propagation of a finite-amplitude elastic pulse in a bar of berea … · 2017-11-27 · propagation...

18
Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea Sandstone: A Detailed Look at the Mechanisms of Classical Nonlinearity, Hysteresis, and Nonequilibrium Dynamics Marcel C. Remillieux 1 , T. J. Ulrich 2 , Harvey E. Goodman 3 , and James A. Ten Cate 1 1 Geophysics Group (EES-17), Los Alamos National Laboratory, Los Alamos, NM, USA, 2 Detonation Science and Technology Group (Q-6), Los Alamos National Laboratory, Los Alamos, NM, USA, 3 Chevron Energy Technology Company, Houston, TX, USA Abstract We study the propagation of a nite-amplitude elastic pulse in a long thin bar of Berea sandstone. In previous work, this type of experiment has been conducted to quantify classical nonlinearity, based on the amplitude growth of the second harmonic as a function of propagation distance. To greatly expand on that early work, a noncontact scanning 3-D laser Doppler vibrometer was used to track the evolution of the axial component of the particle velocity over the entire surface of the bar as functions of the propagation distance and source amplitude. With these new measurements, the combined effects of classical nonlinearity, hysteresis, and nonequilibrium dynamics have all been measured simultaneously. We show that the numerical resolution of the 1-D wave equation with terms for classical nonlinearity and attenuation accurately captures the spectral features of the waves up to the second harmonic. However, for higher harmonics the spectral content is shown to be strongly inuenced by hysteresis. This work also shows data which quantify not only classical nonlinearity but also the nonequilibrium dynamics based on the relative change in the arrival time of the elastic pulse as a function of strain and distance from the source. Finally, a comparison is made to a resonant bar measurement, a reference experiment used to quantify nonequilibrium dynamics, based on the relative shift of the resonance frequencies as a function of the maximum dynamic strain in the sample. 1. Introduction From a mechanical point of view, sedimentary rocks and man-made materials like concrete are complicated solids. They may be described as a disordered network of mesoscopic-sized hardelements (e.g., grains with characteristic lengths ranging from tens to hundreds of microns) cemented together by a softbond system. Such systems belong to a wider class of materials referred to as nonlinear mesoscopic elastic materials (NMEMs) (Guyer & Johnson, 1999). The microscopic-sized imperfections at the interfaces between the hard and soft subsystems are believed to be responsible for several peculiar properties related to nonlinear and nonequilibrium dynamics, including the dependence of elastic parameters and attenuation on strain ampli- tude, slow dynamics, and hysteresis with end point memory (Ostrovsky & Johnson, 2001; TenCate, Smith, & Guyer, 2000; Winkler et al., 1979). Understanding and predicting these properties, especially the cementation (soft bond system), is at the heart of numerous applications including nuclear waste repository, oil and gas exploration, and monitoring of infrastructure integrity. Increasingly, more elaborate techniques have emerged and evolved over the last 20 years to observe and quantify the effects of nonlinear and nonequilibrium dynamics in these materials. Nonlinear elasticity, in the classical sense, is usually modeled by an expansion of the elastic energy as a power series with respect to the strain tensor (Landau & Lifshitz, 1986; Murnaghan, 1937). Practically, in a medium with such nonlinear- ity, a sinusoidal elastic wave experiences distortions, eventually leading to shock formation if damping is sufciently small. In the frequency domain, harmonic generation is observed as some of the energy contained at the source frequency is transferred to other frequencies (i.e., higher harmonics). As part of this process, the wave oscillates around the same equilibrium state as observed in the sample before the source was activated. In the context of elastic wave propagation, classical nonlinearity was studied experimentally (Meegan et al., 1993; Ten Cate et al., 1996) and theoretically (McCall, 1994; Thurston & Shapiro, 1967; Van Den Abeele, 1996) in the early studies on the dynamics of NMEMs. REMILLIEUX ET AL. NONLINEAR PROPAGATION OF ELASTIC PULSE 1 PUBLICATION S Journal of Geophysical Research: Solid Earth RESEARCH ARTICLE 10.1002/2017JB014258 Key Points: The propagation of an elastic pulse is used to decouple the effects of classical nonlinear elasticity and nonequilibrium dynamics The 1-D wave equation with classical nonlinearity and attenuation captures the spectral content of the pulse up to the second harmonic Nonequilibrium dynamics is quantied by tracking the arrival time of an elastic pulse as a function of its amplitude Supporting Information: Supporting Information S1 Movie S1 Correspondence to: M. C. Remillieux, [email protected] Citation: Remillieux, M. C., Ulrich, T. J., Goodman, H. E., & Ten Cate, J. A. (2017). Propagation of a nite-amplitude elastic pulse in a bar of Berea sandstone: A detailed look at the mechanisms of classical nonlinearity, hysteresis, and nonequilibrium dynamics. Journal of Geophysial Research: Solid Earth, 122. https://doi.org/10.1002/2017JB014258 Received 11 APR 2017 Accepted 13 OCT 2017 Accepted article online 18 OCT 2017 ©2017. The Authors. This is an open access article under the terms of the Creative Commons Attribution-NonCommercial-NoDerivs License, which permits use and distri- bution in any medium, provided the original work is properly cited, the use is non-commercial and no modications or adaptations are made.

Upload: others

Post on 18-Mar-2020

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea … · 2017-11-27 · Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea Sandstone: A Detailed Look

Propagation of a Finite-Amplitude Elastic Pulsein a Bar of Berea Sandstone: A Detailed Lookat the Mechanisms of Classical Nonlinearity,Hysteresis, and Nonequilibrium DynamicsMarcel C. Remillieux1 , T. J. Ulrich2, Harvey E. Goodman3, and James A. Ten Cate1

1Geophysics Group (EES-17), Los Alamos National Laboratory, Los Alamos, NM, USA, 2Detonation Science and TechnologyGroup (Q-6), Los Alamos National Laboratory, Los Alamos, NM, USA, 3Chevron Energy Technology Company, Houston, TX,USA

Abstract We study the propagation of a finite-amplitude elastic pulse in a long thin bar of Bereasandstone. In previous work, this type of experiment has been conducted to quantify classical nonlinearity,based on the amplitude growth of the second harmonic as a function of propagation distance. To greatlyexpand on that early work, a noncontact scanning 3-D laser Doppler vibrometer was used to track theevolution of the axial component of the particle velocity over the entire surface of the bar as functions of thepropagation distance and source amplitude. With these new measurements, the combined effects ofclassical nonlinearity, hysteresis, and nonequilibrium dynamics have all been measured simultaneously. Weshow that the numerical resolution of the 1-D wave equation with terms for classical nonlinearity andattenuation accurately captures the spectral features of the waves up to the second harmonic. However, forhigher harmonics the spectral content is shown to be strongly influenced by hysteresis. This work alsoshows data which quantify not only classical nonlinearity but also the nonequilibrium dynamics based on therelative change in the arrival time of the elastic pulse as a function of strain and distance from the source.Finally, a comparison is made to a resonant bar measurement, a reference experiment used to quantifynonequilibrium dynamics, based on the relative shift of the resonance frequencies as a function of themaximum dynamic strain in the sample.

1. Introduction

From a mechanical point of view, sedimentary rocks and man-made materials like concrete are complicatedsolids. They may be described as a disordered network of mesoscopic-sized “hard” elements (e.g., grains withcharacteristic lengths ranging from tens to hundreds of microns) cemented together by a “soft” bond system.Such systems belong to a wider class of materials referred to as nonlinear mesoscopic elastic materials(NMEMs) (Guyer & Johnson, 1999). The microscopic-sized imperfections at the interfaces between the hardand soft subsystems are believed to be responsible for several peculiar properties related to nonlinear andnonequilibrium dynamics, including the dependence of elastic parameters and attenuation on strain ampli-tude, slow dynamics, and hysteresis with end point memory (Ostrovsky & Johnson, 2001; TenCate, Smith, &Guyer, 2000; Winkler et al., 1979). Understanding and predicting these properties, especially the cementation(soft bond system), is at the heart of numerous applications including nuclear waste repository, oil and gasexploration, and monitoring of infrastructure integrity.

