Skip to main content
Log in

Photon Doppler Velocimetry and Simulation of Ejection of Particles from the Surface of Shock-Loaded Samples

  • ATOMS, MOLECULES, OPTICS
  • Published:
Journal of Experimental and Theoretical Physics Aims and scope Submit manuscript

Abstract

The results of photon Doppler velocimetry of ejecta from shock-loaded metal samples are reported. The experiments have been performed with tin and lead samples of a given thickness and a given surface roughness. The direct numerical simulation of the process of mass ejection from the surface of shock-loaded samples is performed for conditions close to experimental by the smoothed particle hydrodynamics method. The areal density and initial velocity distribution of the volume density of ejecta are determined. Using these results, we calculate the time dependence of the profile of the volume density at the expansion of the formed dust cloud to air. Applying an approach based on the transport equation for the correlation function of the scattered field, the main parameters of the velocity distribution of ejecta, areal density of ejecta, etc. are reconstructed from spectral photon Doppler velocimetry data. The experimentally observed temporal dynamics of spectra, which is caused by the drag of dust in air, is described at an appropriately chosen size dispersion of dust particles. The masses of ejecta reconstructed from experimental data are in agreement with the smoothed particle hydrodynamics results.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Fig. 1.
Fig. 2.
Fig. 3.
Fig. 4.
Fig. 5.
Fig. 6.
Fig. 7.
Fig. 8.
Fig. 9.
Fig. 10.
Fig. 11.
Fig. 12.
Fig. 13.
Fig. 14.
Fig. 15.
Fig. 16.
Fig. 17.

Similar content being viewed by others

REFERENCES

  1. V. A. Ogorodnikov, A. G. Ivanov, A. L. Mikhailov, et al., Fiz. Goreniya Vzryva 34, 103 (1998).

    Google Scholar 

  2. W. S. Vogan, W. W. Anderson, M. Grover, et al., J. Appl. Phys. 98, 113508 (2005).

    ADS  Google Scholar 

  3. W. T. Buttler, M. B. Zellner, R. T. Olson, et al., J. Appl. Phys. 101, 063547 (2007).

    ADS  Google Scholar 

  4. M. B. Zellner, M. Grover, J. E. Hammerberg, et al., J. Appl. Phys. 102, 013522 (2007).

    ADS  Google Scholar 

  5. M. B. Zellner, W. Vogan-McNeil, J. E. Hammerberg, et al., J. Appl. Phys. 103, 123502 (2008).

    ADS  Google Scholar 

  6. M. B. Zellner and W. T. Buttler, Appl. Phys. Lett. 93, 114102 (2008).

    ADS  Google Scholar 

  7. V. A. Ogorodnikov, A. L. Mikhailov, V. V. Burtsev, S. A. Lobastov, S. V. Erunov, A. V. Romanov, A. V. Rudnev, E. V. Kulakov, Yu. B. Bazarov, V. V. Glushikhin, I. A. Kalashnik, V. A. Tsyganov, and B. I. Tkachenko, J. Exp. Theor. Phys. 109, 530 (2009).

    ADS  Google Scholar 

  8. V. V. Igonin, G. B. Krasovsky, S. E. Kuratov, et al., Phys. Scr. 142, 014019 (2010).

    Google Scholar 

  9. W. T. Buttler, D. M. Oró, D. L. Preston, et al., J. Fluid Mech. 703, 60 (2012).

    ADS  Google Scholar 

  10. Y. Chen, H. Hu, T. Tang, et al., J. Appl. Phys. 111, 053509 (2012).

    ADS  Google Scholar 

  11. A. L. Mikhailov, V. A. Ogorodnikov, V. S. Sasik, V. A. Raevskii, A. I. Lebedev, D. E. Zotov, S. V. Erunov, M. A. Syrunin, V. D. Sadunov, N. V. Nevmerzhitskii, S. A. Lobastov, V. V. Burtsev, A. V. Mishanov, E. V. Kulakov, A. V. Satarova, et al., J. Exp. Theor. Phys. 118, 785 (2014).

