Sie sind auf Seite 1von 75

Progress in Materials Science 76 (2016) 154228

Contents lists available at ScienceDirect

Progress in Materials Science


journal homepage: www.elsevier.com/locate/pmatsci

Femtosecond laser induced phenomena in


transparent solid materials: Fundamentals and
applications
Dezhi Tan a, Kaniyarakkal N. Sharafudeen b, Yuanzheng Yue c,d,
Jianrong Qiu a,c,
a
State Key Laboratory of Modern Optical Instrumentation, College of Optical Science and Engineering, Zhejiang University,
Hangzhou 310027, China
b
State Key Laboratory of Luminescent Materials and Devices, South China University of Technology, Guangzhou, China
c
Section of Chemistry, Aalborg University, Aalborg 9000, Denmark
d
State Key Laboratory of Silicate Materials for Architecture, Wuhan University of Technology, Wuhan 430070, China

a r t i c l e

i n f o

Article history:
Received 8 April 2014
Received in revised form 20 June 2015
Accepted 22 September 2015
Available online 28 September 2015

a b s t r a c t
The interaction of intense femtosecond laser pulses with transparent materials is a topic that has caused great interest of scientists
over the past two decades. It will continue to be a fascinating field
in the coming years. This is because many challenging fundamental
problems have not been solved, especially concerning the interaction of strong, ultra-short electromagnetic pulses with matter, and
also because potential advanced technologies will emerge due to
the impressive capability of the intense femtosecond laser to create
new material structures and hence functionalities. When femtosecond laser interacts with matter, a large amount of energy will
be released during an ultra-short period of time, resulting in extremely high energy intensity. This opens the avenue to explore new
lightmatter interacting phenomena, investigate the details of the
dynamical processes of the lightmatter interaction, and fabricate
various integrated micro-devices. In recent years we have witnessed exciting development in understanding and applying femtosecond laser induced phenomena in transparent materials. The
interaction of femtosecond laser pulses with transparent materials
relies on non-equilibrium process with photon beams and this provides new access to create materials and micro-devices that cannot

Corresponding author at: State Key Laboratory of Modern Optical Instrumentation, College of Optical Science and
Engineering, Zhejiang University, Hangzhou 310027, China.
E-mail address: qjr@zju.edu.cn (J. Qiu).
http://dx.doi.org/10.1016/j.pmatsci.2015.09.002
0079-6425/ 2015 Elsevier Ltd. All rights reserved.

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

155

be obtained by other means. Understanding of the physical mechanisms of many induced phenomena is extremely challenging. The
aim of this review is to present a critical overview of the current
state of the art in studying femtosecond laser induced various phenomena in transparent materials, including their physical and
chemical mechanisms, the applications and limitations as well as
the future research trends. The first part of the review presents
the basics of femtosecond laser systems, important parameters
influencing the femtosecond laser interaction with transparent
materials, and a brief description of various energy transfer processes in materials during femtosecond laser irradiation. The second part will give an account on various phenomena such as
multiphoton excited upconversion luminescence, long lasting
phosphorescence, formation of color centers, valence state change,
precipitation of nanoparticles and nanocrystals, microvoids,
polarization-dependent and periodic surface structures, refractive
index change, polymerization and air-bubble formation. The third
part describes recently observed anomalous phenomena such
as induced birefringence, nanogratings, nanovoid arrays, migration
of ions, nonreciprocal photosensitivity, high pressure crystalline
phase, and their underlying mechanisms, and their potential prospects as a new tool for photonic technology development. The
final part points out the major challenges and future research
trends in this promising field.
2015 Elsevier Ltd. All rights reserved.

Contents
1.
2.

3.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Femtosecond laser interaction with matter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.
Basic principles of femtosecond laser . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.1.
Generation of femtosecond laser pulse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.2.
Optical parametric amplification and chirped pulse amplification . . . . . . . . . . . . . . . .
2.1.3.
Optical parametric oscillator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.4.
Femtosecond laser systems and parameters. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.
Basic mechanisms of femtosecond laser interaction with matter . . . . . . . . . . . . . . . . . . . . . . .
2.3.
Parameters affecting the femtosecond lasermatter interaction . . . . . . . . . . . . . . . . . . . . . . . .
2.4.
Regimes of femtosecond laser dielectric modification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5.
Femtosecond laser induced various phenomena and applications. . . . . . . . . . . . . . . . . . . . . . .
2.5.1.
Emission . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5.2.
Formation of color centers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5.3.
Valence state change . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5.4.
Precipitation of crystals and nanoparticles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5.5.
Microvoid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5.6.
Polymerization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5.7.
Periodic surface structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5.8.
Refractive index change . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5.9.
Periodic structures induced by interference fields of femtosecond laser . . . . . . . . . . .
2.5.10.
Formation of bubbles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Femtosecond laser induced anomalous phenomena. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.
Femtosecond laser induced nanogratings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.1.
Mechanism of nanograting formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.2.
Applications of femtosecond laser induced nanogratings . . . . . . . . . . . . . . . . . . . . . . .
3.2.
Periodic nanovoid arrays. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.1.
Void formation mechanism and various works . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

156
157
157
158
158
160
160
161
162
167
168
168
172
174
176
179
181
182
185
188
190
192
192
194
197
199
200

156

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

3.2.2.
Applications of femtosecond laser induced periodic nanovoids . . . . . . . . . . . . . . . . . .
Migration of ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Quill writing and nonreciprocal writing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.4.1.
Quill writing, anisotropic photosensitivity and their mechanisms . . . . . . . . . . . . . . . .
3.4.2.
Nonreciprocal photosensitivity in noncentrosymmetric media . . . . . . . . . . . . . . . . . . .
3.4.3.
Mechanism of nonreciprocal photosensitivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.5.
Formation of high pressure crystalline phase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Conclusion and perspective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3.
3.4.

4.

201
203
205
206
208
209
211
213
215
215

1. Introduction
Laser is one of the greatest inventions of the humankind in the last century and has greatly changed
all aspects of our life. Laser leads to the generation of increasingly large optical electric fields,
which have demonstrated a succession of new high intensity optical regimes [1]. Femtosecond
(1 fs = 1015 s) laser is a pulsed laser with the pulse width from 1 to 1000 femtoseconds, and was
developed in the last 80s [1,2]. Femtosecond lasers are a clear technological breakthrough with
exciting potential for many applications and have brought about impressive progress in the study
of lightmatter interaction [1,35]. Ever since the pioneering reports on femtosecond laser micromachining in the mid 90s [6,7], revolutions in laser technology together with photonic material design
have enabled the observation of numerous kinds of new physical and chemical phenomena and
induced nano- and micro-structures, and hence, made possible for researchers and technologists to
control and manipulate light in an unusual and interesting way. The extreme high peak intensities
in the focus of femtosecond laser pulses have offered the possibility for a variety of new applications
ranging from precise scalpels for delicate life science [8] to driving sources for table-top particle
accelerators [9]. One of the most promising applications of femtosecond lasers is micromachining
in transparent materials, such as glasses [10], crystals [11], and polymers [12]. In a transparent
material, no linear absorption of the incident femtosecond laser light occurs. With sufficiently high
energy added into the target material, significant nonlinear absorption takes place, leading to transfer
of electrons from the valence band to the conduction band, as shown in Fig. 1A. There are two classes
of nonlinear excitation mechanisms that play a role in the nonlinear absorption, namely, multiphoton
ionization and avalanche ionization [13]. As a result, permanent damage is created.
The interaction of femtosecond laser and transparent materials is a simple, flexible, versatile, and
relatively low-cost route for the fabrication of efficiently multi-dimensional (3D) index-modified
structures without the need of complex photolithographic processes. Todays advanced femtosecond
laser systems offer a variety of interactions with transparent solid materials, from surface machining,
annealing and ablation to 3D refractive index changes (positive or negative; isotropic or anisotropic).

Fig. 1. (A) Schematic diagram of the nonlinear absorption process of electrons from the valence band to the conduction band in
the transparent materials by femtosecond laser excitation. (B) Typical experimental scheme of femtosecond laser micromachining apparatus.

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

157

These interactions depend on both the laser parameters and the material properties [5,10,14].
Recently, many interesting phenomena induced by femtosecond laser in transparent materials have
been reported, e.g., induction of chirality [15] or nonreciprocity writing [16,17], oxidation or reduction
of dopant [10,18], self-organized nanogratings [19], migration of ions [20], nanovoid formation [21],
etc. Especially, micro-fabrication of arbitrary 3D structures has been achieved with the femtosecond
laser technique. Integrated optical components can be directly inscribed into the bulk of transparent
materials by using the focused femtosecond laser beam due to the locally altered structures and properties in the host materials, as demonstrated in Fig. 1B. The femtosecond laser induced structures
exhibit enormous potentialities for widespread applications of micro-photonic crystals, coupler, 3D
optical data storage, bio-photonic components, multicolor imaging, and so on [10,14,17]. There are
several advantages of femtosecond laser micromachining in transparent materials over other photonic
micro-device fabrication techniques [22]. First, the nonlinear nature of the optical absorption confines
the induced changes to the focal volume, thus a well-defined modified region with minimum collateral damage and heat affected zone can be produced. Second, the absorption process is independent of
the materials, making it possible to fabricate optical micro-devices in substrates of different transparent materials. Third, the excitation caused by the femtosecond pulse is much more deterministic than
that by pulses with longer duration. The optical response of the longer pulse relies statistically on the
number of defect sites or thermally excited electronhole pairs. In the case of femtosecond laser
micromachining, no defect electrons are needed to generate the nonlinear absorption process.
However, despite these exciting prospects for fabrication of different structures and great efforts in
understanding the femtosecond lasermaterials interaction, there are some challenges for large scale
micromachining and for defining the micromachining windows for different types of structural modifications. In other words, the use of the femtosecond laser for fabricating micro-devices is still at its
early stage of development. A key prerequisite for the future widespread use of femtosecond laser
matter interaction is a deep understanding of the influences of the experimental parameters on the
phenomena and the characteristics of the created structures. In the present paper we provide a comprehensive review of the femtosecond laser induced phenomena in transparent materials, and current
understanding of those phenomena. The mechanisms of energy transfer from the laser, underlying
physics of the phenomena, the subsequent structural modifications, and some of the promising applications are discussed. This review will contribute to efficient and optimized creation of high quality
and true 3D photonic structures.

2. Femtosecond laser interaction with matter


2.1. Basic principles of femtosecond laser
Laser is a device that generates light by means of stimulated emission of radiation. The two essential components necessary for laser operation are a laser cavity and a gain medium. The laser cavity
allows the propagation of photons in a restricted narrow frequency range and spatial direction to build
up. The modes of the laser cavity in which photons can propagate are the cavity modes. Build-up of
laser radiation within the cavity is achieved initially via spontaneous emission and then via stimulated
emission from the gain medium in which population inversion has been established in the pair of laser
levels. Light amplification is obtained by stimulated emission. During the laser operation, the population inversion is continuously sustained. The lasers are produced both in continuous wave (CW) and
pulsed mode operations depending on their applications. Since this review mainly deals with femtosecond laser, we will briefly describe the pulsed laser operation and its features. In order to get
nanosecond laser pulses, the well known strategy is Q-switching (Q is the quality factor of the cavity).
The laser cavity may contain a Q-switch, which initially reduces the cavity and then gets switched to
make the cavity Q higher after the lasing level of the gain medium is populated. But to produce
picosecond pulses, the conventional method employed is called mode locking. Without frequencyselective elements inside the laser resonator, the laser generally oscillates simultaneously on many
resonator modes within the spectral gain profile of the active medium. In this multimode operation,
no definite phase relations exist between the different oscillating modes, and the laser output is the

158

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

sum of the intensities of all oscillating modes, and also they are more or less randomly fluctuating in
time. If coupling between the phases of these simultaneously oscillating modes is established, a coherent superposition of the mode amplitudes may be reached, leading to the generation of short output
pulses in the picosecond range. Hence, the laser can be mode-locked by inserting elements into the
cavity, which lock the modes of the cavity in such a way that the output consists of a series of very
short pulses. This mode locking has been achieved by using optical modulators inside the laser
resonator, by pumping the medium with another mode-locked laser (active mode-locking) or by
saturable absorbers (passive mode-locking) or by a combined action of both locking techniques
[1,2326]. Using a mode-locked pump laser, matching of resonator length is realized with the
synchronous pumping technique for ultrafast pulse generation. These conventional mode locking
techniques result in typical pulse duration of picoseconds. To generate even shorter pulses or
femtosecond laser pulses, some special mode-locking techniques are required, and will be discussed
in next section.
2.1.1. Generation of femtosecond laser pulse
The widely adopted techniques for generating femtosecond laser pulses in the high power regime
include Kerr lens mode-locking (KLM), the colliding pulse mode-locking (CPM), and the hybrid
mode-locking. In the case of CPM, which is the first technique to break the picosecond limit, using
an absorber inside a ring resonator, a CW ring dye laser can be passively mode-locked [22]. In this case
the oppositely traveling short-pulses collide in an absorber, and thus the total pulse intensity in the
absorber, where the two pulses collide, is twice that of a single pulse, resulting in larger saturation
and less absorption, and hence a net maximum gain. By properly designing the amplifying gain and
the absorption losses, this situation can be realized and this mode of operation results in short pulses
down to 50 fs. In the hybrid method, a saturable absorbing medium is inserted inside a synchronously
pumped cavity. In this case, the pumping is done with a laser, whose modes are already locked using
the active locking methods. More choices of wavelengths, tunability, and powers can be obtained
through the hybrid locking method than that through a simple passive locking method [25]. More
importantly, a major breakthrough has been achieved owing to the discovery of the self-modelocking in a Ti:sapphire laser by de Spence et al. [27]. That is the development of an oscillator a
key ingredient in todays many commercial laser systems. For the cases of the active or passive locking
method, the nonlinear properties of the amplifying medium are always crucial for the locking process.
For the self-locking modes, the nonlinear properties are crucial for ensuring that the modes may lock,
partially or totally, without any need of an external modulation (active locking) or a saturable absorbing medium (passive locking). To do so, the amplifying medium must enable narrowing the pulse at
each of its round trips through the cavity. KLM is the most widely used self-locking technique, which
takes advantage of the electronic Kerr effect to create an artificial fast saturable absorber [28]. At the
same time, a fiber-based femtosecond technology has been developed. Fiber oscillators using many
mode-locking techniques are commercially available [22].
2.1.2. Optical parametric amplification and chirped pulse amplification
For practical purposes, femtosecond laser with high power and tunable wavelengths should be
applied. Parametric nonlinearities are optical nonlinearities with an instantaneous response based
on the second order and third order nonlinearity of a medium, which cause frequency doubling,
sum and difference frequency generation, parametric amplification and oscillation, and four-wave
mixing. Usually, phase matching is a condition for achieving high efficiency in such processes. This
occurs only in a limited bandwidth. For more applications, the parameters affecting the phase matching should be optimized to keep the wavelengths in the range, where the nonlinear interaction is
strong. Parametric amplification is a phenomenon, for which a signal can be amplified using a parametric nonlinearity and a pump wave. For optical parametric amplifiers (OPAs), either the second
order nonlinearity or the third order nonlinearity can be utilized, as shown in Fig. 2A.
With the rapid development of solid state active materials, new nonlinear optical crystals, and
mode-locking and amplification techniques, ultrafast OPAs are getting more important as a practical
source of femtosecond pulses with tunable wavelengths across the visible and infrared spectral
ranges. The technology of ultrafast pulse generation at numerous wavelengths has rapidly developed,

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

159

Fig. 2. (A) Schematic of an OPA. Left inset: phase-matching geometry for noncollinear OPA. Right inset: energy diagram for the
parametric amplification process. (B) Schematic of a CPA-based laser system. (C) Schematic of an OPCPA system. (Reproduced
with permission from [35,36].)

particularly after the discovery of novel nonlinear optical crystals, such as b-Barium Borate (BBO) and
lithium triborate (LBO) [2932]. This is reflected by considerably improved optical characteristics,
high nonlinear optical coefficient, low group velocity dispersion, broad transparency range, and high
damage threshold.
OPAs are important tools for ultrafast spectroscopy, and much higher light intensity is also necessary for strong field physics and more applications [3,33,34]. One of the most important concepts for
generating high intensity, ultrashort laser pulses is known as chirped pulse amplification (CPA)
[27,33,35,36]. The invention of CPA has advanced the Ti:sapphire laser technology, so that the production of multimillijoule, ultrashort pulses has become standard technology [35]. The CPA process can
lead to both amplitude and phase distortions, preventing the optimal recompression of the laser pulse
and achieving high peak intensity. Stretching and recompression over many orders of magnitude in
pulse duration is a process that requires high accuracies in the design and manufacturing of optical
components and in the construction of the stretcher and compressor (Fig. 2B).
Recently, OPA and CPA have been integrated into optical parametric chirped pulse amplification
(OPCPA, Fig. 2C), and hence the advantages of both techniques can be combined [35,36]. This enables
the generation of high-intensity ultrafast laser pulses in widely different parts of the optical spectrum,
with pulse durations in the few-cycle regime and peak powers reaching the terawatt (TW) level and
beyond. As a result of the phase-preserving properties of parametric amplification, OPCPA has made it
possible to produce intense carrier-envelope-phase-controlled few-cycle laser pulses in various parts
of the spectrum. Compared to conventional CPA, OPCPA is an instantaneous process, in which comparable pulse durations are necessary for the stretched signal and the pump-typically in the order of
nanoseconds in high-energy applications. OPCPA systems exhibit many attractive characteristics, such
as high gain, very broad optical spectrum that only weakly relies on gain, spectral tunability, and negligible thermal load. Those characteristics result in the widespread applications in the preamplifier
stage for high energy short pulse lasers, and in schemes that lead to producing higher peak intensity
through even shorter pulses than those available from CPA systems [36].

160

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

2.1.3. Optical parametric oscillator


A crucial question is how to realize femtosecond laser pulse at various wavelengths. Optical
Parametric Oscillator (OPO) technique is used to generate femtosecond laser pulses with different
wavelength. Usually, the OPO consists essentially of an optical resonator and a nonlinear optical
crystal. Specifically, in the parametric process, a nonlinear medium (usually a crystal) converts the
high energy photon (the pump wave) into two lower energy photons (the signal and idler waves).
The optical resonator serves to make at least one of signal and idler waves resonating. In the nonlinear
optical crystal, the pump, signal and idler waves overlap. The interaction among these three waves
leads to amplitude gain for signal and idler waves (parametric amplification) and a corresponding
deamplification of the pump wave. The exact wavelengths of the signal and idler are determined by
the angle that the pump wave vector makes with respect to the crystal axis. Energy can be efficiently
transferred to the parametric waves if all three waves are traveling at the same velocity. Under most
circumstances, the variation of index of refraction with crystal angle and wavelength allows this
phase matching condition to be met only for a single set of wavelengths for a given crystal angle
and pump wavelength. Thus as the crystal rotates, different wavelengths of light are produced. When
the crystal is contained in a resonant cavity, feedback generates gain in the parametric waves in a
process similar to buildup in a laser cavity. Thus, light output at the resonated wavelength (and other
parametric wavelengths produced simultaneously) occurs. The cavity can either be singly resonant at
either the signal or idler wavelength, or it can be doubly resonant at both wavelengths [3].
2.1.4. Femtosecond laser systems and parameters
Until now, various types of femtosecond laser system have been employed for investigating light
bulk transparent material interactions [37]. The most commonly used systems are regeneratively
amplified Ti:sapphire lasers (800 nm and other wavelengths based on OPA or second harmonic generation (SHG) and third harmonic generation (THG)) with 1500 kHz repetition rate, a few lJ pulse
energy and 50200 femtosecond pulse duration. When the pulse energy is decreased to 20100 nJ
by increasing the repetition rate to 525 MHz in a stretched cavity configuration, a Ti:sapphire oscillator can also be used for micromachining [38,39]. In this case, modifications are performed in the high
frequency regime and the writing speeds are dramatically increased. Besides, innovative ytterbiumbased lasers (1030 nm) have also been adopted to induce various structural phenomena in transparent
materials, which operate in a somewhat intermediate regime with a repetition rate comparable to
(actually slightly lower than) the inverse of the heat diffusion time [4042]. Recently, femtosecond
laser based on Yb3+-doped fiber has been commercially available [43,44].
Femtosecond lasers exhibit several major advantages over conventional lasers: (i) the ultra-short
pulse duration enables a measurement with extremely short temporal resolution on a femtosecond
scale, (ii) the focused ultra-short pulses can generate extremely high energy intensity for frontiers
research in the physics and technology of lightmatter interactions. As a result, processes and induced
phenomena will be varied using lasers with different energy intensity and pulse duration, as displayed
in Fig. 3. There are many important parameters which influence the femtosecond material processing,

Fig. 3. Pulsed lasermatter interaction with different energy intensity (A) and time scale of the physical phenomena associated
with femtosecond lasermatter interaction (B).

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

161

such as the polarization (p), the laser light wavelength (k), the pulse energy at the point of laser
interaction with the transparent materials (E), the pulse duration (s), the pulse repetition rate (f),
the duration of the irradiation, and the numerical aperture (NA) of the focusing lens [45]. Therefore,
a reasonable design of experiments must be based on these parameters and conditions in order to
create new structures and give rise to a specific functionality in transparent materials. More
discussions will be given later.
2.2. Basic mechanisms of femtosecond laser interaction with matter
In the past two decades, femtosecond lasermatter interaction has been extensively studied. The
understanding of the physical processes involved in this interaction and their associated characteristic
time scales is of particular importance and provides an insight why femtosecond laser is a powerful
tool for 3D micromachining applications [5,13,46,47]. Although the physical picture of femtosecond
lasermatter interaction is not completely clear, the following scenario (also see Fig. 3) is generally
accepted. When a single high fluence femtosecond laser pulse irradiates into a material, the laser
energy is first absorbed by the transparent material and this induces generation of photoelectrons.
These electrons then transfer their kinetic energy to the lattice over a picosecond timescale.
Consequently, the heat diffuses, the material melts and fusion or explosion occurs, leaving permanent
structural changes. However, it is only a qualitative description, and the induced phenomena depend
on the chemical natures and physical properties of materials and laser irradiation conditions.
Usually, the femtosecond lasermatter interaction is realized by focusing the laser beam into the
transparent materials with a lens. Then, the beam is moved transversally or longitudinally. As the
gap between the valence and conduction band is larger than the energy of a photon for the light
usually used at 800 nm (Ti:sapphire laser), a single photon cannot induce the band gap transition of
an electron. Thus the linear absorption is forbidden. As suggested above, the extremely high laser peak
intensity can lead to nonlinear absorption, during which the electron can simultaneously absorb the
energy from multiple photons to cause the band gap transition [5,10,13,14]. The multiphoton
absorption (MPA) rate (P(I)) strongly depends on its intensity (I) as described by the following
power law:

PI rk Ik

2:1

where rk is the MPA coefficient for the absorption of k photons. Photoionization and avalanche
ionization provide two possible ways for the multiphoton excitation.
The direct excitation of the electron by the strong femtosecond laser field is called photoionization
(including multiphoton ionization and the tunneling ionization), which occurs in two different
regimes, depending on the laser frequency and intensity [13,46,48,49]. When the femtosecond laser
electric field is sufficiently strong, the Coulomb field that binds a valence electron to its parent atom
will be greatly suppressed. Subsequently, the bound electron can tunnel through the short barrier and
becomes free, and this process is described as tunneling ionization, which plays a dominant role in the
femtosecond lasermatter interaction under strong laser field and low laser frequency, as demonstrated in Fig. 4A.
Multiphoton ionization dominates the nonlinear femtosecond laser process at high laser
frequencies (but still below for linear absorption) through the simultaneous absorption of several
photons by an electron, as shown in Fig. 4C. For this MPA mechanism, the electron must absorb
enough photons to be excited from valance to conduction band. The photoionization rate depends
strongly on femtosecond laser intensity, as suggested by the above nonlinear absorption equation.
Nevertheless, the tunneling ionization rate depends weakly on the laser intensity compared to the
multiphoton rate. The transition between multiphoton ionization and tunneling ionization can be
described by the Keldysh parameter:

x mcne0 Eg
e

1=2
2:2

162

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

Fig. 4. Schematic diagrams of the photoionization excited by femtosecond laser. (A) Tunneling ionization, (B) mixture of
tunneling and multiphoton ionization, (C) multiphoton ionization, and (D, E) avalanche ionization. (Reproduced with
permission from [13].)

where x is the laser frequency, I is the laser intensity at the focus, m and e are the reduced mass and
electron charge, respectively, c is the velocity of light, n is the refractive index of the material, Eg is the
band-gap of the material and e0 is the permittivity of free space [13,48]. When the Keldysh parameter
is larger than about 1.5, photoionization is a multiphoton process. Otherwise, tunneling ionization
happens. There may be an intermediate regime, in which the tunneling and multiphoton ionization
happens simultaneously, as revealed in Fig. 4B.
Moreover, the electron excited to the conduction band can absorb several laser photons sequentially and act as a seed to another process called avalanche ionization, as displayed in Fig. 4D and E
[50,51]. The energy (Eabs), which is absorbed by n electrons in the conduction band, can exceed the
band gap (E), i.e.,

Eabs nhm > E

2:3

where h is the Planck constant and m is the frequency of the electron excitation. In this case, the highly
ionized electron will then release the excess energy to excite another electron to the conduction band
through direct collision, as demonstrated in Fig. 4E. The resultant two electrons in the conduction
band can then repeat this process and ionize more electrons. Finally, an exponential increase of the
free electrons will be achieved, namely avalanche ionization. Such ionization creates highly absorptive
and dense plasma, which facilitates energy transfer from the femtosecond laser pulses to the transparent materials. As a result, many interesting phenomena can be observed during the femtosecond laser
processing.
Following the nonlinear processes, part of the absorbed optical energy is transferred from the electrons to the lattice by electronphonon coupling in the picosecond timescale. Within a couple of
nanoseconds, a pressure or a shock wave separates from the dense, hot focal volume [5254]. On
the microsecond timescale, the injected energy is transported out of the irradiated region by thermal
diffusion. Permanent modifications are produced by melting or non-thermal ionic motion and resolidification at a sufficiently high energy.
2.3. Parameters affecting the femtosecond lasermatter interaction
As described above, many parameters affect the process and resultant phenomena of femtosecond
lasertransparent material interaction, e.g., the polarization (p), the laser light wavelength (k), the
pulse energy at the point of laser interaction with the transparent materials (E), the pulse duration
(s), the pulse repetition rate (f), the duration of the irradiation, the NA and properties of the materials.

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

163

Therefore, in the following section we give a brief review about the experimental conditions and
parameters. This will be helpful for understanding the deterministic nature of the femtosecond
lasermatter interaction, and for tailoring the local microstructure and functionalities of the target
materials by properly adjusting those parameters [5,13,45].
(1) Polarization
As the energy absorbed is polarization-dependent, changes in the structures and properties
induced by femtosecond laser irradiation are strongly dependent on the polarization of the incident
beam [19,45,55,56]. Little et al. suggest that the polarization dependence of photoionization crosssections is responsible for this habit [57]. Therefore, the breakdown threshold is usually significantly
different between the linearly and circularly polarized light [58]. The linearly polarized light is much
more efficient, especially for fourth- and higher-order photon absorption [49,59,60]. In fused silica and
sapphire, polarization-sensitive 6-photon ionization is reported to be the dominant ionization mechanism and the cross-sections of 6-photon ionization for linearly polarized light are significantly larger
than that for circular one [61]. Liu et al. show that the damage threshold for circularly polarized light is
higher than that for linearly polarized one when NA > 0.4, but the former is lower than the latter when
NA < 0.4 [58]. Femtosecond laser induced damage at high NA and the self-focusing induced breakdown
at low NA are proposed to be responsible for this reverse processes. Nanostructures, such as nanogratings and nanocracks, written directly in fused silica by using the femtosecond laser show polarizationdependence [19,55]. These nanogratings are self-organized and periodic with the size and period of 20
and 140 nm, respectively, and orientated in a perpendicular direction to the electric field vector of a
linearly polarized femtosecond laser beam. Shimotsuma et al. suggest that the interference between
the incident light field and the electric field of the bulk electron plasma wave results in a periodic
modulation of electron plasma concentration and permanent structural modifications in the glass
[19]. The evolution of nanoplasmas into disk shaped structures due to high non-linear ionization is
also proposed to generate these nanostructures [55]. Waveguides fabricated in fused silica with circularly polarized light exhibit a greater concentration of 3 member SiO ring structures, resulting in densification and a higher refractive index than those written with linearly polarized light [57]. Ams et al.
demonstrate that the refractive index contrast and propagation losses of the direct written waveguides are dependent on the laser beam polarization [62]. Their results indicate that the circularly
polarized femtosecond beam is more suitable for fabricating curved waveguides with a 1 kHz laser
system. The self-organized and polarization dependent nanogratings have also been fabricated in
the transparent crystal (SrTiO3) by the authors laboratory. This can be done by applying a tightlyfocused linearly polarized femtosecond laser beam to form a single line. This polarized-dependent
behavior is suggested to originate from difference of the polarization-induced bulk damage threshold
[63]. Hnatovsky et al. demonstrate the polarization-sensitive micro-optic structure modification with
the presence of the longitudinal electric field of femtosecond pulses [64]. The local polarization can be
imprinted in the focal region (NA = 0.46), as shown in Fig. 5. The patterns in Fig. 5A and B are created
with pulse energy near the threshold, whereas the significant ablation craters in Fig. 5C and D are generated by increasing the pulse energy. The morphology of the generated craters provides implications
for the 3D structure of the nanocrack patterns, and hence the effect of beam polarization. Obviously,
visible radial [TM, Fig. 5C] and azimuthal [TE, Fig. 5D] nanocrack patterns reflecting the TM and TE
character of the field, respectively, are present. When embedded in bulk material, these polarization
imprints can be easily determined and analyzed in cross-polarized light as shown in Fig. 5E and F
owing to the strong birefringence of the nanopatterns.
(2) Wavelength
The damage threshold intensity, the resultant structures and the properties are strongly
wavelength-dependent [13,6567]. Jia et al. show that the threshold fluence for the visible lasers linearly varies with the wavelength. The threshold fluence becomes nearly a constant at 8002000 nm
[67]. The magnitude of the induced refractive change (Dn) is similar (16  104) for both the fused
silica and borosilicate glass after irradiated by 800 nm pulses. The 400 nm exposure for fused silica

164

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

Fig. 5. Polarization-sensitive micro-optic structures produced inside fused silica substrates with radially (TM) and azimuthally
(TE) polarized femtosecond pulses. (A) and (B) Imprints produced with a focusing objective with an effective NA of about 0.3
after irradiation with multiple 300 nJ pulses. (C) and (D) The same conditions as in (A), (B) except for the pulse energy of 500 nJ.
(E): Optical signature of the two-dimensional array of imprints depicted in (F) as seen in cross-polarized light. The scale bars
shown in the left column apply to the corresponding images in the right column. (Reproduced with permission from [64].)

generates a decrease in refractive index of the order of 5  104, whereas an increase is demonstrated
in the borosilicate glass, similar to the result for fused silica at 800 nm [66]. Shah et al. report that it is
not possible to produce low-loss waveguides using the fundamental wavelength of 1045 nm, while
high quality waveguides with propagation losses below 1 dB/cm at 1550 nm can be produced with
115 nJ/pulse at 1 MHz and 522 nm [65]. Two different types of self-organized, sub-wavelength periodic structures are fabricated in fused silica by a tightly focused, linearly polarized femtosecond laser
beam. The main one with period (KE) is in the direction of the irradiated light polarization and proportional to the wavelength of the writing laser. The second with period (Kk) is in the direction of
the light propagation. In the head of the modified region, the period is approximately comparable
to the wavelength of light [68]. According to Dostovalov et al., the second harmonic (515 nm) is more
efficient than the fundamental harmonic (1030 nm) in terms of amount of absorbed energy, resulting
in a lower inscription threshold. Hence the former harmonic may be more attractive for applications in
femtosecond laser microfabrication [42]. Furthermore, the radius of the beam waist (x0) is dependent
on the laser light wavelength (k), the beam quality factor (M2), the numerical aperture (NA):

x0 2k=M2  p  NA

2:4

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

165

where M2 is a measureable quantity to characterize real mixed-mode beams) [69]. As a result, the
wavelength may influence the light intensity distribution in the focal volume, and change the focusing
strength and the interaction regions and shapes.
The threshold intensity (Ith) for the material damage is determined mainly by three experimental
parameters: the laser pulse duration, s, the pulse energy, E, and NA. The effects of s, E and NA on the
light intensity I at a given wavelength, k, can be described as [5]:

I / E  NA2 =sk2 1  NA2 

2:5

Although the dependence of Ith for micromachining on these parameters deviates from the
expected behavior, the above-mentioned 3 parameters are important for determining the machining
processes and for creating specific structures.
(3) Pulse energy and energy intensity
Femtosecond laser pulses focused in the transparent materials are absorbed through nonlinear
photoionization mechanisms, generating a permanent structural modification in the focal volume.
The minimum E needed for the nonlinear absorption that seeds electrons refers to the threshold
energy. When E is kept close to this threshold, the light absorption causes a change in the index of
refraction within the focal volume. The magnitude of the refractive index change is different from
one type of material to another and both positive and negative index changes are reported. With
low pulse energy, the structural modification in many glasses is accompanied by a smooth change
in refractive index. Under intermediate energy pulse, the birefringent modification occurs. Under high
energy pulse, ultrahigh pressures within the focal volume result in microexplosions and consequently
generating empty voids [70]. Furthermore, as the refractive index is nonlinear and dependent on the
incident energy intensity:

n n0 n2 I

2:6

n is the total refractive index, n0 is the ordinary refractive index and n2 is the nonlinear index [13,71].
Increasing the beam energy can displace the focus and thereby the interaction domain is stretched
prior to the geometrically focusing, and this causes self-focusing. The refractive index variation
depends on the energy, whereas the strength of the self-focusing lens depends only on the peak power
of the pulse. The threshold energy for self-focusing is about one order of magnitude higher than the
damage threshold. Above a critical power (Pcr),

Pcr

3:77k2
8pn0 n2

2:7

due to a balance between the nonlinear increase of the refractive index and the defocusing effect of the
electron plasma, a non-diverging trace beyond the geometrical focus appears [13,72]. In addition,
according to Eq. (2.6), the temporal variation of the laser intensity involves a temporal variation of
the refraction index, which results in the generation of new frequencies in the spectrum of the laser
pulse. This effect is called self-phase modulation [72]. The influences of pulse energy will be further
discussed in the next section.
(4) Pulse duration
The dependence of damage threshold (Ith) on pulse duration (s) has been investigated in a large
range of s (down to 10 fs). But experimental data does not always obey the expected inverse relation
between I and s given by Eq. (2.5). Different research groups even observed different changing trends
of damage threshold with s [6,50,58,7375]. Some groups report that the damage threshold increase
with increasing s in the femtosecond regime, as revealed in Fig. 6A [6,58,74]. In contrast, other groups
report an inverse relation [50,73], which is confirmed by theoretical and experimental studies [67,75],
as revealed in Fig. 6B. In addition, Gawelda et al. demonstrate that extending the pulse duration
improves the spatial distribution of deposited energy by minimizing beam filamentation and prefocal
depletion effects in doped phosphate glasses [56]. More experimental work is needed to uncover the
origin of the differences observed for Ith  s relation.

166

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

Fig. 6. Damage threshold of fused silica vs pulse duration (A) and (B). (C) Dependence of the threshold energy for
micromachining on the NA of the focusing objective for 100 femtosecond pulses in Corning 0211. (Reproduced with permission
from [38,50,58].)

