Skip to main content
Log in

High-Temperature Liquid Metal Infusion Considering Surface Tension-Viscosity Dissipation

  • Published:
Metallurgical and Materials Transactions B Aims and scope Submit manuscript

Abstract

In considering the significant effect of the surface tension-viscosity dissipation driving the fluid flow within a capillary, high-temperature liquid metal infusion was analyzed for titanium, yttrium, hafnium, and zirconium penetrating into a packed bed. A model of the dissipation considers the momentum balance within the capillary to determine the rate of infusion, which is compared with the Semlak-Rhines model developed for liquid metal penetration into a packed bed assumed as a bundle of tubes mimicking the porosity of a packed bed. For liquid Ti, the penetration rate was calculated from 0.2 µs to 1 ms and rose to a maximum of 7 m/s at approximately 1 µs; after which, the rate decreased to 0.7 m/s at 1 ms. Beyond 10 µs, the decreasing trend of the rate of penetration determined by the model of dissipation compared favorably with the Semlak-Rhines equation.

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

Similar content being viewed by others

References

  1. F. Delanny, L. Froyen, and A. Deruyttere: J. Mater. Sci., 1987, vol. 22, pp. 1–16.

    Article  Google Scholar 

  2. N. Eustathopoulos: Acta Mater., 1998, vol. 46, pp. 2319–27.

    Google Scholar 

  3. E. Saiz and A.P. Tomsia: Curr. Opin. Solid State Mater. Sci., 2005, vol. 9, pp. 167–73.

    Article  Google Scholar 

  4. R.E. Loehman, K. Ewsuk, and A.P. Tomsia: J. Am. Ceram. Soc., 1996, vol. 79, pp. 27–32.

    Article  Google Scholar 

  5. W.B. Johnson, A.S. Nagelberg, and E. Breval: J. Am. Ceram. Soc., 1991, vol. 74, pp. 2093–101.

    Article  Google Scholar 

  6. N. Maheswaraiah, A.V. Sandate, and A. Bronson: Int. J. Appl. Ceram. Technol., 2013, vol. 10, pp. 234–44.

    Article  Google Scholar 

  7. E. Saiz and A.P. Tomsia: Nat. Mater., 2004, vol. 3, pp. 903–9.

    Article  Google Scholar 

  8. E. Saiz, C.-W. Hwang, K. Suganuma, and A.P. Tomsia: Acta Mater., 2003, vol. 51, pp. 3185–97.

    Article  Google Scholar 

  9. N. Grigorenko, V. Poluyanskaya, N. Eustathopoulos, and Y.V. Naidich: in Interfacial Science in Ceramic Joining, vol. 58, A. Bellosi, T. Kosmac, and A. P. Tomsia, eds., Springer, Netherlands, 1998, pp. 57–67.

  10. C.H.P. Lupis: Chemical Thermodynamics of Materials, Elsevier Science, New York, 1983.

    Google Scholar 

  11. M. Benhassine, E. Saiz, A.P. Tomsia, and J. De Coninck: Acta Mater., 2011, vol. 59, pp. 1087–94.

    Article  Google Scholar 

  12. T.P. Swiler and R.E. Loehman: Acta Mater., 2000, vol. 48, pp. 4419–24.

    Article  Google Scholar 

  13. P. Delgado and V. Kumar: Appl. Math. Comput., 2015, vol. 266, pp. 328–38.

    Article  Google Scholar 

  14. P. Delgado and V. Kumar: ASME 2013 Fluids Eng. Div. Summer Meet, American Society of Mechanical Engineers, 2013, p. V01CT23A001.

  15. F.N. Rhines and K.A. Semlak: Trans Met. Soc AIME, 1958, vol. 21, pp. 325–31.

    Google Scholar 

  16. G.P. Martins, D.L. Olson, and G.R. Edwards: Metall. Trans. B, 1988, vol. 19B, pp. 95–101.

    Article  Google Scholar 

  17. J.R. Ligenza and R.B. Bernstein: J. Am. Chem. Soc., 1951, vol. 73, pp. 4636–38.

    Article  Google Scholar 

  18. P. Paradis, T. Ishikawa, G. Lee, D. Holland-Moritz, J. Brillo, W. Rhim, and J.T. Okada: Mater. Sci. Eng. R Rep., 2014, vol. 76, pp. 1–53.

