10
Author's personal copy Automatica 44 (2008) 2498–2507 Contents lists available at ScienceDirect Automatica journal homepage: www.elsevier.com/locate/automatica MHD channel flow control in 2D: Mixing enhancement by boundary feedback Eugenio Schuster a,* , Lixiang Luo a , Miroslav Krstić b a Department of Mechanical Engineering and Mechanics, Lehigh University, Bethlehem, PA 18015-1835, United States b Department of Mechanical and Aerospace Engineering, University of California at San Diego, La Jolla, CA 92093-0411, United States article info Article history: Received 5 July 2006 Received in revised form 5 February 2008 Accepted 14 February 2008 Available online 19 September 2008 Keywords: MHD flow control Nonlinear boundary control Active mixing enhancement Distributed parameter systems abstract A nonlinear Lyapunov-based boundary feedback control law is proposed for mixing enhancement in a 2D magnetohydrodynamic (MHD) channel flow, also known as Hartmann flow, which is electrically conducting, incompressible, and subject to an external transverse magnetic field. The MHD model is a combination of the Navier–Stokes PDE and the Magnetic Induction PDE, which is derived from the Maxwell equations. Pressure sensors, magnetic field sensors, and micro-jets embedded into the walls of the flow domain are employed for mixing enhancement feedback. The proposed control law, designed using passivity ideas, is optimal in the sense that it maximizes a measure related to mixing (which incorporates stretching and folding of material elements), while at the same time minimizing the control and sensing efforts. A DNS code is developed, based on a hybrid Fourier pseudospectral-finite difference discretization and the fractional step technique, to numerically assess the controller. © 2008 Elsevier Ltd. All rights reserved. 1. Introduction Recent years have been marked by dramatic advances in active flow control (see Aamo and Krstic (2002) and the references therein), which, if implemented through micro-electro- mechanical sensors and actuators, can become effective in reducing drag and separation over aircraft wings, eliminating in- stabilities in various sections of jet engines (inlet, compressor rotating stall, combustion thermoacoustic oscillations, etc.), reduc- ing jet noise, reducing thermal signature of jet exhaust through actively controlled mixing, and steering the overall vehicle. Up until now active feedback flow control developments have had little impact on electrically conducting fluids moving in elec- tromagnetic fields. Active feedback control in electrically con- ducting flows, implemented through micro-electro-mechanical or micro-electro-magnetic actuators and sensors, can be used to op- timally achieve the desired level of stability (when suppression of turbulence is desired) or instability (when enhancement of mix- ing is desired). As a result, a small amount of active control ap- plied to magnetohydrodynamic (MHD) flows, magnetogasdynamic (MGD) flows, and plasma flows can dramatically change their equi- librium profiles and stability (turbulent fluctuation) properties. This paper was not presented at any IFAC meeting. This paper was recommended for publication in revised form by Associate Editor Denis Dochain under the direction of Editor Frank Allgöwer. Supported by the Pennsylvania Infrastructure Technology Alliance and the NSF CAREER program (ECCS-0645086). * Corresponding author. Tel.: +1 610 758 5253. E-mail address: [email protected] (E. Schuster). These changes influence heat transfer, hydrodynamic drag, pres- sure drop, and the pumping power required to drive the fluid. Prior work in the area of active control of electrically- conducting-fluid flows focuses mainly on electro-magneto-hydro- dynamic (EMHD) flow control for hydrodynamic drag reduction, through turbulence control, in weak electrically conducting fluids such as saltwater. Traditionally two types of actuator designs have been used: one type generates a Lorentz field parallel to the wall in the streamwise direction, while the other type generates a Lorentz field normal to the wall in the spanwise direction. EMHD flow control has been dominated by open-loop strategies that either permanently activate the actuators or pulse them at arbitrary frequencies. However, it has been shown that feedback control schemes can improve the efficiency, by reducing control power, for both streamwise (Spong, Reizes, & Leonardi, 2005) and spanwise (Berger, Kim, Lee, & Lim, 2000; Choi, Moin, & Kim, 1994) approaches. Model-based designs for electromagnetically actuated control for drag reduction have been proposed, using distributed control techniques based on linearization and model reduction, in Baker, Armaou, and Christofides (2002); Singh and Bandyopadhyay (1997). We consider a novel flow control problem that arises when an electrically conducting fluid interacts with a magnetic field in applications that range from liquid metals to plasmas. When an electrically conducting fluid moves in the presence of a transverse magnetic field, it produces an electric field due to charge separation and subsequently an electric current. The interaction between this created electric current and the imposed magnetic field produces a body force, called the Lorentz force, which acts on the fluid itself. Since this force acts in the opposite direction of the 0005-1098/$ – see front matter © 2008 Elsevier Ltd. All rights reserved. doi:10.1016/j.automatica.2008.02.018

eus204/publications/journals/aut08_2D... · 2017-01-04 · Author's personal copy Automatica 44 (2008) 2498 2507 Contents lists available atScienceDirect Automatica journal homepage:

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: eus204/publications/journals/aut08_2D... · 2017-01-04 · Author's personal copy Automatica 44 (2008) 2498 2507 Contents lists available atScienceDirect Automatica journal homepage:

Author's personal copy

Automatica 44 (2008) 2498–2507

Contents lists available at ScienceDirect

Automatica

journal homepage: www.elsevier.com/locate/automatica

MHD channel flow control in 2D: Mixing enhancement by boundary feedbackI

Eugenio Schuster a,∗, Lixiang Luo a, Miroslav Krstić b

a Department of Mechanical Engineering and Mechanics, Lehigh University, Bethlehem, PA 18015-1835, United Statesb Department of Mechanical and Aerospace Engineering, University of California at San Diego, La Jolla, CA 92093-0411, United States

a r t i c l e i n f o

Article history:Received 5 July 2006Received in revised form5 February 2008Accepted 14 February 2008Available online 19 September 2008

Keywords:MHD flow controlNonlinear boundary controlActive mixing enhancementDistributed parameter systems

a b s t r a c t

A nonlinear Lyapunov-based boundary feedback control law is proposed for mixing enhancement ina 2D magnetohydrodynamic (MHD) channel flow, also known as Hartmann flow, which is electricallyconducting, incompressible, and subject to an external transverse magnetic field. The MHD model isa combination of the Navier–Stokes PDE and the Magnetic Induction PDE, which is derived from theMaxwell equations. Pressure sensors, magnetic field sensors, and micro-jets embedded into the walls ofthe flow domain are employed for mixing enhancement feedback. The proposed control law, designedusing passivity ideas, is optimal in the sense that it maximizes a measure related to mixing (whichincorporates stretching and folding of material elements), while at the same timeminimizing the controland sensing efforts. A DNS code is developed, based on a hybrid Fourier pseudospectral-finite differencediscretization and the fractional step technique, to numerically assess the controller.

© 2008 Elsevier Ltd. All rights reserved.

1. Introduction

Recent years have been marked by dramatic advances inactive flow control (see Aamo and Krstic (2002) and thereferences therein), which, if implemented throughmicro-electro-mechanical sensors and actuators, can become effective inreducing drag and separation over aircraft wings, eliminating in-stabilities in various sections of jet engines (inlet, compressorrotating stall, combustion thermoacoustic oscillations, etc.), reduc-ing jet noise, reducing thermal signature of jet exhaust throughactively controlled mixing, and steering the overall vehicle.

