Skip to main content
Log in

On experimental and numerical study of the dynamics of a liquid metal jet hit by a laser pulse

  • Research Article
  • Published:
Experiments in Fluids Aims and scope Submit manuscript

Abstract

In this paper, we present an experimental and numerical study of the laser pulse impact on the liquid metal jet target. The jet motion was recorded with the stroboscopic ultrafast shadow photography. Simulations were carried out in a two-step approach. At the first step, we simulated laser interaction with the target using the radiation hydrodynamics code 3DLINE, which accounts for a number of effects: liquid–gas phase transition, dynamics of ionization, radiation transfer, laser reflection, refraction and absorption. However, this code cannot be used on a deeply refined mesh near the liquid surface and does not account for the surface tension, which strongly affects liquid motion on the microsecond timescale after the laser pulse ends. Therefore, for the second step we employed the OpenFOAM solver, based on the volume-of-fluid method, which overcomes these limitations. The simulated target dynamics is found to be in a fairly good agreement with the experiment.

Graphic abstract

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.

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

Similar content being viewed by others

References

  • Banine VY, Koshelev KN, Swinkels GHPM (2011) Physical processes in euv sources for microlithography. J Phys D Appl Phys 44(25):253001

    Article  Google Scholar 

  • Basko MM, Tsygvintsev IP (2017) A hybrid model of laser energy deposition for multi-dimensional simulations of plasmas and metals. Comput Phys Commun 214:59–70

    Article  MathSciNet  MATH  Google Scholar 

  • Basko MM, Novikov VG, Grushin AS (2015) On the structure of quasi-stationary laser ablation fronts in strongly radiating plasmas. Phys Plasmas 22(5):053111

    Article  Google Scholar 

  • Bogy DB (1979) Drop formation in a circular liquid jet. Annu Rev Fluid Mech 11(1):207–228

    Article  Google Scholar 

  • Brackbill JU, Kothe DB, Zemach C (1992) A continuum method for modeling surface tension. J Comput Phys 100(2):335–354

    Article  MathSciNet  MATH  Google Scholar 

  • Coetzee RV (2005) Volume weighted interpolation for unstructured meshes in the finite volume method. Ph.D. Thesis, North-West University, Potchefstroom Campus

  • Damián S M (2009) Description and utilization of inter Foam multiphase solver. Report, pp 1–64 (2009)

  • Eggers J (1997) Nonlinear dynamics and breakup of free-surface flows. Rev Mod Phys 69(3):865–929

    Article  MATH  Google Scholar 

  • Faik S, Tauschwitz A, Iosilevskiy I (2018) The equation of state package feos for high energy density matter. Comput Phys Commun 227:117–125

    Article  Google Scholar 

  • Ferziger J H, Peric M (2002) Computational methods for fluid dynamics. Springer, New York

    Book  MATH  Google Scholar 

  • Gates DS, Thodos G (1960) The critical constants of the elements. AIChE J 6(1):50–54

    Article  Google Scholar 

  • Hansson BAM, Hertz HM (2004) Liquid-jet laser-plasma extreme ultraviolet sources: from droplets to filaments. J Phys D Appl Phys 37(23):3233

    Article  Google Scholar 

  • Hemberg O, Otendal M, Hertz HM (2003) Liquid-metal-jet anode electron-impact X-ray source. Appl Phys Lett 83(7):1483–1485

    Article  Google Scholar 

  • Hirt CW, Nichols BD (1981) Volume of fluid (VOF) method for the dynamics of free boundaries. J Comput Phys 39(1):201–225

    Article  MATH  Google Scholar 

  • Issa RI (1986) Solution of the implicitly discretised fluid flow equations by operator-splitting. J Comput Phys 62(1):40–65

    Article  MathSciNet  MATH  Google Scholar 

  • Jansson PAC, Hansson BAM, Hemberg O, Otendal M, Holmberg A, de Groot J, Hertz HM (2004) Liquid-tin-jet laser-plasma extreme ultraviolet generation. Appl Phys Lett 84(13):2256–2258

    Article  Google Scholar 

  • Jasak H (1996) Error analysis and estimation for the finite volume method with applications to fluid flows. Ph.D. Thesis, Department of Mechanical Engineering, Imperial College London

