ArticlePDF Available

Rayleigh wave and super-shear evanescent wave excited by laser-induced shock at a soft solid–liquid interface observed by photoelasticity imaging technique

Authors:

Abstract and Figures

We investigated laser-induced shock excitation of elastic surface waves at a free surface and a soft solid–liquid interface using a custom-designed photoelasticity imaging technique. Epoxy-resin and pure water were selected as the solid and liquid media. The elastic surface waves were excited via a shock process induced by focusing a single nanosecond laser pulse on the solid surface. To confirm the experimental observations, the roots of the Rayleigh and Stoneley equations were calculated. For a free surface, we present an entire-field observation of elastic surface waves, which includes a super-shear evanescent wave (SEW) that propagates faster than the shear wave but slower than the longitudinal wave. For a soft solid–liquid interface, we demonstrate the presence of a non-leaky Rayleigh wave that corresponds to a real root of the Stoneley equation. We also evidence the existence of a SEW that propagates 1.7 times faster than the shear speed in the solid and corresponds to a complex conjugate root of the Stoneley equation. These results correct the previously accepted notion that the Scholte wave is the only surface wave that can be generated at a soft solid–liquid interface.
Content may be subject to copyright.
J. Appl. Phys. 131, 123102 (2022); https://doi.org/10.1063/5.0081237 131, 123102
© 2022 Author(s).
Rayleigh wave and super-shear evanescent
wave excited by laser-induced shock at
a soft solid–liquid interface observed by
photoelasticity imaging technique
Cite as: J. Appl. Phys. 131, 123102 (2022); https://doi.org/10.1063/5.0081237
Submitted: 08 December 2021 • Accepted: 05 March 2022 • Published Online: 28 March 2022
Thao Thi Phuong Nguyen, Rie Tanabe-Yamagishi and Yoshiro Ito
COLLECTIONS
Paper published as part of the special topic on Shock Behavior of Materials
This paper was selected as Featured
This paper was selected as Scilight
ARTICLES YOU MAY BE INTERESTED IN
Redshifted biexciton and trion lines in strongly confined (211)B InAs/GaAs piezoelectric
quantum dots
Journal of Applied Physics 131, 123101 (2022); https://doi.org/10.1063/5.0084931
Fiber-form nanostructured tungsten formation by helium arc discharge plasma irradiation
under a gas pressure of 5kPa
Journal of Applied Physics 131, 123301 (2022); https://doi.org/10.1063/5.0085563
Quantum nickelate platform for future multidisciplinary research
Journal of Applied Physics 131, 120901 (2022); https://doi.org/10.1063/5.0084784
Rayleigh wave and super-shear evanescent
wave excited by laser-induced shock at a soft
solidliquid interface observed by photoelasticity
imaging technique
Cite as: J. Appl. Phys. 131, 123102 (2022); doi: 10.1063/5.0081237
View Online Export Citation CrossMar
k
Submitted: 8 December 2021 · Accepted: 5 March 2022 ·
Published Online: 28 March 2022
Thao Thi Phuong Nguyen,
1,2,a)
Rie Tanabe-Yamagishi,
3
and Yoshiro Ito
4
AFFILIATIONS
1
Institute of Research and Development, Duy Tan University, 550000 Danang, Vietnam
2
Faculty of Natural Science, Duy Tan University, 550000 Danang, Vietnam
3
Department of Intelligent Mechanical Engineering, Fukuoka Institute of Technology, 3-30-1 Wajiro-higashi, Higashi-ku,
Fukuoka 811-0295, Japan
4
Department of Mechanical Engineering, Nagaoka University of Technology, 1603-1 Kamitomioka, Nagaoka, Niigata 940-2188,
Japan
Note: This paper is part of the Special Topic on Shock Behavior of Materials.
a)
Author to whom correspondence should be addressed: thaonguyen@duytan.edu.vn
ABSTRACT
We investigated laser-induced shock excitation of elastic surface waves at a free surface and a soft solidliquid interface using a custom-
designed photoelasticity imaging technique. Epoxy-resin and pure water were selected as the solid and liquid media. The elastic surface
waves were excited via a shock process induced by focusing a single nanosecond laser pulse on the solid surface. To confirm the experimen-
tal observations, the roots of the Rayleigh and Stoneley equations were calculated. For a free surface, we present an entire-field observation
of elastic surface waves, which includes a super-shear evanescent wave (SEW) that propagates faster than the shear wave but slower than the
longitudinal wave. For a soft solidliquid interface, we demonstrate the presence of a non-leaky Rayleigh wave that corresponds to a real
root of the Stoneley equation. We also evidence the existence of a SEW that propagates 1.7 times faster than the shear speed in the solid and
corresponds to a complex conjugate root of the Stoneley equation. These results correct the previously accepted notion that the Scholte wave
is the only surface wave that can be generated at a soft solidliquid interface.
Published under an exclusive license by AIP Publishing. https://doi.org/10.1063/5.0081237
INTRODUCTION
The excitation of elastic surface waves by pulsed laser has
received widespread attention in the last few decades.
1
Compared
to conventional methods, the laser beam has proved most useful
for generating and investigating elastic surface waves, thanks to its
non-contact feature and the ability to generate wide-frequency-
band acoustic disturbances.
2
The smallest wavelength of a laser-
excited surface elastic wave can reach a few micrometers, which
corresponds to a frequency approaching 1 GHz and limits the pulse
length to the nanosecond range.
3
This advantage provides opportu-
nities to separate different wave modes propagating on the interface
and, thus, is important to study the interface waves traveling at
only slightly different speeds.
2
The excitation of elastic surface
waves demonstrates great potential for a variety of applications,
including nondestructive testing,
46
investigation of the elastic
properties of surfaces and thin films,
7,8
evaluation of nanostructure
and nanomaterials,
9
and more recently, elastography.
10
Depending on the fluence of the laser pulse, we can classify
the mechanism for the laser excitation of elastic surface waves into
the thermoelastic regime and the ablative regime.
11
Thermoelastic
expansion is the most well-known mechanism for transforming a
laser pulse into acoustic waves. In this regime, low laser irradiation
Journal of
Applied Physics ARTICLE scitation.org/journal/jap
J. Appl. Phys. 131, 123102 (2022); doi: 10.1063/5.0081237 131, 123102-1
Published under an exclusive license by AIP Publishing
induces the localized thermal expansion of the target and generates
surface disturbances.
10,12
In the ablative regime, the laser pulse
ablates the target material to form a high-pressure plasma. Upon
expansion, the plasma induces a shock into the surrounding
media.
13
This laser-induced shock drives an impulse into the solid
and excites four kinds of waves: longitudinal wave (P-wave), shear
wave (S-wave), lateral waves, and elastic surface waves.
14
Compared
to the thermoelastic regime, the shock regime can induce much
stronger acoustic impulses and is a newly established non-contact
elastic surface wave excitation technology.
13
Although the laser-
induced shock process has been thoughtfully investigated in the
last decades, the laser-induced shock excitation of elastic surface-
waves has not been investigated much in the literature.
For a free surface, the Rayleigh wave is the most popular
elastic surface wave that can be excited. The Rayleigh surface wave
is first discovered theoretically as a root of the Rayleigh equation in
the late 19th century and was confirmed later, thanks to many
experimental investigations. In the absence of losses, the Rayleigh
wave is continuous, nondispersive, and exists in the entire fre-
quency range, where the elastic medium can be considered uniform
and isotropic.
15
For studying elastic surface wave excited at the interface
between two media, one must consider the Stoneley wave equation.
In the 1950s, Scholte described a specific case of the Stoneley wave
equation where one medium is liquid and discovered the Scholte
wave.
4
The Scholte wave travels along a solidliquid interface, has
its energy mostly concentrated in the liquid, and propagates
without attenuation at a speed smaller than the Rayleigh speed if
the viscosity of the liquid is negligible.
4,16
For a general solidliquid
interface, a Scholte wave can always be generated. To discuss the
excitation of the Rayleigh wave at a solidliquid interface, we need
to distinguish two situations: a hard solidliquid interface and a
soft solidliquid interface. A hard solidliquid interface refers to
the liquid sound speed smaller than the solid shear speed. Here,
the Rayleigh wave has a speed larger than the sound speed in the
