Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Fluid circulation driven by collectively organized metachronal waves in swimming T.

aceiti nematodes
A. C. Quillen,1, ∗ A. Peshkov,1, † Brato Chakrabarti,2, ‡ Nathan
Skerrett,1, § Sonia McGaffigan,1, ¶ and Rebeca Zapiach3, ∗∗
1
Department of Physics and Astronomy, University of Rochester, Rochester, NY 14627, USA
2
Center for Computational Biology, Flatiron Institute, New York, NY 10010, USA
3
Department of Mechanical Engineering, University of Rochester, Rochester, NY 14627, USA
Recent experiments have shown that the nematode T. aceti can assemble into collectively undu-
lating groups at the edge of fluid drops. This coordinated state consists of metachronal waves and
drives fluid circulation inside the drop. We find that the circulation velocity is about 2 mm/s and
nearly half the speed of the metachronal wave. We develop a quasi two-dimensional hydrodynamics
arXiv:2209.03915v1 [physics.flu-dyn] 8 Sep 2022

model using the Stokes flow approximation. The periodic motion of the nematodes constitute our
moving boundary condition that drives the flow. Our model suggests that large amplitude excursions
of the nematodes tails produce the fluid circulation. We discuss the constraints on containers that
would enhance fluid motion, which could be used in the future design of on demand flow generating
systems.

I. INTRODUCTION emulsion containing droplets of a highly concentrated


aqueous suspension of Bacillus subtilis, the bacterial sus-
Turbatrix aceiti (T. aceiti ) is a type of freely swim- pension organized into a single stable circulating vor-
ming nematode that have been shown to collectively self- tex [10], resulting in fluid pumping. Simulations of tiny
organize at fluid-air interface to form traveling waves swimming particles (microswimmers) show that schools
[1, 2]. Individual T. aceiti nematodes, also called vinegar can corral a volume of liquid much larger than the sum
eels, are self-propelled swimmers that continuously con- of the volumes swept along by each individual [11]. Ex-
sume energy. Thus, a dense suspension of vinegar eels is amples of active matter driven pumps include design of
an example of active matter [3]. microfluidic devices that can guide and control motility
Perhaps the most common example of emergent travel- of self-propelled swimmers resulting in directional flows
ing waves is the much studied problem of ciliary carpets. [12–14].
There, hydrodynamic interactions between actively beat-
ing cilia, spontaneously result in the formation of large
scale waves, known as the metachronal waves [4]. Such An advantage of studying vinegar eels, compared to
organized waves are are critical for the motility of cili- many other active systems is their relatively large size.
ated protists (such as Paramecium [5]), mucus clearance Vinegar eels are visible by eye and 1–2 mm in length, ex-
in mammalian airways [6, 7], and for fluid transport in ceeding the size of flagellates, bacteria and many types of
the brain [8]. cells. The soil nematode Caenorhabditis elegans (C. ele-
What makes our model system of vinegar eels unique gans) belongs to the same order of Rhabditida as Turba-
is that unlike cilia which are affixed to a cell membrane, trix aceti and is widely studied as it is both genomically-
these are freely swimming organisms. They fall under a defined and amenable to genetic manipulation. However,
special class of active agents called ‘swarmalators’ that metachronal waves have not been observed in suspensions
can self-propel and synchronize their phase of locomotion of C. elegans [1].
[9]. It is natural to speculate, can the nematode produced
metachronal wave [1, 2] be harnessed to drive coherent
fluid flows? Key to answering this question are quantita- In analogy to ciliary transport [15], in this paper we
tive measurements of the emergent flow, and this is the seek to understand the relation between the collectively
focus of this paper. organized traveling waves and the generated fluid flows.
Emergence of coherent fluid pumping states has been We combine experimental observations of dense suspen-
reported in both experiments and simulations of other sions of T. aceiti nematodes with hydrodynamic model-
wet active matter systems. For example, with an oil ing. The paper is organized as follows: in section II we
present our experimental measurements that reveal cir-
culating flow driven by the collective organization of the

nematodes in a fluid drop. In section III we compute the
alice.quillen@rochester.edu
† flow field using a vertical average for Stokes flow and a
apeshkov@ur.rochester.edu
‡ bchakrabarti@flatironinstitute.org moving periodic boundary condition. A summary and
§ nskerret@u.rochester.edu discussion follows in section V where we speculate about
¶ smcgaffi@u.rochester.edu the possible biomechanics of the body motion of the ne-
∗∗ rzapiach@u.rochester.edu matodes.
2

FIG. 1: (a) An illustration of the experimental setup. A dilute drop of vinegar contains swimming nematodes and
fluorescent microspheres. It is filmed from above with a video camera. Blue light is absorbed by the microspheres
which fluoresce in yellow-green. The microspheres are used to track circulation induced by collective motion in the
vinegar eel population. (b) A color image from the video which has been annotated to show the direction of
circulation. (c) A postprocessed image from the video shows the nematodes at the boundary.

II. EXPERIMENTAL METHODS matodes appear blue because of the LED lighting. To
best show the nematodes, we subtracted a scaled ver-
A. Sample preparation sion of the red frames from the blue ones. The resulting
image is converted to grayscale and shown in Fig. 1(c).
The nematodes have collected on the boundary, and the
We grow the nematodes in a 1:1 solution of distilled
metachronal wave can be seen in the wave-like features on
water and food grade apple cider vinegar at ∼ 5% con-
the outer edge of the drop. Higher magnification images
centration. For the experiment, 14 ml of the grow culture
([2], also discussed below in section IV A) show that the
containing the nematodes is centrifuged for 3 min at 5000
nematode heads are near the boundary and their bodies
rpm. This causes the nematodes to form a dense clump
are oriented at an angle with respect to the boundary
at the bottom of the centrifuge tube. We extract 300
so that their tails extend into the circulating fluid. The
µl of this dense solution and mix it with 10 ml of wa-
bright oval on the top left of the images is a reflection
ter and 10 µl of a solution containing fluorescent yellow
from the lights used to illuminate the drop and can be
polyethylene microspheres and a Tween 80 biocompat-
ignored.
ible surfactant. The spheres have a diameter of 63–75
µm and a density of 1.00 g/cc. The resulting solution is
centrifuged again. Afterwards 100 µl of the concentrated
solution was extracted from the bottom of the centrifuge C. PIV particle tracking
tube and placed onto a bare glass slide. The resulting
droplet that we filmed for analysis has a diameter of ∼ 1 The fluorescent microspheres were tracked using the
cm and a height h ≈ 1 mm. software package trackpy [16], which implements in
Python the Crocker-Grier algorithm for finding and
tracking single-particle trajectories [17].
B. Imaging In Fig. 2(a) we show tracks traced by fluorescent mi-
crospheres during 1 second of video on top of one of the
The experimental set up for imaging is shown in Fig. 1a video frames. In Fig. 2(b) we show velocity vectors com-
and a video taken with this setup is included as supple- puted from these tracks. In red and with thicker arrows,
mental video A. The fluorescent microsphere markers are we show average velocities computed by fitting a line to
used to measure the flow induced by the collective mo- the trajectory of each microsphere in 1 second of video.
tions of the nematodes. We lit the slide with bright blue This averages over several oscillations of the metachronal
LEDs, causing the microspheres to fluoresce in yellow- wave to better show fluid circulation. Circulation of the
green. The drop was filmed in color at 60 frames per microspheres can also be seen directly from viewing sup-
second and from above with a Blackmagic Pocket 4K plemental video A.
digital camera. Using a polar coordinate system with origin at the cen-
Image frames at a single time from supplemental video ter of the drop, we measure the tangential component of
A are shown in Fig. 1(b) and (c). In Fig. 1(b), we the mean velocity vθ for each tracked microsphere. In
show one of the video frames in color. The micro- Fig. 3(a), the microsphere mean tangential velocities are
spheres appear green as they fluorescence and the ne- shown as a function of distance d from the outer drop
3