Increasingly, more elaborate techniques have emerged and evolved over the last 20 years to observe andquantify the effects of nonlinear and nonequilibrium dynamics in these materials. Nonlinear elasticity, inthe classical sense, is usually modeled by an expansion of the elastic energy as a power series with respectto the strain tensor (Landau & Lifshitz, 1986; Murnaghan, 1937). Practically, in a medium with such nonlinear-ity, a sinusoidal elastic wave experiences distortions, eventually leading to shock formation if damping issufficiently small. In the frequency domain, harmonic generation is observed as some of the energycontained at the source frequency is transferred to other frequencies (i.e., higher harmonics). As part of thisprocess, the wave oscillates around the same equilibrium state as observed in the sample before the sourcewas activated. In the context of elastic wave propagation, classical nonlinearity was studied experimentally(Meegan et al., 1993; Ten Cate et al., 1996) and theoretically (McCall, 1994; Thurston & Shapiro, 1967; VanDen Abeele, 1996) in the early studies on the dynamics of NMEMs.

REMILLIEUX ET AL. NONLINEAR PROPAGATION OF ELASTIC PULSE 1

PUBLICATIONSJournal of Geophysical Research: Solid Earth

RESEARCH ARTICLE10.1002/2017JB014258

Key Points:• The propagation of an elastic pulse isused to decouple the effects ofclassical nonlinear elasticity andnonequilibrium dynamics

• The 1-D wave equation with classicalnonlinearity and attenuation capturesthe spectral content of the pulse up tothe second harmonic

• Nonequilibrium dynamics isquantified by tracking the arrival timeof an elastic pulse as a function of itsamplitude

Supporting Information:• Supporting Information S1• Movie S1

Correspondence to:M. C. Remillieux,[email protected]

Citation:Remillieux, M. C., Ulrich, T. J., Goodman,H. E., & Ten Cate, J. A. (2017).Propagation of a finite-amplitude elasticpulse in a bar of Berea sandstone: Adetailed look at the mechanisms ofclassical nonlinearity, hysteresis, andnonequilibrium dynamics. Journal ofGeophysial Research: Solid Earth, 122.https://doi.org/10.1002/2017JB014258

Received 11 APR 2017Accepted 13 OCT 2017Accepted article online 18 OCT 2017

©2017. The Authors.This is an open access article under theterms of the Creative CommonsAttribution-NonCommercial-NoDerivsLicense, which permits use and distri-bution in any medium, provided theoriginal work is properly cited, the use isnon-commercial and no modificationsor adaptations are made.

Page 2: Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea … · 2017-11-27 · Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea Sandstone: A Detailed Look

Quasi-static stress-strain measurements on rocks and other NMEMs have been made since the early 1900s(Adams & Coker, 1906). These measurements show not only that rocks are nonlinear but also that they arehysteretic. Repeated looping up and down in stress-strain space results in a curved, banana-shaped path(Holcomb, 1981; Zoback & Byerlee, 1975). Attempts to model this quasi-static hysteresis in NMEMs weregoing on simultaneously but independently as the attempts to model pulse propagation in similar samples(Aleshin et al., 2004; Gusev, 2000; Gusev et al., 1997; Van Den Abeele et al., 1997, 2000). Practically, a sinusoidalelastic wave propagating in a medium with pure hysteretic nonlinearity progressively evolves into a triangu-lar wave. In the frequency domain, such evolution is characterized by the growth of the odd harmonics, witha quadratic dependence with respect to the amplitude at the fundamental frequency. The effects of classicaland hysteretic nonlinearity on a sinusoidal waveform are summarized in Figure 1.

Besides hysteresis effects, the classical theory of nonlinearity cannot account for the other peculiar wave phe-nomena observed in NMEMs. The first repeatable observations showing nonclassical effects were made inresonant bar experiments (Johnson et al., 1996; Remillieux et al., 2016; Ten Cate & Shankland, 1996),ultimately leading to a quantification technique known as nonlinear resonant ultrasound spectroscopy(NRUS). In these experiments, a long thin bar with free boundary conditions, which is representative of a1-D unconstrained system, is excited with a harmonic signal at one end while the elastic response is recordedat the opposite end. The source signal sweeps a frequency range around one resonance frequency of the barwith increasing source amplitudes and the variation of the resonance frequency can be tracked as a functionof the maximum strain in the sample. In a linear elastic material, the resonance frequency is not strain depen-dent and the resonant peak does not change with increasing drive amplitude. In NMEMs, the resonance fre-quency starts decreasing as a function of strain, and very noticeably so when the strain goes above a certainlevel, typically near 10�6 for sandstones at ambient conditions.

Figure 1. Effects of pure classical nonlinearity and pure hysteretic nonlinearity on the evolution of a sinusoidal waveform in the time and frequency domains.Reproduction of a figure originally produced by Van Den Abeele et al. (2000).

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014258

REMILLIEUX ET AL. NONLINEAR PROPAGATION OF ELASTIC PULSE 2

Page 3: Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea … · 2017-11-27 · Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea Sandstone: A Detailed Look

Broadly speaking, the drop in resonance frequency indicates material softening with increasing driving strain.However, the dependence of the resonance frequency on strain is not trivial (Pasqualini et al., 2007; TenCateet al., 2004). At very small strains (e.g.,<10�7 for a sandstone at ambient conditions), the material exhibits aninstantaneously reversible decrease of the resonance frequency, behavior dominated by classical nonlinearity.At higher strains (e.g.,>10�6 for a sandstone at ambient conditions), the decrease of the resonance frequencyis governed mostly by a process related to the nonequilibrium dynamics of the evolving strain fields withinthe rock. In summary, under a continuous harmonic excitation, the elastic moduli begin to soften and even-tually reach ametastable steady state; i.e., the elasticity of thematerial reaches a new but unstable equilibriumstate that is tuned to the dynamic strain amplitude induced by the excitation. Conversely, if elastic energy isno longer injected into the system, the material recovers back to its original elastic properties, with a rategenerally proportional to the logarithm of time (TenCate, Smith, et al., 2000; TenCate, Smith, & Guyer, 2000).Notably, conditioning times are much faster than recovery times (TenCate, 2011; TenCate, Smith, et al.,2000), and this temporal asymmetry gives rise to all sorts of interesting observations. For instance, up and downfrequency sweeps around the resonance frequency do not lead to the same resonance curves when thesweeping rates are identical but relatively fast. On the other hand, when the sweeping rates are slow enough,the upward and downward resonance curves merge. In between the two strain regimes described above,classical nonlinearity and nonequilibrium dynamics cannot be disentangled from each other in the contextof a resonance experiment where the effects from these mechanisms are smeared over many dynamic cycles.

Dynamic acoustoelastic testing (DAET) is a technique developed more recently, designed to delineate themechanisms of nonlinear and nonequilibrium dynamics with high sensitivity, regardless of the strain regime.This technique is based on a pump and probe scheme in which high-frequency (HF) pulses are used to probethe various phases of a mechanical system set by a low-frequency (LF) pump (Renaud et al., 2011; Rivièreet al., 2013). Typically, the sample is a long thin bar and the dynamic pump is obtained by exciting the samplenear a resonance mode of vibration, thus approaching the conditions used for NRUS. The probe is obtainedby tracking the relative change in the arrival times of HF pulses traveling across the shortest dimension of thebar, with the assumption that during the short travel of these pulses, the strain field induced by the pump isvarying so slowly that it can be assumed to be constant. Through a careful synchronization of the HF pulseemissions with the pump signal, a stroboscopic effect can be obtained to capture the nonlinear and nonequi-librium dynamics of the material over a full cycle of the LF pump (Lott et al., 2017). The analysis of the relativechange of the arrival times of the HF pulses as a function of the LF strain induced by the pump at the probelocation gives a complete representation of classical nonlinearity, with high sensitivity up to at least thesecond-order terms of nonlinearity, hysteresis, and nonequilibrium dynamics in the conditioning and recov-ery phases (Rivière et al., 2013). In Figure 2 of their paper (but also in Figure 3 of Renaud et al. (2012)), most ofthe drop in elasticity experienced by the rock andmeasured locally by the HF probes during the conditioningphase appears to follow instantaneously the envelope of the LF pump: a process that could be referred to as“fast dynamics.” When the pump reaches a steady state (between 0.05 and 0.1 s in Rivière et al. (2013)), thematerial continues to soften over ten of milliseconds: a process that could be referred to as “slow dynamics.”Likewise, when the pump is switched off and the material recovers back to its original properties, “fast” and“slow” dynamics can be observed clearly in the time series. These findings are consistent with those ofTenCate et al. (2000). Another result of interest is found in the study of Rivière et al. (2015), where the linkand proportionality between the material softening observed in the conditioning phase, the size of thehysteresis, and the frequency shift observed in the nonlinear resonance experiments was clearly established.This result was later confirmed by an independent set of experiments where it was demonstrated that, givenappropriate corrections to account for the fact that the pump and the probe are in orthogonal directions,the material softening obtained from DAET in the conditioning phase is equivalent to that obtained fromNRUS (Lott et al., 2016). Last, Lott et al. (2017) demonstrated experimentally that there are additionalsubtleties to consider that are related to the type of strain field induced by the LF pump (e.g., axial motionversus torsional motion) and the type of waves used as HF probes (e.g., P waves and S waves withvarious polarizations).