(5) Numerical aperture


As discussed above, NA is one of the parameters determining the optical breakdown of the
transparent materials and the beam waist. For diffraction-limited focusing in the presence of weak
self-focusing, the energy required to reach the breakdown intensity related to the NA is described
by Eq. (2.8) [38]:

Eth

Ith
2

sk 2

pNA Ith k2 =Pcr

2:8

where Pcr is the critical power for self-focusing in the material. Therefore, the damage threshold peak
power decreases with an increase of NA, as indicated in Fig. 6C [13,38,58]. Ashcom et al. suggest that
competing nonlinear optical effects are involved in the interaction of femtosecond laser pulses with
transparent materials: self-focusing, supercontinuum generation and multiphoton-induced bulk damage [76]. At low NA (<0.65), self-focusing and supercontinuum generation play an important role in
the energy deposition before the damage takes place, and this is confirmed by optical and ultrasonic
signatures [7678]. At high NA, multiphoton-induced bulk damage predominates the femtosecond
microfabrication process. Furthermore, NA determines the width of the focal volume and then the
resultant feature size. With the NA above 0.6, the induced structures are almost spherically symmetric,
but below this value, the micromachined structures become larger and asymmetric [5].
(6) Repetition rate
Two different material modification regimes can be distinguished depending on whether the duration between the subsequent pulses is longer or shorter than the time scale for heat to diffuse away
from the focal volume. One is the low-frequency regime, in which material modification is generated
by the single pulse, and the other is the high-frequency regime, in which cumulative thermal effects
determine the resultant structures. A variable femtosecond laser repetition rates are applied to
uncover the influence of repetition rate on the properties of the modified structures and the role of
thermal diffusion and heat accumulation effects in forming low-loss optical waveguides in different
glass across a broad range of laser exposure conditions [7982]. At low to moderate repetition rates
(1200 kHz), as thermal diffusion extends from the heated region far outside the focal volume, an
increase in pulse energy results in formation of larger modification structures in the borosilicate glass
[81]. As the repetition rate increases (0.52 MHz), the time between subsequent laser pulses becomes
shorter than the time for the absorbed laser radiation to diffuse out of the focal volume and heat builds
up around the focal volume. Based on the observation of waveguide morphology and thermal modeling discussions, Eaton et al. indicate that strong thermal diffusion effects give way to a weak heat
accumulation effect at 200 kHz repetition rate and 1 lJ pulse energy for generating low loss waveguides in fluoride glasses, while stronger heat accumulation effects above 1 MHz repetition rate offer
possibility to fabricating waveguides exhibiting overall superior guiding performance. The insertion
loss decreases with increasing repetition rate, which is associated with increased heat accumulation
effects from 0.2 to 1.5 MHz that result in stronger refractive index change and smaller mode-field

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

167

diameter for best coupling to optical fibers at 1.5 MHz. Brub et al. report that the refractive index
change induced by the pulse filamentation at repetition rates lower than 50 kHz and low pulse energies is negative at the irradiated volume. At repetition rates above 50 kHz, the refractive index modification process is dominated by a heat accumulation effect which induces glass melting. The
diameters of the refractive index modified regions are functions of repetition rate and translation
speed for various pulse energies, as depicted in Fig. 7 [82]. Contrary to the investigations of Eaton
et al., Osellame et al. reveal that there are no significant differences in the insertion loss performances
of the best waveguides in an erbiumytterbium codoped phosphate glass fabricated by a compact
diode-pumped cavity-dumped Yb:glass laser oscillator at the repetition rate of 505, 685, and
885 kHz, while the quality of all the waveguides becomes significantly lower at frequencies above
1 MHz. They propose that the waveguide uniformity rapidly spoils assigned to the thermal effects
at repetition rates above 1 MHz. This phenomenon becomes obvious in the phosphate glass at repetition rates of 26 MHz, and the nonuniformity prevents from any waveguiding [80]. Different relationships between the repetition rates and the resultant structures, and the different threshold pulse
energy and repetition rate for onset of heat accumulation may originate from different femtosecond
laser systems and different types of glass system [7982].
Other parameters such as the scan speed, the irradiation pulse number, the pulse geometric characteristics, and beam waists also influence the interaction of femtosecond laser with transparent
materials, and hence, the induced structures and their properties, which will be described below
[16,45,70,74,79,8184].
2.4. Regimes of femtosecond laser dielectric modification
Relying on the exposure parameters, material properties and the material transformation stages,
physical and chemical modifications inside various transparent materials induced by different regimes
of femtosecond laser have been identified by comparing the light intensity at the focus [45,70,74,85].
Low intensity in a narrow processing window induces soft positive isotropic refractive index changes
involving changes of electronic configuration and polarizability (type I) [8688]; intermediate
intensity results in birefringent anisotropic zones (type II) [8991]; and high intensity generates voids
in glass (type III) [9294]. Isotropic and anisotropic changes determine specific optical signatures
ranging from optical guiding to polarization sensitivity. Type I modifications represent soft material
transformation, below catastrophic breakdown, with refractive index contrasts in the range of 104
to 103, as shown in Fig. 8A. Low subcritical incident energy intensity, characteristic of low fluences
or loose focusing, usually leads to type I modification. In this case, the relaxation of the electronic
excitation cannot induce high temperature for the local heating, since this temperature is below the
material softening threshold. Under these conditions, the thermomechanical effects seem negligible
and the laser-generated defects determine structural transitions corresponding to a denser packing
in the matrix and resultant refractive index changes [95,96]. The birefringence of type II modification
induced by focused femtosecond laser radiation is attributed either to laser-induced stress, or to the
formation of self-assembled nanogratings with subwavelength periodicity orientated perpendicularly

Fig. 7. Diameter of the refractive index modified outer region for different incident pulse energies as a function of (A) repetition
rate at a translation speed of 50 lm/s and (B) translation speed at a repetition rate of 250 kHz. (Reproduced with permission
from [82].)

168

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

Fig. 8. (A) Phase contrast microscopy images of type I modification traces in static (single and multishot) and longitudinally
scanned conditions; dark and white are positive and negative index changes respectively. (B, C) Phase contrast microscopy
images of type II modification traces in static and longitudinally scanned conditions with nonguiding (NG) and guiding (WG)
properties. Scanning electron microscopy (SEM) of the cross-section of the traces showing the nanoscale arrangement is given
along with corresponding guided modes at 800 nm. Isotropic mode transport for type I traces and polarization type II guiding
for electric field parallel to the nanoplanes are shown in the right. (Reproduced with permission from [88].)

to the laser light polarization leading to anisotropic reflection, as shown in Fig. 8B and C [96]. The type
II modification is reinforced by high incident energy intensity and slightly longer pulse durations,
resulting from a stronger energy confinement assisted by weaker plasma defocusing as compared
to the shorter pulse case. The thermomechanical effects related to pressure waves, compaction,
cavitation, and rarefaction, with signs of low viscosity regions is present in type II traces, leading to
the formation and movement of voids, void agglomeration and reshaping as nanogratings [88,96]. This
facilitates dimensional increase of voids, the seed of hydrodynamic nanoscale rearrangement of
matter. Sub-nanosecond characteristic times are determined for void formation via mechanical
rarefaction with local temperatures in the range of the softening values [97]. The high energy intensity
induced transformations are particularly visible from a dynamic perspective, which are confirmed by
the time-resolved optical transmission imaging and post-mortem spectroscopy [96].
Both type I and type II regimes can be accessed with a single pulse [98,99], implying that the
induced temperature increase is not critical to the resulting structural modifications. However, in
order to induce the type III modifications, several pulses incident within a short period of time
(lss) are necessary so that memory effects take place during the interaction time [100]. Type I has
been used in waveguides and couplers, type II has been used in polarization converters, waveplates,
etc., and the type III finds application in data storage and photonic crystals [74,94,101104]. This
review mainly focuses on type II and type III modifications, since they have not been much explored
till now.

2.5. Femtosecond laser induced various phenomena and applications


2.5.1. Emission
Femtosecond laser induced luminescence is a phenomenon promising many applications. The
major femtosecond laser induced luminescence phenomena observed in transparent solids are as follows: multiphoton excited upconversion luminescence, luminescence related to the valence state
change of active ions induced by femtosecond laser, long-lasting phosphorescence, unusual polarization dependent luminescence, luminescence caused by energy transfer between SHG microcrystals
and rare earth ions.
(1) Multiphoton upconversion luminescence
There are several important mechanisms causing upconversion luminescence, including energy
transfer upconversion (ETU), photon avalanche, MPA, excited state absorption (ESA) or cooperative
upconversion. Compared to the upconversion luminescence based on the other mechanisms, the
MPA upconversion luminescence possesses some important advantages in applications [105,106]. In

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

169

the MPA case, the electrons in the ground state absorb two or more pump photons simultaneously and
get directly transferred to the excited states. On the other hand, the other mechanisms need the help
of interbands, and hence are comparatively more complicated stepwise processes. The relationship
between intensity (I) of fluorescence induced by MPA and the power of incident light (P) can be
described as: I / pn, where n is the absorbed photon number. The value of n can be determined
through the measurement of the relationship between fluorescence intensity and the incident light
power. For the excitation of MPA luminescence, very high energy intensity is required, and it is worth
to note that pulsed lasers are efficient tools to induce such phenomenon. Most upconversion luminescence induced by MPA is pumped by nanosecond laser or picosecond laser in the past. However, they
can hardly generate upconversion luminescence effectively in solid inorganic materials due to weak
nonlinear effects. Due to the higher peak power and shorter pulse duration, hence, by employing femtosecond laser pulses, it is very easy to achieve efficient MPA luminescence. As early as 1999 and 2000,
MPA (three photon absorption, 3PA) upconversion luminescence excited by infrared femtosecond
laser was reported in Ge-doped SiO2 glass by Kazansky et al. [107], and in rare earth doped glass by
Qiu et al. [108]. The 3PA blue emission in Ge-doped SiO2 glass originates from GeO deficient centers.
An anomalous anisotropic light scattering phenomenon in Ge-doped silica glass is observed and
assigned to anisotropic index fluctuations excited by electrons moving along the direction of the light
polarization in the process of photoionization by intense femtosecond laser light at first. These fluctuations scatter strongly in the plane of the light polarization for short wavelength light (similar to
Rayleigh scattering), resulting in scattering peak, particularly the luminescence in the plane of light
polarization and the anisotropic pattern of luminescence [107]. An anisotropic, permanent and blue
luminescent pattern is seen and is modified by another polarized beam in rare earth doped glass
[108]. The blue luminescence arises from 3PA and subsequent relaxation from the 5d level to the
8S7/2 ground state of the Eu2+ ions. In this case, the femtosecond laser not only induces electron
and hole trapping centers in the glass via MPA and multiphoton ionization, but also acts as a driving
force for inducing the distribution of induced defects.
When the driving force in the light polarization direction is larger than that in the perpendicular
direction, the permanent refractive index fluctuations in the light polarization direction will be larger
than those in another direction. Furthermore, Rayleigh scattering in the direction of the light polarization is proportional to the density or refractive index fluctuations. As a result, Qiu et al. suggest that
anisotropic luminescence phenomenon originates from the light scattering of the polarizationdependent permanent microstructure induced by the polarized ultra-short pulsed laser itself. It
should be pointed out that these phenomena are closely related to the formation of polarizationdependent nanogratings which we will discuss further later on. Ever since the authors pioneer work,
there have been extensive studies on femtosecond laser induced multiphoton excited upconversion
luminescence in various transparent materials, including glass [109116], crystals [117123] and
glassceramics [124126]. The authors group observes much stronger near-infrared to visible red
upconversion luminescences induced by near-infrared femtosecond laser in transparent Eu3+-doped
SrOTiO2SiO2 and Sm3+-doped BaOTiO2SiO2 glass ceramics than in the as-prepared glass. X-ray
diffraction and Raman analyses indicate that Sr2TiSi2O8 and Ba2TiSi2O8 microcrystalline particles with
second-order optical nonlinearity are precipitated in the corresponding glass ceramics after heat treatment [124,125]. Qiu et al. have attributed the enhanced emissions to the enhanced absorption of the
second harmonic as a result of precipitation of microcrystalline particles. The intensity of the upconversion luminescence is proportional to the square of the excitation power. The damage threshold of
the glass ceramics also increases greatly compared to the pristine glass.
Besides the defects formed in the glass [107], excitation of the charge transfer state of dopants (e.g.,
Ce3+, Eu2+, Ta5+, Nb5+, Tm3+, Cr3+, Eu3+, Eu2+/Dy3+, Eu3+/Tb3+, Pr3+, and Sm3+) [108126] can also result
in upconversion luminescence, which facilitate the upconversion from near-infrared to visible region.
MPA upconversion emission can also be observed from intrinsic luminescent complex in the crystal
[118]. Ryba-Romanowski et al. report that the intensity of MPA upconversion luminescence related
to the 4S3/24I15/2 transition of Er3+ in yttrium and lutetium vanadate is not dependent on wavelength
of excited femtosecond pulses [122]. The luminescence intensity and relaxation dynamics of vanadate
group is a function of the sample temperature, revealing that a nonradiative energy transfer from
vanadate groups to erbium ions exists. Petit et al. report on the two-photon excited fluorescence in

170

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

the LiY(BO3)3:Eu3+ (LYB:Eu) monoclinic crystal, under excitation of a tightly focused femtosecond laser
beam with the wavelength of 800 nm. They demonstrate the creation of two separate nonlinear voxels
associated with the two pump polarization eigenmodes, which result in light emission in the two fluorescence polarization eigenmodes along the epi-collected signa, as shown in Fig. 9 [123]. Spatial
walk-off propagation gives rise to the spatial discrimination of each of the four polarization schemes,
for both the pump and fluorescence beams, in such biaxial crystal oriented along the crystallographic
c-axis. The third-order nonlinear behavior of the two-photon absorption process is demonstrated with
the quadratic evolution of the fluorescence emission versus the pump irradiance, indicating anisotropic polarization dependence. Strong polarization dependence is revealed, where the extraordinary
pump polarization provides twice more fluorescence excitation than its ordinary counterpart under
the same experimental conditions, indicating the subsequent anisotropy of laser-induced modification
thresholds for higher incident irradiances.
The MPA upconversion luminescence can be employed to realize 3D, solid state, and color display.
However, to do so, more research work needs to be done, e.g., by properly choosing doping species and
engineering microstructure of glass.
(2) Long-lasting phosphorescence
Luminescence can be classified into two categories according to the duration of the luminescence:
fluorescence and phosphorescence. The duration of fluorescence is usually shorter than 103 s, while
the duration of phosphorescence is longer than 103 s. Depending on the decay time of phosphorescence, phosphorescence can be divided into two groups: short-lasting phosphorescence and longlasting phosphorescence.
Qiu et al. have observed long-lasting phosphorescence induced by femtosecond laser irradiation in
various types of glasses and crystals [127132]. After irradiation by a 800 nm femtosecond laser, the
irradiated area of the Ce3+, Tb3+, or Pr3+-doped glasses emits bright and long-lasting blue, green, or red

Fig. 9. Pictures of the two-photon excited epi-fluorescence in polarized light for both the pump laser beam and the collected
beam. Polarizations o and e stand for the ordinary and extraordinary modes, respectively, for both the pump and the epifluorescence beams. (A), (B), (C) and (D): Polarization schemes with oo, oe, ee and eo polarization modes for the pump and the
epi-fluorescence beams, respectively. The cross is at the same place in each picture, and is present for visual help. (Reproduced
with permission from [123].)

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

171

phosphorescence, respectively, which is clearly seen with the naked eyes in the dark even 1 h after the
removal of the femtosecond laser, as displayed in Fig. 10 [127]. The intensity of the phosphorescence
decreases with time. Qiu et al. have proposed that after femtosecond laser irradiation, a part of Ce3+,
Tb3+, and Pr3+ ions are oxidized to Ln4+ (Ln = Ce, Tb or Pr). Ln3+ ions can act as hole trapping centers and
Ln4+ ions as electron trapping centers. After the irradiation by the focused femtosecond laser, free electrons and holes are simultaneously produced in the glass matrix. The holes or electrons are trapped by
defect centers, released by heat at room temperature, and recombine with electrons or holes trapped
by other defect centers. The released energy resulting from the recombination of holes and electrons is
transferred to the rare earth ions and excites the electrons at the ground state to an excited state of the
rare earth ions, finally leading to the characteristic emissions of rare earth ions. Similar mechanism
also attributes blue phosphorescence in the Eu2+-doped aluminosilicate glasses (43CaO:13Al2O3:44SiO2:0.05Eu2O3:0.05Nd2O3 and 43CaO:13Al2O3:44SiO2:0.05Eu2O3) after femtosecond laser irradiation
[128]. Long-lasting phosphorescence has also been reported in the oxygen-deficient Ge-doped silica
glasses, Mn2+-doped alumino-phosphofluoride glass and Ce3+-doped Ca2Al2SiO7 crystal, stemming
from the thermally activated electronhole recombination [130132].
The authors results suggest that it is possible to selectively induce different kinds of defects by
using femtosecond laser irradiation with tunable wavelength, pulse duration, and repetition rate.
The recombination of holes and electrons in various defects may lead to the presence of phosphorescence with various colors in a glass. It is also possible to induce defects that are stable at room temperature by using an femtosecond laser and that are able to be released by another laser. This will be
of importance in the fabrication of rewritable 3D optical memory micro-devices.
(3) Supercontinuum generation
The propagation of intense femtosecond laser light through the bulk transparent materials is
accompanied by the significant modification of its spatial and temporal properties, which results in
extreme spectral broadening, leading to the generation of supercontinuum from ultraviolet (UV) to
mid-infrared (MIR) range [71,72,133,134]. The self-modifications of the pulse shape and spectral
broadening are usually assigned to the strong nonlinear optical interaction of the light field with
the media, and this happens under the conditions of high radiation localization both in space and time.
As suggested above, since the power density is larger than 1013 W/cm2, the nonlinear interaction largely contributes to the refractive index increases of the transparent materials during the femtosecond

Fig. 10. (A) Photograph of the emission states of phosphorescence in glass samples 5 min after the removal of the exciting laser.
The Ce3+, Tb3+, and Pr3+-doped calcium aluminosilicate glass sample shows blue, green and red light emission, respectively. (B)
Photoluminescence, phosphorescence, and excitation spectra of a Tb3+-doped calcium aluminosilicate glass. For the
measurement of the excitation spectrum, the luminescence at 437 nm is monitored. (a) and (b) are phosphorescence spectra
of the glass 350 and 1000 s after the laser irradiation, respectively. (Reproduced with permission from [127].)

172

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

laser irradiation. As a result, self-focusing of the femtosecond laser beam occurs. On the other hand,
the formation of electron plasma due to the high electric field induces a decrease in the real part of
the refractive index and causes self-defocusing of the beam. The balance between the self-focusing
due to the increase of refractive index and self-defocusing due to the plasma formation results in a
phenomenon called self-trapping or filamentation [71]. Consequently, white light supercontinuum
containing Stokes and anti-Stokes waves is generated. Other mechanisms, such as four-wave mixing
[71], pulse splitting [135], self-steepening and generation of optical shocks [136,137], stimulated
Raman scattering [138], are also proposed to be the possible reasons for supercontinuum generation
[139]. It is also reported that the power threshold for supercontinuum generation is consistent with
the calculated critical power for self-focusing, confirming that self-focusing is a dominant initiator
of the sequence of processes generating supercontinuum. [71,133].
Srinivas et al. report supercontinuum generation in a quadratic nonlinear medium (potassium
dihydrogen phosphate, KDP) crystal by using femtosecond laser (100 femtosecond pulses at
790 nm) irradiation [133,140]. An enhanced supercontinuum with the large bandwidth of 385
960 nm is produced by adjusting the phase matching angle of the KDP crystal. For angles away from
the phase matching direction for SHG, an increase in the intensity of the supercontinuum is observed
with reducing the SHG intensity. A conversion efficiency of about 23% over the entire spectral range is
realized when an input energy of about 100 lJ is focused at the center of the KDP crystal. Both Youngs
double slit experiment and Michelson interferometer experiments are further adopted to confirm the
coherent nature of the white light [140]. They also suggest that KDP could be effectively adopted for
generation of sum frequency signals and supercontinuum to produce broadband light with a wavelength range from 350 to 1300 nm. Therefore, the enhancement in the bandwidth of the supercontinuum light toward the shorter wavelength regime (<400 nm) is also achieved. The tunability in the blue
region of the supercontinuum spectrum with angle is demonstrated [133]. Furthermore, the control
over the depolarization properties of supercontinuum generation in a KDP crystal is realized through
changing the plane of polarization of incident light and the orientation of the crystal with respect to
the incident light [141143]. Bradler et al. report a comprehensive investigation of supercontinuum
generation in several single crystals with femtosecond laser irradiation, including yttrium aluminum
garnet (YAG), yttrium vanadate (YVO4), gadolinium vanadate (GdVO4), and potassium-gadolinium
tungstate (KGW). Plateau-like visible and IR spectra with higher IR photon flux are found, as shown
in Fig. 11A and B [144]. Several parameters, like absolute spectral energy density, pulse-to-pulse stability, pump threshold, and beam profile are studied in terms of their dependences on the focusing
conditions, crystal thickness, pump pulse energy, and pump wavelength (7751600 nm). The particular advantages of the above-mentioned materials for use in parametric amplification, femtosecond
spectroscopy, and carrier-envelope phase stabilization are discussed. The effect of numerical aperture
on supercontinuum generation, as well as the supercontinuum generation by femtosecond Gaussian
and Bessel beams has been studied [76,145].
Supercontinuum generation in the MIR range from 2.5 to beyond 10 lm has also been realized in a
bulk crystal and glass recently and increasingly becomes a focus for research [134,146,147]. Liao et al.
report that supercontinuum generation by filamentation can cover from visible to 6 lm in tellurite
glass with the stable conversion efficiency as high as 87% [148]. Silva et al. demonstrate the stable
multi-octave supercontinuum from filamentation of MIR femtosecond pulses in yttrium aluminum
garnet (YAG) crystal (Fig. 11C) [134]. A spectrum spanning 4504500 nm is observed, corresponding
to 3.3 octaves, with a spectral energy density of 2 pJ nm110 nJ nm1, as shown in Fig. 11D, indicating a simple and general method for coherently extending the spectrum of an amplified femtosecond
pulse to an ultra-broad range.
Due to its unique characteristics, the supercontinuum promises an ideal broadband ultrafast light
source for various applications in the field of femtosecond time-resolved spectroscopy, optical pulse
compression for ultra-short pulse generation, optical parametric amplifiers, biomedical applications
and structure modifications [71,139,146,147].
2.5.2. Formation of color centers
A color center is a point lattice defect consisting of a vacant negative ion site and an electron bound
to the site, which can be created in a variety of solids at room temperature by high energy irradiation

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

173

Fig. 11. (A) Black: supercontinuum from 3 mm sapphire with conventional pumping conditions. Green: 3 mm sapphire with
the optimizations described in the text. Blue: 4 mm YAG with improved photon density in the infrared region. (B) Continua
generated in KGW (green), YVO4 (red) and GdVO4 (blue) crystals (all 4 mm). (C) Set-up for supercontinuum generation and
measurement of its angularly resolved spectrum. (D) Supercontinuum generated by 3100 nm, 2.6 lJ pulses in YAG. (Reproduced
with permission from [134,144].)

with ionic beams, x- or c-rays or femtosecond laser pulses [149]. The color centers can selectively
absorb light and make certain transparent materials colored. As mentioned previously, the extremely
high power density allows for strong nonlinear interactions between femtosecond laser pulses and the
transparent materials, which can be used to color various glass and transparent crystals [149154].
Efimov et al. report that short-wavelength component of the supercontinuum is generated in the silicate glass after exposure to the femtosecond laser irradiation (850 nm), which causes photoionization
of silicate glass and finally color center formation [149]. Lonzaga et al. propose a different mechanism
for the color center formation in soda-lime glasses that the coloration can originate from the absorption processes that produce mobile charge carriers [150]. These charged carriers interact to produce
trapped hole centers (H3+) that strongly absorb light at 633 nm. They show that the dependence of
color center formation on pulse energy is extreme and about 15th order, owing to the strong dependence of defect production on the density of excitons. The competition between coloration induced by
the femtosecond pulse irradiation and the subsequent transmission recovery limits the degree of coloration, which can be modulated by controlling adopted pulse energy intensity and repetition rate.
Furthermore, the response of soda-lime glass to ultrafast pulses with the wavelengths of 400 and
267 nm are similar with much lower threshold for darkening. Diffraction gratings can be rapidly
and easily produced in soda-lime glass according to the color center patterns, and also demonstrated
in calcium fluoride crystals (CaF2) by Qius group (Fig. 12A), in lithium fluoride (LiF) by Courrol et al.,
and phosphate glasses by Dekker et al. [155157]. Zhao et al. suggest that free electrons are generated
by the 8 photon absorption of the CaF2 crystal and consequent avalanche ionization when the sample
is irradiated by femtosecond laser with power density of 5.0  1015 W/cm2. The adjacent F vacancies
trap free electrons to form color centers. In the meantime, the avalanche ionization produces high
absorptive and dense plasma and then induces the local melting and material misplacement, leading
to permanent structural changes. Fig. 12B indicates that controllable refractive index change is achievable by changing femtosecond laser irradiation parameters and subsequent annealing temperature
[155]. The absorption and emission spectra indicate that F, F2, F+2 and F+3 color centers (Fig. 12C) are

174

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

Fig. 12. Internal diffraction grating structures (A) and diffraction efficiency and refractive index changes as a function of
annealing temperature (B) for femtosecond laser irradiated CaF2 crystals sample. (C) Absorption spectra of the tracks created in
LiF crystals by 750 lJ, 60 fs laser pulses. (Reproduced with permission from [155,156].)

created through the multiphoton ionization in the LiF and LiYF4 crystals during the femtosecond laser
irradiation. This is confirmed by Pan et al. [156,158,159]. After the irradiation, an electron can be captured by the negative fluorine ions, forming an F center. The other types of color centers are formed by
the aggregation of F centers. Dickinson et al. probe the evolution of color centers in soda lime glass and
single crystal sodium chloride generated by femtosecond laser irradiation on different time scales,
from microseconds to hundreds of seconds. They suggest that the decay of color centers can be well
described in terms of bimolecular annihilation reactions between electron and hole centers in both
samples [154].
Color centers are also produced in Tb3+-doped and Tb3+/Ce3+-codoped heavy germanate glasses
after femtosecond laser irradiation by the authors group [151]. It is recognized that the irradiation
in glasses creates excited electrons, holes and/or bound electronhole pairs (excitons) from trapping,
which leads to formation of color centers. Ce3+ ions are found not only to inhibit formation of color
centers, but also to enhance their recovery. It is reported that formation of color centers resulted from
the moderate local rearrangements of charges inside the material, which is induced by femtosecond
laser irradiation, can be associated with modification of the valence state of active ions. The active ions
can be metallic or rare earth metallic. Therefore, accompanied by the generation of color centers,
valence state manipulation of the doping ions in the glass matrix is also possible, and thereby one
can control the optical properties in a more complex way, which will be discussed next [160162].
2.5.3. Valence state change
Materials with 3D modulated microstructures exhibit widespread potential applications in optical
field. Until now, there have been numerous investigations on the 3D micro-fabrication. We have realized space-selective valence state manipulation of active ions and demonstrated the promising application in 3D optical memory with ultrahigh storage density [160163]. To change the valence state,
the adopted irradiation should be sure to trigger the photo-activate oxidationreduction chemical
reactions through the generation of photoelectrons in the irradiated area following the generic relations: hm + An+ ? A(n+1)+ + e; Bp+ + e ? B(p1)+ [164].
Qiu et al. report the femtosecond laser induced permanent photoreduction of Sm3+ to Sm2+ in the
transparent and colorless Sm3+-doped sodium aluminoborate glass [160]. Upon irradiation by the
focused femtosecond laser with a wavelength of 800 nm, the irradiated area of the glass becomes
orange. The absorption and photoluminescence spectra imply the presence of Sm2+ in the focused
area. As shown in Fig. 13A, new emission peaks at 683, 700, 724, and 760 nm are observed in the emission spectra of the modified glass with these peaks attributed to the 4f4f transitions of Sm2+. Electron
spin resonance (ESR) spectra show that defect centers are created after the laser irradiation. The ESR
signals are assigned to the defect centers of holes trapped by nonbridging oxygen ions and tetrahedral
coordinated boron ions, and electrons trapped by the quasi-F centers. Qiu et al. propose that active
electrons and holes can be created in the glass through multiphoton ionization, Joule heating, and collisional ionization processes. After that, holes are trapped by nonbridging oxygen ions as well as by
tetrahedral coordinated boron atoms, while a part of the electrons may be trapped by the Sm3+,

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

175

Fig. 13. (A) Emission spectra of Sm3+-doped sodium aluminoborate glass before (a) and after (b) femtosecond laser irradiation
(excited by Ar+ laser at 514.5 nm). (B) Photoluminescence images of alphabetical characters recorded on different layers, which
are observed by using a 403 objective lens and the 680 nm emission from Sm2+ with confocal detection implemented (excitated
at 488 nm, 1 mW Ar+ laser). (C) Signal readout by detecting the fluorescence for photoreduction bits with a 200 nm diam. (D)
Example of erasing and rewriting by irradiation with an Ar+ laser (5 mW at 514.5 nm) and a femtosecond laser. (a)
Photoluminescence image before the erasure. (b) Image after Ar+ laser irradiation to photoreduction bit I. (c) Image after Ar+
laser irradiation to bit II. (d) Image after femtosecond laser irradiation to areas I and II. (Reproduced with permission from
[160,163].)

leading to formation of Sm2+. Qiu et al. find that the valence state change is more stable than the
refractive index change in Sm3+ doped fluoroaluminate glass [165].
Qiu et al. also report the photoinduced space-selective reduction of Eu3+ in the fluorozirconate glass
and oxidation of Mn2+ in the Mn and Fe ions codoped silicate glass at room temperature [161,162]. As
suggested above, multiphoton ionization can generate free electrons and holes, subsequently leading
to the reduction of Eu3+ and oxidation of Mn2+. As the induced structure (1.5 mm) is found to be far
longer than that of the Rayleigh length of the focused beam (200 lm), the generation of supercontinuum is also taken into consideration as another mechanism for the femtosecond laser induced photochemical reactions, as discussed previously [149]. The single or two-photon absorption of short
wavelength component of the white supercontinuum light generate photoionization in glass matrix,
leading to the observed photochemical reactions, confirmed by Hiraos group [162,166,167]. The
length of the induced structure is observed to be directly proportional to the square root of the average
power of the laser beam.
The authors group reports that Bi3+ are photoreduced to the lower valence states, Bi2+ and Bi+ in
mesoporous silica glass, after femtosecond laser irradiation. Since there is no absorption for the Bi3
+
-doped mesoporous silica glass in the visible and near-infrared wavelength region and the employed
femtosecond laser acts at a non-resonant wavelength (800 nm), we suggest that photoreduction of Bi3
+
should be a nonlinear process [168]. Compared to the unprocessed area in the Bi3+-doped mesoporous silica glass, the femtosecond laser processed area shows broadband infrared luminescence
with two symmetric bands around 950 and 1235 nm, which is attributed to the 3P1 ? 3P0 electron
transition of unusual Bi+ emission centers.

176

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

The above results indicate that femtosecond laser is a powerful tool for the space-selective valence
state manipulation of various ions inside the transparent materials. Therefore, this technique holds
great potential applications in the fabrication of 3D colored industrial art object, optical memory,
and micro-optical devices, which have already been demonstrated by several groups [163,169,170].
Fig. 13B demonstrates the recording of photoluminescence images of alphabetical characters on
different layers with the monitored emission at 680 nm from Sm2+. Miura et al. suggest that the spacing of 2 lm between alphabetical character planes is sufficient to prevent cross talk in the photoluminescence images [163]. Fig. 13C shows that photoreduction bits with a 200 nm diam could also be read
out clearly by detecting the fluorescence as a signal excited with an Ar+ laser (5 mW at 514.5 nm). The
memory capacity can be as high as 1 Tbit for a glass piece with dimensions of 10  310  31 mm3.
Furthermore, as the Sm2+ can be converted to Sm3+ by photo-oxidation with a CW laser at room
temperature, such as a Ar+ laser or a semiconductor laser, 3D optical memory with rewriting capability
is possible, as displayed in Fig. 13D. In addition, persistent spectral hole burning are observed due to
the photo-oxidation of Sm2+, which may be also useful in the fabrication of optical memory devices
that can store data in both space and wavelength axes, as well as of micro-optical devices
[166,167,171,172].
2.5.4. Precipitation of crystals and nanoparticles
Materials doped with nanoparticles (NPs) and crystals attract much attention, as that can be specifically controlled to exhibit a significant variation in the local structures, refractive index, plasmon
resonance absorption, luminescence and optical nonlinearities [18,164,173,174]. Unfortunately, it is
difficult to control the spatial distribution of NPs in materials using the traditional methods. The
authors work indicates that it is possible to modulate the precipitation of NPs in three dimensions
inside transparent materials by using focused femtosecond laser irradiation.
Miura et al. discover the 3D formation of a SHG crystal, i.e., b-BaB2O4 (BBO) in glass (47.5BaO:5Al2
O3:47.5B2O3) upon irradiation with a nonresonant femtosecond laser [173]. A spherical heated region
is observed with an optical microscope during the focused laser irradiation. It is proposed that the
formation of the spherical domain is caused by the pressure wave and local heating induced by the
focused laser beam. After 10 min irradiation, some crystals are generated near the focal point.
The X-ray diffraction (XRD) pattern and the presence of second harmonic (a blue beam) of the
femtosecond laser confirm the formation of frequency-conversion crystals. The final size of the
modified domain and its spatial distribution can be easily controlled in three dimensions by adjusting
the irradiation conditions such as depth, irradiation time and scanning speed. Low scanning speed
leads to formation of a stable structure with an accommodation layer at the solidliquid interface
between a part of the polycrystal region and the heated zone, indicating that a single crystal or a
crystal with a single-crystal-like structure can be generated in glass by this technique.
Space-selective precipitation of a single LiNbO3 crystal in Li2ONb2O5SiO2 glass system, BaTiO3
and/or Ba2TiSi2O8 phases in Na2OBaOTiO2SiO2 and BaOTiO2SiO2 glass systems, Sr2TiSi2O8 in
SrOTiO2SiO2 glass are also realized [175178]. Raman spectroscopy shows that the structure of
the glass network is destroyed during irradiation by the femtosecond laser and (B3O6)3+, (NbO6)7+,
and (TiO4)4+ anion units as well as crystals are formed in the focal irradiation area [175]. Electron
probe microanalysis (EPMA) in the vicinity of the irradiated region shows that the chemical composition varies radially from the center to the outside, leading to the ring-shaped crystalline phase. The
laser-induced crystallization implies involving the thermal effect. It is suggested that the spaceselective crystallization inside glasses is determined by the atomic diffusion due to the combination
of the thermal effect and the elemental migration associated with the propagation of shock waves
or pressure waves, and this is confirmed by Dai et al. [176,178]. Dai et al. reveal that the threshold time
for inducing crystal formation decreases, and the corresponding dot size increases with increasing
laser power [178]. Ba2TiSi2O8 and TiO2 crystalline grating patterns are written directly inside the
BaOTiO2SiO2 and CaOAl2O3Bi2O3TiO2B2O3 glasses, respectively [177,179]. A periodic structure
with high refractive index contrast is produced, which is confirmed by polarized photographs of the
induced grating [179].
Ba2TiSi2O8 crystal and CaF2 crystalline patterns are also induced in the Er3+-doped BaOTiO2SiO2
glass and the Er3+-doped oxyfluoride glass, respectively [180,181]. The irradiation time required for

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

177

crystallization in Er3+-doped BaOTiO2SiO2 glass is longer than that in the corresponding parent glass
under the same conditions, and this is attributed to the existence of the upconversion luminescence
from Er3+-ions. The diameter of the femtosecond laser-induced dots in the Er3+-doped glass is a function of the irradiation time. The diameter of the dots varies from several lm to 16 lm and increases
rapidly as a function of t1/3 in the initial phase of the laser irradiation (62 s), and then reaches the final
value. This confirms that thermal transfer plays a predominant role in the formation of the dots [180].
Furthermore, by scanning three focused infrared lasers to the precipitated nonlinear optical crystals in
the glass, blue, green and red emissions near the focal point can be obtained, respectively, implying
that this technique may be useful for the solid state, full color 3D display [180]. Confocal upconversion
luminescence spectra show that the precipitated crystals have greatly enhanced upconversion luminescence intensity compared to unmodified glasses, implying the possibility of 3D optical data storage
in the glass [181]. The signal-to-noise ratio (SNR) can be as high as 29 with 800 nm femtosecond laser
excitation. The readout of the crystallization bits can be performed with 980 nm femtosecond laser
excitation. We show that as the laser fluence inducing the crystallization bits increases, the upconversion intensity increases significantly, whereas the size of the bits increases slightly [181]. Zhong et al.
report that multiple crystalline phases of Dy2(MoO4)3 can be generated in Dy2O3MoO3B2O3 glass
upon femtosecond laser irradiation. Their distributions depend mainly on femtosecond laser induced
temperature field, which is asymmetric along the light propagation direction. They propose that an
inhomogeneous intensity distribution of the incident pulse resulting from the self-focusing effect
and spherical aberration effect is responsible for this phenomenon [182].
Zhou et al. demonstrate that femtosecond laser irradiation is an effective strategy to achieve the
energy-transfer control between different active centers by in situ simultaneous tailoring of the phase
evolution and dopant distribution in the glassy phase, as shown in Fig. 14 [183]. By control of the

Fig. 14. (A) Optical microscope image of the induced structure. Photoluminescence is studied along the direction marked by the
white arrow. (B, E, G) and (C, F, H) Emission distribution, typical true-color optical microscope images, and corresponding
spectra of green and blue emission, respectively. (D) Schematic illustration the origin of the alternative greenblue emission
distribution. (Reproduced with permission from [183].)

