scispace - formally typeset
Search or ask a question
Journal ArticleDOI

Review of Second Harmonic Generation Measurement Techniques for Material State Determination in Metals

TL;DR: In this article, the authors present a comprehensive review of the current state of knowledge of second harmonic generation (SHG) measurements, a subset of nonlinear ultrasonic non-destructive evaluation techniques.
Abstract: This paper presents a comprehensive review of the current state of knowledge of second harmonic generation (SHG) measurements, a subset of nonlinear ultrasonic nondestructive evaluation techniques. These SHG techniques exploit the material nonlinearity of metals in order to measure the acoustic nonlinearity parameter, $$\beta $$ . In these measurements, a second harmonic wave is generated from a propagating monochromatic elastic wave, due to the anharmonicity of the crystal lattice, as well as the presence of microstructural features such as dislocations and precipitates. This article provides a summary of models that relate the different microstructural contributions to $$\beta $$ , and provides details of the different SHG measurement and analysis techniques available, focusing on longitudinal and Rayleigh wave methods. The main focus of this paper is a critical review of the literature that utilizes these SHG methods for the nondestructive evaluation of plasticity, fatigue, thermal aging, creep, and radiation damage in metals.

Summary (5 min read)

1 Introduction

  • Nonlinear ultrasonic methods have the powerful ability to characterize microstructural features in materials.
  • Then, the different SHG measurement techniques reported throughout the literature are reviewed, followed by a discussion on how these measurements have been applied to monitor microstructural evolution during material damage.
  • The same general form of A2 has been shown for Rayleigh waves, in terms of dependence on propagation distance, wavenumber, and first harmonic wave.
  • A more accurate expression for the acoustic nonlinearity parameter can be found by accounting for the on-axis diffraction effects of both the first and second harmonic wave.
  • This strain arises from microstructural features such as dislocations and precipitates, and dislocation contributions to β can greatly exceed that of the lattice anharmonicity.

2.3.1 Dislocation Pinning: Hikata et al. Model

  • The dislocation motion contribution to acoustic nonlinearity was first developed by Suzuki et al. [53] and expanded on by Hikata et al. [4], and this model has been used to interpret a multitude of experimental results of second harmonic generation.
  • Note that this stress can be thought of as either an internal residual stress or externally applied stress, but it is small enough such that the dislocation segment does not break away from the pinning points.
  • Assuming the dislocation density is small enough such that bowed dislocations act independently of each other, the line tension, T , of this dislocation segment due to the applied stress was approximated as constant and independent of orientation as T = μb2/2, where b is the Burgers vector and μ is the shear modulus.
  • Further details of this model can be found elsewhere [4,16].
  • It should be specifically noted that the internal stress σ1 in this analysis, as well as the superimposed ultrasonic stress, is assumed to be much smaller than the yield stress of the material, such that dislocation displacement is small.

2.3.2 Dislocation Pinning: Extensions of the Hikata et al. Model

  • Later work by Cash and Cai [55] extended the Hikata et al. model to account for orientation-dependent line energy in the analytical model and verified with dislocation dynamics simulations.
  • At values of Poisson’s ratio greater than about 0.2 (i.e. most metals), the dislocation dynamics simulations and developed analytical model show that relationship between β and the applied stress is not in fact linear, and simulations show that β can even be negative for an edge dislocation at small stresses.
  • If phase information were extracted from the experiments, a negative β could be measured.
  • They validate their model with molecular dynamic simulations.
  • The authors present a procedure to determine S̄′′(τn) without needing to perform the numerical derivatives, and the reader is referred to the paper for more details [59].

2.3.3 Dislocation Dipoles

  • Dislocation dipoles are formed when two dislocations of opposite sign move within some small distance d of each other and become mutually trapped.
  • The force–displacement relation of the dipole is a nonlinear relation, and as such it has been shown that when perturbed by an ultrasonic wave, this feature generates a nonzero component of acoustic nonlinearity [16,45,46,60,61].
  • In fatigue damage, increased cyclic loading causes dislocation substructures to form.
  • The vein structures can further transform into a stable persistent slip band structure (PSBs) that is a ladder-type configuration of vein regions.
  • In planar slip metals, for example the IN100 nickel superalloy studied in [63] (and the references therein), the primary dislocation substructures are planar slip bands and intermittently activated persistent Luders bands (PLBs).

2.3.4 Precipitates

  • Effectively, the precipitate embedded in a surrounding matrix creates a local stress field which is then used as the applied stress on a pinned dislocation segment, as given in Eq. (19).
  • Assuming the precipitate and matrix have different elastic properties, the stress in the matrix at radius r due to this embedded precipitate can then be written as [49,67]: σrr (r) = −4μδ [ 3Bp 3Bp + 4μ ] r3p r3 (24) where Bp is the bulk modulus of the precipitate.
  • Since a dislocation line is assumed to follow a contour of minimum energy, it is assumed that two precipitates a distance L/2 away from each dislocation segment act on the dislocation segment, and contributions from other nearby precipitates are negligible.
  • Note that these results are the same as derived in [49], and based off of other previous work as well [46–48].

2.3.5 Microcracks

  • Higher harmonic generation from crack contacting surfaces as a function of stress was studied by Hirose and Achenbach [68] with the boundary element method.
  • They assumed that the contact of asperities on the crack faces is elastic and the overall nonlinearity originates from the nonlinear stress–strain relationship of the asperity contact.
  • Moreover, the asperity heights are assumed to follow an exponential distribution.
  • Note that Eq. (29) is the famous formula of Greenwood and Williamson [70].