Up until now active feedback flow control developments havehad little impact on electrically conducting fluids moving in elec-tromagnetic fields. Active feedback control in electrically con-ducting flows, implemented through micro-electro-mechanical ormicro-electro-magnetic actuators and sensors, can be used to op-timally achieve the desired level of stability (when suppression ofturbulence is desired) or instability (when enhancement of mix-ing is desired). As a result, a small amount of active control ap-plied tomagnetohydrodynamic (MHD) flows,magnetogasdynamic(MGD) flows, and plasma flows can dramatically change their equi-librium profiles and stability (turbulent fluctuation) properties.

I This paper was not presented at any IFAC meeting. This paper wasrecommended for publication in revised form by Associate Editor Denis Dochainunder the direction of Editor Frank Allgöwer. Supported by the PennsylvaniaInfrastructure Technology Alliance and the NSF CAREER program (ECCS-0645086).∗ Corresponding author. Tel.: +1 610 758 5253.

E-mail address: [email protected] (E. Schuster).

These changes influence heat transfer, hydrodynamic drag, pres-sure drop, and the pumping power required to drive the fluid.

Prior work in the area of active control of electrically-conducting-fluid flows focuses mainly on electro-magneto-hydro-dynamic (EMHD) flow control for hydrodynamic drag reduction,through turbulence control, in weak electrically conducting fluidssuch as saltwater. Traditionally two types of actuator designshave been used: one type generates a Lorentz field parallel to thewall in the streamwise direction, while the other type generatesa Lorentz field normal to the wall in the spanwise direction.EMHD flow control has been dominated by open-loop strategiesthat either permanently activate the actuators or pulse them atarbitrary frequencies. However, it has been shown that feedbackcontrol schemes can improve the efficiency, by reducing controlpower, for both streamwise (Spong, Reizes, & Leonardi, 2005)and spanwise (Berger, Kim, Lee, & Lim, 2000; Choi, Moin, & Kim,1994) approaches. Model-based designs for electromagneticallyactuated control for drag reduction have been proposed, usingdistributed control techniques based on linearization and modelreduction, in Baker, Armaou, and Christofides (2002); Singh andBandyopadhyay (1997).

We consider a novel flow control problem that arises whenan electrically conducting fluid interacts with a magnetic fieldin applications that range from liquid metals to plasmas. Whenan electrically conducting fluid moves in the presence of atransversemagnetic field, it produces an electric field due to chargeseparation and subsequently an electric current. The interactionbetween this created electric current and the imposed magneticfield produces a body force, called the Lorentz force, which acts onthe fluid itself. Since this force acts in the opposite direction of the

0005-1098/$ – see front matter© 2008 Elsevier Ltd. All rights reserved.doi:10.1016/j.automatica.2008.02.018

Page 2: eus204/publications/journals/aut08_2D... · 2017-01-04 · Author's personal copy Automatica 44 (2008) 2498 2507 Contents lists available atScienceDirect Automatica journal homepage:

Author's personal copy

E. Schuster et al. / Automatica 44 (2008) 2498–2507 2499

fluid motion, a high increase of power becomes necessary to drivethe fluid. In addition, this force tends to suppress turbulence andlaminarize the flow, which is undesirable in applications wherea high rate of heat transfer is needed. The heat transfer decreasedue to the laminarization may prevent electrically-conducting-fluid-based cooling systems from producing the heat transferimprovements expected based on the high thermal conductivityof the coolant. Active control can be used to enhance turbulence,mixing, and therefore heat transfer.

We focus in this paper on mixing enhancement by feedbackin MHD flows. We consider the Hartmann flow, an electricallyconducting, incompressible fluid moving between parallel platesthrough an imposed transverse magnetic field, and extendboundary control design ideas for Navier–Stokes equations (Aamo,Krstić, & Bewley, 2003; Balogh, Aamo, & Krstic, 2005) to MHDflows. Micro-jets, pressure sensors, and magnetic field sensorsembedded into the walls of the flow domain would be employedto implement our feedback control law. We develop a directnumerical simulation (DNS) code based on a hybrid Fourierpseudospectral-finite difference discretization scheme and thefractional step technique, and employ it to assess the effectivenessof the proposed controller in a 2D MHD channel flow. The globalmathematical well posedness of MHD equations was establishedin Chen and Wang (2002) for a free boundary problem. The localexact controllability was studied in Barbu, Havarneanu, Popa, andSritharan (2005).

The paper is organized as follows. Section 2 and 3 introducesthe governing equations and their equilibrium solution. Theperturbation equations are introduced in Section 4. The Lyapunovanalysis and the statement of optimality for the boundary controllaw is presented in Section 5. In Section 6 the numerical methodused to simulate the MHD channel flow is described. Results of anextensive simulation study are presented in Section 7. Section 8states the conclusions.

2. Governing equations

Let us consider the flow of an incompressible, conductingfluid between parallel plates where a magnetic field Bo = Boyperpendicular to the channel axis is externally applied. In addition,let us assume the presence of a uniformpressure gradient in the−xdirection. Fig. 1 illustrates the configuration; x and y denote theunit vectors in the x and y directions respectively. This flow wasfirst investigated experimentally and theoretically by Hartmann(1937). The governing equations for the stated problem are thetransport equation of linear momentum

ρ

[∂v∂t

+ (v · ∇)v]

= −∇P + ρν∇2v + j × B, (1)

and the transport equation of magnetic induction

∂B∂t

+ (v · ∇)B =1

µσ∇

2B + (B · ∇)v. (2)

The flow velocity is denoted by v, the magnetic field by B andthe current density by j, while P denotes the pressure, ρ thefluid mass density, ν the kinematic viscosity, µ the magneticpermeability and σ the electrical conductivity. The j × B termrepresents the Lorentz forces. The Lorentz forces couple themechanical and electrodynamic states of the system and act inplanes perpendicular to both current density and magnetic fieldvectors. Coulomb forces qE, where q is the electrical charge andE the electrical field, are negligible in comparison to the Lorentzforces. Themagnetic induction equation is derived fromOhm’s lawj = σ(E + v × B), Faraday’s law ∂B

∂t = −∇ × E, Ampere’s lawµj = ∇ × B, and the fact that B and v are solenoidal ∇ · B = 0,∇ · v = 0.

Fig. 1. Flow between parallel plates in the presence of a transverse magnetic field(Hartmann flow).

Fig. 2. 2D Hartmann flow.

In this work we consider the 2D Hartmann flow. Fig. 2 showsthe geometrical arrangement, where −L ≤ y ≤ L, −∞ < x < ∞.The imposed magnetic field Bo is perpendicular to both planes. Inthis case we can write x = xx + yy, v = v(x, y, t) = U(x, y, t)x +

V (x, y, t)y, B = B(x, y, t) = Bu(x, y, t)x + Bv(x, y, t)y and P =

P(x, y, t).