178

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

precipitation habit of multiple crystalline phases (Ga2O3 and LaF3), active centers (Er3+ and Ni2+) can
be efficiently isolated by selective partitioning into different crystalline phases. The induced structure
shows remarkable alternative green and blue luminescence, as shown in Fig. 14BH. The spatially
defined emission reveals a sharp contrast, with pure blue in the center (Fig. 14F and H) and green
(Fig. 14E and G) color and at the edge of the induced structure, and this is suggested to originate from
the well-controlled distribution of phases and active centers.
NPs are also precipitated in glass matrix through the processes: femtosecond laser irradiation and
subsequent heat treatment. Qiu et al. have demonstrated the space-selective precipitation and control
of noble metal (Ag and Ag) NPs in glass by using a focused infrared femtosecond pulsed laser irradiation at room temperature and further annealing at high temperatures [18,174,184]. Qiu et al. suggest
that nonbridging oxygen acts as the hole trap center and the Ag+ or Au3+ ion acts as an electrontrapping center during the femtosecond laser irradiation resulting from multiphoton ionization, Joule
heating, and collisional ionization, leading to the reduction of Ag+ or Au3+ ions to Ag or Au atoms.
Besides the fundamental wave, white light supercontinuum has also been proved to play an important
role in the formation of metal atoms. The photon-reduced metal atoms aggregate and grow to form
NPs with the size of several nanometers upon heat treatment. The length of the induced structure
is found to be proportional to the square root of the average power of the adopted laser beam, and this
allows controlling the longitudinal spreading of the structurally modified domain from several hundred nm to several millimeters by optimizing the irradiation conditions [174]. The size of deposited
NPs and their spatial distribution can also be controlled.
Furthermore, the precipitated NPs can be space-selectively dissolved by combination of further
femtosecond laser irradiation and annealing processes [18]. Qiu et al. have studied the effect of laser
irradiation conditions on the precipitation of Ag NPs in silicate glass, and found that the quantity and
space distribution of Ag NPs increase with increasing of the incident light intensity, the beam diameter
in the focal plane, the Rayleigh length of the focusing lens and shot numbers of the laser pulse [185].
However, the average size of Ag NPs is observed to be insensitive to the femtosecond laser irradiation
conditions. Effect of other components in the glass, such as Al2O3, CeO2, and PbO, on the precipitation
of Ag or Au NPs in glasses is also investigated [186188]. We have found that a small amount of these
oxides inhibit generation of color centers, the presence of Al2O3 or CeO2 significantly increase the
annealing temperature for the precipitation of Ag NPs. Whereas, PbO accelerates the formation and
growth of Au NPs during the annealing process [186,187], and hence, improves much the nonlinear
absorption of the metal NPs containing glass [188]. The refractive index of the metal NPs depositing
glass varies with annealing temperature because of the formation of color centers and the presence
of metal NPs [185,189192]. This is contrast to unirradiated ones. If the size of NPs is sufficiently
small, NPs or so-called nanoclusters would be luminescent when excited at different wavelengths.
The luminescence can be abrased by high temperature heat treatment [191,193,194]. Ag NPs and
clusters can also be generated by femtosecond laser irradiation in one step [195198]. Besides Ag
and Au NPs, other metal NPs, including Pd NPs [199], Cu NPs [200,201], silicon NPs [202,203], Na
NPs [204], Pb NPs [205], and Ge NPs [206] are also induced by femtosecond laser irradiation without
heat treatment. A strong nonlinear response of NPs doped glass to nanosecond or femtosecond pulses
is demonstrated [188,199,205207].
Nakashima et al. report a space-selective control over the optical properties, magnetism, and
magneto-optical response of a-Fe2O3 and Al or Au NPs co-doped glass upon an infrared femtosecond
laser irradiation [208,209]. The presence of Al NPs leads to a significant increase in the saturation magnetization of a-Fe2O3 at room temperature. An enhancement of the Faraday effect has also been found
together with a negative peak in the magneto-optical spectra at a wavelength (400 nm), which corresponds to the localized surface plasmon resonance (LSPR) peaks of Al NPs in the optical absorption
spectra, indicating a direct coupling between the ferrimagnetic high Faraday rotation and the LSPR
[208]. However, Au NPs induce a positive enhancement of the Faraday Effect, which results from a
coupling of plasmon resonance (519 nm) with diamagnetism of glass matrix [209].
PbS quantum dots (QDs) are also precipitated in glasses by multiple irradiation of the femtosecond
laser pulses followed by heat treatment [210], whereas CdS and PbS QDs in silica xerogel are obtained
using the similar radiation technique, but without conducting heat treatment [211,212]. The size and
the photoluminescence can be tuned by adjusting the laser irradiation conditions. Mardilovich et al.

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

179

have observed the QDs precipitation in CdSxSe1x-doped borosilicate glasses using femtosecond laser
processing [213]. They show that 1 kHz femtosecond laser does not create any structural changes that
can sustain the heat treatment and affect the QDs precipitation, but a 1 MHz pulse repetition rate laser
can induce chemical inhomogeneity across microscopic modifications, resulting in three chemically
distinct regions: sodium and potassium-rich, zinc rich, and silicon rich zones.
Space-selective precipitation of crystals and NPs inside a transparent material by using a focused
femtosecond laser irradiation with/without heat treatment has been realized. This technique promises
the fabrication of 3D multicolored industrial art objects, optical memory, nanograting, waveguides
and integrative waveguide-like optical switches with ultrafast nonlinear response.
2.5.5. Microvoid
Formation of microvoids in transparent materials under multiphoton excitation has been an active
topic due to their promising applications in micro-fabrication and high-density optical storage [92
94,99,214]. The typical void structure is a central volume of less dense material or a hole, surrounded
by a region of higher density material [92,93]. It is reported that microvoids can be generated in various dielectric materials (silica, quartz, and sapphire) [9294,99,215,216] and polymer material
[214,217,218] using a single- or multi-shot regeneratively femtosecond pulsed laser. Glezer et al. propose that tightly focused femtosecond laser pulses can be nonlinearly absorbed by transparent materials, generating highly excited electronion plasma with high temperature and high pressure. These
conditions exist only in a small volume at the focal point. This tight confinement and extreme conditions result in an explosive expansion a microexplosion [92,93,215]. Materials are ejected from the
center and forced to the surrounding volume as a result of the microexplosion, leading to permanent
structural changes, including formation of microvoids surrounded by a region of compacted material.
As the self-focusing significantly reduces the beam waist, voids with the diameter size of 200250 nm
(smaller than the natural beam waist) are observed [93]. Juodkazis et al. indicate that the femtosecond
laser pulse creates an intensity over 1014 W/cm2 converting material within the absorbing volume of
0.2 lm3 into plasma state in a few femtoseconds and the pressure and temperature produced using a
single focused laser pulse (100 nJ, 800 nm, 200 fs) inside a sapphire crystal can be as high as 10 TPa
and 5  105 K, respectively [99]. Strong shock and rarefaction waves can be generated by the extremely high pressure, which result in the formation of a nanovoid surrounded by a shell of shockaffected material inside undamaged area [97,99]. The size of the void and the shock-wave-affected
(densified) region are determined experimentally as functions of the deposited energy and are in
agreement with simulation results, indicating that the experimental results can be understood in
terms of conservation laws and plasma hydrodynamics [99,219221]. The void diameter Dv can be
expressed by Eq. (2.8), through the pulse energy (Ep), the threshold energy for the void formation
(Eth), absorption depth of laser radiation in plasma la, and parameter F, defined as a function of the
compression ratio of the shell.

Dv

la q
3
AEp  Eth
F
1=3

F 1  d1

2:8
2:9

where d is the ratio of the final density of the densified shell (q) to the initial density (q0) of the transparent materials, and d = q/q0 > 1, which is directly derived from the densified shell and void diameters through the law of mass conservation [99,219].
Gu et al. reports on formation of voids in different polymer matrices [214,222,223]. The large
change in refractive index and a void enable confocal reflection microscopy to be used as a detection
method. They show that voids controlled in a multilayered structure can be used for read-only
high-density optical data storage [214]. 3D void-based diamond-lattice and face-centered-cubic lattice photonic crystals are fabricated in transparent bulk polymer. Three gaps are observed in the
[100] direction with a suppression rate of the second gap of up to about 75% for a 32-layer structure
in the former case [222]. A suppression rate of about 70% as well as the second-order stopgaps has
been observed in both [100,111] directions [223]. They show that the dependence of the stopgaps

180

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

on the illumination angle in the [100] direction is significantly different from that in the [111]
direction. Ventura et al. fabricate microvoid channels by local melting in a solidified polymer resin
sample moving perpendicular to the focus of a high numerical-aperture objective under femtosecond
laser (200 fs, 540 nm) irradiation [224]. Channel size, surface quality, and high density channel vicinity
are managed by optimizing the laser energy intensity and scanning speed. Elliptical channel
cross-sections of 0.71.3 lm in lateral diameter and an elongation in the focusing direction of approximately 50% are observed. A sharp peak in reflection and a suppression of infrared transmission in the
stacking direction by 85% at wavelength 4.8 lm with a gap/midgap ratio of 0.11 is demonstrated in a
20 layer woodpile-type photonic crystal with a 1.7 lm layer spacing and a 1.8 lm in-plane channel
spacing. Dense arrays of microvoids are also generated by femtosecond direct laser writing in glass
(Fig. 15A) [225]. Radial birefringence is observed from this microvoid array (Fig. 15B and C). The
micro-based birefringence pattern exhibits the ability to convert the spin angular momentum of light
into orbital optical angular momentum, which is reflected by the production of large arrays of optical
vortex generators with surface densities up to 104 cm2, as shown in Fig. 15DG.
The size and shape of microvoids are also controlled by changing the laser focus inside the polycarbonate, assigned to a combination of heat accumulation and dome formation dynamics, where the
dome acts as a microlens shifting the laser focus within the sample [226].
Microvoids are generated in the Fe:LiNbO3 crystal exhibiting high refractive index via the femtosecond laser induced microexplosion by Zhou et al. [227]. Quasispherical microvoids are generated
using the near threshold power depending on the fabrication depth. Due to the anisotropic property of
the crystal, the effect of both the spherical aberration resulting from the refractive index mismatch
and the chromatic aberration originating from the finite spectral width of the laser in the Y direction
(cutting direction, perpendicular to the crystal axis) is weaker than that in the Z direction (crystal
axis). As a result, the maximum fabrication depth in the Y direction is approximately four times larger
than that in the Z direction [227].
Microvoid arrays can also be fabricated in an optical fiber by using a femtosecond laser for obtaining bending direction sensitive sensors [228]. The size of the microvoids can be controlled by changing
the laser fluence. It is indicated that that the more intense the laser fluence is, the larger the size of
void is. They confirm that microvoids are elliptically shaped in all the cases because the laser beam

Fig. 15. (A) Unpolarized white light imaging of the photomodified glass with increasing fluence values F = 6.28, 7.26, 7.73, 8.15,
8.25, 8.35, 8.41, and 8.46 J cm2 that refer to numbering from 1 to 8, respectively. (B) Crossed linear polarization white light
imaging with inverted grayscale, where LP and LP0 refer to the direction of the linear polarizers. (C) Crossed circular polarization
imaging at 532 nm wavelength, D = 150 lm. (D) Array of 100  100 spin-to-orbital optical angular momentum converters over
1 cm2 area. (E) Enlargement on 10  10 array of optical vortices obtained at the output of the sample at 532 nm wavelength. (F)
Enlargement of panel (E) exhibiting the phase profile of the obtained optical vortices with topological charge two. (G) Zoom on
the phase spatial distribution of a single optical vortex. (Reproduced with permission from [225].)

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

181

is subjected to a cylindrical lens effect of the optical fiber surface. Furthermore, fiber gratings can also
be formed, consisting of periodic microvoids [229,230].
The collapse of microvoids can lead to forming a large cavity with a diameter of several tens of
micrometers, or it is called disruptions [231]. Richter et al. suggest that the formation of these disruptions originates from a fast quenching process of the molten material after the laser irradiation.
Furthermore, they analyze the periodic and non-periodic formation of disruptions, indicating that
the processing parameters strongly influence the formation of disruptions.

2.5.6. Polymerization
Polymerization is a reaction between active monomer molecules to form polymer chains or 3D networks [232]. Nonlinear absorption induced photopolymerization at the focus of a femtosecond laser
beam is a powerful and facile technique for fabricating a variety of micro- and nanostructures
[12,232242]. Compared to the case of one-photon polymerization, the diffraction limit can be
exceeded by nonlinear effects in the multiphoton polymerization to give a subdiffraction limit spatial
resolution [243]. There are two types of materials that are commonly used for femtosecond laser
induced polymerization fabrication: liquid resins and solid photoresists [232,238,240,244246]. In liquid resins, femtosecond laser irradiation leads to an almost instantaneous transition from liquid to
solid phase, and this makes the fabrication of micro- and nanostructures most direct and allows
removal of the unexposed liquid by washing in a proper solvent. Photopolymerization of resins have
a high spatial resolution of higher than 100 nm [247]. Photoinitiators are added for radical generation
upon excitation, and monomers or oligomers act as the main skeleton of micro-nanostructures, and
cross-linkers that ensure insolubility in the developing solvents. Photoresists are initially solid, and
their multiphoton induced polymerization is latent. In the later case, multiple photoinitiators, photosensitizers and other functional molecules or even inorganic doping agents are added to initiate the
polymerizations [248250]. All the femtosecond laser induced polymerizations are confined to a
highly localized area at the focal point due to the quadratic dependence of the two-photon absorption
rate on the laser intensity. When the laser focus is moved in a 3D manner in the polymer matrix, the
polymerization occurs along the trace of the focus, generating 3D structures for various applications.
The sculptures of micro-bulls with the size of 10-lm-long, 7-lm-high are fabricated in resin
(SCR500; JSR, Japan), consisting of urethane acrylate monomers and oligomers as well as photoinitiators via femtosecond laser irradiation by Kawata et al. [12]. A spatial resolution of 120 nm is achieved.
Similar structures are also induced in the single-wall carbon nanotubes (SWCNTs)/polymer composites with the spatial resolution of 200 nm in lateral direction, which is much higher than that reported
for other fabrication processes for the 3D structural composites [251]. Ushiba et al. show that the
resultant composites exhibit higher mechanical and electrical properties, due to the self-alignment
of SWCNTs inside the fabricated matrix. 3D optical data storage and various kinds of photonic crystals
are realized by different groups [233235,237,244,245,249]. Since most of polymer materials used in
the photopolymerization do not possess sufficient third-order nonlinear susceptibility for functional
photonic devices, Gus group adds highly nonlinear QDs into the initial matrix to form nanocomposites
and thereby to enhance their nonlinearity [249,252]. They show that the pure polymer exhibits negligible third-order nonlinearity as evidenced by the Z-scan results. However, for the nanocomposites,
significantly high negative third-order nonlinearity can be observed. In addition, the nanocomposites
give increased third-order nonlinearity with ascending QD concentrations [249].
Recently, a nano-engineered photonic-crystal chiral beam-splitter consisting of a prism featuring a
nanoscale chiral gyroid network are generated by a galvo-dithered femtosecond direct laser writing
method and can separate left- and right-handed circularly polarized light in the wavelength region
around 1.615 lm. This structure will become a useful component for developing integrated photonic
circuits providing a new form of polarization control [253]. Closely packed hexagonal conical microlens arrays are also fabricated by direct femtosecond laser photopolymerization [254]. Fabrication of
high optical quality axicons of 15 lm in radius, having 150, 160, and 170 cone angles, is achieved.
Direct femtosecond laser photopolymerization is also adopted to fabricate high resolution microscopic
spiral phase plates [255]. The total phase change all around their center is generated to be an integer
multiple of 2p for the operating wavelength in the visible range.

182

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

In order to get structures with smaller feature size, many investigations have been carried out to
improve the spatial resolution of photopolymerization [239,243,256258]. Baldacchini et al. and
Fischer et al. pay special attention to the effect of the femtosecond pulse repetition rate on the
photopolymerization processes and the resultant structures [256,257]. By varying the repetition rate,
the ultimate dimensions of the written microstructures depend partly on heat induced polymerization. The widths of two-photon polymerized lines become smaller with decreasing repetition rates
and this is assigned to the localized heat accumulation [256]. In contrast, Fischer et al. report that
there is no influence of the repetition rate on the linewidth scale [257]. They find different nonlinearities for high and low repetition rates consistent with different initiation processes being involved. Cao
et al. show that the photoresin NK Ester BPE-100 with high photosensitivity and mechanical stability
can improve the fabrication resolution based on the single photon photoinhibited polymerization
[258]. Recently, Ma et al. demonstrate experimentally and theoretically that the two-photon absorption probability and polymerization can be efficiently controlled by shaped ultrashort laser pulses, to
reduce the volume of a fabricated microrod to 1/125 of that created by Fourier-transform-limited laser
pulses, i.e., less than the diffraction limit of 1/25 [259]. The super-resolution feature is achieved by
overlapping two laser beams of different wavelengths to enable the wavelength-controlled activation
of photoinitiating and photoinhibiting processes in the polymerization. The initiating beam is used for
starting the polymerization, while the inhibiting beam is used to suppress the polymerization. As the
exposure zone of the inhibiting beam is modulated to a doughnut shape, a diffraction-limit-free polymerization volume can be produced in the exposed center. This approach combines the generation of a
smaller photo-active voxel and the prevention of dark polymerization. 40 nm nanodots are obtained,
which are present only at a short exposure time and a relatively high power level of the initiating
laser. Furthermore, Gan et al. show that using the 3D two-beam optical lithography, 9 nm (l/42 of
the wavelength of the inhibition beam) feature size (Fig. 16A) and 52 nm (l/7) two-line resolution
(Fig. 16B and C), which is the minimum center-to-center distance between the two fabricated lines,
can be created in a newly developed two-photon absorption resin with high mechanical strength
[239].
2.5.7. Periodic surface structures
Femtosecond laser induced periodic surface structures (fs-LIPSSs) on the various materials, such as
metals, semiconductors, and dielectrics, have gained considerable attraction and hold many potential
applications in photonics, plasmonics, optoelectronics, thermal radiation sources and bio-optical
devices [260266]. LIPSSs are also referred to ripples or gratings. Two typical groups of fs-LIPSSs termed as low-spatial-frequency LIPSSs (LSFLs) and high-spatial-frequency LIPSSs (HSFLs) with different
spatial periods and orientations are identified, which depend on the irradiation conditions and materials parameters, such as incident laser fluence and refractive index [263,264,267271].
LSFLs with an orientation perpendicular to their polarization direction consist of spatial periods (K)
close to the irradiation wavelength (k, k > K > k/2). The influence of polarization, angle of incidence,
and wavelength of the incident laser beam implies that LSFLs formation is mainly governed by the
electromagnetic field [272,273]. Therefore, LSFLs are widely considered to be the result of the interference between an incident wave and a surface scattered wave [274276]. Bonse et al. study the dynamics and the femtosecond-LIPSSs with double pulses irradiation, and show that the pulse delay between
subpulses has a strong impact on the formation of nanostructures [263,277,278]. The characteristic
decrease of the LIPSS periods and the ablation crater depths are established with double-pulse delays
of less than 2 ps, and confirmed by a theoretical study [279], which suggests that the energy amount
deposited to the material by the second laser pulse is reduced owing to the increase of the surface
reflectivity induced by the first pulse. Bonse et al. propose that the variation of the absorbed fluence
can lead to strong local changes of the optical properties, promoting the initially dielectric material
locally into a metallic state (as a dense free-electron plasma in the conduction band of the solid)
[264]. The calculations show that the surface reflectivity significantly increases as the material turns
from a dielectric state into a highly absorbing and high reflective metal-like state. The optical properties of the femtosecond laser excited materials significantly depend on the carrier (e.g., free-electron)
density. Consequently, an intensity pattern with modulation length of the order of the wavelengths
can be imprinted by the spatial variations in the absorbed local fluence induced by interference

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

183

Fig. 16. (A) Feature size of free-standing lines versus the intensity of the inhibition beam, under the exposure of the writing
beam with different fluences. Insert: SEM images of points a, b, c, d and e with a scale bar of 100 nm. (B) SEM images of two-line
resolution generated with different intensity of the inhibition laser beam. (C) The cross-sectional profile of image f in (B).
(Reproduced with permission from [239].)

between the incident laser pulse and a surface scattered electromagnetic wave [277]. Furthermore, it
is found that HSFLs begin to appear at a pulse delay of 1.33 ps and completely replace LSFLs at 40 ps,
and this is due to the free-electron plasma and transient changes of the optical properties during the
ablation process, conforming free-electron plasma plays an important role in the formation of LSFLs. A
periodic structure of ripples with a spatial period of 720 nm and an alignment parallel to the electric
field of light on GeS based chalcogenide glass is produced upon irradiation by a focused beam of a
femtosecond laser (1 kHz, 34 fs, 806 nm) [280]. With an increasing number of pulses, from 5 to 50
pulses, a characteristic evolution of ripples is observed from a random structure to a series of
self-organized periodic structures with generally aligned peaks-and-valleys, which is assigned to
the strong temperature gradients combined with interference of the incident laser irradiation and a
scattered surface electromagnetic wave.
For the case of HSFLs, ripples, either orthogonal or parallel to the polarization direction, show a
periodicity significantly smaller than the wavelength of the laser light. The origin of the HSFLs is still
under debate in the literature. Different mechanisms such as SHG, self-organization, interference,
nanoplasmonics and standing wave have been proposed [261,262,265,281]. A threshold effect of
nanograting generation is confirmed by the observation of switching from smooth modification to
appearance of nanogratings [282]. The increase of the intensity and the highly localized nonlinear ionization cause the increase in the fabricated line width. Well-shaped nanogratings can be fabricated by
maintaining the incident laser intensity slightly higher than the threshold. The nanogratings are suggested to be generated by local field enhancement that highly localizes and extends nonlinear ionization in the direction perpendicular to the electric field and then explosively expels electrons, ions, and
neutral atoms out of the heated zone. Liang et al. show that the fabricated nanograting period on the

184

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

surface of silica depends on the repetition rates, but remains constant upon changing the laser pulse
energy originating from the intensity clamping [260,283]. The nanograting period decreases with
increasing the incident pulse fluence and the number of overlapped shots in both stationary and scanning cases, owing to the local intensity distribution and incubation effect [265,283,284]. The interplay
between the local intensity distribution and the incubation effect causes an earlier ablation during the
scanning with less overlapping pulses with an increase in the pulse fluence, and hence, reduces the
sizes of the ablation zones or nanogrooves (less overlapping pulses) and the nanograting period (earlier ablation) [283]. Nanograting formation on the surface of fused silica is further demonstrated to
basically result from a laser ablation process [270]. Based on the nanoplasmonic model and the incubation effect, it is suggested that the laser ablation is initiated at the pre-existing defect sites with the
incident pulse fluence smaller than the ablation threshold. Nanogrooves evolve through the elongation to the randomly induced nanocraters with the increase in the number of shots. With the pulse
fluence slightly above the ablation threshold, the side and inner nanogrooves are generated in pairs.
When the pulse fluence is well above the threshold, a layer of molten material is first removed by
the first few shots leading to the reduction of pulse fluence, the nanogrooves evolve to nanogratings.
The absence of ablation in the nanostructuring of the thin polymer films upon femtosecond laser
irradiation is proposed to account for a self-organization mediated process to generate periodic structures with period lengths similar to the laser wavelength and the direction paralleling to the laser
polarization vector, which does not require surface material removal [285]. In addition, a feedback
process is also confirmed, as repetitive irradiation is needed in order to observe LIPSSs formation.
Recently, laser induced periodic annular surface structures are formed on fused silica irradiated
with multiple femtosecond laser pulses [286]. This surface morphology emerges after the disappearance of the conventional LIPSSs, under more successive laser pulse irradiation, as revealed in Fig. 17A
F. These structures are independent of the laser polarization and universally observed for different
focusing geometries, due to the interference between the reflected laser light field on the surface of
the damage crater and the incident laser light. With the sufficient laser shots, radiation patterns in
the form of an ellipse and a ring with (evolving as shown in Fig. 17GR) a large opening angle (about
100) are observed on the surface of 14 different transparent materials including glasses, crystals, and
polymers, in addition to the directly transmitted laser beam [287]. The elliptic radiation causes a laser
polarization-dependent orientation, whereas the ring-shaped radiation does not. Furthermore, the
ring-shaped radiation consists of numerous colored needlelike substructures with radial orientation
together with those of the central white light and the incident laser pulse (Fig. 17R).

Fig. 17. (A)(F) Evolution of the damage morphology as a function of the laser shots number. Scale bars, 2 lm. (G)(O) Laser
emission patterns as a function of laser shot numbers for horizontally polarized pulses with the same number of laser shots in
each panel. (P) and (Q) Laser emission patterns for 240 laser shots with vertical and circular laser polarizations. (R) Emission
pattern obtained with CaF2 for 580 laser shots, with other experimental conditions identical to (G). (Reproduced with
permission from [286,287].)

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

185

2.5.8. Refractive index change


The possibility of modifying the refractive index of the local region in transparent materials by the
use of focused intense femtosecond near-infrared laser pulses from a Ti:sapphire laser at 800 nm has
attracted a lot of attention during the past several years mainly due to the localized refractive index
change allowing for creating the building blocks of more complex photonic systems embedded in optical materials [70,82,8587,96,98,262]. The first pioneering work was conducted by H. Misawa in
Matsuhara Microphotoconversion Project, Japan Science and Technology Agency. He occasionally
observed formation of a bright spot due to refractive index change in a silicate glass upon irradiation
with ultrashort pulsed laser. Then he proposed a technique for fabrication of 3D optical memory with
ultrahigh storage density [288,289]. K. Miura in Hirao Active Glass Project, Japan Science and Technology Agency first considered and demonstrated the fabrication of optical waveguides by using femtosecond laser direct writing [82,94]. This technique also exhibits the ability to create various
photonic devices, such as couplers, gratings, binary storage elements, and photonic crystals, just by
direct writing in three dimensions [5,10,290295]. The optimal performance of glasses for various
photonic uses, especially the 3D integrated photonics, requires a precise adjustment of the refractive
index in three dimensions, which can be addressed by the nature of high spatial resolution femtosecond laser direct inscription, as the laser induced structural modifications associated with refractive
index changes and the processing degree can be precisely controlled by managing the irradiation conditions [5,10,290295].
As mentioned above, the nonlinear ionization processes such as multiphoton excitation of electrons from the valence band to the conduction band, multiphoton ionization and/or tunneling ionization, as well as impact ionization, are proposed to be responsible for the generation of microplasma in
the focal volume. The high excited plasma transfers its energy to the glass structure, ultimately leading to a change in glass structure and hence, in optical properties such as the index of refraction. However, the intrinsic mechanism of the material modifications and the accompanied refractive index
changes is still an open question and requires further quantification for the benefit of advanced control
options. Depending on the regime of laser interaction, particularly on the incident fluence, pulse duration, wavelength, laser polarization, focusing conditions and scanning speed, femtosecond laser pulses
can induce either positive isotropic refractive index changes (Dn  104 to 103) or void-like rarefaction regions of lower density with compression shells, or self-organized nanoscale layered structures
resulting in formation of birefringence with overall negative index changes [86,95,96,290,296298].
Several mechanisms are considered to be responsible for index changes: (i) thermal mechanisms
(fictive temperature model), when higher density local structures are created after fast quenching
from a high temperature melt [297,299], (ii) non-thermal mechanisms, when index changes originate
from the generation of color centers [66,86,98,157,300,301], (iii) density changes originating from
defects induced structural network reorganization (defects induced densification) [86,297,302], and
(iv) mechanical contributions, when compaction and rarefaction of material result from the pressure
wave release [53,54,303,304].
The densification induced refractive index changes are usually determined by the Raman spectroscopy. A large increase of the oscillation modes of three- and four-membered rings in silica causes
an increase of refraction index [296,297,305]. Shifts of Raman peaks related to the (POP)sym and
(PO2)sym network vibration modes to lower wavenumbers in ErYb doped phosphate glass are
observed in the regions of decreased refractive index induced by cumulative-heating thermal effects
after irradiation by a high repetition rate femtosecond laser, consistent with the results in Ge-based
glass [306,307]. With low repetition rates (1 kHz), positive refractive index is generated resulting from
the formation of color centers in Yb doped phosphate glass, silica, etc., confirmed by Raman
spectroscopy, refractive near-field profilometry, ESR spectroscopy, phase contrast microscopy, photoluminescence of defects and incoherent secondary light emission [87,95,300,301,308]. For the
mechanical contributions, pressure wave is determined by a transient lens (TrL) method, developed
by Sakakura et al. [53,54,303,304]. This method can be used to investigate the temporal and spatial
developments of the refractive index change in a focal region inside the glass irradiated by the femtosecond laser pulse. They indicate that amplitude of the pressure wave increases with increasing
the excitation pulse energy. In addition, Raman spectra at the laser irradiated region reveal that the

186

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

compact silica ring structures increase, suggesting that the photoexcited glass is densified by a shock
due to a pressure wave generation.
Furthermore, different laser wavelengths also lead to different refractive index changes and mechanisms. Saliminia et al. show that though a common origin (densification) is proposed for the mechanism of the laser induced refractive index change in silica glass derived by excitation at both 800 nm
and 1.5 lm wavelength, the index gradient between the core of the irradiated zone and the surrounding region in the 800 nm case is not as sharp (well defined) as that for the 1.5 lm case due to the lower
order of nonlinearity contributing in plasma generation through the multiphoton excitation [297]. No
dependence of the core shape on laser polarization is found by changing the femtosecond beam polarization from vertical to horizontal and circular states. A domain of optical waveguide writing with
high refractive index contrast (0.022) is reported in fused silica by strong focusing of a 522 nm wavelength, 500 kHz repetition rate femtosecond laser with oil-immersion optics [309]. It is suggested that
the effective 2-fold higher fluence provided by green femtosecond laser pulses enables a stronger
interaction with fused silica compared to the fundamental wavelength.
In addition, both the sign of index change and the mechanism underlying such change are also
dependent on the glass compositions [299,310]. Bressel et al. report that densification caused by
the void formation in GeO2 glass is responsible for the changes of refractive index, which is very similar to the changes under hydrostatic pressure. In contrast, the refractive index changes in SiO2 glass
are attributed to the pressure effect or the fictive temperature anomaly, i.e., a resultant smaller specific volume of the glass (a denser phase) at a high thermal quenching rate [299]. Bhardwaj et al. present
a comprehensive study on femtosecond laser induced refractive index modification in a wide variety
of multicomponent glasses such as borosilicate, aluminum-silicate, and heavy-metal oxide glasses
along with lanthanum-borate and sodium-phosphate glasses [310]. They demonstrate that the refractive index modification in multicomponent glasses can be positive, negative, or nonuniform, and exhibits a strong dependence on the glass composition characterized by using high-spatial resolution
refractive index profiling techniques under a wide range of writing conditions. Except some
aluminum-silicate glasses, all other glasses so far studied exhibit a negative/nonuniform index change.
For some glasses with more complex compositions, ions migration and exchange and crystal precipitation can also induce rearrangements of the glass structure accompanied with refractive index
change [20,300,311,312]. As discussed above, the change in the refractive index of phosphate glasses
induced in the high repetition rate regime is different from that in the low repetition rate regime.
However, both repetition rate regimes result in a positive refractive index change in borosilicate
glasses. This indicates that the underlying structural modifications are different in the two types of
glass systems [313]. The refractive index change in borosilicate glass at low repetition rates is assigned
to the formation of non-bridging oxygen atoms, whereas densification and rarefaction of the glass network are the dominating cause for the index change in the high repetition rate regime. The modification of the glass network related to changes in the bridging bonds linking zirconium fluoride
polyhedra is observed in fluorozirconate glass with high repetition rates, which is similar to that
reported elsewhere [307,314]. Although much work has been devoted to this field, more is needed
for further understanding of the femtosecond laser induced refractive index changes.
As suggested above, local refractive index changes originating from electronic and structural modification of matter are the building blocks of laser-induced optical functions in bulk transparent
materials, which provides a well established method for the fabrication of integrated photonic devices
[46,291295,315327]. Recently, great efforts have been devoted, which are of scientific and industrial importance as they offer complex circuitry and hybrid functionality on a small footprint. Compared to other fabrication methods, the direct femtosecond laser writing technique offers the
advantage of a rapid and one step route with the potential for fabrication of 3D photonic architectures
[229]. For example, waveguides and Bragg-gratings can be directly incorporated during the same processing setup to create a monolithic laser with narrow linewidth output [229,316]. As shown in
Fig. 18A and B, waveguide arrays are fabricated by femtosecond laser (400 femtosecond pulses at
520 nm wavelength) waveguide written in fused silica. Adopted writing conditions consist of 300 nJ
pulses delivered at a repetition rate of 20 kHz [293]. Corrielli et al. demonstrate the first experimental
observation of fractional Bloch oscillations (BO), using a photonic lattice (waveguide arrays) as a
model system of a two-particle extended BoseHubbard Hamiltonian [293]. BO is the oscillatory

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

187

Fig. 18. The photonic simulator of correlated BO. (A) Sketch of the waveguide array structure; red-colored waveguides have a
different refractive index change with respect to the others to implement the on-site particle-interaction defect U0. (B)
Section of the fabricated array, imaged with an optical microscope. Scale bar, 100 lm. (C) Experimental and (D) numerical
images of the light intensity distribution in the waveguide lattice main diagonal versus propagation distance z, representing in
Fock space the bloch oscillations for two interacting particles. (E) The evolution of the fidelity. (FH): same as (CE), but for the
single-particle BO in the 1D waveguide array. (Reproduced with permission from [293].)

motion of a quantum particle in a periodic potential driven by a constant force, which constitute one
of the most striking and oldest predictions of coherent quantum transport in periodic lattices [293].
Competition between interactions of particles and their mobility can give rise to novel dynamic
behavior, where particles generate bound states and co-tunnel through the lattice. Although fractional
BO at a frequency twice (or multiple) that of single-particle BO are predicted in the bound states
between few strongly interacting particles, the observation of fractional BO is challenging for
condensed-matter systems. In the photonic simulator, the dynamics of two correlated particles hopping on a one dimensional (1D) lattice is mapped into the motion of a single particle in a two dimensional (2D) lattice with engineered defects and mimicked by light transport in a square waveguide
lattice with a bent axis. At first, Corrielli et al. inject the probe light into the central waveguide and
image its propagation along the waveguide lattice diagonal (Y). Fig. 18C reveals the oscillatory

188

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

behavior of the light dynamics along the waveguide array during experiment, i.e., spreading of light
into several waveguides, until a maximum of the breathing amplitude is reached. Then the light refocuses into the central waveguide and the breathing starts again. This periodic behavior demonstrates
the correlated BO of two interacting particles, seen in the Fock space. A very good agreement with the
numerical simulations of light propagation in the designed structure (Fig. 18D) confirms the accuracy
of the fabricated photonic simulator with an average quantitative estimation value of the agreement
of about 90% (Fig. 18E) provided by the fidelity. The fractional nature of BO for correlated particles is
further investigated by comparing Fig. 18D with the BO pattern for single-particle hopping in the same
lattice (Fig. 18F). In this case, a breathing propagation pattern corresponding to the single-particle BO
is also observed with the position of the maximum breath amplitude corresponding to the refocusing
position for two-particles BO (Fig. 18C), which indicates that the frequency of a two-particle BO is
twice that for a single-particle. A high agreement value between the experimental and theoretical
results is also obtained (Fig. 18G and H).
In addition, monomode waveguides, which have controllable cross-section in shape of triangle,
square, polygon and circle with center-to-vertex distance (or radius for circle) of 18 lm and cladding
thickness of about 6 lm are longitudinally written in phosphate glass by Zhao et al. using femtosecond
laser with the laser modified shell acting as cladding and the encompassed and compressed column as
core [328]. The similar near-field output distribution of waveguide is obtained for orthogonally polarized input light, suggesting symmetric index profile. 1-to-2, 1-to-3 and 1-to-4 splitters are also produced with the adjustable splitting ratio. High contrast cladding waveguides are also fabricated in
various types of nonlinear crystals with different configurations by Chen group [327,329334]. Laser
oscillations at different wavelengths and SHG are realized using the cladding waveguides by this
group.
2.5.9. Periodic structures induced by interference fields of femtosecond laser
As mentioned previously, there are two mechanisms, by which ripples form. First, the interference
between the incident laser waves and the surface scattered waves can lead to the formation of ripples
(LSFLs) with periods close to the wavelengths of radiation lasers [274,276]. Second, the interference
between incident light and excited surface plasmons results in generation of sub-wavelength ripples
(HSFLs) [261,271,274,276,335]. Furthermore, periodic nanostructures are also fabricated in glass
based on the mechanism similar to the second one [19,336]. As the HSFLs promise many applications
in nanophotonics, they have drawn much attention, although it still remains a challenge to theoretically explain or predict the formation of sub-wavelength ripples [337,338]. Therefore, here we introduce the experimental and theoretical results of HSFLs induced by the interference between incident
light and excited surface plasmons.
MPA produces an electrostatically unstable surface leading to an explosive emission of free electrons and positive ions, subsequently to formation of surface plasma and surface plasmons [339].
Because of the spatial and temporal variations of free electron density, reflectivity of the ionized
dielectrics and the excited surface plasmons, the incident laser will be reshaped from the original
Gaussian distribution.
After the critical density of the free electron is created, the laser intensity is partially reflected.
Then, surface plasmons are resonantly excited and then the coupling field intensity is strongly affected
due to the interference between the absorbed laser field and surface plasmons. As a result, the intensity is strongly reprofiled and shows periodic patterns, which locally enhances the field and the ablation, leading to formation of periodic structures [271,276,340]. As this process dominantly occurs in
the intrinsic defects or preformed nanostructures, especially in the nanogroove for the cavity-mode
enhancement of polarization dependence, the polarization dependent nanogratings will be formed
via a positive feedback process [339]. This mechanism suggests that multipulse irradiation is necessary for creating ripples when there are no intrinsic defects, and this is confirmed by Gottmann
et al. [341]. They find that at least five pulses on every surface area are necessary to obtain coherently
continued sub-wavelength ripples by scanning several tracks with an offset on the surface of sapphire
and fused silica. Huang et al. report that the ratio (K/k) of K to the laser wavelength (k) decreases as
irradiation pulse number (N) increases and it is also related to crater diameter (D). The large K/k,
which always appears at small N or in the central area, is associated with shallow grooves. In other

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

189

words, the smooth ripples have large K/k, whereas the ripples with deep grooves have small K/k. This
means that the groove depth (H) is also an important factor for K due to the Gaussian distribution of
the field in the focus [261]. In addition, in the central area K also decreases as N increases and is
accompanied with groove deepening, and this is related to the grating-assisted surface
plasmonslaser coupling. Therefore, they propose that ripples result from the admixture of the
field-distribution effect and the grating-coupling effect: the initial direct surface plasmonslaser
interference and the subsequent grating-assisted surface plasmonslaser coupling.
Jia et al. report that when two collinear femtosecond laser pulses with the wavelengths of 800 nm
and 400 nm, respectively, simultaneously irradiate the surface of ZnSe crystal, regular nanogratings
with period of 180 nm are generated on the whole ablation area because of the interference between
the surface scattered wave of 800 nm lasers and the 400 nm light [342]. The period of the nanogratings
is about k/2n, where n is refractive index of the sample, and k the laser wavelength [281,342]. A 2D
periodic structure including periodic hexagonal lattice of microholes (Fig. 19A), micro-orbicular platforms (Fig. 19B), and microcones is fabricated on the surface of silica glass by MPA using a single shot
of three interfered femtosecond laser pulses without any assistance of diffractive optical elements
[343,344]. The depth and the diameter of the holes in the microhole structures are about 220 nm
and 1.25 lm (Fig. 19E), respectively. The depth and the diameter of the void in the center of the orbicular platform are about 120 nm and 350 nm (Fig. 19F), respectively [344]. The formation of the different microstructures is assigned to the formation of plasma and molten liquid layer on the surface of
bulk silica glass at different energy levels of a pulse. 2D hat-scratch structures are also produced by
three interfering replicas of a single femtosecond laser pulse on silica glass surface, which are directly
related to the intensity of the incident laser beam and a result of the combined laser ablation effects

Fig. 19. Photoinduced periodic microstructures fabricated on the surface of silica glass by three noncoplanar beams with
different pulse energies. (A) Pulse energy is 75 lJ per pulse and (B) energy is 30 lJ per pulse. (C and D) The 3D image of the
microstructures of (A) and (B), respectively. (E and F) Cross-section of the microstructure in the direction of white line showed
in (A and B), respectively. (G) SEM image of complex 2D ZnSe micro/nanostructure. (H) PL spectra of the plane surface and the
2D nanostructures on ZnSe crystal irradiated by 1200 nm fs laser with a pulse energy of 10 lJ. The insets (a) and (b) show the PL
pictures from the 2D nanostructures and plane surface, respectively. (Reproduced with permission from [344,350].)