liquid, thus radiating energy in the liquid and becoming leaky
(so-called leaky Rayleigh wave).
1719
This wave is the least
damped, thus most observable of all the leaking wave types,
20
and
has been thoroughly discussed in the literature. A soft solid
liquid interface refers to the interface where the liquid sound
speed is larger than the solid shear speed. In contrast to the hard
solidliquid interface, the soft solidliquid interface has been
rarely investigated. In 1999, Padilla et al. showed experimentally
and theoretically that for a soft solidliquid interface, the complex
root that gives rise to the leaky Rayleigh wave becomes real and
gives rise to a non-Leaky Rayleigh wave.
18
However, a conflict
result was presented by Glorieux et al.
19
In recent years, the evidence of an inhomogeneous wave that
propagates along a free surface faster than the S-wave but slower
than the P-wave renews the research interest in this topic. This
elastic surface wave has a speed almost twice the S-wave speed and
couples into the S-wave during its propagation. Because of this cou-
pling, it loses energy and decays in its propagating direction.
14,21,22
Throughout the literature, this wave has been called the leaky
wave,
23
the super-shear Rayleigh wave,
21
the pseudo P-wave,
20
non-
geometric P-S wave,
24
or most recently, the super-shear evanescent
wave.
22
We will use super-shear evanescent wave(SEW) when
referring to this wave in this paper since it best describes the leaky
behavior and super-shear propagation speed of the wave.
The existence of SEWs attracts significant interest in seismol-
ogy, engineering, and nondestructive testing. Super-shear rupture, a
rupture that propagates at a speed faster than the shear wave, has
recently been observed in large seismic events.
25
These events relate
to the existence of the SEW and break the theoretical limit that
cracks should not propagate faster than the speed of the Rayleigh
wave.
21
The existence of the SEW can also lead to misidentification
of elastic properties of structures being investigated if one picking
phase velocities of the SEW as the Rayleigh wave phase velocities
for inversion.
23
In recent years, the SEW has also been proposed as
a non-contact tool to measure the elastic properties of soft mate-
rial.
22
Altogether, understanding the process driving SEW genera
tion and propagation characteristics is crucial for both minimizing
surface wave-induced fracture as well as for improving the efficacy of
surface elastic waves in nondestructive testing applications. Although
SEWs have been studied theoretically and experimentally for a free
surface, the existence of a similar wave at a solidliquid interface has
not been discussed that much. In general, a SEW is hard to detect
because it may not separate from the Rayleigh wave until propagating
some distance from its source, at which point the evanescent nature
makes measurement difficult.
22
Moreover, the feasibility of acquiring
this type of wave may require the shear speed in the solid smaller
than the sound speed of the liquid
26
(i.e., soft solidliquid interface),
a condition that has not been discussed much in the literature.
Under these circumstances, it is the purpose of this paper to
examine the laser-induced shock excitation of Rayleigh wave and
SEW at a soft solidliquid interface experimentally. One difficulty
in studying elastic surface waves at a solidliquid interface is the
requirement of a method that can capture the whole-field elastic
disturbances at a resolution sufficient to separate different wave
modes. The piezoelectric transducer is the conventional method for
studying elastic surface waves. In this technique, transducers are
connected to the sample surface and record the arrival time of the
wave trains.
18
The major disadvantage of this method is the mea-
suring accuracy depends on the contact conditions between the
transducer and the specimen.
27,28
Optical detection has been
known as a powerful non-contact method for investigating surface
waves. We can count here, for example, the beam reflection tech-
nique,
29
Brillouin scattering,
30
optical interferometry,
30
and inter-
ferometric method based on the photoelastic effect.
16,31
The
disadvantages of these methods are that they require a high quality
of the surface and rely on the spatial derivative of the surface dis-
placement. Various optical interferometric techniques have been
developed to improve the detection sensitivity and robustness for
industrial applications. For example, the signal quality and reliabil-
ity of laser Doppler vibrometers can be significantly improved by
applying diversity reception technique.
32
By integrating detector
arrays, a classic interferometric design such as a Michelson or
MachZehnder interferometer can be transformed into a robust
system with high sensitivity and can endure a tough industrial envi-
ronment.
33
Another non-contact approach to detect elastic surface
waves with high stability and sensitivity in the harsh environment
is the interferometer schemes employing photorefractive crys-
tals.
34,35
Sagnac-based interferometers have also been developed to
provide an alternative, cost-effective, and easy-to-use scheme for
Journal of
Applied Physics ARTICLE scitation.org/journal/jap
J. Appl. Phys. 131, 123102 (2022); doi: 10.1063/5.0081237 131, 123102-2
Published under an exclusive license by AIP Publishing
industrial inspection of acoustic emission.
36
A fiber-optic Sagnac
interferometer developed by Pelivanov et al. has been demonstrated
to have the detection sensitivity approaching a piezoelectric trans-
ducer with a reduced sensitivity to roughnes,
37,38
thus enabling the
detection of surface acoustic waves propagating over rough surfaces
of composite materials.
39,40
To obtain a direct whole-field observation of the induced
elastic surface wave, we use the approach of dynamic photoelastic
imaging. In our previous works,
4143
we have combined a high-
speed pump-and-probe imaging system and the photoelasticity
technique to develop a custom-designed photoelasticity imaging
system. This system can give a direct and whole-field visualization
of the transient stress distribution inside a solid with high temporal
and spatial resolutions. In addition, it provides a normal shadow-
graph image in the liquid phase at the same time. With these abili-
ties, the photoelasticity imaging technique can serve as a powerful
tool for non-contact, direct, and whole-field visualization of the
elastic surface waves. Recently, Zhang et al. have employed a
similar dynamic photoelastic/shadowgraph imaging system to study
the transient stress waves excited by the incident shock at a fluid
solid boundary.
44
Although the resolution was low because of the
small intensity of the phenomenon under investigation, this tech-
nique provided a whole-field image of surface waves. Their work,
however, studied a hard solidliquid interface and focused on
dynamic fracture mechanisms at the solid surface.
In this research, we selected epoxy-resin and water as solid
and liquid media, as they compose a soft solidliquid interface and
Poisons ratio (ν) of epoxy-resin which is within the domain to
generate the SEW (ν.0:263).
23,24
Furthermore, epoxy-resin has a
high photoelasticity constant that allows clear images of the
induced elastic waves. The elastic surface waves were excited by
focusing a single nanosecond laser pulse on the solid surface. We
first observed the elastic surface waves at a free epoxy-resin surface
to show the ability of the photoelasticity imaging technique as a
potential method for investigating the elastic surface waves. Then,
we used this method to observe the surface disturbances induced at
the interface between epoxy-resin and water. The roots of the
Rayleigh and Stoneley equations were calculated to verify the exper-
imental observations. Our results provide direct and visual evidence
that a non-leaky Rayleigh wave and a SEW can be excited by a
laser beam at a soft solidliquid interface.
CHARACTERISTIC EQUATIONS FOR SURFACE WAVES
The presentation of the Rayleigh equation and Stoneley equa-
tion in this section follows the work of Ansell.
45
Rayleigh equation
The Rayleigh equation for a wave number kalong the interface
can be expressed as
45
R;4uffiffiffiffiffiffiffiffiffiffi
ul
pffiffiffiffiffiffiffiffiffiffiffi
un
pþ(2un)2¼0, (1)
where u¼k2,n¼ω2
c2
s,l¼ω2
c2
p, with c
s
and c
p
are the S-wave speed
and P-wave speed in the solid, respectively, and ωis the angular
frequency of the waves.
The square roots in R give rise to two branch points at u¼l
and u¼n. The top sheets (referred to as + sheets) are defined by
π
2arg ffiffiffiffiffiffiffiffiffiffi
ul
p,arg ffiffiffiffiffiffiffiffiffiffiffi
un
p