3. Equilibrium solution

For channels with constant cross section, as the one depictedin Fig. 2, a fully developed equilibrium flow is established. Inthis case, the flow velocity v = U(y)x has only one component,which depends on the coordinate y (the upper bar denotesequilibrium variables). The magnetic field is decomposed into twocontributions, one due to the external imposed magnetic field andthe other caused by the magnetic field induced by the flow B =

Bo + b = Boy+ b. Substituting this expression for the equilibriummagnetic field B into Eq. (2), and forcing the temporal derivative tozero, shows that the only component of the inducedmagnetic fieldis b = b(y)x. The induction equation reduces then to

0 = µσBodUdy

+d2bdy2

. (3)

Using Ampere’s law it is possible to write the current density j,and consequently the Lorentz force j × B, in terms of b. Then themomentum equation can be written as

0 = −dPdx

+Bo

µ

dbdy

+ ρνd2Udy2

. (4)

We consider viscous fluids with no slip at the fluid-wall interfaceΓ . Therefore the hydrodynamic boundary condition is

v = 0 at Γ , (5)

which means that all the velocity components vanish at the wall.For walls with finite electrical conductivity σw , magnetic perme-ability µw and normal n, the condition that the tangential compo-nent of the electrical field is continuous across the wall interfacecan be expressed in terms of b as Müller and Bühler (2001)

∂ b∂n

−1cb = 0 at Γ , (6)

Page 3: eus204/publications/journals/aut08_2D... · 2017-01-04 · Author's personal copy Automatica 44 (2008) 2498 2507 Contents lists available atScienceDirect Automatica journal homepage:

Author's personal copy

2500 E. Schuster et al. / Automatica 44 (2008) 2498–2507

Fig. 3. Velocity and induced magnetic field profiles for Hartmann flow at Hartmann numbers Ha = 0, 2, 5, 10, 100 for perfectly insulating walls (c = 0).

with the wall conductance ratio defined as c =µwσw tw

µσ L where thewall thickness tw is often small compared to the dimension of thecross section L. Two limiting cases can be considered: i- b = 0 atΓ as c → 0 (perfectly insulating walls), ii- ∂ b

∂n = 0 at Γ as c → ∞

(perfectly conducting walls).Defining the dimensionless variables y∗

=yxo, U∗

=Uvo, b∗

=

bbo, where xo = L, vo =

L2ρν

(− ∂ P∂x ), and bo = µL2

√σρν

(− ∂ P∂x ), we can

rewrite Eqs. (3) and (4) as

HadU∗

dy∗+

d2b∗

dy∗2= 0, Ha

db∗

dy∗+

d2U∗

dy∗2= −1, (7)

with boundary conditions (5) and (6) now expressed as

U∗= 0 at y∗

= ±1

∓db∗

dy∗−

b∗

c= 0 at y∗

= ±1,(8)

where Ha = BoL√

σρν

is the Hartmann number. The solution for

system (7) with boundary conditions (8) is given by

U∗(y∗) =1Ha

c + 1cHa + tanh(Ha)

[1 −

cosh(Ha y∗)

cosh(Ha)

], (9)

b∗(y∗) = −y∗

Ha+

1Ha

c + 1cHa + tanh(Ha)

sinh(Ha y∗)

cosh(Ha). (10)

Fig. 3 shows the velocity and induced magnetic field profiles fordifferent values of theHartmannnumberHa in the case of perfectlyinsulating walls, c = 0.

4. Perturbation equations

Defining the dimensionless variables x∗=

xxo, v∗

=vvo, t∗ =

votxo

, B∗=

Bbo, j∗ =

jσvobo

, with xo, vo, and bo defined in the previoussection, we can rewrite Eqs. (1) and (2) as

∂v∂t

+ (v · ∇)v = −∇P +1R∇

2v +NRm

[(∇ × B) × B] , (11)

∂B∂t

+ (v · ∇)B =1Rm

∇2B + (B · ∇)v, (12)

where R =voLν

is the Reynolds number, N =σ Lb2oρvo

is the Stuartnumber, and Rm = µσvoL is the magnetic Reynolds number. Thestar notation has been dropped for simplicity.

Defining the deviation variables as u = U − U , v = V − V = V ,bu = Bu

− Bu= Bu

− b, bv= Bv

− Bv= Bv

− Bo, p = P − P , wecan write the dimensionless perturbation equations as

∂u∂t

+ (U + u)∂u∂x

+ v∂(U + u)

∂y= −

∂p∂x

+1R

(∂2u∂x2

+∂2u∂y2

)−

NRm

(Bo + bv)

(∂bv

∂x−

∂bu

∂y

)+

NRm

bv ∂ b∂y

, (13)

∂v

∂t+ (U + u)

∂v

∂x+ v

∂v

∂y= −

∂p∂y

+1R

(∂2v

∂x2+

∂2v

∂y2

)+

NRm

(b + bu)(

∂bv

∂x−

∂bu

∂y

)−

NRm

bu∂ b∂y

, (14)

∂u∂x

+∂v

∂y= 0, (15)

∂bu

∂t+ (U + u)

∂bu

∂x+ v

∂(b + bu)∂y

=1Rm

(∂2bu

∂x2+

∂2bu

∂y2

)+ (b + bu)

∂u∂x

+ (Bo + bv)∂u∂y

+ bv ∂U∂y

, (16)

∂bv

∂t+ (U + u)

∂bv

∂x+ v

∂bv

∂y=

1Rm

(∂2bv

∂x2+

∂2bv

∂y2

)+ (b + bu)

∂v

∂x+ (Bo + bv)

∂v

∂y, (17)

∂bu

∂x+

∂bv

∂y= 0, (18)

with initial conditions u(x, y, 0) = uo(x, y), v(x, y, 0) = vo(x, y),bu(x, y, 0) = buo(x, y), b

v(x, y, 0) = bvo(x, y) for −∞ < x < ∞,

−1 < y < 1 and t > 0.

5. Energy analysis, control design, and its inverse optimality

Choosing the energy function as the combination of theperturbed kinetic and magnetic energies of the flow,

E(v, B) =12

∫ 1

−1

∫ d

0k1(u2

+ v2) + k2(bu2+ bv2)dxdy, (19)

we can compute

E(v, B) =

∫ 1

−1

∫ d

0(k1uut + k1vvt + k2bubut + k2bvbv

t )dxdy

Page 4: eus204/publications/journals/aut08_2D... · 2017-01-04 · Author's personal copy Automatica 44 (2008) 2498 2507 Contents lists available atScienceDirect Automatica journal homepage:

Author's personal copy

E. Schuster et al. / Automatica 44 (2008) 2498–2507 2501

= k1

∫ 1

−1

∫ d

0−u

[Uux + uux + vU ′

+ vuy −1R

(uxx + uyy

)+ px

]dxdy

+ k1

∫ 1

−1

∫ d

0−u

[NRm

Bo(bvx − buy

)+

NRm

bv(bvx − buy

)−

NRm

bv b′

]dxdy

+ k1

∫ 1

−1

∫ d

0−v

[Uvx + uvx + vvy −

1R

(vxx + vyy

)+ py

]dxdy

+ k1

∫ 1

−1

∫ d

0−v

[−

NRm

b(bvx − buy

)−

NRm

bu(bvx − buy

)+

NRm

bub′

]dxdy

+ k2

∫ 1

−1

∫ d

0−bu

[Ubux + ubux + vb′

+ vbuy −1Rm

(buxx + buyy

)]dxdy

+ k2

∫ 1

−1

∫ d

0−bu

[−bux − buux − Bouy − bvuy − bvU ′

]dxdy

+ k2

∫ 1

−1

∫ d

0−bv

[Ubv

x + ubvx + vbv

y −1Rm

(bvxx + bv

yy

)]dxdy

+ k2

∫ 1

−1

∫ d

0−bv

[−bvx − buvx − Bovy − bvvy

]dxdy, (20)