    Article  Google Scholar 

  19. T. Ishikawa, P. Paradis, J.T. Okada, and Y. Watanabe: Meas. Sci. Technol., 2012, vol. 23, p. 025305.

    Article  Google Scholar 

  20. T.B. Massalski, H. Okamoto, P.R. Subramanian, and L. Kacpack: Binary Alloy Phase Diagrams, 2nd ed., ASM International, Metals Park, 1990.

    Google Scholar 

  21. P.F. Paradis, T. Ishikawa, and N. Koike: Microgravity Sci. Technol., 2008, vol. 21, pp. 113–18.

    Article  Google Scholar 

Download references

Acknowledgments

This material was based on research sponsored by the Air Force Research Laboratory, under agreement number FA9550-12-1-0242, through the Air Force Office of Scientific Research (AFOSR) and is gratefully acknowledged with the support of Dr. Ali Sayir of AFOSR. The U.S. Government is authorized to reproduce and distribute reprints for governmental purposes notwithstanding any copyright notation thereon. The views and conclusions contained herein are those of the authors and should not be interpreted as necessarily representing the official policies or endorsements, either expressed or implied, of the Air Force Research Laboratory or the U.S. Government. Chris Harris thanks Shell Global Solutions International for supporting his visiting professorship at Imperial College London, under the auspices of which his contribution to this work was carried out.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Arturo Bronson.

Additional information

Manuscript submitted August 18, 2015.

Appendix

Appendix

In this appendix, steps involved in solving the modified-SR equation are presented. We start with Eq. [3] with the appropriate initial conditions:

$$ \begin{aligned} \frac{{d}}{{{{d}}t}}\left( {\left( {\rho \pi R^{2} h + \rho V_{{L}} } \right)\frac{{{{d}}h}}{{{{d}}t}}} \right) = & 2\pi R\sigma \cos \left( \theta \right) - 8\pi \mu \left( {h + h_{0} } \right)\left( {\frac{{{{d}}h}}{{{{d}}t}}} \right) \\ - \rho \left( {\pi R^{2} h + \pi R^{2} h_{{L}} } \right)g - \frac{1}{4}\rho \pi R^{2} \left( {\frac{{{{d}}h}}{{{{d}}t}}} \right)^{2} \\ \end{aligned} $$
(3)
$$ \left. h \right|_{t = 0} = 0;\quad \left. {\frac{{{{d}}h}}{{{{d}}t}}} \right|_{t = 0} = 0. $$
(4)

We simplify Eq. [3] by dividing the left- and right-hand sides by πR 2 and ignoring the gravitational and viscous force contributions resulting from an initial meniscus formation assumption (i.e., right-hand side terms containing h 0 and h L) as they become negligible shortly after the onset of the capillary rise when h >> h L. This yields:

$$ \rho \frac{{d}}{{{{d}}t}}\left( {\{ h + R\tilde{V}_{L} \} \frac{{{{d}}h}}{{{{d}}t}}} \right) = 2\sigma \cos (\theta )/R - 8\mu hR^{ - 2} \frac{{{{d}}h}}{{{{d}}t}} - \rho gh - \frac{1}{4}\rho \left( {\frac{{{{d}}h}}{{{{d}}t}}} \right)^{2} $$
(5)

with \( \tilde{V}_{\text{L}} = V_{\text{L}} /(\pi R^{3} ). \)

The steady-state capillary height (h s) may be obtained by setting all time derivatives to zero in Eq. [5], which then gives:

$$ h_{\text{s}} = \frac{2\sigma \cos (\theta )}{\rho gR}. $$
(6)

We may make use of Eq. [6] to rewrite Eq. [5] as:

$$ \rho \frac{{d}}{{{{d}}t}}\left( {\{ h + R\tilde{V}_{{L}} \} \frac{{{{d}}h}}{{{{d}}t}}} \right) = \rho g(h_{\text{s}} - h) - 8\mu hR^{ - 2} \frac{{{{d}}h}}{{{{d}}t}} - \frac{1}{4}\rho \left( {\frac{{{{d}}h}}{{{{d}}t}}} \right)^{2} . $$
(7)

