Laser-Based Nano Fabrication and Nano Lithography

You are on page 1of 156

Laser-Based Nano

Fabrication and
Nano Lithography
Edited by
Koji Sugioka and Ya Cheng
Printed Edition of the Special Issue Published in Nanomaterials

www.mdpi.com/journal/nanomaterials
Laser-Based Nano Fabrication and
Nano Lithography
Laser-Based Nano Fabrication and
Nano Lithography

Special Issue Editors


Koji Sugioka
Ya Cheng

MDPI • Basel • Beijing • Wuhan • Barcelona • Belgrade


Special Issue Editors
Koji Sugioka Ya Cheng
RIKEN Center for Advanced Photonics East China Normal University
Japan China

Editorial Office
MDPI
St. Alban-Anlage 66
4052 Basel, Switzerland

This is a reprint of articles from the Special Issue published online in the open access
journal Nanomaterials (ISSN 2079-4991) in 2018 (available at: https://www.mdpi.com/journal/
nanomaterials/special issues/Laser Nano Fabrication Lithography)

For citation purposes, cite each article independently as indicated on the article page online and as
indicated below:

LastName, A.A.; LastName, B.B.; LastName, C.C. Article Title. Journal Name Year, Article Number,
Page Range.

ISBN 978-3-03897-410-9 (Pbk)


ISBN 978-3-03897-411-6 (PDF)

Cover image courtesy of Yongfeng Lu.


c 2018 by the authors. Articles in this book are Open Access and distributed under the Creative
Commons Attribution (CC BY) license, which allows users to download, copy and build upon
published articles, as long as the author and publisher are properly credited, which ensures maximum
dissemination and a wider impact of our publications.
The book as a whole is distributed by MDPI under the terms and conditions of the Creative Commons
license CC BY-NC-ND.
Contents

About the Special Issue Editors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii

Preface to ”Laser-Based Nano Fabrication and Nano Lithography” . . . . . . . . . . . . . . . . ix

Seiya Nikaido, Takumi Natori, Ryo Saito and Godai Miyaji


Nanostructure Formation on Diamond-Like Carbon Films Induced with Few-Cycle Laser
Pulses at Low Fluence from a Ti:Sapphire Laser Oscillator
Reprinted from: Nanomaterials 2018, 8, 535, doi:10.3390/nano8070535 . . . . . . . . . . . . . . . . 1

Xin Zheng, Cong Cong, Yuhao Lei, Jianjun Yang and Chunlei Guo
Formation of Slantwise Surface Ripples by Femtosecond Laser Irradiation
Reprinted from: Nanomaterials 2018, 8, 458, doi:10.3390/nano8070458 . . . . . . . . . . . . . . . . 10

Ignacio Falcón Casas and Wolfgang Kautek


Subwavelength Nanostructuring of Gold Films by Apertureless Scanning Probe Lithography
Assisted by a Femtosecond Fiber Laser Oscillator
Reprinted from: Nanomaterials 2018, 8, 536, doi:10.3390/nano8070536 . . . . . . . . . . . . . . . . 20

Ying Liu, John H. Campbell, Ori Stein, Lijia Jiang, Jared Hund and Yongfeng Lu
Deformation Behavior of Foam Laser Targets Fabricated by Two-Photon Polymerization
Reprinted from: Nanomaterials 2018, 8, 498, doi:10.3390/nano8070498 . . . . . . . . . . . . . . . . 32

Florin Jipa, Stefana Iosub, Bogdan Calin, Emanuel Axente, Felix Sima and Koji Sugioka
High Repetition Rate UV versus VIS Picosecond Laser Fabrication of 3D Microfluidic Channels
Embedded in Photosensitive Glass
Reprinted from: Nanomaterials 2018, 8, 583, doi:10.3390/nano8080583 . . . . . . . . . . . . . . . . 52

Yasutaka Nakajima, Shuichiro Hayashi, Akito Katayama, Nikolay Nedyalkov and


Mitsuhiro Terakawa
Femtosecond Laser-Based Modification of PDMS to Electrically Conductive Silicon Carbide
Reprinted from: Nanomaterials 2018, 8, 558, doi:10.3390/nano8070558 . . . . . . . . . . . . . . . . 64

Eugene G. Gamaly, Saulius Juodkazis, Andrei V. Rode


Extreme Energy Density Confined Inside a Transparent Crystal: Status and Perspectives
of Solid-Plasma-Solid Transformations
Reprinted from: Nanomaterials 2018, 8, 555, doi:10.3390/nano8070555 . . . . . . . . . . . . . . . . 73

Xiao-Wen Cao, Qi-Dai Chen, Hua Fan, Lei Zhang, Saulius Juodkazis and Hong-Bo Sun
Liquid-Assisted Femtosecond Laser Precision-Machining of Silica
Reprinted from: Nanomaterials 2018, 8, 287, doi:10.3390/nano8050287 . . . . . . . . . . . . . . . . 82

Dongshi Zhang, Wonsuk Choi, Jurij Jakobi, Mark-Robert Kalus, Stephan Barcikowski,
Sung-Hak Cho and Koji Sugioka
Spontaneous Shape Alteration and Size Separation of Surfactant-Free Silver Particles
Synthesized by Laser Ablation in Acetone during Long-Period Storage
Reprinted from: Nanomaterials 2018, 8, 529, doi:10.3390/nano8070529 . . . . . . . . . . . . . . . . 90

Dongshi Zhang, Wonsuk Choi, Yugo Oshima, Ulf Wiedwald, Sung-Hak Cho, Hsiu-Pen Lin,
Yaw Kuen Li, Yoshihiro Ito and Koji Sugioka
Magnetic Fe@FeOx , Fe@C and α-Fe2 O3 Single-Crystal Nanoblends Synthesized by
Femtosecond Laser Ablation of Fe in Acetone
Reprinted from: Nanomaterials 2018, 8, 631, doi:10.3390/nano8080631 . . . . . . . . . . . . . . . . 107

v
Yoshiki Nakata, Keiichi Murakawa, Noriaki Miyanaga, Aiko Narazaki, Tatsuya Shoji and
Yasuyuki Tsuboi
Local Melting of Gold Thin Films by Femtosecond Laser-Interference Processing to Generate
Nanoparticles on a Source Target
Reprinted from: Nanomaterials 2018, 8, 477, doi:10.3390/nano8070477 . . . . . . . . . . . . . . . . 122

Hsin-hui Huang, Takeshi Nagashima, Wei-hung Hsu, Saulius Juodkazis and Koji Hatanaka
Dual THz Wave and X-ray Generation from a Water Film under Femtosecond Laser Excitation
Reprinted from: Nanomaterials 2018, 8, 523, doi:10.3390/nano8070523 . . . . . . . . . . . . . . . . 131

vi
About the Special Issue Editors
Koji Sugioka, Dr. of Engin, received his B. S., Ms. Eng., and Dr. Eng. degrees in electronics
from Waseda University (Japan) in 1984, 1986, and 1993, respectively. He Joined RIKEN in 1986
and is currently a Team Leader of Advanced Laser Processing Research Team at RIKEN Center
for Advanced Photonics. He is concurrently a guest professor at Osaka University Tokyo Denki
University. He was awarded the degree of Doctor Honoris Causa from University of Szeged,
Hungary in 2018. He has made important contribution to both fundamental research on laser–matter
interactions and diverse applications including practical applications in the said area. He is
internationally renowned for his works on laser micro and nano processing, particularly ultrafast
laser processing technology. He is currently a member of the board of directors of the Laser Institute
of America (LIA) and Japanese Laser Processing Society (JLPS), a council member of the Intl.
Academy of Photonics and Laser Engineering (IAPLE), and a Fellow of SPIE, OSA, LIA, and IAPLE.
He is also an editor-in-chief of Journal of the Laser Micro/Nanoengineering (JLMN) and an editor of
Opto-Electronic Advances (OEA) and Advanced Optical Technologies (AOT).

Ya Cheng, PhD, received his B. S. from Fudan University in 1993 and his PhD in optics from
Shanghai Institute of Optics and Fine Mechanics, Chinese Academy of Sciences in 1998. He is
currently a professor at the Shanghai Institute of Optics and Fine Mechanics and Dean of the School of
Physics and Materials Science, East China Normal University. His research interests include ultrafast
nonlinear optics, femtosecond laser micromachining, and lithium niobate photonics. He is currently
a council member of the Chinese Optical Society and a Fellow of IOP (UK). He is also an editor of
Journal of the Laser Micro/Nanoengineering (JLMN) and an editor of Micromachines.

vii
Preface to ”Laser-Based Nano Fabrication and
Nano Lithography”
The improvement of fabrication resolutions is an eternal challenge for miniaturizing and
enhancing the integration degrees of devices. Laser processing is one of the most widely used
techniques in manufacturing due to its high flexibility, high speed, and environmental friendliness.
The fabrication resolution of laser processing is, however, limited by its diffraction limit. Recently,
much effort has been made to overcome the diffraction limit in nano fabrication. Specifically,
combinations of multiphoton absorption by ultrafast lasers and the threshold effect associated with a
Gaussian beam profile provide fabrication resolutions far beyond the diffraction limit. The use of the
optical near-field achieves nano ablation with feature sizes below 100 nm. Multiple pulse irradiation
from the linearly polarized ultrafast laser produces periodic nanostructures with a spatial period
much smaller than the wavelength. Unlimited diffraction resolutions can also be achieved with
shaped laser beams. In the meanwhile, lasers are also widely used for the synthesis of nano materials
including fullerenes and nano particles.
In view of the rapid advancement of this field in recent years, this Special Issue aims at
introducing the state-of-the-art in nano fabrication and nano lithography, based on laser technologies,
by leading groups in the field. Specifically, this Special Issue consists of two invited feature articles
and ten contributed articles covering relevant topics of the laser-based nano fabrication and nano
lithography, written by internationally recognized experts in the field. It includes the formation of
periodic surface nano structures via the irradiation of intense ultrashort pulses, surface nano
structuring below the diffraction limit using the optical near-field, three-dimensional micro-and
nano-scale additive manufacturing based on two-photon polymerization, three-dimensional micro
and nano fabrication inside transparent materials based on multiphoton absorption using ultrafast
lasers, higher energy density confinement by Bessel ultrashort pulses for micro and nano fabrication,
precision machining by liquid-assisted femtosecond laser ablation, the synthesis of functional nano
particles by laser ablation in liquids, the formation of shape- and size-controlled metallic
nanoparticles by pulsed-laser deposition and laser-induced dot transfer, and dual THz wave and X-
ray generation by femtosecond laser excitation.
I believe that this Special Issue offers a realistic and comprehensive review of the state-of-the-art
in nano fabrication and nano lithography and is beneficial for many researchers including students
and young scientists working not only in the field but also in many other fields.
Last but not least, I would like to thank all of the article authors for their great effort and
wonderful work to complete this informative Special Issue.

Koji Sugioka, Ya Cheng


Special Issue Editors

ix
nanomaterials

Article
Nanostructure Formation on Diamond-Like Carbon
Films Induced with Few-Cycle Laser Pulses at Low
Fluence from a Ti:Sapphire Laser Oscillator
Seiya Nikaido, Takumi Natori, Ryo Saito and Godai Miyaji *
Department of Applied Physics, Tokyo University of Agriculture and Technology, 2-24-16 Nakacho,
Koganei, Tokyo 184-8588, Japan; [email protected] (S.N.);
[email protected] (T.N.); [email protected] (R.S.)
* Correspondence: [email protected]; Tel.: +81-42-388-7153

Received: 2 June 2018; Accepted: 14 July 2018; Published: 16 July 2018

Abstract: This study reports the results of experiments on periodic nanostructure formation on
diamond-like carbon (DLC) films induced with 800 nm, 7-femtosecond (fs) laser pulses at low fluence
from a Ti:sapphire laser oscillator. It was demonstrated that 7-fs laser pulses with a high power
density of 0.8–2 TW/cm2 at a low fluence of 5–12 mJ/cm2 can form a periodic nanostructure with a
period of 60–80 nm on DLC films. The period decreases with increasing fluence of the laser pulses.
The experimental results and calculations for a model target show that 7-fs pulses can produce a
thinner metal-like layer on the DLC film through a nonlinear optical absorption process compared
with that produced with 100-fs pulses, creating a finer nanostructure via plasmonic near-field ablation.

Keywords: femtosecond laser; laser ablation; nanostructure formation; surface plasmon polaritons;
near-field; diamond-like carbon

1. Introduction
Superimposed femtosecond (fs) laser pulses can form a periodic nanostructure (PNS) on solid
surfaces through ablation, where the period size d is typically 10–20% of the laser wavelength
λ [1–6]. There has been considerable interest in this surface phenomenon for application in laser
nanoprocessing, beyond the diffraction limit of light. Numerous studies have been conducted to
understand the mechanism responsible for PNS formation [7–10]. The experimental conditions and
laser parameters for PNS formation have been identified for various target materials, and the dominant
physical mechanisms responsible for nanostructuring have been determined.
Based on a series of experiments and model calculations, Miyazaki and Miyaji found that PNS
formation is induced by fs laser pulses at a moderate fluence F through: a bonding structure change
in the material [11–13]; generation of high-density electrons on the target surface, leading to the
formation of a metal-like layer through linear and nonlinear optical absorption [13–15]; near-field
ablation around the corrugated nanosurface [13–15]; and excitation of standing surface plasmon
polariton (SPP) waves [9,10,15–17]. These laser–matter interaction processes can explain the origin
and growth of PNSs on diamond-like carbon (DLC) [15], Si [16], GaN [10,17], Ti, and stainless steel [9],
and theoretical calculations agree well with the observed nanoperiod, which is much smaller than
λ/2. Based on the physical mechanism, control methods for the PNS shape have been developed,
allowing the formation of homogeneous nanogratings [9,10,17] and a saw-like PNS [18] in air. However,
some important processes for PNS formation are still unknown, and there is no consensus regarding
the detailed mechanism.
For various kinds of material, it has been reported that the d value for a PNS increases with
increasing F for the fs laser pulses at a fluence F of a few 100 mJ/cm2 to a few J/cm2 with a power

Nanomaterials 2018, 8, 535; doi:10.3390/nano8070535 1 www.mdpi.com/journal/nanomaterials


Nanomaterials 2018, 8, 535

density I of a few TW/cm2 [7,8]. Previous studies have concluded that this increase is attributed to the
increasing thickness of the metal-like layer produced on the target material with increasing F [16,17].
However, this has never been experimentally confirmed.
The proposed mechanism of PNS formation suggests that a thin metal-like layer can be produced
by the fs pulses at low F with I ~TW/cm2 via a nonlinear absorption process, allowing confirmation
of the thickness effect for nanostructuring. In this paper, we report the experimental results of PNS
formation on DLC films irradiated with 800 nm, 7-fs laser pulses with a high power density I of
0.8–2 TW/cm2 and a low fluence F of 5–12 mJ/cm2 delivered from a laser oscillator. The results
indicate the formation of a PNS with a period of d = 60–80 nm that decreases with increasing F. Based on
the experimental results and a model calculation, it is shown that the excitation of SPPs at the interface
between the thin metal-like layer and the DLC is certainly responsible for the nanostructuring process,
and that the decrease of d is attributed to the decreasing wavelength of the SPPs with increasing
F through an increase of electron density in the thin metal-like layer.

2. Experimental
Figure 1 shows a schematic diagram of the optical configuration used in the ablation experiments.
As fs laser pulses with a high power density I at low fluence F can produce a thin metal-like layer on
a target surface, the output of a Ti:sapphire laser oscillator was used in the experiments. The pulse
duration Δτ was ~7 fs, the wavelength λ was 680–940 nm, the repetition rate f rep was 80 MHz,
and the pulse energy Upulse was ~5 nJ. The pulses were so-called few-cycle laser pulses, where the
electromagnetic field oscillates for a few cycles [19]. The temporal and spectral profiles of the fs
pulses were monitored with a spectral phase interferometry for direct electric-field reconstruction
(SPIDER) device and a spectrometer, respectively. When measuring the temporal profile, a silver
mirror was inserted to propagate the pulses to the SPIDER device. The output just after the oscillator
had a negative group delay dispersion, which was compensated for to minimize the pulse duration
by passing the beam through a beam splitter (thickness: 1 mm) and a glass plate (thickness: 1 mm).
The laser pulses were spatially expanded with a pair of convex and concave silver mirrors and focused
onto the target surface with a ×40 Schwarzschild-type reflective objective (numerical aperture: 0.50)
to a spot size w0 of ~2 μm (1/e2 radius) on the surface, since the group delay dispersion had to be
suppressed to obtain laser pulses with a high power density. A CMOS camera was used to image the
focused beam on the target surface. The pulse energy Upulse just after the objective was measured with
a pyroelectric detector, and the peak fluence F = 2 Upulse /(π w0 2 ) and the peak power density I = F/Δτ
of the fs laser pulses on the target surface were estimated.

Camera

L RO Target
M
CM
xy stage
Ti:sapphire BS
laser oscillator CM
GP
SPIDER Spectrometer

Figure 1. Schematic diagram of optical configuration for nanostructure formation. CM, convex or
concave mirror; M, mirror; GP, glass plate; BS, beam splitter; L, lens; RO, reflective objective.

As the target, we used a DLC film (thickness: 1.7 μm) that was deposited on a polished
silicon substrate with a plasma-based ion implantation system. The root-mean-square value of
surface roughness was measured to be less than 1 nm with a scanning probe microscope (SPM).
The target was set on an xy motorized stage, which could move at a constant speed v of 0.1–100 μm/s.
The surface morphology was observed using a scanning electron microscope (SEM) and the SPM.

2
Nanomaterials 2018, 8, 535

A two-dimensional Fourier transform was applied to the SPM images to analyze the distribution of
the spatial periodicity in the surface structure along the polarization direction. The bonding structure
of the target surface irradiated with the fs pulses was analyzed using micro-Raman spectroscopy with
a diode-pumped, single-longitudinal-mode, 532 nm laser beam focused with a ×40 objective.

3. Results and Discussion


Figure 2a–c show SEM and SPM images and spatial frequency spectra of DLC films irradiated
with 7 fs pulses with I = 1 TW/cm2 at F = 6 mJ/cm2 for v = 0.1–10 μm/s. For v = 100 μm/s, the surface
was observed to swell and was not ablated because of the small shot number of the laser pulses onto
the target surface. When v was decreased to 10 μm/s (i.e., the shot number increased), the formation
of a PNS with a period d of ~50 nm was observed on the ablated DLC surface, as shown in Figure 2a.
The line-like structure was perpendicular to the direction of polarization. When v was decreased
to 1 μm/s, a PNS with d of ~70 nm formed, as shown in Figure 2b. With a further decrease of v to
0.1 μm/s, deeper ablation traces with d of ~80 nm formed, as shown in Figure 2c. For comparison,
the target surfaces were also irradiated by 100-fs laser pulses with I = 0.1 TW/cm2 at the same F.
These pulses were produced by a glass plate (thickness: 3 mm) positioned just after the laser oscillator.
As shown in Figure 2d, a PNS did not form on the ablated surface under these conditions.

Figure 2. Scanning electron microscopy (SEM) images (top), scanning probe microscopy (SPM)
images (middle), and spatial frequency spectra (bottom) of a diamond-like carbon (DLC) film surface
irradiated with 7-fs pulses, with I = 1 TW/cm2 at F = 6 mJ/cm2 for (a) v = 10 μm/s, (b) v = 1 μm/s,
and (c) v = 0.1 μm/s, and (d) those irradiated with 100-fs pulses with I = 0.1 TW/cm2 at F = 6 mJ/cm2
for v = 0.1 μm/s. E and v denote directions of polarization and laser scanning, respectively.

In previously reported experiments, PNSs formed on DLC films with 100-fs laser pulses with
I = 1–2 TW/cm2 at F = 100–200 mJ/cm2 , delivered from a chirp-pulse amplification Ti:sapphire laser
system [3,11–15]. The results shown in Figure 2 suggest possible laser–matter interaction processes
for PNS formation, as discussed in previous studies [13–16]. As v is decreased, a bonding structure
change—from DLC to glassy carbon (GC)—is induced in the surface layer. This produces nanometer
surface roughness due to swelling of the material, as a thin layer with a high electron density is
produced on the surface through a nonlinear optical absorption process. On the highly curved swollen
metal-like surface, an intense near-field is generated that enhances the incident electric field and
initiates nanoscale ablation. Then, SPPs are transiently excited via coherent coupling of the incident
laser pulses with the corrugated surface, where the GC layer, including high-density electrons, works as

3
Nanomaterials 2018, 8, 535

a thin metal layer between air and the DLC for the excitation of SPPs [20]. The periodic enhancement
of the near-field of SPPs excited in the surface layer induces ablation, which forms a PNS on the surface.
The experimental results shown in Figure 2 indicate that such a process occurs sufficiently when a
DLC film is irradiated with 7-fs pulses with a high density of 1 TW/cm2 at a low fluence of 6 mJ/cm2 .
An increase in F is expected to increase the density of the free electrons produced in the surface
layer, leading to a change in surface morphology. To confirm this, surfaces were ablated with 7-fs pulses
for v = 0.1 μm/s for F = 5–12 mJ/cm2 , corresponding to I = 0.8–2 TW/cm2 . The results are shown in
Figure 3. At the lowest F, multiple shots produced a PNS with d ~85 nm; at the highest F, multiple shots
produced a finer PNS with d ~60 nm. Figure 4 plots the d value obtained from the isolated peak position
in the Fourier spectrum of the SPM images as a function of F and I. With increasing F, d decreases from
about 85 to 60 nm. For irradiation with 100-fs laser pulses with I = 1–4 TW/cm2 at F = 100–400 mJ/cm2 ,
it has been reported that the d value of the PNSs formed on various kinds of material (e.g., DLC, TiN,
stainless steel, Ti, Si, and GaN) increased with increasing F [3,9,16,17], which is opposite to the results
obtained in the present study. This suggests that low-fluence fs pulses with a high power density play
a crucial role in the surface morphological change that leads to nanostructuring.

Figure 3. SEM images (left), SPM images (center), and spatial frequency spectra (right) of DLC film surface
irradiated with 7-fs pulses at v = 0.1 μm/s for (a) I = 0.8 TW/cm2 , F = 5 mJ/cm2 and (b) I = 2 TW/cm2 ,
F = 12 mJ/cm2 . E and v denote directions of polarization and laser scanning, respectively.

Figure 4. Period d of a periodic nanostructure (PNS) on DLC film formed with 7-fs laser pulses as a
function of F and I for v = 0.1 μm/s.

4
Nanomaterials 2018, 8, 535

In a previous study, we reported that PNS formation on a DLC surface is preceded by a change
in the bonding structure, from DLC to GC [13]. The swelling of the target surface observed for
v = 100 μm/s indicates that the onset of ablation at v ≤ 10 μm/s is preceded by a change in the
bonding structure to GC in the target surface. To confirm this, Raman spectra were obtained from
surfaces ablated with 7-fs pulses with I = 1 TW/cm2 at F = 6 mJ/cm2 for v = 0.1 μm/s. The results are
shown in Figure 5, together with spectra of surfaces ablated with 100-fs pulses with I = 0.1 TW/cm2 at
F = 6 mJ/cm2 for v = 0.1 μm/s and non-irradiated DLC for comparison. Each spectrum is normalized
to give a maximum intensity of unity. The asymmetric broad spectrum for the non-irradiated DLC
has a single peak at 1530 cm−1 , which mainly consists of two spectra at peaks at ~1360 cm−1 (D band)
and ~1590 cm−1 (G band) [21]. The D and G bands are attributed to bond angle disorder in sp2
graphite-like micro/nanodomains and bond stretching between pairs of sp2 atoms in both the rings
and chains, respectively. The ratio of the intensities of the D and G peaks (ID /IG ) and the position
of the G peak have been reported to indirectly indicate the composition ratio of sp2 and sp3 bonding
structures in DLC films [22–24]. These reports have shown that an increase in ID /IG and a shift of the
G peak to a higher frequency represent an increase in the amount of sp2 structures. The spectra from
surfaces ablated with 7-fs and 100-fs pulses, shown in Figure 5, clearly show two spectral peaks at
1355 and 1590 cm−1 , respectively, indicating an increase in disordered carbon or GC [25–28]. As shown
in Figure 5b, ID for the surface irradiated with 7-fs pulses is smaller than that for the surface irradiated
with 100-fs pulses. In addition, the position of the G peak for 7-fs pulses is shifted less than that for
100-fs pulses. These results show that less GC existed in the target surface irradiated with 7-fs laser
pulses compared to that which existed with 100-fs pulses, despite the same F.

Figure 5. (a) Raman spectra of non-irradiated DLC film (gray) and DLC films irradiated with 7-fs (red)
and 100-fs (blue) pulses at F = 6 mJ/cm2 for v = 0.1 μm/s; (b) expanded spectra of (a) in the vicinity of
the peaks of D and G bands.

To examine the the bonding structural change and ablation processes in detail, Raman spectra
were obtained from a DLC film irradiated with 7 fs pulses with I = 1 TW/cm2 at F = 6 mJ/cm2 for
various values of v (v = 0.1–100 μm/s). For comparison, spectra were also obtained from a film
irradiated with 100 fs pulses with I = 0.1 TW/cm2 at F = 6 mJ/cm2 . The peak intensities and positions
of the D and G bands in the spectra were identified using a curve-fitting program with the Lorentzian
function [29]. Figure 6a shows ID /IG plotted as a function of v. In the spectrum of the non-irradiated
DLC film, ID /IG was ~1.25. For v = 100 μm/s, the ratio for both 7-fs and 100-fs pulses increased to ~1.5.
With a decrease in v, the ratio monotonically increased, with that for 7-fs pulses being smaller than that
for 100-fs pulses. Figure 6b shows the position of the G peak plotted as a function of v. In the spectrum
of the non-irradiated DLC film, the G peak position was ~1582 cm−1 . For v = 100 μm/s, the position
for both 7-fs and 100-fs pulses shifted to ~1590 cm−1 . With decreasing v, the position monotonically

5
Nanomaterials 2018, 8, 535

shifted to higher frequencies, with that for 7-fs pulses being at lower frequencies than that for 100-fs
pulses. These results show two crucial processes for surface modification and subsequent ablation.
For v = 100 μm/s, where both 7-fs and 100-fs pulses with the same F induced only swelling and no
ablation on the target, the change in the spectra shown in Figure 6 indicates that the amount of GC
at the surfaces irradiated with 7-fs and 100-fs pulses is the same, and that the surface phenomena do
not depend on I. For v ≤ 10 μm/s, where both 7-fs and 100-fs pulses with the same F induced not
only a bonding structure change but also ablation on the target, the experimental results indicate that
7-fs pulses with higher I were strongly absorbed near the target surface through a nonlinear optical
absorption process, forming a thinner GC layer than that produced by 100-fs pulses. The surface of the
layer was then ablated.

Figure 6. (a) Ratio of intensities of D and G peaks (ID /IG ) and (b) position of G peak for DLC films
irradiated with 7-fs laser pulses with I = 1 TW/cm2 at F = 6 mJ/cm2 (red circles) and those irradiated with
100-fs laser pulses with I = 0.1 TW/cm2 at F = 6 mJ/cm2 (blue squares) as function of scanning speed v.

Based on these experimental results and the physical mechanism for nanostructuring [8,15,16,30],
the origin of the decrease in d with increasing F is discussed. The SPP wavelength λspp was calculated
for the model surface illustrated in the inset of Figure 7, where it was assumed that the fs laser pulses
are incident on the target in air, free electrons are produced at the GC surface to form a thin metal-like
layer on the DLC substrate, and SPPs are excited at the interface between the metal-like layer and
the DLC. The calculation method was almost the same as that used in our previous studies [15,16].
Briefly, λspp = 2π/Re[kspp ] was calculated using the following relation between light and SPPs:

kspp = k0 [εDLC ε*/(εDLC + ε*)]1/2 (1)

where k0 is the wavevector of the incident light in vacuum, and ε* and εDLC are the relative dielectric
constants for the metallic GC and the DLC, respectively. As the GC layer is ionized by fs laser pulses,
ε* rapidly changes during the interaction as:

ε* = εGC − [ω p 2 /(ω 2 + iω/τ)] (2)

where εGC is the static dielectric constant for the GC layer, and the second term represents the effect of
free electrons with a density of Ne produced in the GC layer, where ω is the laser frequency in vacuum,
τ = 1 fs is the Drude damping time for free electrons [31,32], and ω p = [e2 Ne /(ε0 m* m)]1/2 is the plasma
frequency, with the dielectric constant of vacuum ε0 , electron charge e, electron mass m, and optical
effective mass of electrons m* = 1. In the calculation, because the wavelength of the 7-fs laser pulse
used in the present experiment was 680–940 nm, the static dielectric constants for DLC and GC were

6
Nanomaterials 2018, 8, 535

used for three wavelengths: εDLC = 6.9 + i3.8 and εGC = 3.0 + i2.8 for λ = 600 nm; εDLC = 8.0 + i2.9 and
εGC = 3.1 + i3.1 for λ = 800 nm; and εDLC = 8.5 + i2.6 and εGC = 3.6 + i4.5 for λ = 1000 nm [33].
Figure 7 shows the period of the PNS, D = λspp /2 = π/(Re[kspp ]), calculated for λ = 600, 800,
and 1000 nm as a function of Ne . The excitation of SPPs at the interface between the metallic GC
layer and the DLC is allowed for Re[ε*] × Re[εDLC ] < 0 [20], which corresponds to the regions of
Ne > 1.0 × 1022 cm–3 for λ = 600 nm, Ne > 6.4 × 1021 cm–3 for λ = 800 nm, and Ne > 5.2 × 1021 cm–3 for
λ = 1000 nm. With increasing Ne , D decreases from ~200 nm to ~100 nm. Because Ne should increase
with increasing I via stronger nonlinear optical absorption, the decrease in D with increasing Ne is in
good agreement with the decrease in d with increasing I for 7-fs laser pulses shown in Figure 4.

Figure 7. Calculated groove period D as function of Ne in the glassy carbon (GC) layer at λ = 600 nm
(blue), 800 nm (red), and 1000 nm (green), where the excitation of surface plasmon polaritons (SPPs) is
allowed in the region (solid curves) of 0 < Re[ε*]. The inset shows a schematic drawing of the initial
target surface modeled for calculation. SPPs (left/right arrows) are excited at the interface between the
DLC and GC layers by a high density of electrons produced by irradiation using high power -density
laser pulses (down arrow).

The present experimental and calculation results show that the period d for a PNS was smaller
than D, and that the d value for a PNS formed with high-fluence 100-fs laser pulses was similar
to D for λ = 800 nm, which is consistent with the results of a previous study [15]. Regarding the
excitation of SPPs on a thin metal film, it has been reported that the wavenumber of the SPPs increases
with decreasing thickness of the film because of an increase in the radiation damping of SPPs [20].
These results suggest that d being smaller than D can be attributed to the excitation of SPPs with
a larger wavenumber by the thinner metallic layer produced with 7-fs laser pulses. A calculation
model for D that includes the effect of the metallic layer thickness will be presented and discussed
in a separate paper. To discuss the formation process of PNS in detail and make a more accurate
model for the nanostructuring, we need to quantitatively measure the amount and thickness of the
GC layer on DLC film by using advanced techniques, such as a grazing-incidence small-angle X-ray
scattering [34–36].

4. Conclusions
This study examined the PNS that formed on a DLC film with 7-fs laser pulses at a low fluence
from a laser oscillator. The results show the formation of a PNS with a period of d = 60–80 nm and
a decrease in d with increasing fluence. Based on the experimental results and a model calculation,
it is shown that the excitation of SPPs at the interface between the thin metal-like layer and the DLC is

7
Nanomaterials 2018, 8, 535

certainly responsible for the nanostructuring process, and that the decrease of d is attributed to the
wavelength of the SPPs decreasing with increasing F due to an increase of electron density in the thin
metal-like layer.

Author Contributions: S.N. and R.S. performed the ablation experiments; T.N. fabricated the micro Raman
spectroscope and analyzed the bonding structure of ablated surfaces; G.M. designed the experiments and made
the physical model; all the authors wrote this paper together.

Funding: This research was supported in part by the Amada Foundation 2015 and Joint
Usage/Research Program on Zero-Emission Energy Research, Institute of Advanced Energy, Kyoto
University (ZE29C-1, 2017).

Acknowledgments: The authors would like to thank K.S. for the preliminary experiments, K.M. for helpful
comments and useful discussions on Raman spectroscopy, and A.H. and Y.H. for the composition analysis of
DLC films.
Conflicts of Interest: The authors declare no conflict of interest.

References and Note


1. Bonse, J.; Sturm, H.; Schmidt, D.; Kautek, W. Chemical, morphological and accumulation phenomena in
ultrashort-pulse laser ablation of TiN in air. Appl. Phys. A Mater. Sci. Process. 2000, 71, 657–665. [CrossRef]
2. Reif, J.; Costache, F.; Henyk, M.; Pandelov, S.V. Ripples revisited: Non-classical morphology at the bottom
of femtosecond laser ablation craters in transparent dielectrics. Appl. Surf. Sci. 2002, 197–198, 891–895.
[CrossRef]
3. Yasumaru, N.; Miyazaki, K.; Kiuchi, J. Femtosecond-laser-induced nanostructure formed on hard thin films
of TiN and DLC. Appl. Phys. A Mater. Sci. Process. 2003, 76, 983–985. [CrossRef]
4. Wu, Q.; Ma, Y.; Fang, R.; Liao, Y.; Yu, Q. Femtosecond laser-induced periodic surface structure on diamond
film. Appl. Phys. Lett. 2003, 82, 1703–1705. [CrossRef]
5. Borowiec, A.; Haugen, H.K. Subwavelength ripple formation on the surfaces of compound semiconductors
irradiated with femtosecond laser pulses. Appl. Phys. Lett. 2003, 82, 4462–4464. [CrossRef]
6. Daminelli, G.; Krüger, J.; Kautek, W. Femtosecond laser interaction with silicon under water confinement.
Thin Solid Films 2004, 467, 334–341. [CrossRef]
7. Bonse, J.; Krüger, J. Femtosecond laser-induced periodic surface structures. J. Appl. Phys. 2012, 24,
042006-1–042006-7. [CrossRef]
8. Miyazaki, K.; Miyaji, G. Mechanism and control of periodic surface nanostructure formation with
femtosecond laser pulses. Appl. Phys. A Mater. Sci. Process. 2014, 114, 177–185. [CrossRef]
9. Miyazaki, K.; Miyaji, G.; Inoue, T. Nanograting formation on metals in air with interfering femtosecond laser
pulses. Appl. Phys. Lett. 2015, 107, 071103. [CrossRef]
10. Miyaji, G.; Miyazaki, K. Fabrication of 50-nm period gratings on GaN in air through plasmonic near-field
ablation induced by ultraviolet femtosecond laser pulses. Opt. Express 2016, 24, 4648–4653. [CrossRef]
[PubMed]
11. Yasumaru, N.; Miyazaki, K.; Kiuchi, J. Glassy carbon layer formed in diamond-like carbon films with
femtosecond laser pulses. Appl. Phys. A Mater. Sci. Process. 2004, 79, 425–427. [CrossRef]
12. Miyazaki, K.; Maekawa, N.; Kobayashi, W.; Kaku, M.; Yasumaru, N.; Kiuchi, J. Reflectivity in
femtosecond-laser-induced structural changes of diamond-like carbon film. Appl. Phys. A Mater. Sci. Process.
2005, 80, 17–21. [CrossRef]
13. Miyaji, G.; Miyazaki, K. Ultrafast dynamics of periodic nanostructure formation on diamond-like carbon
films irradiated with femtosecond laser pulses. Appl. Phys. Lett. 2006, 89, 191902. [CrossRef]
14. Miyaji, G.; Miyazaki, K. Nanoscale ablation on patterned diamond-like carbon film with femtosecond laser
pulses. Appl. Phys. Lett. 2007, 91, 123102. [CrossRef]
15. Miyaji, G.; Miyazaki, K. Origin of periodicity in nanostructuring on thin film surfaces ablated with
femtosecond laser pulses. Opt. Express 2008, 16, 16265–16271. [CrossRef] [PubMed]
16. Miyaji, G.; Miyazaki, K.; Zhang, K.; Yoshifuji, T.; Fujita, J. Mechanism of femtosecond-laser-induced periodic
nanostructure formation on crystalline silicon surface immersed in water. Opt. Express 2012, 20, 14848–14856.
[CrossRef] [PubMed]

8
Nanomaterials 2018, 8, 535

17. Miyazaki, K.; Miyaji, G. Nanograting formation through surface plasmon fields induced by femtosecond
laser pulses. J. Appl. Phys. 2013, 114, 153108. [CrossRef]
18. Miyaji, G.; Miyazaki, K. Shaping of nanostructured surface in femtosecond laser ablation of thin films.
Appl. Phys. A Mater. Sci. Process. 2010, 98, 927–930. [CrossRef]
19. Kärtner, F.X.; Morgner, U.; Schibli, T.; Ell, R.; Haus, H.A.; Fujimoto, J.G.; Ippen, E.P. Few-Cycle Pulses Directly
from a Laser. In Few-Cycle Laser Pulse Generation and Its Applications, 1st ed.; Kärtner, F.X., Ed.; Springer:
Berlin/Heidelberg, Germany, 2004; pp. 73–136. ISBN 978-3-540-20115-1.
20. Raether, H. Surface Plasmons on Smooth and Rough Surfaces and on Gratings; Springer-Verlag: Heidelberg,
Germany, 1988; ISBN 978-3-540-47441-8.
21. Yoshikawa, M.; Katagiri, G.; Ishida, H.; Ishitani, A. Raman spectra of diamondlike amorphous carbon films.
J. Appl. Phys. 1988, 64, 6464–6468. [CrossRef]
22. Zhang, S.; Zeng, X.T.; Xie, H.; Hing, P. A phenomenological approach for the Id/Ig ratio and sp3 fraction of
magnetron sputtered a-C films. Surf. Coat. Technol. 2000, 123, 256–260. [CrossRef]
23. Robertson, J. Diamond-like amorphous carbon. Mater. Sci. Eng. 2002, R37, 129–281. [CrossRef]
24. Tai, F.C.; Lee, S.C.; Wei, C.H.; Tyan, S.L. Correlation between ID /IG Ratio from Visible Raman Spectra and
sp2 /sp3 Ratio from XPS Spectra of Annealed Hydrogenated DLC Film. Mater. Trans. 2006, 47, 1847–1852.
[CrossRef]
25. Tuinstra, F.; Koenig, J.L. Raman Spectrum of Graphite. J. Chem. Phys. 1970, 53, 1126–1130. [CrossRef]
26. Nemanich, R.J.; Solin, S.A. First- and second-order Raman scattering from finite-size crystals of graphite.
Phys. Rev. B 1979, 20, 392–400. [CrossRef]
27. Yoshikawa, M.; Nagai, N.; Matsuki, M.; Fukuda, H.; Katagiri, G.; Ishida, H.; Ishitani, A.; Nagai, I. Raman
scattering from sp2 carbon clusters. Phys. Rev. B 1992, 46, 7169–7174. [CrossRef]
28. Ferrari, A.C.; Robertson, J. Interpretation of Raman spectra of disordered and amorphous carbon. Phys. Rev.
B 2000, 61, 14095–14107. [CrossRef]
29. Wojdyr, M. Fityk: A general-purpose peak fitting program. J. Appl. Crystal. 2010, 43, 1126–1128. [CrossRef]
30. Miyaji, G.; Miyazaki, K. Role of multiple shots of femtosecond laser pulses in periodic surface nanoablation.
Appl. Phys. Lett. 2013, 103, 071910. [CrossRef]
31. Sokolowski-Tinten, K.; Linde, D. Generation of dense electron-hole plasmas in silicon. Phys. Rev. B 2000, 61,
2643–2650. [CrossRef]
32. Because the Drude damping time of the DLC films or GC was never measured, we used the value of Si
reported in Ref. 31. Using the time in a range of 1–10 fs, we confirmed that the calculation results were
almost the same as that with the time of 1 fs.
33. Alterovitz, S.A.; Savvides, N.; Smith, F.W.; Woollam, J.A. Amorphous Hydrogenated “Diamondlike” Carbon
Films and Arc-Evaporated Carbon Films. In Handbook of Optical Constants of Solids; Palik, E.D., Ed.; Academic
Press: San Diego, CA, USA, 1985; pp. 838–852. ISBN 978-0-125-44423-1.
34. Rebollar, E.; Pérez, S.; Hernández, J.J.; Martín-Fabiani, I.; Rueda, D.R.; Ezquerra, T.A.; Castillejo, M.
Assessment and Formation Mechanism of Laser-Induced Periodic Surface Structures on Polymer Spin-Coated
Films in Real and Reciprocal Space. Langmuir 2011, 27, 5596–5606. [CrossRef] [PubMed]
35. Rebollar, E.; Rueda, D.R.; Martín-Fabiani, I.; Rodríguez-Rodríguez, Á.; García-Gutiérrez, M.-C.; Portale, G.;
Castillejo, M.; Ezquerra, T.A. In Situ Monitoring of Laser-Induced Periodic Surface Structures Formation on
Polymer Films by Grazing Incidence Small-Angle X-ray Scattering. Langmuir 2015, 31, 3973–3981. [CrossRef]
[PubMed]
36. Roth, S.V.; Döhrmann, R.; Gehrke, R.; Röhlsberger, R.; Schlage, K.; Metwalli, E.; Körstgens, V.;
Burghammer, M.; Riekel, C.; David, C.; et al. Mapping the morphological changes of deposited gold
nanoparticles across an imprinted groove. J. Appl. Crystallogr. 2015, 48, 1827–1833. [CrossRef]

© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

9
nanomaterials

Article
Formation of Slantwise Surface Ripples by
Femtosecond Laser Irradiation
Xin Zheng 1,2 , Cong Cong 3 , Yuhao Lei 1 , Jianjun Yang 1, * and Chunlei Guo 1,4, *
1 The Guo China-US Photonics Laboratory, State Key Laboratory of Applied Optics,
Changchun Institute of Optics, Fine Mechanics and Physics, Chinese Academy of Sciences,
Changchun 130033, China; [email protected] (X.Z.); [email protected] (Y.L.)
2 College of Materials Science and Opto-Electronic Technology, University of Chinese Academy of Science,
Beijing 100049, China
3 School of the Gifted Young, University of Science and Technology of China, Hefei 230026, China;
[email protected]
4 The Institute of Optics, University of Rochester, Rochester, NY 14627, USA
* Correspondence: [email protected] (J.Y.); [email protected] (C.G.);
Tel.: +86-138-2070-1265 (J.Y.); +1-585-275-2134 (C.G.)

Received: 26 May 2018; Accepted: 19 June 2018; Published: 22 June 2018

Abstract: We report on the formation of slantwise-oriented periodic subwavelength ripple structures


on chromium surfaces irradiated by single-beam femtosecond laser pulses at normal incidence.
Unexpectedly, the ripples slanted in opposite directions on each side the laser-scanned area,
neither perpendicular nor parallel to the laser polarization. The modulation depth was also found to
change from one ripple to the next ripple. A theoretical model is provided to explain our observations,
and excellent agreement is shown between the simulations and the experimental results. Moreover,
the validity of our theory is also confirmed on bulk chromium surfaces. Our study provides insights
for better understanding and control of femtosecond laser nanostructuring.

Keywords: femtosecond laser; laser-induced periodic surface structures; anomalous slanting


ripples; chromium

1. Introduction
During the last several years, the research of femtosecond laser-induced periodic surface structures
(Fs-LIPSSs), or the ripple structures, has attracted tremendous attention because of the abundant
scientific issues involved [1–3]. Fs-LIPSSs have been studied on a variety of materials, including
metals, semiconductors, and dielectrics [4–7]. It has been found that such microstructures have
extensive potential applications, such as magnetic recording media [8], self-cleaning materials [9,10],
anti-reflective metals [11], and solar sensors [12].
In general, the distinct characteristics of the ripple structures are closely dependent on the laser
parameters. When the linearly polarized single-beam femtosecond laser pulses are used to irradiate
materials, the induced ripple structures are either parallel or perpendicular to the direction of the laser
polarization [13–16]. In some cases, however, the ripple structures induced by femtosecond lasers
presented an unusual feature of slantwise orientation, which is neither perpendicular nor parallel to
the laser polarization direction [17–22]. For instance, Qiu et al. [17] reported the slantwise oriented
nanogrooves on a ZnO crystal surface with normal incidence of the single-beam femtosecond laser,
which tended to be perpendicular to the direction of the laser polarization at the increased scanning
speed. By tilting the incident angle of femtosecond laser pulses, Schwarz et al. [19] experimentally
observed the slantwise orientation of the ripple structures on fused silica, and the structure orientation
changed as a function of the laser incident angle. More recently, our research group generated a series

Nanomaterials 2018, 8, 458; doi:10.3390/nano8070458 10 www.mdpi.com/journal/nanomaterials


Nanomaterials 2018, 8, 458

of v-like structures, called a herringbone pattern, on copper [23]. Such anomalous phenomena indicate
the physical complexities during the ripple surface structure formation, which is actually significant
for femtosecond laser nanoprocessing. Nevertheless, a comprehensive underlying mechanisms of
LIPSS orientation is still lacking.
In this work, the formation of slantwise oriented ripple structures is systematically investigated
on chromium surfaces by employing single-beam femtosecond laser pulses at normal incidence.
First, the ripple structures generated on two lateral edges of the laser-scanned area are seen to have
different slantwise orientations with respect to the direction of the laser polarization, and such
behaviors occur even when the laser polarization changes. Secondly, based on the measured
modulation depth of the ripple structures, we develop a theoretical model to elucidate the underlying
mechanisms via the consistent simulations. Finally, additional experiments are performed to confirm
the theory.

2. Materials and Methods


As shown by a schematic illustration of the experimental setup in Figure 1, a commercial
chirped-pulse-amplification of a Ti:sapphire laser system (Spitfire Ace, Spectra Physics, Santa Clara, CA,
USA) was employed as a light source for producing the surface structures, which delivers horizontally
polarized femtosecond laser pulse trains with a repetition rate, a central wavelength, and a time
duration of 1 kHz, 800 nm, and 40 fs, respectively. The maximum energy of each laser pulse was 7 mJ.
Neutral density filters and a half-wave plate were used to control the pulse energy and the direction
of the laser polarization, respectively. The laser beam was focused by an objective lens (plan fluorite
objective, 4×, N.A = 0.13, f = 17.2 mm, Nikon, Tokyo, Japan) at normal incidence. The sample was
mounted on a three-dimensional translation stage (ESP301, Newport Inc., Irvine, CA, USA) that could
be precisely translated via a custom-made computer program. In order to avoid serious ablation of the
material, the sample surface was moved 300 μm away from the focus towards the lens, such that the
focus is located inside the sample, resulting in a laser spot radius of ≈39 μm on the sample surface.
In the experiments, 100 nm-thickness chromium (Cr) films deposited on SiO2 substrate
were chosen as sample materials because of its good physical characteristics, including hardness,
corrosion resistance, high melting point, and adhesiveness, which earn many applications in solar
absorbers, adhesion layers, micro-electromechanical systems devices, etc. [24–26]. Besides, we also
used bulk Cr material to carry out experiments. Based on the method of the previous reports [25,27],
we experimentally obtained the ablation threshold values of a single laser pulse for the film and
bulk of Cr material, which can be given as about 37.5 mJ/cm2 and 248 mJ/cm2 , respectively. All the
experimental performances were carried out in ambient air environment. After the laser irradiation,
the surface morphologies were investigated using a scanning electron microscope (SEM, Phenom,
Eindhoven, Netherlands) and an atomic force microscope (AFM, Bruker, Billerica, MA, USA).

Figure 1. A schematic experimental setup for the formation of slantwise-oriented ripple structures on
chromium surface by femtosecond laser irradiation.

11
Nanomaterials 2018, 8, 458

3. Results and Discussions


Figure 2a exhibits the surface morphology of the Cr film irradiated by single-beam femtosecond
laser pulses at the fluence of F = 56.9 mJ/cm2 with the scanning speed of V = 0.3 mm/s. Due to the
incident Gaussian laser intensity, different surface morphologies could be observed on the area where
the laser scanned; substantial ablation damages occurred in the central area, and there were ripple
structures on the lateral edges, with a spatial period of approximate 650 nm. In particular, the periodic
ripple structures on both edge regions were found to have a slantwise orientation in two different
directions. This behavior is in sharp contrast to the previous reports [13–16], wherein the ripple
orientation is usually either perpendicular or parallel to the direction of the incident laser polarization.
To further characterize the features of such slantwise-oriented ripple structures, we employed
AFM to measure their modulation depths, and the corresponding results are shown in Figure 2b.
Clearly, the measured oscillation curve reveals that the modulation depth of the ripple structures
decreases gradually with increasing the distance from the center to the lateral edges of the laser-scanned
area, which is due to the spatially inhomogeneous distribution of the laser pulse intensity. On the other
hand, the measured peaks suggest that the modulation height of the surface is also varied as a function
of the distance from the center of the laser-scanned area. This can be physically understood as follows:
The film thickness decreased after irradiation of multiple femtosecond laser pulses, leading to the film
thinning at the center of the laser-scanned area with respect to the lateral edge regions. Consequently,
the formation of such slantwise oriented periodic ripple structures is in fact based on the gradient
variation of the film thicknesses.
Noticeably, the measured height of the ripple structures was larger than the thickness of the Cr
film, which may have been due to material reaction with O2 in the ambient atmosphere, leading to
oxide formation on the material surface [28]. More specifically, as shown in Figure 2c, there are
two physical processes happening in the formation of the laser-induced ripple structures: one is the
spatially periodic removal of chromium materials by the modulated laser intensity fringes, and the
other is the growth of chromium oxides at the places where the laser intensity is higher than the
threshold of oxidation. Usually, the laser damage threshold is larger than that of the oxidation process.
Here it should be clear that the periodic femtosecond laser intensity distribution for the ripple structure
formation is originated from interference of the light and its excited surface plasmons [29–32]. Because
the two components possess unequal energies, i.e., the energy of the excited surface plasmons is
usually smaller than that of the incident laser pulse, their interfering intensity patterns, which had a
Gaussian variation profile tend to give a low fringe contrast, or the deconstructive interference fringes
can also hold a certain level of the laser energy. Under such circumstances, the material oxidation can
take place during the formation of the periodic ripple structures, and the resultant additional oxide
layers on the ridge surfaces make the height of the ripple structures become protuberant with respect
to the original film thickness.
Inspired by the anomalous phenomenon of the ripple structures with the slantwise orientation,
we also performed a series of experiments on Cr films by varying the direction of linear polarization of
the femtosecond laser. As shown by the results in Figure 3 (here only the observations on both lateral
edges of the laser-scanned area are shown), for the given laser polarization, the slantwise-oriented
periodic ripple structures are always produced on both lateral edge regions of the laser-scanned
area, being very similar to the observation in Figure 2. Whereas for different laser polarizations,
the slantwise degree of the ripple structures is found to change but still neither perpendicular nor
parallel to the laser polarization direction. Therefore, we can conclude that the formation of slantwise
orientated periodic ripple structures seems to always appear even for different linear polarizations of
femtosecond laser pulses.

12
Nanomaterials 2018, 8, 458

Figure 2. (a) SEM image of the ripple structure formation on a Cr film surface irradiated by single-beam
femtosecond laser irradiation at the energy fluence of F = 56.9 mJ/cm2 with the sample scanning speed
of V = 0.3 mm/s; (b) AFM image with the cross-section profiles of the ripple structures formed on both
lateral edge regions of the laser-scanned area. Arrows of S and E represent directions of the sample
scanning and the laser polarization, respectively; (c) Schematic plots of the periodically distributed
intensity distribution on the Cr surface (upper), and its induced ripple structures (bottom), where Ith
and Iox indicate the threshold intensities of the material damage and oxidation processes, respectively.
The oxidation layer on the top parts of the surface structures are represented by a purple color.

Figure 3. Slantwise-oriented ripple structures on two lateral edge regions (Top and Bottom) of the
laser-scanned area on Cr film surfaces by different linear polarizations of single-beam femtosecond
laser pulses. The angle θ on the upper-left corner of each image represents the direction of the laser
polarization. The blue dash lines identify the orientations of the ripple structures. The angle of γ
indicates an intersection angle between the ripple orientation and the laser polarization direction of
θ = 0◦ . The scale bar is applied to all images in this figure.

13
Nanomaterials 2018, 8, 458

To elucidate our experimental observations, we proposed the following physical scenario:


In our experiments, which are in fact based on multi-pulse femtosecond laser irradiation processes,
the pristine surface of the metal film was modified by the preceding incident femtosecond laser pulses,
leading to a rough, shallow crater with the modulation depth gradually reducing from the beam
center to the peripheral regimes, as shown in Figure 4a, where the inclined surface was created on
the laser irradiation area. After that, for the continuous irradiation of the subsequent femtosecond
laser pulses, the inclining degree of the laser irradiation surface became pronounced (Figure 4b),
and the periodic subwavelength ripple structures were also developed on it, exhibiting the slantwise
orientation with respect to the direction of the laser polarization, as shown in Figure 4c. Noticeably,
due to the higher intensity distribution on the central region of the laser-scanned area, the formation
of the corresponding ripple structures was seriously deteriorated by the accumulating irradiation of
subsequent femtosecond laser pulses.

Figure 4. Schematic diagrams of the physical processes for the formation of slantwise-oriented periodic
ripple structures on the metal surface. E represents the direction of the laser polarization. The different
colors represent variations of the modulation depth, which tended to cause the inclined surface within
the laser irradiation area.

In fact, the effects of the inclined surface on the formation of slantwise-orientated periodic
ripple structures can be theoretically analyzed. According to the previous study of Pham et al. [33],
the presence of the inclined surface on the laser spot area can be described by pX + qY + Z = 1,
within a three-dimensional Cartesian coordinate system X-Y-Z, as shown in Figure 5a. Here p and q are
the geometrical parameters for describing the spatial characteristics of the inclined surface. A normal
component of the inclined surface is represented by n = (p, q, 1). Both the propagation direction and
the electric field vectors of the incident femtosecond laser are defined as Li = (0, 0, 1) and Ei = (cosθ,
sinθ, 0), respectively, wherein θ is an intersection angle between Ei and the X-axis. As shown in
Figure 5a, a plane of the laser incidence (represented by a blue color) is established by the vectors
of n and Li , whose intersection angle is defined by θ i . Moreover, a coordinate system x -y -z is also
built for simplifying the calculation of the electric field on the inclined surface. For the incidence of
femtosecond laser on the inclined surface, its electric field vector Ei is divided into two components
of Ex and Em through its projection onto the x -y and the incident planes, respectively. On the other
hand, the projection of the electric field component Em on the x -y plane is indicated by Ey Finally,
in the x -y plane, the two electric field components Ex and Ey can be developed into a new vector
of Ex y , as shown in Figure 5b, with β being an intersection angle between Ex and Ex y , which is
calculated by the following expression [33]:
⎛
2 ⎞
 2
⎜ n cos θi n2 − sin θi + n3 − n sin θi + κ 2 n2 − sin θi
2 2 2 ⎟
⎜ ⎟
β = arctan⎜  
cot α ⎟ (1)
⎜  2 ⎟
⎝ ⎠
n n − sin θi + n cos θi + n κ cos θi
2 2 3 4 2 2

q cos θ − p sin θ
where θi = arccos √ 1
and α = arcsin √ 2 2 .
1+ p2 + q2 p +q

14
Nanomaterials 2018, 8, 458

Figure 5. (a) A sketch of the laser incidence onto the inclined surface of the material and the
decomposition of the electric field Ei onto different planes; (b) An effective electric field vector on the
x -y plane and its induced periodic ripple structures with the orientation vector k.

In Equation (1), n and κ represent the real and the imaginary parts of the complex refractive
index n of the material, respectively. Accordingly, the orientation vector of the ripple structures on
the inclined surface, k, should be perpendicular to the direction of the electric field Ex y , as shown in
Figure 5b. When the ripple orientation k in the x -y -z coordinate system is transferred into the X-Y-Z
coordinate system, it should be modified into:
 2 
p cos β q cos β p + q2 cos β
k= q sin β −  , − p sin β −  ,  (2)
1 + p2 + q2 1 + p2 + q2 1 + p2 + q2

By considering the actual observation surface happening on the X-Y plane, the orientation vector
k can be re-written as:
 
p cos β q cos β
k = q sin β −  , − p sin β −  , 0 (3)
1 + p2 + q2 1 + p2 + q2

In the experiments, the orientation vector k was obtained by the measurement of angle γ shown in
Figure 3. Thus, the assumed geometric parameters (p, q) of the inclined surface could be calculated by
the non-linear fitting of Equation (3) with the help of the measured values k = (cos γ, sin γ, 0).
For example, with the experimentally measured angles of γ, the achieved p and q values were
(ptop = −0.6399, qtop = −0.8141) and (pbottom = −0.5077, qbottom = 0.6690) for the top and
bottom edges of the laser-scanned area, respectively.
Through combing the above calculated p and q values with the expression of pX + qY + Z = 1,
we could map three-dimensional profiles of the two inclined surfaces, as shown in Figure 6a, where the
left and right surfaces indicate the top and bottom edges of the laser-scanned area, respectively.
Evidently, for each inclined surface, the modulation height is varied as a function of the x-y position.
In addition, we could also calculate the ripple orientation for the single beam femtosecond laser
at different polarization directions. Specifically, because the structure orientation is indicated by
p cos β q cos β
the vector k = (cos γ, sin γ, 0) where cos γ = q sin β − √ 2 2 and sin γ = − p sin β − √ 2 2 ,
1+ p + q 1+ p + q
we could obtain γ values with the available parameters of p and q. Therefore, by changing the laser
polarization from θ = 0◦ to 180◦ , the theoretical fitting of the ripple orientation angle γ on the top
and bottom edges of the laser-scanned area could be obtained, as shown (by red solid curves) in
Figure 6b,c, respectively, wherein the experimental data are given (by blue solid circles) with the
standard deviation. It is seen clearly that the theory and experiment have good consistency in the two
cases. Another feature is that the obtained ripple orientation angle vs the laser polarization direction
had nonlinear variations, which was basically due to the polarization dependent optical absorption.

15
Nanomaterials 2018, 8, 458

Figure 6. (a) Theoretically retrieving the inclined surfaces on the top and bottom edges of the
laser-scanned area in the coordinate system of X-Y-Z, with the help of the calculated parameters
of (ptop = −0.6399, qtop = −0.8141) and (pbottom = −0.5077, qbottom = 0.6690), respectively;
(b,c) compare the simulation results with the experimental data for the ripple orientation angles
on the top and bottom edges of the laser-scanned area, respectively.

