility, and universality.
128 Laser Ablation

Figure 5 The process of laser ablation by photochemistry effect.

Femtosecond laser technique is an effective way of fabricating nanoparticles and large-area regular nanogratings (19,20).
Femtosecond laser-induced periodic surface structures (LIPSSs) have potential applications in the photoelectronics, thermal radi-
ation source, and bio-optic devices fields (21). For example, the so-called black metal prepared by the femtosecond laser ablation
technique can absorb light of any wavelength.

4.06.1.3 Long-Pulsed Laser Ablation


When a laser irradiates the target surface, several physical processes including reflection, dispersion, absorption, and transmission
are involved, as shown in Figure 6. P0, Pb, PR, and Pz are the whole incident energy, the absorbed energy, the reflected energy, and
the transmitted energy, respectively. When laser energy absorbed by the target reaches a specific value (ablation threshold), the
ablation phenomenon emerges. Lasers with long wavelength are not suitable for metal ablation, as most of the laser energy will be
reflected by the metal surface. UV short-pulsed lasers with short wavelength, such as frequency multiplied Nd:YAG laser or excimer
laser, are good candidates for metal ablation. The infrared laser with long wavelength can be effective when ablating some special
materials (22).
According to the energy conservation rule, the energy components in Figure 6 can be related as:
P0 ¼ PR þ Pb þ Pz [2]

or
PR Pb Pz
þ þ ¼1 [3]
P0 P0 P0
R ¼ PR/P0, b ¼ Pb/P0, and z ¼ Pz/P0 are reflectivity, absorbance, and transmittance, respectively. Hence, eqn [3] can be rewritten as
R þ b þ z ¼ 1.
Laser ablation is a special interaction between laser radiation and matter, which depends not only on the laser parameters (e.g.,
the output power, the wavelength, and the radius of irradiation spot) but also on the physical properties of the material (e.g., optical
parameters, such as reflectivity, absorbance, and transmittance, and the thermal properties, such as thermal conductivity and specific
heat capacity). With a low power density laser, the target under irradiation will heat up. When the power density exceeds the
ablation threshold energy, at which evaporation starts at the target surface, the target is ablated.
Here is a detailed physical image of long-pulsed laser ablation (LPLA). When pulsed laser irradiates a solid target, most of the
laser energy is accumulated on the surface of the nontransparent target. Part of the laser power is absorbed by this thin layer under
the irradiated surface, which results in a continuous increase of the surface temperature. Simultaneously, part of the energy is
transported into the inner layer. Thus the thickness of the heated region increases. As the thickness increases, the temperature

Figure 6 During laser irradiation of the target surface, several physical processes, including reflection, absorption, and transmission, occur.
Laser Ablation 129

gradient gets smaller and smaller, leading to a gradually decreased heat conduction rate until zero. That is, the laser heat energy can
only penetrate a thin layer of target. The thickness of this thin layer is defined as the thermal penetration depth.
If the laser energy density is high enough, the excited electrons transfer energy to the lattice by collisions and the target heats up.
Thus, the thermal motion of some of the atoms in the lattice is accelerated. These high-energy atoms can overcome the attraction of
their neighbors. Then, corresponding physical phenomena such as melting, vaporization, and boiling occur. Finally, a complicated
layer structure emerges near the target surface, which indicates the beginning of ablation. The vapor above the surface (including
atoms, molecules, some clusters, micrometer-size droplets, and solid particles) continuously absorbs laser irradiation, until it is
almost ionized. Thus, the ablated material in plasma state ejects from target. Subsequently, a plume is produced near the surface.
The temperature of the surface is dominated by the vaporization mechanism.
The particle density of the plasma plume near the target surface, called the corona region, is very high. The corona region absorbs
about 98% of the laser energy through the inverse bremsstrahlung (IB) absorption effect and photoionization (PI) effect, thus
shielding the target surface from laser irradiation. The heat conduction region is the region outside the corona region where the
particle density of the plasma is relatively low. The laser energy is not effectively absorbed in this region. Intense energy transport
phenomena occur near the target surface, that is, the thermal penetration layer, which is mainly solid-phase, together with liquid-
phase and gas-phase materials. This layered structure due to pulsed laser irradiation will extend deep into the target with time.
The actual physical process of the interaction between the laser and the target is much more complicated than the physical
processes mentioned above. There are complex physical processes such as electron excitation effects (such as, induced electron–hole
pairs), photoelectronic effects, atom or cluster emission, and so on (23,24).
When laser power density increases to 1010 W cm2, the energy of the atoms in the melted target surface is dramatically
increased, which usually leads to the boiling phenomenon. Due to the absence of a vaporization nucleus coupled with the
extremely rapid melting process, the temperature of the melting layer rises suddenly. The target does not boil, although the
temperature exceeds the boiling point. This is the so-called superheating phenomenon. The superheated melt is in a metastable
state. Any small disturbance, such as density perturbation or impurity defect formation, can cause an explosive boiling. This
abnormal boiling phenomenon is called a phase explosion, which is an important topic in pulsed laser ablation (PLA) research
(25–30).
The ablation evaporation is intrinsically different from normal evaporation, since a Knudsen layer is generated in the corona
region during laser ablation (31,32). If the particle density of the vapor is low and the collision between the particles is negligible,
only a normal evaporation phenomenon occurs. However, if the laser power density increases to 109 W cm2 (a typical parameter
for ablation), the density of the vapor particles can reach 1016–1027 cm3. In this case, frequent collisions between the ablated
particles result in the highly preferential distribution of the particles along the perpendicular direction of target surface. The
collisions take place within a few mean free paths away from the surface, which is defined as the Knudsen layer. The velocities of
various particles tend to be the same in this layer, as the particles frequently collide with each other due to the high density. In fact,
this is why the PLD technique can achieve the stoichiometric deposition of films. The presence of the Knudsen layer makes it
possible for approximately the same flight time of different particles in the atmosphere. The related physical process is illustrated in
Figure 2(a).
The physical process of LPLA can be briefly described as followed. When the high-power pulsed laser irradiates the target, the
laser energy is absorbed by the target surface. The surface is then melted and vaporized, which forms a high-temperature, high-
pressure plasma. Therefore, the target under interaction with a pulsed laser can be roughly divided into three separate parts: the
high-temperature and high-pressure plasma, the liquid-phase region, and the solid-phase region, as illustrated in Figure 7.
For various materials, irreversible ablation requires that the laser energy density reaches or exceeds the ablation threshold. For
the interaction between the long-pulsed laser and the material, avalanche ionization brings about a final ablation, which is
mainly determined by doped impurities and various defects. Therefore, the ablation threshold for LPLA is different for different
materials.

Figure 7 Illustration of the regions in a laser-irradiated solid. (a) Unaffected region, (b) laser ablating region, (c) high-temperature and high-density
plasma, and (d) the transparent region for laser.
130 Laser Ablation

4.06.1.4 Femtosecond Laser Ablation


The physical image of femtosecond laser ablation is quite different from nanosecond laser ablation. When the pulse duration time is
in the order of a femtosecond, the impact of the relaxation time of electron–phonon interaction, which is in the order of picosecond,
should be considered. When the femtosecond laser focuses on the surface of a bulk, the photon energy is first absorbed by electrons,
leading to an accelerated thermal motion of the electrons and a dramatic temperature increase of the electron subsystem. However,
during the short pulse, the electrons have no time to transfer the obtained energy to the lattice (or the ions). At this time, the
temperature of the electron gas is very high, while the lattice subsystem’s temperature keeps relatively low. Thus, a ‘cold’ ablation
process occurs. Usually, the thermal equilibrium time of the metal is roughly a few femtoseconds, compared to the order of ns for
semiconductors or insulators. This is in fact the relaxation time of electron–phonon interaction. The relaxation time difference is
a result of the energy gap difference. The larger the energy gap, the longer it takes for the excited electron–hole pairs to reach thermal
equilibrium through generation-recombination (34). In the short period of pulse, there are two temperature subsystems in the
target: the electronic subsystem and the lattice subsystem. These two subsystems can be described using two electron–phonon
coupling thermal conduction equations, named as the two-temperature model for femtosecond laser ablation.
In contrast to LPLA, electrons produced in femtosecond laser ablation depend mainly on multiphoton ionization or tunneling
ionization mechanisms (see Section 4.06.3.1). The dopants and defects in the lattice only have a very small influence on generation-
recombination of electron and vacancy pairs. Therefore, the ablation thresholds for different materials are similar for femtosecond
laser ablation (35). Assuming the same energy density and repetition frequency, the pulse laser with shorter pulse width and hence
higher peak power easily satisfies the ablation threshold of target.
Figure 2(b) demonstrates the main physical processes of femtosecond laser ablation. The features are summarized here:
1. Nonthermal processing. This reduces or even eliminates some defects (the roughness of the ablation edge and large crack
formation) in the traditional long-pulsed laser material processing.
2. High-precision processing. Suppression of the heat conduction effect and the hydrodynamics effect dramatically improves the
working precision of femtosecond laser ablation.
3. Accurate three-dimensional spatial processing. When a femtosecond laser with a focused intensity close to the ablation threshold
irradiates transparent materials, the low-beam intensity at the positions other than the focus point cannot meet the requirement
for the nonlinear absorption of multiphotons. Thus, the laser beam arrives at the focal point inside the material without any
attenuation. The high energy accumulated here induces multiphoton absorption and ionization. Consequently, femtosecond
laser ablation can be precisely controlled at any focused position inside a three-dimensional transparent material, leading to
a highly accurate spatial selection and positioning ability.
4. Universality of femtosecond laser material processing. The extremely high peak intensity induces only multiphoton absorption.
The ablation is dominated by laser intensity, suggesting an invariable ablation threshold. The multiphoton absorption and
ionization threshold mainly depend on the atoms rather than on the density of the electrons in the material. Therefore,
a femtosecond pulse laser can theoretically accomplish precision processing, repair, and treatment for any kind of
materials.

4.06.2 Applications of Laser Ablation in Material Processing I


4.06.2.1 Laser Drilling and Cutting
The two remarkable applications of laser ablation are laser drilling and cutting. The use of lasers for drilling and cutting was first
developed in the 1960s. These two techniques have the longest history and the most mature technology for laser material
processing.
Figure 8 represents the main steps for laser drilling. Two physical processes are involved: (1) The temperature of irradiated
material exceeds the melting point, and surface vaporization occurs; (2) the internal pressure of the vapor is so great that the ablated
material is sputtered. Laser drilling requires incident laser energy above the ablation threshold. Thus, high-power pulsed lasers are
extensively used in industry. The pulsed laser has many advantages, such as short interaction time, small heat effect zone, and
precise dimension control.
The factors that affect the quality of laser drilling are the laser parameters (laser power density, irradiation spot area, incident
angle, pulse width, and pulse repetition frequency), the parameters of the desired hole (diameter, shape, and depth), and the
physical properties of the material (absorption coefficient, reflectivity, and thermal conductivity). To accomplish a high-quality
hole, the material with low-heat conductivity (ensuring less dissipation around the drilled hole), low reflectivity, and high
absorbance (ensuring much laser energy absorbed by target) is preferred. Figure 9 shows micrographs of the laser-drilled holes using
a pulsed laser with the same wavelength and different pulse widths. It is obvious that as the pulse width reduces, the quality of the
holes improves, while the power required decreases (36). The process of Figure 9(c) is the best. This is processed by femtosecond
laser drilling, and it keeps a sharp, clean edge and a regular shape. The worst is the one drilled by nanosecond laser, as shown in
Figure 9(a).
Nowadays, many kinds of materials including metals, semiconductors, ceramic, and polymers dielectric materials can be drilled
by laser. As mentioned earlier, metals are suitable for drilling by short-wavelength lasers, such as excimer lasers (36). For most
ceramic materials with a low reflectivity for infrared lasers, CO2 laser is a good candidate for ceramic drilling (37,38). However, for
Laser Ablation 131

(a) (b)

(c) (d)

Figure 8 Schematic diagram of laser drilling processing: (a) unstructured, (b) Cu opened, (c) dielectric opened, and (d) hole formation.

Figure 9 Micrographs of the laser-drilled holes on steel. The wavelength is 780 nm. The pulse widths are (a) 3.3 ns, (b) 80 ps, and (c) 200 fs,
respectively. Reproduced from Chichkov, B. N.; Momma, C.; Nolte, S.; Alvensleven, F. V.; Tunnermann, A. Appl. Phys. A 1996, 63, 109.

polymers, phase transitions (e.g., solid / melt / vapor) and chemical degradation often occur at high temperatures (39). Unlike
most metals, the polymer materials under laser irradiation can produce a deeper thermal penetration. Lawrence et al. confirmed that
for CO2 laser, the absorption length for Al2O3/SiO2-based refractory material is three orders of magnitude greater than most of
metals (38,40). Generally, both polymers and ceramics have a deeper heat penetration, so the heat transport theoretical model
considering a volumetric heating for drilling is more accurate than a model only considering surface heating (41–43).
Laser cutting is a technology that uses a high-power laser to scan material surface, and the material is heated in an extremely short
time, leading to the temperature rises up to 103 or 104 K, then the melting and vaporization phenomena occur. The ablated material
is blown away by an assisted jet gas, and finally a cut edge with a high quality is left. Figure 10 is a schematic diagram of the

Figure 10 A typical experimental setup of laser cutting technique.


132 Laser Ablation

Figure 11 Configuration of the laser-cutting system with the dual-laser-beam method. Reproduced from Jiao, J. K.; Wang, X. B. Opt. Laser Eng.
2009, 47, 860.

laser-cutting technique. The experimental system usually includes a laser device, a laser delivery system, an assisting gas output
device, and a computer-assisted work-piece or laser movement system.
Initially, laser technology was utilized to cut metals, and gradually its applications extended to the cutting of various materials
such as glass and wood, as well as polymers. A typical laser beam used for cutting needs a power intensity of 1000–2000 W. In the
cutting process, the quality of the cutting edge depends on some parameters at different extent, such as the output mode (pulsed or
CW) of laser, power density, pulse repetition frequency, cutting speed, type and pressure of assistant gas, as well as the type and
thickness of the workpieces (44).
Compared to conventional methods, the laser-cutting technique relies on the thermal rather than mechanical properties of
materials, avoiding mechanical damage of the workpiece. Laser cutting is noncontact, which can deliver a finer finish without
cracks. Therefore, computer-assisted laser cutting is widely applied in industry to continuously and precisely accomplish any
desired shape, although the machining process sensitively depends on the physical properties of materials and laser
parameters.
The most commonly used lasers for cutting are Nd:YAG laser and CO2 laser (45). CO2 lasers with high average power and
better beam quality are suitable for cutting thick materials except for metal workpieces with high reflectivity for the longer
wavelength. Nd:YAG lasers have a shorter wavelength, which is able to process highly reflective materials, such as titanium and
titanium alloy (46,47). Lv. Shanjin et al. (47) investigated the influence of the laser-cutting parameters, for example, pulse
energy, pulse rate, cutting speed, and the type and pressure of gas on the heat-affected zone (HAZ) and surface morphology of
titanium alloy.
The CO2 laser is frequently applied to cut glass. There are two main methods for glass cutting: the controlled-fracture laser-
cutting technique and the evaporative laser-cutting technique. The key problems in glass cutting involve how to improve the control
precision, how to reduce the thermal stress generated during glass cooling, and how to avoid the fracture propagation along an
unexpected path. Recently, a dual-laser-beam method was proposed to cut glass substrates in order to further improve the cutting
quality. In this method, a focused laser beam was used to scribe a straight line on the substrate, and a defocused laser beam was used
to irradiate on the scribing line to generate a tensile stress and separate the substrate. With this method, the fracture propagates
stably, and the glass substrate can be separated along the desired path (48,49). The configuration of the laser-cutting system with the
dual-laser-beam method is illustrated in Figure 11.
The important procedures in laser cutting are: (1) controlling the focus position to obtain an accurate focus and small spot
diameter, and (2) adjusting the nozzle parameters to ensure an appropriate direction and flux of the assisting gas. To avoid the
oxidizing reaction and molten dross formed in the process, high-pressure and high-purity nitrogen or inert gases as assisting cutting
gases have been used in laser cutting. Hong Lei et al. (50) used a CO2 laser with a power of 1800–4000 W to cut silicon steel sheets.
In their experiments, a cyclone slag separator was used to form dross-free cutting kerfs (Figure 12).

Figure 12 Photography of the cutting kerfs: (a) by using a cyclone slag separator and (b) by traditional laser-cutting technology. Reproduced from
Hong, L.; Zhang, Y.; Mi, C. L. Opt. Laser Technol. 2009, 41, 328.
Laser Ablation 133

4.06.2.2 Laser Welding and Laser Modification of Physical Properties of Functional Materials
As a high-efficiency and inexpensive technology, lasers are widely used to weld metals or alloys in industry. In 1970, the CO2 laser
was first utilized to weld metal, but thermal damage was easily generated during welding due to the long wavelength of the CO2
laser and the low absorbance of the material (9). In the following years, the Nd:YAG laser with shorter pulse width and higher peak
power was employed to weld metals. The metal surface can effectively absorb the laser energy to achieve a deeper heat penetration,
and thus the strength of the welded joint is remarkably improved (51). It was found that the Nd:YAG laser could produce a minimal
HAZ and less thermal damage. With the introduction of fibers into laser technology, a more flexible delivery system can produce
a higher-efficiency and better-quality welding.
When a laser beam hits a metal surface, the material at the irradiation spot is heated up to the melting and vaporization
temperature, and then vaporization occurs. As a result, several unexpected blowholes may be left on the metal surface. How to
obtain a desired ‘deep-weld effect’ is a key problem in laser welding technology. Because metals have a temperature-dependent
absorption coefficient and a high reflectivity (40,41), laser energy cannot be effectively absorbed by the workpiece, resulting in
unexpected thermal damage on the surface. Recently, CW laser matched with an appropriate pulsed laser can perform dual-beam
laser welding (51) shown in Figure 13. It was found that dual-beam laser welding could markedly reduce or avoid the formation of
the blowholes in the welded joints.
The welding efficiency 3 can be expressed as follows:
3 ¼ ½ydWDHm =P [4]
where P is the incident laser power, y is the speed of welding, d is the thickness of weldment, W is the laser beam width, and DHm is
thermal enthalpy at the melting point. For thermal penetration welding, the welding efficiency 3 is 0.48, while for heat conduction
welding, 3 is 0.37 (52). The parameters, such as pulse width, power density of laser, irradiation spot area, absorption coefficient, and
the thickness of weld, substantially influence the quality of weld. Compared to the conventional arc welding technique, laser
welding can lead to a deeper penetration, a smaller HAZ, and a rapid cooling rate. However, the same problem in these methods is
still that the internal residual stress caused by heat penetration cannot be fully released, affecting the physical properties of the
workpiece.
In the 1990s, it was found that the physical properties of materials can be modified by laser irradiation. Table 1 lists the recent
progress on the modification of the electrical, optical, and magnetic properties of functional materials, including semiconductors,
conductors, superconductors, and magnetic materials.

4.06.2.3 Laser Surface Modification


Conventional surface modification techniques, for example, ion carburizing, boriding, and nitriding, are conditioned by the
incident energy of ions. The common disadvantages of these techniques are a shallow strengthened layer, a long processing period,
a high required temperature, and an easy transformation of the workpiece. Laser surface modification is noncontact processing by
high-power laser beam heating the material surface, and the subsequent cooling takes place through the heat conduction of the
material itself. The characteristics of laser surface modification are as follows: (1) the convenient delivery of laser energy, which
easily allows selective positioning on the surface for strengthening, and (2) a concentrated laser energy acting on the material, giving

Figure 13 Schematic of the experimental setup of the dual-beam laser welding technique. Reproduced from Yan, S.; Hong, Z.; Watanabe, T.; Tang, J. G.
Opt. Laser Eng. 2010, 48, 732.
134 Laser Ablation

Table 1 Applications of the modification of physical properties of materials by laser irradiation

Irradiated films Modification of properties Wavelength (nm) References

Al2O3 Increasing the resistivity of surface 10640 (53)


YBCO Enhancement of electrical conductivity 248 (54)
ITO Enlarged grain size and reduced resistivity 248 (55)
n-GaAs Decrease of photoconductivity and resistivity 694 (56)
SnO2 Improved Hall mobility and refractive index 1064 (57)
Mg-doped GaN Enhanced fluorescence intensity of blue light 10640 (58)
Doped ferrite Improved magneto-optical effect 632.8 (59)
ZnO Reduced UV emission intensity 248 (60)
n-type ZnO Improved electrical conductivity 248 (61)
Ta2O5 ceramic Enhanced dielectric permittivity 10640 (62)
La0.67Ca0.33MnO3 Improved electrical and ferromagnetic properties 10640 (63)
ZnO Improved UV emission intensity 248 (64)

short processing time, a small HAZ, and a small deformation of the workpiece. This technique is suitable for dealing with some
sheet metals but not with thicker sheets. The laser surface modification process includes surface hardening, laser cladding, and
alloying.

