Fluid Dynamics Research 40 (2008) 737 – 752

Parametric modulation in the Taylor–Couette ferrofluid flow Jitender Singh, Renu Bajaj∗ Centre for Advanced Study in Mathematics, Panjab University, Chandigarh 160014, India Received 4 April 2006; received in revised form 18 April 2008; accepted 23 April 2008 Available online 2 June 2008 Communicated by T. Yoshinaga

Abstract A parametric instability of the Taylor–Couette ferrofluid flow excited by a periodically oscillating magnetic field, has been investigated numerically. The Floquet analysis has been employed. It has been found that the modulation of the applied magnetic field affects the stability of the basic flow. The instability response has been found to be synchronous with respect to the frequency of periodically oscillating magnetic field. © 2008 The Japan Society of Fluid Mechanics and Elsevier B.V. All rights reserved. MSC: 76-XX; 76D17; 76E07 Keywords: Ferrofluid; Couette–Taylor instability; Parametric modulation

1. Introduction Parametric instability occurs due to modulation of some parameter in a dynamical system. The parametric instability of the flows driven by external time periodic forcing, has received much attention due to its practical importance. An example from classical hydrodynamics is the Faraday instability under external periodical modulation. Kumar (1996) has discussed in detail the stability of plane free surface of a viscous liquid on a horizontal plate under vertical periodic oscillation, theoretically. Bajaj and Malik (2001) have studied the parametric instability of the interface between two viscous magnetic fluids, excited by a periodically oscillating magnetic field. Using Floquet theory, they have obtained numerically, the instability zones for harmonic and subharmonic response of the instability entrainment. They have ∗ Corresponding author at: Department of Mathematics, Panjab University, Chandigarh 160014, India.

E-mail addresses: [email protected], [email protected] (R. Bajaj). 0169-5983/$32.00 © 2008 The Japan Society of Fluid Mechanics and Elsevier B.V. All rights reserved. doi:10.1016/j.fluiddyn.2008.04.002

738

J. Singh, R. Bajaj / Fluid Dynamics Research 40 (2008) 737 – 752

found that depending upon the physical parameters, the modulation may destabilize the otherwise stable system or the unstable system may get stabilized by modulation. The Taylor–Couette system (1966) consisting of the flow of a liquid in an annular space between two uniformly, rotating coaxial cylinders can also exhibit parametric instability (e.g. Koschmieder, 1993). The instability entrainment of parametric modulation of the angular velocity of the cylinders in the Taylor–Couette system, has been studied extensively by Donnelly (1964), Carmi and Tustaniwskyj (1981), Riley and Lawrence (1976), Youd et al. (2005), etc. Their investigations show that the onset of Taylorvortex flow is stabilized or destabilized, depending upon the nature of modulation. The value of the critical Reynolds number at parametric modulation of the inner cylinder, is not much different from its corresponding value in the unmodulated flow. Riley and Lawrence (1976) have found numerically that at the onset of the Couette–Taylor instability, subharmonic response is also possible, depending upon the modulation. The subharmonic instability is observed numerically by solving Floquet equation. Marques and Lopez (1997) have investigated stability of the Taylor–Couette flow with axial, periodic oscillations of the inner cylinder. They have found that the axial, parametric modulation of the inner cylinder stabilizes the flow with respect to the centrifugal instabilities. When instability sets in via axisymmetric mode, the new state is synchronous with the basic state. The stabilization is due to waves of azimuthal vorticity propagating out from the boundary layer on the inner cylinder. Walsh and Donnelly (1988) have studied experimentally, the Taylor–Couette flow with the outer cylinder oscillating in axial direction and found the flow to be stabilized. The stability of the Taylor–Couette flow in ferrofluids (Rosenswieg, 1985 and Bashtovoy et al., 1988) in the presence of an axial magnetic field, has been studied by Niklas et al. (1989), Singh and Bajaj (2005, 2006), Odenbach and Gilly (1996), Chang et al. (2003), etc. They have found that the applied magnetic field causes a significant elevation in the critical Taylor number Tc for the onset of Couette–Taylor instability in ferrofluids. Thus, the Taylor–Couette flow in ferrofluids is stabilized by the axially applied magnetic field. The stabilization is due to increase in the rotational viscosity (Shliomis, 1972) of the fluid forced by the action of the magnetic field. The Taylor–Couette ferrofluid flow has found some technological applications such as in making ferrofluid rotary seals. Ferrofluids have application in the electric motors by filling the air-gap between stator and rotor to improve efficiency and thus to save energy (Nethe et al., 2006). Numerous experiments have been conducted to understand the internal flow and the magnetization in the Taylor–Couette ferrofluid flow. Embs et al. (2006) have measured experimentally, the transverse magnetization of a ferrofluid rotating as a rigid body in a constant magnetic field applied perpendicular to the axis of rotation. Leschhorn and Lücke (2006) have investigated the dynamics of a ferrofluid torsional pendulum that is forced periodically to undergo small amplitude oscillations in the presence of a transverse magnetic field. They have found that increase in magnetic field causes damping and pendulum oscillates with small amplitude. In polydisperse ferrofluids the amplitude of oscillation decreases at a faster rate. Kikura et al. (1999) have studied experimentally the Taylor-vortex ferrofluid under the action of magnetic field. They have measured the instantaneous axial velocity of the flow, the critical Reynolds number at the onset of instability, and the rotational viscosity of the fluid using ultra sound velocity profile (UVP) technique. The instantaneous fluid velocity is captured via measuring the instantaneous velocity of the tracer particles along the flow line by means of the Doppler shift in the ultrasound pulse emitted by the transducer attached to the Taylor–Couette apparatus and the echo reflected from the surface of microparticles suspended in the fluid. Information on the position from which the ultrasound signal is reflected is extracted from the time delay after the start of the pulse. Several attempts have been made to measure