Up to this point, the modeling of hysteresis and nonequilibrium dynamics was not addressed. Betweenthe first theoretical description of hysteresis in NMEMs through the Preisach-Mayergoyz formalism byGuyer et al. (1995) and one of the most recent description developed by Pecorari (2015), there has beena considerable number of assumptions and models introduced with most reported and described by

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014258

REMILLIEUX ET AL. NONLINEAR PROPAGATION OF ELASTIC PULSE 3

Page 4: Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea … · 2017-11-27 · Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea Sandstone: A Detailed Look

Guyer and Johnson (2009). Likewise, nonequilibrium dynamics has been described phenomenologicallyusing many models, including the “soft-ratchet” model proposed originally by Vakhnenko et al. (2004) andrecently improved by Berjamin et al. (2017). It is also worth mentioning the damage viscoelastic model usedrecently by Lyakhovsky et al. (2009) to simulate nonlinear resonance experiments. Slow dynamics is notincluded in this model though, and it is not clear if the effects on the wave propagation that are summarizedin Figure 1 can be captured. To the authors’ knowledge, a unified model capable of capturing simultaneouslynonlinear and nonequilibrium effects observed in various types of experiments does not exist.

In this paper, we propose to revisit the experiment conducted by Ten Cate et al. (1996), this time taking dataover many more points on the surface, without contact, using a scanning 3-D laser vibrometer as a receiver.With this extensive high-fidelity data, we will actually be able to isolate and determine the relative contribu-tions of the mechanisms of classical nonlinearity, hysteresis, and nonequilibrium dynamics in a pulse propa-gation experiment. This paper is organized as follows. Section 2 gives the necessary theoretical backgroundon classical nonlinear elasticity, derived from the 1-D elasticity theory. Quantification of classical nonlinearity,without and with accounting for damping in the material, is given in section 3. In this section, the elastic wavepropagation with classical nonlinearity and attenuation is resolved numerically and compared to experimen-tal data to show that classical nonlinearity is not sufficient to model the full elastic behavior of the material.Section 4 examines and quantifies nonequilibrium dynamics in the context of this elastic pulse propagationexperiment. Findings are compared to the resonant bar experiment, which is how nonequilibrium dynamicshas been traditionally quantified. Section 5 concludes.

2. Theoretical Background

In the 1-D theory of classical nonlinear elasticity, the stress (σ) can be expressed as a power series with respectto strain (ε) as

σ ¼ M0ε 1þ βεþ δε2 þ…� �

(1)

where M0 is the linear elastic modulus and β and δ are, respectively, the coefficients of first- and second-order nonlinearity.

In the absence of energy dissipation and external forces, the propagation of an elastic wave in an infinite 1-Dmediumwith first-order nonlinearity (β coefficient only) is governed by the following equation (McCall, 1994):

∂2u x; tð Þ∂t2

� c2∂2u x; tð Þ

∂x2¼ βc2

∂∂x

∂u x; tð Þ∂x

� �2" #

(2)

where u is the displacement and c is the wave speed. Note that geometric nonlinearity is ignored in this equa-tion as it is assumed that material nonlinearity is dominant. In the frequency domain, this results in thegeneration of harmonics. The nonlinear coefficient β can be retrieved from a measurement of the displace-ment amplitude U2 at the second harmonic generated at a distance x from a pure tone source signal as(Meegan et al., 1993)

β ¼ 2U2c2

U21ω2x

(3)

where ω is the angular frequency (ω = 2πf) and U1 is the displacement amplitude at the fundamentalfrequency. In experiments, we often use acceleration data measured with accelerometers or particle-velocitydata measured with laser vibrometers. In the frequency domain, the magnitude of the particle velocity V isrelated to the magnitude of the displacement U as, U = V/ω. Equation (3) can then be adapted to particlevelocity measurements using a factor ω:

β ¼ 2V2c2

V21ωx

(4)

where V1 and V2 are now the magnitudes of the fast Fourier transform (FFT) of the particle velocity at thefundamental frequency and at the second harmonic, respectively.

Energy dissipation can easily be introduced in equation (2) by adding a term used in Burger’s equation(Hamilton & Blackstock, 1998):

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014258

REMILLIEUX ET AL. NONLINEAR PROPAGATION OF ELASTIC PULSE 4

Page 5: Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea … · 2017-11-27 · Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea Sandstone: A Detailed Look

∂2u x; tð Þ∂t2

� b∂2

∂x2∂u x; tð Þ

∂t

� �� c2

∂2u x; tð Þ∂x2

¼ βc2∂∂x

∂u x; tð Þ∂x

� �2" #

(5)

In this equation, b is a damping coefficient expressed as

b ¼ c2

ωQ(6)

where Q is the quality factor (i.e., inverse of loss factor).

The amplitude of a monochromatic wave is attenuated with distance as

V xð Þ ¼ V0 exp �γxð Þ (7)

where γ is the attenuation coefficient. This attenuation coefficient should be measured as a function offrequency, which we did not do in the experiments reported in this paper. The alternative is to use thedamping model proposed in equations (5) and (6). Using this model, the attenuation coefficient can beexpressed as

γ ¼ bω2

2c3(8)

Nonlinear elasticity and damping are competing effects in the context wave propagation. While nonlinearelasticity tends to generate harmonics, damping will attenuate these harmonics since attenuation increaseswith frequency. Such competition can be quantified by the Gol’dberg number (Gol’dberg, 1957; Hamilton &Blackstock, 1998), which is a dimensionless parameter measuring the strength of the nonlinearity relative tothat of dissipation. More specifically, it is the ratio of the absorption length (inverse of absorption coefficient)for a small-amplitude, linear signal at the fundamental frequency to the shock formation distance and can beexpressed using the parameters defined previously as

Γ ¼ βV1ωγc2

(9)

Essentially, if this number is much smaller (respectively, larger) than 1, the propagation is dominated bydamping (respectively, nonlinearity). When this number is close to 1, there is a strong competition betweennonlinearity and damping during the propagation.

3. Quantifying and Modeling Classical Nonlinearity3.1. Experimental Setup

The setup used in the pulse propagation experiments is shown in Figure 2. Experiments were conducted on asample of Berea sandstone (Cleveland Quarries, Amherst, OH) with a length of 1794 mm (70.63 in.) and adiameter of 39.6 mm (1.56 in.), in fact the same sample used by Ten Cate et al. (1996). The sample restedon foam to approach free (unconstrained) boundary conditions. Elastic waves were generated in the sampleusing a PZT-5A (Navy type II) piezoelectric disk with a diameter of 45 mm (1.77 in.) and a thickness of 10 mm(0.394 in.). The transducer and an inertial backload were epoxied onto one flat end of the sample. Figure 3shows the time history and magnitude of the FFT of the normalized source signal used in the experiments.The impulsive signal initially consisted of a sinusoidal wave with a frequency of 22.4 kHz tapered at the startand end by half-Gaussian windows. The output signal from the signal generation card (internal arbitrarywaveform generator in Polytec PSV-3D-500 scanning laser vibrometer) was amplified 20 times by a voltageamplifier (TEGAM 2350) before being used as input signal to the transducer. The propagation of the elasticpulse was recorded at 119 points along a line, on the surface of the sample, with a regular spacing of approxi-mately 5 mm between the points, starting at 10 mm from the source and extending 600 mm from the source.The Cartesian components of the particle velocity were recorded on the surface of the sample with a PolytecPSV-3D-500 scanning laser vibrometer. Data were acquired with a sampling frequency of 2.56 MHz. At eachscanning point, waveforms were averaged 500 times to increase the signal-to-noise ratio. Each signal wasacquired for a duration of 6.4 ms (16,384 samples): response to intermittent sine bursts with approximate

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014258

REMILLIEUX ET AL. NONLINEAR PROPAGATION OF ELASTIC PULSE 5

Page 6: Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea … · 2017-11-27 · Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea Sandstone: A Detailed Look

duration of 0.6 ms and intervals of 5.8 ms. Note that the short duration between acquisitions leads toconditioning effects. However, this experiment focuses on the quantification of classical nonlinearity andconditioning effects will be ignored.