    ADS  Google Scholar 

  12. S. K. Monfared, D. M. Oró, M. Grover, et al., J. Appl. Phys. 116, 063504 (2014).

    ADS  Google Scholar 

  13. W. T. Buttler, D. M. Oró, R. T. Olson, et al., J. Appl. Phys. 116, 103519 (2014).

    ADS  Google Scholar 

  14. V. A. Ogorodnikov, A. L. Mikhailov, V. S. Sasik, S. V. Erunov, M. A. Syrunin, A. V. Fedorov, N. V. Nevmerzhitskii, E. V. Kulakov, O. A. Kleshchevnikov, M. V. Antipov, I. V. Yurtov, A. V. Rudnev, A. V. Chapaev, A. S. Pupkov, E. D. Sen’kovskii, et al., J. Exp. Theor. Phys. 123, 357 (2016).

    ADS  Google Scholar 

  15. V. A. Ogorodnikov, A. L. Mikhailov, S. V. Erunov, M. V. Antipov, A. V. Fedorov, M. A. Syrunin, E. V. Kulakov, O. A. Kleshchevnikov, I. V. Yurtov, A. A. Utenkov, S. A. Finyushin, E. A. Chudakov, D. A. Kalashnikov, A. S. Pupkov, A. V. Chapaev, et al., J. Exp. Theor. Phys. 125, 985 (2017).

    ADS  Google Scholar 

  16. M. V. Antipov, V. A. Arinin, A. B. Georgievskaya, et al., J. Dynam. Behav. Mater. 3, 300 (2017).

    Google Scholar 

  17. O. T. Strand, D. R. Goosman, C. Martinez, et al., Rev. Sci. Instrum. 77, 083108 (2006).

    ADS  Google Scholar 

  18. B. J. Jensen, D. B. Holtkamp, P. A. Rigg, et al., J. Appl. Phys. 101, 013523 (2007).

    ADS  Google Scholar 

  19. A. R. Valenzuela, G. Rodriguez, S. A. Clarke, et al., Rev. Sci. Instrum. 78, 013101 (2007).

    ADS  Google Scholar 

  20. E. A. Moro, M. E. Briggs, L. M. Hull, et al., Appl. Opt. 52, 08661 (2013).

    ADS  Google Scholar 

  21. W. J. Warren, E. A. Moro, M. E. Briggs, et al., Appl. Opt. 53, 4661 (2014).

    ADS  Google Scholar 

  22. M. V. Astashkin, V. K. Baranov, A. B. Georgievskaya, A. G. Golubinskii, D. A. Irinichev, E. E. Meshkov, S. N. Stepushkin, A. Yu. Syundyukov, and V. Yu. Khatunkin, JETP Lett. 99, 146 (2014).

    ADS  Google Scholar 

  23. G. Prudhomme, P. Mercier, L. Berthe, et al., J. Phys.: Conf. Ser. 500, 142022 (2014).

    Google Scholar 

  24. G. Prudhomme, P. Mercier, and L. Berthe, J. Phys.: Conf. Ser. 500, 142027 (2014).

    Google Scholar 

  25. A. V. Fedorov, A. L. Mikhailov, S. A. Finyushin, D. A. Kalashnikov, E. A. Chudakov, E. I. Butusov and I. S. Gnutov, J. Exp. Theor. Phys. 122, 685 (2016).

    ADS  Google Scholar 

  26. A. V. Fedorov, A. L. Mikhailov, L. K. Antonyuk, and I. V. Shmelev, Combust. Explos., Shock Waves 52, 476 (2016).

    Google Scholar 

  27. A. V. Fedorov, A. L. Mikhailov, S. A. Finyushin, D. A. Kalashnikov, E. A. Chudakov, E. I. Butusov, and I. S. Gnutov, Combust. Explos., Shock Waves 52, 482 (2016).

    Google Scholar 

  28. A. V. Andriyash, M. V. Astashkin, V. K. Baranov, A. G. Golubinskii, D. A. Irinichev, A. N. Kondrat’ev, S. E. Kuratov, V. A. Mazanov, D. B. Rogozkin, S. N. Stepushkin, and V. Yu. Khatunkin, J. Exp. Theor. Phys. 122, 970 (2016).