J. Singh, R. Bajaj / Fluid Dynamics Research 40 (2008) 737 – 752

739

experimentally, the internal flow field of the Taylor-vortex ferrofluid flow (see also Kikura et al., 2005; Ito et al., 2005). All these experiments are concerned with the consideration of transverse magnetic field applied orthogonally to the axis of rotation. An axial magnetic field can also be applied to the Taylor–Couette system to measure the flow field and magnetization inside the ferrofluid at the onset of the Taylor-vortex flow. An axial magnetic field can be generated by stacking a current carrying solenoid around the outer cylinder. Further the Taylor–Couette ferrofluid flow can be modulated parametrically, by applying periodically oscillating axial magnetic field. This can be done using AC current in the solenoid. It thus becomes natural to consider the response of the Couette–Taylor instability to the parametric modulation of a periodically oscillating, axially applied magnetic field. In the present exposition, we have investigated this important stability problem numerically, under weak field limits. A basic modulated solution has been obtained in closed form and its stability has been discussed using the classical Floquet theory (e.g. Jordan and Smith, 1988; Farkas, 1994). This stability problem may give some insight in understanding the changes occurring in the flow field and magnetization inside the ferrofluid at the onset of the Taylor-vortex ferrofluid flow, in presence of an axial time periodic oscillating magnetic field to facilitate the further study. 2. Mathematical formulation Consider an incompressible, viscous, Newtonian ferrofluid flow in between two concentric cylinders of infinite aspect ratio with radii r1 and r2 , (r1 < r2 ) rotating with constant angular speeds 1 and 2 , respectively, about the vertical axis. A time periodic magnetic field h0 (t) = (0, 0, H0 (t)), oscillating about the steady value (0, 0, h 0 ), is applied to the system such that H0 (t) = h 0 (1 +  sin(t)), where h 0  0,  0,  > 0 and t  0. As a result the ferrofluid exhibits a non-zero magnetization m. The flow is governed by the following equations (Shliomis, 1972), ju jt jm jt

1   + u · ∇u = − ∇ p + ∇ 2 u + 0 m · ∇h + 0 ∇ × (m × h),   2

(2.1)

+ u · ∇m = 21 (∇ × u) × m − (m − m∗ (t)) − m × (m × h),

(2.2)

∇ · u = 0,

(2.3)

∇ × h = 0,

(2.4)

∇ · (m + h) = 0,

(2.5)

where, u is the velocity and h is the magnetic field in ferrofluid inside the annulus at any time t; , p and  are the density, pressure and kinematic viscosity of ferrofluid, respectively; 0 is the permeability constant,  = kb Tb /3Vh , = 0 /6 , where kb , Tb and Vh are Boltzmann constant, temperature of ferrofluid and hydrodynamic volume of each ferrocolloid particle, respectively. is the volume fraction of ferromagnetic particles and m∗ (t) = (0, 0, m ∗ (t)), is the instantaneous equilibrium magnetization (see Shliomis and Morozov, 1994) of ferrofluid at B =0, where, B is the Brownian time constant for rotational diffusion of ferromagnetic particles in ferrofluid. m ∗ (t) is given by m ∗ (t) = nm[coth( f (t)) − 1/( f (t))],

(2.6)

740

J. Singh, R. Bajaj / Fluid Dynamics Research 40 (2008) 737 – 752

where f (t)=1+  sin t, n is the volume density of ferromagnetic particles in ferrofluid, m is the magnetic moment of each ferromagnetic particle and = 0

mh 0 , kb Tb

(2.7)

is the magnetic field parameter or Langevin parameter. It is a dimensionless measure of the steady part of the applied magnetic field. The boundary conditions for the velocity field u, the magnetic induction field b, and the magnetic field h are u|r =r1 = (0, r1 1 , 0), u|r =r2 = (0, r2 2 , 0),  n · b|r =r1 ,r2 = 0,  n × h|r =r1 ,r2 = (0, −H0 (t), 0),

(2.8)

where,  n denotes an unit outward drawn normal to the curved surface of the outer cylinder. 2.1. A basic solution In the weak field limits for the applied magnetic field, i.e.  f (t) < 1, the Langevin relation (2.6) approximates to m ∗ (t) ≈ nm f (t)/3, and the instantaneous equilibrium magnetization m∗ (t) of ferrofluid becomes m∗ (t) = h0 (t), where = 0

nm 2 . 3kb Tb

(2.9)

Then the system of equations (2.1)–(2.5) along with the boundary conditions (2.8) admits a basic solution in form, ⎫ u0 = (0, r (r ), 0), ⎪ ⎬ h0 (t) = (0, 0, H0 (t)), (2.10) m0 (t) = (0, 0, M0 (t)), ⎪ ⎭ p0 =  r 2 dr, where  = A + Br −2 , ∗ = 2 /1 ,

and

A=

1 (∗ − 2 )

1 > 0,

 M0 (t) = h 0 1 +

1 − 2

,

B=

r12 1 (1 − ∗ ) , 1 − 2

 = r1 /r2

 2 + 2







( sin t −  cos t) + M0 (0) − h 0 1 −

 2 + 2

exp(−t).

We are interested in time periodic, basic modulated solutions. For this, we take that the time t = 0 corresponds to

 M0 (0) = h 0 1 − 2 , (2.11)  + 2

J. Singh, R. Bajaj / Fluid Dynamics Research 40 (2008) 737 – 752

so that,

 M0 (t) = h 0 1 +

 2 + 2

741

( sin t −  cos t) .

