5
A three-dimensional model for the effective viscosity of bacterial suspensions Brian M. Haines, 1 Andrey Sokolov, 2, 3 Igor S. Aranson, 3 Leonid Berlyand, 1 and Dmitry A. Karpeev 4 1 Department of Mathematics, Penn State University, McAllister Bldg, University Park, PA 16802 2 Illinois Institute of Technology, 3101 South Dearborn Street, Chicago, IL 60616 3 Materials Science Division, Argonne National Laboratory, 9700 South Cass Avenue, Argonne, IL 60439 4 Mathematics and Computer Science Division, Argonne National Laboratory, 9700 South Cass Avenue, Argonne, IL 60439 (Dated: August 4, 2009) We derive the effective viscosity of dilute suspensions of swimming bacteria from the microscopic details of the interaction of an elongated body with the background flow. An individual bacterium propels itself forward by rotating its flagella and reorients itself randomly by tumbling. Due to the bacterium’s asymmetric shape, interactions with a prescribed generic (such as pure shear or planar shear) background flow cause the bacteria to preferentially align in directions in which self-propulsion produces a significant reduction in the effective viscosity, in agreement with recent experiments on suspensions of Bacillus subtilis. Recently, there has been much interest in studying the dynamical properties of interacting self-propelled organ- isms such as flocking birds, schooling fish, swarming bac- teria, etc [1–6]. Such systems, relying on non-local inter- actions, exhibit collective behavior absent from passive systems, such as suspensions of inert particles. Experi- ments on bacteria suspended in liquid films have demon- strated enhanced diffusion of tracer particles [1] and col- lective flow speeds that exceed the speed of an individual bacterium by an order of magnitude [7]. In our experi- ments (detailed in [8]), the swimming of suspensions of Bacillus subtilis was observed to cause a decrease in the effective viscosity by an order of magnitude. Whereas phenomenological arguments relating the viscosity of sus- pensions to the activity of particles [9] and studies of vis- cosity in two-dimensional geometry [10] have been pre- sented, the issue of viscosity in suspensions of swimmers still lacks conceptual clarity. In [9], the order parame- ter used to characterize the local ordering of the system (i.e., the local alignment of swimming particles) is bor- rowed from the theory of systems with nematic ordering and is inherently phenomenological. In particular, the relationship of the evolution of the order paramter to the microscopic alignment dynamics has not been clari- fied, and the very possibility of arriving at macroscopic expressions for the effective viscosity from first-principle arguments has not been established. In this Letter, we study the rheology of dilute suspen- sions of swimming bacteria. In the dilute regime (when the volume fraction φ 2%), bacteria interact only with the background flow and not each other, simplifying the problem dramatically. Nevertheless, producing an ad- equate model has proved challenging. Numerical sim- ulations in [11] using spherical bacteria with imposed boundary velocities modeling self-propulsion were able to predict a decrease in the effective viscosity only in the presence of a gravitational field. In [10], modeling bacte- ria as self-propelled discs (where drag on the flagellum is neglected), we obtained an asymptotic expression for the effective viscosity in terms of a given orientation distri- bution function P . This showed that there is a decrease in the effective viscosity due to self-propulsion when bac- teria align along one of the principal axes of the hydro- dynamic rate of strain tensor. This alignment can only be achieved by asymmetric bacteria. In this work, we significantly generalize and extend the work in [10]. We consider asymmetric bacteria in a three- dimensional geometry and include the effects of tumbling (random reorientation). We have derived analytical ex- pressions for all components of the deviatoric stress ten- sor and explicit expressions for the effective viscosity in several relevant cases. Most significantly, we base our analysis on microscopic considerations, and, hence, ex- tend and complement the phenomenological theory of nematic ordering. In particular, we quantify the depen- dence of the local ordering and the effective viscosity on the microscopic parameters: the shear rate, vorticity of the flow and the tumbling rate. Our analysis shows a general trend of decreasing effec- tive viscosity with increasing bacterial concentration and activity (speed of swimming). The effect is best demon- strated in Fig 1 where selected results on experiments with aerobic motile bacteria Bacillus subtilus confined in a free-standing liquid film are presented. A detailed ac- count of the experimental work will be reported in the companion paper [8]. The viscosity η is inferred from the decay rate ν of a macroscopic vortex of size L 2 mm in- duced in a thin film containing a bacterial suspension by a moving microprobe. The experiments were performed for a broad range of bacterial volume fraction φ. As one can see from the Figure, an almost seven-fold drop of the decay rate ν is measured while the density is increased from φ = 0 (no bacteria) to φ 2%. Since the viscos- ity η νL 2 , this indicates a significant decrease of the viscosity. Bacteria come in a variety of shapes, including spheres, rods, and spirals and employ many different forms of motility. We consider bacteria similar to the B. sub- tilis used in the experiments in [7]. This is a rod-shaped unicellular microorganism that propels itself through the