where U ′ and b′ denote Uy and by respectively.We assumeperiodicboundary conditions in the streamwise direction, i.e., v(x = 0) =

v(x = d), B(x = 0) = B(x = d) and p(x = 0) = p(x = d). Weapply control only in the wall normal direction, i.e.,

u(x, −1, t) = u(x, 1, t) = 0, (21)v(x, −1, t) = v(x, 1, t) = vwall(x, t), (22)

where the control vwall(x, t), to be employed on both the top andthe bottom walls, is to be designed. Note that (22) ensures thatthe net mass flux through the walls be zero. We measure the wallnormal component of the induced magnetic field,

bu(x, −1, t) = bu(x, 1, t) = 0, (23)

bv(x, −1, t) = bvbot_wall(x, t), b

v(x, 1, t) = bvtop_wall(x, t), (24)

where bvbot_wall(x, t) and bv

bot_wall(x, t) are measured on the bottomand top wall, respectively, and (23) follows from assumingperfectly insulating walls.

Lemma 1. Taking into account boundary conditions (21)–(24) thetime derivative of E(v, B) along the trajectories can be written as

E(v, B) = −1Rm(v, B) −

∫ d

0vwall

k1∆p + k2∆

(bv2

)2

dx

+ g(v, B), (25)

where

m(v, B) = k1

∫ 1

−1

∫ d

0(u2

x + u2y + v2

x + v2y )dxdy

+ k2RRm

∫ 1

−1

∫ d

0

((bux)

2+ (buy)

2+ (bv

x )2+ (bv

y)2) dxdy, (26)

g(v, B) = −k1

∫ 1

−1

∫ d

0U ′uvdxdy (27)

− k2

∫ 1

−1

∫ d

0b′buvdxdy (28)

+ k2

∫ 1

−1

∫ d

0U ′bubvdxdy (29)

+ k1

∫ 1

−1

∫ d

0

NRm

b′(ubv

− vbu)dxdy (30)

+ k1

∫ 1

−1

∫ d

0

NRm

b(bvx − buy

)vdxdy (31)

− k1

∫ 1

−1

∫ d

0

NRm

Bo(bvx − buy

)udxdy (32)

+ k2

∫ 1

−1

∫ d

0b(buux + bvvx

)dxdy (33)

+ k2

∫ 1

−1

∫ d

0Bo

(buuy + bvvy

)dxdy (34)

+ k1

∫ 1

−1

∫ d

0

NRm

bu(bvx − buy

)vdxdy (35)

− k1

∫ 1

−1

∫ d

0

NRm

bv(bvx − buy

)udxdy (36)

+ k2

∫ 1

−1

∫ d

0bubuuxdxdy (37)

+ k2

∫ 1

−1

∫ d

0bubvvxdxdy (38)

+ k2

∫ 1

−1

∫ d

0bvbuuydxdy (39)

+ k2

∫ 1

−1

∫ d

0bvbvvydxdy, (40)

∆p = P(x, 1, t) − P(x, −1, t), (41)

(bv2

)= (bv(x, 1, t))2 − (bv(x, −1, t))2. (42)

This lemma, proved in Appendix A, provides a relationshipbetween the time derivative of E(v, B) and the function m(v, B),which appears to be connected to mixing. A number of inherentlydifferent processes is called mixing. Ottino (1989) distinguishessub-problems of mixing: (i) mixing of a single fluid (or similarfluids) governed by the stretching and folding of materialelements; (ii) mixing governed by diffusion or chemical reactions;and (iii) mixing of different fluids governed by the breakup andcoalescence of material elements. Of course, all processes may bepresent simultaneously. In this work, we are interested in the firstsub-problem. The measure (26) is related to mixing due to thedirect correspondence between stretching of material elementsand the spatial gradients of the flow field. Folding is presentimplicitly in (26) due to the boundedness of the flow domainand the fact that v satisfies the Navier–Stokes equation. In thisfirst sub-problem, the interfaces between the fluids are passive(Aref & Tryggvason, 1984), and the mixing may be determinedby studying the movement of a passive tracer, or dye, in ahomogeneous fluid flow. The intuitive correspondence betweenstretching of material elements and the spatial gradients of theflow field will be further reinforced with our dye simulations.The presence of the spatial derivatives of the induced magneticfield b in (26) is motivated by the direct relationship betweenthe perturbed induced magnetic field and the perturbed velocityfield. The incorporation of the spatial derivatives of the inducedmagnetic field in (26) is also consistent with the incorporation ofthe perturbedmagnetic energy in (19), and allows for the existenceof an elegant solution to the optimal control problem as stated inTheorem 1.

Lemma 2. The function g(v, B) satisfies

|g(v, B)| ≤ g1m(v, B) + g2m2(v, B) + g3

∫ d

0v2walldx +

12q(v, B),

where q(v, B) = g4n+g5n2, n(v, B) =∫ d0 (bv

top_wall)2+(bv

bot_wall)2dx

and g1, g2, g3, g4 and g5 are nonnegative constants which depend onlyon the flow parameters.

Page 5: eus204/publications/journals/aut08_2D... · 2017-01-04 · Author's personal copy Automatica 44 (2008) 2498 2507 Contents lists available atScienceDirect Automatica journal homepage:

Author's personal copy

2502 E. Schuster et al. / Automatica 44 (2008) 2498–2507

This lemma, proved in Appendix B, provides a bound onthe crossterm, involving both the perturbation and equilibriumvariables, that originates from the nonlinear terms in the MHDequations. This term is similar to the so-called instantaneousproduction term in the fluid mechanics literature.

Our goal is to design a feedback control law, in terms of suctionand blowing of fluid normally to the channel wall (achievable bymicro-electro-mechanical (MEM) jets (Lofdahl & el Hak, 1999)),that is optimal with respect to some meaningful cost functionalrelated to m(v, B). The control solution is presented in thefollowing theorem.

Theorem 1. The cost functional

J(vwall) = limt→∞

[2βE(v(t), B(t)) +

∫ t

0h(v(τ ), B(τ ))dτ

], (43)

where β > g3 is a positive constant and

h(v, B) =2βR

m(v, B) − 2βg(v, B) − β

∫ d

0v2walldx

− β

∫ d

0

k1∆p + k2∆

(bv2

)2

2

dx, (44)

is maximized by the control

vwall = −

k1∆p + k2∆

(bv2

)2

. (45)

Moreover, for arbitrary values of control vwall, solutions of system(13)–(18) satisfy

h(v, B) ≤ l1m(v, B) + l2m2(v, B) + βq(v, B)

− l3

∫ d

0v2walldx − β

∫ d

0

k1∆p + k2∆

(bv2

)2

2

dx, (46)

l1 = 2β(1R

+ g1

), l2 = 2βg2, l3 = β − g3. (47)

Proof. By Lemma 1, we can write Eq. (44) as

h(v, B) = −2βE(v, B) − β

∫ d

0

k1∆p + k2∆

(bv2

)2

2

dx

− 2β∫ d

0vwall

k1∆p + k2∆

(bv2

)2

dx − β

∫ d

0v2walldx

= −2βE(v, B) − β

∫ d

0

vwall +

k1∆p + k2∆

(bv2

)2

2

dx,

(48)

and the cost functional can be written as

J(vwall) = limt→∞

2βE(v(t), B(t)) − 2β∫ t

0E(v(τ ), B(τ ))dτ

−β

∫ t

0

∫ d

0

vwall +

∆p +

(bv2

)2

2

dxdτ

= 2βE(v(0), B(0))

− β limt→∞

∫ t

0

∫ d

0

vwall +

k1∆p + k2∆

(bv2

)2

2

dxdτ .