4.06.2.3.1 Surface Hardening


Surface hardening, also known as surface heat treatment, includes transformation hardening and shock hardening. Transformation
hardening is a heat-treatment process by high-power laser beam irradiation on the surface of metals and other materials. The
temperature of the heated material reaches the phase transition point, but melting does not occur. After the end of the laser pulse,
the metal begins to quench due to the cooling of the material itself. Compared to conventional heat treatment, the laser modifi-
cation process induces an increase of 15–20% in the hardness of the material. For shock hardening, a high-strength shock wave or
stress wave is generated by the pulsed laser acting on the material surface, and then a strong plastic deformation occurs on the metal
surface. In the laser-impacted zone, the microstructure of the material shows a network of dislocation tangles, which is similar to the
substructure produced by an explosive shock or fast plane shock, significantly improving the surface hardness, yield strength, and
fatigue life of the material. In fact, transformation hardening relies on the thermal properties of materials, while shock hardening
relies on the dynamics properties of materials.

4.06.2.3.2 Laser Cladding


Under laser beam irradiation, a coating material on the substrate surface will melt. Then, when the beam is removed, the molten
coating material will rapidly cool (the cooling rate up to 103–108 K s1) and then be solidified and combined with the matrix
material. In the cladding process, the crystal grain is refined, the composition segregation decreases, and the defect formation is
reduced. A higher compressive stress exists in the solidified structure, which greatly leads to surface wear resistance and antifatigue
properties. The laser cladding process is used mainly for cast iron and alloy materials. For example, Ni-based powder was selected as
laser-surface-modifying material to improve the wear resistance of copper by CW CO2 laser (65).

4.06.2.3.3 Alloying
When a stoichiometric alloy coating on a metal substrate surface is irradiated by a high-energy laser beam, then a mixed melting
occurs within a very short time to form a new surface alloy. Because of a fast heating, the composition, structure, and properties of
the workpiece only have a slight difference in the melted and heat-affected zone, and the deformation of the workpiece also is small.
Finally, the elements of the alloy coating are fully dissolved in the substrate surface layer. Hence, the modified layer has a uniform
composition, low porosity, and dense structure. Experiments indicate that the energy and the moving speed of the laser beam on the
workpiece surface can influence the composition of the alloy (66).
An excimer laser having a high photon energy is often used for surface modification of various materials, such as pure metals or
alloys, glass, ceramics, plastics, and many wide-gap materials (67,68). For many wide-gap materials, an ultraviolet laser is more
suitable for surface modification than other lasers. Reference (66) reported that titanium nitride was modified by the laser beams,
with an energy density of 20.0 J cm2 (TEA CO2) and 2.4 J cm2 (XeCl laser), respectively. The results show that the energy absorbed
from the CO2 laser is mainly converted into thermal energy, causing melting, vaporization, and shock waves in the vapor. Energy
absorbed from the XeCl laser leads primarily to a quick and intense target evaporation. Thus, XeCl laser-induced target effects are
localized and confined to the target surface and its vicinity. With the use of the excimer laser treatment, the corrosion resistance of
the Ti-6Al-4V alloy is significantly improved (69). Experiments show that the reflectivity of metal or metal alloy is affected by laser
wavelength, leading to a different ablation threshold. Figure 14 indicates that the reflectance spectrum of the Ti-6Al-4V alloy surface
and a comparative sum-up of effects produced by the three laser wavelengths 1064, 534, and 266 nm of Nd:YAG laser is given in
Table 2 (70). According to the results, it can be seen that laser surface modification of metals significantly depends on laser
wavelength.
Laser Ablation 135

65

60
1064 nm
55

50

Reflectance (%)
45 532 nm

40

35
266 nm
30

25
1000 800 600 400 200
Wavelength (nm)

Figure 14 Reflectance spectrum of the nonirradiated Ti-6Al-4V surface, with laser wavelengths indicated. Reproduced from Milovanovic, D. S.;
Radak, B. B.; Gakovic, B. M.; Batani, D.; Momcilovic, M. D.; Trtica, M. S. J. Alloy Compd. 2010, 501, 89.

Table 2 A comparative sum-up of effects produced by the three laser wavelengthsa

Reflection Damage threshold Periodic surface structures


.
Laser at 1064 nm 57.00% 0.9 J cm2 Nanometer scale: parallel waves, tto .
E , period: 800 nm after 30 pulses
Laser at 532 nm 42.70% 0.25 J cm2 Nanometer scale: parallel waves, tto .
E , period: 400 nm after 50 pulses
Laser at 266 nm 29.00% 0.12 J cm2 Nanometer scale: parallel waves, tto E , period: 200 nm after 8 pulses
a
Reproduced from Milovanovic, D. S.; Radak, B. B.; Gakovic, B. M.; Batani, D.; Momcilovic, M. D.; Trtica, M. S. J. Alloy Compd. 2010, 501, 89

The integrated surface modification technology by laser combined with other modified methods is growing rapidly. Reference
(71) reported that titanium alloy matrix can be treated by glow discharge plasma nitriding and laser remelting processing at the
same time. Composite gradient coatings containing TiN and Ti2N were prepared, and thus the surface morphology of titanium was
remarkably improved and cracks were reduced.

4.06.2.4 Femtosecond Laser-Induced Surface Periodic Structure and Nanogratings


Femtosecond LIPSSs have become a focus in this field (72–74). The ripples or nanogrooves have a period on the scale of the incident
laser wavelength, and the orientation of these periodic structures is perpendicular to the polarization of the incident laser. The
mechanism of LIPSSs has been extensively studied by many scientists in the past several decades, and some theoretical models have
been established, such as the uneven energy distribution model (75,76), general surface scattering (GSS) model (77,78), Coulomb
explosion model (79), photon resonance absorption model (80), Boson condensation model (81), and energy accumulation effect
model (82).Of these models, the most famous one is the GSS model.
The GSS model to explain the nature of the observed periodic structure was proposed by M. Oron and co-workers (77,78). The
incident laser light is partially scattered by surface defects and produces a tangential wave propagating along the surface. The formation
mechanism of LIPSSs is the interference at the air-material (or vacuum-material, etc.) interface between the incident laser light and the
matter polarization associated with laser-induced surface excitation. Predictions given by the GSS model were in good agreement
with experimental observations up to nanosecond pulse duration, but for the femtosecond pulse laser, a morphology emerged that
was not predicted by the theory (74), implying that other processes were taking place. In 1992, M. Bonch-Bruevich et al. (83) modified
the GSS model and showed that LIPSS is induced by interference between the incident femtosecond laser with the excited plasmons
on the surface. In the modified model, the separation of period structure for linear polarization laser can be expressed as:
l
d¼ [5]
h  sin q
where h ¼ Re[3/(3 þ 1)]1/2 is the real part of the effective refractive index of the surface plasmon, and 3 is the dielectric constant
of the metal.
136 Laser Ablation

In 2002, Jurgen Reif et al. (79) studied femtosecond laser ablation of wide band gap insulators and presented a Coulomb
explosion model. They observed a complex structure of fine ripples at the bottom of the ablated crater, and they assumed that the
ripples structures were due to self-organized structure formation during the relaxation of the highly nonequilibrium surface after
explosive positive ion emission. This model explained that for dielectric material, LIPSS with a period less than laser wavelength
scale was produced by multipulse ablation with pulse energy less than the threshold energy.
Three-dimensional, periodic nanowriting on diamond clusters was reported (84). Periodic ripples have been observed on single-
crystal and polycrystalline diamond surfaces, shown in Figure 15. Further, it has been experimentally shown that the periodicity of
these corrugated two- and three-dimensional structures is shorter than that of the laser wavelength used (248 nm for the excimer
femtosecond laser and 825 nm for the Ti: sapphire femtosecond laser). In 2006, B. Tan et al. (85) found that a femtosecond LIPSS
was formed on the surface-polished crystalline silicon. Unlike the patterns formed by a continuous or nanosecond pulsed laser, the
spacing of the ripple formed by femtosecond pulses was not affected by the incident angle of the laser beam. Also, the pulse
repetition rate had no impact on the ripple formation on crystalline silicon. The orientation of these periodic structures was
perpendicular to the vector of the electric field of the laser beam.
Wang and Guo et al. (86,87) investigated periodic structures on the surfaces of three different noble metals, Cu, Ag, and Au,
following femtosecond laser radiation. Under identical experimental conditions, laser-induced surface patterns show a higher
morphological clearness on the metal with a larger electron–phonon energy coupling coefficient. Angular dependence of the period
of these structures was studied, and the results indicated that these structures were formed by the interference between the incident
light and the excited surface plasmons.
A. Y. Vorobyev et al. (88) performed a detailed study of the formation of LIPSSs on tungsten at near-damage threshold fluences
and found a unique type of LIPSS entirely covered with nanostructures, illustrated in Figure 16.
In Figure 17, periodic nanostructures were observed on the surface of ZnSe after irradiation by a femtosecond Ti: sapphire laser,
which was aligned perpendicularly to the laser polarization direction (89). With the laser polarization parallel to the moving
direction, long-range gratings by slowly moving the crystal were produced. Huang and Xu et al. reported that by the simple scanning
technique with appropriate irradiation conditions, arbitrary size of uniform nanograting could be produced on wide band-gap
materials as well as graphite (90).
In 2009, Eric Mazur et al. induced periodic linear grooves with 40 nm wide, 500 nm deep, and up to 0.3 mm long synthetic
single-crystal diamond with femtosecond pulses at 800 nm (20). L. Sudrie et al. (91) prepared permanent birefringent

5 μm 2 μm

Figure 15 Three-dimensional periodic ripples induced by a femtosecond laser on single-crystal and polycrystalline diamond surfaces. Reproduced
from Ozkan, A. M.; Malshe, A. P.; Railkar, T. A. Appl. Phys. Lett. 1999, 75, 3716.

(a) (b)
542 nm

2 μm 500 nm
λ =800 nm λ =800 nm

Figure 16 Femtosecond laser-induced periodic surface structures on W surface. Reproduced from Vorobyev, A. Y.; Guo, C. J. Appl. Phys. 2008,
104, 063523.
Laser Ablation 137

(a) (b) (c)

NONE SEI 10.0kV X 20.000 1μm WD 8.2mm

Figure 17 SEM images of gratings at the angle 0 (a), 45 (b) and 90 to the moving direction of laser and its polarization direction.

(a) (b)

d = 120 μm d = 430 μm

Figure 18 Photographs of the titanium samples. (a) The sample with a groove period of 120 mm. (b) The sample with a groove period of 430 mm.
Reproduced from Vorobyev, A. Y.; Topkov, A. N.; Gurin, O. V.; Svich, V. A.; Guo., C. L. Appl. Phys. Lett. 2009, 95, 121106.

structures with controllable microscopic dimensions inscribed in pure fused silica platelets. The birefringence properties of
transmission gratings and of a quasi-uniform layer have been established. S. H. Cho et al. fabricated an internal diffraction
grating with photo-induced refractive index modification in planar silica plates by a high-intensity femtosecond (150 fs) Ti:
0
sapphire laser (lp ¼ 790 nm). The low-density plasma formation causes the refractive index modification with a SiE center
defect (92).
Self-organized nanograting periodical structures were induced by an ultrashort intense pulsed laser, which effectively changed
the absorption and reflective properties of metal surfaces. Guo et al. investigated femtosecond laser-induced grating structures with
different periods on titanium surfaces (72–74,93). They found that by using a femtosecond laser structuring technique, near-
perfectly reflective metals are transformed to highly absorptive over an ultrabroad electromagnetic spectrum, ranging from ultra-
violet to terahertz, indicated as Figure 18. The so-called ‘black metal’ phenomenon can be applied to the metal shell surface of
military weapons, allowing weapons to achieve effective stealth.

4.06.3 Applications of Laser Ablation in Material Processing II


4.06.3.1 Microimage of Femtosecond Laser Interaction with Transparent Material
From a microscopic point of view, laser–solid interaction is achieved via the electrons absorbing energy from photons and being
excited from the initial equilibrium state. The absorption mechanisms include single-photon resonance transition, two-photon and
higher-order multiphoton transitions, tunneling ionization, and above-barrier ionization (94). Typically, the single-photon tran-
sition mechanism plays a dominant role due to a larger cross-sectional area.
The interaction between femtosecond laser pulses and transparent materials has a unique nature. The key lies in the transparent
dielectric material usually having a wider band gap, which is much larger than the visible and near-infrared single-photon energy, so
the linear absorption of solid from a low-intensity laser is weak. For example, a wavelength of an 800 nm fs laser, single-photon
138 Laser Ablation

energy is 1.5498 eV, far less than the band gap of some transparent materials, such as glass. It is not possible for electronic transition
from the valence band to the conduction band by single-photon absorption, that is, linear absorption, and consequently color
centers cannot be produced in glass. Femtosecond laser-induced color centers in glass are attributed to nonlinear effects caused by
multiphoton absorption, which requires the intensity of a laser exceeding w1 GW cm2.
When a high-intensity femtosecond laser interacts with transparent dielectrics, the bound electrons in solid absorb incident
multiphoton energy and their kinetic energy exceeds the ionization potential energy to become free electrons by jumping from the
valence band to the conduction band; this process is called multiphoton ionization, namely, multiphoton absorption, as shown in
Figure 19. Multiphoton ionization is a multi-order and nonlinear process, and the reaction cross section is very small. Thus, only an
intense laser with a high-photon density could give rise to such a reaction. Furthermore, if this ionization process shows avalanche
characteristics, this process will be called avalanche ionization.
The mechanisms of femtosecond laser interaction with transparent materials can be described by several theoretical models,
such as, the micro-explosion model (95), the Coulomb explosion model (96–98), the avalanche breakdown model (99), and the
threshold model (100). Early studies suggest that avalanche ionization is the main reason for ablation (101). However, femto-
second laser ablation of dielectric materials relies on multiphoton ionization or avalanche ionization, and there is some dispute
about the mechanism (102,103). Most researchers believe that the multiphoton ionization process first provides seed electrons in
the ablation process and subsequently leads to avalanche ionization to accelerate ablation (104). This is the physical meaning of the
so-called multiphoton collision ionization.
In 2001, Chris B. Schaffer of Harvard University presented that in femtosecond laser irradiation, the nonlinear ionization of
transparent materials includes three processes: multiphoton absorption, tunneling effect, and avalanche ionization. The three
processes can be determined by the Keldysh parameter (eqn [6]). They measured the ablation threshold of transparent materials
with a 110 fs pulse laser, and the damage threshold was found to depend on the laser wavelength and band gap of the material itself:
For the center wavelength of 400 nm, multiphoton ionization plays a dominant role in ablation, while for the wavelength of
800 nm, the tunneling effect does; for wide gap materials, the generation of free electrons is largely due to the role of avalanche
ionization; for narrow gap materials, PI provides a large number of free electrons (105). The Keldysh parameter g is defined as (106):

uð2m Eg Þ1=2
g¼ [6]
eE
where m and e are effective mass and charge of electron, respectively, u and E are angular frequency and amplitude of the incident
electric field, respectively, and Eg is the band gap of solid dielectric. If the width of the band gap is narrow and the incident electric
field is strong, the Keldysh parameter g w1.5 and then tunneling ionization dominates the ablation process; if g w1.5, multi-
photon ionization does; if g w1.5, a mixture mechanism of tunneling ionization and multiphoton ionization. Three ionization
modes of femtosecond laser ablation of transparent material are indicated in Figure 20.

Figure 19 Schematic diagram of multiphoton ionization.

Figure 20 Photon ionization mechanism with different Keldysh parameters.


Laser Ablation 139

4.06.3.2 Applications of Femtosecond Laser Ablation of Transparent Materials


A variety of nonlinear effects produced by ultrashort pulse laser interaction with transparent dielectric material led to many
microstructures induced in materials, suggesting that the ultrashort laser has a broad prospect for development of high-density
three-dimensional data optical storage, optical waveguides, photonic crystals, and so on.
The Mazur group at Harvard University first studied the micro-explosion process and the scale of the microcavity formed in
femtosecond laser interaction with transparent materials (94). Intense laser beam is focused inside the transparent material, and
then ultrahigh-temperature, high-pressure plasma is produced near the focus, causing micro-explosions and microcavity in the body
of the material, and the surrounding material becomes compact because of compression. The diameter of the microcavity formed in
explosions is less than the optical diffraction limit (94). Through multiphoton absorption, this femtosecond laser process induced
a microstructure with an ultra-diffraction limit, mainly applicable in optical storage, as shown in Figure 21.
The femtosecond laser can achieve three-dimensional precise positioning in the transparent material ablation. The principle is as
follows: The femtosecond laser with the intensity near-damage threshold is focused at any internal position of transparent material.
Due to the low-energy beam intensity, it cannot meet the material requirements for multiphoton nonlinear absorption, so the laser
beam has almost no attenuation before it reaches the focal point. Therefore, an accurate adjustment of the focal point is necessary
for more energy absorption to complete multiphoton absorption and ionization, accomplish ultra-precision machining of any parts
of the internal three-dimensional space of the transparent material, and induce the expected microstructure. For example,
a femtosecond laser can engrave a ‘micro bull’ inside transparent materials, with approximately 7  106 pixels with a spatial
resolution of about 150 nm, as shown in Figure 22 (107). Femtosecond laser-engraved ‘micro spider’ and ‘m-dragon’ models are
given in Figure 23 (108,109). Figure 24 illustrates that micro-tweezers with submicron probe tips are fabricated by a two-photon
microstereolithography system developed with a Ti: sapphire femtosecond laser (110).
When the laser beam is focused on amorphous materials, such as quartz glass and a variety of doped glasses, the material will
absorb more laser energy at the focal point, causing changes of lattice structure and the refractive index. This process is helpful for the
successful preparation of three-dimensional optical waveguides (105,111,112).