3 Experimental Techniques

  • Second harmonic generation measurements of the acoustic nonlinearity parameter can be conducted using multiple wave types, different generation and detection methods, and a variety of experimental set-ups.
  • This section gives a detailed overview of current experimental methods, measurement techniques, set-ups, and post-processing used for both longitudinal and Rayleigh wave SHG measurements of β.
  • An absolute measure of β is possible with longitudinal waves using either capacitive transducers [12,73] or contact piezoelectric transducers using a calibration procedure [92,93] in which the absolute displacement amplitude of the first and second harmonic waves can be measured.
  • In the simplification of a relative measure of nonlinearity, the voltage amplitudes of the first and second harmonic are instead measured and the relative acoustic nonlinearity parameter is calculated, which is defined as: β ′ = A v 2 (Av1) 2 8 xκ2 (31) where recall that κ is the wavenumber as defined in Eq. (5).

3.1.1 Piezoelectric Methods: Relative Amplitude Measurements

  • A longitudinal wave relative measurement of β (meaning a measure of only the voltage amplitudes of A1 and A2) has been widely utilized throughout the recent literature [49,58,74–76,87,96–101].
  • Transducers must be bonded (or coupled using a liquid coupling agent) to the sample surface which can introduce measurement variation if conditions are not accounted for or not kept consistent [96].
  • While this receiving transducer will simultaneously detect the first and second harmonic amplitude, tuning it to the second harmonic frequency is crucial since that amplitude will be a few orders of magnitude smaller than the first harmonic wave amplitude.
  • The exact values of input amplitude for these experiments can be crucial.
  • Above this stress amplitude, β has shown to be relatively constant, unless extremely high stress amplitudes are excited such that dislocations break away from pinning points.

3.1.2 Piezoelectric Methods: Absolute Amplitude Measurements

  • Absolute measurements ofβ, through measuring the absolute displacement of the fundamental and second harmonic amplitudes of the received wave, are possible with piezoelectric contact transducers through a reciprocity-based calibration procedure.
  • Then, a measured current signal can be directly converted to the displacement in the following way: A(ω) = H(ω)I ′(ω) (32) where A(ω) is the displacement amplitude and I ′(ω) is the current signal measured during the nonlinear measurement.
  • After the calibration, the transmitting transducer is attached to the opposite side of the sample and aligned with the receiving transducer.
  • The transmitting transducer is excited with a tone burst signal at the fundamental frequency, and the receiving transducer is used to measure the output current I ′(ω).

3.1.3 Capacitive Methods

  • Absolute measurements of β first became possible through the development of the capacitive receiver for SHG measurements by Gauster and Breazeale [11].
  • The wave propagates through sample, and the capacitive transducer is held a small distance (about 1–10 µm) away from the opposite side of the sample.
  • Note that the factor of 2 in the denominator of Eq. (34) is to account for the doubling of the displacement Fig. 5 Experimental schematic for SHG calibration and absolute measurement of β using piezoelectric transducers (adapted from [33] with permission from Springer) Fig. 6 Schematic of experimental setup for capacitive dection for SHG measurements.
  • While this method offers a direct way of measuring absolute displacements of the first and second harmonic waves compared to piezoelectric transducers (which require a series of calibrations for these absolute measurements [92]) sample preparation is cumbersome, requiring an optically flat and parallel sample surface over the receiver area and a small gap of only a few microns.

3.1.4 Laser Methods

  • Laser ultrasonic SHG methods have the unique advantage of being a noncontact measurement, with the capability of detecting an absolute, point-like displacement measurement.
  • As such, measurements utilizing Rayleigh waves have received considerable attention throughout the literature—for example using a variety of wedge-contact generation and/or reception [38,42,43,79, 80,115–118], laser interferometer detection [38,81,115], air coupled detection [64], comb transducer generation and detection [39], and electromagnetic acoustic transducer (EMAT) detection [119,120].
  • Fitting models based on a least-squares fit have been applied to nonlinear Rayleigh wave measurements to more accurately extract β ′ by directly incorporating attenuation and diffraction [39].
  • Then, the source transducer is removed completely and remounted to repeat the entire measurement set multiple times to achieve statistically significant data.
  • The following sections discuss in detail some of these configurations.

3.2.1 Wedge-Contact Methods

  • Wedge-contact methods utilize an intermediary layer between the transducer and the sample to satisfy phase matching conditions.
  • The wedge material is typically made of acrylic or other plastic material, and is designed to mount the transducer at the angle required to excite the Rayleigh surface wave in the sample.
  • Acoustic coupling is necessary between both the transducer and wedge, as well as between the wedge and sample.
  • Consistent clamping force and uniform contact are difficult to secure and this is the major source of scatter in this method.

3.2.2 Air Coupled Detection

  • Using air-coupled transducers for ultrasonic measurements is not new, e.g. [121–125], but only recently have they been applied to second harmonic generation measurements [41,64].
  • The air-coupled transducer detects a longitudinal wave in air that is leaked from the propagating Rayleigh wave in the sample.
  • 3 SHG Measurement Variations and Corrections Variations in SHG measurements can arise from multiple sources [102].
  • For longitudinal waves, losses such as attenuation must be accounted for if they are non-negligible in the material and for the frequencies and propagation distances measured.
  • In relative measurements, this is needed for example if different transducers are used in different measurements—the frequency response of transducers will bias the first and second harmonic differently.

4 Applications

  • There has been significant work in the past few decades aimed at using second harmonic generation as an NDE technique for early damage detection, by relating the acoustic nonlinearity parameter to different microstructural features.
  • Numerous experiments have shown that β increases with increasing number of fatigue cycles and increasing cumulative plastic strain e.g. [62,74,86,134].
  • The key aspect to interpreting these dislocation-based SHG measurements, as shown particularly in the numerous SHG measurements over fatigue damage, is that specific dislocation substructures and evolution in the material system measured is crucial for accurate SHG measurement evaluation.
  • Results consistently showed a drastic decrease in β followed by a gradual increase, over different aging temperatures, which the authors correlated to changes in dislocation density and second phase distributions.
  • Baby et al. [97] measured the acoustic nonlinearity parameter in a titanium alloy subjected to creep damage, which produced an increase of volume fraction of voids in the microstructure.