    ADS  Google Scholar 

  29. J.-E. Franzkowiak, G. Prudhomme, P. Mercier, et al., Rev. Sci. Instrum. 89, 033901 (2018).

    ADS  Google Scholar 

  30. A. V. Andriyash, M. V. Astashkin, V. K. Baranov, et al., J. Appl. Phys. 123, 243102 (2018).

    ADS  Google Scholar 

  31. A. V. Fedorov, I. S. Gnutov, and A. O. Yagovkin, J. Exp. Theor. Phys. 126, 76 (2018).

    ADS  Google Scholar 

  32. J. J. Monaghan, J. Comput. Phys. 136, 298 (1997).

    ADS  MathSciNet  Google Scholar 

  33. S. P. Marsh, LANL Shock Hugoniot Data (Univ. California Press, Berkeley, 1980).

  34. O. Durand and L. Soulard, J. Appl. Phys. 114, 194902 (2013).

    ADS  Google Scholar 

  35. O. Durand and L. Soulard, J. Appl. Phys. 117, 165903 (2015).

    ADS  Google Scholar 

  36. L. D. Landau and E. M. Lifshitz, Course of Theoretical Physics, Vol. 6: Fluid Mechanics (Nauka, Moscow, 2000; Pergamon, New York, 1987).

  37. L. D. Cloutman, Am. J. Phys. 56, 643 (1988).

    ADS  Google Scholar 

  38. A. B. Georgievskaya and V. A. Raevsky, AIP Conf. Proc. 1426, 1007 (2012).

    ADS  Google Scholar 

  39. A. B. Georgievskaya and V. A. Raevsky, J. Dynam. Behav. Mater. 3, 321 (2017).

    Google Scholar 

  40. G. Dimonte, G. Terrones, F. J. Cherne, et al., J. Appl. Phys. 113, 024905 (2013).

    ADS  Google Scholar 

  41. J.-L. Shao, P. Wang, and A.-M. He, J. Appl. Phys. 116, 073501 (2014).

    ADS  Google Scholar 

  42. S. Dyachkov, A. Parshikov, and V. Zhakhovsky, AIP Conf. Proc. 1793, 100024 (2017).

    Google Scholar 

  43. Ya. B. Zel’dovich and Yu. P. Raizer, Physics of Shock Waves and High-Temperature Hydrodynamic Phenomena (Academic, New York, 1966, 1967; Fizmatlit, Moscow, 2008).

