Hostname: page-component-8448b6f56d-cfpbc Total loading time: 0 Render date: 2024-04-20T04:10:13.770Z Has data issue: false hasContentIssue false

The inertial orientation dynamics of anisotropic particles in planar linear flows

Published online by Cambridge University Press:  04 April 2018

Navaneeth K. Marath
Affiliation:
Engineering Mechanics Unit, Jawaharlal Nehru Centre for Advanced Scientific Research, Jakkur, Bangalore, 560 064, India
Ganesh Subramanian*
Affiliation:
Engineering Mechanics Unit, Jawaharlal Nehru Centre for Advanced Scientific Research, Jakkur, Bangalore, 560 064, India
*
Email address for correspondence: sganesh@jncasr.ac.in

Abstract

In the Stokes limit, the trajectories of neutrally buoyant torque-free non-Brownian spheroids in ambient planar linear flows are well known. These flows form a one-parameter family, with the velocity gradient tensor given by $\unicode[STIX]{x1D735}\boldsymbol{u}^{\infty \dagger }=\dot{\unicode[STIX]{x1D6FE}}(\mathbf{1}_{x}^{\prime }\mathbf{1}_{y}^{\prime }+\unicode[STIX]{x1D706}\mathbf{1}_{y}^{\prime }\mathbf{1}_{x}^{\prime })$. The parameter $\unicode[STIX]{x1D706}$ is related to the ratio of the vorticity to the extension (given by $(1-\unicode[STIX]{x1D706})/(1+\unicode[STIX]{x1D706})$), and ranges from $-1$ to 1, with $\unicode[STIX]{x1D706}=1\,,0$ and $-1$ being planar extensional flow, simple shear flow and solid-body rotation respectively. The unit vectors $\mathbf{1}_{x}^{\prime }$ and $\mathbf{1}_{y}^{\prime }$ are unit vectors along the flow and gradient axes of the simple shear flow ($\unicode[STIX]{x1D706}=0$). The trajectories, as described by a unit vector along the spheroid symmetry axis, are closed orbits for $\unicode[STIX]{x1D706}<\unicode[STIX]{x1D706}_{crit}$, where $\unicode[STIX]{x1D706}_{crit}=\unicode[STIX]{x1D705}^{2}(1/\unicode[STIX]{x1D705}^{2})$ for an oblate (a prolate) spheroid of aspect ratio $\unicode[STIX]{x1D705}$. We investigate analytically the orientation dynamics of such a spheroid in the presence of weak inertial effects. The inertial corrections to the angular velocities at $O(Re)$ and $O(St)$, where $Re$ and $St$ are the Reynolds ($Re=\unicode[STIX]{x1D70C}_{f}\dot{\unicode[STIX]{x1D6FE}}L^{2}/\unicode[STIX]{x1D707}$) and Stokes numbers ($St=\unicode[STIX]{x1D70C}_{p}\dot{\unicode[STIX]{x1D6FE}}L^{2}/\unicode[STIX]{x1D707}$) respectively, are derived using a reciprocal theorem formulation. Here, $L$ is the semimajor axis of the spheroid, $\unicode[STIX]{x1D707}$ is the viscosity of the suspending fluid, $\dot{\unicode[STIX]{x1D6FE}}$ is the shear rate, and $\unicode[STIX]{x1D70C}_{p}$ and $\unicode[STIX]{x1D70C}_{f}$ are the particle and fluid densities respectively. A spheroidal harmonics formalism is then used to evaluate the reciprocal theorem integrals and obtain closed-form expressions for the inertial corrections. The detailed examination of these corrections is restricted to the aforementioned Stokesian closed-orbit regime ($\unicode[STIX]{x1D706}<\unicode[STIX]{x1D706}_{crit}$). Here, even weak inertia, for asymptotically long times, of $O(1/(\dot{\unicode[STIX]{x1D6FE}}Re))$ or $O(1/(\dot{\unicode[STIX]{x1D6FE}}St))$, will affect the leading-order orientation distribution on account of the indeterminate nature of the distribution across orbits in the Stokes limit. For $\unicode[STIX]{x1D706}<\unicode[STIX]{x1D706}_{crit}$, inertia