(49)

The cost functional (43) is maximized when the last integral in(49) is zero. Therefore the control (45) is optimal. In addition, byLemma 2 we can write

h(v, B) ≤2βR

m(v, B) + βq(v, B) − β

∫ d

0v2walldx

−β

∫ d

0

k1∆p + k2∆

(bv2

)2

2

dx

+ 2β(g1m(v, B) + g2m2(v, B) + g3

∫ d

0v2walldx

)≤ l1m(v, B) + l2m2(v, B) + βq(v, B)

− l3

∫ d

0v2walldx − β

∫ d

0

k1∆p + k2∆

(bv2

)2

2

dx. �

The goal of the control law (45) is to increase the value ofm(v, B). It is clear from inequality (46),which gives anupper boundon h(v, B) in terms of m(v, B), that this goal is targeted in the costfunctional (43). Inequality (46) implies that h(v, B) cannot bemadelarge without making the mixing measure m(v, B) large, so thecost functional (43) is meaningful with respect to our goal. Notingthat β > g3 implies that l3 is positive (which is not a designchoice because β is just an analysis constant in the cost functional,whereas the gains k1 and k2 can have arbitrary positive values), weobserve that the control law (45) maximizes J(vwall), and thereforeh(v, B), and consequently mixing, with minimal control (vwall)

and sensing (∆p, ∆(bv2)) effort (the cost function (43) also putspenalty on the control and sensing effort through h(v, B)). The onlyterm whose role in (46) is not obvious is q(v, B). This term is notrelated tomixing in an obviousway but it is a perturbation variableand, as such, its growth indicates a growth of instability, whichcontributes to mixing.

The feedback (45) is independent of the parameters of the flow,and thus robust to parameter uncertainties. It requires sensing (ofpressure and induced magnetic field) only at the boundary and isdecentralized.

6. Numerical method

A direct numerical simulation is performed based on thefull MHD equations, to allow the measurement of the inducedmagnetic field at the boundary, as required by the controllaw (45). Although past research exists on the simulation ofthe MHD equations for compressible flows, results for unsteadyincompressible flows are scarce, due to inherent challenges. Thefirst difficulty is in the multiple time scales—while the momentumequation has R � 1, the induction equation has Rm � 1.Secondly, the MHD equations become stiffer as the magneticReynolds number decreases. Based on the similar structures ofthe Navier–Stokes and Magnetic Induction equations, our firstapproach to the problem was to integrate the equations withdifferent integration steps on a staggered grid within a periodicchannel flow geometry using a hybrid Fourier pseudospectral—finite difference discretization and the fractional step technique.Taking advantage of the periodic boundary conditions in thestreamwise (x) direction, this direction is discretized using Fourier

Page 6: eus204/publications/journals/aut08_2D... · 2017-01-04 · Author's personal copy Automatica 44 (2008) 2498 2507 Contents lists available atScienceDirect Automatica journal homepage:

Author's personal copy

E. Schuster et al. / Automatica 44 (2008) 2498–2507 2503

pseudospectral methods (Canuto, 1998), while the wall-normal(y) direction is discretized using central finite differences on anon-uniform staggered grid (Morinishi, Lund, Vasilyev, & Moin,1998). The equations are integrated in time using a fractionalstep method (Dukowicz & Dvinsky, 1992), designed to ensure thefulfillment of the divergence-free conditions, based on a hybridRunge–Kutta/Crank–Nicolson time discretization (Bewley, 1999).Nonlinear terms are integrated explicitly using a fourth-order,low-storage Runge–Kutta method, while linear terms are treatedimplicitly using the Crank–Nicolson method.

7. Simulation results

In this section, the laminarization property of the imposedmagnetic field and the effectiveness of the proposed control law(45) for mixing enhancement are studied numerically on the flowdomain −1 < y < 1, 0 < x < 4π , with NX = 150 grid pointsin the x direction and NY = 128 grid points in the y direction, andwith fixed flow-rateQ = 1.5. The tests follow a specific procedure:first, a fully established hydrodynamic flow (no magnetic field)is calculated; second, a magnetic field is imposed on the fullyestablished flow, which leads to another fully established MHDflow with lower perturbation energy or even to a linearly stableMHD flow; finally, boundary feedback is applied to the MHD flowand an increase in the flow complexity is observed and confirmedby the evolution of dye blobs in the flow.

7.1. Hydrodynamic channel flow

When Bo = 0, the momentum equation (11) reduces to thewell-known Navier–Stokes equation. The two-dimensional chan-nel flow, also known as the Poiseuille flow, is frequently citedas a paradigm for transition to turbulence, and has drawn exten-sive attention through the history of fluid dynamics. This is a clas-sical flow control problem that has been studied in Aamo andKrstic (2002) and the references therein assuming the availabilityof an array of pressure sensors on the walls and an array of MEMSmicro-jet actuators (also distributed along the walls) capable ofblowing/suction in thewall-normal direction. Incompressible con-ventional flows in 2D channels can be stable for lowReynolds num-bers, as infinitesimal perturbations in the flow field are dampedout. The flows turn linearly unstable for high Reynolds numbersR > 5772 (Panton, 1996). Such flows usually reach statisticallysteady states, which we call fully established flows. The full MHDcode is capable of simulating 2D pure hydrodynamic channel flowsby simply setting B0 = 0, which means that no magnetic field isimposed. Simulation results are presented in Fig. 4 to show howa channel flow (R = 7500) develops to a fully established flow.The initial velocity profile is the parabolic equilibrium solutionof the Navier–Stokes equation, which is linearly unstable for thisReynolds number. Fig. 4 shows how the vorticity map evolves intime until reaching a fully established flow when the initial equi-librium velocity profile is infinitesimally perturbed at t = 0.

7.2. Stabilization effect of the imposed magnetic field in MHD channelflows

When Bo 6= 0, Fig. 3 shows that the equilibrium profileis flattened in the center of the channel. In addition, Fig. 5(a)shows the effect of the imposed transverse magnetic field on thestability properties of the flow. Vorticity maps obtained throughdirect numerical simulation studies show the stabilizing effect ofthe imposed magnetic field on the 2D Hartmann flow at t =

140, 285, 374. The magnetic field is imposed at t = 0 withthe fully established flow (R = 7500) achieved in Fig. 4 forthe pure hydrodynamic channel (Section 7.1). Magnetic fields

Fig. 4. Vorticity maps for R = 7500 at t = 0, 1262, 1682, 4485, for a purehydrodynamic channel flow (Bo = 0).

of three different levels of strength (B0 = 0.1, 0.2, 0.3) areimposed on the fully established flow at time t = 0. Themagnetic Reynolds number is Rm = 0.1 and the Stuart numberis N = 0.01 in all cases. Observing the vorticity maps, it isinteresting to note that weak magnetic fields (Ha < 3) havesignificant stabilization effects on the fully established flows. Flowswith lower Reynolds numbers, with a stronger tendency towardsstability, are more easily stabilized by the magnetic fields. Theperturbation energy of the velocity field, E(v) =

12d

∫ 1−1

∫ d0 (u2

+

v2)dxdy, is used to quantify the level of stability/instability of theflow. The time evolutions of perturbation energy are shown inFig. 5(b). In all cases, the perturbation energy is reduced by theimposed magnetic field, and another fully established flow profilewith lower perturbation energy is reached.