  • Kemp AJ, Meyer-ter Vehn J (1998) An equation of state code for hot dense matter, based on the qeos description. Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 415(3):674–676

  • Kim DA, Vichev IY (2017) Modeling of liquid tin target deformation by laser pulse (122):19 (IPM Preprints)

  • Koshelev K, Krivtsun V, Ivanov V, Yakushev O, Chekmarev A, Koloshnikov V, Snegirev E, Medvedev V (2012) New type of discharge-produced plasma source for extreme ultraviolet based on liquid tin jet electrodes. J Micro/Nanolithogr MEMS MOEMS 11(2):021103–1

    Google Scholar 

  • Krukovskiy AYu, Novikov VG, Tsygvintsev IP (2017) 3d simulation of the impact made by a noncentral laser pulse on a spherical tin target. Math Models Comput Simul 9(1):48–59

    Article  MathSciNet  Google Scholar 

  • Kurilovich D, Klein AL, Torretti F, Lassise A, Hoekstra R, Ubachs W, Gelderblom H, Versolato OO (2016) Plasma propulsion of a metallic microdroplet and its deformation upon laser impact. Phys Rev Appl 6(1):014018

    Article  Google Scholar 

  • Kurilovich D, Pinto TP, de Faria Torretti F, Schupp R, Scheers J, Stodolna AS, Gelderblom H, Eikema KSE, Witte S, Ubachs W, Hoekstra R, Versolato OO (2018) Expansion dynamics after laser-induced cavitation in liquid tin microdroplets. Phys Rev Appl 10(5):054005

    Article  Google Scholar 

  • Kurilovich D, Basko MM, Kim DA, Torretti F, Schupp R, Visschers JC, Scheers J, Hoekstra R, Ubachs W, Versolato OO (2018) Power-law scaling of plasma pressure on laser-ablated tin microdroplets. Phys Plasmas 25(1):012709

    Article  Google Scholar 

  • Larsson DH, Takman PAC, Lundström U, Burvall A, Hertz HM (2011) A 24 kev liquid-metal-jet X-ray source for biomedical applications. Rev Sci Instrum 82(12):123701

    Article  Google Scholar 

  • Leroux S, Dumouchel C, Ledoux M (1996) The stability curve of newtonian liquid jets. At Sprays 6(6):623–647

    Article  Google Scholar 

  • Leroux S, Dumouchel C, Ledoux M (1997) The break-up length of laminar cylindrical liquid jets. Modification of Weber’s theory. Int J Fluid Mech Res 24(1–3):428–438

    Article  Google Scholar 

  • Lin SP, Lian ZW (1990) Mechanisms of the breakup of liquid jets. AIAA J 28(1):120–126

    Article  Google Scholar 

  • McCarthy MJ, Molloy NA (1974) Review of stability of liquid jets and the influence of nozzle design. Chem Eng J 7(1):1–20

    Article  Google Scholar 

  • Mucha PB (2003) On Navier-Stokes equations with slip boundary conditions in an infinite pipe. Acta Applicandae Mathematicae 76:1–15

    Article  MathSciNet  MATH  Google Scholar 

  • Nikiforov AF, Novikov VG, Uvarov VB (2005) Quantum-statistical models of hot dense matter. Methods for computation opacity and equation of state. Birkhäuser, Switzerland

    MATH  Google Scholar 

  • Novikov VG, Solomyannaya AD (1998) Spectral characteristics of plasma consistent with radiationf. High Temp 36(6):835–841

    Google Scholar 

  • Rajyaguru C, Higashiguchi T, Koga M, Kawasaki K, Hamada M, Dojyo N, Sasaki W, Kubodera S (2005) Parametric optimization of a narrow-band 13.5-nm emission from a li-based liquid-jet target using dual nano-second laser pulses. Appl Phys B 80(4):409–412

    Article  Google Scholar 

  • Schäfer M (2006) Computational engineering: introduction to numerical methods. Springer, New York

    Google Scholar 

  • Sterling AM, Sleicher CA (1975) The instability of capillary jets. J Fluid Mech 68(3):477–495

    Article  MATH  Google Scholar 

  • Torrisi L, Margarone D (2006) Investigations on pulsed laser ablation of sn at 1064 nm wavelength. Plasma Sour Sci Technol 15(4):635