results in a drift across the closed orbits in Stokes flow, and this orbital drift is characterized using a multiple time scale analysis. The orbits stabilized by the inertial drift, at $O(Re)$ and $O(St)$, are identified in the $\unicode[STIX]{x1D706}{-}\unicode[STIX]{x1D705}$ plane. For the majority of ($\unicode[STIX]{x1D706},\unicode[STIX]{x1D705}$) combinations, the stabilized orbit is either one confined to the plane of symmetry (the flow-gradient plane) of the ambient flow (the tumbling orbit) or one where the spheroid is aligned with the ambient vorticity vector (the spinning orbit). However, for some ($\unicode[STIX]{x1D706},\unicode[STIX]{x1D705}$) combinations, depending on the initial orientation, the orbit stabilized can be either the spinning or the tumbling orbit, since both orbits have non-trivial basins of attraction, separated by a pair of unstable (repelling) limit cycles, on the unit sphere of orientations. A stochastic orientation decorrelation mechanism in the form of rotary Brownian motion, characterized by a Péclet number, $Pe_{r}$ ($Pe_{r}=\dot{\unicode[STIX]{x1D6FE}}/D_{r}$, where $D_{r}$ is the rotary Brownian diffusivity), is included to eliminate the aforementioned dependence on the initial orientation distribution for certain ($\unicode[STIX]{x1D706}$, $\unicode[STIX]{x1D705}$) combinations. The unique steady-state orientation distribution determined by the combined effect of Brownian motion and inertia is obtained by solving a closed-orbit-averaged drift–diffusion equation. The steady-state orientation dynamics of an inertial spheroid in a planar linear flow, in the presence of weak thermal orientation fluctuations, has similarities to the thermodynamic description of a one-component system. Thus, we identify a tumbling–spinning transition in a $C{-}\unicode[STIX]{x1D705}{-}Re\,Pe_{r}$ space. Here, $C$ is the orbital coordinate that acts as a label for the closed orbits in the Stokes limit. This transition implies hysteretic orientation dynamics in certain regions in the $C$$\unicode[STIX]{x1D705}$$Re\,Pe_{r}$ space, although the hysteretic volume shrinks rapidly on either side of simple shear flow. In the hysteretic region, one requires exceedingly large times to achieve the unique steady-state distribution (underlying the thermodynamic interpretation), and for durations relevant to experiments, the system may instead attain an initial-condition-dependent metastable distribution.

Type
JFM Papers
Copyright
© 2018 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.)

References

d’Avino, G., Hulsen, M. A., Greco, F. & Maffettone, P. L. 2014 Bistability and metabistability scenario in the dynamics of an ellipsoidal particle in a sheared viscoelastic fluid. Phys. Rev. E 89 (4), 043006.Google Scholar
Babcock, H. P., Teixeira, R. E., Hur, J. S., Shaqfeh, E. S. G. & Chu, S. 2003 Visualization of molecular fluctuations near the critical point of the coil–stretch transition in polymer elongation. Macromolecules 36 (12), 45444548.10.1021/ma034073pGoogle Scholar
Bender, C. M. & Orszag, S. A. 1999 Advanced Mathematical Methods for Scientists and Engineers I. Springer.10.1007/978-1-4757-3069-2Google Scholar
Bentley, B. J. & Leal, L. G. 1986 An experimental investigation of drop deformation and breakup in steady, two-dimensional linear flows. J. Fluid Mech. 167, 241283.10.1017/S0022112086002811Google Scholar
