• Aucun résultat trouvé

Finite element modeling of surface acoustic wave propagation in polycrystalline aluminium: effective phase velocity

N/A
N/A
Protected

Academic year: 2021

Partager "Finite element modeling of surface acoustic wave propagation in polycrystalline aluminium: effective phase velocity"

Copied!
7
0
0

Texte intégral

(1)

HAL Id: hal-03240228

https://hal.archives-ouvertes.fr/hal-03240228

Submitted on 13 Jun 2021

HAL is a multi-disciplinary open access

archive for the deposit and dissemination of

sci-entific research documents, whether they are

pub-lished or not. The documents may come from

teaching and research institutions in France or

abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est

destinée au dépôt et à la diffusion de documents

scientifiques de niveau recherche, publiés ou non,

émanant des établissements d’enseignement et de

recherche français ou étrangers, des laboratoires

publics ou privés.

Finite element modeling of surface acoustic wave

propagation in polycrystalline aluminium: effective

phase velocity

Martin Ryzy, Tomas Grabec, Istvan Veres

To cite this version:

Martin Ryzy, Tomas Grabec, Istvan Veres. Finite element modeling of surface acoustic wave

prop-agation in polycrystalline aluminium: effective phase velocity. Forum Acusticum, Dec 2020, Lyon,

France. pp.1833-1838, �10.48465/fa.2020.0165�. �hal-03240228�

(2)

FINITE ELEMENT MODELING OF SURFACE ACOUSTIC WAVE

PROPAGATION IN POLYCRYSTALLINE ALUMINUM: EFFECTIVE

PHASE VELOCITY

Martin Ryzy

1

Tom´aˇs Grabec

2

Istv´an A. Veres

3 1

Research Center for Non-Destructive Testing GmbH, 4040 Linz, Austria

2

Nuclear Physics Institute, 25068 ˇ

Reˇz, Czech Republic

3

Qorvo Inc., Apopka, FL 32703, USA

martin.ryzy@recendt.at

ABSTRACT

In this work we use a finite element method (FEM) to solve the elastodynamic problem of multiple scattering of sur-face acoustic Rayleigh-waves (SAWs) in polycrystalline aluminum. A realistic grain structure is generated by using a Laguerre tessellation algorithm. We calculate an ensem-ble average (coherent) response by repeated simulations and present two methods to extract the effective Rayleigh-wave phase velocity. Both methods give identical results in a frequency range from 12 MHz to 100 MHz. They re-veal a microstructure induced phase velocity transition at a frequency were the Rayleigh-wavelength is approximately one third of the average grain size.

1. INTRODUCTION

Grain boundary scattering in polycrystalline metals is not only an interesting use case of the more general class of problems dealing with wave propagation in random, het-erogeneous media, but also of great interest for nondestruc-tive evaluation (NDE) applications, as the grain scatter-ing is related to the microstructural properties of common engineering metals (e. g. mean grain size). Usually, lon-gitudinal bulk acoustic waves are studied in this context, and a multitude of experimental [1–9], analytical [10–17] and numerical studies [17–23] (mostly based on the finite-element method) exist, which investigate scattering in-duced changes of the effective phase velocity and/or atten-uation. The effect of grain boundary scattering on surface acoustic Rayleigh waves is much less studied [21, 24], al-though this wave-mode is easier experimentally accessible as it propagates along the materials’ surface and may al-low a more accurate determination of phase velocity and attenuation in experiments.

We recently closed this gap by demonstrating a broad-band laser-ultrasound measurement of the effective sur-face acoustic wave phase velocity and attenuation in poly-crystalline aluminum in a frequency range from around 10 MHz to 130 MHz [8]. In an ongoing work, we aim to complement/verify these experiments by modeling the underlying elastodynamic problem with a finite element model (FEM) in a synthetic polycrystal. Here, we present

first results of this study, related to the behavior of the ef-fective phase velocity of Rayleigh-waves.

2. FINITE ELEMENT MODELING OF SURFACE ACOUSTIC WAVE PROPAGATION

Figure 1. Sketch of the synthetic polycrystal used for the finite element SAW-propagation study. The dimensions of the domain are Lx× Ly× Lz= 1 × 3 × 0.45 mm3. The

location of the line-source for wave excitation is indicated as red solid line.