π
2. The Riemann surface for R
has four sheets designated by a group of two signs (+,+). Each of
the signs stands for the Riemann sheet of ffiffiffiffiffiffiffiffiffiffi
ul
p,ffiffiffiffiffiffiffiffiffiffiffi
un
p

. The
Rayleigh equation, thus, can take two forms on the four sheets,
4uffiffiffiffiffiffiffiffiffiffi
ul
pffiffiffiffiffiffiffiffiffiffiffi
un
pþ(2un)2¼0
and
þ4uffiffiffiffiffiffiffiffiffiffi
ul
pffiffiffiffiffiffiffiffiffiffiffi
un
pþ(2un)2¼0:
On squaring (1), filtering out the frequency dependence by writing
the equation intern of wave speed c¼ω
k, and extracting the irrele-
vant root c¼0, the Rayleigh equation becomes the well-known
polynomial,
46
Ψ¼c2
c2
s

3
8c2
c2
s

2
þ24 16 c2
s
c2
p
!
c2
c2
s16 1 c2
s
c2
p
!
¼0:(2)
Equation (2) always has three roots for cthat have the real
part larger than zero.
47
One of these three roots is always real and
gives rise to the well-known Rayleigh wave.
45,47
Depending on
Poissons ratio of the solid medium, the other two roots can be real
or complex.
45,47
For Poissons ratio larger than 0.263, the two roots
are complex conjugates of each other and give rise to the SEW.
22,23
Stoneley equation
The form of the Stoneley equation given for the wave number
kalong the interface is
45
S;ffiffiffiffiffiffiffiffiffiffiffiffi
um
p4uffiffiffiffiffiffiffiffiffiffi
ul
pffiffiffiffiffiffiffiffiffiffiffi
un
pþ(2un)2
hi
þρn2ffiffiffiffiffiffiffiffiffiffi
ul
p¼0,
(3)
where
u¼k2,ρ¼ρliquid
ρsolid
,n¼ω2
c2
s
,m¼ω2
c2
l
,l¼ω2
c2
p
,
where csand cpare the S-wave and P-wave speeds in the solid and
clis the sound speed in the liquid. ρliquid and ρsolid are the densities
of the liquid and solid, respectively. kis the wave number in the
horizontal direction, and ωis the angular frequency of the waves.
Three square roots in S give rise to three branch points in the
complex uplane at u¼l,u¼m, and u¼n. The branch cuts
were taken as (1,l) for ffiffiffiffiffiffiffiffiffiffi
ul
p,(1,m) for ffiffiffiffiffiffiffiffiffiffiffiffi
um
p, and
(1,n) for ffiffiffiffiffiffiffiffiffiffiffi
un
p. The top Riemann sheet (referred to as the
+ sheet) for each of the square roots is given by
π
2(arg ffiffiffiffiffiffiffiffiffiffi
ul
p,arg ffiffiffiffiffiffiffiffiffiffiffiffi
um
p,arg ffiffiffiffiffiffiffiffiffiffiffi
un
pπ
2:
The Riemann surface for the function S has eight sheets, each
of which are defined by a combination of sheets
ffiffiffiffiffiffiffiffiffiffi
ul
p,ffiffiffiffiffiffiffiffiffiffiffiffi
um
p,ffiffiffiffiffiffiffiffiffiffiffi
un
p

. And so, the general equation (3) can
Journal of
Applied Physics ARTICLE scitation.org/journal/jap
J. Appl. Phys. 131, 123102 (2022); doi: 10.1063/5.0081237 131, 123102-3
Published under an exclusive license by AIP Publishing
be written as eight separate equations,
45
S1;ffiffiffiffiffiffiffiffiffiffiffiffi
um
p4uffiffiffiffiffiffiffiffiffiffi
ul
pffiffiffiffiffiffiffiffiffiffiffi
un
pþ(2un)2
hi
þρn2ffiffiffiffiffiffiffiffiffiffi
ul
p¼0,
S2;ffiffiffiffiffiffiffiffiffiffiffiffi
um
p4uffiffiffiffiffiffiffiffiffiffi
ul
pffiffiffiffiffiffiffiffiffiffiffi
un
pþ(2un)2
hi
ρn2ffiffiffiffiffiffiffiffiffiffi
ul
p¼0,
S3;ffiffiffiffiffiffiffiffiffiffiffiffi
um
pþ4uffiffiffiffiffiffiffiffiffiffi
ul
pffiffiffiffiffiffiffiffiffiffiffi
un
pþ(2un)2
hi
ρn2ffiffiffiffiffiffiffiffiffiffi
ul
p¼0,
S4;ffiffiffiffiffiffiffiffiffiffiffiffi
um
pþ4uffiffiffiffiffiffiffiffiffiffi
ul
pffiffiffiffiffiffiffiffiffiffiffi
un
pþ(2un)2
hi
þρn2ffiffiffiffiffiffiffiffiffiffi
ul
p¼0,
S1;S1¼0, S2;S2¼0, S3;S3¼0, S4;S4¼0:
(4)
To solve the set of equations (4), we use the function T intro-
duced by Ansell,
45
where T is defined by T ¼S1S2S3S4,
T¼n4(ul)ρ2(um)þ4uffiffiffiffiffiffiffiffiffiffi
ul
pffiffiffiffiffiffiffiffiffiffiffi
un
pþ(2un)2
no
2

n4(ul)ρ2(um)4uffiffiffiffiffiffiffiffiffi
ul
pffiffiffiffiffiffiffiffiffiffiffi
un
pþ(2un)2
no
2