Dividing Eq. [7] by the density ρ, we acquire:

$$ \frac{{d}}{{{{d}}t}}\left( {\{ h + R\tilde{V}_{{L}} \} \frac{{{{d}}h}}{{{{d}}t}}} \right) = g(h_{\text{s}} - h) - \frac{8\mu h}{{\rho R^{2} }}\frac{{{{d}}h}}{{{{d}}t}} - \frac{1}{4}\left( {\frac{{{{d}}h}}{{{{d}}t}}} \right)^{2} . $$
(8)

We introduce the following nondimensional capillary height and time:

$$ \tilde{h} = h/h_{\text{s}} ;\quad \tau = t\sqrt {g/h_{\text{s}} } $$
(9)

Then, Eq. [8] can be transformed into Eq. [10], which has the nondimensional capillary height and time:

$$ \frac{{d}}{{{{d}}\tau }}\left( {\{ \tilde{h} + (R/h_{\text{s}} )\tilde{V}_{\text{L}} \} \frac{{{{d}}\tilde{h}}}{{{{d}}\tau }}} \right) = (1 - \tilde{h}) - \frac{{8\mu h_{\text{s}} \tilde{h}}}{{\rho R^{2} \sqrt {gh_{\text{s}} } }}h\frac{{{{d}}\tilde{h}}}{d\tau } - \frac{1}{4}gh_{s} \left( {\frac{{{{d}}\tilde{h}}}{{{{d}}\tau }}} \right)^{2} . $$
(10)

Equation [10] can be written in nondimensionalized form by introducing a nondimensional viscosity (\( \tilde{\mu } \)):

$$ \frac{{d}}{{{{d}}\tau }}\left( {\{ \tilde{h} + (R/h_{\text{s}} )\tilde{V}_{\text{L}} \} \frac{{{{d}}\tilde{h}}}{{{{d}}\tau }}} \right) = (1 - \tilde{h}) - 8\tilde{\mu }\tilde{h}\frac{{{{d}}\tilde{h}}}{{{{d}}\tau }} - \frac{1}{4}\left( {\frac{{{{d}}\tilde{h}}}{{{{d}}\tau }}} \right)^{2} $$
(11)

with

$$ \tilde{\mu } = \frac{\mu }{{\rho R^{2} \sqrt {\frac{g}{{h_{\text{s}} }}} }}. $$
(12)

We define nondimensional rate of penetration \( \tilde{v} = {{d}}\tilde{h}/{{d}}\tau \)and use the fact that:

$$ \frac{{{{d}}^{2} \tilde{h}}}{{{{d}}\tau^{2} }} = \frac{{{{d}}\tilde{v}}}{{{{d}}\tau }} = \tilde{v}\frac{{{{d}}\tilde{v}}}{{{{d}}\tilde{h}}} = \frac{{d}}{{{{d}}\tilde{h}}}\left( {\frac{1}{2}\tilde{v}^{2} } \right). $$
(13)

Equation [11] can be written as:

$$ (\tilde{h} + \alpha )\frac{{d}}{{{{d}}\tilde{h}}}\left( {\frac{1}{2}\tilde{v}^{2} } \right) + \frac{5}{4}\tilde{v}^{2} = (1 - \tilde{h}) - 8\tilde{\mu }\tilde{h}\tilde{v}, $$
(14)

where \( \alpha = R\tilde{V}_{\text{L}} /h_{\text{s}} \).

Further simplification of the left-hand side of Eq. [14] yields:

$$ \frac{{d}}{{{{d}}\tilde{h}}}\left\{ {(\tilde{h} + \alpha )^{5/2} f} \right\} = \left\{ {(1 - \tilde{h}) - 8\tilde{\mu }\tilde{h}\sqrt {2f} } \right\}(\tilde{h} + \alpha )^{3/2} $$
(15)

with

$$ f(\tilde{h}) = \frac{1}{2}\tilde{v}^{2} . $$
(16)

Equation [15] is subject to the initial condition:

$$ f(0) = 0 $$
(17)

Special Case \( \tilde{\mu } = 0 \)

This case would be true during the initiation of the flow (for Ti, t < 0.1 µs). In this case Eq. [14] may be integrated explicitly by using Eq. [17] to obtain:

$$ (\tilde{h} + \alpha )^{5/2} f = \frac{2}{5}(1 + \alpha )\left\{ {(\tilde{h} + \alpha )^{5/2} - \alpha^{5/2} } \right\} - \frac{2}{7}\left\{ {(\tilde{h} + \alpha )^{7/2} - \alpha^{7/2} } \right\}. $$
(18)