A three-dimensional model for the effective viscosity of bacterial suspensions

Embed Size (px)

DESCRIPTION

bacterial suspension

Citation preview

Page 1: A three-dimensional model for the effective viscosity of bacterial suspensions

A three-dimensional model for the effective viscosity of bacterial suspensions

Brian M. Haines,1 Andrey Sokolov,2, 3 Igor S. Aranson,3 Leonid Berlyand,1 and Dmitry A. Karpeev4

1Department of Mathematics, Penn State University, McAllister Bldg, University Park, PA 168022Illinois Institute of Technology, 3101 South Dearborn Street, Chicago, IL 60616

3Materials Science Division, Argonne National Laboratory, 9700 South Cass Avenue, Argonne, IL 604394Mathematics and Computer Science Division, Argonne National Laboratory, 9700 South Cass Avenue, Argonne, IL 60439

(Dated: August 4, 2009)

We derive the effective viscosity of dilute suspensions of swimming bacteria from the microscopicdetails of the interaction of an elongated body with the background flow. An individual bacteriumpropels itself forward by rotating its flagella and reorients itself randomly by tumbling. Due to thebacterium’s asymmetric shape, interactions with a prescribed generic (such as pure shear or planarshear) background flow cause the bacteria to preferentially align in directions in which self-propulsionproduces a significant reduction in the effective viscosity, in agreement with recent experiments onsuspensions of Bacillus subtilis.

Recently, there has been much interest in studying thedynamical properties of interacting self-propelled organ-isms such as flocking birds, schooling fish, swarming bac-teria, etc [1–6]. Such systems, relying on non-local inter-actions, exhibit collective behavior absent from passivesystems, such as suspensions of inert particles. Experi-ments on bacteria suspended in liquid films have demon-strated enhanced diffusion of tracer particles [1] and col-lective flow speeds that exceed the speed of an individualbacterium by an order of magnitude [7]. In our experi-ments (detailed in [8]), the swimming of suspensions ofBacillus subtilis was observed to cause a decrease in theeffective viscosity by an order of magnitude. Whereasphenomenological arguments relating the viscosity of sus-pensions to the activity of particles [9] and studies of vis-cosity in two-dimensional geometry [10] have been pre-sented, the issue of viscosity in suspensions of swimmersstill lacks conceptual clarity. In [9], the order parame-ter used to characterize the local ordering of the system(i.e., the local alignment of swimming particles) is bor-rowed from the theory of systems with nematic orderingand is inherently phenomenological. In particular, therelationship of the evolution of the order paramter tothe microscopic alignment dynamics has not been clari-fied, and the very possibility of arriving at macroscopicexpressions for the effective viscosity from first-principlearguments has not been established.

In this Letter, we study the rheology of dilute suspen-sions of swimming bacteria. In the dilute regime (whenthe volume fraction φ 2%), bacteria interact only withthe background flow and not each other, simplifying theproblem dramatically. Nevertheless, producing an ad-equate model has proved challenging. Numerical sim-ulations in [11] using spherical bacteria with imposedboundary velocities modeling self-propulsion were ableto predict a decrease in the effective viscosity only in thepresence of a gravitational field. In [10], modeling bacte-ria as self-propelled discs (where drag on the flagellum isneglected), we obtained an asymptotic expression for theeffective viscosity in terms of a given orientation distri-

bution function P . This showed that there is a decreasein the effective viscosity due to self-propulsion when bac-teria align along one of the principal axes of the hydro-dynamic rate of strain tensor. This alignment can onlybe achieved by asymmetric bacteria.