:
(5)
By writing equation T ¼0 in term of wave speed c¼ω
k,wehave
c
cs

8
1c
cp

2
!
ρ21c
cl

2
!
þ4ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1c
cs

2
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1c
cp

2
sþ2c
cs

2
!
2
8
<
:9
=
;
2
2
43
5
c
cs

8
1c
cp

2
!
ρ21c
cl

2
!
4ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1c
cs

2
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1c
cp

2
sþ2c
cs

2
!
2
8
<
:9
=
;
2
2
43
5¼0:(6)
The Stoneley wave equation can give rise to either real roots
or complex roots.
45
The slowest real root of the Stoneley equa-
tion for a solidliquid interface corresponds to the Scholte
wave.
48
For a hard solidliquid interface, one of the complex
roots gives rise to the leaky Rayleigh wave.
17
While the roots that
correspond to the Scholte wave and leaky Rayleigh wave are rec-
ognized and treated extensively in the literature, the other roots
have not been discussed this much and have usually been con-
sidered trivial or extraneous.
EXPERIMENTAL METHODS
Laser-induced shock excitation of elastic surface
waves
We excited the elastic surface waves by focusing a single
1064-nm laser pulse with a full width at half maximum (FWHM)
of 13 ns through a objective lens onto an epoxy-resin surface
to a spot approximate 50 μm in radius. When interacting with the
target, the laser beam gasifies and ionizes the target material to
form a plasma. This plasma continues to absorb the laser pulse
energy to get a high-temperature, high-pressure status. Upon
expansion, this plasma drives a shock wave into the upper
medium and an impulse on the target surface that excites the
elasticwavesintothesolid.Figures 1(a) and 1(b) show the two
configurations of the experiment in this research. In the free
surface configuration [Fig. 1(a)],thetargetisputintheair.
We covered the epoxy-resin surface with a thin layer of black
paint to enhance the absorptivity because laser ablation of
epoxy-resin in air generates weak elastic waves that cannot be
observed.
41,49
In the solidliquid configuration [Fig. 1(b)], the
target is put in a glass vessel filled with pure water. The waterair
interface is 5 mm above the target. The laser beam is directed
through the water and focused on the target surface. In this
experiment, we did not apply an absorptive coating to the target
because the impulse was significantly increased, thanks to the
confinement effect of the liquid.
41,50
The target was put on an
XY translational stage and was moved after each shot to refresh
the surface. The pulse energy was 60 mJ.
Photoelasticity imaging system
We used a custom-designed time-resolved photoelasticity
imaging technique to observe the laser-induced elastic surface
waves. In this technique, a circular polariscope and fast imaging
system are used to achieve a whole-field visualization of the tran-
sient stress inside the solid [Fig. 1(c)]. The imaging system was
based on a pump-and-probe system with an intensified charge-
coupled device (ICCD) camera operated in a gated mode to
capture a single-shot image for each ablation pulse.
49
This
system captured photoelasticity images for epoxy-resin block
while it gave normal shadowgraph images for non-birefringent
media, water phase, in our case. We used a Nd:YAG laser
(Powerlite 8000) to induce the elastic waves, and another
Nd:YAG laser (NY 82) to provide the illuminating light
(532 nm, FWHM = 6 ns). We defined the delay time as the inter-
val between the pump pulse and the probe pulse. Although this
technique has a high temporal resolution of nanoseconds, it
allows only single-shot observations where one image is cap-
tured per ablation pulse,
49
thus measurements with varying
delay durations must be conducted to capture the whole time-
dependent evolution of the elastic waves.
Journal of
Applied Physics ARTICLE scitation.org/journal/jap
J. Appl. Phys. 131, 123102 (2022); doi: 10.1063/5.0081237 131, 123102-4
Published under an exclusive license by AIP Publishing
RESULTS AND DISCUSSION
Elastic surface waves observed at a free surface
In Fig. 2, we present a visualization of laser-induced elastic
surface waves at a free surface. Figure 2(a) shows the evolution of a
shock wave in the air and elastic waves induced in the solid,
observed from 100 to 1900 ns after irradiation. Figure 2(b) shows
detailed images of the waves observed at 700 ns. Although each
frame was taken for a different event, the propagation of excited
waves can be well followed, thanks to the good reproducibility of
the events.
As early as 100 ns after irradiation, we can observe a shock
wave that expanded fast into the air. At the laser fluence used in
our experiment, the induced plasma expands following a
laser-supported-radiation-wave model, and thus the shock front
does not have a semi-circular shape but is slightly elongated in the
vertical direction.
50
Inside the shock wave in the air, there was a
shadow of the ablated material. Inside the solid, we could observe
elastic wavefronts. From 300 ns, the longitudinal wave (P-wave)
and the shear wave (S-wave) can be distinguished. From 700 ns,
two surface waves are visible [Fig. 2(b)]. The 1st surface wave is
characterized as a dark pattern propagating along the interface
slightly slower than the S wave. The 2nd surface wave is character-
ized as a tangent line to the S-wave front with one end is on the
interface. This wave vanished fast and its location on the interface
is difficult to be determined after 900 ns. From 1000 ns, the P-wave
was almost undetected, but the S-wave, 1st, and 2nd surface waves
can be observed until 1900 ns.
We tracked the distance traveled by the P-wave, S-wave, and
the 1st surface wave with time. To measure the distances traveled
by the P-wave and S-wave, we centered a circle at the laser-focal
point and fit it to the wavefronts. The distances traveled by these
wavefronts were elucidated from the radius of the circle. The results
are presented in Fig. 3 with the plots presenting the mean value
measured at each delay time, and the average deviation at each
delay time is within 1%. To estimate the wave speeds, we fitted a
linear regression model to each data set. High goodness of fit
(R
2
> 0.99) confirmed the reproducibility of the result. We found
that the P- and S-wave propagated at constant speeds of 2550 +40
and 1110 +20 ms1, respectively. The 1st surface wave propagated
at a constant speed of 1030 +20 ms1, which is about 0:93 times
the S-wave speed.
The 2nd surface wave had one end traveling along the
solidair boundary and appeared as a tangent line to the S-wave
front. The propagation of the 2nd surface wave on the interface was
difficult to track because it vanished at a large delay time and its
position on the boundary was not clear to locate. On the other
hand, the angle θbetween it and the interface showed good
stability and was measured to be θ¼32+1. Thus, from the
S-wave speed of cs¼1110 +20 ms1, we estimated that the 2nd
surface wave propagates along the free surface at the speed
c2¼cs
sinθ¼2100 +100 ms1, which approximates 1.9 times the
S-wave speed.
To confirm the nature of the observed elastic surface waves,
we solved the Rayleigh equation using Eq. (2) for epoxy-resin with
cs¼1110 ms1and cp¼2550 ms1. The results return a real root
c¼0:94cs, which is close to the speed of the 1st wavefront, and a
complex conjugate root c0¼(1:92 +0:37i)cs, of which the real
part is close to the speed of the 2nd wavefront.
FIG. 1. Simplified schematic of the experimental setup. Laser-induced shock excitation of elastic waves at a free surface (a) and at a soft solidliquid interface (b).
(c) Schematic diagram of the photoelasticity imaging system (BE: beam expander, P: polarizer, Q: quarter wave plate, A: analyzer, ND: neutral density filter, and BP: band-
pass filter).
Journal of
Applied Physics ARTICLE scitation.org/journal/jap
J. Appl. Phys. 131, 123102 (2022); doi: 10.1063/5.0081237 131, 123102-5
Published under an exclusive license by AIP Publishing
It is well known that the real root of the Rayleigh equation
gives rise to the Rayleigh wave speed. We, thus, can conclude that
the 1st wavefront is the Rayleigh wave.
The complex conjugate root of the Rayleigh equation has been
demonstrated theoretically to give rise to a SEW which propagates
faster than the shear wave and is weakly transverse in polariza-
tion.
14,20
Due to matching tangential wave vectors at the surface, it
couples into the plane shear wave and decays as it propagates along
the surface.
14
Although the physical admissibility of this wave has
been previously debated, current experimental observations in the
near field of excitation source confirm its existence.
21,22
It is also
demonstrated that the non-geometric wave that was reported in
many shallow high-frequency seismic surveys is a SEW.
23
In our
observation, the 2nd surface wave corresponds to a complex conju-
gate root of the Rayleigh equation. It travels faster than the shear
wave, thus radiating the energy into the solid at the S-wave speed
and appearing in the image as a tangent line to the S-wave front.
We, thus, conclude that the 2nd wavefront observed in our image is
the SEW.
Although the existence of SEW has been reported,
2123
this
result presents a direct whole-field visualization of the SEW at a
free surface. Such a whole-field image of SEW was only provided
by simulation before. Thus, we propose the photoelasticity imaging
technique as an effective and unique method for whole-field and
direct monitoring of elastic surface waves.
Elastic surface waves observed at a soft solidliquid
interface
Figure 4 presents the propagation of elastic waves observed at
the soft solidliquid interface. Because of the confinement effect of
the liquid, the impulse has been significantly increased. As a result,
we observed a complicated fringe system inside the solid.
41
In the
liquid medium, we can observe the primary shock wave, reflected
wave, and the cavitation bubble, which had been described in our
previous work.
43
In the solid phase, we can identify the P-wave,
which is the fastest wavefront. Inside this wavefront are photoelastic
fringes that semi-quantitatively show the stress distribution in the