5 Conclusions and Future Outlook

  • Second harmonic generation measurement techniques have the unique capability to detect microstructural changes in metals prior to macroscopic cracking.
  • While these methods have been studied for many decades, they have received considerable attention in the recent literature in efforts to address safe and effective operation of the aging infrastructure of transportation, energy industries, and defense systems.
  • Applications of β to fatigue and other dislocation-based material damage, thermal aging, creep damage, and radiation damage are discussed.
  • Limitations should be kept in mind with any SHG measurement technique.
  • At the same time, it is necessary to continue to push the controlled laboratory-type measurements to better understand SHG in terms of microstructural changes and effects, and to continue to develop physics-based materials models to relate nonlinear ultrasonic wave propagation to material microstructural changes.

Did you find this useful? Give us your feedback

Figures (8)

Content maybe subject to copyright    Report

J Nondestruct Eval (2015) 34:273
DOI 10.1007/s10921-014-0273-5
Review of Second Harmonic Generation Measurement Techniques
for Material State Determination in Metals
K. H. Matlack · J.-Y. Kim · L. J. Jacobs · J. Qu
Received: 11 July 2014 / Accepted: 12 November 2014 / Published online: 25 November 2014
© Springer Science+Business Media New York 2014
Abstract This paper presents a comprehensive review of
the current state of knowledge of second harmonic genera-
tion (SHG) measurements, a subset of nonlinear ultrasonic
nondestructive evaluation techniques. These SHG techniques
exploit the material nonlinearity of metals in order to mea-
sure the acoustic nonlinearity parameter, β. In these measure-
ments, a second harmonic wave is generated from a propagat-
ing monochromatic elastic wave, due to the anharmonicity
of the crystal lattice, as well as the presence of microstruc-
tural features such as dislocations and precipitates. This arti-
cle provides a summary of models that relate the different
microstructural contributions to β, and provides details of the
different SHG measurement and analysis techniques avail-
able, focusing on longitudinal and Rayleigh wave methods.
The main focus of this paper is a critical review of the liter-
ature that utilizes these SHG methods for the nondestructive
evaluation of plasticity, fatigue, thermal aging, creep, and
radiation damage in metals.
Keywords Second harmonic generation · Microstructural
evolution in metal · Nonlinear ultrasonic measurements
K. H. Matlack (
B
)
Department of Mechanical and Process Engineering, Swiss Federal
Institute of Technology (ETH Zurich), Zurich, Switzerland
e-mail: matlackk@ethz.ch
J.-Y. Kim · L. J. Jacobs
School of Civil and Environmental Engineering,
Georgia Institute of Technology, Atlanta, GA, USA
J. Qu
Department of Civil and Environmental Engineering,
Northwestern University, Evanston, IL, USA
1 Introduction
Nonlinear ultrasonic methods have the powerful ability to
characterize microstructural features in materials. Compared
to more conventional linear ultrasonic methods that can
detect cracks or features on the order of the wavelength
of the ultrasonic wave, nonlinear methods are sensitive to
microstructural features that are orders of magnitude smaller
than the wavelength. Second harmonic generation (SHG) is a
type of nonlinear ultrasonic method that has been shown to be
capable of detecting and monitoring microstructural changes
in metals. The physical mechanism of this is as follows: as
a sinusoidal ultrasonic wave propagates through a mater-
ial, the interaction of this wave with microstructural features
generates a second harmonic wave. This effect is quantified
with the measured acoustic nonlinearity parameter, β.SHG
measurement methods have received significant focus and
attention in the literature in recent decades, as the reliability
and integrity of structural components becomes increasingly
important to ensure safe operation of critical structures in,
for example, the energy, transportation, and aviation indus-
try. This paper presents a review of SHG measurements and
their applications, to provide a comprehensive summary of
these techniques, to thoroughly explain these measurements
to new comers in this field of research, and to show where
advances in this field of research are needed.
SHG methods were first reported on back in the 1960s,
with a series of papers by Breazeale et al. [1,2], work by
Gedroitz and Krasilnikov [3], and another series by Hikata
et al. [4,5]. Some of the first reported SHG measurements of
anharmonicity were conducted by both Gedroitz and Krasil-
nikov [3], as well as Breazeale and Thompson [1]. In the
latter, the second harmonic wave was measured in polycrys-
talline aluminum over increasing source voltage of the fun-
damental wave with quartz crystal transducers, showing the
123