We generated a synthetic polycrystal with the dimen-sions of Lx × Ly × Lz = 1 × 3 × 0.45 mm3 with the

polycrystal generation software NEPER [25] (https:// neper.info) as illustrated in Fig. 1. The large sample-height (Lx) was chosen to avoid any Lamb-like modes

and back reflections from the lower surface. The ob-tained grain structure is a Laguerre-tessellation with a log-normal distribution of equivalent spherical grain diame-ters d (see Fig. 2). By fitting the according probabil-ity densprobabil-ity function p(d) ∝1/(2πσd) exp −(ln d−µ)2

/2σ2

(see red solid line in Fig. 2) we obtain an expecta-tion value for the equivalent spherical diameter of ¯d =

(3)

Figure 2. Distribution of the equivalent spherical grain diameter d of the synthetic polycrystal (blue bins). The red solid line is a fit of a log-normal probability density function with an expectation value of 134.1 µm (indicated as vertical black dashed line) and a standard deviation of 50.4 µm.

exp µ + 0.5σ2 = 134.1 µm and a standard deviation of

sd(d) =pexp (2µ + σ2) (exp (µ2) − 1) = 50.4 µm. The

domain was discretized into a structured cubic mesh with a lattice spacing of 1.25 µm which gives 2.0736 billion de-grees of freedom in total. The structured mesh type has been successfully applied in previous FEM grain scatter-ing studies and was shown to perform well when its spac-ing is smaller than a tenth of the mean grain size and a tenth of the wavelength [18, 20, 23], which is the case for the investigated structure and range of frequencies. A ran-domly rotated stiffness tensor Cijkl with the single

crys-tal elastic constants of aluminum (C11 = 107.6 GPa,

C12 = 61.80 GPa, C44 = 28.40 GPa) [26] was assigned

to each grain by using a method by Arvo [27], as described previously [23], which leads to a uniform distribution of crystallite orientations and thus to an untextured, macro-scopically isotropic polycrystal. Surface acoustic waves were excited by applying a uniform pressure load in y-direction along a line located on the top surface (x = 0) at y = 0 (indicated by the red line Fig. 1). The load had a Gaussian spatial distribution with a full width at half maximum (FWHM) of 18.32 µm, and a temporal distri-bution which corresponds to a laser-source with a Gaus-sian pulse length of 25 ns (FWHM) [23, 28]. Boundary conditions were chosen to sustain plane surface acoustic wave propagation and to maximally suppress unwanted wave-modes (Lamb waves, reflected bulk waves). Thus we used symmetry boundary conditions on the lateral sam-ple faces (z = 0 - plane and z = Lz - plane) and in

the y = 0 - plane, absorbing boundaries at the domain-bottom x = −Lx and at the face opposite to the

excita-tion (y = Ly), and a free boundary in the plane of

wave-propagation x = 0. An overall simulation time of 1050 ns was chosen to allow the Rayleigh-wave to propagate to the opposite sample face and a time step of ∆t ≈ 0.92 ns

was chosen to fulfill the Courant–Friedrichs–Lewy condi-tion [29]. The model was solved with the commercial fi-nite element time-domain software PZFLEX (Weidlinger Associates, Mountain View, CA) on an Intel Xeon E5620 computer (2 × 4 cores, 2.4 GHz) with typical computation times of 50 hours per simulation.

3. MODEL VERIFICATION AND EXTRACTION OF THE EFFECTIVE RAYLEIGH-WAVE PHASE

VELOCITY

To validate the FEM-model we first performed a simula-tion in a homogenized version of the polycrystal by assign-ing the same Voigt-averaged (i. e. isotropic) elastic con-stants (C11V = 112.0 GPa, C44V = 26.2 GPa) to each

crystallite. A sketch of the obtained normal surface dis-placement ux at a time t = 450 ns after wave excitation

is shown in Fig. 3(a). As expected, two wavefronts are visible. The first bipolar pulse at y ≈ 1.3 mm corre-sponds to the Rayleigh surface acoustic wave and has a much higher magnitude as the second one at y ≈ 2.8 mm, which stems from the surface skimming longitudinal wave (SSLW). Both modes have a planar wavefront and the cov-ered distances comply with their Voigt-averaged wave ve-locities of cV

R = 2911.4 ms−1 and cVL = 6440.6 ms−1 (a

density of ρ = 2700 kgm−3 was used), respectively. We thus infer that the FEM-model supports plane Rayleigh-wave propagation sufficiently.

