You are on page 1of 19

How many ways a cell can move: the modes of self-

propulsion of an active drop


Aurore Loisy a∗ , Jens Eggers a and Tanniemola B. Liverpool a‡
arXiv:2001.03970v2 [cond-mat.soft] 4 Mar 2020

Numerous physical models have been proposed to explain how cell motility emerges from internal
activity, mostly focused on how crawling motion arises from internal processes. Here we offer a
classification of self-propulsion mechanisms based on general physical principles, showing that
crawling is not the only way for cells to move on a substrate. We consider a thin drop of active
matter on a planar substrate and fully characterize its autonomous motion for all three possible
sources of driving: (i) the stresses induced in the bulk by active components, which allow in
particular tractionless motion, (ii) the self-propulsion of active components at the substrate, which
gives rise to crawling motion, and (iii) a net capillary force, possibly self-generated, and coupled to
internal activity. We determine travelling-wave solutions to the lubrication equations as a function
of a dimensionless activity parameter for each mode of motion. Numerical simulations are used
to characterize the drop motion over a wide range of activity magnitudes, and explicit analytical
solutions in excellent agreement with the simulations are derived in the weak-activity regime.

1 Introduction
vide the source of motion. The asymmetry may be in the drop
To perform essential biological functions such as wound healing shape, resulting from an imbalance in surface tension, typically
and immune response, but also in pathological processes such as due to imposed chemical or thermal gradients which provide a
cancer metastasis, eukaryotic cells adapt their mode of migration non-zero flux leading to motion even for a passive drop 16–18 . A
to the geometrical and physicochemical properties of their envi- drop of active matter, in contrast, generates fluxes and asymme-
ronment while relying on the same machinery, the actomyosin try all by itself due to energy input from its components 19–23 that
cytoskeleton 1–3 . In view of the complexity of cell motility, one can cause the drop to move spontaneously 10–13,15,24–27 . Several
may want to ask first: what are the physical requirements for au- studies have shown propulsion of active drops on a surface with
tonomous motion, and what are the possible ways to move? Here a number of related models 8,9,13,14,28,29 . However the complexity
we answer these questions by taking a deformable drop of ac- of the underlying dynamics means identifying similarities and dif-
tive matter (such as the cytoskeleton) and classifying the possible ferences between them is difficult, leading to an ongoing debate
mechanisms for self-propulsion on a substrate. about mechanisms.
Motion on a hard surface is a particularly important class of
motility, because it is the first step towards understanding the self- The hydrodynamic theory of active matter provides a now well-
propulsion of cells in the tissue of multicellular organisms, and in accepted description of active liquids in terms of a limited number
vitro experimental investigations of cell motility often involve the of coupled nonlinear governing equations for conserved fields and
study of cells in contact with a solid substrate 1,4–7 . However, how broken-symmetry fields 19,20,22,23 . One way to study the problem
such self-propulsion emerges from the components of living cells of a moving active drop is through direct numerical simulations of
remains a subject of debate 7–14 . those equations in a domain with moving boundaries 13,25–27,30,31 .
A minimal system to study motility is provided by a deformable While those provide valuable information, they are computation-
drop of material with anisotropic components that consume en- ally expensive and they fail at providing a simple picture of the
ergy (active matter) on a flat rigid surface 13–15 . For a drop of soft mechanisms at play. Another approach, which we shall follow
material to self-propel, two things are required: an asymmetry to here, takes advantage of the geometry of the problem: assuming
give a direction of motion and a mechanical energy flux to pro- that the drop is characterized by a small height-to-width ratio,
one can use the disparity of length scales to reduce the full set
of governing equations and boundary conditions to a single evo-
a
School of Mathematics, University of Bristol - Bristol BS8 1UG, UK.
lution equation much easier to analyze and comprehend. This

E-mail: aurore.loisy@bristol.ac.uk. framework, known as the lubrication (or long-wave) theory 32,33 ,

E-mail: t.liverpool@bristol.ac.uk. has been exploited extensively for the study of thin films and

1–19 | 1
tractionless tank-treading driven by active stresses
achieved without exerting traction on the surface, a remarkable
,
+ property which has been the subject of a recent communication 43 .

The second possible source of motion is the self-advection term


,

wn that arises if the active units propel themselves at a speed w


!

along their own tangent. When coupled to strong enough fric-


!V
tion with the substrate, self-advection allows a drop to “crawl”
along the surface 7–10,13,44,45 . Crawling driven by self-advection
tangential traction on the substrate encompasses much prior work on motile active drops on hard sur-
0
faces 7–10,13,29,44,45 , and is revisited here within our simple frame-
work.
crawling driven by active self-advection
The third way to move is due to the action of a net capillary
force, as would result from (possibly self-induced) thermal or
chemical gradients. This mechanism has been exploited exten-
sively to create self-propelled passive droplets 16–18,46–50 , and here
w we address the effect of coupling it to internal activity.
strong adhesion
2 Model of a thin active drop
sliding driven by capillarity
Our model, illustrated in Fig. 2, consists of a 2D drop of viscous,
active, nematic liquid on a rigid substrate and confined by sur-
Ffriction Fcapillary face tension. The director is strongly anchored at the boundaries,
and the interaction of the liquid with the substrate is modelled
?1 ?2 < ?1 by a partial slip boundary condition. The number density of ac-
tive units is assumed uniform: motility induced by density gra-
dients 11,12,24 is not considered here. We further assume a drop
Fig. 1 Classification of the modes of motion of an active drop. In trac- geometry with a small height-to-width ratio and use the lubrica-
tionless tank-treading driven by active stresses, here drawn in the drop
frame of reference, motion arises from the internal net flow (blue arrows)
tion approximation to reduce the original problem to a nonlin-
generated by active stresses (∝ α), and is achieved without exerting any ear third-order ordinary differential equation for the drop shape
traction anywhere on the substrate except near the contact line. In crawl- which involves the drop velocity as an unknown constant and
ing driven by self-advection, macroscopic motion arises from the self- with prescribed contact angles as boundary conditions. It is ob-
advection (∝ w) of polarized active units, provided that adhesion with the
tained from the balance of activity, viscosity and surface tension
substrate is strong enough to transmit momentum effectively. In sliding
driven by capillarity, the drop is pulled by a net capillary force (due to, in a regime where the director field minimizes the free energy
e.g., an asymmetry in contact angles φ1 and φ2 ), the driving mechanism (no backcoupling to the flow). In the following subsections we
is an external or a self-generated gradient of surface tension or energy. outline each of the ingredients that go into our model and analy-
sis. The reader not interested in the details of the model and the
derivation can find the thin drop problem we solve summarized
droplets of passive nematic liquid crystals 34–39 and has recently in Section 2.6.
been extended to active liquids with (nematic or polar) orienta-
tional order 14,29,40–42 . Prior work has been concerned with thin
film stability 40 , dewetting 29 , and drop spreading 41 . But to the 2.1 Height equation
best of our knowledge the question of motility has only been tack-
We consider a drop moving on a substrate in the x-direction. At
led superficially 14,29 , mostly due to the difficulty in obtaining a
steady-state, the drop shape is described by the height function
closed form for the evolution equation.
h(x), and the constant drop velocity is denoted V (both being
In this paper, we present a unifying description of a thin active
unknown). In the co-moving frame of reference, the flux through
drop on a planar substrate in terms of a single ODE. We show that
a cross section must vanish. This reads
its available modes of motion fall into three distinct classes which
Z h
can be identified based on general principles, independent of the
(ux + wnx − V ) dz = 0 (1a)
details of the model (Fig. 1). 0
A first way to generate motion is through the extra stresses
where u is the fluid velocity inside the drop (with ∇ · u = 0) and
generated by the active components in the bulk. At the contin-
wn describes the additional transport due to the self-advection at
uum scale, these active stresses yield an extra contribution to the
speed w of active units whose orientations are characterised by a
stress tensor σ a = −αnn where n is the director (a unit vector
local orientation n (a unit vector).
that describes the local orientation of the active units). We show
that the motion of a drop that originates from active stresses is The height function, defined on the domain x ∈ [−L/2, L/2],
controlled by the global topology of the director field, and can be must satisfy Eq. (1a) together with four boundary conditions at