Bretherton, F. P. 1962 The motion of rigid particles in a shear flow at low Reynolds number. J. Fluid Mech. 14 (2), 284304.10.1017/S002211206200124XGoogle Scholar
Candelier, F., Einarsson, J., Lundell, F., Mehlig, B. & Angilella, J.-R. 2015a Role of inertia for the rotation of a nearly spherical particle in a general linear flow. Phys. Rev. E 91 (5), 053023.10.1103/PhysRevE.91.053023Google Scholar
Candelier, F., Einarsson, J., Lundell, F., Mehlig, B. & Angilella, J.-R. 2015b Role of inertia for the rotation of a nearly spherical particle in a general linear flow (erratum). Phys. Rev. E 92, 059901.Google Scholar
Candelier, F., Einarsson, J. & Mehlig, B. 2016 Angular dynamics of a small particle in turbulence. Phys. Rev. Lett. 117 (20), 204501.10.1103/PhysRevLett.117.204501Google Scholar
Chandrasekhar, S. 1943 Stochastic problems in physics and astronomy. Rev. Mod. Phys. 15 (1), 1.10.1103/RevModPhys.15.1Google Scholar
Chwang, A. T. 1975 Hydromechanics of low-Reynolds-number flow. Part 3. Motion of a spheroidal particle in quadratic flows. J. Fluid Mech. 72, 1734.10.1017/S0022112075002911Google Scholar
Chwang, A. T. & Wu, T. Y.-T. 1974 Hydromechanics of low-Reynolds-number flow. Part 1. Rotation of axisymmetric prolate bodies. J. Fluid Mech. 63, 607622.10.1017/S0022112074001819Google Scholar
Chwang, A. T. & Wu, T. Y.-T. 1975 Hydromechanics of low-Reynolds-number flow. Part 2. Singularity method for Stokes flows. J. Fluid Mech. 67, 787815.10.1017/S0022112075000614Google Scholar
Dabade, V., Marath, N. K. & Subramanian, G. 2015 Effects of inertia and viscoelasticity on sedimenting anisotropic particles. J. Fluid Mech. 778, 133188.10.1017/jfm.2015.360Google Scholar
Dabade, V., Marath, N. K. & Subramanian, G. 2016 The effect of inertia on the orientation dynamics of anisotropic particles in simple shear flow. J. Fluid Mech. 791, 631703.10.1017/jfm.2016.14Google Scholar
De Gennes, P. G. 1974 Coil–stretch transition of dilute flexible polymers under ultrahigh velocity gradients. J. Chem. Phys. 60 (12), 50305042.10.1063/1.1681018Google Scholar
Einarsson, J., Angilella, J. R. & Mehlig, B. 2014 Orientational dynamics of weakly inertial axisymmetric particles in steady viscous flows. Physica D 278, 7985.10.1016/j.physd.2014.04.002Google Scholar
Einarsson, J., Candelier, F., Lundell, F., Angilella, J. R. & Mehlig, B. 2015 Rotation of a spheroid in a simple shear at small Reynolds number. Phys. Fluids 27 (6), 063301.10.1063/1.4921543Google Scholar
Ennis, G. J., Okagawa, A. & Mason, S. G. 1978 Memory impairment in flowing suspensions. II. Experimental results. Can. J. Chem. 56 (22), 28242832.10.1139/v78-466Google Scholar
Hinch, E. J. 1974 Mechanical models of dilute polymer solutions for strong flows with large polymer deformations. Colloq. Intl CNRS 233, 241247.Google Scholar
Hinch, E. J. & Leal, L. G. 1972 The effect of Brownian motion on the rheological properties of a suspension of non-spherical particles. J. Fluid Mech. 52, 683712.10.1017/S002211207200271XGoogle Scholar
Hoffman, B. D. & Shaqfeh, E. S. G. 2007 The dynamics of the coil–stretch transition for long, flexible polymers in planar mixed flows. J. Rheol. 51 (5), 947969.10.1122/1.2754293Google Scholar
Hudson, S. D., Phelan, F. R. Jr., Handler, M. D., Cabral, J. T., Migler, K. B. & Amis, E. J. 2004 Microfluidic analog of the four-roll mill. Appl. Phys. Lett. 85 (2), 335337.10.1063/1.1767594Google Scholar