4.06.3.3 Applications of Femtosecond Pulsed Laser in Nanoparticle Formation


Droplets are often generated in the laser ablation stage of the PLD technique, leading to a reduced surface smoothness of films,
which is not expected in the film growth process. However, in the preparation of granular materials, especially for nanoparticles, the
PLA technique has great development potential. Since the mid-1990s, nanoparticles have been successfully synthesized in a gas
medium using laser ablation in the laboratory (113–119). This preparation technology has rapidly developed since the advent of the
femtosecond laser (120–123). Currently, according to the type of medium, the laser methods for the laser preparation of nano-
particles can be classified into two categories: one is the synthesis of nanoparticles in the atmosphere, caused by condensation of the
vapor during laser ablation of targets, and the other is the preparation of nanoparticles in a liquid medium. Many theoretical
mechanisms have been discussed to make the formation of nanoparticles clear, and most of them show that the nanoparticles are
formed during the plasma expansion and flight in gas or liquid medium. Of course, it is possible to synthesize particles after the
plasma reaches the substrate surface, and even directly produce nanoparticles in the jet of the ablation process (117,124,125). In
2000, F. Mafuné in the Clusters Research Institute of Japan’s Toyota investigated in detail the dynamics of metal nanoparticle
formation and consequently presented a theoretical model (126,127).
The synthesis process of nanoparticles by the laser technique can be simply described as follows. In the atmosphere or liquid
medium, the plasma generated during laser ablation continuously absorbs the laser energy and expands in the medium. Subse-
quently, various ions, atoms, or clusters interact with the atmospheric species in the deposition chamber, leading to the combi-
nation of the larger particles under certain technical conditions (Figure 25). The characteristic parameters of plasma such as
temperature and density are dependent on the thermophysical properties of the target material and the laser parameters, such as
laser fluence and ambient atmosphere (128,129).
The earliest technology for nanoparticle preparation using the laser is accomplished in the gas phase. When a pulsed laser
irradiates the target, a high-temperature and high-density plasma plume is formed in the target surface, and the plume rapidly

Figure 21 Diagram of femtosecond laser direct writing in transparent solid materials within the layered three-dimensional optical data storage
medium.
140 Laser Ablation

Figure 22 Application examples of femtosecond laser ablation of transparent materials. (a) A micro-gearwheel, (b) a micro-chain, (c) and (d) illustrate
two different scanning modes: raster scanning and profile scanning, and (e) a microbull. Reproduced from Tanaka, T.; Sun, H. B.; Kawata, S. Appl.
Phys. Lett. 2002, 80, 312.

expands due to continuous energy absorption from the laser in the gas medium. Ashfold et al. estimated that the density of neutral
species nn ¼ 1018 cm3, the temperature of plasma T ¼ 4500 K, and the density of ions ni ¼ 1013 cm3; if the expansion speed of
plasma is 20 km s1, there will be 1015 atoms in a 0.13 mm3 plume and the generated pressure will be several times atmospheric
pressure (130).

Figure 23 A scanning electron microscopic image of a microspider-array (Reproduced from Chichkov, B. N.; Fadeeva, E.; Koch, J., et al. Proc. SPIE 2006,
6106, 610612.) and m-dragon model (b) (Reproduced from Juodkazis, S.; Nishimura, K.; Misawa, H. Chin. Opt. Lett. 2007, 5, 198.) fabricated by
femtosecond laser machining of transparent materials.
Laser Ablation 141

Figure 24 SEM image of the micro tweezers, with submicron probe tips fabricated by two-photon microstereolithography. Reproduced from
Maruo, S.; Ikuta, K.; Korogi, H. Appl. Phys. Lett. 2003, 82, 133.

Figure 25 Illustration of nanoparticle formation in plasma.

After a laser pulse, the plasma persistently and rapidly expands, and temperature and pressure sharply decline, resulting in
collisions and aggregations between the atoms, electrons, and ions, and then new condensed matter, or a phase transition to form
a new substance, is generated. The experiments show that neutral species play a leading role in the phase transition process: The
condensation between the species helps to form the nuclei, and subsequently the plasma rapidly cools and the condensation occurs
in a different way. If the condensed process is completed on the substrate surface, this is just the principle of PLD of thin films. If the
particles of the plasma freely coalesce, nanoparticles or other shaped material will be formed, accompanied by a large number of
complex physical and chemical processes. The properties of the substrate (such as surface structure and temperature) as well as the
surrounding gas medium will affect the quality of the final synthesis.
In 1998, Geohegan et al. (115) reported the first time-resolved measurements of photoluminescence from gas-suspended
nanoparticles and utilized gated intensified CCD-array imaging of this PL to reveal dramatically different Si-nanoparticle formation
and propagation dynamics in He and Ar. Ar (1.0 Torr) stops and reflects the Si plume, resulting in a stationary, uniformly distributed
nanoparticle cloud. He (10 Torr) slows the silicon plume, angularly segregating most of the nanoparticles to a turbulent smoke ring
that propagates at w10 m s1 through the chamber. This experiment confirmed that the nanoparticles can be condensed in the gas
phase.
The first essential factor is the type and pressure of the atmosphere. The background gas pressure is one of the most important
parameters affecting the size of the nanoparticles (131,132). The experiments (131), in the laser ablation system filled with inert
buffer gas, show that the uniform and dispersed nanoparticles can be successfully prepared. As the buffer gas pressure increases, the
size of the nanoparticles is increased accordingly. Grigoriu et al. discussed the relation between the average size and the ambient
pressure by an inertia fluid model (133). In 2000, Ozawa et al. synthesized nanometer-size particles of tungsten W using a Q-switch
Nd:YAG laser (134). Particle size with the maximum particle generation increased from less than 10 nm to more than 80 nm,
depending on the increase of the ambient pressure. In 3.8 J cm2, the number of nanoparticles generated larger than about 80 nm
does not depend on the pressure. On the other hand, the number of particles less than 80 nm in size depends on the ambient
pressure. The measurement results are shown in Figure 26.
The type of atmosphere has a direct impact on the preparation of nanoparticles. N. Okubo et al. (135) have fabricated nano-
particles of titanium oxide by ablating a Ti target with pulsed CO2 laser in an Ar-diluted oxygen environment. In the Ar gas envi-
ronment, the Ti atoms collide with gaseous atoms and create the ions and electrons, that is, a plasma plume. The formation of the
plasma indicates that Ti atoms have chances to form TiO2 nanoparticles in the O2 as reactive gas. The mean size of NPs fabricated at
142 Laser Ablation

Figure 26 The comparison of size distributions with different fluences, (a) 1.9 J cm2 and (b) 3.8 J cm2. Reproduced from Ozawa, E.;
Kawakami, Y.; Seto, T. Scr. Mater. 2001, 44, 2279.

the lower pressure are small (5 nm) and cohesive to each other, and the major part of the deposit is amorphous, suggesting that the
oxygen supply is not high enough to grow crystalline oxide particles. On the contrary, at the higher pressure, the average size of the
independent NPs significantly increases (28 nm). Under this condition, the oxygen is plentiful, and the generated nanoparticles of
anatase TiO2 crystal and amorphous hybrid structure. It should be emphasized that the NPs are spherical crystals of single anatase, if
sufficient environmental gas is supplied. No rutile phases were observed in any of the conditions studied (Figure 27).
In a liquid medium, pulsed laser synthesis of nanoparticles is a more effective technique. The first attempt to prepare metal
nanoparticles in liquid medium by laser ablation technique was made by German scientists A. Henglein and A. Fojtik. In 1993,
a glass-supported gold and nickel film in the solution was irradiated by the ruby laser with 694 nm wavelength, and gold and nickel
nanoparticles were prepared. The results show that the average particle size was dependent on laser intensity (136). In the same year,
T. M. Cotton and his colleagues developed this technology. In their experiments, a 1064 nm Nd:YAG pulsed laser was used to
synthesize different sizes and distributions of gold, silver, copper, platinum, and other metal nanoparticles, that is, nano-metal
colloids, by controlling the experimental conditions in different solutions (137).
In liquid medium, the better preparation efficiency of NPs is not accidental. In liquid medium, plasma was rapidly formed on
the surface zone of a solid target irradiated by laser. Some researchers (138–142) show that the liquid medium has a greater
influence on the evolution of the plasma than does gas. As the laser ablation proceeds, plasma continues to absorb the laser energy,
and simultaneously the ablated target continues to supply new ablation products for the plume, which prompts the rapid expansion
of plasma at supersonic speed, thus forming a shock wave. The shock wave generated in the plasma produces an additional pressure,
leading to an increase in the temperature of the plasma. Therefore, the liquid medium limiting the shock waves elevates the
temperature, pressure, and density of the plasma.

Figure 27 TEM and SAD images of nanoparticles fabricated by PLD at a total pressure of 1.5  104 Pa (a) and 6.7  104 Pa (b), where flow rates of Ar
and O2 gas for both cases are 83 cm3 s1 and 8.3 cm3 s1, respectively. Reproduced from Okubo, N.; Nakazawa, T.; Katano, Y.; Yoshizawa, I. Appl.
Surf. Sci. 2002, 197–198, 679.
Laser Ablation 143

Figure 28 TEM images of nanoparticles at fluence: (a) 10, (b) 20, and (c) 35 J cm2 and the particle distribution. Reproduced from Kumar, B.;
Yadav, D.; Thareja, R. K. J. Appl. Phys. 2011, 110, 074903.

For example, Berthe et al. investigated the pressure (2–2.5  109 Pa) of laser-induced Al-plasma in a water-confinement regime by
a XeCl excimer laser with the intensity of 1–2  109 W cm2 and the pulse width of 50 ns (139–141). Peyre et al. found that higher-
pressure plasma (1010 Pa) was induced by short pulse laser (3 ns) (140–142). Laser wavelength and pulse width will affect the
pressure of induced plasma. In Sakka’s experiments (143,144), a 532 nm Nd:YAG pulsed laser was used to produce graphite plasma
in water, and the density 1022–1023 cm3, the temperature 4000–5000 K, and the pressure 1010 Pa of C-plasma could be obtained.
The presence of a high-temperature, high-pressure, and high-density plasma zone is advantageous to growing high-temperature
and high-pressure nuclei, and thus the metastable phase is formed at room temperature. A variety of chemical reactions may occur
during the evolution of the plasma. The last stage of the evolution of plasma in the liquid limit is cooling and condensation. Different
types of nucleation have different applications in materials processing. After plasma quenching, part of the plasma is deposited on the
surface of a solid target due to fluid restriction, which will lead to the coating of the target surface, applicable to the surface treatment
of materials (145,146). Another part of the plasma is condensed and dispersed into the liquid medium, and small particles can
generally be suspended or floating on the liquid surface. This can be used for the preparation of nanoparticles themselves.
The average size and size distribution of laser-fabricated nanoparticles in liquid medium are affected by the physical properties of
the target material, laser parameters (laser power density, pulse number, and pulse duration time) and liquid concentration (147).
Figure 28 (148) shows the TEM image of NPs synthesized using PLA in deionized water and their particle size distribution at laser
fluence of 10, 20, and 35 J cm2, respectively. NPs have a spherical shape with a narrow size distribution. The particles are mainly
distributed in the range w5–35 nm with an average of 10, 13, and 15 nm at the fluence of 10, 20, and 35 J cm2, respectively.
Laser ablation synthesis of nanoparticles in liquid medium technology continues to develop, succeeding in the preparation not
only of elemental nanoparticles, but also the alloy nanoparticles. I. Lee et al. (149) fabricated nanometer-sized gold-silver alloy
nanoparticles by ablating gold-silver alloy with an ideal ratio in solution. Nanometer-sized alloy particles have unique physical,
chemical, and catalytic properties, and thus the laser ablation preparation method has a great market potential in the fields of
microelectronic materials, catalysts, and biological engineering applications.

4.06.4 Pulsed Laser Ablation and Pulsed Laser Deposition Technology


4.06.4.1 Physical Picture of Pulsed Laser Deposition
The PLD technique is now widely used in film preparation, which is one of the most important applications in the field of laser
ablation. In the 1970s, with the advent of the excimer laser and the popularization of the electronic Q-switched laser, the PLD
technique began to attract a significant amount of attention (150). The excimer laser has a high-power density and a very short pulse
144 Laser Ablation

Absorption
Thermal Conduction

Surface Melting

Vaporization
Plasma Production

Plasma Shielding
Inverse Bremsstrahlung
Absorption

Figure 29 Illustration of the pulsed laser deposition process.

duration, which could lead to a lower ablation depth and hence the droplets caused by the melting effect are reduced greatly during
the ablation process. Therefore, the excimer laser deposition technique can remarkably improve the quality of the film. In 1987,
D. Dijkkamp et al. at Bell Laboratory first successfully prepared the high-temperature superconducting thin films by using the high-
energy KrF excimer laser (151). Also, it was discovered that PLD can be used for fabrication of epitaxial thin films and multi-
component ceramic oxides, nitrides, metallic multilayers, as well as a variety of superlattice materials (152). Therefore, the PLD
technique has undergone rapid development in the following years (153).
In the late 1990s, femtosecond laser was utilized in the PLD technique, and diamond-like carbon (DLC) films were successfully
prepared (154,155). The power density of the femtosecond laser is extremely high (up to 1021 W cm2) within a very short pulse
duration time (1014–1015 s). The laser ablation process shows several new features, such as nonthermal effects and high-precision
micro-machining to achieve clean ablation during preparation of the targets, which can greatly improve the quality of the films and
promote the development of PLD technology. Research on the mechanisms of femtosecond laser ablation and deposition is
increasing, and some theoretical work has been carried out. For example, the ultrashort pulse laser interaction with the target should
obey the non-Fourier heat conduction rule; the high-energy photon colliding with the target will significantly affect the physical
parameters of the target; the electron–electron collision effect and electronic density of states (DOSs) effect should be considered in
heat transport process; and so on (33,156).
The experimental setup is shown in Figure 4, and the whole preparation process includes three regimes (7) as shown in Figure 29:
(1) Laser ablation and plasma formation
A focused high-energy pulsed laser irradiates the target surface, and then the target absorbs the laser energy, resulting in surface
temperature up to the vaporization point. Subsequently, the vapor is ionized, and local high-density plasma is formed.
(2) Isothermal and adiabatic expansion of plasma
Plasma generated on the target surface will continue to absorb the laser energy and gain further ionization, resulting in a rapid
rise of temperature and internal pressure of the plasma. The elliptical plasma, that is, the plasma plume, normal to the target surface
is formed, and the plasma plume expands and travels in vacuum or atmosphere.
(3) Deposition of thin film
When the expanding plasma plume arrives at the substrate surface, the gaseous species meet each other, combine, and finally
aggregate together.

4.06.4.2 Introduction to Plasma Expansion in Pulsed Laser Deposition


Figure 30 indicates that under the high-energy laser irradiation, a high-temperature and high-pressure plasma is sputtered from the
target surface. Figure 31 shows experimentally the plasma plume observed under conditions of different times and different oxygen
pressures (157).
Laser Ablation 145

For nanosecond laser ablation, the expansion evolution of plasma plume in the space can be roughly divided into two stages: (1)
isothermal plasma expansion. When t < s (pulse duration time), the temperature of the plasma tends to decrease due to expansion,
but at the same time the plasma still absorbs the incident laser energy. The two opposing processes roughly cancel each other out;
thus, the plasma temperature remains constant in this stage (158). (2) Adiabatic plasma expansion. After the termination of the
laser pulse, the rapid expansion of the plasma gives rise to a temperature decrease. At this stage, the heat exchange of the plasma with
its surroundings is negligible, so the expansion is in an adiabatic regime.
Regarding the plasma as an ideal gas and using fluid mechanics, the dynamic equation of plasma expansion is established
according to the overall mass and momentum conservation law (7,159,160). However, modified dynamics models have been
developed based on the local mass and momentum conservation law (161–164), and the theoretical results obtained are in
agreement with the experimental results. The models can naturally describe the characteristics of plasma expansion confirmed by
experiments (165–168): (1) the velocity distribution of the plasma is self-similar, (2) the plasma density at the ablation surface is
approximately constant, and (3) the degree of ionization has an effect on the evolution of the plasma.
The laser ablation process determines that the number density distributions in radial and longitudinal directions have distinct
characteristics. In the direction perpendicular to the target surface, a large-density gradient and different initial velocities of sputtered
atoms lead to the maximal number density of plasma existing in the region adjacent to the target surface rather than on the surface.

Substrate
Thin film
Plasma flow

Focus lense

Target

Figure 30 A sketch of plasma expansion.

Figure 31 Photos of plasma plume at different oxygen pressures. Reproduced from Harilal, S. S.; Bindhu, C. V.; Tillack, M. S.; Najmabadi, F.;
Gaeris, A. C. J. Appl. Phys. 2003, 93, 2380.
146 Laser Ablation

In the radial direction parallel to the target surface, the plasma expansion is mainly caused by the large density gradient, so the
maximal number density locates just on the target surface. Near the surface, the number density in several mean free paths
approximately retains a constant value in the whole ablation process.
Studies have shown that the plasma generated by each pulse may be considered to expand independently and finally reaches
the substrate surface. The state of the plasma is critical for the thin-film growth process of PLD. Based on the above mentioned,
the whole plasma as a single pulse eventually arrives at the substrate. Because the plasma concentration in the spatial distri-
bution is inhomogeneous, the incident species flux changes with time, which provides a theoretical basis for investigation of
thin-film growth in PLD. The evolution of a femtosecond laser-produced plasma is a highly nonequilibrium and nonlinear
process, and the physical conclusions based on nanosecond laser deposition cannot be simply applied to the femtosecond laser
deposition.

4.06.4.3 Introduction to Film Growth in Pulsed Laser Deposition


After isothermal and adiabatic expansions, the plasma plume is collected on a substrate on which thin films grow. The deposition
process has a direct impact on the quality of the thin films. First, the species impinge on the substrate and begin to gather together,
and then the nuclei are continuously formed. In the following deposition process, a number of so-called island structures appear on
the substrate. The existing islands gradually capture diffusing and later-arriving incident species so that the sizes of islands increase
quickly. Finally, the coalescence of existing islands completes a layer of continuous film. The deposition conditions can be
reasonably controlled to achieve a continuous and layer-by-layer growth, until the required film thickness is achieved. The film
growth is a very complex process, and some factors, including the interaction between the ablated species (atoms, molecules, ions,
clusters, etc.) and the substrate surface, the interaction between various species, the substrate temperature, and the incident kinetic
energy of species, play an important role in the growth process.
For the evolution of film growth, the formation of surface defects, adatom migration and diffusion on the substrate, thermo-
dynamics of the adatoms gathered into a nucleus as well as the thin-film growth mode, and so on, are very important research
topics. The island nucleation in the early stage is particularly important for thin-film growth. Two methods are commonly used to
study the film growth process: the continuity equation theory and numerical simulation (169,170). Considering thin-film growth as
a random dynamics process, the Monte-Carlo method is widely used in this field.
Compared with other deposition techniques, the thin-film growth process of PLD has some distinct characteristics: (1) energetic
deposition. The kinetic energy of the species generated by PLA is in the order of 100–103 eV. High-energy species reach the substrate
surface and have a bombardment effect on the deposited film and substrate so that the adatoms obtain an instantaneous mobility
that enhances their diffusion speed on the surface. Also, energetic particle bombardment can cause atomic bond breaking and
structural reconstruction (171). The energetic species impinge from the plume to a surface and transfer their kinetic energies to
surface atoms, leading to the formation of point defects such as vacancies or interstitials. This defect structure is likely to attract the
following incident species to form nuclei. Consequently, the density of islands increases, and the layer-by-layer growth of thin film
is accelerated. (2) Pulsed deposition. The production of PLA reaches the substrate in pulsed mode. After the termination of one
pulsed laser, a large number of species almost simultaneously impinge on the substrate. Thus, a high instantaneous concentration of
species increases the chance for the adatoms to meet each other and aggregate, which promotes layer-by-layer film growth.
Therefore, the interval of two pulses needs to ensure that the incident species have sufficient time to find a suitable growing point.
Namely, pulse repetition frequency for the PLD process is an important parameter (172,173). (3) Stoichiometric deposition. The
most significant advantage of the PLD technique is to keep identical stoichiometry between the deposited film and the target (174–
176). The reason is that a large number of energetic species reach the substrate surface, freely diffuse, and condense to a film on the
substrate. The principle can be found in Section 4.06.2.
The quality of the films prepared by PLD is closely related to laser parameters (laser energy density, wavelength, pulse repetition
frequency, and pulse width), type and pressure of atmosphere in the chamber, the geometric parameters of the experimental setup
(the position of on-axis and off-axis, the distance between target and substrate), as well as substrate parameters (material type,
temperature, and quality of surface).
Here, the influences of laser parameters on film growth are emphasized. Laser energy density can directly affect the type and
concentration of the species in the plasma plume (177,178), which can change the stoichiometric ratio of films. The higher the laser
energy, the greater the bombardment intensity of plasma, the more the number of point defects on the substrate surface (179,180)
and the higher the concentration of the active adatoms. Adjusting the laser energy density sometimes makes it possible to change the
film growth mode (181), for example, such that vertical growth changes to two-dimensional growth (182).
It must be pointed out that laser energy density should be near the ablation threshold in order to avoid the generation of droplets
and improve the quality of the films (183). The experiments show (132,184) that, when the laser energy density exceeds the ablation
threshold, the ablation plume with a large number of droplets or particles reaches the substrate surface, and the surface roughness of
the film remarkably increases. Figure 32 (185) shows an SEM micrograph of an Mg:Nb film deposited at low-energy density of the
laser beam, 8 J cm2 (Figure 32(a)), and at high-energy density, 47 J cm2 (Figure 32(b)). In Figure 32(a), the left and right sides of
the micrographs show a 2500 and 5000 magnification, respectively. In Figure 32(b), the left and right sides of the micrographs
show a 1000 and 4000 magnification, respectively. The most important difference between the low- and high-energy deposited
samples is the size distribution of the spheres. At low-energy densities of the laser beam, these spheres have a diameter lower than
2 mm while at high-energy densities diameters larger than 10 mm can be observed.
Laser Ablation 147

Figure 32 SEM micrograph of the Mg:Nb thin-film samples deposited at different energy densities 8 J cm2 (a) and 47 J cm2 (b) of laser beam.