In order to confirm the above theoretical analyses, we carried out further experiments on the
surface of Cr bulk material. As shown in Figure 7a, when the femtosecond laser energy fluence was
given at F = 40.6 mJ/cm2 , the subwavelength ripple structures with orientation perpendicular to the
laser polarization can still be formed in the central parts of the laser-scanned area. While on both
lateral edges (marked by the red dot frames) of the laser-scanned area, the obtained ripple structures
exhibit different slantwise orientations. From the corresponding AFM image, as shown in Figure 7b,
we can also find that the modulation of the ripple depth is decreased with larger distances from
the center of the laser scribed area, which is attributed to the Gaussian beam profile distribution.
Clearly, the measured varying tendencies of the ripple depth on the Cr bulk material are very similar
to the observations on Cr films.

Figure 7. (a) SEM image of the ripple structure formation on the surface of Cr bulk material irradiated
by single-beam femtosecond laser pulses at the energy fluence of F = 40.6 mJ/cm2 ; (b) AFM image
with the cross-section profiles for the ripple structures on two edge regions of the laser-scanned area.

On the other hand, when the femtosecond laser energy fluence was decreased to approximately
F = 25.7 mJ/cm2 , the ripple structures turned out to be oriented perpendicular to the direction of

16
Nanomaterials 2018, 8, 458

the laser polarization, especially on both lateral edges of the laser-scanned area, shown in Figure 8a,
being similar to many previous reports [13–16]. As a matter of fact, this situation can be maintained
even for the laser energy fluences of about F = 29.6 mJ/cm2 , as shown in Figure 8b, where the
ripple-covered region seemed to be enlarged with the orientation still perpendicular to the direction of
the laser polarization. Based on the experimental comparisons, it is revealed that the spatial alignment
of the ripple structures can be transferred from the slantwise tendency into the direction perpendicular
to the laser polarization, if the femtosecond laser energy fluence is weak enough. In other words,
the incident larger energy fluence of the Gaussian laser beam was a key factor for the formation of
the inclined surface during the multi-pulse laser irradiation, which finally resulted in the slantwise
oriented ripple structures on the lateral edges of the laser-scanned area.
Such a ripple orientation-transferring process can be understood as follows: For a Gaussian laser
pulse irradiation on the material with a damage threshold intensity of Ith , the resultant ablation depth
varied as a function of the distance away from the beam center. Therefore, the gradient surfaces were
likely to be modified on the ablation
edges, with an incline angle ϕ proportional to the variation rate
of the laser intensity, i.e., ϕ ∝ ln II0 , where I0 was the peak intensity of the laser pulse. Evidently,
th
with increasing the laser energy fluence, the higher peak intensity could result in a larger incline degree
of the ablation surface, as shown in Figure 8c, and the consequent formation of the slantwise oriented
ripple structures. Whereas for a femtosecond laser with lower energy fluences, the peak intensity
caused a smaller incline degree of the ablation surface, providing negligible influence on orientation of
the ripple structures.

Figure 8. SEM images of the Cr bulk surfaces irradiated by single-beam femtosecond laser pulses with
the different energy fluences. (a) F = 25.7 mJ/cm2 ; (b) F = 29.6 mJ/cm2 ; (c) Variation rates (dot curves)
of the laser intensity at the damage threshold (Ith ) for two cases of different pulse energy fluences,
where I0 and I0 are the peak intensities of two laser pulses.

4. Conclusions
In conclusion, we comprehensively studied the generation of slantwise-oriented subwavelength
periodic ripple structures on the surfaces of chromium material by normal incidence of linearly
polarized femtosecond laser pulses. Our experimental results on chromium films demonstrated that
the ripple structures formed on two lateral edges of the laser-scanned area are slantwise-oriented in
two different directions, being neither perpendicular nor parallel to the laser polarization. When the

17
Nanomaterials 2018, 8, 458

laser polarization direction is changed, the slantwise ripple structures were still observable but with
different orientations, AFM measurements suggested that the modulation height of the ripple-covered
surface exhibited gradual variations with the distance from the center to the lateral edges, which is due
to the inhomogeneous distribution of the Gaussian laser intensity. A physical model was proposed by
considering the inclined ablation surfaces after multi-pulse irradiations. The agreement between the
simulations and the measured results confirmed the validity of our theory.
The above mentioned slantwise ripple structures were also generated on bulk Cr surfaces. For the
reduced laser fluence, however, the slantwise observations began to transfer into the commonly
observed ripple structures with an orientation perpendicular to the direction of the laser polarization,
which indicated the negligible influence of the inclined ablation surface. Our investigation provides
a comprehensive understanding of femtosecond laser-material interactions, which may help us
design and fabricate uniform subwavelength and even nanoscale structures and devices for the
future applications.

Author Contributions: C.G. and J.Y. conceived and designed the experiments; X.Z. and Y.L. performed the
experiments; C.C. performed the simulations; X.Z., Y. L., and C.C. analyzed the data; X.Z. and J.Y. wrote the paper.
Funding: The authors would like to acknowledge the support from National Key R&D Program of China
(2017YFB1104700); National Natural Science Foundation of China (11674178, 61774155,); Natural Science
Foundation of Tianjin City (17JCZDJC37900); Jilin Provincial Science & Technology Development Project
(20180414019GH).
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Krüger, J.; Kautek, W. The femtosecond pulse laser: A new tool for micromachining. Laser Phys. 1999, 9,
30–40.
2. Vorobyev, A.Y.; Guo, C.L. Direct femtosecond laser surface nano/microstructuring and its applications.
Laser Photonics Rev. 2013, 7, 385–407. [CrossRef]
3. Sugioka, K.; Cheng, Y. Ultrafast lasers—Reliable tools for advanced materials processing. Light Sci. Appl.
2014, 3, e149. [CrossRef]
4. Wang, J.C.; Guo, C.L. Ultrafast dynamics of femtosecond laser-induced periodic surface pattern formation
on metals. Appl. Phys. Lett. 2005, 87, 251914. [CrossRef]
5. Xue, L.; Yang, J.J.; Yang, Y.; Wang, Y.; Zhu, X. Creation of periodic subwavelength ripples on tungsten surface
by ultra-short laser pulses. Appl. Phys. A 2012, 109, 357–365. [CrossRef]
6. Borowiec, A.; Haugen, H.K. Subwavelength ripple formation on the surfaces of compound semiconductors
irradiated with femtosecond laser pulses. Appl. Phys. Lett. 2003, 82, 4462–4464. [CrossRef]
7. Hnatovsky, C.; Taylor, R.S.; Rajeev, P.P.; Simova, E.; Bhardwaj, V.R.; Rayner, D.M.; Corkum, P.B. Pulse duration
dependence of femtosecond-laser-fabricated nanogratings in fused silica. Appl. Phys. Lett. 2005, 87, 014104.
[CrossRef]
8. Stupakiewicz, A.; Szerenos, K.; Afanasiev, D.; Kirilyuk, A.; Kimel, A.V. Ultrafast nonthermal photo-magnetic
recording in a transparent medium. Nature 2017, 542, 71. [CrossRef] [PubMed]
9. Wu, B.; Zhou, M.; Li, J.; Ye, X.; Li, G.; Cai, L. Superhydrophobic surfaces fabricated by microstructuring of
stainless steel using a femtosecond laser. Appl. Surf. Sci. 2009, 256, 61–66. [CrossRef]
10. Vorobyev, A.Y.; Guo, C.L. Multifunctional surfaces produced by femtosecond laser pulses. J. Appl. Phys.
2015, 117, 033103. [CrossRef]
11. Vorobyev, A.Y.; Makin, V.S.; Guo, C.L. Brighter light sources from black metal: Significant increase in
emission efficiency of incandescent light sources. Phys. Rev. Lett. 2009, 102, 234301. [CrossRef] [PubMed]
12. Hwang, T.Y.; Vorobyev, A.Y.; Guo, C.L. Enhanced efficiency of solar-driven thermoelectric generator with
femtosecond laser-textured metals. Opt. Express 2011, 19 (Suppl. 4), A824–A829. [CrossRef] [PubMed]
13. Qi, L.T.; Nishii, K.; Namba, Y. Regular subwavelength surface structures induced by femtosecond laser
pulses on stainless steel. Opt. Lett. 2009, 34, 1846–1848. [CrossRef] [PubMed]

18
Nanomaterials 2018, 8, 458

14. He, W.L.; Yang, J.J.; Guo, C.L. Controlling periodic ripple microstructure formation on 4h-sic crystal with
three time-delayed femtosecond laser beams of different linear polarizations. Opt. Express 2017, 25, 5156–5168.
[CrossRef] [PubMed]
15. Jia, X.; Jia, T.Q.; Zhang, Y.; Xiong, P.X.; Feng, D.H.; Sun, H.Y.; Qiu, J.R.; Xu, Z.Z. Periodic nanoripples in the
surface and subsurface layers in zno irradiated by femtosecond laser pulses. Opt. Lett. 2010, 35, 1248–1250.
[CrossRef] [PubMed]
16. Tang, Y.F.; Yang, J.J.; Zhao, B.; Wang, M.W.; Zhu, X.N. Control of periodic ripples growth on metals by
femtosecond laser ellipticity. Opt. Express 2012, 20, 25826–25833. [CrossRef] [PubMed]
17. Liu, J.K.; Jia, T.Q.; Zhou, K.; Feng, D.H.; Zhang, S.A.; Zhang, H.X.; Jia, X.; Sun, Z.R.; Qiu, J.R. Direct writing
of 150 nm gratings and squares on zno crystal in water by using 800 nm femtosecond laser. Opt. Express
2014, 22, 32361–32370. [CrossRef] [PubMed]
18. Yang, Y.; Yang, J.J.; Liang, C.; Wang, H.; Zhu, X.; Kuang, D.; Yang, Y. Sub-wavelength surface structuring of
niti alloy by femtosecond laser pulses. Appl. Phys. A 2008, 92, 635–642. [CrossRef]
19. Schwarz, S.; Rung, S.; Hellmann, R. One-dimensional low spatial frequency lipss with rotating orientation
on fused silica. Appl. Surf. Sci. 2017, 411, 113–116. [CrossRef]
20. Jiang, L.; Shi, X.S.; Li, X.; Yuan, Y.P.; Wang, C.; Lu, Y.F. Subwavelength ripples adjustment based on electron
dynamics control by using shaped ultrafast laser pulse trains. Opt. Express 2012, 20, 21505–21511. [CrossRef]
[PubMed]
21. Pan, A.; Dias, A.; Gomez-Aranzadi, M.; Olaizola, S.M.; Rodriguez, A. Formation of laser-induced periodic
surface structures on niobium by femtosecond laser irradiation. J. Appl. Phys. 2014, 115, 173101. [CrossRef]
22. Petrović, S.M.; Gaković, B.; Peruško, D.; Stratakis, E.; Bogdanović-Radović, I.; Čekada, M.; Fotakis, C.;
Jelenković, B. Femtosecond laser-induced periodic surface structure on the ti-based nanolayered thin films.
J. Appl. Phys. 2013, 114, 233108. [CrossRef]
23. Garcell, E.M.; Lam, B.; Guo, C. Femtosecond laser-induced herringbone patterns. Appl. Phys. A 2018, 124,
405. [CrossRef]
24. Kotsedi, L.; Nuru, Z.Y.; Mthunzi, P.; Muller, T.F.G.; Eaton, S.M.; Julies, B.; Manikandan, E.; Ramponi, R.;
Maaza, M. Femtosecond laser surface structuring and oxidation of chromium thin coatings: Black chromium.
Appl. Surf. Sci. 2014, 321, 560–565. [CrossRef]
25. Banerjee, S.P.; Fedosejevs, R. Single-shot ablation threshold of chromium using UV femtosecond laser pulses.
Appl. Phys. A 2014, 117, 1473–1478. [CrossRef]
26. Sue, J.A.; Chang, T.P. Friction and wear behavior of titanium nitride, zirconium nitride and chromium nitride
coatings at elevated temperatures. Surf. Coat. Technol. 1995, 76–77, 61–69. [CrossRef]
27. Saghebfar, M.; Tehrani, M.K.; Darbani, S.M.R.; Majd, A.E. Femtosecond pulse laser ablation of chromium:
Experimental results and two-temperature model simulations. Appl. Phys. A 2017, 123, 28. [CrossRef]
28. Öktem, B.; Pavlov, I.; Ilday, S.; Kalaycıoğlu, H.; Rybak, A.; Yavaş, S.; Erdoğan, M.; Ilday, F.Ö. Nonlinear laser
lithography for indefinitely large-area nanostructuring with femtosecond pulses. Nat. Photonics 2013, 7,
897–901. [CrossRef]
29. Huang, M.; Zhao, F.L.; Cheng, Y.; Xu, N.S.; Xu, Z.Z. Origin of laser-induced near-subwavelength ripples:
Interference between surface plasmons and incident laser. ACS Nano 2009, 3, 4062–4070. [CrossRef] [PubMed]
30. Sakabe, S.; Hashida, M.; Tokita, S.; Namba, S.; Okamuro, K. Mechanism for self-formation of periodic grating
structures on a metal surface by a femtosecond laser pulse. Phys. Rev. B 2009, 79, 033409. [CrossRef]
31. Garrelie, F.; Colombier, J.P.; Pigeon, F.; Tonchev, S.; Faure, N.; Bounhalli, M.; Reynaud, S.; Parriaux, O.
Evidence of surface plasmon resonance in ultrafast laser-induced ripples. Opt. Express 2011, 19, 9035–9043.
[CrossRef] [PubMed]
32. Okamuro, K.; Hashida, M.; Miyasaka, Y.; Ikuta, Y.; Tokita, S.; Sakabe, S. Laser fluence dependence of periodic
grating structures formed on metal surfaces under femtosecond laser pulse irradiation. Phys. Rev. B 2010, 82,
165417. [CrossRef]
33. Pham, K.X.; Tanabe, R.; Ito, Y. Laser-induced periodic surface structures formed on the sidewalls of
microholes trepanned by a femtosecond laser. Appl. Phys. A 2012, 112, 485–493. [CrossRef]

© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

19
nanomaterials

Article
Subwavelength Nanostructuring of Gold Films by
Apertureless Scanning Probe Lithography Assisted by
a Femtosecond Fiber Laser Oscillator
Ignacio Falcón Casas ∗ and Wolfgang Kautek ∗
Department of Physical Chemistry, University of Vienna, Währinger Strasse 42, Vienna A-1090, Austria
* Correspondence: [email protected] (I.F.C.); [email protected] (W.K.);
Tel.: +43-664-8175230 (W.K.)

Received: 5 June 2018; Accepted: 11 July 2018; Published: 16 July 2018

Abstract: Optical methods in nanolithography have been traditionally limited by Abbe’s diffraction
limit. One method able to overcome this barrier is apertureless scanning probe lithography assisted
by laser. This technique has demonstrated surface nanostructuring below the diffraction limit.
In this study, we demonstrate how a femtosecond Yb-doped fiber laser oscillator running at high
repetition rate of 46 MHz and a pulse duration of 150 fs can serve as the laser source for near-field
nanolithography. Subwavelength features were generated on the surface of gold films down to a
linewidth of 10 nm. The near-field enhancement in this apertureless scanning probe lithography setup
could be determined experimentally for the first time. Simulations were in good agreement with
the experiments. This result supports near-field tip-enhancement as the major physical mechanisms
responsible for the nanostructuring.

Keywords: near-field; femtosecond laser; nanolithography; subwavelength; tip-enhancement; AFM

1. Introduction
Optical methods in lithography are limited by Abbe’s diffraction limit. Apertureless scanning
probe nanolithography assisted by a laser is a method able to overcome this limitation [1–9].
In this technique, a sharp scanning probe microscope (SPM) tip is placed a few nanometres above
the surface of a substrate. The tip is irradiated by a laser and a strong enhanced field may be
generated in the proximity of the apex of the tip. The evanescent near-field decays exponentially in
both lateral and vertical axes, which leads to a confinement of the electromagnetic field. This may
surpass the modification fluence threshold of the substrate. The combination of tip-enhancement and
confinement can be employed to produce sub-wavelength surface nanostructuring. However, although
near-field enhancement has been predicted in simulations [10–13] and observed experimentally [14–20],
a number of unresolved issues still exist. An example is how much thermal effects are contributing.
On one hand, a laser-irradiated tip can reach a high temperature [21–23], leading to the destruction
or modification of the tip itself. Even if the tip is not modified, heat transfer from the hot tip
to the substratee—either by conduction or radiation—may lead to the melting of the substrate.
Recent research has found huge near-field heat transfer coefficient values, whose origin has not been
clarified yet [24–26]. On the other hand, laser irradiation can produce a fast bending and expansion of
cantilevers due to thermal diffusion, leading to mechanical indentation or scratching of the surface
of the substrate [23]. This problem can be handled by using low spring constant cantilevers [4] or
keeping the tip in non-contact mode [8].
Femtosecond lasers have demonstrated excellent performance in material ablation because the
heat affected zone can be minimized to a few nanometres and can avoid laser–plasma interactions
completely. Continuous-wave lasers have also been employed in scanning probe optical lithography,

Nanomaterials 2018, 8, 536; doi:10.3390/nano8070536 20 www.mdpi.com/journal/nanomaterials


Nanomaterials 2018, 8, 536

but the use of ultrashort pulsed lasers allows for reducing the average power applied to the tip,
while producing very high intensities. Although these high intensities can be beneficial regarding
substrate structuring, laser intensities above the MW/cm2 level might lead to tip damage. Special care
has to be taken with metal-coated tips, as it has been reported in apertureless scanning near-field
microscopy (aSNOM) experiments [27].
Apertureless scanning probe near-field lithography assisted by femtosecond laser has
demonstrated subwavelength surface structuring of metal and polymer films, reaching linewidths
down to 10–15 nm [4,8]. These experiments were performed with Ti:Sa laser sources, at laser
wavelengths λ ≈ 790 nm. Theoretical calculations have shown that, under some conditions, longer laser
wavelengths might lead to higher near-field enhancement [28,29]. A combination of high repetition
rate and low energy per pulse helps to reduce thermomechanical instabilities of the tip-cantilever [30].
In the present study, apertureless scanning probe near-field lithography was employed to write
nanofeatures on the surface of gold films. A home-built femtosecond Ytterbium-doped fiber laser
oscillator based on a novel design [31] was applied. A repetition rate of 46 MHz and a low pulse
energy was chosen to allow a better thermomechanical equilibrium of the cantilever in contrast to laser
sources with kHz repetition rates and higher energy per pulse [30]. The generated subwavelength
grooves exhibited typical linewidths of 40–60 nm, down to 10 nm, thus surpassing the diffraction limit.
The near-field enhancement could be determined experimentally for the first time, and was compared
with simulations.

2. Materials and Methods


We employed a scanning probe microscope NTEGRA (NT-MDT, Moscow, Russia) working in
atomic force microscope (AFM) contact mode to scan the surface of the samples. AFM measurements
were done in ambient conditions. A closed-loop configuration of the scanner stage was enabled
to increase the lateral positioning accuracy. We used silicon probes with a radius of curvature of
10 nm (CSG10, NT-MDT) at the apex of the tip. Contact mode low spring constant cantilevers were
chosen to avoid mechanical deformation of the gold films. The measured resonance frequency of the
cantilever was 27 KHz, 30 μm width and 225 μm length. By using these values, and following the
Sader’s method [32], we obtained a spring constant of the cantilever k = 0.4 N/m. We determined the
load force applied against the sample from the spring constant value of the cantilever. We performed
force spectroscopy on a silicon calibration grating (TGZ1, NT-MDT, Moscow, Russia) in order to obtain
a linear relationship between cantilever deflection (in nA) and cantilever height (in nm). A deflection
setpoint value of 2 nA was chosen when scanning in contact mode, corresponding to a load force of
Fload ≈ 20 nN. This low value of the load force was chosen to avoid the possibility of mechanical surface
modification. The scanning probe microscope data were evaluated with the software Gwyddion.
Gold films were prepared by thermal evaporation. Two films were deposited on mica
(15 and 30 nm thickness) and another on a glass substrate (30 nm thickness). Details of AFM surface
characterization of the samples are shown in Appendix B.
A scheme of the experimental setup is shown in Figure 1. The laser source for the experiments was a
home-built femtosecond Ytterbium-doped fiber laser oscillator, designed according to [31]. The output of
the laser cavity was directed to a pair of diffraction gratings that compensate the dispersion and compress
the laser pulse. The laser beam consisted of laser pulses with temporal length τ = 150 fs, 1 nJ energy per
pulse, central wavelength λ = 1040 nm, 46 MHz repetition rate, and linear polarization. The polarization
was controlled by a half-wave plate. The laser power was adjusted by a combination of another half-wave
plate and a polarizing beamsplitter cube. After passing through a periscope, the laser beam was focused
onto the SPM tip by an achromatic lens (AC254-060-B-ML, Thorlabs, Newton, NJ, USA) with a focal
length f = 60 mm. The laser spot diameter at the focal plane was 50 μm and the angle of incidence θ = 88◦
with respect to the tip axis. Under this condition, the tip with a height of ca. 15 μm was entirely irradiated.
The laser peak intensity was adjusted during the experiments from a minimum of I = 0.15 × 108 W/cm2
to a maximum value of I = 2.90 × 108 W/cm2 .

21
Nanomaterials 2018, 8, 536

Figure 1. A scheme of the experimental setup. The p-polarized laser beam is focused onto an atomic
force microscope (AFM) tip by an achromatic lens. The tip height is about 15 μm, so the laser focal
spot size (50 μm) illuminates the tip entirely. A half-wave plate allows for controlling the laser
polarization angle.

The nanostructuring procedure consisted of three steps. First, an AFM scan of the sample surface
was performed to obtain the topography before laser irradiation. The stage scanned one line along the
fast axis (y-axis) and then moved along the slow axis (x-axis) to the next line position, where the scan
along the fast axis was repeated. In the second step, we scanned the same area and, at a certain position
of the slow scanning axis, the scanning in the slow direction (x-axis) was paused. Then, the laser beam
was engaged and the tip was irradiated, while moving along the fast axis (y-axis). After a number of
laser pulses, the laser beam was turned off and the scan continued. Finally, we scan again the same
area to observe any change on the surface.
Near-field simulations were performed using a MNPBEM17 toolbox, which is based on a
boundary element method (BEM) [33,34]. We introduced the values of the refractive indexes of
silicon at 300 K [35] and 25 nm thick gold films [36]. The simulations were performed taking into
account the retardation of the electromagnetic fields, by solving the full Maxwell equations.

3. Results

3.1. Near-Field Enhancement Simulations


Near-field enhancement simulations were performed for a gold substrate and two silicon tip
geometries, a sphere and a rod (Figure 2). The enhancement factor γ can be defined as γ = |E|/|E0 |,
where E is the induced electric field on the tip-substrate and E0 is the initial laser electric field.
A p-polarized electromagnetic plane wave irradiated the tip at an angle of incidence θ = 88◦ (referred
to the axis normal to the substrate). The wavelength of the incoming laser light was set at λ = 1040 nm,
which corresponds to the central wavelength of our fiber laser. We obtained near-field enhancement
factors of γ ≈ 5–15 for the sphere, depending on the sphere’s radius of curvature and the tip-substrate
distance (Figure 2a). The enhancement factor of the rod was higher than for the sphere and changed
with the length L of the rod. Typical values ranged from γ = 12 (L = 70 nm, r = 10 nm) to γ = 220
(L = 200 nm, r = 10 nm) (Figure 2b). In both cases, the maximum field enhancement was located at the
particle-substrate interspace and decreased exponentially with distance.

22
Nanomaterials 2018, 8, 536

(a) (b)

Figure 2. Near-field enhancement between a gold nanofilm and: (a) a spherical particle (radius = 10 nm);
(b) a rod (r = 10 nm, length = 70 nm) both placed 1 nm above a gold surface. A p-polarized
electromagnetic plane wave irradiates at λ = 1040 nm, with an angle of incidence θ = 88◦ . The white
lines represent the direction and magnitude of the Poynting vector of far-field scattered laser light.

3.2. Scanning Probe Near-Field Nanolithography


In this section, we present the results of the laser irradiation and nanostructuring of gold nanofilms.
In order to identify the irradiation parameters in the far-field which lead to irreversible modifications
of the nanofilms, the threshold fluence as a function of the angle of incidence θ and the number of
pulses N were determined (without the AFM tip engaged).
The influence of the number of pulses N was studied on a 15 nm thick gold film on mica. An AFM
image of the surface of the sample before laser irradiation can be seen on Figure 3a. The gold surface
is composed of nanocrystalline islands with typical sizes of about 50–100 nm, similar to gold film
surfaces reported in [37]. Far-field laser irradiation was performed at an angle of incidence of θ = 86◦
and an intensity of I = 2.4 × 108 W/cm2 . The scanning direction was set from left to right, at a scanning
speed of 1 μm/s. A substantial morphological modification was observed after an irradiation time
of 90 seconds, corresponding to a number of pulses N = 4.2×109 (Figure 3b). There, the gold film is
strongly deformed and forms bumps that elevate to heights up to 100 nm. This experiment served to
identify morphological surface changes produced by far-field irradiation.

23
Nanomaterials 2018, 8, 536

(a) (b)

Figure 3. Morphology changes produced by the far-field laser. Gold on mica (15 nm thickness)
(a) before laser irradiation; (b) after a laser irradiation time of 90 seconds, corresponding to
N = 4.2 × 109 . Angle of incidence θ = 86◦ , I = 2.4 × 108 W/cm2 . Scan area 1 × 1 μm2 .

To analyze the influence of the angle of incidence, a similar experiment was conducted on a 30 nm
thick gold film on glass at θ = 80◦ . Far-field surface modifications similar to Figure 3b were observed
(not shown here), even for a lower number of pulses N = 0.19 × 109 . Therefore, we inferred that the
angle of incidence has a drastic effect on far-field surface modifications, more significant than the
number of pulses. In order to produce near-field nanolithography, far-field surface modifications need
to be avoided. Therefore, the laser intensity was decreased by changing the angle of incidence to
θ = 88◦ . At this angle of incidence, no far-field surface modifications were observed, even for a high
number of pulses N = 4.2 × 109 . Figure 4a shows the surface of the sample (30 nm thick gold film
on glass) before laser irradiation. Similar gold grain patterns can be observed before and after laser
irradiation (e.g., green boxes in Figure 4a,b). This indicates that the far-field laser did not affect the
surface morphology and only the areas below the tip were modified during irradiation.

(a) (b)

Figure 4. Nanolithography dependence on the number of scans at tip illumination. Gold on glass
(30 nm thickness) (a) before laser irradiation and (b) after laser irradiation at three lines (slow axis
scanning from left to right stopped). Number of scans: 7 (line 1), 10 (line 2) and 14 (line 3),
scanning speed 0.62 μm/s, angle of incidence θ = 88◦ , I = 2.4 × 108 W/cm2 .

Figure 4b shows the effect produced by repeated tip passes on the same line under laser
illumination. The tip was irradiated at three lines (slow axis scanning stopped) during 30, 60 and 90 s
(N = 1.2, 2.4 and 4.2 ×109 , respectively) (from left to right), corresponding to a number of tip passes on

24
Nanomaterials 2018, 8, 536

each line of 7, 10 and 14 times, respectively. Three vertical lines were structured on the gold surface
(Figure 4b). The linewidth and depth both increased with the number of passes (Table 1).

Table 1. Dependence of width (FWHM – Full Width at Half Maximum) and depth of lines on the
number of passes (in Figure 4b, from left to right). Data are averaged from 256 profile lines.

Line 1 Line 2 Line 3


Number of passes 7 10 14
Width (nm) 36 45 76
Depth (nm) 0.4 0.5 1.0

The dependence on the laser intensity was investigated on a 30 nm thick gold film on mica
(Figure 5a). The tip was irradiated at four vertical lines (slow axis scanning from left to right stopped)
at increasing laser intensities I = 0.2, 0.3, 0.4, 1.0 × 108 W/cm2 (from left to right). The scanning speed
was set at 0.38 μm/s. Laser intensities below 0.4 ×108 W/cm2 produced a very small effect. The depth
of lines increased with the laser intensity. A line with an averaged FWHM width of 10 nm (Figure 5b)
was produced at I = 1.0 × 108 W/cm2 . The line profile was obtained by averaging the green rectangle
area marked in Figure 5a.

(a) (b)

Figure 5. Nanolithography depending on laser intensity. Gold on mica (30 nm thickness) (a) after
laser irradiation at four vertical lines (slow axis scanning from left to right stopped). I = 0.2, 0.3, 0.4,
1.0 × 108 W/cm2 , angle of incidence θ = 88◦ , scanning speed 0.38 μm/s; (b) averaged line profile of
the area marked in Figure 5a. The FWHM width of the line is 10 nm. The image was obtained in AFM
contact error mode, where a constant force is applied to the tip and the variations of the cantilever’s
deflection are recorded.

Laser intensity dependence was also studied on a 30 nm thick gold film on glass. Figure 6a
shows five vertical lines irradiated at increasing laser power I = 0.7, 1.0, 2.0, 2.7 and 2.9 × 108 W/cm2
(from left to right) and scanning speed of 0.38 μm/s. The profile line was averaged (taking the full
area in Figure 6a), to reduce the effect of the gold roughness. The FWHM width and depth of lines
(Figure 6b) increased with the laser intensity (Table 2). The writing performance at lines 1, 4 and 5 was
affected by 1–3 nm height irregularities of the gold surface.

25
Nanomaterials 2018, 8, 536

(a) (b)

Figure 6. Nanolithography depending on the laser intensity. Gold on glass (30 nm thickness (a)
after laser irradiation at five lines (slow axis scanning from left to right stopped). Laser intensity
I = 0.7 (line 1), 1.0 (line 2), 2.0 (line 3), 2.7 (line 4) and 2.9 × 108 W/cm2 , scanning speed 0.38 μm/s,
angle of incidence θ = 88◦ ; (b) line profile obtained after taking an average of 256 horizontal profile
lines. Widths and depths are indicated in Table 2.

Table 2. Laser intensity effect on width (FWHM) and depth of lines in Figure 6b (from left to right).
Data are averaged from 256 profile lines.

Line Number 1 2 3 4 5
I (×108 W/cm2 ) 0.7 1.0 2.0 2.7 2.9
Width (nm) 71 47 41 60 52
Depth (nm) 0.48 0.52 0.70 0.74 0.77

A summary of the laser far-field parameters is provided in Table 3.

Table 3. Far-field laser irradiation parameters for gold film on mica (15 nm thick) and on glass
(30 nm thick): angle of incidence θ, laser peak intensity I, pulse fluence F, number of pulses N.

θ I (×108 W/cm2 ) F (μJ/cm2 ) N (×109 )


80◦ 2.4 37 0.2
86◦ 2.4 37 1.4–4.2
88◦ 0.15–2.9 2–44 1.4–4.2

4. Discussion
Nanostructures were generated on the surface of gold films by using apertureless scanning
near-field lithography. The structured lines formed on the gold films show typical linewidths of
40–70 nm and depths of 0.4–1.0 nm. The width of lines increased with the number of repeated tip
scans and laser intensity. The smallest linewidth measured was 10 nm. The surface roughness of the
gold samples affected the structuring performance (for instance, fourth and fifth lines in Figure 5).
A threshold for surface modification at I ≈ 0.4 × 108 W/cm2 (Figure 5a) was observed at a scanning
speed of 0.38 μm/s. The angle of incidence (i.e., the laser intensity absorbed by the gold film) has a
drastic effect on the far-field irradiation, as it can be seen in Figure 3b. At high angles of incidence
(near 90◦ ), the contrast between the near-field fluence near the tip and the far-field fluence on the
illuminated substrate is sufficiently high so only nanostructures are generated without modification of
the substrate (Figure 4b).
A comparatively low threshold fluence of Fth ≈ 6 μJ/cm2 for surface modification on gold was
observed. This is in contrast to experiments with moderate repetition rates (1 kHz, Fth = 12 mJ/cm2 ) [4],

26
Nanomaterials 2018, 8, 536

with single pulses (Fth = 2 mJ/cm2 ) [38,39] and calculations with single pulses (Fth = 4.5 mJ/cm2 ) [40].
Near-field surface modification was also achieved at high repetition rates but lower energy per pulse
(e.g., 80 MHz, 1 nJ) [8]. This seems to indicate that the repetition rate has a strong influence on the
modification of the gold surface. The so-called cool-ablation [41] has been observed at very high repetition
rates of GHz using bursts of laser pulses. Threshold fluences can be reduced when the repetition rate is
increased above 1–10 MHz [41,42].
Two main mechanisms have been proposed in the context of apertureless scanning probe near-field
lithography assisted by laser: near-field tip-enhancement and thermomechanical effects. Simulations
in Figure 2 show the presence of an enhanced electromagnetic field between the tip and the substrate.
The major question is whether this enhancement factor is high enough to raise the laser field above the
modification threshold of the substrate. An experimental estimation of the near-field enhancement,
based on the observation of far-field morphological surface changes of the gold films, was undertaken.
In the experiments shown in the Figure 3, we observed evidence of complete melting of the gold film
after a high number of pulses N = 4.2 × 109 (Figure 3b). Based on this observation, the enhancement
factor γ was determined by varying the angle of incidence on the 30 nm thick gold films. At an angle
of incidence θ = 80◦ , the surface morphology was modified (not shown here), similarly to Figure 3b.
Accordingly, the far-field fluence was above the modification threshold. An increase to an angle of
incidence θ = 88◦ resulted in a reduction of the absorbed laser energy by a factor of I88◦ /I80◦ ≈ 0.02
(see Appendix A). Under these conditions, one observes near-field modification without far-field
modification (see Figure 4). In a first approximation, one could conclude that the enhancement is at
least I80◦ /I88◦ ≈ 50.
This result is now compared with a simulated enhancement factor. For the spherical tip shown in
Figure 2a, the electric field enhancement is γ ≈ 7. The intensity enhancement is γ2 = |E|2 /|E0 |2 = 49,
in good agreement with the intensity threshold obtained experimentally (I80◦ /I88◦ ≈ 50).

5. Conclusions
Nanostructuring of gold films by apertureless scanning probe near-field lithography using a novel
Ytterbium-doped fiber oscillator as the laser source was demonstrated. Lines written on gold show
depths of 0.5–1.0 nm and typical lateral sizes of 40–70 nm, down to 10 nm. A near-field enhancement
factor was determined experimentally and compared with simulations. A good agreement with the
experimental results supports a mechanism based mainly on near-field enhancement. This approach
provides direct access to the study of femtosecond laser-matter interaction in the near-field nano-regime.
The compact setup, the high repetition rate, the possibility of working in ambient conditions and
comparatively reduced cost make this an appealing technique for sub-100 nm lithography.

Author Contributions: Conceptualization, I.F.C and W.K.; Formal analysis, I.F.C; Investigation, I.F.C; Resources,
W.K.; Software, I.F.C; Supervision, W.K.; Visualization, I.F.C; Writing – original draft, I.F.C; Writing – review and
editing, I.F.C and W.K.

Funding: This research received no external funding

Acknowledgments: Open access funding was provided by the University of Vienna. We thank Stefan Hummel
for the preparation of one of the gold films.
Conflicts of Interest: The authors declare no conflict of interest.

Appendix A. Laser Absorbance Change versus Angle of Incidence for a Gold Nanofilm
The laser intensity threshold for surface modification was estimated by calculating the ratio
between the absorbed laser intensity at two angles of incidence: θ = 80◦ (far-field laser irradiation
produced morphology changes of the gold surface) and 88◦ (no far-field changes, only near-field when
the tip was engaged). The two differences considered are the variation of both intensity and absorption
with the incident angle. The first one can be calculated through the relationship

27
Nanomaterials 2018, 8, 536

I (θ ) = I0 cos2 (θ ), (A1)

obtaining

I (80◦ )
≈ 25. (A2)
I (88◦ )
Then, the change in the absorbed laser intensity by the gold surface at the two different angles of
incidence can be calculated. By using the refractive index at λ = 1040 nm for 25 nm thick gold films [36]
and the Fresnel equations for p-polarization, one can calculate the transmittance T

n2 cosθi − n1 cosθt
rp = , (A3)
n1 cosθt + n2 cosθi

T = 1 − R = 1 − |r p |2 , (A4)

where n1 , n2 are the refractive indexes of air and gold, θi , θt are the incident and transmission angles,
r p is the Fresnel reflection coefficient for p-polarization and R is the reflectivity. For simplicity,
we assume that the transmitted laser light is completely absorbed by the gold substrate.
This assumption is a good approximation because the penetration depth of light in gold ( δ ≈ 12 nm at
normal incidence ) is less than the gold film thickness. The transmittance values obtained at the angles
of incidence considered are

T (80◦ )
≈ 2. (A5)
T (88◦ )
A total absorbed intensity ratio is obtained by the multiplication of the ratios of
Equations (A2) and (A5). This means that, when the angle of incidence is changed from θ = 80◦
to θ = 88◦ , the absorbed laser intensity is reduced by a factor of 50.

Appendix B. Morphology Characterization of the Surface of Au Films


The characterization of the topography of the Au films by AFM is presented here. Constant force
and deflection error modes were recorded simultaneously by the AFM software (Nova, NT-MDT).
Posterior analysis and calculation of average roughness values Ra were performed with the software
Gwyddion. The average roughness Ra is defined as
n
1
Ra =
n ∑ | h i |, (A6)
i =1

where h is the height of each pixel and n is the number of pixels. In addition, 256 × 256 pixels (full
images) were included for the calculations of R a .

Figure A1. Topography of the 15 nm thickness Au film on mica imaged in AFM constant force mode;
average surface roughness Ra = 4.31 nm, scan size 3 × 3 μm2 .

28
Nanomaterials 2018, 8, 536

(a) (b)

(c) (d)

Figure A2. Topography of the 30 nm thickness Au film on mica imaged in AFM constant force
mode (a,c) and error deflection mode (b,d). Average surface roughness Ra = 1.71 nm (a), 20.84 pA (b),
1.75 nm (c) and 7.55 pA (d), scan size 5 × 5 μm2 (a,b) and 2 × 2 μm2 (c,d).

(a) (b)

(c) (d)

Figure A3. Topography of the 30 nm thickness Au film on glass imaged in AFM constant force mode
(a,c) and error deflection mode (b,d). Average surface roughness Ra = 1.12 nm (a), 18.88 pA (b),
1.53 nm (c) and 21.45 pA (d), scan size 5 × 5 μm2 (a,b) and 2 × 2 μm2 (c,d).

29
Nanomaterials 2018, 8, 536

References
1. Jersch, J.; Demming, F.; Dickmann, K. Nanostructuring with laser radiation in the nearfield of a tip from
a scanning force microscope. Appl. Phys. A 1996, 64, 29–32. [CrossRef]
2. Dickmann, K.; Jersch, J.; Demming, F. Focusing of Laser Radiation in the Near-field of a Tip (FOLANT) for
Applications in Nanostructuring. Surf. Interface Anal. 1997, 25, 500–504. [CrossRef]
3. Yin, X.; Fang, N.; Zhang, X.; Martini, I.B.; Schwartz, B.J. Near-field two-photon nanolithography using an
apertureless optical probe. Appl. Phys. Lett. 2002, 81, 3663–3665. [CrossRef]
4. Chimmalgi, A.; Choi, T.Y.; Grigoropoulos, C.P.; Komvopoulos, K. Femtosecond laser aperturless near-field
nanomachining of metals assisted by scanning probe microscopy. Appl. Phys. Lett. 2003, 82, 1146–1148.
[CrossRef]
5. H’dhili, F.; Bachelot, R.; Rumyantseva, A.; Lerondel, G.; Royer, P. Nano-patterning photosensitive polymers
using local field enhancement at the end of apertureless SNOM tips. J. Microsc. 2003, 209, 214–222. [CrossRef]
[PubMed]
6. Kirsanov, A.; Kiselev, A.; Stepanov, A.; Polushkin, N. Femtosecond laser-induced nanofabrication in the
near-field of atomic force microscope tip. J. Appl. Phys. 2003, 94, 6822–6826. [CrossRef]
7. Haefliger, D.; Stemmer, A. Writing subwavelength-sized structures into aluminium films by thermo-chemical
apertureless near-field optical microscopy. Ultramicroscopy 2004, 100, 457–464. [CrossRef] [PubMed]
8. Milner, A.A.; Zhang, K.; Prior, Y. Floating Tip Nanolithography. Nano Lett. 2008, 8, 2017–2022. [CrossRef]
[PubMed]
9. Falcón Casas, I.; Kautek, W. In Laser Micro-Nano-Manufacturing and 3D Microprinting; Chapter Apertureless
Scanning Near-Field Optical Lithography; Hu, A., Ed.; Springer: New York, NY, USA, 2018; in print.
10. Novotny, L.; Sánchez, E.J.; Xie, X.S. Near-field optical imaging using metal tips illuminated by higher-order
Hermite–Gaussian beams. Ultramicroscopy 1998, 71, 21–29. [CrossRef]
11. Martin, Y.C.; Hamann, H.F.; Wickramasinghe, H.K. Strength of the electric field in apertureless near-field
optical microscopy. J. Appl. Phys. 2001, 89, 5774–577. [CrossRef]
12. Esteban, R.; Vogelgesang, R.; Kern, K. Simulation of optical near and far fields of dielectric apertureless
scanning probes. Nanotechnology 2006, 17, 475–482. [CrossRef]
13. Esteban, R.; Vogelgesang, R.; Kern, K. Tip-substrate interaction in optical near-field microscopy. Phys. Rev. B
2007, 75, 195410. [CrossRef]
14. Huang, S.M.; Hong, M.H.; Lu, Y.F.; Luk’yanchuk, B.S.; Song, W.D.; Chong, T.C. Pulsed-laser assisted
nanopatterning of metallic layers combined with atomic force microscopy. J. Appl. Phys. 2002, 91, 3268–3274.
[CrossRef]
15. Bouhelier, A.; Beversluis, M.; Novotny, L. Applications of field-enhanced near-field optical microscopy.
Ultramicroscopy 2004, 100, 413–419. [CrossRef] [PubMed]
16. Roth, R.M.; Panoiu, N.C.; Adams, M.M.; Osgood, R.M.; Neacsu, C.C.; Raschke, M.B. Resonant-plasmon field
enhancement from asymmetrically illuminated conical metallic-probe tips. Opt. Express 2006, 14, 2921–2931.
[CrossRef] [PubMed]
17. Hartschuh, A. Tip-Enhanced Near-Field Optical Microscopy. Angew. Chem. Int. Ed. 2008, 47, 8178–8191.
[CrossRef] [PubMed]
18. Huth, F.; Chuvilin, A.; Schnell, M.; Amenabar, I.; Krutokhvostov, R.; Lopatin, S.; Hillenbrand, R. Resonant
Antenna Probes for Tip-Enhanced Infrared Near-Field Microscopy. Nano Lett. 2013, 13, 1065–1072. [CrossRef]
[PubMed]
19. Wang, L.; Xu, X.G. Scattering-type scanning near-field optical microscopy with reconstruction of vertical
interaction. Appl. Phys. Lett. 2015, 6, 8973. [CrossRef] [PubMed]
20. Møller, S.H.; Vester-Petersen, J.; Nazir, A.; Eriksen, E.H.; Julsgaard, B.; Madsen, S.P.; Balling, P. Near-field
marking of gold nanostars by ultrashort pulsed laser irradiation: Experiment and simulations. Appl. Phys. A
2018, 124, 210. [CrossRef]
21. Gerstner, V.; Thon, A.; Pfeiffer, W. Thermal effects in pulsed laser assisted scanning tunneling microscopy.
J. Appl. Phys. 2000, 87, 2574–2580. [CrossRef]
22. Milner, A.A.; Zhang, K.; Garmider, V.; Prior, Y. Heating of an Atomic Force Microscope tip by femtosecond
laser pulses. Appl. Phys. A 2010, 99, 1–8. [CrossRef]

30
Nanomaterials 2018, 8, 536

23. Huber, C.; Prior, Y.; Wolfgang, K. Laser-induced cantilever behaviour in apertureless scanning near-field
optical microscopes. Meas. Sci. Technol. 2014, 25, 075604. [CrossRef]
24. Kim, K.; Song, B.; Fernández-Hurtado, V.; Lee, W.; Jeong, W.; Cui, L.; Thompson, D.; Feist, J.; Reid, M.T.H.;
García-Vidal, F.J.; et al. Radiative heat transfer in the extreme near field. Nature 2015, 528, 387–391. [CrossRef]
[PubMed]
25. Song, B.; Ganjeh, Y.; Sadat, S.; Thompson, D.; Fiorino, A.; Fernández Hurtado, V.; Feist, J.; García Vidal, F.J.;
Cuevas, J.C.; Reddy, P.; et al. Enhancement of near-field radiative heat transfer using polar dielectric thin
films. Nat. Nanotechnol. 2015, 10, 253–258. [CrossRef] [PubMed]
26. Kloppstech, K.; Könne, N.; Biehs, S.A.; Rodriguez, A.W.; Worbes, L.; Hellmann, D.; Kittel, A. Giant heat
transfer in the crossover regime between conduction and radiation. Nat. Commun. 2017, 8. [CrossRef]
[PubMed]
27. Dutoit, B.; Zeisel, D.; Deckert, V.; Zenobi, R. Laser-Induced Ablation through Nanometer-Sized Tip Apertures:
Mechanistic Aspects. J. Phys. Chem. B 1997, 101, 6955–6959. [CrossRef]
28. Zhang, W.; Cui, X.; Martin, O.J.F. Local field enhancement of an infinite conical metal tip illuminated by
a focused beam. J. Raman Spectrosc. 2009, 40, 1338–1342. [CrossRef]
29. Mihaljevic, J.; Hafner, C.; Meixner, A.J. Simulation of a metallic SNOM tip illuminated by a parabolic mirror.
Opt. Express 2013, 21, 25926–25943. [CrossRef] [PubMed]
30. Huber, C.; Trügler, A.; Hohenester, U.; Prior, Y.; Kautek, W. Optical near-field excitation at commercial
scanning probe microscopy tips: a theoretical and experimental investigation. Phys. Chem. Chem. Phys. 2014,
16, 2289–2296. [CrossRef] [PubMed]
31. Verhoef, A.; Zhu, L.; Israelsen, S.M.; Grüner-Nielsen, L.; Unterhuber, A.; Kautek, W.; Rottwitt, K.; Baltuška, A.;
Fernández, A. Sub-100 fs pulses from an all-polarization maintaining Yb-fiber oscillator with an anomalous
dispersion higher-order-mode fiber. Opt. Express 2015, 23, 26139–26145. [CrossRef] [PubMed]
32. Sader, J.E.; Borgani, R.; Gibson, C.T.; Haviland, D.B.; Higgins, M.J.; Kilpatrick, J.I.; Lu, J.; Mulvaney, P.; Shearer, C.J.;
Slattery, A.D.; et al. A virtual instrument to standardise the calibration of atomic force microscope cantilevers.
Rev. Sci. Instrum. 2016, 87, 093711. [CrossRef] [PubMed]
33. Hohenester, U.; Trügler, A. MNPBEM—A Matlab toolbox for the simulation of plasmonic nanoparticles.
Comput. Phys. Commun. 2012, 183, 370–381. [CrossRef]
34. Available online: http://physik.uni-graz.at/mnpbem/ (accessed on 3 March 2018).
35. Green, M.A. Self-consistent optical parameters of intrinsic silicon at 300K including temperature coefficients.
Sol. Energy Mater. Sol. Cells 2008, 92, 1305–1310. [CrossRef]
36. Yakubovsky, D.I.; Arsenin, A.V.; Stebunov, Y.V.; Fedyanin, D.Y.; Volkov, V.S. Optical constants and structural
properties of thin gold films. Opt. Express 2017, 25, 25574–25587. [CrossRef] [PubMed]
37. Karakouz, T.; Holder, D.; Goomanovsky, M.; Vaskevich, A.; Rubinstein, I. Morphology and Refractive Index
Sensitivity of Gold Island Films. Chem. Mater. 2009, 21, 5875–5885. [CrossRef]
38. Link, S.; Burda, C.; Nikoobakht, B.; El-Sayed, M.A. Laser-Induced Shape Changes of Colloidal Gold
Nanorods Using Femtosecond and Nanosecond Laser Pulses. J. Phys. Chem. B 2000, 104, 6152–6163.
[CrossRef]
39. González-Rubio, G.; Guerrero-Martínez, A.; Liz-Marzán, L.M. Reshaping, Fragmentation, and Assembly of
Gold Nanoparticles Assisted by Pulse Lasers. Acc. Chem. Res. 2016, 49, 678–686. [CrossRef] [PubMed]
40. Lin, Z.; Leveugle, E.; Bringa, E.M.; Zhigilei, L.V. Molecular Dynamics Simulation of Laser Melting of
Nanocrystalline Au. J. Phys. Chem. C 2010, 114, 5686–5699. [CrossRef]
41. Kerse, C.; Kalaycıoğlu, H.; Elahi, P.; Çetin, B.; Kesim, D.K.; Akçaalan, Ö.; Yavaş, S.; Aşık, M.D.; Öktem, B.;
Hoogland, H.; et al. Ablation-cooled material removal with ultrafast bursts of pulses. Nature 2016, 537, 84–88.
[CrossRef] [PubMed]
42. Finger, J.; Reininghaus, M. Effect of pulse to pulse interactions on ultra-short pulse laser drilling of steel
with repetition rates up to 10 MHz. Opt. Express 2014, 22, 18790–18799. [CrossRef] [PubMed]

c 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

31
nanomaterials

Article
Deformation Behavior of Foam Laser Targets
Fabricated by Two-Photon Polymerization
Ying Liu 1 , John H. Campbell 2 , Ori Stein 3 , Lijia Jiang 1 , Jared Hund 3 and Yongfeng Lu 1, *
1 Department of Electrical and Computer Engineering, University of Nebraska-Lincoln, Lincoln,
NE 68588-0511, USA; [email protected] (Y.L.); [email protected] (L.J.)
2 Material Science Solutions, 2136 Westbrook Lane, Livermore, CA 94550, USA; [email protected]
3 Schafer Livermore Lab, 303 Lindbergh Avenue, Livermore, CA 94551, USA; [email protected] (O.S.);
[email protected] (J.H.)
* Correspondence: [email protected]; Tel.: +402-472-8323

Received: 31 May 2018; Accepted: 3 July 2018; Published: 6 July 2018

Abstract: Two-photon polymerization (2PP), which is a three-dimensional micro/nano-scale additive


manufacturing process, is used to fabricate component for small custom experimental packages
(“targets”) to support laser-driven, high-energy-density physics research. Of particular interest is the
use of 2PP to deterministically print millimeter-scale, low-density, and low atomic number (CHO)
polymer matrices (“foams”). Deformation during development and drying of the foam structures
remains a challenge when using certain commercial acrylic photo-resins. Acrylic resins were chosen
in order to meet the low atomic number requirement for the foam; that requirement precludes the use
of low-shrinkage organic/inorganic hybrid resins. Here, we compare the use of acrylic resins IP-S
and IP-Dip. Infrared and Raman spectroscopy are used to quantify the extent of the polymerization
during 2PP vs. UV curing. The mechanical strength of beam and foam structures is examined,
particularly the degree of deformation that occurs during the development and drying processes.
The magnitude of the shrinkage is quantified, and finite element analysis is used in order to simulate
the resulting deformation. Capillary drying forces during development are shown to be small and
are likely below the elastic limit of the foam log-pile structures. In contrast, the substantial shrinkage
in IP-Dip (~5–10%) causes large shear stresses and associated plastic deformation, particularly near
constrained boundaries and locations with sharp density transitions. Use of IP-S with an improved
writing procedure results in a marked reduction in deformation with a minor loss of resolution.

Keywords: two-photon polymerization; low-density foam structures; laser targets; structure


deformation; acrylate resin; Raman microspectroscopy

1. Introduction
Two-photon polymerization (2PP) is a direct-write technology that has recently been used to create
millimeter-scale laser target components to support the Department of Energy’s (DOE) High Energy
Density (HED) research programs [1–4]. In the first published work in this area, Bernat et al. [5] and
Jiang et al. [6] report the use of 2PP to print simulated fill tubes and low-density foam-like structures,
respectively. More recently, Jiang et al. [7,8], Stein et al. [9], and Oakdale et al. [10] discuss details of
the design, fabrication, characterization, and assembly of low-density foam targets.
Details of the 2PP process and technology have been reviewed recently [11]. In brief,
polymerization is initiated by the simultaneous absorption of two photons by a photoinitiator in
a reactive monomer/oligomer resin, and it thus depends on the square of the laser irradiance.
In practice, an initiator is selected that has negligible absorption at the incident fundamental laser
frequency but measurable two-photon absorption at the second harmonic. Because two-photon
absorption cross-sections are very low, the probability of reaction initiation is negligible except near

Nanomaterials 2018, 8, 498; doi:10.3390/nano8070498 32 www.mdpi.com/journal/nanomaterials


Nanomaterials 2018, 8, 498

the laser focus. Therefore, photopolymerization only occurs at the peak of the focal irradiance and
generates a volumetric polymer dot (“voxel”) that is generally smaller than the diffraction-limited
spot size.
Typically, voxels range from 200–400 nm for fabricated structures that are similar to the ones
reported here [6–8]. Structures are created by moving the repetition (rep)-rated laser beam (kHz to
MHz) through the resin, thus generating overlapping voxels that, with proper scanning control, are
built into the computer-aided design (CAD) three-dimensional (3D) shapes. Any unreacted resin is
later removed during a post-writing development process, leaving behind a polymeric replica of the
CAD design. Figure 1 shows examples of 2PP log-pile foam structures generated in our laboratory for
laser target fabrication applications.
Structures described in this paper were made using one of two commercial resins: IP-Dip or IP-S.
The IP-series are a family of proprietary acrylic resins plus initiator marketed by Nanoscribe GmbH for
use with their commercial 2PP writing system. Using commercial resins is appealing, as it eliminates
the need for custom formulation. In addition, acrylics are acceptable target materials as the atomic
composition is largely carbon and hydrogen with minor amounts of oxygen. Many HED physics
experiments, using foam or other polymer structures, require materials comprised of low atomic
number elements because of the well-known variation in X-ray absorption (opacity) with atomic
numbers [12]. Thus, the use of other common resins, such as hybrid organic/inorganic resins, for
example, Ormocers, SZ2080 [13] or thiol-based resins [14], is not acceptable. Acrylic based low-density
components and targets produced by 2PP are now being shot at major HED national research centers in
the USA, including Lawrence Livermore National Laboratory (LLNL), Laboratory for Laser Energetics
(LLE), Naval Research Lab (NRL), etc.

Figure 1. Examples of (a) foam plate and (b) a rod of two-photon polymerization (2PP) fabricated
log-pile structures for laser target applications. The foam plate: 1.5 × 1.5 × 0.10 mm3 with
a 4 × 4 × 2 μm3 beam lattice structure (density ~0.2 g/cm3 ). The foam rod: 2.0 × 0.25 × 0.35 mm3
with a 6.2 × 6.2 × 1.0 μm3 beam lattice structure (density ~0.1 g/cm3 ).

Acrylic resins, like IP-S and IP-Dip, are commonly used for 2PP fabrication of nano/microstructures.
However, they are often limited by various materials and processing issues:

• inherent strength of the polymerized resin [14,15];


• strong sensitivity to writing conditions (peak irradiance and shots per site, i.e., dose) [16,17];
• low photo-conversion of resin to polymer [18–20];
• shrinkage during photopolymerization or development or both [6,9,20];
• stresses due to capillary forces during drying [6,9,21]; and,
• control of adhesion to the substrate [6,9].

33
Nanomaterials 2018, 8, 498

Consequently, the user must consider each of these issues when selecting a given resin, the writing
conditions, and the development method for a particular application. In this paper, we examine these
issues by a combination of experiments and modeling, and suggest some possible methods for
controlling the negative impacts of each in IP-Dip and IP-S resins. IP-Dip is particularly problematic.
In contrast, IP-S shows significant improvement over IP-Dip, while still maintaining the required high
CH content for target applications. The trade-off is in the line resolution, which is better in IP-Dip than
in IP-S.
Shrinkage and deformation problems with the use of IP-Dip have been shown to stem largely
from the low resin-to-polymer conversion (<50%) and an associated low modulus and yield strength of
the polymer. Infrared and Raman spectroscopy are used to quantify the resin-to-polymer conversion
by following the signature alkene vibrational bands. Finite element analysis (FEA) is used to simulate
the degree of shrinkage and the resulting plastic strains that occur during the development and drying
of 2PP log-pile foam structures.

2. Experimental

2.1. Photo-Resins and Properties


IP-Dip and IP-S commercial negative-tone, acrylate-based photoresists were used in this work
(Nanoscribe GmbH, Eggenstein-Leopoldshafen, Germany), as they meet the low atomic number
requirement for the proposed HED application. These resins are designed for use with Nanoscribe’s
Dip-in Laser Lithography (DiLL) technology and they serve as both the immersion and photosensitive
material. Specifically, the resins match the refractive index of the final focusing lens and achieve the
highest numerical aperture (i.e., the best resolution) at a given magnification.
IP-Dip has a low viscosity and is recommended for use in high-resolution applications requiring
narrow line width. In contrast, IP-S is more viscous and designed for mesoscale printing at larger
line widths. Elemental compositions and key properties of the resins are summarized in Table 1.
The elemental compositions were determined using the procedure as reported in [6]. The resins
are predominately CHx with small amounts of O and traces of N thus satisfying the low atomic
number requirement.

Table 1. Composition and key properties of IP-Dip and IP-S resins. Unless otherwise noted, the
physical and the mechanical properties are from Nanoscribe GmbH.

Elemental Analysis of Resins


Carbon Hydrogen Empirical
Resin Nitrogen (at.%) Oxygen (at.%)
(at.%) (at.%) Formula
IP-Dip 40.2 46 0.04 13.7 CH2 N0.001 O0.34
IP-S 31.5 54.1 5.8 11.8 CH1.72 N0.086 O0.37
Physical and Mechanical Properties
Density (liq) Density Young’s Modulus Hardness Poisson’s
Resin Refractive Index
(g/cm3 ) (s)(g/cm3 ) * (GPa) (MPa) Ratio ***
IP-Dip 1.14–1.19 1.2 0.75–2.5 **, 4.5 152 0.35 1.52
IP-S 1.16–1.19 1.2 4.6 160 0.35 1.48
* [6], ** [16], *** [22].

2.2. 2PP Microfabrication


Micro-fabrications were carried out using a Photonic Professional GT system (Nanoscribe
GmbH [19]). Two-photon excitation was accomplished using the 780 nm frequency-doubled output
from an Er-fiber laser (1580 nm, TEM00 , M2 < 1.2, Toptica Photonics AG, Germany) operating at 80 MHz
with a temporal pulse length ~100 fs. An integrated set of beam transport optics directed the laser
output with circular polarization to a final focusing objective that dipped directly into the photoresist.