2| 1–19
z < " m = .5m Appendix B):
m ∂j σij = 0 (3a)
V
r"u=0 where σij is the stress tensor
h(x) r" < = 0 n a
n σij = −pδij + η(∂j ui + ∂i uj ) + σij + σij (3b)
?1 ux = `u <xz =2 ?2
a
x and p is the pressure, η is the viscosity, σij is the contribution
!L=2 L=2 n
to the stress arising from activity, and σij is the contribution to
the stress arising from its nematic elasticity 51 . The active stress
Fig. 2 Model of a 2D drop of active fluid moving at velocity V on a rigid
reads 20
surface. The fluid motion inside the drop is governed by the incompress- a
ible Stokes flow equations, with u the velocity and σ the stress tensor, σij = −αni nj (3c)
which includes an active contribution σ a = −αnn where n is the direc-
and is due to the forces exerted by the active units on the sur-
tor field. The mechanical interaction with the substrate is modeled by a
partial slip boundary condition (ℓu is the slip length, η is the viscosity) rounding fluid. It can be derived from modeling active units as
and a free surface boundary condition is applied at the interface (γ is the force dipoles 52 and subsequent coarse-graining. The magnitude
surface tension coefficient and κ is the curvature). The drop shape is of α is proportional to the strength of the force pair and the den-
described by the height function h(x) on the domain x ∈ [−L/2, L/2]
sity of units, and the sign of α depends on whether the induced
where L is the drop width. Contact angles φ1 and φ2 are prescribed on
each side of the drop. flow is extensile (α > 0) or contractile (α < 0). For α = 0, one
recovers the standard momentum balance for passive nematic liq-
uid crystals. Since thin films and drops of passive nematics have
been studied extensively (e.g. 34–39 ), and since we are chiefly con-
the contact lines:
cerned here by α 6= 0, we will first work in a regime where ne-
n
h(− L2 ) = 0, h( L2 ) = 0, matic stresses σij can be neglected (see Appendix B). They will
(1b) be included later on in Appendix E.
h′ (− L2 ) = φ1 , h′ ( L2 ) = −φ2 ,
At the solid/liquid interface we use a partial slip boundary con-
where φ1,2 are the contact angles on each side of the drop. The dition:
drop velocity V enters as a constant which must be determined ℓu σxz
ux = at z = 0 (4a)
as part of the solution. The drop width L is also unknown and is η
determined by the volume constraint where ℓu is a slip length (no-slip is obtained for ℓu = 0). At the
Z L/2
gas/liquid interface we use a free surface boundary condition:
h dx = Ω (1c)
−L/2 σ · m = γκm at z = h (4b)

where Ω is the (prescribed) drop volume. To close the problem where m is the unit outward vector normal to the free surface,
described by Eqs. (1a–c), one must now determine an explicit γ is the uniform surface tension, and κ = −∇ · m is the signed
expression of the integral on the left-hand-side of Eq. (1a) in curvature.
terms of h.
The director n = (cos θ, sin θ), which describes the coarse-
2.2 Self-advection of active units grained orientation of the active units, is determined by minimiz-
ing the free energy of a nematic liquid crystal in the strong elastic
The self-advection velocity wn in Eq. (1a) accounts for the
limit 51 :
ability of polarized active components, such as motile bacteria
∇2 θ = 0. (5)
or cytoskeletal filaments undergoing polymerization and tread-
milling, to propel themselves along their own tangent. Such self- Hence the effect of the director on the flow is taken into account,
advection is confined close to the substrate, and to facilitate com- but the back-coupling of the flow on the director is negligible in
parison with prior work we assume the same following functional this regime (see Appendix B).
form as in 13 :  
z As for boundary conditions, we assume strong anchoring (fixed
w = w0 exp − (2)
ℓw angle relative to the surface orientation) at both the substrate
where w0 is a characteristic self-advection speed and ℓw is the and the free surface. Restricting to situations where anchoring
characteristic height over which the self-advection term decays in is either parallel or normal to the surfaces, and remarking that a
the direction normal to the substrate. rotation of n by π/2 is equivalent to a change of sign of α, we
assume without loss of generality that the director is anchored
2.3 Hydrodynamics of an active liquid parallel to the substrate:
The equations of motion for an active liquid are well- θ=0 at z = 0. (6a)
established 19,20,22,23 . Inside the drop, the velocity field is solution
of the momentum conservation equation (neglecting inertia, see At the free surface, we assume that the anchoring angle with re-

1–19 | 3
spect to the surface tangent is ωπ/2 (ω ∈ Z), which reads We introduce several dimensionless groups that reflect the
π physics at play: C = γ/(ηU ) is an inverse capillary number which
θ=ω + arctan(h′ ) at z = h. (6b) compares surface tension to viscous stresses, A = (αL)/(ηU ) is
2
the ratio of active stresses to viscous ones, and W = w0 /U con-
2.4 Force and traction on the drop trols the strength of self-advection compared to the internal fluid
flow.
Before going further it is useful to write down, without any sim-
At leading order in ǫ, Eq. (5) reduces to ∂z̃2 θ = 0. Integrating
plifying assumptions, the force balance for a drop on a substrate.
twice and using the anchoring conditions [Eq. (6)], we find the
It reads (as shown in Appendix C)
expression of the orientation field:
Ffriction + Fcapillary = 0 (7a)  ωπ  z̃
θ=m + ǫh̃′ (10)
2 h̃
with
where ω is effectively a winding number which measures the
Fcapillary = γ (cos φ2 − cos φ1 ) (7b) number of quarter-turns of the director across the drop height,
Z and where
Ffriction = −σxz |z=0 dx (7c) h2
m= 2 (11)
substrate h + ℓ2θ
where Fcapillary is a driving force due to an imbalance in surface is an ad hoc regularizing function, borrowed from 36,37 , and intro-
energies (with φ1 and φ2 the contact angles on each side of the duced to alleviate the conflict of strong anchoring conditions for
drop), and Ffriction is the opposing force (of frictional nature) ex- h → 0. Here ℓθ is a characteristic small length scale such that for
erted by substrate on the drop. If the φ1 6= φ2 , |Fcapillary | > 0: the h ≫ ℓθ , one retrieves the strong anchoring limit (m = 1) and for
drop is pulled by the net capillary force, and its velocity is deter- h ≪ ℓθ , the anchoring constraint is relaxed (m = 0).
mined by the balance with friction (leading to capillarity-driven Then, we have to distinguish between two situations:
sliding). If the contact angles are the same, Fcapillary = Ffriction = 0:
1. ω 6= 0 implies θ = O(1), therefore no rescaling is needed
while a passive drop would necessarily remain static, this is not
(θ̃ = θ) and at leading order the director is not coupled to
the case in the presence of activity (leading to tractionless tank-
the drop shape;
treading or crawling).
Besides, the mechanical interaction of the drop with the sub- 2. ω = 0, the director remains aligned with the bounding sur-
strate can be characterized by the spatial distribution of the trac- faces (deviations from the aligned state are due to defor-
tion, the latter being defined as the local force per unit area ex- mations of the free interface), θ = O(ǫ) so we rescale the
erted by the substrate on the drop. The tangential component of director orientation as θ̃ = θ/ǫ.
the traction, denoted σsubstrate/drop , is
The expression of the (rescaled) director orientation is, at lead-
σsubstrate/drop = ex · σ · (−ez )|z=0 = −σxz |z=0 (8) ing order in ǫ,