FIG. 2. Laser-induced shock excitation of elastic surface waves at a free surface observed by the photoelasticity imaging technique. (a) Evolution of elastic waves induced
in the solid. (b) Detailed image of elastic waves observed at 700 ns after irradiation.
Journal of
Applied Physics ARTICLE scitation.org/journal/jap
J. Appl. Phys. 131, 123102 (2022); doi: 10.1063/5.0081237 131, 123102-6
Published under an exclusive license by AIP Publishing
solid phase.
41
Because of the presence of many fringes, the
S-wavefront cannot be distinguished. However, the P-S lateral wave
can be identified. The angle at which the P-S lateral wavefront
makes to the interface can be used to calculate the S-wave speed.
Our observation at 1000ns shows that the P-S lateral wave made
an angle 28to the solidliquid boundary, revealing that the
S-wave speed was 1180 ms
1
(with P-wave speed was 2550 ms
1
).
This speed is higher than the S-wave speed observed for the free
surface. This difference can be explained by the fact that that the
shear wave speed increased as the induced stress was increased.
52
In the interval of 12002000 ns, we can separate three waves at
the boundary. We named them #1, #2, and #3 wavefronts as stated
in Fig. 4(a). The #1 wavefront has the shape of near-vertical fringes.
The #2 wavefront was recognized as concentric semicircle fringes.
The #3 wavefront was recognized as fringes that were inclined to
the solidliquid interface. In the photoelastic images, the number
of fringes semi-quantitatively shows the strength of the wave, and
the width of the fringe group in the propagation direction can
show the wavelength. We, thus, used the center point of each fringe
group as the track point to measure the distance traveled by each
wavefront. A closer look into the near-boundary area at earlier
delay time revealed the #4 wavefront that appeared between P-wave
and S-wave. The #4 wavefront was distinguished from the #3 wave-
front in that it travels at a higher speed. We show a magnified
observation of this wavefront from 600 to 900 ns in Fig. 4(b).
Within this period, the #4 wavefront is well distant from the group
of #1, #2, and #3 wavefronts. Before 500 ns, the #4 wave could not
be separated from other surface waves. After 900 ns, this wave was
getting dim and its position on the interface was difficult to locate.
Figure 4(c) shows the travel distance over time of the observed
wavefronts. The plots present the mean value measured at each
delay time. The average deviation at each point is within 3%. By
applying a linear fit to each data set, we can estimate the speed of
these waves. The #1 wavefront travels at 880 +40 ms1, which is
about 0.75 times the S-wave speed. The #2 wavefront travels at
1150 +60 ms1, which is slightly slower than the S-wave speed.
The #3 wavefront travels at 1590 +40 ms1, which is about 1.35
times the S-wave speed. The #4 wavefront travels at 2000 +100 ms1,
which is about 1.7 times the S-wave speed.
To discuss the nature of these observed waves, we solved the
Stoneley equation [Eq. (6)] for the interface between water and
epoxy-resin, with cl¼1480 ms1and ρliquid ¼998 kgm3for water
and cs¼1180 ms1,cp¼2550 ms1,andρsolid ¼1100 ms1kg m3
for epoxy-resin. The result gave two real roots (ignore the negative
ones) at c1¼0:779csand c2¼0:996cs. When comparing to our
observation, we found that the #1 wavefront fit root c1.The#2wave-
front fit root c2,asshowninFig. 4(c).
Besides the two real roots, Eq. (6) also gave three complex
conjugates (ignore the negative ones): c3¼(1:36 +0:313i)cs,
c4¼(1:66 +0:296i)cs, and c5¼(2:50 +0:197i)cs. In a complex
conjugate root, the real part reveals propagating speed, and the
imaginary part reveals the attenuation along the propagation direc-
tion of the wave. Among the three above conjugates, we temporar-
ily consider c5as an extraneous root since it results in a wave speed
that is larger than the P-wave speed. We then compared the roots
c3and c4with the observed wavefronts and found that wavefront
#3 fit root c3and wavefront #4 fit root c4[Fig. 4(c)].
Root c1of the Stoneley equation has been well recognized as
the Scholte wave,
2,18
and, thus, we can confirm that the #1 wave-
front observed in our photoelastic images is the Scholte wave.
When the viscosity of the liquid is negligible (like water in our
case), the Scholte wave travels along a solidliquid interface
without attenuation.
4,16
In our photoelasticity images, the Scholte
wave is represented by a group of parallel fringes. Because the wave
was not attenuated, the number of fringes was almost unchanged
with time within the observation period.
It has been well known that for a hard solidliquid configura-
tion, a complex conjugate root of the Stoneley equation gives rise to
a leaky Rayleigh wave. In 1999, Padilla et al. proposed that for a
soft solidliquid interface, this complex root becomes a real root
and indicates the existence of a non-leaky Rayleigh wave on the
solidliquid interface.
18
Our observation agrees with the result of
Padilla et al. that real roots c2of the Stoneley equation correspond
to the speed of a physical wave. We, thus, conclude that the #2
wavefront observed in our images is the Rayleigh wave. Because
this wave is not leaky (corresponding to a real root), it attenuates
weakly along the propagation direction and can still be detected in
our images until several hundred nanoseconds after irradiation.
The #4 wavefront was well fit by root c4and shows the same
properties of a SEW that is recorded at a free boundary. (1) It
arises from a complex root of the wave equation. (2) It travels faster
than the S-wave but slower than the P-wave. (3) It is tangent to the
S-wave front. (4) It attenuates quickly that could not be detected
FIG. 3. Distances traveled by the P-wave, S-wave, and the 1st surface wave
plotted against time, observed for a free surface. The plots present the mean
value measured at each delay time. The average deviation at each delay time is
within 1%. The dashed lines show the linear fit applied to each data set.
Journal of
Applied Physics ARTICLE scitation.org/journal/jap
J. Appl. Phys. 131, 123102 (2022); doi: 10.1063/5.0081237 131, 123102-7
Published under an exclusive license by AIP Publishing
after a short interval. We, thus, propose that the #4 wavefront is
a SEW.
Although the #3 wavefront fit root c3of the wave equation, the
fact that it is much stronger than the #4 wavefront and still can be
detected until 2000 ns after irradiation makes us hesitate to con-
clude that the #3 wavefront corresponds to root c3, which should
give rise to an evanescent or leaky wave that attenuates quickly. We
noticed that the speed of the #3 wavefront is very close to the speed
of the primary shock wave that is induced in the liquid.
50
We, thus,
propose a possibility that the #3 wavefront is the combination of a
lateral wave that is caused by the energy irradiation of the primary
shock wave into the solid and a surface wave that arises from the c3
FIG. 4. (a) Laser-induced shock excitation of elastic surface waves at a soft solidliquid interface. (b) Magnified observation of the #4 wavefront. (c) A comparison
between experimentally determined (symbols) and theoretically calculated (lines) distance traveled by each elastic surface wave. c1,c2,c3, and c4are the roots of the
Stoneley equation. The plots present the mean value measured at each delay time. The average deviation at each delay time is within 3%.
Journal of
Applied Physics ARTICLE scitation.org/journal/jap
J. Appl. Phys. 131, 123102 (2022); doi: 10.1063/5.0081237 131, 123102-8
Published under an exclusive license by AIP Publishing
root. This combination may explain why the #3 wavefront is strong
and attenuates very slow. From the experiments presented in this
paper, we cannot give a solid conclusion about the #3 wavefront
and suggest that more research is needed to confirm its origin.
In short, our observation evidence that a Rayleigh wave and a
SEW can be excited by a laser beam at a soft solidliquid interface.
For a solidliquid interface, it is a widespread belief that only
two typical types of elastic surface waves may exist: Scholte wave
and leaky Rayleigh wave.
48,52
While the Scholte wave always exists
for a solidliquid interface, the leaky Rayleigh wave is considered to
exist only for a hard solidliquid interface.
17
In 1999, Padilla et al.
showed that for a soft solidliquid interface, a Rayleigh wave can
exist.
18
The admissibility of this wave was debated later by Glorieux
et al.,
19
who concluded that a Rayleigh-type wave could only exist
when the sound speed in the liquid is at least 1.3 times smaller
than the shear speed in the solid. In the works of Padilla et al.
18
and Glorieux et al.,
19
plexiglass was used as solid, and water
was used as the liquid medium. Plexiglass has a shear speed of
1430 m/s and sound speed in water can change from 1430 ms
1
at
5 °C to 1480 ms
1
at 20 °C.
54
This selection of interface is within
the unstable region where csapproximate cl, and, thus, the results
can be conflict. In our experiment, the shear speed in the solid is
well under the value of sound speed in the water and, therefore,
can confirm the existence of a non-leaky Rayleigh wave at a soft
solidliquid interface where the liquid sound speed is larger than
the solid shear speed. Because this wave is not leaky, it is difficult
to be detected by methods based on reflection at the interface or
transmission in the liquid phase.
18
This can explain why a Rayleigh
wave was normally not detected in the previous reports.
For a liquidsolid interface, the generation of a wave that
arrives between the S-wave and the P-wave has been theoretically
predicted. In 1961, Phinney described this wave mode as an intrin-
sic leaky vibration of a solid half surface that does not seethe
liquid half-space in a significant wave.
20
This wave may not always
be physically separable on experimental records because of the
close association with the body waves.
20
Moreover, this wave is
hard to observe because the acquisition of them is challenging.
22
In
fact, the records on surface waves that arrive between the P-wave
and the S-wave on a solidliquid interface are very limited. In seis-