34
Nanomaterials 2018, 8, 498

In the case of the IP-Dip resin, a 63X objective with a 1.4 numerical aperture was used for printing;
whereas, with IP-S, the objective was 25X with a numerical aperture of 0.8. Table 2 provides a summary
of the 2PP writing conditions used to fabricate the structures reported here.

Table 2. Top level summary of typical 2PP writing conditions used to prepare low-density structures
reported in this work.

Parameter Units IP-DIP IP-S


Final focusing power 63X 25X
Numerical aperture (NA) 1.4 0.8
Refractive index 1.52 1.48
Wavelength μm 0.78 0.78
Beam waist (calculated) μm 0.27 0.46
Focal spot area (calculated) μm2 0.23 0.66
Pulse energy nJ 0.19 0.21
Pulse length fs 100 100
Pulse peak power kW 1.9 2.1
Peak irradiance kW/μm2 8.2 3.2
Pulse repetition rate MHz 80 80
Average power mW 15 17
Scan speed μm/s 10,000 10,000
Line width (at 1cm/s scan) μm 0.4 0.65
Shots/micron scanned ~8000 ~8000

The average incident laser power was measured by a photodiode that was located at the input
to the focusing objective. The passive losses in beam propagation to the sample plane were assumed
to be constant and accounted for in Nanoscribe’s as-built system calibration. The laser output power
was controlled by an acousto-optic modulator that can be adjusted over a range of approximately 0 to
50 mW (average power). The beam diameter at focus, Db , was calculated by:

2nλ
Db = , (1)
πN A
where NA is the numerical aperture, n is the refractive index and λ is the wavelength.
Structures were created using the Nanoscribe built-in software package, DESCRIBE™ 2.5
(Nanoscribe GmbH, Germany), which generates General Writing Language (GWL) files directly.
The Photonic Professional GT system (Nanoscribe GmbH, Germany) uses both a piezo stage and
two coupled galvanic mirrors to write the structure. The galvanic mirrors allow for rapid x-y scanning
at up to 10–20 mm/s over an area 200 μm in diameter when using the 63X final focusing objective or
400 μm in diameter when using the 25X objective lens. The vertical (z) motion is controlled by the
piezo stage and the built-in z-drive of the focusing objective, which ranges up to several mm in height.
The system can print structures with an area of up to 25 × 25 mm2 by using the motorized stage and
“stitching” the structures together. The largest dimension of the structures that were fabricated in this
application was 2 mm. The stitching accuracy is typically 1–4 microns [6].
The photoresist was deposited as a drop on a 25 × 25 × 0.7 mm3 glass substrate that was mounted
in an aluminum sample tray. The tray, with substrate and resist, was inserted into the Nanoscribe GT
housing that is attached to a precision piezoelectric-driven stage. All of the operations were carried
out under yellow room lighting to avoid polymerization by single-photon absorption.

2.3. Structure Development, Drying, and Characterization


After exposure, the sample substrate with IP resin was removed from the holder and was
developed at room temperature for 1 h in 50 mL of propylene glycol monomethyl ether acetate
(PGMEA, Sigma-Aldrich, St. Louis, MO, USA), followed by a 1 h soak in 25 mL of isopropyl alcohol
(IPA, Sigma-Aldrich, St. Louis, MO, USA). If a release layer was used, then the substrate plus the

35
Nanomaterials 2018, 8, 498

structure were removed from the IPA and immersed in the Microchem-specified release agent (Remover
PG™, MicroChem Corp., Newton, MA, USA). The IPA was removed for either air or supercritical
drying. Supercritical drying was accomplished using carbon dioxide (CO2 ) and a commercial drying
system (SPI-DRY™ , SPI Supplies, Inc., West Chester, PA, USA). Completion of the IPA solvent exchange
with CO2 was determined using gas chromatography. The exchange was terminated when the residual
IPA attained a level of 0.03% in the monitored CO2 effluent.
The surface morphology was characterized by optical and scanning electron microscopy (SEM).
SEM images were obtained using a Hitachi model S4700 (Hitachi, Ltd., Tokyo, Japan). To obtain
high-quality images, the samples were vapor coated with ~5 nm of chromium or gold. The imaging
voltage was kept low (<10 kV) to avoid damaging the structures.
Raman spectra were recorded using a Raman microscope (Renishaw, InVia™ H 18415, UK)
operating at an excitation wavelength of 785 nm and was focused onto the sample through a 50X
objective lens (NA 0.75). Raman scattering was collected using the same lens. The average laser power
and accumulation time used to record the Raman spectra were 10 mW and 10 s, respectively. Fourier
transform infrared (FTIR) spectra were recorded on resins and polymerized thin films between 400
and 4000 cm−1 using a FTIR spectrometer (Nicolet™ iS50, Thermo Fisher Scientific Co., Waltham,
MA, USA) equipped with diamond attenuated total reflection (ATR). Polymerized thin films (6 μm)
were prepared by spin coating the resin on fused silica substrates, and then curing by single-photon
polymerization at 395 nm for 10 min at 12 mW/cm2 .

2.4. Finite Element Analysis


Finite element analysis was used to simulate the shrinkage and deformation of the log-pile
structures and foam rods via COMSOL Multiphysics® 5.3 software (COMSOL, Inc., Burlington, MA,
USA), assuming a linear elastic response. Mesh configurations were created using COMSOL’s built
in “fine mesh” to ensure solution convergence and computational efficiency. Constrained (fixed)
boundary conditions were chosen to simulate the adhesion of the polymerized resin to the substrate.
Input material properties are given in Table 1. The effective Young’s modulus for the open cell
foam, Ef , was estimated using the correlation, as reported by Ashby [23]:

2
ρf
E f = Es , (2)
ρs

where, ρf is the density of foam; and, ρs and Es are the density and Young’s modulus, respectively, for
the polymerized resin.
Shrinkage was simulated by an equivalent thermal contraction of the structure using a stepped
temperature drop and a user-defined thermal expansion coefficient and heat capacity for the foam
and top layer. In contrast, a standard solid mechanics treatment was used to model the deformation
caused by capillary forces. Further details are given in Section 3.5.

3. Results and Discussion

3.1. Structural Resolution of ID-Dip and IP-S Resin


The feature size of a microstructure fabricated by 2PP is determined by the size of the voxels,
which is related to the induced photon intensity and sequent chemical reactions. The absorption of
photons depends on the square of the light intensity, and the use of ultrashort pulses can start intense
nonlinear processes at relatively low average power [24]. Theoretical studies have been established by
several groups to investigate the dependence of linewidth that is based on nonlinear absorption [25,26].
In experiment, measurements of 2PP line widths versus laser power have been reported for three
Nanoscribe resins (IP-DIP, -L780, and -G780) [6]. The work was carried out using the same Nanoscribe
Professional GT system used here and at scan rates of 10 and 20 mm/s. A simple engineering model

36
Nanomaterials 2018, 8, 498

was used to predict the line characteristics vs. laser power and scan rate. Here, similar measurements
and treatment are reported for IP-S.
A set of support bars was first printed followed by a series of suspended lines normal to the bars
(Figure 2a). The lines were printed using different laser powers and suspended to avoid complications
due to interactions at the resin-to-glass interface. Each laser pulse above the threshold power initiated
some degree of polymerization in an ellipsoidal-shaped voxel at laser focus. The fast laser repetition
rate generated a continuous line of polymer comprised of closely overlapping voxels, each having
an effective volume, Vvox . The typical spacing between successive shots was ~0.1 nm at a scan rate of
10 mm/s and a laser repetition rate of 80 MHz, so the volume within a typical effective voxel received
~103 –104 laser shots (Table 2).

Figure 2. (a) Suspended line structures used to quantify 2PP line width vs. laser power for IP-S. Line
widths were measured by scanning electron microscopy (SEM) (at normal incidence; see inset image)
for lines printed in 2.0 mW stepped-increments of laser power; (b) Measured and calculated effective
voxel volume vs. laser power2 and (c) linewidth vs. laser power for IP-S resin. The lines were calculated
using the model described in Equations (3)–(6).

The results are plotted in Figure 2b,c in terms of the effective voxel volume and linewidth
as a function of the average laser power. The data were analyzed using a simplified engineering
treatment that was initially suggested by Leatherdale and DeVoe [27], and more recently used by
Thiel et al. [28]. These authors relate the absorbed dose in an initiated voxel volume, Vi (nm3 ), to the
laser operating conditions:
Vi ∼ k ( Pa − Pt )2 texp , (3)

where Pa is the laser average power (mW), Pt is the threshold power (mW), texp is the exposure time (s),
and k is a proportional constant. The square dependence on power is due to the two-photon nature of
the process and, thus directly proportional to the square of the per-pulse peak laser irradiance above
the threshold. Note that we report the results in terms of the system average output power, rather than

37
Nanomaterials 2018, 8, 498

peak irradiance to simplify comparison with the typical system operational parameters. The average
laser power was monitored and controlled during system operation.
Making use of the fact that the exposure time was inversely proportional to the scan rate (Rs )
and assuming the effective polymerized voxel volume, Vvox (nm3 ), was proportional to the initiated
volume, Vi , led to:
Vvox ∼ k ( Pa − Pt )2 /Rs , (4)

Experiments showed the polymerized voxel was ellipsoidal with diameter, D, and length, Z,
perpendicular and parallel to the beam propagation direction, respectively, giving a geometric volume:

πD2 Z πD3 Ar
Vvox = = , (5)
3 3
where Ar is the line aspect ratio, Z/D, which for IP-Dip and IP-S was measured at 2.5 and 5.4,
respectively. Note that the ratio of the aspect ratios for IP-S/IP-Dip is 2.2, in good agreement with the
value of 1.8 for the ratio of the numerical apertures that were used for printing in IP-S (NA = 1.4) and
IP-Dip (NA = 0.8).
Combining Equations (4) and (5) and recognizing that the linewidth (Lw ) equals the effective
voxel diameter (D) under constant scan rate conditions leads to the useful correlation for linewidth vs.
operating laser power:
 1
3( Pa − Pt )2
3
Lw = k , (6)
π Ar R s

The measured and calculated effective voxel volume and linewidth for IP-S are plotted vs.
(Pa − Pt )2 and (Pa − Pt ) in Figure 2b,c, respectively. In general, the agreement is reasonable given the
indicated error in linewidth measurements. Similar reasonably good agreement has been reported in
prior tests using IP-Dip and other resins [6].
A threshold power (Pt ) of 6 mW was assumed for IP-S, which is equivalent to the value for IP-Dip
and other resins that were determined at very low scan rates (~0.1 mm/s) [20]. This threshold agreed
with the lack of detectable polymerization (i.e., lines) below 12.5 mW at the much greater scan rates
used here (10 mm/s). Clearly, some degree of polymerization (gelation) occurs at powers between
6–12 mW; but the dose is insufficient to generate a structure that is capable of surviving development.
In certain cases, the polymerization rate can be varied by a change of laser power and scan speed.
Some authors have reported changes in polymerization propagation and termination rates due to
temperature gradients formed around the focal point during 2PP [29,30]. However, in situ temperature
measurements have not revealed a significant heating effect on the polymerization process when
working at close-to-threshold conditions [31]. Therefore, the effect of localized thermal accumulation
on 2PP fabricated structures is not included in this fitting model.
The primary benefit of the analysis reported here is as an engineering tool that, by interpolation,
can reliably predict the line dimensions at different laser operating conditions in a given resin.
The major limitation is that the analysis is largely an empirical treatment and it does not address the
details of the excitation and complex polymerization chemistry of the process.

3.2. Plastic Strain in Simple Beam Structures Written in IP-Dip and IP-S Resins
Polymers characteristically have low elastic moduli and yield strengths but can accommodate
significant plastic strain before ultimate failure [32]. These characteristics are an advantage in many 2PP
applications. For example, many photo-resins undergo some shrinkage during polymer conversion, as
evidenced by the greater density of the polymer vs. resin phase. Typically, polymer shrinkage is less
than 2%. Consequently, polymers can generally accommodate small amounts of plastic strain without
failure, thus leaving the desired structure fully intact.

38
Nanomaterials 2018, 8, 498

Problems tend to arise in 2PP fabrication when there are large strains, particularly in the cases of
significant shrinkage or differential shrinkage during development. Examples of this are shown in
Figure 3 for two log-pile like foam blocks fabricated in IP-Dip resin. The blocks were designed to be
50 × 50 × 50 μm3 ; yet, after drying, both had shrunk by ~10% to ~45 μm in width. The extent of the
shrinkage was clearly visible in the rows at the base of the block where the fabricated lines contacted
and adhered to the substrate. This degree of shrinkage is consistent with the shrinkage measured in
other foam-like log-pile structures that were fabricated here and reported elsewhere [6,9,11,14].
Both structures in Figure 3 accommodated the shrinkage without evidence of plastic strain in
the central portion of the structure. In such cases, one could attempt to compensate for shrinkage by
simply designing and fabricating a proportionally larger structure.
Problems due to shrinkage were most evident at the boundaries of IP-Dip log-pile structures
(Figure 3b). Large shear stresses developed at the fixed boundary between the polymer and the
substrate. Also, certain structural elements, such as cantilever-type beams or simple beams that span
long unsupported distances, exhibited large plastic deformation. Such deformations are visible in the
structure in Figure 3b, but are noticeably absent for the shorter spans in Figure 3a.

Figure 3. Log-pile structures with (a) 3 × 3 × 1 μm3 and (b) 6 × 6 × 1 μm3 cell size fabricated in IP-Dip
resin with ~300–400 nm line width. Note the lack of observable plastic deformation at the smaller cell
size in (a) in contrast to the visible bending in the simply-supported and cantilever beam sub-elements
at the larger cell size in (b).

Similar log-pile structures were written in IP-S (Figure 4). The linewidth and height were greater
than those in IP-Dip because of the larger numerical aperture, as discussed in Section 3.1. Consequently,
IP-S structures had to be fabricated using a bigger cell size (6 × 6 × 3 μm3 , Figure 4a) to achieve
foam densities that are equivalent to IP-Dip. The beams were laterally offset in successive layers
with repeating alignment on every fourth vertical layer (Figure 4a). Each beam in the log-pile was
fabricated using vertically offset and partially overlapping (50%) double scans to achieve the 3 μm
height (Figure 4a inset). In general, the IP-S log-pile structures exhibited significantly less deformation
than the similar structures that were written in IP-Dip (Figure 4b–d).
To better compare the strengths of structures written in IP-Dip vs. IP-S, we fabricated, developed,
and air dried a series of simply supported and cantilever beams in the two resins (Figures 5 and 6).
Except for the beam length, all of the structures were designed and fabricated in the same way.
The beam cross-sections were designed to be 3 × 3 μm2 and were fabricated using a 10-wide × 6-high
scan grid, specifically 10 lateral scans at 0.3 μm line spacing and six vertical scans at 0.5 μm layer
spacing. The average laser power was 15 mW, and the scanning speed was 10 mm/s. The beam
structures were developed, rinsed, and air dried, as described in Section 2.3.

39
Nanomaterials 2018, 8, 498

Figure 4. (a) The design of a log-pile structure with a 6 × 6 × 3 μm3 cell size fabricated using 50%
overlapping double scans as described in the text. SEM images of (b) the 250 × 250 × 100 μm3 foam
block fabricated in IP-S resin and in magnified views from (c) the top showing the linewidth and
horizontal lattice spacing and (d) the side indicating the repeating overlap of every fourth layer, i.e.,
4 × 3 um = 12 um.

Figure 5. Simply supported beam structures fabricated in (a,c) IP-Dip and (b,d) IP-S resin. The scan
direction was from right to left, as indicated by the arrow. The average laser power was 15 mW, and the
scanning speed was 10 mm/s; see the text for further details.

40
Nanomaterials 2018, 8, 498

Figure 6. Cantilever beam structures of varying lengths with an integrated end support printed in
(a) IP-Dip and (b) IP-S resin. The printed beam width is 3 μm with a lateral spacing between beams of
~20 μm and vertically suspended above the base substrate by ~20 μm. The “critical length” for collapse
under capillary drying forces is indicated by the dashed line.

The final objectives used for IP-S and IP-Dip were 25X (NA = 0.8) and 63X (NA = 1.4), respectively
(Table 2), with associated fabricated linewidths of ~0.4 and ~0.65 μm. Thus, adjacent scan lines
overlapped more in IP-S than in IP-Dip.
Simply supported beam structures that were fabricated in both IP-S and Dip showed no
measurable plastic deformation after development and air drying (Figure 5). The only difference
in performance between the two resins was (a) unevenness in the vertical thickness of the longest
beam fabricated in IP-Dip (140 μm) and (b) greater overall vertical beam thickness as achieved in IP-S.
The latter effect was due to the greater depth of field (Rayleigh range).
Figure 6 shows cantilever beams that were fabricated in IP-S and IP-Dip and then developed,
rinsed in IPA, and air dried. Some of the longer beams plastically deformed to such an extent that they
were connected in pairs as well as to the substrate. This is not surprising as the liquid meniscus that
drives the capillary forces would be expected to span the spaces between the beams, as well as connect
the beams to the substrate.
The effect of capillary drying forces on deformation in microscale cantilever and simply supported
beams has been rather extensively studied because of the common use of air drying for solvent removal
in many microfabrication processes (for example, [33–36]). The extent of deformation is generally
characterized by a “critical length”, which refers to the distance from the beam attachment at the end
support to the point of beam adhesion to a neighboring beam, the substrate, or both. For example,
the critical length that was observed for the cantilever beams that were fabricated in IP-Dip was
~55–60 μm, whereas for IP-S, the value was ~153–173 μm (Figure 6).
Liu et al. [33] and Mastrangelo and Hsu [35] both provide closed-form solutions for estimating the
critical length based on the polymer properties, beam dimensions, and inter-beam spacing. Although
their mathematical approaches differ, they arrive at the same relationship:
 1/4
3Ew3 d2
Lc = (7)
8γ cos(θ )

where Lc is the critical length (μm), E Young’s modulus (GPa), w beam width (μm), d beam spacing
(μm), γ solvent surface tension (22 dyne/cm, IPA), and θ the wetting angle. Here, we assumed the
structure was fully wetted (θ ~0◦ ). Using the reported Young’s modulus for IP-S of 4.6 GPa (Table 3)
gives a critical length of 157 um, which agrees well with the measured value. Repeating the same
calculation for IP-Dip is problematic as Young’s modulus depends strongly on the writing speed
and laser power (i.e., energy dose, J/cm3 ). For example, Lemma et al. [16] report a linear increase
of 0.35 GPa/mW in Young’s modulus from ~0.75 to 3.6 GPa over a range in average laser power

41
Nanomaterials 2018, 8, 498

from 5–13 mW. The writing speed was 100 μm/s. In the work reported here, the writing speed
was 10,000 um/s at a laser power of 15 mW. Therefore, Young’s modulus was expected to be lower.
Equation (7) was used to estimate a Young’s modulus of ~0.1 GPa based on the observed cantilever
beam critical length of ~60 μm (Figure 6a).

Table 3. Summary of FTIR peak intensities (normalized to the C=O peak) for CH2 =CH- stretching and
bending vibrational modes.

Band Group IP-S: Peak Intensity IP-Dip: Peak Intensity


(cm−1 ) and Mode Resin UV-Cured Film DC Resin UV-Cured Film DC
C=C
~1635 0.06 0 100 0.07 0.02 71.43
stretch
~1405 C=C bend 0.03 0 100 0.34 0.08 76.47
~940 C=C bend 0.11 0 100 N.D. N.D. N.D.
~810 C=C bend 0.1 0 100 0.41 0.07 82.93

Mastrangelo [35] also treats the case of capillary collapse for a simply supported beam (i.e., a beam
clamped at both ends). Using his results, we predicted the critical span to be ~160 and 400 μm for
IP-Dip and IP-S, respectively. This agreed with the lack of collapse that was observed for the beam
structures in Figure 5.
Polarization of the laser beam has been reported to affect the intensity distribution and thermal
gradients around the focal spot thus leading to different polymerization rates, which can, in certain
cases, affect the feature size and introduce small changes (~20%) in some mechanical properties [30].
We believe that impact of polarization effect on mechanical properties is likely to be small for our
application compared to other effects. For example, Young’s modulus was estimated to be ~0.1 GPa
for the IP-Dip polymerized structures written here, while the fully polymerized IP-Dip photoresist
has a Young’s modulus of 4.5 GPa (Table 1). Also, at the employed high writing speed (10,000 μm/s),
anisotropy in heat flow would be a second order effect for enhancing the mechanical stability of the
foam targets. Besides, the beam is circular polarized for our writing process. Other research work has
shown that the circular polarization of incident light could ensure a more spherical voxel within the
xy-plane [37]. This avoids polarization-dependent linewidth between separate log-pile layers where
the scan directions are perpendicular to each other.
Yoshimoto et al. [34] offers a different approach for describing plastic yield in micro-cantilever
beams, with the resulting expression for the critical length in terms of the yield strength:
 1/2
σy d
Lc = w (8)
6γ cos(θ )

where δy is the yield strength (MPa) and the other variables are the same as given above. To our
knowledge, the yield strength for IP-Dip and IP-S has not been reported, so Equation 7 and the results
in Figure 6 provide a means to estimate these values, specifically, δy ~3 MPa for IP-Dip and ~20 MPa
for IP-S.

3.3. Fourier Transform Infrared and Micro-Raman Vibrational Spectroscopy of Resin Conversion
FTIR and micro-Raman vibrational spectroscopy were used to monitor the degree of
polymerization in the IP-S and IP-Dip acrylic resins. Other recent studies have shown these techniques
provide a wealth of molecular detail at the nano to microscale, about the extent of monomer/oligomer
photo-conversion (see, for example, [11,14,20,38,39]). The characteristic vibrational bands that
were associated with the CH2 =CH-, C=O, and C-O groups that comprise the two resins are well
known [40–43] and are clearly detected in both the FTIR and micro-Raman spectra (Figure 7).
FTIR bands are due to linear optical absorption by an oscillating dipole associated with the
vibrations of a particular molecule or functional group [41,42]. In contrast, Raman bands are scattering

42
Nanomaterials 2018, 8, 498

phenomena and relate to the polarizability of the molecule or the molecular group. Specifically, Raman
bands are associated with the radiation from an oscillating dipole that was induced by the incident
laser electric field [40,41]. Thus, the two methods are complementary in that vibrational bands that
are weak or not detected by one method may be detected by the other; this is often the case for
our application.

Figure 7. Fourier transform infrared (FTIR) spectra of the resin and fully cured film of (a) IP-S and (b)
IP-Dip over the fingerprint region of 700–1800 cm−1 . The bands associated with the terminal CH2 =CH-
stretching and bending modes are indicated on the spectra. (c) Raman spectra of IP-Dip and IP-S 2PP
cured photoresists.

In general, the FTIR spectra provide greater structural detail across the so-called molecular
fingerprint region (~700 to 1800 cm−1 ), which includes characteristic stretching and bending modes
of CH2 =CH-, C=O, and C-O [42,43]. In addition, the method is insensitive to fluorescence from the
initiator in the resin. The major drawback is that the FTIR spectrometer can only probe macroscopic
samples. Raman microspectroscopy, on the other hand, has the advantage of being able to probe
small volumes (<0.5 μm dia.) and it has greater sensitivity to the CH2 =CH- stretching vibration at
~1600–1640 cm−1 [44]. This bond has a low dipole moment and it gives only weak FTIR bands, while
the Raman signal is strong due to the large polarizability of the C=C bond. The main drawback to
micro-Raman for our application was the interference caused by fluorescence from the initiator in
the resin.
FTIR spectra of unreacted resins and UV-cured films of IP-S and IP-Dip are shown in Figure 7.
Table 3 summarizes the measured strength of the C=C stretching band and the three bending modes at
~1635, 1405, ~810, and ~940 cm−1 , respectively. The intensity was normalized to the C=O band intensity
because that group concentration is expected to remain constant in a given sample. The degree of

43
Nanomaterials 2018, 8, 498

conversion (DC) was calculated by comparing C=C stretching or bending mode intestines to a reference
C=O band before and after UV photopolymerization [15,20]:
  

DC = 1 − ( AC=C /AC=O )/ AC =C /AC =O × 100, (9)

where AC=C , AC=O , A C=C , and A C=O are the integrated intensity of corresponding peaks in the
polymerized and the unpolymerized resins. The IP-S spectra showed the expected result of complete
reaction of the terminal alkene group after UV exposure. In contrast, the UV-exposed IP-Dip sample
still contained ~17–29% unreacted C=C based on the ratios of the bands at ~810, ~1405, and 1635 cm−1
before and after UV exposure.
We next carried out micro-Raman measurements of lines that were written by 2PP in both resins
using the conditions that are summarized in Table 2. The samples were developed and then examined
using Raman microscopy, as described in Section 2.3.
Figure 7c shows the Raman bands for the C=O and CH2 =CH- stretching modes after 2PP exposure.
Both of the resins showed significant amounts of unreacted CH2 =CH-. IP-Dip in particular showed
low conversion, which is consistent with the trend that was observed in the UV-exposure results in
Figure 7b. The reason for the difference in vinyl conversion of IP-Dip vs IP-S resins is difficult to assess
without detailed knowledge of the chemical structure of these proprietary resins. Nevertheless, the IR
and Raman data coupled with the elemental resin composition do offer a few hints. We assume here
that the vinyl conversion continues (propagates) until the well-known free radical termination by
oxygen (O2 ) [45]. Note that, based on the IR spectra and elemental composition results, IP-Dip is a fully
vinyl-acrylate resin, whereas IP-S does show the presence of some amine functionality (amine bands at
~3300–3500 cm−1 ). We suspect that IP-Dip conversion becomes sterically hindered early on. In other
words, molecular rearrangement is too slow (diffusion limited) to permit access to the unreacted vinyl
groups before the small, highly mobile O2 terminates the reaction. In contrast, the full UV conversion of
IP-S suggests adequate mobility to achieve full reaction before termination. The incomplete conversion
of IP-S under 2PP is possibly due to insufficient initiation. Of course, use of other common organic
chemistry structural analysis tools, for example, C13 and H1 nuclear magnetic resonance (NMR), gas
chromatography-mass spectrometry (GC-MS), size exclusion chromatography (SEC), etc., could fully
elucidate the structures of both resins. We have chosen not to do that here. Hopefully, the resin vendor
will soon publish the structure making such analyses unnecessary.
The results suggest that low conversion is a major factor contributing to the large shrinkage
and low modulus and yield strength of the IP-Dip foam structures. This agrees with prior work by
Jiang et al. [12,15], who also reported low conversion by 2PP in custom resins with a resulting loss in
mechanical integrity.
It is also probable that the C=C conversion varies through the width of the line, being greatest
at the point of peak irradiance at the center of the focal spot and lower near the gaussian beam
edges. Thus, the effective thickness of the line would be less than the observed thickness. Because the
mechanical strength varies as approximately the cube of the line thickness, and then small negative
changes in the effective line thickness would significantly weaken the part.

3.4. Fabrication of Foam Rods in IP-Dip and IP-S Resins


Figures 8 and 9 show foam rods (2.0 × 0.25 × 0.35 mm3 ) fabricated in IP-Dip and IP-S, respectively,
using the writing conditions in Table 2. Both foam structures have a design density of 0.10 g/cm3 .
One of the IP-Dip rods was fabricated as a 100% foam log-pile structure (Figure 8a), whereas the other
had a 15-μm-thick cap layer that functioned as a laser ablator (Figure 8b). Only IP-S rods with the
15 μm cap layer were fabricated (Figure 9).
The IP-Dip and IP-S rods were comprised of a series of 125 × 125 × 100 μm3 and
250 × 250 × 100 μm3 sub-blocks, respectively. The ability to print larger sub-blocks for the IP-S
rods is a direct result of the numerical aperture for printing (IP-S 0.8 vs. 1.4 for IP-Dip). The IPS

44
Nanomaterials 2018, 8, 498

sub-block width was designed to match the rod width, thereby reducing the number of stitching
boundaries by eightfold over that for rods that were written in IP-Dip. The foam cell size for the IP-S
rods was also designed to be larger (6 × 6 × 3 μm3 ) to offset the mass of the thicker beams and still
achieve the 0.10 g/cm3 density goal. Each IP-S beam was fabricated with vertically offset and partially
overlapping double scans (50%) to achieve a taller and stiffer (more reacted) structure (Figure 4a).

Figure 8. SEM images of 2 × 0.25 × 0.3 mm3 foam rod with x, y, z cell dimensions of 6.2 × 6.2 × 1 μm3
fabricated in IP-Dip (a) without and (b) with a 15-um-thick fully dense cap layer.

Figure 9. SEM images of (a) 2 × 0.25 × 0.3 mm3 foam rod with x, y, z cell dimensions of 6 × 6 × 3 μm3
and showing areas at higher magnification to illustrate (b) structure and (c) stitching boundary quality.

45
Nanomaterials 2018, 8, 498

The IP-Dip rods showed the maximum deformation of ~6–7% at the interfaces of the foam with
the substrate and with the solid cap layer (Figure 8). In contrast, the IP-S foam rods showed much
less shrinkage and deformation in these interfaces (Figure 9). The largest defects/deformations in the
IP-S rod occurred at the corners of the stitching boundaries (Figure 9b,c) and were consistent in size
and location throughout the structure. Because these defects occurred at the outer surface of the rod,
they did not impact the region of the target that was irradiated by the laser. Nevertheless, work is
continuing on improving the process to eliminate these defects.

3.5. Analysis of Shrinkage and Deformation in IP-Dip Foam Rods


Greater detail on the deformation of the IP-Dip rods at the interfaces between the foam-substrate
boundary and the full-density cap layer are shown in Figure 10. Specifically, the images in Figure 10a,c
provide magnified views of the rod end regions within the dashed boxes in Figure 8.

Figure 10. Details of the structural deformation of IP-Dip foam rods, specifically showing the regions
within the dashed boxes in Figure 8. The SEM images show the ends of the rod (a) without and (b)
with a 15-um-thick fully dense cap layer. Resin shrinkage during development and drying produced
residual axial shear stresses and associated plastic strains in both the foam only and the foam with
top cap, as shown schematically in (c,d) and also indicated in the SEM images; the arrow lengths are
notional representations of the relative magnitude of the axial shear stresses.

Comparison of the designed and measured dimensions of the IP-Dip foam rods indicated ~10%
maximum axial shrinkage at the top of the foam rod. This is consistent with the measured shrinkage

46
Nanomaterials 2018, 8, 498

in other foam-like structures that were fabricated here and reported elsewhere [6,9]. The images in
Figure 10 show that the axial shear stresses (and strains) due to shrinkage were greatest near “fixed”
boundaries that were constrained by adhesion of foam to the glass substrate and to the full density top
cap. Also, the stitching boundaries (Figure 10a,c) are regions of inherently lower strength that in some
cases have been observed to cause layer-to-layer delamination under shrinkage-induced shear strains.
The axial shear stresses are indicated schematically in the SEM images in Figure 10a,b where the
arrow lengths are notional representations of the relative magnitude of the stresses. Shear stresses in
the lateral direction also exist, but are smaller.
Deformation of IP-Dip foam rods during the development process may be due to capillary forces
and/or internal shrinkage due to the complete or partial removal of unconverted resin within the
fabricated lines. Simulations were used to separately investigate the magnitude of these two effects by
FEA. The geometry and the mesh configuration of a unit lattice and the foam rods with and without
the cap layer are shown in Figures 11a and 12a,b, respectively. The base of the lattice and the foam were
constrained to simulate adhesion to an infinitely stiff substrate. In the case of the rod, the top cap was
also assumed to fully adhere to the foam, although each can elastically respond to the applied stress.
We first considered the effects of meniscus drying forces. The capillary pressure, P, due to
meniscus forces during air drying, can be estimated using the Young–Laplace equation:

P ∼ 4γ cos(θ )/L (10)

where L is the effective pore diameter (or inter-beam distance), γ is the solvent surface tension, and θ is
the wetting angle. Applying Equation (10) to the 6.2 × 6.2 × 1 μm3 foam unit cell and using a surface
tension for IPA of γ = 22 dyne/cm and assuming a fully wetted structure (θ ~0◦ ) gives an estimated
maximum capillary pressure of ~8 kPa (~12 psi). The FEA of the foam rods predicted a maximum
three-dimensional (3D) deformation of only ~130 nm over the entire length of the rod structure. This
was less than half of the 2PP line width. Consequently, the deformation was expected to be well within
the elastic limit with no permanent plastic strain, even at the very low estimated foam modulus value
of ~0.1 MPa.
In contrast to the capillary drying effects, FEA simulations of the impact of shrinkage showed
deformations that closely matched observations. Figure 11 shows show the FEA mesh configuration
and the simulations results for a 24 × 24 × 8 μm3 beam lattice representing the microscale details
of an individual building block of the foam rod. The lattice architecture had a horizontal beam
(x, y) spacing of 6.2 μm, a beam height (z) of 1 μm, and a beam width (i.e., line width) of 0.4 μm.
The simulation showed that shrinkage of the log-pile structures leads to uniform compression in the
center region and plastic deformation at the end, as is consistent with the behavior shown in Figure 3.
The vertical strain gradient (Figure 11b,c) was ~6–7 nm/micron over the 8 μm height and 24 × 24 μm
base of this lattice structure. This predicted an expected total deformation of ~35 μm over the ~60 μm
thick first layer of the 2 mm foam rod (Figure 10a,b).

Figure 11. (a) Mesh configuration and (b,c) finite element analysis (FEA) simulation of
shrinkage-induced deformation for 24 × 24 × 8 μm3 log-pile block fabricated in IP-Dip with 6.2 μm
line spacing. See text for details.

47
Nanomaterials 2018, 8, 498

The simulations of the magnitude of rod 3D deformation that is caused by shrinkage (Figure 12c,d)
were also in reasonable agreement with the observed behavior in Figures 8 and 10. The main differences
between the simulations and observations were for the 100% foam rod. Specifically, the simulation
showed a vertical expansion at the rod ends (Figure 12c), which is most likely due to the effect of the
material Poisson ratio upon contraction. In the real rods, this did not occur, probably because of the
shearing that was observed at the horizontal stitching boundaries, as shown in Figure 10a. In the case
of the rod with a cap layer, the simulations compared most closely with the observations. These rods
showed reduced shear at the stitching boundaries (Figure 10d) due to the axial constraint of the top
layer. Moreover, the simulations correctly predicted the bending of the top cap at the ends of the rod
where the shrinkage-induced stresses were the greatest. In this work, we assumed that the observed
shrinkage was due to the effects of the unconverted resin, as discussed in Section 3.3. Other earlier
studies of shrinkage of 2PP structures reached a similar conclusion [44,46]. However, a quantitative
description of shrinkage at the molecular level remains elusive and is a subject of continued interest.

Figure 12. Mesh configuration used for the FEA simulations of foam rods fabricated in IP-Dip resin
(a) without and (b) with a 15-μm-thick fully dense cap layer and (c,d) the computed deformation due
to shrinkage.

4. Summary and Conclusions


A series of low-atomic number (CHO) millimeter-scale foam laser targets with a 4–6 μm cell size
were fabricated using 2PP. The targets are used to support HED physics research, thus driving the
requirement that the foam contain only low-atomic number elements. The targets were comprised of
a full-density 15 um cap layer at a 0.10 g/cm3 foam base. The cap layer served as an ablator.
Two commercial acrylic resins were evaluated for preparing the foam targets: (i) IP-Dip, a low
viscosity resin designed for high-resolution printing with a large numerical aperture objective and
(ii) IP-S a high viscosity resin for mesoscale printing using a lower numerical aperture. A fabricated
linewidth in IP-S for different irradiance conditions was reported and compared to prior measurement
on IP-Dip and also analyzed using a simple engineering model.
Infrared and Raman spectroscopy were used to measure the extent of 2PP polymerization by
monitoring the C=C bond conversion. Although both IP-Dip and IP-S showed significant unconverted
material, IP-Dip was worse; the results for IP-Dip agreed with observations that were reported by
others. It was proposed that full or partial removal of unreacted resin from within the beams during
development was the primary cause of the shrinkage and the resulting deformation observed in the
final structures.
Simple end-supported beam test structures were fabricated in each resin to assess the polymer
strength and the impact of capillary drying forces. The results showed that simple beams and foams

48
Nanomaterials 2018, 8, 498

fabricated with these resins were strong enough to support typical capillary drying forces of ~5–10 kPa
(~0.7–1.5 psi) without plastic deformation. However, polymer linear shrinkage of up to 6–7% or more
during resin development led to large structural plastic deformation in IP-Dip. Finite element analysis
was used to simulate the effects of both capillary drying forces and the polymer shrinkage. Drying
forces produced elastic deformations <0.5 μm, whereas shrinkage generated ~100× greater axial plastic
deformation μm for these target structures.
A significant reduction in shrinkage-induced deformation and improvements in the structure
strength and rigidity were achieved by using IP-S resin. Initial tests showed great improvements
in the fabricated rod dimensional stability, with up to 4× fewer stitching boundaries and ~1/10th
the shrinkage.

Author Contributions: Y.L., J.H.C., O.S. and Y.L. conceived and designed the experiments; Y.L. and O.S. performed
the experiments; Y.L., J.H.C., O.S., L.J., J.H., and Y.L. contributed the analysis; Y.L. and J.H.C. wrote the paper.
All authors discussed the results and commented on the manuscript at all stages.
Acknowledgments: The authors gratefully acknowledge the financial support of the U.S. Dept. of Energy under
contract No. DE-NA0001385. We greatly appreciate the technical support of our many colleagues at the University
of Nebraska-Lincoln Laser Assisted Nano-Engineering Laboratory and at Schafer Livermore Laboratory.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. National Research Council. Frontiers in High Energy Density Physics: The x-Games of Contemporary Science,
1st ed.; The National Academies Press: Washington, DC, USA, 2003.
2. McCrory, R.; Meyerhofer, D.; Betti, R.; Craxton, R.; Delettrez, J.; Edgell, D.; Glebov, V.Y.; Goncharov, V.;
Harding, D.; Jacobs-Perkins, D.; et al. Progress in direct-drive inertial confinement fusion. Phys. Plasmas
2008, 15, 055503. [CrossRef]
3. Rambo, P.K.; Smith, I.C.; Porter, J.L.; Hurst, M.J.; Speas, C.S.; Adams, R.G.; Garcia, A.J.; Dawson, E.;
Thurston, B.D.; Wakefield, C. Z-Beamlet: A multikilojoule, terawatt-class laser system. Appl. Opt. 2005, 44,
2421–2430. [CrossRef] [PubMed]
4. Spaeth, M.L.; Manes, K.; Kalantar, D.; Miller, P.; Heebner, J.; Bliss, E.; Spec, D.; Parham, T.; Whitman, P.;
Wegner, P. Description of the NIF laser. Fusion Sci. Technol. 2016, 69, 25–145. [CrossRef]
5. Bernat, T.; Campbell, J.; Petta, N.; Sakellari, I.; Koo, S.; Yoo, J.-H.; Grigoropoulos, C. Fabrication of
micron-scale cylindrical tubes by two-photon polymerization. Fusion Sci. Technol. 2016, 70, 310–315.
[CrossRef]
6. Jiang, L.; Campbell, J.; Lu, Y.; Bernat, T.; Petta, N. Direct writing target structures by two-photon
polymerization. Fusion Sci. Technol. 2016, 70, 295–309. [CrossRef]
7. Jiang, L.; Campbell, J.; Lu, Y.; Bernat, T.; Petta, N. Precision fabrication of laser targets: Development
of 2-photon polymerization as a next-generation tool. In Proceedings of the International Congress on
Applications of Lasers & Electro-Optics, Atlanta, GA, USA, 18–22 October 2015.
8. Jiang, L.J.; Maruo, S.; Osellame, R.; Xiong, W.; Campbell, J.H.; Lu, Y.F. Femtosecond laser direct writing in
transparent materials based on nonlinear absorption. MRS Bull. 2016, 41, 975–983. [CrossRef]
9. Stein, O.; Liu, Y.; Streit, J.; Campbell, J.; Lu, Y.; Aglitskiy, Y.; Petta, N. Fabrication of low-density
shock-propagation targets using two-photon polymerization. Fusion Sci. Technol. 2018, 73, 153–165.
[CrossRef]
10. Oakdale, J.S.; Smith, R.F.; Forien, J.B.; Smith, W.L.; Ali, S.J.; Bayu Aji, L.B.; Willey, T.M.; Ye, J.;
van Buuren, A.W.; Worthington, M.A. Direct laser writing of low-density interdigitated foams for plasma
drive shaping. Adv. Funct. Mater. 2017, 27, 1702425. [CrossRef]
11. Baldacchini, T. Three-Dimensional Microfabrication Using Two-Photon Polymerization: Fundamentals, Technology,
and Applications, 1st ed.; William Andrew: Waltham, MA. USA, 2015.
12. Lindl, J. Inertial confinement fusion: The quest for ignition and energy gain using indrect drive. Nucl. Fusion
1999, 39, 825.

49
Nanomaterials 2018, 8, 498

13. Mačiulaitis, J.; Deveikytė, M.; Rekštytė, S.; Bratchikov, M.; Darinskas, A.; Šimbelytė, A.; Daunoras, G.;
Laurinavičienė, A.; Laurinavičius, A.; Gudas, R. Preclinical study of SZ2080 material 3D microstructured
scaffolds for cartilage tissue engineering made by femtosecond direct laser writing lithography. Biofabrication
2015, 7, 015015. [CrossRef] [PubMed]
14. Jiang, L.; Xiong, W.; Zhou, Y.; Liu, Y.; Huang, X.; Li, D.; Baldacchini, T.; Jiang, L.; Lu, Y. Performance
comparison of acrylic and thiol-acrylic resins in two-photon polymerization. Opt. Express 2016, 24,
13687–13701. [CrossRef] [PubMed]
15. Žukauskas, A.; Matulaitienė, I.; Paipulas, D.; Niaura, G.; Malinauskas, M.; Gadonas, R. Tuning the refractive
index in 3D direct laser writing lithography: Towards grin microoptics. Laser Photonics Rev. 2015, 9, 706–712.
[CrossRef]
16. Lemma, E.D.; Rizzi, F.; Dattoma, T.; Spagnolo, B.; Sileo, L.; Qualtieri, A.; De Vittorio, M.; Pisanello, F.
Mechanical properties tunability of three-dimensional polymeric structures in two-photon lithography.
IEEE Trans. Nanotechnol. 2017, 16, 23–31. [CrossRef]
17. König, K.; Uchugonova, A.; Straub, M.; Zhang, H.; Licht, M.; Afshar, M.; Feili, D.; Seidel, H. Sub-100 nm
material processing and imaging with a sub-15 femtosecond laser scanning microscope. J. Laser Appl. 2012,
24, 042009. [CrossRef]
18. Oakdale, J.S.; Ye, J.; Smith, W.L.; Biener, J. Post-print uv curing method for improving the mechanical
properties of prototypes derived from two-photon lithography. Opt. Express 2016, 24, 27077–27086.
[CrossRef] [PubMed]
19. O’Brien, A.K.; Cramer, N.B.; Bowman, C.N. Oxygen inhibition in thiol-acrylate photopolymerizations.
J. Polym. Sci. Pol. Chem. 2006, 44, 2007–2014. [CrossRef]
20. Jiang, L.J.; Zhou, Y.S.; Xiong, W.; Gao, Y.; Huang, X.; Jiang, L.; Baldacchini, T.; Silvain, J.-F.; Lu, Y.F.
Two-photon polymerization: Investigation of chemical and mechanical properties of resins using raman
microspectroscopy. Opt. Lett. 2014, 39, 3034–3037. [CrossRef] [PubMed]
21. Paz, V.F.; Emons, M.; Obata, K.; Ovsianikov, A.; Peterhänsel, S.; Frenner, K.; Reinhardt, C.; Chichkov, B.;
Morgner, U.; Osten, W. Development of functional sub-100 nm structures with 3D two-photon polymerization
technique and optical methods for characterization. J. Laser Appl. 2012, 24, 042004. [CrossRef]
22. Greaves, G.N.; Greer, A.; Lakes, R.; Rouxel, T. Poisson’s ratio and modern materials. Nat. Mater. 2011, 10,
823. [CrossRef] [PubMed]
23. Ashby, M. The properties of foams and lattices. Philos. Trans. R. Soc. Lond. A Math. Phys. Eng. Sci. 2006, 364,
15–30. [CrossRef] [PubMed]
24. Hsu, W.-H.; Masim, F.C.P.; Balčytis, A.; Juodkazis, S.; Hatanaka, K. Dynamic position shifts of x-ray emission
from a water film induced by a pair of time-delayed femtosecond laser pulses. Opt. Express 2017, 25,
24109–24118. [CrossRef] [PubMed]
25. Lee, K.S.; Yang, D.Y.; Park, S.H.; Kim, R.H. Recent developments in the use of two-photon polymerization in
precise 2D and 3D microfabrications. Polym. Adv. Technol. 2006, 17, 72–82. [CrossRef]
26. Serbin, J.; Egbert, A.; Ostendorf, A.; Chichkov, B.; Houbertz, R.; Domann, G.; Schulz, J.; Cronauer, C.;
Fröhlich, L.; Popall, M. Femtosecond laser-induced two-photon polymerization of inorganic–organic hybrid
materials for applications in photonics. Opt. Lett. 2003, 28, 301–303. [CrossRef] [PubMed]
27. Leatherdale, C.A.; DeVoe, R.J. Two-photon microfabrication using two-component photoinitiation systems:
Effect of photosensitizer and acceptor concentrations, In Nonlinear Optical Transmission and Multiphoton
Processes in Organics. In Proceedings of the Optical Science and Technology, Spie’s 48th Annual Meeting,
San Diego, CA, USA, 3–8 August 2003.
28. Thiel, M.; Fischer, J.; von Freymann, G.; Wegener, M. Direct laser writing of three-dimensional submicron
structures using a continuous-wave laser at 532 nm. Appl. Phys. Lett. 2010, 97, 221102. [CrossRef]
29. Takada, K.; Kaneko, K.; Li, Y.-D.; Kawata, S.; Chen, Q.-D.; Sun, H.-B. Temperature effects on pinpoint
photopolymerization and polymerized micronanostructures. Appl. Phys. Lett. 2008, 92, 041902. [CrossRef]
30. Rekštytė, S.; Jonavičius, T.; Gailevičius, D.; Malinauskas, M.; Mizeikis, V.; Gamaly, E.G.; Juodkazis, S.
Nanoscale precision of 3D polymerization via polarization control. Adv. Opt. Mater 2016, 4, 1209–1214.
[CrossRef]
31. Mueller, J.B.; Fischer, J.; Mange, Y.J.; Nann, T.; Wegener, M. In-situ local temperature measurement during
three-dimensional direct laser writing. Appl. Phys. Lett. 2013, 103, 123107. [CrossRef]

50
Nanomaterials 2018, 8, 498

32. Juodkazis, S.; Mizeikis, V.; Seet, K.K.; Misawa, H.; Wegst, U.G. Mechanical properties and tuning of
three-dimensional polymeric photonic crystals. Appl. Phys. Lett. 2007, 91, 241904. [CrossRef]
33. Liu, J.-L.; Feng, X.-Q.; Xia, R.; Zhao, H.-P. Hierarchical capillary adhesion of microcantilevers or hairs. J. Phys.
D: Appl. Phys. 2007, 40, 5564. [CrossRef]
34. Yoshimoto, K.; Stoykovich, M.P.; Cao, H.; de Pablo, J.J.; Nealey, P.F.; Drugan, W.J. A two-dimensional model
of the deformation of photoresist structures using elastoplastic polymer properties. J. Appl. Phys. 2004, 96,
1857–1865. [CrossRef]
35. Mastrangelo, C.; Hsu, C. Mechanical stability and adhesion of microstructures under capillary forces. Ii.
Experiments. J. Microelectromech. Syst. 1993, 2, 44–55. [CrossRef]
36. Mastrangelo, C. Adhesion-related failure mechanisms in micromechanical devices. Tribol. Lett. 1997, 3,
223–238. [CrossRef]
37. Guney, M.; Fedder, G. Estimation of line dimensions in 3D direct laser writing lithography.
J. Micromech. Microeng. 2016, 26, 105011. [CrossRef]
38. Shukla, A.K.; Palani, I.; Manivannan, A. Laser-assisted dry, wet texturing and phase transformation of
flexible polyethylene terephthalate substrate revealed by raman and ultraviolet-visible spectroscopic studies.
J. Laser Appl. 2018, 30, 022008. [CrossRef]
39. Suzuki, T.; Morikawa, J.; Hashimoto, T.; Buividas, R.; Gervinskas, G.; Paipulas, D.; Malinauskas, M.;
Mizeikis, V.; Juodkazis, S. Thermal and optical properties of sol-gel and su-8 resists. In Advanced Fabrication
Technol. Micro/Nano Optics Photonics V, 82490K; International Society for Optics and Photonics: Bellingham,
WA, USA, 2012.
40. Harris, D.C.; Bertolucci, M.D. Symmetry and Spectroscopy: An Introduction to Vibrational and Electronic
Spectroscopy; Chapter 3; Dover Publication: New York, NY, USA, 1989.
41. Colthup, N.B.; Daly, L.H.; Wiberley, S.E. Introduction Infrared Raman Spectroscopy; Elsevier, Academic Press:
San Diego, CA, USA, 2012.
42. Smith, B.C. Infrared Spectral Interpretation: A Systematic Approach; CRC Press: Boca Raton, FL, USA, 1998.
43. Socrates, G. Infrared and Raman Characteristic Group Frequencies: Tables and Charts; John Wiley & Sons:
Chichester, UK, 2001.
44. Ovsianikov, A.; Shizhou, X.; Farsari, M.; Vamvakaki, M.; Fotakis, C.; Chichkov, B.N. Shrinkage of
microstructures produced by two-photon polymerization of Zr-based hybrid photosensitive materials.
Opt. Express 2009, 17, 2143–2148. [CrossRef] [PubMed]
45. LaFratta, C.N.; Baldacchini, T. Two-photon polymerization metrology: Characterization methods of
mechanisms and microstructures. Micromachines 2017, 8, 101. [CrossRef]
46. Ovsianikov, A.; Viertl, J.; Chichkov, B.; Oubaha, M.; MacCraith, B.; Sakellari, I.; Giakoumaki, A.; Gray, D.;
Vamvakaki, M.; Farsari, M.; et al. Ultra-low shrinkage hybrid photosensitive material for two-photon
polymerization microfabrication. ACS Nano 2008, 2, 2257–2262. [CrossRef] [PubMed]

© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

51
nanomaterials

Article
High Repetition Rate UV versus VIS Picosecond
Laser Fabrication of 3D Microfluidic Channels
Embedded in Photosensitive Glass
Florin Jipa 1 , Stefana Iosub 1 , Bogdan Calin 1 , Emanuel Axente 1 , Felix Sima 1,2, *
and Koji Sugioka 2, *
1 Center for Advanced Laser Technologies (CETAL), National Institute for Laser, Plasma and Radiation
Physics (INFLPR), 409 Atomistilor, Magurele RO-77125, Romania; florin.jipa@inflpr.ro (F.J.);
stefana.iosub@inflpr.ro (S.I.); bogdan.calin@inflpr.ro (B.C.); emanuel.axente@inflpr.ro (E.A.)
2 RIKEN Center for Advanced Photonics, 2-1 Hirosawa, Wako, Saitama 351-0198, Japan
* Correspondence: felix.sima@inflpr.ro (F.S.); [email protected] (K.S.); Tel.: +4021-4574-491 (F.S.)

Received: 6 July 2018; Accepted: 26 July 2018; Published: 31 July 2018

Abstract: Glass is an alternative solution to polymer for the fabrication of three-dimensional (3D)
microfluidic biochips. Femtosecond (fs) lasers are nowadays the most promising tools for transparent
glass processing. Specifically, the multiphoton process induced by fs pulses enables fabrication of
embedded 3D channels with high precision. The subtractive fabrication process creating 3D hollow
structures in glass, known as fs laser-assisted etching (FLAE), is based on selective removal of the
laser-modified regions by successive chemical etching in diluted hydrofluoric acid solutions. In this
work we demonstrate the possibility to generate embedded hollow channels in photosensitive Foturan
glass volume by high repetition rate picosecond (ps) laser-assisted etching (PLAE). In particular,
the influence of the critical irradiation doses and etching rates are discussed in comparison of two
different wavelengths of ultraviolet (355 nm) and visible (532 nm) ranges. Fast and controlled
fabrication of a basic structure composed of an embedded micro-channel connected with two open
reservoirs, commonly used in the biochip design, are achieved inside glass. Distinct advantages
such as good aspect-ratio, reduced processing time for large areas, and lower fabrication cost
are evidenced.

Keywords: picosecond laser processing; 3D microfluidic channels; photosensitive glass

1. Introduction
Microfluidic systems typically consisting of three-dimensionally (3D) embedded channels
connected to open micro-reservoirs are useful tools for many biological and medical studies, since
they are basic elements for biochips such as lab-on-a-chip devices and micro-total-analysis-systems
that can perform reaction, detection, analysis, separation, and synthesis of biochemical materials with
high-efficiency, high-speed, high-sensitivity, low reagent consumption, and low waste production [1,2].
The unique 3D geometries offer flexibilities and specific functionalities for fabrication of the
biochips [3,4] or even organs-on-a-chip systems [5]. Such microfluidic devices for biomedical
applications are generally fabricated based on PDMS with casting technologies [6,7]. However,
although they exhibit indubitable advantages such as biocompatibility, good optical quality, and easy
to use, some drawbacks including non-reusability, and adsorption of organic compounds, as well as
the requirement of multiple stacking and sealing processes in the fabrication procedure push us
to find an alternative [8]. During the last decade, femtosecond (fs) laser fabrication has proven
to be a powerful tool for 3D internal modification of transparent glass materials and fabrication of
embedded channels [9–12]. Thus, glass is a good alternative to PDMS for specific biological applications,

Nanomaterials 2018, 8, 583; doi:10.3390/nano8080583 52 www.mdpi.com/journal/nanomaterials


Nanomaterials 2018, 8, 583

which allow creating robust, easy to clean and reusable devices. Among many glasses, Foturan is
one of the most suitable materials for fabrication of microfluidic systems. Foturan is a photosensitive
glass with photoreactive properties due to the addition of a photoactive agent (photo-sensitizer) and
metal ions (nucleation agent) in the glass matrix. The photoactive agent is cerium (<0.04 wt% Ce2 O3 ),
while the nucleation ion is silver (0.05–0.15 wt% Ag2 O) [13]. Unlike other transparent glasses these
agents allow this glass to be processed in a 3D manner by space selective control of the precipitation
process [14]. Photoactivation takes place at wavelengths shorter than 350 nm, and then a successive
annealing treatment induces silver clustering that converts a latent image of irradiated region into
an observable one. During the thermal treatment, a metasilicate crystalline phase is grown around
the formed silver clusters which, by an isotropic chemical etching, can be selectively removed to
create 3D hollow micropatterns in the glass matrix [15]. Masuda et al. used a high-intensity fs laser
emitting light of 150 fs pulse width at 775 nm wavelength, 1 kHz repetition rate and 0.4 μJ, to fabricate
complex 3D microfluidic structures inside Foturan glass with a high spatial resolution [16]. In this case,
the photo-reaction of near-infrared fs pulses with glass takes place by two-step excitation of electrons
with three photon absorptions each, resulting in a six-photon process [17].
This technique enabling the fabrication of 3D microfluidic structures has been termed femtosecond
laser-assisted etching (FLAE). FLAE of Foturan glass was then applied to fabricate specific biochips
called nano-aquariums for monitoring of continuous motion of Euglena gracilis [18] and evaluation of
gliding mechanism of Phormidium cyanobacteria [19]. Meanwhile, picosecond (ps) lasers, which are also
categorized as ultrafast lasers, are becoming more common tools for practical use due to higher power
and higher reliability as compared to fs lasers. At this time-scale, the deposition of laser energy is still
typically faster than the electron-phonon coupling processes (which are material-dependent), enabling
the minimization of heat-affected zone and, thus, high-quality micro- and nano-fabrication [20].
In addition, the high peak power (Ppeak = E/τ, E—pulse energy, τ—pulse duration) can induce
nonlinear absorption processes in materials which do not absorb the laser wavelength, thus allowing
the processing not only of the surface, but also of the inside of transparent materials similarly to the
fs lasers [9,10]. Therefore, the ps laser may be able to replace the fs lasers for 3D microfabrication of
Foturan glass, which is beneficial in terms of the high throughput process.
Veiko et al. have demonstrated the possibility to modify Foturan glass by a ps Nd:YAG laser with
a pulse width of 30 ps at 532 nm wavelength and a repetition rate of 10 Hz [21]. A mechanism based
on two-photon absorption was proposed. Thus, microstructures attributed to the phase transition
from amorphous to crystalline inside the Foturan glass-ceramic material were fabricated by means of
local laser modification followed by subsequent thermal treatment. It has been further shown that
using the same laser source it is possible to create 3D channels in bulk by hydrofluoric acid (HF)
etching of crystallized regions developed by thermal treatment using a CO2 laser [22]. Even though the
proof of concept was demonstrated, the potential of ps lasers for the structuring and fabrication of 3D
embedded channels in Foturan glass is still not fully explored and many challenges remained unmet.
In this study we demonstrate the successful fabrication of 3D microfluidic structures in Foturan
glass by high repetition rate ps laser-assisted etching (PLAE) using either the second or third harmonics
(visible (VIS) 532 nm or ultraviolet (UV) 355 nm wavelengths) of a Nd:YVO4 laser. Critical irradiation
doses and etching ratios were examined for both wavelengths to optimize the experimental conditions
of PLAE. Controlled fabrication of the microfluidic structures consisting of an embedded channel
connected with two open micro-reservoirs is achieved for both cases. Our study evidenced that
transparent materials processing with high repetition rate ps laser pulses based on multi-photon
absorption could be a viable alternative to classical fs micro-fabrication. More importantly, large-area
3D micro/nanofabrication with considerably reduced processing time and production costs will offer
great advantages for manufacturing with high repetition rate ps laser pulses.