7.3. Simulations of controlled MHD flow

The controller is started at t = 0with the fully establishedMHDflow (R = 7500, Rm = 0.1, N = 0.01) shown in Fig. 5(a). Thegains of the control are the same for all cases (kv = 0.1, kb =

10,000). The time evolution of the vorticity map is shown forB0 = 0.3 in Fig. 6(a). Fig. 6(b) shows the perturbation energy E(v)

and the control effort C(v) =1d

∫ d0 v(x, −1, t)2 + v(x, 1, t)2dx,

for magnetic fields of different strength. The ratio between thekinetic energy of the boundary control flow and the perturbationkinetic energy, C(v)/E(v), is less than 1%, which suggests thatsmall control can result in considerable mixing effect. Fig. 7 showsthe evolution in time ofm(v, B), ourmixingmeasurement. In Fig. 7,the magnetic field is imposed at t = 0 with the fully establishedflow (R = 7500) achieved in Fig. 4 for the pure hydrodynamicchannel. The controller is started at around t = 6000 with thefully established flow (R = 7500, Rm = 0.1, N = 0.01) achievedin Fig. 5(a). We can observe once again the negative and positiveeffect on mixing produced by the magnetic field and the boundarycontrol respectively. An intuitive representation of the controlmechanism in this case can be seen from the boundary zoom-in(Fig. 8). The velocity vectors show that boundary control is pushing,by blowing, the nearby vortex into the center of the flow.

The mixing governed by the stretching and folding of materialelements, as the one considered in this work, can be determinedby studying the movement of a passive tracer, or dye, in ahomogeneous fluid flow. The location of the dye as a function oftime completely describes the mixing. A particle tracking analysisis carried out to further visualize the mixing effectiveness of the

Page 7: eus204/publications/journals/aut08_2D... · 2017-01-04 · Author's personal copy Automatica 44 (2008) 2498 2507 Contents lists available atScienceDirect Automatica journal homepage:

Author's personal copy

2504 E. Schuster et al. / Automatica 44 (2008) 2498–2507

(a) Flow being stabilized by magnetic field (B0 = 0.3, t = 140, 285, 374). (b) Perturbation energy as function of time (B0 = 0.1, 0.2, 0.3).

Fig. 5. Uncontrolled flow at R = 7500, Rm = 0.1, N = 0.01.

(a) Flow being destabilized by boundary control (B0 = 0.3, t = 47, 140, 327). (b) Perturbation energy and control effort as functions of time (B0 = 0.3).

Fig. 6. Controlled flow at R = 7500, Rm = 0.1, N = 0.01.

Fig. 7. Evolution ofm(v, B) (R = 7500, Rm = 0.1, N = 0.01, B0 = 0.3).

control law. At t = 0 several blobs are distributed along thecenterline of the channel, concentrated on several circular regions,as shown in Fig. 9(a). The 100,000 particles used in this simulationstudy are assumed to exactly follow the fluid motion. Fig. 9(b)shows the evolution in time of the dye blobs in the uncontrolledcase, whereas Fig. 9(c) shows the particle map evolution for thecontrolled flow. In both cases, the tracking starts with the fullyestablished MHD flow (R = 7500, Rm = 0.1, N = 0.01, B0 =

0.3) shown in Fig. 5(a). The difference in complexity between theuncontrolled and controlled cases is manifested.

Fig. 8. Controlled flow at R = 7500, Rm = 0.1, N = 0.01. Pressure and velocityzoom at the boundary (B0 = 0.3).

8. Conclusions

Using the L2-norm of first-order spatial derivatives of thevelocity and magnetic field perturbations as a measure of mixing(that incorporates stretching and folding of material elements),a feedback law that maximizes this measure and minimizes thecontrol and sensing efforts was designed for a 2D Hartman flow.The controller does not drive the states (or the control inputs)unbounded but it does locally destabilize the system, leadingto bounded unsteadiness, and, indirectly, to enhanced mixing.The controller effectiveness is demonstrated in a full MHD code,showing flow patterns considerably more complex than in thefully established uncontrolled flow, despite a small control effort,

Page 8: eus204/publications/journals/aut08_2D... · 2017-01-04 · Author's personal copy Automatica 44 (2008) 2498 2507 Contents lists available atScienceDirect Automatica journal homepage:

Author's personal copy

E. Schuster et al. / Automatica 44 (2008) 2498–2507 2505

Fig. 9. (a) Initial particle distribution (t = 0), (b) Particle distribution foruncontrolled flow (R = 7500, Rm = 0.1, N = 0.01, B0 = 0.3, t = 55), (c) Particledistribution for controlled flow (R = 7500, Rm = 0.1, N = 0.01, B0 = 0.3,t = 47, 140, 327).

compared to the reference flow velocity. Improved mixing isconfirmed with dye blob simulations in the flow. Consideringthe plant dimension, it is remarkable that the control is astatic output feedback (i.e., proportional, decentralized), yieldingimplementability in MEMS hardware (Balogh, Liu, & Krstic, 2001).

Acknowledgements

The authors want to thank Professor Thomas Bewley at UCSDfor helpful discussions and for providing the Navier–Stokes flowsolver which was used as the starting point for the development ofthe full MHD flow solver.

Appendix A. Proof of Lemma 1

Integrating by parts the first integral in expression (20), andrecalling the boundary conditions (21)–(24), we have∫ 1

−1

∫ d

0

[−U

12(u2)x −

12u(u2)x − uvU ′

−12v(u2)y

×1Ru

(uxx + uyy

)− upx

]dxdy

= −

∫ 1

−1U12u2

|d0 dy︸ ︷︷ ︸

=0

∫ 1

−1

12uu2

|d0 dy︸ ︷︷ ︸

=0

+

∫ 1

−1

∫ d

0

12u2uxdxdy

∫ 1

−1

∫ d

0U ′uvdxdy −

∫ d

0

12vu2

|1−1 dx︸ ︷︷ ︸

=0

+

∫ 1

−1

∫ d

0

12u2vydxdy

+1R

∫ 1

−1uux |

d0 dy︸ ︷︷ ︸

=0

−1R

∫ 1

−1

∫ d

0(ux)

2dxdy +1R

∫ d

0uuy |

1−1 dx︸ ︷︷ ︸

=0

−1R

∫ 1

−1

∫ d

0(uy)

2dxdy −

∫ 1

−1up |

d0 dy︸ ︷︷ ︸

=0

+

∫ 1

−1

∫ d

0uxpdxdy

= −

∫ 1

−1

∫ d

0U ′uvdxdy −

1R

∫ 1

−1

∫ d

0((ux)

2+ (uy)

2)dxdy

+

∫ 1

−1

∫ d

0uxpdxdy. (A.1)

Following identical approach we can rewrite the third integral in(20) as

−1R

∫ 1

−1

∫ d

0(v2

x + v2y )dxdy +

∫ 1

−1

∫ d

0vypdxdy −

∫ d

0vwall∆pdx,

(A.2)

wherewe have taken into account that vy|y=±1 = ux+vy|y=±1 = 0(u|y=±1 = 0 ⇒ ux|y=±1 = 0). Similarly, we can rewrite the fifthintegral in (20) as

∫ 1

−1

∫ d

0b′vbudxdy −

1Rm

∫ 1

−1

∫ d

0

[(bux)

2+ (buy)

2] dxdy, (A.3)

and the seventh integral in (20) as

∫ d

0

12vwall∆

(bv2

)dx −

1Rm

∫ 1

−1

∫ d

0

[(bv

x )2+ (bv

y)2] dxdy, (A.4)

wherewe have taken into account that bvy |y=±1 = bux+bv

y |y=±1 = 0(bu|y=±1 = 0 ⇒ bux |y=±1 = 0).