Fig. 3(b) shows the surface displacement uxof a

poly-crystalline simulation at the same instance of time as the homogenized case. As before, Rayleigh-wave and SSLW ballistic wavefronts are clearly visible, but this time they are slightly distorted due to the grain boundary scattering. In addition to the wavefront distortion, the grain scattering causes diffuse wavefield-fluctuations, mainly in the back-ward direction. Differences in the wave speeds (compared to the homogeneous case) are not evident in this view.

In order to analyze the effective phase velocities, the wavefields were spatially averaged along the z-direction in both cases. In addition, the simulations in the poly-crystalline sample were repeated 28 times by generating 4 tessellations with the same grain statistics (equivalent to the one shown in Fig. 1) and running 7 simulations in each tessellation while using different sets of random grain rotations in each run. This approach was previously shown to be sufficient to obtain an ensemble average re-sponse [20]. The resulting averaged surface displacement fields are shown in the center graph of Fig. 4 for the homo-geneous (upper graph) and the polycrystalline case (lower graph). In both cases, the coherent Rayleigh-wave (more pronounced peak) and the SSLW are clearly visible. The effect of grain scattering induces an effective attenuation of the averaged Rayleigh-wave and a cancellation of the grain noise. This is most evident in the amplitude drop of the Rayleigh-wave peak in the polycrystalline case, which is illustrated with the three cross sections at y = 0.75, 1.5 and 2.25 mm. Moreover, the peaks are broadened compared to the homogeneous case. This can be seen in the insets of the center graphs of Fig. 4), which show magnified views

(4)

Figure 3. Out-of-plane displacement uxat the sample surface x = 0 for a homogeneous material model (a) and a

poly-crystalline material model (b) at a time t = 450 ns after wave excitation, calculated with a time-domain FEM-model. Two ballistic plane wavefronts are visible in both cases: One belongs to the Rayleigh surface acoustic wave at y ≈ 1.3 mm and the other one to the surface skimming longitudinal wave (SSLW) at y ≈ 2.8 mm. Please note the non-linear scaling of the color bars, which is useful to improve the visibility of the SSLW.

of the Rayleigh-peak time-traces at y = 1.5 mm.

For the extraction of the frequency dependent effective Rayleigh-wave phase velocity cR(f ) we applied two

dif-ferent data processing procedures:

The first one is based on evaluating the phase of the temporal Fourier transform Ft{ux(y, t)} = ˜ux(y, f ) of

the initial signal, whose magnitude |˜ux(y, f )| and phase

arg(˜ux(y, f )) are shown in the left graphs of Fig. 4 for a

frequency of f = 60 MHz. Please note that the magnitude is mainly determined by the Rayleigh-mode and the SSLW appears as a beating of these two wave-modes (see oscilla-tions) and dominantly decays due to geometric spreading, whereas a plane Rayleigh-wave does not suffer geometric attenuation. In contrast to the homogeneous case, the mag-nitude in the polycrystalline case decays in an exponen-tial manner due to the induced effective attenuation. The effective Rayleigh phase velocity cR was obtained by

fit-ting (minimization of the root mean square error (RMSE)) a straight line φ(y, f ) = φ0(f ) + kR(f )y (green lines in

the insets) to the (unwrapped) phase arg(˜ux(y, f )) (blue

lines), and by using the fundamental dispersion relation 2πf = cRkRfor each frequency in a range from 12 MHz to

100 MHz. For the fits a range of y ∈ [1, 2.9]mm was cho-sen to avoid interference from the SSLW for small y and er-rors induced by Fourier-transforming a partially truncated

Rayleigh-pulse at large y.