190

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

including ionization, shock wave, plasma expansion, and phase explosion [345]. Nanogratings are
generated by a single shot of two pulses interfered with each other [346].
The femtosecond laser induced classical ripples on dielectrics, semiconductors, and conductors can
be achieved, which exhibit a prominent non-classical characteristic with the periods significantly
shorter than laser wavelengths [261]. The material surface irradiated by femtosecond laser with
damage-threshold fluence behaves metallic, no matter for metals, semiconductors, or dielectrics
[261,347]. Wagner et al. show that the ripples extend coherently regardless of the orientation of
the polarization, and hence a scanning perpendicular to the polarization results in parallel grooves.
The ripple spacing on fused silica does not noticeably depend on the pulse spacing [347]. The simulation results show that uniform structures, in terms of ablation shapes and subwavelength ripples, can
be easily generated with a lower fluence or subpulse energy ratio of 1:1 with a selected fluence [338].
The incident fluence, angle and wavelength also influence the periodicity [271,285,335,348,349]. More
discussions about periodic structures in transparent materials induced by interference fields of femtosecond laser will be given below.
Complex 2D micro/nanostructures with 2D spots forming a hexagonal arrangement and short periodic nanostructures embed in the hexagonal microstructures are fabricated on ZnSe crystal surface by
the interference of three 800 nm femtosecond laser beams (Fig. 19G) [350]. Compared with the initial
surface of ZnSe crystal, the intensity of the near band-edge (NBE) emission of 2D nanostructures is
enhanced by tens to hundreds times, and the SHG is compressed to 1323% under the excitation of
femtosecond laser with the wavelength from 1200 to 1600 nm (Fig. 19H). Pan et al. suggest that
the enhancement of NBE emission and the compression of SHG are induced by several factors, e.g.,
the increase in optical absorption, the reabsorption of the SHG on the nanostructures, and the mismatch of the polarization direction dependence of the SHG on the nanostructures.
2.5.10. Formation of bubbles
Besides the normal birefringence structures and voids, bubbles are also observed in the femtosecond laser induced modifications [351354]. The behavior of bubbles is not only ubiquitous but also
complex in a multitude of fluid system, which is intensively studied and widely used in contemporary
science and technology [353,355]. The development of laser brings a promising prospect for the study
and application of bubbles [355,356]. Long-lived bubbles with the size of 13 lm at the boundary of
the irradiated region in transparent materials have been induced by the tightly focused femtosecond
laser, which are evident from the cross-sectional image and the corresponding top view images of the
line structure [351,352]. The color of bubbles is visible under white light illumination in the optical
microscope attributed to the FabryPerot effect [352]. In situ observation reveals that the bubble
moves against the direction of propagation of laser pulses as far as 2 lm in CaF2 [351]. Yang et al. find
that the bubbles are shifted perpendicular to the writing direction [352]. The bubbles locate at about
5 lm toward one side of this center, not toward the region with the highest intensity (which is in the
center of a Gaussian beam). The asymmetry of bubble formation and its dependence on writing direction are demonstrated and assigned to the light pressure at the tilted front of the pulse. The authors
report that permanent micro-bubbles with varied size and number density are generated in borosilicate glasses by adjusting the focusing depth (d) of a tightly focused femtosecond laser [353]. Different
from the previous report, in the authors case, most of bubbles locate at center of the modified region.
Increasing d from 80 to 220 lm, the average size of generated bubbles experiences an increasedecrease process, determined by the balance between the incident laser fluence and absorption of
plasma (Fig. 20). When d = 80 lm, numerous small bubbles appear irregularly within the laserwriting track (Fig. 20A). With d in the range of 100180 lm, the induced bubbles basically consist
of two parts: well aligned bigger bubbles at the center of the laser-writing line and random smaller
bubbles around the bigger bubbles (Fig. 20BF). With d as 200 or 220 lm, smaller bubbles with a
homogeneous size are observed (Fig. 20G and H). However, the number density of generated bubbles
experiences an opposite changing process compared to the change of the size. Bellouard and Hongler
study the influence of the scanning speed on the structures formed in the fused silica when exposed to
low-energy pulses at the cumulative regime (repetition rate 9.4 MHz) [354]. At low speeds, erratic
patterns are produced characterized by randomly sized bubbles, which are separated by random space
intervals. As the speed increases (typically above 10 mm/s), self-organized patterns are observed with

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

191

Fig. 20. The bubbles induced at different focusing depths. (Reproduced with permission from [353].)

varying complexity and periodicity with the lines behaving as waveguides. These patterns are characterized by an outer shell of modified material, inside which well-defined cavities exist with a dark
appearance due to total internal reflection associated with bottom-light illumination. The complexity
of these patterns gradually decreases and ultimately converges to form a single spherical cavity. A
strong dependence of these patterns along the writing direction is observed, which is assigned to
the presence of pulse front tilt (PFT, defined as a tilt in the intensity distribution in the front of the
pulse) or to the imperfect symmetry of the energy distribution of the laser spot [352]. The smallest
bubbles are found at greater depths and formed before the larger ones. A threshold of pulse energy
is determined to be about 210 nJ for the formation of periodic line patterns. As the pulse energy is
increased, the outer shell of the laser-affected-zones gets wider and the bubble size increases, and this
is consistent with a temperature driven process. 2D and 3D patterns are also generated to fabricate
bubble crystals.
Yang et al. propose that bubbles can be formed on the axis of the focused laser beam as a result of
microexplosion subsequently moving from the molten center to the boundary of irradiated region and
also formed directly at the boundary region of a light affected zone, where tensile stress is responsible
for the formation of cavitation bubbles when the rupture strength in the molten glass is exceeded
[352]. We suggest that the unique conditions with extreme temperature and pressure and heat
accumulation effect can induce intensive physical and chemical changes in the melted region, such
as boiling, evaporation, plasma formation, bond breaking and reforming [353]. Under such conditions,
the nucleation of bubbles can be induced from several possible reasons, such as phase explosion, the
plasma formation and vaporized glass. Furthermore, during the bubbles generation, they suffer the
optical radiation pressure from the incident laser, and gain the initial momentum along the laser propagation axis. Then, the bubbles can be driven away from the laser irradiated region by the radial shock

192

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

wave. On the other hand, with the increase of laser shots, the laser induced temperature difference
causes a radial pressure gradient driving bubbles away from the center of the melted region. As a
result, the above three actions overcome buoyancy and the viscous force, and push the bubbles to
move from the focal center to the bottom of the melted region [353]. The previously generated smaller
bubbles can be coalesced to form big bubbles [354]. The bubble growth rate as a result of everincreasing coalescence events may exceed the scanning speed up to a point at which the laser beam
propagation is altered, stopping the underlying nonlinear absorption process. This causes a cooling
down below the bubble nucleation threshold. When the laser beam has moved away from the bubble
zone, the nonlinear absorption starts again leading to the formation of the periodic structures. The
mechanism of bubble formation is a puzzle and the mechanism for the formation and evolution of
bubbles induced by femtosecond lasers in solid transparent materials needs further investigation.
These femtosecond laser induced bubbles may hold promising applications in optical storage, 3D
optical micro-devices. Birefringence and waveguiders are demonstrated, which are based on a chain
of connected bubble structures [352,357].

3. Femtosecond laser induced anomalous phenomena


3.1. Femtosecond laser induced nanogratings
As discussed above, different regimes of structural changes can be generated depending mainly on
the incident fluence of the femtosecond laser pulses. At intermediate energies, birefringent index
changes are induced (type II). About a decade ago, a peculiar self-organization behavior, which manifests as nanogratings with a width of about 20 nm and a tunable period of 140320 nm, is observed in
the local volume upon the femtosecond laser irradiation, demonstrated in silica glass [19]. These
structures consist of thin regions with refractive index of n1 characterized by a strong oxygen deficiency, surrounded by larger regions with refractive index of n2 [358]. Typically, femtosecond laser
induced nanogratings in the glass exhibit refractive index difference for extraordinary and ordinary
rays of about 24  103 with a typical period of about k/2n, n being the refractive index and k the
writing wavelengths [359,360]. Since this self-organization behavior was discovered, inducing
nanogratings by femtosecond laser in transparent materials, with their ability to locally control polarization state of transmitted light, has been proven to be a facile and powerful technique for numerous
applications [101,103] ranging from anisotropic microreflectors [361], polarization-sensitive waveguides [88,362], polarization converters [363] to multidimensional optical memory [364,365] and
polarization selective holograms [366]. Intensive research has been devoted to understanding the formation mechanism of the nanogratings [19,74,103,365,367369]. It is reported that the homogeneous
periodic layered medium behaves like an uniaxial birefringent crystal [370], where an electric field
oscillating parallel (TE mode) and perpendicular (TM mode) to the layered medium leads to different
extents of the phase shift. Thus, birefringence is generated in an isotropic glass just due to the presence of the nanogratings.
Femtosecond laser induced nanogratings usually exhibit two periodic structures: one perpendicular to the polarization and the other parallel to light propagation direction [68]. The first grating has a
period shorter than the wavelength of the incident light depending on experimental conditions. The
second period is growing from the head of the structure to the tail with the initial period close to
the ratio (k/n) of the light wavelength (k) and the material refractive index (n). Nanoporous structures
are observed in the nanoplanes of the nanogratings, indicating possible chemical decomposition of
glass during the irradiation [371]. Long-range Bragg-like gratings with the planes thickness of smaller
than 10 nm are produced with the laser polarization parallel to the scan direction [367]. Nanogratings
also lead to a local increase of etching rate, which strongly depends on the polarization. This may be
assigned to the oriented crack structures or nanopores [55,101,371373]. Therefore after chemical
etching, microfluidic channels are embedded in the glass. Champion et al. report that for nanogratings,
a volume expansion occurs and the energy deposition has an influence on the volume variation that
can be correlated with etching rate, Raman spectra and optical properties [373,374]. Nanograting orientation influences the stress distribution induced in the material. As the deposited energy increases,

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

193

nanopores are produced, leading to the inflation of the lamellae volume, and hence to localized stress
generation and eventually the collapse of the nanogratings structure when the stress becomes too
high [374]. A combination of small angle X-ray scattering measurements and direct observations with
focused ion beam milling and SEM confirms the presence of nanopores or cavities, and hence, reveals
that the grating planes are composed of isolated narrow, sheet-like cavities with thicknesses around
30 nm and transverse diameters of 200300 nm [375,376].
Hnatovsky et al. and Richter et al. find the following aspects [55,377,378]. First, the writing speed of
10100 pulses/lm is necessary for triggering nanoplanes organization. Second, the energy threshold
for the nanograting formation decreases with an increase of the adopted repetition rate. The writing
speed and repetition rate affect the modification of nanogratings implying a cumulative effect works.
The grating period continuously decreases with increasing the number of irradiated laser pulses and is
nearly independent of the pulse energy [377,378].
Polarization dependent self-organized nanogratings are fabricated inside SrTiO3 crystal and fused
silica by Qiu et al. using a low repetition rate femtosecond laser [63,379381]. The groove orientations
of those gratings can be controlled by adjusting the irradiation pulse number per unit scanning length,
i.e., either through adjusting the scanning velocity at a fixed pulse repetition rate or through changing
the pulse repetition rate at a fixed scanning velocity [379,381]. A void-moving model is suggested to
qualitatively explain the formation mechanism of the grating as well as the variation of the groove orientation. Dai et al. report that self-organized nanogratings fabricated by femtosecond laser can be
orderly rotated in three-dimension inside fused silica by controlling the laser polarization direction
(Fig. 21A and B) [380]. Fig. 21D shows that the lines vary gradually from bright to dark depending
upon polarization plane azimuth h, indicating that the difference Dn between the ordinary and
extraordinary indices of refraction will change from positive to negative. A nearly perfect symmetric
distribution of the birefringence signal on both sides with the center at 90 can be seen in Fig. 21E.
Furthermore, the scanning directions strongly influence the birefringence signal intensity. This
observed non-reciprocal writing phenomenon indicates that the incident pulses possess front tilt.
Therefore, the authors propose that the incident electric field can project a sub-vector along the light
propagation direction, and this sub-vector can force an oscillation in the excited plasma wave, resulting in the rotation of self-organized periodic nanogratings. More discussions will be given about the
front tilt of femtosecond pulses later. In addition, Song et al. suggest that the polarization dependent
fine structures may originate from the polarization-induced bulk damage threshold [63]. We also
investigate that the dependence of the nanostructure length and period on the polarization plane azimuth. Zhang et al. observe that the optical birefringence increases with increasing laser pulse energy,
and then becomes saturated at the laser pulse energy of about 2 lJ [381].
The birefringence of the femtosecond laser induced nanogratings in silica increases with an
increase in inscription scan and energy until it reaches the constant value of 102 above 0.7 lJ
[382]. Birefringent modification can be characterized by two parameters: the retardance and the azimuth of the slow axis, which can be independently controlled during the writing experiment as the
retardance is a function of incident fluence and the azimuth of the slow axis is defined by polarization
[359,364]. Beresna et al. show that the threshold for nanogratings formation is independent of the
writing speed and slightly higher for 1030 nm than that for 515 nm. Strong retardance is generated
above 2 lJ (400 mW) with 1030 nm and 1.5 lJ (300 mW) with 515 nm irradiation. Once the nanogratings are induced, the retardance rapidly reaches a certain saturation value and weakly depends on the
irradiation laser power, and this is confirmed by Poumellec et al. [383]. In contrast, Richter et al. report
that the retardance increases with increase of incident pulse energy from 0.5 lJ to 0.8 lJ owing to an
elongation of the modified region, which leads to an increase of the birefringent volume, confirmed by
the recent report [377,378,384]. In addition, retardance slowly decreases with increase of the writing
speed and the incident pulses, and this is contrast to a previous study [383]. Increasing repetition rates
leads to decrease of retardance [384]. A retardance decrease is observed at average power higher than
800 mW attributed to the Rayleigh scattering of the stronger material damage [359]. The polarization
dependent losses are introduced by nanopores observed inside the nanogratings with characteristic
dimensions of few tens of nanometers [371,382]. The measured retardance dependence on the wavelength reveals a steady increase in the spectral region from 200 to 680 nm and a long plateau tail from
680 to 2100 nm. Poumellec et al. report that the retardance is proportional to the length of the

194

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

Fig. 21. (A and B) SEM images of self-organized nanogratings in the transversal cross-section of written lines with varied
polarization plane azimuth. The writing direction of line is along S in (A), and along S0 in (B). In these two cases, the samples
move in the same direction with light propagating in opposite directions (referred to:  and ). The red bar denotes the
orientation of nanograting. (C) The pulse energy was increased to 2.2 lJ to get a better pattern and the writing direction of line
was along S0 . (DF) Optical microscope images, two lines of each group are independently written with opposite scanning
directions. (D) The illumination was linear polarization light. (E) Cross-polarizer images recorded the birefringence
phenomenon in the same region in (D). (F) The intensity variation of birefringence signal from the written lines with varied
polarization plane azimuth. (Reproduced with permission from [380].)

nanogratings and the number of planes within the affected volume [383]. Birefringence and index
contrast increase linearly with the pulse energy until reaching a saturation value, i.e., the threshold
for self-focusing [377,378]. The polarization contrast intensity of nanogratings decreases with increasing the annealing temperature up to 1150 C.
3.1.1. Mechanism of nanograting formation
The physical mechanism responsible for the generation of nanogratings induced by femtosecond
laser irradiation has not been fully understood. But inspiringly, many researchers have been making
great efforts to do so. Some physical models have been proposed in the past decades
[19,74,103,336,365]. Shimotsuma et al. propose that there is a coupling between the inscription laser
light and the electric field of the induced bulk electron plasma wave propagating in the plane of light
polarization, when the electron plasma is produced after MPA. The initial interaction can be invoked
by inhomogeneities induced by electrons moving in the plane of light polarization. The coupling is
enhanced by a periodic structure created via the interference between the incident light field and
the electric field of the bulk electron plasma wave, leading to the periodic modulation of the electron

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

195

plasma concentration and the structural changes in glass. An exponential growth of the periodic structures perpendicular to the light polarization is caused by a positive gain coefficient for the plasma
wave, resulting in frozen-in of periodic nanogratings in the materials [19,364]. This concept is consistent with the theoretical calculations. In addition, as the plasma electrons are created in the process of
breaking of SiOSi bonds, SiSi bonds, nonbridging oxygen-hole centers (NBOHCs) and interstitial
oxygen atoms (Oi) are produced, which are evidenced by the photoluminescence and ESR spectra.
Such oxygen atoms are mobile and can diffuse from the regions of high concentration, even to form
O2, leading to low oxygen concentration in the nanogratings, and this is confirmed by the presence
of nanoporous in the nanoplanes inscribed in the silica [371,382]. Double femtosecond laser pulses
experiments have also verified this suggestion [364]. Unfortunately, the coupling mechanism mediating the collaborative action of subsequent laser pulses is still not fully clarified. In addition, the actual
plasma lifetime (about 150 femtoseconds) is many orders of magnitude too short to explain any influence on the interaction with the subsequent pulse [377]. Furthermore, the Nolte group suggests that
two distinct mechanisms are involved: short lived self trapped excitons (STEs) and more stable
dangling-bond type defects at longer time scales [378]. Excitons are produced when the nonlinear
absorption of a femtosecond laser pulse has generated free electrons within the focal volume resulting
from the interactions with small lattice distortions. NBOHCs can be generated due to the nonradiative
decay of STEs with life time of about 400 ps, indicating that the enhancement of the grating formation
at small pulse duration is associated with the presence of STEs. Excited STEs can act as the seeds for
collisional ionization and enable an increased absorption of the subsequent pulses, resulting in more
energy available to promote nanograting formation [385]. When the pulse separation exceeds the life
time of the STEs, the resulting point defects will play a crucial role in this process.
Taylor et al. propose a nanoplasmonic model based on the understanding of the formation and orientation of nanoplanes in terms of local field enhancement and this is in agreement with the modulation results [74,101,367,386]. In this case, nanogratings can be referred to as a set of oriented cracks
and self-organization emerges due to cumulative effect and, therefore it likely requires low ionization
rates [386]. As glass is in a disordered metastable solid state, it always possesses microscopic heterogeneities (due to local chemical and structural variations or density fluctuation or actual voids or gas
inclusions), which could lead to an inhomogeneity in the nonlinear response [386]. Therefore, ionized
hot-spots can be created by the focused high power femtosecond laser pulses in transparent dielectrics due to localized inhomogeneous multiphoton ionization at defects or color centers. These hotspots generated by a single laser pulse can evolve into spherically-shaped nanoplasmas over several
pulses caused by a feedback mechanism based upon memory of previous nonlinear ionization
(Fig. 22A) [367]. This memory effect leads to an increase in the local ionization rate and thus enhances
ionization of material on the following pulse. This is accomplished without producing any significant
material damage that reduces the low intensity light transmission through the sample. These underdense nanoplasmas with electron plasma density Ne smaller than the critical density Ncr grow during
successive pulses irradiation under the influence of the laser field. Field enhancement at the edges of
the nanoplasma leads to asymmetric growth of the initially spherically-shaped plasma, in the direction perpendicular to the laser light field polarization (Fig. 22A), to create ellipsoidally-shaped plasmas, and finally plasma disks when Ne approaches Ncr and the asymmetric growth accelerates. The
evolution of many nanoplasmas into plasma disks is demonstrated schematically in Fig. 22BE. Field
enhancement at the boundary of the nanoplasma disks promotes their merging into nanoplanes
(Fig. 22D). The growth from the edges is so strong that when one row of nanoplanes is partially
inscribed over, the existing planes coherently link with the new planes to force them to line up
together, resulting in the growth of the planes with successive laser pulses. The electron plasma density inside nanoplanes can exceed the critical density during irradiation. As a result, the planes become
quasi-metallic and subsequently influence light propagation of the laser pulses and interference of
scattered and incident light will emerge. An array of planar waveguides can be generated near the
top of the carrot-shaped modified zones. The lowest order optical mode whose field distribution reinforces the growth of the nanoplanes into quasi-metallic waveguides dictates that the planes can
assemble not closer to each other than half of the femtosecond laser wavelength in the medium.
Therefore, well-defined planes form first at the top of the carrot structures and eventually grow
to fill the entire modified zone. In addition, due to destructive interference of scattered and incident

196

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

Fig. 22. (A) Asymmetric filed enhancement at two locations on a nanoplasma under the influence of the laser electric field E.
The relative permittivity e is the ratio of the real part of the nanoplasma permittivity to the dielectric function of the medium
that surrounds the nanoplasma. (BE) Evolution of nanoplasmas into nanoplanes. Randomly distributed underdense
nanoplasma droplets (droplet size is a few tens of nanometers) grow asymmetrically in the presence of the laser field over
hundreds of laser pulses to become ellipsoidal and finally flatten and merge to become micrometer-sized nanoplanes.
(Reproduced with permission from [101].)

light, ionization is suppressed directly adjacent to each plasma plane and is enhanced at a distance of
about k/2n and this is consistent with the experimental observation [386].
For the above two models, the condition above or near critical plasma concentration is necessary.
However, the applied laser energy may be not sufficient to satisfy the required plasma temperature
and density in some cases [336,365,387,388], so that one can hardly expect significant plasmon features responsible for nanograting formation. The Kazansky group proposes an excitonpolariton-medi
ated self-organization mechanism for fabrication of nanogratings in silica glass under intense femtosecond light irradiation [365]. Excitonpolaritons are mixed lightmatter quasiparticles responsible
for fascinating nonlinear optical effects in semiconductors including polariton lasing and Bose
Einstein condensation, and exhibit potential applications in so-called polaritronics. Interference and
dipoledipole interaction of polaritons lead to formation of gratings of dielectric polarization. Due
to an ultrafast exciton self-localization into a quasicrystal structure, the polariton gratings remain
frozen in the transparent matrix and a permanent 3D imprint of excitonpolariton gas is formed.
The period of this grating is observed to increase with the distance from the front of the laser induced
structure, starting from a value close to the laser wavelength due to the dependence of the splitting
between two interfering excitonpolariton modes on the group velocity of the excitonpolaritons.
This is consistent with theoretical simulation results and other studies [68,74]. The critical concentration of excitonpolaritons for periodic excitonpolariton nanograting formation is determined to be
about 1015 cm3, below which no nanogratings can be formed and induced modification does not
exhibit anisotropy. This might explain phase transition from type I modifications to type II.
Excitonpolariton interaction mechanism may clarify the nanograting formation at low densities of
low energy free electrons. At higher incident fluence, the excitonpolariton interaction should be
heavily screened by the electron plasma as it begins to efficiently absorb laser energy and to develop
secondary electrons. The nanoplasma model of nanograting formation may be established between
the two extremes, plasma wave interference with laser light and excitonpolariton interaction, with
evolution to the critical density at nanoplane sites.
In addition, Cheng et al. have accessed to the snapshots of morphologies in the laser-affected
regions in a porous glass, which demonstrate the evolution of the formation of nanogratings with
an increasing number of laser pulses [389]. Excitation of plasma waves at the interfaces between
zones modified and unmodified by the femtosecond laser irradiation is suggested to induce a periodic
distribution of the defects (e.g., nanovoids), which is evidenced by the experimental observation.

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

197

Afterward, nanovoids emerge in low viscosity phases since the free void can easily expand due to
lower strength of the topological constraints. Finally the nanovoids develop into nanogratings
[96,389]. Besides, the formation of nanopores or nanovoids may also be driven by an anomalous thermodynamic behavior of silica glass [390]. The free volume of silica glass is larger (low density) than
that of the liquid silica (high density). Canning et al. suggest that a rapid cooling of the molten state
generated by femtosecond laser irradiation hinders crystallization of the liquid silica, and results in a
highly metastable solid state (corresponding to the softened state of silica) in terms of potential
energy. This metastable state is associated with an unusually dense, highly strained liquid-like structure, and hence should undergo a volume expansion rather than contraction. Subsequently, enormous
negative pressure occurs, which triggers the formation of nanoporous or nanovoids [390]. However,
there are still some unanswered questions. For example, is the difference in the silica volume (density)
such that the internal potential energy before and after irradiation high enough to initiate this
behavior?
Anyway, much work should be done to get the full picture of the mechanisms of nanograting formation. Especially, further research is needed to provide a more detailed description of the thermodynamic and hydrodynamic aspects, including the free-carrier dynamics, carrier heating, density, and
temperature-dependent changes in the collision frequency [386]. Such picture will help to open
new ways for controllable fabrication of sub-wavelength nanogratings in various transparent materials for applications in optical technologies [336].
3.1.2. Applications of femtosecond laser induced nanogratings
Due to the unique properties of the nanogratings inscribed in transparent materials by femtosecond laser irradiation, various kinds of photonic devices and applications are demonstrated based on
the nanogratings. The Kazansky group reports the first diffractive elements fabrication by femtosecond direct writing: a strongly birefringent Fresnel zone plate, which is an attractive optical component
due to their compactness and focusing abilities [391]. Alternate zone rings can be produced directly by
inducing a local refractive index modification of the order of about 102. The embedded zone plates
exhibit efficiencies that vary by a factor of about 6 for orthogonal polarizations. Directly written anisotropic microreflectors are also demonstrated, which strongly reflect blue light only along the polarization axis of the incident writing beam [361]. This anisotropic effect is caused by a periodic
modulation of refractive index of amplitude Dn  102 with a characteristic period K of 150 nm over
a spot size of about 1.5 lm. Polarization-sensitive diffractive nanogratings are generated in bulk silica
glass by Beresna and Kazansky [392]. Manipulation of the induced birefringence is achieved by
controlling the polarization azimuth of the writing laser beam. The diffraction-based circularpolarization beam splitter is demonstrated and the behavior agrees with the theoretical prediction.
They suggest that the operation of space-variant subwavelength gratings of beam polarization converters can extend to the visible range. Sensitivity of polarization nanogratings to incident circular
and radially polarization is exploited for polarimetric measurements [393]. Combined with a linear
polarizer, a CCD sensor and the nanograting arrays, a polarization imaging device is fabricated, which
enables real-time polarimetric measurement of circular and radially polarization beams (Fig. 23)
[393]. The imaging device can potentially be used for medical applications, such as skin diagnostics.
A space variant polarization converter is fabricated for generation of optical vortices with radial or
azimuthal polarization in silica [363]. The converter allows switching from radial to azimuthal polarization by controlling the handedness of incident circular polarization.
As the birefringence caused by the nanogratings can be described by two independent parameters:
retardance and slow axis direction, it is possible to control these two parameters with exposure and
polarization of irradiation independently, and this is demonstrated by simultaneously recording of
two data sets [364,365]. The 4th dimension of optical data storage, i.e., an orientation of slow axis,
can be recorded with a resolution of several degrees in the range from 0 to 180. In addition, the retardance can be expanded range over 100 nm and determine eight discrete levels. Therefore, fivedimensional optical data storage is possible (Fig. 24) [364,365]. Furthermore, since the nanogratings
can be erased by another femtosecond laser and new nanogratings with different orientations can
be created, the optical data are rewritable [101,364]. Shimotsuma et al. show that the storage capacity
is 300 Gbit cm3, which is as high as the 10 times of the usual 12 cm BlueRay disk capacity. This

198

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

Fig. 23. Measurement of the radially polarized light and the high-order cylindrical vector (CV) beam with the high resolution
waveplate mask. (A) and (B) Image of the mask while illuminated with the high order CV beam. (C) and (D) Measured
polarization distribution of the radially polarized light and the high-order CV beam. (Reproduced with permission from [393].)

simple and easily available technique exhibits the advantage over other types of optical memory, e.g.,
holographic memory, fluorescence 3D memory, and spectra hole-burning (SHB) memory.
Polarization selective computer-generated holograms (PSCGHs) for visible light operation are
fabricated in glass by a femtosecond laser [366]. Arrays of tailored micro-waveplates are created by
controlling the phase retardation and orientation of nanogratings embedded in fused silica. A detour
of each micro-waveplate, combined with the orientation of its principal optical axis, simultaneously
realizes a different phase function for each polarization. PSCGHs are attractive for integration with
other free-space and waveguides embedded in glass.
Fiber Bragg gratings (FBGs) can be fabricated by direct femtosecond laser writing through a
point-by-point way in fibers and glass system [229,394,395]. These FBGs have been used as optical
waveguides, waveguide retarders, polarization beam splitters, optical edge filters, and temperaturecompensated fiber-optic 3D shape sensor [229,321,323,396400].
A polarization-dependent light attenuator is fabricated by the authors group using the femtosecond laser induced self-organized nanogratings, i.e., a plane consisting of lines written with a laser
polarization plane azimuth fixed at 45 [381]. The powers of the zeroth-, first-, and negative firstorder diffraction lights are measured after the probe laser pass through the fabricated structure from
the normal incidence. Rotating the polarization plane azimuth of the probe laser with a half-wave
plate, the detected power varies regularly with a period of 180, which can be deduced to the birefringence effect of the nanogratings with a period far smaller than the laser wavelength.
Regularly arranged hollow nanograting structures are written inside porous glass in a controllable
manner using a linearly polarized femtosecond laser [401]. Channels of nanofluidic are formed by
incorporating the single nanovoids into smooth continuous structures in the porous glass. After further annealing, the nanochannels can confine and transport fluorescent solutions without leakage
and clogging. Nanochannels with narrower widths are achievable by optimization of the pore sizes,
providing useful tools for applications ranging from nanofluidic research to biomolecular analysis.
Nanochannels can also produced by chemical etching of nanogratings for nanofluidic applications
[55,101,372,402,403].

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

199

Fig. 24. Images of the Small World Map taken with optical (A) and polarization (B-azimuth angle, C-retardance) microscopes.
The structure is imprinted in silica glass using femtosecond laser beam modulated with LCOS-SLM. Actual size of the structure is
3.4 mm  1.8 mm. (Reproduced with permission from [364].)