Adding the first, third, fifth, and seventh integrals in (20), andtaking into account (A.1)–(A.4) and (15) we obtain

−1Rm(v, B) −

∫ d

0vwall

k1∆p + k2∆

(bv2

)2

dx

−k1

∫ 1

−1

∫ d

0U ′uvdxdy − k2

∫ 1

−1

∫ d

0b′buvdxdy.

Adding the second, fourth, sixth, and eighth integrals in (20) weobtain (29)–(40). Consequently, we can write the time derivativeof E(v, B) as (25). �

Appendix B. Proof of Lemma 2

Following Balogh et al. (2001), we can write

v(x, y, t) = v(x, 1, t) −

∫ 1

yvy(x, y, t)dy

= vwall(x, t) −

∫ 1

yvy(x, y, t)dy, (B.1)

and therefore

v2(x, y, t) ≤ (1 + 2b)v2wall +

(1 +

2b

) (∫ 1

yvydy

)2

, (B.2)

where we use Young’s inequality (b > 0) to write

vwall

∫ 1

yvydy ≤ bv2

wall +1b

(∫ 1

yvydy

)2

.

By Schwartz inequality we can write(∫ 1

yvydy

)2

=

(∫ 1

y1vydy

)2

(∫ 1

y12dy

) (∫ 1

yv2ydy

)= (1 − y)

∫ 1

yv2ydy ≤ (1 − y)

∫ 1

−1v2ydy,

Page 9: eus204/publications/journals/aut08_2D... · 2017-01-04 · Author's personal copy Automatica 44 (2008) 2498 2507 Contents lists available atScienceDirect Automatica journal homepage:

Author's personal copy

2506 E. Schuster et al. / Automatica 44 (2008) 2498–2507

and conclude that∫ 1

−1

∫ d

0v2dxdy ≤ 2(1 + 2b)

∫ d

0v2walldx

+ 2(1 +

2b

) ∫ 1

−1

∫ d

0v2ydxdy. (B.3)

Similar derivations for u, bv , and bu provide∫ 1

−1

∫ d

0u2dxdy ≤ 2

∫ 1

−1

∫ d

0u2ydxdy, (B.4)∫ 1

−1

∫ d

0(bv)2dxdy ≤ 2(1 + 2b)

∫ d

0(bv

top_wall)2+ (bv

bot_wall)2dx

+ 2(1 +

2b

) ∫ 1

−1

∫ d

0(bv

y)2dxdy, (B.5)∫ 1

−1

∫ d

0(bu)2dxdy ≤ 2

∫ 1

−1

∫ d

0(buy)

2dxdy. (B.6)

Considering that, given w(x, y),∫ 1

−1

∫ d

0w4dxdy ≤ 2

∫ 1

−1

∫ d

0w2dxdy

∫ 1

−1

∫ d

0

[w2

x + w2y

]dxdy,

taking into account (B.4), (B.5), (B.6), and considering that 2pq <p2 + q2, we can write∫ 1

−1

∫ d

0(bu)4dxdy ≤ 2

{(∫ 1

−1

∫ d

0(buy)

2dxdy)2

+

(∫ 1

−1

∫ d

0

[(bux)

2+ (buy)

2] dxdy)2}

, (B.7)

∫ 1

−1

∫ d

0u4dxdy ≤ 2

{(∫ 1

−1

∫ d

0u2ydxdy

)2

+

(∫ 1

−1

∫ d

0

[u2x + u2

y

]dxdy

)2}

, (B.8)

∫ 1

−1

∫ d

0(bv)4dxdy ≤ 2 (1 + 2b)

{(∫ d

0(bv

top_wall)2dx

)2

+

(∫ 1

−1

∫ d

0

[(bv

x )2+ (bv

y)2] dxdy)2

}

+ 2(1 +

2b

) {(∫ 1

−1

∫ d

0(bv

y)2dxdy

)2

+

(∫ 1

−1

∫ d

0

[(bv

x )2+ (bv

y)2] dxdy)2

}. (B.9)

With these preliminary results, and by applying Young’sinequality, we can now find bounds for each one of the terms (27)–(40) of g(v, B) to obtain

|g(v, B)| ≤ h1

∫ 1

−1

∫ d

0u2xdxdy + h2

∫ 1

−1

∫ d

0u2ydxdy

+ h3

∫ 1

−1

∫ d

0v2xdxdy + h4

∫ 1

−1

∫ d

0v2ydxdy

+ h5

∫ 1

−1

∫ d

0(bv

x )2dxdy + h6

∫ 1

−1

∫ d

0(buy)

2dxdy

+ h7

∫ 1

−1

∫ d

0(bv

y)2dxdy + h8

(∫ 1

−1

∫ d

0u2ydxdy

)2

+ h8

(∫ 1

−1

∫ d

0

[u2x + u2

y

]dxdy

)2

+ h9

(∫ 1

−1

∫ d

0(buy)

2dxdy)2

+ h10

(∫ 1

−1

∫ d

0(bv

y)2dxdy

)2

+ h11

(∫ 1

−1

∫ d

0

[(bv

x )2+ (bv

y)2] dxdy)2

+ h12

(∫ 1

−1

∫ d

0

[(bux)

2+ (buy)

2] dxdy)2

+ h13

∫ 1

−1

∫ d

0

[(bv

x )4+ (buy)

4] dxdy + h14

∫ d

0v2walldx

+ h15

∫ d

0(bv

top_wall)2+ (bv

bot_wall)2dx

+ h16

(∫ d

0(bv

top_wall)2+ (bv

bot_wall)2dx

)2

,

where

h1 =k2a8

|b|max +k2a14

,

h2 = k1|U ′|max

2a1

+ k1NRm

|b′|max

2a4

+ k1NRm

2a7Bo2

+k2a10

Bo +k2a17

,

h3 = k21a9

|b|max +k2a15

,

h4 = k1|U ′|max2a1

(1 +

2b1

)+ k2|b′

|max2a2

(1 +

2b2

)+ k1

NRm

|b′|max2a5

(1 +

2b5

)+ k1

NRm

2(1 +

2b12

)+

k2a11

Bo + k1NRm

2a6|b|max2(1 +

2b6

)+

k2a16

,

h5 = k1NRm

1a6

|b|max + k1NRm

1a7

Bo,

h6 = k2|b′|max

2a2

+ k2|U ′|max

2a3

+ k1NRm

|b′|max

2a5

+ k1NRm

1a6

|b|max + k1NRm

1a7

Bo + k2a8|b|max2 + k2a10Bo2,

h7 = k2|U ′|max2a3

(1 +

2b3

)+ k1

NRm

|b′|max2a4

(1 +

2b4

)+ k2a9|b|max2

(1 +

2b9

)+ k2a11Bo2

(1 +

2b11

)+ k1

NRm

2(1 +

2b13

),

h8 = k1NRm

2a13c13

2,

h9 = k1NRm

2a12c12

2 + k22a14 + k22a15c15

+ k22a17c17

,

h10 = k22a15c15

(1 +

2b15

)+ k22a16

(1 +

2b16

)+ k22a17c17

(1 +

2b17

),

Page 10: eus204/publications/journals/aut08_2D... · 2017-01-04 · Author's personal copy Automatica 44 (2008) 2498 2507 Contents lists available atScienceDirect Automatica journal homepage:

Author's personal copy

E. Schuster et al. / Automatica 44 (2008) 2498–2507 2507

h11 = k22a15c15

(2 + 2b15 +

2b15

)+ k22a16

(2 + 2b16 +

2b16

)+ k22a17c17

(2 + 2b17 +

2b17

),

h12 = k1NRm

2a12c12

2 + k22a14 + k22a15c15

+ k22a17c17

,

h13 = k1NRm

a12c12 + k1NRm

a13c13,

h14 = k12a1|U ′|max(1 + 2b1) + k22a2|b′

|max(1 + 2b2)

+ k1NRm

2a5|b′|max(1 + 2b5) + k1

NRm

2a6|b|max2(1 + 2b6)

+ k1NRm

2a12

2(1 + 2b12),

h15 = k22a3|U ′|max(1 + 2b3) + k1

NRm

2a4|b′|max(1 + 2b4)

+ k2a9|b|max2(1 + 2b9) + k2a11Bo2(1 + 2b11)

+ k1NRm

2a13

2(1 + 2b13),

h16 = k22a15c15(1 + 2b15) + k22a16(1 + 2b16)+ k22a17c17(1 + 2b17).

Defining g1 = max(h1, h2, h3, h4, h5, h6, h7), g2 = 6max(h8,h9, h10, h11, h12, h13), g3 = h14, g4 = h15, g5 = h16, we finallyarrive at the inequality in Lemma 2. �

References

Aamo, O., & Krstic, M. (2002). Flow control by feedback. Springer.Aamo, O. M., Krstić, M., & Bewley, T. R. (2003). Control of mixing by boundary

feedback in 2d channel flow. Automatica, 39, 1597–1606.Aref, H., & Tryggvason, G. (1984). Vortex dynamics of passive and active interfaces.

Physica D, 12D, 59–70.Baker, J., Armaou, A., & Christofides, P. (2002). Drag reduction in transitional

linearized channel flow using distributed control. International Journal ofControl, 75(15), 1213–1218.

Balogh, A., Aamo, O.M., & Krstic,M. (2005). Optimalmixing enhancement in 3Dpipeflow. IEEE Transactions on Control Systems Technology, 13(1), 27–41.

Balogh, A., Liu,W.-J., & Krstic,M. (2001). Stability enhancement by boundary controlin 2D channel flow. IEEE Transactions on Automatic Control, 46(11), 1696–1711.

Barbu, V., Havarneanu, T., Popa, C., & Sritharan, S. (2005). Local exact controllabilityfor the magnetohydrodynamic equations revisited. Advances in DifferentialEquations, 10(5), 481–504.

Berger, T., Kim, J., Lee, C., & Lim, J. (2000). Turbulent boundary layer control utilizingthe lorentz force. Physics of Fluids, 12, 631.

Bewley, T. (1999). Optimal and robust control and estimation of transition, convection,and turbulence. Stanford University thesis.

Canuto, C. (1998). Spectral methods in fluid dynamics. Springer Verlag.Chen, G.-Q., &Wang, D. (2002). Global solutions of nonlinearmagnetohydrodynam-

ics with large initial data. Journal of Differential Equations, 182, 344–376.Choi, H., Moin, P., & Kim, J. (1994). Active turbulence control for drag reduction in

wall-bounded flows. Journal of Fluid Mechanics, 262, 75.

Dukowicz, J., & Dvinsky, A. (1992). Approximate factorization as a high order split-ting for the implicit incompressible flow equations. Journal of ComputationalPhysics, 102(2), 336–347.

Hartmann, J. (1937). Theory of the laminar flow of an electrically conductiveliquid in a homogeneous magnetic field. Det Kgl. Danske Videnskabernes Selskab.Mathematisk-fysiske Meddelelser , XV (6), 1–27.

Lofdahl, L., & el Hak,M. G. (1999).MEMS applications in turbulence and flow control.Progress in Aerospace Sciences, 35, 101–203.

Morinishi, Y., Lund, T., Vasilyev, O., &Moin, P. (1998). Fully conservative higher orderfinite difference schemes for incompressible flow. Journal of ComputationalPhysics, 143(1), 90–124.

Müller, U., & Bühler, L. (2001). Magnetofluiddynamics in channels and containers.Springer.

Ottino, J. M. (1989). The kinematics of mixing: Stretching, chaos, and transport.Cambridge University Press.

Panton, R. L. (1996). Incompressible flow (2nd ed.). New York: Wiley.Singh, S., & Bandyopadhyay, P. (1997). Linear feedback control of boundary layer

using electromagnetic microtiles. Transactions of ASME, 119, 852–858.Spong, E., Reizes, J., & Leonardi, E. (2005). Efficiency improvements of electromag-

netic flow control. Heat and Fluid Flow, 26, 635–655.

Eugenio Schuster is Assistant Professor of MechanicalEngineering and Mechanics at Lehigh University. Heholds undergraduate degrees in Electronic Engineering(Buenos Aires University, Argentina, 1993) and NuclearEngineering (Balseiro Institute, Argentina, 1998). Heobtained his M.S. (2000) and Ph.D. (2004) degrees atUniversity of California SanDiego. Schuster is the recipientof the NSF Career Award. His research interests are indistributed parameter andnonlinear control systems,withapplications that include fusion reactors, plasmas, andmagnetohydrodynamic flows.

Lixiang Luo received his B.S. degree from TsinghuaUniversity, China, in 2003, and hisM.S. degree from LehighUniversity in 2006, where he is now working towards aPh.D. His research interests include computational fluiddynamics, MHD (magnetohydrodynamics), reduced-ordermodeling and nonlinear control.

Miroslav Krstić is the Sorenson Professor and Directorof the Center for Control Systems and Dynamics at UCSan Diego. He received his Ph.D. in 1994 from UC SantaBarbara andwas Assistant Professor at University ofMary-land until 1997. He is a coauthor of the books Non-linear and Adaptive Control Design (1995), Stabilizationof Nonlinear Uncertain Systems (1998), Flow Control byFeedback (2002), Real-time Optimization by ExtremumSeeking Control (2003), Control of Turbulent and Mag-netohydrodynamic Channel Flows (2007), and BoundaryControl of PDEs: A Course on Backstepping Designs (2008).

Krstic received the Axelby and Schuck paper prizes, NSF Career, ONR Young Inves-tigator, and PECASE awards, the UCSD Research Award, and is a Fellow of IEEE. Hiseditorial service includes IEEE TAC, Automatica, SCL, and Int. J. Adaptive Contr. Sig.Proc. He was VP Technical Activities of IEEE Control Systems Society.