3.2. Experimental Results and Data Analysis

Figure 4 shows three selected time histories andmagnitudes of the FFT of the axial component of the particlevelocities for a source signal with a peak amplitude of 16 Vpp (before amplification) at various distances fromthe source. The time series have been truncated to show only the direct pulse; the reflection from the far end,opposite to the source, is also recorded but does not overlap with the direct pulse. The time series have beensynchronized so that they can be plotted on the same time scale, and the calculation of the FFT is not affectedby the arrival time of the pulse at a measurement point. A full movie of the pulse propagation (both axial andradial components) recorded over the 119 measurement points is provided in the supporting information.The first waveform and corresponding FFT show that the vibrational response measured at 10 mm fromthe source is similar to the source signal, with some minor differences caused by the frequency responseof the piezoelectric transducer—even though the transducer is operated well away from its first resonancefrequency, its frequency response is not flat and modulates the source signal. More importantly, the vibra-tional response at this location does not exhibit a significant harmonic content, which attests to the linearityof the source and quality of the bond between the source and the sample. The next two sets of plots show

Figure 2. Experimental setup for monitoring the propagation of an elastic pulse in the sample of Berea sandstone.

Figure 3. Normalized electrical source signal used in the pulse propagation experiments: (a) time history and (b) magnitude of the FFT. The signal consists of a sinewave with frequency of 22.4 kHz, tapered at the start and end by half-Gaussian windows.

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014258

REMILLIEUX ET AL. NONLINEAR PROPAGATION OF ELASTIC PULSE 6

Page 7: Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea … · 2017-11-27 · Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea Sandstone: A Detailed Look

how the sinusoidal waveform distorts with propagation distance, exhibiting growth of the second harmonic(i.e., peaks observed at 44.8 kHz) and increasing higher-frequency content (i.e., between 60 and 80 kHz). Theevolution of the waveform with propagation distance is consistent with analytical predictions on the effectsof classical nonlinearity and hysteretic nonlinearity: the β term in the classical theory of nonlinearity inequation (1) induces a steepening of the waveform (eventually leading to an N wave in the absence ofdamping), whereas hysteretic nonlinearity induces a triangulation of the waveform, as shown in Figure 1.Note that Ten Cate et al. (1996) used a higher-order term of classical nonlinearity, i.e., δ term in equation (1),to model the observed triangulation of the waveform. At relatively small amplitudes, the distortion producedby this term results in what appears to be a triangulation. However, as the amplitude increases, the originalsinusoidal waveform is bowed and not triangulated (Kadish, TenCate, & Johnson, 1996). With this term, theamplitude at the third harmonic (3f0) depends on the cube of the amplitude at the fundamental frequency(f0), whereas, as will be shown in section 4, we observe a quadratic dependence, as expected for hystereticnonlinearity (Van Den Abeele et al., 2000). In 1996, theories of hysteretic nonlinearity and nonequilibriumdynamics in geomaterials (and experimental evidence supporting these theories) were just emerging andcould not be used in the analysis so it seemed natural to use the classical description of nonlinearity to modelthe experimental observations.

Figure 4. Axial component of the particle velocity measured at representative locations along the propagation path of the impulsive wave for a source signal at22.4 kHz. (left column) Time histories and (right column) magnitudes of the FFT. (top row) 10 mm from the source. (middle row) 335 mm from the source.(bottom row) 590 mm from the source.

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014258

REMILLIEUX ET AL. NONLINEAR PROPAGATION OF ELASTIC PULSE 7

Page 8: Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea … · 2017-11-27 · Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea Sandstone: A Detailed Look

Classical nonlinearity was quantified by reducing the spectral data at all 119 measurement points with theexpression given in equation (4). Figure 5a shows the variation of 2V2c

2 as a function of V12ωx, based on a

wave speed of 2010 m/s. First, note that there is an offset at V12ωx = 0. This offset is not due to the nonlinear-

ity of the system but to the finite duration of the pulse (i.e., bandwidth). The term 2V2c2 is calculated from the

spectrum amplitude at 44.8 kHz (second harmonic amplitude) at various distances from the source. For thesource signal (see Figure 3), which is a linear signal, the amplitude at the second harmonic (i.e., at44.8 kHz) is 2 orders of magnitude smaller than the amplitude at the fundamental frequency but is not zero.The same applies to the signals recorded on the sample very close to the source. For this analysis, what is ofinterest is not the offset at V1

2ωx = 0 but the growth of the second harmonic amplitude with propagationdistance, which is quantified by the slope of the data reported in Figure 5a. The slope of the linear fit throughthis data set is β = 370. Experiments conducted years earlier on the same sample by Ten Cate et al. (1996) ledto β = 400. This close agreement is remarkable given that the source, source signals, and receivers were dif-ferent in the two experiments. In the work of Ten Cate et al. (1996), the elastic pulse generated by the sourcewas centered at 12.4 kHz and the vibrational signals were recorded by a small number of accelerometersalong the pulse propagation. This agreement also indicates that the material properties of the sample havenot changed significantly over the long time separating the two experiments. In both studies, however, theeffects of damping are ignored in the quantification of classical nonlinearity.

It is possible to correct the measured spectra by the amount of damping in the system using the model givenin equations (5)–(8). Figure 5b shows the variation of 2V2c

2 as a function of V12ωx for the 119 scanning points,

where the amplitudes at the fundamental frequency and second harmonic have been corrected for dampingbased on a quality factorQ = 60. The quality factor was estimated from the width (full width at half maximum)of the resonance curve of a purely longitudinal mode near 20 kHz in the current experiment. The slope of thelinear fit of this corrected data is β = 660.

Last, it is worth mentioning the high variability of the data around the linear fit for V12ωx> 5 in Figure 5a and

V12ωx> 10 in Figure 5b. In this experiment, the Gol’dberg number is Γ = 0.6 (using c = 2010 m/s, V1 = 15 mm/

s near the source, f0 = 22.4 kHz, Q = 60, and β = 660). This value is close to 1 and suggest that the competitionbetween nonlinearity and damping is strong, which results in data variability after a certain distance. Suchvariability would likely be reduced if the experiment had been conducted in a sample with much smallerdamping (e.g., metallic component with distributed damage).

Up to this point, dispersion and its effects on the wave propagation were not discussed. The movie of thewave propagation provided in the supporting information shows that double peaks appear and disappearin the waveform along the propagation path. It seems that not all frequency components are traveling atthe same speed. This is most likely caused by dispersion, which affects the wave propagation in thefrequency range of the third harmonic and above. This is also why the amplitude spectra do not show a clearpeak around the third harmonic. To mitigate the dispersion effects, the experiment described above wasrepeated but this time using a source signal with a lower frequency, namely, 15.7 kHz. The time historyand corresponding magnitude of the FFT of the axial component of the particle velocity measured at590 mm from the source, for this source frequency, are shown in Figure 6. The amplitude spectrum now

Figure 5. The variation of 2V2c2 as a function of V1

2ωx, the slope of which is β. This is a postprocessing of the data set corresponding to the spectral data reported inFigure 4. (a) Without any correction for damping, (b) where the amplitudes at the fundamental frequency and second harmonic have been corrected for damping,based on a quality factor Q = 60.