(2.12)

The velocity field of the basic flow (2.10) is the circular Couette flow in which each ferrofluid particle in the annulus describes circular path about the vertical axis of the rotating cylinders. 2.2. Perturbation equations The stability of the basic solution (2.10) has been considered by imposing small axisymmetric disturbances on it in the following normal modes, ⎫ u = (R/)u0 + (u r (r, t), u  (r, t), u z (r, t)) exp(ikz), ⎪ ⎬ h = Hh0 (t) + (h r (r, t), h  (r, t), h z (r, t)) exp(ikz), (2.13) m = Hm0 (t) + (m r (r, t), m  (r, t), m z (r, t)) exp(ikz), ⎪ ⎭ 2 2 p = (R / ) p0 + p1 (r, t) exp(ikz), where the quantities u, h, m and p are dimensionless. Here we have used R = r2 − r1 as the distance scale and R 2 / as the time scale for non-dimensionalization purpose, H = R(0 /2)1/2 /,

t ∈ [0, ∞),

r ∈ [r1∗ , r2∗ ], r1∗ = r1 /R, r2∗ = r2 /R, k ∈ R and z ∈ (−∞, ∞). The perturbations satisfy the following linearized system of equations,  ⎫ j 2 k 2 R 2 ∗ 2 ∗ 2 ⎪ (DD −k ) DD −k − u  +iH H0 (t)k 3 m r +ik H(M0 (t)+H0 (t)) ⎪ ur = ⎪ ⎪ jt  ⎪ ⎪ ⎪ ∗ 2 ⎪ DD − k )h , ×( r ⎪  ⎪ ⎪ j ⎪ ∗ 2 ∗ 2 ⎪ ⎪ DD − k − u  = D (r  R /)u r − ik H H0 (t)m  , ⎪ ⎪ jt ⎪ ⎪  2 ⎬ 2 j H M0 (t) R r D R ∗ 2 mr + (DD − k )u r ( + M0 (t)H0 (t)) + m = ⎪  jt 2 2ik ⎪ ⎪ ⎪ ⎪ R 2 M0 (t)2 ⎪ ⎪ + hr , ⎪ ⎪  ⎪  2 ⎪ 2 ⎪ ⎪ j H M0 (t)ik R r D R ⎪ ⎪ ( + M0 (t)H0 (t)) + mr = m − u, ⎪ ⎪  jt 2 2 ⎪ ⎭ ∗ 2 2 (DD − k )(m r + h r ) = −k m r ,

(2.14)

where, D≡

j jr

, D∗ ≡

j

1 + . jr r

The boundary conditions (2.8) reduce to, u r = D∗ u r = u  = m r + h r = D∗ (m r + h r ) = 0 at r = r1∗ , r2∗ .

(2.15)

742

J. Singh, R. Bajaj / Fluid Dynamics Research 40 (2008) 737 – 752

The stability characteristics have been expressed in terms of the ratio of the inertial forces to the viscous forces in the Taylor–Couette flow, called the Taylor number T defined by T = −4 A1 R 4 /2 ,

(2.16)

which represents a dimensionless measure of the angular speed of the inner cylinder.

3. Solution We expand the perturbation variables, u r , u  , and m r + h r in terms of orthogonal functions as, u r (r, t) =



A j (t)u(a j r ),

(3.1a)

B j (t)v(b j r ),

(3.1b)

j=1

u  (r, t) =

∞ j=1

m r (r, t) + h r (r, t) =



C j (t)u(a j r ),

(3.1c)

j=1

where the complex functions A j (t), B j (t) and C j (t) of the real variable t, are to be determined. Also u(a j r ) = J1 (a j r ) + d j Y1 (a j r ) + e j I1 (a j r ) + f j K 1 (a j r )

(3.2)

v(b j r ) = J1 (b j r ) + g j Y1 (b j r ).

(3.3)

and

Here, J1 , Y1 , I1 and K 1 are the Bessel functions of order one, a j and b j are the solutions of certain transcendental equations, which have been given in Appendix. The coefficients, d j , e j , f j and g j have been evaluated using the boundary conditions satisfied by the functions u(a j r ) and v(b j r ), which are the solutions of following characteristic value problems, respectively (Chandrasekhar, 1966), (D D ∗ )2 F = a 4 F,

F = D F = 0 at r = r1∗ , r2∗ ,

D D ∗ G = −b2 G, G = 0 at r = r1∗ , r2∗ ,

(3.4a) (3.4a)

where, D≡

1 d d and D ∗ ≡ + . dr dr r

Eliminating m  from the system of equations (2.14), substituting equations (3.1) in it, multiplying the resulting system of equations throughout by r u(aq r ) for each q = 1, 2, 3, . . . , and integrating it under the

J. Singh, R. Bajaj / Fluid Dynamics Research 40 (2008) 737 – 752

743

limits of r , we obtain the following system of first order ordinary differential equations, ⎫ ∞ ∞

(11) d A j (11) (12) (13) ⎪ q j {Q q j A j + Q q j B j + Q q j C j }, = ⎪ ⎪ ⎪ dt j=1 j=1 ⎪ ⎪ ⎪ ∞ ∞ ⎪

(22) dB j

(21) ⎪ (22) (23) (24) ⎪ q j {Q q j A j + Q q j B j + Q q j C j + Q q j D j }, = ⎪ ⎬ dt j=1 ∞

j=1



dC j ⎪ ⎪ q j q j D j , = ⎪ ⎪ ⎪ dt j=1 j=1 ⎪ ⎪   ⎪ ∞ ∞ ⎪

(41) d A j

(41) ⎪ (44) dD j (42) (43) (44) ⎭ q j {Q q j A j + Q q j B j + Q q j C j + Q q j D j }. ⎪ + q j = dt dt j=1 j=1 (1)

(2)

(4)

(11)

(22)

(41)

(3.5)

(44)

The coefficients Q q j for 1    3, Q q j and Q q j for 1    4, q j , q j , q j , q j and q j are given in Appendix. To solve the system (3.5) numerically, we have truncated each series after a suitable number of N terms where N is a positive integer. System (3.5) after truncation, can be represented in the following matrix form, B(t)

dX(t) = A(t)X(t), 0  t < ∞, dt

(3.6)

where X = ( A1 , A2 , . . . , A N , B1 , B2 , . . . , B N , C1 , C2 , . . . , C N , D1 , D2 , . . . , D N ) is the column matrix of order 4N × 1; the symbol denotes matrix transpose. B(t) and A(t) are non-singular square matrices of order 4N . All the entries in these matrices are time periodic with period 2/0 and are functions of the dimensionless parameters, ∗ , , , 0 , , k and T , where 0 =  R 2 / is the dimensionless frequency of modulation. 3.1. Floquet analysis As (3.6) is a system of ordinary differential equations with periodic coefficients, Floquet analysis (Jordan and Smith, 1988; Farkas, 1994) has been applied to solve it. A fundamental matrix U(t) for the system (3.6) satisfies B(t)

dU(t) = A(t)U(t), 0  t < ∞. dt

(3.7)