(in Section 3 we report instead σdrop/substrate = −σsubstrate/drop as mωπz̃/(2h̃) if ω 6= 0,
this is what one would measure experimentally). θ̃ = (12)
 ′
mh̃ z̃/h̃ if ω = 0.
From Eq. (7) one can remark that, for Fcapillary = 0, we neces-
sarily have Z The derivation of the thin drop equation then proceeds as follows.
σsubstrate/drop dx = 0 (9) The z-component of the Stokes flow equation [Eq. (3a)] gives, at
substrate
leading order in ǫ, ∂z̃ p̃ = 0, and using the normal component of
but σsubstrate/drop does not have to be identically zero. In other the free surface boundary condition [Eq. (4b)] we find
words, autonomous propulsion driven by active processes is nec-
essarily force-free (in the sense that Fcapillary = Ffriction = 0) but is p̃ = −Cǫ3 h̃′′ for any ω. (13)
not, in general, traction-free (σdrop/substrate is not zero everywhere).
The x-component of Eq. (3a) gives, at leading order, ∂z̃ σ̃xz =
∂x̃ p̃. This can be integrated once in z̃, and using the tangential
2.5 Lubrication approximation component of Eq. (4b) we find the expression of the shear stress:
We consider a geometry where the drop characteristic height
3 ′′′

H is much smaller than its characteristic width L. We intro- −Cǫ h̃ (z̃ − h̃) if ω 6= 0,
duce a small parameter ǫ = H/L ≪ 1 and work in the frame- σ̃xz = (14)
−Cǫ3 h̃′′′ (z̃ − h̃) − Aǫ2 h̃′ if ω = 0.

work of lubrication theory 32,33 . Following the usual procedure
(e.g., 14,36,37,39–42 ), we rescale the coordinates and variables as Substituting the definition of σ̃xz [Eq. (3b)] into Eq. (14) and
follows: t̃ = (tU )/L, x̃ = x/L, z̃ = z/(ǫL), h̃ = h/(ǫL), integrating once in z̃ with the partial slip boundary condition [Eq.
ℓ̃u,w = ℓu,w /(ǫL), ũx = ux /U , ũz = uz /(ǫU ), p̃ = (pǫ2 L)/(ηU ), (4a)] yields the parallel component of the fluid velocity:
σ̃ij = (σij L)/(ηU ) where U is a characteristic velocity scale in
the x-direction for the internal flow. ũx = ũcx + ũax (15a)

4| 1–19
with ũcx the capillary flow and defining

z̃ 2 ηV
 
ũcx = −Cǫ3 − (z̃ + ℓ̃u )h̃ h̃′′′ (15b) Ṽ = , (18)
2 γǫ3

and ũax the active flow αL



 if ω 6= 0,
 2πωγǫ2



(1 − cos 2θ̃) Ã = (19)
Aǫ h̃ if ω 6= 0,



 2πωm

 αL
a
 
 if ω = 0,
ũx = (15c) γǫ
mz̃ 2
  ′


2

Aǫ − (z̃ + ℓ̃u )h̃ if ω = 0. ηw0


2 h̃ W̃ = (20)
γǫ3
Averaging the flow over the drop height we find
we can write the problem as
Z h̃  
1 h̃
ũcx dz̃ = Cǫ3
 
+ ℓ̃u h̃h̃′′′ (16a) h̃
+ ℓ̃u h̃h̃′′′ + Ãf˜α (h̃) + W̃ f˜w (h̃) = Ṽ
h̃ 0 3 3
and 

if ω 6= 0,

1
 

Aǫ h̃ if ω 6= 0, m

˜α

 2πωm f (h̃) =


Z 
1
ũax
 
dz̃ = (16b) 
(3 − m)h̃ (21a)
+ ℓ̃u h̃′ if ω = 0,

h̃ −
  
(3 − m)h̃
0  
−Aǫ2

 + ℓ̃u h̃′ if ω = 0. 6
6

not considered 6 0,
if ω =
Using the expression of w given by Eq. (2) we can also write the



mean flow due to self-advection ˜w
f (h̃) =
 ℓ̃w 1 − exp −h̃/ℓ̃w
 h  i

 if ω = 0,
not considered
 if ω 6= 0, h̃
Z h̃ 
1

w̃ñx dz̃ = h  i (16c) where Ṽ is the dimensionless rescaled drop velocity, to be deter-
h̃ 0 1 − exp −h̃/ℓ̃w
mined as part of the solution, and m is a regularizing function,


W ℓ̃w
 if ω = 0.
h̃ defined by Eq. (11), that relaxes the strong anchoring boundary
Equation (16) closes Eq. (1a) which, in rescaled variables, can be conditions for h → 0. This ODE is supplemented by four bound-
written as ary conditions
1 h̃ c
Z
Ṽ = (ũx + ũax + w̃ñx ) dz̃ (17) h̃(− L̃2 ) = 0, h̃( L̃2 ) = 0,
h̃ 0
(21b)
h̃′ (− L̃2 ) = φ̃1 , h̃′ ( L̃2 ) = −φ̃2 ,

Note that we must have C ∼ ǫ−3 such that surface tension where φ̃1,2 are the contact angles on each side of the drop, and
enters at leading order, A ∼ ǫ−1 for ω 6= 0 and A ∼ ǫ−2 for where the drop width L̃ is determined from
ω = 0 such that active stresses play a role at leading order, and Z L̃/2
W ∼ 1 to have the effect of self-advection at leading order. h̃ dx̃ = Ω̃ (21c)
−L̃/2

where Ω̃ is a prescribed drop volume.

To characterize the local mechanical interaction of the drop


2.6 Thin drop equation
with the rigid surface, we also introduce

L
σ̃drop/substrate = σxz |z=0 (22)
To summarize, the steady-state shape h of a thin active drop mov- γǫ3
ing at constant (unknown and possibly zero) velocity V along the which is the (rescaled dimensionless) local traction exerted by
substrate is the solution of a third-order nonlinear ODE. The form the drop on the surface in the x-direction. It can be expressed in
of this ODE depends on the winding number ω, defined as the terms of the local drop shape and reads:
number of quarter-turns of the director across the drop height  ′′′
imposed by the anchoring boundary conditions. h̃h̃ if ω 6= 0,
σ̃drop/substrate = (23)
 ′′′
h̃h̃ − Ãh̃′ if ω = 0.