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014258

REMILLIEUX ET AL. NONLINEAR PROPAGATION OF ELASTIC PULSE 8

Page 9: Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea … · 2017-11-27 · Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea Sandstone: A Detailed Look

exhibits a clear peak at the third harmonic. Also of interest is the fact that the amplitude at the third harmonicgrows faster than the amplitude at the second harmonic. The piezoelectric transducer is less efficient at15.7 kHz than at 22.4 kHz, and consequently, the amplitude of the elastic wave generated in the sampleusing the same input voltage is about twice as small. Because of the smaller frequency and smaller waveamplitude in this experiment, the Gol’berg number decreased to Γ = 0.2 (using c = 2010 m/s, V1 = 7.5 mm/snear the source, f0 = 15.7 kHz, Q = 60, and β = 660) and damping tends to dominate the propagation. Theelastic wave does not have a sufficient amplitude for efficient second harmonic generation, which wouldindicate that classical nonlinearity has been activated, but it is sufficient to activate hysteretic nonlinearity.We could then study the effects of amplitude and frequency on the activation of the various mechanismsbut this is beyond the scope of this paper. As a more modest achievement, we repeated the experiment at17.1 kHz. The axial component of the particle velocity measured at 590 mm from the source at thisfrequency is shown in Figure 7. The peak at the third harmonic is not as clear as previously becausedispersion starts to play a role in the wave propagation but the amplitude at the second harmonic is largerthan previously thanks to the larger source amplitude and higher frequency.

3.3. Modeling the 1-D Wave Propagation With Classical Nonlinearity and Damping

The simple 1-D elastic wave equation with damping given in equation (5) is now used to model the experi-ments to see how well it fits the data. We use the following parameters in the model: a wave speedc = 2010 m/s, a damping coefficient b = 0.525 (based on a quality factor Q = 60 and a center frequency ofthe impulsive source signal f = 22.4 kHz), and a nonlinear parameter β = 660. The wave equation is projectedonto a finite-element space using the “Mathematics, General Form PDE” module of the software packageCOMSOL Multiphysics version 5.2a. The geometry consists of a line with a length of 1794 mm. One boundaryis imposed a flux condition (source) while the opposite boundary is free and so imposed a Neumann condi-tion, i.e., ∂u(x, t)/∂x= 0. The impulsive source signal is the measured signal depicted in Figure 4 for a receivertaken at 10 mm from the source. This signal is normalized to have a peak amplitude of 40.5. This approachallows us to take into account the response of the transducer in the 1-D model. The computational domain

Figure 6. Axial component of the particle velocity measured at 590mm from the source for a source signal at 15.7 kHz. (left) Time histories and (right) magnitudes ofthe FFT.

Figure 7. Axial component of the particle velocity measured at 590mm from the source for a source signal at 17.1 kHz. (left) Time histories and (right) magnitudes ofthe FFT.

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014258

REMILLIEUX ET AL. NONLINEAR PROPAGATION OF ELASTIC PULSE 9

Page 10: Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea … · 2017-11-27 · Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea Sandstone: A Detailed Look

is discretized into Lagrange quadratic elements with a uniform length of 3 mm, which ensures at least sixelements per wavelength up to 110 kHz. The solution is integrated in time with a time step of 100 ns usingthe generalized α method (Chung & Hulbert, 1993), which is a fully implicit, unconditionally stable, andsecond-order accurate method with control over the dissipation of spurious high-frequency modes. The timestep is small enough to satisfy the Courant-Friedrichs-Lewy stability condition.

Figure 8 (left) shows six typical measured and simulated time histories of the axial component of the particlevelocity at various positions along the sample. The agreement between the simulations and the experimentsis excellent at all points. The model captures the arrival times, amplitudes, and overall shapes of the wave-forms observed in the experiments, for both the direct and reflected pulses. Some small discrepancies areobserved and expected in the later portion of the reflected pulses at some positions because the modelignores the presence of the source and backload and resulting interaction with the reflected pulse. The timeseries depicted in Figure 8 (left) are now truncated and synchronized to retain only the direct, first arrivalpulses. The measured and simulated magnitudes of the FFT of the direct pulses are shown in Figure 8 (right).The FFTs again show just how well the simple model works; the growth of the second harmonic is very wellcaptured. In fact, after correcting the amplitudes at the fundamental frequency and second harmonic fordamping, the variation of 2V2c

2 as a function of V12ωx for the six simulated signals are aligned perfectly with

the linear fit shown in Figure 5b, as expected. However, it is quite clear that very soon (~65 mm), the modelfails to capture the third harmonic and higher-frequency behavior between 60 and 80 kHz. This is expectedsince the model does not include terms related to hysteretic nonlinearity and does not include the effectsof dispersion.

Practically, at 22.4 kHz, the dynamic behavior of the rock is dominated by classical nonlinearity and attenua-tion while hysteresis and other nonequilibrium effects are far less important. As shown in the previoussection, this is not necessarily true at lower frequencies where the source amplitude is smaller. With thequality of the data taken in these experiments, it is instructive to see at what point the effects of nonequili-brium dynamics and hysteresis do start to become apparent.

4. Quantifying Nonequilibrium Dynamics4.1. Experimental Setup

An alternative to the experiment reported in the previous section is to monitor the evolution of the waveformat a fixed position on the sample for various source amplitudes. In essence, this is a similar experiment to thatreported by Cabaret et al. (2015), where the propagation of torsional pulses is studied in a 1-D chain of 70beads. Disordered bead packs and granular lattices serve as simplified paradigms for understanding someof the mechanisms responsible for nonlinear and nonequilibrium dynamics. In their study, the physicalmechanism at stake is the friction between the beads. Consequently, the wave propagation is mostly affectedby hysteretic nonlinearity and is very well captured by a model based on the Preisach-Mayergoyz formalism.Here we focus on the propagation of a longitudinal elastic wave in a disordered consolidated granularmaterial. The nonlinear and nonequilibrium effects involved in the wave propagation in such a mediumare not only produced by friction between the constituents but also possibly by clapping and othermicroscopic interactions.

The same experimental setup and source signal (see Figures 2 and 3) as in the previous section are used forthis experiment. Waveforms were also averaged 500 times but the duration between acquisitions was set to320 ms to allow the sample to start recovering back to its original elastic properties between each average.Figure 9 shows the time histories of the axial component of the particle velocity measured at 500 mm fromthe source for 20 source amplitudes ranging from 1 to 20 Vpp in steps of 1 Vpp (before amplification). Theoriginal sinusoidal waveform observed at the lowest amplitudes progressively evolves into a triangular wave,as a result of hysteresis, with additional distortion caused by classical nonlinearity and dispersion, asdiscussed in the previous section. Most notably, the waveforms experience a significant delay in arrival timeas the source amplitude increases. This delay is observed not only at the extrema (peaks and troughs) but alsoat the zero crossings. Classical nonlinearity would not induce any delay at the zero crossings (where the strainis equal to zero) because it produces an instantaneous variation of the modulus without time constantsinvolved. This delay is a direct consequence of nonequilibrium dynamics, specifically the strain-inducedmaterial softening experienced during the conditioning phase. Note that, because of the material

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014258

REMILLIEUX ET AL. NONLINEAR PROPAGATION OF ELASTIC PULSE 10

Page 11: Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea … · 2017-11-27 · Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea Sandstone: A Detailed Look

Figure 8. Measured and simulated axial components of the particle velocity at various positions along the sample: (left) time histories and (right) magnitudes of theFFT. In the waveforms, the time window is long enough to show the direct and reflected pulses. Only the direct pulses are used to compute the FFTs.

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014258

REMILLIEUX ET AL. NONLINEAR PROPAGATION OF ELASTIC PULSE 11

Page 12: Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea … · 2017-11-27 · Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea Sandstone: A Detailed Look

softening and the corresponding decrease of the wave speed from the source to the measurement point, thepulse is stretched. This stretch is quantified by monitoring the fundamental frequency f0, which decreases by0.53% from the smallest to the largest source amplitude.