Therefore, dU(t) = B(t)−1 A(t)U(t), U(0) = U0 . dt

(3.8)

We take U(0)=I4N , the identity matrix of order 4N . The system (3.8) has been solved numerically, using a numerical method given in Nelson (1969) and Farkas (1994), with N =8. The interval [0, 2/0 ] is divided into s equal parts by t0 = 0 < t1 < t2 < · · · < ts = 2/0 , i.e. t = t j − t j−1 = 2/(0 s), j = 1, 2, . . . , s. Let F(t) = B(t)−1 A(t), then F(t j−1 + t) ≈ F(t j−1 ), for all t ∈ [t j−1 , t j ), for t sufficiently small so that an approximation to the solution of (3.8) at t j is, U(t j ) = U(t j−1 ) exp(tF(t j−1 )).

(3.9)

744

J. Singh, R. Bajaj / Fluid Dynamics Research 40 (2008) 737 – 752

Using this iteration process, the approximate solution at t = 2/0 is given by U(2/0 ) = U(0)

s 

exp(tF(t j−1 )).

(3.10)

j=1

To discuss the stability of the basic flow (2.10), we have calculated the eigenvalues  j ’s of U(2/0 ) numerically. The Floquet exponents  j and the Floquet multipliers  j are related by  j = exp(2 j /0 ),

1  j  4N .

(3.11)

 j ’s are the functions of the dimensionless parameters: the modulation frequency 0 , the modulation amplitude , the magnetic field parameter , the angular velocity ratio ∗ , the radius ratio parameter ,

the Taylor number T and the axial wave number k. The marginal state of the basic modulated system (2.10) is determined by setting max (real( j )) = 0,

1  j  4N

(3.12)

The basic modulated Taylor–Couette flow is stable for max1  j  4N (real( j )) < 0, and unstable for max1  j  4N (real( j )) > 0. In the marginal state, fixing the values of the parameters ∗ , , , 0 , and , a trial value of k is taken and the Taylor number T is varied until Eq. (3.12) is approximately satisfied. In this way a root (k, T ) of Eq. (3.12) is obtained numerically. The above procedure is repeated for different values of k until a root of (3.12) with a minimum value of T is obtained. The minimum value of T is called the critical Taylor number denoted by Tc , and the corresponding value of the axial wave number k is called the critical axial wave number denoted by kc . The flow is stable for T < Tc and unstable for T > Tc . The instability sets in first at the critical Taylor number Tc . If a Floquet exponent satisfying (3.12) is identically zero, then the disturbance in the marginal state oscillates periodically with the forcing frequency 0 , and the instability response is called synchronous or harmonic. On the other hand, if the imaginary part of the Floquet exponent satisfying (3.12), is equal to 0 /s, for some positive integer s > 1, the disturbance in the marginal state oscillates with a frequency 0 /s, and the instability response is called subharmonic of order 1/s. Not all of these actually occur (Jordan and Smith, 1988).

4. Results and discussion We have considered the modulation of applied magnetic field in weak field limits. So to obtain the results, the numerical values of magnetic field parameter , i.e. the dimensionless measure of the steady part of the applied magnetic field, and the amplitude of modulation , have been varied such that 1+  sin(0 t) < 1. Thus, at =0.1, 0.2, 0.3 and 0.5, the permissible values of  are: 0  < 9, 0  < 4, 0  < 2.6 and 0  < 1, respectively. The numerical results have been obtained for a diester based ferrofluid of magnetite. The saturation magnetization of magnetite is Ms f = 480 × 103 amp m−1 . The average magnetic moment of single magnetite particle is m = 2.247 × 10−19 amp m2 . For the ferrofluid under consideration, we have taken, Tb = 298 K, = 0.2,  = 1.614 × 103 kg m−3 ,  = 0.09235 Ns m−2 , −1 = 9.51942 × 10−5 s (particle size 13.9 nm), and Vh = 1.413 × 10−24 m3 . These values have been taken from the reference (Bashtovoy et al., 1988). For this range of parameters, the Brownian relaxation

J. Singh, R. Bajaj / Fluid Dynamics Research 40 (2008) 737 – 752

2450

745

3800

2400

3750

 = 0.5

 = 0.5

 = 0.2

3700

 = 0.2

 = 0.1 Tc

Tc

2350

 = 0.1

3650

2300 3600 2250

3550

2200 0

2

4

6

8

3500

10

0

2

4

8

6

6450

8000

6400

 = 0.5

7800

 = 0.5

6350

Tc

 = 0.1

Tc

 = 0.2 7600

 = 0.2  = 0.1

6300

7400 6250 7200

6200 6150

7000 0

2

4

6

8

0

2

4

6

8

Fig. 1. Variation of the critical Taylor number Tc with the amplitude of modulation  at: (a) The radius ratio  = 0.95, the angular speed ratio ∗ = 0.5, the gap width R = 1 mm, the volume fraction = 0.2 and the frequency 0 = 1. (b) The radius ratio  = 0.95, the angular speed ratio ∗ = 0, the gap width R = 1 mm, the volume fraction = 0.2 and the frequency 0 = 1. (c) The radius ratio  = 0.95, the angular speed ratio ∗ = −0.5, the gap width R = 1 mm, the volume fraction = 0.2 and the frequency 0 = 1. (d) The radius ratio  = 0.50, the angular speed ratio ∗ = 0, the gap width R = 1 mm, the volume fraction

= 0.2 and the frequency 0 = 1.

mechanism dominates (Rosenswieg, 1985; Bashtovoy et al., 1988). All numerical calculations except for Fig. 1(d) and Fig. 2(a), have been performed at the radius ratio  = 0.95 and the gap width R = 1 mm between the cylinders. Fig. 1(a) shows the variation of the critical Taylor number with the amplitude of modulation , for various values of the magnetic field parameter , and the angular velocity ratio of the cylinders ∗ = 0.5. At a given strength of the applied magnetic field, Tc rises with increase of . Thus, the increase of the amplitude of modulation has a stabilizing effect on the basic modulated flow. A similar variation of Tc with  has been observed when the outer cylinder is held fixed and the inner cylinder is allowed to