FIG. 2: (a) Tracks of individual fluorescent microspheres from supplemental video A. The spheres were tracked
using 60 video frames during 1 second of video. The tracks are shown on top of the first video frame from the video.
(b) Velocities of the fluorescent spheres averaged over 1 second of video are shown with red arrows on top of the first
video frame in the sequence. The thinner brown arrows show instantaneous velocities of the same spheres computed
from positions in consecutive video frames.

boundary. As indicated by the dotted line on the fig- TABLE I: Measurements


ure, the circulation velocity decays as a function of dis-
tance from the boundary and goes to zero at the center Radius of drop Rd 5.25 mm
of the drop. The black dotted line in on Fig. 3(a) shows Volume of drop Vdrop 100 µl
the curve vθ (d) = uc e−(d−d0 )/hs with parameters uc , d0 Metachronal frequency fM W 5.45 ± 0.16 Hz
and decay length hs . The offset d0 is used to describe a Metachronal wavelength λM W 0.93 ± 0.09 mm
peak distance where circulation is highest and uc gives
Metachronal wave speed VM W 5.1 ± 0.5 mm/s
the peak circulation velocity. Such an exponential decay
results from hydrodynamic screening caused flow near Velocity of circulating flow uc 2 ± 1 mm/s
boundaries [18–20]. In the following section, we will show Exponential decay length of circulation hs 0.7 ± 0.2 mm
that the decay length hs is consisted with hydrodynamic
screening in a shallow drop.
In Fig. 3(b,c) we plot the instantaneous radial and time and then sum the pixel values in the entire prod-
tangential velocity components as a function of distance uct image. The peaks in the sum occur at multiples
from the drop edge. The exponential curve from Fig. 3(a) of the metachronal wave period. We similarly measure
is overlayed on Fig. 3(b) for comparison and illustrates the metachronal wavelength λM W , by rotating an indi-
that the instantaneous velocity can be larger than the vidual frame about the center of the drop, multiplying
time-averaged circulation speed. The velocities we mea- it by the original frame and summing over the prod-
sure for the microspheres can exceed the forward swim uct image. Peaks in the sum occur at rotations that
speed of the nematodes that are involved in the wave. are the metachronal wavelength divided by the drop ra-
The nematodes forming the wave advance along the bor- dius. Errors in these quantities are estimated from the
der at a much slower speed, ∼ 0.1, mm/s than both the strength and widths of the peaks in these sums. These
metachronal wave and the circulation velocity [1, 2], high- measurements are summarized in Table I. We also list a
lighting the emergence of large scale coherent transport range of values for the peak circulation speed uc and de-
through collective self-organization. cay length hs consistent with mean tangential velocities
shown Fig. 3(a).
The wavelength of the metachronal wave is λM W ∼
1 mm, similar to that of our previous measurements
D. Metachronal wave measurements [2]. The metachronal wave frequency is fM W ≈ 5.45
Hz. Together these give a metachronal wave velocity
We characterize the kinematics of the emerging vM W = λM W fM W ≈ 5.1 mm/s. Near the collective
metachronal wave. To this end, we measure the wave wave, the microspheres have a mean circulation velocity
frequency fM W with a cross correlation technique, as of about uc ∼ 2 mm/s. The ratio of microsphere mean
described by Quillen et al. [2]. We compute the prod- circulation velocity to metachronal wave speed is about
uct of two image frames separated by an interval of uc /vM W ∼ 2/5.
4

Stokes equation as follows:


∇ · U = 0, (1)
1
− ∇p + ν∇2 U = 0, (2)
ρ
where p is the pressure and ν is the kinematic viscosity
of the fluid. The velocity field U is a function of (xk , z),
where xk = {x, y} spans the plane on which the drop re-
sides and the height of the drop is given as 0 ≤ z ≤ h(xk )
with h(xk ) being the height of the drop. The character-
istic drop thickness h ∼ 1 mm is much smaller compared
to the typical drop radius of R ∼ 5 mm. This natural
scale seperation allows us to invoke the lubrication ap-
proximation for the problem [21]. Scaling xk ∼ R and
z ∼ h we find W ∼ {U, V }h/R. In the limit h/R  1 the
vertical velocity W = 0 to the leading order and ∂z p = 0.
This smallness of the aspect ratio h/R can now be ex-
ploited to separate the wall normal dependance (i.e., in
the direction z) of all the velocity fields from its in-plane
averaged value as is done in classical Hele-Shaw problems.
For this prupose, we introduce the following ansatz for
the velocity field
U = u(xk )f (z), (3)
where f (z) is a single scalar function which describes the
velocity profile over the height of the cell and u(xk ) =
ux̂ + vẑ is a 2D depth averaged velocity field defined as
FIG. 3: Microsphere velocities. (a) Average tangential
1 h
Z
velocity vθ of tracked microspheres as a function of u(xk ) = U dz. (4)
distance from the outer drop edge. The dotted line h 0
shows an exponentially decaying function with function The ansatz of equation 3 is reasonable because the verti-
and parameters written on the top right of the panel. cal velocity vanishes to the leading order of the problem.
(b) Instantaneous tangential velocity of microspheres as The precise nature of f (z) depends on the boundary con-
a function of distance from the drop outer edge. The ditions of the problem. For flow in the droplet with a free
dotted line shows the same exponentially decaying interface and no-slip bottom surface we take
function as in panel (a). (c) Microsphere instantaneous  
radial velocity component as a function of distance from 3  z 2
f (z) = 1− 1− . (5)
the drop outer edge. 2 h
On the other hand, for flow between two flat plates
with no-slip boundary
 conditions we could use f (z) =
III. HYDRODYNAMIC THEORY FOR
6 z/h − z 2 /h2 . The ansatz of equation 3 along with
COHERENT FLOWS the definition for u gives
u
∂zz U = −αz 2 , (6)
h
To understand the underlying mechanics of the coher-
ent transport driven by the metachronal waves we build where αz = 3 for a free interface and αz = 12 for flow
a quasi 2D model hydrodynamic model for the fluid flow. between two no-slip boundaries. In our above formula-
tion we have ignored curvature effects of the drop height
near the boundaries. This amounts to assuming that the
surface tension effects and the forces from the nematodes
in determining the drop shape is negligible [1].
We now average Eq. (1) and (2) over depth to obtain
A. Hele-Shaw approximation for the quasi 2D flow
the quasi 2D approximation for the evolution of the ve-
locity field:
We consider a three-dimensional incompressible flow ∇k · u = 0, (7)
field in the drop U = {U, V, W } and restrict ourselves in
the limit of low Reynolds number. In this limit, the evo- 1 h αz i
− ∇k p + ν ∇2k − 2 u = 0, (8)
lution of the flow-field is governed by the incompressible ρ h
5