3.2. Periodic nanovoid arrays


Spatially modulated refractive index and structure changes can be generated in the bulk transparent material induced by focused femtosecond laser irradiation. High fluence leads to the localized
formation of lower-density cavity-like structures, nanovoids, depressed structure surrounded by
compacted matter through the microexplosion [92,93]. Periodic nanovoid structures in transparent
materials are vital for nanophotonics such as photonic crystals and nanogratings due to large refractive index contrast (Dn). Kanehira et al. first report fabrication of periodic nanosized voids inside
borosilicate glass by a single femtosecond laser pulse writing [21]. The spherical nanovoids are aligned
spontaneously with a period along the propagation direction of the writing laser beam. The neighboring two nanovoids are independent of each other. Both the diameter of the voids and the period
decrease gradually with the closing of the bottom surface of the glass sample, and approach limiting
values at a distance of about 90 lm from the bottom surface. This length of the periodic part at a range
of about 90 lm from the bottom surface is independent of the pulse energy of the femtosecond laser
and can increase to 130 10 lm with an increase of incident pulse numbers. The void period can be
controlled from 1.6 to 3.2 lm and the void size increases from 600 nm to 1.0 lm as the pulse energy
varies from 10 to 40 lJ with 1000 pulses irradiation. Toratani et al. investigate the effect of NA on the
resultant structure of nanovoids with NA varied from 0.25 to 0.9 [404]. With the higher NA lens, the
void is observed to be longer, while with the small 0.25 NA lens only the spherical void is created
apparently due to the small energy above the void formation threshold energy of 0.2 lJ, which

200

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

Fig. 25. Laser beam with pulse energy of 25 lJ is focused at 909 lm beneath the surface. From left to right, the irradiation time
is (A) 2 s, (B) 1 s, (C) 0.5 s, and (D) 0.25 s, and the pulse number launched into the sample is 2000, 1000, 500, and 250,
respectively. (E) The length of the void array vs the laser focal depth beneath the entrance surface. (Reproduced with permission
from [406].)

contradicts the previous study, where high NA is necessary for the formation of nanoviods [404,417].
Since then, the authors group has fabricated periodic nanoviod arrays in various materials including
borosilicate glass, fused silica glass, CaF2 crystal, Al2O3 crystal and SrTiO3 crystal [405412]. The period
increases with an increase of the pulse numbers [405,413]. Hu et al. fabricate 2D quasiperiodical void
structures and show that the length of the nanovoid array first increases and then deceases with an
increase of the focal depth (Fig. 25), and this phenomenon is observed in different modified materials,
being attributed to the change of effective energy used to induce nanovoids formation [406,409,411].
The period between consecutive voids increases as scanning speed increases [414,415]. Thomas et al.
report that the periodicity of the nanovoids is readily controlled by the energy influence and the periodicity patterns written at fixed fluence but with opposite writing directions is significantly different
[416]. Two nanovoid arrays in perpendicular directions, parallel (Z axis) and vertical (X axis) to the
laser propagation direction, respectively, are produced simultaneously using single femtosecond laser
beam inside CaF2 crystals [417]. We find that high laser power and small value of D (defined as the
distance from the focal point to the sample profile on ZY plane) are the dominant factors for achieving regular and long void arrays along X-axis. Only when the laser pulse energy is higher than 50 lJ
and D is less than 100 lm, two void arrays can be created at the same time. We show that when
the laser focused is far from the sample profile, the light spot will be roundish. Therefore, the void
array along X axis is induced by the light pass through side facet of the sample.
3.2.1. Void formation mechanism and various works
Though considerable research has been done to fabricate periodic nanovoid structures in transparent materials, so far the underlying formation mechanism is still under debate. Kanehira et al. find that
in order to create the self-organized and periodic void structures, the femtosecond laser beam must be
focused at a deep position near the bottom surface of the glass and the filament path must also reach
the bottom surface [21]. Therefore, they suggest that under the irradiation of the femtosecond laser
pulses, the refractive index is changed at the focal point, and then the light filament propagates
toward the bottom surface of the glass. When the temperature of the filament line core is high enough,
this core gets highly absorbing. Then microexplosion occurs firstly around the bottom surface to create
a nanovoid due to the lower breakdown threshold of the glass at the surface. When the next femtosecond laser pulse propagates from the focal point, it is trapped in the heated region around the former
void leading to a new high temperature region and the formation of the next void. This process is
repeated many times and the periodic nanovoid arrays formation is completed at the focal point. In
this case, the observed void size at the focal point is different from that near the bottom surface
assigned to differences in the absorbed energy. As the pulse energy of the incident laser is increased,
heating can occur even at the regions far from the void because of thermal diffusion, which induces an
increase in the void period.

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

201

In contrast to the above observation, the authors work implies that the periodic nanovoids are not
necessarily generated first near the bottom surface of the sample and void arrays can also be formed
inside glasses without the filamentation path approaching the bottom of the glass. Therefore, we propose a different mechanism responsible for the formation of periodic nanostructures. Sun et al. suggest that a standing electron plasma wave (EPW) created by the interference of a femtosecond laser
driven electron wave and its reflected wave is responsible for the formation of the periodic void arrays,
similar to the model describing the generation of nanogratings in the glasses and nanoripples on the
surface [405]. The free electrons generated by the leading part of the femtosecond laser pulse can
interact with the successive pulses, forming an EPW. When the laser pulses propagate deeply into
the glass, the losses of the pulse energies combined with the diffraction will reduce the fluence, resulting in the termination of EPW production. Therefore, a large plasma density gradient will be created
after a certain propagation distance of the incident light. When the plasma wave propagates into this
region, it will be reflected and then interfere with the forward-traveling wave, creating a standing
wave pattern. As the optical breakdown in glass is sensitive to the free electron density, the standing
EPW can induce periodic microexplosions leading to the periodic void array structure. In this case, the
electron density required for forming an array of a period of 2 lm can be estimated to be about
1019 cm3, which is consistent with the previously reported results in femtosecond laser irradiation
of dielectrics [387,388]. In addition, the periodic refocusing phenomenon resulting from the competition between the defocusing effect induced by the diffraction and the self-focusing effect caused by
both the Kerr self-focusing and the parabolic index profile in the transverse direction is also proposed
to generate periodic void array, and this is verified by the agreement between the Zernike-type positive optical phase-contrast microscopy (PCM) and the Fresnel propagation results (Fig. 26A and B)
[406,407,409,411,413,418,419].
A physical model in combination of the nonlinear effects of femtosecond laser pulses and the interface spherical aberration effects is employed to analyze the mechanisms of periodic nanoviods production by Song et al. [410412]. Interface spherical aberration can be induced by the refractive
index contrast between environment medium and sample. A simulation of fluence distribution of
the light field is conducted based on a nonparaxial nonlinear Schrodinger equation incorporating
the effect of interface spherical aberration. The simulation result shows that the laser fluence around
the Gaussian focal point exhibits quasiperiodic alternations between maximum and minimum
(Fig. 26E). Since the electron density generated by multiphoton ionization is nearly proportional to
|E|6, more electrons are produced at the local maximums than that at the local minimums. As a result,
a serial of microexplosions occur at the positions with the local maximum laser fluence. Finally, the
ablated materials are ejected to the surrounding volume and then a string of voids are left at the positions with the local maximums. Reasonable agreements between the experimental results and the
theoretical analyses confirm the interface spherical aberration to be the origin of the self-formation
of the void array. Therefore, it is possible to control the formation of the self-organized void array
by tuning the amount of the spherical aberration.
Terakawa et al. have systematically investigated void structures fabricated inside various kinds of
transparent materials by femtosecond laser irradiation [420]. They show that a long void array can be
produced in a material with a low critical power for self-focusing. Both the fabrication process and the
void shape probably depend on the heat effect and fluidity of the surrounding structure which is thermally melted during irradiation. The coefficient of thermal expansion plays an important role in fabricating a precise void with a clear boundary, based on the comparison of void shape in fused silica to
those in other materials.
3.2.2. Applications of femtosecond laser induced periodic nanovoids
The technique of fabricating periodic nanovoids in transparent materials by femtosecond laser has
promising applications in photonic crystals. 2D and 3D quasiperiodical structures can be easily fabricated in glasses, crystals, and other transparent materials [21,415,420].
A single-mode waveguide consisting of 90 bend void arrays in fused silica is fabricated by femtosecond laser serving as an optical mirror first [421]. The light propagating through the waveguide
is reflected by the refractive index difference of a void array reflector and the interference of the scattered light from the voids is not optimally designed. The void arrays are used to constitute waveguides

202

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

Fig. 26. (A and B) PCM pictures (top) of single pulse irradiation effects in silica and BK7 and the comparison with the Fresnel
propagation results (bottom). The dashed lines show the correspondence between the position of the dots and the calculated
fluence peaks. (C and D) Horizontal axial cross-section of the PCM picture (red symbols) and of the numerical results (black line)
for silica and BK7. (E) The simulated fluence distribution around the focal point of the objective lens with NA = 0.9 in the
presence of both interface spherical aberration and nonlinear effects. (Reproduced with permission from [410,420].)

limited by void-like damage zones with very loose coupling among adjacent guides, thus allowing the
excitation of a single one. This shows the possibility of using created void-like structures for both the
fabrication of integrated optical devices as well as for the control of previously induced refractive
index change [422]. The corresponding diffractive patterns of microvoid arrays are demonstrated in

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

203

three-dimension, indicating that the fabricated periodic microvoid arrays have a good performance as
diffractive beam splitters [415]. Though according to the Babinet principle, the void array with normal
incidence has similar diffractive pattern to 2D grating, the actual diffractive patterns are similar to the
diffractive patterns of 1D gratings, rather than 2D gratings. Wang et al. find that the void arrangement
has a slight horizontal deviation due to the moving and positioning precision of moving stages, and in
vertical direction the voids are distributed relatively uniformly. Terakawa et al. propose and design a
MachZehnder interferometer composed of optical waveguides and photonic crystals by use of void
array to demonstrate the potentiality to fabricate compact optical circuits, as shown in Fig. 27
[420]. Simulation results of the optical propagation in the MachZehnder interferometer indicate that
the photonic crystals using periodic void arrays have potential to fabricate compact optical circuits.
Guidance of 632 and 830 nm light through the void arrays based wave-guides are demonstrated in
BK7 glass [416]. The cross-section of the pattern shows the evidence of a double structure, with inner
and external modified regions that take the form of a comet, and the formation of this structure is
assigned to the flow of glass and some form of viscoelastic deformation.
A radial polarizer based on light refraction on transparent isotropic spheric nanovoids is also produced by the Kazansky group [423]. They demonstrate theoretically and experimentally that the circularly polarized light impinging on the spheric nanovoid arrays produces double charged optical
vortex. Therefore this technique offers a practical alternative to conventional radial polarizers and a
flexible way to fabricate dense arrays of optical vortex generators which could be used for integrated
quantum optics and optofluidics.
Ahmed et al. propose an idea of fast cutting a display glass plate that is pre-processed by micromachining single shot rear-surface and internal void arrays aligned on working plane prior to glass cleaving [414]. Rear surface glass cutting is demonstrated through working line and post glass cleaving. An
additional void array down to the front surface on the same plane increases reliability of glass cutting.
3.3. Migration of ions
The success of constructing 3D micro optical components or devices inside transparent materials is
highly dependent on the ability to modify materials local structures and properties. Especially, the
space-selective manipulation of element distribution is highly desirable as most of optical properties
of the materials such as refractive index and luminescence are related to element distribution
strongly, which is also highly challenging. Space-selective manipulation of element distribution in
transparent materials has been achieved by the Qiu group and Miura group recently, induced by
the high repetition rate (250 kHz) femtosecond laser [20,424].
A spherical modified zone with the size much larger than the laser focus size can be observed in
glass after femtosecond laser irradiation. Each zone has an eye-like area with the size nearly equaling
to the focus size in the center [424]. Micro-Raman spectra show that the ratio of the integrated intensity of the two peaks at 806 cm1 (assigned to the localized breathing motions of the oxygen atoms

Fig. 27. Schematic of model for simulation of optical propagation in a MachZehnder interferometer consisting of a void array
calculated by FDTD method. (A) Obliquely-directed view of the model. (B) Top view of the model. (Reproduced with permission
from [420].)

204

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

inside the boroxol ring) and 774 cm1 (assigned to the six-membered ring with one or two BO4 tetrahedra) first decreases and then increases with increasing the distance from the laser modified zone
center. Electron dispersive X-ray spectra reveal that a portion of Na+ and O2 ions migrate from the
vicinity of focal point to the boundary of the laser modified area after the femtosecond laser irradiation. More work on the multicomponent glass system characterized by EPMA shows that the relative
concentrations of network formers of the glass are higher in the central area and lower in the periphery of the modified region compared with the unirradiated areas, and the relative concentrations of
network modifiers or doped ions are as opposed to that of network formers, migrating to the boundary
of the modified rings [20,189,425428]. Besides the cross-section perpendicular to the laser propagation axis, similar phenomena are also observed in the cross-section along the laser propagation axis
[429,430]. By using water as well as 1-bromonaphthalene as immersion liquids, the opposite elemental distribution is observed after femtosecond laser irradiation, which is assigned to the interfacial
spherical aberration effect [430]. The diameter of the laser modified region is reported to be a function
of the irradiation conditions [426,428]. At first, the diameter of the laser modified region grows rapidly
with increasing the irradiation time, due to heat accumulative effect. Then, the size of the structure
remains nearly constant after a threshold of irradiation time. A simultaneous control over the precipitation of multiple crystalline phases (Ga2O3 and LaF3) and active ions (Er3+ and Ni2+) migration is also
realized, as discussed above [183]. Furthermore, Sakakura et al. report a method for controlling the
shape of the elemental distribution more flexibly by simultaneous irradiation at multiple spots using
a spatial light modulator (Fig. 28B and G) [431,432]. The accumulation of thermal energy is induced by
focusing 250 kHz femtosecond laser pulses at a single spot inside the glass, and the transient temperature distribution is modulated by focusing 1 kHz laser pulses at multiple spots in the same glass. The
shape of the resulting modification can be controlled as be triangle and square in glass with different
compositions.
No element migration is observed using 1 kHz femtosecond laser irradiating the same glass, indicating that the heat accumulation effect of the high repetition rate femtosecond laser play an important role in this process [424]. When the high repetition rate femtosecond laser is focused inside a
glass sample, the energy deposit rate is so high that the heat accumulation occurs and the temperature
at the focal point can reach higher than 3000 K which enables localized melting and bond breaking. A
sharp temperature gradient produced by the thermal diffusion process drives diffusion of ions in the

Fig. 28. Transmission optical microscope images of the modification in SiO2B2O3Al2O3CaO glass during (left) and after
(right) irradiation with 250 kHz femtosecond laser pulses at the center (A) and multispot irradiation (B, 250 kHz at the center
and 1 kHz at the surrounding four spots). (C) and (D) Distributions of elements by EPMA in the modifications after single spot
and multiple spot irradiation, respectively. Schematic illustration of elemental distribution changes and heat modification after
femtosecond laser irradiation at a high repetition rate in the case of (E) 250 kHz irradiation at the center and (F) 250 kHz
irradiation at the center and 1 kHz irradiation at the surrounding four points. (G) Triangle, square, and hexagonal shapes of
molten regions produced inside an alumino-borosilicate glass by irradiation with 250 kHz femtosecond laser pulses at the
center and 1 kHz femtosecond laser pulses at three, four, and six spots, respectively. (Reproduced with permission from [431].)

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

205

laser modified region [429,433]. Because of the higher diffusion coefficient, glass network modifiers
will move out of the central area of the laser irradiated region to the low temperature zone. Since glass
network formers have a stronger binding energy with oxygen than the network modifiers, they are apt
to migrate to the hot region filling the vacancies resulted from migration of other ions. The ions with
higher diffusion coefficient tend to move out of the central area with high temperature to diminish the
concentration gradient. The diffusion coefficient is determined by the local temperature (T), and the
ion activation energy (Q), described as Eq. (3.1) [424,426,428].

D D0 eQ=RT

3:1

In this dynamic process, at first, the elements are forced to diffuse away from the center of the laser
affected zone due to the laser induced temperature and pressure gradient, and then the elemental distribution is frozen though some of elements may diffuse toward the center of the laser affected zone
due to concentration gradient. Therefore, the relative concentrations of network modifiers of the glass
are higher at the margin of the inner structure and lower in the central area, while the relative distribution of network formers has an opposite tendency. This proposition is confirmed by the numerical
simulation results [431,433,434]. In addition, a simulation of the mean diffusion length of molten glass
reveals that the transient diffusion of ions under heat accumulation and repeated temperature elevation at multiple spots are responsible for the controlled shape of the distribution [431].
Accompanied with the ion migration, the optical properties of the local zone are also modified,
enabling us to fabricate multifunctional optical devices in glass. Due to the relative higher concentration, the fluorescence intensity increases by 20% in the Eu3+ enriched region in the glass compared
with that for the original glass [425]. Due to the ion migration assisted inscription of high refractive
index contrast, fabrication of waveguiders is demonstrated [311,432,435]. More applications will be
developed, such as shaped waveguides.

3.4. Quill writing and nonreciprocal writing


Since the first demonstration of laser micromachining, there are usually two common beliefs about
the interaction between femtosecond laser and matter. One belief is that a Gaussian femtosecond laser
beam interacting with an isotropic medium can generate only centrosymmetric material modifications. Another is that the photosensitivity and corresponding light-induced phenomena do not change
on the reversal of light propagating direction in a homogeneous medium. However, the quill and nonreciprocal writing experiments indicate that this is not always true. Recently, two remarkable phenomena have been reported: a quill writing effect revealing strong dependence of femtosecond
laser induced phenomena in glass on the orientation of writing direction relative to the direction of
the PFT [16,352,436,437] and nonreciprocal photosensitivity manifesting itself as dependence of femtosecond laser induced phenomena on light propagating direction in the non-centrosymmetric crystal
[17]. In these effects, the dependence of the imprinted structures on the laser beam polarization has
been established, which originates from the intrinsic anisotropy of the experiment associated with the
beam movement. Anisotropic photosensitivity of an isotropic homogeneous medium under uniform
femtosecond laser irradiation is also observed, characterizing that the modification of glass depends
on the mutual orientation of the light polarization azimuth and the PFT [438]. The Kazansky group
shows that the directional asymmetry of femtosecond laser induced modifications in glass (quill
effect) is independent of the orientation of a medium or the direction of light propagation and can
be controlled by the PFT. In contrast to the glass case, the directional dependence of writing in lithium
niobate is dependent on the orientation of a crystal with respect to direction of the beam movement
and the light propagating direction, and a PFT is not necessary [16,17,352,436]. Groups of line structures are also written in a LiNbO3 sample using a picoseconds laser system (1064 nm wavelength, 9 ps
pulse duration, 250 kHz repetition rate), but this does not give evidence of the nonreciprocal writing
phenomenon. In contrast, quill effect, which is featured by the dependence of the laser damage on the
direction of writing, occurs also for picosecond laser machining [384]. This indicates that ultrashort
pulses are necessary for the creation of the nonreciprocal writing, possibly owing to the higher

206

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

intensities that can be achieved with femtosecond pulses without inducing strong damage in the
sample [17].
3.4.1. Quill writing, anisotropic photosensitivity and their mechanisms
A remarkable phenomenon in femtosecond laser processing of transparent materials has been
observed by the Kazansky group, i.e., a change in material modification by reversing the writing direction. The writing technique is something like writing with a quill pen with the anisotropic tip shape
(namely quill effect). The observed phenomenon is attributed to the breaking of the laser beam symmetry due to the presence of PFT (Fig. 29) [16]. Based on the experiments of characterization and control of PFT, they provide the experimental evidence that the tilt in intensity front of an ultrashort pulse
can imprint itself directly in laser induced material modifications [352]. Since then, PFT is considered
as one of the laser parameters defining the lightmatter interactions in material processing and many
studies confirm the quill writing effect causing writing direction dependent structures
[352,380,436,437]. The lines written in both directions at low energies are the same (150 femtosecond
pulse duration, 250 kHz repetition rate, 800 nm wavelength, 0.55 numerical aperture) [16]. As the
irradiation energy increases to above some threshold, the direction dependence in the written lines
is observed clearly, particularly in the birefringence of the lines. This dependence can also be seen
in the morphology (texture) of the lines written in opposite directions, with a line written in one direction being rougher than a line written in the reversed direction. Another intriguing result is that lasers
with polarizations perpendicular or parallel to the movement of the sample create different textures
in the modified zone in one direction and the same textures when writing in the opposite direction,
and these phenomena are also observed in the SEM images of the cross-sections of the lines and confirmed by Gecevicius et al. [439]. Especially, the nanogratings with the period of about 300 nm can be
created only in the initial part of cross-sections of lines written in one of two directions, followed by
one with a collateral damage due to thermal effect (Fig. 29B, top). In contrast, nanogratings along the
direction of light polarization with the period of about 250 nm together with the additional periodicity, along the direction of light propagation, of about 720 nm appear in almost entire cross-sections of
the lines, written in opposite direction (Fig. 29B, bottom). These lines demonstrate no evidence of the
collateral thermal damage and much stronger birefringence [16]. Anisotropic bubble formation and an
unusual transition from the regime of bubble formation to self-organized formation of birefringence
are also observed at high pulse energies and can be controlled by adjusting writing directions
[352]. The quill writing effect depends strongly on the focus depth of the laser irradiation beneath
the sample surface and the scanning speed [352,436]. Scanning direction also affects the etching rate
of the inscribed lines and luminescence of the femtosecond laser induced color centers [440,441].
The PFT can be characterized using a GRENOUILLE device and tuned by using the pulse compressor
or the spatio-temporal focusing with a low numerical aperture or a liquid crystal phase-only spatial
light modulator (SLM) [352,436,442]. Especially, the simultaneous spatial and temporal focusing

Fig. 29. (A) Crossed-polarized (CP) and Nomarski-differential interference contrast (DIC) images of the lines written with
orthogonal polarizations. The difference in texture for two polarizations is observed only for one writing direction. The tilted
front of the pulse along writing direction is shown. (B) SEM images of cross-sections of lines written with polarization
perpendicular to writing direction are also shown. The region of collateral damage is marked with a black dashed line.
(Reproduced with permission from [16]).

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

207

(SSTF) lateral spatial chirping can be used to form a frequency-distributed array of low numerical
aperture (NA) beamlets to enhance the PFT significantly and dynamic control of both the PFT and
the intensity profile of the beam using a SLM can induce or remove directional effects [436].
The presence of a spatial frequency chirp and related PFT is very common in femtosecond laser systems, which can be enhanced in a dispersive media, as in the case of electron plasma close to plasma
frequency, which is produced through multiphoton ionization of glass in the focus of the beam [443].
In the presence of intensity gradients, the charges (e.g., electrons) can be expelled and accelerated by
the ponderomotive force (light pressure) from the region of high intensity and this tends to push electrons in front of the laser pulse, as a kind of snow-plough effect. Therefore, Kazansky et al. suggest
that after femtosecond laser writing, the generated electron plasma will experience the ponderomotive force along the direction of the intensity gradient resulting from the tilt of the intensity distribution [16]. By moving the beam, the ponderomotive force in the front of the pulse will trap and displace
the electrons along the direction of movement of the beam and only in one direction corresponding to
the tilt in the intensity distribution (quill effect), and this is in agreement with the numerical simulation results [336]. In detail, according to the simulation results, the plasma can be excited by the very
front of the laser pulse, which scatters most of the rest of the pulse [336]. As a result, even a small PFT
will give rise to free electron plasma with maximum density slightly shifted with respect to the beam
axis. Asymmetry of the rest of the pulse will be further increased by the asymmetric scattering of the
PFT-generated electron plasma, resulting in the more defect states created off-axis that is important
for excitation by the subsequent laser pulses. Therefore, if the laser beam moves to the direction of
the density maximum created by the prior laser pulses, plasma triggered by the subsequent pulses
starts well before the geometric focus owing to the previously accumulated defects. Then, beam
energy depletion happens due to both the energy expenditure for electron excitation and the plasma
scattering of the rest laser beam, thus leading to reduced energy penetration into the focal region and,
hence, to a soft modification. In contrast, when the beam is directed to the opposite direction, to the
less pretreated material sites, deeper penetration to the focal region can be obtained, accompanied
by more energy absorption and stronger structural modification. As a conclusion, the anisotropic trapping and displacement of the energy and electrons with the movement of the beam influence the
interference of plasma waves, and the creation of birefringence. The periodic structure, with the period of the wavelength of light, along the direction of light propagation can be produced resulting from
the interference between plasma waves and plasma oscillation. Trapping of the electron plasma
damps plasma oscillation and related interference lead to the generation of longitudinal periodic
structure. This mechanism is confirmed by the observation of different textures of modified material
written with light polarizations parallel and perpendicular to the movement of the sample in one of
the writing directions, caused by the difference in boundary conditions for two orthogonal polarizations at the interface of the tilted pulse front along the writing direction.
However, Salter et al. report that using a SLM, the quill writing may be either spatio-temporal in
nature, arising from PFT or time-invariant [442]. They show that PFT may be not necessary for quill
writing, and instead an intensity gradient without PFT is able to do it. Therefore, Poumellec et al. propose a new tentative interpretation based on space-charge built from ponderomotive force and stored
in the dielectric inducing an asymmetric stress field [383].
Another interesting PFT related phenomenon is also reported by the Kazansky group, i.e., anisotropic photosensitivity of an isotropic homogeneous medium under uniform illumination. Modification
of the transparent isotropic glass by intense ultrashort laser pulse depends on the polarization azimuth of the laser beam, referred to as ultrafast light blade, drawing an analogy between strong material modifications produced by ultrashort light pulses polarized along a tilted front and material
cutting with sharp blade [438]. Kazansky et al. attribute this new phenomenon to the anisotropy of
the lightmatter interaction caused by spacetime couplings in ultrashort light pulses. Fig. 30A reveals
that the internal structure of the light modified region is dependent on both the PFT value and the
beam polarization azimuth angle a. The correlation between the orientation of the pulse front with
respect to the polarization azimuth of the writing laser beam and the direction associated with the
strongest induced absorption is shown in Fig. 30B. The position of the bubbles is also correlated with
the direction of the PFT. This experiment unambiguously demonstrates that the tilt in the intensity
front of the pulse is responsible for the observed anisotropic photosensitivity of the isotropic glass.

208

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

Fig. 30. (A) Dots written with orthogonal polarizations and different PFT values. (B) Transmitted light optical microscope
images of modified regions produced by three different orientations of PFT. White and red arrows indicate directions of the
writing laser polarization and the PFT. (C) The red tetragon: pulse with tilted intensity front. An electromagnetic wave with
wave vector k is incident on a planar density gradient produced by tilted intensity front at a nonzero angle of incidence u and
has the electric field vector E lying at an angle a to the plane determined by k. (Insert) Typical measured GRENOUILLE trace. The
black line indicates zero delay. The shift of the trace center in delay axis indicates the PFT. (D) Optical microscope images of
modified regions along beam propagation direction for writing beam polarized along (a = 0) and perpendicular (a = 90) to the
pulse intensity front for exposure time 2 s (bottom) and 8 s (top). (Reproduced with permission from [438].)

Cross-sections along the beam propagating direction exhibit that the strongest modifications take
place in the pre-focus region and that the morphology differs significantly between the region modified by the beam polarized along the PFT and the one perpendicular to it (Fig. 30D). When the beam is
polarized along the PFT, the region of strong coloration locates in the tail and the colored frozen jet is
in the head of the imprinted structure. An extraordinary, colored flower shaped structure is created by
the beam polarized perpendicular to the PFT. This PFT induced anisotropy of the femtosecond laser
matter interaction manifests itself in the modification of the glass generated even by a single shot,
implying that it stems from electronic rather than thermal mechanism. This finding opens up an
interesting opportunity to control the photon flux interacting with a target submerged into condensed
isotropic medium.
3.4.2. Nonreciprocal photosensitivity in noncentrosymmetric media
In a homogeneous medium, it is commonly believed that photosensitivity and the corresponding
light-induced material modifications are the same when reversing the direction of light propagation.
However, Yang et al. report that in a non-centrosymmetric crystal (LiNbO3), modifications can differ
when an untilted light beam moves in opposite directions within the crystal and, moreover, when
light propagates in opposite directions (as illustrated in Fig. 31A and B), both experimentally and
theoretically [17]. Similar to the quill writing, there is also an energy threshold (about 2 lJ) for the
generation of nonreciprocal ultrafast laser writing.
Fig. 31C and D shows the groups of structures written by the laser beam propagating along the +z
and the z axes of the LiNbO3 crystal, respectively, which reveals the different created structures for

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

209

Fig. 31. Illustrations of non-reciprocal ultrafast laser writing with the laser beam propagating along the z (A) and the +z axes
(B). Quantitative phase microscope images of lines written along the y axis with 2.4 lJ (top) and 2 lJ (bottom) pulse energies.
The lines are written by propagating the laser beam along the +z axis (C) and the z axis (D) of the crystal. (Reproduced with
permission from [17].)

writing directions along +y and y axis of the sample. Furthermore, a mirror change can also be
observed in the structural modifications when the propagating direction of the writing beam is
reversed. In addition, in contrast to the lines written (2.4 mJ) along the +y axis by the beam propagating along the z axis, which show optical damage features (Fig. 31C), no damage is observed in the
lines fabricated with the same parameters by the beam propagating along z axis (Fig. 31D). Yang
et al. confirm that the change in structural modifications between these lines is only created by reversing the light propagation direction of the focused laser with respect to the z axis of the crystal. The
modification features, produced along the +y axis by focusing below the z face of the crystal, can only
be found in the line written along the y axis when focusing below the +z face. A mirror change of the
modifications in two similar structures written along the y axis with the reversal of the beam propagation direction along the z axis can be observed. A similar mirror change of phase profiles is also
observed for the line structures with pulse energy of 2 lJ (bottom). The modification of the crystal
structure is also determined by the orientation of the writing direction with respect to the y axis of
the crystal (the direction of the beam movement). No anisotropy is seen with the beam translating
along the x axis.

3.4.3. Mechanism of nonreciprocal photosensitivity


How does this unusual nonreciprocal femtosecond laser writing happen? In order to investigate the
mechanism of nonreciprocal photosensitivity, theoretical discussions are given by Yang et al. [17]. The

210

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

linewidth increases when the scan speed is reduced, indicating that a heat accumulation effect takes
place during the writing process. Upon femtosecond laser irradiation, a heat current J (Eq. (3.2)), generated by the ponderomotive force and the photon drag effect, is carried by the electrons of the plasma
created by the femtosecond laser pulses.

J i gijklmn Ej E k rn El E m ifijklmn Ej Ek E l E m kn

3:2

where subscripts are cartesian indices, E is the complex amplitude of the light electric field, and Ek E l is
proportional to the light intensity and responsible for heating through plasma absorption. The first
and second terms on the right-hand side of Eq. (3.2) denote pressure created by the front of the pulse
and the photon drag effect, respectively, with gijklmn and fijklmn = fikjlmn = filmjkn as sixth rank tensors
describing the material asymmetry; and k is the wavevector. When the y-polarized laser beam propagates along the z axis of the LiNbO3 crystal, the current along the y axis can be described as Eq. (3.3),
which comprises symmetry allowed components of tensors gijklmn and fijklmn.

Jy

gyyyyyz

!
@jEy j2
ikfyyyyyz jEy j2 jEy j2
@z

3:3

To highlight that this heat current can be excited even under homogeneous illumination in a homogeneous noncentrosymmetric medium, Yang et al. refer to this phenomenon as the photothermal
effect in non-centrosymmetric media, or the bulk photothermal effect. The laser field, driving the heat
current, as demonstrated by Eq. (3.2), lasts about 150 fs, which leads to an anisotropic energy distribution. This can be imprinted in the anisotropy of lattice temperature across the irradiated area.
Specifically, the rate of the average heat production within the focus area can be written as:

@Q
@t

asp  f I

3:4

homogeneous

where I is the laser intensity, sp is the pulse duration, f is the pulse repetition rate and a is determined
by the absorption coefficient, the irradiated volume and so on. The heating homogeneously increases
the temperature within the whole focal zone. However, in the LiNbO3 focal area, there is an average
heat current (J y ) along the y axis (Eq. (3.3)),

J y bsp  f I2

3:5

where b is determined by relevant non-zero components of the material tensors. As the light beam
propagates along the z axis of the crystal, the heat flow will push the heat along the y axis at the transfer rate of

@Q
@t

AJ y

3:6

anisotropic

with A as the cross-section of the irradiated area in the XZ plane, and hence, will cause a temperature
gradient (DT) between opposite sides of the beam.

d @Q
DT
kA @t

anisotropic

d
Jy
k

3:7

DT depends on the relevant component of the material tensor, which is responsible for the observed
effect. In particular, if the laser beam is y-polarized and Im{nyyyyyyz} > 0, then Jy < 0. Therefore, according to Eq. (3.5), when the y axis is horizontal along negative direction to the right and the anisotropic
heating happens, the temperature of the right-hand side of the irradiated zone will be higher than that
of the left-hand side (DT = dJy/k), as shown by a simplified model (Fig. 32), where the beam movement
is set to be discontinuous; that is, the beam movement along the y axis consists of jumps equal to the
beam diameter in length. When the beam jumps to the right (the negative direction of the y axis),
the temperature of the beam left side increases to T + DT, and the temperature of the right side of

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

211

Fig. 32. Illustration of the bulk photothermal effect induced differential heating of the LiNbO3 crystal. Heat flows (black arrow)
along the y direction of the crystal. The temperature of the crystal increases until saturation when the beam is displaced in the
y direction (red arrow) and oscillates near the level defined by isotropic heating (green dashed line) when displacement is
opposite to the heat flow (blue arrow). The large circles show the laser beam and increasingly darker color reflects increasing
temperature of the sample in the position of the beam. (Reproduced with permission from [17].)

the beam is T. The anisotropic heating mechanism is switched on, resulting in temperature increase of
the bean right side with DT higher than the left side. Consequently, the temperature of the right side
of the beam will be T + 2DT. The same temperature increase will go on with more jumps. After m
jumps in the direction coinciding with the direction of the anisotropic heat flow, the temperature
of the rightmost irradiated area will be:

T max
parallel T 0

md
J
k y

3:8

Furthermore, when the light beam moves in the positive direction along the y axis, the temperature
of the left side of the beam remains unchanged after each jump. Therefore, although the anisotropic
heating gives rise to an increase of the temperature of the beam right side, the temperature of the
opposite side does not increase (Fig. 32):

d
T max
opposite T 0 J y
k

3:9

As a result, the anisotropic heating will lead to a very different scenario for the laser writing, even
to shock-induced damage (Fig. 31B and C). Thus, when the beam is translated along the y axis, an inplane heat flow can be created either parallel or antiparallel to the beam velocity. As the heating of the
crystal is stronger when the direction of the heat flow coincides with the direction of the beam movement, modification of the non-centrosymmetric crystal exhibits a pronounced directional dependence.
In addition, when the light beam propagates along the z axis, no thermal current is produced along the
x axis, and that is why that the crystal modification is not sensitive to the light beam translation along
the x axis.
3.5. Formation of high pressure crystalline phase
At extreme pressure and temperature, common materials will tend to form new dense phases (or
namely metastable phases) with compacted atomic arrangements and unusual physical properties. In
addition, the synthesis and study of new metastable phases of matter under pressure above 100 GPa
and temperature above 104 K may reveal the details of planet and star interiors, and lead to materials
with extraordinary properties. Many phases have been theoretically predicted, which may be produced under appropriate formation conditions [444447]. As said above, it is possible to establish

212

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

extreme pressure and temperature conditions in table-top laboratory experiments with femtosecond
laser pulses highly focused inside a transparent material, with intensity in the focal spot well above
the threshold for optical breakdown [93,99,220]. As a 100 fs pulse with 100 nJ energy is tightly focused
deep inside a crystal, an energy density of several MJ cm3 can be injected into a submicron volume,
which is several times higher than the strength of any material and superheat solid to a plasma. The
confined plasma explodes to generate a powerful shock wave that expands out of the focal volume and
compresses the surrounding pristine material [93,99,220]. The shock wave of the expanding plasma
can induce extreme pressure of over 1 TPa, causing dramatic changes in the material properties. For
example, microexplosions in sub-micrometer sized regions of sapphire can be induced by tightly
focused femtosecond laser pulses with a temporal length of about 100 fs and a pulse energy of approximately 100 nJ. Fast, explosive expansion of photogenerated high density plasma creates strong heating and pressure transients with peak temperature and pressure of about 105 K and 10 TPa,
respectively [448]. Recently, a new superdense stable phase of body-centered-cubic aluminum (bcc
Al), predicted by first-principles theories to exist at pressure above 380 GPa, is synthesized, which
is confirmed by the Synchrotron XRD microanalysis (lXRD) (Fig. 33A and C) [444]. According to a
detailed examination, Vailionis et al. show that the measured peak positions in Fig. 33C can be perfectly indexed as a bcc Al phase (space group Im3m) with a lattice constant a = 2.866 (11.77 3
per atom), which matches the predicted value [449], but not yet experimentally observed, high pressure Al phase. The mean crystallite diameter size of the generated bcc Al phase is determined to be
about 18 nm using Rietveld refinement fitting and Scherrers algorithm.
The spatial separation of Al and O ions under conditions of complete confinement is proposed to be
responsible for the generation of high pressure bcc Al [444,448,450,451]. The femtosecond laser pulses
with intensity above the optical breakdown threshold break bonds of Al2O3 and ionize Al and O atoms
into plasma, which exists until the material cools down to a temperature below the thermal ionization
threshold. In this plasma, the Al and O ions diffuse and scatter at different rates, enabling spatial separation of the ions to occur (Fig. 33D). As the spatial separation through diffusion can only take place
in a hot plasma state, where atoms are ionized and ionic collisions are governed by the Rutherford
scattering with cross-section inversely proportional to the squared relative energy and reduced mass
of the interacting particles. A model based on the Rutherford cross-section indicates that spatial separation by tens-of-nm between oxygen and aluminum is possible if the diffusion length of oxygen
exceeds 100 nm, leading to the formation of bcc Al [448]. Specifically, the spatial separation of oxygen

Fig. 33. lXRD image acquired (A) in the center of the shockwave compressed area, and (B) outside the shockwave compressed
area. (C) Comparison between a radially integrated experimental lXRD profile obtained from the experimental data shown in
(A), and theoretical lXRD profile expected for bcc-Al refined using materials studio package. For comparison, simulated profiles
of host sapphire (vertically offseted gray) and native fcc-Al (purple) are also shown. (D) schematic explanation of spatial and
temporal transformations in the volume of ionized sapphire affected by a microexplosion, leading to spatial separation of Al and
O ions and quenching of the bcc Al. (Reproduced with permission from [444].)