Introducing appropriate nondimensionalization and rescaling


(denoted by a tilde), such that all rescaled quantities are O(1), Finally the flow inside the drop is, at leading order, parallel to

1–19 | 5
the wall. Redefining self-propulsion driven by
ηux active stresses self-advection capillarity
ũx = (24)
γǫ3 ω 6= 0 0 0
the rescaled fluid velocity is given as a function of the drop shape à =6 0 0 0 and 6= 0
by W̃ 0 6= 0 0
    φ2 − φ1 0 0 6= 0
c h̃ ωπz̃
ũ + Ã 1 − cos if ω 6= 0,

 x

m

 h̃ Table 2 The three basic modes of motion: (i) tractionless tank-treading
ũx = (25) driven by active stresses (∝ Ã) and controlled by the winding number ω
2
 ′
(if ω = 0 the drop is static), (ii) crawling driven by self-advection (∝ W̃),

mz̃ h̃


ũcx + Ã − (z̃ + ℓ̃u )h̃ if ω = 0,


2 (iii) sliding driven by a capillary force Fcapillary = γ(cos φ2 − cos φ1 ) and

possibly modulated by activity.
where ũc is the usual capillary parabolic flow
 2 

ũcx = − − (z̃ + ℓ̃u )h̃ h̃′′′ . (26) static (we comment on this at the end of this subsection). There-
2 fore we assume ω 6= 0, that is, we enforce a winding of the
director through anchoring conditions at the bounding surfaces
2.7 Numerical methods and parameters [Fig. 3(a)]. The governing ODE for the drop shape reduces to
Stable solutions to Eqs. (21a–c) and presented in Section 3 were  
h̃ h̃
obtained numerically as steady solutions to the time-dependent + ℓ̃u h̃h̃′′′ + Ã = Ṽ. (27)
3 m
problem (presented in Appendix A and given by Eq. (40)) in the
thin drop approximation. One can readily see that the drop shape and velocity are con-
Our time integration algorithm is based on a Crank-Nicolson trolled by the dimensionless parameter à = (αL)/(2πωγǫ2 ),
scheme with adaptive time-stepping. For space discretization, we where α and ω can be of either sign.
use second-order finite difference schemes on a uniform grid. At Since Ã(α, ω) = Ã(−α, −ω), changing the direction in which
each time step, the resulting nonlinear system of equations was the director winds (from counter-clockwise to clockwise) is equiv-
solved using the Matlab nonlinear system solver. The solution was alent to changing the sign of activity (from extensile to contrac-
advanced in time until the steady-state was reached, correspond- tile). It is also interesting to note that if {h̃(x̃), V} is a solution
ing to the sought-after travelling-wave solution. for à then {h̃(−x̃), −Ṽ} is a solution for −Ã: reversing the sign
Numerical parameters used in the simulations are summarized of à simply reverses the direction of motion. Therefore in the
in Table 1. The volume (surface area) of the drop was kept con- following we will only consider à > 0.
stant across all the simulations and set to Ω̃ = 1. The evolution of the drop shape and velocity with à is shown
in Fig. 3(b,c). Overall, the drop becomes thinner and faster as
self-propulsion driven by activity increases. Solutions are however qualitatively different
active stresses self-advection capillarity at low and high Ã.
Ngrid 800 400 200
In the limit of small Ã, the drop shape is close to a parabola
Ω̃ 1 1 1
(the equilibrium shape for a passive drop), and its velocity can be
ℓ̃u 0.05 0.01 0.05
computed analytically at linear order in à (Appendix D.1):
ℓ̃θ 0.05 0.01 0.05
φ̃1 1 1 10
p p
φ̃L̃0 b̃ − (b̃ − 1) arctanh(1/ b̃)
φ̃2 1 1 5 Ṽ = Ã (28)
4
p
arctanh(1/ b̃)
Table 1 Numerical parameters used in the simulations (unless men- q
tioned otherwise): number of grid points (Ngrid ), drop volume (Ω̃), slip where b̃ = 1 + 12ℓ̃u /(L̃0 φ̃) and L̃0 = 6Ω̃/φ̃. The first correc-
length (ℓ̃u ), characteristic thickness for strong anchoring relaxation (ℓ̃θ , tion for the drop width is quadratic, so at linear order L̃ = L̃0 =
set equal to ℓ̃u ), and contact angles (φ̃1 and φ̃2 ). q
6Ω̃/φ̃ and the mean drop height is H̃ = H̃0 = Ω̃/L̃0 where Ω̃
is the drop volume (kept constant across simulations). Compari-
3 Results son to the numerical solution (dash-dotted lines in Fig. 3(c) (left
panels) is excellent and shows that this solution remains valid up
Three distinct driving mechanisms (active stresses, self-advection,
to à ≈ 0.1.
and capillary forces due to different contact angles) are embed-
For à > 1, the drop is locally flat (the numerical value of this
ded in Eqs. (21a–c), leading to the three modes of motion sum-
threshold depends on ℓ̃u and φ, so the fact it is unity here is coin-
marized in Table 2 and that we will analyze separately in the
cidental). The extent of the flat region rapidly increases with Ã:
following.
analysis of numerical data indicates that, for à > 1, the frac-
tion of the drop which is not flat first decreases as Ã−1 . For
3.1 Self-propulsion driven by active stresses à & 10, more than 90% of the drop is bounded by a flat free
We consider the motion of a drop arising solely from active surface [Fig. 3(c), top right panel]. In this regime the drop veloc-
stresses (W̃ = 0, φ̃1 = φ̃2 ). We found that if ω = 0, the drop is ity is exactly given by Ṽ = Ãh̃flat where h̃flat is the height of the

6| 1–19
(a) z~ ~ x)
h(~
fre ,L
es A~ =
ur
fa c 2:!.02
e
V~

?~1 partial slip ?~2 = ?~1


x
~
~
!L=2 ~
L=2

(b)

(c) (d)
1 100
~ .at )=L
~

/ A~!1
~ H~0

~ !L
H=

0.5 10!1
(L

(locally) .at
0 10!2
0.3
0.4
~ V~
~ ! V)=

0.2
~ A~
V=

0.2
(A~H

analysis 0.1
numerical
H~
0 0
10!2 10!1 100 101 102 10!2 10!1 100 101 102
A~ A~

(e) !V

! = +1 n ux H ! = +2

!V !V

Fig. 3 Self-propulsion driven by active stresses: the motion of a drop endowed with active stresses is controlled by the global topology of the director
field and can be achieved without exerting traction locally on the surface. (a) Model of a thin active drop with active stresses and a winded director. The
drop shape and velocity are controlled by à = (αL)/(2πωγǫ2 ) where ω is the winding number. (b) Numerical profiles of the drop shape and tangential
traction exerted on the substrate. (c) Effect of à on the drop shape and velocity, H̃ is the mean height and the analytical solution for Ṽ is given by Eq.
(28). (d) Effect of slip on the drop shape and velocity (Ã = 1). (e) Sketch of a tractionless flat drop moving at velocity V = (αH)/(2πωη): director and
velocity fields in the co-moving frame of reference (solution for ω = 2 is also valid for a drop confined between two walls). Colored symbols in (b-d)
mark corresponding state points across panels.