In the second method, we separated the Rayleigh-wave from the SSLW in the time domain, by restricting ux(y, t)

to regions where both wave-modes are clearly separated (see semi-transparent yellow region in the center graphs of Fig. 4 where y ∈ [0.45, 2.95]mm). Then a Tukey-window with a length of 56.8 ns and a rise time of 4.66 ns (red line in the insets of the center graphs of Fig. 4) was applied around the center of the Rayleigh-peak (blue line) and the obtained traces were zero-padded. Thereafter, fast Fourier transforms were applied along y and t to get the response ˜

ux(k, f ) in the reciprocal space (see right graphs in Fig. 4),

where only the Rayleigh-mode is left. Then, for each frequency f , Lorentzian functions of the form g(k, f ) ∝

α2+ (k R− k)2

−0.5

where fit (RMSE minimization) to the magnitude |˜ux(k, f )| to obtain to Rayleigh-peak

cen-ter wavenumbers kRand the phase velocity cR= 2πf k−1R .

Examples of the fits are shown in the insets of the right graphs of Fig. 4 for a frequency of 60 MHz (indicated by the red line in the main graph). Please note that a broad-ening (see particularly the side-lobes in the homogeneous case) of the Rayleigh-peaks occurs due to the applied zero-padding. However, as we are only interested in the peak-center positions, this will not influence the obtained results.

(5)

Figure 4. Averaged surface normal displacement fields of FEM-simulations from a homogeneous sample (upper row of graphs) and from a polycrystalline sample (lower row of graphs). The center graphs show the wavefields in the physical domain, the insets time traces of the the Rayleigh-wave peaks at y = 1.5 mm and the Tukey-window used for its separation from the SSLW. The graphs on the left show the magnitude and (unwrapped) phase (inset) of the temporal Fourier-transform of the initial data at a frequency of 60 MHz. The green line is a linear fit to the phase and overlaps with the data. The graphs on the right side show the magnitude of the signals in the reciprocal space (kf - space) obtained after separation of the Rayleigh-wave from the SSLW by taking data from the semi-transparent yellow areas in the center graphs and applying a Tukey-window (as shown in the inset of the center graph) prior to double Fourier-transformation. The insets in the right graphs show parts of the cross sections around the Rayleigh-wave (blue line) used for fitting Lorentzian functions (green lines) at a frequency of f = 60 MHz.

4. RESULTS AND DISCUSSION

The extracted effective Rayleigh-wave phase velocities are shown in Fig. 5 for a frequency range from 12 MHz to 100 MHz. For the homogeneous case, both data pro-cessing methods (blue and green solid lines) give virtu-ally identical results and maximvirtu-ally deviate by around 10 ms−1(≈ 0.34 %) from the Voigt-averaged phase veloc-ity cV

R= 2911.40 ms−1 (red line). The observed

system-atic deviation is caused by numerical dispersion (‘mesh-scattering’) and is comparable to the numerical phase ve-locity dispersion typically observed in bulk-wave FEM-studies [23]. The ‘oscillations‘ at lower frequencies ob-served in the phase-fitting-method may be caused by re-maining influences from the SSLW, which is more effec-tively rejected at higher frequencies when the width of the excitation pulse starts to exceed its (half) wavelength.

The phase velocity of the polycrystalline sample is somewhat lower than the Voigt-velocity. This can be ex-pected, as Voigt-averaged elastic constants (and so the wave velocities) are upper boundaries for effective elas-tic constants [30]. Again, both fitting methods give virtu-ally identical results, except for small deviations at high frequencies and the previously discussed oscillations

ob-served in the phase-fitting method. Please note that like in the homogeneous case, the numerical dispersion also here leads to a steady drop of the phase velocity with rising frequency. However, in contrast to the homogenized sam-ple, there is a pronounced velocity drop of around 10 ms−1 (≈ 0.35 %) at approximately 70 MHz, where the grain-size to wavelength ratioλ/d¯= 0.31 ≈1/3and the normalized

wavenumber amounts to kV

Rd ≈ 20. This transition is very¯

similar to that predicted by Stanke & Kino [11] for trans-verse waves in polycrystalline aluminum, were the phase velocity drops around 3.3 % at kdmfp ≈ 30 (see Fig. 7

in reference [11])1. For this reason, and because both

of the applied data-processing methods give the same sults, we are confident that the observed transition is re-lated to the grain structure here as well. Hence, an accu-rate measurement of the Rayleigh-wave phase velocity dis-persion may be used for microstructure characterization in future NDE-applications. Further numerical studies with synthetic polycrystals with different microstructures (e. g.