where ∇k = ∂x x̂+∂z ẑ is the 2D gradient and ∇2k is the as- periodic in x, with period λM W = 2π/kM W , and decays
sociated 2D Laplacian. The above set of equation differs at larger y has the following stream function
from the two-dimensional incompressible Stokes equa- ∞
"
tion by the additional dissipative term −αz /h2 which −y/hs
X
ψ(x, y) =a0 e + (12)
accounts for the average friction force between the sur-
j=1
face and the droplet interface. By omiting the term ∇2k u
we would obtain the so-called Darcy approximation of (acj cos(jkM W x)+asj sin(jkM W x))e−βj jkM W y
#
the Navier-Stokes equation, which is often used to model
flows in porous media and viscous fingering in Hele-Shaw (bcj cos(jkM W x)+bsj sin(jkM W x))e−jkM W y ,
cells and describes potential flow. In our case, the flow
need not be in general a potential flow [22, 23]. The
with coefficients a0 , acj , asj , bcj , bsj where
above set of equations are often termed as the Brinkman
correction to Hele-Shaw flows [24]. It is worth point- r
αz
ing out that, the Green’s function associated with the βj ≡ 1 + . (13)
(jkM W h)2
full Stokes equation is non-trivial [25] and involves the
use of the method of images to calculate the flow field The characteristic length scale for the exponential decay
[18, 25]. One advantage of our 2D Brinkman approxi- associated with the constant circulating term with coef-
mation is that it circumvents this challenge and allows ficient a0 is
us to compute tractable solutions in a spatially periodic
domain, as we discuss next. √
hs ≡ h/ αz . (14)
Since the depth-averaged velocity u acts like a two di-
mensional incompressible fluid, we can describe the two- The terms with coefficients bcj , bsj have zero vorticity as
dimensional flow with a stream function ψ(xk ), where they satisfy Laplace’s equation ∇2k ψ = 0.
The above solution also makes it evident that the term
(u, v) = (∂y ψ, −∂x ψ) . (9) proportional to αz in equation 8 gives rise to hydro-
dynamic screening with an exponential decay length of
The vorticity ω is related to the Laplacian of the stream √
hs = h/ αz typical in Brinkman correction [26]. With
function a drop thickness of about h ∼ 1 mm and αz = 3, consis-
tent with a fixed, no-slip lower boundary and stress free
ω = ∂x v − ∂y u = −∂xx ψ − ∂yy ψ = −∇2k ψ. (10) upper interface, the decay length hs ∼ 0.6 mm. This is
consistent with the decay length measured in the induced
Upon taking the curl of Eqn. 8 we obtain the evolution circulation in section II.
of the vorticity as
αz
∇2k ω = ω. (11) B. The metachronal wave boundary condition
h2
With negative αz , this is known as the homogeneous The nematodes engaged in the metachronal wave are
Helmholtz equation, and with positive αz , it is known densely packed. Their heads are near the outer edge of
as the homogeneous screened Poisson equation. the boundary and the tails touch moving fluid. This sug-
gests that the coherent circulating flow is driven by the
motion of the tails of the nematodes involved in the wave.
To incorporate this in our model, we develop a kinematic
description of the tail motion. We assume that the tails
act like a moving, continuous no-slip boundary for the
fluid flow. The trajectory of each point on the bound-
ary is described by a periodic function associated with
FIG. 4: Coordinate directions. The boundary is on the the oscillatory motions in the wave. The metachronal
bottom and the direction to the center of the drop is wave gives a delay between the trajectories of neighbor-
upward. The metachronal wave travels to the right. We ing boundary points.
ignore the curvature of the drop boundary. Neglecting curvature of the drop edge, the boundary
is spatially periodic in x with wavelength λM W and with
We set our x direction along the azimuthal direction of metachronal wave traveling in the positive x direction
the drop and the y direction is aligned with radius from with velocity VM W . We use the coordinate s ∈ [0, λM W )
the drop center and is increasing with distance from the to describe positions along the boundary. The trajectory
drop edge (see Fig. 4). Since the flow-field is driven by of a material point at s = 0 is described by a displace-
the motions of the nematode tails we expect the solutions ment function δ 0 (t) which is temporally periodic with
to be periodic along x over a metachronal wavelength. frequency fM W . Due to the propagating traveling wave,
For positive αz , a general solution of equation 11 that is a point on the boundary that is displaced horizontally
6

from the reference point at s = 0 undergoes the same tra- TABLE II: Parameters for flow models
jectory but delayed in time. With xb = (xb , yb ), points
on the boundary evolve as Metachronal wavelength λM W 1 mm
Metachronal frequency fM W 5 Hz
xb (s, t) = δ 0 (t − s/VM W ) + sx̂ + xm . (15)
Screening parameter αz 3
In the above description, the boundary is a 1 dimensional Drop thickness h 1 mm
space curve that is described by a displacement vector Screening length hs = √h 0.6 mm
αz
function δ 0 (t − s/VM W ). This yields a wave traveling
with the metachronal wave velocity VM W . The constant
xm sets the position of the boundary at s = t = 0. The The key parameters for our boundary model are listed
velocity points on the boundary: in Table II. We chose αz = 3 corresponding to a fixed
lower boundary and a stress free upper boundary as dis-
d
Vb (s, t) = [δ 0 (t − s/VM W )] . (16) cussed in section III A. We take drop thickness h = 1
dt mm which yields the screening length of the solution as
The no-slip boundary condition for the fluid velocity im- hs = 0.6 mm consistent with the our experimental mea-
plies: surements.