With an increase in the laser irradiation energy density, the grain size becomes larger (186,187). An explanation of this
phenomenon could be that by increasing the energy density, both the plasma density and the plasma kinetic energy increase. When
the species of the plasma arrive at the substrate surface, the kinetic energy of the deposited adatoms creates higher diffusion, which
results in the coalescence of grains and particles (186).
At fixed pulse intensity, pulse repetition frequency has a great influence on film growth and surface morphology (188,189).
Hydroxyapatite coating on a metal substrate has been prepared by the ArF PLD technique (190). When the pulse repetition rate
increases from 20 to 80 Hz, the structural order of the coatings decreases, and, eventually, at 100 Hz the coatings become
amorphous. CeO2 films at pulse repetition frequency 1, 3, and 5 Hz were prepared, and it was found that with decreasing pulse
frequency, the island density on the surface was significantly reduced and characteristically smoother films were achieved using
a low deposition rate (191). Ag films were deposited at a pulse repetition frequency of 10–100 Hz, and the experimental results
showed that the average size of the island decreased but the island density increased with the pulse frequency increase (192). In the
growth process, a slower deposition rate, that is, a lower pulse frequency at fixed pulse intensity, means that the atoms on the
island can much more easily reach the island edge and the nuclei will be given more time to ripen, avoiding a three-dimensional
island growth (193,194). In a word, adjusting the frequency of pulse repetition has a significant influence on the preparation of
films (195).
Experimental and theoretical simulations show that there are several scaling laws in the thin-film growth process (196–199).
Reference (196) reported that Ag thin films were deposited on an insulator substrate with substrate temperatures of 93  C and
135  C, respectively. Both experimental and theoretical results confirmed that the electrical percolation thickness of the films
depended on pulse repetition frequency, with a scaling law with a single exponent of 0.31 and 0.34, as shown in Figure 33.
Many theoretical models have been developed to describe the thin-film growth process, including the diffusion-limited aggregation
(DLA) model (200), the improved DLA model (201,202), the Bruschi model considering substrate temperature (203), the Taylor
model (204), the Kuzma model (205), and the pulsed kinetic Monte-Carlo model, considering pulse frequency and incident particle
energy (170).

4.06.5 Thermodynamics of Laser Ablation

Laser ablation includes some complex physical phenomena: heat transfer process, plasma effects, various radiation effects,
mechanical effects, and so on. The physical nature is determined by the thermodynamics of the pulse ablation process. Namely, the
study of the laser ablation mechanism is based on thermodynamics in order to study the quantitative transfer law and the related
physical effects in the laser ablation process and to analyze the relationship between various technical parameters and ablation
148 Laser Ablation

135 °C
60

Thickness dperc at Percolation (nm)


50

40 f -0.31

30
(a)

50 93 °C

40

f -0.34
30

20 (b)

2 5 10 20 50
Pulse frequency (Hz)

Figure 33 Effect of pulse repetition frequency on Ag film thickness at the electrical percolation transition. The data are well fit by a power law with
a single exponent of about 0.33 for (a) and (b). Reproduced from Warrender, J. M.; Aziz, M. J. Phys. Rev. B 2007, 76, 045414.

results. Investigation of the laser ablation mechanism can advance the development of laser ablation technology and improve the
quality of ablated workpieces or fabricated materials (such as films and nanoparticles).

4.06.5.1 Theoretical Framework of the Thermodynamics of Long-Pulsed Laser Ablation – The Basic Equation
and Plasma Shielding Effect
The thermodynamics of LPLA is physically based on the local equilibrium heat transport process. In this process, the interaction of
laser and target depends mainly on the electron–phonon interaction. The electron–phonon relaxation time is tens of a picosecond,
that is to say, the time taken by the electrons to transfer their kinetic energy, obtained from the laser, to the lattice. For a long-pulse
laser such as the nanosecond laser, the electrons have sufficient time to deliver their energy to the lattice, and then thermal equi-
librium is easily established between the electrons and the lattice subsystem before the termination of the laser pulse.
After the target absorbs the laser energy, the temperature of the target surface will increase up to a phase transition point. If the
required heat energy for phase change is satisfied, the target will melt. The heating and melting processes of pulsed laser irradiation
on material constitute a three-dimensional heat flow problem. However, even if a long-pulsed laser is utilized in the ablation
process, that is, nanosecond or picosecond, the thermal diffusion distance is short, too. The dimension of the laser irradiation area
(in millimeters) is usually several orders of magnitude larger than the thermal diffusion distance in the vertical direction of the target
surface. Thus, if the target surface is regarded as an infinite plane, the corresponding heat conduction problem can be approximately
treated as the transport of laser energy along the direction perpendicular to the target surface. This is why the laser ablation thermal
conduction phenomenon is theoretically simplified into a one-dimensional heat flow model.
In Figure 34, a schematic of the laser–solid interaction is given. During the time for which a long-pulsed laser irradiates a target
surface, the target can be roughly divided into three separate regions and five parts (33,206): (1) Solid phase region A without

A B C D E

Laser

(1) (2) (3)


x 0

Figure 34 Illustration of long-pulsed laser ablation of target.


Laser Ablation 149

melting, but absorbing the laser energy, (2) liquid-phase region generated by the melted target (including normal liquid phase
region B and superthermal liquid phase region C), and (3) high-temperature and high-pressure plasma cloud D and plume E.
A basic theoretical framework of the LPLA is based on the basic heat conduction equation and the corresponding boundary and
initial conditions.
Only considering the coexistence of a solid and liquid phases in the ablation stage, a basic thermodynamics equation, that is,
a classical Fourier conduction law, can be expressed as:
 
vTi ðx; tÞ v vTi ðx; tÞ
ri ðTÞcpi ðTÞ ¼ ki ðTÞ ð0 < t  s; 0 < x  dÞ [7]
vt vx vx
where i ¼ 1, 2 represents the solid and liquid phases, respectively, r(T) is the temperature-dependent density of target, and cp(T) is
the temperature-dependent specific heat capacity under constant pressure, s is pulse width, d is the target thickness, and x refers to
the vertical direction of the target. The right side of eqn [7] presents the transferred heat flow caused by the existing temperature
gradient in the target, and the left side presents target temperature changes due to the heat conduction. This equation does not take
into account the heat source term, that is, the laser energy radiation during the heat conduction process. During laser irradiation, the
formed plasma plume can absorb 98% of the laser energy by IB and PI mechanisms, and the remainder (only 2%) can transmit to
the target surface. As an approximation, the heat diffusion equation, without considering a heat source term, can be used to describe
the laser–target interaction.
For long-pulse ablation, thermal conductivity ki is assumed to be a constant due to the temperature only having a small effect on
it. The thermal diffusion coefficient ai ¼ ki/(rcp), so eqn [7] can be changed into:

v2 Ti ðx; tÞ 1 vTi ðx; tÞ


¼ ð0 < t  s; 0 < x  dÞ [8]
vx2 ai vt
However, to adequately describe the physical nature of the laser–material interaction, the heat diffusion equation with the heat
source term should be considered. Here are two evolution stages of plasma: (1) one is that at the beginning of ablation, the vapor is
not significantly ionized and plasma cannot be formed, (2) the other is that plasma rapidly expands and consequently its density
decreases, which makes plasma transparent to the incident laser beam. For the two stages, the pulsed laser can directly illuminate the
target surface, and the laser energy absorbed by the corona zone will be transferred to the target surface by plasma radiation. In order
to deeply investigate the temperature evolution of the target, the heat diffusion equation with a heat source term can be expressed as:
 
vTi ðx; tÞ v vTi ðx; tÞ
ri ðTÞci ðTÞ ¼ ki ðTÞ þ Sðx; tÞ ð0 < t  s; 0 < x  dÞ [9]
vt vx vx
where S(x,t) is the heat source term, usually regarded as the Gaussian profile (207–209), that is, S(x,t) ¼ (1  R)I0(x,t)$bebx. Here,
I0(x,t) is the output power density of laser, R is the reflectance, and b is the absorption coefficient of the target.
In addition to considering the source term, the shielding effect of plasma and the evaporation effect should be considered (210).
The shielding effect of plasma is caused by two methods – IB radiation and PI – which absorb laser energy. High-temperature and
high-pressure plasma formed on the target surface absorbs the incident laser radiation and prevents the laser energy from reaching
the surface. This blocking effect is called the plasma shielding effect. For infrared laser ablation, the shielding effect is caused mainly
by IB radiation. For UV laser ablation, the shielding effect of plasma is based on the PI mechanism (211). It is assumed that the
plasma radiation can be regarded as blackbody radiation.
Assuming that plasma is an ideal gas, the ionization process of the vapor is accomplished through IB and PI absorption
mechanisms. The IB process is usually described by the inverse absorption length aIB (cm1) (212):
   
N0
aIB z 1:25  1046 l3 Te0:5 þ Ne Ne ¼ sIB Ne [10]
200
where l is the laser wavelength, Te is the electron temperature, N0 and Ne are the density of the neutral atom and electron,
respectively, and sIB is the cross section of IB absorption. For the excimer laser ablation of metals, such as iron, the plasma particle
density is in the range of 5  1018 cm3, and the electron temperature is 3 eV (211). In eqn [10] sIB represents the total cross section
of the IB process, and the contributions of both neutral and ion in IB processes are included (213,214), that is, sIB ¼ sIB,neu þ sIB,ion.
The cross section of neutral sIB,neu equals approximately 2.75  1023 cm2. A laser fluence of 1–6 J cm2 is equivalent to
0.125  1019–0.75  1019 photons cm2, thus, 0.0034–0.02% photons can be absorbed by IB due to neutrals, which is too small to
significantly attenuate the laser or heat plasma.
In the UV laser ablation process, direct PI of excited atoms plays an important role in the shielding effect of the vapor. The cross
section of PI, sPI can be expressed by (212):

ðEI  E Þ2:5 IH
0:5
sPI z7:9  1018 cm2 [11]
ðhvÞ3
where IH is the ionization potential of hydrogen, EI is the ionization potential of the atom, and E* is the energy of the excited
level that can be photoionized. For example, the photon energy of the 248 nm excimer laser is 5 eV, making
sPI z 1.3  1017 cm2, which are several orders of magnitude higher than the IB cross section. If a laser fluence of 1–6 J cm2 is
150 Laser Ablation

equivalent to 0.125  1019–0.75  1019 photons cm2, the probability of an excited atom absorbing an UV photon equals
approximately 100%; that is, the ionized condition is formed. As a consequence, the vapor ionization degree is readily
enhanced, and the PI is the most important absorption mechanism in this case.
During the pulse duration time, the total energy E absorbed by the target can be expressed as:
Z s
E¼ E0 ðtÞdt [12]
0

where E0 (t)is the laser energy transiently absorbed by the target. For convenience, the time tth is defined as the time taken by the
target to absorb laser energy to reach the ablation threshold Eth. So, when 0 < t < tth, the transient power density absorbed by the
target I(t) is as follows:
IðtÞ ¼ ð1  RL ÞI0 ðtÞ ð0 < t < tth Þ [13a]
Here RL is the target reflectivity for the laser. For tth < t < s, I(t) can be given as:
IðtÞ ¼ ð1  RL ÞI0 ðtÞexp½ðaIB þ sPI nz ÞH þ ð1  RP ÞIem ðtth < t < sÞ [13b]
z
Here, I0(t) is the laser pulse intensity profile, n is the number density of the excited neutrals, and H is the dimension of the plasma
perpendicular to target. RP is the target reflectivity for the plasma emission. Iem is the plasma-emitted intensity. The second term in
the right side of eqn [13a] indicates the absorbed laser power density by the target after plasma irradiation. Equation [13] presents
the heat source term in the heat conduction equation considering the shielding effect of the plasma.
It is assumed that the target ablation threshold Eth is a constant. According to the Lambert–Beer law (215):
EðxÞ ¼ Einc expðbxÞ [14]
where E(x) is the energy density at target depth x and Einc is the total incident energy density. Substituting E(x) ¼ Eth and eqn [12]
into eqn [14], then the maximum ablation depth x can be expressed as:
Z s 
1
x ¼ ln E0 ðtÞdt=Eth [15]
b 0

4.06.5.2 Theoretical Framework of the Thermodynamics of Long-Pulsed Laser Ablation – Dynamic Physical
Parameters and the Vaporization Effect
Normally, the thermophysical parameters of the target such as the absorbance and the absorption coefficient are all temperature
dependent, which affects the heat conduction process (216,217). For simplicity, the ablation of transparent material and the plasma
radiation effect are not considered in the following discussions.
The absorption by the material of energy from the laser can be regarded as a ‘secondary’ energy source on the surface and a variety
of physical effects occur due to this energy source (218,219). The experiments show that for ordinary metals, the resistivity of the
melted metal linearly increases with temperature (220) and is two times larger than that of solid metal below the melting point.
Thus, the absorption coefficient rises, and the reflectivity declines after the melting. Reference (221) shows that the absorbance of
material is proportional to the resistivity. For semiconductors, the number of carriers increases with the temperature, so the lattice
scattering effect is enhanced and then the mobilities of electrons and holes decrease. As a result, the resistivity of such material also
increases with temperature; consequently, the absorbance also does (222).
According to the time-dependent surface temperature in Refs. (216,223), the dynamic absorptivity can be written in the
following explicit form:
b ¼ b0 þ A1 ½Ct þ ðC2 t 2 þ DtÞ1=2  [16]
where b0 is the absorbance at room temperature, and A1 is a constant varying with the material too. The value of A1 for Fe target is
0.85  105 K1 (223). Based on Maxwell’s equations and the Lambert–Beer law, the temperature-dependent absorption coefficient
is given as:
rffiffiffiffiffiffiffiffiffiffiffi sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
4pk 4ps 4ps0
bðTÞ ¼ ¼ ¼ [17]
l0 30 l0 c 30 l0 c½1 þ aðT  T0 Þ

and the temperature-dependent absorptivity can be expressed by:


rffiffiffiffiffiffiffiffiffiffiffiffiffi sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
4py30 4pc30 ð1 þ aðT  T0 ÞÞ
bðTÞ ¼ 2 ¼ [18]
s l0 s0

where s0 is the target conductance at initial temperature T0, 30 is the permittivity of vacuum 30 ¼ 8.85  1012 C2 N1 m2, c is the
laser propagation velocity in vacuum, and a is the target temperature coefficient of resistance. For a general metal, a is in the range
400  105–700  105 K1 (224).
The following discussion is about the effect of vaporization on laser ablation. Vaporization is the species (atoms or
molecules) emitting from the surface of target, due to the phase transition from condensed matter to vaporized
Laser Ablation 151

phase. Vaporization phenomena can occur at any laser influence and pulse duration time, which does not need to satisfy
the temperature threshold, but energy threshold (if E < Eth, the vaporization cannot occur). For the low-power laser ablation,
the vaporization effect on laser ablation is not obvious. However, for the high-power laser, the vaporization
effect is significant due to the sufficient absorption of laser energy, which should be considered in the heat diffusion
equation (225):
 
vTðx; tÞ vT v Tðx; tÞ
rl ðTÞCl ðTÞ  Cl rl yr ¼ kl ðTÞ þ Sðx; tÞ ðsm < t < s; 0 < x < sðtÞÞ [19]
vt vx vx vx
where the subscript l means all parameters are under the condition of liquid state. The heat loss of the vaporized species per area is
vT
Cl rl yr , and the velocity of vaporization (or the velocity of surface recession) is yr, which is determined by the Hertz–Knudsen
vx
equation (29):
p
yr ¼ Cs [20]
rl ð2pkB T=ml Þ1=2

where p is the gas pressure, Cs is the sticking coefficient (226), and ml is the particle average mass. Under the limit Vliq << Vgas, Vlip
and Vgas are the molar volumes of liquid and gas; the relation between the equilibrium vapor pressure p and the temperature T may
be computed from the Clausius–Clapeyron equation (227):
  
DHv ðTb Þm 1 1
p ¼ pb exp  [21]
kB Tb T

where T is the temperature of the target surface, Pb is 1 atm, DHv(Tb) is the latent heat of vaporization, and Tb is the boiling
temperature at 1 atm.
It is assumed that for the time sv taken by the target to fulfill the requirement for the phase transition of vaporization, the
ablation energy threshold is:
Eth ¼ ð1  RÞI0 sv [22]
According to the vaporization relaxation time sv (228), eqn [22] can be changed into:

Eth ¼ 3½1  2expð1Þk1 cl rl ðTv =ð1  RÞÞ2 =I0 [23]


where rl is mass density of liquid material, cl is specific heat capacity per mass, and Tv is the vaporization temperature at
1 atm. Equation [23] indicates that Eth depends on target material and laser parameters. The evolution of ablation interface
(229) is:

ð1  RÞ2 I0 ðt  sv Þ
xt ¼ ðsv < t < sÞ [24]
ðrl cl $DT þ Lv Þ
DT is the biggest temperature difference and Lv is the latent heat of ablation. For sv < t < s, the ablation thickness, which is
proportional to time, increases with the incident laser power density. The ablation surface moves forward at an even speed,
which is determined by the parameters of the laser and the target (159,229,230). The maximum ablation thickness xt can be
expressed as:
ð1  RÞðE  Eth Þ
xt ¼ [25]
r1 c1 $DT þ Lv
When pulse width is reduced on the order of a picosecond, this is roughly equal to the electro–phonon interaction relaxation
time. For the nanosecond laser ablation process, the heat conduction of laser energy is a diffusion process, obeying the Fourier
conduction law. In this case, pulse width is much greater than the time required for electrons to deliver energy to the lattice, which is
effectively instantaneous relative to the pulse duration time. However, for picosecond laser ablation, the heat conduction is not
a diffusion process but a wave-type propagation, which can be described by a hyperbolic heat process model. Different from the
conventional heat transfer law, ultrashort pulse laser ablation has the characteristics of nonequilibrium heating and nonequilibrium
phase transition. For example, when a pulsed laser irradiates the target with an ultrashort pulse width of 150 ps and energy density
of more than 1013 W m2, a very high-temperature gradient will be produced on the target surface, and the heat penetration depth is
less than several nanometers, which is equivalent to the mean free path of electrons. For uniform material, the thermal equilibrium
relaxation time of the target in laser heating (electron–phonon interaction time s0) is approximately between 108 and 1014
(231–233). In other words, the heat conduction process does not obey Fourier law, but rather non-Fourier heat conduction law
(231,234–236). At the end of the 1940s, Cattaneo first presented the non-Fourier heat conduction law, which is called the general
Fourier law (237):
vq
q þ s0 ¼ kVT [26]
vt
152 Laser Ablation

It is assumed that the target is irradiated by an Nd:YAG laser with power density 109 W cm2 and a pulse width of 150 ps (238,239).
The non-Fourier heat conduction considering the heat source term is:
vT v2 T v2 T Sðx; tÞ
þ s0 2 ¼ a 2 þ [27]
vt vt vx rc
When pulse duration time decreases on the order of a few picoseconds, the classical Fourier law will be transformed into non-
Fourier law as eqn [27]. In Section 4.06.5.3, for femtosecond laser ablation, two-temperature systems exist in the target, suggesting
that the above-mentioned heat conduction theory is not applicable.