1–19 | 7
flat region. In practice, the drop velocity is well approximated by This geometrically constrained setup is perhaps the easiest to
[Fig. 3(c), bottom panels] control experimentally: one can imagine confining a drop of
bacterial suspension 56,57 or of microtubule-kinesin mixture 15 be-
Ṽ = ÃH̃ (29) tween two surfaces, one used for imaging the traction maps 58–60
and the other designed to ensure appropriate anchoring (through,
with H̃ the mean drop height: the error on Ṽ is less than 5 % for
e.g., manipulation of the surface chemistry or architecture 61–65 ).
à > 1 and goes to zero as à → ∞. The mechanical interaction
Traction maps would show a zero traction on the channel walls
with the wall (modeled through the slip length ℓu ) does not ap-
everywhere except at the drop edges, where the traction magni-
pear in Eq. (29), but it enters indirectly through the dependence
tude and sign would only depend on the wettability of the walls.
of H̃ on ℓ̃u (Fig. 3d): more slip yields thicker (and hence faster)
It is important to emphasize that tractionless motion controlled
droplets.
by ω 6= 0 is not related to the spontaneous flow transition in active
At high Ã, the drop is flat everywhere except near the con-
nematics films 66 . Our analysis describes a drop of active nematic
tact lines, and the local tangential traction σ̃drop/substrate induced by
in the strong elastic limit, that is, in a regime where K > O(1)
this drop on the substrate is identically zero almost everywhere
(see Appendix B), where K = (ΓK)/(U L) with 1/Γ the rotational
(Fig. 3b, right panel). This is remarkable: while autonomous
viscosity and K the nonequilibrium analog of an elastic constant.
propulsion driven by active processes is necessarily force-free, it
In this limit, there is no internal flow (and hence no drop mo-
is not, in general, tractionless (see Section 2.4). Strictly speaking
tion) for ω = 0. It is well-known that for ω = 0, internal flows
σ̃drop/substrate = 0 where h′′′ = 0 (from Eq. (23)), that is, every-
can occur spontaneously in thin films of active nematics beyond
where except at the drop edges. Integrating σ̃drop/substrate over an
a critical height due to a splay (or bend) instability 66 . Within
edge yields a force of magnitude φ̃2 /2 and directed inward: as
our framework and with our notations, this instability requires
seen from the substrate, the drop effectively acts as a contrac-
K 6 O(ǫ), in other words, it requires a drop thicker than the one
tile force dipole, independent of activity and due to finite contact
we consider here. Whether the spontaneous flow transition for
angle φ̃.
active films 66 results in a “spontaneous tractionless motion tran-
A sketch of the tractionless motion of a flat drop is provided in
sition” for active drops remains an open question. To answer it,
Fig. 3(e), where we also illustrate the role played by the wind-
one must first integrate the full dynamic equation for the direc-
ing number ω. The winding of the director (green rods) induces
tor [Eq. (41)] rather than its strong elastic limit [Eq. (5)]. This
an active stress in the liquid which must be balanced by the vis-
problem, significantly more intricate, is left to future work.
cous stress such that the total shear stress vanishes. The internal
In any case it is of fundamental importance to note that this
fluid flow thereby generated is sinusoidal, rather than parabolic
kind of motion is only possible for active matter driven in the
in other modes of motion (blue arrows, plotted in the co-moving
bulk and cannot happen for propulsion due to driving at or near
frame of reference). Going back to the original dimensional vari-
boundaries (the other two modes considered in this paper).
ables, the fluid velocity reads, in the laboratory frame of refer-
ence, h  ωπz i
ux = V 1 − cos (30) 3.2 Self-propulsion driven by self-advection
H
Crawling is a mode of cell motility well-characterized experi-
where H is the drop height and where
mentally 5,6,67 and captured by various physical models 7–9,13,14,28 .
αH Crawling motility is usually understood as follows: polymeriza-
V = . (31)
2πωη tion of actin filaments in a thin protrusion at the leading edge gen-
erates a pushing force against the cell membrane, which, when
The net flow is not zero and causes the drop to move at a velocity
combined with anchoring to the substrate via focal adhesions,
V in a tank-treading fashion while exerting no tangential traction
causes the cell to move forward.
on the surface. The winding number controls the number of fluid
A simple way to account for this mechanism, illustrated in
circulation cells, which is exactly equal to |ω|. The drop speed is
Fig. 4(a), consists in adding a self-advection term to the mass
maximum for |ω| = 1, which corresponds to antagonist anchoring
conservation equation (which describes the net polymerization of
conditions for the director at the wall and at the free surface.
filaments in a given direction), while adhesion is controlled by
The case |ω| = 2, while less favored energetically, generates
the amount of slip at the substrate (here through the slip length
a flow which is symmetric with respect to the drop midplane.
ℓu ). To ease comparison with prior work 13 we chose an advection
In particular, the solution has zero fluid velocity and zero shear
velocity which is maximum at the substrate (denoted w0 ) and de-
stress at both boundaries, therefore it also solves the problem
cays exponentially over a characteristic length ℓw in the direction
of a drop squeezing through a narrow channel [right panel in
normal to the substrate, as described in Section 2.2.
Fig. 3(e)]. In this configuration the drop motion is completely
We emphasize that what generates motion here is a flux of mat-
independent of the amount of slip at the walls, since the drop
ter: crawling can be obtained solely from self-advection, in the
height is geometrically constrained. This solution is reminis-
absence of active stresses (Ã = 0) or mismatch in the contact
cent of contraction-based amoeboid motility such as exhibited
angles (φ̃1 = φ̃2 ). The governing equation then reduces to
by leukocyte and human breast cancer cells squeezing through
complex 3D extracellular geometries 53,54 and by confined cells 


ℓ̃w h  i
migrating in microchannels 2,55 . + ℓ̃u h̃h̃′′′ + W̃ 1 − exp −h̃/ℓ̃w = Ṽ. (32)
3 h̃

8| 1–19
(a)

(b)

(c) (d)
1
4 0.8
~ protrusion =L
~

0.6
~0

no no
~ L

protrusion protrusion
L=

2 0.4
L

0.2
protrusion protrusion
0 0
0.7
numerical
~ protrusion =`~w

analysis 1.4
0.6
~ W~

1.2
V=

1
H

0.5
0.8
0 0.2 0.4 0.6 0 0.2 0.4 0.6
~
W ~
W
Fig. 4 Self-propulsion driven by self-advection: a drop with directed self-advection of active units (due to e.g. polymerization toward the front) develops
a frontal protrusion and crawls more effectively as slip is reduced. (a) Model of a thin active drop with self-advection close to the substrate. The effect
of self-advection on the drop shape and velocity is controlled by W̃ = (w0 η)/(γǫ3 ) where w0 is the characteristic self-advection speed of the active
units. The height over which the strength of self-advection decays was kept constant across simulations (ℓ̃w = 0.1). (b) Numerical profiles of the drop
shape and tangential traction exerted on the substrate. (c) Effect of W̃ on the drop shape and velocity, the analytical solution is given by Eq. (33). (d)
Effect of slip on the drop shape and velocity (top and middle panels, W̃ = 0.25), and on the transition between non-protruded and protruded drops
(bottom panel). Colored symbols in (b-d) mark corresponding state points across panels.