746

J. Singh, R. Bajaj / Fluid Dynamics Research 40 (2008) 737 – 752

7

8000

6

= 8.99,  = 0.1 7800 7600

4

= 2.5,  = 0.3

Tc

Tc × 10−4

5

3

7400

2 7200

1 0 0.2

0.4

0.6 

0.8

0.99

7000 50

100 0

150

200

Fig. 2. Variation of the critical Taylor number Tc with: (a) The radius ratio parameter  at the magnetic field parameter = 0.3, the angular speed ratio ∗ = 0, the gap width R = 10 mm, the volume fraction = 0.2, the amplitude of modulation  = 0.2 and the frequency 0 = 5. (b) The frequency of oscillation 0 at the radius ratio  = 0.95, angular speed ratio ∗ = −0.5, the gap width R = 1 mm and the volume fraction = 0.2.

rotate (Figs. 1(b)), and, when the cylinders are counter-rotating (Figs 1(c)). For ordinary fluids, when the Taylor–Couette flow is modulated by oscillating the azimuthal velocity of the inner cylinder, at a fixed frequency, the critical Taylor number decreases with increase in the amplitude of modulation, and thus the flow is destabilized (Riley and Lawrence, 1976). Marques and Lopez (1997) have investigated numerically, the stability of the basic modulated fluid flow under the axial time periodic oscillation of the inner cylinder and keeping the outer cylinder fixed. They have found that the modulation stabilizes the flow. Fig. 1(d) shows the variation of the critical Taylor number Tc with the amplitude of modulation  at the radius ratio  = 0.50, the angular speed ratio ∗ = 0, the gap width R = 1 mm, the volume fraction

= 0.2 and the frequency 0 = 1. The critical Taylor number rises with the amplitude of modulation and the onset of instability is delayed if compared to the variation of Tc with  at higher radius ratio (compare Fig. 1(d) with Fig. 1(b)). At a fixed gap width between the cylinders, the instability in the basic ferrofluid flow depends upon the radius ratio of the cylinders irrespective of their radii. Also, in the weak field limits, the critical Taylor number does not vary much with change in the gap width R (Singh and Bajaj, 2005). We have observed numerically, that at a lower radius ratio of the cylinders, the instability sets in first at a higher Taylor number than the corresponding Taylor number obtained at higher radius ratio. As the radius ratio approaches 1 for the fixed value of gap width R, the radii of the cylinders reach infinity. We have calculated the critical Taylor number upto  = 0.99. Beyond this value due to singularity, the numerical solution cannot be obtained. Fig. 2(a) shows the variation of the critical Taylor number with the radius ratio of the cylinders at the onset of the instability for the fixed values = 0.2, ∗ = 0, R = 10 mm, = 0.2,  = 0.3 and 0 = 5, for 0.1   0.99.

J. Singh, R. Bajaj / Fluid Dynamics Research 40 (2008) 737 – 752

747

T

10000

9000 0 = 1 8000 0 = 200 7000 2

2.5

3 k

3.5

4

Fig. 3. A stability diagram in the (k, T ) plane drawn at the radius ratio  = 0.95, the angular speed ratio ∗ = −0.5, the gap width R = 1 mm, the volume fraction = 0.2, the Langevin parameter = 0.1 and the amplitude of modulation  = 8.99.

The variation of the critical Taylor number with the frequency 0 of modulation has been drawn in Fig. 2(b). At a given amplitude of modulation and the strength of the applied magnetic field, the increase of the modulation frequency has a destabilizing effect on the basic modulated flow. For ordinary fluids, Riley and Lawrence (1976) have found numerically that in the modulation of the azimuthal velocity of the inner cylinder and for small amplitude ratio, the critical Taylor number increases monotonically with the frequency, approaching the steady value asymptotically. They have also observed that the instability response can be harmonic or subharmonic, depending upon the amplitude and frequency of modulation. However, when the flow is modulated by the axial oscillation of the inner cylinder, at the onset of instability against axisymmetric disturbances, the frequency response has been found to be harmonic in all cases (Marques and Lopez, 1997). In the present study also, we have observed numerically that with the parametric modulation of the axially applied magnetic field, the instability response is harmonic and nowhere subharmonic for the considered parametric values. Fig. 3 shows a stability diagram for the stability of the basic modulated flow in the (k, T ) plane. Considering the curve corresponding to 0 = 1, the basic modulated flow has been observed to be stable in the parametric region below this curve and unstable in the parametric region above it. For the curve corresponding to 0 = 200, the basic modulated flow is stable below it and unstable above it. Thus, the region between these curves in the (k, T ) plane which was stable at 0 = 1, becomes unstable when the modulation frequency 0 is increased to 200. Fig. 4 shows a stability diagram for the stability of a supercritical modulated flow in the (k, ) plane for different values of the modulation frequency 0 . In the absence of modulation, the basic flow is stable for k < 2.28 and k > 4.178, and it is unstable in the interval 2.28 < k < 4.178. With the parametric modulation of the basic modulated flow, there is a curve in the (k, ) plane that separates the region of stability from the region of instability. With an increase of the modulation frequency the corresponding curves in (k, ) plane lie above the one drawn at a lower value of 0 . Thus, with an increase of 0 the region of instability in the (k, T ) plane for a supercritical modulated flow, widens.

748

J. Singh, R. Bajaj / Fluid Dynamics Research 40 (2008) 737 – 752 9

(c) (b)

8 (a) 7 6 5 4 3 2 1 0

2.2

2.5

3

3.5

4

k

Fig. 4. A stability diagram for a supercritical basic modulated flow in the (k, ) plane drawn for: (a) 0 = 1, b 0 = 100 and c  = 121, at radius ratio  = 0.95, the angular speed ratio ∗ = 0, the gap width R = 1 mm, the volume fraction = 0.2, the Langevin parameter = 0.1 and the Taylor number T = 4000.