4.06.5.3 Main Theoretical Results of the Thermodynamics of Long-Pulsed Laser Ablation


To solve the thermodynamic equation, suitable boundary and initial conditions are necessary. Using Si as an example, the ther-
modynamics of LPLA is investigated. The heat conduction eqn [9] can be transformed into:

v2 Ti ðx; tÞ ð1  RÞI0 ðx; tÞbebx 1 vTi ðx; tÞ


þ ¼ ð0 < t  s; 0 < x  dÞ [28]
vx2 ki ai vt
Based on the energy conservation rule and an adiabatic approximation, the boundary conditions before melting can be given as
follows:
vTs ðx; tÞ 
ks ¼ ð1  RÞI0 ð0 < t < sm Þ [29a]
vx x¼0

vTs ðx; tÞ 
ks ¼0 ð0 < t < sm Þ [29b]
vx x¼d

where d is the heat diffusion distance, which is time-dependent d ¼ 3.37(ait)0.5 (228). After melting, there are solid and liquid
phases in target; thus the boundary conditions at the solid–liquid interface can be satisfied by an energy balance equation and
continuous-temperature condition:
vT1  vTs   vs
k1  ks ¼ ð1  RÞIðtÞx¼s  Lm [30]
vx x¼s vx x¼s vt

Tl ðx; tÞ ¼ Ts ðx; tÞ ¼ Tm ðx ¼ sðtÞÞ [31]


The boundary conditions at the back surface can be given by:


Ts  ¼ T0 [32]
x¼xn

During pulse duration time, the temperature at the frontier of the liquid target remains at the vaporization point Tq, so:

T1 x¼0 ¼ Tq ðsm < t  sÞ [33]

When a pulse is finished, we assume the adiabatic hypothesis at the ablation surface, that is:
vT1 
k1 ðs < t < s þ t0 Þ [34]
vx x¼0
where s þ t0 is the pulse period.
Based on the above conditions, the time sm taken by the target to reach the melting point is obtained (33).

3ðTm  T0 Þ2 Cs rs Ks
sm ¼ [35]
4Is2

Based on the above-mentioned heat conduction equation and the boundary and initial conditions, the space- and time-
dependence of target temperature before target melting is represented in Figure 35.
For the melted target, the temperature and the dynamic interface position can be given analytical solutions. The evolutions of
temperature in liquid and solid phases are:
ð1  RÞI0
Tm  Tq  ð1  ebs Þ
bKl pffiffiffiffiffiffi ð1  RÞI0
Tl ðx; tÞ ¼ Tq þ erf ðx=2 al t Þ þ ð1  ebx Þ [36]
erf ðlÞ bKl

ð1  RÞI0 bs
Tm  T0 þ e pffiffiffiffiffiffiffi
bKs ð1  RÞI0 bx
Ts ðx; tÞ ¼ T0 þ  1=2 ! erfcðx=2 as t Þ 
bKs
e [37]
a
erfc l l
as
Laser Ablation 153

Figure 35 Temperature distribution of target at different times and positions. Reproduced from Zhang, D. M.; Li, Z. H.; Zhong, Z. C.; Li, X. G.;
Guan, L. Dynamic Principle of Pulsed Laser Deposition, 66; Science Press: Beijing, 2011 (in Chinese).

where al and as are the thermal diffusivity of the liquid and solid phases, respectively. l depends on the liquid–solid interface
pffiffiffiffiffiffi
l ¼ sðtÞ=2 al t ; here s(t) satisfies the differential equation:
     2  
s2 Is s Is
exp $ Kl ðTm  Tq Þ  ð1  ebs Þ exp $ Ks ðTm  T0 Þ  ebs
4al t b vs 4as t b
  þ Lm ¼ Is $expðbsÞ    [38]
pffiffiffiffiffiffiffiffiffiffi s vt pffiffiffiffiffiffiffiffiffiffi s
pal t $erf pffiffiffiffiffiffi pas t $erfc pffiffiffiffiffiffiffi
2 al t 2 as t

Usually, if the corresponding physical properties of the target and technique parameters are given, one can directly solve the
heat conduction equation by analytical or numerical methods to obtain accurate results of target temperature at different times
and positions, and the evolution of the dynamic interface. This research method for LPLA is quite universal.
According to eqns [9], [13], and [19], the shielding effect of plasma, the vaporization effect, and dynamic absorptivity can be
investigated for the thermodynamics of laser ablation (207,208,217). In Ref. (217), according to the Saha equation and for the
example of Fe, the temperature evolution of the target is obtained from three different models: (1) neglecting the effects of
vaporization and plasma shielding, (2) taking only the vaporization effect into account, and (3) considering the two effects of
vaporization and plasma shielding. The results of three models are represented by curve (1), curve (2), and curve (3), respectively.
The temperature dependence of the absorption coefficient and absorptivity are taken into account in all curves (Figure 36).
In Figure 37, the calculated and measured ablation depths depend on the laser fluence shown, where the experimental
data come from Ref. (211). Only considering the dynamic absorptivity, Lunney et al. simulated the blue dashed curve.
The red dot curve is simulated to describe the thermal phenomena before and after melting (217). The satisfactory agreement
between model predictions and experimental data confirms that the improved thermal model is correct and reasonable.
The Saha equation is usually used to study the plasma-shielding effect in the ablation process. However, for a multielemental
target, the mean ionization of crystals is difficult to define. Reference (208) presents a theoretical model to describe the high-power
nanosecond LPA of multielemental oxide superconductors YBa2Cu3O7 by considering both the vaporization effect and the plas-
ma-shielding effect. In this model, a mean ionization of crystals hU0 i is reasonably defined to analyze the plasma-shielding effect.
hU0 i is the weighted mean of the first ionization energy of each atom (240):
P
Xi U0i
hU0 i ¼ i P [39]
Xi
i
154 Laser Ablation

Figure 36 Comparison of the evolution of temperature at d ¼ 0.00 mm and d ¼ 0.05 mm obtained from three different models: (1) neglecting the effects
of vaporization and plasma shielding, (2) taking only the vaporization effect into account, and (3) considering the two effects of vaporization and
plasma shielding. Reproduced from Fang, R. R.; Zhang, D. M.; Li, Z. H.; Yang, F. X.; Li, L.; Tan, X. Y.; Sun, M. Solid State Commun. 2008, 145, 556.

Figure 37 Fluence dependence of the calculated and measured ablation depths. Reproduced from Fang, R. R.; Zhang, D. M.; Li, Z. H.; Yang, F. X.;
Li, L.; Tan, X. Y.; Sun, M. Solid State Commun. 2008, 145, 556. The experimental data is from Lunney, J. G.; Jordan, R. Appl. Surf. Sci. 1998,
127–129, 941.

where the subscript i refers to the four kinds of atom in YBa 2Cu 3O 7 , U 0i is the first ionization energy of the ith type of
atom, and X i is the number of the ith type atom. From eqn [39], we can obtain the mean ionization hU0 i of YBa2 Cu 3 O 7 as
10.398 eV.
The ratio of IB absorption and PI absorption in the shielding effect can be estimated. If each evaporated
atom in local thermal equilibrium has only one excited electron, the free electron density N e is given by the Saha
equation (241):

Ne2 ¼ 3  1021 ðkB TP Þ3=2 expðhU0 iÞ=kB TP [40]


Laser Ablation 155

Figure 38 Mass removal per pulse as a function of laser fluence. Reproduced from Fang, R. R.; Zhang, D. M.; Li, Z. H.; Li, L.; Tan, X. Y.; Yang, F. X. Phys.
Status Solidi A 2007, 204, 4241. The experimental data is from Bulgakova, N. M. and Bulgakov, A. V. Appl. Phys. A 2001, 73, 199.

From eqn [40], Ne ¼ 1.44  1020, at which the IB by electron–ion collisions becomes the primary mechanism of photon
absorption (sIB,ion w1.25  1019 cm2). A laser fluence of 1–15 J cm2 is equivalent to 0.5  1019–8.1  1019 photons cm2; the
product of the cross section and photon density clearly indicates a probability of 62.5–100% for the incoming photons to be
absorbed by the IB mechanism.
According to the cross section of PI in eqn [11], the photon energy of an infrared laser with wavelength of 1064 nm is 1.16 eV;
thus the cross section sPI z 2.7  1022 cm2. The absorption from an infrared laser by excited atoms is 0.135–2.187%. Through
the IB mechanism, 2.5–40% of the incident energy is absorbed by electron–neutral collisions, while 65–100% by electron–ion
collision. Therefore, IB absorption plays a prominent role in infrared laser ablation.
Figure 38 presents the numerical results of two theoretical models: considering the effects of both vaporization and plasma
shielding and only taking the vaporization effect into account. We can see from Figure 38 that the two curves nearly overlap in the
range of the laser fluence from 1 to 3 J cm2 and begin to separate gradually when the laser fluence is above 3 J cm2. This
phenomenon can be explained by the fact that with the laser fluence increasing gradually, the plasma-shielding effect becomes more
and more marked. The numerical results considering the effects of vaporization and plasma shielding produce a good agreement
with the experimental data. It is clear that the modified model considering the vaporization effect and the plasma-shielding effect
can describe the ablation process more accurately.
Based on the non-Fourier heat conduction eqn [27], the picosecond laser ablation process was investigated using an Al target as
an example. The laser power density of 1013 W m2 and pulse duration time of 150 ps were used in the calculations. Table 3
provides the parameters of the Al target.
With the energy conservation law and adiabatic conditions, the boundary conditions can be depicted as:
vT 
k ¼ bI0 ð0 < t < tm Þ
vx x¼0
vT  [41]
¼ 0 ð0 < t < tm Þ
vx x¼N
lim T ¼ 0
x/N

where tm is the time at which the target begins to melt. The initial conditions are:

T x¼0 ¼ T0
vT  [42]
¼0
vt t¼0

Table 3 Thermal and optical properties of Al targeta

ks (J s1 cm  C) rs (g cm3) cs (J g1  C) s0(s) R

2.388 2.6991 0.86 1011 0.9


a
Reproduced from Rohsenow, W. M.; Hartnett, J. P.; Ganic, E. N. Heat Transfer Hand Book; McGraw-Hill: New York, 1985; p 236.
156 Laser Ablation

700
Non-Fourier conduction model
Fourier conduction model
600

500

Temperature/°C 400

300

200

100

0
0 .0 5 .0 ×1 0–12 1 .0 ×1 0–11 1 .5 ×1 0–11 2 .0 ×1 0–11 2 .5 ×1 0–11 3 .0 ×1 0–11 3 .5 ×1 0–11
Time/s

Figure 39 The surface temperature evolution obtained from the Fourier and non-Fourier models. Reproduced from Zhang, D. M.; Li, L.; Li, Z. H.;
Guan, L.; Tan, X. Y. Phys. B 2005, 364, 285.

where T0 is the initial temperature of the target. Under the boundary and initial conditions, the target temperature depending on
depth and time was obtained (238,239).
Z  
bI0 pffiffiffiffiffiffiffiffiffiffi t 1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Tðx; tÞ ¼ a=s0 pffiffiffiffiffiffiffiffi J0 t 02  s0 x2 =a expðt 0 =2s0 Þdt 0
k x s0 =a 2s0
Z t sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2bI0 ba 0 1 þ 4ab2 s0
þ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi expðbxÞ pffiffiffiffiffiffiffiffi expððt  t Þ=2s0 Þsin h ðt  t 0 Þ [43]
k 1 þ 4ab 2 x s0 =a 4s20
 
1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 J0 t 02  s0 x2 =a expðt 0 =2s0 Þdt 0
2s0
where J0(t) is the zero-rank Bessel function.
Under the same conditions, the dependence of Fourier temperature on depth x and time t can be expressed in the form:
Z t
pffiffiffiffiffiffiffiffiffi 1
Tðx; tÞ ¼ I0 a=p=k pffiffiffi expðx=ð4at 0 ÞÞdt 0
0 t0
Z t  
pffiffiffiffiffiffiffiffiffi 1 x2
 I0 ab a=p=k pffiffiffi expðab2 ðt  t 0 ÞÞexp  dt 0 [44]
0 t0 4at 0
Z
I0 a t
þ expðab2 t 0 Þdt 0
k 0
The calculation results based on eqn [43] and eqn [44] are shown in Figure 39 and Figure 40, respectively. The comparison
shows that the effect of non-Fourier heat conduction cannot be ignored. When t > 5  1012 s, the non-Fourier heat has an obvious
effect on the temperature of the target surface; that is, the non-Fourier temperature is distinctly higher than that of the Fourier heat
conduction model. There is a delay of heat conduction in the inner target, and the deeper the target is, the longer the delay time is.

4.06.5.4 Femtosecond Laser Ablation Models: Classic and Improved Two-Temperature Equations
When the pulse width exceeds 100 ps, the surface temperature of the material is determined by thermal conductivity; thus laser
ablation can be regarded as an equilibrium process, obeying the Fourier conduction rule. If the laser pulse width is reduced, the non-
Fourier heat conduction will be applicable.
However, if the pulse width is reduced on the order of a subpicosecond or femtosecond, the heat transfer process of laser
ablation shows entirely different characteristics. When the laser energy is absorbed by the electrons, their kinetic energy sharply
increases and then the temperature of the electron subsystem suddenly rises. In a short period of pulse, the energy cannot be
transferred from the hot electrons to the target ions (delivery time roughly from subpicosecond to tens of a picosecond). After the
termination of a pulse, the temperature of the ion subsystem in the target slowly rises, while the temperature of the electron
subsystem declines, until the equilibrium temperature is achieved.
Laser Ablation 157

450

400 Fourier model


Non-Fourier model
350

300

Temperature/°C
250

200

150

100

50

0
0.0 2.0× 10–8 4.0× 10–8 6.0× 10–8 8.0× 10–8 1.0× 10–8 1.2× 10–7 1.4× 10–7 1.6× 10–7
Position/m
11
Figure 40 Temperature distributions for two models at t ¼ 2  10 s. Reproduced from Zhang, D. M.; Li, L.; Li, Z. H.; Guan, L.; Tan, X. Y. Phys.
B 2005, 364, 285.

Soviet scientists first paid attention to the special heat conduction phenomenon. In 1957, Kaganov et al. found that an energy
exchange exists between electrons and lattice, and presented a theoretical model to describe the nonequilibrium heat transport
process. Also, they explained the electron–phonon coupling coefficient (243).
In 1974, Anisimov et al. put forward the classic two-temperature model (TTM) (244). Based on the one-dimensional nonsteady-
state heat conduction equation, the two different interaction processes, that is, photon and electron, and electron and lattice were
considered. The temperature evolvements of the electron and lattice subsystems can be given by two differential equations:
v v
Ce Te ¼ ke 2 Te  gðTe  Tl Þ þ Sðx; tÞ
vt vx
[45]
v
Cl Tl ¼ gðTe  Tl Þ
vt
where Te and Tl are the electron and lattice temperature, respectively, Cl and Ce are the specific heat capacity of lattice and electron,
respectively, and g is the electron–phonon coupling coefficient. The first equation describes the electronic temperature,
and g(Te  Tl) represents the transferred energy from electrons to lattice by electron–phonon collision; the heat source term S(x,t)
is the energy absorbed by electrons from the laser by photon–electron collision. The second equation indicates the temperature
evolvement of the lattice subsystem, and the temperature rises due to the electron–phonon interaction g(Te  Tl).
In 1984, J. G. Fujimoto et al. found that when high-intensity 75 fs optical pulses illuminate a tungsten metal surface, an
anomalous heating mechanism appears and a transient nonequilibrium temperature difference occurs between the electrons and
lattice. Pump–probe measurements indicated an electron–phonon energy relaxation time of several hundred femtoseconds (245).
In 1987, Elsayed-Ali et al. (246) used 150–300 fs laser pulses to monitor the thermal modulation of the transmissivity of thin
copper films. The experimental results show that at the beginning of the laser pulse, a larger temperature difference between the
electron and lattice existed, which proved that TTM is reasonable. They also found that the process of electron–phonon energy
transfer (electron–lattice coupling relaxation time) was observed to be 1–4 ps, increasing with the laser fluence.
In 1987, Allen (247) presented an electron–phonon coupling coefficient equation. In 1990, Brorson et al. (248) reported the first
systematic femtosecond pump–probe measurements of the electron–phonon coupling constant in some metallic films, and their
experiments confirmed Allen’s theoretical predictions. In 1996, Chichkov et al. (249) studied the laser ablation of solid targets by
0.2–5000 ps Ti: sapphire laser system. They resolved the TTM equation to obtain analytical solutions.
Both before and after the 1990s, with the development of the femtosecond pulsed laser, the TTM equation has been extensively
investigated. In 1988, Corkum et al. (250) obtained an analytic solution of the TTM and compared the theoretical results with their
experimental measurements of ultrashort laser pulse ablation of molybdenum and copper targets. Optical damage caused by laser
pulses with a duration s  ns can be understood only with a TTM of metals. In 1998, A. P. Kanavin (251) studied different regimes of
heat propagation in metals irradiated by subpicosecond laser pulses on the basis of TTM and obtained analytical solutions cor-
responding to the different temperature dependences of electron thermal conductivity.
In 2002, Gamaly et al. (252) investigated the transient reflectivity of gallium films induced by 150 fs laser pulses by the pump–
probe technique. The experimental results showed that the transient electron–phonon collision rate is a strong function of
158 Laser Ablation

temperature and laser intensity, which is drastically different from that observed under equilibrium conditions. In 2005, Gamaly
et al. (253) carried out experiments on the ablation of metals such as Cu, Al, Fe, and Pb in both air and vacuum and revealed that the
presence of air results in a significant reduction in the ablation threshold. The temperature in the skin-layer during a single-laser
pulse is calculated using a two-temperature approximation while the heat conduction can be neglected. Thus, TTM equations can be
transformed into:
v
Ce Te ¼ gðTe  Ti Þ þ ð1  ab ÞI0 ðtÞ expðab xÞ
vt
[46]
v
Ci Ti ¼ gðTe  Ti Þ
vt
Additionally, the TTM model is combined with thermoelasticity dynamics and the electron blast model to investigate the
mechanisms of ultrashort pulse laser ablation. Hetnarski and Ignaczak (254,255) have summarized five nonclassical approaches
that include the relaxation time effects in dynamic thermoelasticity. Chen et al. (256) extended the dual-hyperbolic two-temperature
and hot-electron blast models to investigate the deformation in metal films subjected to ultrashort laser heating. The major driving
force for nonthermal damage is the so-called hot-electron blast force, which is generated by nonequilibrium hot electrons. Readers
can find related works on femtosecond laser ablation in a review paper in Ref. (257).
Here, an improved TTM model is introduced in order to illustrate the general theoretical method of investigating femtosecond
laser ablation of the target. Moreover, related theoretical work, such as a unified TTM model and electronic DOS effect, are discussed
(208,217,258,259).
When the electron temperature Te is higher than 4000 K, the electron–electron collision plays a very important role in the
femtosecond laser ablation (260–262), causing a change in the thermophysical parameters. In this case, the TTM equation cannot
reasonably describe the thermophysical mechanism. Also, it is necessary to take into account the electron–phonon and electron–
electron collisions affecting the thermophysical parameters, such as electronic thermal conductivity and the optical parameters of
materials.
The thermophysical parameters such as specific heat capacity and thermal conductivity are temperature dependent. When
4000 K < Te < 10 000 K, the electron heat capacity is usually a linear function of the electron temperature (263), given by:
ce ðTe Þ ¼ ce0 Te [47]

where c0e is a constant. The electron thermal conductivity can be described as:
kðTe ; T1 Þ ¼ ce vF2 s=3 [48]
where Tl is the lattice temperature, vF is the Fermi velocity, and s is the electron–phonon relaxation time.
For good conductors, such as metals, the electron–electron collision frequency may be determined by nee ¼ ATe2 ; whereas the
electron–phonon collision frequency is proportional to Tl, namely ne–ph ¼ BTl. Here A and B are constants, and both contribute to
the electron collision frequency n.
For 4000 K < Te < 10 000 K, a relationship between the electron–phonon coupling relaxation time s and the electron–electron
and electron–phonon collision frequency for electron temperature below the Fermi temperature is given by (264):
1
¼ n ¼ nee þ neph ¼ ATe2 þ BT1 [49]
s
Substituting eqn [49] into the expression of the electron thermal conductivity, we can obtain the following equation:

vF2 c0e Te 1 BTe


kðTe ; T1 Þ ¼ ¼ k0 2 [50]
3 ATe2 þ BT1 ATe þ BT1

where k0 ¼ vF2 c0e =ð3BÞ is the electron heat conductivity at room temperature (Te ¼ Tl ¼ 300 K).
Now, consider the variation of the optical parameters with temperature. The experiments confirmed that the optical parameters
of target directly influence the laser energy absorption, which has an important impact on femtosecond laser ablation (21). At lower
electron temperatures, it is allowed to assume that the optical parameter is constant (265–269). However, when the electron
temperature is rising, the optical parameters such as the absorption coefficient and the absorptivity will remarkably change with the
electron temperature.
According to classic Maxwell theory, the optical properties, such as the absorption coefficient and the absorptivity with electron
temperature, are obtained as shown in eqn [17] and eqn [18] (270). Substituting eqns [17], [18], [47], and [48] into the classic TTM
model, the following improved two-temperature model, which considers the effect of electron temperature on the heat capacity,
thermal conductivity, absorption coefficient, and absorptivity of the electron (270), is obtained.

v BTe v2
c0e Te Te ¼ k0 2 Te  gðTe  Tl Þ þ Sðx; tÞ [51]
vt ATe þ BT1 vx2

v
c1 T1 ¼ gðTe  Tl Þ [52]
vt
Laser Ablation 159

Table 4 The thermal and optical properties of copper

Coefficient of the electron heat capacity (Ce0 ) 96.6 J m3 K2


Lattice heat capacity (Cl) 3.5  106 J m3 K
e–ph coupling coefficient (g) 1  1017 W m3 K
Electron heat conductivity at room temperature (k0) 400 W m1 K
Coefficient of e–e collision frequency (A) 1.75  107 K2 S1
Coefficient of e–ph collision frequency (B) 1.98  1011 K1 S1
Target temperature coefficient of resistance (a) 4.3  103 K1

where the heat source S(x,t) is as follows:


!
4pI0 ð4ln 2Þt 2
Sðx; tÞ ¼ bbIð0; tÞ expðbxÞ ¼ exp 
l0 ðsp Þ2
  0:5 
4ps0
 exp  x [53]
e0 l0 c½1 þ aðTe ðx; tÞ  T0 Þ
" #   0:5
ð4ln 2Þt 2 D
¼ C exp   exp  x
ðsp Þ2 1 þ aðTe ðx; tÞ  T0 Þ

where C ¼ 4pI0/l0, D ¼ 4ps0/(30l0c).