This yields:

$$ f = \frac{2}{5}(1 + \alpha ) - \frac{2}{7}(\tilde{h} + \alpha ) - \beta /(\tilde{h} + \alpha )^{5/2} $$
(19)

with

$$ \beta = \frac{2}{5}\alpha^{5/2} (1 + \frac{2}{7}\alpha ). $$
(20)

Note that, as \( \tilde{h} \) rises from zero, f increases from zero to a maximum and then falls back to zero again, at the maximum value of \( \tilde{h} \), Since α is very small in comparison with unity, we have, approximately:

$$ \tilde{h}_{\hbox{max} } = \frac{7}{5} + \frac{2}{5}\alpha . $$
(21)

This expression is obtained by simply neglecting the last term in the right-hand side of Eq. [19] and setting it to zero. A better approximation is obtained by substituting Eq. [21] into the last term on the right-hand side of Eq. [19] before setting the equation to zero. The result is:

$$ \tilde{h}_{\hbox{max} } = \frac{7}{5} + \frac{2}{5}\alpha - \frac{7}{5}\alpha^{5/2} \frac{{1 + \frac{2}{7}\alpha }}{{\left\{ {\frac{7}{5}(1 + \alpha )} \right\}^{5/2} }} = \frac{7}{5} + \frac{2}{5}\alpha - \frac{5}{7}\sqrt {\frac{5}{7}} \alpha^{5/2} \frac{{1 + \frac{2}{7}\alpha }}{{(1 + \alpha )^{5/2} }}. $$
(22)

Noting that \( \tilde{h}_{\hbox{max} } \) is larger than the steady-state value at \( \tilde{h}_{\hbox{max} } = 1 \), we see that the system is underdamped. We now compute the maximum value of f and, hence, the fluid velocity. Let \( \tilde{h} = \tilde{h}_{m} \)when f is maximized. The value of \( \tilde{h}_{m} \)is obtained by setting the derivative of \( f({{d}}f/{{d}}\tilde{h}) \) to zero:

$$ \frac{{{{d}}f}}{{{{d}}\tilde{h}}}_{{\tilde{h} = \tilde{h}_{m} }} = - \frac{2}{7} + \frac{5}{2}\beta /(\tilde{h}_{m} + \alpha )^{7/2} = 0. $$
(23)

Hence,

$$ \beta /(\tilde{h}_{m} + \alpha )^{5/2} = \frac{4}{35}(\tilde{h}_{m} + \alpha ). $$
(24)

Substituting Eq. [24] back into Eq. [19], with \( \tilde{h} \) set to \( \tilde{h}_{m} \), we acquire:

$$ f_{\hbox{max} } = \frac{2}{5}(1 - \tilde{h}_{m} ). $$
(25)

Now, substituting Eq. [20] into Eq. [24] and rearranging, we obtain:

$$ (\tilde{h}_{m} + \alpha )^{7/2} = \frac{7}{2}\alpha^{5/2} (1 + \frac{2}{7}\alpha ). $$

Then

$$ \tilde{h}_{m} = \left( {(1 + \frac{2}{7}\alpha )\frac{7}{2}\alpha^{5/2} } \right)^{2/7} - \alpha $$
(26)

General μ

In this case, Eq. [15] needs to be solved numerically, subject to the initial condition in Eq. [4]. By dividing the range of \( \tilde{h} \) into intervals \( [\tilde{h}_{k - 1} ,\tilde{h}_{k} ] \); k = 1, 2 … N, with \( \tilde{h}_{0} = 0 \), \( \tilde{h}_{N} = \tilde{h}_{\text{s}} \). Let

$$ f_{k} = f(\tilde{h}_{k} ). $$
(27)

For

$$ \tilde{h}_{k} \le \tilde{h} < \tilde{h}_{k + 1} $$
(28)

we assume that \( f = f_{k} \) is constant on the right-hand side of Eq. [15] during the integration process in the interval \( \tilde{h}_{k} \le \tilde{h} < \tilde{h}_{k + 1} \). Equation [15] can therefore be approximated by:

$$ \frac{{d}}{{{{d}}\tilde{h}}}\left\{ {(\tilde{h} + \alpha )^{5/2} f} \right\} = (1 - \tilde{h})(\tilde{h} + \alpha )^{3/2} - 8\tilde{\mu }\tilde{h}(\tilde{h} + \alpha )^{3/2} \sqrt {2f_{k} } ;\quad \tilde{h}_{k} \le \tilde{h} < \tilde{h}_{k + 1} . $$
(29)