    Article  Google Scholar 

  • Tsygvintsev IP, Krukovskiy AYu, Gasilov VA, Novikov VG, Popov IV (2016) Mesh-ray model and method for calculating the laser radiation absorption. Math Models Comput Simul 8(4):382–390

    Article  Google Scholar 

  • Versteeg HK, Malalasekera W (1995) An introduction to computational fluid dynamics. The finite volume method. Longman, England

    Google Scholar 

  • Vinokhodov A, Krivokorytov M, Sidelnikov Yu, Krivtsun V, Medvedev V, Bushuev V, Koshelev K, Glushkov D, Ellwi S (2016) Stable droplet generator for a high brightness laser produced plasma extreme ultraviolet source. Rev Sci Instrum 87(10):103304

    Article  Google Scholar 

  • Yu Y, Wang Q, Yi L, Liu J (2014) Channelless fabrication for large-scale preparation of room temperature liquid metal droplets. Adv Eng Mater 16(2):255–262

    Article  Google Scholar 

  • Yu Y, Wang Q, Wang XL, Wu YH, Liu J (2016) Liquid metal soft electrode triggered discharge plasma in aqueous solution. RSC Adv 6(115):114773–114778

    Article  Google Scholar 

Download references

Acknowledgements

This work was funded by the Russian Science Foundation through Grant No. 14-11-00699. Calculations have been performed at HPC MVS-10P (JSCC RAS) and HPC K100 (KIAM RAS). We thank Mikhail Basko for proofreading the manuscript and for fruitful discussions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ilia Vichev.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: 3DLINE CODE

In the current simulations, we used a one-fluid one-temperature ideal plasma model with quasi-stationary ionization. The general set of differential equations can be written in the form

$$\begin{aligned} \begin{aligned} \frac{\mathrm{d} \rho }{\mathrm{d}t}&=-\rho \nabla \cdot \vec v,\\ \rho \frac{\mathrm{d}\vec v}{\mathrm{d}t}&= -\nabla P,\\ \frac{\mathrm{d}\varepsilon }{\mathrm{d}t}&=-P \frac{\mathrm{d}(1/\rho )}{\mathrm{d}t} - \frac{1}{\rho }\nabla \cdot \vec W + G^\mathrm{rad} + G^\mathrm{las},\\ \end{aligned} \end{aligned}$$
(1)

where \(\mathrm{d}/\mathrm{d}t=\partial /\partial t + \vec v\cdot \nabla\) is the material derivative, \(\vec v\) is the velocity, \(\rho\) is the density, \(P(\rho ,T)\) is the pressure, \(\varepsilon (\rho ,T)\) is the internal energy per unit mass, \(\vec W=-\chi \nabla T\) is the thermoconductive heat flux, \(\chi (\rho ,T)\) is the coefficient of thermal conductivity, \(G^\mathrm{rad}\) is the source/sink of energy per unit mass due to the absorption/generation of thermal radiation and \(G^\mathrm{las}\) is the laser energy deposition per unit mass. The thermodynamic properties were calculated as a function of density and temperature by using the FEOS code (Faik et al. 2018; Kemp and Meyer-ter Vehn 1998), based on the Thomas–Fermi model with quasi-empirical corrections in the low-temperature region.

For the radiation transport, we used the multi-group diffusion approach in quasi-stationary approximation (without the \(\displaystyle \frac{1}{c}\frac{\mathrm{d}}{\mathrm{d}t}\) term which is negligible for \(T\ll 1\) keV):

$$\begin{aligned} \begin{aligned} \rho G^\mathrm{rad}&=-\sum \limits _\nu \left( 4\pi j_\nu -c\varkappa ^P_{\nu } U_{\nu }\right) ,\\ \nabla \cdot \vec W_{\nu }&=4\pi j_\nu -c\varkappa ^P_{\nu } U_{\nu },\\ \vec W_{\nu }&=-\frac{c}{3\varkappa ^R_{\nu }}\nabla U_\nu . \end{aligned} \end{aligned}$$
(2)

Here, index \(\nu\) denotes the spectral group, \(j_\nu\) is the group emissivity, c is the speed of light, \(\varkappa ^{P,R}_{\nu }\) is the mean Plank or Rosseland opacity, \(U_{\nu }\) is the local radiation energy density in group \(\nu\), and \(\vec W_{\nu }\) is the corresponding energy flux. In this work, we used the interpolation (Novikov and Solomyannaya 1998) of the opacity and emissivity coefficients between the two tables for the limiting cases of a transparent plasma and of an optically thick plasma, both generated by the THERMOS code (Nikiforov et al. 2005).