273 Page 2 of 23 J Nondestruct Eval (2015) 34 :273
linear dependence of the generated second harmonic wave on
the squared amplitude of the fundamental wave. Other ini-
tial experimental studies primarily focused on single crystals
(e.g. copper [2,610], germanium [11], aluminum [5,12]),
and on fused silica [1215].
There are a multitude of other nonlinear ultrasonic
NDE techniques. Acousto-elasticity measurements are based
on the phenomenon that changes in the stress state of
a nonlinear medium cause a change in the wave veloc-
ity [16], and measurements exploiting the acousto-elastic
effect have shown the sensitivity of material nonlinearity
to fatigue microcracks [17]. Nonlinear elastic wave spec-
troscopy (NEWS) methods—such as nonlinear resonant
ultrasound spectroscopy (NRUS) and nonlinear wave modu-
lation spectroscopy (NWMS) techniques—have been exten-
sively used to characterize a variety of materials such as geo-
materials, rock, and concrete [18]. These techniques exploit
the nonlinear hysteretic nature of these materials. NRUS
techniques look at a resonance frequency shift with increas-
ing amplitude excitation [1820]. NWMS techniques typi-
cally look at modulation frequencies as a result of mixing
a very low frequency (sometimes induced by a shaker or
mechanical tapping) and a high frequency ultrasonic wave
[21,22]. Nonlinear time reversal methods have been used
in geophysics applications and detecting surface defects
[23,24]. Nonlinear mixing techniques [25,26] utilize two
input waves at different ultrasonic frequencies, and have
recently gained attention in the literature. Material nonlinear-
ity causes the generation of sum or difference frequencies,
depending on input wave polarities, where the two waves
interact. Nonlinear mixing techniques have the potential to
be isolated from equipment and system nonlinearities inher-
ent in harmonic generation experiments.
Second harmonic generation measurements have shown
to be very applicable in detecting microstructural changes in
metals prior to macroscopic damage and/or microcracking.
Some experimental techniques have the unique advantage
of simplicity in utilizing commercial transducers and stan-
dard ultrasonic testing equipment. Much of the current work
on SHG techniques is focused on how these measurements
can be applied to interrogate real materials under realistic
loading conditions. A variety of different measurement tech-
niques have been developed to interrogate different geome-
tries and damage types. This review is intended to extend pre-
vious reviews [27,28], and will specifically focus on second
harmonic generation experiments and applications explored
particularly within the recent years. This review begins with
a theoretical overview of SHG of longitudinal and Rayleigh
waves, followed by a review of different microstructural con-
tributions to β. Then, the different SHG measurement tech-
niques reported throughout the literature are reviewed, fol-
lowed by a discussion on how these measurements have been
applied to monitor microstructural evolution during material
damage. Finally, an outlook on potential future directions of
SHG measurement research is given.
2 Review of SHG Theory
2.1 Longitudinal Waves
Consider longitudinal wave propagation through an isotropic
medium with a quadratic nonlinearity, which results from a
non-quadratic interatomic potential in crystalline materials.
The equation of motion, simplified to one-dimension is:
ρ
2
u
t
2
=
∂σ
xx
x
(1)
where ρ is the material density, u is the particle displacement,
σ
xx
is the normal stress in the x-direction, x is the material
coordinate, and t is time. The constitutive equation for a
quadratic nonlinearity is given as:
σ
xx
= σ
0
+ E
1
u
x
+
1
2
E
2
u
x
2
+ ... (2)
where E
1
and E
2
are the appropriate second- and third-order
elastic constants. The nonlinear wave equation can thus be
derived as
2
u
t
2
= c
2
1 β
u
x
2
u
x
2
(3)
where β is the nonlinearity parameter, and c is the longitudi-
nal wave velocity in the material. For a material in its virgin
state, β is equivalent to the lattice anharmonicity component,
β
0
, and is a function of second- and third-order elastic con-
stants of the material:
β
0
=−
3C
11
+ C
111
σ
0
+ C
11
(4)
where C
11
and C
111
are the second- and third-order Brugger
elastic constants, respectively, written in Voigt notation, and
σ
0
is the initial stress in the material. Equation (4) assumes
wave propagation in the (100) direction, and is also an exact
solution for isotropic materials. Note that β depends on the
crystalline structure and symmetry of the material, which
was shown in [29] through calculations of β for pure mode
propagation for various single-crystals.
The time harmonic solution to Eq. (3), assuming plane
wave propagation, has the form:
u = A
1
sin x ωt) +
β A
2
1
xκ
2
8
cos(2κ x 2ωt) + ...
(5)
123

J Nondestruct Eval (2015) 34 :273 Page 3 of 23 273
where ω(= 2π f )is the radial frequency of the wave at fre-
quency f , κ(= ω/c) is the wavenumber of the propagating
wave, A
1
is the amplitude of the first harmonic wave. The
coefficient in front of the second term is A
2
, the amplitude
of the second harmonic wave, which assumes the absence of
attenuation, diffraction, scattering, and assumes plane wave
propagation. By simply rearranging this amplitude term, the
nonlinearity parameter can be expressed in terms of acoustic
quantities, i.e.:
β =
8A
2
A
2
1
xκ
2
(6)
When written in this form, β is generally referred to as
the acoustic nonlinearity parameter. Thus by measuring the
second harmonic wave amplitude, along with the first har-
monic amplitude, wavenumber, and propagation distance,
one can determine the acoustic nonlinearity parameter, β.
This derivation can be expanded to three dimensions [30],
and has been derived for Rayleigh waves [31] and explored
for Lamb waves [32]. The same general form of A
2
has
been shown for Rayleigh waves, in terms of dependence on
propagation distance, wavenumber, and first harmonic wave.
Experimental results have shown the same relationship of
A
2
proportional to A
2
1
holds true for Lamb waves as well
[8285].
Note that the energy transfer from the first to second har-
monic wave in SHG is very small compared to the energy of
the propagating first harmonic wave, such that the decrease in
A
1
due to the energy transfer is insignificant for small propa-
gation distances. Further, in most experiments, the amplitude
A
2
is orders of magnitude smaller than A
1
. In the present
considerations, we assume the wave propagation distance is
small enough such that the energy loss of A
1
is negligible
compared to the total energy of the propagating first har-
monic wave.
For real materials and finite propagation distances,
attenuation (dissipation, scattering, diffraction) will further
decrease the amplitudes of the first and second harmonic
waves with increasing propagation distance. The effect on the
generated second harmonic wave when attenuation effects
are non-negligible is derived elsewhere for longitudinal
waves [5,16,33]. Attenuation effects are non-negligible at
larger propagation distances, i.e. when x
2
2α
1
)<<1
does not hold, and the acoustic nonlinearity parameter for
this case is given as [5,16]:
β
atten
= β
x
2
2α
1
)
{
1 exp
[
2
2α
1
)x
]
}
(7)
where α
1
is the attenuation coefficient at the first harmonic
frequency, and α
2
is the attenuation coefficient at the second
harmonic frequency. Note that in Eq. (7)itisassumedthat
SHG and attenuation effects occur independently, and in the
limit α
2
2α
1
, the measured β equals the actual β.
A more accurate expression for the acoustic nonlinearity
parameter can be found by accounting for the on-axis dif-
fraction effects of both the first and second harmonic wave.
A diffraction correction
D
β
to β was introduced by Hurley
et al. [14]as:
D
β
=
|
D(ω)
|
2
|
D(2ω)
|
(8)
where
|
D(ω)
|
and
|
D(2ω)
|
are the diffraction corrections
to the first and second harmonic waves, respectively. The
acoustic nonlinearity parameter scaled by this diffraction cor-
rection is thus β
D
= β D
β
.
The linear diffraction correction, i.e. the diffraction cor-
rection for the propagating first harmonic wave, has been
derived in full previously [34] for a piston source such that
the amplitude is constant across the source. This diffraction
correction is given by:
D, x, a) =1exp(iκa
2
/x)
J
0
a
2
/x)+iJ
1
a
2
/x)
(9)
where a is the transducer radius, and J
0
and J
1
are Bessel
functions of the first kind. In actuality, transducers are not a
perfect piston source and there is some spatial distribution
of amplitude over the surface of the transducer face, which
could potentially approximate a Gaussian distribution [35].
The diffraction of the second harmonic wave is spatially
different than that of the first harmonic. The wave is gener-
ated not by the transducer (in a perfect system at least), but
by the propagating first harmonic wave, which is diffracting
over propagation distance. This nonlinear diffraction can be
physically interpreted as follows: at each instance that a por-
tion of the second harmonic wave is generated, that portion
will then diffract linearly over the remainder of the propa-
gation distance to the receiving transducer. This nonlinear
diffraction effect has been derived as [14,36]:
D(2ω, x, a) =
x
0
[
D, x σ/2, a)
]
2
dσ
x
(10)
where D, x, a) is the linear diffraction correction given in
Eq. (9).
2.2 Rayleigh Surface Waves
Consider a Rayleigh surface wave propagating in the positive
x direction in an isotropic infinite half-space, with direction
z pointing into the half-space, with a weak quadratic nonlin-
earity. The Rayleigh wave motion along a stress-free surface,
123