53
Nanomaterials 2018, 8, 583

2. Materials and Methods


Foturan used in this study is a photo-structurable glass ceramic manufactured by Schott North
America Inc., Corporate Office, Elmsford, NY, USA. It is a photo-sensitive alkali-aluminosilicate glass
material consisting of SiO2 (75%–85%), Li2 O (7%–11%), K2 O and Al2 O3 (3%–6%), Na2 O (1%–2%), ZnO
(0%–2%), Sb2 O3 (0.2%–0.4%), Ag2 O (0.05%–0.15%), and Ce2 O3 (0.01%–0.04%). In our experiments,
we cut Foturan glass wafers to 10 × 10 × 2 mm3 dimensions to create 3D microfluidic structures.
Prior to use, the samples were successively cleaned in baths of acetone, alcohol, and deionized water.
The laser direct writing was conducted by a ps laser beam (Lumera, Coherent) delivering pulses
of duration below 10 ps at 500 kHz repetition rate, coupled with a customized workstation (Figure 1).
The second (532 nm—VIS) and the third (355 nm—UV) harmonics of a Nd:YVO4 laser were used in
experiments with laser power ranging from tens of mW for critical doses up to 500 mW for material
processing using VIS wavelength and from values below 1 mW for critical doses up to 10 mW for
structuring in UV. For both wavelengths, the scanning speed was varied from 0.1 mm/s to 1 mm/s
in order to determine the optimum value of interline spaces which is an important parameter for a
line-by-line scanning process. The beam was focused using an aspheric lens of 15 mm focal length
producing a circular focal spot of 4 μm in diameter. The entire writing process was monitored with a
CCD camera, which was also used to control the focusing position on the sample surface and inside
the volume. The samples were placed on an X-Y motorized translation stage (PlanarDL, Aerotech Inc.,
Pittsburgh, PA, USA) with computer-control which provided 200 mm travel range on each axis with
±500 nm step accuracy and ±100 nm precision. In addition, the stage has the ability to trigger the
laser firing for precise synchronization, increasing thus the accuracy of the imprinted pattern.

Figure 1. Scheme and photo of the workstation used for high repetition rate ps laser irradiation of
Foturan glass.

The exposed samples were then annealed in a furnace (model MTF M1238-250 from Carbolite Gero
Limited, Hope Valley, UK) controlled with the following program: heating with a slope of 5 ◦ C/min up
to 500 ◦ C, then keeping the temperature constant for 1 h in order to grow Ag nanoparticles, increasing
the temperature again with a slope of 3 ◦ C/min to 605 ◦ C, and keeping the temperature constant
for another hour to obtain the crystalline phase of Li metasilicate. The process was followed by
chemical etching in 8% HF solution under ultrasonic condition. During the etching, the crystalline
phase grown around Ag nanoparticles was selectively removed. Profilometry analysis was carried out
with a stylus profiler XP2 from Ambios Technology, 0.01 mm/s speed and 1 mm working distance.
Optical interrogation of the samples was performed in transmission mode with a microscope, model
DM4000 B Led from Leica Microsystems, Wetzlar, Germany. Scanning electron microscopy studies
were employed with an FEI Co. microscope, model Inspect S, Hillsboro, OR, USA.

54
Nanomaterials 2018, 8, 583

3. Results

3.1. Critical Dose Evaluation


In a first step, the experimental procedure was devoted to the evaluation of the critical dose Dc
for each irradiation wavelength, in order to find optimum parameters for laser processing. By using a
model proposed by Fuqua et al. [23] as the critical laser fluence Fc , the critical irradiation dose can be
expressed as:
Dc = Fcm × N (1)

N is the number of laser pulses necessary to induce photoreaction and Ag nucleation and m
represents the number of photons in the multiphoton absorption process for electron generation. Fc is
defined as the lowest fluence at which a sufficiently high density of nuclei (Ag nanoclusters) is able to
form an interconnected network of the metasilicate crystalline phase in photo-structurable glass by the
thermal treatment for selective removal. Here, the density of nuclei ρ generated at a fluence of F can be
expressed by ρ = K × Fm × N (K is a proportionality constant).
The determination of UV and VIS critical doses was performed by irradiating Foturan glass surface
with different laser powers and exposure times. For fixed power values, several exposure times from
0.5 to 3 s were applied, corresponding to 2.5 × 105 to 15 × 105 pulses, resulting in the accumulation of
different laser doses in distinct areas. The exposure reproducibility was demonstrated by generating a
matrix of identical patterns for each area with a separation distance of 10 μm. This approach allowed
easier optical evaluation of the samples and an improved accuracy in Dc evaluation.
For each laser wavelength, several power densities were first applied in order to define the optimal
power interval for the Dc evaluation. For a good statistical analysis, we have irradiated 25 identical
points using the same laser power and the experiment was repeated twice. No visible differences
were observed between experiments. In the case of UV, irradiation the laser power was systematically
varied from 0.6 mW to 1 mW with 0.1 mW step (Figure 2a). The same procedure was used for the
VIS wavelength for which the laser power was adjusted in the 30–50 mW range with a 5 mW step
(Figure 2c).

Figure 2. Identification of threshold irradiation parameters for ps laser irradiation of Foturan glass.
(a,b) Optical images of the exposure map written on glass surface using 355 nm ps laser pulses (a) and
corresponding critical fluence determination (b); (c,d) Optical images of the exposure map written on
glass surface using 532 nm ps laser pulses (c) and corresponding critical fluence determination (d).
The inset images represent detailed views of areas marked with square dots.

55
Nanomaterials 2018, 8, 583

After irradiation, the glass samples were submitted to a classical thermal treatment protocol
described in the previous section. Upon the thermal treatment, brownish crystalline phase became
visible on the samples by optical microscopy when the accumulated dose exceeded Dc . The analyses of
the irradiated patterns revealed a dose-dependent modification of samples for both laser wavelengths.
We identified the irradiation parameters (power as a function of exposure time or number of pulses),
at which the glass surface suffered visible modification by laser irradiation followed by thermal
treatment (see stars marked in Figure 2a,c) for selective chemical etching, revealing Dc . Fc is a function
of the number of pulses for both 355 nm (Figure 2b) and 532 nm (Figure 2d) laser wavelengths.
The fluences were estimated for each laser power considering the spot size of 4 μm in diameter.
By linearly fitting the log-log representation data, in Figure 2b,d we determined the critical fluences
of 1.742 J/cm2 (m = 3) for VIS and 0.635 J/cm2 (m = 2) for UV. By introducing these values into
Equation (1), the critical dose values for both wavelengths were determined to be Dc = 5.28 J3 /cm6 for
VIS and Dc = 0.4032 J2 /cm4 for UV, respectively.
The evaluation of Dc is essential in order to further correlate with the translation stage
programming for optimizing writing speeds of various complex 3D shapes. In addition, the subsequent
wet chemical etching for the subtractive process is critically dependent on the laser irradiation dose for
the high precision fabrication of embedded microfluidic channels.

3.2. Etching Rate Estimation


Using the estimated critical doses, we further determined the etching rate dependence for ps
UV and VIS lasers, respectively. In HF solution, the contrast ratio in etching selectivity between the
unexposed and the laser exposed Foturan glass was found to be 1:50, and was dependent on both the
exposure dose and the HF concentration. This ratio was almost coincident with that determined when
using the fs laser [24].
We have determined the etching rate of unexposed Foturan glass starting from an initial thickness
of 2 mm after immersion in 48 mL solution of 8% HF concentration. At different time intervals,
the glass was extracted from solution to measure its thickness with an electronic micrometer for up to
1 h of total exposure. The dependence of the glass thickness on the etching time is plotted in Figure 3a.
The evaluation of the linear fitting evidenced an etching rate of about 1.6 μm/min which is in line
with previously reported values, such as the study by Helvajian et al. (1.3 μm/min for a slightly lower
concentration of 5% HF solution) [25].

Figure 3. Etching of Foturan glass in 8% HF solution: (a) etched thickness as a function of etching time
for unexposed glass; (b,c) width of open channels after 5 min of etching of the linear pattern written by
UV (b) and VIS (c) lasers as a function of writing speed. Each linear regression curve corresponds to
different laser power.

The etching rates for VIS and UV lasers exposed samples were estimated by evaluating depths
obtained after etching for different time intervals. In both cases, at moderate laser exposures (just above
the modification thresholds), ~10 ± 2 μm/min etching rates were obtained. Consequently, the etching
ratios were found to be at 1:10. These ratios are dependent on the laser exposure doses. Specifically,
an etching rate of 20 μm/min can be achieved when increasing the laser power, corresponding to
an etching ratio of up to 1:25. Comparing to etching ratios reported in other publications, where

56
Nanomaterials 2018, 8, 583

identical wavelength but different pulse widths were used to expose the glass, the values obtained
in our study are similar while taking account of processing parameters such as irradiation doses and
HF concentration.
The etching characteristics of glass modified by the high repetition rate ps laser were investigated
by profilometry measurements. For both laser wavelengths, sets of five identical lines with 1 mm
length and 100 μm interspace were written on the glass surface using different irradiation doses.
To this purpose, predetermined scanning speeds ranging from 0.1 mm/s to 1 mm/s were employed at
different laser powers, followed by annealing treatment. The channel widths created after 5 min of HF
etching were evaluated, as presented in Figure 3b,c, for UV and VIS ps laser processing, respectively.
As a general observation, one can notice that the channel widths are dependent on the laser dose,
as expected (Figure 3b,c). Indeed, for both wavelengths, a higher dose revealed the formation of wider
channels (>10 μm), while only few micrometer widths are obtained for lower doses. In particular,
one can obtain 11 μm width channels with UV ps laser pulses of 10 mW power at a writing speed of
0.1 mm/s, and narrower than 4 μm width, with 8 mW average laser power and 1 mm/s writing speed.
Similarly, channels with widths wider than 12 μm are obtained by VIS ps laser pulses of 300 mW
power and writing speed of 0.1 mm/s, and narrower than 3 μm width, at 200 mW laser power and
1 mm/s writing speed. These results allow us to predict the optimum laser processing parameters and
scanning regimes for choosing appropriate interspaces between linear patterns in order to create more
complex 3D structures.

3.3. Fabrication of Microfluidic Embedded Channels


Further, the capability of generating embedded microfluidic channels in photosensitive glass using
either UV or VIS high repetition rate ps laser pulses was demonstrated. By the new PLAE technique,
we propose fabricating a simple microfluidic structure consisting of two open micro-reservoirs
connected by an embedded channel (Figure 4). Specifically, square reservoirs of 1 × 1 mm2 area
are connected by a 1 mm length and 300 μm width channel. The whole structure was designed and
fabricated by irradiating the glass in a two-layer configuration (see Figure 4a): (i) the first layer (upper
part) of the reservoirs (opening) were formed by scanning the focused laser beam on the glass surface
and then (ii) the second layer (bottom part) of the reservoirs and connecting channel were formed
by changing the laser focusing position from the glass surface into the volume at specific depths.
Each layer consists of parallel lines written by linear scanning of the focused laser beam at a speed of
0.5 mm/s with lateral sliding at a 5 μm step between each line. The PLAE process was then followed by
a two-step treatment at 500 ◦ C for one hour followed by one at 605 ◦ C for another hour (Figure 4b) and
successive chemical etching in HF (Figure 4c) in order to obtain the designed structure. The etching
time depends on the exposure parameters, which were varied between 50 and 60 min, similar with
times used for FLAE.

Figure 4. PLAE process of Foturan glass. (a) Sketch of the proposed design (with two open reservoirs
connected by an embedded channel) and laser irradiation using two-layer configuration; and (b,c) partial
view by optical microscopy of the obtained structure after irradiation with the VIS laser followed by (b)
thermal treatment and (c) chemical etching in 8% HF solution. The scale bar represents 0.5 mm.

57
Nanomaterials 2018, 8, 583

We created two structures written with the identical writing scheme using VIS-PLAE at 500 mW
laser power, 1 μJ energy, and 10 μm interline spaces, but at different scanning speeds to investigate
influence of the exposure dose on the created structures. The laser irradiation was carried out on
the surface (first layer) and at 250 μm depth (second layer) with writing speeds fixed at 0.1 mm/s
(Figure 5a,b) and 0.9 mm/s (Figure 5c) respectively. SEM analyses have revealed that a lower irradiation
dose (correlated with faster writing speed and smaller number of applied laser pulses) can generate
embedded microfluidic channels in glass volume after 50 min of wet chemical etching (Figure 5b,c).
Contrarily, when slower speed (higher irradiation dose) is applied, the channel rooftop disappeared
due to second layer overexposure (Figure 5a). Indeed, by increasing the irradiation dose inside glass,
the volume affected by multiphoton absorption becomes larger. Consequently, if the position of the
second layer is close enough to the surface, the channel rooftop can be removed during the chemical
etching. It is worth mentioning that fabrication of the structure obtained at the scanning speed of
0.1 mm/s (Figure 5a) required a processing time of about 70 min, while the structure fabricated at
0.9 mm/s (Figure 5c) could be written in only 10 min. One may further adjust the thickness of channel
rooftop by tailoring the scanning speed.

Figure 5. SEM images of structures fabricated by VIS-PLAE using a ps laser at power of 500 mW and a
design of 10 μm interline spaces, at two different scanning speeds: 0.1 mm/s (a,b) and 0.9 mm/s (b,c).
The laser irradiation was carried out with two-layer configuration: on the surface (first layer) and at
250 μm depth (second layer).

Fabrication of the similar structure was also attempted with the same writing scheme by ps laser
processing at a wavelength of 355 nm. After finding optimal parameters, embedded channels could be
also fabricated by PLAE using a UV ps laser (UV-PLAE). In particular, by applying a power of 12 mW,
energy of 0.024 μJ, a 5 μm interline step and a writing speed of 0.5 mm/s the structure with two open
reservoirs connected by embedded channel was successfully created. With these processing conditions,
the entire structure was written in less than 20 min and an etching time of 60 min.
In Figure 6, we present SEM images of twin structures separated by 2 mm written in Foturan
glass by UV-PLAE. Pairs of two open square reservoirs of 1 × 1 mm2 area are connected by 1 mm
length and 335 μm width embedded channels.

58
Nanomaterials 2018, 8, 583

Figure 6. SEM images of twin structures fabricated by UV-PLAE using a ps laser of 12 mW laser power
and 5 μm interline spaces, at 0.5 mm/s scanning speed. The laser irradiation was carried out using
two-layer configuration: (on the surface (first layer) and at 500 μm depth (second layer)).

4. Discussion
During interaction with solid materials, the pulsed laser beam is depositing its energy, inducing
different phenomena dependent on the pulse energy, duration, and focusing optics [26]. It is rather
a thermal process for long pulses (nanosecond regime), while a physical aspect predominates
for ultrashort pulses (less than a few ps pulses that are shorter than electron-phonon coupling
time in materials) [27]. Consequently, in the case of ultrashort pulses, the heat-affected zone is
minimized [20,28,29]. In most of the studies, the photo-physical and photo-chemical mechanisms
involved during glass processing with ultrashort pulses are evidenced for femtosecond laser pulses and
extensively addressed, from both fundamental and applicative points of view [12,30–32]. Origins of
material modification under interaction with fs pulses could be related to the densification induced by
pressure wave and/or fast heating-cooling processes [33,34]. As a result, electrons in the conduction
band are heated by the laser pulse very quickly so that they do not have time to diffuse out from the
irradiated volume or to recombine. The photoionization is then responsible of seeding electrons for
the subsequent avalanche ionization. It was found out that the electron density increases by avalanche
ionization until its plasma frequency reaches the critical plasma density [35]. On the other hand,
the avalanche ionization is more efficient for longer pulses since it allows more time for increasing the
electron density. In our study, ps laser irradiance values of 3.73 × 1013 and of 3.57 × 1012 Wcm−2 for
VIS and UV, respectively, were determined, below values of 1014 Wcm−2 for fs laser irradiances [16].
These values for ps laser irradiances corresponds to a Keldysh parameter above 1.5 which indicates
the multi-photon absorption is more dominant rather than tunneling ionization in the case of both UV
and VIS irradiation [36]. In contrast, pulses longer than a few tens of ps do not reach enough intensity
to directly photoionize the electrons.
In the first studies using ultrafast laser reported in literature, Foturan photosensitive glasses
were exposed to 150 fs laser pulses at 775 nm wavelength [16]. A second thermal treatment step
followed by isotropic etching conducted to a preferential material removal from irradiated regions.
In this case, the experimental investigation of Fc on the number of pulses allowed the calculation of
a critical dose of Dc = 1.3 × 10−5 J6 /cm12 with m = 6 for the 775 nm wavelength fs laser. Interaction
mechanisms of laser pulses with Foturan glass were further explored by employing: (i) fs laser (150 fs,
775 nm, 1 kHz), and ns lasers of (ii) 266 nm; (iii) 355 nm; and (iv) 308 nm laser wavelengths [17].
A significant increase in absorption spectrum of the exposed samples around 360 nm was found for fs
irradiated glasses corresponding to absorption from oxygen deficient centers which originated from
the interband excitation of electrons. The absorption at 315 nm related to Ce ions was not observed,
suggesting that Ce3+ ions do not contribute to electron generation for the reduction to Ag atoms in the
case of fs-irradiated samples at 775 nm. Thus, it was concluded that free electrons were generated by

59
Nanomaterials 2018, 8, 583

two-step interband excitation through the defect levels with three-photon absorption each resulting in
six photons in total for photoreaction. A similar photoreaction mechanism by successive interband
excitation was evidenced for 266 nm ns laser-irradiated samples, but by a linear two-photon process,
i.e., two-step excitation by single-photon absorption each. On the other hand, in the case of 355 nm
laser-irradiated samples, free electrons are generated by Ce3+ , while in the case of the 308 nm laser
both absorption by Ce3+ (single-photon absorption) and interband excitation (the linear two-photon
process) was found. In this study, we evaluated the possibility to obtain embedded microfluidic
channels in Foturan glass by using a high repetition rate (500 kHz) laser of <10 ps pulse duration at
532 nm and 355 nm wavelengths. A nonlinear absorption process is evident at the interaction of both
532 nm and 355 nm laser pulses with Foturan glass, since the absorption edge of Foturan is shorter than
350 nm. The Foturan glass also has an absorption peak around 315 nm ascribed to Ce3+ absorption,
which corresponds to photon energy of 3.93 eV. Therefore, for 532 nm with the photon energy of 2.33 eV,
two photons are necessary to generate free electrons from Ce3+ . Another channel for free electron
generation is interband excitation, for which the two-step excitation model through the intermediate
state has been proposed [17]. For this excitation, the photon energy of 3.49–4.66 eV is required for each
step, indicating four photons in total for the 532 nm beam. We have found m = 3 photon process for
VIS pulses. Thus, it is likely that the free electrons are generated by both the Ce3+ absorption and the
interband excitation through the intermediate state similarly to the case of the 308 nm ns laser [17].
In the case of UV pulses, at 355 nm, with the photon energy of 3.49 eV, we found m = 2 photon process
which should correspond to free electron generation by two-photon absorption by Ce3+ similarly to
the case of the 355 nm ns laser [17]. In addition, we consider that, in our experimental conditions,
the high repetition rate of the ps laser pulses can increase the nonlinear absorptivity into Foturan glass
material [37]. The alteration of the surface and volume morphologies that can be observed above
the embedded channel in Figure 6 (white dotted regions) may support the hypothesis that, for high
repetition rate UV ps laser pulses, the heat effect becomes more important when the laser beam is
focused above a critical intensity.
During interaction with a Gaussian laser beam (fs or ps durations) with Foturan glass,
the absorption profile affects the shape of modification volume to develop an elliptical cross-sectional
shape of the crystallized area. In case of a high intensity fs laser pulse, the main advantage resides
in high density, compact modification at very low power, which can be used for high resolution
micro-processing. On the other hand, ps lasers can represent a viable alternative for large-scale
processing both on surface and in volume due to high average power compensation since longer,
high-energy, pulses can modify more material per pulse. Indeed, a larger energy can be deposited in
the material since the time is longer. Thus, this energy allows more time for growth of electron density
during the laser irradiation and in consequence an increase of formed Ag atoms. These Ag atoms
are then responsible of the formation of larger Ag clusters and a larger glass crystalline area during
thermal treatment. As a result, one may consider that the area of the crystalline phase is dependent
on pulse duration. Using the same focusing optics, ps laser beams can consequently decrease the
processing times as compared with femtosecond lasers. As a direct comparison, to create a similar
structure, irradiation time was 1.5 h for laser pulses of 360 fs at 522 nm (2 mm/s scanning speed and
250 KHz laser repetition rate) while, in our case, it was of approximately 10 min at a scanning speed of
0.9 mm/s and 500 KHz (SI in [38]). On the other hand, in case of shorter pulses one needs less energy
to achieve the intensity for optical breakdown allowing the achievement more precise machining with
femtosecond lasers rather than the longer pulse lasers. Thus, higher resolution processing is achieved
with fs pulses than ps pulses. In case of ps laser pulses, a more efficient process could be achieved for
UV pulses since the critical dose is one order of magnitude less than in case of VIS pulses compensating
the laser power conversion efficiency (1:2 in case of VIS and 1:3 for UV).
The high repetition rate ps laser pulses could thus stand as a prospective processing benefit,
attractive for several applications that require high speed and cost-effective manufacturing. On the
other hand, by controlling the irradiation dose with respect to the laser wavelength and etching

60
Nanomaterials 2018, 8, 583

parameters one can explore the unique characteristics in 3D microfabrication for a wide range
of applications.
We have finally demonstrated that PLAE is a suitable and very fast fabrication method of 3D
embedded microfluidic channels with good aspect ratio and sharp edges without any cracks by using
either ps laser pulses in UV or VIS.

5. Conclusions
High repetition rate ps laser processing at both 532 and 355 nm wavelengths was applied for
fabrication of 3D microfluidic structures in Foturan glass. A three-photon process with a critical fluence
of 1.742 J/cm2 for the VIS case and a two-photon process with 0.635 J/cm2 for the UV case were found.
Critical dose values of 5.28 J3 /cm6 for VIS and 0.4032 J2 /cm4 for UV cases were calculated.
Straight lines of 1 mm length and 100 μm interspace were then written on the glass surface at
different scanning speeds using different laser powers. Open channels with widths ranging from
3 to 13 μm were developed by thermal treatment and HF etching depending on irradiation doses.
Based on a subtractive fabrication process consisting of selective removal of laser-modified regions by
chemical etching, we could further fabricate 3D hollow structures in glass by PLAE at both 355 nm
and 532 nm laser wavelengths. Due to high power, high repetition rate laser pulses which increase
multiphoton absorption, we could apply the laser irradiation process very fast for fabrication of
embedded structures. A simple configuration consisting of two micro-reservoirs connected by an
embedded channel can be achieved in less than 10 min of laser irradiation by the PLAE technique
using either 355 or 532 nm wavelengths.

Author Contributions: Conceptualization: F.S.; methodology: F.J., S.I., and B.C.; investigation: F.J., S.I., and E.A.;
writing—original draft preparation: F.S.; writing—review and editing: F.S. and K.S.; project administration: F.S.

Funding: This research was funded by UEFISCDI grants 131PED/2017, 148/PED2017 and TE7/2018.

Acknowledgments: The authors are grateful to Iuliana Iordache for profilometry measurements and to Gianina
Popescu-Pelin for SEM support.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Sackmann, E.K.; Fulton, A.L.; Beebe, D.J. The present and future role of microfluidics in biomedical research.
Nature 2014, 507, 181–189. [CrossRef] [PubMed]
2. Yeo, L.Y.; Chang, H.C.; Chan, P.P.; Friend, J.R. Microfluidic devices for bioapplications. Small 2011, 7, 12–48.
[CrossRef] [PubMed]
3. Sima, F.; Serien, D.; Wu, D.; Xu, J.; Kawano, H.; Midorikawa, K.; Sugioka, K. Micro and
nano-biomimetic structures for cell migration study fabricated by hybrid subtractive and additive 3D
femtosecond laser processing. In Proceedings of the Laser-Based Micro- and Nanoprocessing XI (LBMP),
San Francisco, CA, USA, 31 January–2 February 2017; Volume 10092, pp. 1009207–1009215.
4. Sima, F.; Xu, J.; Wu, D.; Sugioka, K. Ultrafast laser fabrication of functional biochips: New avenues for
exploring 3D micro-and nano-environments. Micromachines 2017, 8, 40. [CrossRef]
5. Caplin, J.D.; Granados, N.G.; James, M.R.; Montazami, R.; Hashemi, N. Microfluidic organ-on-a-chip
technology for advancement of drug development and toxicology. Adv. Healthc. Mater. 2015, 4, 1426–1450.
[CrossRef] [PubMed]
6. Fujii, T. Pdms-based microfluidic devices for biomedical applications. Microelectron. Eng. 2002, 61, 907–914.
[CrossRef]
7. Sia, S.K.; Whitesides, G.M. Microfluidic devices fabricated in poly (dimethylsiloxane) for biological studies.
Electrophoresis 2003, 24, 3563–3576. [CrossRef] [PubMed]
8. Bhatia, S.N.; Ingber, D.E. Microfluidic organs-on-chips. Nat. Biotechnol. 2014, 32, 760–772. [CrossRef]
[PubMed]

61
Nanomaterials 2018, 8, 583

9. Itoh, K.; Watanabe, W.; Nolte, S.; Schaffer, C.B. Ultrafast processes for bulk modification of transparent
materials. MRS Bull. 2006, 31, 620–625. [CrossRef]
10. Gattass, R.R.; Mazur, E. Femtosecond laser micromachining in transparent materials. Nat. Photonics 2008, 2,
219–225. [CrossRef]
11. Sugioka, K.; Cheng, Y. Femtosecond laser processing for optofluidic fabrication. Lab Chip 2012, 12, 3576–3589.
[CrossRef] [PubMed]
12. Sima, F.; Sugioka, K.; Vázquez, R.M.; Osellame, R.; Kelemen, L.; Ormos, P. Three-dimensional femtosecond
laser processing for lab-on-a-chip applications. Nanophotonics 2017, 7, 613–634. [CrossRef]
13. Livingston, F.; Adams, P.; Helvajian, H. Influence of cerium on the pulsed uv nanosecond laser processing of
photostructurable glass ceramic materials. Appl. Surf. Sci. 2005, 247, 526–536. [CrossRef]
14. Stookey, S. Catalyzed crystallization of glass in theory and practice. Ind. Eng. Chem. 1959, 51, 805–808.
[CrossRef]
15. Livingston, F.; Helvajian, H. Variable UV laser exposure processing of photosensitive glass-ceramics:
Maskless micro-to meso-scale structure fabrication. Appl. Phys. A 2005, 81, 1569–1581. [CrossRef]
16. Masuda, M.; Sugioka, K.; Cheng, Y.; Aoki, N.; Kawachi, M.; Shihoyama, K.; Toyoda, K.; Helvajian, H.;
Midorikawa, K. 3D microstructuring inside photosensitive glass by femtosecond laser excitation. Appl. Phys.
A: Mater. Sci. Process. 2003, 76, 857–860. [CrossRef]
17. Hongo, T.; Sugioka, K.; Niino, H.; Cheng, Y.; Masuda, M.; Miyamoto, I.; Takai, H.; Midorikawa, K.
Investigation of photoreaction mechanism of photosensitive glass by femtosecond laser. J. Appl. Phys.
2005, 97, 063517. [CrossRef]
18. Hanada, Y.; Sugioka, K.; Kawano, H.; Ishikawa, I.S.; Miyawaki, A.; Midorikawa, K. Nano-aquarium for
dynamic observation of living cells fabricated by femtosecond laser direct writing of photostructurable glass.
Biomed. Microdevices 2008, 10, 403–410. [CrossRef] [PubMed]
19. Hanada, Y.; Sugioka, K.; Shihira-Ishikawa, I.; Kawano, H.; Miyawaki, A.; Midorikawa, K. 3D microfluidic
chips with integrated functional microelements fabricated by a femtosecond laser for studying the gliding
mechanism of cyanobacteria. Lab Chip 2011, 11, 2109–2115. [CrossRef] [PubMed]
20. Le Harzic, R.; Huot, N.; Audouard, E.; Jonin, C.; Laporte, P.; Valette, S.; Fraczkiewicz, A.; Fortunier, R.
Comparison of heat-affected zones due to nanosecond and femtosecond laser pulses using transmission
electronic microscopy. Appl. Phys. Lett. 2002, 80, 3886–3888. [CrossRef]
21. Ageev, E.; Kieu, K.; Veiko, V.P. Modification of photosensitive glass-ceramic foturan by ultra short laser
pulses. In Proceedings of the Fundamentals of LaserAssisted Micro and Nanotechnologies, St Petersburg,
Russia, 5–8 July 2010.
22. Sergeev, M.; Veiko, V.; Tiguntseva, E.; Olekhnovich, R. Picosecond laser fabrication of microchannels inside
foturan glass at CO2 laser irradiation and following etching. Opt. Quantum Electron. 2016, 48, 485. [CrossRef]
23. Fuqua, P.D.; Janson, S.W.; Hansen, W.W.; Helvajian, H. Fabrication of true 3D microstructures in
glass/ceramic materials by pulsed uv laser volumetric exposure techniques. In Proceedings of the Laser
Applications in Microelectronic and Optoelectronic Manufacturing IV, San Jose, CA, USA, 1999; Volume 3618,
pp. 213–221.
24. Sugioka, K.; Cheng, Y. Fabrication of 3d microfluidic structures inside glass by femtosecond laser
micromachining. Appl. Phys. A 2014, 114, 215–221. [CrossRef]
25. Hansen, W.W.; Janson, S.W.; Helvajian, H. Direct-write uv-laser microfabrication of 3D structures in
lithium-aluminosilicate glass. In Proceedings of the Laser Applications in Microelectronic and Optoelectronic
Manufacturing II, San Jose, CA, USA, 1997; Volume 2991, pp. 104–112.
26. Allmen, M.; Blatter, A. Laser-Beam Interactions with Materials: Physical Principles and Applications;
Springer Science & Business Media: Berlin/Heidelberg, Germany, 2013; Volume 2.
27. Chichkov, B.N.; Momma, C.; Nolte, S.; Von Alvensleben, F.; Tünnermann, A. Femtosecond, picosecond and
nanosecond laser ablation of solids. Appl. Phys. A 1996, 63, 109–115. [CrossRef]
28. Singh, J.P.; Thakur, S.N. Laser-Induced Breakdown Spectroscopy; Elsevier: Amsterdam, The Netherlands, 2007.
29. Harilal, S.S.; Freeman, J.R.; Diwakar, P.K.; Hassanein, A. Femtosecond laser ablation: Fundamentals and
applications. In Laser-Induced Breakdown Spectroscopy; Springer: Berlin/Heidelberg, Germany, 2014;
pp. 143–166.
30. Sugioka, K.; Cheng, Y. Femtosecond laser three-dimensional micro-and nanofabrication. Appl. Phys. Rev.
2014, 1, 041303. [CrossRef]

62
Nanomaterials 2018, 8, 583

31. Malinauskas, M.; Žukauskas, A.; Hasegawa, S.; Hayasaki, Y.; Mizeikis, V.; Buividas, R.; Juodkazis, S. Ultrafast
laser processing of materials: From science to industry. Light Sci. Appl. 2016, 5, e16133. [CrossRef]
32. Jiang, L.J.; Maruo, S.; Osellame, R.; Xiong, W.; Campbell, J.H.; Lu, Y.F. Femtosecond laser direct writing in
transparent materials based on nonlinear absorption. MRS Bull. 2016, 41, 975–983. [CrossRef]
33. Schaffer, C.B.; Brodeur, A.; García, J.F.; Mazur, E. Micromachining bulk glass by use of femtosecond laser
pulses with nanojoule energy. Opt. Lett. 2001, 26, 93–95. [CrossRef] [PubMed]
34. 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, 5674–5686. [CrossRef] [PubMed]
35. Wu, A.Q.; Chowdhury, I.H.; Xu, X. Femtosecond laser absorption in fused silica: Numerical and experimental
investigation. Phys. Rev. B 2005, 72, 085128. [CrossRef]
36. Schaffer, C.B.; 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, 1784. [CrossRef]
37. Miyamoto, I.; Cvecek, K.; Schmidt, M. Evaluation of nonlinear absorptivity in internal modification of bulk
glass by ultrashort laser pulses. Opt. Express 2011, 19, 10714–10727. [CrossRef] [PubMed]
38. Wu, D.; Wu, S.Z.; Xu, J.; Niu, L.G.; Midorikawa, K.; Sugioka, K. Hybrid femtosecond laser microfabrication
to achieve true 3D glass/polymer composite biochips with multiscale features and high performance:
The concept of ship-in-a-bottle biochip. Laser Photonics Rev. 2014, 8, 458–467. [CrossRef]

© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

63
nanomaterials

Article
Femtosecond Laser-Based Modification of PDMS to
Electrically Conductive Silicon Carbide
Yasutaka Nakajima 1 , Shuichiro Hayashi 2 , Akito Katayama 1 , Nikolay Nedyalkov 3 and
Mitsuhiro Terakawa 1,2, *
1 School of Integrated Design Engineering, Keio University, 3-14-1, Hiyoshi, Kohoku-ku, Yokohama 223-8522,
Japan; [email protected] (Y.N.); [email protected] (A.K.)
2 Department of Electronics and Electrical Engineering, Keio University, 3-14-1, Hiryoshi, Kohoku-ku,
Yokohama 223-8522, Japan; [email protected]
3 Institute of Electronics, Bulgarian Academy of Sciences, Tzarigradsko shouse 72, Sofia 1784, Bulgaria;
[email protected]
* Correspondence: [email protected]; Tel.: +81-45-566-1737

Received: 28 June 2018; Accepted: 19 July 2018; Published: 22 July 2018

Abstract: In this paper, we experimentally demonstrate femtosecond laser direct writing of conductive
structures on the surface of native polydimethylsiloxane (PDMS). Irradiation of femtosecond laser
pulses modified the PDMS to black structures, which exhibit electrical conductivity. Fourier-transform
infrared (FTIR) and X-ray diffraction (XRD) results show that the black structures were composed
of β-silicon carbide (β-SiC), which can be attributed to the pyrolysis of the PDMS. The electrical
conductivity was exhibited in limited laser power and scanning speed conditions. The technique
we present enables the spatially selective formation of β-SiC on the surface of native PDMS only by
irradiation of femtosecond laser pulses. Furthermore, this technique has the potential to open a novel
route to simply fabricate flexible/stretchable MEMS devices with SiC microstructures.

Keywords: femtosecond laser; silicon carbide; polydimethylsiloxane; laser direct writing

1. Introduction
Polydimethylsiloxane (PDMS) is a widely used polymer in various applications, including
wearable/implantable devices and microfluidics, owing to its biocompatibility, optical transparency,
flexibility, and elasticity [1]. Recently, PDMS has attracted considerable attention as a soft material to
be utilized for flexible/stretchable electrical devices, such as stretchable displays [2] and stretchable
strain sensors [3]. In such flexible/stretchable electrical devices, precise micro- or nano-sized structures
composed of electrically conductive materials, e.g., metals or carbon materials, or semiconductor
materials are essential. Photolithography has been used for the fabrication of microstructures composed
of electrically conductive materials on the surface of or inside PDMS; however, the method requires
multiple steps for the fabrication [4–6]. Multi-photon photoreduction of metal ions induced by
femtosecond laser pulses enables the fabrication of metal structures on the surface of/or inside a soft
material [7,8]. A method based on the photoreduction of a metal ink was also reported to fabricate
metal structures on the surface of a soft material [9]. These methods require the additional doping of
electrically conductive materials, including metal nanoparticles and metal ions.
It is challenging to form a conductive structure by laser irradiation without doping an additional
material to polymers. Carbonization of polymers by laser irradiation enables the direct writing
of an electrically conductive structure on polymers [10,11]. Rahim et al. reported the spatially
selective fabrication of an electrically conductive carbon structure by the carbonization of a polyimide
by laser irradiation [10]. The carbon structures fabricated on the surface of the polyimide were
transferred to the surface of PDMS to fabricate a PDMS-based stretchable strain sensor. For direct

Nanomaterials 2018, 8, 558; doi:10.3390/nano8070558 64 www.mdpi.com/journal/nanomaterials


Nanomaterials 2018, 8, 558

modification of PDMS, the formation of carbon materials by irradiating 800-nm femtosecond laser
pulses [12] or ultraviolet nanosecond laser pulses [13] to PDMS was reported. A limited number
of papers reported the formation of semiconductor structures on the surface of or inside PDMS
using a laser. By the irradiation of 527-nm or 1064-nm femtosecond laser pulses [12,14] or 532-nm or
1064-nm nanosecond laser pulses [13,14], c-silicon was formed on the surface of PDMS by chemical
modification. However, the fabrication of electrically conductive structures on the surface of PDMS
by direct modification of PDMS using a laser has not been reported, despite the demands for various
PDMS-based electrical devices.
In this study, we present the formation of electrically conductive structures on the surface of PDMS
by irradiation with femtosecond laser pulses. The electrical conductivity of the formed structures is
measured. Analytical Fourier-transform infrared (FTIR) spectroscopy and X-ray diffraction (XRD)
results showed that the formed structures were composed of β-silicon carbide (β-SiC). To the best of
our knowledge, this is the first demonstration of the direct modification of native PDMS to a conductive
material composed of SiC. The presented method enables a direct fabrication of conductive lines in
a biocompatible polymer with potential applications in microelectromechanical systems (MEMS)
and electro-bioimplants.

2. Materials and Methods


Liquid photo-curable PDMS (KER-4690A/B, Shin-Etsu Chemical Co., Ltd., Tokyo, Japan) in a
mold was illuminated using an ultraviolet lamp at a wavelength of 365 nm for 30 min. The polymerized
PDMS was rinsed with ethanol, which prevented the adhesion of PDMS to a cover glass on which
PDMS was placed during the laser irradiation. Laser pulses with a central wavelength of 522 nm (the
second harmonic wave of a 1045-nm femtosecond laser (High Q-2, Spectra-Physics, Santa Clara, CA,
USA)), a pulse duration of 192 fs, and a repetition rate of 63 MHz were used for laser direct writing.
Femtosecond laser pulses focused by an objective lens (numerical aperture (NA) of 0.4, Olympus,
Tokyo, Japan) were irradiated to the lower surface of the PDMS from the bottom in air. The lower
surface of the PDMS had contact with the cover glass. The beam diameter d at the focal point was
assumed to be ~1.6 μm, according to the formula d = 1.22λ/NA. Laser power used for proof-of-concept
experiments was 150 mW. Using a xyz-translation stage, the samples were scanned on the xy-plane.
In the scanned area, adjacent lines were sufficiently overlapped so that the scanned area was assumed
to be fully modified in the area. In experiments for the characterization of structures formed under
different irradiation conditions, laser pulses were irradiated to the lower surface of the PDMS with a
140-μm air-gap between the surface of PDMS and the surface of the cover glass, which was performed
in order to exclude the effect of the cover glass on the fabrication of structures on the lower surface
of the PDMS. For this experiment, laser power was varied from 70 mW to 350 mW. The fabrication
process was monitored in real time with a CMOS camera (Thorlabs, Newton, NJ, USA).
The structures formed by laser irradiation were observed by optical microscopy and scanning
electron microscopy (SEM, Inspect F50, FEI, Hillsboro, OR, USA). Also, the formed structures were
examined with FTIR spectroscopy (ALPHA-E, Bruker, Billerica) and XRD (D8 Discover, Bruker,
Billerica, MA, USA). FTIR was performed for the wavenumbers, 400 to 4000 cm−1 . XRD was
performed for 2θ, 12.4◦ to 100◦ . For XRD, a generation voltage of 40 kV was used. Current–voltage
curves of the fabricated structures were obtained in the range from 0 to 10 V in 0.1 V steps by
two-terminal measurement using a digital source meter (2401, Keithley, Cleveland, OH, USA).
Probes were set 6 mm apart from each other on the surface of the fabricated structures for all the
experiments. Average resistance was determined by calculating the resistance at each voltage, using the
obtained measurements.

65
Nanomaterials 2018, 8, 558

3. Results and Discussion

3.1. Formation of Conductive Structures on PDMS by Femtosecond Laser Pulse Irradiation


Multiple line structures were fabricated with a line–to–line interval of 20 μm by focused
femtosecond laser pulses at a scanning speed of 2 mm/s in the x-direction. The lengths of the scanned
area were 5 mm in both x- and y-directions. A photographic image of the structure fabricated on the
surface of PDMS is shown in Figure 1a. The irradiated surface changed from optically transparent to
black-colored; however, no obvious laser ablation was visually observed. Microscale surface roughness
was observed on the black structure with SEM (Figure 1b). The direction of the observed ripple
structures corresponded to the scanning direction. The line-to-line interval, i.e., the scanning interval,
was 25 μm, which was comparable to the period of the formed ripple structures. The peak laser intensity
at the focal point under the experimental condition is estimated to be 6.2 × 1011 J/cm2 , which is lower
than the estimated peak intensity used for laser ablation of PDMS under the conditions of a laser
wavelength of 527 nm, pulse duration of 300 fs, and repetition rate of 33 Hz [12]. We performed the
experiment at a repetition rate of 63 MHz; therefore, the formation of the black structures was possibly
attributed to heat accumulation by the laser pulse train. The peak intensity in the present study is
comparable to the peak intensity that induced a change of the refractive index of PDMS under the
conditions of a laser wavelength of 800 nm, pulse duration of 130 fs, and repetition rate of 1 kHz,
reported in a previous study [15]. Therefore, the femtosecond laser pulse irradiation may induce the
scission of chemical bonds.

Figure 1. (a) Photographic image of the structure fabricated on the surface of native PDMS. Multiple
lines were fabricated with a line–to–line interval of 20 μm by moving the sample at a scanning speed of
2 mm/s in the x-direction. The laser power was 150 mW. The size of the scanned area was 5 mm ×
5 mm. (b) SEM image of the irradiated area on the PDMS surface. The white double-headed arrow
shows the scanning direction.

For the measurement of the electrical conductivity of black structures fabricated by irradiation
with femtosecond laser pulses, structures with lengths of 8 mm in the x-direction and 2 mm in the
y-direction were fabricated. The structures were fabricated at a scanning speed of 2 mm/s in the
x-direction. Multiple lines with a line–to–line interval of 25 μm were fabricated, which was sufficient
to obtain fully-overlapped black structures. Figure 2 shows the current–voltage curve of the structures.
Probes were set 6 mm apart from each other on the surface of the black structures. The current
increased linearly with the applied voltage. The average resistance was calculated to be 4.8 kΩ.
By assuming that the line structures have a half-circle shape in the cross-section and that they are
partially overlapped, the volume of the structure and resistivity are estimated to be 4.0 × 10−5 cm3
and 5.3 Ω cm, respectively. These results clearly demonstrate that PDMS was modified to electrically
conductive material by laser irradiation.

66
Nanomaterials 2018, 8, 558

Figure 2. Current–voltage curve of the structures fabricated by laser pulse irradiation. Multiple lines
were formed with a line–to–line interval of 25 μm by moving the sample at a scanning speed of 2 mm/s
in the x-direction. The size of the scanned area was 8 mm in the x-direction and 2 mm in the y-direction.
The laser power was 150 mW.

3.2. Analytical Results of FTIR Spectroscopy and XRD


To investigate the chemical composition of the black structures, FTIR spectroscopy was carried
out. Figure 3 shows the FTIR spectra of native PDMS (Figure 3a) and PDMS irradiated by femtosecond
laser pulses under the same laser conditions as those in Figure 1 (Figure 3b). In the spectrum of the
native PDMS, sharp peaks corresponding to C–H (2950 and 2900 cm−1 ), CH2 deformation (1400 cm−1 ),
Si–O (1080 cm−1 ), Si–CH3 rocking (820 cm−1 ), and Si–O–Si deformation (460 cm−1 ) were observed.
In the spectrum of the PDMS irradiated by laser pulses, wide peaks of Si–O (1080 cm−1 ), Si–CH3
rocking (820 cm−1 ), and Si–O–Si deformation (460 cm−1 ) were observed; no sharp peak was observed
(Figure 3b). The peaks of C–H (2950 and 2900 cm−1 ) and CH2 deformation (1400 cm−1 ), which are
typical bonds between carbon and hydrogen, were not observed. The disappearance of the peaks
corresponding to C–H (2950 and 2900 cm−1 ) and CH2 deformation (1400 cm−1 ) suggests that scission
of the corresponding chemical bonds was induced by the laser pulse irradiation, leading to the
release of carbon and hydrogen in the bonds as gaseous species, including hydrocarbon gas and CO2
gas [16]. Typical peaks of a carbon material, the D band (1350 cm−1 ) and G band (1598 cm−1 ) [17],
were not observed, showing that the formation of carbon materials is negligible in this study. On the
other hand, a peak corresponding to Si–O (1080 cm−1 ) was observed after the laser pulse irradiation,
which indicates the possible formation of SiO2 and SiO.

Figure 3. FTIR spectra of (a) native PDMS and (b) PDMS irradiated by femtosecond laser pulses at 150 mW.

67
Nanomaterials 2018, 8, 558

In order to identify the material of the black structures, XRD analyses were performed. Figure 4
shows XRD patterns of the native PDMS (Figure 4a) and PDMS irradiated by femtosecond laser pulses
(Figure 4b). The laser conditions correspond to the case of Figure 1. For the native PDMS, no significant
diffraction peak was observed. On the other hand, diffraction peaks were observed around 2θ = 36◦ ,
60◦ , and 72◦ for the PDMS irradiated by laser pulses (Figure 4b). These peaks correspond to the (111),
(220), and (311) diffraction planes of crystalline β-SiC, demonstrating the laser direct modification
from native PDMS to SiC.

Figure 4. XRD patterns of (a) native PDMS and (b) PDMS irradiated by femtosecond laser pulses at 150 mW.

The estimated resistivity of the black structures, 5.3 Ω cm (Figure 2), is approximately 40 times
larger than the resistivity of bulk SiC, 0.13 Ω cm [18]. The lower electrical conductivity is attributed
to the roughness of the formed structures, as well as the possible formation of secondary products,
including silicon oxides and silicon carbide oxides. Thermal annealing, including laser-based methods,
would improve the electrical conductivity of the formed SiC structures.
Scission of chemical bonds and crystallization are necessary for the formation of β-SiC from PDMS
or other siloxanes. Thermal treatments at appropriate temperatures such as oven annealing are reported
to be necessary to form SiC [19,20]. It has been widely recognized that consecutive femtosecond laser
pulses with a high repetition rate induce heat accumulation [21]. In this study, we used a femtosecond
laser pulse train at a repetition rate of 63 MHz, which is considered to be effective in increasing the
temperature by heat accumulation on the surface of native PDMS. The formation of SiC from native
PDMS by laser irradiation has not been reported previously; however, the formation of β-SiC by laser
irradiation of polycarbosilane (PCS) by the sintering of PCS powders using a CO2 laser [22] or by the
pyrolysis of PCS thin films on Si and SiO2 substrates using a millisecond pulsed laser [23] has been
reported. The formation of carbon materials by the irradiation of PDMS (pyrolysis of PDMS using
an infrared continuous wave (CW) laser [17]) has also been reported by other groups. In addition
to the thermal effect, the photochemical effect induced by the non-linear optical interaction with the
high-intensity femtosecond laser pulse should be considered when discussing the scission of chemical
bonds and crystallization in this study. For example, the wavelengths of 800 nm and 400 nm led to
different results of bond scission with the femtosecond laser ablation, although the sizes of the laser
ablation craters were comparable [24].

3.3. Characterization of Structures Formed under Different Irradiation Conditions


Investigation of parameter dependence on the formation of β-SiC by the irradiation of
femtosecond laser pulses to native PDMS is crucial for the elucidation of its formation mechanism,
leading to the practical application of the presented technique. Figure 5 shows the XRD patterns of
structures fabricated by laser pulse irradiation to native PDMS under different irradiation conditions.
Note that the XRD analyses shown in Figure 5 were performed for structures formed on PDMS
with a 140-μm air-gap between the surface of PDMS and the surface of the cover glass. This setup
was performed in order to clarify that Si, required for the formation of SiC, is derived not from the

68
Nanomaterials 2018, 8, 558

cover glass, but from the PDMS. Under all irradiation conditions, black structures, similar to the
structures formed when the PDMS was directly placed on the cover glass, were formed. In Figure 5a–c,
the scanning speed was fixed at 2 mm/s and laser power was varied. Diffraction peaks of β-SiC
were weak for 70 mW (Figure 5a), while peaks were clear for 150 mW (Figure 5b) and 250 mW
(Figure 5c). When laser power was fixed at 150 mW (Figure 5d–f), diffraction peaks of β-SiC were
weak for 0.5 mm/s (Figure 5d), while peaks were clear for 1 mm/s (Figure 5e) and 5 mm/s (Figure 5f).
The weak diffraction peaks of β-SiC for a lower scanning speed indicate that the exceeding laser pulses
may have induced the modification of formed β-SiC to amorphous SiC or other materials.

Figure 5. XRD patterns of structures fabricated by laser pulse irradiation to native PDMS under
different irradiation conditions. (a–c) had a fixed scanning speed of 2 mm/s. Laser power was 70 mW
(a), 150 mW (b), and 250 mW (c), respectively. (d–f) had a fixed laser power of 150 mW. Scanning speed
was 0.5 mm/s (d), 1 mm/s (e), and 5 mm/s (f), respectively.

Figure 6 shows SEM images of structures fabricated by laser pulse irradiation to native PDMS
under the same irradiation conditions as Figure 5, respectively. Under the condition of 70 mW at
2 mm/s, where diffraction peaks of β-SiC were weak (Figure 5a), structures aligned along the scanning
direction were observed (Figure 6a). Since the laser power was low, it is possible that unmodified
PDMS remained and/or that modification might not have been sufficient to form β-SiC. For 150 and
250 mW at 2 mm/s, structures aligned along the scanning direction were also observed, but partially
connected (Figure 6b,c). Since the enhanced optical field is generated between adjacent structures
on a sub-micro scale [25], localized melting possibly occurred by succeeding laser pulses at the area
above the melting temperature of SiC [26] to connect structures. In Figure 6d–f, laser power was
fixed at 150 mW, and scanning speed was varied. For 150 mW at 0.5 mm/s, bumps as well as cracks
were observed (Figure 6d). The number of pulses per beam spot, i.e., pulse overlap, is calculated to
be 2 × 105 pulses at the scanning speed of 0.5 mm/s. For 1 mm/s and 5 mm/s, the surfaces were
comparably smooth (Figure 6e,f).
Figure 7 shows the laser power dependence of average resistance for formed structures (Figure 7a)
and scanning speed dependence of average resistance for formed structures (Figure 7b). Probes were
set 6 mm apart from each other on the surface of fabricated SiC structures. A voltage range of
0 to 10 V was applied and resistance was measured in 0.1 V steps. The average resistance was
plotted in the figure. As shown in Figure 7a, average resistance varied with changes in laser power.
Among the presented conditions, the lowest resistance was obtained at 150 mW. For higher laser
powers, cracks were formed on the surface of SiC structures (Figure 6), which could increase the
resistance of the structures. When the laser power was fixed at 150 mW, the lowest resistance was
obtained at 2 mm/s. Resistance for scanning speed of 0.5 mm/s was not plotted since hardly any

69
Nanomaterials 2018, 8, 558

current flowed. With slow scanning speeds, 0.5 mm/s and 1 mm/s, the formation of cracks (Figure 6d,e)
and modification of formed β-SiC, possibly due to excessive heat effects, occur, which could decrease
the conductivity. For 4 mm/s and 5 mm/s, average resistances of 262.5 and 740.9 kΩ, respectively,
were measured. The XRD result shows that the formation of β-SiC is possible even at a scanning
speed of 5 mm/s (Figure 5). However, the resistances were higher for the cases of 4 mm/s and
5 mm/s compared to the case of 2 mm/s. Melting of the fabricated structures, which could improve
the electrical connection between the formed β-SiC, was assumed to be less than the case for lower
scanning speeds, resulting in an increase in resistance. Since the width of the SiC structure depends
on the laser power and the scanning speed, further optimization of the scanning interval would be
effective in improving the conductivity, as well as forming uniform structures.

Figure 6. SEM images of SiC surface fabricated on PDMS under different irradiation conditions.
(a–c) had a fixed scanning speed of 2 mm/s. Laser power was 70 mW (a), 150 mW (b) and 250 mW (c).
(d–f) had a fixed laser power of 150 mW. Scanning speed was 0.5 mm/s (d), 1 mm/s (e) and 5 mm/s (f).

Figure 7. (a) Laser power dependence on average resistance of formed structures. Scanning speed was
2 mm/s. (b) Scanning speed dependence on average resistance of formed structures. Laser power was
150 mW. Hardly any current flowed at 0.5 mm/s.

4. Conclusions
In conclusion, the black structures were formed by the irradiation of femtosecond laser pulses to
native PDMS. The formed black structures exhibit electrical conductivity. The FTIR and XRD results
showed that the formed black structures were composed of β-SiC. To the best of our knowledge, this is
the first demonstration of the direct modification of native PDMS to SiC only by laser irradiation.
The technique we present enables simple direct writing of conductive lines on the surface of PDMS,
which can be utilized for the fabrication of flexible/stretchable devices.

70
Nanomaterials 2018, 8, 558

Author Contributions: Y.N. and S.H. performed experiments. Y.N., S.H., and A.K. prepared the setups for the
experiments and measurements. Y.N., N.N., and M.T. designed the study and the project. Y.N. and M.T. wrote the
main text of the manuscript. All authors contributed to the discussion and reviewed the manuscript.

Funding: This study was partially supported by a grant from the Amada Foundation. Y.N. is grateful
for a Grant-in-Aid for Research Fellow of the Japan Society for the Promotion of Science (JSPS).

Conflicts of Interest: The authors declare no conflict of interest.

References
1. Isiksacan, Z.; Guler, M.T.; Aydogdu, B.; Bilican, I.; Elbuken, C. Rapid fabrication of microfluidic PDMS
devices from reusable PDMS molds using laser ablation. J. Micromech. Microeng. 2016, 26, 035008. [CrossRef]
2. Sekitani, T.; Nakajima, H.; Maeda, H.; Fukushima, T.; Aida, T.; Hata, K.; Someya, T. Stretchable active-matrix
organic light-emitting diode display using printable elastic conductors. Nat. Mater. 2009, 8, 494–499.
[CrossRef] [PubMed]
3. Park, H.; Jeong, Y.R.; Yun, J.; Hong, S.Y.; Jin, S.; Lee, S.J.; Zi, G.; Ha, J.S. Stretchable Array of Highly Sensitive
Pressure Sensors Consisting of Polyaniline Nanofibers and Au-Coated Polydimethylsiloxane Micropillars.
ACS Nano 2015, 9, 9974–9985. [CrossRef] [PubMed]
4. Martinez, V.; Stauffer, F.; Adagunodo, M.O.; Forro, C.; Vörös, J.; Larmagnac, A. Stretchable
Silver Nanowire-Elastomer Composite Microelectrodes with Tailored Electrical Properties. ACS Appl.
Mater. Interfaces 2015, 7, 13467–13475. [CrossRef] [PubMed]
5. Guo, L.; Deweerth, S.P. An effective lift-off method for patterning high-density gold interconnects on an
elastomeric substrate. Small 2010, 6, 2847–2852. [CrossRef] [PubMed]
6. Gou, H.; Xu, J.; Xia, X.; Chen, H. Air Plasma Assisting Microcontact Deprinting and Printing for Gold thin
Film and PDMS Patterns. ACS Appl. Mater. Interfaces 2010, 2, 1324–1330. [CrossRef] [PubMed]
7. He, G.C.; Zheng, M.L.; Dong, X.Z.; Jin, F.; Liu, J.; Duan, X.M.; Zhao, Z.S. The Conductive Silver Nanowires
Fabricated by Two-beam Laser Direct Writing on the Flexible Sheet. Sci. Rep. 2017, 7, 41757. [CrossRef]
[PubMed]
8. Terakawa, M.; Torres-Mapa, M.L.; Takami, A.; Hienemann, D.; Nedyalkov, N.N.; Nakajima, Y.; Hordt, A.;
Ripken, T.; Hiesterkamp, A. Femtosecond laser direct writing of metal microstructure in a stretchable
poly(ethylene glycol) diacrylate (PEGDA) hydrogel. Opt. Lett. 2016, 41, 1392–1395. [CrossRef] [PubMed]
9. Liu, Y.; Lee, M. Laser Direct Synthesis and Patterning of Silver Nano/Microstructures on a Polymer Substrate.
ACS Appl. Mater. Interfaces 2014, 6, 14576–14582. [CrossRef] [PubMed]
10. Rahimi, R.; Ochoa, M.; Yu, W.; Ziaie, B. Highly stretchable and sensitive unidirectional strain sensor via laser
carbonization. ACS Appl. Mater. Interfaces 2015, 7, 4463–4470. [CrossRef] [PubMed]
11. Rahimi, R.; Ochoa, M.; Ziaie, B. Direct Laser Writing of Porous-Carbon/Silver Nanocomposite for Flexible
Electronics. ACS Appl. Mater. Interfaces 2016, 8, 16907–16913. [CrossRef] [PubMed]
12. Atanasov, P.A.; Stankova, N.E.; Nedyalkov, N.N.; Fukata, N.; Hirsch, D.; Rauschenbach, B.; Amoruso, S.;
Wang, X.; Kolev, K.N.; Valova, E.I.; et al. Fs-laser processing of medical grade polydimethylsiloxane (PDMS).
Appl. Surf. Sci. 2016, 374, 229–234. [CrossRef]
13. Stankova, N.E.; Atanasov, P.A.; Nikov, R.G.; Nikov, R.G.; Nedyalkov, N.N.; Stoyanchov, T.R.; Fukata, N.;
Kolev, K.N.; Valova, E.I.; Georgieva, J.S.; et al. Optical properties of polydimethylsiloxane (PDMS) during
nanosecond laser processing. Appl. Surf. Sci. 2016, 374, 96–103. [CrossRef]
14. Stankova, N.E.; Atanasov, P.A.; Nedyalkov, N.N.; Stoyanchov, T.R.; Kolev, K.N.; Valova, E.I.; Georgieva, J.S.;
Armyanov, S.A.; Amoruso, S.; Wang, X.; et al. Fs- and ns-laser processing of polydimethylsiloxane (PDMS)
elastomer: Comparative study. Appl. Surf. Sci. 2015, 336, 321–328. [CrossRef]
15. Cho, S.H.; Chang, W.S.; Kim, K.R.; Hong, J.W. Femtosecond laser embedded grating micromachining of
flexible PDMS plates. Opt. Commun. 2009, 282, 1317–1321. [CrossRef]
16. Hasegawa, Y.; Iimura, M.; Yajima, S. Synthesis of continuous silicon carbide fibre—Part 2 Conversion of
polycarbosilane fibre into silicon carbide fibres. J. Mater. Sci. 1980, 15, 720–728. [CrossRef]
17. Alcántara, J.C.C.; Zorrilla, M.C.; Cabriales, L.; Rossano, L.M.L.; Hautefeuille, M. Low-cost formation of bulk
and localized polymer-derived carbon nanodomains from polydimethylsiloxane. Beilstein J. Nanotechnol.
2015, 6, 744–748. [CrossRef] [PubMed]

71
Nanomaterials 2018, 8, 558

18. Samsonov, G.V. Plenum Press Handbooks of High Temperature Materials No. 2 Properties Index; Plenum Press:
New York, NY, USA, 1965; pp. 1–635, ISBN 978-1-4899-6405-2.
19. Al-Ajrah, S.; Lafdi, K.; Liu, Y.; Le Coustumer, P. Fabrication of ceramic nanofibers using polydimethylsiloxane
and polyacrylonitrile polymer blends. J. Appl. Polym. Sci. 2018, 135, 45967. [CrossRef]
20. Burns, G.T.; Taylor, R.B.; Xu, Y.; Zangvil, A.; Zank, G.A. High-Temperature Chemistry of the Conversion of
Siloxanes to Silicon Carbide. Chem. Mater. 1992, 4, 1313–1323. [CrossRef]
21. Eaton, S.; Zhang, H.; Herman, P.; Yoshino, F.; Shah, L.; Bovatsek, J.; Arai, A. Heat accumulation effects
in femtosecond laser-written waveguides with variable repetition rate. Opt. Express 2005, 13, 4708–4716.
[CrossRef] [PubMed]
22. Jakubenas, K.; Marcus, H.L. Silicon Carbide from laser Pyrolysis of Polycarbosilane. J. Am. Chem. Soc. 1995,
2263–2266. [CrossRef]
23. Colombo, P.; Martucci, A.; Fogato, O.; Villoresi, P. Silicon Carbide Films by Laser Pyrolysis of Polycarbosilane.
J. Am. Ceram. Soc. 2001, 26, 224–226. [CrossRef]
24. Shibata, A.; Machida, M.; Kondo, N.; Terakawa, M. Biodegradability of poly(lactic-co-glycolic acid) and
poly(l-lactic acid) after deep-ultraviolet femtosecond and nanosecond laser irradiation. Appl. Phys. A 2017,
123, 438. [CrossRef]
25. Terakawa, M.; Nedyalkov, N.N. Near-field optics for nanoprocessing. Adv. Opt. Technol. 2016, 5, 17–28.
[CrossRef]
26. Ashwath, P.; Xavior, M.A. Processing methods and property evaluation of Al2 O3 and SiC reinforced metal
matrix composites based on aluminium 2xxx alloys. J. Mater. Res. 2016, 31, 1201–1219. [CrossRef]

© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

72
nanomaterials

Article
Extreme Energy Density Confined Inside a
Transparent Crystal: Status and Perspectives
of Solid-Plasma-Solid Transformations
Eugene G. Gamaly 1, *, Saulius Juodkazis 2 and Andrei V. Rode 1, *
1 Laser Physics Centre, Research School of Physics and Engineering, The Australian National University,
Canberra ACT 2601, Australia
2 Centre for Micro-Photonics, Swinburne University of Technology, Hawthorn VIC 3122, Australia;
[email protected]
* Correspondence: [email protected] (E.G.G.); [email protected] (A.V.R.);
Tel.: +61-2-6125-8659 (E.G.G.); +61-2-6125-4637 (A.V.R.)