mology, there were field reports on an inhomogeneous wave with
the real part of the horizontal slowness lying in the range
1/cP,p,1/cS. These waves were observed in the studies of the
non-geometric P-S wave conversion at water bottom.
54,55
The theo-
retical analysis verified that this particular mode of conversion is
excited only when the acoustic source is close to the interface and
the solid S-wave speed is lower than the liquid sound speed,
26
which is equivalent to a soft solidliquid configuration.
In many experiments of elastic surface waves generated at a
solidliquid interface, the SEW has not been recorded (see, for
example, Refs. 16,18,19,29,44, and 56). In these works, the SEW
can be neglected because of their exponentially decaying amplitude.
It would not separate from the other elastic surface waves until it
had traveled some distance from its sources, at which point the
evanescent nature of it makes measurement difficult.
22
Moreover,
the feasibility of acquiring this type of wave is characterized by
cs,cl,cPand a high ratio of cp/cs. Thus, a proper selection of
liquidsolid pairs is essential to detect this wave. To the best of our
knowledge, our result verifies for the first time the existence of the
SEW excited by a laser beam on a soft solidliquid interface, thus
correcting the previously accepted notion that the Scholte wave is
the only surface wave that can be excited at such a boundary. We
propose that the dynamical feature of SEWs on a solidliquid inter-
face should be investigated more because it has the potential to
infer S-wave information and would be useful for applications in
seismology, engineering, and nondestructive testing.
In this research, we used Rayleigh and Stoneley equations for
a plane interface problem to compute the speed of surface distur-
bances induced by a laser-induced shock impulse. This approach
may not fully account for all the interface waves which propagate in
the near field of the excitation source. Future efforts should follow
a more efficient approach using Greens function to define all wave
features propagating at the speed between S-wave and P-wave.
CONCLUSIONS
We investigated the laser-induced shock excitation of elastic
surface waves at a free surface and a soft solidliquid interface
using a custom-designed photoelasticity imaging technique.
For the free surface, we provided a whole-field image of the
super-shear evanescent wave and the Rayleigh waves excited by a
laser-induced shock. Such direct and whole-field visualization has
not been presented in the literature, thus demonstrating photoelas-
ticity imaging technique as a unique and potential tool for analyz-
ing elastic surface waves.
We differentiate three elastic surface waves that are excited by
the laser-induced shock at the soft solidliquid interface. (1) The
Scholte wave travels without attenuation along the interface at a
speed equal to 0.75 times the solid shear speed. (2) The Rayleigh
wave propagates slightly slower than the shear wave and attenuates
weakly along the propagation direction. It corresponds to a real
root of the Stoneley equation and, thus, is not in a leaky mode.
(3) The super-shear evanescent wave travels 1.7 times faster than the
solid shear speed and couples into the shear wave plane. This wave
correlates to a complex conjugate root of the Stoneley equation.
It attenuates quickly and can only be observed in a brief interval.
Our result confirmed that there is a non-leaky Rayleigh wave at a
soft solidliquid interface, the existence of which has been debated in
the literature. We also provide the first direct evidence of a super-
shear evanescent wave excited by a pulsed laser at a solidliquid
interface. This work gives new insight into the elastic surface waves
generated at a soft solidliquid interface by a laser-induced shock,
suggesting that more theoretical and experimental studies are needed
to clarify the dynamical features of these phenomena.
ACKNOWLEDGMENTS
The authors wish to acknowledge Dr. Nguyen Van Yen and
Dr. Tina Mai from Duy Tan University for their inspiring discus-
sions and valuable comments.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Journal of
Applied Physics ARTICLE scitation.org/journal/jap
J. Appl. Phys. 131, 123102 (2022); doi: 10.1063/5.0081237 131, 123102-9
Published under an exclusive license by AIP Publishing
Ethics Approval
Ethics approval is not required.
DATA AVAILABILITY
The data that support the findings of this study are available
from the corresponding author upon reasonable request.
REFERENCES
1
P. A. Doyle, J. Phys. D: Appl. Phys. 19, 1613 (1986).
2
V. Gusev, C. Desmet, W. Lauriks, C. Glorieux, and J. Thoen, J. Acoust. Soc.
Am. 100, 1514 (1996).
3
P. Hess, Phys. Today 55,4247 (2002).
4
F. Jenot, M. Ouaftouh, M. Duquennoy, and M. Ourak, J. Appl. Phys. 97,
094905 (2005).
5
M. Ochiai, J. Phys. Conf. Ser. 278, 012009 (2011).
6
F. Mirzade, J. Appl. Phys. 110, 064906 (2011).
7
D. Schneider, in IEEE International Ultrasonics Symposium (IUS) (IEEE, 2012),
p. 269.
8
H. Zhang, A. Antoncecchi, S. Edward, P. Planken, and S. Witte, Phys. Rev. B
104, 205416 (2021).
9
G. Chow, E. Uchaker, G. Cao, and J. Wang, Mech. Mater. 91, 333 (2015).
10
S. Das, A. Schill, C.-H. Liu, S. Aglyamov, and K. V. Larin, J. Biomed. Opt. 25,
035004 (2020).
11
P. Grasland-Mongrain, Y. Lu, F. Lesage, S. Catheline, and G. Cloutier, Appl.
Phys. Lett. 109, 221901 (2016).
12
C. Desmet, V. Gusev, W. Lauriks, C. Glorieux, and J. Thoen, Appl. Phys. Lett.
68, 2939 (1996).
13
N. Hosoya, A. Yoshinaga, A. Kanda, and I. Kajiwara, Int. J. Mech. Sci. 140,
486 (2018).
14
C. T. Schröder and W. R. Scott, J. Acoust. Soc. Am. 110, 2867 (2001).
15
M. K. Teshaev, I. I. Safarov, N. U. Kuldashov, M. R. Ishmamatov, and
T. R. Ruziev, J. Vib. Eng. Technol. 8, 579 (2020).
16
Q. Han, M. Qian, and H. Wang, J. Appl. Phys. 100, 093101 (2006).
17
J. Zhu, J. S. Popovics, and F. Schubert, J. Acoust. Soc. Am. 116, 2101 (2004).
18
F. Padilla, M. de Billy, and G. Quentin, J. Acoust. Soc. Am. 106, 666 (1999).
19
C. Glorieux, K. Van de Rostyne, K. Nelson, W. Gao, W. Lauriks, and J. Thoen,
J. Acoust. Soc. Am. 110, 1299 (2001).
20
R. A. Phinney, Bull. Seismol. Soc. Am. 51, 527 (1961).
21
A. Le Goff, P. Cobelli, and G. Lagubeau, Phys. Rev. Lett. 110, 1 (2013).
22
J. J. Pitre, M. A. Kirby, L. Gao, D. S. Li, T. Shen, R. K. Wang, M. ODonnell,
and I. Pelivanov, Appl. Phys. Lett. 115, 083701 (2019).
23
L. Gao, J. Xia, and Y. Pan, Geophys. J. Int. 199, 1452 (2014).
24
M. Roth and K. Holliger, Geophys. J. Int. 140, F5 (2000).
25
S. Das, Supershear earthquake rupturesTheory, methods, laboratory experi-
ments and fault superhighways: An update,in Perspectives on European
Earthquake Engineering and Seismology, Geotechnical, Geological and
Earthquake Engineering Vol. 39, edited by A. Ansal (Springer, Cham, 2015), pp.
120.
26
N. Allouche and G. Drijkoningen, Geophys. Prospect. 64, 543 (2016).
27
Z. Cai, S. Liu, and C. Zhang, AIP Conf. Proc. 1806, 050015 (2017).
28
V. Gusev and P. Hess, Nondestruct. Test. Eval. 11, 313 (1994).
29
C. Desmet, V. Gusev, W. Lauriks, C. Glorieux, and J. Thoen, Opt. Lett. 22,69
(1997).
30
H. Coufal, K. Meyer, R. K. Grygier, P. Hess, and A. Neubrand, J. Acoust. Soc.
Am. 95, 1158 (1994).
31
A. Neubrand and P. Hess, J. Appl. Phys. 71, 227 (1992).
32
A. Dräbenstedt, AIP Conf. Proc. 1600, 263 (2014).
33
B. Pouet, A. Wartelle, and S. Breugnot, Nondestruct. Test. Eval. 26, 253
(2011).
34
J. Xiong, X. Xu, C. Glorieux, O. Matsuda, and L. Cheng, Rev. Sci. Instrum. 86,
053107 (2015).
35
A. A. Kamshilin, R. V. Romashko, and Y. N. Kulchin, J. Appl. Phys. 105,
031101 (2009).
36
T. A. Carolan, R. L. Reuben, J. S. Barton, and J. D. C. Jones, Appl. Opt. 36,
380 (1997).
37
I. Pelivanov, T. Buma, J. Xia, C.-W. Wei, and M. ODonnell, J. Appl. Phys.
115, 113105 (2014).
38
I. Pelivanov, T. Buma, J. Xia, C.-W. Wei, and M. ODonnell, Photoacoustics 2,
63 (2014).
39
J. Spytek, A. Ziaja-Sujdak, K. Dziedziech, L. Pieczonka, I. Pelivanov, and
L. Ambrozinski, NDT E Int. 112, 102249 (2020).
40
P. Pyzik, A. Ziaja-Sujdak, J. Spytek, M. ODonnell, I. Pelivanov, and
L. Ambrozinski, Photoacoustics 21, 100226 (2021).
41
T. T. P. Nguyen, R. Tanabe, and Y. Ito, Appl. Phys. Lett. 102, 124103
(2013).
42
T. T. P. Nguyen, R. Tanabe-Yamagishi, and Y. Ito, Appl. Surf. Sci. 470, 250
(2019).
43
T. T. P. Nguyen, R. Tanabe-Yamagishi, and Y. Ito, Appl. Phys. Express 14,
042005 (2021).
44
Y. Zhang, C. Yang, H. Qiang, and P. Zhong, Phys. Rev. Res. 1, 33068 (2019).
45
J. H. Ansell, Pure Appl. Geophys. 94, 172 (1972).
46
M. Hayes and R. S. Rivlin, Z. Angew. Math. Phys. 13, 80 (1962).
47
L. Tsang, J. Acoust. Soc. Am. 63, 1302 (1978).
48
V. G. Mozhaev and M. Weihnacht, Ultrasonics 40, 927 (2002).
49
T. T. P. Nguyen, R. Tanabe, and Y. Ito, Appl. Phys. A 116, 1109 (2014).
50
T. T. P. Nguyen, R. Tanabe, and Y. Ito, Opt. Laser Technol. 100, 21 (2018).
51
J. A. Martin, D. G. Schmitz, A. C. Ehlers, M. S. Allen, and D. G. Thelen,
J. Biomech. 90, 9 (2019).
52
K. Shukla, J. M. Carcione, J. S. Hesthaven, and E. Lheureux, J. Acoust. Soc.
Am. 147, 3136 (2020).
53
J. Lubbers and R. Graaff, Ultrasound Med. Biol. 24, 1065 (1998).
54
J. Ivanov, C. B. Park, R. D. Miller, J. Xia, J. A. Hunter, R. L. Good, and
R. A. Burns, SEG Technical Program Expanded Abstracts 2000 (Society of
Exploration Geophysicists, 2000), Abstract No. 1307-1310.
55
N. Allouche, G. G. Drijkoningen, W. Versteeg, and R. Ghose, Geophysics 76,
T1 (2011).
56
C. Bakre, P. Rajagopal, and K. Balasubramaniam, AIP Conf. Proc. 1806,
020004 (2017).
Journal of
Applied Physics ARTICLE scitation.org/journal/jap
J. Appl. Phys. 131, 123102 (2022); doi: 10.1063/5.0081237 131, 123102-10
Published under an exclusive license by AIP Publishing
... where c ij is the elastic constants of KDP crystals. The Rayleigh wave (R wave) expression is shown in Eq. (10) [41,42]: ...
... In addition, the research on the speed of waves by Nguyen et al. [42] showed that the angle θ between the material surface and the tangent line of the S wave presents a fixed value. Thus, there is the following relationship between the speed of the H wave c h and S wave c s : ...
Article
Full-text available