The transport and absorption of the laser light were calculated with the hybrid model, described in detail in Refs. Basko and Tsygvintsev (2017), Tsygvintsev et al. (2016). The detailed description of the discretization scheme and the numerical algorithm can be found in Ref. Krukovskiy et al. (2017).

Appendix B: OpenFOAM PACKAGE

Unlike the 3DLINE code, the interDyMFOAM solver is based on the incompressible fluid model and does not include temperature. That is, it neglects any processes governing the temperature dynamics (such as thermal conduction, radiative heat transfer and heating by compression) as well as the temperature dependence of material properties such as its normal density, surface tension and phase transitions. At the same time, the dynamic effects of the surface tension and viscosity are fully accounted for. Gravity could also be included, but, being of negligible influence in the current setting, it was ignored in order to preserve the reflection symmetry with respect to the horizontal plane \(x=0\).

Because the liquid–gas phase transition is not described by the model, the density distribution, received from the 3DLINE code, was interpreted as a mixture of two immiscible fluids, namely liquid tin and a tin vapor at low density. During the OpenFOAM simulation, these fluids are considered as one effective fluid, whose physical properties are calculated as weighted averages according to the volume-of-fluid (VOF) method (Hirt and Nichols 1981). The fluid density \(\rho\) and its dynamic viscosity \(\mu\) are obtained as

$$\begin{aligned} \rho = \rho _l\gamma + \rho _g(1 - \gamma ). \end{aligned}$$
$$\begin{aligned} \mu = \mu _l\gamma + \mu _g(1 - \gamma ), \end{aligned}$$

where the subscripts l and g denote the liquid and the gaseous phases, respectively. The phase fraction \(\gamma\) can take values within the range \(0\le \gamma \le 1\), with the values of zero and one corresponding to the regions accommodating only one phase, i.e., \(\gamma =0\) for gas and \(\gamma =1\) for liquid. The same approach applies to the fluid velocity

$$\begin{aligned} \vec {v} = \vec {v}_l\gamma + \vec {v}_g(1 - \gamma ). \end{aligned}$$

The evolution of the phase fraction is described by the equation

$$\begin{aligned} \frac{\partial \gamma }{\partial t} + \nabla \cdot (\vec {v}\gamma ) + \nabla \cdot [\vec {v}_r \gamma (1 - \gamma )] = 0, \end{aligned}$$
(3)

where

$$\begin{aligned} \vec {v}_r = \vec {v}_l - \vec {v}_g \end{aligned}$$
(4)

is the relative velocity vector, which should be orthogonal to the phase interface,

$$\begin{aligned} \vec {v}_r \times \nabla \gamma = 0. \end{aligned}$$
(5)

The surface tension at the liquid–gas interface generates an additional pressure gradient resulting in a force, which is evaluated per unit volume using the continuum surface force (CSF) model (Brackbill et al. 1992). For a constant surface tension \(\sigma\), it can be written as

$$\begin{aligned} \vec {f}_\sigma = \sigma \kappa \nabla \gamma , \end{aligned}$$

where \(\kappa\) is the mean curvature of the free surface determined from the expression

$$\begin{aligned} \kappa = - \nabla \cdot \left( \frac{\nabla \gamma }{\left| \nabla \gamma \right| }\right) . \end{aligned}$$

For incompressible fluid, the continuity equation has the form

$$\begin{aligned} \nabla \cdot \vec {v} = 0. \end{aligned}$$
(6)

Finally, with the account of everything mentioned above, the momentum equation in the Cartesian index notation takes the form

$$\begin{aligned} \begin{aligned} \frac{\partial }{\partial t}(\rho v_i) +&\frac{\partial }{\partial r_k}(\rho v_i v_k) - \frac{\partial }{\partial r_k}\left( \mu \frac{\partial v_i}{\partial r_k}\right) - \frac{\partial v_k}{\partial r_i}\frac{\partial \mu }{\partial r_k} =\\ =&-\frac{\partial p}{\partial r_i} + \sigma \kappa \frac{\partial \gamma }{\partial r_i}. \end{aligned} \end{aligned}$$
(7)

Here, p is the pressure, which is derived from the discretized form of the continuity and momentum equations (Issa 1986).