273 Page 4 of 23 J Nondestruct Eval (2015) 34 :273
assuming plane wave propagation, can be decomposed in
terms of its shear and longitudinal wave components, which
for a propagating sinusoidal wave is given as:
u
x
= A
1
e
pz
2 ps
κ
2
R
+ s
2
e
sz
e
i
R
xωt)
(11)
u
z
= iA
1
p
κ
R
e
pz
2κ
2
R
κ
2
R
+ s
2
e
sz
e
i
R
xωt)
(12)
where p
2
= κ
2
R
κ
2
p
, s
2
= κ
2
R
κ
2
s
, and κ
R
, κ
P
, and κ
S
are
the wavenumbers for the Rayleigh, longitudinal, and shear
waves, respectively. The second harmonic wave displace-
ment components can be approximated at a sufficiently far
distance as [37,38]:
u
x
(2ω) = A
2
e
2 pz
2 ps
κ
2
R
+ s
2
e
2sz
e
i2
R
xωt)
(13)
For isotropic materials, the acoustic nonlinearity due to shear
waves vanishes due to symmetry conditions, so the gener-
ated second harmonic wave is purely due to the longitudi-
nal wave component of the Rayleigh wave. Herrmann et al.
[38] derived the acoustic nonlinearity parameter in terms of
out-of-plane components of the first and second harmonic
amplitude as:
β =
¯u
z
(2ω)
¯u
2
z
(ω)x
i8 p
κ
2
P
κ
R
1
2κ
2
R
κ
2
R
+ s
2
(14)
where the overbar indicates the displacement is evaluated at
z = 0.
As can be seen in Eq. (14), the dependence of β on first
and second harmonic amplitude, as well as on propagation
distance is the same for nonlinear Rayleigh waves as it is for
nonlinear longitudinal waves, at least under the plane wave
assumption. The dependence on propagation distance is gen-
erally exploited in nonlinear Rayleigh wave measurements—
the wave propagation distance is varied over multiple mea-
surements of first and second harmonic wave amplitude, and
a relative measure of β can thus be made.
Since nonlinear Rayleigh waves are typically used for
longer propagation distances, accounting for attenuation [39]
and diffraction effects [35,39,40] are crucial for accurate
measurements of β. Shull et al. [35] explored diffraction
effects of nonlinear Rayleigh waves both theoretically and
numerically for a general forcing function, a uniform line
source, and a Gaussian source. For example, a line source
with a Gaussian distribution along the y-axis is defined by:
w(y) = v
0
e
(y/a)
(15)
where v
0
is the source amplitude and a is the half-length of
the line source. The first harmonic wave velocity component
propagation distance [m]
Sample/source width [m]
−0.05 0 0.05
0
0.1
0.2
0.3
0.4
0.5
0
0.2
0.4
0.6
0.8
1
propagation distance [m]
Sample/source width [m]
−0.05 0 0.05
0
0.1
0.2
0.3
0.4
0.5
0
1
2
3
4
5
x 10
−3
Fig. 1 Amplitude of first harmonic from Eq. (16)(left) and second har-
monic from Eq. (17)(right) Rayleigh wave with propagation distance
and sample width, including diffraction and attenuation effects, at y=0
[40]. Results normalized by input first harmonic amplitude (reproduced
from [40] with permission from Elsevier)
for this forcing function has been derived as [35]:
v
1
(x, y) =
v
0
e
1
x)
1 + ix/x
0
exp(
(y/a)
2
1 + ix/x
0
) (16)
where x
0
= κ
R
a
2
/2 is the Rayleigh distance that marks the
transition from the near field to the far field diffraction region
of the source, where κ
R
is the Rayleigh wavenumber. The
second harmonic wave velocity component can be written
as:
v
2
(x, y) = β
11
v
2
0
D
R
(2ω) (17)
where β
11
is a nonlinearity parameter defined elsewhere
[35,37], and D
R
(2ω) is the complex diffraction coefficient
for the propagating and generated second harmonic Rayleigh
wave, which is given in full in [35]. The first and second har-
monic Rayleigh wave amplitude with propagation distance
including diffraction and attenuation effects, i.e. Eqs. (16)
and (17), are plotted in Fig. 1 [40]. These plots are normal-
ized to the input first harmonic amplitude at x=0. In these
plots, it is assumed that α
n
= n
4
α
1
, and note that the prop-
agation distance, x, is plotted on the vertical axis on these
plots, and the sample/source width, y, is plotted on the hori-
zontal axis.
The proportionality of the generated second harmonic
wave remains the square of the fundamental at a given propa-
gation distance x, i.e. v
2
(y) v
2
1
(y), though the dependence
of v
2
(y)on propagation distance is obscured. While the linear
dependence of v
2
(y) on propagation distance has been shown
to be a good approximation for short propagation distances
in many metals [38,4143], the full solution incorporating
attenuation, diffraction, and source effects will result in a
more accurate determination of β.
123