The residual crack defects on the surface of potassium dihydrogen phosphate (KDP) crystals are the bottleneck that limits the improvement of laser damage resistance in the application of high-power laser devices. The multiple stress waves introduced by these residual surface lateral cracks on crystals under laser irradiation are the main inducement for damage extension and reduction of laser damage resistance. However, the coupling of these stress waves complicates their propagation in the crystal, and the interaction mechanism between each stress wave and laser damage has not been quantitatively characterized. Herein, a laser damage dynamic model for surface lateral cracks is constructed to reproduce the dynamic behaviors of the evolution of micro-defects to sub-millimeter damage pits under laser irradiation. Combined with the time-resolved pump and probe technique, the distribution of stress waves induced by lateral cracks was detected in situ to determine the type of stress waves. Then, the initiation and extension of laser damage were analyzed quantitatively to establish the correlations between different stress waves and damage extension. It is found that the longitudinal, shear, and Rayleigh waves induced by lateral cracks lead to large crush zones on the surface of KDP crystals, as well as butterfly-like damage sites accompanied by a large number of cracks at the bottom in the longitudinal section. The scale of the damage site can reach up to approximately 150 µm for lateral crack defects with large surface widths. This study ultimately reveals the physical mechanism of damage evolution induced by lateral cracks, providing effective guidance for developing control standards of surface crack defects during optical ultra-precision machining processes. This is of great significance for the improvement of laser damage resistance of KDP crystals in high-power laser systems.
... Several previous papers have used this surface wave-based OCE technique to evaluate the biomechanical properties of phantoms, ex vivo human corneas, and human skin tissue, demonstrating the feasibility of this approach. [34][35][36] In the present, the supershear Rayleigh waves (SRWs) are proposed to characterize the behavior of elastic waves in brain tissue. The theoretical derivation of the SRW is developed and demonstrates that the propagation direction is along the brain tissue surface and that the SRW velocity is considerably faster than the bulk transverse wave velocity. ...
Article
Full-text available
Recently, supershear Rayleigh waves (SRWs) have been proposed to characterize the biomechanical properties of soft tissues. The SRWs propagate along the surface of the medium, unlike surface Rayleigh waves, SRWs propagate faster than bulk shear waves. However, their behavior and application in biological tissues is still elusive. In brain tissue elastography, shear waves combined with magnetic resonance elastography or ultrasound elastography are generally used to quantify the shear modulus, but high spatial resolution elasticity assessment in 10 μm scale is still improving. Here, we develop an air-coupled ultrasonic transducer for noncontact excitation of SRWs and Rayleigh waves in brain tissue, use optical coherent elastography (OCE) to detect, and reconstruct the SRW propagation process; in combing with a derived theoretical model of SRWs on a free boundary surface, we quantify the shear modulus of brain tissue with high spatial resolution. We first complete validation experiments using a homogeneous isotropic agar phantom, and the experimental results clearly show the SRW is 1.9649 times faster than the bulk shear waves. Furthermore, the propagation velocity of SRWs in both the frontal and parietal lobe regions of the brain is all 1.87 times faster than the bulk shear wave velocity. Finally, we evaluated the anisotropy in different brain regions, and the medulla oblongata region had the highest anisotropy index. Our study shows that the OCE system using the SRW model is a new potential approach for high-resolution assessment of the biomechanical properties of brain tissue.
... The shock behavior of materials must be carefully characterized and anticipated for practical applications in various fields, including defense (armor design, ballistics, etc.), safety issues (shock protection, debris shielding, shock mitigation, etc.), and engineering processes (improved fatigue life after shock peening, adherence testing, explosive welding, etc.). Such industrial needs motivate both experimental and theoretical research work dedicated, for example, to better understanding laser-matter interactions, 57 to optimize laser shock peening of metallic alloys, 58,59 to design graded density impactors in order to shape pressure loading, 60 to assess damage tolerance through consecutive loading tests, 61 to investigate pore collapse and subsequent shock mitigation in foams, 62 to model drop impacts in the context of cold spray, 63 friction at compressed metal-metal interfaces, 64 or shock-induced bonding in processes like explosive welding. 65 ...
... This so-called supershear (SS) surface wave (or supershear evanescent wave, SEW, see for example Pitre et al., 2019) allows the surface elastic energy to propagate away with a greater speed and is likely responsible for the supershear dynamics at the surface, such as the supershear crack propagation and supershear earthquake (Bhat et al., 2007;Das, 2007;Passelègue et al., 2013;Socquet et al., 2019). Besides the free surface, a very recent study by Nguyen et al. (2022) suggest the SEW can also be supported at the soft solid-fluid interface. ...
Article
Surface waves play important roles in many fundamental and applied areas from seismic detection to material characterizations. Supershear surface waves with propagation speeds greater than bulk shear waves have recently been reported, but their properties are not well understood. Here we describe theoretical and experimental results on supershear surface waves in rubbery materials. We find that supershear surface waves can be supported in viscoelastic materials with no restriction on the shear quality factor. Interestingly, the effect of prestress on the speed of the supershear surface wave is opposite to that of the Rayleigh surface wave. Furthermore, anisotropy of material affects the supershear wave much more strongly than the Rayleigh surface wave. We offer heuristic interpretation as well as theoretical verification of our experimental observations. Our work points to the potential applications of supershear waves for characterizing the bulk mechanical properties of soft solid from the free surface.
Article
Full-text available
Light-induced acoustic waves can be used as sensitive probes, providing a pathway toward microscopic imaging and metrology in optically inaccessible media. The ability to detect such waves depends on the interaction of an optical probe pulse with the acoustic waves in the topmost layers of the structure. Therefore, the interplay between optoacoustic coupling and material boundaries, combined with the properties of acoustic waves near free surfaces is of prime importance. Here we show an approach toward optimized optical detection of such laser-excited acoustic waves. We explore the physics underlying this detection, finding that the presence of a free surface actually reduces the optoacoustic interaction, and subsequently enhancing this interaction via adding transparent nanolayers on the free surface. Our work uncovers an important yet rarely explored aspect in optical detection of strain waves via free surfaces and may lead to strategies for signal enhancement in the imaging and characterization of subsurface structures using laser-excited strain waves.
Article
Full-text available
We studied the dynamics of nanosecond-pulsed laser ablation of graphite-coated and black-paint-coated targets in liquids using a custom-designed time-resolved photoelasticity imaging technique. We presented the first demonstration of a planar head wave that was almost parallel to the target surface. In the solid, we observed a planar stress wave that was a counterpart of the planar head wave. This planar stress wave distorted the typical stress distribution induced by pulsed laser ablation in liquid. The planar head wave and stress wave traveled at the acoustic speed in the corresponding medium. These wavefronts were stronger as the number of shots increased.
Article
Full-text available
Adhesively bonded metals are increasingly used in many industries. Inspecting these parts remains challenging for modern non-destructive testing techniques. Laser ultrasound (LU) has shown great potential in high-resolution imaging of carbon-reinforced composites. For metals, excitation of longitudinal waves is inefficient without surface ablation. However, shear waves can be efficiently generated in the thermo-elastic regime and used to image defects in metallic structures. Here we present a compact LU system consisting of a high repetition rate diode-pumped laser to excite shear waves and noncontact detection with a highly sensitive fiber optic Sagnac interferometer to inspect adhesively bonded aluminum plates. Multiphysics finite difference simulations are performed to optimize the measurement configuration. Damage detection is performed for a structure consisting of three aluminum plates bonded with an epoxy film. Defects are simulated by a thin Teflon film. It is shown that the proposed technique can efficiently localize defects in both adhesion layers.
Article
Full-text available