Vb (s, t) = u(xb (s, t)). (17)


D. Role of different boundary motions on fluid
circulation
C. Finding a flow field consistent with the velocity
on the boundary
Motivated by our prior work on a phase oscillator
model for the collective motion [2], we consider the ve-
Once a set of positions on the boundary and the ve-
locity field that would arise from a boundary where each
locities of these positions are specified, we try to ob-
point has a small amplitude of oscillation compared to
tain a two-dimensional fluid flow field consistent with
the metachronal wavelength λM W . We describe the dis-
the boundary condition. For this we use a minimiza-
placement vector of points on the boundary with the
tion algorithm to find the coefficients in solution of the
function
stream-function in Eq. (12), as illustrated below.
Using Eq. (17) for the boundary condition and inte-
δ 0 (t) =Ab cos(ωM W t)ŷ+ (20)
grating over the boundary, we construct a non-negative
function [Bb cos(ωM W t) − Cb sin(ωM W t)] x̂
Z Lb
1 with ωM W = 2πfM W and three constant coefficients
g(δ 0 , ψ) = |Vb (s) − u(s)|2 ds, (18)
Lb 0 Ab , Bb , Cb . We include only three terms because we can
adjust the phase of the wave to remove a term that is
where Lb is the length of the boundary corresponding to proportional to sin(ωM W t)ŷ. Positions on the boundary
a single wavelength of the metachronal wave. The above xb (s, t) are then generated from δ 0 (t) using Eq. (15). To
function is zero for a velocity field u(xk ) derived from compute the circulation speed we take ym to be the dis-
a stream function ψ(xk ) using equation 9 that is con- tance between the mean y component of δ 0 (t) and the
sistent with the velocity boundary condition described drop edge.
by the displacement function δ 0 (t). We use a multivari- Depending on the choice of the constants in Eq. (20),
ate minimization algorithm to minimize this function for we obtain qualitatively different models for the bound-
different values of the coefficients in Eq. (12). For the ary motions. We explore two models, denoted Model A
present problem we retained j = 20 Fourier modes for and Model B, with properties summarized in Table III.
the stream-function which results in 81 free parameters. Model A has Cb = 0 and each particle on the boundary
To carry out the minimization we used Nelder-Mead Sim- moves back and forth along a line segment. This model
plex method available through python’s scipy.optimize is similar to the phase oscillator model [2] developed pre-
package. The quality of the result is measured by com- viously where the points on the nematode bodies moved
puting the standard deviation σv of the difference be- back and forth, but at an angle with respect to the outer
tween flow velocity and boundary velocity integrated on drop edge. Model B has Cb 6= 0 and each particle on the
the boundary. This is equivalent
p to the square root of the boundary particles has a loop shaped trajectory. Fig-
minimization function, σv = g(δ 0 , ψ). Once the fitting ure 5(a, c) display the boundary position at a given time
is done, the coefficient a0 of the stream function allows instant for the two different boundary models with the
us to compute and characterize the circulation speed in arrows indicating the instantaneous velocity of the mate-
the model at y = ym as rial points. The blue line on the left shows the trajectory
a0 −ym /hs of a single point on the boundary over one oscillation
uc (ym ) = − e . (19) period.
hs
7

TABLE III: Oscillating Boundary Models of the nematodes involved in the metachronal wave.

Model A Model B Tail Tr.


Amplitude (mm) Ab 0.10 0.10 -
Amplitude (mm) Bb -0.05 -0.07 -
Amplitude (mm) Cb 0.0 0.10 -
Circulation (mm/s) uc (ym ) -0.6 1.9 2.6 A. Tail trajectories
Visc. dissipation per λM W (pW) 70 86 830
Power in circulation 1% 9% 2%
To make accurate measurements of the eel tails we use
the 10x magnification and high speed videography (1057
fps) described previously in [2] (see supporting video B).
The associated velocity fields for these two boundary
For a few nematode tails, we marked their positions on
models are depicted in Fig. 5(b,d) and their circulation
every frame using the software package ImageJ. The re-
speeds are listed in Table III. Interestingly, Model A, with
sulting tail trajectories are shown in Fig. 6 and they are
back and forth oscillation of boundary material points,
also marked on supplemental video B. In Fig. 6(a), tail tip
yields a small circulation speed uc and predicts the in-
positions are plotted for about an oscillation period with
correct direction for circulation with uc < 0. In contrast,
thick colored lines on top of a gray scale image from sup-
model B provides a circulation speed that is consistent
plemental video B. The tail tip velocities are shown with
with that we observed experimentally. The computed
small arrows. Thin colored segments highlights tangent
solution captures features of the experimental flow field,
directions for the tail. In Fig. 6(b) we display smoothed
including the back and forth motions associated with the
versions of the same six tail tip trajectories. The tra-
metachronal wave. Both Model A and B flow fields have
jectories are shifted horizontally so that their mean x
rms velocity difference of σv = 0.3 mm/s between flow
positions are the same. The similarity between the six
and boundary (the square root of the minimization func-
trajectories gives us confidence that the measured tra-
tion in equation 18).
jectories are associated with the collective motion, are
similar and periodic.
E. Energetics of the flow Based on our previous study on phase oscillator mod-
els for the formation of the metachronal wave [2], we
It is natural to ask how much energy is required to had expected the nematode tails to lie near their bodies
maintain these streaming flows or what fraction of the throughout the metachronal wave. The phase oscillator
energy injected by the nematode body motions is used model assumed that material points on nematode bod-
by the coherent flows. To answer this we estimate the ies are oscillating back and forth with an amplitude of
viscous dissipation from the velocity fields computed in about 0.1 mm and do not involve large amplitude ellip-
Fig. 5(b,d) (see appendix A for details). Using viscos- tical motions. However, as shown in Fig. 6, and seen in
ity µ = 0.9 mPa s we find that the estimated dissipa- supplemental video B, the tracked tail tip trajectories are
tion rate is about 80 pW per wavelength for both models not ellipses but figure 8 shapes. Excursions away from
of the boundary motion (see Table III). With about 20 the drop edge (in the y direction on Fig. 6) are about 0.8
vinegar eels per metachronal wavelength this corresponds mm which is almost as large as the metachronal wave-
to 4 pW per vinegar eel. Remarkably, we find that for length of 1 mm, and exceeds the back and forth motions
model B, only about 9% of this power goes into main- of the rest of the nematode bodies and their heads. The
taining fluid circulation, corresponding to the constant tail tips extend well outside the region near the bound-
or a0 term in the stream function given in Eq. (12) (see ary where the nematodes are densely packed and their
Table III). velocities can be as high as ∼ 10 mm/s.
We found that some tails could not be tracked for a
whole period because they went out of focus during part
IV. ROLE OF TAIL MOTIONS IN DRIVING of the video. We note that supplement video B is of a
FLUID FLOW shallow drop, with a depth about half that of the drop
shown in supplemental video A. The shallower depth fa-
The computed flow models for the oscillating bound- cilitates viewing individual nematodes, as they are less
aries suggest that elliptical trajectories for points on a likely to go in and out of focus, an issue at higher magnifi-
boundary are necessary to drive the level of circulation cation. However, as discussed in section III A, a shallower
in the fluid that we saw in supplement video A. As the drop depth reduces the screening length hs . We have no
nematode heads are trapped at the drop edge, the mo- reason to suspect that the depth affects the motions of ne-
tion is primarily imparted by the nematode tails moving matodes participating in the metachronal wave, though
within the circulating fluid. To better understand how the drop shape can influence whether metachronal wave
the circulation is driven we examine the behavior of tails can form [1].
8