In this work, we significantly generalize and extend thework in [10]. We consider asymmetric bacteria in a three-dimensional geometry and include the effects of tumbling(random reorientation). We have derived analytical ex-pressions for all components of the deviatoric stress ten-sor and explicit expressions for the effective viscosity inseveral relevant cases. Most significantly, we base ouranalysis on microscopic considerations, and, hence, ex-tend and complement the phenomenological theory ofnematic ordering. In particular, we quantify the depen-dence of the local ordering and the effective viscosity onthe microscopic parameters: the shear rate, vorticity ofthe flow and the tumbling rate.

Our analysis shows a general trend of decreasing effec-tive viscosity with increasing bacterial concentration andactivity (speed of swimming). The effect is best demon-strated in Fig 1 where selected results on experimentswith aerobic motile bacteria Bacillus subtilus confined ina free-standing liquid film are presented. A detailed ac-count of the experimental work will be reported in thecompanion paper [8]. The viscosity η is inferred from thedecay rate ν of a macroscopic vortex of size L ∼ 2 mm in-duced in a thin film containing a bacterial suspension bya moving microprobe. The experiments were performedfor a broad range of bacterial volume fraction φ. As onecan see from the Figure, an almost seven-fold drop of thedecay rate ν is measured while the density is increasedfrom φ = 0 (no bacteria) to φ 2%. Since the viscos-ity η ∝ νL2, this indicates a significant decrease of theviscosity.

Bacteria come in a variety of shapes, including spheres,rods, and spirals and employ many different forms ofmotility. We consider bacteria similar to the B. sub-tilis used in the experiments in [7]. This is a rod-shapedunicellular microorganism that propels itself through the

Page 2: A three-dimensional model for the effective viscosity of bacterial suspensions

2

FIG. 1: Semi-logarithmic plot illustrating the decay of thetypical velocity v of a macroscopic vortex normalized by themaximum velocity vmax for different values of the volume frac-tion φ. The dashed lines show a fit to v ∼ exp(−νt).

rotation of several helical flagella distributed through-out its body. It moves in two distinct modes (see, e.g.[12, 13]): forward movement (swimming) and tumbling(random reorientation). It spends most of its time mov-ing forward, with its flagella bundled together so thatthey rotate as one strand. At random times (dependingon conditions, the average time between tumbling eventsvaries from 1 to 60 sec—see [13]), the flagella unbundleand rotate separately, causing the bacterium to reorient,until the next bundling event. This effectively resultsin a random reorientation of the bacterium that can bemodeled as a random walk on the unit sphere [14].

We model a bacterium as a rigid prolate spheroid withsemi-axes a, b (see Fig. 2) subject to no-slip boundaryconditions in a Stokesian fluid of viscosity η whose ve-locity is denoted by u and pressure by p. Self-propulsionis modeled by means of a rigidly attached point force ofmagnitude c and direction d located at xf , displaced byb(1 + λ)d from the center of the bacterium xc. In orderto model tumbling, the bacterium exerts a torque τ onthe fluid selected to make ˙

d contain a white noise pro-cess. The dynamics of each bacterium is determined byenforcing a balance of hydrodynamic forces and torqueswith those exerted by the bacterium on the water:

∫∂B

σ · ndx + cd = 0∫∂B σ · n × (x − xc) dx + τ = 0,

(1)

where ∂B is the surface of the bacteria, σij := −pδij +

2ηeij is the stress, and eij := 12

(∂ui

∂xj+ ∂uj

∂xi

)is the rate

of strain.In the dilute limit, the fluid problem is posed for a

single particle in an infinite domain. We obtained theexact solution by taking the Green’s function for Stokes’

FIG. 2: Schematic representation of a bacterium (left), bac-terium with orientation (α, β), alternatively described by the

unit vector d (right).

equation and canceling the flow at the surface of the bac-terium using Faxen’s law for the prolate spheroid (see[15]). Assuming the bacterium is placed in a backgroundflow described by a constant rate of strain E and vor-ticity Ω, we plug the hydrodynamic solutions into (1)to obtain the equations of motion for a single bacteriumwith linear velocity v = vd and angular velocity ω:

c =6πbηvN (2a)

τi =8πηb3[XCdidj + Y C (δij − didj)

](Ωj − ωj)

− 8πηb3Y HεijmdmdkEjk, (2b)

XC , Y C , Y H are scalar functions of the eccentricity e =√b2 − a2/b and N is a scalar function of e, λ [16].Since in isotropic media the effective viscosity does not

depend on the locations of individual bacteria, we do notneed to solve eq. (2a) explicitly. Introducing P (d, t),which represents the probability that any bacterium, attime t, has orientation d, eq. (2b) leads to the Fokker-Planck equation