  44. J. R. Asay and M. Shahinpoor, High-Pressure Shock Compression of Solids (Springer, New York, 1993).

    MATH  Google Scholar 

  45. C. Roland, T. de Rességuier, A. Sollier, et al., J. Dynam. Behav. Mater. 3, 156 (2017).

    Google Scholar 

  46. D. S. Sorenson, P. Pazuchanics, R. P. Johnson, et al., Tech. Report LA-UR-14-24722 (Los Alamos Natl. Labor., 2014).

  47. D. S. Sorenson, G. A. Capelle, M. Grover, et al., J. Dynam. Behav. Mater. 3, 233 (2017).

    ADS  Google Scholar 

  48. S. K. Monfared, W. T. Buttler, D. K. Frayer, et al., J. Appl. Phys. 117, 223105 (2015).

    ADS  Google Scholar 

  49. M. M. Schauer, W. T. Buttler, D. K. Frayer, et al., J. Dynam. Behav. Mater. 3, 217 (2017).

    Google Scholar 

  50. D. S. Sorenson, R. W. Minich, J. L. Romero, et al., J. Appl. Phys. 92, 5830 (2002).

    ADS  Google Scholar 

  51. A. Ishimaru, Propagation and Scattering of Waves in Random Media (Academic, New York, 1978; Mir, Moscow, 1981).

  52. V. V. Sobolev, Light Scattering in Planetary Atmospheres (Nauka, Moscow, 1972; Pergamon, Oxford, 1975).

  53. H. C. van de Hulst, Multiple Light Scattering (Academic, New York, 1980).

    Google Scholar 

  54. P. Laven, Appl. Opt. 42, 436 (2003). http://www.philiplaven.com.

    ADS  Google Scholar 

  55. E. D. Palik and G. Ghosh, Handbook of Optical Constants of Solids (Academic, San Diego, 1998).

    Google Scholar 

  56. D. M. Oró, J. E. Hammerberg, W. T. Buttler, et al., AIP Conf. Proc. 1426, 1351 (2012).

    ADS  Google Scholar 

  57. L. A. Dombrovsky, Comp. Therm. Sci. 4, 297 (2012).

    Google Scholar 

  58. L. I. Sedov, Similarity and Dimensional Methods in Mechanics (Cleaver Hume, London, 1959; Nauka, Moscow, 1977).

  59. A. N. Parshikov and S. A. Medin, J. Comput. Phys. 180, 353 (2002).

    ADS  Google Scholar 

  60. J. K. Dukowicz, J. Comput. Phys. 61, 119 (1985).

    ADS  MathSciNet  Google Scholar 

  61. H. Wendland, Adv. Comput. Math. 4, 389 (1995).

    MathSciNet  Google Scholar 

  62. M. S. Egorova, S. A. Dyachkov, A. N. Parshikov, et al., Comput. Phys. Commun. 234, 112 (2019).

    ADS  Google Scholar 

  63. J.-E. Franzkowiak, P. Mercier, G. Prudhomme, et al., Appl. Opt. 57, 2766 (2018).

    ADS  Google Scholar 

Download references

Funding

V.V.Zh. acknowledges the support of the Russian Science Foundation (project no. 19-19-00697).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to A. N. Kondratev.

Additional information

Translated by R. Tyapaev

Appendices

Simulation of the Initial Stage of Ejection by the Contact Smoothed Particle Hydrodynamics Method

The smoothed particle hydrodynamics (SPH) method is a method for the interpolation of a function f(r) whose values are specified on a discrete set of N points of space r1, …, rN. Using these known values and the smoothing function (kernel) W(r, h), one can approximate the following function at a certain point r of space:

$$f({\mathbf{r}}) = \sum\limits_{j = 1}^N {{{f}_{j}}\Delta {{V}_{j}}W({\text{|}}{\mathbf{r}} - {{{\mathbf{r}}}_{j}}{\text{|}},h),} $$
(A.1)

where ΔVj is the corresponding volume element and h is the smoothing distance. The smoothing kernel satisfies the conditions

$$\int\limits_{|{\mathbf{r}} - {\mathbf{R}}| < h}^{} {W({\text{|}}{\mathbf{r}} - {\mathbf{R}}{\text{|}},h){{d}^{3}}r} = 1,$$
(A.2)
$$\mathop {\lim }\limits_{h \to 0} W({\text{|}}{\mathbf{r}} - {\mathbf{R}}{\text{|}},h) = \delta ({\mathbf{R}}),$$
(A.3)

where δ(R) is the Dirac delta function. This representation is convenient because the gradient of the function f(r) is expressed in terms of the smoothing kernel W(r, h) calculated for different points of space:

$$\nabla f({\mathbf{r}}) = \sum\limits_{j = 1}^N {{{f}_{j}}\Delta {{V}_{j}}\nabla W({\text{|}}{\mathbf{r}} - {{{\mathbf{r}}}_{j}}{\text{|}},h)} ,$$
(A.4)
$$\nabla W({\text{|}}{\mathbf{r}} - {{{\mathbf{r}}}_{j}}{\text{|}},h) = W{\kern 1pt} '({\text{|}}{\mathbf{r}} - {{{\mathbf{r}}}_{j}}{\text{|}},h)\frac{{{\mathbf{r}} - {{{\mathbf{r}}}_{j}}}}{{{\text{|}}{\mathbf{r}} - {{{\mathbf{r}}}_{j}}{\text{|}}}}.$$
(A.5)

Thus, the derivative of the function at a given point of space is approximated by simple summation using its nearest environment in the smoothing distance.