0 (d) -0.2 (c) (b)

ur

-0.4

-0.6 (a)

-0.8

-1 0

10

20

30

t

Fig. 5. Time evolution of the normalized radial component u r of the perturbations in the velocity field of the fluid at the radius ratio  = 0.95, the angular speed ratio ∗ = −0.5, the gap width R = 1 mm, the volume fraction = 0.2, the Langevin parameter = 0.1 and the frequency 0 = 0.5. The curves (a)–(d) correspond to  = 1, 3, 6 and 8.99, respectively.

We have observed numerically, that as 0 is increased to 200, the flow is stable for k < 2.28 and k > 4.178, and unstable for 2.28 < k < 4.178 no matter how much large  may be in its considered range. This stability behavior of the flow is thus similar to that if there is no modulation. The time evolution of the normalized perturbations u r , br and h z , drawn at r = (r1∗ + r2∗ )/2, that corresponds to the middle of the gap between the cylinders, at axial position z = 0, 0  t  6/0 , at the onset of instability in the basic modulated flow has been illustrated in Figs. 5, 6(a) and (b), respectively.

J. Singh, R. Bajaj / Fluid Dynamics Research 40 (2008) 737 – 752

1

1

0.5

0.5

749

(d) (a)

(c)

hz

br

(a) 0

0

(d) (b) (c)

-0.5

(b)

-0.5

-1 0

10

20 t

30

0

10

20 t

30

Fig. 6. Time evolution of the profiles at the radius ratio  = 0.95, the gap width R = 1 mm, the angular speed ratio ∗ = −0.5, the Langevin parameter = 0.1, the volume fraction = 0.2 and the frequency 0 = 0.5. The curves (a)–(d) correspond to  = 1, 3, 6 and 8.99, respectively: (a) The normalized radial component br of the perturbations in the magnetic induction field. (b) The normalized axial component h z of the perturbations in the magnetic field.

The instability response has been found to be harmonic. It is evident from Fig. 5 that with increase of the amplitude of the modulation at the onset of the instability, the fluctuations in the normalized radial disturbance of the velocity field, increase. Similar variation of the normalized components u  , and u z , has been observed with t. The origin of Taylor-vortices is the first stage on the route to the passage of circular into turbulent motion as the Taylor number increases beyond its critical value. Development of turbulence in the Taylor-vortex flow starts for the Taylor number much higher than that of the formation of Taylor-vortices. In the present study, the critical Taylor numbers obtained in the weak field limit are less than the supercritical Taylor numbers at which the onset of nonlinear instabilities such as wavy vortex flow, chaotic Taylor-vortex flow, and the turbulent Taylor-vortex flow starts. The present analysis is not sufficient to capture the nature and onset of these instabilities as they essentially require nonlinear analysis.

5. Conclusion We have investigated numerically, the stability of a modulated Taylor–Couette ferrofluid flow between the two uniformly rotating cylinders via parametric excitation of a periodically oscillating, axially applied magnetic field, in the weak field limits, against axisymmetric disturbances. The stability of a basic modulated flow has been discussed using the Floquet theory. The increase in the frequency of modulation of the applied magnetic field has a destabilizing effect on the basic flow, however, an increase in the amplitude of modulation of magnetic field, has a stabilizing effect on the basic flow. This effect increases by increasing the strength of the steady part of the applied magnetic field. The time periodically oscillating magnetic field induces the time periodic oscillations in the ferrofluid velocity at the onset of the instability. The instability response has been found to be harmonic. Under

750

J. Singh, R. Bajaj / Fluid Dynamics Research 40 (2008) 737 – 752

the weak field limits, the instability response has been found to be nowhere subharmonic. The onset of instability in the basic flow also depends upon the radius ratio of the cylinders. The results of the present study are qualitatively similar to those related to the axial periodic oscillation of the inner cylinder, studied by Marques and Lopez (1997). The degree of stabilization in the basic flow by the modulation of applied magnetic field, under the weak field limits, is comparable to the degree of stabilization in the modulated Taylor–Couette flow by axial wall oscillations of the inner cylinder. Though the magnetic fluid and ordinary fluid have their own characteristics, we find an analogy between the magnetic field oscillations in the Taylor–Couette ferrofluid flow and the axial wall oscillations in the inner cylinder, in the Taylor–Couette flow of an ordinary fluid. Both have stabilizing action on the respective basic flow to almost same extent. From the experimental point of view, however, the Taylor–Couette ferrofluid flow is free from geometrical disturbances as caused by the wall oscillations, and more strict comparisons with the theoretical results are expected with respect to the onset of nonlinear instabilities such as wavy vortex flow, chaotic Taylor-vortex flow, and finally, the turbulent Taylor-vortex flow. Acknowledgment Authors are thankful to the referees for their valuable suggestions, constructive criticism, and helpful comments, which helped in improvement of the paper. Appendix The coefficients a j and b j which have been used in system (3.1a), are the solutions of following transcendental equations, respectively    J1 a j r1∗ ) Y1 (a j r1∗ ) I1 (a j r1∗ ) K 1 (a j r1∗ )       J1 (b j r ∗ ) Y1 (b j r ∗ )   J1 (a j r2∗ ) Y1 (a j r2∗ ) I1 (a j r2∗ ) K 1 (a j r2∗ )  1 1   = 0.   = 0,  J1 (b j r2∗ ) Y1 (b j r2∗ )   J0 (a j r1∗ ) Y0 (a j r1∗ ) I0 (a j r1∗ ) −K 0 (a j r1∗ )    J0 (a j r2∗ ) Y0 (a j r2∗ ) I0 (a j r2∗ ) −K 0 (a j r2∗ ) The following coefficients have been used in the system of (3.5), equations (3.5),  r∗ 2 (11) q j = r u(aq r )(DD∗ − k 2 )u(a j r ) dr , r1∗

(22) q j

R2 = 2



r2∗

r1∗

r u(aq r )v(b j r )r D dr ,

q j is the Kronecker delta function, (41) q j

=−

(11)