Ivanov, Y., Van de Ven, T. G. M. & Mason, S. G. 1982 Damped oscillations in the viscosity of suspensions of rigid rods. I. Monomodal suspensions. J. Rheol. 26 (2), 213230.10.1122/1.549664Google Scholar
Jeffery, G. B. 1922 The motion of ellipsoidal particles immersed in a viscous fluid. Proc. R. Soc. Lond. A 102, 161179.Google Scholar
Karnis, A., Goldsmith, H. L. & Mason, S. G. 1963 Axial migration of particles in Poiseuille flow. Nature 200 (4902), 159160.10.1038/200159a0Google Scholar
Kim, S. & Karrila, S. J. 1991 Microhydrodynamics: Principles and Selected Applications. Butterworth–Heinemann.Google Scholar
Krishnamurthy, D.2014 Heat transfer from drops in shearing flows and collective motion in micro-scale swimmer suspensions. Master’s thesis, Jawaharlal Nehru Centre for Advanced Scientific Research, Bangalore, India.Google Scholar
Krishnamurthy, D. & Subramanian, G. 2018a Heat or mass transport from drops in shearing flows. Part 1. The open streamline. J. Fluid Mech. (submitted).Google Scholar
Krishnamurthy, D. & Subramanian, G. 2018b Heat or mass transport from drops in shearing flows. Part 2. Inertial effects on transport. J. Fluid Mech. (submitted).Google Scholar
Kushch, V. I. 1997 Microstresses and effective elastic moduli of a solid reinforced by periodically distributed spheroidal particles. Intl J. Solids Struct. 34, 13531366.10.1016/S0020-7683(96)00078-9Google Scholar
Kushch, V. I. 1998 Elastic equilibrium of a medium containing a finite number of arbitrarily oriented spheroidal inclusions. Intl J. Solids Struct. 35, 11871198.10.1016/S0020-7683(97)00098-XGoogle Scholar
Leal, L. G. 1979 The motion of small particles in non-Newtonian fluids. J. Non-Newtonian Fluid Mech. 5, 3378.10.1016/0377-0257(79)85004-1Google Scholar
Leal, L. G. 2004 Flow induced coalescence of drops in a viscous fluid. Phys. Fluids 16 (6), 18331851.10.1063/1.1701892Google Scholar
Leal, L. G. & Hinch, E. J. 1971 The effect of weak Brownian rotations on particles in shear flow. J. Fluid Mech. 46, 685703.10.1017/S0022112071000788Google Scholar
Lee, J. S., Dylla-Spears, R., Teclemariam, N. P. & Muller, S. J. 2007a Microfluidic four-roll mill for all flow types. Appl. Phys. Lett. 90 (7), 074103.Google Scholar
Lee, J. S., Shaqfeh, E. S. G. & Muller, S. J. 2007b Dynamics of DNA tumbling in shear to rotational mixed flows: pathways and periods. Phys. Rev. E 75 (4), 040802.10.1103/PhysRevE.75.040802Google Scholar
Liou, K.-N. 1986 Influence of cirrus clouds on weather and climate processes: a global perspective. Mon. Weath. Rev. 114 (6), 11671199.10.1175/1520-0493(1986)114<1167:IOCCOW>2.0.CO;22.0.CO;2>Google Scholar
Marath, N. K., Dwivedi, R. & Subramanian, G. 2017 An orientational order transition in a sheared suspension of anisotropic particles. J. Fluid Mech. 811, R3.10.1017/jfm.2016.779Google Scholar
Marath, N. K. & Subramanian, G. 2017 The effect of inertia on the time period of rotation of an anisotropic particle in simple shear flow. J. Fluid Mech. 830, 165210.10.1017/jfm.2017.534Google Scholar
Marath, N. K. & Subramanian, G.2017 The orientation dynamics of anisotropic particles in shearing flows. PhD thesis, Jawaharlal Nehru Centre for Advanced Scientific Research, Bangalore, India.Google Scholar
Morse, P. M. & Feshbach, H. 1953 Methods of Theoretical Physics. McGraw-Hill.Google Scholar