∂P

∂t= −∇d ·

(P

˙dD

)+ D∆dP, (3)

where ∆d is the spherical Laplacian, ∇d is the spherical

gradient, and ˙dD is the deterministic contribution to ˙

d(due to interaction with the background flow only—i.e.,the solution of eq. (2b) with τ = 0). We assume therelevant distribution for calculating bulk quantities is thesteady state distribution P∞ obtained from eq. (3) bysetting ∂P∞

∂t = 0.The bulk deviatoric stress is defined by

Σij := nV

∫S2

∫V

(σij(d) − δijσkk(d)

)P∞(d)dxdS,

where nV is the volume density of bacteria, V is anyvolume containing the entire suspension (the value of Σ isindependent of V ), and σ(d) is the hydrodynamic stressdue to an inclusion centered at the origin with orientationd. Plugging in u and p, we obtain our first result,

Page 3: A three-dimensional model for the effective viscosity of bacterial suspensions

3

Σij =ηEij + φ

[b2

a2η

(5

∫S2

ΛijklP∞dSEkl − 3BY H

∫S2

(εikldj + εjkldi)dlεkmndmdnP∞dSEmn

+3Y HD

∫S2

(εikldj + εjkldi)dlεkmndm(∂nP∞)dS

)+

c

16πa2K

∫S2

(δij − 3didj)P∞dS

]+ O(φ2),

(4)

where Λ is a function of shape and orientation and Kis a scalar function of e and λ, both given in [16], andB := b2−a2

b2+a2 . The first term in eq. (4) is the standardNewtonian stress. The second and third terms are due tothe presence of passive spheroids and are taken from [15].The fourth term is the diffusive stress, obtained from [17].The last term is due to self-propulsion. Since K > 0,the sign of the last term is determined by the sign of c(c > 0 for bacteria) and the distribution of orientationsof the bacteria with respect to the background flow. Thisterm is proportional to δij − 3didj , the same form as theterm in [18] derived for ensembles of dipoles. It can beinterpreted as an active stress in the notation of [19].To obtain further insight into the rheology of bacterialsuspension, we must determine P∞.

Without loss of generality, we choose planar shearingand straining flows in three cases, allowing us to calcu-late viscosities valid in a variety of relevant physical sit-uations. First, we consider a pure straining flow, whichhas a tendency to align elongated particles and hencebacteria. Such a flow is representative of the flows bacte-ria would experience when swimming in groups, as theytend to align in these cases. Next, we calculate the ef-fective viscosity for planar shearing flows. This allowsfor comparison with previous analytical work on the vis-cosity of passive suspensions that use such flows (e.g.,[17, 20, 21]). Finally, we consider an oscillatory shearflow. This type of flow is used in experiments where vis-cosity is measured. The shearing and straining flows aredescribed by E with E12 = E21 = γ

2 and all other com-

ponents Eij = 0 with Ω =(0, 0,− γ

2

)and Ω = (0, 0, 0),

respectively. For the oscillatory case, we replace γ withγ sin ω0t. We define the effective viscosity by η := Σ12

γ .It is not possible to solve the Fokker-Planck equation an-alytically for any of our background flows, so we must doso asymptotically in various parameters and numericallyin the general case. The parameters used depend on theform of the Fokker-Planck equation for each flow. Thenumerics are performed using a finite difference schemeon a uniform mesh with 80,400 points. The resultinglinear system is solved using the method of bi-conjugategradients. Numerical viscosities are then calculated usingeq. (4) by setting the various parameters to values estab-lished in the literature. It is assumed that the bacteriahave dimensions b = 4µm (see [7]), b

a = 5.7 (except forthe small eccentricity asymptotics, where it is assumed

that B = 0.01) and λ = 0.5. We assume a bacterialswimming speed v = 50µm s−1 and a rotational diffu-sion constant D = 0.017s−1 (see [8]).

In the planar straining flow, bacteria tend to align withthe steady axis of the flow, corresponding to angles α =π4 , 3π

4 with β = π2 . Assuming full alignment with this

axis, we calculate an effective viscosity

η

η= 1 + φ

[S − 3bv

8a2γNK

]+ O(φ2), (5)

where S is a scalar function of e, given in [16]. The secondterm in eq. (5) represents the passive contribution andthe third the active contribution. This formula extendsthe result of [10] to 3D. Since K > 0, there is a decreasein the effective viscosity due to self-propulsion. Note thateq. (5) is only valid for γ D, because the assumptionof full alignment is invalid otherwise. Eq. (5) is plottedagainst γ with the corresponding numerics and a plot ofone numerical solution P∞ in Fig. 3.