Received: 26 June 2018; Accepted: 19 July 2018; Published: 21 July 2018

Abstract: It was demonstrated during the past decade that an ultra-short intense laser pulse
tightly-focused deep inside a transparent dielectric generates an energy density in excess of several
MJ/cm3 . Such an energy concentration with extremely high heating and fast quenching rates leads
to unusual solid-plasma-solid transformation paths, overcoming kinetic barriers to the formation
of previously unknown high-pressure material phases, which are preserved in the surrounding
pristine crystal. These results were obtained with a pulse of a Gaussian shape in space and in time.
Recently, it has been shown that the Bessel-shaped pulse could transform a much larger amount of
material and allegedly create even higher energy density than what was achieved with the Gaussian
beam (GB) pulses. Here, we present a succinct review of previous results and discuss the possible
routes for achieving higher energy density employing the Bessel beam (BB) pulses and take advantage
of their unique properties.

Keywords: light-matter interaction; ultra-short laser pulses; high-pressure/density conditions;


phase transitions

1. Microexplosion Studies with Gauss-Shaped Beam


The studies of confined microexplosions during the last decade revealed the major features of
this complicated phenomenon where the processes of electro-magnetic field/dielectric interaction,
plasma formation and high-pressure hydrodynamics are intertwined. The concise description of these
processes is as follows. The tight focusing of the laser beam deep inside a transparent crystal allows
achieving the absorbed energy density in excess of the strength of any material in a sub-micron volume
surrounded by the pristine solid. After energy transfer from hot electrons to ions, the expanding strong
shock wave accompanied by the rarefaction wave starts propagating outside of this volume. After the
shock decelerating and stopping, the void, surrounded by a shell of compressed and pressure modified
material converted to the novel phases, is formed. All transformed material remains confined inside
the bulk of undamaged material ready for further studies. These studies employed the short intense
laser beam with the Gaussian spatial and temporal intensity profile [1–4].
The short intense laser pulse with the Gaussian spatial and temporal intensity profiles tightly
focussed inside a transparent crystal generates an energy density of several MJ/cm3 . The pressure
produced is in excess of a few TPa, which is higher than the strength of any existing material
(diamond has the highest Young’s modulus of 1 TPa = 1 MJ/cm3 ). The laser pulse, 150 fs, 100–200 nJ,
800 nm, tightly-focussed inside sapphire with a microscope lens (N A = 1.4) creates the solid

Nanomaterials 2018, 8, 555; doi:10.3390/nano8070555 73 www.mdpi.com/journal/nanomaterials


Nanomaterials 2018, 8, 555

density plasma at the temperature of a few tens of electron Volts (∼5 × 105 K) with the record-high
heating rate of 1018 K/s [1,2]. It was found that the novel (previously unobserved) high-pressure
phases of aluminium and silicon were formed [3,4] following the ultrashort laser-induced confined
microexplosion. Pressure/temperature conditions created in the microexplosion are similar to those
in hot cores of stars and planets (“primeval soup” or warm dense matter). The material converted to
high pressure/temperature solid density plasma is then transformed into the novel solid phase during
the ultra-fast cooling and re-structuring. The major difference from the core-star conditions is the
record-fast cooling (∼1016 K/s) from plasma state to solid state. In the previous experiments, the study
of the pressure-affected materials was produced postmortem, well after the end of the pulse when
transformed material was cooled down to the ambient conditions. The structure of laser-transformed
material was determined by the synchrotron X-ray diffraction [3] and with the electron diffraction [4]
in transmission electron microscopy (TEM) [4].

1.1. Novelty of the Phase Transformation Path during and after Confined Microexplosion
The solid transforms to a solid-density plasma state (Te ∼ 50 eV) during a pulse time shorter
than all energy relaxation times. A strong shock wave (SW) starts propagating from the energy
deposition region several picoseconds after the pulse due to energy transfer from electrons to the
ions. The shock wave decelerates and converts into a sound wave in the surrounding cold pristine
crystal. The phenomenon is similar, but not identical to an underground nuclear explosion: the
massless energy carriers (photons) deliver the energy inside a transparent crystal without changing
the atomic and mass content of a material. All laser-affected material is expelled from the energy
deposition area by the combined action of shock and rarefaction waves, forming a void surrounded
by the shell of material compressed against the surrounding cold pristine crystal. The material
returns from the high-pressure plasma state (high entropy, chaotic) to the ambient conditions at room
temperature/pressure, however attaining a phase state different from the initial solid state. In all
known methods of high pressure phase formation, the initial crystalline structure is re-structured,
i.e., the atoms are moved from the initial arrangement to the new positions under the action of high
pressure. During the transformation path under confined microexplosion, the initial state of a crystal
is completely destroyed and forgotten. The irradiated material is converted into a chaotic mixture
of ions and electrons at high temperature. Therefore relaxation to the ambient conditions occurs
along the unknown paths going through the metastable intermediate equilibrium potential minima.
The theoretical (computational, modified DFT-studies) during the last decade searched for the possible
paths of material transformations under high pressure from the initially chaotic (stochastic) state [5].
These studies uncovered many physically allowed paths for the formation of multiple novel phases
(including incommensurable phases) from the initially chaotic state. The confined microexplosion
method now is the only practically realised way for the formation of novel material phases from
the plasma state, preserving the transformed material confined inside the pristine crystal for further
structural studies.

1.2. Limitations of the Confined Microexplosion Method with the GB


There are limitations to the energy density and amount of laser-affected material in confined
microexplosion generated by the tightly-focused Gauss beam. The main limitation is imposed by
diffraction: the radius of the diffraction-limited focal spot is [2,6]: r Airy = 0.61λ/N A; which defines
the central Airy disk at the focus. For λ = 800 nm and N A = 1.4, one gets r f oc = 0.35 μm and a
focal area of 0.38 μm2 . The absorption length in dense plasma equals ∼30 nm, giving the energy
deposition volume ∼10−14 cm−3 . With absorbed energy around 100 nJ, the absorbed energy density
amounts to 107 J/cm3 = 10 TPa. The number of laser-affected atoms constitutes around 1011 atoms
(a few picograms), making structural studies extremely difficult. Therefore, the questions arises: is
it possible to increase the absorbed energy density and/or increase the amount of the laser-affected
material and thus the amount of the novel phase? Preliminary studies have shown that it is very

74
Nanomaterials 2018, 8, 555

difficult to overcome the energy density of several MJ/cm3 (several TPa of pressure) and increase
the amount of laser-affected material using a tightly-focused Gauss beam. First, the ionisation wave
moving towards the laser pulse with increasing intensity increases the absorbing volume and limits
the energy density [7]. Moreover, the experiments with increasing laser pulse energy demonstrated
that at the energy per 150-fs pulse of 200 nJ, the cracks surrounding the focal area destroyed the regular
void formation [2]. Diffraction-free Bessel beams (BB) raised the hope of achieving a higher energy
density and larger amounts of the material affected. Below, we describe the recent progress made with
these studies. Then, we describe some effects (and unresolved problems), the solutions of which may
lead to a further increase of the absorbed energy density.

2. Status of the BB-Transparent Crystal Interactions


It was demonstrated recently that the BB (150 fs, 2 μJ) focused inside sapphire produced a
cylindrical void of 30 μm in length and 300 nm in diameter [8]. The void volume, V = 30 μm × πr2 =
2.12 × 10−12 cm−3 , appears to be two orders of magnitude larger than that generated by the GB.
The conclusions based solely on the void size measurements and on the energy and mass conservation
laws without any ad hoc assumptions about the interaction process are the following [9]. The material
initially filling the void was expelled and compressed into a shell by the high-pressure shock wave.
The work necessary to remove material with the Young modulus Y from volume V equals at least
Y × V = 0.848 μJ (Y = 4 × 105 J/cm3 , the Young modulus of sapphire). This is evidence of strong
(>40%) absorption of the pulse energy. In order to generate a strong shock wave capable of expelling
such an amount of material, the absorbed energy should be concentrated in the central spike with
a much smaller diameter than that of the void (the absorbed energy density is still not known,
theoretically nor experimentally).
The unique features of the diffraction-free Bessel beam spatial distribution of intensity in the
focal area allow one to understand some of the experimental findings and indicate new problems and
opportunities. The spatial distribution of intensity across the cylindrical focal volume in a transparent
medium unaffected by light and observed experimentally is close to Durnin’s solution, J02 (kr r ) [10]: the
central spike surrounded by circular bands with the maximum of intensity on the axis approximately
five-times higher than in the next band.
The parameters of the quasi non-diffracting BB created by any device from the incoming cw-laser
pulse in air (axicon, circular slit, spatial light modulator (SLM), etc.) with the cone angle θ are the
following (Figure 1): radius of the incoming beam before the BB-creating device, R; zmax = R/ tan(θ );
kr = k × sin(θ ); k z = k × cos(θ ) [10]. Building the intensity distribution in the low intensity short
pulse BB occurs in a similar way to that as with the cw-laser, as was demonstrated experimentally [8].
It is worth noting that in these experiments, during the pulse time, the beam propagates a distance
comparable to the length of the elongated cylindrical focus (Zmax ), L pulse = t pulse × c/n (n is the
refractive index in a transparent crystal unaffected by a laser). For example, a 150-fs (800 nm) low
intensity pulse in sapphire propagates ∼30 microns, which is close to t pulse × c/n = 30 μm [8].

75
Nanomaterials 2018, 8, 555

Spherical explosion

Gaussian beam
E
k z
void

Cylindrical explosion
E
kz kr
Bessel beam

k
void
id 4

Figure 1. Schematic presentation of radial microexplosion-driven inside a transparent material by a


focused linearly-polarised (E-field) Gaussian beam (GB) and Bessel beam (BB)with projection Ez along
the optical axis (z-axis); θ is the angle with the optical axis (wavevector k = k2r + k2z is shown on the
upper half of the conical wave). Resonant absorption of the Er component (along the radial direction
kr ) allows a higher energy density deposition for the cylindrical microexplosion. With the central void
diameter in sapphire comparable for the GB [1] and BB [8] pulses, the volume was 180-times larger in
the case of BB.

In the non-absorbing media, the length of the focus (the distance where diffraction is strongly
suppressed) apparently is independent of the pulse duration. On the axis of the BB (in the focal area),
the time of the interaction of the electromagnetic wave with matter could be shorter than the pulse
duration. Therefore, a beam of any duration allegedly propagates the same distance Zmax allowed by
the focusing device. This seemingly obvious statement should be confirmed experimentally.
Under the action of intense pulse, the ionization breakdown occurs early in the pulse time near
the central spike where intensity is maximum. The studies of the interaction process of intense BB at an
intensity above the ionization threshold are absent to the best of our knowledge. Estimates, suggestions
and problems relating to the formation of the intensity distribution and interaction process based
on the studies of confined microexplosion and intense short pulse interactions with dielectrics are
presented below. Experimental observation of the void formation by GB and BB pulses is shown in
Figure 2. The BB pulses are used to dice transparent materials [11] and to inscribe high efficiency
optical gratings in silica [12].

76
Nanomaterials 2018, 8, 555

D IVSXOVH *DXVVLDQ IVSXOVH %HVVHO E


$ 2
$O
$ 2
$O
HQWUDQFH

aQP

YRLG DPRUSKRXV

YRLG
PLGGOH

YRLG
aQP

YRLG H[LW

PP QP

Figure 2. SEM side-view images of the voids made with ultra-short Gaussian [1,13] (a) and Bessel [8]
(b) single pulses in sapphire. Focusing of the 800-nm/130-fs Gaussian pulses of ∼150 nJ of energy was
carried out with an objective lens of numerical aperture N A = 1.3 and was stacked into a vertical plane
of the void-structures [13]. This plane was used to split the sapphire sample for the side-view SEM
observation. The voids made at larger depth were affected by spherical aberration, which reduced the
void and elongated amorphous region. The 800-nm/140-fs, 2 μJ of energy, Bessel pulses were used to
make cylindrical voids of a diameter of ∼ 300 nm revealed by focussed ion beam (FIB) milling [8].

Control of Energy Deposition by BB Pulses


In short intense BB interaction with the initially transparent medium, the ionization threshold
is reached at the axis of the focal volume where intensity attains the maximum value. It occurs early
in the pulse time close to the beginning of the elongated focal region. A narrow cylindrical plasma
region is created along the axis. Incident light starts absorbing in plasma. Let us take the incident field
structure near the axis as the following E (Er , E ϕ = 0; Ez ); H (0; H ϕ , 0). Then, the Poynting vector
reads, S = 4πc
(E × H). Therefore, the energy flows are generated inward along the radius and along
the z-axis in direction of the beam propagation: Sr = 4π c
( H ϕ · Ez ) and Sz = 4π
c
( H ϕ · Er ) [J/(cm2 s)].
Thus, by changing the cone angle, one can control the radial and axial energy flows.
The interaction mode of intense BB with a transparent crystal dramatically changes after the ionization
threshold is achieved. The surface, where the real part of the permittivity is zero, ε re = 0, separates
the dielectric (ε re > 0) and plasma (ε re < 0) regions. The gradient of the permittivity is directed along
the radius of a cylinder. The energy flow goes inward in the radial direction. Thus, the incident
wave splits into the evanescent and reflection waves. The resonance absorption occurs in the vicinity
of the zero-epsilon surface, creating a plasma wave (plasmon) propagating along the radius in the
direction to the axis of the cylindrical focal region. The evanescent wave decays along the radius in
the same direction. Thus, the zero-permittivity surface generates simultaneously coherent plasmons
and evanescent waves coming together (focusing) to the axis of the cylindrical focal region. One may
expect that coupling of evanescent waves and plasmons also contributes to the increase of the intensity

77
Nanomaterials 2018, 8, 555

and energy density near the axis in a way similar to that discovered in the studies of extraordinary
optical transmission (EOT) through sub-wavelength hole arrays [14].
The ideal diffraction free beam is the monochromatic Bessel beam [10], created via superposition
of plane waves, the wave vectors of which are evenly distributed over the surface of a cone. The Bessel
function of the first kind zero order, J0 , is a sum of the Hankel functions of the first and second kind [15],
where the inward energy flow is balanced by the outward flow.
It was suggested [16] that the quasi diffractionless BB can be presented as the result of the
interference of two conical running Hankel beams, carrying equal amounts of energy towards and
outwards from the beam axis, yielding no net transversal energy flux in the BB. The interference
of two Hankel beams with different amplitudes creates unbalanced BB where the net radial energy
flux appears. Unbalancing creates the inward radial energy flux from the conical tails of the beam.
The study of stability in the frame of the non-linear Shrodinger equation (NLSE) equation revealed
that the Bessel-like solutions in pure Kerr media are unstable [17].
In the interaction of intense short pulse BB with a transparent dielectric at the intensity below
the ionization threshold, the BB apparently retains its balanced structure. After the plasma formation,
the energy flow directed inward to the axis is created due to absorption leading to destruction of this
balance. One may argue that after the ionization threshold, the Hankel function of the first kind might
be considered as an appropriate approximation of the field distribution near the axis of cylindrical
focus, being the exact solution of the Bessel equation describing the electric field increasing while
focusing. One may conjecture that the BB becomes unstable, tending to focus onto the cylindrical axis,
thus creating an energy density higher than a tightly-focused, but diffraction-limited Gaussian beam.
Let us now consider the relation between the pulse duration, absorption, focal region length
and laser-affected area length. In short intense BB interaction with the initially transparent medium,
the ionization threshold is reached at the axis of the focal volume where the intensity is maximum.
This occurs early in the pulse time close to the beginning of the elongated focal region. The intense pulse
converts the initially transparent material into strongly absorbing plasma practically at the moment of
its arrival at some space point. Therefore, the plasma region gradually increases along the axis as the
pulse proceeds until the end of the pulse. The last portion of light arrives after travelling through the
transparent crystal a distance t p × c/n. The laser-affected distance then reads Llas = (t p × c/n) cos θ
(θ is the half-cone angle; Figure 1). One can see now the difference between the BB-affected area in
a transparent medium (diffraction-free focus) and the laser-affected area in an intense short pulse
laser/crystal interaction. For sufficiently short pulses, the laser-affected area might be shorter than the
diffraction-free zone, Llas < Zmax . Thus, laser pulse duration might be another lever (along with the
cone angle) to control the energy deposition volume.
Experiments demonstrated that short intense BB could affect a much larger amount of material
producing solid-plasma-solid transformation (direct measurements) at allegedly a pressure of several
TPa (conclusions on the basis of the analysis of the experiments) [8,9]. J.Hu [18] measured the average
speed of the shock wave, vsw ≈ 60 km/s, during the cylindrical microexplosion, generated by the
BB in sapphire, by the pump-probe technique. The estimate of the driving pressure based on this
measurement, Psw = 0 v2sw = 14.4 TPa (0 is the initial mass density of sapphire), gives the direct
experimental evidence of the extreme energy density created by the BB in the focal volume.
There are indications from theoretical studies [9] that the originally stable diffraction-free BB
at high intensity in the presence of strong ionization nonlinearity may become unstable. Now, it is
difficult to conclude if this may happen in a way similar to the self-focusing instability with Kerr-like
non-linearity (rather, not because the paraxial approximation is invalid in this case) or similar to
the instability of two unbalanced Hankel beams, which seems more relevant to the case (again, the
ionization non-linearity should be accounted for).
The oblique incidence, inherent for the formation of the BB and long focus, implies the possibility
of the surface wave (plasmon) formation and propagation along the zero-real-permittivity surface
at the same time with the plasmon moving radially due to the resonance absorption. The plasma

78
Nanomaterials 2018, 8, 555

wave may converge to the axis, contributing to the increase in the absorbed energy density. One may
conjecture if it might be relevant for some kind of Langmuir collapse.
It would be crucially important to find the electric field distribution up to the central axis in
order to determine the absorbed energy density. This requires a solution of the Maxwell equations in
cylindrical geometry coupled to material equations accounting for the change in the permittivity
(electrons” number density and collision rate) in accord with the intensity in any space/time
point. This is a formidable task; however, it can be clearly formulated for the numerical solution.
Different approximations may also be discussed.
It was demonstrated experimentally [18] that the Bessel beam-induced microexplosion in sapphire,
producing open-ended channel, proceeds as an axial-symmetric cylindrical explosion, and a mass
conservation was experimentally validated [19]. Therefore, the direct theoretical modeling of the
cylindrical explosion after the energy deposition of the BB beam inside a narrow on-axis cylinder also
can be performed in the frame of two-temperature plasma hydrodynamics in cylindrical geometry in a
way similar to as was done with the Gauss beam in spherical geometry [2].

3. Conclusions and Outlook


In conclusion, we should state that further progress in achieving and steering the high energy
density strongly depends on the future pump-probe experiments, which will register with time/space
resolution the history of the BB-generated microexplosion, processes of returning to the ambient
state and new phases’ formation. It is worth showing the time and space scales for the succession of
events comprising such a history that might in some approximation be extracted from the previous
studies [2,4,9].
Let us suggest the BB, 2 μJ, 800 nm, 150 fs, impinges a sapphire crystal several tens of microns
thick, creating a focal region of ∼30 μm long at the ten microns depth from the outer surface of
a sample. The stages of successive transformations are the following; the time count starts at the
beginning of the pump pulse:

1. The low intensity stage before ionisation threshold lasts a few fs at the beginning of the pulse;
2. As the ionisation threshold is attained, the cylindrical plasma region is created at the axis of the
focal region with a diameter less than a micron. One should note that the full length of the focal
region of 30 μm is reached at the end of the pulse, assuming that light propagates as in unaffected
sapphire with a speed of c/n ∼ 2 × 1010 cm/s;
3. The cylinder diameter of the energy absorption region to the end of the pulse allegedly might be
around the doubled absorption length in a dense plasma ∼60 nm;
4. The shock wave is created after the energy transfer from electrons to ions in a 7–10-ps time span;
5. The shock wave propagates during another 4–6 ps until it is converted into the acoustic wave,
effectively stopped by the cold pressure of the crystal (∼Young modulus of sapphire). The void
surrounded by the shell of compressed material is formed by the rarefaction wave;
6. The thermal wave of conventional heat conduction spreads into the laser-unaffected crystal,
cooling the laser-affected area down to the ambient conditions during tens of nanoseconds.
The material re-structuring occurs most probably during Stages 5 and 6. The whole area affected
by the heat from the laser-heated region is a cylinder with a length of around 32–34 microns with
a diameter of about 2–4 micrometers.

Thus, the whole area affected by the shock and heat waves from the energy deposition region
is a cylinder 30 microns long and a few microns in diameter. The time span for the whole process of
material transformation is around tens of nanoseconds. The recent arrival of X-ray free electron lasers
(XFEL) with a pulse duration as short as 7–15 fs and a photon energy of 8–10 keV currently available at
EuroXFEL at DESY in Hamburg and in SACLA XFEL at Spring-8 at Riken Institute in Japan creates
new opportunities for uncovering the mechanism of the formation of the new states of matter. Up to
17-keV pulses expected in the near future at the SLAC National Acceleration Laboratory at Stanford.

79
Nanomaterials 2018, 8, 555

All these new sources or coherent ultra-short X-ray radiation will be used to uncover the processes
involved in formation of such unusual material states. For such experiments, the tailored axial intensity
distribution of the optical BB pulses can be prepared using diffraction optical elements [20], which can
be made with a central hole for the co-axial fs-optical-pump and fs-X-ray-probe. To conclude, a light
(or X-ray) probe with a sub-picosecond duration and sub-micron spatial resolution may shed light
on the unusual formation of novel high-pressure phases starting from the “primeval soup” (warm
dense matter) to the solid state at the ambient conditions, being preserved and confined inside a bulk
of pristine crystal ready for further structural studies [7].

Author Contributions: E.G.G. conceived the idea, A.V.R. and S.J. carried out experiments. E.G.G. wrote the first
draft. All authors edited the final version.

Funding: The Australian Research Council Discovery project DP170100131.

Acknowledgments: This research was supported by the Australian Government through the Australian Research
Council’s Discovery Projects scheme.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Juodkazis, S.; Nishimura, K.; Tanaka, S.; Misawa, H.; Gamaly, E.E.; Luther-Davies, B.; Hallo, L.; Nicolai, P.;
Tikhonchuk, V. Laser-induced microexplosion confined in the bulk of a sapphire crystal: Evidence of
multimegabar pressures. Phys. Rev. Lett. 2006, 96, 166101. [CrossRef] [PubMed]
2. Gamaly, E.E.; Juodkazis, S.; Nishimura, K.; Misawa, H.; Luther-Davies, B.; Hallo, L.; Nicolai, P.;
Tikhonchuk, V. Laser-matter interaction in a bulk of a transparent solid: confined microexplosion and void
formation. Phys. Rev. B 2006, 73, 214101. [CrossRef]
3. Vailionis, A.; Gamaly, E.G.; Mizeikis, V.; Yang, W.; Rode, A.; Juodkazis, S. Evidence of super-dense
Aluminum synthesized by ultra-fast microexplosion. Nat. Commun. 2011, 2, 445. [CrossRef] [PubMed]
4. Rapp, L.; Haberl, B.; Pickard, C.J.; Bradby, J.E.; Gamaly, E.G.; Williams, J.S.; Rode, A.V. Experimental
evidence of new tetragonal polymorphs of silicon formed through ultrafast laser-induced confined
microexplosion. Nat. Commun. 2015, 6, 7555. [CrossRef] [PubMed]
5. Shi, X.; He, C.; Pickard, C.J.; Tang, C.; Zhong, J. Stochastic generation of complex crystal structures
combining group and graph theory with application to carbon. Phys. Rev. B 2018, 97, 014104. [CrossRef]
6. Born, M.; Wolf, E. Principles of Optics; Cambridge University Press: Cambridge, UK, 2003.
7. Gamaly, E.G.; Rapp, L.; Roppo, V.; Juodkazis, S.; Rode, A.V. Generation of high energy density by fs-laser
induced confined microexplosion. New J. Phys. 2013, 15, 025018. [CrossRef]
8. Rapp, L.; Meyer, R.; Giust, R.; Furfaro, L.; Jacquot, M.; Lacourt, P.A.; Dudley, J.M.; Courvoisier, F.
High aspect ratio microexplosions in the bulk of sapphire generated by femtosecond Bessel beams. Sci. Rep.
2016, 6, 34286. [CrossRef] [PubMed]
9. Gamaly, E.G.; Rode, A.V.; Rapp, L.; Giust, R.; Furfaro, L.; Lacourt, P.A.; Dudley, J.M.; Courvoisier, F.;
Juodkazis, S. Interaction of the ultra-short Bessel beam with transparent dielectrics: Evidence of high-energy
concentration and multi-TPa pressure. arXiv 2017, arXiv:1708.07630.
10. Durnin, J.; Miceli, J.J.; Eberly, J.H. Diffraction-free beams. Phys. Rev. Lett. 1987, 58, 1499. [CrossRef]
[PubMed]
11. Marcinkevičius, A.; Juodkazis, S.; Matsuo, S.; Mizeikis, V.; Misawa, H. Application of Bessel Beams for
microfabrication of dielectrics by femtosecond laser. Jpn. J. Appl. Phys. 2001, 40, L1197–L1199. [CrossRef]
12. Mikutis, M.; Kudrius, T.; Šlekys, G.; Paipulas, D.; Juodkazis, S. High 90% efficiency Bragg gratings formed
in fused silica by femtosecond Gauss-Bessel laser beams. Opt. Mat. Express 2013, 11, 1862–1871. [CrossRef]
13. Buividas, R.; Gervinskas, G.; Tadich, A.; Cowie, B.; Mizeikis, V.; Vailionis, A.; de Ligny, D.; Gamaly, E.G.;
Rode, A.; Juodkazis, S. Phase transformation in laser-induced microexplosion in olivine (Fe,Mg)2 SiO4 .
Adv. Eng. Mater. 2014, 16, 767–773. [CrossRef]
14. Liu, H.; Lalanne, P. Microscopic theory of the extraordinary optical transmission. Nature 2008, 452, 728–731.
[CrossRef] [PubMed]
15. Morse, P.; Feshbach, H. Methods of Theoretical Physics; McGraw Hill: New York, NY, USA, 1953; Volume 1–2.

80
Nanomaterials 2018, 8, 555

16. Salo, J.; Fagerholm, J.; Friberg, A.T.; Salomaa, M.M. Unified description of nondiffracting X and Y waves.
Phys. Rev. E 2000, 62, 4261. [CrossRef]
17. Porras, M.A.; Parola, A.; Faccio, D.; Dubietis, A.; Trapani, P.D. Nonlinear Unbalanced Bessel Beams:
Stationary Conical Waves Supported by Nonlinear Losses. Phys. Rev. Lett. 2004, 93, 153902. [CrossRef]
[PubMed]
18. Hu, J. High throughput micro/nano manufacturing by femtosecond laser temporal pulse shaping.
In Proceedings of the 9th International Conference on Information Optics and Photonics, Harbin, China,
17–20 July 2017.
19. Wang, G.; Yu, Y.; Jiang, L.; Li, X.; Xie, Q.; Lu, Y. Cylindrical shockwave-induced compression mechanism in
femtosecond laser Bessel pulse micro-drilling of PMMA. Appl. Phys. Lett. 2017, 110, 161907. [CrossRef]
20. Dharmavarapu, R.; Bhattacharya, S.; Juodkazis, S. Diffractive optics for axial intensity shaping of Bessel
beams. J. Opt. 2018, 20, 085606. [CrossRef]

c 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

81
nanomaterials

Article
Liquid-Assisted Femtosecond Laser
Precision-Machining of Silica
Xiao-Wen Cao 1 , Qi-Dai Chen 2 , Hua Fan 2 , Lei Zhang 1 , Saulius Juodkazis 3,4
and Hong-Bo Sun 2,5, *
1 State Key Laboratory of Integrated Optoelectronics, School of Mechanical Science and Engineering,
Nanling Campus, Jilin University, Changchun 130025, China; [email protected] (X.-W.C);
[email protected] (L.Z.)
2 State Key Laboratory of Integrated Optoelectronics, College of Electronic Science and Engineering,
Jilin University, Changchun 130012, China; [email protected] (Q.-D.C.); [email protected] (H.F.)
3 Centre for Micro-Photonics, Faculty of Science, Engineering and Technology, Swinburne University of
Technology, Hawthorn, VIC 3122, Australia; [email protected]
4 Melbourne Centre for Nanofabrication, ANFF, 151 Wellington Road, Clayton, VIC 3168, Australia
5 State Key Laboratory of Precision Measurement Technology and Instruments, Department of Precision
Instrument, Tsinghua University, Haidian, Beijing 100084, China
* Correspondence: [email protected]; Tel: +86-010-6279-8249

Received: 12 April 2018; Accepted: 23 April 2018; Published: 28 April 2018

Abstract: We report a systematical study on the liquid assisted femtosecond laser machining of
quartz plate in water and under different etching solutions. The ablation features in liquid showed
a better structuring quality and improved resolution with 1/3~1/2 smaller features as compared
with those made in air. It has been demonstrated that laser induced periodic structures are present
to a lesser extent when laser processed in water solutions. The redistribution of oxygen revealed
a strong surface modification, which is related to the etching selectivity of laser irradiated regions.
Laser ablation in KOH and HF solution showed very different morphology, which relates to the
evolution of laser induced plasma on the formation of micro/nano-features in liquid. This work
extends laser precision fabrication of hard materials. The mechanism of strong absorption in the
regions with permittivity (epsilon) near zero is discussed.

Keywords: femtosecond laser; silica; Laser materials processing; nonlinear optics at surfaces

1. Introduction
Femtosecond (fs) laser has proved to be an efficient tool for micro/nanomachining [1] in
micro-optics [2], micromechanics [3], microfluidics [4], organic light-emitting diode (OLED) display [5],
and micro-sensing [6]. Based on the nonlinear nature of light–matter interaction via multi-photon
and avalanche absorption [4,7], fs-laser machining is independent of the material’s hardness and has
been demonstrated on a wide range of metals, semiconductors, and dielectrics [8–10]. By intense
laser pulses focused with an objective lens, structures and patterns of 2D and 3D morphology have
been realized [2,8]. However, it also has disadvantages. The laser–matter interaction at the surface is
inevitably affected by the evolving ablation pattern of the fabricated structure due to the scattering
and absorption, also chemical modification [11]. The laser-induced ripples generated by an imprint of
a plasmonic wave [12,13] would greatly increase the surface roughness which can be desired depending
on application. Also, laser ablated debris randomly falling on the surface could enhance the light
absorption and scattering [9,14]. Especially when laser induced thermal effects are pronounced [15,16],
debris is difficult to wash out after laser fabrication. To solve these problems, fs-laser-assisted etching
method [17,18] and liquid-assisted fs-laser machining [19,20] have been proposed as two possible

Nanomaterials 2018, 8, 287; doi:10.3390/nano8050287 82 www.mdpi.com/journal/nanomaterials


Nanomaterials 2018, 8, 287

solutions. The former method uses laser-induced selective etching to remove the modification area
without leaving debris. While the latter uses the laser induced shock wave and water plasma generating
cavitation bubbles to wash the debris. When a laser beam is focused on the sample surface through
the liquid, subsequent laser pulses are scattered by the bubbles which is a disadvantage for precision
machining [10]. For geometry when light is focused through a transparent substrate onto an interface
with liquid, light-induced backside wet etching (LIBWE) is realized as originally developed for
optical projection processing [21,22]. We adopted LIBWE for the direct laser writing. LIBWE delivers
a better surface morphology of fs-laser processed areas in liquid. However, the surface quality
(smoothness) after laser-assisted etching would suffer from loss in resolution and departure from the
initial design due to chemical enhancement of material removal. Although debris can be removed in
the liquid-assisted fs-laser machining, surface roughness is affected by the periodic structures [23].
In this paper, we report a systematical study of liquid assisted fs-laser machining of quartz plates
in water, KOH, and HF solutions as explored in LIBWE geometry. The obtained micro-holes in liquid
have a better morphology and a smaller diameter (1/3~1/2) compared with those fabricated in air.
It has been found that the formation of laser induced periodic structures was decreased in solutions.
Due to the redistribution of oxygen at the laser ablation site, chemical reactions affected formation of
the induced periodic structures and provided better control over surface modification.

2. Materials and Methods


The schematics of the employed experimental laser microfabrication system are shown in Figure 1.
A regeneratively amplified Ti:sapphire laser was used, which delivered pulses with a duration of
100 fs (FWHM), center wavelength of 800 nm, and repetition rate of 500 Hz. The diameter of the
beam was approximately 6 mm, and then expanded to 18 mm with a beam expander, which was
composed of two lenses with focal lengths of −50 mm and 150 mm. The laser beam was focused
onto the upper surface (in contact with solution) through the sample using an objective lens (50×
magnification, numerical aperture NA of 0.7 and aperture of 9 mm). The pulse energy was measured
at the exit pupil of the objective lens. A quartz plate, 1-mm-thick, was mounted on a 3D translation
stage with a positioning resolution of 0.1 μm (Figure 1). The upper surface was in contact with various
solutions: deionized water, KOH (5% and 40% by mass ratio), and HF solution (2% or 5 mol/L). In this
LIBWE case, the focused spot would not be influenced by bubble formation during laser fabrication.
The laser fluence of the incident pulses was continuously tuned using a variable attenuator, while the
laser pulse number was controlled by a shutter with a temporal resolution of 1 ms, which made it
possible to obtain a single pulse.

Figure 1. Schematics of the liquid-assisted femtosecond laser precision-machining system. The 50×
magnification objective lens was uses, RF is the reflector. A CCD (charge-coupled device) was used for
real-time observation of laser processing.

83
Nanomaterials 2018, 8, 287

After laser fabrication, the sample was cleaned in an ultrasonic bath with KOH solution (40%,
mass concentration) at room temperature, which removed all the debris from the surface thoroughly.
After cleaning, the morphology of the ablated region was obtained with a scanning electron microscope
(SEM, JSM-7500F, JEOL Ltd., Akishima-shi, Japan). At each machining parameter, a 5 × 5 hole-array
was fabricated and all the results were the average of these 25 dots.

3. Results and Discussion


At first, the morphology of the holes fabricated with different laser pulse energies and pulse
numbers were investigated after fabrication in air and deionized water. As shown in Figure 2, the laser
pulse energy was set from 68 nJ to 107 nJ and the pulse number was changing from N = 1 to 100.
Apparently, it indicates that the size of the holes ablated in deionized water was much smaller than
those produced in air ambient (in LIBWE geometry). The interior of holes made in deionized water
were relatively cleaner. In air, an unstructured ablation hole was observed at one pulse regardless
of the pulse energy. Grating-like ripples with an orientation perpendicular to the laser polarization
(the normal ripples) [24] were found around the irradiated regions at exposures above five pulses and
68 nJ/pulse energy. It is worth noting that a smaller ablation hole with diameter of about 120 nm was
observed in the center of the irradiation region at 5–10 pulses of 68 nJ. There was still a pronounced
ablation hole at the center for exposure by five pulses of 94 nJ and 107 nJ, which may be due to the
competition between laser ablation and ripple formation as discussed later in more detail. When the
pulse number was more than 10, laser induced ripples dominated the morphology regardless of the
laser pulse energy.

Figure 2. Scanning electron microscope (SEM) images of holes induced in air (a) and water (b) by
different number of pulses varied from 1 to 100 and pulse energy ranging from 68 nJ to 107 nJ. Scale bars
are 2 μm.

In water, the ablated products may interact and undergo chemical modifications in a short time
after the laser pulse, especially inside the cavitation bubble [25]. Compared with air, the atom density
of water is three orders of magnitude higher and confines the laser shock energy and plasma plume

84
Nanomaterials 2018, 8, 287

to a smaller region. As aforementioned (Figure 2a), ablation holes were observed with a smaller
diameter after the first pulse regardless of the pulse energy. However, the hole formation by the
ablation continued till 100 pulses at 68 nJ, 40 pulses at 81 nJ, 40 pulses at 94 nJ, and 10 pulses at 107 nJ
and dominated laser structuring and material removal. Competition between direct hole formation
by laser ablation and ripple formation was refined on larger areas and larger numbers of laser pulses.
Ripples were printed directly on the surface as the ablation hole changed from circular, to elliptical,
to linear and, finally, along the linear central feature (Figure 2b). This revealed a systematic change of
surface modifications when all focusing conditions are same.
The diameter of the ablated area with the pulse number in air (a) and in water (b) is shown in
Figure 3. In air, the diameter increased quickly with the pulse number below 30 pulses and then
maintained saturation until 100 pulses. In water, the diameter changed much slower as the pulse
increased at energies of 68 nJ, 81 nJ, and 94 nJ. However, it went up linearly with pulse accumulation at
107 nJ, which could also be observed in Figure 2b. The ablation threshold fluences of the quartz plate
in air and water were evaluated by investigating the dependency of the ablated size (squared diameter,
D2 ) on the irradiation pulse energy. By fitting the data according to the equation (assuming Gaussian
intensity profile) [26]
D2 = 2ω02 ln( E p /Eth ) (1)

where ω 0 is the beam waist (radius) at the focal plane and Eth is the threshold energy. The ln(Eth ) could
be obtained as asymptotic value when D2 →0. The threshold fluences, φth , could be calculated for the
peak amplitude
φth = 2Eth /πω02 (2)

For a single shot, the calculated threshold pulse energies were about 46.4 nJ and 35.7 nJ in air and
in water, respectively, which corresponded to the fluences of 2.3 J/cm2 and 5.3 J/cm2 , respectively.

Figure 3. Diameter of the ablated area with pulse number in air (a) and water (b). Diameter is defined
as a recognizable surface damage cross section in SEM images.

Laser induced periodic structures were found depressed in water as shown in Figure 2 with more
details in Figure 4. Taking the period at 68 nJ, for example, the period observed at the ablation center
at 5 or 10 pulses is around 150 nm. It quickly increased to around 260 nm at 30 pulses. Then it kept the
saturation value up to 100 pulses. If the pulse energy increased, as shown in the Figure 4b, the center
period at 50 pulses was around 289 nm at 81 nJ. While as a case in water, the period at 45 pulses and
68 nJ was around a slightly larger period of 300 nm. This is attributed to the stronger ablation which
induced structures deep into the sub-surface volume. The induced nanostructures were formed by the
accepted model of a standing wave at the interface of the active plasma (excited) layer and silica host
(the refractive index around 1.48) rather than at the silica–air interface. The ambient refractive index

85
Nanomaterials 2018, 8, 287

has been demonstrated to be an important factor to the period of the induced structures; the smaller
the refractive index, the larger the period [27,28].

Figure 4. Laser induced periodic structures in air. (a) The period evolution at 68 nJ; (b) period at
50 pulses and 81 nJ in air; (c) period at 50 pulses and 81 nJ in water.

From the cross-section, it can be seen that the depth of periodic structures extended a micrometer
into silica. Interestingly, the structures become squeezed with smaller periods at larger depths,
expandable by the effective medium theory, i.e., a larger effective refractive index more far from the
ablated interface.
The redistribution of the oxygen revealed lower concentration in the center but higher at the
ablation border by energy disperse spectrum (EDS) measurement, as shown in Figure 5a. In the water,
the morphology in cross section was different. The ablation hole without nanostructures extended
deep into the bulk. However, the oxygen presented a similar distribution as that in the air ablated
samples (Figure 5b). Ablation of silica evolves via building up of the electrostatic field between the fast
removed electrons from the surface and lagging ions. High temperature and pressure conditions are
created [29], which would cause phase changes and ion separation during the stage of hydrodynamic
movement after an ultra-short laser pulse.

Figure 5. O and Si content along the depth of ablated surface (a) in air and (b) in water. The background
SEM image has the same axial and lateral scales as the compositional plot.

The laser ablation process caused elemental redistribution, which is the most probable reason for
selective etching of the ablated area and non-ablated area [30]. For example, the formed nanogratings

86
Nanomaterials 2018, 8, 287

in air would become different with larger nanovoids but could be made smooth after HF etching for
10 min, as shown in the second line in Figure 6. Slow etching at room temperature becomes accelerated
by the ablation which is known also in LIBWE to cause cavitation (hence surface disintegration under
negative tensile pressure). The ablation holes become rounder and deeper at the same pulse energies
as compared with those in water (lines one and two in Figure 6).

Figure 6. Comparison between ablations in different solutions. The pulse energy for each is 68 nJ and
the scale bar denotes 2 μm.

The morphologies of laser ablation in the etching solutions were very different for the HF and
KOH: the surface becomes cleaner without laser induced ripples. Nanovoids of elliptical shape similar
with those water but with a nanogaps extending perpendicular with the laser polarization were
observed for KOH and HF. This can be attributed to the electronic heat conductivity enhancement
along the electrical component of the oscillating light E-field which facilitates chemical reaction and
thermal diffusion [31]. E-field control of electronic transport during laser irradiation (rather diffusional
all directional spreading) is expected to manifest itself as enhanced chemical etching of material
deposition and can be linked to the convectional mechanisms in liquid environment [32].
Femtosecond laser machining is discussed as ‘cold processing’ [33] due to having a smaller heat
affected zone. However, this is achieved due to well-controlled energy deposition and high intensity
with temperatures reaching more than 500 ◦ C [29] at the irradiated area, especially in the cavitation
and bubble formation mode. High temperature enhances chemical reactions such as etching in KOH
and HF solutions via the standard Arrhenius activation mechanism. The reaction in HF acid is more
effective in material removal as compared with KOH since reaction product K2 SiO3 is less dissolvable
compared to SiF4 .
A unique feature of the laser ablated patterns is the well-centered hole formation especially
recognizable with smaller numbers (N = 5) of pulses (see Figures 2 and 6). The pronounced central
deeper ablation holes can be explained by the energy deposition. The strongest light absorption takes
place at the regions where permittivity (epsilon) is near zero (ENZ) [34]. This occurs at the very central
part and can explain the formation of stronger ablation which is localized into the sub-diffraction
area. Since ENZ regions depend on the permittivities of the host as well as the free electron plasma,
this feature of strong localization of ablation can be engineered and will be the focus of future studies.

4. Conclusions
In summary, we have systematically investigated the interaction of fs-laser pulses with fused
quartz (silica) in air, water, KOH, and HF solutions in LIBWE geometry. Smaller holes and smooth

87
Nanomaterials 2018, 8, 287

surface were obtained in water and water solutions. The threshold in water had a similar decreasing
trend with pulse accumulation. In addition, laser induced periodic structures were less pronounced
in water. Etching enhancement along the E-field of a linearly polarized laser beam was observed in
the laser ablation in solution, which is attributed to the thermal effect enhanced chemical reaction
and E-field enhanced electronic conductivity. Our systematical investigation opens up prospects for
a better controlled high precision nano-/micro-scale fabrication of hard and chemically inert materials.
The mechanism of energy deposition into the ENZ regions is discussed.

Author Contributions: X.-W.C. performed the experiments; Q.-D.C. and H.-B.S. conceived and designed the
experiments; S.J. analyzed the data; H.F. and L.Z. contributed materials/analysis tools; all the authors wrote this
paper together.
Acknowledgments: This work was supported by the National Key R&D Program of China (2017YFB1104600),
National Natural Science Foundation of China (NSFC) (61590930, 51335008, 61435005, 61605055), and the Program
for JLU Science and Technology Innovative Research Team (JLUSTIRT) (2017TD-21).
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Malinauskas, M.; Žukauskas, A.; Hasegawa, S.; Hayasaki, Y.; Mizeikis, V.; Buividas, R.; Juodkazis, S.
Ultrafast Laser Processing of Materials: From Science to Industry. Light Sci. Appl. 2016, 5, e16133. [CrossRef]
2. Ni, J.; Wang, C.; Zhang, C.; Hu, Y.; Yang, L.; Lao, Z.; Xu, B.; Li, J.; Wu, D.; Chu, J. Three-Dimensional Chiral
Microstructures Fabricated by Structured Optical Vortices in Isotropic Material. Light Sci. Appl. 2017, 6, e17011.
[CrossRef]
3. Zhang, Y.L.; Chen, Q.D.; Xia, H.; Sun, H.B. Designable 3D Nanofabrication by Femtosecond Laser Direct
Writing. Nano Today 2010, 5, 435–448. [CrossRef]
4. Juodkazis, S.; Mizeikis, V.; Matsuo, S.; Ueno, K.; Misawa, H. Three-Dimensional Micro- and Nano-Structuring
of Materials by Tightly Focused Laser Radiation. Bull. Chem. Soc. Jpn. 2008, 81, 411–448. [CrossRef]
5. Guo, F.; Karl, A.; Xue, Q.-F.; Tam, K.C.; Forberich, K.; Brabec, C.J. The Fabrication of Color-Tunable Organic
Light-Emitting Diode Displays via Solution Processing. Light Sci. Appl. 2017, 6, e17094. [CrossRef]
6. Guo, L.; Xia, H.; Fan, H.T.; Zhang, Y.L.; Chen, Q.D.; Zhang, T.; Sun, H.B. Femtosecond Laser Direct Patterning
of Sensing Materials toward Flexible Integration of Micronanosensors. Opt. Lett. 2010, 35, 1695–1697.
[CrossRef] [PubMed]
7. Tanaka, T.; Sun, H.B.; Kawata, S. Rapid Sub-Diffraction-Limit Laser Micro/Nanoprocessing in a Threshold
Material System. Appl. Phys. Lett. 2002, 80, 312–314. [CrossRef]
8. Sugioka, K.; Cheng, Y. Ultrafast Lasers—Reliable Tools for Advanced Materials Processing. Light Sci. Appl.
2014, 3, e149. [CrossRef]
9. Shaheen, M.E.; Gagnon, J.E.; Fryer, B.J. Femtosecond Laser Ablation Behavior of Gold, Crystalline Silicon,
and Fused Silica: A Comparative Study. Laser Phys. 2014, 24, 106102. [CrossRef]
10. Juodkazis, S.; Nishi, Y.; Misawa, H. Femtosecond Laser-Assisted Formation of Channels in Sapphire Using
Koh Solution. Phys. Status Solidi RRL 2008, 2, 275–277. [CrossRef]
11. Ahsan, M.S.; Lee, M.S.; Hasan, M.K.; Noh, Y.-C.; Sohn, I.-B.; Ahmed, F.; Jun, M.B.G. Formation Mechanism of
Self-Organized Nano-Ripples on Quartz Surface Using Femtosecond Laser Pulses. Optik 2015, 126, 5979–5983.
[CrossRef]
12. Huang, M.; Zhao, F.; Cheng, Y.; Xu, N.; Xu, Z. Origin of Laser-Induced near-Subwavelength Ripples:
Interference between Surface Plasmons and Incident Laser. Acs Nano 2009, 3, 4062–4070. [CrossRef]
[PubMed]
13. Wang, L.; Chen, Q.-D.; Yang, R.; Xu, B.-B.; Wang, H.-Y.; Yang, H.; Huo, C.-S.; Sun, H.-B.; Tu, H.-L.
Rapid Production of Large-Area Deep Sub-Wavelength Hybrid Structures by Femtosecond Laser Light-Field
Tailoring. Appl. Phys. Lett. 2014, 104, 031904. [CrossRef]
14. Li, Q.K.; Yu, Y.H.; Wang, L.; Cao, X.W.; Liu, X.Q.; Sun, Y.L.; Chen, Q.D.; Duan, J.A.; Sun, H.B. Sapphire-Based
Fresnel Zone Plate Fabricated by Femtosecond Laser Direct Writing and Wet Etching. IEEE Photonics Technol. Lett.
2016, 28, 1290–1293. [CrossRef]

88
Nanomaterials 2018, 8, 287

15. Richter, S.; Heinrich, M.; Döring, S.; Tünnermann, A.; Nolte, S. Formation of Femtosecond Laser-Induced
Nanogratings at High Repetition Rates. Appl. Phys. A 2011, 104, 503–507. [CrossRef]
16. Varkentina, N.; Dussauze, M.; Royon, A.; Ramme, M.; Petit, Y.; Canioni, L. High Repetition Rate Femtosecond
Laser Irradiation of Fused Silica Studied by Raman Spectroscopy. Opt. Mater. Express 2015, 6, 79–90.
[CrossRef]
17. Li, Q.K.; Cao, J.J.; Yu, Y.H.; Wang, L.; Sun, Y.L.; Chen, Q.D.; Sun, H.B. Fabrication of an Anti-Reflective
Microstructure on Sapphire by Femtosecond Laser Direct Writing. Opt. Lett. 2017, 42, 543–546. [CrossRef]
[PubMed]
18. Liu, X.Q.; Chen, Q.D.; Guan, K.M.; Ma, Z.C.; Yu, Y.H.; Li, Q.K.; Tian, Z.N.; Sun, H.B. Dry-Etching-Assisted
Femtosecond Laser Machining. Laser Photonics Rev. 2017, 11, 1600115. [CrossRef]
19. Kaakkunen, J.J.J.; Silvennoinen, M.; Paivasaari, K.; Vahimaa, P. Water-Assisted Femtosecond Laser Pulse
Ablation of High Aspect Ratio Holes. Phys. Procedia 2011, 12, 89–93. [CrossRef]
20. Li, Y.; Qu, S.L. Water-Assisted Femtosecond Laser Ablation for Fabricating Three-Dimensional Microfluidic
Chips. Curr. Appl. Phys. 2013, 13, 1292–1295. [CrossRef]
21. Wang, J.; Niino, H.; Yabe, A. One-Step Microfabrication of Fused Silica by Laser Ablation of an Organic
Solution. Appl. Phys. A Mater. 1999, 68, 111–113. [CrossRef]
22. Böhme, R.; Braun, A.; Zimmer, K. Backside Etching of UV-Transparent Materials at the Interface to Liquids.
Appl. Surf. Sci. 2002, 186, 276–281. [CrossRef]
23. Ameer-Beg, S.; Perrie, W.; Rathbone, S.; Wright, J.; Weaver, W.; Champoux, H. Femtosecond Laser
Microstructuring of Materials. Appl. Surf. Sci. 1998, 127, 875–880. [CrossRef]
24. Buividas, R.; Mikutis, M.; Juodkazis, S. Surface and Bulk Structuring of Materials by Ripples with Long and
Short Laser Pulses: Recent Advances. Prog. Quantum Electron. 2014, 38, 119–156. [CrossRef]
25. Yan, Z.J.; Chrisey, D.B. Pulsed Laser Ablation in Liquid for Micro-/Nanostructure Generation. J. Photochem. Photobiol.
C 2012, 13, 204–223. [CrossRef]
26. Baudach, S.; Bonse, J.; Krüger, J.; Kautek, W. Ultrashort Pulse Laser Ablation of Polycarbonate and
Polymethylmethacrylate. Appl. Surf. Sci. 2000, 154, 555–560. [CrossRef]
27. Wang, L.; Xu, B.-B.; Cao, X.-W.; Li, Q.-K.; Tian, W.-J.; Chen, Q.-D.; Juodkazis, S.; Sun, H.-B. Competition
between Subwavelength and Deep-Subwavelength Structures Ablated by Ultrashort Laser Pulses. Optica
2017, 4, 637–642. [CrossRef]
28. Wang, L.; Chen, Q.-D.; Cao, X.-W.; Buividas, R.; Wang, X.; Juodkazis, S.; Sun, H.-B. Plasmonic Nano-Printing:
Large-Area Nanoscale Energy Deposition for Efficient Surface Texturing. Light Sci. Appl. 2017, 6, e17112.
[CrossRef]
29. Chen, C.; Sun, H.B.; Guo, J.C.; Wang, L.; Chen, Q.D.; Yang, R.; Yu, Y.S. Monitoring Thermal Effect in
Femtosecond Laser Interaction with Glass by Fiber Bragg Grating. J. Lightw. Technol. 2011, 29, 2126–2130.
[CrossRef]
30. Juodkazis, S.; Yamasaki, K.; Mizeikis, V.; Matsuo, S.; Misawa, H. Formation of Embedded Patterns in Glasses
Using Femtosecond Irradiation. Appl. Phys. A Mater. 2004, 79, 1549–1553. [CrossRef]
31. Rekštytė, S.; Jonavičius, T.; Gailevičius, D.; Malinauskas, M.; Mizeikis, V.; Gamaly, E.G.; Juodkazis, S.
Nanoscale Precision of 3D Polymerization via Polarization Control. Adv. Opt. Mater. 2016, 4, 1209–1214.
[CrossRef]
32. Louchev, O.A.; Juodkazis, S.; Murazawa, N.; Wada, S.; Misawa, H. Coupled Laser Molecular Trapping,
Cluster Assembly, and Deposition Fed by Laser-Induced Marangoni Convection. Opt. Express 2008, 16,
5673–5680. [CrossRef] [PubMed]
33. Joglekar, A.P.; Liu, H.-H.; Meyhöfer, E.; Mourou, G.; Hunt, A.J. Optics at Critical Intensity: Applications to
Nanomorphing. Proc. Natl. Acad. Sci. USA 2004, 101, 5856–5861. [CrossRef] [PubMed]
34. Gamaly, E.G.; Rode, A.V. Ultrafast Re-Structuring of the Electronic Landscape of Transparent Dielectrics:
New Material States (Die-Met). Appl. Phys. A 2018, 124, 278. [CrossRef]

© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

89
nanomaterials

Article
Spontaneous Shape Alteration and Size Separation of
Surfactant-Free Silver Particles Synthesized by Laser
Ablation in Acetone during Long-Period Storage
Dongshi Zhang 1,† , Wonsuk Choi 1,2,3,† , Jurij Jakobi 4 , Mark-Robert Kalus 4 ,
Stephan Barcikowski 4 , Sung-Hak Cho 2,5 and Koji Sugioka 1, *
1 RIKEN Center for Advanced Photonics, 2-1 Hirosawa, Wako, Saitama 351-0198, Japan;
[email protected] (D.Z.); [email protected] (W.C.)
2 Department of Nano-Mechatronics, Korea University of Science and Technology (UST),
217 Gajeong-Ro, Yuseong-Gu, Daejeon 34113, Korea; [email protected]
3 Department of Nano-Manufacturing Technology, Korea Institute of Machinery and Material (KIMM),
156 Gajeongbuk-Ro, Yuseong-Gu, Daejeon 34103, Korea
4 Technical Chemistry I and Center for Nanointegration Duisburg-Essen (CENIDE), University of
Duisburg-Essen, Universitaetsstrasse 7, 45141 Essen, Germany; [email protected] (J.J.);
[email protected] (M.-R.K.); [email protected] (S.B.)
5 Department of Laser & Electron Beam Application, Korea Institute of Machinery and Material (KIMM),
156 Gajeongbuk-Ro, Yuseong-Gu, Daejeon 34103, Korea
* Correspondence: [email protected]; Tel.: +81-(0)48-467-9495
† These authors contributed equally to this work.

Received: 5 July 2018; Accepted: 11 July 2018; Published: 13 July 2018

Abstract: The technique of laser ablation in liquids (LAL) has already demonstrated its flexibility
and capability for the synthesis of a large variety of surfactant-free nanomaterials with a high purity.
However, high purity can cause trouble for nanomaterial synthesis, because active high-purity
particles can spontaneously grow into different nanocrystals, which makes it difficult to accurately
tailor the size and shape of the synthesized nanomaterials. Therefore, a series of questions arise
with regards to whether particle growth occurs during colloid storage, how large the particle size
increases to, and into which shape the particles evolve. To obtain answers to these questions, here,
Ag particles that are synthesized by femtosecond (fs) laser ablation of Ag in acetone are used as
precursors to witness the spontaneous growth behavior of the LAL-generated surfactant-free Ag
dots (2–10 nm) into different polygonal particles (5–50 nm), and the spontaneous size separation
phenomenon by the carbon-encapsulation induced precipitation of large particles, after six months of
colloid storage. The colloids obtained by LAL at a higher power (600 mW) possess a greater ability
and higher efficiency to yield colloids with sizes of <40 nm than the colloids obtained at lower power
(300 mW), because of the generation of a larger amount of carbon ‘captors’ by the decomposition
of acetone and the stronger particle fragmentation. Both the size increase and the shape alteration
lead to a redshift of the surface plasmon resonance (SPR) band of the Ag colloid from 404 nm to
414 nm, after storage. The Fourier transform infrared spectroscopy (FTIR) analysis shows that the
Ag particles are conjugated with COO– and OH– groups, both of which may lead to the growth
of polygonal particles. The CO and CO2 molecules are adsorbed on the particle surfaces to form
Ag(CO)x and Ag(CO2 )x complexes. Complementary nanosecond LAL experiments confirmed that
the particle growth was inherent to LAL in acetone, and independent of pulse duration, although
some differences in the final particle sizes were observed. The nanosecond-LAL yields monomodal
colloids, whereas the size-separated, initially bimodal colloids from the fs-LAL provide a higher
fraction of very small particles that are <5 nm. The spontaneous growth of the LAL-generated metallic
particles presented in this work should arouse the special attention of academia, especially regarding
the detailed discussion on how long the colloids can be preserved for particle characterization and
applications, without causing a mismatch between the colloid properties and their performance.