Using this representation of the spatial dependence of the function, one can construct a set of finite-difference SPH approximations of Eqs. (5)–(7). In this work, we use a scheme with the solution of the Riemann problem at the contact boundary between a pair of particles [59], which ensures the monotonicity of functions near the discontinuity point. In this case, the system of Eqs. (5)–(7) has the form

$$\frac{{d{{\rho }_{i}}}}{{dr}} = 2{{\rho }_{i}}\sum\limits_{j = 1}^N {\frac{{{{m}_{j}}}}{{{{\rho }_{j}}}}({{{\mathbf{U}}}_{i}} - {\mathbf{U}}_{{ij}}^{*})} \cdot \nabla {{W}_{{ij}}},$$
(A.6)
$$\frac{{d{{{\mathbf{U}}}_{i}}}}{{dt}} = - \frac{2}{{{{\rho }_{i}}}}\sum\limits_{j = 1}^N {\frac{{{{m}_{j}}}}{{{{\rho }_{j}}}}P_{{ij}}^{*}\nabla {{W}_{{ij}}},} $$
(A.7)
$$\frac{{d{{e}_{i}}}}{{dr}} = \frac{2}{{{{\rho }_{i}}}}\sum\limits_{j = 1}^N {\frac{{{{m}_{j}}}}{{{{\rho }_{j}}}}P_{{ij}}^{*}} ({{{\mathbf{U}}}_{i}} - {\mathbf{U}}_{{ij}}^{*}) \cdot \nabla {{W}_{{ij}}},$$
(A.8)

where \({\mathbf{U}}_{{ij}}^{*}\) and \({\mathbf{P}}_{{ij}}^{*}\) are the contact velocity and pressure for the ith and jth particles, respectively, which are obtained by solving the one-dimensional Riemann problem along the lines connecting their centers. The solution presented in [60] is used to calculate the contact values. The Wendland-C2 polynomial [61] is used as the smoothing function.

A large number of particles (high spatial resolution) is required for SPH calculations with a necessary accuracy. The parallel computation algorithm used in this case should be optimized in the computational time. In this work, we use the dynamic division of the medium into Voronoi cells [62]; the automated redistribution of particles between these cells ensures the balance of computational load. The essence of the approach is illustrated in Fig. 17. The simulated sample consists of a large number of particles (several millions), which are processed by parallel-operating processors. Each processor is applied only to a local set of data on particles (usually, 50 000–100 000 particles per processor core), which correspond to one Voronoi cell in the simulated space. The Voronoi diagram is unambiguously determined by the positions of the centers of cells. Consequently, the variation of the positions of these centers in order to improve the balance of the computational load leads to a change in the diagram and to the corresponding redistribution of particles between operating processors. Thus, parallel decomposition can be rapidly adapted to varying conditions. The algorithm was described in detail in [62].

The representation of the medium in the form of a set of SPH particles has a number of advantages over mesh methods. Eulerian mesh schemes require control of contact surfaces and free boundaries, which is the main source of errors at simulation by these methods. Special procedures of the rearrangement of the mesh in the region of formation of a jet at ejection are needed for Lagrangian mesh schemes; consequently, a significant numerical diffusion of the momentum and energy occurs in the rearrangement region and, as a result, small-scale processes accompanying jet formation are lost. The simulation by the meshless SPH method does not require overcoming the indicated technical difficulties. For this reason, it is optimal for the simulation of various physical phenomena such as jet formation, rotational and shear flows of compressible materials, and fragmentation of the computation region into parts because of effects of destruction.

Distribution of Dust Particles Decelerating in a Gas

The initial velocity and coordinate distribution of the ejected mass varies with time because of the deceleration of dust particles in air. The distribution function of particles with a certain size d at a given time tm can be written in the form

$$\begin{gathered} N({v},z,{{t}_{{\text{m}}}}) = \frac{{{{N}_{{\text{a}}}}}}{{{{{v}}_{{{\text{fs}}}}}}}\int {d{{{v}}_{0}}F\left( {\frac{{{{{v}}_{0}}}}{{{{{v}}_{{{\text{fs}}}}}}}} \right)} \\ \, \times \delta (z - z({{{v}}_{0}},{{t}_{{\text{m}}}}))\delta ({v} - {v}({{{v}}_{0}},{{t}_{{\text{m}}}})), \\ \end{gathered} $$
(B.1)

where Na is the number of particles ejected from unit area, (1/\({{{v}}_{{{\text{fs}}}}}\))F(\({{{v}}_{0}}\)/\({{{v}}_{{{\text{fs}}}}}\)) is the initial normalized velocity distribution of dust particles (see Eq. (2)), \({v}\)(\({{{v}}_{0}}\), tm) and z(\({{{v}}_{0}}\), tm) = \(\int_0^{{{t}_{{\text{m}}}}} {dt{v}} \)(\({{{v}}_{0}}\), t) are the velocity and coordinate of a particle with the initial velocity \({{{v}}_{0}}\) at the time tm. The dependence \({v}\)(\({{{v}}_{0}}\), tm) is determined by means of the numerical integration of Eq. (1) taking into account a change in the drag law behind the shockwave front in air [36, 37]. The distribution given by (B.1) is obtained from the solution of the one-dimensional kinetic equation, where the drag force is taken into account and collisions between particles are neglected.