Equation [29] is solved with the condition:

$$ f(h_{k} ) = f_{k} . $$
(30)

Note that Eq. [29] may be written in abbreviated form as:

$$ \frac{{d}}{{{{d}}\tilde{h}}}\left\{ {(\tilde{h} + \alpha )^{5/2} f} \right\} = (1 - B_{k} \tilde{h})(\tilde{h} + \alpha )^{3/2} ;\quad \tilde{h}_{k} \le \tilde{h} < \tilde{h}_{k + 1} $$
(31)

with

$$ B_{k} = 1 + 8\tilde{\mu }\sqrt {2f_{k} } . $$
(32)

Integrating Eq. [31] over the interval \( \left[ {\tilde{h}_{k} ,\tilde{h}_{k + 1} } \right] \) yields the following result:

$$ \begin{aligned} \left( {\tilde{h}_{k + 1} + \alpha } \right)^{5/2} f_{k + 1} - \left( {\tilde{h}_{k} + \alpha } \right)^{5/2} f_{k} = & \frac{2}{5}(1 + B_{k} \alpha )\left( {\tilde{h}_{k + 1} + \alpha } \right)^{5/2} - \frac{2}{7}B_{k} \left( {\tilde{h}_{k + 1} + \alpha } \right)^{7/2} \\ - \frac{2}{5}(1 + B_{k} \alpha )\left( {\tilde{h}_{k} + \alpha } \right)^{5/2} + \frac{2}{7}B_{k} \left( {\tilde{h}_{k} + \alpha } \right)^{7/2} . \\ \end{aligned} $$
(33)

Note that the only assumption made to arrive at the solution as in Eq. [33] is that B k = constant inside the integral interval \( \tilde{h}_{k} \le \tilde{h} < \tilde{h}_{k + 1} \). Also, since the nondimensional depth of penetration has been discretized, values of all \( \tilde{h}_{k} \)’s are known. Equation [33], therefore, can be used to solve for the nondimensional rate of penetration (\( \tilde{v} = \sqrt {2f} \)) by using the following equation:

$$ f_{k + 1} = \left( {\tilde{h}_{k + 1} + \alpha } \right)^{ - 5/2} \left[ {\left( {\tilde{h}_{k} + \alpha } \right)^{{\frac{5}{2}}} f_{k} + \frac{2}{5}\left( {1 + B_{k} \alpha } \right)\left( {\left( {\tilde{h}_{k + 1} + \alpha } \right)^{5/2} } \right) - \frac{2}{7}B_{k} \left( {\left( {\tilde{h}_{k + 1} + \alpha } \right)^{{\frac{7}{2}}} } \right) - \frac{2}{5}\left( {1 + B_{k} \alpha } \right)\left( {\left( {\tilde{h}_{k} + \alpha } \right)^{{\frac{5}{2}}} } \right) + \frac{2}{7}B_{k} \left( {\left( {\tilde{h}_{k} + \alpha } \right)^{7/2} } \right)} \right] . $$
(34)

The nondimensional time is evaluated by using Eq. [9]. By using the solution from Eq. [34] along with those from Eqs. [14] and [16], the relative importance of inertial \( \left( {F_{\text{inertia}} } \right) \), gravitational (\( F_{\text{g}} \)), viscous (\( F_{{{\mu }}} \)), and end-drag (\( F_{{\text{end}}{\text{-}}{\text{drag}}} \)) forces with respect to the surface tension (\( F_{\sigma } \)) force can be computed as \( \sqrt {2f} + \left( {\tilde{h} + {{\alpha }}} \right){{d}}f/{{d}}\tilde{h} \) (inertial/surface tension forces), \( \tilde{h} \) (gravity/surface tension forces), \( 8\tilde{\mu }\tilde{h}\sqrt {2f} \) (viscous/surface tension forces), and \( f/2 \) (end-drag/surface tension forces), respectively.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kumar, V., Harris, C.K., Bronson, A. et al. High-Temperature Liquid Metal Infusion Considering Surface Tension-Viscosity Dissipation. Metall Mater Trans B 47, 108–115 (2016). https://doi.org/10.1007/s11663-015-0518-4

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11663-015-0518-4

Keywords

Navigation