The data shown in Figure 9 can be further reduced to quantify nonequilibrium dynamics, by estimating thetime delay as a function of strain amplitude at the zero crossings of the waveform, as in the work of Cabaretet al. (2015). The waveform measured at the lowest source amplitude is used as a reference. By definition, atthe zero crossings, the strain amplitude is equal to 0. For the analysis, the strain amplitude experienced by thesample at the strain extremum (peak or trough) preceding (1/4 of cycle before) the zero crossing of interest isconsidered. A total of 28 sets of zero crossing points are considered. In each set, 20 source amplitudes areused for the analysis, with the lowest source amplitude used as a reference and the higher amplitudescompared to this reference.

The relative time delay between the signals also provides the relative change in the longitudinal wave speed,Δci/0/c0, where the superscript “i” denotes the ith source amplitude and “0” the lowest source amplitude forwhich the sample is assumed to behave linearly. Finally, at the perturbation level, the relative change in theYoung’s modulus E (the modulus involved in the propagation of a longitudinal wave in a long thin bar) isrelated to the relative change in the speed of the longitudinal wave as

ΔEi=0=E0 ¼ 2Δci=0=c0 (10)

The relative change in the elastic modulus over the propagation path of the waveform can be tracked as afunction of the maximum strain amplitude at the measurement point. Along the axial direction (x direction),the strain component of interest can be expressed as

εxx ¼ vxc

(11)

Reduced data from Figure 9 are shown in Figure 10. Three distinct regimes can be identified in the waveform:(i) the earlier portion of the waveform, where the dynamic strain at the measurement point transitions fromzero to steady state; (ii) the center portion of the waveform, where the absolute value of the peak amplitudeof the dynamic strain is quasi-constant (i.e., steady state); and (iii) the later portion of the waveform where thedynamic strain at the measurement point transitions from steady state back to zero, i.e., a path similar to theearlier portion of the waveform but in reverse. In the earlier portion of the waveform, the elastic modulusdecreases progressively from its undisturbed equilibrium value to a value 4.1% smaller. The decrease ofthe elastic modulus is an instantaneous effect produced by the hysteresis and referred to as fast dynamics insection 1. It was already evidenced in DAET experiments (e.g., Rivière et al., 2013), in the propagation of atorsional elastic pulse in a 1-D granular chain (Cabaret et al., 2015), and in the interaction between two elasticwaves in a rock sample (TenCate et al., 2016). In the center portion of the waveform, the evolution of thematerial softening with the strain amplitude is essentially independent of the wave cycle selected for the

Figure 9. Time histories of the axial component of the particle velocity measured at 500 mm from the source for 20 amplitudes of a pulse centered at 22.4 kHz.

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014258

REMILLIEUX ET AL. NONLINEAR PROPAGATION OF ELASTIC PULSE 12

Page 13: Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea … · 2017-11-27 · Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea Sandstone: A Detailed Look

analysis. In other words, the pulse is not long enough to observe the effects of slow dynamics in theconditioning phase. In the conditioning phase, the elasticity of the material reaches a new (but unstable)equilibrium state where the modulus is 4.9% smaller than in the undisturbed (stable) equilibrium. Notethat, in this steady state regime, the dependence of material softening seems to intensify once the strainreaches a value of 4 to 5 microstrains. In the later portion of the waveform, the evolution of the elasticmodulus follows a different path from that observed in the earlier portion: recovery of the undisturbedequilibrium elastic modulus occurs at a much slower rate than material softening. This result from a pulsepropagation experiment is in excellent qualitative agreement with the results obtained in resonant barexperiments by TenCate, Smith, et al. (2000); the two experiments will be compared quantitatively in thenext section. The experiment was repeated at other positions on the sample. Raw and reduced data at650 mm from the source are shown in Figures 11 and 12. These data lead to the same conclusions, bothqualitatively and quantitatively, as those reached at 500 mm from the source.

The same experiment was repeated at 15.7 kHz to measure the delays at the zero crossings as a function ofthe strain amplitude but also to quantify the growth of the odd-harmonic amplitudes with respect to theamplitude at the fundamental frequency. The time histories of the axial component of the particle velocityfor 20 source amplitudes at 500mm from the source are shown in Figure 13. The corresponding reduced dataare shown in Figure 14. Unlike previously, in this experiment where the strain amplitude is much smaller

Figure 10. Relative change in elasticity as a function of strain estimated from the propagation of a pulse centered at 22.4 kHz and measured at 500 mm from thesource, for 20 source amplitudes. The red star, blue circle, and green cross symbols correspond to the first five, 9th to 21st, and last five zero-crossings of thewaveforms, respectively, as indicated by the inset. A total of 23 curves are plotted in the figure from the reduced data. Each curve consists of 19 points (20 amplitudeswith one amplitude, the lowest, used as a reference for the calculation of the time delays).

Figure 11. Time histories of the axial component of the particle velocity measured at 650 mm from the source for 20 amplitudes of a pulse centered at 22.4 kHz.

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014258

REMILLIEUX ET AL. NONLINEAR PROPAGATION OF ELASTIC PULSE 13

Page 14: Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea … · 2017-11-27 · Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea Sandstone: A Detailed Look

throughout the pulse duration, there is no clear distinction between the early, steady state, and late portionsof the waveform. In other words, the effects of slow dynamics are not visible. The peak strain amplituderemains below 4 microstrains and the material softens by up to 1.8%. It is likely that strain needs to reacha threshold amplitude for the material to exhibit unambiguously the effects of slow dynamics. The same dataset is now analyzed in the frequency domain, with amplitude spectra shown in Figure 15. The spectra showthat there is almost no contribution from the even (second and fourth) harmonics but a significant growth ofthe third and fifth harmonics. The dependence of the odd harmonics with the amplitude at the fundamentalfrequency is a quadratic one, which indicates that the behavior we observe in this particular experiment ismostly due to hysteretic nonlinearity (see Figure 1).

4.2. Quantitative Comparison With a Resonant Bar Experiment

To see how well the pulse propagation measurements can quantify nonequilibrium dynamics, a standardresonant bar experiment was performed for comparison. The arrangement of the resonant bar experimentis depicted in Figure 16. The piezoelectric transducer described previously was digitally swept up with asequence of harmonic voltage signals. Each harmonic signal was applied for 55 ms and the transient vibra-tional response was recorded during the last 40 ms of the source signal to ensure that steady state conditionshad been reached. Vibrational spectra were constructed from the harmonic responses, using heterodyne

Figure 12. Relative change in elasticity as a function of strain estimated from the propagation of a pulse centered at 22.4 kHz and measured at 650 mm from thesource, for 20 source amplitudes. The red star, blue circle, and green cross symbols correspond to the first five, 9th to 21st, and last five zero-crossings of thewaveforms, respectively, as indicated by the inset. A total of 23 curves are plotted in the figure from the reduced data. Each curve consists of 19 points (20 amplitudeswith one amplitude, the lowest, used as a reference for the calculation of the time delays).

Figure 13. Time histories of the axial component of the particle velocity measured at 500 mm from the source for 20 amplitudes of a pulse centered at 15.7 kHz.

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014258

REMILLIEUX ET AL. NONLINEAR PROPAGATION OF ELASTIC PULSE 14

Page 15: Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea … · 2017-11-27 · Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea Sandstone: A Detailed Look

processing. The experiment was conducted for frequencies ranging between 0.3 and 7 kHz in steps of 2.5 Hzand at 22 excitation amplitudes ranging between 0.25 and 10 Vpp (before amplification). The source signalswere generated by a function generator (National Instrument PXI-5406) and amplified 20 times by a voltageamplifier (TEGAM 2350). The axial component of the acceleration was measured on the flat end opposite tothe source by an accelerometer (PCB Piezotronics 352A60) connected to a signal conditioner (PCBPiezotronics 482C Series). The signal measured by the accelerometer was digitized with a sampling rate of10 MHz (National Instrument PXI-5122).