1Please note that d

mfpin [11] is the mean free path length in the

mi-crostructure and not identical with the equivalent spherical grain diameter expectation value ¯d as used here. Moreover, the microstructure morpholo-gies are different, which will affect both dmfpand ¯d. Thus, the values for

(6)

20 40 60 80 100

f

(MHz)

2870 2880 2890 2900 2910 2920 2930

p

h

a

se

v

el

o

ci

ty

c

R

!

m

s

− 1

"

Voigt-average |˜ux(k, f)|, Lorentzian arg(˜ux(y, f)), linear

fit-type

homogeneous

polycrystalline

Figure 5. Effective Rayleigh-wave phase velocities ex-tracted from finite element simulations of a homogeneous and a polycrystalline material model. Two data processing methods are compared, one based on a linear fit of the sig-nal phase in yf -space (blue solid lines) and the other one on a Lorentzian fit of the signal magnitude in kf -space (green solid lines). The red line indicates the theoretical phase velocity of cV

R= 2911.4 ms−1 obtained from

Voigt-averaged elastic constants.

varying grain size distributions) or the development of ana-lytical mean-field models for the prediction of the effective Rayleigh-wave velocity in polycrystalline media will help to elucidate this issue further.

5. CONCLUSION

In summary, we developed a finite-element model to solve the elastodynamic problem of surface acoustic plane-wave propagation in a polycrystalline aluminum sample by using a Laguerre-tessellation to emulate a realistic microstruc-ture. The model accounts for multiple scattering events of any order, and can be generalized to other materials with lower crystal symmetry, texture, and different grain morphologies. We demonstrated and compared two data-processing methods for the determination of the effective Rayleigh-wave phase velocity. Results from both meth-ods are in excellent agreement and reveal a phase-velocity transition at a frequency were the Rayleigh-wavelength is approximately one third of the grain size. Due to the simi-larity of this transition to that observed for transverse wave, we conclude that it is caused by the grain structure and may thus be used for its characterization in future NDE appli-cations.

6. ACKNOWLEDGEMENTS

This work was funded by the Austrian research funding association (FFG) under the scope of the COMET pro-gram within the research project “Photonic Sensing for Smarter Processes (PSSP)” (Contract No. 871974), by the project “multimodal and in-situ characterization of inho-mogeneous materials” (MiCi) by the federal government of Upper Austria and the European Regional Develop-ment Fund (EFRE) in the framework of the EU-program IWB2020 and by OP RDE, MEYS (ESS-Scandinavia-CZ-OP, CZ.02.1.01/0.0/0.0/16 013/0001794).

7. REFERENCES

[1] W. P. Mason and H. J. McSkimin, “Attenuation and Scattering of High Frequency Sound Waves in Metals and Glasses,” The Journal of the Acoustical Society of America, vol. 19, p. 10, May 1947.

[2] L. G. Merkulov, “Investigation of ultrasonic scattering in metals,” Soviet Physics - Technical Physics, vol. 1, pp. 59–69, 1956.

[3] F. E. Stanke, “Inversion of Attenuation Measurements in Terms of a Paramterized Autocorrelation Function,” in NDE of Microstructure for Process Control: Pro-ceedings of a Symposium, (Metals Park, Ohio), pp. 55– 60, 1985.