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

213

and aluminum ions is mainly assigned to the difference in both their diffusion coefficients and diffusion time. Plasma lifetime limits the diffusion time. The oxygen gains energy from electrons earlier
and starts to move in about 327 ps after the pulse, and for aluminum the related time is about
545 ps [450]. Therefore, oxygen ions move away from colder Al ions about several tens of nanometers before the Al ions starts moving. The total separation is about 32 nm between the elements, which
is sufficient for formation of the Al nanocrystals. Especially, Gamaly et al. reveal that with laser fluence
up to 50 times higher than the ionization threshold, the energy can be effectively absorbed in the bulk
of the material, resulting in an enhanced ion separation in the non-ideal plasma of microexplosion
[451]. Furthermore, the separation of the constituents is confirmed by the observation of molecular
oxygen [371,382,452]. In addition, after the laser pulse is gone and energy dissipation begins, thermal
ionization by electron impact maintains the ionization process at almost the same level as during the
action of laser pulse. The superheated solid is droven to be a plasma leading to generation of a powerful shock wave, and the plasma expands from the focal volume and compresses material against the
cold bulk solid into a shell with increased density, and hence, voids form as well.
Femtosecond laser induced microexplosion offers a new strategy for the synthesis of high pressure
phases of superdense and superhard materials in useful amounts under table-top laboratory
conditions.

4. Conclusion and perspective


Progress in high power ultrashort pulse lasers has opened new frontiers in science and technology
of lightmatter interactions [453]. The availability of high repetition femtosecond lasers, precise fast
scanning stages and galvano mirrors allows for an increasing number of the applications. In particular,
femtosecond lasermatter interaction has been demonstrated to be a versatile and powerful technique for the high precision material removal, deposition and modification. More recently, various
phenomena induced by femtosecond laser in transparent materials have attracted considerable interest due to a wide range of potential applications, such as laser surgery, integrated optics, optical data
storage, and 3D micro- and nanostructuring. An important and key feature of femtosecond laser
matter interaction is its extremely high precision owing to the intrinsic multiphoton absorption
and the efficient suppression of the heat diffusion to the surrounding regions of the processed local
volume, which imposes very stringent requirements on the spatial and temporal characteristics of
the laser pulses. Another feature is that the irradiation parameters can be easily controlled to tailor
the materials local structure and functionalities, and to discover new physical and chemical
phenomena [17,438].
Further investigations are a necessity in order to clarify new phenomena induced by femtosecond
laser, and to realize more precise, faster micromachining in larger scale for fabricating high quality
functional materials. In particular, the mechanism of the femtosecond lasermatter interaction
including multiphoton absorption and ionization and the subsequent energy dispersion should be
better understood. Here we propose the following subjects that could be of importance and priority
for future fundamental and application research in terms of femtosecond laser interaction with transparent materials.
Various phenomena, such as generation of nanogratings, refractive index change and bubbles, have
been observed during and after the interaction of femtosecond laser with transparent materials.
Recently, new phenomena, e.g., formation of nanoporous structure in silica glass [371], and SHG in
Ag-doped phosphate glass [454] induced by femtosecond laser have been observed and discussed.
Despite these extensive investigations there is still a lack of a clear picture about the dynamic processes of the observed phenomena. These phenomena cannot be well explained by the existing theories since they depend on numerous factors such as chemical nature of materials and experimental
conditions. In addition, many linear and nonlinear effects co-exist during the femtosecond laser irradiation. Therefore, much more theoretical and experimental studies need to be conducted in order to
clarify the effects of spatial distribution of femtosecond laser beam, self-focusing, plasma formation
and recombination on the materials microstructure and functionalities, and to exactly determine temperature, pressure and shock wave, and their generation and decay processes. Physical modeling of the

214

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

femtosecond irradiation dynamics should be performed due to its complexity. For instance, the energetic excitation in material via femtosecond laser pulses is actually induced by a combined effect of
strong-field excitation (multiphoton and tunnel excitation), collisional excitation (likely leading to
an avalanche process), and absorption in the plasma consisting of the electrons excited to the conduction band. The dynamic models should take the co-action of these parallel processes into account by
means of various rate-equation models in combined with descriptions of the excited electrons. The
optical properties of the highly excited dielectric undergo a rapid change during the laser pulse, which
must be included in a detailed modeling of the excitations [22].
In order to clarify mechanisms of various femtosecond laser induced phenomena, it is also necessary to further develop various time-resolved femtosecond laser techniques. Recently, Schultze et al.
reported control of dielectrics with few-cycle femtosecond laser [455]. Usually, the electric and optical
properties of semiconductors are controlled with microwave fields which form the basis of modern
electronics, information processing and optical communications. Such control can be extended to optical frequencies for wideband materials such as dielectrics using strong electric fields. Few-cycle femtosecond laser pulses permit damage-free exposure of dielectrics to electric fields of several volts per
angstrom and significant modifications in their electronic system. Fields of such strength and temporal confinement can turn a dielectric from an insulating state to a conducting state within an optical
period. Schulze et al. study the underlying electron processes with sub-femtosecond solid-state
spectroscopy, and hence, reveal the feasibility of manipulating the electronic structure and electric
polarizability of a dielectric reversibly with the electric field of light. They irradiate fused silica with
a waveform-controlled near-infrared few-cycle light field of several volts per angstrom and probe
changes in extreme-ultraviolet absorptivity and near-infrared reflectivity on a timescale of approximately a hundred attoseconds to a few femtoseconds. The field-induced changes follow, in a highly
nonlinear fashion, the turn-on and turn-off behavior of the driving field, in agreement with the predictions of a quantum mechanical model. The ultrafast reversibility of the effects implies that the
physical properties of a dielectric can be controlled with the electric field of light, offering the potential for petahertz-bandwidth signal manipulation. Therefore, development of various femtosecond
laser techniques including femtosecond pomp-probe spectroscopy technique, femtosecond X-ray
absorption spectroscopy and imaging technique will allow us to clarify various dynamics of ultrafast
processes and realize precise control of physicochemical properties of materials.
Up to now, various femtosecond laser techniques have been developed for modification of transparent materials. Usually, they are only based on control of parameters of single femtosecond laser
beam, such as pulse energy, pulse repetition rate, pulse width, and polarization. To achieve effective
structural modification, novel techniques based on multi-beam femtosecond laser interference have
been proposed. Especially, multi-beam interference technique has been demonstrated to be powerful
for the precise control of 2D and 3D refractive index change patterns [342,350459]. In addition, it is
also useful to realize novel properties of materials based on nonlinear femtosecond laser field
[460,461].
Recently, femtosecond pulse shaping has received much attention for studying femtosecond laser
interaction with matter. Femtosecond pulse shaping refers to manipulations with temporal profile of
an ultrashort laser pulse. Pulse shaping can be used to shorten/prolong the duration of optical pulse, or
to generate complex pulses. Temporal pulse shaping is efficient to precisely control the dynamics of
energy deposition in materials, significantly improving the fabrication quality and precision, owing
to the highly tunable temporal profiles of ultrafast laser pulses with temporal resolutions far
exceeding the characteristic energy transfer time [259,462]. Quantum coherent control of the
resonance-mediated two-photon absorption in rare-earth ions has been demonstrated by using the
phase-shaped femtosecond laser pulse. The control efficiency depends on the laser repetition rate
due to the long lifetime and the short decoherence time of the excited state, and the higher laser repetition rate yields lower control efficiency [463]. The techniques based on pulse shaping are promising
for controlling the femtosecond laser induced processes and structures. Furthermore, the optical axial
elongation of the fabricated structure is always a major problem that has thus far limited design
flexibility, especially for the direction along the optical axis. This problem has been solved recently
by controlling the light intensity distribution profile and using the adequate focal length of the
hologram, leading to homogeneous and elongation-free 3D microfabrication [464].

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

215

Furthermore, the femtosecond laserglass interaction could become a potential powerful tool for
understanding the most complicated and interesting unresolved problems in condensed matter
science, namely, the nature of glass state and glass transition [465]. Due to its enormous energy density, the femtosecond laser can generate extremely high pressure and temperature field in a focused
nano-domain of glass, and hence, can breakdown all the possible chemical bonds and re-forming a
new glass phase. Thus, a glass-in-glass scenario is created. Compare to the original glass, the extremely formed glass should possess a lower density, higher fictive temperature (Tf) and a lower fictive
pressure (Tp), whereas the surrounding glassy domain should exhibit has a higher density, lower Tf and
higher Tp due to an enormous impact of the sudden shock wave. This provides us with a unique opportunity to clarify current glass problems like glass formation, glass transition [466], liquidliquid transition, polyamorphism [467] and structural heterogeneity [468] occurring in nano-confinement under
the otherwise not reachable conditions. This will be done by combining other techniques such as
microstructural probe, nano-calorimetry, and newly emerging tools. The structural evolution induced
by femtosecond laser and the subsequent relaxation process could be recorded as a function of laser
intensity and glass composition. The recorded results should be compared with the molecular
dynamic simulation results. From derived structural and dynamic information, one can infer what
occurs thermodynamically during the femtosecond laserglass interaction. The derived knowledge
will also be valuable for designing the femtosecond laserglass interaction process to obtain targeted
functionalities of glassy materials.
The underlying goal of this review is to document and systematize progress in experimental and
theoretical exploitations of femtosecond laser induced phenomena in transparent materials. In this
overview, we mainly pay attention to the new phenomena observed recently, the related mechanisms
and their applications. We also give a comprehensive review over the fundamental concepts, the characteristics, and the typical systems for ultrafast laser, the controlling parameters such as wavelength,
intensity, exposure time, and pulse duration, light polarization, pulse geometric characteristics. All
these aspects influence the resulting structures and properties in transparent materials irradiated
by femtosecond laser. Femtosecond laser interaction with transparent materials is an exciting
research area. More new phenomena will be discovered and more breakthroughs in scientific understanding and technological applications will be achieved concerning light interaction with matter.
Such research will contribute to development of the condensed matter physics, materials science,
and new laser processing technique.
Acknowledgements
This work was financially supported by the National Natural Science Foundation of China (Grant
Nos. 51072054, 51132004, and 51102209), Open Fund of the State Key Laboratory of High Field Laser
Physics (Shanghai Institute of Optics and Fine Mechanics), and the National Basic Research Program of
China (2011CB808100).
References
[1] Steinmeyer G, Sutter DH, Gallmann L, Matuschek N, Keller U. Frontiers in ultrashort pulse generation: pushing the limits
in linear and nonlinear optics. Science 1999;286:150712.
[2] Valdmanis JA, Fork RL. Design considerations for a femtosecond pulse laser balancing self phase modulation, group
velocity dispersion, saturable absorption, and saturable gain. IEEE J Quantum Electron 1986;22:1128.
[3] Sdmeyer T, Marchese SV, Hashimoto S, Baer CRE, Gingras G, Witzel B, et al. Femtosecond laser oscillators for high-field
science. Nat Photonics 2008;2:599604.
[4] Ditmire T, Zweiback J, Yanovsky VP, Cowan TE, Hays G, Wharton KB. Nuclear fusion from explosions of femtosecond laserheated deuterium clusters. Nature 1999;398. 489-2.
[5] Gattass RR, Mazur E. Femtosecond laser micromachining in transparent materials. Nat Photonics 2008;2:21925.
[6] Du D, Liu X, Korn G, Squier J, Mourou G. Laser-induced breakdown by impact ionization in SiO2 with pulse widths from 7
ns to 150 fs. Appl Phys Lett 1994;6:30713.
[7] Pronko PP, Dutta SK, Squier J, Rudd JV, Du D, Mourou G. Machining of submicron holes using a femtosecond laser at 800nm. Opt Commun 1995;114:10610.
[8] Ratkay-Traub I, Juhasz T, Horvath C, Suarez C, Kiss K, Ferincz I, et al. Ultra-short pulse femtosecond laser surgery.
Ophthalmol Clin North Am 2001;14:34755.
[9] Mangles SPD, Murphy CD, Najmudin Z, Thomas AGR. Monoenergetic beams of relativistic electrons from intense laser
plasma interactions. Nature 2004;431:5358.

216

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

[10] Qiu JR, Miura K, Hirao K. Femtosecond laser-induced microfeatures in glasses and their applications. J Non-Cryst Solids
2008;354:110011.
[11] Thomas J, Heinrich M, Burghoff J, Nolte S, Ancona A, Tnnermann A. Femtosecond laser-written quasi-phase-matched
waveguides in lithium niobate. Appl Phys Lett 2007;91:151108.
[12] Kawata S, Sun HB, Tanaka T, Takada K. Finer features for functional nanodevices. Nature 2001;412:6978.
[13] Schaffer CB, Brodeur A, Mazur E. Laser-induced breakdown and damage in bulk transparent materials induced by tightly
focused femtosecond laser pulses. Meas Sci Technol 2001;12:178494.
[14] Itoh K, Watanabe W, Nolte S, Schaffer C. Ultrafast processes for bulk modification of transparent materials. MRS Bull
2006;31:6205.
[15] Poumellec B, Sudrie L, Franco M, Prade B, Mysyrowicz A. Femtosecond laser irradiation stress induced in pure silica. Opt
Express 2003;11:10709.
[16] Kazansky P, Yang W, Bricchi E, Bovatsek J, Arai A, Shimotsuma Y, et al. Quill writing with ultrashort light pulses in
transparent materials. Appl Phys Lett 2007;90:151120.
[17] Yang WJ, Kazansky PG, Svirko YP. Non-reciprocal ultrafast laser writing. Nat Photonics 2008;2:99104.
[18] Qiu JR, Jiang XW, Zhu CS, Shirai M, Si JH, Jiang N, et al. Manipulation of gold nanoparticles inside transparent materials.
Angew Chem Int Ed 2004;43:22304.
[19] Shimotsuma Y, Kazansky PG, Qiu JR, Hirao K. Self-organized nanogratings in glass irradiated by ultrashort light pulses.
Phys Rev Lett 2003;91:247405.
[20] Kanehira S, Miura K, Hirao K. Ion exchange in glass using femtosecond laser irradiation. Appl Phys Lett 2008;93:023112.
[21] Kanehira S, Si JH, Qiu JR, Fujita K, Hirao K. Periodic nanovoid structures via femtosecond laser irradiation. Nano Lett
2005;5:15915.
[22] Balling P, Schou J. Femtosecond-laser ablation dynamics of dielectrics: basics and applications for thin films. Rep Prog
Phys 2013;76:036502.
[23] Band YB. Light and matter. New York: John Wiley & Sons Inc; 2006.
[24] Demtroder W. Laser spectroscopy. 3rd ed. Springer; 2003.
[25] Rulliere C. Femtosecond laser pulses. Springer; 2005.
[26] Brabec T, Krausz F. Intense few-cycle laser fields: frontiers of nonlinear optics. Rev Mod Phys 2000;72:54591.
[27] Spence D, Kean PN, Sibbett W. 60 fsec pulse generation from a self-mode-locked Ti-sapphire laser. Opt Lett 1991;16:424.
[28] Keller U. Recent developments in compact ultrafast lasers. Nature 2003;424:8318.
[29] Eimerl D, Davis L, Velsko S, Graham EK, Zalkin A. Optical, mechanical, and thermal properties of barium borate. J Appl
Phys 1987;62:196883.
[30] Chen C, Wu Y, Jiang A, Wu B, You G, Li R, et al. New nonlinear-optical crystal: LiB3O5. J Opt Soc Am B 1989;6:61621.
[31] Zhang JY, Huang JY, Shen YR, Chen C. Optical parametric generation and amplification in barium borate and lithium
triborate crystals. J Opt Soc Am B 1993;10:175864.
[32] Borsutzky A, Brnger R, Huang C, Wallenstein R. Harmonic and sum-frequency generation of pulsed laser radiation in
BBO, LBO, and KDP. Appl Phys B Photophys Laser Chem 1991;52:5562.
[33] Perry Michael D, Mourou Gerard. Terawatt to petawatt subpicosecond lasers. Science 1994;264:91724.
[34] Miller DE, Zapata LE, Ripin DJ, Fan TY. Sub-picosecond pulses at 100 W average power from a Yb:YLF chirped-pulse
amplification system. Opt Lett 2012;37:27002.
[35] Witte S, Eikema KSE. Ultrafast optical parametric chirped-pulse amplification. IEEE J Sel Top Quantum Electron
2012;18:296307.
[36] Jovanovic I. Chirped-pulse amplification: ultrahigh peak power production from compact short-pulse laser systems. Opt
Photon 2010;5:303.
[37] Della Valle G, Osellame R, Laporta P. Micromachining of photonic devices by femtosecond laser pulses. J Opt A: Pure Appl
Opt 2009;11:013001.
[38] Schaffer CB, Brodeur A, Garcia JF, Mazur E. Micromachining bulk glass by use of femtosecond laser pulses with nanojoule
energy. Opt Lett 2001;26:935.
[39] Minoshima K, Kowalevicz AM, Hartl I, Ippen EP, Fujimoto JG. Photonic device fabrication in glass by use of nonlinear
materials processing with a femtosecond laser oscillator. Opt Lett 2001;26:15168.
[40] Killi A, Morgner U, Lederer MJ, Kopf D. Diode-pumped femtosecond laser oscillator with cavity dumping. Opt Lett
2004;29:128890.
[41] Osellame R, Chiodo N, Della Valle G, Taccheo S, Ramponi R, Cerullo G, et al. Optical waveguide writing with a diodepumped femtosecond oscillator. Opt Lett 2004;29:19002.
[42] Dostovalov A, Babin S, Dubov M, Baregheh M, Mezentsev V. Comparative numerical study of energy deposition in
femtosecond laser microfabrication with fundamental and second harmonics of Yb-doped laser. Laser Phys
2012;22:9306.
[43] Wise FW, Chong A, Renninger WH. High-energy femtosecond fiber lasers based on pulse propagation at normal
dispersion. Laser Photonics Rev 2008;2:5873.
[44] Richardson DJ, Nilsson J, Clarkson WA. High power fiber lasers: current status and future perspectives. J Opt Soc Am B
2010;27:B6392 [Invited].
[45] Poumellec B, Lancry M, Chahid-Erraji A, Kazansky PG. Modification thresholds in femtosecond laser processing of pure
silica: review of dependencies on laser parameters. Opt Mater Express 2011;1:76682 [Invited].
[46] Ams M, Marshall GD, Dekker P, Dubov M, Mezentsev VK, Bennion I, et al. Investigation of ultrafast laser-photonic material
interactions: challenges for directly written glass photonics. IEEE J Sel Top Quantum Electron 2008;14:137081.
[47] Mao SS, Qur F, Guizard S, Mao X, Russo RE, Petite G, et al. Dynamics of femtosecond laser interactions with dielectrics.
Appl Phys A 2004;79:1695709.
[48] Keldysh LV. Zh Eksp Teor Fiz 1964;47:1945 [Engl transl Sov Phys JETP 1965;20:1307].
[49] Joglekar AP, Liu H, Meyhfer E, Mourou G, Hunt AJ. Optics at critical intensity: applications to nanomorphing. Proc Natl
Acad Sci USA 2004;101:5856.

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

217

[50] Stuart C, Feit MD, Rubenchik AM, Shore BW, Perry MD. Laser-induced damage in dielectrics with nanosecond to
subpicosecond pulses. Phys Rev Lett 1995;74:2248.
[51] Rethfeld B. Unified model for the free-electron avalanche in laser-irradiated dielectrics. Phys Rev Lett 2004;92:187401.
[52] Schaffer CB, Nishimura N, Glezer EN, Kim AMT, Mazur E. Dynamics of femtosecond laser-induced breakdown in water
from femtoseconds to microseconds. Opt Express 2002;10:196203.
[53] Sakakura M, Terazima M. Initial temporal and spatial changes of the refractive index induced by focused femtosecond
pulsed laser irradiation inside a glass. Phys Rev B 2005;71:024113.
[54] Sakakura M, Terazima M, Shimotsuma Y, Miura K, Hirao K. Observation of pressure wave generated by focusing a
femtosecond laser pulse inside a glass. Opt Express 2007;15:567486.
[55] Hnatovsky C, Taylor RS, Simova E, Rajeev PP, Rayner DM, Bhardwaj VR, et al. Fabrication of microchannels in glass using
focused femtosecond laser radiation and selective chemical etching. Appl Phys A 2006;84:4761.
[56] Gawelda W, Puerto D, Siegel J, Ferrer A, Ruiz de la Cruz A, Fernndez H, et al. Ultrafast imaging of transient electronic
plasmas produced in conditions of femtosecond waveguide writing in dielectrics. Appl Phys Lett 2008;93:121109.
[57] Little DJ, Ams M, Dekker P, Marshall GD, Dawes JM, Withford MJ. Femtosecond laser modification of fused silica: the effect
of writing polarization on SiO ring structure. Opt Express 2008;16:2002937.
[58] Liu D, Li Y, Liu M, Yang H, Gong Q. The polarization-dependence of femtosecond laser damage threshold inside fused
silica. Appl Phys B 2008:915979.
[59] Reiss HR. Polarization effects in high-order multiphoton ionization. Phys Rev Lett 1972;29:112931.
[60] Lompr L, Mainfray G, Manus C, Thebault J. Multiphoton ionization of rare gases by a tunable-wavelength 30-psec laser at
1.06 lm. Phys Rev A 1977;15:160412.
[61] Temnov VV, Sokolowski-Tinten K, Zhou P, El-Khamhawy A, von der Linde D. Multiphoton ionization in dielectrics:
comparison of circular and linear polarization. Phys Rev Lett 2006;97:237403.
[62] Ams M, Marshall GD, Withford MJ. Study of the influence of femtosecond laser polarisation on direct writing of
waveguides. Opt Express 2006;14:1315863.
[63] Song J, Ye JY, Qian MD, Lin X, Bian HD, Dai Y, et al. Polarization dependence of the self-organized microgratings induced in
SrTiO3 crystal by a single femtosecond laser beam. Opt Express 2013;21:184618.
[64] Hnatovsky C, Shvedov V, Krolikowski W, Rode A. Revealing local field structure of focused ultrashort pulses. Phys Rev Lett
2011;106:123901.
[65] Shah L, Arai A, Eaton S, Herman P. Waveguide writing in fused silica with a femtosecond fiber laser at 522 nm and 1 MHz
repetition rate. Opt Express 2005;13:19992006.
[66] Alexander M, Streltsov AM, Borrelli NF. Study of femtosecond-laser-written waveguides in glasses. J Opt Soc Am B
2002;19:2496504.
[67] Jia TQ, Chen HX, Huang M, Zhao FL, Li XX, Xu SZ, et al. Ultraviolet-infrared femtosecond laser-induced damage in fused
silica and CaF2 crystals. Phys Rev B 2006;73:054105.
[68] Yang WJ, Bricchi E, Kazansky PG, Bovatsek J, Arai AY. Self-assembled periodic sub-wavelength structures by femtosecond
laser direct writing. Opt Express 2006;14:1011724.
[69] Skupin S, Berg L. Self-guiding of femtosecond light pulses in condensed media: plasma generation versus chromatic
dispersion. Phys D 2006;220:1430.
[70] Lancry M, Poumellec B, Chahid-Erraji A, Beresna M, Kazansky PG. Dependence of the femtosecond laser refractive index
change thresholds on the chemical composition of doped-silica glasses. Opt Mater Express 2011;1:71123.
[71] Couairona A, Mysyrowicz A. Femtosecond filamentation in transparent media. Phys Rep 2007;441:47189.
[72] Couairon A, Sudrie L, Franco M, Prade B, Mysyrowicz A. Filamentation and damage in fused silica induced by tightly
focused femtosecond laser pulses. Phys Rev B 2005;71:125435.
[73] Tien AC, Backus S, Kapteyn H, Murnane M, Mourou G. Short-pulse laser damage in transparent materials as a function of
pulse duration. Phys Rev Lett 1999;82:38836.
[74] Hnatovsky C, Taylor R, Rajeev P, Simova E, Bhardwaj V, Rayner D, et al. Pulse duration dependence of femtosecond-laserfabricated nanogratings in fused silica. Appl Phys Lett 2005;87:014104.
[75] Shcheblanov NS, Itina TE. Femtosecond laser interactions with dielectric materials: insights of a detailed modeling of
electronic excitation and relaxation processes. Appl Phys A 2013;110:57983.
[76] Ashcom JB, Gattass RR, Schaffer CB, Mazur E. Numerical aperture dependence of damage and supercontinuum generation
from femtosecond laser pulses in bulk fused silica. J Opt Soc Am B 2006;23:231722.
[77] DesAutels GL, Brewer CD, Walker MA, Juhl SB, Finet MA, Powers PE. Femtosecond micromachining in transparent bulk
materials using an anamorphic lens. Opt Express 2007;15:1313948.
[78] Mizeikis V, Juodkazis S, Balinas T, Misawa H, Kudryashov SI, Zvorykin VD, et al. Optical and ultrasonic signatures of
femtosecond pulse filamentation in fused silica. J Appl Phys 2009;105:123106.
[79] Eaton SM, Zhang HB, Herman PR, Yoshino F, Shah L, Bovatsek J, et al. Heat accumulation effects in femtosecond laserwritten waveguides with variable repetition rate. Opt Express 2005;13:470816.
[80] Osellame R, Chiodo N, Della Valle G, Cerullo G, Ramponi R, Laporta P, et al. Waveguide lasers in the C-band fabricated by
laser inscription with a compact femtosecond oscillator. IEEE J Sel Top Quantum Electron 2006;12:27785.
[81] Eaton SM, Zhang H, Ng ML, Li J, Chen WJ, Ho S, et al. Transition from thermal diffusion to heat accumulation in high
repetition rate femtosecond laser writing of buried optical waveguides. Opt Express 2008;16:944358.
[82] Brub JP, Bernier M, Valle R. Femtosecond laser-induced refractive index modifications in fluoride glass. Opt Mater
Express 2013;3:598611.
[83] Jamshidi-Ghaleh K, Abdolahpour D, Mansour N. Laser fluence and shot number dependence of laser-induced optical
properties modification of transparent materials. Laser Phys Lett 2006;3:5737.
[84] Arabanian AS, Massudi R. Comparison on pulse propagation and plasma generation in contact and noncontact geometries
for femtosecond micromachining of silica glass. Opt Eng 2013;52:024303.
[85] Bricchi E, Klappauf BG, Kazansky PG. Form birefringence and negative index change created by femtosecond direct
writing in transparent materials. Opt Lett 2004;29:11921.
[86] Davis H, Miura K, Sugimoto N, Hirao K. Writing waveguides in glass with a femtosecond laser. Opt Lett 1996;21:1729.

218

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

[87] Mishchik K, Ferrer A, Ruiz de la Cruz A, Mermillod-Blondin A, Mauclair C, Ouerdane Y, et al. Photoinscription domains for
ultrafast laser writing of refractive index changes in BK7 borosilicate crown optical glass. Opt Mater Express
2013;3:6785.
[88] Mishchik K, Cheng G, Huo G, Burakov IM, Mauclair C, Mermillod-Blondin A, et al. Nanosize structural modifications with
polarization functions in ultrafast laser irradiated bulk fused silica. Opt Express 2010;18:2480924.
[89] Sudrie L, Franco M, Prade B, Mysyrowicz A. Study of damage in fused silica induced by ultra-short IR laser pulses. Opt
Commun 2001;191:3339.
[90] Smelser CW, Mihailov SJ, Grobnic D. Formation of Type I-IR and Type II-IR gratings with an ultrafast IR laser and a phase
mask. Opt Express 2005;13:537786.
[91] Lu P, Grobnic D, Mihailov SJ. Characterization of the birefringence in fiber Bragg gratings fabricated with an ultra-fast
infrared laser. J Lightwave Technol 2007;25:77986.
[92] Glezer EN, Milosavljevic M, Huang L, Finlay RJ, Her TH, Callan JP, et al. Three-dimensional optical storage inside
transparent materials. Opt Lett 1996;21:20235.
[93] Glezer EN, Mazur E. Ultrafast-laser driven micro-explosions in transparent materials. Appl Phys Lett 1997;71:8824.
[94] Sun HB, Xu Y, Matsuo S, Misawa H. Microfabrication and characteristics of two-dimensional photonic crystal structures in
vitreous silica. Opt Rev 1999;6:3968.
[95] Mishchik K, DAmico C, Velpula PK, Mauclair C, Ouerdane Y, Boukenter A, et al. Ultrafast laser-induced electronic and
structural modifications in bulk fused silica. J Appl Phys 2013;114:133502.
[96] Stoian R, Mishchik K, Cheng G, Mauclair C, DAmico C, Colombier JP, et al. Investigation and control of ultrafast laserinduced isotropic and anisotropic nanoscale-modulated index patterns in bulk fused silica. Opt Mater Express
2013;3:175568.
[97] Mermillod-Blondin A, Bonse J, Rosenfeld A, Hertel IV, Meshcheryakov YP, Bulgakova NM, et al. Dynamics of femtosecond
laser induced voidlike structures in fused silica. Appl Phys Lett 2009;94:041911.
[98] Miura K, Qiu JR, Inouye H, Mitsuyu T, Hirao K. Photowritten optical waveguides in various glasses with ultrashort pulse
laser. Appl Phys Lett 1997;71:332931.
[99] Juodkazis S, Nishimura K, Tanaka S, Misawa H, Gamaly EG, Luther-Davies B, et al. Laser-induced microexplosion confined
in the bulk of a sapphire crystal: evidence of multimegabar pressures. Phys Rev Lett 2006;96:166101.
[100] Rajeev PP, Gertsvolf M, Simova E, Hnatovsky C, Taylor RS, Bhardwaj VR, et al. Memory in nonlinear ionization of
transparent solids. Phys Rev Lett 2006;97:253001.
[101] Taylor R, Hnatovsky C, Simova E. Applications of femtosecond laser induced self-organized planar nanocracks inside fused
silica glass. Laser Photonics Rev 2008;2:2646.
[102] Kazansky PG, Shimotsuma Y. Self- assembled sub wavelength structures and form birefringence created by femtosecond
laser writing in glass: properties and applications. J Ceram Soc Jpn 2008;116:105262.
[103] Beresna M, Gecivicius M, kazansky PG. Polarization sensitive elements fabricated by femtosecond laser nanostructuring
of glass. Opt Mat Express 2011;1:78395.
[104] Richter S, Heinrich M, Dring S, Tnnermann A, Nolte S. Formation of femtosecond laser-induced nanogratings at high
repetition rates. Appl Phys A 2011;104:5037.
[105] Tan DZ, Yamada Y, Zhou SF, Shimotsuma Y, Miura K, Qiu JR. Photoinduced luminescent carbon nanostructures with ultrabroadly tailored size ranges. Nanoscale 2013;5:120927.
[106] Tan DZ, Zhou SF, Qiu JR. Comment on Upconversion and downconversion fluorescent graphene quantum dots: ultrasonic
preparation and photocatalysis. ACS Nano 2012;6:65301.
[107] Kazansky PG, Inouye H, Mitsuyu T, Miura K, Qiu JR, Hirao K, et al. Anomalous anisotropic light scattering in Ge-doped
silica glass. Phys Rev Lett 1999;82:2199.
[108] Qiu JR, Kazansky PG, Si JH, Miura K, Mitsuyu T, Hirao K, et al. Memorized polarization-dependent light scattering in rareearth-ion-doped glass. Appl Phys Lett 2000;77:19402.
[109] You HP, Nogami M. Three-photon-excited fluorescence of Al2O3SiO2 glass containing Eu3+ ions by femtosecond laser
irradiation. Appl Phys Lett 2004;84:2076.
[110] You HP, Hayakawa T, Nogami M. Upconversion luminescence of Al2O3SiO2: Ce3+ glass by femtosecond laser irradiation.
Appl Phys Lett 2004;85:3432.
[111] Meng XG, Tanaka KK, Murai S, Fujita K, Miura K, Hirao K. Two-photon-excited fluorescence from silicate glass containing
tantalum ions pumped by a near-infrared femtosecond pulsed laser. Opt Lett 2006;3:28679.
[112] Xu SQ, Wang W, Zhou SF, Zhu B, Qiu JR. Highly efficient red, green and blue upconversion luminescence of Eu3+/Tb3+codoped silicate by femtosecond laser irradiation. Chem Phys Lett 2007;442:4925.
[113] Qiao YB, Chen DP, Ren JJ, Wu BT, Qiu JR, Akai T. Blue emission from Eu2+-doped high silica glass by near-infrared
femtosecond laser irradiation. J Appl Phys 2008;103:023108.
[114] Yu LX, Nogami M. Upconversion luminescence properties of europium in ZnOSiO2 glasses by femtosecond laser
excitation. Mater Chem Phys 2008;107:1868.
[115] Zeng HD, Song J, Chen DP, Yuan SL, Jiang XW, Cheng Y, et al. Three-photon-excited upconversion luminescence of
niobium ions doped silicate glass by a femtosecond laser irradiation. Opt Express 2008;16:65026.
[116] Xu YS, Chen DP, Zhang Q, Wang W, Zeng HD, Shen C, et al. Two-photon excited red upconversion luminescence of thulium
ions doped GeS2In2S3CsI glass. Chem Phys Lett 2009;472:1046.
[117] Yang LY, Dong YJ, Chen DP, Wang C, Hua X, Da N, et al. Three-photon-excited upconversion luminescence of Ce3+: YAP
crystal by femtosecond laser irradiation. Opt Express 2005;14:2437.
[118] Yang L, Wang C, Dong Y, Da N, Hu X, Chen DP, et al. Three-photon-excited upconversion luminescence of YVO4 single
crystal by infrared femtosecond laser irradiation. Opt Express 2005;13:1015762.
[119] Yang LY, Dong YJ, Chen DP, Wang C, Da N, Jiang XW, et al. Upconversion luminescence from 2E state of Cr3+ in Al2O3
crystal by infrared femtosecond laser irradiation. Opt Express 2005;13:78938.
[120] Dong YJ, Xu J, Zhou GQ, Zhao GJ, Jie MY, Yang LY, et al. Blue upconversion luminescence generation in Ce3+:Gd2SiO5
crystals by infrared femtosecond laser irradiation. Opt Express 2006;14:1899904.