After integration over \({{{v}}_{0}}\), Eq. (B.1) can be represented in the form

$$\begin{gathered} N({v},z,{{t}_{{\text{m}}}}) = \frac{{{{N}_{{\text{a}}}}}}{{{{{v}}_{{{\text{fs}}}}}}}\sum\limits_i^{} {F\left( {\frac{{{v}_{0}^{{(i)}}}}{{{{{v}}_{{{\text{fs}}}}}}}} \right)} \\ \, \times {{\left| {\frac{{dz}}{{d{{{v}}_{0}}}}({v}_{0}^{{(i)}},{{t}_{{\text{m}}}})} \right|}^{{ - 1}}}\delta ({v} - {v}({v}_{0}^{{(i)}},{{t}_{{\text{m}}}})), \\ \end{gathered} $$
(B.2)

where summation is performed over the roots \({v}_{0}^{{(i)}}\) = \({v}_{0}^{{(i)}}\)(z, tm) of the equation z = z(\({{{v}}_{0}}\), tm). The distribution N(\({v}\), z, tm) on the (\({v}\), z) plane is nonzero on the \({v}\) = \({v}\)(z) curve (see Fig. 11), which is obtained by excluding \({{{v}}_{0}}\) from the system of equations \({v}\) = \({v}\)(\({{{v}}_{0}}\), tm) and z = z(\({{{v}}_{0}}\), tm). Each point on this curve corresponds to a certain initial velocity \({{{v}}_{0}}\). Summation in Eq. (B.2) is equivalent to summation over groups of particles corresponding to monotonic segments in the dependence \({v}\)(z).

Formulas (B.1) and (B.2) describe the evolution of the distribution of particles with a given diameter. To calculate the volume density distribution of the ejected mass ρ(\({v}\), z, tm) taking into account the dispersion of sizes d of dust particles, it is necessary to multiply Eq. (B.1) or (B.2) by the mass of a single particle md = ρ0πd3/6 and to integrate over d with the distribution n(d). The result can be expressed in terms of the areal density of ejecta, using the relation Na = ρa/\(\overline {{{m}_{{\text{d}}}}} \), where \(\overline {{{m}_{{\text{d}}}}} \) is the average mass of the particle. In this work, it is assumed that the particle size distribution n(d) is independent of their initial velocity (according to [47, 49], this is valid at \({{{v}}_{0}}\)/\({{{v}}_{{{\text{fs}}}}}\) > 1.1, i.e., for a larger fraction of particles involved in the multiple scattering of the probe signal).

If integration is replaced by discrete summation over sizes, the expression for the volume density ρ(\({v}\), z, tm) is reduced to summation over groups of particles each corresponding to a monotonic segment of the dependence \({v}\)(z) for particles with a given size. The calculations in this work were performed in the 16-group approximation. The distribution n(d) was modeled by an ensemble of particles with four different sizes; four monotonic segments can be separated on the \({v}\)(z) curve for each of four groups of particles (see Fig. 11). We approximated the log-normal distribution given by Eq. (14) with the width σ = 0.5 by a set of four discrete sizes d/dm = 0.8, 1.3, 1.8, 2.3 with respective weight factors of 0.50, 0.32, 0.13, 0.05. An additional increase in the number of discrete values d/dm hardly affects the results of calculation of the volume density and transport scattering coefficient.