[4] X.-G. Zhang, W. A. Simpson, J. M. Vitek, D. J. Barnard, L. J. Tweed, and J. Foley, “Ultrasonic atten-uation due to grain boundary scattering in copper and copper-aluminum,” The Journal of the Acoustical So-ciety of America, vol. 116, pp. 109–116, July 2004. [5] T. Garcin, J. H. Schmitt, and M. Militzer, “In-situ

laser ultrasonic grain size measurement in superalloy INCONEL 718,” Journal of Alloys and Compounds, vol. 670, pp. 329–336, June 2016.

[6] F. Dong, X. Wang, Q. Yang, H. Liu, D. Xu, Y. Sun, Y. Zhang, R. Xue, and S. Krishnaswamy, “In-situ measurement of Ti-6Al-4V grain size distribution us-ing laser-ultrasonic technique,” Scripta Materialia, vol. 154, pp. 40–44, Sept. 2018.

[7] X. Bai, Y. Zhao, J. Ma, Y. Liu, and Q. Wang, “Grain-Size Distribution Effects on the Attenuation of Laser-Generated Ultrasound in α-Titanium Alloy,” Materi-als, vol. 12, no. 1, 2018.

[8] M. Ryzy, T. Grabec, J. A. ¨Osterreicher, M. Hettich, and I. A. Veres, “Measurement of coherent surface acous-tic wave attenuation in polycrystalline aluminum,” AIP Advances, vol. 8, p. 125019, Dec. 2018.

[9] J. L. Blackshire, “Model-Assisted Analysis of Local Attenuation in Ti-6Al-4V Materials,” in 2019 IEEE In-ternational Ultrasonics Symposium (IUS), (Glasgow, United Kingdom), pp. 932–935, IEEE, Oct. 2019.

(7)

[10] S. Hirsekorn, “The scattering of ultrasonic waves by polycrystals,” The Journal of the Acoustical Society of America, vol. 72, pp. 1021–1031, Sept. 1982.

[11] F. E. Stanke and G. S. Kino, “A unified theory for elas-tic wave propagation in polycrystalline materials,” The Journal of the Acoustical Society of America, vol. 75, pp. 665–681, Mar. 1984.

[12] R. L. Weaver, “Diffusivity of ultrasound in polycrys-tals,” Journal of the Mechanics and Physics of Solids, vol. 38, no. 1, pp. 55–86, 1990.

[13] J. A. Turner, “Elastic wave propagation and scattering in heterogeneous, anisotropic media: Textured poly-crystalline materials,” The Journal of the Acoustical Society of America, vol. 108, p. 12, Aug. 1999. [14] M. Calvet and L. Margerin, “Velocity and attenuation

of scalar and elastic waves in random media: A spectral function approach,” The Journal of the Acoustical So-ciety of America, vol. 131, pp. 1843–1862, Mar. 2012. [15] C. M. Kube and J. A. Turner, “Acoustic attenuation

co-efficients for polycrystalline materials containing crys-tallites of any symmetry class,” The Journal of the Acoustical Society of America, vol. 137, pp. EL476– EL482, June 2015.

[16] G. Sha, M. Huang, M. J. S. Lowe, and S. I. Rokhlin, “Attenuation and velocity of elastic waves in polycrys-tals with generally anisotropic grains: Analytic and nu-merical modeling,” The Journal of the Acoustical Soci-ety of America, vol. 147, pp. 2442–2465, Apr. 2020. [17] X. Bai, B. Tie, J.-H. Schmitt, and D. Aubry,

“Com-parison of ultrasonic attenuation within two- and three-dimensional polycrystalline media,” Ultrasonics, vol. 100, p. 105980, Jan. 2020.

[18] A. Van Pamel, C. R. Brett, P. Huthwaite, and M. J. S. Lowe, “Finite element modelling of elastic wave scat-tering within a polycrystalline material in two and three dimensions,” The Journal of the Acoustical Society of America, vol. 138, pp. 2326–2336, Oct. 2015.

[19] D. Dobrovolskij, S. Hirsekorn, and M. Spies, “Simu-lation of Ultrasonic Materials Evaluation Experiments Including Scattering Phenomena due to Polycrystalline Microstructure,” Physics Procedia, vol. 70, pp. 644– 647, 2015.