0 1 2 3 4 5 6 7 8−150

−100

−50

0

γ (s−1)

( η η−

1) φ−

1

NumericsSmall µ asymptoticsAligned AsymptoticsNumeric fit2ππα0

0

π

π2

β

FIG. 3:“

ηη− 1

”φ−1 vs. γ for the case of no vorticity

(Ω = (0, 0, 0)) evaluated numerically along with small µasymptotics (eq. (6)), aligned asymptotics (eq. (5)), anda plot of P∞ for γ = 0.14s−1.

For the same flow, we now assume the dimensionlessparameter µ := γB

D 1 (i.e., bacteria have a weak ten-dency to align with the background flow due to beingnearly-spherical or weak advection). In this regime, wecalculate an asymptotic solution to the Fokker-Planckequation, given by P∞ = 1

[1 + µ 1

4 sin 2α sin2 β]

+O(µ2). Using eq. (4), we derive the effective viscosity

η

η= 1 + φ

[M − 3b

80a2γNKvµ + O (

µ2)]

+ O(φ2), (6)

where M is a scalar function of e given in [16]. Thesecond term is the passive contribution and the third

Page 4: A three-dimensional model for the effective viscosity of bacterial suspensions

4

term is the active contribution. This formula indicatesthat the suspension is very weakly non-Newtonian (sinceη = C + O(γ2)). A striking feature of the formula is thefact that the active contribution does not disappear in thelimit γ → 0. This is counterintuitive because as γ → 0,P∞ → 1

4π and hence the active contribution to the bulkdeviatoric stress Σa averages to zero. However, for smallγ, P∞ = 1

4π +Cγ + . . . and Σa acquires a contribution ∝γ. When calculating Σa from P∞, the constant term (inγ) averages to zero, but the linear term remains. Hence,the active contribution to the effective viscosity ηa :=Σa

γ contains a non-vanishing term. However, in reality,the time it takes for a suspension to reach the steadystate P∞ increases as γ → 0. Thus, for small enough γ,diffusion will dominate advection and the actual P∞ willbe closer to 1

4π , which will produce ηa = 0. Eq. (6) isplotted against γ along with corresponding numerics anda plot of one numerical solution P∞ in Fig. 3.

We next consider the planar shear flow. P∞ has beencalculated for small eccentricity e asymptotically in thesmall parameter ε := b

a − 1 in [17]. This yields

η

η= 1+φ

[52− 27vbDN(5λ2 + 10λ + 2)

10a2(36D2 + γ2)(1 + λ)4ε + O(ε2)

]+O(φ2).

(7)The second and third terms are the passive and activecontribution, respectively. As in [17], the effect due toeccentricity differs from that of a sphere at O(ε2). Eq. (7)is plotted against γ along with corresponding numericsand a plot of one numerical solution P∞ in Fig. 4.

10−6

10−4

10−2

100

102

−160

−140

−120

−100

−80

−60

−40

−20

0

20

γ (s−1)

( η η−

1) φ−

1

NumericsSmall γ asymptotics

10−4

10−2

100

102

2.0

2.2

2.4

2.6

γ (s−1)

( η η−

1) φ−

1

NumericsSmall ε asymp.

2ππα00

π

π2

β

2ππα00

π

π2

β

FIG. 4:“

ηη− 1

”φ−1 vs. γ for the case with vorticity. The in-

set is for B = b2−a2

b2+a2 = .01 along with small ε = ba− 1 asymp-

totics (eq. (7)) along with a plot of P∞ for γ = 0.017s−1 ,while the main plot is for b

a= 5.7 along with small γ

Dasymp-

totics (eq. (8)) and a plot of P∞ for γ = 0.14s−1.

We now place our particles in a weak ( γD 1) oscilla-

tory shear flow with frequency ω0. We then calculate thesolution to the Fokker-Planck equation asymptotically inγD : P (α, β, t) = 3

8πγBD sin 2α sin2 β6D sin ω0t

36D2+(ω0)2 + . . .. Usingthis, we can calculate an effective viscosity

η

η= 1 + φ

[M − K

27bvBDN

20a2 (36D2 + (ω0)2)

]+ . . . (8)

The second and third terms are the passive and activecontributions, respectively. Eq. (8) is plotted againstγ along with corresponding numerics and a plot of onenumerical solution P∞ in Fig. 4.