Qq j =

H M0 ik

2



r2∗

r1∗



r2∗

r1∗

r u(aq r )(DD∗ − k 2 )u(a j r ) dr, q j = q j ,

r u(aq r )(DD∗ − k 2 )2 u(a j r ) dr ,

(42)

(11)

J. Singh, R. Bajaj / Fluid Dynamics Research 40 (2008) 737 – 752 (12)

Qq j = −

2k 2 R 2 



r2∗

r1∗

751

r u(aq r )v(b j r ) dr ,

iH(M0 + H0 ) (11) Qq j , k  ∗ H2 M0 H0 (11) A R 4 r2 2 (21) q j − 2 r u(aq r )u(a j r )D dr , Qq j = 2  r1∗ (13)

(11)

Q q j = −ik H M0 q j −

R2 2

(22)

Qq j =

=

r1∗

r u(aq r )(DD∗ − k 2 )v(b j r ) dr ,





Q q0 j

r2∗

iH k R 2 H0 M02

(23)

Qq j = where



r2∗

r1∗

Q q0 j +

iH H0 R 2 (11) ( + M0 H0 + M02 )q j , k

r u(aq r )u(a j r ) dr ,

i Hk R 2 dM0 (11) 2 =  M0 + M0 H0 + q j , 2 dt

M0 k 2 R 4 iHk 3 M0 (22) dM0 (42) (43) M0 ( + M0 H0 ) + 2 q j , Q q j = − Qq j = Q q0 j 2 2 dt

d(M0 H0 ) dM0 (11) R4 2 2 + 2 M0 − 2 ( + M0 H0 ) + M0 ( + M0 H0 ) + q j  dt dt  ∗ R 4 r2 − 2 r u(aq r )(r D)2 (DD∗ − k 2 )u(a j r ) dr , ∗ 4 r 1

(24) Qq j

iH H0 (11) = q j , k

(44)

M02 k 2 R 2

Qq j = −



(41) Qq j

Q q0 j −

R2 

(11)

(2 + 2 M0 H0 + M02 )q j .

References Bajaj, R., Malik, S.K., 2001. Parametric instability of the interface between two viscous magnetic fluids. J. Magn. Magn. Matter 253, 35–44. Bashtovoy, V.G., Berkowsky, B.M., Vislovich, A.N., 1988. Introduction to Thermomechanics of Magnetic Fluids. Springer, Berlin. Carmi, S., Tustaniwskyj, J.I., 1981. Stability of modulated finite-gap cylindrical Couette flow: linear theory. J. Fluid Mech. 108, 19–42. Chandrasekhar, S., 1966. Hydrodynamic and Hydromagnetic Stability. Oxford University Press, Oxford. Chang, M.H., Chen, C.K., Weng, H.C., 2003. Stability of ferrofluid flow between concentric rotating cylinders with an axial magnetic field. Int. J. Eng. Sci. 41, 103–121. Donnelly, R.J., 1964. Experiments on the stability of viscous flow between rotating cylinders III. Enhancement of hydrodynamic stability by modulation. Proc. R. Soc. London A 281, 130–139.

752

J. Singh, R. Bajaj / Fluid Dynamics Research 40 (2008) 737 – 752

Embs, J.P., May, S., Wagner, C., Kityk, A.V., Leschhorn, A., Lücke, M., 2006. Measuring the transverse magnetization of rotating ferrofluids. Phys. Rev. E 73, 036302–036308. Farkas, M., 1994. Periodic Motions. Springer, New York. Ito, D., Kikura, H., Aritomi, M., Takeda, Y., 2005. Mode bifurcation control of a magnetic fluid on Taylor–Couette vortex flow with small aspect ratio. J. Phys. Conf. Ser. 14, 35–41. Jordan, D.W., Smith, P., 1988. Nonlinear Ordinary Differential Equations. Clarendon Press, Oxford. Kikura, H., Takeda, Y., Durst, F., 1999. Velocity profile measurement of the Taylor vortex flow of a magnetic fluid using the ultrasonic Doppler method. Exp. Fluids 26, 208–214. Kikura, H., Aritomi, M., Takeda, Y., 2005. Velocity measurement on Taylor–Couette flow of a magnetic fluid with small aspect ratio. J. Magn. Magn. Matter 289, 342–345. Koschmieder, E.L., 1993. Bènard Cells and Taylor Vortices. Cambridge University Press, Cambridge. Kumar, K., 1996. Linear theory of Faraday instability in viscous liquids. Proc. R. Soc. London A 452, 1113–1126. Leschhorn, A., Lücke, M., 2006. Periodically forced ferrofluid pendulum: effect of polydispersity. Z. Phys. Chem. 220, 89–96. Marques, F., Lopez, J.M., 1997. Taylor–Couette flow with axial oscillations of the inner cylinder: Floquet analysis of the basic flow. J. Fluid Mech. 348, 153–175. Nelson, E., 1969. Topics in Dynamics I: Flows. Princeton University Press and The University of Tokyo Press, Princeton, NJ. Nethe, A., Scholz, T., Stahlmann, H.D., 2006. Improving the efficiency of electric machines using ferrofluids. J. Phys.: Condens. Matter 18, S2985–S2998. Niklas, M., Krumbhaar, H.M., Lücke, M.H., 1989. Taylor vortex flow of ferrofluids in the presence of general magnetic fields. J. Magn. Magn. Matter 81, 29–38. Odenbach, S., Gilly, H., 1996. Taylor vortex flow of magnetic fluids under the influence of an azimuthal magnetic field. J. Magn. Magn. Matter 152, 123–128. Riley, P.J., Lawrence, R.L., 1976. Linear Stability of modulated circular Couette flow. J. Fluid Mech. 75, 625–646. Rosenswieg, R.E., 1985. Ferrohydrodynamics. Cambridge University Press, Cambridge. Shliomis, M.I., 1972. Effective viscosity of magnetic suspensions. Sov. Phys. JETP 34, 1291–1294. Shliomis, M.I., Morozov, K.I., 1994. Negative viscosity of ferrofluid under alternating magnetic field. Phys. Fluids 6, 2855–2861. Singh, J., Bajaj, R., 2005. Couette flow in ferrofluids with magnetic field. J. Magn. Magn. Matter 294, 53–62. Singh, J., Bajaj, R., 2006. Stability of ferrofluid flow in rotating porous cylinders with radial flow. Magnetohydrodynamics 42, 46–56. Walsh, T.J., Donnelly, R.J., 1988. Taylor–Couette flow with periodically corotated and counterrotated cylinders. Phys. Rev. Lett. 60 (8), 700–703. Youd, A.J., Willis, A.P., Barenghi, C.F., 2005. Non-reversing modulated Taylor–Couette flows. Fluid Dyn. Res. 36, 61–73.