FIG. 5: (a) In blue we show the position of a single point on the boundary during a full oscillation period for a
sinusoidal oscillating boundary denoted Model A and with parameters listed in Table III. In red we show the
boundary at single moment in time. Velocity vectors are shown with black arrows. (b) A velocity field for Model A
that was found by minimizing the function of Eqn. 18 over the boundary shown in (a). The stream function is
shown as a red image and with black contours. Velocity vectors in the fluid are shown with green arrows. The
velocity vectors on the boundary are shown with blue arrows. (c) Similar to a) but for sinusoidal oscillating
boundary model B. (d) The associated velocity field for model B.

B. Flow field computed from tail trajectories (computed from Eq. (19)) and viscous dissipation rate
for this flow model are listed in the rightmost column
We proceed to compute the fluid flow resulting from of Table III. The quality of fit, estimated from equation
the observed tail trajectories. We use a single tail trajec- 18, gives rms σv ∼ 2 mm/s which is poorer than for the
tory to generate a boundary from a sequence of tails un- flows associated with Model A and B. The higher value
dergoing the metachronal wave using equation Eq. (15). is probably due to the higher velocities on the boundary
On Fig. 7(a) we show the trajectory of the rightmost and the unphysical overlap region.
tail of Fig. 6(a) with a dotted line. The associated wave The circulation velocity uc = 2.6 mm/s, for the flow
boundary is shown with a solid thick line. The generated shown in Fig. 7(b), is sufficiently high to be consistent
boundary shows that the tail would overlap with neigh- with the experimental values reported in Fig. 3(a). In-
boring nematodes at different times which is indeed con- terestingly we find that the power required to drive the
firmed from the examination of the supplemental video flow is about 40 pW per nematode in the wave with
B that highlights the large tail tip excursions. To gener- only 2% of this power being used for fluid circulation.
ate a boundary, we chose the rightmost trajectory, shown This estimated power is about 10 times higher than the
in pink in Fig. 6(a), because it has the least oscillatory 3 pW power estimated power for propulsion of C. ele-
motion in the x direction and provides the least over- gans [27], suggesting that vinegar eels expend more en-
lap. Reduced overlap in the boundary condition is desir- ergy while they are in the metachronal wave than they
able since it is unphysical to have different fluid veloci- would while freely swimming. We speculate that this
ties at a single point in our 2D fluid model. Using the could be because: (i) tail trajectories for nematodes in
boundary shown in Fig. 7(a), we generate an associated the metachronal (extending a maximum distance about
two-dimensional fluid flow with the method described in 0.8 mm peak to peak) are larger than those of a freely
section III C. The flow is shown Fig. 7(b). Circulation swimming eel (which are about 0.3 mm peak to peak, see
9

FIG. 6: Tail trajectories as seen in a higher magnification high speed video. (a) Six nematode tail tips are tracked
and plotted with a thick line. Each tail tip is plotted with a different color. Arrows show the velocity of the tail tip.
Thin colored segments show tangent directions for the tail at a subset of times with each segment ending at a tail
tip location. The underlying gray scale image shows a frame from the high speed video (supplemental video B). A
scale bar is shown in red on the upper left. The yellow arrow shows that the metachronal wave travels to the right.
The heads of vinegar eels involved in the metachronal wave are located at the bottom of the image, near the edge of
the drop. (b) Smoothed versions of the same six tracked tail trajectories. Horizontal x positions have been shifted so
that the mean position is near zero. Vertical y positions give the distance from the drop edge.