J Nondestruct Eval (2015) 34 :273 Page 5 of 23 273
2.3 Microstructural Contributions to β
The parameter β depends on the crystalline structure of the
material, and also on localized strain present in the mate-
rial. This strain arises from microstructural features such
as dislocations and precipitates, and dislocation contribu-
tions to β can greatly exceed that of the lattice anhar-
monicity. This section provides a comprehensive review
of different microstructural contributions to the magnitude
of the acoustic nonlinearity parameter. Theoretical deriva-
tions of effects of dislocation pinning [4,16,44], dislocation
dipoles [45,46], precipitate-pinned dislocations [16,4749],
and cracks [5052] are reviewed.
2.3.1 Dislocation Pinning: Hikata et al. Model
The dislocation motion contribution to acoustic nonlinearity
was first developed by Suzuki et al. [53] and expanded on by
Hikata et al. [4], and this model has been used to interpret a
multitude of experimental results of second harmonic gener-
ation. The model is based on the dislocation string vibration
model of Granato and Lücke [54], and considers dislocation
bowing as a line segment pinned between two points, a dis-
tance 2L apart. These pinning points can be grain boundaries,
other dislocations, or point defects in the material. Assume a
small but non-zero longitudinal stress, σ, with shear compo-
nent τ such that τ = Rσ where R is the resolving shear factor,
is then applied to this dislocation segment such that it bows
out between the two pinning points. An approximation of this
geometry is depicted in Fig. 2, where the radius of curvature,
r, of the bowed segment and angle θ are annotated. Note
that this stress can be thought of as either an internal resid-
ual stress or externally applied stress, but it is small enough
such that the dislocation segment does not break away from
the pinning points. An ultrasonic stress wave sets the pinned
dislocation segment into a vibrational motion. However, the
dynamics of the dislocation motion is neglected by assum-
ing that the mass density of the dislocation line is negligi-
ble. Further, the dislocation dynamics that may occur during
Fig. 2 a Diagram showing geometry of bowed dislocation segment of
length 2L between two pinning points and under an applied shear stress
τ
i
, in terms of radius of curvature r and angle θ. b Diagram showing
movement of dislocation segment with superimposed ultrasonic stress
on top of initial stress τ
transient plastic deformation is not considered since this is
not of the source of ultrasonic nonlinearity considered here.
Assuming the dislocation density is small enough such
that bowed dislocations act independently of each other, the
line tension, T, of this dislocation segment due to the applied
stress was approximated as constant and independent of ori-
entation as T = μb
2
/2, where b is the Burgers vector and μ is
the shear modulus. In the Hikata et al. model, the area swept
out by the dislocation is approximated as a circular arc with
a constant radius.
The total strain in the material can be written as a summa-
tion of the lattice strain plus the strain due to the dislocation
motion, i.e. ε = ε
1
+ γ , where is the conversion factor
between shear and longitudinal strain and the shear strain γ
due to a distribution of bowed dislocations. Assuming the
same form of the stress–strain relation as in Eq. (2)forthe
internal stress, the resulting stress–strain relationship due to
the total strain in the material is:
ε =
1
E
1
+
2L
2
R
3μ
σ +
E
2
E
3
1
σ
2
+
4L
4
R
3
5μ
3
b
2
σ
3
(18)
with bowed dislocations density , dislocation loop length
L. Assuming a small ultrasonic stress σ is superimposed on
the internal stress σ
1
, and assuming the lattice contribution
of the nonlinearity parameter is β
0
=−E
2
/E
1
, the change
in nonlinearity parameter due to dislocation pinning can be
written as:
β
pd
=
24
5
L
4
R
3
C
2
11
μ
3
b
2
σ
1
(19)
Further details of this model can be found elsewhere [4,16].
Note that E
1
is defined as σ
0
+ C
11
in Eq. (4), but here
it is assumed that the initial stress σ
0
is zero. It should be
specifically noted that the internal stress σ
1
in this analysis,
as well as the superimposed ultrasonic stress, is assumed to
be much smaller than the yield stress of the material, such
that dislocation displacement is small.
2.3.2 Dislocation Pinning: Extensions of the Hikata et al.
Model
Later work by Cash and Cai [55] extended the Hikata et al.
model to account for orientation-dependent line energy in
the analytical model and verified with dislocation dynamics
simulations. Results showed that the Hikata et al. model can
accurately predict β only for small values of Poisson’s ratio
for both screw and edge dislocations. However, at values
of Poisson’s ratio greater than about 0.2 (i.e. most metals),
the dislocation dynamics simulations and developed analyt-
ical model show that relationship between β and the applied
stress is not in fact linear, and simulations show that β can
123