Parametric modulation in the Taylor–Couette ferrofluid ...

Jun 2, 2008 - A basic modulated solution has been obtained in closed form and its ... u0 = (0,r (r),0), h0(t) = (0,0, H0(t)), m0(t) = (0,0, M0(t)), p0 = ∫ r 2 dr,. ⎫.

393KB Sizes 2 Downloads 54 Views

Recommend Documents

Temperature modulation in ferrofluid convection
tude modulation prevail, i.e., when →0, a general solution .... describe the problem for the simplest case s=0 when N=1 and L=1 to obtain some analytic results.

STABILITY OF FERROFLUID FLOW IN ROTATING ...
STABILITY OF FERROFLUID FLOW IN ROTATING. POROUS CYLINDERS WITH RADIAL FLOW. J. Singh, R. Bajaj. Centre for Advanced Study in Mathematics, ...

Decision Threshold Modulation in the Human Brain
Oct 27, 2010 - all had normal or corrected-to-normal visual acuity. None of them was ... After 5 s of target display, the back- ground turned black .... Thus, our hypothesis was that both behavioral and fMRI data would be better explained by an ...

Seasonal modulation of seismicity in the Himalaya ... - Semantic Scholar
Apr 26, 2007 - 3Geology and Planetary Sciences Department, California Institute of. Technology, Pasadena, California, USA. 4Department of Mines and ...

Mechanisms underlying the noradrenergic modulation ...
Design, Cambridge, UK) and DATAVIEW Analysis Software (W.J.. Heitler, University of St Andrews, UK). N indicates the number of. Fig. 2. Caudal motor neurons ...

Comparison of Parametric and Non-Parametric ...
Systems (ITS) applications, in particular for safety and energy management ... Technology, China. R. Bajcsy is with the ... Section III introduces the models implemented ... 1As an alternative to iterative prediction, the GMR and ANN approaches.

Sensory Modulation and Affective Disorders in Children ...
Descriptive statistics and the Pearson product-moment coefficient of correlation calculations were used for data analysis. RESULTS. The results indicated that ...

amplitude modulation pdf
There was a problem previewing this document. Retrying... Download. Connect more apps... Try one of the apps below to open or edit this item. amplitude ...

temperature modulation in rayleigh–bénard convection
periodic, sinusoidally varying free-boundary temperature, have been investigated ... They found that the critical Rayleigh number increases with modulation.

density of electron states in modulation-doped ...
(IDEG), has been intensively studied both experimentally and theoretically The ... distributed in the wire section, which is clearly irrelevant to realistic QWR's.

Modulation of Medium-Chain Fatty Acid Synthesis in Synechococcus ...
May 26, 2016 - These efforts can be combined with advances in next-generation ... MCFA, leads to MCFA accumulation in transgenic hosts (Jones et al., 1995; Leonard et .... into Blast2Go software v 3.0.10 for functional annotation and statistical ....

pdf amplitude modulation
Sign in. Page. 1. /. 1. Loading… Page 1 of 1. File: Pdf amplitude modulation. Download now. Click here if your download doesn't start automatically. Page 1 of 1.

Bandlimited Intensity Modulation - IEEE Xplore
Abstract—In this paper, the design and analysis of a new bandwidth-efficient signaling method over the bandlimited intensity-modulated direct-detection (IM/DD) ...

Non-Parametric Econometrics
A multiple regression model can be defined as: y = m(x1 ... We consider the partial linear model: ... standard parametric models (spatial autocorrelation models).

Effect of impurity correlation in modulation-doped ...
the electron mobility1–5 and the density of states DOS.6–9. It should be mentioned that all existing theories1–9 of the disorder effect from impurity doping in ...

Oriented Modulation for Watermarking in Direct Binary ... - IEEE Xplore
watermark embedding, while maintaining high image quality. This technique is ... extracted features and ultimately to decode the watermark data. Experimental ...

5-HT Modulation of Multiple Inward Rectifiers in ...
used for data acquisition and off-line current measurements. Current ... were made at room temperature. Correction of the voltage ...... C: box-and-whisker plots.

Simultaneous Inference in General Parametric Models
Simultaneous inference is a common problem in many areas of application. If multiple null hypotheses are tested ...... Stefanski, L. A. and Boos, D. D. (2002).

Entropic Graph Regularization in Non-Parametric ... - NIPS Proceedings
Most of the current graph-based SSL algorithms have a number of shortcomings – (a) in ... clude [7, 8]) attempt to minimize squared error which is not optimal for classification problems [10], ..... In this section we present results on two popular

The Beat Goes On: Rhythmic Modulation of Cortical ...
A good illustration of the type of information conveyed by tagging frequencies .... each trial, 20 brief (50 ms) stimuli were presented serially at a constant rate of 1 per 840 ms ..... is consistent with a greater effect on sensorimotor than posteri

Entropic Graph Regularization in Non-Parametric Semi ...
We also showed that this objective can be minimized using alternating minimization (AM), and can ... For example, on the Internet alone we create 1.6 ... We use the phone classification problem to demonstrate the scalability of the algorithm.

Reportage Gauquelin modulation d'azote.pdf
le spécialiste en images satellitaires radars Telespazio. “Appliqué au blé, cet outil d'aide à la décision s'adressera. aux agriculteurs, aux conseillers, semenciers.