the right hand side of Fig. 1 by Quillen et al. [2]); (ii) flow eel is a slender object, we model its hydrodynamics using
velocities may be overestimated because of overlaps in the local slender body theory (SBT) [29, 30], which relates
tail trajectories and (iii) due to confinement, nematodes the viscous forces fv (s) per unit length to the centerline
involved in the metachronal wave would expend part of velocity as
their energy pushing against each other. The power re-
quired to both drive both the metachronal wave and the 8πµ∂t x(s, t) = −L[fv ]. (22)
associated circulation per nematode could be up to an or- Here, µ is the fluid viscosity, and L is the local mobil-
der of magnitude higher than that expended while freely ity operator that accounts for drag anisotropy along the
swimming. Improved hydrodynamic modeling would be body and is given by
needed to better estimate the power required to drive the  
flow and modeling taking into account steric interactions 1 1
would be needed to estimate the power dissipated within L[fv ](s) = n̂(s)n̂(s) + t̂(s)t̂(s) · fv (s), (23)
ξ⊥ ξk
the nematode bodies.
where ξ⊥ = (2 − c)−1 and ξk = −(2c)−1 are resistance
coefficients in the normal and tangential directions, and
C. Mechanical model for a single tail motion c = ln(2 e) < 0. In the above expression, {t̂(s), n̂(s)}
are the tangent and normal vectors along the centerline,
Our analysis of the nematode tail trajectories in the respectively. The elastic forces are related to the viscous
previous section revealed a characteristic figure of 8 shape forces following the force and moment balance equations;
for the tail tips. We seek to develop a mechanical model fv + ∂s F = 0, (24)
that sheds light on such an emergent pattern. We build
a simplified model for the backbone of an isolated ne- Ms + ∂s x × F = 0. (25)
matode. We model the centerline dynamics of an indi- Here F is the elastic contact forces of an inextensible
vidual nematode as a planar, inextensible Euler elastica rod, and M is the bending moment of the filament given
with bending rigidity B. The centerline is parameterized by M(s) = −B(κ(s, t) − κ0 (s, t))xs × xss (here xs refers
by arc-length s and identified by a Lagrangian marker to the derivative ∂s x). Scaling length by L and time
x(s, t). Following [28] the actuation in the nematode that with the relaxation time τ = 8πµL4 /B of an elastic fila-
produces periodic traveling undulation is modeled using ment, the nematode body shapes are governed by three
a preferred time-dependent curvature given by: dimensionless numbers: (i) A/L that sets the amplitude
As of the target curvature, (ii) kw L that sets the dimension-
κ0 (s, t) = sin (kw s − 2πΩ0 t) . (21) less wave number and (iii) W r = Ω0 τ , the worm number
L
[28] that compares the frequency of wave propagation to
Here kw is the wave number of the target curvature, A the elastic relaxation time of the nematode body. Using
is the amplitude and Ω0 is the frequency. We consider L = 2 mm, B = 10−14 N m2 [31], and f0 = 2πΩ0 = 5 Hz
the nematode head at s = 0 to be fixed in space. as we obtain W r ≈ 40. We solve the coupled set of PDEs
the heads of nematodes participating in the metachronal numerically as outlined in [30, 32].
wave do not move very far (see Fig. 5 [2]). The tail at The shapes of the nematode body and the tail trajec-
s = L is both force and moment free. Since the vinegar tories at x(s = L) are outlined in Fig. 8 and reveal the
10

(a) Modeled shape of the eel (b) Path traced by the tail tip

y/L
y/L
<latexit sha1_base64="9tqJ/0lo79ARo4yPqIS3nERDP5Y=">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</latexit>
<latexit sha1_base64="9tqJ/0lo79ARo4yPqIS3nERDP5Y=">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</latexit>

x/L
<latexit sha1_base64="pi08hRQzxryptjZZkOugfhssweo=">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</latexit>
x/L
<latexit sha1_base64="pi08hRQzxryptjZZkOugfhssweo=">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</latexit>

FIG. 8: (a) Shape of an isolated nematode as computed


from the preferred-curvature model. (b) The
characteristic figure of 8 shape of the tail tips emerging
from the inextensibility of the nematode body.

body, as required for locomotion (e.g., [33]), we find that


a natural outcome is a figure of 8 shaped tail tip tra-
jectory. However, a more complex model than explored
here would be needed to better match the observed body
shapes.

V. SUMMARY AND DISCUSSION


FIG. 7: (a) The trajectory of the rightmost tail in
Fig. 6 is shown as a dotted pink line on the left, after Using fluorescent microspheres we have measured the
smoothing. Points on the trajectory are shifted and circulation velocity of flow in a 1 cm diameter and 1 mm
delayed using equation 15 to give the solid pink line deep drop that is driven by collective motion in a con-
which shows the boundary position at a single moment centration of swimming T. aceiti nematodes, which are
during propagation of the metachronal wave. Arrows commonly called vinegar eels. The mean circulation ve-
show velocity vectors of points on the boundary. (b) locity is about 2 mm/s, and located within a mm of the
The flow field associated with a boundary that is outer edge of the drop. We find that the ratio of circu-
approximated by the tail tip trajectories. The boundary lating flow velocity to metachronal wave speed is about
condition is given by points and velocities on the solid 2/5 and decays rapidly as a function of distance from the
line in (a). The color map shows the stream function outer drop edge.
and its contours are plotted with black lines. The thick Perturbative models [34–36] for causing steady flow
orange line shows the boundary. Blue vectors show by wave-like ripples on a surface have been developed
velocities on the boundary. Green vectors show fluid for three-dimensional flows in the Stokes flow or low
velocities. The y axis shows y − ym where ym is Reynolds number limit. These predict that the driven
approximately the mean y position of a boundary point. flow depends on the square of the amplitude of surface
perturbations in units of the wavelength times the sur-
face wave speed. This scaling suggests that the ratio of
flow speed to wave speed should be low. In this con-
natural emergence of the figure of 8 trajectory shape at text, the ratio of 2/5 for the ratio of circulation speed to
the tail tips. Fig. 8 also shows that the amplitude of metachronal wave speed that we measured in the dilute
motion reaches a maximum near the tail tip, and this is vinegar drop is unexpectedly high.
consistent with observations of both freely swimming ne- Unlike the planar sheet model [34, 35] where the fluid
matode and those participating in the metachronal wave. lies in an infinite 3 dimensional half-space, our system is
However, these simulations do not give tail amplitudes as shallow, with a fixed lower boundary (the slide that our
large as we see in either setting. The body shapes Fig. 8 drop lies on) and free upper boundary (the drop surface).
have decreasing wavelength near the tail tip, opposite We take into account the shallow drop depth by integrat-
to what we observe, suggesting that the tails are stiffer ing the velocity vertically as a function of depth. Depth
or remain straighter than the rest of the nematode bod- averaged velocities give a two-dimensional hydrodynamic
ies. When body curvature propagates down a nematode model. Neglecting variation of the vertical velocity pro-
11

files and drop thickness on horizontal position, the fluid remain to be explored.
obeys the homogeneous screened Poisson equation, typi- A flow model generated with a boundary generated
cal of Hele-Shaw flow [24]. from tail motions exhibits sufficient circulation to be con-
Using the trajectory of a single nematode tail, we con- sistent with the circulation we observed experimentally.
struct a boundary condition for the metachronal wave by The flow model suggests that the motions of the vinegar
delaying the tail-tip trajectory at neighboring positions. eel tails are important for driving fluid circulation.
This model gives us the velocity and location of particles The high ratio of circulation speed to metachronal
along a moving 1 dimensional boundary. With a general wave speed we measure suggests that systems of ne-
solution to the 2D homogeneous screened Poisson equa- matodes could be engineered to drive flows. What
tion that decays at large distances from the boundary and type of container would best give circulation? Border-
using an optimization routine, we find the stream func- taxis of the nematodes should be facilitated, otherwise
tion that minimizes the difference between fluid velocity the nematodes will not enter a collective state giving a
and particle velocities on the boundary. The screening metachronal wave. The container edge where the eels
in the screened Poisson equation is sensitive to the drop would be corralled could be shallow (less than 1 mm
depth, with shallower drops having circulating flows that thick) or beveled, similar to the contact angle caused by
decay more rapidly as a function of distance from the surface tension in a drop (with contact angle below 70◦ ;
boundary than circulating flows in deeper drops. We find see Peshkov et al. [1]). The top surface could be covered,
that the circulation velocity is sensitive to the boundary rather than open, though this would affect the screening
particle trajectory shape, with ellipse trajectories more length scale. We assumed an open surface giving αz ∼ 3

