Hostname: page-component-7c8c6479df-p566r Total loading time: 0 Render date: 2024-03-28T18:05:54.427Z Has data issue: false hasContentIssue false

The generalized Onsager model for a binary gas mixture

Published online by Cambridge University Press:  23 July 2014

V. Kumaran*
Affiliation:
Department of Chemical Engineering, Indian Institute of Science, Bangalore 560 012, India
S. Pradhan
Affiliation:
Chemical Technology Group, Bhabha Atomic Research Centre, Mysore 571 130, India
*
Email address for correspondence: kumaran@chemeng.iisc.ernet.in

Abstract

The Onsager model for the secondary flow field in a high-speed rotating cylinder is extended to incorporate the difference in mass of the two species in a binary gas mixture. The base flow is an isothermal solid-body rotation in which there is a balance between the radial pressure gradient and the centrifugal force density for each species. Explicit expressions for the radial variation of the pressure, mass/mole fractions, and from these the radial variation of the viscosity, thermal conductivity and diffusion coefficient, are derived, and these are used in the computation of the secondary flow. For the secondary flow, the mass, momentum and energy equations in axisymmetric coordinates are expanded in an asymptotic series in a parameter $\def \xmlpi #1{}\def \mathsfbi #1{\boldsymbol {\mathsf {#1}}}\let \le =\leqslant \let \leq =\leqslant \let \ge =\geqslant \let \geq =\geqslant \def \Pr {\mathit {Pr}}\def \Fr {\mathit {Fr}}\def \Rey {\mathit {Re}}\epsilon = (\Delta m/ m_{av})$, where $\Delta m$ is the difference in the molecular masses of the two species, and the average molecular mass $m_{av}$ is defined as $m_{av}= (\rho _{w1} m_1 + \rho _{w2} m_2)/\rho _w$, where $\rho _{w1}$ and $\rho _{w2}$ are the mass densities of the two species at the wall, and $\rho _w = \rho _{w1} + \rho _{w2}$. The equation for the master potential and the boundary conditions are derived correct to $O(\epsilon ^2)$. The leading-order equation for the master potential contains a self-adjoint sixth-order operator in the radial direction, which is different from the generalized Onsager model (Pradhan & Kumaran, J. Fluid Mech., vol. 686, 2011, pp. 109–159), since the species mass difference is included in the computation of the density, viscosity and thermal conductivity in the base state. This is solved, subject to boundary conditions, to obtain the leading approximation for the secondary flow, followed by a solution of the diffusion equation for the leading correction to the species mole fractions. The $O(\epsilon )$ and $O(\epsilon ^2)$ equations contain inhomogeneous terms that depend on the lower-order solutions, and these are solved in a hierarchical manner to obtain the $O(\epsilon )$ and $O(\epsilon ^2)$ corrections to the master potential. A similar hierarchical procedure is used for the Carrier–Maslen model for the end-cap secondary flow. The results of the Onsager hierarchy, up to $O(\epsilon ^2)$, are compared with the results of direct simulation Monte Carlo simulations for a binary hard-sphere gas mixture for secondary flow due to a wall temperature gradient, inflow/outflow of gas along the axis, as well as mass and momentum sources in the flow. There is excellent agreement between the solutions for the secondary flow correct to $O(\epsilon ^2)$ and the simulations, to within 15 %, even at a Reynolds number as low as 100, and length/diameter ratio as low as 2, for a low stratification parameter $\mathcal{A}$ of 0.707, and when the secondary flow velocity is as high as 0.2 times the maximum base flow velocity, and the ratio $2 \Delta m / (m_1 + m_2)$ is as high as 0.5. Here, the Reynolds number $\mathit{Re}= \rho _w \varOmega R^2 / \mu $, the stratification parameter $\mathcal{A}= \sqrt{m \varOmega ^2 R^2 / (2k_BT)}$, $R$ and $\varOmega $ are the cylinder radius and angular velocity, $m$ is the molecular mass, $\rho _w$ is the wall density, $\mu $ is the viscosity and $T$ is the temperature. The leading-order solutions do capture the qualitative trends, but are not in quantitative agreement.

Type
Papers
Copyright
© 2014 Cambridge University Press 

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

Footnotes

This work forms part of the PhD Thesis of Sahadev Pradhan, Indian Institute of Science, 2014.

References

Berger, M. H. 1987 Finite element analysis of flow in a gas-filled rotating annulus. Intl J. Numer. Meth. Fluids 7, 215231.CrossRefGoogle Scholar
Bird, G. A. 1994 Molecular Gas Dynamics and the Direct Simulation of Gas Flows. Clarendon Press.CrossRefGoogle Scholar
Chapman, S. & Cowling, T. G. 1970 The Mathematical Theory of Non-Uniform Gases. Cambridge University Press.Google Scholar
Gunzburger, M. D. & Wood, H. G. 1982 A finite element method for the Onsager pancake equation. Comput. Meth. Appl. Mech. Engng 31, 4359.CrossRefGoogle Scholar
Gunzburger, M. D., Wood, H. G. & Jordan, J. A. 1984 A finite element method for gas centrifuge problem. SIAM J. Sci. Stat. Comput. 5, 7894.CrossRefGoogle Scholar
Lahargue, J. P. & Soubbaramayer, 1978 A numerical model for the investigation of the flow and isotope concentration field in an ultracentrifuge. Comput. Meth. Appl. Mech. Engng 15 (2), 259273.CrossRefGoogle Scholar
Olander, D. R. 1981 The theory of uranium enrichment by the gas centrifuge. Prog. Nucl. Energy 8, 133.CrossRefGoogle Scholar
Pradhan, S. & Kumaran, V. 2011 The generalized Onsager model for the secondary flow in a high-speed rotating cylinder. J. Fluid Mech. 686, 109159.CrossRefGoogle Scholar
Roberts, W. W.1985 Three dimensional stratified gas flows past impact probes and scoops: $N$ -body Monte Carlo calculations. In Proc. Sixth Workshop on Gases in Strong Rotation, Tokyo (ed. Y. Takashima), p. 115.Google Scholar
Roberts, W. W. & Hausman, M. A. 1988 Hypersonic, stratified gas flows past an obstacle: direct simulation Monte Carlo calculations. J. Comput. Phys. 77, 283317.CrossRefGoogle Scholar
Soubbaramayer,  & Billet, J. 1980 A numerical method for optimizing the gas flow field in a centrifuge. Comput. Meth. Appl. Mech. Engng 24, 165185.CrossRefGoogle Scholar
Viecelli, J. A. 1983 Exponential difference operator approximation for the sixth order Onsager equation. J. Comput. Phys. 50, 162170.CrossRefGoogle Scholar
Wood, H. G., Jordan, J. A. & Gunzburger, M. D. 1984 The effect of curvature on the flow field in rapidly rotating gas centrifuges. J. Fluid Mech. 140, 373395.CrossRefGoogle Scholar
Wood, H. G., Mason, T. C. & Soubbaramayer, 1996 Multi-isotope separation in a gas centrifuge using Onsager’s pancake model. Sep. Sci. Technol. 31 (9), 11851213.CrossRefGoogle Scholar
Wood, H. G. & Morton, J. B. 1980 Onsager’s pancake approximation for the fluid dynamics of a gas centrifuge. J. Fluid Mech. 101, 131.CrossRefGoogle Scholar
Wood, H. G. & Sanders, G. 1983 Rotating compressible flows with internal sources and sinks. J. Fluid Mech. 127, 299311.CrossRefGoogle Scholar