1–19 | 9
Besides ℓ̃w (which is kept constant here, ℓ̃w = 0.1) and ℓ̃u (set yields
here to ℓ̃u = 0.01), the drop shape and velocity are controlled by    
a single dimensionless group: W̃ = (ηw0 )/(γǫ3 ). h̃ (3 − m)h̃
+ ℓ̃u h̃h̃′′′ − Ã + ℓ̃u h̃′ = Ṽ (34a)
3 6
Numerical solutions of Eq. (32), computed for a range of W̃,
are presented in Fig. 4(b,c). They reveal the existence of a critical with different contact angles imposed at the boundaries (as may
value of W̃, denoted W̃c , above which a protrusion develops at arise, from, e.g., a gradient of surface energy):
the front. The numerical value of W̃c increases linearly with ℓ̃u ϕ̃ ϕ̃
[Fig. 4(d), bottom panel], and the transition is sharper for larger φ̃1 = φ̃ + , φ̃2 = φ̃ − (34b)
2 2
ℓ̃u , in agreement with prior work 13 .
with ϕ̃ = φ̃1 − φ̃2 the contact angle difference, as depicted in
For W̃ < W̃c , the drop profile is nearly parabolic and it is possi-
Fig. 5(a). The effect of activity on the drop shape and velocity is
ble to derive an analytical expression of the drop velocity at linear
controlled by the dimensionless parameter à = (αL)/(γǫ). The
order in W̃ (Appendix D.2):
sign of à depends on whether active units, modelled as force
dipoles 52 , induce an extensile flow (α > 0, e.g. certain bacteria)
(
2ℓ̃w 1
Ṽ = W̃ p or a contractile flow (α < 0, e.g. the actin-myosin complex).
L̃0 φ̃ (b̃ − 1) arctanh(1/ b̃)
p
2 b̃d˜2 F2 ({1, 1}, {3/2, 2}, −d)
˜

˜   p −1 
+ 2 e(b̃−1)d − 1 ln 1 + 2 cot arcsin(1/ b̃)/2 − 1
  

)
p p p i
(b̃−1)d˜
h
˜ i/

−e 2 arctanh( b̃) + iπ 1 + 4T ( 2b̃d, b̃)
Numerical solutions to Eq. (34) for various à are presented in
(33)
Fig. 5(b,c). With respect to the passive case, the drop base is nar-
with i2 = −1, b̃ = 1 + 12ℓ̃u /(L̃0 φ̃), d = (φ̃L̃0 )/(4ℓ̃w ), and where rower (wider) with contractile (extensile) activity [Fig. 5(b,c)],
T (χ, c) is Owen’s T function 68 and 2 F2 ({a1 , a2 }, {b1 , b2 }, ζ) is as is the case for static symmetric drops 41 . At high −Ã (high con-
the generalized hypergeometric function 69 (these functions are tractility), the drop breaks up; this phenomenon is outside the
implemented in Mathematica). The first correction to the drop scope of this paper and its analysis is left to future work. Activity
also influences the drop speed: contractile (extensile) drops are
q
width is quadratic, so at linear order L̃ = L̃0 = 6Ω̃/φ̃. Compar-
faster (slower) than their passive counterpart [Fig. 5(c)] (note
ison to the numerical solution is shown in Fig. 4(c) (left panels):
that with normal anchoring, the extensile drop would be nar-
agreement is excellent nearly up to W̃c .
rower and faster). Increasing slip results in greater drop veloc-
For W̃ > W̃c , the drop has a frontal protrusion of thickness ities, as for a passive drop since friction hinders sliding motion
∼ ℓ̃w which grows in length upon increasing W̃ [Fig. 4(c), right [Fig. 5(d)].
panels]. The drop velocity magnitude is of the order of W̃, but its
growth with W̃ is faster than linear. Increasing slip reduces the
drop velocity [Fig. 4(d), middle panel]: as expected crawling is
most effective when the substrate provides strong adhesion.
We finally emphasize that these results, here obtained under
the thin drop approximation, are in very good agreement with
prior full numerical simulations 13 (Fig. 1 and S1 therein).

The dependence of Ṽ and L̃ on à is approximately linear over


3.3 Self-propulsion driven by capillarity and modulated by a rather large range of Ã, and can be computed exactly from a
activity perturbation analysis in the limit of small |ϕ̃| and small |Ã|. We
find (Appendix D.3) that the drop width is
We consider an asymmetric drop moving under the action of a
net capillary force (φ1 6= φ2 implies |Fcapillary | > 0), as depicted L̃ = L̃0 + ÃL̃α + O(Ã2 , ϕ̃2 ) (35a)
in Fig. 5(a). The director field is chosen to be nearly aligned
(ω = 0), as larger distortions would essentially lead back to Sec- with (Ω is the drop volume)
tion 3.1 (motion driven by active stresses and controlled by the s
winding of the director). Here we ask: can activity facilitate (or 6Ω̃ Ω̃
L̃0 = , L̃α = (35b)
impede) the drop motion, and does the sign of activity (extensile φ̃ 2φ̃2
or contractile) matter? The answer: Yes, and yes.
and the drop velocity is
A minimal mathematical description of capillarity-driven slid- h i
ing is obtained by setting W̃ = 0 and ω = 0 in Eq. (21a), which Ṽ = ϕ̃ Ṽϕ,L0 + Ã(Ṽϕ,Lα + Ṽϕ,α ) + O(Ã2 , ϕ̃2 ) (36a)

10 | 1–19
(a)
z~ ~ x)
h(~ extensile (, > 0) contractile (, < 0)
,L
fre A~ =
es .0
u rf
ace
V~

?~1 partial slip ?~2 < ?~1


x
~
~
!L=2 ~
L=2

(b)

(c) (d)
30

25
s lip
si ng
rea
20 in c

15 A~ = !50
V~

10
A~ = 0

5
A~ = 50

0
10!3 10!2 10!1
`~u

Fig. 5 Self-propulsion driven by capillarity and modulated by activity: the motion of an asymmetric drop is enhanced by contractility and hindered by
extensility. (a) Model of a thin active drop driven by the capillary force that results from the difference in contact angles (here φ̃1 = 10 and φ̃2 = 5)
that may arise from, e.g., a gradient of surface energy. Parallel anchoring of the director is prescribed at both bounding surfaces. The effect of activity
on the drop shape and velocity is controlled by à = (αL)/(γǫ), whose sign depends on whether the active stress is extensile (α > 0) or contractile
(α < 0). (b) Numerical profiles of the drop shape and tangential traction exerted on the substrate. (c) Effect of à on the drop width and velocity with
respect to the passive case, analytical expressions are given by Eq. (35) and Eq. (36). (d) Effect of slip on the drop velocity. Colored symbols in (b-d)
mark corresponding state points across panels.