Nanomaterials 2018, 8, 529; doi:10.3390/nano8070529 90 www.mdpi.com/journal/nanomaterials


Nanomaterials 2018, 8, 529

The spontaneous size separation phenomenon may help researchers to realize a more reproducible
synthesis for small metallic colloids, without concern for the generation of large particles.

Keywords: laser ablation in liquids; particle growth; carbon encapsulation; polygonal particle;
core-shell particle; surfactant-free

1. Introduction
Laser ablation in liquids (LAL) is increasingly gaining worldwide attention as a newly emerging
bottom-up and top-down combined technique for novel nanomaterial synthesis, offering diverse
applications including optics and biology [1–4]. In particular, the LAL-synthesized nanomaterials
possess a higher purity than the counterparts synthesized by wet-chemistry techniques [1], because
surfactants or stabilizers are not required for colloid synthesis, which makes them very attractive and
competitive for catalytic applications [5–8]. High-purity nanomaterial means that a large amount
of active sites are exposed to the surrounding reactants or nanoparticles (NPs), which is actually a
double-edged sword in nanoscience [7]. This is because spontaneous particle growth [9–12] may occur,
which can increase the difficulty in accurately controlling the sizes [13–16] and the shapes [17–21] of
the as-prepared nanomaterials.
Six mechanisms have been proposed for the growth of the LAL-generated nanomaterials,
including LaMer-like growth [9], coalescence [22], Ostwald ripening [10], particle (oriented)
attachment [18,23–25], adsorbate-induced growth [18,26], and reaction-induced growth [17]. Unlike
the wet-chemistry synthesis method, where particle growth terminates when the seed concentration
drops below the critical concentration [27], the particle growth can be sustained during the entire
period of LAL, because every pulse ablation produces new seeds for particle growth. This is the
reason novel football-like AgGe microspheres [17], with sizes up to 7 μm, can be obtained by LAL.
Another difference from the conventional wet-chemistry techniques is the possibility of inducing
multiple growth processes to generate different nanostructures. For example, large (>100 nm) hollow
Mn3 O4 spheres are formed from ca. 20 nm cubic particles that are assembled from smaller 5–8 nm
smaller particles after a long-period LAL [18]. Previously, the investigation on particle growth has
mainly been on the fragmentation of large metallic particles, such as Pt [10] and Au [11] into 2~3 nm
particles, and then their evolution over time is observed. No attempt has been made to study the
growth of the metallic particles just after laser irradiation, which occupies the dominant position in all
three liquid-phase laser synthesis methodologies, LAL, laser fragmentation in liquids (LFL), and laser
melting in liquids (LML) [1].
LAL is currently most frequently implemented in a batch chamber where the ablation is often
accompanied by LFL when the particles enter the beam path in the liquids [1]. The fragmentation
behavior of the particles may become dominant when implementing the LAL for a long period.
Long-period LAL means that the LAL should last more than half an hour, in order to generate highly
saturated colloids in a limited volume of liquids [9]. In this case, a large amount of active surfactant-free
particles are generated by LFL, which increases the chance for the particles to encounter each other by
Brownian motion, and then to grow overtime during storage, transportation, and when employed
in applications. If growth does occur, then it should attract special attention from academia, because
until now, the characterization and application of metallic particles such as Ag NPs [28–31] have
been seldom described in the literature. Therefore, it is not clear whether the size information shown
in the literature has been obtained from either freshly prepared colloids or after a certain period of
growth. The particle growth of the metallic particles is also accompanied by the shape transformation
into nonspherical nanostructures, such as nanowires [10,11]. Tsuji’s group reported that the post
laser irradiation of LAL-generated Ag colloids in either a citrate [32], polyvinylpyrrolidone (PVP)
aqueous solution [33], water [34], or acetone-water mixed solution [35] may cause the formation of

91
Nanomaterials 2018, 8, 529

nanoplates and nanorods during colloidal aging [36]. However, it is still unknown whether the Ag
particles synthesized by long-period LAL can still grow into polygonal products, and the maximal
size at which the particle growth terminates is also unknown, especially in the case of LAL in organic
solvents, where the solvent decomposition definitely occurs to generate a large amount of carbon or
hydrocarbons, which may precipitate on the active particles to terminate the particle growth.
In this paper, one hour long-period LAL of Ag in acetone, using a femtosecond laser (fs-LAL),
was implemented to investigate the particle growth behavior. Laser powers of 300 mW and 600 mW
were used to change the productivity of the Ag colloids, as well as the amount of both ultra-small
particles and large particles with sizes less than and greater than 10 nm, respectively. The variation
in the surface plasmon resonance (SPR) band of the Ag colloids during storage was recorded using
UV-VIS spectroscopy, while the particles’ sizes before and after the six month storage were analyzed
using transmission electron microscopy (TEM), both of which provide evidence for particle growth
and newly discovered size separation phenomena. The surface chemistry of the Ag NPs was analyzed
using Fourier transform infrared spectroscopy (FTIR), to identify the chemical groups that may be
responsible for the particle growth. This study was complemented by nanosecond (ns) laser ablation
to demonstrate that particle growth is not limited to fs-LAL. Finally, the scenarios for both the particle
growth and the size separation phenomena are proposed.

2. Results and Discussion

2.1. Femtosecond Laser Ablation


It has been reported that the redshift of the SPR peaks for the Ag NPs is often indicative of
an increase of the particle size [37] via either particle ripening or coalescence, thus giving indirect
evidence for particle growth. In our experiments, the variations in the absorption spectra of the fresh
Ag colloids synthesized by LAL at laser powers of 300 mW and 600 mW, were recorded within 32 h,
with a time interval of 30 min, as shown in Figure 1a,b. As the aging time increases, in both cases,
the colloidal SPR peaks redshift from 404 nm to 410 nm, which suggests the spontaneous particle
growth during colloidal storage. The higher SPR peak intensity of Ag colloids synthesized at 600 mW
compared with those synthesized at 300 mW, indicates that the LAL at a higher laser power gives rise
to a slightly higher productivity of the colloids. A slight increase in the SPR peak intensity with aging
is observed in both cases, which is attributed to the increased concentration of the Ag NPs, caused by
the evaporation of acetone. After six months of storage, during which time the growth was already
terminated, the absorption spectra of the colloids were characterized again, as shown in Figure 1c,d.
It is clear that the evaporation of the acetone further increases the colloidal concentrations, even though
the colloids were sealed with acetone in glass containers. To compensate for the evaporated acetone,
the colloids were diluted with additional acetone, and characterized (pink and blue curves in Figure 1c).
The magnified SPR peaks in the wavelength range of 360–500 nm show a further redshift to 414 nm
in both cases, which indicates that particle growth should continue even after the 32 h of storage.
The inset figures in Figure 1c show photographs of the colloids in acetone after six months of storage.
Both of the Ag colloids are orange colored. The higher concentration of Ag colloids prepared at a
higher power is evident from the darker orange color. These colors are different from the light yellow
color of the Ag colloids reported previously [38], probably due to the much higher concentration of
Ag colloids presented in this work. Particle precipitation was also observed at the bottom of the glass
containers. Therefore, a subsequent analysis of both of the stable colloids in liquids, and the particles
precipitated at the bottom of the glass containers, was performed.

92
Nanomaterials 2018, 8, 529

Figure 1. Absorption spectra for Ag nanoparticles (NPs) synthesized by laser ablation of Ag in


acetone, and then stored for various periods. Upper: spectra for samples prepared at a laser power of
(a) 300 mW and (b) 600 mW, and stored within 32 h (insets: enlarged spectra in the wavelength range
of 360 nm to 500 nm for samples stored for 0 and 32 h). Lower: spectra for (c) samples prepared at laser
powers of 300 mW and 600 mW, and stored for six months (inset: photos of aged Ag NPs in acetone),
and (d) enlarged spectra in the wavelength range of 360 nm to 500 nm (black and red curves: as-aged
colloidal samples; pink and blue curves: diluted as-aged samples prepared by the addition of pure
acetone to the colloids).

Figures 2 and 3 show the TEM images and the calculated size distributions of the fresh and six
month aged Ag particles generated by LAL, at the laser powers of 300 mW and 600 mW, respectively.
In accordance with the redshift of the SPR peaks, shown in Figure 1a,b, particle growth indeed occurs
(Figures 2a–f and 3a–f), especially for ultrasmall particles with sizes less than 10 nm. The average
sizes of the fresh Ag particles synthesized at 300 mW and 600 mW were evaluated to be 5.9 ± 7.6 nm
and 5.9 ± 12.2 nm (Figure 3g), respectively, which confirmed that the average particle size was almost
independent of the laser power. However, a greater amount of particles larger than 10 nm were
generated at a higher laser power of 600 mW (Figure 2a vs. Figure 3a), as indicated by the increased
ratio between the big and small particles (Figure 2g vs. Figure 2h). The majority of the Ag NPs are in
the range of 1–10 nm, occupying ca. 90% of the total amount for both of the cases. A further subdivision
of the size distributions of the Ag NPs shows that most of the particles are ca. 2 nm (Figures 2c
and 3c). After six months of storage, the average sizes of the Ag colloids increase to 7.4 ± 7.6 nm and
7.8 ± 8.2 nm, respectively. In the case of the colloids synthesized at 600 mW, a significant decrease in
the large particles was observed (Figure 3a,d).

93
Nanomaterials 2018, 8, 529

Figure 2. TEM images (a–f) and size distributions (g,h) of fresh (a–c,g) and six month (d–f,h) aged Ag
particles, synthesized by laser ablation of Ag in acetone, at a laser power of 300 mW.

In both cases, the number ratios of the as-aged colloids with sizes in the range of 10–20 nm
increased to 12–14% after 6 months of storage. The number ratios of the as-aged colloids with a
diameter of 20~30 nm were almost the same as those of the fresh colloids (Figure 2g,h and Figure 3g,h).
The number ratio of the as-aged colloid with a diameter of 30~40 nm increased slightly for the as-aged
300 mW LAL colloid, but decreased for the as-aged 600 mW LAL colloid. The number ratios of the Ag
NPs with diameters larger than 40 nm were less than 1%, which is almost negligible for both cases,
so it is difficult to quantify the variation in the number ratios of the colloids. However, a comparison
of the particle morphologies of the fresh and as-aged colloids clearly shows that the number ratio of
the particles larger than 40 nm did not change significantly (Figure 2a vs. Figure 2d) after six months
of storage for the 300 mW LAL colloid, but did decrease significantly for the as-aged 600 mW LAL
colloid (Figure 3a vs. Figure 3d).

94
Nanomaterials 2018, 8, 529

Figure 3. TEM images (a–f) and sizes distributions (g,h) of fresh (a–c,g) and six month (d–f,h) aged Ag
particles, synthesized by laser ablation of Ag in acetone at laser power of 600 mW.

In more detailed observations of large particles, regardless of them being fresh (Figure 4a–c)
or six months aged colloids (Figure 4d–f) prepared at laser powers of 300 mW and 600 mW,
the carbon-encapsulated particles with core sizes ranging from 25 nm to ca. 200 nm, and a carbon shell
thickness of 3~10 nm are all found, as shown in Figure 4. Some smaller particles were captured by the
carbon shells, which led to the formation of Ag@C-Ag truffle-like aggregates. The crystalline structure
of the Ag cores was confirmed by X-ray diffraction (XRD) characterization, as shown in Figure 5a.
The featured peaks of the colloids fit well with the standard card for Ag (ICCD No. 03-065-2871).
The presence of the carbon shells was verified by Raman spectroscopy observation of both the D-band
(1360 cm−1 ) and G-band (1582 cm−1 ) peaks of carbon (Figure 5b). The ratio of the D-band to the
G-band peak intensities was calculated to be 0.98. The G-band is associated with the ordered graphite
(sp2 ) structure, while the D-band is related to the disordered graphite layers, such as soot, chars,
glassy carbon, and evaporated amorphous carbon [39]. Consequently, it can be concluded that the
carbon shells possess high ratios of disorders. According to Robertson [39], the carbon shells that
are composed of both crystalline graphite and amorphous carbon can be assigned to diamond-like
carbon (DLC), a metastable form of carbon. Because of the presence of abundant hydrogen, which are
generated from the LAL-induced acetone decomposition, it is highly possible that other products,

95
Nanomaterials 2018, 8, 529

such as hydrogenated amorphous carbon (a-C: H) and tetrahedral amorphous carbon (ta-C) [39] also
form during LAL. Nevertheless, their existence cannot be confirmed by the Raman spectrum, shown
in Figure 5b. The identification and quantification of different carbon disorders will be a focus of our
future studies.

Figure 4. TEM images of the Ag@C core-shell particles observed in (a–c) fresh and (d–f) six month
aged colloids synthesized by laser ablation in liquids (LAL), at laser powers of 300 mW (a,d) and
600 mW (b,e). The TEM images in (c,f) show the Ag@C-Ag composites from the fresh and the six
month aged colloids.

Figure 5. (a) XRD pattern and (b) Raman spectrum of the LAL-synthesized Ag NPs.

Some polygonal Ag nanocrystals, such as triangular (Figure 6d), pentagonal (Figure 6c,f),
hexagonal (Figure 6a,d,f), octagonal (Figure 6b,e), and some spherical particles with sharp edges,
(Figure 6f) were observed, of which the sizes were in the range of 5–50 nm, similar to the Ag
nanocrystals prepared by the reduction of AgNO3 [40]. The maximal sizes of the newly grown
Ag nanocrystals are much smaller than the 100~500 nm crystals obtained by the post-irradiation of the

96
Nanomaterials 2018, 8, 529

LAL-generated Ag spheres [33], which indicates that the surrounding carbon atoms may cover the
newly formed nanocrystals and inhibit their further growth into larger particles. The following results
provide evidence for the spontaneous growth of small particles into nonspherical Ag nanocrystals:
(1) the amount of the ultrasmall particles, of 2–3 nm in diameter, is significantly decreased after storage
(Figure 2g,h and Figure 3g,h); (2) polygonal particles are not surrounded by ultrasmall particles
(Figure 6), unlike the spherical Ag NPs encapsulated by carbon shells (Figure 4); and (3) previous
reports show that the size of the Ag NPs that correspond to an SPR band of 414 nm is around 30 nm [41],
which is much larger than the ca. 8 nm (Figures 2h and 3h) of the six month aged Ag colloids with
the same SPR position (Figure 1c,d) in this work. Considering that the polygonal Ag nanostructures,
such as the hexagonal and triangular Ag nanoplates, often have a higher SPR band position than the
Ag spheres with the same sizes [42], it is reasonable to deduce that the formation of the polygonal
particles is the main contribution to the redshift of the SPR band during the storage of the colloids,
rather than a change of the particle size. From the occurrence of particle growth, the building blocks
(fresh ultrasmall Ag NPs) should be surfactant-free, which facilitates the attachment and ripening of
ultrasmall particles into various polygonal nanoparticles, with the growth direction governed by the
surface bindings [9,18], which will be discussed below.

Figure 6. (a–f) Polygonal particles obtained after six months of storage of the colloids.

To determine the surface groups that may direct the particle growth into the polygonal particles,
the six month aged Ag particles were analyzed using FTIR, and the results are shown in Figure 7.
The vibration bands at 1165 cm−1 and 1250 cm−1 are assigned to the wagging and the rocking
vibrations of CH2 , respectively, while the peaks at 1375 cm−1 and 1734 cm−1 are assigned to the
stretching vibrations of C–H alkane [43] and C=O [44], respectively. The broad peak between
3000 cm−1 and 3500 cm−1 is because of the OH stretching [44]. The broad bands of 2061~2158 cm−1
and 2325~2343 cm−1 are related to the stretching vibrations of the adsorbed CO [45,46] and CO2 [47]
molecules, respectively. Therefore, the CO-metal [48] and CO2 -metal complexes [49], such as Ag(CO)n
(n = 1~3) [50] and Ag–O–C–O, may be generated during LAL (Figure 7b). The three vibration peaks
at 2902 cm−1 , 2932 cm−1 , and 2960 cm−1 , are due to the CH stretching from the alkyl groups [44].

97
Nanomaterials 2018, 8, 529

The FTIR analysis indicates that the acetone molecules are decomposed into CO2 , CO, alkanes, and OH–
groups during LAL, which may either strongly adsorb onto the Ag NPs or interact with the Ag NPs to
form Ag(CO)n or Ag(CO2 )n complexes. As a result, the colloids are endowed with ultrahigh stability,
which enables them to self-stabilize, even after six months of storage. Tsuji’s group confirmed that
shape transformation from the Ag spheres into prisms is not induced by acetone molecules, but should
be caused by water molecules. In the present case, considering the absence of water molecules in the
liquids, the OH– formed by the acetone decomposition and reconstruction is one possible candidate
to cause the anisotropic growth of Ag particles [35] via the Ostwald ripening process [33]. Chemical
adsorbed carboxylate (COO–) groups may be another candidate to trigger the selective Ag crystal facet
growth during ripening [51].

Figure 7. (a) FTIR spectrum of the six month aged Ag NPs and the (b) magnified spectrum where the
peaks of Ag(CO)n (n = 1–3) complexes are located.

The optical images of the six month aged colloids show that some particles have already
precipitated during storage (inset images in Figure 8c,f). In comparison with the particle morphologies
of the fresh and six month aged colloids synthesized by LAL at a laser power of 600 mW (Figure 3a,d),
it is also clear that the amount of particles with diameters larger than 40 nm decreased dramatically
after a long-period storage (Figure 3g,h), which is due to the precipitation of larger particles. Figure 8
shows the TEM images of the precipitated particles prepared at laser powers of 300 mW and 600 mW
after six months of storage. The precipitated particles are mainly particles larger than 40 nm, which
are encapsulated by carbon to form particle networks, which also encapsulate a certain amount of
particles with sizes less than 40 nm, so that the number ratios of the ultrasmall colloids (1–10 nm)
decrease (Figures 2h and 3h). Despite the precipitation induced by the carbon capture, the number
ratios of the colloids with sizes of 10–20 nm still increase, and the number ratios of the colloids with
sizes of 20–30 nm remain almost unchanged. This means that the continuous supplement of the Ag
NPs with sizes of 10–30 nm as a result of the particle growth of the ultrasmall surfactant-free Ag
NPs occurs. An analysis of the change in the colloidal size distribution after six months of storage
(Figure 2g,h and Figure 3g,h) led us to conclude that the colloids synthesized by LAL at higher powers
possess a higher ability to separate the Ag colloids with sizes less than 40 nm. This is because more
carbonaceous substances are generated at 600 mW, and these can then easily capture the large particles
to form aggregate and precipitate over time, whereas the 300 mW LAL generates fewer carbon clusters,
so that the size separation ability decreases, making it less efficient to separate the particles larger
than 40 nm. The carbon ‘captors’ that originate from the LAL-induced decomposition of acetone have
three forms, the carbon shells of Ag@C particles that capture small Ag particles to form truffle-like
aggregates (Figures 4c and 8c), the active wandering carbon clusters that gradually precipitate on
particles to induce the formation of particle networks (Figure 8b,e), and the carbon nanosheets (Figure 9)

98
Nanomaterials 2018, 8, 529

that mainly capture the ultrasmall Ag particles with sizes less than 10 nm (Figure 9d–f). Recently,
Escobar-Alarcón showed that the LAL of graphite in water could produce carbon nanosheets [52],
in which they proposed that graphene exfoliation was the formation mechanism for carbon nanosheets.
However, in our case, carbon comes from the decomposition of acetone molecules during LAL. Hence,
the carbon nanosheets should form by the self-assembly of carbon clusters inside liquids, which can also
explain the irregular structure shapes of nanosheets shown in Figure 9 and in Escobar-Alarcón et al. [52].

Figure 8. Precipitated particles after the six month storage of the Ag colloids, synthesized by
femtosecond LAL at laser powers of (a–c) 300 mW and (d–f) 600 mW, respectively. The arrows shown
in the optical images in the insets (c,f) indicate the particle precipitation after six months storage.

Figure 9. (a–f) TEM images of the carbon nanosheet formed during storage of the colloids.

99
Nanomaterials 2018, 8, 529

Overall, the particle growth followed by the size separation phenomena was determined to
occur during storage. The scenario is summarized in Figure 10. After long-period LAL, a large
amount of Ag (orange colored spheres) and C (black colored spheres) clusters together to generate
large Ag@C core-shell particles (Figure 10a). During the colloidal storage, the growth of ultrasmall
surfactant-free Ag particles of 1–25 nm in diameter occurs through the Ostwald ripening mechanism
(Figure 10b,c). Meanwhile, the soft carbon shells of large particles capture the surrounding small
particles to form Ag@C-Ag aggregates. The aggregated particles precipitate at the bottom of the
container (Figure 10b,c). The precipitation of the large particles cause the size distribution of the Ag
particles in the supernatant to narrow by carbon encapsulation, which is termed size separation in this
work. The self-size separation, with the aid of the large particle precipitation by carbon-encapsulation,
can offset the disadvantage of the increased amount of large particles that are formed at a higher
laser power, thus allowing a higher-productivity synthesis of metallic colloids with the uniform size
distribution with that synthesized by the LAL at a lower laser power. The sediments may be technically
separated by centrifugation or filtering to yield stable, monomodal Ag colloids in acetone.

Figure 10. (a–c) Schematic of particle growth into polygonal nanocrystals, followed by the size
separation of ultrasmall particles by large particle precipitation, due to carbon encapsulation. In the
liquids, Ag and carbon are denoted by orange and black colored spheres, respectively.

2.2. Nanosecond Laser Ablation


Nanosecond laser ablation of Ag was performed in acetone for comparison, so as to investigate
whether the spontaneous size separation and growth evolution into the polygonal crystals are specific
for the fs-LAL, or whether they occur at a significantly longer pulse duration. Figure 11 shows the
absorbance spectrum of the Ag NPs synthesized by the ns laser ablation of Ag in acetone, and that of
the colloid after storage for five weeks. The SPR position of the Ag colloids was slightly redshifted
from 400 nm to 407 nm after five weeks of storage, accompanied by the broadening of the SPR peak.
Both indicate the change of the colloidal properties. Figure 12 shows the morphologies of the fresh
and the aged Ag NPs, and their size distributions. Compared to the fresh Ag colloids with an average
size of 8.93 ± 2.7 nm, the Ag NPs size was slightly increased to 11.1 ± 3.9 nm after storage for five
weeks. Regarding the ultrasmall Ag NPs of 1–10 nm, the average size is increased from 7.5 ± 1.5 nm
to 8.2 ± 1.3 nm after long-term storage. A significant decrease in the number ratio of the 1–10 nm
particles from 68% to 47%, and the dramatic increase in the number ratio of the 10–20 nm particles
from 22% to 50% (Figure 12g,h), provide evidence for the gradual growth of hte ns laser-synthesized
Ag NPs in acetone during storage. The observation of more polygonal nanocrystals from the five
week aged Ag colloid (Figure 12f) compared with the fresh colloid (Figure 12c) indicates that particle
growth is always accompanied by the shape alteration of the metallic particles, regardless of the pulse
duration used for colloid synthesis. Particle growth and simultaneous shape alternation are considered
to constitute the main reason for the redshift and broadening of the SPR band (Figure 11).

100
Nanomaterials 2018, 8, 529

Figure 11. Absorption spectra for Ag NPs synthesized by ns laser ablation of Ag in acetone and then
stored for five weeks.

Figure 12. TEM images (a–f) and sizes distributions (g,h) of fresh (a–c,g) and five week (d–f,h) aged
Ag particles synthesized by ns laser ablation of Ag in acetone at a pulse energy of 150 mJ.

101
Nanomaterials 2018, 8, 529

However, compared to those particles synthesized by fs-LAL (Figures 2a–f and 3a–f), the carbon
shells were almost negligible for the ns-LAL generated Ag NPs (Figure 12a–f). No big particles larger
than 40 nm were observed, which indicates that the fs-LAL, at high fluences (170 and 191 J/cm2 for
300 mW and 600 mW, respectively), causes more severe degradation of the acetone molecules than
the ns-LAL at a low laser fluence (3 J/cm2 ). An FTIR study on the LAL in tetrahydrofuran showed
that the carbonylic compounds were dominantly generated during ns-LAL, whereas the more olefinic
species were dominantly generated during fs and picosecond (ps) laser ablation; therefore, the fs-LAL
created more hydrophobic species [53]. If such hydrophobic species are also preferably formed during
fs-LAL in acetone, then this could explain why their lower solubility triggers phase separation
(into soft carbon shells, adsorbed on hydrophobic metal Ag), which supports the aggregation and
sedimentation processes.
Regarding the polydispersity of the colloid, fs-LAL at high intensities leads to the formation
of large Ag droplets jetting off the molten metal layer. This jetting is caused by Rayleigh instability,
with the droplets evident even in front of an expanding cavitation bubble boundary [54]. These droplets
solidify as large particles, so that the size distributions from the fs-LAL show stronger bimodality
(even large spheres) compared to the ns-LAL.
Therefore, the ns-LAL in acetone yields colloids that are far less polydisperse than the fs-LAL.
However, after the sedimentation or size separation of the aged fraction, the final colloid in the
supernatant of the fs-LAL colloid has significantly smaller primary particles. Comparing the number
ratios of 1–10 nm ultrasmall particles (80–90% for fs-LAL, Figures 2g and 3g, vs. 68% for ns-LAL,
Figure 12g), it can be easily deduced that fs-LAL is more efficient than ns-LAL for the generation of
ultrasmall particles, at the expense of the mass yield lost during the size separation. In particular,
the histograms from the TEM measurements show that the fs-LAL supernatant will contain a
significantly larger portion of very small particles (<5 nm).

3. Materials and Methods


Silver colloids were synthesized by laser ablation of Ag in acetone by fs laser (FGPA μ
Jewel D-1000-UG3, IMRA America Inc., Ann Arbor, MI, USA), with a pulse duration, wavelength,
and repetition rate of 457 fs, 1045 nm, and 100 kHz, respectively. Two laser powers of 300 mW and
600 mW were adopted for the LAL. The spot sizes of the 300 mW LAL and 600 mW LAL were 15 and
20 μm, which gave the laser fluences were 170 of 191 J/cm2 for the 300 mW LAL and 600 mW LAL,
respectively. A circular Ag plate with a diameter of 10 mm and a thickness of 1 mm was placed inside
a glass dish filled with 8 mL acetone. The liquid thickness was ca. 5 mm above the target surface.
The fs laser beam was then focused on the Ag plate surface using a 20× objective lens (NA = 0.45,
Mitutoyo, Japan). An area of 3.5 × 3.5 mm2 was scanned using the parallel-line scanning method,
described by the authors of [55–57], with an adjacent line interval of 5 μm. The scanning speed was set
at 1 mm/s. Each ablation lasted 1 h to ensure long-period LAL [9].
The colloids that were just synthesized by the LAL, termed fresh colloids, were directly deposited
onto the TEM grids (EMJapan, U1015, Tokyo, Japan, 20 nm thick carbon films on copper grids) and then
characterized using TEM (Jeol, JEM-1230, Tokyo, Japan). UV-VIS spectroscopy (Shimadzu, UV-3600
Plus, Kyoto, Japan) was used to measure the absorption spectra during the colloidal aging every 30 min.
More than 500 particles were measured by ImageJ to calculate the average particle sizes of the colloids.
For the XRD (Rigaku, CuKα radiation (40 kV-30 mA), SmartLab-R 3kW, Tokyo, Japan) measurement,
the colloids were centrifuged by a centrifuge (Eppendorf, Centrifuge 5430, Hamburg, Germany) at a
rotation speed of 14,000 rpm for 10 min, and then deposited on a silicon wafer. The colloids were stored
and sealed in glass containers at room temperature. After six months of storage, both the colloids
in the liquids and the precipitated particles at the bottom of the glass container were characterized
using TEM and UV-VIS spectroscopy. The surface chemistry of the Ag particles was analyzed using
FTIR (Shimadzu, Prestige-21, Kyoto, Japan) and Raman spectrometers (LabRAM, He-Ne laser, 632 nm,
0.686 mW, Tokyo, Japan).

102
Nanomaterials 2018, 8, 529

The nanosecond laser synthesis of the Ag colloids was performed using a Nd:YAG ns-laser
(SpitLight DPSS250-100, InnoLas Laser GmbH, Krailling, Germany) at a pulse duration, wavelength,
repetition rate, and pulse energy of 11 ns, 1064 nm, 100 Hz, and 150 mJ, respectively. The detected
spot size on the target surface after ablation was 2.5 mm in diameter. The laser fluence is calculated to
be 3 J/cm2 . The Ag target was placed inside a glass cuvette filled with 3 mL of acetone. The liquid
thickness above the target corresponded to the optical path (10 mm) of the glass cuvette used.
The experiments were performed under continuous stirring of the liquid, with irradiation of the
unfocused laser beam for 60 s. The colloids were characterized using UV-VIS spectroscopy (Thermo
Scientific Evolution 201, Tokyo, Japan) and TEM (Zeiss, Type EM 910, Oberkochen, Germany), directly
after synthesis and after five weeks of storage in sealed polypropylene tubes at room temperature.

4. Conclusions
The investigation of Ag colloids synthesized by the fs-LAL of Ag in acetone, at laser powers
of 300 mW and 600 mW, confirmed the spontaneous growth of the ultrasmall particles and the
spontaneous size separation during long-period storage. The FTIR spectroscopy analysis showed
that the CO and CO2 molecules are adsorbed on the Ag particles to form Ag(CO)n and Ag(CO2 )n
complexes, which may contribute to the high stability of the supernatant Ag NPs. Carboxylate (COO–)
and hydroxyl (OH–) species also conjugate on the Ag NPs, which may be the reason for the anisotropic
growth of the Ag particles into the polygonal nanocrystals over time. Both the particle size increase
and the shape transformation into the polygonal nanocrystals caused a redshift of the SPR bands
from 404 nm to 414 nm. The carbonaceous species generated from the acetone decomposition and
pyrolysis during the LAL gradually adsorbed onto the large particles, and the soft carbon shells of the
Ag@C particles also captured smaller particles, to form Ag@C-Ag aggregates. Particle aggregation
and the formation of Ag–C networks significantly compromise the stability of the large Ag particles
and cause their gradual precipitation during long-period storage, which leaves ultrasmall particles
behind in the liquid supernatant. A higher size separation ability is endowed with the colloids
obtained by fs-LAL at a higher power (600 mW), which possibly benefits from a larger amount of
carbon captors, because of the stronger decomposition or pyrolysis of the solvent, which possibly
creates olefinic species with a low solubility in acetone. Such phase-separating, carbon-induced size
separation induced by the colloids themselves may be helpful to conquer the challenging issue of
the wide (i.e., bimodal) size distribution of the metallic particles generated by ultrashort-pulsed-LAL.
When the growth is terminated and mechanically fractionated from the supernatant as a precipitate,
the resulting supernatant of the fs-LAL-derived colloids bears a significantly higher fraction of very
small particles (≤5 nm), compared to the ns-LAL. On the other hand, the ns-LAL of the Ag in acetone
yields a monomodal particle size distribution with lower polydispersity, and the particle size also
grows during storage, but without size focusing via precipitation, because no thick organic shells were
observed by the TEM observation.
The particle growth is often neglected for the LAL-generated metallic particles. This must be taken
into account when characterizing the particle state by TEM and before using these particles for practical
applications. If the characterization and application were conducted at different times and particle
growth had occurred, then the mismatch of the particle properties and their performances, particularly
for size-sensitive applications such as catalysis and biology, would mislead the interpretation of all of
the experimental results.

Author Contributions: D.Z. and K.S. conceived and designed the experiments; W.C. performed the experiments;
D.Z. characterized the particles and analyzed the data; S.B., J.J., and M.-R.K. contributed with the ns-LAL work;
D.Z. and K.S. wrote the paper; and all of the authors read and revised the manuscript.
Acknowledgments: We would like to thank Materials Characterization Support Unit, RIKEN CEMS for providing
access to the SEM and TEM microscopes as well as XRD and FTIR spectrometers.
Conflicts of Interest: The authors declare no conflict of interest.

103
Nanomaterials 2018, 8, 529

References
1. Zhang, D.; Gökce, B.; Barcikowski, S. Laser synthesis and processing of colloids: Fundamentals and
applications. Chem. Rev. 2017, 117, 3990–4103. [CrossRef] [PubMed]
2. Zeng, H.; Du, X.W.; Singh, S.C.; Kulinich, S.A.; Yang, S.; He, J.; Cai, W. Nanomaterials via laser
ablation/irradiation in liquid: A review. Adv. Funct. Mater. 2012, 22, 1333–1353. [CrossRef]
3. Xiao, J.; Liu, P.; Wang, C.X.; Yang, G.W. External field-assisted laser ablation in liquid: An efficient strategy
for nanocrystal synthesis and nanostructure assembly. Prog. Mater. Sci. 2017, 87, 140–220. [CrossRef]
4. Zhang, J.; Claverie, J.; Chaker, M.; Ma, D. Colloidal metal nanoparticles prepared by laser ablation and their
applications. ChemPhysChem 2017, 18, 986–1006. [CrossRef] [PubMed]
5. Zhang, J.; Chen, G.; Chaker, M.; Rosei, F.; Ma, D. Gold nanoparticle decorated ceria nanotubes with
significantly high catalytic activity for the reduction of nitrophenol and mechanism study. Appl. Catal. B
2013, 132, 107–115. [CrossRef]
6. Hebié, S.; Holade, Y.; Maximova, K.; Sentis, M.; Delaporte, P.; Kokoh, K.B.; Napporn, T.W.; Kabashin, A.V.
Advanced electrocatalysts on the basis of bare au nanomaterials for biofuel cell applications. ACS Catal.
2015, 5, 6489–6496. [CrossRef]
7. Zhang, D.; Liu, J.; Li, P.; Tian, Z.; Liang, C. Recent advances in surfactant-free, surface charged and defect-rich
catalysts developed by laser ablation and processing in liquids. ChemNanoMat 2017, 3, 512–533. [CrossRef]
8. Zhang, J.; Chaker, M.; Ma, D. Pulsed laser ablation based synthesis of colloidal metal nanoparticles for
catalytic applications. J. Colloid Interface Sci. 2017, 489, 138–149. [CrossRef] [PubMed]
9. Zhang, D.; Liu, J.; Liang, C. Perspective on how laser-ablated particles grow in liquids. Sci. China Phys.
Mech. Astron. 2017, 60, 074201. [CrossRef]
10. Jendrzej, S.; Gökce, B.; Amendola, V.; Barcikowski, S. Barrierless growth of precursor-free, ultrafast
laser-fragmented noble metal nanoparticles by colloidal atom clusters—A kinetic in situ study. J. Colloid
Interface Sci. 2016, 463, 299–307. [CrossRef] [PubMed]
11. Poletti, A.; Fracasso, G.; Conti, G.; Pilot, R.; Amendola, V. Laser generated gold nanocorals with broadband
plasmon absorption for photothermal applications. Nanoscale 2015, 7, 13702–13714. [CrossRef] [PubMed]
12. Zhou, L.; Zhang, H.; Bao, H.; Liu, G.; Li, Y.; Cai, W. Onion-structured spherical mos2 nanoparticles induced
by laser ablation in water and liquid droplets’ radial solidification/oriented growth mechanism. J. Phys.
Chem. C 2017, 121, 23233–23239. [CrossRef]
13. Kabashin, A.V.; Meunier, M. Synthesis of colloidal nanoparticles during femtosecond laser ablation of gold
in water. J. Appl. Phys. 2003, 94, 7941–7943. [CrossRef]
14. Rehbock, C.; Merk, V.; Gamrad, L.; Streubel, R.; Barcikowski, S. Size control of laser-fabricated surfactant-free
gold nanoparticles with highly diluted electrolytes and their subsequent bioconjugation. Phys. Chem. Chem. Phys.
2013, 15, 3057–3067. [CrossRef] [PubMed]
15. Liu, J.; Liang, C.; Tian, Z.; Zhang, S.; Shao, G. Spontaneous growth and chemical reduction ability of Ge
nanoparticles. Sci. Rep. 2013, 3, 1741. [CrossRef]
16. Zhang, D.; Lau, M.; Lu, S.; Barcikowski, S.; Gökce, B. Germanium sub-microspheres synthesized by
picosecond pulsed laser melting in liquids: Educt size effects. Sci. Rep. 2017, 7, 40355. [CrossRef] [PubMed]
17. Zhang, D.; Gökce, B.; Notthoff, C.; Barcikowski, S. Layered seed-growth of agge football-like microspheres
via precursor-free picosecond laser synthesis in water. Sci. Rep. 2015, 5, 13661. [CrossRef] [PubMed]
18. Zhang, D.; Ma, Z.; Spasova, M.; Yelsukova, A.E.; Lu, S.; Farle, M.; Wiedwald, U.; Gökce, B. Formation
mechanism of laser-synthesized iron-manganese alloy nanoparticles, manganese oxide nanosheets and
nanofibers. Part. Part. Syst. Charact. 2017, 34, 1600225. [CrossRef]
19. Liang, C.; Sasaki, T.; Shimizu, Y.; Koshizaki, N. Pulsed-laser ablation of mg in liquids: Surfactant-directing
nanoparticle assembly for magnesium hydroxide nanostructures. Chem. Phys. Lett. 2004, 389, 58–63.
[CrossRef]
20. Zhang, H.; Duan, G.; Li, Y.; Xu, X.; Dai, Z.; Cai, W. Leaf-like tungsten oxide nanoplatelets induced by laser
ablation in liquid and subsequent aging. Cryst. Growth Des. 2012, 12, 2646–2652. [CrossRef]
21. Niu, K.Y.; Yang, J.; Kulinich, S.A.; Sun, J.; Li, H.; Du, X.W. Morphology control of nanostructures via surface
reaction of metal nanodroplets. J. Am. Chem. Soc. 2010, 132, 9814–9819. [CrossRef] [PubMed]
22. Schaumberg, C.A.; Wollgarten, M.; Rademann, K. Metallic copper colloids by reductive laser ablation of non
metallic copper precursor suspensions. J. Phys. Chem. A 2014, 118, 8329–8337. [CrossRef] [PubMed]

104
Nanomaterials 2018, 8, 529

23. Liu, J.; Liang, C.; Zhu, X.; Lin, Y.; Zhang, H.; Wu, S. Understanding the solvent molecules induced
spontaneous growth of uncapped tellurium nanoparticles. Sci. Rep. 2016, 6, 32631. [CrossRef] [PubMed]
24. Wu, C.-H.; Chen, S.-Y.; Shen, P. Special grain boundaries of anatase nanocondensates by oriented attachment.
CrystEngComm 2014, 16, 1459–1465. [CrossRef]
25. Wang, H.; Odawara, O.; Wada, H. Facile and chemically pure preparation of YVO4 : Eu3+ colloid with novel
nanostructure via laser ablation in water. Sci. Rep. 2016, 6, 20507. [CrossRef] [PubMed]
26. Huang, C.-C.; Yeh, C.-S.; Ho, C.-J. Laser ablation synthesis of spindle-like gallium oxide hydroxide
nanoparticles with the presence of cationic cetyltrimethylammonium bromide. J. Phys. Chem. B 2004,
108, 4940–4945. [CrossRef]
27. Xia, Y.; Xiong, Y.; Lim, B.; Skrabalak, S.E. Shape-controlled synthesis of metal nanocrystals: Simple chemistry
meets complex physics? Angew. Chem. Int. Ed. 2009, 48, 60–103. [CrossRef] [PubMed]
28. Tsuji, T.; Kakita, T.; Tsuji, M. Preparation of nano-size particles of silver with femtosecond laser ablation in
water. Appl. Surf. Sci. 2003, 206, 314–320. [CrossRef]
29. Pyatenko, A.; Shimokawa, K.; Yamaguchi, M.; Nishimura, O.; Suzuki, M. Synthesis of silver nanoparticles by
laser ablation in pure water. Appl. Phys. A 2004, 79, 803–806. [CrossRef]
30. Streubel, R.; Bendt, G.; Gökce, B. Pilot-scale synthesis of metal nanoparticles by high-speed pulsed laser
ablation in liquids. Nanotechnology 2016, 27, 205602. [CrossRef] [PubMed]
31. Tilaki, R.M.; Mahdavi, S.M. Stability, size and optical properties of silver nanoparticles prepared by laser
ablation in different carrier media. Appl. Phys. A 2006, 84, 215–219. [CrossRef]
32. Tsuji, T.; Tsuji, M.; Hashimoto, S. Utilization of laser ablation in aqueous solution for observation of photoinduced
shape conversion of silver nanoparticles in citrate solutions. J. Photochem. Photobiol. A 2011, 221, 224–231.
[CrossRef]
33. Tsuji, T.; Mizuki, T.; Ozono, S.; Tsuji, M. Laser-induced silver nanocrystal formation in polyvinylpyrrolidone
solutions. J. Photochem. Photobiol. A 2009, 206, 134–139. [CrossRef]
34. Tsuji, T.; Higuchi, T.; Tsuji, M. Laser-induced structural conversions of silver nanoparticles in pure
water—Influence of laser intensity. Chem. Lett. 2005, 34, 476–477. [CrossRef]
35. Tsuji, T.; Kikuchi, M.; Kagawa, T.; Adachi, H.; Tsuji, M. Morphological changes from spherical silver
nanoparticles to cubes after laser irradiation in acetone–water solutions via spontaneous atom transportation
process. Colloids Surf. A 2017, 529, 33–37. [CrossRef]
36. Tsuji, T.; Nakanishi, M.; Mizuki, T.; Ozono, S.; Tsuji, M.; Tsuboi, Y. Preparation and shape-modification of
silver colloids by laser ablation in liquids: A brief review. Sci. Adv. Mater. 2012, 4, 391–400. [CrossRef]
37. Kőrösi, L.; Rodio, M.; Dömötör, D.; Kovács, T.; Papp, S.; Diaspro, A.; Intartaglia, R.; Beke, S. Ultra-small,
ligand-free Ag nanoparticles with high antibacterial activity prepared by pulsed laser ablation in liquid.
J. Chem. 2016, 2016, 4143560. [CrossRef]
38. Tiedemann, D.; Taylor, U.; Rehbock, C.; Jakobi, J.; Klein, S.; Kues, W.A.; Barcikowski, S.; Rath, D. Reprotoxicity
of gold, silver, and gold–silver alloy nanoparticles on mammalian gametes. Analyst 2014, 139, 931–942.
[CrossRef] [PubMed]
39. Robertson, J. Diamond-like amorphous carbon. Mater. Sci. Eng. 2002, 37, 129–281. [CrossRef]
40. Sengan, M.; Veeramuthu, D.; Veerappan, A. Photosynthesis of silver nanoparticles using durio zibethinus
aqueous extract and its application in catalytic reduction of nitroaromatics, degradation of hazardous dyes
and selective colorimetric sensing of mercury ions. Mater. Res. Bull. 2018, 100, 386–393. [CrossRef]
41. Bastús, N.G.; Merkoçi, F.; Piella, J.; Puntes, V. Synthesis of highly monodisperse citrate-stabilized silver
nanoparticles of up to 200 nm: Kinetic control and catalytic properties. Chem. Mater. 2014, 26, 2836–2846.
[CrossRef]
42. An, J.; Tang, B.; Ning, X.; Zhou, J.; Xu, S.; Zhao, B.; Xu, W.; Corredor, C.; Lombardi, J.R. Photoinduced
shape evolution: From triangular to hexagonal silver nanoplates. J. Phys. Chem. C 2007, 111, 18055–18059.
[CrossRef]
43. Yu, B.; Shi, Y.; Yuan, B.; Qiu, S.; Xing, W.; Hu, W.; Song, L.; Lo, S.; Hu, Y. Enhanced thermal and flame
retardant properties of flame-retardant-wrapped graphene/epoxy resin nanocomposites. J. Mater. Chem. A
2015, 3, 8034–8044. [CrossRef]
44. Mansur, H.S.; Sadahira, C.M.; Souza, A.N.; Mansur, A.A.P. Ftir spectroscopy characterization of poly
(vinyl alcohol) hydrogel with different hydrolysis degree and chemically crosslinked with glutaraldehyde.
Mater. Sci. Eng. C 2008, 28, 539–548. [CrossRef]

105
Nanomaterials 2018, 8, 529

45. Yajima, T.; Uchida, H.; Watanabe, M. In-situ ATR-FTIR spectroscopic study of electro-oxidation of methanol
and adsorbed CO at Pt–Ru alloy. J. Phys. Chem. B 2004, 108, 2654–2659. [CrossRef]
46. Pritchard, J.; Catterick, T.; Gupta, R.K. Infrared spectroscopy of chemisorbed carbon monoxide on copper.
Surf. Sci. 1975, 53, 1–20. [CrossRef]
47. Dong, C.; Wirasaputra, A.; Luo, Q.; Liu, S.; Yuan, Y.; Zhao, J.; Fu, Y. Intrinsic flame-retardant and thermally
stable epoxy endowed by a highly efficient, multifunctional curing agent. Materials 2016, 9, 1008. [CrossRef]
[PubMed]
48. Liang, B.; Zhou, M.; Andrews, L. Reactions of laser-ablated Ni, Pd, and Pt atoms with carbon monoxide:
Matrix infrared spectra and density functional calculations on M(CO)n (n = 1–4), M(CO)n − (n = 1–3),
and M(CO)n + (n = 1–2), (M = Ni, Pd, Pt). J. Phys. Chem. A 2000, 104, 3905–3914. [CrossRef]
49. Ramis, G.; Busca, G.; Lorenzelli, V. Low-temperature CO2 adsorption on metal oxides: Spectroscopic
characterization of some weakly adsorbed species. Mater. Chem. Phys. 1991, 29, 425–435. [CrossRef]
50. Liang, B.; Andrews, L. Reactions of laser-ablated Ag and Au atoms with carbon monoxide: Matrix infrared
spectra and density functional calculations on Ag(CO)n (n = 2, 3), Au(CO)n (n = 1, 2) and M(CO)n + (n = 1–4;
M = Ag, Au). J. Phys. Chem. A 2000, 104, 9156–9164. [CrossRef]
51. Mikhlin, Y.L.; Vorobyev, S.A.; Saikova, S.V.; Vishnyakova, E.A.; Romanchenko, A.S.; Zharkov, S.M.;
Larichev, Y.V. On the nature of citrate-derived surface species on Ag nanoparticles: Insights from X-ray
photoelectron spectroscopy. Appl. Surf. Sci. 2018, 427, 687–694. [CrossRef]
52. Escobar-Alarcón, L.; Espinosa-Pesqueira, M.E.; Solis-Casados, D.A.; Gonzalo, J.; Solis, J.; Martinez-Orts, M.;
Haro-Poniatowski, E. Two-dimensional carbon nanostructures obtained by laser ablation in liquid: Effect of
an ultrasonic field. Appl. Phys. A 2018, 124, 141. [CrossRef]
53. Van’t Zand, D.D.; Nachev, P.; Rosenfeld, R.; Wagener, P.; Pich, A.; Klee, D.; Barcikowski, S. Nanocomposite
fibre fabrication via in situ monomer grafting and bonding on laser-generated nanoparticles. J. Laser
Micro/Nanoeng. 2012, 7, 21–27. [CrossRef]
54. Shih, C.-Y.; Streubel, R.; Heberle, J.; Letzel, A.; Shugaev, M.; Wu, C.; Schmidt, M.; Gokce, B.; Barcikowski, S.;
Zhigilei, L. Two mechanisms of nanoparticle generation in picosecond laser ablation in liquids: The origin of
the bimodal size distribution. Nanoscale 2018, 10, 6900–6910. [CrossRef] [PubMed]
55. Zhang, D.; Chen, F.; Fang, G.; Yang, Q.; Xie, D.; Qiao, G.; Li, W.; Si, J.; Hou, X. Wetting characteristics on
hierarchical structures patterned by a femtosecond laser. J. Micromech. Microeng. 2010, 20, 075029. [CrossRef]
56. Zhang, D.; Chen, F.; Yang, Q.; Si, J.; Hou, X. Mutual wetting transition between isotropic and anisotropic on
directional structures fabricated by femotosecond laser. Soft Matter 2011, 7, 8337–8342. [CrossRef]
57. Zhang, D.; Chen, F.; Yang, Q.; Yong, J.; Bian, H.; Ou, Y.; Si, J.; Meng, X.; Hou, X. A simple way to achieve
pattern-dependent tunable adhesion in superhydrophobic surfaces by a femtosecond laser. ACS Appl.
Mater. Interfaces 2012, 4, 4905–4912. [CrossRef] [PubMed]

© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

106
nanomaterials

Article
Magnetic Fe@FeOx, Fe@C and α-Fe2O3 Single-Crystal
Nanoblends Synthesized by Femtosecond Laser
Ablation of Fe in Acetone
Dongshi Zhang 1,† , Wonsuk Choi 1,2,3,† , Yugo Oshima 4 , Ulf Wiedwald 5 , Sung-Hak Cho 2,6 ,
Hsiu-Pen Lin 7,8 , Yaw Kuen Li 8 , Yoshihiro Ito 7,9 and Koji Sugioka 1, *
1 RIKEN Center for Advanced Photonics, 2-1 Hirosawa, Wako, Saitama 351-0198, Japan;
[email protected] (D.Z.); [email protected] (W.C.)
2 Department of Nano-Mechatronics, Korea University of Science and Technology (UST), 217 Gajeong-Ro,
Yuseong-Gu, Daejeon 34113, Korea; [email protected]
3 Department of Nano-Manufacturing Technology, Korea Institute of Machinery and Material (KIMM),
156 Gajeongbuk-Ro, Yuseong-Gu, Daejeon 34103, Korea
4 Condensed Molecular Materials Laboratory, RIKEN Cluster for Pioneering Research, 2-1 Hirosawa, Wako,
Saitama 351-0198, Japan; [email protected]
5 Faculty of Physics and Center for Nanointegration Duisburg-Essen (CENIDE),
University of Duisburg-Essen, 47057 Duisburg, Germany; [email protected]
6 Department of Laser & Electron Beam Application, Korea Institute of Machinery and Material (KIMM),
156 Gajeongbuk-Ro, Yuseong-Gu, Daejeon 34103, Korea
7 Emergent Bioengineering Materials Research Team, RIKEN Center for Emergent Matter Science,
2-1 Hirosawa, Wako, Saitama 351-0198, Japan; [email protected] (H-P.L.); [email protected] (Y.I.)
8 Department of Applied Chemistry, National Chiao Tung University, Science Building 2, 1001 Ta Hsueh Road,
Hsinchu 300, Taiwan; [email protected]
9 Nano Medical Engineering Laboratory, RIKEN Cluster for Pioneering Research, 2-1 Hirosawa, Wako,
Saitama 351-0193, Japan
* Correspondence: [email protected]; Tel.: +81-(0)48-467-9495
† These authors contributed equally to this work.

Received: 24 July 2018; Accepted: 18 August 2018; Published: 20 August 2018

Abstract: There are few reports on zero-field-cooled (ZFC) magnetization measurements for Fe@FeOx
or FeOx particles synthesized by laser ablation in liquids (LAL) of Fe, and the minimum blocking
temperature (TB ) of 120 K reported so far is still much higher than those of their counterparts
synthesized by chemical methods. In this work, the minimum blocking temperature was lowered to
52 K for 4–5 nm α-Fe2 O3 particles synthesized by femtosecond laser ablation of Fe in acetone.
The effective magnetic anisotropy energy density (Keff ) is calculated to be 2.7–5.4 × 105 J/m3 ,
further extending the Keff values for smaller hematite particles synthesized by different methods.
Large amorphous-Fe@α-Fe2 O3 and amorphous-Fe@C particles of 10–100 nm in diameter display a soft
magnetic behavior with saturation magnetization (Ms ) and coercivities (Hc ) values of 72.5 emu/g
and 160 Oe at 5 K and 61.9 emu/g and 70 Oe at 300 K, respectively, which mainly stem from
the magnetism of amorphous Fe cores. Generally, the nanoparticles obtained by LAL are either
amorphous or polycrystalline, seldom in a single-crystalline state. This work also demonstrates
the possibility of synthesizing single-crystalline α-Fe2 O3 hematite crystals of several nanometers
with (104), (113), (116) or (214) crystallographic orientations, which were produced simultaneously
with other products including carbon encapsulated amorphous Fe (a-Fe@C) and Fe@FeOx core-shell
particles by LAL in one step. Finally, the formation mechanisms for these nanomaterials are proposed
and the key factors in series events of LAL are discussed.

Keywords: hematite α-Fe2O3; core-shell; blocking temperature; superparamagnetism; laser ablation in liquids;
femtosecond laser; single-crystalline

Nanomaterials 2018, 8, 631; doi:10.3390/nano8080631 107 www.mdpi.com/journal/nanomaterials


Nanomaterials 2018, 8, 631

1. Introduction
The newly emerged technique of laser ablation in liquids (LAL) [1–5] has proven to be valid
for the synthesis of a large variety of colloids resulting from material removal of the substances [6].
With specific targets, simply changing the liquids for LAL allows the easy alteration of colloidal
properties such as sizes [7,8] and phases [9–11]. For the synthesis of magnetic particles by LAL,
the most frequently investigated material is Fe, whose results have turned out to be very diverse
depending on experimental conditions [2], in which nanosecond (ns) lasers were typically used.
For example, Vahabzadeh and Torkamany made use of ns lasers at fundamental and 2nd harmonic
wavelengths (1064 nm and 532 nm) to ablate Fe in water for the formation of Fe3 O4 and FeO
particles, whose saturation magnetizations (Ms ) and coercivity (Hc ) were 22.5 emu/g and 11.5 Oe
and 14.8 emu/g and 22 Oe [12], respectively. These values were much lower than those of bulk
magnetite (Ms = 92 emu/g and Hc = 500–800 Oe). Zeng et al. prepared FeO nanoparticles by ns
laser fragmentation in liquid (LFL) of Fe using a water/poly(vinyl pyrrolidone) (PVP) solution [13].
Amendola et al. obtained FeOx (Fe3 O4 , FeO, α-Fe) nanoparticles with an Ms of 100 emu/g by ns LAL
in water [14]. Pandey et al. found that the ns LAL of commercial Fe2 O3 powders in doubly distilled
water improved the hematite particle crystallinity and increased Ms from 0.024 to 3.41 emu/g [15].
Svetlichnyi et al. found that the Hc of the FeOx (Fe, Fe2 O3 and Fe3 O4 ) nanoparticles synthesized by ns
LAL of Fe in water increased from 144 Oe to 370 Oe when the measurement temperature was reduced
from 300 K to 77 K [16]. Ismail et al. showed that the Ms values of 16.3–20.3 emu/g for the magnetic
iron oxide (Fe3 O4 , α-Fe2 O3 , FeO and ε-Fe2 O3 ) nanoparticles synthesized by ns LAL in SDS aqueous
solution were larger than 13.8–16.2 emu/g for FeOx (Fe3 O4 , α-Fe2 O3 , ε-Fe2 O3 ) particles synthesized
by ns LAL in dimethylformamide (DMF) at the same laser energies [17]. Meanwhile, Kanitz et al.
employed a femtosecond (fs) laser for LAL of Fe and reported that the products changed depending
on the adopted solutions: α-iron, wüstite and magnetite were synthesized in water, α-iron, cementite
and FeOx in methanol, amorphous-Fe and α-Fe mixture in ethanol and acetone, and amorphous-Fe@C
core-shell particles in toluene [18]. Their Ms and Hc values were measured to be 23, 80, 60, 67 and
14 emu/g, and 77, 92, 65, 56, and 52 Oe at 300 K, respectively. Santillán et al. synthesized FeOx (Fe3 O4 ,
γ-Fe2 O3 or α-Fe) by fs laser (120 fs, 1 kHz, 800 nm) ablation of Fe in water, by which Ms , number density,
magnetic radius and total radius of 49.3 emu/g, 2.9 × 1018 , 1.1 nm, 1.9 nm, respectively, were obtained,
which were different from the magnetic properties (26.7 emu/g, 3468 μB and 0.7 × 1018 , 1.5 nm, 3.2 nm)
of the particles obtained by LAL in a trisodium citrate aqueous solution [19]. Most of these works
mainly focused on the hysteresis curves to reveal the dependence of magnetic properties on the phases
of LAL-prepared FeOx or Fe/FeOx particles. Little attention has been paid to the zero-field-cooled
(ZFC) and field-cooled (FC) curves including the information about blocking temperature (TB ) which is
closely related to the product of particle size and magnetic anisotropy energy density essentially defining
the energy barrier between two easy axes of magnetization [20].
Amendola et al. observed a blocking temperature TB = 200 K of FeOx particles synthesized by
the ns-LAL of Fe in water [14]. Franzel et al. reported that Fe3 O4 and Fe3 C synthesized by picosecond
(ps) LAL of Fe in ethanol had a blocking temperature TB = 120 K [21], much lower that of FeOx particles
synthesized by ns-LAL, which was due to the generation of higher ratios of smaller particles by ps-LAL.
As is well known, a low TB value (<100 K) is a good indicator for superparamagnetism arising from
small particles [22,23]. The endowment of superparamagnetism to the as-prepared magnetic particles
often requires the particle size to be around or less than 10 nm [24]. To date, the minimum blocking
temperature of LAL-generated Fe@FeOx or FeOx particles is still above 100 K [21]. Due to the fact
that smaller FeOx particles often possess lower blocking temperatures [20], the synthesis of ultrasmall
Fe@FeOx and FeOx particles by LAL is essential to lower the blocking temperature below 100 K.
To this end, fs-LAL is a better choice than ps- and ns-LAL because of the phase/Coloumb explosion
mechanism for fs-LAL rather than the thermal ablation mechanism for ns-LAL [25].
Despite many reports on the fs-LAL of Fe and studies on the magnetic properties (Ms , Mr
and Hc ) of the products [11,18], the ZFC/FC curves were not measured. To fill this gap and

108
Nanomaterials 2018, 8, 631

to further lower the blocking temperature of Fe@FeOx possessing superparamagnetic properties,


the fs-LAL of Fe in acetone was performed. XRD, high-resolution transition electron microscropy
(HRTEM), energy-dispersive X-ray (EDX), selected area electron diffraction (SAED), fast Fourier
transform (FFT), X-ray photoelectron spectroscopy (XPS) and Raman characterizations were performed
to clarify the composition of the as-prepared particles. TEM analysis was performed to display
the particle morphologies and calculate the size distribution of the colloids. Both ZFC/FC and hysteresis
curves of the synthesized magnetic particles were measured, from which TB , Ms , Mr and Hc values
were determined.