Thus, the physical model consists of the phase fraction evolution equation (3), the continuity equation (6), the momentum equation (7) and condition (5) for the “compression velocity” (4). The numerical algorithm for solving this system is based on the PISO cycle (Issa 1986), which provides the pressure–velocity coupling (Damián 2009). This coupling is achieved in the following way. At each iteration cycle, first the velocity is predicted by using the pressure field from the previous iteration. Then, by using this velocity field, the new pressure is found by solving the momentum equation, and the mass flow is recovered. At last, the velocity field is corrected according to the new mass flow. The numerical solution is performed using the finite volume method (Jasak 1996; Versteeg and Malalasekera 1995; Ferziger and Peric 2002).

Appendix C: Boundary conditions

As was mentioned before, due to the symmetry of the problem the simulations were carried out in 1/4 of the full volume, using the reflective boundary conditions in the planes \(x = 0\) and \(y = 0\), and the free outflow conditions at the other boundaries. For equations (1), governing matter motion and the conduction heat flow in the 3DLINE code, these conditions have the form

$$\begin{aligned} \left. P\right| _\mathrm{free\ outflow}&=0,\\ \left. \vec n\cdot \vec W\right| _\mathrm{free\ outflow}&=0,\\ \left. \vec n\cdot \vec v\right| _\mathrm{reflective}&=0,\\ \left. \vec n\cdot \vec W\right| _\mathrm{reflective}&=0,\\ \end{aligned}$$

where \(\vec n\) is the unit outward normal to the bounding surface. Conductive heat flow through the domain boundaries can be neglected, because the energy is mainly carried out by the convective term \(\vec v\cdot \nabla \varepsilon\) and the radiative transfer term \(\vec W_\nu\). The boundary conditions for the radiation transfer equation (2) are written in the Marshack form

$$\begin{aligned} \left. \vec n\cdot \vec W_{\nu }\right| _\mathrm{reflective}&=0,\\ \left. \vec n\cdot \vec W_{\nu }\right| _\mathrm{free\ outflow}&=\frac{1}{2}c U_{\nu }.\\ \end{aligned}$$

For the laser transfer, we considered a parallel incoming beam in the \(+z\) direction with the Gaussian intensity distribution

$$\begin{aligned} \left. I\right| _{z=z_{min}}=I_0(t)\cdot \exp \left( -8\frac{x^2+y^2}{D_{bw}}\right) , \end{aligned}$$

where \(D_{bw}=100~\upmu\)m is the diameter at the \(1/e^2\) level and \(I_0(t)\) is the intensity on the axis. The traced geometric rays were reflected from the \(x=0\) and \(y=0\) planes, and freely exited at all the other boundaries.

In the OpenFOAM, we used the most appropriate option from the available set of implemented boundary conditions. For the flow velocity \(\vec {v}\), at both reflective boundaries the “zeroGradient” condition of

$$\begin{aligned} \left. \vec {v}\cdot \vec {n}\right| _\mathrm{reflective} = 0 \end{aligned}$$
(8)

was applied. For any other flow variable \(\phi\) (like the phase fraction or the pressure), the “symmetryPlane” condition (Mucha 2003)

$$\begin{aligned} \left. \nabla \phi \cdot \vec {n}\right| _{y=0}= 0 \end{aligned}$$
(9)

was used in the \(y=0\) plane. And in the plane \(x=0\), we used the “slip” condition (Schäfer 2006)

$$\begin{aligned} \left. \vec {\tau }\cdot T \cdot \vec {n}\right| _{x=0} = 0, \end{aligned}$$
(10)

where T is the deviatoric viscous stress tensor and \(\vec {\tau }\) is a unit tangential vector. In our setting, the boundary conditions (8), (9) and (8), (10) have the same physical meaning, but different numeric realization.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Iartsev, B., Vichev, I., Tsygvintsev, I. et al. On experimental and numerical study of the dynamics of a liquid metal jet hit by a laser pulse. Exp Fluids 61, 119 (2020). https://doi.org/10.1007/s00348-020-02952-4

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00348-020-02952-4

Navigation