The initial condition is:
Te ðx; 0Þ ¼ T1 ðx; 0Þ ¼ T0 [54]
And the boundary conditions can be expressed as:
vTe 
ke ¼ ð1  RÞI0 ðtÞ
vx x¼0

vTe 
ke ¼0 [55]
vx d
where T0 ¼ 300 K is the initial temperature, which is uniform across the target, and d is the depth of the target. The thermal and
optical properties of copper in Table 4 are adopted (246,271).
Under the initial condition eqn [54] and the boundary conditions eqn [55], the improved TTM equation can be resolved and the
calculation results are indicated in Figure 41. Figure 41(a) and 41(b) show the temporal evolution of the electron and lattice
temperature at the surface for the copper target irradiated by a 100 fs, 800 nm pulse at 0.2, 0.3, and 0.4 J cm2, respectively.
In Figure 41(a), the maximum temperature of the electron temperature at the surface is in the range 5500–8000 K for the laser
fluence varying from 0.2 to 0.4 J cm2. The surface electron and lattice temperatures increase with increasing laser fluence. For a fixed
laser fluence, the electron temperature rapidly increases with the ablation time, while suddenly decreasing after it reaches the
maximum, as shown in Figure 41(a). The temperature of the lattice rises gradually; see Figure 41(b).

Figure 41 Temporal evolution of electron temperature (a) and lattice temperature (b) at the surface for the copper target.
160 Laser Ablation

Figure 42 Dependence of the ablation rate on the laser fluence; numerical data 1 represents the improved model.

Figure 42 shows the dependence of the ablation rate on the laser fluence. Curve 1 represents how the ablation rate depends on
laser fluence based on the improved TTM model as eqns [51] and [52]. Curve 2 stands for the model without considering electron
temperature-dependent heat capacity, thermal conductivity of the electron, absorption coefficient, and absorptivity. The black dot
represents the experimental data (272). It can be seen that the results calculated based on the improved TTM model are in good
agreement with the experimental data. In a word, the effects of physical properties on femtosecond laser ablation should be
considered (246,248,272).

4.06.5.5 Femtosecond Laser Ablation Models: Unified Two-Temperature Equations and Density of State Effect
Pulsed laser ablation from nanosecond to femtosecond laser has been extensively investigated (273–275). The first stage of ablation
is the absorption of the laser energy through photon–electron interactions. It takes a few femtoseconds for the electrons to rees-
tablish the Fermi distribution. The second stage is the heat energy diffusion to the lattice through electron–phonon interactions, and
the characteristic time scale of this stage is called the electron–phonon coupling time, typically on the order of tens of picoseconds
(275). Experiments and theoretical research reveal that two ablation mechanisms exist in the laser ablation process: (1) equilibrium
ablation (208,209,225,276–278) and (2) nonequilibrium ablation (122,279–281). We define the electron–phonon coupling time
sR and the laser pulse width sL. When sR/sL >> 1, it is the nonequilibrium ablation; when sR/sL << 1, it is the equilibrium ablation;
and when sR/sL w1, it can be called a mixed ablation.
As for the picosecond laser ablation process, the ablation mechanism is very complex due to the coexistence of equilibrium and
nonequilibrium mechanisms. The general non-Fourier law mentioned above is suitable to describe the heat conduction law of
picosecond laser ablation. However, until now, no perfect heat conduction model exists that describes the thermophysical
phenomenon with the laser pulse width ranging from nanoseconds to femtoseconds. A unified model would be helpful for the
experimental investigations on laser ablation.
sR
Defining a factor exp  a , the following equations give a unified thermal model, which can describe the thermophysical
sL
phenomenon of the laser ablation process with laser pulse width ranges from nanosecond to femtosecond:
 
v sR v2
ce Te ¼ ke exp  a Te  gðTe  Tl Þ þ ð1  RÞbI0 ðtÞexp ðbxÞ [56]
vt sL vx2

v
c1 T1 ¼ gðTe  Tl Þ [57]
vt

Equations [56] and [57] describe the heat conduction process of electrons and lattice subsystem, respectively. In eqn [56], a is an
undetermined coefficient, which varies for different materials and is obtained by fitting theoretical results to experimental data.
When Te is below 3000–4000 K, the influence of the thermal excitation of electrons on the ce and ke is very small (265). Therefore,
ce and ke can be regarded as constants.
Laser Ablation 161

The ratio of the electron–phonon coupling time and laser pulse width is an important criterion. As sR >> sL, that is,
sR
exp  a /0, the ablation is nonequilibrium ablation. The unified model changes into the simplified two-temperature model:
sL
v
Ce Te ¼ gðTe  Tl Þ þ ð1  RÞbI0 ðtÞexpðbxÞ; [58]
vt

v
Cl T ¼ gðTe  Tl Þ [59]
vt l
This is the classic TTM model.
For the long pulses, when sR << sL, the ablation is a thermal equilibrium process, that is, T ¼ Te ¼ Tl, eqns [56] and [57] can be
changed into one equation:
v v2
Cn T ¼ k 2 T þ ð1  RÞbI0 ðtÞexpðbxÞ; [60]
vt vx
where Cn ¼ Ce þ Cl, k, and T are the specific heat, the thermal conductivity, and the temperature of the target, respectively. Equation
[60] becomes the classical equation of one-dimensional heat conduction. The first term of the right side in eqn [60] is the heat
conduction term of the target.
sR
When 0 < exp  a < 1, it is the mixed process, including equilibrium and nonequilibrium ablations. Therefore, the unified
sL
thermal model can well describe the whole thermal conduction phenomenon, with laser pulse width ranging from the nanosecond
to femtosecond time scale.
For eqns [56] and [57], the initial condition is:
Te ðx; 0Þ ¼ Tl ðx; 0Þ ¼ T0 [61]
and the boundary conditions can be expressed as follows:
vTe 
ke ¼ ð1  RÞI0 ðtÞ; [62]
vx x¼0

vTe 
ke ¼ 0; [63]
vx d
where T0 ¼ 300 K is the initial temperature, which is uniform across the target, and d is the thermal diffusion depth. In the low-
fluence ablation case, the number density of the hot electrons is considered low enough that the energy transfer occurs only within
the area characterized by the skin depth d ¼ 1/ab. However, if the laser fluence is very high, the electron thermal diffusion length
rffiffiffiffiffiffiffiffiffiffiffi
ke
should be d ¼ sR (282).
Ce
Under the initial condition eqn [61] and the boundary conditions eqns [62] and [63], the heat conduction eqns [56] and [57] are
numerically solved by a finite difference scheme to investigate damage threshold fluences versus the pulse width. The thermal and
optical properties of gold in Table 5 are adopted.
Assuming that pulse width 101–103 ps, the value of a can be determined as 0.67 through fitting the experimental results with
computed results. The melting point of gold is 1337.33 K (275). The computed modeling results and experimental results are shown
in Figure 43.
In Figure 43, for the pulse duration shorter than 2  102 ps, the threshold fluence slowly increases, and for the longer pulse
duration, the threshold fluence rapidly increases with pulse duration. The results are more reasonable than those obtained by the
simple one-dimensional heat conduction equation (274).
The absorbance b ¼ 1  R of a pure metal consists of two parts, b ¼ b1 þ b2. b1 is the intrinsic absorbance, and b2 is the
contribution due to surface roughness. For an optically smooth metal surface, the order of b2 is about 1.2% of b1, while the role of
b2 is enhanced as the surface roughness increases. In fact, the damage threshold influence is an average value obtained by multipulse
ablation. For low pulse duration (i.e., less than 10 ps), the modified surface is obscured and the value of b2 is very small. For pulse

Table 5 Thermal and optical properties of gold

Wavelength (l) 1053 nm


Thermal conductivity of electron (ke) 318 W m1 K
Specific heat of electron (Ce) 67.7 J m3 K
Specific heat of lattice (Cl) 2.30  106 J m3 K
Absorption coefficient (ab) 7.88  105 m1
Reflectance (R) 0.94
e–ph coupling coefficient (g) 2.1  1016 W m3 K
The electron–phonon coupling time (sR) 6 ps
162 Laser Ablation

Figure 43 Damage threshold fluences versus pulse width for the gold target irradiated by 1053 nm pulse. The experimental data is from Stuart, B. C. and
Feit, M. D. J. Opt. Soc. Am. B 1996, 13, 459.

duration longer than 10 ps, the value of b2 becomes bigger (21). In our simulation, we only take into account the intrinsic
absorbance b1 without considering b2. So for pulse duration longer than 10 ps, the theoretical data are a little bit higher than the
experimental ones.
The rapid development of the femtosecond laser has increased the demand for quantitative and predictive modeling of the fast
and highly nonequilibrium processes occurring in the target material by laser excitation, especially for electron temperatures higher
than 104 K. When the electron temperature is higher than 104 K, the electron DOSs and the energy band structure will change,
leading to the corresponding thermophysical parameters (including electron–phonon coupling and electron heat capacity) of the
target varying with temperature. The specific heat capacity of the electron is small, so its temperature can suddenly increase up to the
Fermi temperature after laser irradiation. When the electron temperature reaches 104 K, the thermal excitation of the electron below
the Fermi surface can significantly influence the thermophysical properties of the target (258–260,263,265). Consequently, using
TTM equations to describe such a process, we must consider the temperature-dependent parameters, such as the electron–phonon
coupling coefficient, the specific heat capacity of the electron, the absorption coefficient, and absorptivity. Also, investigation of the
DOS effect on the physical parameters is essential to quantitatively study the ultrahigh-energy femtosecond laser ablation target
(265–268).
The electronic structure of the target directly determines the effect of thermal excitation of electrons on the thermal physical
parameters because the different electronic structures can induce different DOS effects. For noble metals such as Au, the thermal
excitation from the low-laying d band results in the positive deviation of the heat capacity from the linear dependence at sufficiently
high electron temperatures. However, for transition metals, such as Ni, d band electrons through the Fermi surface lead to a high
density of electrons at the Fermi surface. Therefore, the d band electrons of Ni metal with high-density electronic states can be excited
to the s band, which has a low density of electronic states. This different electronic structure produces smaller thermal physical
parameters than the commonly used parameters at higher temperatures.
When the electron temperature is lower, the optical parameters (absorption coefficient and absorptivity) can be considered as
constant (263,265,267–269). However, experiments have confirmed that both the absorption coefficient and the absorptivity are
temperature dependent. If electron temperature is over 4000 K, both the thermal parameters vary with the electron temperature,
while if the electron temperature exceeds 10 000 K, the temperature can much more significantly affect the parameters (283–285). In
a word, since the DOSs of electrons significantly determine the thermophysical and optical parameters, an improved TTM equation
describing the heat conduction process at the electron temperature 10 000 K is expected.
Several theoretical studies (265–268) show that for electron temperature 10 000 K, the heat capacity of electrons as a function of
temperature can be expressed as:
Z N
vf ð3; m; Te Þ
Ce ðTe Þ ¼ ð3  3F Þ gð3Þde [64]
N vTe " #
 1
ð3  mÞ
where g(3) is the DOS of target, m is the chemical potential at Te, and the Fermi distribution f ð3; m; Te Þ ¼ exp þ1 .
kB Te
Using the calculated electronic structure by Vienna Ab-initio Simulation Package (VASP) software (265–268), the DOS affecting the
Laser Ablation 163

heat capacity of electrons and the electron–phonon coupling coefficient of Ni and Au can be investigated to obtain the relationship
between the parameters and electron temperature. In Figure 44, the great difference between the dashed and solid lines fully shows
that when the electron temperature is high, the DOS effect must not be ignored.
In Ref. (258), a simple method, that is, fitting the solid line calculated by VASP software in Figure 44, was used to obtain fitting
expressions of electron heat capacity with temperature. For the Ni target,
 
54:2
ce ðTe Þ ¼ 30:72  Te 0:0745  105 J m3 K1 [65]
1 þ e 0:4
and for the Au target,
 
38:27
ce ðTe Þ ¼  5:05  Te 0:8  105 J m3 K1 [66]
1þe 0:38

Similarly, the electron–phonon coupling coefficient with temperature also can be obtained by the above-mentioned method
(258). For the Ni target,
 
183:6
Ge ðTe Þ ¼ 1:31 þ Te þ0:39  1017 W m3 K1 [67]
1 þ e 0:14
and for the Au target,
 
2:07
Ge ðTe Þ ¼ 2:08 þ Te 0:8  1017 Wm3 K1 [68]
1þe 0:28

Substituting eqns [65]–[68] into the classic TTM model, an improved TTM model can be given as:
v v
Ce ðTe Þ ¼ ke Te  GðTe ÞðTe  Tl Þ þ Sðx; tÞ [69]
vx vx

v
Cl T ¼ GðTe ÞðTe  Tl Þ [70]
vt l
The initial condition is:

Te ðx; 0Þ ¼ Tl ðx; 0Þ ¼ T0 [71]


and the boundary conditions can be expressed as follows:
vTe 
ke ¼ ð1  RÞI0 ðtÞ [72]
vx x¼0

vTe 
ke ¼0 [73]
vx d

Figure 44 Electron temperature dependences of the electron heat capacity of Ni(a) and Au(b). Solid lines show the results of the calculations performed
with DOS obtained from VASP. Dashed lines show the commonly used approximations of the thermophysical material properties.
164 Laser Ablation

Table 6 Physical parameters and optical parameters of Ni and Au

Metal ke (W m1 K1) Cl (J m3 K1) a (K1)

Ni 91 4.1 4.3  103


Au 318 2.5 6.5  103

Using the improved TTM model with the initial and the boundary conditions, the surface temperature evolutions of Ni and Au
targets during laser ablation are investigated. Table 6 gives the physical and optical parameters of Ni and Au. And the laser
parameters of pulse width 100 fs, wavelength 800 nm, and laser energy density 0.65 J cm2 are used in the calculations. Figure 45
indicates the time-dependent temperatures of the electron and the lattice of the Ni-target surface.
From Figure 45, it can be seen that the electron temperature suddenly increases to 10 000 K and then slowly declines. The lattice
temperature gradually increases to a final equilibrium temperature of about 2500 K. The results suggest the electron–phonon
coupling coefficient of laser ablation of Ni sep z 2 ps.
Figure 46(a) and 46(b) indicate that the obtained threshold fluence of Ni (a) and Au (b) in TTM simulations performed with
thermophysical parameters calculated with DOS is obtained from VASP (dot lines), and the commonly used approximations
(dashed dot lines) and the experimental data from (286) for the melting thresholds. The theoretical results from the new improved
TTM equation and the experimental data are more consistent than the results from the traditional two-temperature equation. For

12000
11000
10000
electron temperature at 0.65J cm–2
9000 lattice temperature at 0.65J cm–2
8000
Temperature(K)

7000
6000
5000
4000
3000
2000
1000

0 1000 2000 3000


Time (fs)

Figure 45 The time-dependent temperatures of the electron and the lattice of the Ni-target surface.

Figure 46 Threshold fluence of Ni(a) and Au(b) for surface melting as a function of film thickness. The dot lines are the results considering the
DOS effect.
Laser Ablation 165

Au, the theoretical values of the new improved TTM equation at film thickness larger than about 350 nm tend to a constant; the
theoretical and experimental data of Ni and Au show that the melting threshold increases rapidly with the film thickness and finally
tends to a constant value. The saturation threshold fluence of Ni obtained in the TTM simulations considering the DOS effect,
230 J m2, is a little higher than the experimental data 220 J m2 (286). The calculated threshold fluence with the commonly used
parameters, 275 J m2, is not in good agreement with the experimental data. The saturation threshold fluence at large film thick-
nesses obtained in the TTM simulations performed with the commonly used parameters, 1100 J m2, is significantly lower than the
experimental value, 1800 J m2 (286), whereas the value predicted in TTM simulations performed with the new thermophysical
parameters, 1600 J m2, is in a better agreement with the experimental data. The comparisons confirm that the DOS effect on
femtosecond laser ablation is significant.
Ultrashort-pulsed laser ablation technology is developing in leaps and bounds, and the femtosecond laser ablation mechanism
is becoming more complex. It is a pity that we cannot give a deeper and more comprehensive introduction to this research field in
the present chapter. The interested reader can refer to the related references, such as a recent review paper in Ref. (257).