2. Materials and Methods


Colloids were synthesized by laser ablation of a Fe sheet (99.45 wt % Fe, 0.42 wt % O, 0.13 wt
% C) using a fs laser system (FGPA μJewel D-1000-UG3, IMRA America Inc., Ann Arbor, MI, USA).
The pulse duration, wavelength and repetition rate of the laser system were 457 fs, 1045 nm and 100 kHz,
respectively. An Fe sheet with dimensions of 20 mm × 20 mm × 1 mm was placed inside a glass
container and then immersed in 8 mL acetone for LAL. The liquid thickness above the target surface
was kept at 5 mm. Then, a fs laser beam was focused on the Fe sheet surface by a 20× objective lens
(numerical aperture (NA) = 0.4, Mitutoyo, Kawasaki, Japan) and scanned over an area of 3.5 × 3.5 mm2
using the scan method described in [26–28] with a line interval of 5 μm and a scan speed of 1 mm/s to
ablate the Fe sheet. The ablation process lasted around 1 h. The average laser power was set to 600 mW.
The spot size was 26 μm. The peak irradiance and laser fluence were calculated to be 1.13 × 109 W/m2
and 113 J/cm2 , respectively.
The colloids were directly deposited onto TEM grids (EMJapan, U1015, Tokyo, Japan, 20 nm thick
carbon films on copper grids) after LAL without any pre-treatment and then characterized using TEM
(Jeol, JEM-1230, Tokyo, Japan) operating at 80 kV. HRTEM and STEM-EELS (scanning transmission
electron microscopy–electron energy loss spectroscopy) were performed with a JEM-ARM200F
(Jeol, Tokyo, Japan) equipped with third-order aberration correctors for both illuminating and imaging
lens systems operated at 200 kV. For XRD and magnetic property measurements, the colloids with liquids
were centrifuged by a centrifuge (Eppendorf, Centrifuge 5430, Hamburg, Germany) at a rotation speed of
14,000 rpm for 10 min. The precipitated particles were then collected in a cuvette and dried in a freeze
dryer (Rikakikai, S-1000, Eyela, Tokyo, Japan). The dried particles were deposited on an amorphous
glass plate (10 mm × 10 mm × 1 mm) for XRD, XPS and Raman characterizations. The composition of
the particles was analysed using XRD (Rigaku, CuKα radiation (40 kV-30 mA), SmartLab-R 3kW, Tokyo,
Japan). The surface chemistry of the particles was analysed by XPS (Thermo Scientific, ESCALAB 250,
Tokyo, Japan) and Raman spectroscopy (LabRAM, Hiriba, He-Ne laser, 632 nm, 0.686 mW, Tokyo, Japan).
A zeta-potential and particle size analyzer (ELSZ-2PL, Photal, Osaka, Japan) was used to measure the zeta
potential of the fresh colloid and the colloid stored after 3 weeks. UV-vis spectroscopy (Shimadzu, UV-3600
Plus, Tokyo, Japan) was used to measure the absorption spectra of colloids.
Magnetic properties of the particles were measured in He gas atmosphere using
the superconducting quantum interference device (SQUID) magnetometer (Quantum Design, MPMS
XL7, San Diego, USA). The dried 26.5 mg Fe@α-Fe2 O3 particle powder was filled into a capsule,
which was then placed into the magnetometer. Zero-field-cooled (ZFC) magnetization was measured
by cooling samples in a zero magnetic field and then increasing the temperature from 5 K to 300 K
with magnetization-temperature data recorded every 5 K at an applied field of 50 Oe. Field-cooled
(FC) curves were recorded by cooling the samples from 300 K to 5 K with a constant field of 50 Oe.
The field dependence of the magnetization (hysteresis loop) was recorded up to ±70 kOe at T = 5 K
and ±10 kOe at T = 300 K, respectively.

109
Nanomaterials 2018, 8, 631

3. Results

3.1. Material Property


After drying on TEM grids, the particles synthesized in acetone form a particle network
(Figure 1a–c), connected by a large amount of small clusters, which is a typical phenomenon after
ferro-fluidic colloid drying [29–34]. The average size of the particles is estimated to be 5–6 nm
(Figure 1d). Small particles with sizes of less than 10 nm occupy more than an 87% number frequency
of all of the particles. In particular, small particles with a ~90% number frequency of 1–10 nm are in the
majority, with the highest number frequency at 4–5 nm (Figure 1f–i). The large particles are in the form
of core-shell particles with a shell thickness of ca. 6 nm (Figure 1c). No crystalline peak was observed
in the XRD spectrum (Figure 1e), similar to the case of the nanomaterials obtained by LAL of Cu in
acetone [35]. This is either due to the low amount of particle powders used for the XRD spectrum or
due to the too-small crystallites of the particles [35]. Actually, we used several mg particles for XRD
characterization. With the same amount, a well featured XRD spectrum of Ag particles synthesized by
LAL of Ag in acetone has been observed [36], which indicates that the particles synthesized by LAL of
Fe in acetone are in very low crystallinity.

Figure 1. (a–c) TEM images of the particles synthesized by the laser ablation of Fe in acetone at 400 mW.
(d) Size distribution of the synthesized particles. The inset figure shows the detailed size distribution
in the range of 0–10 nm. (e) XRD pattern of the synthesized particles, where no peaks were detected,
probably due to the low crystallinity of the particles and a large amount of carbon clusters. (f,g) and
(h,i) black field and white field scanning transmission electron microscopy (STEM) images of small
particles, respectively.

To confirm the compositions of small clusters and core-shell particles, the distribution of Fe, C,
O elements in the nanoblends was analyzed by EDX as shown in Figures 2 and 3. As indicated by
TEM image (Figure 2a) and the Fe and O elemental distributions (Figure 2c,e), the small clusters
were identified as FeOx . Besides this, a certain amount of carbon was also detected (Figure 2b,d).
However, given that the particles were deposited on the carbon membrane of a TEM grid, it is difficult
to differentiate whether the detected carbon comes from the particles or not.
Apparent evidence for the generation of carbon during LAL was witnessed by HRTEM
characterization (Figure 3a,f,g) and EDX analysis (Figure 3b–e). Both amorphous carbon (Figure 3b,d,f)
and onion-like carbon (Figure 3b,d,f,g) were discovered, which encapsulated amorphous Fe particles
to form the amorphous-Fe@carbon (a-Fe@C) core-shell particles (a representative particle is shown
with an arrow marked in Figure 3b). Facilitated by the adhesion of different carbon shells, a-Fe@C
core-shell particles gradually evolve into a particle network (Figures 1a–c and 3g). This phenomenon is

110
Nanomaterials 2018, 8, 631

consistent with previous reports that LAL in organic solvents often causes the decomposition of solvent
molecules and results in the formation of carbon-encapsulated particle networks [10,37]. Amorphous
FeOx clusters (Figure 3a–e) were also generated, which surrounded big core-shell particles to facilitate
the formation of particle network. Figures 2a–e and 3a–e also indicate that, besides a-Fe@C core-shell
particles, large Fe@FeOx core-shell particles with diameters of tens of nm are produced by LAL of Fe
in acetone.

Figure 2. (a) TEM image and (b–e) EDX mapping of small clusters. (b) TEM image of mixed elements
of (c) Fe, (d) C and (e) O images.

Figure 3. (a) TEM image and (b–e) EDX mapping of a Fe@FeOx core-shell particle. (b) TEM image of
mixed elements of (c) Fe, (d) C and (e) O images. (f,g) TEM images of the Fe@C particles.

To better understand the compositions of Fe@FeOx core-shell particles, a core-shell particle was
selected as the representative particle for HRTEM, SAED and FFT characterizations, as shown in
Figure 4a–c. To further clarify the crystallinity in different regions, HRTEM images of seven domains
of FeOx shells and one larger domain of an Fe core were displayed in Figure 4d–k. The SAED pattern
of the core-shell particle indicates that the core-shell particle has a low crystallinity since only two
diffraction rings were observed (Figure 4b), which fits well with the (104) and (214) planes of α-Fe2 O3
(ICSD No. 01-089-0597). FFT analysis gives more information about the crystallinity of the core-shell
particle, which indicates that two more planes of the (113) and (116) planes of α-Fe2 O3 are also present.
Figure 4d–h display the HRTEM images of different crystal domains in the FeOx shell which possess

111
Nanomaterials 2018, 8, 631

(104), (113), (116) and (214) planes of α-Fe2 O3 with interplanar distances of 0.270, 0.217, 0.170 and
0.150 nm, respectively. Besides α-Fe2 O3 , another diffraction ring belonging to a crystal plane with an
interplanar distance of 0.300 nm was also detected, which can be assigned to the (220) plane of Fe3 O4
(ICSD No. 01-089-0950). It is noteworthy that (i) the crystals in the FeOx shell are mainly single crystalline;
and (ii) the quality of the single-crystallinity is not particularly good because many defects among
the crystal planes are obvious (Figure 4d–g,l–o). One domain of the FeOx shell is almost completely
amorphous (Figure 4j). As concluded from Figure 4d–h, the FeOx shell is composed of many α-Fe2 O3
single-crystalline crystals with different crystallographic orientations. Regarding the Fe core, it is totally
amorphous (Figure 4k). That is why no peak was detected during XRD characterization (Figure 1e).
Regarding the dominant small particles, four crystalline domains outside the Fe@FeOx particle are shown
in Figure 4l–o. The interplanar distances of 0.217, 0.217, 0.217 and 0.270 nm of these crystalline domains
are well indexed to (113), (113), (113) and (104) planes of α-Fe2 O3 . As indicated by HRTEM analysis,
it is clear that single-crystalline α-Fe2 O3 nanocrystals with different crystallographic orientations are
abundant in the nanoblends.

Figure 4. (a) High-resolution transition electron microscropy (HRTEM) image of a Fe@FeOx core-shell
particle, (b) selected area electron diffraction (SAED) and (c) FFT analysis of the core-shell particle.
HRTEM images of (d–j) α-Fe2 O3 shells, (i) Fe3 O4 shell and (j) amorphous FeOx shell and (k) amorphous
Fe core, respectively. (l–o) Single-crystalline α-Fe2 O3 crystals. Inset images in (k,n) show the FFT
image indicating an amorphous structure and the inverse FFT image of the structure showing clearer
crystal planes, respectively. The scale bar in the inset image of (n) is 0.5 nm.

112
Nanomaterials 2018, 8, 631

Raman spectroscopy shows seven peaks at 220.9, 239.5, 286.0, 401.4, 495.9, 606.7, 659.4, 812.0, 1050,
1099.4, 1297.7 and 1603.0 cm−1 (Figure 5) associated with α-Fe2 O3 , in accordance with the conclusion from
HRTEM analysis (Figure 4) that α-Fe2 O3 is the main crystalline product. Besides the peaks corresponding
to α-Fe2 O3 , another small peak is also observed at 1584.7 cm−1 , which can be assigned to the G-band of
carbon and therefore indicates the presence of a large amount of carbon in the particles, in accordance with
the HRTEM images shown in Figure 3f–g and XPS analysis which show that the outermost 5 nm-thick
surfaces of the as-prepared particles are composed of 72.22% C, 1.29% Fe and 26.49% O. The high C ratio
and low Fe and O ratios suggest that a higher ratio of carbon/carbon-byproduct clusters are generated by
the laser-induced decomposition of acetone molecues. High-resolution XPS Fe 2p, O 1s and C 1s spectra
are shown in Figure 6. After peak fitting, the ratio of sp2 /sp3 -C was calculated to be 0.89. The sp2 and
sp3 carbons correspond to an ordered graphite (sp2 ) structure and disordered graphite layers (e.g., soot,
chars, glassy carbon, and evaporated amorphous carbon [36,38]), respectively. Thus, more than half of
the C that precipitates on Fe@Fe2 O3 particles is crystalline because the sp2 /sp3 ratio is less than 1, in
accordance with the HRTEM images shown in Figure 3f–g, where onion-like carbons with some defects
appear as shells to embed a-Fe particles inside. When both sp2 and sp3 C states are mixed in particles,
diamond-like carbon (DLC) [39] structures are considered to be generated. DLC structures should be
a typical product of LAL in organic solvents since they were also observed from other nanomaterials
obtained by the LAL of different metals (e.g., Ag [36], Mo [39], Ti, [40], Ta [41], Nb [41], Hf [41], Mo [41]
and Co [37]) in organic solvents. Peak fitting of C 1s (Figure 6c) shows that 15.91% and 21.90% of the
carbon have C=O and C–O bonding, respectively, which is due to adsorbed acetone molecules and their
decomposition byproducts. From the XPS Fe 2p spectrum (Figure 6a), only the binding energies that
correspond to Fe3+ (711.2 eV and 724.9 eV) and Fe0 (707.3 eV and 720.1 eV) are observed, which come
from α-Fe2 O3 and a-Fe, respectively. Because of abundant C–O and C=O bindings on the particle surfaces,
only a small amount of Fe-O binding can be deconvoluted from the O 1s spectrum (Figure 6b). Both XPS
and Raman spectra support the conclusion from HRTEM analysis that α-Fe2 O3 is the dominant phase of
the crystalline particles.

Figure 5. Raman spectrum of the particles obtained by the laser ablation in liquids (LAL) of Fe in acetone.

113
Nanomaterials 2018, 8, 631

Figure 6. High resolution XPS (a) Fe 2p, (b) O 1s and (c) C 1s spectra from the colloids synthesized by
laser ablation of Fe in acetone.

The synthesized Fe@α-Fe2 O3 and α-Fe2 O3 particles were unstable with gradual particle precipitation
at the bottom of the glass container (right optical image inset in Figure 7a) during the colloid storage.
As a result, the absorbance spectrum of the colloid downshifted (Figure 7a) and the optical transparency
of the colloid increased after 3-week storage (left and middle optical images inset in Figure 7a).
Additionally, the zeta potential values of the colloid decreased from −34.88 mV to −27.95 mV after
3-week storage, which indicates the decreased stability of the colloid. The zeta potential is indicative
of the difference in the electric potentials between the charges of the species which strongly adsorb on
the particle surface and those (with the opposite sign) of the diffuse layer in the dispersing medium [42].
It is often considered that the colloids with a zeta potential value smaller than −30 mV or larger than
30 mV are stable, while those with zeta potential values in the range of −30~30 mV are unstable [3].
Hence, in principle, one would expect that the stable particles with larger charges remain well dispersed
while those with lower charges precipitate. However, in our case, the zeta potential value of the colloids
decreased over time, which indicated that the gradual aggregation of nanoparticles occurred during
colloidal storage. The magnetic properties among magnetic particles and the “capture” behavior of
both carbon shells (Figure 3a) and free carbon clusters [36] cause colloidal aggregation and precipitation
during storage. Considering the excellent long-term (six-month) stability of Ag colloid produced by
LAL in acetone [36], it is highly possible that the magnetostatic interaction among magnetic particles is
the main reason to be responsible for the colloidal aggregation and precipitation.

Figure 7. (a) Absorption spectra for the fresh colloid (red curve) synthesized by laser ablation of Fe in
acetone at 600 mW and (b) the colloid stored for 3 weeks (black curve), respectively. Inset images in (a) are
optical images of the fresh colloid (left) and the colloid stored for 3 weeks (middle), where the precipitation
of the colloid (as indicated by white arrows in the right optical image) causes the downshift of
the absorbance spectra. (b) Zeta potential curves of the fresh colloid and the colloid stored for 3 weeks.

Compared with the techniques of laser target evaporation in gases and laser ablation in air
whereby metal-oxide [43,44] is produced, LAL is better at the synthesis of Fe@C and Fe@FeOx core-shell
particles and further indicates a new way to synthesize single-crystalline iron oxide particles.

114
Nanomaterials 2018, 8, 631

3.2. Magnetic Properties


The magnetic properties of the Fe@α-Fe2 O3 particles prepared in acetone are presented in
Figure 8. To reveal the relationship between magnetization and temperature, ZFC and FC curves of
the as-prepared nanomaterials were measured. The ZFC curve in Figure 8a shows a broad maximum
peaking at 52 K, followed by a plateau and a subsequent gradual increase from 250–300 K suggesting
two different fractions of particles. The FC branch exhibits an almost linear increase of the magnetic
moment with decreasing temperature from 300 K to 5 K. Figure 8b,c show the magnetic hysteresis
loops at 5 K and 300 K with saturation magnetization MS = 72.5 emu/g and 61.9 emu/g, respectively.
The magnification around zero field delivers coercivities of HC = 160 Oe at 5 K and 70 Oe at 300 K.

Figure 8. Magnetic characterization of the Fe@α-Fe2 O3 particles synthesized by laser ablation in


acetone. (a) ZFC/FC curves in 50 Oe, (b,c) hysteresis curves at 5 and 300 K. (d) Magnified hysteresis
curves at 5 and 300 K.

The structural and morphological studies above reveal a mixture of Fe@α-Fe2 O3 , Fe@C, and α-Fe2 O3
particles (Figures 2–4) which can be split into two size regimes. The sizes of large Fe@α-Fe2 O3 and
Fe@C core-shell particles with a number frequency of about 10% are in the range of 10–100 nm,
with the maximum diameter of the distribution being 30 nm (Figure 1d). α-Fe2 O3 particles are
significantly smaller, with a medium size of 4–5 nm (Figure 1d), at a number frequency of about 90%.
For the magnetometry of magnetic powders, magnetization is the ratio of the magnetic moment with
respect to mass of particles in different size regimes. Taking identical mass densities as a rough estimate,
the relative mass fraction of the large particles is calculated to be 87%–93%. Further consideration of
the significantly smaller magnetization of α-Fe2 O3 of 1 emu/g as compared to amorphous Fe with MS
> 100 emu/g [45] further reduces the relative magnetic signal from α-Fe2 O3 . This means that the vast
majority of the magnetic signal stems from the larger particles while the smaller are expected contributing
less than 1% in the saturated state.
Therefore, the saturation magnetization of MS = 72.5 emu/g and 61.9 emu/g at 5 K and 300 K,
respectively, is considered to mainly stem from the amorphous Fe cores of large a-Fe@FeOx core-shell
particles. This deduction is also helpful in understanding the magnetic properties (MS = 67 emu/g at
300 K) of a-Fe@C and Fe3 O4 nanoblends synthesized by fs-LAL of Fe in acetone under other conditions
(35 fs, 800 nm, 5 kHz, 800 μJ/pulse) [18]. The remanence to saturation ratio (Mr /Ms ), also called

115
Nanomaterials 2018, 8, 631

the saturation magnetization ratio, was calculated to be ca. 0.1, which was smaller than that (0.5) expected
theoretically for randomly oriented single domain grains [46], which indicates the presence of a significant
amount of superparamagnetic small particles, domain walls in large particles, and the occurrence of
antiferromagnetic interactions [47]. Here, all the above may add to an overall low remanence. It is well
known that α-Fe2 O3 is weakly ferromagnetic or antiferromagnetic [48], and its presence as the shell
material endows antiferromagnetic properties to the Fe@α-Fe2 O3 particles while the small α-Fe2 O3
particles are expected to be superparamagnetic at 300 K and thermally blocked at 5 K (see discussion
below). The interfacial magnetic interactions between ferromagnetic cores and antiferromagnetic
shells [49,50] may also endow high orbital magnetic moment to LAL-synthesized a-Fe@α-Fe2 O3 particles.
For the metallic a-Fe core, Grinstaff et al. have confirmed that glassy a-Fe is a soft ferromagnetic material
with MS = 152 emu/g and Hc = 160 Oe at T = 5 K [45]. While Hc fits well to the present results,
the lower MS can be explained by the mixture of amorphous Fe, the weakly ferromagnetic α-Fe2 O3 ,
and the unknown but significant amount of C in the sample. Thus, the remanent magnetization mainly
originates from large Fe@α-Fe2 O3 core-shell ferromagnetic particles with multi-domains, which cannot
rapidly demagnetize by domain formation in the absence of an applied field.
A more interesting phenomenon is the magnetic signature of small α-Fe2 O3 particles, which results
in different ZFC/FC curves as compared to those of a-Fe@C particles synthesized by fs-LAL of Fe in
acetone under other conditions [18]. The broad peak with a maximum at 52 K is ascribed to the blocking
behavior of α-Fe2 O3 particles on top of an almost constant signal in the interval of 5–250 K arising from
the larger Fe@α-Fe2 O3 particles. The smallest particles of the Fe@α-Fe2 O3 particles gradually cross their
blocking temperature TB with the temperature increasing above 250 K. At 300 K, however, only a minor
fraction of Fe@α-Fe2 O3 is superparamagnetic, explaining the monotonous increase in the FC branch;
moreover, the interparticle interactions [51] among LAL-generated particles [43] of the magnetic core-shell
particles are supposedly not strong enough to cause collective magnetic freezing to enter a spin-glass state.
We do not observe any sharp change of the magnetization, which indicates the absence of
the Morin transition in α-Fe2 O3 . Previous reports have shown that in small α-Fe2 O3 particles with
diameters below 19 nm (cf. Figure 1a–d), the Morin transition occurring in bulk α-Fe2 O3 is smeared
out over a wide temperature range or even completely suppressed [52,53]. This leads to weakly
ferromagnetic α-Fe2 O3 for all considered temperatures. The blocking behavior with TB = 52 K
at the peak position can be translated to an effective magnetic anisotropy energy density Keff via
Keff ·V = 25 kB TB with V the particle volume and kB Boltzmann’s constant. The factor of 25 is the
natural logarithm of the product of the measurement time window of SQUID magnetometry (10 s)
and the intrinsic attempt frequency of about 1010 Hz [53]. For 4–5 nm hematite nanospheres, as in
our cases, Keff is calculated to be 2.7–5.4 × 105 J/m3 . Bödker et al. extracted an energy barrier of
300–600 K for 16 nm which translates to Keff = 0.5–1.1 × 105 J/m3 [53] and a strongly increasing Keff
when the particle size is reduced, reaching a maximum value of 2.4 × 105 J/m3 for 5.9 nm particles
at the smallest investigated diameter [54]. In this light, the obtained results for the 4–5 nm α-Fe2 O3
particles convincingly extend the size dependence to smaller diameters.
Despite a very broad size distribution, the α-Fe2 O3 particles synthesized by fs-LAL possess the lowest
blocking temperature of 52 K among all α-Fe2 O3 particles synthesized by LAL [2]. The blocking
temperature of FeOx particles synthesized by the ns-LAL of Fe in water was ca. 220 K, which corresponds
to the particle size of 15 nm [14]. The ps-LAL of Fe in ethanol gave rise to the formation of Fe3 O4 /Fe3 C
mixture colloids which had a bimodal size distribution with maxima at ca. 3 nm and ca. 12 nm [21].
Because of the generation of a greater amount of small particles by ps-LAL, the blocking temperature
down-shifted to 120 K when the applied field was 50 Oe [21]. According to the previously reported
relationship between the blocking temperatures and particle sizes [20], it is estimated that the average size
of Fe3 O4 /Fe3 C mixture colloid obtained by ps-LAL in ethanol was ca. 9 nm. In the case of fs-LAL shown
in this work, the blocking temperature is further lowered to 52 K, which corresponds to the particle size
of ca. 4–5 nm for α-Fe2 O3 .

116
Nanomaterials 2018, 8, 631

3.3. Formation Mechanism


Considering the advantage of fs-LAL over ps-LAL, and ns-LAL enabling the synthesis of ultrasmall
magnetic particles with lower blocking temperatures, the formation mechanism of both Fe@C, Fe@FeOx
and ultrasmall α-Fe2 O3 particles is here proposed to show the uniqueness of the fs-LAL process. The large
size difference between ultrasmall α-Fe2 O3 clusters of several nm and large core-shell particles with sizes
ranging from tens of nm to 130 nm indicates that large particles do not form through particle growth
mechanism but form through ejection of large Fe particles during LAL [55]. The ultrasmall α-Fe2 O3
particles less than 10 nm (Figure 5g) should form due to phase/Coloumb explosion mechanism for fs-LAL.
In contrast, owing to the thermal ablation mechanism, the sizes of the majority of the small particles
inside the cavitation bubble are already 12 nm for ns-LAL [56]. The particle growth after bubble collapse
often leads to a further increase in the particle sizes. Therefore, despite a small amount of large Fe particle
ejection due to thermal effects during fs-LAL, the main “cold” process of fs-LAL is more efficient at
generating ultrasmall particles than both ps- and ns-LAL.
Due to the plasma-induced decomposition of acetone molecules and the dissociation of the dissolved
oxygen (Figure 9a) [57], O radicals are generated during fs-LAL of Fe in acetone, which may react with
the surrounding Fe atoms (generated from plasma-induced target material atomization) to form FeOx
clusters (Figure 9b). However, due to the existence of limited oxygen, the main products generated from
the plasma phase are pure Fe clusters. The sizes of Fe and FeOx clusters may increase slightly during
bubble expansion (Figure 9c) by coalescence. Inside the cavitation bubble, reductive gases [58] such as H2 ,
CO and CH4 reduce FeOx , which results in the formation of pure Fe particles (Figure 8d). After bubble
collapses (Figure 8e), the outer parts of large Fe particles are oxidized into FeOx shells containing a large
amount of α-Fe2 O3 domains (Figure 9g), while the ultrasmall Fe particles are oxidized into α-Fe2 O3
particles (Figure 9f,g). Due to the difference in local temperature and pressure as well as the variation in
oxygen abundance around Fe particles, Fe particles crystallize into α-Fe2 O3 or Fe3 O4 single crystals along
different crystallographic orientations. It is also possible that (1) reductive gases in the cavitation bubbles
inhibit complete oxidation of Fe into FeOx and their further polycrystallization; (2) during bubble collapses,
the shock waves [59] render small α-Fe2 O3 crystals with high kinetic energy to make them quickly eject
towards the already formed Fe@FeOx particles to be captured by FeOx shells as single domains.

Figure 9. (a–f) Schematic of formation mechanism for Fe@α-Fe2 O3 particles by fs-LAL of Fe in acetone.
Oxygen radicals that react with Fe atoms to from FeOx come from fs laser induced decomposition of
acetone and dissolved oxygen. Fe: grey color, C: black color, α-Fe2 O3 : red color, other FeOx phases: green
and blue color. Note that the shockwaves [59] generated during bubble collapse can push the already
formed α-Fe2 O3 ultrasmall particles towards Fe@FeOx particles to be captured by the FeOx shells.

117
Nanomaterials 2018, 8, 631

Extremely superfast cooling of the molten Fe droplets inhibits their crystallization so that a-Fe
rather than crystalline Fe particles form after LAL. The cooling rate required for a-Fe formation ranges
from 105 to 107 K/s [60], which can be easily obtained during the ultrafast quenching of LAL-generated
plasma (thousands of Kelvin quenching within a submicro-second interval [61]). During the ejection
of the molten Fe particles from the ablated target, they interact with acetone molecules, as a case of
the electric explosion of steel in carbon-rich liquids [62,63], resulting in the formation of carbon atoms
and other carbonaceous byproducts, which then precipitate on Fe particles to form C-shells. During their
precipitation, Fe particles with high surface activity act as catalysts to facilitate the formation of onion-like
carbon shells. The presence of carbons on the Fe particles inhibits particle growth and coalescence [4] and
prevent surface oxidation, leading to the formation of Fe@C core-shell particles (Figure 3f–g).
Under the impact of the high-temperature and high-pressure environment in the plasma phase of
fs-LAL, the atomization/ionization of carbon impurities in the iron substrates and the decomposition
of the acetone molecules [64,65] occur simultaneously (Figure 9a), which leads to the formation of free
C clusters. Besides pure carbon, polycyclic structures may be also generated through the interaction
between Fe particles and organic solvents [63], which may precipitate on the already particles to make
them evolve into networks. In consequence, the stability of the colloids decreases during storage and
results in the precipitation of particles (Figure 7).

4. Conclusions
This work has demonstrated the capability of synthesizing large Fe@α-Fe2 O3 and small α-Fe2 O3
particles which are practically completely split in the size histogram. Four to five nanometer α-Fe2 O3
particles exhibit a low blocking temperature of 52 K by fs-LAL in acetone, among the lowest ever achieved
by LAL. From superparamagnetic blocking, an effective magnetic anisotropy Keff = 2.7–5.4 × 105 J/m3
has been estimated which extends previous investigations convincingly towards smaller hematite
particle sizes. Surprisingly, most small α-Fe2 O3 particles were single-crystalline, and so the possibility
of synthesizing single-crystalline particles by LAL was demonstrated. Because of the dominant mass
of large Fe@α-Fe2 O3 and Fe@C particles (10–100 nm), all nanoblends show a soft magnetic behavior
with saturation magnetization (Ms ) and coercivities (Hc ) values of 72.5 emu/g and 160 Oe at 5 K
and 61.9 emu/g and 70 Oe at 300 K, respectively, which mainly originate from amorphous Fe core
particles. Previously, ZFC/FC curves were seldom investigated as compared to hysteresis curves for
LAL-generated magnetic particles. Here, it was shown that the blocking temperatures in the ZFC curves
can be used to estimate the sizes of small magnetic particles.

Author Contributions: D.Z., K.S. conceived and designed the experiments; W.C. performed the experiments;
D.Z. characterized the particles and analyzed the data; Y.O. and D.Z. did the magnetic measurement;
U.W. analyzed the magnetic properties; H.-P.L. and Y.I. measured the zeta potential of the colloid, D.Z. and
K.S. wrote the paper, all authors read and revised the manuscript.
Acknowledgments: The authors acknowledge Aiko Nakao from the Institute of Physical and Chemical Research
(RIKEN) for her help with XPS measurement and analysis. Also, we would also like to thank the Materials
Characterization Support Unit, RIKEN CEMS for providing access to the TEM microscopy, XRD, Raman and XPS
instruments and the support from Masaki Takeguchi from NIMS (National Institute for Materials Science) for
HRTEM characterization and analysis.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Zhang, D.; Gökce, B. Perspective of laser-prototyping nanoparticle-polymer composites. Appl. Surf. Sci. 2017,
392, 991–1003. [CrossRef]
2. Zhang, D.; Gökce, B.; Barcikowski, S. Laser synthesis and processing of colloids: Fundamentals and applications.
Chem. Rev. 2017, 117, 3990–4103. [CrossRef] [PubMed]
3. Zhang, D.; Liu, J.; Li, P.; Tian, Z.; Liang, C. Recent advances in surfactant-free, surface charged and defect-rich
catalysts developed by laser ablation and processing in liquids. ChemNanoMat 2017, 3, 512–533. [CrossRef]

118
Nanomaterials 2018, 8, 631

4. Zhang, D.; Liu, J.; Liang, C. Perspective on how laser-ablated particles grow in liquids. Sci. China Phys.
Mech. Astron. 2017, 60, 074201. [CrossRef]
5. Amendola, V.; Meneghetti, M. What controls the composition and the structure of nanomaterials generated
by laser ablation in liquid solution? Phys. Chem. Chem. Phys. 2013, 15, 3027–3046. [CrossRef] [PubMed]
6. Zhang, D.; Gökce, B.; Sommer, S.; Streubel, R.; Barcikowski, S. Debris-free rear-side picosecond laser ablation
of thin germanium wafers in water with ethanol. Appl. Surf. Sci. 2016, 367, 222–230. [CrossRef]
7. Letzel, A.; Gökce, B.; Wagener, P.; Ibrahimkutty, S.; Menzel, A.; Plech, A.; Barcikowski, S. Size quenching during
laser synthesis of colloids happens already in the vapor phase of the cavitation bubble. J. Phys. Chem. C 2017,
121, 5356–5365. [CrossRef]
8. Rehbock, C.; Merk, V.; Gamrad, L.; Streubel, R.; Barcikowski, S. Size control of laser-fabricated surfactant-free gold
nanoparticles with highly diluted electrolytes and their subsequent bioconjugation. Phys. Chem. Chem. Phys. 2013,
15, 3057–3067. [CrossRef] [PubMed]
9. Zhang, D.; Ma, Z.; Spasova, M.; Yelsukova, A.E.; Lu, S.; Farle, M.; Wiedwald, U.; Gökce, B. Formation mechanism
of laser-synthesized iron-manganese alloy nanoparticles, manganese oxide nanosheets and nanofibers. Part. Part.
Syst. Charact. 2017, 34, 1600225. [CrossRef]
10. Amendola, V.; Riello, P.; Meneghetti, M. Magnetic nanoparticles of iron carbide, iron oxide, iron@iron oxide, and
metal iron synthesized by laser ablation in organic solvents. J. Phys. Chem. C 2011, 115, 5140–5146. [CrossRef]
11. Lasemi, N.; Bomati Miguel, O.; Lahoz, R.; Lennikov, V.; Pacher, U.; Rentenberger, C.; Kautek, W. Laser-assisted
synthesis of colloidal fewx oy and fe/fex oy nanoparticles in water and ethanol. Chemphyschem 2018. [CrossRef]
[PubMed]
12. Vahabzadeh, E.; Torkamany, M.J. Iron oxide nanocrystals synthesis by laser ablation in water: Effect of
laser wavelength. J. Clust. Sci. 2014, 25, 959–968. [CrossRef]
13. Liu, P.; Cai, W.; Zeng, H. Fabrication and size-dependent optical properties of feo nanoparticles induced by
laser ablation in a liquid medium. J. Phys. Chem. C 2008, 112, 3261–3266. [CrossRef]
14. Amendola, V.; Riello, P.; Polizzi, S.; Fiameni, S.; Innocenti, C.; Sangregorio, C.; Meneghetti, M. Magnetic
iron oxide nanoparticles with tunable size and free surface obtained via a “green” approach based on laser
irradiation in water. J. Mater. Chem. 2011, 21, 18665–18673. [CrossRef]
15. Pandey, B.K.; Shahi, A.K.; Shah, J.; Kotnala, R.K.; Gopal, R. Optical and magnetic properties of fe2 o3 nanoparticles
synthesized by laser ablation/fragmentation technique in different liquid media. Appl. Surf. Sci. 2014, 289, 462–471.
[CrossRef]
16. Svetlichnyi, V.A.; Shabalina, A.V.; Lapin, I.N.; Goncharova, D.A.; Velikanov, D.A.; Sokolov, A.E.
Characterization and magnetic properties study for magnetite nanoparticles obtained by pulsed laser
ablation in water. Appl. Phys. A 2017, 123, 763. [CrossRef]
17. Ismail, R.A.; Sulaiman, G.M.; Abdulrahman, S.A.; Marzoog, T.R. Antibacterial activity of magnetic iron oxide
nanoparticles synthesized by laser ablation in liquid. Mater. Sci. Eng. C 2015, 53, 286–297. [CrossRef] [PubMed]
18. Kanitz, A.; Hoppius, J.S.; del Mar Sanz, M.; Maicas, M.; Ostendorf, A.; Gurevich, E.L. Synthesis of magnetic
nanoparticles by ultrashort pulsed laser ablation of iron in different liquids. Chemphyschem 2017, 18, 1155–1164.
[CrossRef] [PubMed]
19. Santillán, J.M.J.; Muñetón Arboleda, D.; Coral, D.F.; Fernández van Raap, M.B.; Muraca, D.; Schinca, D.C.;
Scaffardi, L.B. Optical and magnetic properties of fe nanoparticles fabricated by femtosecond laser ablation
in organic and inorganic solvents. Chemphyschem 2017, 18, 1192–1209.
20. Jongnam, P.; Eunwoong, L.; Nong-Moon, H.; Misun, K.; Chul, K.S.; Yosun, H.; Je-Geun, P.; Han-Jin, N.;
Jae-Young, K.; Jae-Hoon, P.; et al. One-nanometer-scale size-controlled synthesis of monodisperse magnetic
iron oxide nanoparticles. Angew. Chem. 2005, 117, 2932–2937.
21. Franzel, L.; Bertino, M.F.; Huba, Z.J.; Carpenter, E.E. Synthesis of magnetic nanoparticles by pulsed laser ablation.
Appl. Surf. Sci. 2012, 261, 332–336. [CrossRef]
22. Khan, U.; Adeela, N.; Irfan, M.; Ali, H.; Han, X.F. Temperature mediated morphological and magnetic phase
transitions of iron/iron oxide core/shell nanostructures. J. Alloys Compd. 2017, 696, 362–368. [CrossRef]
23. Pardoe, H.; Chua-anusorn, W.; St. Pierre, T.G.; Dobson, J. Structural and magnetic properties of nanoscale iron
oxide particles synthesized in the presence of dextran or polyvinyl alcohol. J. Magn. Magn. Mater. 2001, 225, 41–46.
[CrossRef]
24. Qiang, Y.; Antony, J.; Sharma, A.; Nutting, J.; Sikes, D.; Meyer, D. Iron/iron oxide core-shell nanoclusters for
biomedical applications. J. Nanopart. Res. 2006, 8, 489–496. [CrossRef]

119
Nanomaterials 2018, 8, 631

25. Sugioka, K.; Cheng, Y. Femtosecond laser three-dimensional micro- and nanofabrication. Appl. Phys. Rev. 2014,
1, 041303. [CrossRef]
26. Zhang, D.; Chen, F.; Fang, G.; Yang, Q.; Xie, D.; Qiao, G.; Li, W.; Si, J.; Hou, X. Wetting characteristics on
hierarchical structures patterned by a femtosecond laser. J. Micromech. Microeng. 2010, 20, 075029. [CrossRef]
27. Zhang, D.; Chen, F.; Yang, Q.; Si, J.; Hou, X. Mutual wetting transition between isotropic and anisotropic on
directional structures fabricated by femotosecond laser. Soft Matter 2011, 7, 8337–8342. [CrossRef]
28. Zhang, D.; Chen, F.; Yang, Q.; Yong, J.; Bian, H.; Ou, Y.; Si, J.; Meng, X.; Hou, X. A simple way to
achieve pattern-dependent tunable adhesion in superhydrophobic surfaces by a femtosecond laser. ACS Appl.
Mater. Interfaces 2012, 4, 4905–4912. [CrossRef] [PubMed]
29. Moussa, S.; Atkinson, G.; El-Shall, M.S. Laser-assisted synthesis of magnetic fe/fe2 o3 core: Carbon-shell
nanoparticles in organic solvents. J. Nanopart. Res. 2013, 15, 1470. [CrossRef]
30. Zufía-Rivas, J.; Morales, P.; Veintemillas-Verdaguer, S. Effect of the sodium polyacrylate on the magnetite
nanoparticles produced by green chemistry routes: Applicability in forward osmosis. Nanomaterials 2018, 8, 470.
31. Nguyen, V.; Gauthier, M.; Sandre, O. Templated synthesis of magnetic nanoparticles through the self-assembly of
polymers and surfactants. Nanomaterials 2014, 4, 628. [CrossRef] [PubMed]
32. Maneeratanasarn, P.; Khai, T.V.; Kim, S.Y.; Choi, B.G.; Shim, K.B. Synthesis of phase-controlled iron oxide
nanoparticles by pulsed laser ablation in different liquid media. Phys. Status Solidi A 2013, 210, 563–569. [CrossRef]
33. Sukhov, I.A.; Aleksandr, V.S.; Georgii, A.S.; Viau, G.; Garcia, C. Formation of nanoparticles during laser
ablation of an iron target in a liquid. Quantum Electron. 2012, 42, 453. [CrossRef]
34. De Bonis, A.; Lovaglio, T.; Galasso, A.; Santagata, A.; Teghil, R. Iron and iron oxide nanoparticles obtained
by ultra-short laser ablation in liquid. Appl. Surf. Sci. 2015, 353, 433–438. [CrossRef]
35. Marzun, G.; Bönnemann, H.; Lehmann, C.; Spliethoff, B.; Weidenthaler, C.; Barcikowski, S. Role of dissolved
and molecular oxygen on cu and ptcu alloy particle structure during laser ablation synthesis in liquids.
Chemphyschem 2017, 18, 1175–1184. [CrossRef] [PubMed]
36. Zhang, D.; Choi, W.; Jakobi, J.; Kalus, M.-R.; Barcikowksi, S.; Cho, S.-H.; Sugioka, K. Spontaneous shape
alteration and size separation of surfactant-free silver particles synthesized by laser ablation in acetone
during long-period storage. Nanomaterials 2018, 8, 529. [CrossRef] [PubMed]
37. Zhang, H.; Liang, C.; Liu, J.; Tian, Z.; Shao, G. The formation of onion-like carbon-encapsulated cobalt carbide
core/shell nanoparticles by the laser ablation of metallic cobalt in acetone. Carbon 2013, 55, 108–115. [CrossRef]
38. Robertson, J. Diamond-like amorphous carbon. Mater. Sci. Eng. 2002, 37, 129–281. [CrossRef]
39. Madrigal-Camacho, M.; Vilchis-Nestor, A.R.; Camacho-López, M.; Camacho-López, M.A. Synthesis of
moc@graphite nps by short and ultra-short pulses laser ablation in toluene under n2 atmosphere.
Diamond Relat. Mater. 2018, 82, 63–69. [CrossRef]
40. De Bonis, A.; Santagata, A.; Galasso, A.; Laurita, A.; Teghil, R. Formation of titanium carbide (tic) and tic@c
core-shell nanostructures by ultra-short laser ablation of titanium carbide and metallic titanium in liquid. J. Colloid
Interfaces Sci. 2017, 489, 76–84. [CrossRef] [PubMed]
41. Zhang, H.; Liu, J.; Tian, Z.; Ye, Y.; Cai, Y.; Liang, C.; Terabe, K. A general strategy toward transition
metal carbide/carbon core/shell nanospheres and their application for supercapacitor electrode. Carbon 2016,
100, 590–599. [CrossRef]
42. Giorgetti, E.; Muniz-Miranda, M.; Marsili, P.; Scarpellini, D.; Giammanco, F. Stable gold nanoparticles obtained in
pure acetone by laser ablation with different wavelengths. J. Nanopart. Res. 2012, 14, 648. [CrossRef]
43. Safronov, A.; Beketov, I.; Komogortsev, S.; Kurlyandskaya, G.; Medvedev, A.; Leiman, D.; Larrañaga, A.; Bhagat, S.
Spherical magnetic nanoparticles fabricated by laser target evaporation. AIP Adv. 2013, 3, 052135. [CrossRef]
44. Osipov, V.V.; Platonov, V.V.; Uimin, M.A.; Podkin, A.V. Laser synthesis of magnetic iron oxide nanopowders.
Tech. Phys. 2012, 57, 543–549. [CrossRef]
45. Grinstaff, M.W.; Salamon, M.B.; Suslick, K.S. Magnetic properties of amorphous iron. Phys. Rev. B 1993, 48, 269.
[CrossRef]
46. Goya, G.; Berquo, T.; Fonseca, F.; Morales, M. Static and dynamic magnetic properties of spherical
magnetite nanoparticles. J. Appl. Phys. 2003, 94, 3520–3528. [CrossRef]
47. Hadjipanayis, G.; Sellmyer, D.J.; Brandt, B. Rare-earth-rich metallic glasses. I. Magnetic hysteresis. Phys. Rev. B 1981,
23, 3349. [CrossRef]
48. Teja, A.S.; Koh, P.-Y. Synthesis, properties, and applications of magnetic iron oxide nanoparticles. Prog. Cryst.
Growth Charact. Mater. 2009, 55, 22–45. [CrossRef]

120
Nanomaterials 2018, 8, 631

49. Wiedwald, U.; Spasova, M.; Salabas, E.; Ulmeanu, M.; Farle, M.; Frait, Z.; Rodriguez, A.F.; Arvanitis, D.; Sobal, N.;
Hilgendorff, M. Ratio of orbital-to-spin magnetic moment in co core-shell nanoparticles. Phys. Rev. B 2003, 68, 064424.
[CrossRef]
50. Han, L.; Wiedwald, U.; Biskupek, J.; Fauth, K.; Kaiser, U.; Ziemann, P. Nanoscaled alloy formation from
self-assembled elemental co nanoparticles on top of pt films. Beilstein J. Nanotechnol. 2011, 2, 473. [CrossRef]
[PubMed]
51. Xu, Y.Y.; Rui, X.F.; Fu, Y.Y.; Zhang, H. Magnetic properties of α-fe2 o3 nanowires. Chem. Phys. Lett. 2005, 410, 36–38.
[CrossRef]
52. Wu, C.; Yin, P.; Zhu, X.; OuYang, C.; Xie, Y. Synthesis of hematite (α-fe2 o3 ) nanorods: Diameter-size and
shape effects on their applications in magnetism, lithium ion battery, and gas sensors. J. Phys. Chem. B 2006,
110, 17806–17812. [CrossRef] [PubMed]
53. Bødker, F.; Hansen, M.F.; Koch, C.B.; Lefmann, K.; Mørup, S. Magnetic properties of hematite nanoparticles.
Phys. Rev. B 2000, 61, 6826.
54. Bødker, F.; Mørup, S. Size dependence of the properties of hematite nanoparticles. Europhys. Lett. 2000, 52, 217.
55. Shih, C.-Y.; Streubel, R.; Heberle, J.; Letzel, A.; Shugaev, M.; Wu, C.; Schmidt, M.; Gokce, B.; Barcikowski, S.;
Zhigilei, L. Two mechanisms of nanoparticle generation in picosecond laser ablation in liquids: The origin of
the bimodal size distribution. Nanoscale 2018, 10, 6900–6910. [CrossRef] [PubMed]
56. Ibrahimkutty, S.; Wagener, P.; Rolo, T.d.S.; Karpov, D.; Menzel, A.; Baumbach, T.; Barcikowski, S.; Plech, A. A
hierarchical view on material formation during pulsed-laser synthesis of nanoparticles in liquid. Sci. Rep. 2015,
5, 16313. [CrossRef] [PubMed]
57. Semaltianos, N.G.; Friedt, J.-M.; Chassagnon, R.; Moutarlier, V.; Blondeau-Patissier, V.; Combe, G.;
Assoul, M.; Monteil, G. Oxide or carbide nanoparticles synthesized by laser ablation of a bulk hf target in
liquids and their structural, optical, and dielectric properties. J. Appl. Phys. 2016, 119, 204903. [CrossRef]
58. Kalus, M.-R.; Barsch, N.; Streubel, R.; Gokce, E.; Barcikowski, S.; Gokce, B. How persistent microbubbles
shield nanoparticle productivity in laser synthesis of colloids—Quantification of their volume, dwell
dynamics, and gas composition. Phys. Chem. Chem. Phys. 2017, 19, 7112–7123. [CrossRef] [PubMed]
59. Lauterborn, W.; Vogel, A. Shock wave emission by laser generated bubbles. In Bubble Dynamics and Shock Waves;
Springer: Berlin, Germany, 2013; pp. 67–103.
60. Suslick, K.S.; Choe, S.-B.; Cichowlas, A.A.; Grinstaff, M.W. Sonochemical synthesis of amorphous iron.
Nature 1991, 353, 414. [CrossRef]
61. Dell’Aglio, M.; Gaudiuso, R.; De Pascale, O.; De Giacomo, A. Mechanisms and processes of pulsed laser
ablation in liquids during nanoparticle production. Appl. Surf. Sci. 2015, 348, 4–9.
62. Lázár, K.; Varga, L.K.; Kovács Kis, V.; Fekete, T.; Klencsár, Z.; Stichleutner, S.; Szabó, L.; Harsányi, I. Electric
explosion of steel wires for production of nanoparticles: Reactions with the liquid media. J. Alloys Compd. 2018,
763, 759–770.
63. Beketov, I.V.; Safronov, A.P.; Bagazeev, A.V.; Larrañaga, A.; Kurlyandskaya, G.V.; Medvedev, A.I. In situ
modification of fe and ni magnetic nanopowders produced by the electrical explosion of wire. J. Alloys Compd. 2014,
586, S483–S488. [CrossRef]
64. Liu, P.; Cui, H.; Yang, G. Synthesis of body-centered cubic carbon nanocrystals. Cryst. Growth Des. 2008, 8, 581–586.
[CrossRef]
65. Seyedeh Zahra, M.; Parviz, P.; Ali, R.; Soghra, M.; Rasoul, S.-B. Generation of various carbon nanostructures in
water using ir/uv laser ablation. J. Phys. D Appl. Phys. 2013, 46, 165303.

© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

121
nanomaterials

Article
Local Melting of Gold Thin Films by Femtosecond
Laser-Interference Processing to Generate
Nanoparticles on a Source Target
Yoshiki Nakata 1, *, Keiichi Murakawa 1 , Noriaki Miyanaga 1 , Aiko Narazaki 2 , Tatsuya Shoji 3
and Yasuyuki Tsuboi 3
1 Institute of Laser Engineering, Osaka University, 2-6 Yamadaoka, Suita, Osaka 565-0871, Japan;
[email protected] (K.M.); [email protected] (N.M.)
2 National Institute of Advanced Industrial Science and Technology, Central 5, Higashi 1-1-1, Tsukuba,
Ibaraki 305-8565, Japan; [email protected]
3 Graduate School of Science, Osaka City University, 3-3-138 Sugimoto Sumiyoshi-ku, Osaka 558-8585, Japan;
[email protected] (T.S.); [email protected] (Y.T.)
* Correspondence: [email protected]; Tel.: +81-6-6879-8729

Received: 15 June 2018; Accepted: 23 June 2018; Published: 28 June 2018

Abstract: Shape- and size-controlled metallic nanoparticles are very important due to their wide
applicability. Such particles have been fabricated by chemosynthesis, chemical-vapor deposition,
and laser processing. Pulsed-laser deposition and laser-induced dot transfer use ejections of molten
layers and solid-liquid-solid processes to fabricate nanoparticles with a radius of some tens to
hundreds of nm. In these processes, the nanoparticles are collected on an acceptor substrate. In the
present experiment, we used laser-interference processing of gold thin films, which deposited
nanoparticles directly on the source thin film with a yield ratio. A typical nanoparticle had
roundness f r = 0.99 and circularity f circ = 0.869, and the radius was controllable between 69 and
188 nm. The smallest radius was 82 nm on average, and the smallest standard deviation was 3 nm.
The simplicity, high yield, and ideal features of the nanoparticles produced by this method will
broaden the range of applications of nanoparticles in fields such as plasmonics.

Keywords: femtosecond laser; interference; laser processing; melt; nanoparticle; gold; thin film

1. Introduction
Shape- and size-controlled metallic nanoparticles have a variety of applications in the fields of
plasmonics, catalysts, biology, etc. Such nanoparticles have been fabricated by chemosynthesis [1,2],
chemical-vapor deposition [3], electron-beam lithography [4], laser processing, etc. In laser processing,
pulsed-laser deposition (PLD) [5–7] and local melting of thin films have been employed, with the latter
technique using melting and re-solidification of the film [8–11]. For area processing by a Gaussian
beam, the particle sizes are dispersed due to the Rayleigh instability criterion [8]. On the other
hand, localization of the melt area by patterned illumination using a mask to avoid the effects of the
Rayleigh instability results in a relatively uniform size distribution [9–11]. In this case, the thin film is
ablated through the mask, and the rest of the source thin film melts, shrinks, and forms nanoparticles.
The laser-interference processing technique has also been applied to metallic thin films. Using this
technique, liquid structures such as nanodrops [12–15], nanobumps, and nanowhiskers [13–17] have
been fabricated for cases where the thermal-diffusion length is short when compared to the period of
the interference pattern. In these processes, motions of the liquid source metal are induced periodically
in space by the interference pattern, and nano-sized structures freeze simultaneously after photon
and emission of radiation. We call this mechanism the solid-liquid-solid (SLS) [16], in contrast to the

Nanomaterials 2018, 8, 477; doi:10.3390/nano8070477 122 www.mdpi.com/journal/nanomaterials


Nanomaterials 2018, 8, 477

vapor-liquid-solid (VLS) mechanism that is used to generate one-dimensional (1D) nanomaterials [18].
In nanowhisker generation, nanodrops detach from the nanowhiskers and remain on the source target,
but they blow off before morphological measurements can be carried out. In some other experiments,
a single-focus spot has been employed in the laser-induced dot transfer (LIDT) technique [19–22].
This is a variation of the laser-induced forward transfer (LIFT) technique [23–25] where a nanoparticle
from a source film is caught on an acceptor substrate.
In the present experiment, we collected gold nanoparticles directly on the source target without
requiring an acceptor substrate (Laser-Induced Dot caught On Source target; LIDOS). We measured
the morphologies of the nanoparticles using scanning electron microscopy (SEM) without any cleaning
of the target. We analyzed the dependence of the size distributions of the nanoparticles on parameters
such as the film thickness and fluence. The results that we obtained are supported by a model that
explains the formation mechanism.

2. Materials and Methods


Our experimental setup is shown in Figure 1. We used a femtosecond laser beam (IFRIT, Cyber
Laser, Tokyo, Japan) operating at 785 nm with a 240 fs (full-width at half-maximum, FWHM) pulse
width. The beam was split by a diffractive optical element, which was optimized to the wavelength
(the diffraction efficiency was approximately 60%) to generate four first-order diffracted beams.
The four beams were focused on the surface of the target via a demagnification system, for which
the focal lengths of the achromatic lenses were f 1 = 300 mm and f 2 = 40 mm. The original beam
diameter was 6.0 mm (FWHM at 1/e2 ), which was de-magnified to 0.8 mm. The zeroth-order beam
was dumped. The correlation angle was θ = 16.7◦ , and the period of the interference pattern was
Λ = 19.3 μm. For this experiment, we chose a gold thin-film target, which has surface-plasmon
resonance [26] and stability. We deposited the films on silica-glass substrates by magnetron sputtering
and was electrically insulated from the earth. All of the experiments were performed using a single
shot of laser irradiation at atmospheric conditions. We imaged the surface structures of the films using
SEM (JEOL, JSM-7400FS, Tokyo, Japan).

Figure 1. Experimental setup. The pulse width of the laser was 240 fs and the wavelength was 785 nm.
DOE (diffractive-optical element) split four 1st order diffracted beams. f 1 = 300 mm and f 2 = 40 mm.

3. Results and Discussion


Figure 2 shows a top view of the structures fabricated with different film thickness. The target-film
thickness, the average fluence over the interference pattern, the average radius and its standard
deviation (s.d.) are summarized in Table 1. The precise lattice structures with Λ = 19.3 μm period
were determined by the interference pattern. For the thinner film of 40 nm thickness, the nanoparticles
were found to lie on the source target at random locations, as shown in Figure 2a-1. The circular black
holes are the areas from which the particles were ejected and through which the substrate surface
appeared. Figure 2a-2 is a magnified view of Figure 2a-2, and demonstrates that particles with good
roundness could be fabricated. The average radius of the nanoparticles was 188 nm, and the standard
deviation was 7 nm. The yield ratio, which is the number of nanoparticles relative to the number of
holes, was 83% in the field of view in Figure 2a-1. Figure 3 shows a high-resolution SEM (HR-SEM)

123
Nanomaterials 2018, 8, 477

image of the same field with 100,000× magnification. The surface of the nanoparticle was smooth,
and some nanostructures smaller than 10 nm adhered to the surface; these may have been smaller
nanoparticles that had condensed from the gold vapor. The radius of this nanoparticle was 174.8 nm.
The circularity f circ and roundness f r of the nanoparticle were defined by the following equations:

f circ = 4πS/P2 (1)

f r = 4S/(2a)2 , (2)

where S is the surface area of the nanoparticle; P is the perimeter; and 2a is the length of the major axis,
assuming the shape to be an ellipsoid. For this nanoparticle, we obtained f circ = 0.869 and f r = 0.99,
so it was a fairly round sphere with a slightly rough surface. Note that f circ is affected by the resolution
of the SEM.

Figure 2. Scanning electron microscopy (SEM) images of the target surface with different film thickness.
(a-1) 40 nm; (b-1) 50 nm; (c-1) 60 nm; and (d-1) 100 nm, and their magnified images on right column:
(a-2) 40 nm; (b-2) 50 nm; (c-2) 60 nm; and (d-2) 100 nm. The experimental parameters and results are
summarized in Table 1. The white bars in the images on the left and right columns represent 5 μm and
500 nm, respectively.

Table 1. Parameters and results for the experiments shown in Figure 2, Figure 3, and Figure 6. In the
case of Figure 2c, nanoparticles should adhere to the edge of holes of the film and not be separated.
In the case of Figure 2d, the nano-structure was nano-projecting, not nanoparticle.

Figure 2(a-1,a-2),
Parameter Figure 2(b-1,b-2) Figure 2(c-1,c-2) Figure 2(d-1,d-2) Figure 6(a-1,a-2) Figure 6(b-1,b-2) Figure 6(c-1,c-2)
Figure 3
film thickness (nm) 40 50 60 100 30 40 40
fluence (mJ/cm2 ) 73.2 73.2 73.2 73.2 58.6 58.6 169.9
averaged radius (nm) 188 170 82 69 n.d. 137 138
standard d. (nm) 7 3 5 9 n.d. 10 27

124
Nanomaterials 2018, 8, 477

Figure 3. Observation with high-resolution scanning electron microscopy (HR-SEM) of a single


nanoparticle shown in Figure 2a-1. The film thickness was 40 nm, and the fluence was 73.2 mJ/cm2 .
The radius was 175 nm. The white bar in the image represent 100 nm.

When the film thickness was 50 nm, smaller nanoparticles were fabricated, as shown in
Figure 2(b-1,b-2). Most remained on the lower left sides of the holes, but the reason for this is
not known. The yield ratio was 93%. The standard deviation of the radius was 3 nm, which was the
smallest dispersion in this experiment. When the film thickness was 60 nm, smaller nanoparticles
always remained on the lower right edges of the holes, as shown in Figure 2(c-1,c-2). The yield ratio
of the field of view was 100%. When the thickest film of 100 nm was used, only the textured surface
of the gold film was seen, as shown in Figure 2(d-1,d-2). White spots, which show the existence of
projecting structures, could be seen at the centers of the depressions in the processed area. Owing to
the thickness of this film, none of these depressions punched through to expose the substrate.
The radius as a function of film thickness is summarized in Figure 4. The smallest film thickness
was 40 nm. The radius of the nanoparticles was smaller than 200 nm in all cases. It is apparent that the
radius of the nanoparticles decreased as the film thickness increased from 40 to 60 nm. The results
that we obtained are supported by a model that explains the formation mechanism, as shown in
Figure 5 [16,27,28]. For every thickness, the films were almost opaque at 785 nm. As the thermal
conductivity of gold at room temperature is far higher than that of silica glass, the temperature
distribution along the substrate is normal inside the gold film levels immediately. As a result, the rise
in temperature is inversely proportional to the mass of the film, as illustrated in the top curves in the
figure. For the thinner film shown in Figure 5a, the region at temperatures above the melting point
(Tm.p. = 1064 ◦ C) was wider than that for the thicker films, as illustrated in Figure 5b,c. The molten layer
was launched from the substrate by the reaction to volumetric expansion during the solid-to-liquid
phase transition, as shown in Figure 5a-i. It was then squeezed and formed a nanoparticle due to
surface tension, as in Figure 5a-ii. It solidified because of cooling by photon emission, and could then
be pulled by electrostatic forces and adhered to the target surface, as shown in Figure 5a-iii. Most of
the nanoparticles adhered to the metal surface because of the metallic bond. However, when the film
was too thin, as shown in Figure 6(a-1,a-2), the region simply boiled away.