Okagawa, A., Cox, R. G. & Mason, S. 1973 The kinetics of flowing dispersions. VII. Oscillatory behavior of rods and discs in shear flow. J. Colloid Interface Sci. 45, 303329.10.1016/0021-9797(73)90271-3Google Scholar
Petrich, M. P., Chaouche, M., Koch, D. L. & Cohen, C. 2000 Oscillatory shear alignment of a non-Brownian fiber in a weakly elastic fluid. J. Non-Newtonian Fluid Mech. 91 (1), 114.10.1016/S0377-0257(99)00092-0Google Scholar
Prager, S. 1957 Stress–strain relations in a suspension of dumbbells. Trans. Soc. Rheol. 1 (1), 5362.10.1122/1.548808Google Scholar
Rosen, T., Do-Quang, M., Aidun, C. K. & Lundell, F. 2015a Effect of fluid and particle inertia on the rotation of an oblate spheroidal particle suspended in linear shear flow. Phys. Rev. E 91 (5), 053017.10.1103/PhysRevE.91.053017Google Scholar
Rosen, T., Einarsson, J., Nordmark, A., Aidun, C. K., Lundell, F. & Mehlig, B. 2015b Numerical analysis of the angular motion of a neutrally buoyant spheroid in shear flow at small Reynolds numbers. Phys. Rev. E 92 (6), 063022.10.1103/PhysRevE.92.063022Google Scholar
Rosen, T., Lundell, F. & Aidun, C. K. 2014 Effect of fluid inertia on the dynamics and scaling of neutrally buoyant particles in shear flow. J. Fluid Mech. 738, 563590.10.1017/jfm.2013.599Google Scholar
Saffman, P. G. 1956 On the motion of small spheroidal particles in a viscous liquid. J. Fluid Mech. 1 (5), 540553.10.1017/S0022112056000354Google Scholar
Schroeder, C. M., Babcock, H. P., Shaqfeh, E. S. G. & Chu, S. 2003 Observation of polymer conformation hysteresis in extensional flow. Science 301 (5639), 15151519.10.1126/science.1086070Google Scholar
Schroeder, C. M., Teixeira, R. E., Shaqfeh, E. S. G. & Chu, S. 2005 Characteristic periodic motion of polymers in shear flow. Phys. Rev. Lett. 95 (1), 018301.10.1103/PhysRevLett.95.018301Google Scholar
Shaqfeh, E. S. G. 2005 The dynamics of single-molecule DNA in flow. J. Non-Newtonian Fluid Mech. 130 (1), 128.10.1016/j.jnnfm.2005.05.011Google Scholar
Subramanian, G. & Brady, J. F. 2004 Multiple scales analysis of the Fokker–Planck equation for simple shear flow. Physica A 334 (3), 343384.10.1016/j.physa.2003.10.055Google Scholar
Subramanian, G. & Koch, D. L. 2005 Inertial effects on fibre motion in simple shear flow. J. Fluid Mech. 535, 383414.10.1017/S0022112005004829Google Scholar
Subramanian, G. & Koch, D. L. 2006a Centrifugal forces alter streamline topology and greatly enhance the rate of heat and mass transfer from neutrally buoyant particles to a shear flow. Phys. Rev. Lett. 96 (13), 134503.10.1103/PhysRevLett.96.134503Google Scholar
Subramanian, G. & Koch, D. L. 2006b Inertial effects on the orientation of nearly spherical particles in simple shear flow. J. Fluid Mech. 557, 257296.10.1017/S0022112006009724Google Scholar
Subramanian, G. & Koch, D. L. 2006c Inertial effects on the transfer of heat or mass from neutrally buoyant spheres in a steady linear velocity field. Phys. Fluids 18 (7), 073302.10.1063/1.2215370Google Scholar
Subramanian, G. & Koch, D. L. 2007 Heat transfer from a neutrally buoyant sphere in a second-order fluid. J. Non-Newtonian Fluid Mech. 144 (1), 4957.10.1016/j.jnnfm.2007.02.013Google Scholar
Taylor, G. I. 1923 The motion of ellipsoidal particles in a viscous fluid. Proc. R. Soc. Lond. A 103 (720), 5861.Google Scholar
Taylor, G. I. 1934 The formation of emulsions in definable fields of flow. Proc. R. Soc. Lond. A 146 (858), 501523.Google Scholar