effective than linear trajectories at driving circulation. and a screening exponential length hs = h/ αz (as in
equation 14) where h is the depth of the container. With
We had expected to find that the tails of nematodes
a fixed upper boundary (for an enclosed container) we
participating the metachronal wave undergo ellipse tra-
would expect αz ∼ 12. A container twice as deep would
jectories. However, upon examination, we found that the
be needed to give the same screening length in a closed
tails of the nematodes undergoing collective motion have
container as one with an open surface. A beveled edge
figure of 8 trajectories and undergo large 0.8 mm excur-
might be the best compromise as it might corral the eels,
sions away from the drop edge. These are about twice
facilitating collective motion, while simultaneously allow-
as large as tail motions exhibited by a freely swimming
ing a larger screening exponential decay length for the
nematode or most of nematode body that is participat-
circulating flow. Large amplitude motions in the nema-
ing in the metachronal wave. To probe the mechanical
tode tails would help drive circulation, so the distance
origins of such motions we constructed a simple preferred
between outer and inner container edges should be at
curvature model for an isolated nematode. This model
least a few mm.
captures the characteristic figure of 8 shape of the tail
and shows that this shape is naturally caused by a cur- The model we used for fluid flow is two-dimensional,
vature wave propagating down an elastic body. While ignores variations in fluid depth, and is only appropriate
this model captures the characteristic figure of 8 shape in the Stokes or low Reynolds number limit. With a
of the tail and an increase in amplitude toward the tail flow velocity of 1 mm/s, a length scale of 1 mm and
tip, the simulated results are not in great agreement for dynamic viscosity of water ν ∼ 10−6 m2 s−1 = 1 mm2
an isolated nematode. This can be possibly due to subtle s−1 , the Reynolds number is about 1, so inertia could be
structural features associated within the nematode such important in the velocity field.
as varying cross-sectional width along its backbone and We observed nematode bodies overlapping other ne-
variations in how such structural features alter the inter- matodes in our high magnification video and there are
nal force-generation. The associated biomechanics will be numerous nematodes that are not engaged in the collec-
considered in the future. Understanding collective effects tive motion. Our 2D flow and boundary model neglects
poses another set of challenges in modeling. For example, motions of overlapping nematodes and those that do not
in contrast with the experimental measurements for ne- participate in the collective motion. Extending the fi-
matodes participating in the metachronal wave, our sim- delity and capability of the hydrodynamic modeling is
ple mechanical model does not predict large amplitude challenging but would improve understanding of the rela-
oscillations of the tail tip. Since hydrodynamic interac- tion between collective motion in a dense concentrations
tions are screened, we believe that such emergent dynam- of oscillating swimming organisms and associated flows
ics of the nematode tails could be associated with local resulting from their collective motion.
steric interactions. Our prior work suggested that steric
interactions were required for metachronal wave forma-
tion in the vinegar eel system and caused the waves trav-
eling down the nematode bodies to deviate from a pure
sinusoidal function [2]. Recently, it has been proposed
that steric interactions alone can lead to the emergence
of collective behaviors in arrays of cilia and alter their iso-
lated waveforms [37]. The potential role of such effects
12

ACKNOWLEDGMENTS helpful discussions and suggestions. We thank the Flat-


iron Institute, Center for Computational Biology for their
This material is based upon work supported in part by welcome and hospitality in Spring 2022. We thank Maya
National Science Foundation Grant No. PHY-1757062, Parada for many interesting discussions on the subject of
and National Science Foundation Grant No. DMR- designing and manufacturing containers for the vinegar
1809318. eels.
We thank Michael J. Shelley and Adam Lamson for