Citations
More filters
Journal ArticleDOI
TL;DR: In this paper, the authors review the recent advances in the theory of nonlinear guided waves, as well as the numerical simulations and experiments that demonstrate their utility, including the application of higher harmonic generation of ultrasonic guided wave modes for nondestructive evaluation.
Abstract: Interest in using the higher harmonic generation of ultrasonic guided wave modes for nondestructive evaluation continues to grow tremendously as the understanding of nonlinear guided wave propagation has enabled further analysis. The combination of the attractive properties of guided waves with the attractive properties of higher harmonic generation provides a very unique potential for characterization of incipient damage, particularly in plate and shell structures. Guided waves can propagate relatively long distances, provide access to hidden structural components, have various displacement polarizations, and provide many opportunities for mode conversions due to their multimode character. Moreover, higher harmonic generation is sensitive to changing aspects of the microstructures such as to the dislocation density, precipitates, inclusions, and voids. We review the recent advances in the theory of nonlinear guided waves, as well as the numerical simulations and experiments that demonstrate their utility.

107 citations


Additional excerpts

  • ...EQ-TARGET;temp:intralink-;e015;63;284H 1⁄4 H1 þH2 with kH2k ≪ kH1k; (15)...

    [...]

Journal ArticleDOI
Yi Yang1, Ching-Tai Ng1, Andrei Kotousov1, Hoon Sohn2, Hyung Jin Lim2 
TL;DR: In this paper, the second harmonic generation due to non-linear interaction of Lamb waves with a fatigue crack was investigated in 3D finite element simulations and experimental studies were carried out to provide physical insight into the mechanism of second harmonic generator.

107 citations


Cites background from "Review of Second Harmonic Generatio..."

  • ...Recently, several studies have focused on the nonlinearity of Lamb waves [32]....

    [...]

Journal ArticleDOI
TL;DR: In this paper, nonlinear Lamb wave-mixing was applied in the detection of micro-cracks in plates, and the analysis of the nonlinear interaction of these waves with cracks of various lengths and widths was performed using finite-element simulations.
Abstract: In engineering practice, failures due to fatigue cracks in metallic structures have always been difficult to predict. In our study, nonlinear Lamb wave-mixing was applied in the detection of micro-cracks in plates. The analysis of the nonlinear interaction of these waves with cracks of various lengths and widths was performed using finite-element simulations. The simulation results showed that the sideband at the sum frequency provide a sensitive means for micro-crack detection. Moreover, the sideband amplitudes show a monotonic increase with micro-crack length, but a decrease with micro-crack width. Experiments using Lamb wave-mixing were conducted on plates with fatigue cracks. The experimental results show that a proposed acoustic nonlinearity parameter related to the sideband at the sum frequency is sensitive to micro-cracks in plate, and is well correlated with the damage degree.

101 citations

Journal ArticleDOI
TL;DR: In this paper, a wave vector analysis is conducted that enables selection of primary waves traveling in any direction that generate phase matched secondary waves, and the authors have tabulated many sets of primary wave and secondary wave sets.
Abstract: The extraordinary sensitivity of nonlinear ultrasonic waves to the early stages of material degradation makes them excellent candidates for nondestructive material characterization. However, distinguishing weak material nonlinearity from instrumentation nonlinearity remains problematic for second harmonic generation approaches. A solution to this problem is to mix waves having different frequencies and to let their mutual interaction generate sum and difference harmonics at frequencies far from those of the instrumentation. Mixing of bulk waves and surface waves has been researched for some time, but mixing of guided waves has not yet been investigated in depth. A unique aspect of guided waves is their dispersive nature, which means we need to assure that a wave can propagate at the sum or difference frequency. A wave vector analysis is conducted that enables selection of primary waves traveling in any direction that generate phase matched secondary waves. We have tabulated many sets of primary waves and ...

72 citations

Journal ArticleDOI
TL;DR: This study theoretically reveals that the low frequency S0 mode of Lamb waves can be used as the fundamental waves to quantitatively identify micro‐cracks in thin plates.

56 citations


Cites background from "Review of Second Harmonic Generatio..."

  • ...A large number of experiments have demonstrated that the nonlinear ultrasonic detection technique based on second harmonics is an effective tool to detect the change of the microstructure of the metal materials [37]....

    [...]

References
More filters
MonographDOI
01 Jan 1985

5,957 citations

Journal ArticleDOI
TL;DR: In this article, the authors proposed a new theory of elastic contact, which is more closely related to real surfaces than earlier theories, and showed how the contact deformation depends on the topography of the surface, and established the criterion for distinguishing surfaces which touch elastically from those which touch plastically.
Abstract: It is usually assumed that the real area of contact between two nominally flat metal surfaces is determined by the plastic deformation of their highest asperities. This leads at once to the result that the real area of contact is directlyproportional to the load and independent of the apparent area-a result with many applications in the theories of electric contacts and friction. Archard pointed out that plastic deformation could not be the universal rule, and introduced a model which showed that, contrary to earlier ideas, the area of contact could be proportional to the load even with purely elastic contact. This paper describes a new theory of elastic contact, which is more closely related to real surfaces than earlier theories. We show how the contact deformation depends on the topography of the surface, and establish the criterion for distinguishing surfaces which touch elastically from those which touch plastically. The theory also indicates the existence of an 'elastic contact hardness', a composite quantity depending on the elastic properties and the topography, which plays the same role in elastic contact as the conventional hardness does in plastic contact. A new instrument for measuring surface topography has been built; with it the various parameters shown by the theory to govern surface contact can be measured experimentally. The typical radii of surface asperities have been measured. They were found, surprisingly, to be orders of magnitude larger than the heights of the asperities. More generally we have been able to study the distributions of asperity heights and of other surface features for a variety of surfaces prepared by standard techniques. Using these data we find that contact between surfaces is frequently plastic, as usually assumed, but that surfaces which touch elastically are by no means uncommon in engineering practice.