Any resonance mode can be selected to quantify nonequilibrium dynamics as long as the mode type (eitherpurely longitudinal or purely torsional) is unchanged, as recently demonstrated by Remillieux et al. (2016).Here we focus on the longitudinal modes to be consistent with the pulse propagation experiments. Thevibrational spectra for this experiment are shown in Figure 17. Material softening is observed when the driveamplitude of the source becomes sufficiently large. For a given drive amplitude, such softening can be quan-tified by plotting the relative frequency shift as a function of the maximum strain in the sample, the slope ofwhich is the nonlinear parameter α. The maximum strain in the sample may be inferred analytically from themeasured vibrational response. For the longitudinal modes, the strain component of interest is εxx, where x isthe direction given by the axis of symmetry of the bar. For a system with a 1-D-like geometry and

Figure 14. Relative change in elasticity as a function of strain estimated from the propagation of a pulse centered at 15.7 kHz and measured at 500 mm from thesource, for 20 source amplitudes. The red star, blue circle, and green cross symbols correspond to the first five, 9th to 21st, and last five zero-crossings of thewaveforms, respectively, as indicated by the inset. A total of 23 curves are plotted in the figure from the reduced data. Each curve consists of 19 points (20 amplitudeswith one amplitude, the lowest, used as a reference for the calculation of the time delays).

Figure 15. Magnitudes of the FFT of the axial component of the particle velocity measured at 500 mm from the source for 20 amplitudes of a pulse centered at15.7 kHz.

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014258

REMILLIEUX ET AL. NONLINEAR PROPAGATION OF ELASTIC PULSE 15

Page 16: Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea … · 2017-11-27 · Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea Sandstone: A Detailed Look

unconstrained boundaries, the expression given in equation (11) can also be used in the context of a reso-nance experiment to relate the maximum amplitude of the axial component of the particle velocity at thefree end of the sample (where data are acquired) to the maximum amplitude of the axial component ofthe strain in the sample.

Figure 18 shows the shift in resonance frequency at the jth source amplitude with respect to the smallestsource amplitude as a function of the amplitude of the axial strain εxx. This evolution is very similar to thatobserved in Figures 10 and 12 in the context of a pulse propagation experiment. The relative frequency shiftvaries almost linearly with the maximum strain beyond 4 microstrains. This linear dependence is moreobvious when higher strain amplitudes can be reached in the experiment (typically on the order of 15 to20 microstrains). When a mode has enough points beyond this strain value of 4 microstrains, the pointsare fitted linearly and the slope of this fit is calculated. It appears that material softening converges to a singlevalue (all curves superimpose) of α = 5260 ± 160 for the modes L6 through L12. Below the sixth resonancemode, the elastic response does not have a sufficiently large amplitude to reach the threshold strain valueof 4 microstrains.

The slope of the relative change of the resonance frequency is approximately twice as small as the relativechange of the Young’s modulus observed in the pulse propagation experiment, which is consistent with

Figure 16. Experimental setup for conducting nonlinear resonant ultrasound spectroscopy (NRUS) on the sample of Berea sandstone.

Figure 17. Vibrational spectra measured on the sample of Berea sandstone at 22 source amplitudes. The first 12 modes of longitudinal vibration L1 through L12 areindicated. The red star symbols denote the location of the resonances for all source amplitudes. The resonance frequencies are not computed for the first threemodes because of the poor signal-to-noise ratio.

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014258

REMILLIEUX ET AL. NONLINEAR PROPAGATION OF ELASTIC PULSE 16

Page 17: Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea … · 2017-11-27 · Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea Sandstone: A Detailed Look

the analytical relationship between Young’s modulus and resonance frequency of a longitudinal mode at theperturbation level:

ΔE=E0 ¼ 2Δf=f 0 (12)

5. Conclusions

We demonstrated that monitoring the propagation of a finite-amplitude elastic pulse in a rock can be used toisolate the mechanisms of classical nonlinearity, hysteresis, and nonequilibrium dynamics at time scales lessthan 1 ms. Such experiment contrasts with the complexity of the pump and probe scheme in a dynamicacousto-elastic experiment and with the incomplete description of nonlinearity in a resonant bar experiment.Note that there are still some great benefits to conducting these experiments, for instance, to study slowdynamics at much longer time scales (say, minute to hours). From a modeling point of view, the significanceof this experiment lies in its simplicity: a 1-D-like geometry with only one source frequency (or time scale)involved. It is our hope that the experimental results reported in this paper will be useful to the developmentand validation of a truly universal model of the NMEM dynamics.

ReferencesAdams, F. D., & Coker, E. G. (1906). An investigation into the elastic constants of rocks, more especially with reference to cubic compressibility,

Publ. (Vol. 46). Washington, DC: Carnegie Inst. of Washington.Aleshin, V., Gusev, V., & Zaitsev, V. Y. (2004). Propagation of acoustics waves of nonsimplex form in a material with hysteretic quadratic

nonlinearity: Analysis and numerical simulations. Journal of Computational Acoustics, 12(03), 319–354. https://doi.org/10.1142/S0218396X04002328

Berjamin, H., Favrie, N., Lombard, B., & Chiavassa, G. (2017). Nonlinear waves in solids with slow dynamics: An internal-variable model.Proceedings of the Royal Society A, 473(2201), 20170024. https://doi.org/10.1098/rspa.2017.0024

Cabaret, J., Béquin, P., Theocharis, G., Andreev, V., Gusev, V. E., & Tournat, V. (2015). Nonlinear hysteretic torsional waves. Physical ReviewLetters, 115(5), 054301. https://doi.org/10.1103/PhysRevLett.115.054301

Chung, J., & Hulbert, G. M. (1993). A time integration algorithm for structural dynamics with improved numerical dissipation: Thegeneralized-α method. Journal of Applied Mechanics, 60(2), 371–375. https://doi.org/10.1115/1.2900803

Gol’dberg, Z. A. (1957). On the propagation of plane waves of finite amplitude. Soviet Physics: Acoustics, 3, 340–347.Gusev, V. (2000). Propagation of acoustic pulses in material with hysteretic nonlinearity. The Journal of the Acoustical Society of America,

107(6), 3047–3058. https://doi.org/10.1121/1.429333Gusev, V., Glorieux, C., Lauriks, W., & Thoen, J. (1997). Nonlinear bulk and surface shear acoustic waves in materials with hysteresis and end-

point memory. Physics Letters A, 232(1-2), 77–86. https://doi.org/10.1016/S0375-9601(97)00357-5Guyer, R. A., & Johnson, P. A. (1999). Nonlinear mesoscopic elasticity: Evidence for a new class of materials. Physics Today, 52(4), 30–36. https://

doi.org/10.1063/1.882648Guyer, R. A., & Johnson, P. A. (2009). Nonlinear mesoscopic elasticity: The complex behavior of granular media including rocks and soil (p. 410).

Weinheim: Wiley-VCH Verlag. https://doi.org/10.1002/9783527628261Guyer, R. A., McCall, K. R., & Boitnott, G. N. (1995). Hysteresis, discrete memory, and nonlinear wave propagation in rock: A new paradigm.

Physical Review Letters, 74(17), 3491–3494. https://doi.org/10.1103/PhysRevLett.74.3491Hamilton, M. F., & Blackstock, D. T. (Eds.) (1998). Nonlinear acoustics: Theory and applications (455 pp). San Diego, CA: Academic Press.

Figure 18. Relative frequency shift as a function of the strain component εxx for the longitudinal modes L4 through L12.

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014258

REMILLIEUX ET AL. NONLINEAR PROPAGATION OF ELASTIC PULSE 17

AcknowledgmentsWe wish to thank Martin Lott, Pierre-Yves Le Bas, and Paul Johnson for theirvaluable discussions and suggestions.The current policies of Los AlamosNational Laboratory do not allow thepublic release of experimental datapresented in this study. For more infor-mation, please contact the authors.