[20] A. Van Pamel, G. Sha, S. I. Rokhlin, and M. J. S. Lowe, “Finite-element modelling of elastic wave propagation and scattering within heterogeneous media,” Proceed-ings of the Royal Society A: Mathematical, Physical and Engineering Science, vol. 473, Jan. 2017.

[21] T. Grabec, M. Ryzy, and I. A. Veres, “Numerical modeling of surface elastic wave scattering in poly-crystalline materials,” in 2017 IEEE International Ul-trasonics Symposium (IUS), (Washington, DC, USA), pp. 1–4, IEEE, Sept. 2017.

[22] P. Lhuillier, B. Chassignole, M. Oudaa, S. Kerherv´e, F. Rupin, and T. Fouquet, “Investigation of the ultra-sonic attenuation in anisotropic weld materials with finite element modeling and grain-scale material de-scription,” Ultrasonics, vol. 78, pp. 40–50, July 2017. [23] M. Ryzy, T. Grabec, P. Sedl´ak, and I. A. Veres,

“Influ-ence of grain morphology on ultrasonic wave attenua-tion in polycrystalline media with statistically equiaxed grains,” The Journal of the Acoustical Society of Amer-ica, vol. 143, pp. 219–229, Jan. 2018.

[24] I. M. Kaganova and A. A. Maradudin, “Surface acous-tic waves on a polycrystalline substrate,” Physica Scripta, vol. T44, pp. 104–112, Feb. 1992.

[25] R. Quey, P. Dawson, and F. Barbe, “Large-scale 3D random polycrystals for the finite element method: Generation, meshing and remeshing,” Computer Meth-ods in Applied Mechanics and Engineering, vol. 200, pp. 1729–1745, Apr. 2011.

[26] W. P. Mason and H. J. McSkimin, “Energy Losses of Sound Waves in Metals Due to Scattering and Diffu-sion,” Journal of Applied Physics, vol. 19, pp. 940– 946, Oct. 1948.

[27] J. Arvo, “Fast Random Rotation Matrices,” Graphic Gems, vol. 3, pp. 117–120, 1992.

[28] A. G. Every, Z. N. Utegulov, and I. A. Veres, “Laser thermoelastic generation in metals above the melt threshold,” Journal of Applied Physics, vol. 114, p. 203508, Nov. 2013.

[29] R. Courant, K. Friedrichs, and H. Levy, “On the Par-tial Difference Equations of Mathematical Physics,” IBM Journal of Research and Development, vol. 11, pp. 215–234, Mar. 1967.

[30] A. Morawiec, “Review of Deterministic Methods of Polycrystal Elastic Constants,” Textures and Mi-crostructures, vol. 22, no. 3, pp. 139–167, 1994. [31] A. Van Pamel, G. Sha, M. J. S. Lowe, and S. I. Rokhlin,

“Numerical and analytic modelling of elastodynamic scattering within polycrystalline materials,” The Jour-nal of the Acoustical Society of America, vol. 143, pp. 2394–2408, Apr. 2018.

Références

Documents relatifs

In conclusion, we have proposed a new error analysis method based on radiative transfer equation in terms of energy quantity: the radiative transfer equation is derived in

Finally, as it seems that there is no analytical solution for the 2D wave propagation in unsaturated soils in the literature, to verify the accuracy of this

  Modèle : Classes métier   Vue : Pages JSP (ou facelet)   Contrôle : Beans managés.   Séparer le contrôle applicatif des classes

A survey for potential similarly bright or brighter methanol masers, as described in our full survey paper, will allow us to determine more than one proper motion and derivation of

Tordjman , Construction and analysis of higher order finite ele- ments with mass lumping for the wave equation, in Proceedings of the Second International Conference on Mathematical

De même, le concepteur doit être en mesure d'apprécier la portée et les limites des résultats d'essais normalisés pour connaître les critères qu'il doit appliquer avant d'utiliser

Comparison of PCET kinetics for 1:3 and 1:4 enable an analysis of charge transfer as mediated by differing proton configurations within a hydrogen bonded interface.. Previous

In the context of our study, the hypothesis we stated checks the effect of one I V (Prior knowledge activation through brainstorming) on the D V (reading comprehension). Thus,