5,371 citations

BookDOI
01 Jan 1982
TL;DR: The Review of Progress in Quantitative NDE (ROPQN) as mentioned in this paper is the world's leading conference in reporting annually new research and development results in quantitative NDE and promotes communication between the research and engineering communities and emphasize current reporting of work in progress.
Abstract: The Review of Progress in Quantitative NDE is the world's leading conference in reporting annually new research and development results in quantitative NDE. The conference reports on both fundamental and applied advances in NDE and promotes communication between the research and engineering communities and emphasize current reporting of work in progress. Attendees include representatives of academia (including students), industry, and government with approximately one-half coming from the United States and the other half from overseas. This volume represents the best report of ongoing work that is available anywhere. Connections and overlap with the medical diagnostic community are highlighted.

1,989 citations

Journal ArticleDOI
TL;DR: In this article, a quantitative theory of damping and modulus changes due to dislocations is developed, and it is found that the model used by Koehler of a pinned dislocation loop oscillating under the influence of an applied stress leads to two kinds of loss, one frequency dependent and the other not.
Abstract: A quantitative theory of damping and modulus changes due to dislocations is developed. It is found that the model used by Koehler of a pinned dislocation loop oscillating under the influence of an applied stress leads to two kinds of loss, one frequency dependent and the other not. The frequency dependent loss is found to have a maximum in the high megacycle range. The second type of loss is a hysteresis loss which proves to be independent of frequency over a wide frequency range which includes the kilocycle range. This loss has a strain‐amplitude dependence of the type observed in the kilocycle range. The theory provides a quantitative interpretation of this loss.

1,754 citations


"Review of Second Harmonic Generatio..." refers methods in this paper

  • ...The model is based on the dislocation string vibration model of Granato and Lücke [54], and considers dislocation bowing as a line segment pinned between two points, a distance 2L apart....

    [...]

Book
01 Jan 1967

983 citations


"Review of Second Harmonic Generatio..." refers background in this paper

  • ...The stress, σrr (r), in the radial direction at radius r > ra for this scenario are given in Eringen [65], as: σrr (r) = −p0 r3p r3 (23) Assuming the precipitate and matrix have different elastic properties, the stress in the matrix at radius r due to this embedded precipitate can then be written as [49,67]: σrr (r) = −4μδ [ 3Bp 3Bp + 4μ ] r3p r3 (24) where Bp is the bulk modulus of the precipitate....

    [...]

  • ...The stress, σrr (r), in the radial direction at radius r > ra for this scenario are given in Eringen [65], as:...

    [...]

Frequently Asked Questions (17)
Q1. What is the FFT used to determine the amplitude of the first and second harmonic waves?

Then the fast Fourier transform (FFT) is calculated to determine the amplitude of the signal as a function of frequency, in order to extract the relative amplitudes of the first and second harmonic waves. 

The amplifier should be inherently linear, and have the capability of exciting a high-powered ultrasonic wave, for example 1,200 Vpp on a 50 load. 

The linear diffraction correction, i.e. the diffraction correction for the propagating first harmonic wave, has been derived in full previously [34] for a piston source such that the amplitude is constant across the source. 

The phenomenon of contact acoustic nonlinearity, directly related to nonlinearity generated from crack interfaces, has been utilized to interrogate material interfaces [50,128,157] and bonds [158,159]. 

The wedge material is typically made of acrylic or other plastic material, and is designed to mount the transducer at the angle required to excite the Rayleigh surface wave in the sample. 

The force–displacement relation of the dipole is a nonlinear relation, and as such it has been shown that when perturbed by an ultrasonic wave, this feature generates a nonzero component of acoustic nonlinearity [16,45,46,60,61]. 

The dependence on propagation distance is generally exploited in nonlinear Rayleigh wave measurements— the wave propagation distance is varied over multiple measurements of first and second harmonic wave amplitude, and a relative measure of β can thus be made. 

Recent work shows that the acoustic nonlinearity parameter is sensitive to microstructural changes in reactor pressure vessel steels induced by increasing neutron fluence, different irradiation temperatures, and different material compositions [75,76]. 

The key aspect to interpreting these dislocation-based SHG measurements, as shown particularly in the numerous SHG measurements over fatigue damage, is that specific dislocation substructures and evolution in the material system measured is crucial for accurate SHG measurement evaluation. 

A technique to minimize coupling effects on the nonlinearity measurement has been proposed in [96], in which the authors report a decrease in measurement variation by half using light oil coupling. 

A damped oscillatory behavior of β has been shown in regions above the Buck hook, which is a result of the Peierls–Nabarro barrier stress associated with dislocation motion [89,105], but this oscillatory behavior is greatly damped in polycrystalline metals due to the random orientations of slip systems (and thus random values of the Schmid factor) [89]. 

This detection system has a reported sensitivity of 10−16 m, which is more than sufficient to detect the displacement amplitudes of the propagated first and second harmonic wave [108]. 

A shift in focus to this type of monitoring with SHG could help advance the possibility of in-service applications of SHG measurements. 

Note in comparison to the irradiated copper single crystals, the neutron fluence for reactor pressure vessel materials is about 1–3 × 1019 n/cm2 (E > 1 MeV) after 40 years of operation, and significantly higher for internals materials closer to the fuel rods. 

the dislocation dynamics that may occur duringtransient plastic deformation is not considered since this is not of the source of ultrasonic nonlinearity considered here. 

They pointed out that the change in β due to fatigue was much greater than changes due to precipitates, but that precipitate structures could greatly affect the dislocation structure and interactions during fatigue. 

The authors show with a simulation experiment that their solution can track β for changing dislocation lengths, but the crucial link to real SHG experiments has yet to be realized.