Solution of the Transport Equation for the Multigroup Description of the Ejecta Cloud

To solve the transport equation (19) in the case of scattering of waves from the ensemble of particles, which is characterized an ambiguous dependence of the velocity of particles on the z coordinate, we use the multigroup representation of the scattering coefficient

$$\langle {{\sigma }_{{{\text{tr}}}}}\exp [ - i{{k}_{0}}(\mu - \mu {\kern 1pt} '{\kern 1pt} ){v}t]\rangle $$

in the form of expansion (21) into partial contributions. The substitution of Eq. (21) transforms the transport equation (19) to the form

$$\begin{gathered} \left( {\mu \frac{\partial }{{\partial z}} + {{\sigma }_{{{\text{tr}}}}}(z) + \kappa (z)} \right)I(z,\mu ,t) \\ = \frac{1}{{4\pi }}\sum\limits_k^{} {\sigma _{{{\text{tr}}}}^{{(k)}}(z)\exp [ - i{{k}_{0}}\mu {{{v}}_{k}}(z)t]{{\Phi }_{k}}(z,t),} \\ \end{gathered} $$
(C.1)

Here, summation is performed over groups each corresponding to a monotonic segment of the z dependence of the velocity \({{{v}}_{k}}\)(z) for particles with a given diameter,

$${{\Phi }_{k}}(z,t) = 2\pi \int\limits_{ - 1}^1 {d\mu \exp [i{{k}_{0}}\mu {{{v}}_{k}}(z)t]I(z,\mu ,t).} $$
(C.2)

is the partial density, and σtr(z), κ(z), \(\sigma _{{{\text{tr}}}}^{{(k)}}\)(z) and \({{{v}}_{k}}\)(z) parametrically depend on the motion time tm.

Transforming Eq. (C.1) to the integral form and integrating it over the angular variable μ, we obtain the following equation for the partial density Φi:

$${{\Phi }_{i}}(\zeta ,t)\, = \,\sum\limits_k^{} {\int\limits_0^{{{\tau }_{{{\text{tot}}}}}} {d\zeta {\kern 1pt} '{\kern 1pt} {{h}_{{ik}}}(\zeta ,\zeta {\kern 1pt} '{\kern 1pt} ,t){{\Phi }_{k}}(\zeta {\kern 1pt} '{\kern 1pt} ,\zeta )\, + \,\Phi _{i}^{{(0)}}(\zeta {\kern 1pt} ,t)} ,} $$
(C.3)

where the optical depth

$$\zeta = \int\limits_z^\infty {dz{\kern 1pt} '[{{\sigma }_{{{\text{tr}}}}}(z{\kern 1pt} ') + \kappa (z{\kern 1pt} ')].} $$
(C.4)

is introduced instead of the variable z. The transport scattering coefficient σtr(z) in Eq. (C.4) is determined by the sum of partial contributions:

$${{\sigma }_{{{\text{tr}}}}}(z) = \sum\limits_k^{} {\sigma _{{{\text{tr}}}}^{{(k)}}(z).} $$
(C.5)

The total optical thickness of the ejecta cloud is τtot = ζ(z = \({{{v}}_{{{\text{fs}}}}}\)tm). In Eq. (C.3), \(\Phi _{i}^{{(0)}}\)(ζ, t) corresponds to the partial density from a source in the form of a monochromatic plane wave incident on the medium,

$$\Phi _{i}^{{(0)}}(\zeta ,t) = \exp [ - \zeta - i{{k}_{0}}{{{v}}_{i}}(\zeta )\,t].$$
(C.6)

The kernels hik(ζ, ζ′, t) in the system of integral equations (C.3) are given by the expressions

$$\begin{gathered} {{h}_{{ik}}}(\zeta ,\zeta {\kern 1pt} ',t) = \frac{\Lambda }{2}\frac{{\sigma _{{{\text{tr}}}}^{{(k)}}(\zeta {\kern 1pt} ')}}{{\sigma _{{{\text{tr}}}}^{{}}(\zeta {\kern 1pt} ')}}\int\limits_{ - 1}^1 {\frac{{d\mu }}{{{\text{|}}\mu {\text{|}}}}\eta \left( {\frac{{\zeta {\kern 1pt} '\, - \zeta }}{\mu }} \right)} \\ \times \exp \left( { - \frac{{\zeta {\kern 1pt} '\, - \zeta }}{\mu } - i{{k}_{0}}\mu [{{{v}}_{k}}(\zeta {\kern 1pt} ') - {{{v}}_{i}}(\zeta )]\,t} \right), \\ \end{gathered} $$
(C.7)

where it is assumed that the single scattering albedo Λ = σtr/(σtr + κ) is the same for all groups of particles and is independent of the coordinate. The Λ values is determined only by the material of the sample and the wavelength of radiation. Calculations within the Mie theory [54] for the wavelength λ = 1.55 μm, optical constants from [55], and particle sizes 1–5 μm give Λ = 0.89 and 0.83 for Pb and Sn, respectively.