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

219

[121] Wang XS, Qiu JR, Song J, Xu J, Liao Y, Sun HY, et al. Simultaneous three-photon absorption induced ultraviolet
upconversion in Pr3+:Y2SiO5 crystal by femtosecond laser irradiation. Opt Commun 2008;28:299302.
[122] Ryba-Romanowski W, Macalik B, Strzep A, Lisiecki R, Solarz P, Kowalski RM. Investigation of visible emission induced by
infrared femtosecond pulses in erbium-doped YVO4 and LuVO4 single crystals. J Lumin 2013;144:21722.
[123] Petit Y, Royon A, Marquestaut N, Dussauze M, Fargues A, Veber P, et al. Two-photon excited fluorescence in the LYB: Eu
monoclinic crystal: towards a new scheme of single-beam dual-voxel direct laser writing in crystals. Opt Express
2013;12:82233.
[124] Zhu B, Zhang SM, Zhou SF, Jiang N, Qiu JR. Enhanced upconversion luminescence of transparent Eu3+-doped glassceramics containing nonlinear optical microcrystals. Opt Lett 2007;32:6535.
[125] Zhu B, Zhang SM, Lin G, Zhou SF, Qiu JR. Enhanced multiphoton absorption induced luminescence in transparent Sm3+doped Ba2TiSi2O8 glass-ceramics. J Phys Chem C 2007;111:1711821.
[126] Zeng HD, Lin ZY, Zhang Q, Chen DP, Liang XL, Xu YS, et al. Green emission from Eu2+/Dy3+ codoped SrOAl2O3B2O3 glassceramic by ultraviolet light and femtosecond laser irradiation. Mater Res Bull 2011;46:31922.
[127] Qiu JR, Miura K, Inouye H, Kondo Y, Mitsuyu T, Hirao K. Femtosecond laser-induced three-dimensional bright and longlasting phosphorescence inside calcium aluminosilicate glasses doped with rare earth ions. Appl Phys Lett 1998;73:1763.
[128] Qiu JR, Miura K, Inouye H, Mitsuyu T, Hirao K. Blue emission induced in Eu2+-doped glasses by an infrared femtosecond
laser. J Non-Cryst Solids 1999;244:1858.
[129] Qiu JR, Jiang X, Zhu CS, Si JH, Li C, Su Q, et al. Photostimulated long-lasting phosphorescence in rare-earth-doped glasses.
Chem Lett 2003;32:7501.
[130] Qiu JR, Gaeta AL, Hirao K. Long-lasting phosphorescence in oxygen-deficient Ge-doped silica glasses at room temperature.
Chem Phys Lett 2001;333:23641.
[131] Jiang XW, Qiu JR, Fan YY, Hu HF, Zhu CS. Long-lasting phosphorescence and photostimulated long-lasting
phosphorescence in Mn2+-doped alumino-phosphofluoride glasses irradiated by a femtosecond laser. J Mater Res
2003;18:6169.
[132] Qiu JR, Kodama N, Yamaga M, Miura K, Mitsuyu T, Hirao K. Infrared femtosecond laser pulse-induced three-dimensional
bright and long-lasting phosphorescence in a Ce3+-doped Ca2Al2SiO7 Crystal. Appl Opt 1999;38:72025.
[133] Kumar RSS, Harsha SS, Rao DN. Broadband supercontinuum generation in a single potassium di-hydrogen phosphate
(KDP) crystal achieved in tandem with sum frequency generation. Appl Phys B 2007;86:61521.
[134] Silva F, Austin DR, Thai A, Baudisch M, Hemmer M, Faccio D, et al. Multi-octave supercontinuum generation from midinfrared filamentation in a bulk crystal. Nat Commun 2012;3:807.
[135] Ranka JK, Schirmer RW, Gaeta AL. Observation of pulse splitting in nonlinear dispersive media. Phys Rev Lett
1996;77:37836.
[136] Gaeta AL. Catastrophic collapse of ultrashort pulses. Phys Rev Lett 2000;84:35825.
[137] Barviau B, Kibler B, Picozzi A. Wave-turbulence approach of supercontinuum generation: influence of self-steepening and
higher-order dispersion. Phys Rev A 2009;79:063840.
[138] Kards TM, Ratajska-Gadomska B, Gadomski W, Lapini A, Righini R. The role of stimulated Raman scattering in
supercontinuum generation in bulk diamond. Opt Express 2013;21:242019.
[139] Smirnov SV, Ania-Castanon JD, Ellingham TJ, Kobtsev SM, Kukarin S, Turitsyn SK. Optical spectral broadening and
supercontinuum generation in telecom applications. Opt Fiber Technol 2006;12:12247.
[140] Srinivas NKMN, Harsha SS, Rao DN. Femtosecond supercontinuum generation in a quadratic nonlinear medium (KDP).
Opt Express 2005;13:32249.
[141] Kumar RSS, Deepak KLN, Rao DN. Control of the polarization properties of the supercontinuum generation in a
noncentrosymmetric crystal. Opt Lett 2008;33:1198200.
[142] Kumar RSS, Deepak KLN, Rao DN. Depolarization properties of the femtosecond supercontinuum generated in condensed
media. Phys Rev A 2008;78:043818.
[143] Yu J, Jiang HB, Yang H, Gong QH. Depolarization of white light generated by femtosecond laser pulse in KDP crystals. J Opt
Soc Am B 2011;28:156670.
[144] Bradler M, Baum P, Riedle E. Femtosecond continuum generation in bulk laser host materials with sub-lJ pump pulses.
Appl Phys B 2009;97:561574.
[145] Majus D, Dubietis A. Statistical properties of ultrafast supercontinuum generated by femtosecond Gaussian and Bessel
beams: a comparative study. J Opt Soc Am B 2013;30:9949.
[146] Liao M, Chaudhari C, Qin G, Yan X, Suzuki T, Ohishi Y. Tellurite microstructure fibers with small hexagonal core for
supercontinuum generation. Opt Express 2009;17:1217482.
[147] Yu Y, Gai X, Wang T, Ma P, Wang RP, Yang ZY, et al. Mid-infrared supercontinuum generation in chalcogenides. Opt Mater
Express 2013;3:107586.
[148] Liao M, Gao W, Cheng T, Duan Z, Xue X, Kawashima H, et al. Ultrabroad supercontinuum generation through
filamentation in tellurite glass. Laser Phys Lett 2013;10:036002.
[149] Efimov OM, Glebov LB, Grantham S, Richardson M. Photoionization of silicate glasses exposed to IR femtosecond pulses. J
Non-Cryst Solids 1995;191:94100.
[150] Lonzaga JB, Avanesyan SM, Langford SC, Dickinson JT. Color center formation in soda-lime glass with femtosecond laser
pulses. J Appl Phys 2003;94:4332.
[151] Chen GR, Yang YX, Qiu JR, Jiang XW, Hirao K. Formation of infrared femtosecond laser induced colour centres in Tb3+doped and Tb3+/Ce3+-codoped heavy germanate glasses. Chin Phys Lett 2003;20:19972000.
[152] Courrol LC, Samad RE, Gomes L, Ranieri IM, Baldochi SL, Freitas AZD, et al. Color center production by femtosecond pulse
laser irradiation in LiF crystals. Opt Express 2004;12:28893.
[153] Pan SK, Jiang BX, Jiang XW, Qiu JR, Zhu CS. Near infrared ultra-fast intense laser-induced colour centres in KCl crystal. J
Cryst Growth 2004;263:6489.
[154] Dickinson JT, Orlando S, Avanesyan SM, Langford SC. Color center formation in soda lime glass and NaCl single crystals
with femtosecond laser pulses. Appl Phys A 2004;79:85964.

220

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

[155] Zhao QZ, Qiu JR, Jiang XW, Zhao CJ, Zhu CS. Fabrication of internal diffraction gratings in calcium fluoride crystals by a
focused femtosecond laser. Opt Express 2004;12:7426.
[156] Courrol LC, Samad RE, Gomes L, Ranieri IM, Baldochi SL, de Freitas AZ, et al. Color center production by femtosecond pulse
laser irradiation in LiF crystals. Opt Express 2004;12:28893.
[157] Dekker P, Ams M, Marshall GD, Little DJ, Withford MJ. Annealing dynamics of waveguide Bragg gratings: evidence of
femtosecond laser induced colour centres. Opt Express 2010;18:327483.
[158] Samad RE, Courrol LC, Gomes L, Ranieri IM, Baldochi SL, de Freitas AZ, et al. Fluorescence properties of colour centres
produced by ultrashort laser irradiation in LiF crystals. J Phys: Conf Ser 2010;249:012009.
[159] Pan SK, Jiang BX, Li HJ, Cheng QX, Jiang XW, Qiu JR, et al. Space-selective laser function defects induced by ultra-fast
intense laser in LiF crystal. J Cryst Growth 2005;275:e23337.
[160] Qiu JR, Miura K, Suzuki T, Mitsuyu T, Hirao K. Permanent photoreduction of Sm3+ to Sm2+ inside a sodium aluminoborate
glass by an infrared femtosecond pulsed laser. Appl Phys Lett 1999;74:102.
[161] Qiu JR, Kojima K, Miura K, Mitsuyu T, Hirao K. Infrared femtosecond laser pulse induced permanent reduction of Eu3+ to
Eu2+ in a fluorozirconate glass. Opt Lett 1999;24:7868.
[162] Qiu JR, Zhu CS, Nakaya T, Si JH, Kojima K, Ogura F, et al. Space-selective valence state manipulation of transition metal
ions inside glasses by a femtosecond laser. Appl Phys Lett 2001;79:35679.
[163] Miura K, Qiu JR, Fujiwara S, Sakasuchi S, Hirao K. Three-dimensional optical memory with rewriteable and ultrahigh
density using the valence-state change of samarium ions. Appl Phys Lett 2002;80:22635.
[164] Royon A, Petit Y, Papon G, Richardson M, Canioni L. Femtosecond laser induced photochemistry in materials tailored with
photosensitive agents [Invited]. Opt Mater Express 2011;1:86682.
[165] Yang LY, Da N, Chen DP, Zhao QZ, Jiang XW, Zhu CS, et al. Valence state change and refractive index change induced by
femtosecond laser irradiation in Sm3+ doped fluoroaluminate glass. J Non-Cryst Solids 2008;354:13525.
[166] Fujita K, Nishi M, Hirao K. Ultrashort-laser-pulse-induced persistent spectral hole burning of Eu3+ in sodium borate
glasses. Opt Lett 2001;26:16813.
[167] Fujita K, Yasumoto C, Hirao K. Photochemical reactions of samarium ions in sodium borate glasses irradiated with nearinfrared femtosecond laser pulses. J Lumin 2002;98:31723.
[168] Zhou SF, Lei WQ, Jiang N, Hao JH, Wu E, Zeng HP, et al. Space-selective control of luminescence inside the Bi-doped
mesoporous silica glass by a femtosecond laser. J Mater Chem 2009;19:46038.
[169] Lim J, Lee M, Kim E. Three-dimensional optical memory using photoluminescence change in Sm-doped sodium borate
glass. Appl Phys Lett 2005;86:191105.
[170] Lee S, Lee M, Lim K. Femtosecond laser induced PL change in Sm-doped sodium borate glass and 3D optical memory. J
Lumin 2007;122123:9902.
[171] Qiu JR, Miura K, Nouchi K, Suzuki T, Kondo Y, Mitsuyu T, et al. Valence manipulation by lasers of samarium ion in
micrometer-scale dimensions inside transparent glass. Solid State Commun 2000;113:3414.
[172] Park GJ, Hayakawa T, Nogami M. Spectral hole burning and fluorescence in femtosecond laser induced Sm2+-doped
glasses. J Lumin 2004;106:1038.
[173] Miura K, Qiu JR, Mitsuyu T, Hirao K. Space-selective growth of frequency-conversion crystals in glasses with ultrashort
infrared laser pulses. Opt Lett 2000;25:40810.
[174] Qiu JR, Shirai M, Nakaya T, Si JH, Jiang XW, Zhu CS, et al. Space-selective precipitation of metal nanoparticles inside
glasses. Appl Phys Lett 2002;81:30402.
[175] Yu BK, Chen B, Yang XY, Qiu JR, Jiang XW, Zhu CS, et al. Study of crystal formation in borate, niobate, and titanate glasses
irradiated by femtosecond laser pulses. J Opt Soc Am B 2004;21:837.
[176] Yonesaki Y, Miura K, Araki R, Fujita K, Hirao K. Space-selective precipitation of non-linear optical crystals inside silicate
glasses using near-infrared femtosecond laser. J Non-Cryst Solids 2005;351:88592.
[177] Dai Y, Zhu B, Qiu JR, Ma HL, Lu B, Cao SX, et al. Direct writing three-dimensional Ba2TiSi2O8 crystalline pattern in glass
with ultrashort pulse laser. Appl Phys Lett 2007;90:181109.
[178] Dai Y, Zhu B, Qiu JR, Ma HL, Lu B, Yu BK. Space-selective precipitation of functional crystals in glass by using a high
repetition rate femtosecond laser. Chem Phys Lett 2007;443:2537.
[179] Liu Y, Zhu B, Wang L, Dai Y, Ma HL, Lakshminarayana G, et al. Femtosecond laser direct writing of TiO2 crystalline patterns
in glass. Appl Phys B 2008;93:6137.
[180] Zhu B, Dai Y, Ma HL, Zhang S, Lin G, Qiu JR. Femtosecond laser induced space-selective precipitation of nonlinear optical
crystals in rare-earth-doped glasses. Opt Express 2007;15:606974.
[181] Liu Y, Zhu B, Dai Y, Qiao XS, Ye S, Teng Y, et al. Femtosecond laser writing of Er3+-doped CaF2 crystalline patterns in glass.
Opt Lett 2009;34:34335.
[182] Zhong MJ, Du YY, Ma HL, Han YM, Lu B, Dai Y, et al. Crystalline phase distribution of Dy2(MoO4)3 in glass induced by 250
kHz femtosecond laser irradiation. Opt Mater Express 2012;2:115664.
[183] Zhou SF, Jiang N, Miura K, Tanabe S, Shimizu M, Sakakura M, et al. Simultaneous tailoring of phase evolution and dopant
distribution in the glassy phase for controllable luminescence. J Am Chem Soc 2010;132:1794552.
[184] Qiu JR, Jiang XW, Zhu CS, Inouye H, Si Jh, Hirao K. Optical properties of structurally modified glasses doped with gold ions.
Opt Lett 2004;29:3702.
[185] Zeng HD, Qiu JR, Jiang XW, Zhu CS, Gan FX. The effect of femtosecond laser irradiation conditions on precipitation of silver
nanoparticles in silicate glasses. J Phys: Condens Matter 2004;16:29016.
[186] Zeng HD, Qiu JR, Jiang XW, Zhu CS, Gan FX. Effect of Al2O3 on the precipitation of Ag nanoparticles in silicate glasses. J
Cryst Growth 2004;262:2558.
[187] Dai Y, Qiu JR, Hu X, Yang L, Jiang XW, Zhu CS, et al. Effect of cerium oxide on the precipitation of silver nanoparticles in
femtosecond laser irradiated silicate glass. Appl Phys B 2006;84:5015.
[188] Zeng HD, Chen GR, Qiu JR, Jiang XW, Zhu CS, Gan FX. Effect of PbO on precipitation of laser-induced gold nanoparticles
inside silicate glasses. J Non-Cryst Solids 2008;354:11558.
[189] Dai Y, Yu G, He M, Ma H, Yan X, Ma G. High repetition rate femtosecond laser irradiation-induced elements redistribution
in Ag-doped glass. Appl Phys B 2011;103:6637.

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

221

[190] Zhao QZ, Qiu JR, Jiang XW, Zhao CJ, Zhu CS. Mechanisms of the refractive index change in femtosecond laser-irradiated
Au3+-doped silicate glasses. J Appl Phys 2004;96:71225.
[191] Zhao QZ, Qiu JR, Jiang XW, Zhao CJ, Zhu CS. Controllable precipitation and dissolution of silver nanoparticles in ultrafast
laser pulses irradiated Ag+-doped phosphate glass. Opt Express 2004;12:403540.
[192] Tu ZF, Teng Y, Zhou JJ, Zhou SF, Qiu JR. Micro-modification of luminescence property in Ag+-doped phosphate glass using
femtosecond laser irradiation. J Non-Cryst Solids 2014;383:1614.
[193] Dai Y, Hu X, Wang C, Chen DP, Jiang XW, Zhu CS, et al. Fluorescent Ag nanoclusters in glass induced by an infrared
femtosecond laser. Chem Phys Lett 2007;439:814.
[194] Hua B, Qiu JR, Shimotsuma Y, Fujita K, Hirao K. Photoreduction of Ag+ in aluminoborate glasses induced by irradiation of a
femtosecond laser. J Mater Res 2005;20:6448.
[195] Takeshima N, Kuroiwa Y, Narita Y, Tanaka S, Hirao K. Precipitation of silver particles by femtosecond laser pulses inside
silver ion doped glass. J Non-Cryst Solids 2004;336:2346.
[196] Royon A, Bourhis K, Bellec M, Papon G, Bousquet B, Deshayes Y, et al. Silver clusters embedded in glass as a perennial high
capacity optical recording medium. Adv Mater 2010;22:52826.
[197] Maurel C, Cardinal T, Bellec M, Canioni L, Bousquet B, Treguer M, et al. Luminescence properties of silver zinc phosphate
glasses following different irradiations. J Lumin 2009;129:15148.
[198] Bourhis K, Royon A, Papon G, Canioni L, Makria N, Petit Y, et al. Luminescence properties of micrometric structures
induced by direct laser writing in silver containing phosphate glass. J Non-Cryst Solids 2013;377:1425.
[199] Zeng HD, Yang YX, Jiang XW, Chen GR, Qiu JR, Gan FX. Preparation and optical properties of silicate glasses containing Pd
nanoparticles. J Cryst Growth 2005;280:51620.
[200] Teng Y, Qian B, Jiang N, Liu Y, Luo FF, Ye S, et al. Light and heat driven precipitation of copper nanoparticles inside Cu2+doped borate glasses. Chem Phys Lett 2010;485:914.
[201] Teng Y, Zhou JJ, Luo FF, Lin G, Qiu JR. Controllable space selective precipitation of copper nanoparticles in borosilicate
glasses using ultrafast laser irradiation. J Non-Cryst Solids 2011;357:23803.
[202] Miura K, Hirao K, Shimotsuma Y, Sakakura M, Kanehira S. Formation of Si structure in glass with a femtosecond laser.
Appl Phys A 2008;93:1838.
[203] Lin G, Pan HH, Dai Y, He F, Chen DP, Cheng Y, et al. Formation of Si nanocrystals in glass by femtosecond laser
micromachining. Mater Lett 2011;65:35447.
[204] Jiang N, Su D, Qiu JR, Spence JCH. On the formation of Na nanoparticles in femtosecond-laser irradiated glasses. J Appl
Phys 2010;107:064301.
[205] Lin G, Pan HH, Qiu JR, Zhao QZ. Nonlinear optical properties of lead nanocrystals embedding glass induced by thermal
treatment and femtosecond laser irradiation. Chem Phys Lett 2011;516:18691.
[206] Lin G, Luo FF, He F, Teng Y, Tan WJ, Si JH, et al. Space-selective precipitation of Ge crystalline patterns in glasses by
femtosecond laser irradiation. Opt Lett 2011;36:2624.
[207] Qu SL, Zhang YW, Li HJ, Qiu JR, Zhu CS. Nanosecond nonlinear absorption in Au and Ag nanoparticles precipitated glasses
induced by a femtosecond laser. Opt Mater 2006;28:25965.
[208] Nakashima S, Sugioka K, Tanaka K, Shimizu M, Shimotsuma Y, Miura K, et al. Plasmonically enhanced Faraday effect in
metal and ferrite nanoparticles composite precipitated inside glass. Opt Express 2012;20:281919.
[209] Nakashima S, Sugioka K, Tanaka K, Midorikawa K, Mukai K. Optical and magneto-optical properties in Fe-doped glasses
irradiated with femtosecond laser. Appl Phys B 2013;113:4516.
[210] Liu C, Kwon YK, Heow J, Kim BH, Sohn IB. Controlled precipitation of lead sulfide quantum dots in glasses using the
femtosecond laser pulses. J Am Ceram Soc 2010;93:12214.
[211] Hamzaoui EH, Bernard R, Chahadih A, Chassagneux F, Bois L, Jegouso D, et al. Laser-induced direct space-selective
precipitation of CdS nanoparticles embedded in a transparent silica xerogel. Nanotechnology 2010;21:134002.
[212] Chahadih A, Hamzaoui HE, Bernard R, Boussekey L, Bois L, Cristini O, et al. Direct-writing of PbS nanoparticles inside
transparent porous silica monoliths using pulsed femtosecond laser irradiation. Nanoscale Res Lett 2011;6:542.
[213] Mardilovich P, Fletcher LB, Troy NW, Yang L, Huang H, Risbud SH. Ultrafast laser fabrication of hybrid micro- and nanostructures in semiconductor-doped borosilicate glasses. Int J Appl Glass Sci 2013;4:8799.
[214] Day D, Gu M. Formation of voids in a doped polymethylmethacrylate polymer. Appl Phys Lett 2002;80:24046.
[215] Schaffer CB, Glezer EN, Nishimura N, Mazur E. Ultrafast laser induced microexplosions: explosive dynamics and submicrometer structures. Proc SPIE 1998;3269:3645.
[216] Watanabe W, Toma T, Yamada K, Nishii J, Hayashi K, Itoh K. Optical seizing and merging of voids in silica glass with
infrared femtosecond laser pulses. Opt Lett 2000;25:166971.
[217] Yamasaki K, Juodkazis S, Watanabe M, Sun HB, Matsuo S, Misawa H. Recording by microexplosion and two-photon
reading of three-dimensional optical memory in polymethylmethacrylate films. Appl Phys Lett 2000;76:10002.
[218] McGrane SD, Grieco A, Ramos KJ, Hooks DE, Moore DS. Femtosecond micromachining of internal voids in high explosive
crystals for studies of hot spot initiation. J Appl Phys 2009;105:073505.
[219] Juodkazis S, Misawa H, Hashimoto T, Gamaly EG, Luther-Davies B. Laser-induced microexplosion confined in a bulk of
silica: formation of nanovoids. Appl Phys Lett 2006;88:201909.
[220] Gamaly EG, Juodkazis S, Nishimura K, Misawa H, Luther-Davies B, Hallo L, et al. Laser matter interaction in the bulk of a
transparent solid: confined microexplosion and void formation. Phys Rev B 2006;73:214101.
[221] Hashimoto T, Juodkazis S, Misawa H. Void formation in glasses. New J Phys 2007;9:253.
[222] Zhou GY, Ventura MJ, Vanner MR, Gu M. Use of ultrafast-laser-driven microexplosion for fabricating three-dimensional
void-based diamond-lattice photonic crystals in a solid polymer material. Opt Lett 2004;29:22402.
[223] Zhou G, Ventura MJ, Vanner MR, Gu M. Fabrication and characterization of face-centered-cubic void dots photonic
crystals in a solid polymer material. Appl Phys Lett 2005;86:011108.
[224] Ventura MJ, Straub M, Gu M. Void channel microstructures in resin solids as an efficient way to infrared photonic crystals.
Appl Phys Lett 2003;82:1649.
[225] Brasselet E, Royon A, Canioni L. Dense arrays of microscopic optical vortex generators from femtosecond direct laser
writing of radial birefringence in glass. Appl Phys Lett 2012;100:181901.

222

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

[226] Meunier T, Villafranca AB, Bhardwaj R, Weck A. Mechanism for spherical dome and microvoid formation in polycarbonate
using nanojoule femtosecond laser pulses. Opt Lett 2012;37:316870.
[227] Zhou G, Gu M. Anisotropic properties of ultrafast laser-driven microexplosions in lithium niobate crystal. Appl Phys Lett
2005;87:241107.
[228] Gouya K, Watanabe K. Micro-void arrays in an optical fiber machined by a femto-second laser for obtaining bending
direction sensitive sensors. Proc SPIE 2013;8677:86770Q.
[229] Thomas J, Voigtlnder C, Becker RG, Richter D, Tnnermann A, Nolte S. Femtosecond pulse written fiber gratings: a new
avenue to integrated fiber technology. Laser Photonics Rev 2012;6:70923.
[230] Kondo Y, Nouchi K, Mitsuyu T, Watanabe M, Kazansky PG, Hirao K. Fabrication of long-period fiber gratings by focused
irradiation of infrared femtosecond laser pulses. Opt Lett 1999;24:6468.
[231] Richter S, Dring S, Burmeister F, Zimmermann F, Tnnermann A, Nolte S. Formation of periodic disruptions induced by
heat accumulation of femtosecond laser pulses. Opt Express 2013;21:1545263.
[232] Malinauskasa M, Farsarib M, Piskarskasa A, Juodkazis S. Ultrafast laser nanostructuring of photopolymers: a decade of
advances. Phys Rep 2013;533:131.
[233] Maruo S, Nakamura O, Kawata S. Three-dimensional microfabrication with two-photon-absorbed photopolymerization.
Opt Lett 1997;22:1324.
[234] Cumpston BH, Ananthavel SP, Barlow S, Dyer DL, Ehrlich JE, Erskine LL, et al. Two-photon polymerization initiators for
three-dimensional optical data storage and microfabrication. Nature 1999;398:514.
[235] Sun HB, Matsuo S, Misawa H. Three-dimensional photonic crystal structures achieved with two-photon-absorption
photopolymerization of resin. Appl Phys Lett 1999;74:7869.
[236] Deubel M, Freymann GV, Wegener M, Pereira S, Busch K, Soukoulis CM. Direct laser writing of three-dimensional
photonic-crystal templates for telecommunications. Nat Mater 2004;3:4447.
[237] Zhou WH, Kuebler SM, Braun KL, Yu TY, Cammack JK, Ober CK, et al. An efficient two-photon-generated photoacid applied
to positive-tone 3D microfabrication. Science 2002;296:11069.
[238] Juodkazis S, Mizeikis V, Misawa H. Three-dimensional microfabrication of materials by femtosecond lasers for photonics
applications. J Appl Phys 2009;106:051101.
[239] Gan ZS, Cao YY, Evans RA, Gu M. Three-dimensional deep sub-diffraction optical beam lithography with 9 nm feature size.
Nat Commun 2013;4:2061.
[240] Hossain MM, Gu M. Fabrication methods of 3D periodic metallic nano/microstructures for photonics applications. Laser
Photonics Rev 2014;2:23349.
[241] Soukoulis CM, Wegener M. Past achievements and future challenges in the development of three-dimensional photonic
metamaterials. Nat Photonics 2011;5:52330.
[242] Lee W, Pruzinsky SA, Braun PV. Multi-photon polymerization of waveguide structures within three-dimensional photonic
crystals. Adv Mater 2002;14:2714.
[243] Fischer J, Wegener M. Three-dimensional optical laser lithography beyond the diffraction limit. Laser Photonics Rev
2013;7:2244.
[244] Malinauskas M, Zukauskas A, Bickauskaite G, Gadonas R, Juodkazis S. Mechanisms of three-dimensional structuring of
photo-polymers by tightly focussed femtosecond laser pulses. Opt Express 2010;18:1020921.
[245] Seet KK, Mizeikis V, Matsuo S, Juodkazis S, Misawa H. Three-dimensional spiral-architecture photonic crystals obtained
by direct laser writing. Adv Mater 2005;17:5415.
[246] Sekkat Z, Kawata S. Laser nanofabrication in photoresists and azopolymers. Laser Photonics Rev 2014;1:126.
[247] Korte F, Koch J, Serbin J, Ovsianikov A, Chichkov BN. Three-dimensional nanostructuring with femtosecond laser pulses.
IEEE Trans Nanotechnol 2004;3:46872.
[248] Zhang YL, Chen QD, Xia H, Sun HB. Designable 3D nanofabrication by femtosecond laser direct writing. Nano Today
2010;5:43548.
[249] Jia BH, Buso D, van Embden J, Li JF, Gu M. Highly non-linear quantum dot doped nanocomposites for functional threedimensional structures generated by two-photon polymerization. Adv Mater 2010;22:24637.
[250] Serbin J, Egbert A, Ostendorf A, Chichkov BN, Houbertz R, Domann G, et al. Femtosecond laser-induced two-photon
polymerization of inorganicorganic hybrid materials for applications in photonics. Opt Lett 2003;28:3013.
[251] Ushiba S, Shoji S, Masui K, Kuray P, Kono J, Kawata S. 3D microfabrication of single-wall carbon nanotube/polymer
composites by two-photon polymerization lithography. Carbon 2013;59:2838.
[252] Ventura MJ, Bullen C, Gu M. Direct laser writing of three-dimensional photonic crystal lattices within a PbS quantum-dotdoped polymer material. Opt Express 2007;15:181722.
[253] Turner MD, Saba M, Zhang QM, Cumming BP, Schrder-Turk GE, Gu M. Miniature chiral beamsplitter based on gyroid
photonic crystals. Nat Photonics 2013;7:8015.
[254] Zukauskas A, Malinauskas M, Reinhardt C, Chichkov BN, Gadonas R. Closely packed hexagonal conical microlens array
fabricated by direct laser photopolymerization. Appl Opt 2012;51:49955003.
[255] Brasselet E, Malinauskas M, Zukauskas A, Juodkazis S. Photopolymerized microscopic vortex beam generators: precise
delivery of optical orbital angular momentum. Appl Phys Lett 2010;97:211108.
[256] Baldacchini T, Snider S, Zadoyan R. Two-photon polymerization with variable repetition rate bursts of femtosecond laser
pulses. Opt Express 2012;20:298909.
[257] Fischer J, Mueller JB, Kaschke J, Wolf TJA, Unterreiner AN, Wegener M. Three-dimensional multi-photon direct laser
writing with variable repetition rate. Opt Express 2013;21:2624460.
[258] Cao YY, Gan ZS, Jia BH, Evans RA, Gu M. High-photosensitive resin for super-resolution direct-laser-writing based on
photoinhibited polymerization. Opt Express 2011;19:1948694.
[259] Ma J, Cheng WJ, Zhang S, Feng DH, Jia TQ, Sun ZR, et al. Coherent quantum control of two-photon absorption and
polymerization by shaped ultrashort laser pulses. Laser Phys Lett 2013;10:085304.
[260] Sun Q, Liang F, Valle R, Chin SL. Nanograting formation on the surface of silica glass by scanning focused femtosecond
laser pulses. Opt Lett 2008;33:27135.

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

223

[261] Huang M, Zhao FL, Cheng Y, Xu NS, Xu ZZ. Origin of laser-induced near-subwavelength ripples: interference between
surface plasmons and incident laser. ACS Nano 2009;3:406270.
[262] Kim D, Jang W, Kim T, Moon A, Lim KS, Lee M. Nanostructure and microripple formation on the surface of sapphire with
femtosecond laser pulses. J Appl Phys 2012;111:093518.
[263] Bonse J, Krger J, Hhm S, Rosenfeld A. Femtosecond laser-induced periodic surface structures. J Laser Appl
2012;24:042006.
[264] Hhm S, Rosenfeld A, Krger J, Bonse J. Femtosecond laser-induced periodic surface structures on silica. J Appl Phys
2012;112:014901.
[265] Liang F, Valle R, Chin SL. Mechanism of nanograting formation on the surface of fused silica. Opt Express
2012;2:438996.
[266] Straub M, Weigand B, Afshar M, Feili D, Seidel H, Knig K. Periodic subwavelength ripples on lithium niobate surfaces
generated by tightly focused sub-15 femtosecond sub-nanojoule pulsed near-infrared laser light. J Opt 2013;15:055601.
[267] Bonse J, Munz M, Sturm H. Structure formation on the surface of indium phosphide irradiated by femtosecond laser
pulses. J Appl Phys 2005;97:013538.
[268] Buividas R, Rektyte S, Malinauskas M, Juodkazis S. Nano-groove and 3D fabrication by controlled avalanche using
femtosecond laser pulses. Opt Mater Express 2013;3:167486.
[269] Shi XS, Jiang L, Li X, Wang SM, Yuan YP, Lu YF. Femtosecond laser-induced periodic structure adjustments based on
electron dynamics control: from subwavelength ripples to double-grating structures. Opt Lett 2013;38:37436.
[270] Liang F, Valle R, Gingras D, Chin SL. Role of ablation and incubation processes on surface nanograting formation. Opt
Mater Express 2011;1:124450.
[271] Das SK, Messaoudi H, Debroy A, McGlynn E, Grunwald R. Multiphoton excitation of surface plasmon-polaritons and
scaling of nanoripple formation in large bandgap materials. Opt Mater Express 2013;3:170515.
[272] Han YH, Zhao XL, Qu SL. Polarization dependent ripples induced by femtosecond laser on dense flint (ZF6) glass. Opt
Express 2011;19:191505.
[273] Jiang L, Shi XS, Li X, Yuan YP, Wang C, Lu YF. Subwavelength ripples adjustment based on electron dynamics control by
using shaped ultrafast laser pulse trains. Opt Express 2012;20:2150511.
[274] Zhou GS, Fauchet PM, Siegman AE. Growth of spontaneous periodic surface structures on solids during laser illumination.
Phys Rev B 1982;26:536681.
[275] Skolski JZP, Rmer GRBE, Obona JV, Ocelik V, Huis in t Veld AJ, De Hosson JThM. Laser-induced periodic surface
structures: Fingerprints of light localization. Phys Rev B 2012;85:075320.
[276] Yuan YP, Jiang L, Li X, Wang C, Xiao H, Lu YF, et al. Formation mechanisms of sub-wavelength ripples during femtosecond
laser pulse train processing of dielectrics. J Phys D Appl Phys 2012;45:175301.
[277] Rosenfelda A, Rohloffa M, Hhma S, Krgerb J, Bonse J. Formation of laser-induced periodic surface structures on fused
silica upon multiple parallel polarized double-femtosecond-laser-pulse irradiation sequences. Appl Surf Sci
2012;258:92336.
[278] Hhm S, Rohloff M, Rosenfeld A, Krger J, Bonse J. Dynamics of the formation of laser-induced periodic surface structures
on dielectrics and semiconductors upon femtosecond laser pulse irradiation sequences. Appl Phys A 2013;110:5537.
[279] Du G, Yang Q, Chen F, Bian H, Meng X, Si JH. Ultrafast electron dynamics manipulation of laser induced periodic ripples via
a train of shaped pulses. Laser Phys Lett 2013;10:026003.
[280] Messaddeq SH, Valle R, Soucy P, Bernier M, El-Amraoui M, Messaddeq Y. Self-organized periodic structures on Ge-S
based chalcogenide glass induced by femtosecond laser irradiation. Opt Express 2012;20:298829.
[281] Buividas R, Rosa L, liupas R, Kudrius T, lekys G, Datsyuk V, et al. Mechanism of fine ripple formation on surfaces of
(semi)transparent materials via a half-wavelength cavity feedback. Nanotechnology 2011;22:055304.
[282] Liang F, Sun Q, Gingras D, Valle R, Chin SL. The transition from smooth modification to nanograting in fused silica. Appl
Phys Lett 2010;96:101903.
[283] Liang F, Valle R, Chin SL. Pulse fluence dependent nanograting inscription on the surface of fused silica. Appl Phys Lett
2012;100:251105.
[284] Liang F, Valle R, Chin SL. Physical evolution of nanograting inscription on the surface of fused silica. Opt Mater Express
2012;2:9006.
[285] Rebollar E, Vzquez de Aldana JR, Martn-Fabiani I, Ma Hernndez, Rueda DR, Ezquerra TA, et al. Assessment of
femtosecond laser induced periodic surface structures on polymer films. Phys Chem Chem Phys 2013;15:1128798.
[286] Liu Y, Brelet Y, He ZB, Yu LW, Forestier B, Deng YK, et al. Laser-induced periodic annular surface structures on fused silica
surface. Appl Phys Lett 2013;102:251103.
[287] Liu Y, Brelet Y, He ZB, Yu LW, Mitryukovskiy S, Houard A, et al. Ciliary white light: optical aspect of ultrashort laser
ablation on transparent dielectrics. Phys Rev Lett 2013;110:097601.
[288] Misawa H. Japan Patent 08-220688; 1996.
[289] Watanabe M, Sun HB, Juodkazis S, Takahashi T, Matsuo S, Suzuki Y, et al. Three-dimensional optical data storage in
vitreous silica. Jpn J Appl Phys 1998;37. L1527-30.
[290] Fernandes LA, Grenier JR, Herman PR, Aitchison JS, Marques PVS. Stress induced birefringence tuning in femtosecond
laser fabricated waveguides in fused silica. Opt Express 2012;20:2410314.
[291] Crespi A, Ramponi R, Osellame R, Sansoni L, Bongioanni I, Sciarrino F, et al. Integrated photonic quantum gates for
polarization qubits. Nat Commun 2011;2:566.
[292] Spagnolo N, Vitelli C, Aparo L, Mataloni P, Sciarrino F, Crespi A, et al. Three-photon bosonic coalescence in an integrated
tritter. Nat Commun 2013;4:1606.
[293] Corrielli G, Crespi A, Valle GD, Longhi S, Osellame R. Fractional Bloch oscillations in photonic lattices. Nat Commun
2013;4:1555.
[294] Tillmann M, Dakic B, Heilmann R, Nolte S, Szameit A, Walther P. Experimental boson sampling. Nat Photonics
2013;7:5404.
[295] Crespi A, Osellame R, Ramponi R, Brod DJ, Galvo EF, Spagnolo N, et al. Integrated multimode interferometers with
arbitrary designs for photonic boson sampling. Nat Photonics 2013;7:5459.