Page 18: Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea … · 2017-11-27 · Propagation of a Finite-Amplitude Elastic Pulse in a Bar of Berea Sandstone: A Detailed Look

Holcomb, D. J. (1981). Memory, relaxation, and microfracturing in dilatant rock. Journal of Geophysical Research, 86(B7), 6235–6248. https://doi.org/10.1029/JB086iB07p06235

Johnson, P. A., Zinszner, B., & Rasolofosaon, P. N. J. (1996). Resonance and elastic nonlinear phenomena in rock. Journal of GeophysicalResearch, 101(B5), 11,553–11,564. https://doi.org/10.1029/96JB00647

Kadish, A., TenCate, J. A., & Johnson, P. A. (1996). Frequency spectra of nonlinear elastic pulse-mode waves. The Journal of the AcousticalSociety of America, 100(3), 1375–1382. https://doi.org/10.1121/1.415984

Landau, L. D., & Lifshitz, E. M. (1986). Theory of elasticity. In Course of theoretical physics (Vol. 7, Chap. 3, pp. 87–107). New York: Pergamon Press.Lott, M., Remillieux, M. C., Garnier, V., Le Bas, P.-Y., Ulrich, T. J., & Payan, C. (2017). Nonlinear elasticity in rocks: A comprehensive three-

dimensional description. Physical Review Materials, 1(2), 023603. https://doi.org/10.1103/PhysRevMaterials.1.023603Lott, M., Remillieux, M. C., Le Bas, P.-Y., Ulrich, T. J., Garnier, V., & Payan, C. (2016). From local to global measurements of nonclassical nonlinear

elastic effects in geomaterials. The Journal of the Acoustical Society of America, 140(3), EL231–EL235. https://doi.org/10.1121/1.4962373Lyakhovsky, V., Hamiel, Y., Ampuero, J.-P., & Ben-Zion, Y. (2009). Non-linear damage rheology and wave resonance in rocks. Geophysical

Journal International, 178(2), 910–920. https://doi.org/10.1111/j.1365-246X.2009.04205.xMcCall, K. R. (1994). Theoretical study of nonlinear elastic wave propagation. Journal of Geophysical Research, 99, 99(B2), 2591–2600. https://

doi.org/10.1029/93JB02974Meegan, G. D. Jr., Johnson, P. A., Guyer, R. A., & McCall, K. R. (1993). Observations of nonlinear elastic wave behavior in sandstone. The Journal

of the Acoustical Society of America, 94(6), 3387–3391. https://doi.org/10.1121/1.407191Murnaghan, F. D. (1937). Finite deformation of an elastic solid. American Journal of Mathematics, 59(2), 235–260. https://doi.org/10.2307/

2371405Ostrovsky, L. A., & Johnson, P. A. (2001). Dynamic nonlinear elasticity in geomaterials. Rivista del Nuovo Cimento della Societa Italiana di Fisica,

24S4(7), 1–46.Pasqualini, D., Heitmann, K., TenCate, J. A., Habib, S., Higdon, D., & Johnson, P. A. (2007). Nonequilibrium and nonlinear dynamics in Berea and

Fontainebleau sandstones: Low-strain regime. Journal of Geophysical Research, 112, B01204. https://doi.org/10.1029/2006JB004264Pecorari, C. (2015). A constitutive relationship for mechanical hysteresis of sandstone materials. Proceedings of the Royal Society A, 471(2184),

20150369. https://doi.org/10.1098/rspa.2015.0369Remillieux, M. C., Guyer, R. A., Payan, C., & Ulrich, T. J. (2016). Decoupling nonclassical nonlinear behavior of elastic wave types. Physical Review

Letters, 116(11), 115501. https://doi.org/10.1103/PhysRevLett.116.115501Renaud, G., Le Bas, P.-Y., & Johnson, P. A. (2012). Revealing highly complex elastic nonlinear (anelastic) behavior of Earth materials applying a

new probe: Dynamic acoustoelastic testing. Journal of Geophysical Research, 117, B06202. https://doi.org/10.1029/2011JB009127Renaud, G., Talmant, M., Callé, S., Defontaine, M., & Laugier, P. (2011). Nonlinear elastodynamics in micro-inhomogeneous solids observed by

head-wave based dynamic acoustoelastic testing. The Journal of the Acoustical Society of America, 130(6), 3583–3589. https://doi.org/10.1121/1.3652871

Rivière, J., Renaud, G., Guyer, R. A., & Johnson, P. A. (2013). Pump and probe waves in dynamic acousto-elasticity: Comprehensive descriptionand comparison with nonlinear elastic theories. Journal of Applied Physics, 114(5), 054905. https://doi.org/10.1063/1.4816395

Rivière, J., Shokouhi, P., Guyer, R. A., & Johnson, P. A. (2015). A set of measures for the systematic classification of the nonlinear elasticbehavior of disparate rocks. Journal of Geophysical Research: Solid Earth, 120, 1587–1604. https://doi.org/10.1002/2014JB011718

TenCate, J. A. (2011). Slow dynamics of Earth materials: An experimental overview. Pure and Applied Geophysics, 168(12), 2211–2219. https://doi.org/10.1007/s00024-011-0268-4

TenCate, J. A., Malcolm, A. E., Feng, X., & Fehler, M. C. (2016). The effect of crack orientation on the nonlinear interaction of a Pwave with an Swave. Geophysical Research Letters, 43(12), 6146–6152. https://doi.org/10.1002/2016GL069219

TenCate, J. A., Pasqualini, D., Habib, S., Heitmann, K., Higdon, D., & Johnson, P. A. (2004). Nonlinear and nonequilibrium dynamics in geo-materials. Physical Review Letters, 93(6), 065501. https://doi.org/10.1103/PhysRevLett.93.065501

Ten Cate, J. A., & Shankland, T. J. (1996). Slow dynamics in the nonlinear elastic response of Berea sandstone. Geophysical Research Letters,23(21), 3019–3022. https://doi.org/10.1029/96GL02884

TenCate, J. A., Smith, E., Byers, L. W., & Shankland, T. J. (2000). Slow dynamics experiments in solids with nonlinear mesoscopic elasticity. AIPConf. Proc., 524(1), 303–306. https://doi.org/10.1063/1.1309228

TenCate, J. A., Smith, E., & Guyer, R. A. (2000). Universal slow dynamics in granular solids. Physical Review Letters, 85(5), 1020–1023. https://doi.org/10.1103/PhysRevLett.85.1020

Ten Cate, J. A., Van Den Abeele, K. E. A., Shankland, T. J., & Johnson, P. A. (1996). Laboratory study of linear and nonlinear elastic pulse pro-pagation in sandstone. The Journal of the Acoustical Society of America, 100(3), 1383–1391. https://doi.org/10.1121/1.415985

Thurston, R. N., & Shapiro, M. J. (1967). Interpretation of ultrasonic experiments on finite-amplitude waves. The Journal of the AcousticalSociety of America, 41(4B), 1112–1125. https://doi.org/10.1121/1.1910443

Vakhnenko, O. O., Vakhnenko, V. O., Shankland, T. J., & Ten Cate, J. A. (2004). Strain-induced kinetics of intergrain defects as the mechanism ofslow dynamics in the nonlinear resonant response of humid sandstone bars. Physical Review E, 70(1), 015602. https://doi.org/10.1103/PhysRevE.70.015602

Van Den Abeele, K. E.-A. (1996). Elastic pulsed wave propagation in media with second- or higher-order nonlinearity. Part I. Theoreticalframework. The Journal of the Acoustical Society of America, 99(6), 3334–3345. https://doi.org/10.1121/1.414890

Van Den Abeele, K. E.-A., Johnson, P. A., Guyer, R. A., & McCall, K. R. (1997). On the quasi-analytic treatment of hysteretic nonlinear response inelastic wave propagation. The Journal of the Acoustical Society of America, 101(4), 1885–1898. https://doi.org/10.1121/1.418198

Van Den Abeele, K. E.-A., Johnson, P. A., & Sutin, A. (2000). Nonlinear elastic wave spectroscopy (NEWS) techniques to discern materialdamage, part I: Nonlinear wave modulation spectroscopy (NWMS). Research in Nondestructive Evaluation, 12(1), 17–30. https://doi.org/10.1007/s001640000002

Winkler, K., Nur, A., & Gladwin, M. (1979). Friction and seismic attenuation in rocks. Nature, 277(5697), 528–531. https://doi.org/10.1038/277528a0

Zoback, M. D., & Byerlee, J. D. (1975). The effect of cyclic differential stress on dilatancy in Westerly granite under uniaxial and triaxial con-ditions. Journal of Geophysical Research, 80(11), 1526–1530. https://doi.org/10.1029/JB080i011p01526

Journal of Geophysical Research: Solid Earth 10.1002/2017JB014258

REMILLIEUX ET AL. NONLINEAR PROPAGATION OF ELASTIC PULSE 18