[1] A. Peshkov, S. McGaffigan, and A. C. Quillen, Soft Mat- mans, and T. N. Shendruk, Journal of Fluid Mechanics
ter 18, 1174 (2022). 806, 35 (2016).
[2] A. C. Quillen, A. Peshkov, E. Wright, and S. McGaffigan, [26] H. C. Brinkman, Flow, Turbulence and Combustion 1,
Phys. Rev. E 104, 014412 (2021). 27 (1949).
[3] M. C. Marchetti, J. F. Joanny, S. Ramaswamy, T. B. [27] J. Sznitman, X. Shen, R. Sznitman, and P. E. Arratia,
Liverpool, J. Prost, M. Rao, and R. A. Simha, Reviews Physics of Fluids 22, 121901 (2010).
of Modern Physics 85, 1143 (2013). [28] T. Majmudar, E. E. Keaveny, J. Zhang, and M. J. Shel-
[4] B. Chakrabarti, S. Furthauer, and M. J. Shelley, Pro- ley, Journal of the Royal Society Interface 9, 1809 (2012).
ceedings of the National Academy of Sciences 119, [29] J. B. Keller and S. I. Rubinow, Journal of Fluid Mechan-
e2113539119 (2022). ics 75, 705 (1976).
[5] S. L. Tamm, The Journal of Cell Biology 55, 250 (1972). [30] A.-K. Tornberg and M. J. Shelley, Journal of Computa-
[6] M. A. Sleigh, J. R. Blake, and N. Liron, American Review tional Physics 196, 8 (2004).
of Respiratory Disease 137, 726 (1988). [31] M. Backholm, W. S. Ryu, and K. Dalnoki-Veress, Pro-
[7] B. A. Afzelius, Journal of Pathology 204, 470 (2004). ceedings of the National Academy of Sciences, 110, 4528
[8] R. Faubel, C. Westendorf, E. Bodenschatz, and (2013).
G. Eichele, Science 353, 176 (2016). [32] B. Chakrabarti and D. Saintillan, Physical Review Fluids
[9] K. P. O’Keeffe, H. Hong, and S. H. Strogatz, Nature 4, 043102 (2019).
Communications 8, 1504 (2017), URL https://doi. [33] Q. Wen, M. D. Po, E. Hulme, S. Chen, X. Liu, S. W.
org/10.1038%2Fs41467-017-01190-3. Kwok, M. Gershow, A. M. Leifer, V. Butler, C. Fang-
[10] H. Wioland, F. G. Woodhouse, J. Dunkel, J. O. Kessler, Yen, et al., Neuron 76, 750 (2012).
and R. E. Goldstein, Phys. Rev. Lett. 110, 268102 [34] G. I. Taylor, Proceedings of the Royal Society of London.
(2013). Series A. Mathematical and Physical Sciences 209, 447
[11] C. Jin, Y. Chen, C. C. Maass, and A. J. Mathijssen, (1951), URL https://doi.org/10.1098%2Frspa.1951.
Phys. Rev. Lett. 127, 088006 (2021). 0218.
[12] M. S. Davies Wykes, X. Zhong, J. Tong, T. Adachi, [35] C. Brennan, Journal of Fluid Mechanics 65, 799 (1974).
Y. Liu, L. Ristroph, M. D. Ward, M. J. Shelley, and [36] J. Koiller, K. Ehlers, and R. Montgomery, J. Nonlinear
J. Zhang, Soft Matter 13, 4681 (2017). Sci. 6, 507 (1996).
[13] P. Galajda, J. Keymer, P. Chaikin, and R. Austin, J. [37] R. Chelakkot, M. F. Hagan, and A. Gopinath, Soft mat-
Bacteriology 189, 8704 (2007). ter 17, 1091 (2021).
[14] A. Guidobaldi, Y. Jeyaram, I. Berdakin, V. V.
Moshchalkov, C. A. Condat, V. I. Marconi, L. Giojalas,
and A. Silhanek, Physics Review E, 89, 032720 (2014).
[15] L. Xu and Y. Jiang, Cells 8, 736 (2019).
[16] D. Allan, T. Caswell, N. Keim, and C. van der
Wel, Trackpy v0.3.2, URL https://doi.org/10.5281/
zenodo.60550.
[17] J. C. Crocker and D. G. Grier, Journal of Colloid Inter-
face Science 179, 298 (1996).
[18] N. Liron and S. Mochon, Journal of Engineering Mathe-
matics 10, 287 (1976).
[19] M. P. Brenner, Physics of Fluids 11, 754 (1999).
[20] B. Cui, H. Diamant, and B. Lin, Phys. Rev. Letters 89
(2002).
[21] C. K. Batchelor and G. Batchelor, An introduction to
fluid dynamics (Cambridge university press, 2000).
[22] W. Boos and A. Thess, Journal of Fluid Mechanics 352,
305 (1997).
[23] J. W. Bush, Journal of Fluid Mechanics 352, 283 (1997).
[24] J. Zeng, Y. C. Yortsos, and D. Salin, Physics of Fluids
15, 3829 (2003).
[25] A. J. T. M. Mathijssen, A. Doostmohammadi, J. M. Yeo-
13

tion
f (z) f (z)
σxx = ∂x U = ¯ ∂x u = ¯ ∂xy ψ
f f
f (z) f (z)
σyy = ∂x V = ¯ f (z)∂y v = − ¯ ∂xy ψ
f f
Appendix A: Computing viscous dissipation σzz = ∂z W = 0
1 f 0 (z) f 0 (z)
σxz = (∂x W + ∂z U ) = u == ∂y ψ
The viscous dissipation rate in an incompressible flow 2 2f¯ 2f¯
depends on the traceless part of the velocity gradient 1 f 0 (z) f 0 (z)
σyz = (∂y W + ∂z V ) = v = − ∂x ψ
2 2f¯ 2f¯
1 f (z)
1 σxy = ∂x V + ∂y U ) = (∂x v + ∂y u)
σij = (∂x ui + ∂xi uj ). (A1) 2 2f¯
2 j
f (z)
= (−∂xx ψ + ∂yy ψ). (A7)
2f¯
The viscous dissipation rate per unit volume We integrate the squares of the components of the ve-
locity gradient over depth;
Z h Rh
X f (z)2 dz
˙ = µ trσ 2 = µ (σij )2 (A2) 2
dz σxx = 0 ¯ 2 (∂xy ψ)2
ij 0 ( f )
hc0
= ¯ 2 (∂xy ψ)2
(f )
where the dynamic viscosity µ = ρν, and ν is the kine- Z h
2 hc0
matic viscosity. dz σyy = ¯ 2 (∂xy ψ)2
0 (f)
We integrate over z to find the viscous dissipation rate Z h
per unit area. 2
dz σzz =0
0
It is useful to compute dimensionless constants that h
Rh
f (z)2 dz 1
Z
depend on the assumed form for the velocity field as a dz 2
σxy = 0
(∂yy ψ − ∂xx ψ)2
function of depth 0 (f¯)2 4
hc0 1
= ¯ 2 (∂yy ψ − ∂xx ψ)2
(f ) 4
Z h
1 Rh 0 2
f¯ = f (z)dz (A3) h
f (z) dz 1
Z
2
h 0 dz σxz = 0 ¯2 (∂y ψ)2
Z h 0 (f ) 4
1
c0 = f (z)2 dz (A4) c1 1
h = (∂y ψ)2
Z
0
h
h(f¯)2 4
c1 = h f 0 (z)2 dz. h
c1 1
Z
(A5) 2
0 dz σyz = (∂x ψ)2 . (A8)
0 h(f¯)2 4
In terms of the stream function, the viscous dissipation
rate per unit area is
The full three-dimensional velocity field (U, V, W ) is re- Z h
lated to the vertically averaged two-dimensional velocity
ėA (x, y) = dz (x,
˙ y, z)
field (u, v) with 0
"  
c0 2 1 2
=µh ¯ 2 2(∂xy ψ) + (∂yy ψ − ∂xx ψ)
1 (f ) 2
U (x, y, z, t) = ¯u(x, y, t)f (z)
f
#
c1 2 2

1 + 2 ¯ 2 (∂y ψ) + (∂x ψ) . (A9)
V (x, y, z, t) = ¯v(x, y, t)f (z) 2h (f )
f
W (x, y, z, t) = 0. (A6) Using f (z) from equation 5, corresponding to a fixed base
and free surface, and with equations A5, we find f¯ = 2/3,
c0 = 8/15, c1 = 4/3, c0 /(f¯)2 = 1.2, and c1 /(f¯)2 = 3
From U, V, W , we compute the different components of which are used in equation A9.
the velocity gradient tensor in terms of the stream func- To estimate the viscous dissipation rate, we integrate
the dissipation rate per unit area with equation A9, over
an area with x ranging from 0 to λM W and y ranging
from the boundary to y = ym + λM W where ym (defined
in equation 15) is approximately the mean y value for the
boundary.

You might also like