1–19 | 11
where the coefficients are given below: tion, a mode of cell motility characterized by strong cell-substrate
adhesions, with self-advection provided by actin polymerization
β̃0 φ̃ in a leading edge protrusion 5,67 .
Ṽϕ,L0 = (36b)
6 ln(β̃p/m )

Finally the third mode motion is, unlike the other two, not
φ̃2 L̃α [β̃0 L̃0 + 6ℓ̃u ln(β̃p/m )] driven by active processes but by a net capillary force, as can
Ṽϕ,Lα = − (36c)
6L̃20 β̃0 [ln(β̃p/m )]2 be induced by (external or self-generated) thermal or chemical
gradients. By coupling this driving with internal activity, one can
( further tune the drop velocity and create droplets faster than their
L̃0 β̃p/m
Ṽϕ,α = (36d) passive counterparts.
24β̃0 [ln(β̃p/m )]2

+ (2β̃02 − β̃m
2
) dilog(β̃p/m ) Our 2D model is expected to be valid for any 3D drop where
variations in the additional spatial dimension are much slower
− (2β̃02 + β̃m
2
) dilog(2φ̃/β̃p ) than in the other two. Beyond that, while extending the thin
drop formulation to 3D is rather straightforward, solutions may
− β̃02 [ln(β̃p )]2 − [ln(β̃m )]2

be far more complex (for example, based on the form of Eq. (27),
we expect a fingering instability for a 3D drop driven by active
+ [2β̃02 + β̃m
2 2

] ln(2φ̃) − β̃m [1 + ln(β̃p )] ln(β̃p/m ) stresses).
)
π2 β̃0 φ̃
+ (2β̃02 + β̃m
2
)−2 Our present attempt to provide a generic classification of self-
6 β̃p/m propulsion mechanisms, one of them being the extensively stud-
q ied treadmilling-driven crawling, led us to introduce an en-
with β̃0 = φ̃ 1 + (12ℓ̃u )/(L̃0 φ̃), β̃p = β̃0 + φ̃, β̃m = β̃0 − φ̃, tirely new class (tractionless tank-treading driven by bulk active
Z y
ln(t) stresses) and to give a new twist to an old mechanism (self-
β̃p/m = β̃p /β̃m and with dilog(y) = dt. Compari-
1 1 −t propulsion driven by gradients): we hope those will trigger fur-
son with the numerical solution is excellent [bottom panels in
ther theoretical investigations and experimental realizations.
Fig. 5(c)].
Embedding an active suspension into an otherwise self-
propelled passive droplet, driven by a gradient of surface energy A Unsteady height evolution
or surface tension 47–50 , could be a rather straightforward way to
realize experimentally this activity-modulated, capillarity-driven
self-propulsion. Consider the general (unsteady) case of a drop shape described
by a height function h∗ (x∗ , t) in the laboratory frame of reference
4 Conclusions (denoted by a star). It is related to the flow at the free surface by
To summarize, we have obtained a generic and unified descrip- a kinematic boundary condition 32,70
tion of a thin 2D drop of active liquid moving on a solid substrate
∂t h∗ + (ux + wnx )∂x∗ h∗ − (uz + wnz ) = 0 (37)
that consists of a single ODE. We have analyzed, using a combi-
nation of numerical simulations and asymptotic analysis, the au- with u the fluid velocity and wn the self-advection at speed w
tonomous propulsion of this drop induced by three possible driv- of active units with local orientations n = (cos θ, sin θ). At the
ing sources (summarized in Fig. 1): active stresses (Fig. 3), active free surface, the anchoring angle is θ = ω π2 + arctan(∂x∗ h∗ ). It
self-advection (Fig. 4), and a (possibly self-generated) capillary follows that
force (Fig. 5).  ωπ  p
Motion driven by active stresses does not require a shape asym- nx ∂x∗ h∗ − nz = − sin 1 + (∂x∗ h∗ )2 . (38)
2
metry, is efficient even in the presence of slip and allows self-
propulsion without the need to exert traction anywhere on the Besides, using flow incompressibility and wall impermeability,
surface, giving rise to “tractionless tank treading”. This new one can show that
mode of motion, driven in the bulk rather than at the boundaries, Z h∗
is topologically protected and is particularly suited for moving u x ∂ x ∗ h∗ − u z = ∂ x ∗ ux dz. (39)
0
rapidly through tiny pores. Therefore it provides a robust physi-
cal mechanism for efficient cell migration in tissues. Therefore the kinematic boundary condition Eq. (37) can be
In contrast, motion driven by the self-advection of polarized rewritten as
active units at the substrate, known as crawling, is character- Z h∗  ωπ  p
ized by a strong shape anisotropy and is most efficient in the ab- ∂ t h∗ + ∂ x ∗ ux dz = w(h∗ ) sin 1 + (∂x∗ h∗ )2 (40)
0 2
sence of slip. Therefore this mode of self-propulsion is particularly
suited for moving on 2D surfaces which provides strong anchoring In this paper we restrict to either w = 0 or ω = 0, so the right-
points. A prominent example of crawling is mesenchymal migra- hand-side of Eq. (40) is zero.

12 | 1–19
B Regime of validity with
Z
Fsubstrate/drop = σ · m ds (43b)
The evolution of the director field in a nematic liquid crystal is ∂Dsol/liq
governed by
Z
Ffree surface/drop = σ · m ds (43c)
∂Dgas/liq
∂t ni + [(uj + wnj )∂j ]ni − Ωij nj =
where D is the domain occupied by the drop, ∂D is its boundary
T
δij (λEjk nk + ΓK∇2 nj ) (41) (decomposed into solid/liquid and gas/liquid interfaces), and m
T is the unit normal directed outward the boundary.
where δij = δij − ni nj is a transverse projection operator, Eij =
(∂i uj + ∂j ui )/2 and Ωij = (∂j ui − ∂i uj )/2 are the strain-rate and
rotation-rate tensors, λ is the flow alignment parameter, 1/Γ is The x-component of the force exerted by the substrate on the
the rotational viscosity and K is the nonequilibrium analog of a drop is
Frank constant.
Z
Fsubstrate/drop,x = −σxz |z=0 dx
∂Dsol/liq
Momentum conservation reads, including inertial terms,
Z
= −η(∂z ux + ∂x uz ) dx (44)
ρ(∂t + uj ∂j )ui = ∂j σij (42)
∂Dsol/liq

where ρ is the mass density of the fluid and where σij is given by
n
Eq. (3b) with σij = −K∂i nk ∂j nk . ≡ Ffriction
a
Since the active contribution σxz |z=0 vanishes for parallel or nor-
In addition to ǫ and φ1,2 , the dynamics of the drop is controlled
mal anchoring of the director, Fsubstrate/drop,x is purely frictional
by seven dimensionless groups that can be constructed from those
and we denote it Ffriction in the main text.
equations: λ, W = w0 /U , K = (ΓK)/(U L), A = (αL)/(ηU ),
C = γ/(ηU ), N = K/(ηU L) and R = ρU L/η. The equations
we solve here, presented in Section 2, are obtained when the di- The force exerted by the free surface on the drop is, using Eq.
mensionless groups satisfy the conditions summarized in Table 3. (4b), Z
They describe a thin drop whose dynamics is determined from the Ffree surface/drop = γκm ds (45)
∂Dgas/liq
balance of active, viscous and surface tension forces in a regime
where inertia and nematic stresses are negligible and where the and in 2D we have
director field is not back-coupled to the flow. 1
m= √ (−h′ , 1)
1 + h′2