References

1. Yang, J. J. Laser Optoelect. Prog. 2004, 41, 44 (in Chinese).


2. Maiman, T. H. Nature 1960, 187, 493.
3. Bromberg, J. The Laser in America, 1950–1970; MIT Press: Cambridge, MA, 1991; p 202.
4. Hohig, R. H.; Woolston, J. R. Appl. Phys. Lett. 1963, 2, 138.
5. Ready, J. F. Appl. Phys. Lett. 1963, 3, 11.
6. Smith, H. M.; Turner, A. F. Appl. Opt. 1965, 4, 147.
7. Singh, R. K.; Narayan, J. Phys. Rev. B 1990, 41, 8843.
8. Norton, D. P. Pulsed Laser Deposition of Complex Material: Progress Towards Applications. In Pulsed Laser Deposition of Thin Films; Eason, R., Ed.; Wiley-Interscience: New
York, 2007; p 3.
9. Maddox, S. J. British Welding Inst. Report, 1970, p 36.
10. Locke, E. V. U.S. Patent 3848104.
11. Liu, G. Y.; Toncich, D. J.; Harvey, E. C. Opt. Laser Eng. 2004, 42, 639.
12. Lee, S. K.; Na, S. J. Appl. Phys. A 1999, 68, 417.
13. Ruddock, I. S.; Bradley, D. J. Appl. Phys. Lett. 1976, 29, 296.
14. Fork, R. L.; Greene, B. I.; Shank, C. V. Appl. Phys. Lett. 1981, 38, 671.
15. Asaki, M. T.; Huang, C. P.; Garvey, D.; Zhou, J.; Kapteyn, H. C.; Murnane, M. M. Opt. Lett. 1993, 18, 977.
16. Keller, U. Nature 2003, 424, 831.
17. Lozhkarev, V. V.; Freidman, G. I.; Ginzburg, V. N.; Katin, E. V.; Khazanov, E. A.; Kirsanov, A. V.; Luchinin, G. A.; Mal’shakov, A. N.; Martyanov, M. A.; Palashov, O. V.;
Poteomkin, A. K.; Sergeev, A. M.; Shaykin, A. A.; Yakovlev, I. V.; Garanin, S. G.; Sukharevi, S. A.; Rukavishnikov, N. N.; Charukhchev, A. V.; Gerke, R. R.; Yashin, V. E. Opt.
Express 2006, 14, 446.
18. Tajima, T.; Mourou, G. Phys. Rev. Spec. Top. Accel. Beams 2002, 5, 03130121.
19. Huang, M.; Zhao, F. L.; Cheng, Y.; Xu, N. S.; Xu, Z. Z. Opt. Express 2008, 16, 19354.
20. Shinoda, M.; Gattass, R. R.; Mazur, E. J. Appl. Phys. 2009, 105, 053102.
21. Vorobyev, A. Y.; Guo, C. L. Phys. Rev. B 2005, 72, 195422.
22. Bubb, D. M.; Papantonakis, M. R.; Horwitz, J. S.; Haglund, R. F., Jr.; Toftmann, B.; McGill, R. A.; Chrisey, D. B. Chem. Phys. Lett. 2002, 352, 135.
23. Bloembergen, N.; Kurz, H.; Liu, J. M.; Yen, R. In Laser and Electron-Beam Interactions with Solids; Appleton, B. R., Celler, G. K., Eds.; North-Holland: New York,
1982; pp 3–60.
24. Garrison, B. J.; Itina, T. E.; Zhigilei, L. V. Phys. Rev. E 2003, 68, 041501.
25. Perez, D.; Lewis, L. J. Phys. Rev. Lett. 2002, 89, 255504.
26. Yoo, J. H.; Jeong, S. H.; Mao, X. L.; Greif, R.; Russo, R. E. Appl. Phys. Lett. 2000, 76, 783.
27. Pakhomov, A. V.; Thompson, M. S.; Gregory, D. A. J. Phys. D Appl. Phys. 2003, 36, 2067.
28. Kudryashov, S. I.; Allen, S. D. J. Appl. Phys. 2003, 93, 4306.
29. Bulgakova, N. M.; Bulgakov, A. V. Appl. Phys. A 2001, 73, 199.
30. Yoo, J. H.; Borisov, O. V. Anal. Chem. 2001, 73, 2288.
31. Peterlongo, A.; Miotello, A.; Kelly, R. Phys. Rev. E 1994, 50, 4716.
32. Kelly, R.; Miotello, A. Phys. Rev. E 1999, 60, 2616.
33. Zhang, D. M.; Li, Z. H.; Zhong, Z. C.; Li, X. G.; Guan, L. Dynamic Principle of Pulsed Laser Deposition, 66; Science Press: Beijing, 2011 (in Chinese).
34. Allmen, M.v.; Blatter, A. Laser-Beam Interactions with Materials; Springer: Berlin, 1998, 187.
35. Liu, X. D.; Mourou, G. IEEE J. Quantum Electron. 1997, 33, 1706.
36. Chichkov, B. N.; Momma, C.; Nolte, S.; Alvensleven, F. V.; Tunnermann, A. Appl. Phys. A 1996, 63, 109.
37. Hector, L. G.; Hetnarski, R. B. In Thermal Stresses, Hetnarski, R. B., Ed.; Elsevier: Amsterdam, 1996; p 453.
38. Lawrence, J.; Minami, K.; Li, L.; Edwards, R. E.; Gale, A. W. Appl. Surf. Sci. 2002, 186, 162.
39. Godovsky, Y. K. Thermophysical Properties of Polymers, 28; Spring-Verlag: New York, 1992.
40. Lawrence, J.; Li, L. Appl. Surf. Sci. 2000, 168, 71.
41. Li, J. F.; Li, L.; Stott, F. H. Int. J. Heat Mass Transfer 2004, 47, 1159.
42. Zhang, C.; Salama, I. A.; Quick, N. R. J. Appl. Phys. 2006, 99, 113530.
43. Zhang, C.; Quick, N. R.; Kar, A. J. Appl. Phys. 2008, 103, 014909.
44. Choudhury, I. A.; Shirley, S. Opt. Laser Technol. 2010, 42, 503.
45. Sharma, A.; Yadava, V. Opt. Laser Technol. 2012, 44, 159.
46. Yilbas, B. S.; Zhuja, S. Z.; Hashmi, M. S. J. J. Mater. Process. Technol. 2003, 136, 12.
47. Shanjin, Lv; Yang, W. Opt. Laser Eng. 2006, 44, 1067.
48. Triantafyllidis, D.; Bernstein, J. R.; Li, L.; Stott, F. H. J. Laser Appl. 2003, 15, 49.
49. Jiao, J. K.; Wang, X. B. Opt. Laser Eng. 2009, 47, 860.
50. Hong, L.; Zhang, Y.; Mi, C. L. Opt. Laser Technol. 2009, 41, 328.
166 Laser Ablation

51. Yan, S.; Hong, Z.; Watanabe, T.; Tang, J. G. Opt. Laser Eng. 2010, 48, 732.
52. Duley, W. W., Ed. Laser Welding; John Wiley & Sons: NewYork, 1999; p 1.
53. Stolz, B.; Backes, G.; Gillne, A.; Kreutz, E. W. Appl. Surf. Sci. 1997, 109–110, 242.
54. Konopka, J.; Lewandowski, S. J.; Okunev, V. D.; Samoilenko, Z. A.; Abal’oshevet, A. Phys. C 2000, 341–348, 2147.
55. Adurodija, F. O.; Izumi, H. Appl. Surf. Sci. 2001, 177, 114.
56. Gnatyuk, V. A. SPIE 2000, 3890, 464.
57. Dawar, A. L. J. Mater. Sci. 1993, 28, 639.
58. Lai, W. C.; Yokoyama, M. J. Appl. Phys. 2000, 39, L1138.
59. Katsnelson, E. J. Appl. Phys. 1995, 77, 4604.
60. Aohi, T.; Hatanka, Y.; Look, D. C. Appl. Phys. Lett. 2000, 76, 3257.
61. Oh, M. S.; Hwang, D. K.; Lim, J. H.; Choi, Y. S.; Park, S. J. Appl. Phys. Lett. 2007, 91, 042109.
62. Ji, L. F.; Jiang, Y. J. Appl. Phys. Lett. 2004, 85, 1577.
63. Chang, L.; Jiang, Y. J. Appl. Phys. Lett. 2007, 90, 082505.
64. Zhao, Y.; Jiang, Y. J. J. Appl. Phys. 2008, 103, 114903.
65. Trtica, M. S.; Tarasenko, V. F.; Gakovic, B. M.; Fedenev, A. V.; Petkovska, L. T.; Radak, B.; Lipatov, E. I.; Shulepov, M. A. Appl. Surf. Sci. 2005, 252, 474.
66. Gower, M. C. In Laser Processing in Manufacturing, Crafer, R. C., Oakley, P. J., Eds.; Chapman and Hall: London 1993. Vol. 189.
67. Cappelli, E.; Orlando, S.; Sciti, D.; Montozzi, M.; Pandolfi, L. Appl. Surf. Sci. 2000, 154, 682.
68. Yue, T. M.; Yu, J. K.; Mei, Z.; Man, H. C. Mater. Lett. 2002, 52, 206.
69. Yymada, K.; Moristita, S.; Kutsuna, M.; Miyamoto, I. First International Symposium on High-Power Laser Macroprocessing[C]; Isamu, Miyamoto, 2003, p 65.
70. Milovanovic, D. S.; Radak, B. B.; Gakovic, B. M.; Batani, D.; Momcilovic, M. D.; Trtica, M. S. J. Alloy. Compd. 2010, 501, 89.
71. Golebiewski, M.; Kluezl, G.; Major, R.; Mróz, W.; Wierzchon, T.; Ebner, R.; Major, B. Mater. Chem. Phys. 2003, 81, 315.
72. Vorobyev, Y.; Guo, C. Phys. Rev. B 2005, 72, 195422.
73. Vorobyev, Y.; Guo, C. Opt. Photonics News 2007, 18, 43.
74. Vorobyev, Y.; Guo, C. J. Appl. Phys. 2008, 103, 43513.
75. Sipe, J. E.; Young, J. F.; Preston, J. S.; Driel, H. M.v Phys. Rev. B 1983, 27, 1141.
76. Young, J. F.; Preston, J. S.; Driel, H. M.v.; Sipe, J. E. Phys. Rev. B 1983, 27, 1155.
77. Oron, M.; Sorensen, G. Appl. Phys. Lett. 1979, 35, 782.
78. Ehrlich, D. J.; Brueck, S. R. J.; Tsao, J. Y. Appl. Phys. Lett. 1982, 41, 1.
79. Reif, J.; Costache, F.; Henyk, M.; Pandelov, S. V. Appl. Surf. Sci. 2002, 197–198, 891.
80. Walser, R. M.; Baecker, M. F.; Ambrose, J. G.; Sheng, D. Y. In Laser and Electron Beam Solid Interactions and Materials Processing, Vol. 177; Elsevier: New York, 1981.
81. Vechten, J. A.v. Solid State Commun. 1981, 39, 285.
82. Li, Z. H.; Li, P. N.; Fan, J. Q.; Fang, R. R.; Zhang, D. M. Opt. Laser Eng. 2010, 48, 64.
83. Bonch-Bruevich, M.; Libenson, M. N.; Makin, V. S.; Trubaev, V. V. Opt. Eng. (Bellingham) 1992, 31, 718.
84. Ozkan, A. M.; Malshe, A. P.; Railkar, T. A. Appl. Phys. Lett. 1999, 75, 3716.
85. Tan, B.; Venkatakrishnan, K. J. Micromech. Microeng. 2006, 16, 1080.
86. Wang, J. C.; Guo, C. L. J. Appl. Phys. 2005, 87, 251914.
87. Wang, J. C.; Guo, C. L. J. Appl. Phys. 2006, 100, 023511.
88. Vorobyev, A. Y.; Guo, C. J. Appl. Phys. 2008, 104, 063523.
89. Ma, H. L.; Guo, Y.; Zhong, M. J.; Li, R. X. Appl. Phys. A 2007, 89, 707.
90. Huang, M.; Zhao, F. L.; Xu, Z. Z.; Xu, N. S. Opt. Express 2008, 16, 19354.
91. Sudrie, L.; Franco, M.; Prade, B.; Mysyrowicz, A. Opt. Commun. 1999, 171, 279.
92. Cho, S. H.; Kumagai, H.; Midorikawa, K. Nucl. Instrum. Methods Phys. Res. B 2002, 197, 73.
93. Vorobyev, A. Y.; Topkov, A. N.; Gurin, O. V.; Svich, V. A.; Guo., C. L. Appl. Phys. Lett. 2009, 95, 121106.
94. Mevel, E.; Breger, P.; Trainhan, R.; Petite, G.; Agostini, P. Phys. Rev. Lett. 1993, 70, 406.
95. Glezer, E. N.; Mazur, E. Appl. Phys. Lett. 1997, 71, 882.
96. Bitenskii, I. S.; Murakhmetov, M. N.; Parilis, E. S. Sov. Phys. Tech. Phys. 1979, 24, 618.
97. Cheng, H. P.; Gillaspy, J. D. Phys. Rev. B 1997, 55, 2628.
98. Henyk, M.; Mitzner, R.; Wolfframm, D.; Reif, J. Appl. Surf. Sci. 2000, 154–155, 249.
99. Yablonovitch, E.; Bloembergen, N. Phys. Rev. Lett. 1972, 29, 907.
100. Pronko, P. P.; Dutta, S. K.; Du, D.; Singh, R. K. J. Appl. Phys. 1995, 78, 6233.
101. Gamaly, E. G.; Rode, A. V.; Luther-Davies, B.; Tikhonchuk, V. T. Phys. Plasmas 2002, 9, 949.
102. Stuart, B. C.; Feit, M. D.; Rubenchik, A. M.; Shore, B. W.; Perry, M. D. Phys. Rev. Lett. 1995, 74, 2248.
103. Li, M.; Menon, S.; Nibarger, J. P.; Gibson, G. N. Phys. Rev. Lett. 1999, 82, 2394.
104. Pronko, P. P.; VanRompay, P. A.; Horvath, C.; Loesel, F.; Juhasz, T.; Liu, X.; Mourou, G. Phys. Rev. B 1998, 58, 2387.
105. Schaffer, C. B.; Brodeur, A.; Mazur, E. Meas. Sci. Technol. 2001, 12, 1784.
106. Keldysh, L. V. Sov. Phys. JETP 1965, 20, 1307.
107. Tanaka, T.; Sun, H. B.; Kawata, S. Appl. Phys. Lett. 2002, 80, 312.
108. Chichkov, B. N.; Fadeeva, E.; Koch, J., et al. Proc. SPIE 2006, 6106, 610612.
109. Juodkazis, S.; Nishimura, K.; Misawa, H. Chin. Opt. Lett. 2007, 5, 198.
110. Maruo, S.; Ikuta, K.; Korogi, H. Appl. Phys. Lett. 2003, 82, 133.
111. Schaffer, C. B.; Garcia, J. F.; Mazur, E. Appl. Phys. A Mater. Sci. Proc. 2003, 76, 351.
112. Schaffer, C. B.; Jamison, A. O.; Mazur, E. Appl. Phys. Lett. 2004, 84, 1441.
113. El-Shall, M. S.; Li, S.; Turkki, T.; Graiver, D.; Pernisz, U. C.; Baraton, M. I. J. Phys. Chem. B 1995, 99, 17805.
114. Burr, T. A.; Seraphin, A. A.; Werwa, E.; Kolenbrander, K. D. Phys. Rev. B 1997, 56, 48l8.
115. David, B.; Geohegan, A.; Puretzky, A.; Gerd, D.; Pennycook, S. J. Appl. Phys. Lett. 1998, 72, 2987.
116. Geohegan, D. B.; Puretzky, A. A.; Duscher, G.; Stephen, J. P. Appl. Phys. Lett. 1998, 73, 438.
117. Hata, K.; Fujita, M.; Yoshida, S.; Makimura, T.; Murakami, K.; Shigekawa, H.; Mizutani, W.; Tokumoto, H. Appl. Phys. Lett. 2001, 79, 692.
118. Belomoin, G.; Therrien, J.; Smith, A.; Rao, S.; Twesten, R.; Chaieb, S.; Nayfeh, M. H.; Wagner, L.; Mitas, L. Appl. Phys. Lett. 2002, 80, 841.
119. Nakata, Y.; Muramoto, J.; Okada, T.; Maeda, M. J. Appl. Phys. 2002, 91, 1640.
120. Eliezer, S.; Eliaz, N.; Grossman, E.; Fisher, D.; Gouzman, I.; Henis, Z.; Pecker, S.; Horovitz, Y.; Fraenkel, M.; Maman, S.; Lereah, Y. Phys. Rev. B 2004, 69, 144119.
121. Amoruso, S.; Bruzzese, R.; Spinelli, N.; Velotta, R.; Vitiello, M.; Wang, X.; Ausanio, G.; Iannotti, V.; Lanotte, L. Appl. Phys. Lett. 2004, 84, 4502.
122. Amoruso, S.; Ausanio, G.; Bruzzese, R.; Vitiello, M.; Wang, X. Phys. Rev. B 2005, 71, 033406.
123. Albert, O.; Roger, S.; Glinec, Y.; Loulergue, J. C.; Etchepare, J.; Boulmer –Leborgne, C.; Perriere, J.; Milton, E. Appl. Phys. A Mater. Sci. Proc. 2003, 76, 319.
Laser Ablation 167