224

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

[296] Chan JW, Huster TR, Risbud SH, Krol DM. Modification of the fused silica glass network associated with waveguide
fabrication using femtosecond laser pulses. Appl Phys A 2003;76:36772.
[297] Saliminia A, Nguyen NT, Chin SL, Valle R. Densification of silica glass induced by 0.8 and 1.5 m intense femtosecond laser
pulses. J Appl Phys 2006;99:093104.
[298] Gross S, Ams M, Palmer G, Miese CT, Williams RJ, Marshall GD, et al. Ultrafast laser inscription in soft glasses: a
comparative study of athermal and thermal processing regimes for guided wave optics. Int J Appl Glass Sci
2012;3:33248.
[299] Bressel L, de Ligny D, Sonneville C, Martinez V, Mizeikis V, Buividas R, et al. Femtosecond laser induced density changes in
GeO2 and SiO2 glasses: fictive temperature effect. Opt Mater Express 2011;1:60513 [Invited].
[300] Lin G, Luo FF, He F, Chen QX, Chen DP, Cheng Y. Different refractive index change behavior in borosilicate glasses induced
by 1 kHz and 250 kHz femtosecond lasers. Opt Mater Express 2011;1:72431.
[301] Little DJ, Ams M, Dekker P, Marshall GD, Withford MJ. Mechanism of femtosecond-laser induced refractive index change
in phosphate glass under a low repetition-rate regime. J Appl Phys 2010;108:033110.
[302] Ponader CW, Schroeder JF, Streltsov AM. Origin of the refractive-index increase in laser-written waveguides in glasses. J
Appl Phys 2008;103:063516.
[303] Sakakura M, Terazima M. Oscillation of the refractive index at the focal region of a femtosecond laser pulse inside a glass.
Opt Lett 2004;29:154850.
[304] Sakakura M, Terazima M, Shimotsuma Y, Miura K, Hirao K. Thermal and shock induced modification inside a silica glass
by focused femtosecond laser pulse. J Appl Phys 2011;109:023503.
[305] Chan JW, Huser T, Risbud S, Krol DM. Structural changes in fused silica after exposure to focused femtosecond laser
pulses. Opt Lett 2001;26:17268.
[306] Fletcher LB, Witcher JJ, Reichman WB, Arai A, Bovatsek J, Krol DM. Changes to the network structure of ErYb doped
phosphate glass induced by femtosecond laser pulses. J Appl Phys 2009;106:083107.
[307] Petit L, Carlie N, Anderson T, Richardson M, Richardson K. Progress on the photoresponse of chalco-genide glasses and
films to near-infrared femtosecond laser irradiation: a review. IEEE J Sel Top Quantum Electron 2008;14:132334.
[308] Cho SH, Kumagai H, Midorikawa K. In situ observation of dynamics of plasma formation and refractive index modification
in silica glasses excited by a femtosecond laser. Opt Commun 2002;207:24353.
[309] Eaton SM, Ng ML, Osellame R, Herman PR. High refractive index contrast in fused silica waveguides by tightly focused,
high-repetition rate femtosecond laser. J Non-Cryst Solids 2011;357:238791.
[310] Bhardwaj VR, Simova E, Corkum PB, Rayner DM, Hnatovsky C, Taylor RS, et al. Femtosecond laser-induced refractive index
modification in multicomponent glasses. J Appl Phys 2005;97:083102.
[311] Toney Fernandez T, Haro-Gonzlez P, Sotillo B, Hernandez M, Jaque D, Fernandez P, et al. Ion migration assisted
inscription of high refractive index contrast waveguides by femtosecond laser pulses in phosphate glass. Opt Lett
2013;38:524851.
[312] Tikhomirov VK, Seddon AB, Koch J, Wandt D, Chichkov BN. Fabrication of buried waveguides and nanocrystals in Er3+doped oxyfluoride glass. Phys Stat Sol 2005;202:R735.
[313] Little DJ, Ams M, Gross S, Dekker P, Miese CT, Fuerbach A, et al. Structural changes in BK7 glass upon exposure to
femtosecond laser pulses. J Raman Spectrosc 2011;42:7158.
[314] Gross S, Lancaster DG, Ebendorff-Heidepriem H, Monro TM, Fuerbach A, Withford MJ. Femtosecond laser induced
structural changes in fluorozirconate glass. Opt Mater Express 2013;3:57483.
[315] Ams M, Dekker P, Marshall GD, Withford MJ. Monolithic 100 mw Yb waveguide laser fabricated using the femtosecondlaser direct-write technique. Opt Lett 2009;34:2479.
[316] Palmer G, Gross S, Fuerbach A, Lancaster DG, Withford MJ. High slope efficiency and high refractive index change in
direct-written Yb-doped waveguide lasers with depressed claddings. Opt Express 2013;21:1741320.
[317] Gross S, Arriola A, Palmer G, Jovanovic N, Spaleniak I, Meany TD, et al. Ultrafast laser inscribed 3D integrated photonics.
Proc SPIE 2013;8876:887604.
[318] Crespi A, Osellame R, Ramponi R, Giovannetti V, Fazio R, Sansoni L, et al. Anderson localization of entangled photons in an
integrated quantum walk. Nat Photonics 2013;7:3228.
[319] Spagnolo N, Vitelli C, Sansoni L, Maiorino E, Mataloni P, Sciarrino F, et al. General rules for bosonic bunching in multimode
interferometers. Phys Rev Lett 2013;111:130503.
[320] Spagnolo N, Aparo L, Vitelli C, Crespi A, Ramponi R, Osellame R, et al. Quantum interferometry with three-dimensional
geometry. Sci Rep 2012;2:862.
[321] Lee KKC, Mariampillai A, Haque M, Standish BA, Yang VXD, Herman PR. Temperature-compensated fiber-optic 3D shape
sensor based on femtosecond laser direct-written Bragg grating waveguides. Opt Express 2013;21:2407686.
[322] Fernandes LA, Grenier JR, Marques PVS, Aitchison JS, Herman PR. Strong birefringence tuning of optical waveguides with
femtosecond laser irradiation of bulk fused silica and single mode fibers. J Lightwave Technol 2012;31:35639.
[323] Grenier JR, Fernandes LA, Herman PR. Femtosecond laser writing of optical edge filters in fused silica optical waveguides.
Opt Express 2013;21:4493502.
[324] Heinrich M, Szameit A, Dreisow F, Keil R, Minardi S, Pertsch T, et al. Observation of three-dimensional discrete-continuous
X waves in photonic lattices. Phys Rev Lett 2009;103:113903.
[325] Keil R, Heinrich M, Dreisow F, Pertsch T, Tnnermann A, Nolte S, et al. All-optical routing and switching for threedimensional photonic circuitry. Sci Rep 2011;1:94.
[326] Dreisow F, Longhi S, Nolte S, Tnnermann A, Szameit A. Vacuum instability and pair production in an optical setting. Phys
Rev Lett 2012;109:110401.
[327] Chen F, de Aldana JRV. Optical waveguides in crystalline dielectric materials produced by femtosecond-laser
micromachining. Laser Photonics Rev 2014;2:25175.
[328] Dong MM, Wang CW, Wu ZX, Zhang Y, Pan HH, Zhao QZ. Waveguides fabricated by femtosecond laser exploiting both
depressed cladding and stress-induced guiding core. Opt Express 2013;21:155229.
[329] Liu HL, Jia YC, Vzquez de Aldana JR, Jaque D, Chen F. Femtosecond laser inscribed cladding waveguides in Nd:YAG
ceramics: fabrication, fluorescence imaging and laser performance. Opt Express 2012;20:186209.

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

225

[330] He RY, An Q, Vzquez de Aldana JR, Lu QM, Chen F. Femtosecond-laser micromachined optical waveguides in Bi4Ge3O12
crystals. Appl Opt 2013;52:37138.
[331] Liu HL, Jia YC, Chen F, Vzquez de Aldana JR. Continuous wave laser operation in Nd:GGG depressed tubular cladding
waveguides produced by inscription of femtosecond laser pulses. Opt Mater Express 2013;3:27883.
[332] Jia YC, Vzquez de Aldana JR, Lu QM, Jaque D, Chen F. Second harmonic generation of violet light in femtosecond-laserinscribed BiB3O6 cladding waveguides. Opt Mater Express 2013;3:127984.
[333] Ren YY, Chen F, Vzquez de Aldana JR. Near-infrared lasers and self-frequency-doubling in Nd:YCOB cladding
waveguides. Opt Express 2013;21:115627.
[334] Jia YC, Vzquez de Aldana JR, Che F. Efficient waveguide lasers in femtosecond laser inscribed double-cladding
waveguides of Yb:YAG ceramics. Opt Mater Express 2013;3:64550.
[335] Wagner R, Gottmann J. Sub-wavelength ripple formation on various materials induced by tightly focused femtosecond
laser radiation. J Phys: Conf Ser 2007;59:3337.
[336] Bulgakova NM, Zhukov VP, Meshcheryakov YP. Theoretical treatments of ultrashort pulse laser processing of transparent
materials: toward understanding the volume nanograting formation and quill writing effect. Appl Phys B
2013;113:43749.
[337] ktem B, Pavlov I, Ilday S, Kalaycolu H, Rybak A, Yavas S, et al. Nonlinear laser lithography for indefinitely large-area
nanostructuring with femtosecond pulses. Nat Photonics 2013;7:897901.
[338] Yuan YP, Jiang L, Li X, Wang C, Qu LT, Lu YF. Simulation of rippled structure adjustments based on localized transient
electron dynamics control by femtosecond laser pulse trains. Appl Phys A 2013;111:8139.
[339] Huang M, Zhao FL, Cheng Y, Xu NS, Xu ZZ. Mechanisms of ultrafast laser-induced deep-subwavelength gratings on
graphite and diamond. Phys Rev B 2009;79:125436.
[340] Shinoda M, Gattass RR, Mazur E. Femtosecond laser-induced formation of nanometer-width grooves on synthetic singlecrystal diamond surfaces. J Appl Phys 2009;105:053102.
[341] Gottmann J, Wortmann D, Hrstmann-Jungemann M. Fabrication of sub-wavelength surface ripples and in-volume
nanostructures by fs-laser induced selective etching. Appl Surf Sci 2009;255:56416.
[342] Jia TQ, Chen HX, Huang M, Zhao FL, Qiu JR, Li RX, et al. Formation of nanogratings on the surface of a ZnSe crystal
irradiated by femtosecond laser pulses. Phys Rev B 2005;72:125429.
[343] Guo Z, Qu S, Han Y, Liu S. Multi-photon fabrication of two-dimensional periodic structure by three interfered
femtosecond laser pulses on the surface of the silica glass. Opt Commun 2007;280:236.
[344] Guo Z, Qu S, Ran L, Han Y, Liu S. Formation of two-dimensional periodic microstructures by a single shot of three
interfered femtosecond laser pulses on the surface of silica glass. Opt Lett 2008;33:23835.
[345] Han Y, Qu S. Two-dimensional hat-scratch periodic structures induced by a single femtosecond laser pulse on silica
glass. Appl Surf Sci 2009;255:75248.
[346] Guo Z, Qu S, Liu S, Lee JH. Periodic microstructures induced by interfered femtosecond laser pulses. Proc SPIE
2010;7657:76570K.
[347] Wagner R, Gottmann J, Horn A, Kreutz EW. Formation of sub-wavelength laser induced periodic surface structures by
tightly focused femtosecond laser radiation. Proc SPIE 2004;5662:16872.
[348] Wu QH, Ma YR, Fang RC, Liao Y, Yu QX, Chen XL, et al. Femtosecond laser-induced periodic surface structure on diamond
film. Appl Phys Lett 2003;82:17035.
[349] Huang M, Zhao FL, Cheng Y, Xu NS, Xu ZZ. Large area uniform nanostructures fabricated by direct femtosecond laser
ablation. Opt Express 2008;16:1935465.
[350] Pan J, Jia TQ, Huo YY, Jia X, Feng DH, Zhang S, et al. Great enhancement of near band-edge emission of ZnSe twodimensional complex nanostructures fabricated by the interference of three femtosecond laser beams. J Appl Phys
2013;114:093102.
[351] Watanabe W, Itoh K. Motion of bubble in solid by femtosecond laser pulses. Opt Express 2002;10:6038.
[352] Yang WJ, Kazansky PG, Shimotsuma Y, Sakakura M, Miura K, Hirao K. Ultrashort-pulse laser calligraphy. Appl Phys Lett
2008;93:171109.
[353] Luo FF, Lin G, Sun HY, Zhang G, Li L, Chen DP, et al. Generation of bubbles in glass by a femtosecond laser. Opt Commun
2011;284:45925.
[354] Bellouard Y, Hongler MO. Femtosecond-laser generation of self-organized bubble patterns in fused silica. Opt Express
2011;19:680721.
[355] Brenner MP, Hilgenfeldt S, Lohse D. Single-bubble sonoluminescence. Rev Mod Phys 2002;74:42584.
[356] Lim K, Quinto-Su PA, Klaseboer E, Khoo BC, Venugopalan V, Ohl C. Nonspherical laser-induced cavitation bubbles. Phys
Rev E 2010;81:016308.
[357] Graf R, Fernandez A, Dubov M, Brueckner HJ, Chichkov BN, Apolonski A. Pearl-chain waveguides written at megahertz
repetition rate. Appl Phys B 2007;87:217.
[358] Bricchia E, Kazansky PG. Extraordinary stability of anisotropic femtosecond direct-written structures embedded in silica
glass. Appl Phys Lett 2006;88:111119.
[359] Beresna M, Geceviius M, Lancry M, Poumellec B, Kazansky PG. Broadband anisotropy of femtosecond laser induced
nanogratings in fused silica. Appl Phys Lett 2013;103:131903.
[360] Mauclair C, Zamfirescu M, Colombier JP, Cheng G, Mishchik K, Audouard E, et al. Control of ultrafast laser-induced bulk
nanogratings in fused silica via pulse time envelopes. Opt Express 2012;20:129973005.
[361] Mills JD, Kazansky PG, Bricchi E, Baumberg JJ. Embedded anisotropic microreflectors by femtosecond-laser
nanomachining. Appl Phys Lett 2002;81:1968.
[362] Cheng G, Mishchik K, Mauclair C, Audouard E, Stoian R. Ultrafast laser photoinscription of polarization sensitive devices in
bulk silica glass. Opt Express 2009;17:951525.
[363] Beresna M, Gecevicius M, Kazansky PG, Gertus T. Radially polarized optical vortex converter created by femtosecond laser
nanostructuring of glass. Appl Phys Lett 2011;98:201101.
[364] Shimotsuma Y, Sakakura M, Kazansky PG, Beresna M, Qiu JR, Miura K, et al. Ultrafast manipulation of self-assembled form
birefringence in glass. Adv Mater 2010;22:403943.

226

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

[365] Beresna M, Gecevicius M, Kazansky PG, Taylor T, Kavokin AV. Exciton mediated self-organization in glass driven by
ultrashort light pulses. Appl Phys Lett 2012;101:053120.
[366] Cai WJ, Libertun AR, Piestun R. Polarization selective computer-generated holograms realized in glass by femtosecond
laser induced nanogratings. Opt Express 2006;14:378591.
[367] Bhardwaj VR, Simova E, Rajeev PP, Hnatovsky C, Taylor RS, Rayner DM, et al. Optically produced arrays of planar
nanostructures inside fused silica. Phys Rev Lett 2006;96:057404.
[368] Rajeev P, Gertsvolf M, Hnatovsky C, Simova E, Taylor R, Corkum P, et al. Transient nanoplasmonics inside dielectrics. J
Phys B Atom Mol Opt Phys 2007;40. S273-82.
[369] Umran FA, Liao Y, Elias MM, Sugioka K, Stoian R, Cheng GH, et al. Formation of nanogratings in a transparent material
with tunable ionization property by femtosecond laser irradiation. Opt Express 2013;21:1525967.
[370] Born M, Wolf E. Principles of optics: electromagnetic theory of propagation, interference and diffraction of light. 7th
ed. Cambridge: Cambridge University Press; 1999. 986.
[371] Lancry M, Poumellec B, Canning J, Cook K, Poulin JC, Brisset F. Ultrafast nanoporous silica formation driven by
femtosecond laser irradiation. Laser Photonics Rev 2013;7:95362.
[372] Hnatovsky C, Taylor RS, Simova E, Bhardwaj VR, Rayner DM, Corkum PB. Polarization-selective etching in femtosecond
laser-assisted microfluidic channel fabrication in fused silica. Opt Lett 2005;30:18679.
[373] Champion A, Bellouard Y. Direct volume variation measurements in fused silica specimens exposed to femtosecond laser.
Opt Mater Express 2012;2:78998.
[374] Champion A, Beresna M, Kazansky PG, Bellouard Y. Stress distribution around femtosecond laser affected zones: effect of
nanogratings orientation. Opt Express 2013;21:2494251.
[375] Richter S, Plech A, Steinert M, Heinrich M, Dring S, Zimmermann F, et al. On the fundamental structure of femtosecond
laser-induced nanogratings. Laser Photonics Rev 2012;6:78792.
[376] Zimmermann F, Plech A, Richter S, Dring S, Tnnermann A, Nolte S. Structural evolution of nanopores and cracks as
fundamental constituents of ultrashort pulse-induced nanogratings. Appl Phys A 2014;114:759.
[377] Ramirez LPR, Heinrich M, Richter S, Dreisow F, Keil R, Korovin AV. Tuning the structural properties of femtosecond-laserinduced nanogratings. Appl Phys A 2010;100:16.
[378] Richter S, Heinrich M, Dring S, Tnnermann A, Nolte S, Peschel U. Nanogratings in fused silica: formation, control, and
applications. J Laser Appl 2012;24:042008.
[379] Song J, Wang XH, Hu X, Xu J, Liao Y, Sun H, et al. Orientation-controllable self-organized microgratings induced in the bulk
SrTiO3 crystal by a single femtosecond laser beam. Opt Express 2007;15:145249.
[380] Dai Y, Wu GR, Lin X, Ma GH, Qiu JR. Femtosecond laser induced rotated 3D self-organized nanograting in fused silica. Opt
Express 2012;20:180728.
[381] Zhang FT, Yu YZ, Cheng C, Dai Y, Qiu JR. Fabrication of polarization-dependent light attenuator in fused silica using a lowrepetition-rate femtosecond laser. Opt Lett 2013;38:22124.
[382] Canning J, Lancry M, Cook K, Weickman A, Brisset F, Poumellec B. Anatomy of a femtosecond laser processed silica
waveguide. Opt Mater Express 2011;1:9981008 [Invited].
[383] Poumellec B, Lancry M, Desmarchelier R, Herv E, Brisset F, Poulin JC. Asymmetric orientational writing in glass with
femtosecond laser irradiation. Opt Mater Express 2013;3:158699.
[384] Corbari C, Champion A, Gecevicius M, Beresna M, Bellouard Y, Kazansky PG. Femtosecond versus picosecond laser
machining of nano-gratings and micro-channels in silica glass. Opt Express 2013;21:394658.
[385] Bulgakova NM, Zhukov VP, Meshcheryakov YP, Gemini L, Brajer J, Rostohar D, et al. Pulsed laser modification of
transparent dielectrics: what can be foreseen and predicted by numerical simulations? J Opt Soc Am B 2014;31:C8C14.
[386] Buschlinger R, Nolte S, Peschel U. Self-organized pattern formation in laser-induced multiphoton ionization. Phys Rev B
2014;89:184306.
[387] Sun Q, Jiang HB, Liu Y, Wu ZX, Yang H, Gong QH. Measurement of the collision time of dense electronic plasma induced by
a femtosecond laser in fused silica. Opt Lett 2005;30:3202.
[388] Papazoglou DG, Zergioti I, Tzortzakis S. Plasma strings from ultraviolet laser filaments drive permanent structural
modifications in fused silica. Opt Lett 2007;32:20557.
[389] Liao Y, Ni J, Qiao L, Huang M, Bellouard Y, Sugioka K, et al. High-fidelity visualization of formation of volume nanogratings
in porous glass by femtosecond laser irradiation. Optica 2015;2:32934.
[390] Canning J, Lancry M, Cook K, Poumellec B. New theory of femtosecond induced changes and nanopore formation. Proc
SPIE 2013;8351:83512M.
[391] Bricchi E, Mills JD, Kazansky PG, Klappauf BG, Baumberg JJ. Birefringent Fresnel zone plates in silica fabricated by
femtosecond laser machining. Opt Lett 2002;27:22002.
[392] Beresna M, Kazansky PG. Polarization diffraction grating produced by femtosecond laser nanostructuring in glass. Opt
Lett 2010;35:16624.
[393] Gecevicius M, Beresna M, Kazansky PG. Polarization sensitive camera by femtosecond laser nanostructuring. Opt Lett
2013;38:40969.
[394] Mihailov SJ, Smelser CW, Lu P, Walker RB, Grobnic D, Ding H. Fiber Bragg gratings made with a phase mask and 800-nm
femtosecond radiation. Opt Lett 2003;28:9957.
[395] Mihailov SJ, Grobnic D, Smelser CW, Lu P, Walker RB, Ding HM. Bragg grating inscription in various optical fibers with
femtosecond infrared lasers and a phase mask. Opt Mater Express 2011;1:75465.
[396] Zhang HB, Eaton SM, Herman PR. Single-step writing of Bragg grating waveguides in fused silica with an externally
modulated femtosecond fiber laser. Opt Lett 2007;32:255961.
[397] Fernandes LA, Grenier JR, Herman PR, Aitchison JS, Marques PVS. Femtosecond laser writing of waveguide retarders in
fused silica for polarization control in optical circuits. Opt Express 2011;19:18294301.
[398] Fernandes LA, Grenier JR, Herman PR, Aitchison JS, Marques PVS. Femtosecond laser fabrication of birefringent directional
couplers as polarization beam splitters in fused silica. Opt Express 2011;19:119929.
[399] Li JZ, Ho S, Haque M, Herman PR. Nanograting Bragg responses of femtosecond laser written optical waveguides in fused
silica glass. Opt Mater Express 2012;2:156370.

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

227

[400] Mikutis M, Kudrius T, lekys G, Paipulas D, Juodkazis S. High 90% efficiency Bragg gratings formed in fused silica by
femtosecond Gauss-Bessel laser beams. Opt Mater Express 2013;3:186271.
[401] Liao Y, Shen YL, Qiao LL, Chen DP, Cheng Y, Sugioka K, et al. Femtosecond laser nanostructuring in porous glass with sub50 nm feature sizes. Opt Lett 2013;38:1879.
[402] Liao Y, Cheng Y, Liu CN, Song JX, He F, Shen YL, et al. Direct laser writing of sub-50 nm nanofluidic channels buried in glass
for three-dimensional micro-nanofluidic integration. Lab Chip 2013;13:162631.
[403] Xu J, Wu D, Hanada Y, Chen C, Wu SZ, Cheng Y, et al. Electrofluidics fabricated by space-selective metallization in glass
microfluidic structures using femtosecond laser direct writing. Lab Chip 2013;13:460816.
[404] Toratani E, Kamata M, Obara M. Self-fabrication of void array in fused silica by femtosecond laser processing. Appl Phys
Lett 2005;87:171103.
[405] Sun HY, Song J, Li CB, Xu J, Wang XS, Cheng Y, et al. Standing electron plasma wave mechanism of void array formation
inside glass by femtosecond laser irradiation. Appl Phys A 2007;88:2858.
[406] Hu X, Dai Y, Yang LY, Song J, Zhu CS, Qiu JR. Self-formation of quasiperiodic void structure in CaF2 induced by femtosecond
laser irradiation. J Appl Phys 2007;101:023112.
[407] Dai Y, Hu X, Song J, Yu BK, Qiu JR. Self-assembled quasi-periodic voids in glass induced by a tightly focused femtosecond
laser. Chin Phys Lett 2007;24:19414.
[408] Song J, Sun HY, Wang XS, Xu J, Qiu JR, Xu Z. Self-organized void strings induced in SrTiO3 crystal by a femtosecond laser.
Chin Phys Lett 2007;5:S21821.
[409] Hu X, Song J, Zhou QL, Yang LY, Wang XD, Zhu CS, et al. Self-formation of void array in Al2O3 crystal by femtosecond laser
irradiation. Chin Opt Lett 2008;6:38890.
[410] Song J, Wang XS, Hu X, Dai Y, Qiu JR, Cheng Y. Formation mechanism of self-organized voids in dielectrics induced by
tightly focused femtosecond laser pulses. Appl Phys Lett 2008;92:092904.
[411] Song J, Wang XS, Xu J, Sun HY, Xu ZZ, Qiu JR. Microstructures induced in the bulk of SrTiO3 crystal by a femtosecond laser.
Opt Express 2007;15:23417.
[412] Song J, Luo FF, Hu X, Zhao QZ, Qiu JR, Xu ZZ. Mechanism of femtosecond laser inducing inverted microstructures by
employing different types of objective lens. J Phys D Appl Phys 2011;44:495402.
[413] Jang W, Kim D, Kim T, Moon A, Lim KS, Lee M, et al. Formation of void array inside transparent and absorptive glasses by
femtosecond laser irradiation. J Nanosci Nanotechnol 2012;12:4798802.
[414] Ahmed F, Lee MS, Sekita H, Sumiyoshi T, Kamata M. Display glass cutting by femtosecond laser induced single shot
periodic void array. Appl Phys A 2008;93:18992.
[415] Wang XH, Chen F, Yang Q, Liu HW, Bian H, Si JH, et al. Fabrication of quasi-periodic micro-voids in fused silica by single
femtosecond laser pulse. Appl Phys A 2011;102:3944.
[416] Thomas J, Bernard R, Alti K, Dharmadhikari AK, Dharmadhikari JA, Bhatnagar A, et al. Pattern formation in transparent
media using ultrashort laser pulses. Opt Commun 2013;304:2938.
[417] Hu X, Qian B, Zhang P, Wang X, Su L, Qiu JR, et al. Self-organized microvoid array perpendicular to the femtosecond laser
beam in CaF2 crystals. Laser Phys Lett 2008;5:3947.
[418] Mauclair C, Mermillod-Blondin A, Rosenfeld A, Hertel IV, Audouard E, Miyamoto I, et al. Multipoint focusing of single
ultrafast laser Pulses. J Laser Micro Nanoeng 2011;6:23944.
[419] Mauclair C, Mermillod-Blondin A, Landon S, Huot N, Rosenfeld A, Hertel IV, et al. Single-pulse ultrafast laser imprinting of
axial dot arrays in bulk glasses. Opt Lett 2011;36:3257.
[420] Terakawa M, Toratani E, Shirakawa T, Obara M. Fabrication of a void array in dielectric materials by femtosecond laser
micro-processing for compact photonic devices. Appl Phys A 2010;100:10417.
[421] Toratani E, Kamata M, Obara M. Self-organization of nano-void array for photonic crystal device. Microelectron Eng
2006;83:17825.
[422] Mndez C, Vzquez de Aldana JR, Torchia GA, Roso L. Optical waveguide arrays induced in fused silica by void-like defects
using femtosecond laser pulses. Appl Phys B 2007;86:3436.
[423] Beresna M, Gecevicius M, Bulgakova NM, Kazansky PG. Twisting light with micro-spheres produced by ultrashort light
pulses. Opt Express 2011;19:1898996.
[424] Liu Y, Zhu B, Wang L, Dai Y, Ma H, Qiu JR. Femtosecond laser induced coordination transformation and migration of ions in
sodium borate glasses. Appl Phys Lett 2008;92:121113.
[425] Liu Y, Shimizu M, Zhu B, Dai Y, Qian B, Qiu JR, et al. Micromodification of element distribution in glass using femtosecond
laser irradiation. Opt Lett 2009;34:1368.
[426] Teng Y, Zhou JJ, Lin G, Hua JJ, Zeng HP, Zhou SF, et al. Ultrafast modification of elements distribution and local
luminescence properties in glass. J Non-Cryst Solids 2012;358:11859.
[427] Wang X, Sakakura M, Liu Y, Qiu JR, Shimotsuma Y, Hirao K, et al. Modification of long range order in germanate glass by
ultra fast laser. Chem Phys Lett 2011;511:2669.
[428] Tu ZF, Teng Y, Zhou JJ, Zhou SF, Zeng HP, Qiu JR. Raman spectroscopic investigation on femtosecond laser induced residual
stress and element distribution in bismuth germinate glasses. J Raman Spectrosc 2013;44:30711.
[429] Luo FF, Qian B, Lin G, Xu J, Liao Y, Song J, et al. Redistribution of elements in glass induced by a high-repetition-rate
femtosecond laser. Opt Express 2010;18:62629.
[430] Luo FF, Song J, Hu X, Sun HY, Lin G, Pan HH, et al. Femtosecond laser-induced inverted microstructures inside glasses by
tuning refractive index of objectives immersion liquid. Opt Lett 2011;36:21257.
[431] Sakakura M, Kurita T, Shimizu M, Yoshimura K, Shimotsuma Y, Fukuda N, et al. Shape control of elemental distributions
inside a glass by simultaneous femtosecond laser irradiation at multiple spots. Opt Lett 2013;38:493942.
[432] Miura K, Shimizu M, Sakakura M, Kurita T, Shimotsuma Y, Hirao K. Formation mechanism and applications of laser
induced elemental distribution in glasses. PIERS Proc 2012:1923.
[433] Shimizu M, Sakakura M, Kanehira S, Nishi M, Shimotsuma Y, Hirao K, et al. Formation mechanism of element distribution
in glass under femtosecond laser irradiation. Opt Lett 2011;36:21613.
[434] Shimizu M, Sakakura M, Nishi M, Shimotsuma Y, Hirao K, Miura K. Control of element distribution in glass with
femtosecond laser. Proc SPIE 2012;8244:82440O.

228

D. Tan et al. / Progress in Materials Science 76 (2016) 154228

[435] Shimizu M, Miura K, Yasuda N, Sakakura M, Kanehira S, Nishi M, et al. Formation of elemental distribution in glass using
thermal accumulation with femtosecond laser. Mater Res Soc Symp Proc 2009;1230 (1230-MM07-04).
[436] Vitek DN, Block E, Bellouard Y, Adams DE, Backus S, Kleinfeld D, et al. Spatio-temporally focused femtosecond laser pulses
for nonreciprocal writing in optically transparent materials. Opt Express 2010;18:246738.
[437] Poumellec B, Lancry M, Poulin JC, Ani-Joseph S. Non reciprocal writing and chirality in femtosecond laser irradiated silica.
Opt Express 2008;16:1835461.
[438] Kazansky PG, Shimotsuma Y, Sakakura M, Beresna M, Gecevicius M, Svirko Y, et al. Photosensitivity control of an isotropic
medium through polarization of light pulses with tilted intensity front. Opt Express 2011;19:2065764.
[439] Gecevicius M, Beresna M, Zhang J, Yang WJ, Takebe H, Kazansky PG. Extraordinary anisotropy of ultrafast laser writing in
glass. Opt Express 2013;21:395968.
[440] Matsuo S, Umeda Y, Tomita T, Hashimoto S. Laser-scanning direction effect in femtosecond laser-assisted etching. J Laser
Micro Nanoeng 2013;8:358.
[441] Lancry M, Poumellec B, Desmarchelier R, Bourguignon B. Oriented creation of anisotropic defects by IR femtosecond laser
scanning in silica. Opt Mater Express 2012;2:180921.
[442] Salter PS, Booth MJ. Dynamic control of directional asymmetry observed in ultrafast laser direct writing. Appl Phys Lett
2012;101:141109.
[443] Akturk S, Kimmel M, OShea P, Trebino R. Measuring pulse-front tilt in ultrashort pulses using GRENOUILLE. Opt Express
2003;11:491501.
[444] Vailionis A, Gamaly EG, Mizeikis V, Yang W, Rode AV, Juodkazis S. Evidence of superdense aluminium synthesized by
ultrafast microexplosion. Nat Commun 2011;2:445.
[445] Tan DZ, Lin G, Liu Y, Teng Y, Zhuang YX, Zhu B, et al. Synthesis of nanocrystalline cubic zirconia using femtosecond laser
ablation. J Nanopart Res 2011;13:118390.
[446] Tan DZ, Zhou SF, Xu BB, Chen P, Shimotsuma Y, Miura K, et al. Simple synthesis of ultra-small nanodiamonds with tunable
size and photoluminescence. Carbon 2013;62:37481.
[447] Tan DZ, Zhou SF, Qiu JR, Khusro N. Preparation of functional nanomaterials with femtosecond laser ablation in solution. J
Photochem Photobiol C 2013;17:5068.
[448] Mizeikis V, Vailionis A, Gamaly EG, Yang W, Rode A, Juodkazis S. Synthesis of super-dense phase of aluminum under
extreme pressure and temperature conditions created by femtosecond laser pulses in sapphire. Proc SPIE
2013;8249:82490A.
[449] Moriarty JA, McMahan AK. High-pressure structural phase transitions in Na, Mg, and Al. Phys Rev Lett 1982;48:80912.
[450] Gamaly EG, Vailionis A, Mizeikis V, Yang W, Rode AV, Juodkazis S. Warm dense matter at the bench-top: fs-laser-induced
confined micro-explosion. High Energy Density Phys 2012;8:137.
[451] Gamaly EG, Rapp L, Roppo V, Juodkazis S, Rode AV. Generation of high energy density by fs-laser-induced confined
microexplosion. New J Phys 2013;15:025018.
[452] Bressel L, de Ligny D, Gamaly EG, Rode AV, Juodkazis S. Observation of O2 inside voids formed in GeO2 glass by tightlyfocused fs-laser pulses. Opt Mater Express 2011;1:11507.
[453] Tan DZ, Liu XF, Dai Y, Ma GH, Meunier M, Qiu JR. A universal photochemical approach to ultra-small, well-dispersed
nanoparticle/reduced graphene oxide hybrids with enhanced nonlinear optical properties. Adv Opt Mater
2015;3:83641.
[454] Papon G, Petit Y, Marquestaut N, Royon A, Dussauze M, Rodriguez V, et al. Fluorescence and second-harmonic generation
correlative microscopy to probe space charge separation and silver cluster stabilization during direct laser writing in a
tailored silver-containing glass. Opt Mater Express 2013;3:185561.
[455] Schultze M, Bothschafter EM, Sommer A, Holzner S, Schweinberger W, Fiess M, et al. Controlling dielectrics with the
electric field of light. Nature 2013;493:758.
[456] Kawamura K, Sarukura N, Hirano M, Ito N, Hosono H. Periodic nanostructure array in crossed holographic gratings on
silica glass by two interfered infrared-femtosecond laser pulses. Appl Phys Lett 2001;79:122830.
[457] Si JH, Qiu JR, Zhai JF, Shen YQ, Hirao K. Photoinduced permanent gratings inside bulk azodye-doped polymers by the
coherent field of a femtosecond laser. Appl Phys Lett 2002;80:35961.
[458] Si JH, Meng ZC, Kanehira S, Qiu JR, Hua B, Hirao K. Multiphoton-induced periodic microstructures inside bulk azodyedoped polymers by multibeam laser interference. Chem Phys Lett 2004;399:2769.
[459] Qu SL, Qiu JR, Zhao CJ, Jiang XW, Zeng HD, Zhu CS, et al. Metal nanoparticle precipitation in periodic arrays in Au2O-doped
glass by two interfered femtosecond laser pulses. Appl Phys Lett 2004;84:20468.
[460] Si JH, Kitaoka K, Qiu JR, Mitsuyu T. Optically encoded second-harmonic generation in germanosilicate glass by a
femtosecond laser. Opt Lett 1999;24:9113.
[461] Qiu JR, Si JH, Hirao K. Photoinduced stable second-harmonic generation in chalcogenide glasses. Opt Lett 2001;26:9146.
[462] Sugioka K, Cheng Y. A tutorial on optics for ultrafast laser materials processing: basic microprocessing system to beam
shaping and advanced focusing methods. Adv Opt Technol 2012;1:35364.
[463] Zhang SA, Lu CH, Jia TQ, Qiu JR, Sun ZR. Coherent phase control of resonance-mediated two-photon absorption in rareearth ions. Appl Phys Lett 2013;103:194104.
[464] Yamaji M, Kawashima H, Suzuki J, Tanaka S, Shimizu M, Hirao K, et al. Homogeneous and elongation-free 3D
microfabrication by a femtosecond laser pulse and hologram. J Appl Phys 2012;111:083107.
[465] Anderson PW. Through a glass lightly. Science 1995;267:1615.
[466] Angell CA. Formation of glasses from liquids and biopolymers. Science 1995;267:192435.
[467] Greaves G, Wilding MC, Fearn S, Langstaff D, Kargl F, Cox S, et al. Detection of first-order liquid/liquid phase transitions in
yttrium oxide-aluminum oxide melts. Science 2008;322:56670.
[468] Zhang YF, Yang G, Yue YZ. Calorimetric signature of structural heterogeneity in a ternary silicate glass. J Am Ceram Soc
2013;96:30357.

Das könnte Ihnen auch gefallen