We have calculated the effective viscosity of bacte-rial suspension and performed experiments in qualitativeagreement. The dilute limit allowed us neglect the in-teractions between bacteria and thus to carry out thecalculations fully analytically. While it is not obviousthat the interactions are negligible in experiments, theresults are nevertheless in qualitative agreement with ex-periment. We have demonstrated and observed that bac-terial self-propulsion decreases the effective viscosity ofan ambient fluid. For weak enough background flows,this decrease outweighs the passive increase in viscositydue to the suspension to produce a net decrease in theviscosity of the fluid. For strong background flows theeffect of self-propulsion becomes negligible and an activesuspension becomes indistinguishable from a passive one.

The reduction in viscosity due to self-propulsion pre-dicted in our model relies on the bacteria being “push-ers” (i.e., c > 0). However, our results are still valid for“pullers” (e.g., some motile algae) which can be modeledin the same way but with c < 0. In this case, there is anincrease in viscosity due to self propulsion. That there isa fundamental difference in the physics of “pushers” and“pullers” was previously observed in [22, 23].

The work of I.S. Aranson, A. Sokolov, and D. Karpeevwas supported by the DOE grant DE-AC02-06CH11357.The work of B. Haines and L. Berlyand was supportedby the DOE grant DE-FG02-08ER25862 and NSF grantDMS-0708324.

[1] X.-L. Wu and A. Libchaber, Phys. Rev. Lett. 84, 3017(2000).

[2] M. J. Kim and K. S. Breuer, Phys. Fluids 16 (2004).[3] C. Dombrowski et al., Phys. Rev. Lett. 93, 098103

(2004).[4] I. H. Riedel, K. Kruse, and J. Howard, Science 309

(2005).[5] J. Buhl et al., Science 312 (2006).[6] D. Saintillan and M. J. Shelley, Phys. Rev. Lett. 100,

178103 (2008).[7] A. Sokolov, I. S. Aranson, J. O. Kessler, and R. E. Gold-

stein, Phys. Rev. Lett. 98, 158102 (2007).[8] A. Sokolov and I. S. Aranson, Submitted to Phys. Rev.

Lett. (2009), reduction of viscosity in suspension of swim-ming bacteria.

[9] Y. Hatwalne, S. Ramaswamy, M. Rao, and R. A. Simha,Phys. Rev. Lett. 92, 118101 (2004).

[10] B. M. Haines, I. S. Aranson, L. Berlyand, and D. A.Karpeev, Phys. Biol. 5 (2008).

[11] T. Ishikawa and T. J. Pedley, J. Fluid Mech. 588, 399(2007).

[12] H. C. Berg and D. A. Brown, Nature 239, 500 (1972).[13] L. Turner, W. S. Ryu, and H. C. Berg, J. Bacteriol. 182,

Page 5: A three-dimensional model for the effective viscosity of bacterial suspensions

5

2793 (2000).[14] M. Wu, J. W. Roberts, S. Kim, D. L. Koch, and M. P.

Delisa, Applied and Environmental Microbiology 72,4987 (2006).

[15] S. Kim and S. Karrila, Microhydrodynamics (Dover Pub-lications, New York, 1991).

[16] See EPAPS document for table of functions.[17] E. J. Hinch and L. G. Leal, J. Fluid Mech. 52, 683 (1972).[18] R. A. Simha and S. Ramaswamy, Phys. Rev. Lett. 89,

058101 (2002).

[19] K. Kruse, J. F. Joanny, F. Julicher, J. Prost, and K. Seki-moto, Phys. Rev. Lett. 92, 078101 (2004).

[20] G. B. Jeffery, R. Soc. London Ser. A 102, 161 (1922).[21] L. G. Leal and E. J. Hinch, J. Fluid Mech. 46, 685 (1971).[22] J. P. Hernandez-Ortiz, C. G. Stoltz, and M. D. Graham,

Phys. Rev. Lett. 95, 204501 (2005).[23] V. T. Gyrya, I. S. Aranson, L. V. Berlyand, and D. A.

Karpeev, Bull. Math. Biol. (to appear).