Significance: Shear wave optical coherence elastography is an emerging technique for characterizing tissue biomechanics that relies on the generation of elastic waves to obtain the mechanical contrast. Various techniques, such as contact, acoustic, and pneumatic methods, have been used to induce elastic waves. However, the lack of higher-frequency components within the elastic wave restricts their use in thin samples. The methods also require moving parts and/or tubing, which therefore limits the extent to which they can be miniaturized. Aim: To overcome these limitations, we propose an all-optical approach using photothermal excitation. Depending on the absorption coefficient of the sample and the laser pulse energy, elastic waves are generated either through a thermoelastic or an ablative process. Our study aimed to experimentally determine the boundary between the thermoelastic and the ablative regimes for safe all-optical elastography applications. Approach: Tissue-mimicking graphite-doped phantoms and chicken liver samples were used to investigate the boundary between thermoelastic and ablative regimes. A pulsed laser at 532 nm was used to induce elastic waves in the samples. Laser-induced elastic waves were detected using a line field low coherence holography instrument. The shape of the elastic wave amplitude was analyzed and used to determine the transition point between thermoelastic and ablative regimes. Results: The transition from the thermoelastic to the ablative regime is accompanied by the nonlinear increase in surface wave amplitude as well as the transformation of the wave shape. Correlation between the absorption coefficient and the transition point energy was experimentally determined using graphite-doped phantoms and applied to biological samples ex vivo. Conclusions: Our study described a methodology for determining the boundary region between thermoelastic and ablative regimes of elastic wave generation. These can be used for the development of a safe method for completely noncontact, all-optical microscale assessment of tissue biomechanics using laser-induced elastic waves.
Article
Full-text available
We investigate the generation and propagation characteristics of leaky Rayleigh waves (LRWs) caused by a spherical shock wave incident on a water-glass boundary both experimentally and numerically. The maximum tensile stress produced on the solid boundary is attributed to the dynamic interaction between the LRWs and an evanescent wave generated concomitantly along the boundary. The resultant tensile stress field drives the initiation of crack formation from pre-existing surface flaws and their subsequent extension along a circular trajectory, confirmative with the direction of the principal stress on the boundary. We further demonstrate that this unique ringlike fracture pattern, prevalent in damage produced by high-speed impact, can be best described by the Tuler-Butcher criterion for dynamic failure in brittle materials. The orientation of the ring fracture extension into the solid also follows closely with the trajectory of the local maximum tensile stress distribution.
Article
Full-text available
Background In this article, to the development of the theory and methods for calculating vibrations of dissipative mechanical systems consisting of solids, which include both deformable and non-deformable bodies is discussed. Theoretically, the problem in the mathematical aspect, there are not enough developed solution methods and algorithms for dissipative mechanical systems has not yet been posed. The problems of choosing a nucleus and its rheological parameters, their influence on the frequency and damping coefficient systems have not been studied Purpose The goal of the work is to formulate the statement, develop the solution methods and the algorithm for studying the problems of the dynamics of dissipative mechanical systems consisting of thin-walled plates (or shells) with attached masses and point development theory. Methods A method and algorithm for solving problems of eigen and forced vibrations of dissipative mechanical systems consisting of rigid and deformable bodies, based on the methods of Muller, Gauss, Laplace and Runge integral transform, are developed. Results To describe the dissipative properties of the system as a whole, the concept of a global damping coefficient (GDC) is introduced. In the case of a dissipatively homogeneous system of the GDC, it is determined by the imaginary part of the first modulo complex natural frequency. In the role of GDC in the case of a dissipatively inhomogeneous system, are the imaginary parts of both the first and second frequencies. Moreover, the “Change of Roles” occurs with the characteristic value of the stiffness coefficient of the deformable elements; the real parts of the first and second frequencies are closest. At the indicated characteristic value of the deformable elements, the global damping coefficient has a pronounced maximum. A change in the parameter, on which the global damping coefficient so substantially depends, can be achieved by varying physical properties or geometrical dimensions, thereby opening up the promising possibility of effectively controlling the damping properties of dissipative-inhomogeneous mechanical systems. Conclusions The developed solution methods, algorithms and programs allow determining the dissipative properties of a mechanical system depending on various physic-mechanical parameters, geometrical dimensions and boundary conditions. A method for estimating the dissipative properties of the system as a whole (with forced vibrations) depending on the instantaneous values of the deformable elements (shock absorbers) has been developed. The developed method allows to reduce (several times) the amplitudes of displacements and stresses.
Article
An accurate solution of the wave equation at a fluid-solid interface requires a correct implementation of the boundary condition. Boundary conditions at fluid-solid interface require continuity of the normal component of particle velocity and traction, whereas the tangential components vanish. A main challenge is to model interface waves, namely, the Scholte and leaky Rayleigh waves. This study uses a nodal discontinuous Galerkin (dG) finite-element method with the medium discretized using an unstructured uniform triangular meshes. The natural boundary conditions in the dG method are implemented by (1) using an explicit upwind numerical flux and (2) by using an implicit penalty flux and setting the modulus of rigidity of the acoustic medium to zero. The accuracy of these methods is evaluated by comparing the numerical solutions with analytical ones, with source and receiver at and away from the interface. The study shows that the solutions obtained from the explicit and implicit boundary conditions provide the correct results. The stability of the dG scheme is determined by the numerical flux, which also implements the boundary conditions by unifying the numerical solution at shared edges of the elements in an energy stable manner.
Article
Inspection of adhesively bonded metallic plates, commonly used in aircraft structures, remains challenging for modern non-destructive testing (NDT) techniques. When a probing ultrasound (US) wave interacts with the plate boundaries, it produces multiple propagating guided waves (or Lamb modes). Analysis of these waves can be complicated due to a strong geometrical dispersion. However, recent studies showed that for specific frequencies zero-group velocity (ZGV) modes exist. Any changes in a bounded structure, like delaminations, drastically alter the conditions for zero-group velocity waves, thus making this method highly sensitive for NDT applications. Laser ultrasound (LU) provides a very broad bandwidth of the generated waves, thus it is a feasible tool for a spectroscopy-based investigation of the ZGV modes. In this paper, we demonstrate an application of high-repetition rate LU with a non-contact fiber-optic Sagnac interferometer on receive for high resolution imaging of delaminations in a structure consisting of 3 epoxy-bonded aluminum plates. The investigation is supported by numerical analysis of Lamb waves existing in the structure to determine ZGV modes sensitive to delaminations at particular bonding interfaces. Tracking the selected modes permits imaging and identification of defects. We also show that mean frequency estimation of the ZGV modes can improve the contrast-to-noise ratio compared to an amplitude of the single ZGV frequency.
Article
We describe surface wave propagation in soft elastic media at speeds exceeding the bulk shear wave speed. By linking these waves to the elastodynamic Green's function, we derive a simple relationship to quantify the elasticity of a soft medium from the speed of this supershear evanescent wave (SEW). We experimentally probe SEW propagation in tissue-mimicking phantoms, human cornea ex vivo, and skin in vivo using a high-speed optical coherence elastography system. Measurements confirm the predicted relationship between SEW and bulk shear wave speeds, agreeing well with both theoretical and numerical models. These results suggest that SEW measurements may be a robust method to quantify elasticity in soft media, particularly in complex, bounded materials where dispersive Rayleigh-Lamb modes complicate measurements.
Article
It has recently been shown that shear wave speed in tendons is directly dependent on axial stress. Hence, wave speed could be used to infer tendon load provided that the wave speed-stress relationship can be calibrated and remains robust across loading conditions. The purpose of this study was to investigate the effects of loading rate and fluid immersion on the wave speed-stress relationship in ex vivo tendons, and to assess potential calibration techniques. Tendon wave speed and axial stress were measured in 20 porcine digital flexor tendons during cyclic (0.5, 1.0 and 2.0 Hz)or static axial loading. Squared wave speed was highly correlated to stress (r ²avg = 0.98)and was insensitive to loading rate (p = 0.57). The constant of proportionality is the effective density, which reflects the density of the tendon tissue and additional effective mass added by the adjacent fluid. Effective densities of tendons vibrating in a saline bath averaged 1680 kg/m ³ and added mass effects caused wave speeds to be 22% lower on average in a saline bath than in air. The root-mean-square error between predicted and measured stress was 0.67 MPa (6.7% of maximum stress)when using tendon-specific calibration parameters. These errors increased to 1.31 MPa (13.1% of maximum stress)when calibrating based on group-compiled data from ten tendons. These results support the feasibility of calculating absolute tendon stresses from wave speed squared based on linear calibration relationships.