124. Li, S.; Silvers, S. J.; El-Shall, M. S. J. Phys. Chem. B 1997, 101, 1794.
125. Movtchan, I. A.; Marine, W.; Dreyfus, R. W.; Le, H. C.; Sentis, M.; Autric, M. Appl. Surf. Sci. 1996, 96–98, 251.
126. MafunÓ, F.; Kohno, J. Y.; Takeda, Y.; Kondow, T.; Sawabe, H. J. Phys. Chem. B 2000, 104, 9111.
127. MafunÓ, F.; Kohno, J. Y.; Takeda, Y.; Kondow, T. J. Phys. Chem. B 2001, 105, 9050.
128. Kumar, B.; Thareja, R. K. J. Appl. Phys. 2010, 108, 064906.
129. Kazakevich, P. V.; Simakin, A. V.; Shafeev, G. A.; Viau, G.; Soumare, Y.; Verduraz, F. B. Appl. Surf. Sci. 2007, 253, 7831.
130. Ashfold, M. N.; Claeyssens, F.; Fuge, G. M.; Henley, S. J. Chem. Soc. Rev. 2004, 33, 23.
131. Suzuki, N.; Makino, T.; Yamada, Y.; Yoshida, T.; Onari, S. Appl. Phys. Lett. 2000, 76, 1389.
132. Murugesan, S.; Kuppusami, P.; Parvathavarthini, N.; Mohandas, E. Surf. Coat. Technol. 2007, 201, 7713.
133. Grigoriu, C.; Nicolae, I.; Ciupinaa, V.; Prodan, G.; Suematsu, H.; Yatsui, K. J. Optoelectron. Adv. Mater. 2004, 6, 825.
134. Ozawa, E.; Kawakami, Y.; Seto, T. Scr. Mater. 2001, 44, 2279.
135. Okubo, N.; Nakazawa, T.; Katano, Y.; Yoshizawa, I. Appl. Surf. Sci. 2002, 197–198, 679.
136. Fojtik, A.; Henglein, A.; Bunsen-Ges, B. Phys. Chem. 1993, 97, 252.
137. Neddersen, J.; Chumanov, G.; Cotton, T. M. Appl. Spectrosc. 1993, 47, 1959.
138. Peyrer, P.; Berthe, L.; Scherpereel, X.; Fabbro, R. J. Mater. Sci. 1998, 33, 1421.
139. Beahe, L.; Fabbro, R.; Peyre, P.; Bartnicki, E. J. Appl. Phys. 1999, 85, 7552.
140. Beahe, L.; Sollier, A.; Peyre, P.; Fabbro, R.; Bartnicki, E. J. Phys. D Appl. Phys. 2000, 33, 2142.
141. Peyer, P.; Berthe, L.; Fabbro, R.; Sollier, A. J. Phys. D Appl. Phys. 2000, 33, 498.
142. Sollier, A.; Berthe, L.; Fabbro, R. Eur. Phys. J. Appl. Phys. 2001, 16, 131.
143. Sakka, T.; Takatani, K.; Ogate, Y. H.; Mabuchi, M. J. Phys. D Appl. Phys. 2002, 35, 65.
144. Saito, K.; Sakka, T.; Ogata, Y. H. J. Appl. Phys. 2003, 94, 5530.
145. Simakin, A. V.; Obraztsova, E. D.; Shafeev, G. A. Chem. Phys. Lett. 2000, 332, 231.
146. Inoue, W.; Okoshi, M.; Inoue, N. Appl. Phys. A 2004, 79, 1457.
147. Kazakevich, P. V.; Voronov, V. V.; Simakin, A. V.; Shafeev, G. A. Quantum Electron. 2004, 34, 951.
148. Kumar, B.; Yadav, D.; Thareja, R. K. J. Appl. Phys. 2011, 110, 074903.
149. Lee, I.; Han, S. W. Chem. Commun. 2001, 1782.
150. Basov, N. G.; Danilychev, V. A.; Popov, Y.; Khodkevich, D. D. Zh. Eksp. Fiz. i Tekh. Pis’ma. Red. 1970, 12, 473.
151. Dijkkamp, D.; Venkateasan, T.; Wu, X. D. Appl. Phys. Lett. 1987, 51, 619.
152. Chrisey, D. B., Hubler, G. K., Eds. Pulsed Laser Deposition of Thin Films; Wiley: New York, 1994.
153. Greer, J. A.; Tabal, M. D. J. Vac. Sci. Technol. A 1995, 13, 1175.
154. Qian, F.; Singh, R. K.; Dutta, S. K.; Pronko, P. P. Appl. Phys. Lett. 1995, 67, 3120.
155. Okoshi, M.; Higuchi, S.; Hanabusa, M. J. Appl. Phys. 1999, 86, 1768.
156. Eason, R., Ed., Vol. 100; Wiley-Interscience: New York, 2007.
157. Harilal, S. S.; Bindhu, C. V.; Tillack, M. S.; Najmabadi, F.; Gaeris, A. C. J. Appl. Phys. 2003, 93, 2380.
158. Harilal, S. S.; O’Shay, B.; Tillack, M. S.; Mathew, M. V. J. Appl. Phys. 2005, 98, 013306.
159. Zhang, D. M.; Li, Z. H.; Yu, B. M.; Huang, M. T.; Guan, L.; Zhong, Z. C.; Li, G. D. Sci. China Ser. A Math. 2001, 44, 1485.
160. Zhang, D. M.; Guan, L.; Li, Z. H.; Zhong, Z. C.; Hou, S. P.; Yang, F. X. Acta Phys. Sin. 2003, 52, 242 (in Chinese).
161. Tan, X. Y.; Zhang, D. M.; Li, Z. H.; Fang, R. R. J. Phys. D Appl. Phys. 2008, 41, 035210.
162. Tan, X. Y.; Zhang, D. M.; Feng, S. Q.; Li, Z. H.; Fang, R. R.; Liu, G. B.; Sun, M. Chin. Phys. Lett. 2008, 25, 198.
163. Chen, K. R.; King, T. C.; Hes, J. H.; Leboeuf, J. N.; Geohegan, D. B.; Wood, R. F.; Puretzky, A. A.; Donato, J. M. Phys. Rev. B 1999, 60, 8383.
164. Chen, K. R.; Leboeuf, J. N.; Wood, R. F.; Geohegan, D. B.; Donato, J. M.; Liu, C. L.; Puretzky, A. A. Phys. Rev. Lett. 1999, 75, 4706.
165. Shibkov, V. M.; Aleksandrov, A. F.; Ershov, A. P.; Timofeev, I. B.; Chernikov, V. A.; Shibkova, L. V. Plasma Phys. Rep. 2005, 31, 895.
166. Harilal, S. S.; O’Shay, B.; Tao, Y. Z.; Tillack, M. S. J. Appl. Phys. 2006, 99, 083303.
167. Zhigilei, L. V.; Garrison, B. J. J. Appl. Phys. 2000, 88, 1281.
168. Geohegan, D. B. Appl. Phys. Lett. 1992, 60, 2832.
169. Wang, E. G. Prog. Phys. 2003, 23, 1 (in Chinese).
170. Guan, L.; Zhang, D. M.; Li, L.; Li, Z. H. Nucl. Instrum. Methods Phys. Res. B 2008, 266, 57.
171. Mayr, S. G.; Moske, M.; Samwer, K.; Taylor, M. E.; Atwater, H. A. Appl. Phys. Lett. 1999, 75, 4091.
172. Troyan, V. I.; Nevolin, V. N.; Maximov, G. A.; Filatov, D. O.; Lægsgaard, E. Phys. Rev. B 2002, 65, 073406.
173. Ohresser, P.; Shen, J.; Barthel, J.; Zheng, M.; Mohan, Ch. V.; Klaua, M.; Kirschner, J. Phys. Rev. B 1999, 59, 3696.
174. Ingen, R. P. v.; Fastenau, R. H. J.; Mittemeijer, E. J. J. Appl. Phys. 1994, 76, 1871.
175. Antoni, F.; Fogarassy, E.; Fuchs, C.; Grob, J. J.; Prevot, B.; Stoquert, J. P. Appl. Phys. Lett. 1995, 67, 2072.
176. Riet, E.; van de, Kools, J. C. S.; Dieleman, J. J. Appl. Phys. 1993, 73, 8290.
177. Voevodin, A. A.; Donley, M. S. Surf. Coat. Technol. 1996, 82, 199.
178. Stevefelt, J.; Collins, C. B. J. Phys. D Appl. Phys. 1991, 24, 2149.
179. Chey, S. J.; Cahill, D. G. Surf. Sci. 1997, 380, 377.
180. Pomeroy, J. M.; Jacobsen, J.; Hill, C. C.; Cooper, B. H.; Sethna, J. P. Phys. Rev. B 2002, 66, 235412.
181. Schiechl, H.; Rauchbauer, G.; Biedermann, A.; Schmid, M.; Varga, P. Surf. Sci. 2005, 594, 120.
182. Shin, B.; Leonard, J. P.; McCamy, J. W.; Michael, J. A. Appl. Phys. Lett. 2005, 87, 181916.
183. Agura, H.; Suzuki, A.; Matsushita, T.; Aoki, T.; Okuda, M. Thin Solid Films 2003, 445, 263.
184. Blank, D. H. A.; IJsselsteijn, R. P. J.; Out, P. G.; Kuiper, H. T. J.; Floksura, J.; Rogalla, H. Mater. Sci. Eng. B 1992, 13, 67.
185. Mosaner, P.; Bazzanella, N.; Bonelli, M.; Checchetto, R.; Miotello, A. Mater. Sci. Eng. B 2004, 108, 33.
186. Tsoutsouvab, M. G.; Panagopoulosb, C. N.; Kompitsasa, M. Appl. Surf. Sci. 2011, 257, 6314.
187. Chen, H. J.; Jia, C. H.; Zhang, X. N.; Zhang, W. F. Vacuum 2010, 85, 193.
188. Warrender, J. M.; Aziz, M. J. Phys. Rev. B 2007, 75, 085433.
189. Dale, D.; Fleet, A.; Suzuki, Y.; Brock, J. D. Phys. Rev. B 2006, 74, 085419.
190. Arias, J. L.; Mayor, M. B.; Pou, J.; León, B.; Pérez-Amor, M. Vacuum 2002, 67, 653.
191. Develos, K. D.; Kusunoki, M.; Mukaida, M.; Ohshima, S. Phys. C 1999, 320, 21.
192. Warrender, J. M.; Aziz, M. J. Appl. Phys. A 2004, 79, 713.
193. Willmott, P. R. Prog. Surf. Sci. 2004, 76, 163.
194. Zhu, Z.; Zheng, X. J.; Li, W. J. Appl. Phys. 2009, 106, 054105.
195. Moholka, A. V.; Shinde, S. S.; Babar, A. R.; Sim, K.-U.; Kwon, Ye-bin; Rajpure, K. Y.; Patil, P. S.; Bhosale, C. H.; Kim, J. H. Sol. Energy 2011, 85, 1354.
196. Warrender, J. M.; Aziz, M. J. Phys. Rev. B 2007, 76, 045414.
168 Laser Ablation

197. Hinnemann, B.; Hinrichsen, H.; Wolf, D. E. Phys. Rev. Lett. 2001, 87, 135701.
198. Hinnemann, B.; Hinrichsen, H.; Wolf, D. E. Phys. Rev. E 2003, 67, 011602.
199. Lam, P.; Liu, S. J.; Woo, C. H. Phys. Rev. B 2002, 66 (4), 045408–045413.
200. Witten, T. A.; Sander, L. M. Phys. Rev. Lett. 1981, 47, 1400.
201. Zhang, Z. Y.; Chen, X.; Lagally, M. G. Phys. Rev. Lett. 1994, 73, 1829.
202. Zhang, Z. Y.; Chen, X.; Lagally, M. G. Science 1997, 276, 377.
203. Bruschi, P.; Cagnoni, P.; Nannini, A. Phys. Rev. B 1997, 55, 7955.
204. Taylor, M. E.; Atwater, A. H. Appl. Surf. Sci. 1998, 127–129, 159.
205. Kuzma, M.; Bester, M.; Pyziak, L.; Stefaniuk, I.; Virt, I. Appl. Surf. Sci. 2000, 168, 132.
206. Groot, J. S. D.; Estabrook, K. G.; Kruer, W. L.; Drake, R. P.; Mizuno, K.; Cameron, S. M. Phys. Fluids B 1992, 4, 701.
207. Li, L.; Zhang, D. M.; Li, Z. H.; Guan, L.; Tan, X. Y. Phys. B 2006, 383, 194.
208. Fang, R. R.; Zhang, D. M.; Li, Z. H.; Li, L.; Tan, X. Y.; Yang, F. X. Phys. Status Solidi A 2007, 204, 4241.
209. Zhang, D. M.; Liu, D.; Li, Z. H.; Guan, L.; Tan, X. Y.; Li, L.; Fang, R. R.; Hu, D. Z.; Liu, G. B. Appl. Surf. Sci. 2007, 253, 6144.
210. Amoruso, S. Appl. Surf. Sci. 1999, 138–139, 292.
211. Lunney, J. G.; Jordan, R. Appl. Surf. Sci. 1998, 127–129, 941.
212. Zel’dovich, Ya. B.; Raizer, Yu. P. In Physics of Shock Waves and High-Temperature Hydrodynamics Phenomena; Academic Press: New York, 1966;. Vol. 1.
213. Amoruso, S.; Toftmann, B.; Schou, J. Phys. Rev. E 2004, 69, 056403.
214. Amoruso, S. Appl. Phys. A 1999, 69, 323.
215. Inam, A.; Wu, X. D.; Venkatesan, T.; Ogale, S. B.; Chang, C. C.; Dijkkamp, D. Appl. Phys. Lett. 1987, 51, 1112.
216. Li, L.; Zhang, D. M.; Li, Z. H.; Guan, L.; Tan, X. Y.; Liu, D.; Fang, R. R. Phys. Status Solidi A 2006, 203, 906.
217. Fang, R. R.; Zhang, D. M.; Li, Z. H.; Yang, F. X.; Li, L.; Tan, X. Y.; Sun, M. Solid State Commun. 2008, 145, 556.
218. Huang, Y. L.; Yang, F. H.; Liang, G. Y.; Su, J. Y. Chin. J. Lasers 2003, 30, 449 (in Chinese).
219. Yang, L. B.; Liu, X. F.; Zhang, K. X.; Zhang, G. R.; Wang, C. Y. High Power Laser Part. Beams 1994, 6, 99 (in Chinese).
220. Lu, Y. W.; Lu, Q. S. Laser Technology in Military Applications, 31; National Defence Industry Press: Beijing, 1999 (in Chinese).
221. Zheng, Q. G.; Gu, J. H. Laser Interaction with Matter, 13; Huazhong University Press: Wuhan, 1996 (in Chinese).
222. Lu, J. Laser Interaction with Matter, 18; China Machine Press: Beijing, 1996 (in Chinese).
223. Hassan, A.; Elnicklawy, M.; Eladawi, M.; Hemida, A. Opt. Laser Technol. 1993, 25, 155.
224. Billings, B. H.; Frederikse, H. P. R. American Institute of Physics Handbook. 3rd ed; McGraw-Hill: New York, 1972; p 346.
225. Zhang, D. M.; Tan, X. Y.; Li, Z. H. Phys. B 2005, 357, 348.
226. Xu, X. F.; Willis, D. A. J. Heat Transfer 2002, 124, 293.
227. Miotello, A.; Kelly, R. Appl. Phys. Lett. 1995, 67, 3535.
228. Harrach, R. J. J. Appl. Phys. 1976, 47, 2473.
229. Singh, R. K. J. Non-Cryst. Solids 1994, 178, 199.
230. Zhang, D. M.; Li, Z. H.; Huang, M. T.; Zhang, M. J.; Guan, L.; Zou, M. Q.; Zhong, Z. C. Acta Phys. Sin. 2001, 50, 914.
231. Saidane, A.; Pulko, S. H. Microelectron. Eng. 2000, 51–52, 469.
232. Luikov, A. V. Int. J. Heat Mass Transfer 1966, 9, 139.
233. Taitel, Y. Int. J. Heat Mass Transfer 1972, 15, 369.
234. Sieniutycz, S. Int. J. Heat Mass Transfer 1977, 20, 1221.
235. Jiang, R. Q. Heat Conduction, Mass Diffusion and Transient Impact Effect of Momentum Transfer; Science Press: Beijing, 1997, 154.
236. Jiang, F. M.; Liu, D. Y. Adv. Mechanics 2002, 32, 128.
237. Joseph, D. D.; Presiosi, L. Rev. Mod. Phys. 1989, 61, 41.
238. Zhang, D. M.; Li, L.; Li, Z. H.; Guan, L.; Tan, X. Y. Phys. B 2005, 364, 285.
239. Zhang, D. M.; Li, L.; Li, Z. H.; Guan, L.; Tan, X. Y.; Liu, D. Eur. Phys. J. Appl. Phys. 2006, 33, 91.
240. Villars, P.; Phillips, J. C. Phys. Rev. B 1988, 37, 2345.
241. Sun, C. W.; Lu, Q. S.; Fan, Z. X.; Chen, Y. Z. The Effect of Laser Irradiation, 93; National Defence Industry Press: Beijing, 2002 (in Chinese).
242. Rohsenow, W. M.; Hartnett, J. P.; Ganic, E. N. Heat Transfer Hand Book; McGraw-Hill: New York, 1985; p 236.
243. Kaganov, M. I.; Lifshita, L. M.; Tanatarov, L. V. Sov. Phys. JETP 1957, 4, 173.
244. Anisimov, S. I.; Kapeliovich, B. L.; Perelman, T. L. Sov. Phys. JETP 1974, 39, 375.
245. Fujimoto, J. G.; Liu, J. M.; Ippen, E. P. Phys. Rev. Lett. 1984, 53, 1837.
246. Elsayed-Ali, H. E.; Norrisal, T. B.; Pessot, M. A.; Mourou, G. A. Phys. Rev. Lett. 1987, 58, 1212.
247. Allen, P. B. Phys. Rev. Lett. 1987, 59, 1460.
248. Brorson, S. D.; Kazeroonian, A.; Moodera, J. S.; Face, D. W.; Cheng, T. K.; Ippen, E. P.; Dresselhaus, M. S.; Dresselhaus, G. Phys. Rev. Lett. 1990, 64, 2172.
249. Chichkov, B. N.; Momma, C.; Nolte, S.; Alvensleben, F. V.; Tünnermann, A. Appl. Phys. A 1996, 63, 109.
250. Corkum, P. B.; Brunel, F.; Sherman, N. K. Phys. Rev. Lett. 1988, 61, 2886.
251. Kanavin, A. P.; Smetanin, I. V.; Isakov, V. A.; Afanasiev, Y. V.; Chichkov, B. N.; Wellegehausen, B.; Nolte, S.; Momma, C.; Tünnermann, A. Phys. Rev. B 1998, 57,
14698.
252. Gamaly, E. G.; Rode, A. V.; Uteza, O.; Samoc, M.; Luther-Davies, B. Appl. Surf. Sci. 2002, 197–198, 730.
253. Gamaly, E. G.; Madsen, N. R. Laser Part. Beams 2005, 23, 167.
254. Hetnarski, R. B.; Ignaczak, J. J. Therm. Stresses 1999, 22, 451.
255. Hetnarski, R. B.; Ignaczak, J. Int. J. Solids Struct. 2000, 37, 215.
256. Chen, J. K.; Grimes, L. E. Int. J. Solids Struct. 2002, 39, 3199.
257. Gamaly, E. G. Phys. Rep. 2011, 508, 91.
258. Fang, R. R.; Zhang, D. M.; Wei, H.; Hu, D. Z.; Li, Z. H.; Tan, X. Y.; Sun, M.; Yang, F. X. Eur. Phys. J. Appl. Phys. 2008, 42, 229.
259. Fang, R. R.; Hua, W.; Li, Z. H.; Zhang, D. M. Solid State Commun. 2012, 152, 108.
260. Fann, W. S.; Storz, R.; Tom, H. W. K. Phys. Rev. B 1992, 46, 13592.
261. Sun, C. K.; Vallée, F.; Acioli, L.; Ippen, E. P.; Fujimoto, J. G. Phys. Rev. B 1993, 48, 12365.
262. Sun, C. K.; Vallée, F.; Acioli, L. H.; Ippen, E. P.; Fujimoto, J. G. Phys. Rev. B 1994, 50, 15337.
263. Wang, J. C.; Guo, C. L. Appl. Phys. Lett. 2005, 87, 251914.
264. Wang, X. Y.; Riffe, D. M.; Lee, Y. S.; Downer, M. C. Phys. Rev. B 1994, 50, 8016.
265. Lin, Z. B.; Zhigilei, L. V. Appl. Surf. Sci. 2007, 253, 6295.
266. Povarnitsyn, M. E.; Itina, T. E.; Sentis, M.; Khishchenko, K. V.; Levashov, P. R. Phys. Rev. B 2007, 75, 235414.
267. Lin, Z. B.; Zhigilei, L. V. Phys. Rev. B 2008, 77, 075133.
268. Lin, Z. B.; Zhigilei, L. V. Proc. SPIE 2006, 6261, 62610U–1.
Laser Ablation 169

269. Tsuchiya, T.; Kawamura, K. Phys. Rev. B 2002, 66, 094115.


270. Fang, R. R.; Zhang, D. M.; Hua, W.; Li, Z. H.; Yang, F. X.; Gao, Y. H. Laser Part. Beams 2010, 28, 157.
271. Schafer, C.; Urbassek, H. M.; Zhigilei, L. V. Phys. Rev. B 2002, 66, 115404.
272. Hirayama, Y.; Obara, M. J. Appl. Phys. 2005, 97, 064903.
273. Amoruso, S.; Ausanio, G.; Bruzzese, R.; Gragnaniello, L.; Lanotte, L.; Vitiello, M.; Wang, X. Appl. Surf. Sci. 2006, 252, 4863.
274. Stuart, B. C.; Feit, M. D. J. Opt. Soc. Am. B 1996, 13, 459.
275. Jiang, L.; Tsai, H. L. J. Heat Transfer 2005, 127, 1167.
276. Lu, Q. M.; Mao, S. S.; Mao, X. L.; Russo, R. E. Appl. Phys. Lett. 2002, 80, 3072.
277. Lu, Q. M. Phys. Rev. E 2003, 67, 016410.
278. Amoruso, S.; Armenante, M.; Bruzzese, R.; Spinelli, N.; Velotta, R.; Wang, X. Appl. Phys. Lett. 1999, 75, 7.
279. Guo, X. D.; Li, R. X.; Hang, Y.; Xu, Z. Z.; Yu, B. K.; Ma, H. L.; Lu, B.; Sun, X. W. Mater. Lett. 2009, 62, 1769.
280. Zeng, X.; Mao, X. L.; Greif, R.; Russo, R. E. Appl. Phys. A 2005, 80, 237.
281. Crouch, C. H.; Carey, J. E.; Warrender, J. M.; Aziz, M. J.; Mazur, E.; Génin, F. Y. Appl. Phys. Lett. 2004, 84, 1850.
282. Furusawa, K.; Takahashi, K.; Kumagai, H.; Midorikawa, K.; Obara, M. Appl. Phys. A 1999, 69, S359.
283. Huang, C. H.; Shen, H. Y.; Zeng, Z. D. Opt. Laser Technol. 1990, 22, 345.
284. Huang, Y. L.; Yang, F. H.; Liang, G. Y.; Su, J. Y. Chin. J. Lasers 2003, 30, 449.
285. Fisher, D.; Fraenkel, M.; Zinamon, Z.; Henis, Z.; Moshe, E.; Horovitz, Y.; Luzon, E.; Maman, S.; Eliezer, S. Laser Part. Beams 2005, 23, 391.
286. Wellershoff, S. S.; Hohlfeld, J.; Güdde, J.; Matthias, E. Proc. SPIE 1998, 3343, 378.

You might also like