The system of integral equations (C.3) is a multigroup generalization of the well-known Milne equation [53] to the case of moving particles. Equation (C.1) was previously considered in [30] in the single-group approximation (i.e., for the case of a single-valued function \({v}\)(z)). In this work, we use the 16-group approximation (the ensemble of particles with four different sizes for each of which four monotonic segments can be distinguished on the \({v}\)(z) curve (see Appendix B)).

The system of equations (C.3) is numerically integrated on the grid of ζn and μm values, which are roots of Legendre polynomials. For each time t entering into the Doppler phase factors in Eqs. (C.6) and (C.7), we solve the system of corresponding algebraic equations and, thus, form a set of density values Φkn, t).

The backscattered signal intensity I(μ = 1, ζ = 0, t) is calculated by the formula

$$I(t)\, = \,\frac{\Lambda }{{4\pi }}\sum\limits_k^{} {\int\limits_0^{{{\tau }_{{{\text{tot}}}}}} {d\zeta \frac{{\sigma _{{{\text{tr}}}}^{{(k)}}(\zeta )}}{{\sigma _{{{\text{tr}}}}^{{}}(\zeta )}}} } \exp [ - \zeta {\kern 1pt} - \,i{{k}_{0}}{{{v}}_{k}}(\zeta )t]{{\Phi }_{k}}(\zeta ,t).$$
(C.8)

The Doppler spectrum of the backscattered signal I(ω) is then calculated by means of the discrete Fourier transform of Eq. (C.8) in the variable t.

Result (C.8) concerns the case of the backscattering of waves from the free layer (i.e., it is assumed that the reflection of waves from the free surface is absent).

Formula (C.8) can be generalized to the case of a reflecting free surface by introducing the corresponding boundary condition (see [30]) and can be represented in the form of an expansion of the solution in successive reflections from the boundary. This expansion can be summed for the case of a diffusely reflecting surface [30]. The simplest result is obtained under the assumption that the reflection coefficient of the surface is unity and reflection is specular [28, 63]. In this case, the backscattering intensity can be expressed in terms of the solution for a free layer with a doubled optical thickness. The intensity is given by Eq. (C.8) after the substitution of Φk(ζ, t) calculated for the layer with the thickness 2τtot, the substitution of 2τtot for the upper limit of the integral, and the replacement of one exponential factor by the sum

$$\begin{gathered} \exp [ - \zeta - i{{k}_{0}}{{{v}}_{k}}(\zeta )t] \\ + \exp [ - 2{{\tau }_{{{\text{tot}}}}} + \zeta - i{{k}_{0}}(2{{{v}}_{{{\text{fs}}}}} - {{{v}}_{k}}(\zeta ))t]. \\ \end{gathered} $$

This result makes it possible to estimate the maximum effect of reflection of waves from the surface on the backscattered signal.

The assumption of specular reflection from the surface [28, 63] is valid for two reasons. First, the character of reflection (specular or diffuse) is insignificant for multiply scattered radiation [52, 53]. Second, the reflection coefficient of radiation from metallic samples is close to unity (in the case of normal incidence, r = 0.92 and 0.87 for Pb and Sn, respectively [55], and the angular-averaged coefficients are 0.9 and 0.85, respectively).

For experimentally implemented optical thicknesses of the ejecta cloud τtot ≫ 1 and Λ values corresponding to Pb and Sn particles, the effect of reflection of waves from the free surface on the backscattered signal spectrum is insignificant (even for channel 2 in experiment 1).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Andriyash, A.V., Dyachkov, S.A., Zhakhovsky, V.V. et al. Photon Doppler Velocimetry and Simulation of Ejection of Particles from the Surface of Shock-Loaded Samples. J. Exp. Theor. Phys. 130, 338–357 (2020). https://doi.org/10.1134/S1063776120030097

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1134/S1063776120030097

Navigation