125
Nanomaterials 2018, 8, 477

Figure 4. The radii of the nanoparticles as a function of film thickness. The fluence was 73.2 mJ/cm2 .
The error bars reflect the s.d.

For the film thickness of 50 nm, the temperature was lower, which resulted in a less-effective
launch, as shown in Figure 5b-i. A nanoparticle is formed at the top center of the molten layer because
of surface tension, as illustrated in Figure 5b-ii. If it freezes at this time, it can form a nanodrop on a
hollow bump [12], or a nanowhisker [16,17], which is left after the nanoparticle detaches. The resulting
nanoparticles are left on the processed area or squeezed to the edges of the holes by surface tension,
as illustrated in Figure 5b-iii and shown in Figure 2b,c. For a film thickness of 100 nm, a hole could
not be punched through the gold film at the given fluence, as shown in Figure 2(d-1,d-2). In this case,
the motion of the molten layer resulted in a textured film surface, as shown in Figure 5c-i–c-iii.

Figure 5. Schematic illustration of the solid-liquid-solid (SLS) process. Each panel illustrates the
temporal change of a spot in an interference pattern. The correspondence with the results shown in
Figure 2 is noted on the bottom. From (a–c), the thickness of the film is listed as following: (a), 40 nm;
(b), 50 and 60 nm; (c), 100 nm.

126
Nanomaterials 2018, 8, 477

Figure 6. SEM images of the target surface. (a-1) film thickness was 30 nm, and fluence was 58.6 mJ/cm2 ;
(b-1) film thickness was 40 nm, and fluence was 58.6 mJ/cm2 ; (c-1) film thickness was 40 nm, and fluence
was 169.9 mJ/cm2 . Pictures on the right column are the corresponding magnified images: (a-2) film
thickness was 30 nm, and fluence was 58.6 mJ/cm2 ; (b-2) film thickness was 40 nm, and fluence
was 58.6 mJ/cm2 ; (c-2) film thickness was 40 nm, and fluence was 169.9 mJ/cm2 . The experimental
parameters and results are summarized in Table 1. The white bars in the images on the left and right
columns represent 5 μm and 500 nm, respectively.

The radius of the nanoparticles as a function of the fluence are summarized in Figure 7 for a
film thickness of 40 nm. The radius was smaller than 200 nm in all cases. It is interesting that mixed
nanoparticles such as singles and twins were produced, as shown in Figure 6(b-1,b-2). In Figure 7,
data for single particles are plotted. It is interesting that the formation of twins or multiple droplets
in a string is seen also in the behavior of water solutions [29]. The process of drop detachment was
controlled in those experiments by dissolving polymeric molecules, thus changing the viscoelasticity of
the solution. In our case, it may be possible to control multiple-nanoparticle formation by changing the
material, e.g., using silver, chromium and alloys to obtain different values of viscoelasticity. When the
fluence was at the highest value that we employed, 169.9 mJ/cm2 , the process caused more spattering.
As can be seen in Figure 6(c-1,c-2), the standard deviation of the radius was highest, 27 nm.

Figure 7. Radius of nanoparticles as a function of fluence. The film thickness was 40 nm. The error
bars reflect the s.d.

127
Nanomaterials 2018, 8, 477

We now compare this method with others. The roundness of the nanoparticles produced in this
experiment was excellent, and the radius controllable between 69 and 188 nm. In PLD, the radii of the
nanoparticles that condensed from the gas phase ranged from a few nm to some tens of nm [5,6]. On the
other hand, the droplets fabricated by PLD were some hundreds of nm in radius [7]. Using LIDT,
nanoparticles with radii smaller than 250 nm can be fabricated [19,30]. Electron-beam lithography is
utilized for fabricating aligned nanoparticles a few nm in diameter [4]. Chemosynthesis can be used
to fabricate nanoparticles and nanorods with radii of a few nm in both structures [1]. In summary,
this technique is a good alternative to PLD or LIDT, and can generate nanoparticles with radii of tens
or hundreds of nm via the SLS process.

4. Conclusions
Using the SLS process, we successfully fabricated gold nanoparticles with radii of tens to
hundreds of nm and good roundness-which were subsequently deposited on the source substrate-using
irradiation through the interference pattern of a fs laser. The film thickness and fluence were the
key parameters for controlling the sizes of the nanoparticles, with a thinner film resulting in a larger
nanoparticle radius. The smallest radius was 82 nm on average, and the smallest standard deviation
was 3 nm. A typical nanoparticle was found to have roundness f r = 0.99.
Compared with the methods such as chemosynthesis, VLS and PLD, SLS is useful for fabricating
pure and uniform nanomaterials. No catalyst or chemosynthetic solution is required, and a more
uniform size distribution will result in better plasmonic resonance properties. We anticipate that
different source materials such as metals, alloys, and non-metals with plasticity can produce
nanoparticles using the SLS mechanism. The process is very simple and does not require cleaning,
temperature control, evacuation, purification, etc. These advantages will broaden the range of
applications for such nanoparticles.

Author Contributions: Conceptualization, Y.N.; Methodology, Y.N.; Validation, Y.N.; Formal Analysis, Y.N.
Investigation, K.M.; Resources, N.M.; Data Curation, K.M. and Y.N.; Writing-Original Draft Preparation, Y.N.;
Writing-Review & Editing, Y.N.; Visualization, Y.N.; Supervision, Y.N.; Project Administration, Y.N.; Funding
Acquisition, Y.N., A.N., T.S., Y.T.

Funding: This research was funded by the Japan Society for the Promotion of Science (JSPS) (23360035,
A16H038850).

Conflicts of Interest: The authors declare no conflict of interest.

References
1. Jana, N.R.; Gearheart, L.; Murphy, C.J. Wet Chemical Synthesis of High Aspect Ratio Cylindrical Gold
Nanorods. J. Phys. Chem. B 2001, 105, 4065–4067. [CrossRef]
2. Silver, P.; Pietrobon, B.; Mceachran, M.; Kitaev, V. Synthesis of Size-Controlled Faceted Tunable Plasmonic
Properties and Self- Assembly of These Nanorods. ACS Nano 2009, 3, 21–26. [CrossRef]
3. Okumura, M.; Tsubota, S.; Iwamoto, M.; Haruta, M. Chemical Vapor Deposition of Gold Nanoparticles on
MCM-41 and Their Catalytic Activities for the Low-temperature Oxidation of CO and of H2 . Chem. Lett.
1998, 27, 315–316. [CrossRef]
4. Vieu, C.; Carcenac, F.; Pepin, A.; Chen, Y.; Mejias, M.; Lebib, A.; Manin-Ferlazzo, L.; Couraud, L.; Launois, H.
Electron beam lithography: Resolution limits and applications. Appl. Surf. Sci. 2000, 164, 111–117. [CrossRef]
5. Guczi, L.; Horváth, D.; Pászti, Z.; Tóth, L.; Horváth, Z.E.; Karacs, A.; Petõ, G. Modeling Gold Nanoparticles:
Morphology, Electron Structure, and Catalytic Activity in CO Oxidation. J. Phys. Chem. B 2000, 104, 3183–3193.
[CrossRef]
6. Muramoto, J.; Sakamoto, I.; Nakata, Y.; Okada, T.; Maeda, M. Influence of electric field on the behavior of Si
nanoparticles generated by laser ablation. Appl. Phys. Lett. 1999, 75, 751–753. [CrossRef]
7. Uetsuhara, H.; Goto, S.; Nakata, Y.; Vasa, N.; Okada, T.; Maeda, M. Fabrication of a Ti:sapphire planar
waveguide by Pulsed Laser Deposition. Appl. Phys. A Mater. Sci. Process. 1999, 69, S719–S722. [CrossRef]

128
Nanomaterials 2018, 8, 477

8. Henley, S.J.; Carey, J.D.; Silva, S.R.P. Pulsed-laser-induced nanoscale island formation in thin metal-on-oxide
films. Phys. Rev. B Condens. Matter Mater. Phys. 2005, 72, 195408. [CrossRef]
9. Mäder, M.; Höche, T.; Gerlach, J.W.; Böhme, R.; Zimmer, K.; Rauschenbach, B. Large area metal dot matrices
made by diffraction mask projection laser ablation. Phys. Status Solidi Rapid Res. Lett. 2008, 2, 34–36.
[CrossRef]
10. Höche, T.; Böhme, R.; Gerlach, J.W.; Rauschenbach, B.; Syrowatka, F. Nanoscale laser patterning of thin gold
films. Philos. Mag. Lett. 2006, 86, 661–667. [CrossRef]
11. Nakata, Y.; Okada, T.; Maeda, M. Holographic fabrication of micron structures using interfered femtosecond
laser beams split by diffractive optics. Proc. SPIE Int. Soc. Opt. Eng. 2003, 4977. [CrossRef]
12. Nakata, Y.; Okada, T.; Maeda, M. Nano-sized hollow bump array generated by single femtosecond laser
pulse. Jpn. J. Appl. Phys. 2003, 42, L1452–L1454. [CrossRef]
13. Nakata, Y.; Momoo, K.; Hiromoto, T.; Miyanaga, N. Generation of superfine structure smaller than 10 nm by
interfering femtosecond laser processing. In Proceedings of the SPIE the International Society for Optical
Engineering, San Francisco, CA, USA, 22–27 January 2011; Volume 7920.
14. Nakata, Y.; Tsuchida, K.; Miyanaga, N.; Furusho, H. Liquidly process in femtosecond laser processing.
Appl. Surf. Sci. 2009, 255, 9761–9763. [CrossRef]
15. Nakata, Y. Frozen water drops in the nanoworld. SPIE Newsroom 2009, 2, 1–2. [CrossRef]
16. Nakata, Y.; Miyanaga, N.; Momoo, K.; Hiromoto, T. Solid-liquid-solid process for forming free-standing
gold nanowhisker superlattice by interfering femtosecond laser irradiation. Appl. Surf. Sci. 2013, 274, 27–32.
[CrossRef]
17. Nakata, Y.; Miyanaga, N.; Momoo, K.; Hiromoto, T. Template free synthesis of free-standing silver
nanowhisker and nanocrown superlattice by interfering femtosecond laser irradiation Template free synthesis
of free-standing silver nanowhisker and nanocrown superlattice by interfering femtosecond laser irradiation.
Jpn. J. Appl. Phys. 2014, 53, 096701.
18. Wagner, R.S.; Ellis, W.C. Vapor-liquid-solid mechanism of single crystal growth. Appl. Phys. Lett. 1964,
4, 89–90. [CrossRef]
19. Narazaki, A.; Sato, T.; Kurosaki, R.; Kawaguchi, Y.; Niino, H. Nano- and microdot array formation of FeSi2 by
nanosecond excimer laser-induced forward transfer. Appl. Phys. Express 2008, 1, 0570011–0570013. [CrossRef]
20. Narazaki, A.; Sato, T.; Kurosaki, R.; Kawaguchi, Y.; Niino, H. Nano- and microdot array formation by
laser-induced dot transfer. Appl. Surf. Sci. 2009, 255, 9703–9706. [CrossRef]
21. Kuznetsov, A.I.; Evlyukhin, A.B.; Reinhardt, C.; Seidel, A.; Kiyan, R.; Cheng, W.; Ovsianikov, A.;
Chichkov, B.N. Laser-induced transfer of metallic nanodroplets for plasmonics and metamaterial applications.
J. Opt. Soc. Am. B 2009, 26, B130. [CrossRef]
22. Willis, D.A.; Grosu, V. Microdroplet deposition by laser-induced forward transfer. Appl. Phys. Lett. 2005,
86, 1–3. [CrossRef]
23. Levene, M.L.; Scott, R.D.; Siryj, B.W. Material Transfer Recording. Appl. Opt. 1970, 9, 2260. [CrossRef]
[PubMed]
24. Bohandy, J. Metal deposition from a supported metal film using an excimer laser. J. Appl. Phys. 1986,
60, 10–12. [CrossRef]
25. Nakata, Y.; Okada, T. Time-resolved microscopic imaging of the laser-induced forward transfer process.
Appl. Phys. A Mater. Sci. Process. 1999, 69, S275–S278. [CrossRef]
26. Jain, P.; Lee, K.; El-Sayed, I.; El-Sayed, M. Calculated absorption and scattering properties of gold
nanoparticles of different size, shape, and composition: Applications in biological imaging and biomedicine.
J. Phys. Chem. B 2006, 110, 7238–7248. [CrossRef] [PubMed]
27. Wu, C.; Zhigilei, L.V. Nanocrystalline and Polyicosahedral Structure of a Nanospike Generated on Metal
Surface Irradiated by a Single Femtosecond Laser Pulse. J. Phys. Chem. C 2016, 120, 4438–4447. [CrossRef]
28. Ivanov, D.S.; Lin, Z.; Rethfeld, B.; O’Connor, G.M.; Glynn, T.J.; Zhigilei, L.V. Nanocrystalline structure of
nanobump generated by localized photoexcitation of metal film. J. Appl. Phys. 2010, 107. [CrossRef]

129
Nanomaterials 2018, 8, 477

29. Clasen, C.; Bico, J.; Entov, V.M.; McKinley, G.H. “Gobbling drops”: The jettingdripping transition in flows of
polymer solutions. J. Fluid Mech. 2009, 636, 5–40. [CrossRef]
30. Zhigunov, D.M.; Evlyukhin, A.B.; Shalin, A.S.; Zywietz, U.; Chichkov, B.N. Femtosecond Laser Printing
of Single Ge and SiGe Nanoparticles with Electric and Magnetic Optical Resonances. ACS Photonics 2018,
5, 977–983. [CrossRef]

© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

130
nanomaterials

Article
Dual THz Wave and X-ray Generation from a Water
Film under Femtosecond Laser Excitation
Hsin-hui Huang 1 , Takeshi Nagashima 2, *, Wei-hung Hsu 1 , Saulius Juodkazis 3,4, * and
Koji Hatanaka 1,5,6, *
1 Research Center for Applied Sciences, Academia Sinica, Taipei 115, Taiwan;
[email protected] (H.-h.H.); [email protected] (W.-h.-H.)
2 Faculty of Science and Engineering, Setsunan University, 17-8 Ikeda-Nakamachi, Neyagawa,
Osaka 572-8508, Japan
3 Nanotechnology Facility, Center for Micro-Photonics, Swinburne University of Technology, Hawthorn,
VIC 3122, Australia
4 Melbourne Centre for Nanofabrication, the Victorian Node of the Australian National Fabrication Facility,
Clayton, VIC 3168, Australia
5 College of Engineering, Chang Gung University, Taoyuan 33302, Taiwan
6 Department of Materials Science and Engineering, National Dong-Hwa University, Hualien 97401, Taiwan
* Correspondence: [email protected] (T.N.); [email protected] (S.J.);
[email protected] (K.H.); Tel.: +81-72-813-1167 (T.N.); +61-3-9214-8718 (S.J.);
+886-2-2787-3132 (K.H.)

Received: 25 June 2018; Accepted: 11 July 2018; Published: 13 July 2018

Abstract: Simultaneous emission of the THz wave and hard X-ray from thin water free-flow was
induced by the irradiation of tightly-focused femtosecond laser pulses (35 fs, 800 nm, 500 Hz) in
air. Intensity measurements of the THz wave and X-ray were carried out at the same time with
time-domain spectroscopy (TDS) based on electro-optic sampling with a ZnTe(110) crystal and a
Geiger counter, respectively. Intensity profiles of the THz wave and X-ray emission as a function of
the solution flow position along the incident laser axis at the laser focus show that the profile width of
the THz wave is broader than that of the X-ray. Furthermore, the profiles of the THz wave measured
in reflection and transmission directions show different features and indicate that THz wave emission
is, under single-pulse excitation, induced mainly in laser-induced plasma on the water flow surface.
Under double-pulse excitation with a time separation of 4.6 ns, 5–10 times enhancements of THz
wave emission were observed. Such dual light sources can be used to characterise materials, as well
as to reveal the sequence of material modifications under intense laser pulses.

Keywords: femtosecond laser; intense laser; water; THz wave; time-domain spectroscopy; X-ray;
ablation; double-pulse excitation; plasma; z-scan; intensity enhancement

1. Introduction
An intense (>1013 W/cm2 ) femtosecond laser at near-IR (photon energy ∼1 eV) and matter
interaction [1–3] induce highly-nonlinear optical processes and result in photon conversion to an X-ray
of several-keV [4], as well as to a THz wave at meV [5] in association with the white light continuum
in the visible spectral range [6]. Strong laser ablation of solid and solution targets also occurs, which
is usually associated with the production of laser plasma [7]. Studies of such photon conversion
mechanisms for different applications [8–10] have been carried out separately for widely different
wavelengths. Experimental techniques based on the X-ray or THz wave have been contributing
immensely to basic physics and chemistry/material science [10–13], have expanded the understanding
of the basic mechanisms of intense laser-matter interaction and have contributed to laser-material
processing/printing [14].

Nanomaterials 2018, 8, 523; doi:10.3390/nano8070523 131 www.mdpi.com/journal/nanomaterials


Nanomaterials 2018, 8, 523

Gases [15–18], atomic clusters [19–21] and solids [22,23] have been used as targets for THz wave
emission from the laser-induced plasmas. Fast electron motions accelerated by laser ponderomotive
forces [15,18,19] and the four-mixing process induced by two-colour excitation [17] have been proposed
as THz emission mechanisms from laser-induced plasmas. Although such laser-induced plasmas of the
solid targets generate a powerful THz wave because of the high plasma densities, there are problems
with the instability of the intensity and a limitation in the successive delivery of laser irradiation
to the target. The use of liquids or solutions as targets is promising since solutions have moderate
atomic density (∼1022 cm−3 ) and ease of the delivery, but there have been no reports with liquid
targets (water) until recently [5,24]. As for intense laser-induced X-ray emission, it has been widely
studied with solids [15,25–28] and solutions [29–38]. One advantage of using solutions, e.g., water, as
a target sample is that a fresh and smooth surface can be easily prepared for each laser irradiation even
under intense laser irradiation at high repetition rates. Furthermore, the addition of solute chemical
agents such as electrolyte, CsCl for instance [39], with different concentrations or nano-particles, gold
for instance [40], with different shapes and sizes into water makes it possible to explore different
excitation/absorption mechanisms and to control the spectral characteristics of X-ray sources by
utilizing solute-dependent monochromatic characteristic X-ray lines and broad components mainly
based on bremsstrahlung.
Simultaneous emission of the THz wave and X-ray and combined usages of THz wave and
X-ray pulses have recently been introduced as readily-available table-top realisations [41]. To date,
experiments on such simultaneous emission with a Al-coated glass substrate [15] and gas targets,
He [15] or Ar [20], in vacuum chambers were reported. This type of work, though the numbers of such
reports are still quite limited, will be the first step for the discussion of the conversion of near-IR (eV)
laser photon energy into photons at the two ends of the spectrum, several-keV (X-ray) and meV (THz
wave), and will contribute well to other fields like material sciences. One important parameter for the
synchronized emission of the X-ray and THz wave for practical applications is their intensities. It has
been reported that chirped femtosecond laser pulses enhance THz wave emission from water [24] and
the X-ray from aqueous solutions [42]. In cases of X-ray emission from aqueous solutions, double-pulse
excitations [37,43] and the addition of solutes such as electrolytes [37] or gold nano-particles [40] to
water are also reported to be effective for the X-ray intensity enhancements. For THz wave emission,
one recent experimental study with double-pulse excitation to gas-clusters in a vacuum chamber [21]
was reported, but the inter-pulse delay time was limited to 0.5 ns. One advantage of the solution targets,
in addition to that described above, is the possibility of a transient solution surface roughening from
its original nano-smooth surface by a pre-pulse irradiation. Dynamic changes induced on the liquid
surface by plasma formation, capillary transient surface roughening instabilities and mist/droplet
formation associated with shock-wave expansion [44] contribute favourably to increased interaction
volume and augmented X-ray intensity up to more than an order of magnitude [37,43].
In this study, a simultaneous emission of the THz wave and hard X-ray in air using distilled water
as a target irradiated by tightly-focused near-IR femtosecond laser pulses and THz wave emission
enhancements under double-pulse excitation are presented.

2. Experimental Section
The experimental setup is shown in Figure 1a. A flat solution flow of distilled water with a
thickness smaller than 20 μm was prepared using a metal nozzle (Flatjet Nozzle LARGE, Metaheuristic,
Okayama, Japan), and the flow rate was regulated by a circulation pump (PMD-211, SANSO, Hyogo,
Japan) controlled by a conventional voltage regulator. The flow rate was set at <70 mL/min. The
nozzle was mounted to a rotational and 3D-automatic stages (KS701-20LMS, Suruga Seiki, Shizuoka,
Japan), and its position was finely set by a home-made LabView code, as reported previously [43].
Transform-limited femtosecond laser pulses (λ = 800 nm, >35 fs, 1 kHz, linearly-polarized, Mantis,
Legend Elite HE USP, Coherent, Inc., Santa Clara, CA, USA) were separated as two beams with different
polarizations by a half-wave plate (65-906, Edmund Optics, Barrington, NJ, USA) and a polarization

132
Nanomaterials 2018, 8, 523

beam splitter (47-048, Edmund Optics, Barrington, NJ, USA). Horizontally- and vertically-polarized
pulses were defined as the excitation pulse for X-ray/THz wave generation and the probe pulse for
the time-domain spectroscopy (TDS) for THz wave measurements [45,46], respectively. The repetition
rate of the excitation pulses was modulated by a wheel chopper (500 Hz, 3502 Optical Chopper,
New Focus, CA, USA) for TDS measurements, and the pulses were tightly-focused in air onto the
solution flow surface by using an off-axis parabolic mirror (OAPM, 1-inch diameter, the effective
focus length f = 50.8 mm, the reflection angle of 90 degrees and numerical aperture N A = 0.25,
47-097, Edmund Optics, Santa Clara, CA, USA). The incident angle of the excitation pulses along the
Z -axis to the solution normal was fixed at 60 degrees for the highest X-ray emission, as reported
previously [42]. Under these experimental conditions, each excitation pulse at a 500-Hz repetition rate
irradiates the fresh and flat solution flow. Experiments on double-pulse excitations with a pre-pulse
(vertically-polarized, 0.1 mJ/pulse, 4.6 ns in advance of the main excitation pulse) were also carried
out using an optical delay line (SGSP46-800, Sigma Koki, Tokyo, Japan).

Figure 1. (a) The experimental setup for the simultaneous measurements of THz wave (time-domain
spectroscopy (TDS)) and X-ray (Geiger counter). The femtosecond laser pulses (<35 fs, 800 nm, 500 Hz,
horizontally-polarised to the solution flow surface, i.e., p-pol.) were focused onto the solution flow;
ODL, optical delay line for TDS, L-plano-convex lens ( f = 50 cm). The thickness of ZnTe(110) crystal
for TDS was 1 mm; C, optical chopper (500 Hz) for TDS measurements. The parent focal lengths for
off-axis parabolic mirrors (OAPMs) (off-axis parabolic mirrors) were f = 50.8 mm (OAPM1, 1-inch
diameter), 101.6 mm (OAPM2, 2-inch), 152.4 mm (OAPM3, 2-inch) and 101.6 mm (OAPM4, 2-inch
with a hole in its centre for the probe), respectively. The distance between the laser focus and Geiger
counter was 12 cm. FM, flip-folding mirror for TDS measurements in the transmission direction. The
inset shows the close-up of the solution surface, the laser-water interaction region; (b) Representative
TDS signals from water flow when the laser intensity is 0.4 mJ/pulse in reflection and transmission
directions at the Z-position for the highest X-ray intensity and (c) their normalized spectra.

X-ray intensity was measured by a Geiger counter (SS315, Radhound, southern scientific, West
Sussex, UK). All the measurements were carried out in air under atmospheric pressure (1 atm) at
room temperature (RT = 296 K). Its observation angle was 15 degrees to the solution normal towards
the excitation side, and its distance from the laser focus was 12 cm. Therefore, it is certain that the
Geiger counter detects only X-ray, neither α- nor β-ray. THz wave signals were collected in reflection
(30 degrees to the solution normal, 90 degrees to the laser incident direction) and transmission
(along the excitation Z-axis) directions with two independent OAPMs (the reflected focal length

133
Nanomaterials 2018, 8, 523

f = 101.6 mm and 152.4 mm, the off-axis angle of 90 degrees, MPD249-M01, MPD269-M01, ThorLabs,
Newton, NJ, USA). As conventional TDS measurements, THz wave and the probe pulses after n
variable optical delay (TSDM60-20, OptoSigma, Tokyo, Japan) were focused to a 1 mm-thick ZnTe(110)
crystal (Nippon mining & metals Co., Ltd., Tokyo, Japan) by an OAPM (the reflected focal length
f = 101.6 mm, the off-axis angle of 90 degrees, MPD249H-M01, ThorLabs, Newton, NJ, USA) and
a plano-convex lens ( f = 50 cm), respectively. Lock-in measurements were carried out with a
balanced photo-diode (Model2307, New Focus, CA, USA) and a lock-in amplifier (SR830, Stanford
Research System, Sunnyvale, CA, USA). For measurements in the transmission direction, an additional
flip-folding mirror (FM) was set for the THz wave path to be bent; therefore, the total number of metal
mirrors for the transmission measurements was larger by one compared to that for the reflection.
One representative THz wave signal of water in reflection/transmission directions by the EO
sampling in TDS measurements is shown in Figure 1b. Detection efficiencies for the reflection and
the transmission directions, though their optical paths were shared partly, have not been calibrated;
therefore, the THz wave intensities and converted FFT spectra are not comparable quantitatively
between the signals in the two observation directions. A single cycle of the electric field oscillation was
clearly observed, and the vibration structure afterwards was also very distinct. This is reflected in their
Fourier-transformed emission spectra shown in Figure 1c as absorption bands at 1.1 and 1.4 THz due
to water vapour, as reported elsewhere [47]. Note that the spectra shown in Figure 1c are normalized.
This absorption is considered to be due to the water vapour in the atmosphere of the laboratory and
long-living micro-droplets (mist) formation in the vicinity of the water film induced by the laser
irradiation every 2 ms. The central frequencies of the observed THz wave were around 0.9–1 THz
for the reflection and the transmission. As discussed in detail below, the THz wave was emitted
from the area in the vicinity of the upstream-side of the air/water interface. In the transmission, the
THz wave transmitted through the water film. Water absorption in the THz wave region was well
studied recently, and a transmission experiment of the THz wave through 0.5 mm-thick water films
was recently demonstrated using TDS for precise measurements of temperature [48], the discussion of
which can be used for characterization of light-water interaction at high intensities. Under conventional
transmission conditions thorough 20 μm-thick water toward the transmission direction, but with the
incident angle at 60 degrees, the transmittance of the THz wave electric field intensity at 1 THz can
be estimated to be 0.64. Since water shows higher absorption at higher frequencies [48,49], the high
frequency components in the transmission are expected to be reduced compared with those in the
reflection. However, the observed spectra in the transmission showed a slight blue shift compared
to that in the reflection. It should be pointed out that the small shift can be due to an extrinsic effect
since the spectra obtained in the transmission and the reflection directions have not been calibrated in
this study. In the following, the relative intensities of the THz wave obtained in the identical direction
are discussed.
Under these conditions, X-ray and THz wave emission was induced at the same time and
measured simultaneously. Experiments were performed with different excitation laser intensities,
different solution positions along the Z-axis and under the double-pulse excitation. All the experiments
were carried out at atmospheric pressure at RT conditions.

3. Results and Discussion

3.1. Laser Intensity Dependences


Figure 2 shows the intensities of the X-ray and THz wave in the reflection and the transmission as
a function of the excitation laser intensity. The solution flow position along the Z-axis was optimized
for the highest X-ray intensity for each laser intensity. The highest intensity of X-ray was obtained when
the highest electron temperature, Te , was reached, hence for the strongest absorption of femtosecond
laser pulses. The X-ray emission in Figure 2a shows the slope γ = 2 scaling for this maximized
absorbed intensity. The absorbed energy density, Wabs , (per volume) is the relevant quantity that

134
Nanomaterials 2018, 8, 523

should be considered, since the free electron density, ne , is approaching the critical density, ncr , (plasma
reflection range) during the light absorption and Wabs ∝ nncre Fp , where Fp is the pulse fluence [43].
The absorbed energy density was defined by the mechanism of electron generation ne ∝ I pm ∝ Fpm
and determined by the corresponding exponent m (m = 1 for the linear absorption). The strongest
absorption took place when permittivity  ⇒ 0 when material was transforming from dielectric to
metal-like [50]. Under these conditions, direct light absorption m = 1 was dominant regardless of the
initial nonlinearity required to excite free carriers and to decrease the real part of the permittivity. The
experimentally observed slope γ = 2 shown in Figure 2a of the X-ray intensity as a function of the
excitation laser intensity energy E p (or I p , Fp ) was expected. This signifies a decreasing volume where
light was absorbed at the increasing electronic excitation. This was confirmed by femtosecond laser
ablation and etching where strong localization of modification took place at the very centre of the laser
irradiated spot [51].

Figure 2. Laser intensity dependencies of (a) X-ray and THz wave intensities (b) in the reflection and
(c) in the transmission. The laser polarization was horizontal to the solution flow, i.e., p-pol. During the
measurements, the solution flow position along the Z-axis was finely optimized for the highest X-ray
intensity at each laser intensity.

THz wave intensities measured in the reflection and the transmission are shown in Figure 2b,c,
respectively. For the 20 μm-thick solution flow, it showed different behaviours: THz wave emission in
the reflection was almost independent of the pulse intensity, while that in transmission was linearly
increasing ETHz ∝ E p for the p-polarized laser excitation. At the used N A = 0.25, the geometric focus
diameter was d = 1.22λ/N A ∼ 3.9 μm, which for the smallest excitation laser intensity E p = 0.1 mJ
defined the intensity I p = E p /(π (d/2)2 t p ) = 24 PW/cm2 . The actual pulse intensity was considerably
lower than this calculation due to air breakdown and intensity clamping known in the filamentation of
femtosecond pulses [43]. Such high irradiance conditions are usually not explored in the studies of
THz wave emission, and low irradiance scaling is usually PTHz ∝ Pl2 [52].
There have been controversial and various discussions on THz wave emission mechanisms from
laser-induced plasmas. As the first step for such discussions on water, however, it is meaningful to
learn discussions on similar experimental conditions with metal targets [52] to those in this paper
with water; under a grazing angle laser excitation, with flat solution flow and the THz wave radiation
from the flow surface. THz wave emission from solid flat surfaces under a grazing angle excitation,
α, (angle between the surface and the laser beam) was usually found to be much longer at ∼ 1 ps
(1 THz) than the excitation ultra-short laser pulse of tens of femtoseconds; the angle of incidence was
(α ∼ π/2) [53]. The excitation laser pulse was travelling on the surfaces at the velocity v = c/ cos α,
and the heated electrons emitted the THz field ETHz ∝ ∂Te /∂t ∝ E⊥ / sin α, where E⊥ is the light field
component perpendicular to the solid surface (p-component of the laser E-field), Te is the electron
temperature and c and t are the speed of light and time, respectively [52]. This Cherenkov-type
synchronism [52] explains THz wave emission in specular reflection, its polarization (defined by E⊥ )

135
Nanomaterials 2018, 8, 523

and the angular dependence of the THz wave emission pattern on the light incidence angle for the s
and p−polarisations. The electron scattering mechanisms and their temperature dependence define
4/(2−n)
the strongly nonlinear scaling of the emitted THz power PTHz ∝ Pl , where Pl is the laser pulse
power and n (n < 2) is the exponent defining the temperature dependence of the effective electron
collision frequency ν = ν0 + βT n [52]. Arguably, in the case of water, the plasma breakdown on its
surface acts like a metal mirror for the THz wave generated in the air breakdown region, as well as
from a plasma skin depth travelling on the flow surface at the slanted irradiation, as discussed above
(see the inset in Figure 1a). The transmission of the THz wave was weaker, but was linearly increasing
with the excitation laser intensity, as shown in Figure 2c. This is an indication of either a larger excited
volume or a higher temperature of electrons (considering that there was no strong difference in electron
scattering behaviour). The larger excited volume emitting the THz wave was one probable cause, since
the dependence of the THz wave emission on the axial position of optical femtosecond laser excitation
on the solution flow showed a wider axial width, as discussed next.

3.2. Z-Scan for THz Wave and X-ray Emission


Figure 3 shows the X-ray and THz wave intensity profiles at different solution positions along the
Z-axis when the excitation laser intensities were 0.2 and 0.7 mJ/pulse. The X-ray profile (Figure 3a)
showed an asymmetric feature with a longer tail to the upstream side. This reflects that the X-ray
emission mechanism was mainly related to laser-induced plasma formation [25]. When the solution
was set at the downstream side, the incident laser light was partly reflected by the self-induced plasma,
which resulted in the instantaneous degradation of X-ray intensity observed in the downstream side.
When the excitation laser intensity increased to 0.7 mJ/pulse (Figure 3c), the profile width became
broader from 44 μm to 50 μm with a broader tail, and this may indicate that the X-ray source size
became enlarged. X-ray emission spectra in the hard X-ray region from water under these experimental
conditions [38,39] showed a broadband emission due to bremsstrahlung with no characteristic lines
since such bands of oxygen and hydrogen are far in the longer wavelength region [54] and the estimated
electron temperature, Te , was up to 2 keV at the highest [38,39].
The peak position of THz wave emission in the reflection direction (Figure 3b) was almost the
same as that of X-ray emission though its width of THz wave emission was apparently broader at
225 μm compared with X-ray emission. This implies that there is a mechanism to enhance the THz
emission from the water plasma with low electron density and temperatures when the focal point
is far from the water film. One possible mechanism is the four-wave mixing/optical rectification
process [17] in which the second harmonic (400 nm) component of the white light continuum was
generated in air at the focal point by self-phase modulation and the residual excitation laser pulse
was mixed, which resulted in the enhancement of THz wave emission from the water plasma at the
far sides from the peak position. On the other hand, the profile width when the laser intensity was
0.7 mJ/pulse (Figure 3d) became as narrow as the width of the X-ray emission, which indicates that
the mechanism of laser-induced plasma formation dominated more in THz wave emission when the
laser intensity increased. THz wave emission from only air was negligibly small; however, when the
solution position was set far from the laser focus z > 200 μm at the downstream side, the emission
became relatively dominant, as shown in Figure 3b.
In the transmission, THz wave emission changed its nature from that in the reflection. When the
laser intensity was 0.2 mJ/pulse, intensity degradation at the peak position for THz wave emission
in the reflection was clearly observed (Figure 3b). In addition, local peaks at the upstream side
(z = −135 μm) and the downstream side (z = 24 μm) were observed as indicated by red arrows.
This feature may be assigned to the laser-induced plasma especially, at Z-positions close to the THz
wave emission peak in reflection; the laser-induced plasma reflected THz wave radiation. Under
this hypothesis, at the positions indicated
 by the red arrows, the plasma density reached its critical
density for THz wave, defined as ω p = ne e2 /me ε 0 , where ω p , ne , e, me and ε 0 are the plasma cyclic
frequency, electron density, its mass and permittivity. The critical electron density ne = ncr was

136
Nanomaterials 2018, 8, 523

1.23 × 1016 cm−3 for 1 THz. The Debye length, λ D , could be also estimated to be about 2.3 μm with Te
at 2 keV [38,39]. Similarly, when the laser pulse energy increased to 0.7 mJ/pulse (Figure 3d), such
features were clearly observed, and the position range along the Z-axis with the plasma density higher
than critical ncr = 1.23 × 1016 cm−3 was changed from Δz1 = 159 μm (0.2 mJ/pulse) to Δz2 = 229 μm
(0.7 mJ/pulse) along the Z-axis. As in the discussion of Figure 3b described above, the broad tails
at the upstream and the downstream sides can be assigned to optical coherent processes such as
four-wave mixing. Its relative intensity at the tails to the intensity close to zero-position was increasing
if compared to the case with 0.2 mJ/pulse (Figure 3b), which was considered to be the result of the
dominant plasma effect at the zero-position when the laser intensity was higher. Furthermore, similarly
to the case of the reflection shown in Figure 3b, THz wave emission from air became dominant when
the solution position was far from the laser focus and was observed as the constant THz wave intensity
at the solution position z > 250 μm at the downstream side. For all the excitation laser intensities, the
high reflectivity correlated with low transmittance of the water film target.

Figure 3. X-ray (a) and THz wave (b) intensities as a function of the flow position along the Z-axis when
the laser intensity was 0.2 mJ/pulse and 0.7 mJ/pulse (c,d); Red arrows in (b,d) for the transmission
indicate the Z-positions at (b) −135 μm and 24 μm (Δz1 = 159 μm) and (d) −167 μm and 62 μm
(Δz2 = 229 μm), respectively.

3.3. THz Wave Intensity Enhancements under Double-Pulse Excitation


One experiment for THz wave emission enhancements was performed under a double-pulse
excitation. Figure 4 shows X-ray emission (a) and THz wave emission (b,c) with the pre-pulse
(0.1 mJ/pulse, vertically-polarized, s-pol.) with the delay time of 4.6 ns in advance of the main pulse
(0.4 mJ/pulse, horizontally-polarized, p-pol.) irradiation. X-ray emission apparently showed an
intensity enhancement under the double-pulse excitation, as expected [37,38]. One additional peak
at the downstream side was also clearly discernible, as reported recently [43]. With a time delay of
4.6 ns between the pulses, the initial processes of water film ablation and transient surface roughening
under action of capillary forces and micro-droplet formation (mist) at the close position to the initial
location of the solution surface, all induced by the pre-pulse irradiation, caused a more effective

137
Nanomaterials 2018, 8, 523

coupling of the main pulse with such a modified solution surface. The enhancement was caused by
multiple scattering, local refocusing of light by droplets and the perturbed surface, which resulted in
the X-ray intensity enhancement, which can reach an order of magnitude and is useful for practical
applications. THz wave emission in the reflection (Figure 4b) was also enhanced about five times,
and the profile width under the double-pulse excitation became narrower at 52 μm as compared with
the 106 μm width under the single-pulse excitation. The profile of THz wave emission along the
Z-axis under the double-pulse excitation showed only a single peak at the centre as in the case of the
single-pulse excitation, which was different from the profile of X-ray emission with the additional peak
at the downstream side. This is consistent with requirement of thermal gradients lasting ∼1 ps for
∼1 THz emission, which are less likely on a fragmented water film, while X-ray emission is maintained
by hot plasma and geometrical factors are less important. A detailed investigation is needed using
time-resolved shadowgraphy to reveal the geometrical evolution of the disintegrating water surface. In
the case of THz wave emission in the transmission, the profile showed an apparent change (Figure 4c),
namely the intensity was enhanced more than ten times, and the profile along the Z-axis changed to a
single peak from the profiles with a dip at its centre, as shown in Figure 3b,d. Under the pre-pulse
irradiation condition, transient surface roughness, droplet (mist) formation and hole formation on
the solution flow were expected at a delay time of 4.6 ns [37]. These initial processes of laser ablation
induced by the pre-pulse irradiation may cause the enhancements of THz wave emission especially in
the transmission direction.

reflection
X-ray intensity (arb. unit)

transmission
electric field (arb. unit)

electric field (arb. unit)

Z position (μm) Z position (μm) Z position (μm)

Figure 4. Z-position-dependent intensities of the X-ray in log-scale (a), the THz wave in the reflection
(b) and the transmission (c) under the double-pulse excitation condition. Filled circles are under the
single-pulse excitation condition, and open circles are under the double-pulse excitation condition,
where the laser intensities for the main excitation pulse (horizontally-polarized, p-pol.) and the
pre-pulse (vertically-polarized, s-pol., 4.6 ns in advance of the main pulse) were 0.4 mJ/pulse and
0.1 mJ/pulse, respectively.

4. Conclusions and Outlook


This study reports the demonstration of the dual X-ray and THz wave simultaneous emission
from water flow irradiated by focused femtosecond laser pulses in air. Different characteristic features
of THz wave emission, which are associated with X-ray emission, in the reflection and the transmission
under the single-pulse excitation were clearly revealed. Enhancements of THz wave emission under
the double-pulse excitation up to 5–10 times were also shown and indicated that further enhancements
of THz wave emission are expected by optimizing the laser excitation conditions. Another option as
laser parameters to enhance THz wave emission is laser chirp [24] or double-colour excitation with
the fundamental and the second harmonic, expecting efficient augmentation of energy delivery to the
target via an optical process such as four-wave-mixing [17]. Various solution samples can be utilised as
targets, since X-ray emission from water is also enhanced by the addition of electrolyte [39] and gold
nano-particles [40]. Other materials such as bismuth [55] and copper [56] can be also used as targets
at the tested high-irradiance conditions. Interaction of the X-ray or THz wave with matter originates
with electrons at keV or meV, in other words with electrons bound in inner-shells or with structural

138
Nanomaterials 2018, 8, 523

absorption resonances, respectively. Based on the basics, X-ray and THz wave science and technology
for spectroscopy and imaging have made their progress independently [57,58]. Under experiments
in water or at atmospheric pressure, ultrasound emission is also expected [59], and super-resolution
photoacoustic imaging is also further developed [60]. With synchronized X-ray and THz wave emission
as introduced in this paper, not only for the basic mechanism study on THz wave emission from
aqueous solutions based on laser-plasma dynamics, but also combined synchronous usages of X-ray
and THz wave or ultrasound are expected to contribute well to studies on nanomaterials from the
nano-scale viewpoints to macro-scales.

Author Contributions: H.-h.H., T.N. and K.H. performed the experiments. W.-h.H. was involved in the system
setup. H.-h.H., T.N., S.J. and K.H. analysed the data and wrote the paper.
Acknowledgments: K.H. acknowledges the Japan Science and Technology Agency (JST) PRESTO (Precursory
Research for Embryonic Science and Technology) Program (SAKIGAKE, Innovative use of light and materials/life)
for its partial support of this research. S.J. is grateful for the support via the Australian Research Council
DP170100131 grant.

Funding: This research received no external funding.

Conflicts of Interest: The authors declare no conflict of interest. The founding sponsors had no role in the design
of the study; in the collection, analyses or interpretation of data; in the writing of the manuscript; nor in the
decision to publish the results.

References
1. Bagratashvili, V.N.; Letokhov, V.S.; Makarov, A.A.; Ryabov, E.A. (Eds.) Multiple Photon Infrared Laser
Photophysics and Photochemistry; Harwood Academic Publishers: Reading, UK, 1985.
2. Nakajima, K.; Deguchi, M. (Eds.) Science of Superstrong Field Interactions; American Institute of Physics:
College Park, MD, USA, 2002.
3. Yamanouchi, K.; Midorikawa, K. (Eds.) Multiphoton Processes and Attosecond Physics; Springer:
New York, NY, USA, 2012.
4. Corde, S.; Phuoc, K.T.; Lambert, G.; Fitour, R.; Malka, V.; Rousse, A.; Beck, A.; Lefebvre, E. Femtosecond X
rays from laser-plasma accelerators. Rev. Mod. Phys. 2013, 85, 1–48. [CrossRef]
5. Dey, I.; Jana, K.; Fedorov, V.Y.; Koulouklidis, A.D.; Mondal, A.; Shaikh, M.; Sarkar, D.; Lad, A.D.;
Tzortzakis, S.; Couairon, A.; et al. Highly efficient broadband terahertz generation from ultrashort laser
filamentation in liquids. Nat. Commun. 2017, 8, 1184. [CrossRef] [PubMed]
6. Fork, R.L.; Shank, C.V.; Hirlimann, C.; Yen, R.; Tomlinson, W.J. Femtosecond white-light continuum pulses.
Opt. Lett. 1983, 8, 1–3. [CrossRef] [PubMed]
7. Rethfeld, B.; Ivanov, D.S.; Garcia, M.E.; Anisimov, S.I. Modelling ultrafast laser ablation. J. Phys. D Appl. Phys.
2017, 50, 193001. [CrossRef]
8. Helliwell, J.R.; Rentzepis, P.M. (Eds.) Time-Resolved Diffraction; Oxford Science Publications:
Oxford, UK, 1986.
9. Hoffmann, M.C.; Fülöp, J.A. Intense ultrashort terahertz pulses: Generation and applications. J. Phys. D
Appl. Phys. 2011, 44, 083001. [CrossRef]
10. Tonouchi, M. Cutting-edge terahertz technology. Nat. Photonics 2007, 1, 97–105. [CrossRef]
11. Kampfrath, T.; Tanaka, K.; Nelson, K.A. Resonant and nonresonant control over matter and light by intense
terahertz transients. Nat. Photonics 2013, 7, 680–690. [CrossRef]
12. Wenz, J.; Schleede, S.; Khrennikov, K.; Bech, M.; Thibault, P.; Heigoldt, M.; Pfeiffer, F.; Karsch,
S. Quantitative X-ray phase-contrast microtomography from a compact laser-driven betatron source.
Nat. Commun. 2015, 6, 7568. [CrossRef] [PubMed]
13. Baierl, S.; Hohenleutner, M.; Kampfrath, T.; Zvezdin, A.K.; Kimel, A.V.; Huber, R.; Mikhaylovskiy, R.V.
Nonlinear spin control by terahertz-driven anisotropy fields. Nat. Photonics 2016, 10, 715–718. [CrossRef]
14. Phillips, K.C.; Gandhi, H.H.; Mazur, E.; Sundaram, S.K. Ultrafast laser processing of materials: A review.
Adv. Opt. Photonics 2015, 7, 684–712. [CrossRef]
15. Hamster, H.; Sullivan, A.; Gordon, S.; White, W.; Falcone, R.W. Subpicosecond, electromagnetic pulses from
intense laser-plasma interaction. Phys. Rev. Lett. 1993, 71, 2725–2728. [CrossRef] [PubMed]

139
Nanomaterials 2018, 8, 523

16. Löffler, T.; Jacob, F.; Roskos, H.G. Generation of terahertz pulses by photoionization of electrically biased air.
Appl. Phys. Lett. 2000, 77, 453–455. [CrossRef]
17. Cook, D.J.; Hochstrasser, R.M. Intense terahertz pulses by four-wave rectification in air. Opt. Lett. 2000,
25, 1210–1212. [CrossRef] [PubMed]
18. D’Amico, C.; Houard, A.; Franco, M.; Prade, B.; Mysyrowicz, A.; Couairon, A.; Tikhonchuk, V.T. Conical
forward THz emission from femtosecond-laser-beam filamentation in air. Phys. Rev. Lett. 2007, 98, 235002.
[CrossRef] [PubMed]
19. Nagashima, T.; Hirayama, H.; Shibuya, K.; Hangyo, M.; Hashida, M.; Tokita, S.; Sakabe, S. Terahertz pulse
radiation from argon clusters. Opt. Express 2009, 17, 8907. [CrossRef] [PubMed]
20. Balakin, A.V.; Dzhidzhoev, M.S.; Gordienko, V.M.; Esaulkov, M.; Zhvaniya, I.A.; Ivanov, K.A.; Kotelnikov, I.;
Kuzechkin, N.A.; Ozheredov, I.A.; Panchenko, V.Y.; et al. Interaction of high-intensity femtosecond radiation
with gas cluster beam: Effect of pulse duration on joint terahertz and X-ray emission. IEEE Trans. Terahertz
Sci. Technol. 2016, 7, 70–79. [CrossRef]
21. Mori, K.; Hashida, M.; Nagashima, T.; Li, D.; Teramoto, K.; Nakamiya, Y.; Inoue, S.; Sakabe, S.
Directional linearly polarized terahertz emission from argon clusters irradiated by noncollinear double-pulse
beams. Appl. Phys. Lett. 2017, 111, 241107. [CrossRef]
22. Sagisaka, A.; Daido, H.; Nashima, S.; Orimo, S.; Ogura, K.; Mori, M.; Yogo, A.; Ma, J.; Daito, I.;
Pirozhkov, A.S.; et al. Simultaneous generation of a proton beam and terahertz radiation in high-intensity
laser and thin-foil interaction. Appl. Phys. B 2008, 90, 373–377. [CrossRef]
23. Tokita, S.; Sakabe, S.; Nagashima, T.; Hashida, M.; Inoue, S. Strong sub-terahertz surface waves generated
on a metal wire by high-intensity laser pulses. Sci. Rep. 2015, 5, 8268. [CrossRef] [PubMed]
24. Qi, J.; Yiwen, E.; Williams, K.; Dai, J.; Zhang, X.-C. Observation of broadband terahertz wave generation
from liquid water. Appl. Phys. Lett. 2017, 111, 071103.
25. Attwood, D. Soft X-rays and Extreme Ultraviolet Radiation; Cambridge University Press: Cambridge, UK, 1999.
26. Turcu, I.C.E.; Dance, J.B. X-rays from Laser Plasmas; WILEY: Hoboken, NJ, USA, 1998.
27. Yoshida, M.; Fujimoto, Y.; Hironaka, Y.; Nakamura, K.G.; Kondo, K.; Ohtani, M.; Tsunemi, H. Generation of
picosecond hard x rays by tera watt laser focusing on a copper target. Appl. Phys. Lett. 1998, 73, 2393–2395.
[CrossRef]
28. Hatanaka, K.; Yomogihata, K.; Ono, H.; Nagafuchi, K.; Fukumura, H.; Fukushima, M.; Hashimoto, T.;
Juodkazis, S.; Misawa, H. Hard X-ray generation using femtosecond irradiation of PbO glass.
J. Non-Cryst. Solids 2008, 354, 5485–5490. [CrossRef]
29. Hansson, B.; Rymell, L.; Berglund, M.; Hertz, H. A liquid-xenon-jet laser-plasma x-ray and EUV source.
Microelectron. Eng. 2000, 53, 667–670. [CrossRef]
30. Vogt, U.; Stiel, H.; Will, I.; Nickles, P.V.; Sandner, W.; Wieland, M.; Wilhein, T. Influence of laser intensity and
pulse duration on the extreme ultraviolet yield from a water jet target laser plasma. Appl. Phys. Lett. 2001,
79, 2336–2338. [CrossRef]
31. Düsterer, S.; Schwoerer, H.; Ziegler, W.; Ziener, C.; Sauerbrey, R. Optimization of EUV radiation yield from
laser-produced plasma. Appl. Phys. B 2001, 73, 693–698. [CrossRef]
32. Hansson, B.A.M.; Berglund, M.; Hemberg, O.; Hertz, H.M. Stabilization of liquified-inert-gas jets for
laser–plasma generation. J. Appl. Phys. 2004, 95, 4432–4437. [CrossRef]
33. Rajyaguru, C.; Higashiguchi, T.; Koga, M.; Sasaki, W.; Kubodera, S. Systematic optimization of the extreme
ultraviolet yield from a quasi-mass-limited water-jet target. Appl. Phys. B 2004, 79, 669–672. [CrossRef]
34. Hansson, B.A.M.; Hemberg, O.; Hertz, H.M.; Choi, M.B.H.J.; Jacobsson, B.; Janin, E.; Mosesson, S.; Rymell, L.;
Thoresen, J.; Wilner, M. Characterization of a liquid-xenon-jet laser-plasma extreme-ultraviolet source.
Rev. Sci. Instrum. 2004, 75, 2122–2129. [CrossRef]
35. Hsu, W.H.; Masim, F.C.P.; Porta, M.; Nguyen, M.T.; Yonezawa, T.; Balčytis, A.; Wang, X.; Rosa, L.;
Juodkazis, S.; Hatanaka, K. Femtosecond laser-induced hard X-ray generation in air from a solution flow
of Au nano-sphere suspension using an automatic positioning system. Opt. Express 2016, 24, 19994–20001.
[CrossRef] [PubMed]
36. Hatanaka, K.; Miura, T.; Fukumura, H. White X-ray pulse emission of alkali halide aqueous solutions
irradiated by focused femtosecond laser pulses: A spectroscopic study on electron temperatures as functions
of laser intensity, solute concentration, and solute atomic number. Chem. Phys. 2004, 299, 265–270. [CrossRef]

140
Nanomaterials 2018, 8, 523

37. Hatanaka, K.; Ono, H.; Fukumura, H. X-ray pulse emission from cesium chloride aqueous solutions when
irradiated by double-pulsed femtosecond laser pulses. Appl. Phys. Lett. 2008, 93, 064103. [CrossRef]
38. Hatanaka, K.; Fukumura, H. X-ray emission from CsCl aqueous solutions when irradiated by intense
femtosecond laser pulses and its application to time-resolved XAFS measurement of I in aqueous solution.
X-ray Spectrom. 2012, 41, 195–200. [CrossRef]
39. Hatanaka, K.; Miura, T.; Fukumura, H. Ultrafast X-ray pulse generation by focusing femtosecond infrared
laser pulses onto aqueous solutions of alkali metal chloride. Appl. Phys. Lett. 2002, 80, 3925–3927. [CrossRef]
40. Masim, F.C.P.; Porta, M.; Hsu, W.H.; Nguyen, M.T.; Yonezawa, T.; Balčytis, A.; Juodkazis, S.; Hatanaka, K.
Au nanoplasma as efficient hard X-ray emission source. ACS Photonics 2016, 3, 2184–2190. [CrossRef]
41. Zhang, X.C.; Shkurinov, A.; Zhang, Y. Extreme terahertz science. Nat. Photonics 2017, 11, 16–18. [CrossRef]
42. Hatanaka, K.; Ida, T.; Ono, H.; Matsushima, S.; Fukumura, H.; Juodkazis, S.; Misawa, H. Chirp effect in
hard X-ray generation from liquid target when irradiated by femtosecond pulses. Opt. Express 2008, 16,
12650–12657. [CrossRef] [PubMed]
43. Hsu, W.H.; Masim, F.C.P.; Balčytis, A.; Juodkazis, S.; Hatanaka, K. Dynamic position shifts of X-ray
emission from a water film induced by a pair of time-delayed femtosecond laser pulses. Opt. Express 2017,
25, 24109–24118. [CrossRef] [PubMed]
44. Hatanaka, K.; Tsuboi, Y.; Fukumura, H.; Masuhara, H. Nanosecond and Femtosecond Laser Photochemistry
and Ablation Dynamics of Neat Liquid Benzenes. J. Phys. Chem. B 2002, 106, 3049–3060. [CrossRef]
45. Wu, Q.; Zhang, X. Free-space electro-optic sampling of terahertz beams. Appl. Phys. Lett. 1995, 67, 3523–3525.
[CrossRef]
46. Lee, Y.S. Principles of Terahertz Science and Technology; Springer: New York, NY, USA, 2009.
47. Exter, M.V.; Fattinger, C.; Grischkowsky, D. Terahertz time-domain spectroscopy of water vapour. Opt. Lett.
1989, 14, 1128–1130. [CrossRef] [PubMed]
48. Novelli, F.; Chon, J.W.M.; Davis, J.A. Terahertz thermometry of gold nanospheres in water. Opt. Lett. 2016,
41, 5801–5804. [CrossRef] [PubMed]
49. Thrane, L.; Jacobsen, R.H.; Uhd Jepsen, P.; Keiding, S.R. THz reflection spectroscopy of liquid water.
Chem. Phys. Lett. 1995, 240, 330–333. [CrossRef]
50. Gamaly, E.G.; Rode, A.V. Ultrafast re-structuring of the electronic landscape of transparent dielectrics: New
material states (Die-Met). Appl. Phys. A 2018, 124, 278. [CrossRef]
51. Cao, X.W.; Chen, Q.D.; Fan, H.; Juodkazis, L.Z.S.; Sun, H.B. Liquid-Assisted Femtosecond Laser
Precision-Machining of Silica. Nanomaterials 2018, 8, 287. [CrossRef] [PubMed]
52. Oladyshkin, I.V. Diagnostics of the electron scattering in metals in terms of a terahertz response to
femtosecond laser pulse. JETP Lett. 2016, 103, 435–439. [CrossRef]
53. Oladyshkin, I.V.; Fadeev, D.A.; Mironov, V.A. Thermal mechanism of laser induced THz generation from a
metal surface. J. Opt. 2015, 17, 075502. [CrossRef]
54. Kirz, J.; Attwood, D.; Henke, B.L.; Howells, M.R.; Kennedy, K.D.; Kim, K.J.; Kortright, J.B.; Perera, R.C.;
Pianetta, P.; Riordan, J.C.; et al. X-ray Data Booklet; Lawrence Berkeley National Laboratory, University of
California: Berkeley, CA, USA, 1986.
55. Ilyakov, I.E.; Shishkin, B.V.; Fadeev, D.A.; Oladyshkin, I.V.; Chernov, V.V.; Okhapkin, A.I.; Yunin, P.A.;
Mironov, V.A.; Akhmedzhanov, R.A. Terahertz radiation from bismuth surface induced by femtosecond
laser pulses. Opt. Lett. 2016, 41, 4289–4292. [CrossRef] [PubMed]
56. Lu, X.; Ishida, Y.; Mishina, T.; Nguyen, M.T.; Yonezawa, T. Enhanced Terahertz Emission from CuxO/Metal
Thin Film Deposited on Columnar-Structured Porous Silicon. Bull. Chem. Soc. Jpn. 2015, 88, 1385–1387.
[CrossRef]
57. Sakdinawat, A.; Attwood, D. Nanoscale X-ray imaging. Nat. Photonics 2010, 4, 840–848. [CrossRef]
58. Mittleman, D.M. Twenty years of terahertz imaging. Opt. Express 2018, 26, 9417–9431. [CrossRef] [PubMed]

141
Nanomaterials 2018, 8, 523

59. Masim, F.C.P.; Hsu, W.H.; Liu, H.L.; Yonezawa, T.; Balčytis, A.; Juodkazis, S.; Hatanaka, K.
Photoacoustic signal enhancements from gold nano-colloidal suspensions excited by a pair of time-delayed
femtosecond pulses. Opt. Express 2017, 25, 19497–19507. [CrossRef] [PubMed]
60. Chaigne, T.; Arnaland, B.; Vilov, S.; Bossy, E.; Katz, O. Super-resolution photoacoustic imaging via
flow-induced absorption fluctuations. Optica 2017, 4, 1397–1404. [CrossRef]

c 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

142
MDPI
St. Alban-Anlage 66
4052 Basel
Switzerland
Tel. +41 61 683 77 34
Fax +41 61 302 89 18
www.mdpi.com

Nanomaterials Editorial Office


E-mail: [email protected]
www.mdpi.com/journal/nanomaterials
MDPI
St. Alban-Anlage 66
4052 Basel
Switzerland
Tel: +41 61 683 77 34
Fax: +41 61 302 89 18
www.mdpi.com ISBN 978-3-03897-411-6

You might also like