h′′ (46)
self-propulsion driven by κ=
active stresses self-advection capillarity [1 + h′2 ]3/2
ω 6= 0 ω=0 ω=0 p
θ = O(1) θ = O(ǫ) θ = O(ǫ) ds = 1 + h′2 dx
W = w0 /U 6 O(ǫ) = O(1) 6 O(ǫ)
It is straightforward to show that in the x-direction,
A = (αL)/(ηU ) = O(ǫ−1 ) 6 O(ǫ−1 ) = O(ǫ−2 )
−3 −3
C = γ/(ηU ) = O(ǫ ) = O(ǫ ) = O(ǫ−3 ) Ffree surface/drop,x = γ (cos φ2 − cos φ1 )
−1
N = K/(ηU L) 6 O(ǫ) 6 O(ǫ ) 6 O(ǫ−1 ) (47)
R = ρU L/η 6 O(ǫ−1 ) 6 O(ǫ−1 ) 6 O(ǫ−1 ) ≡ Fcapillary
K = (ΓK)/(U L) > O(1) > O(ǫ−1 ) > O(ǫ−1 )
λ 6 O(1) 6 O(1) 6 O(1) where φ1 and φ2 are the contact angles on each side of the drop.
Since Ffree surface/drop,x originates purely from surface tension, we
Table 3 Range of validity of our analysis in terms of the dimensionless
groups governing the drop dynamics.
refer to it as Fcapillary in the main text.

D Perturbation analysis
C Force balance on the drop
In this section we will derive, using a perturbation analysis, the
first effect of activity à or of self-advection W̃ on the drop shape
The sum of the forces exerted on the drop, denoted Ftotal , satisfies
h̃, velocity Ṽ, and width L̃.
Ftotal = 0 (as follows from integrating Eq. (3) over the drop) and
is defined by
To facilitate the derivation we first rescale the x-coordinates by
Ftotal = Fsubstrate/drop + Ffree surface/drop (43a) introducing the change of variable ỹ = ξ˜x̃ with ξ˜ = 2/L̃, and we

1–19 | 13
set m = 1. The governing ODE for g̃(ỹ) = h̃(x̃) is then presented in 71 . We introduce a test function t(y) which satisfies
  t(±1) = 0. We can write

ξ˜3 + ℓ̃u g̃g̃ ′′′ + Ãf˜α (g̃) + W̃ f˜w (g̃) = Ṽ 1 1 1
3
Z Z Z
tgδ′′′ dy = Vδ tR1 dy + tR2 dy (52)
 −1 −1 −1

 g̃ if ω 6= 0,
The left-hand-side can be integrated by part three times to yield

f˜α (g̃) =  

−ξ˜ + ℓ̃u g̃ ′ if ω = 0, (48a)

 Z 1 Z 1
 1
3 tgδ′′′ dy = −c t′ −1 − t′′′ gδ dy (53)
−1 −1

not considered if ω 6= 0,
Choosing an adequate test function which satisfies t′′′ = 0, we



f˜w (g̃) = can determine Vδ from
 ℓ̃w 1 − exp −g̃/ℓ̃w
 h  i
 if ω = 0,
g̃  1
Z 1 Z 1
− c t′ −1 = Vδ tR1 dy + tR2 dy. (54)
with boundary conditions −1 −1

On the other hand, choosing a test function with t′′′ = δ yields


g̃(±1) = 0, ˜ ′ (±1) = ∓φ̃ + ϕ̃ ,
ξg̃ (48b)
2 Z 1 Z 1 Z 1
 1
− c t′ −1 − δ gδ dy = Vδ tR1 dy + tR2 dy (55)
where ϕ̃ = φ̃1 − φ̃2 is the difference between contact angles on −1 −1 −1
each side of the drop, and with the volume constraint
From the volume constraint one has
Z 1
g̃ dx̃ = ξ˜Ω̃.
Z 1 Z 1
(48c)
−1
δ gδ dy = ξΩ − g0 dy (56)
−1 −1

In the following all quantities are scaled in the lubrication frame-


therefore we can determine ξ (and then L = 2/ξ) from
work and we drop the tilde in the remaining of this section.
Z 1 Z 1 Z 1
We start from the exact analytical solution, denoted by sub-  1
− c t′ −1 − ξΩ + g0 dy = Vδ tR1 dy + tR2 dy (57)
script 0, for a symmetric passive drop (ϕ = 0, A = 0, W = 0): −1 −1 −1

g0 = φ(1 − y 2 )/(2ξ) (49a)


p
L0 = 6Ω/φ (49b)

V0 = 0. (49c)
D.1 Self-propulsion driven by active stresses
We will then expand the solution to Eq. (48) as a perturbation
power series of the relevant parameters for each mode of motion:

g = g0 + δgδ + O(δ 2 ) (50a)


We consider here the case of motion solely driven by active
L = L0 + δLδ + O(δ 2 ) (50b) stresses (ω 6= 0, ϕ = 0, A 6= 0, W = 0), so the ODE reduces
to
V = V0 + δVδ + O(δ 2 ), (50c) g 
ξ3 + ℓu gg ′′′ + Ag = V. (58)
3
where δ is the small parameter in which we expand. We write the perturbation solution as
As we shall see later on, substituting this expansion into the
governing ODE and matching terms at order δ yield an ODE of g = g0 + Agα + O(A2 ) (59a)
the form
gδ′′′ (y) = Vδ R1 (y) + R2 (y) (51) L = L0 + ALα + O(A2 ) (59b)

with boundary conditions V = V0 + AVα + O(A2 ). (59c)

gδ (±1) = 0 gδ′ (±1) = c At linear order the correction is solution of


V − g0
where c is a constant which depends on the mode of motion con- ′′′
gα =  gα  (60a)
0
sidered. ξ3 + ℓ u g0
3
Conveniently Vδ and Lδ can be computed without solving Eq. ′
(51) by using a solvability condition in the spirit of the approach with gα (±1) = 0, gα (±1) = 0, (60b)

14 | 1–19
and we find after integration with b = 1 + 6ℓu ξ/φ and d = φ/(2ξℓw ). A series of manipulations
 allowed us to obtain an explicit expression of I, which reads
3
gα = √ (61a) (
2φξ 2 (b − 1) arctanh(1/ b) 1 √
I= √ 2 bd 2 F2 ({1, 1}, {3/2, 2}, −d)
(b − 1) b
c (1 + y)2 ln(1 + y) − (1 − y)2 ln(1 − y) − (22 ln 2)y
 

  √ −1 
h√ √ √ √ + 2 e(b−1)d − 1 ln 1 + 2 cot arcsin(1/ b)/2 − 1
i   
− ( b + y)2 ln( b + y) − ( b − y)2 ln( b − y)
)
h√ √ √ √ i  (b−1)d
h √  √
√ i
2 2
+ ( b + 1) ln( b + 1) − ( b − 1) ln( b − 1) y −e 2 arctanh( b) + iπ 1 + 4T ( 2bd, i/ b)

(67)
Lα = 0 (61b)
where i2 = −1, where T (χ, c) is Owen’s T function 68 defined by
φ c
Vα = √ (61c) 
1

2ξ arctanh(1/ b) Z c exp − χ2 (1 + t2 )
1 2
√ √ T (χ, c) = dt
where b = 1 + 6ℓu ξ/φ and c = b − (b − 1) arctanh(1/ b). 2π 0 1 + t2

and where 2 F2 ({a1 , a2 }, {b1 , b2 }, ζ) is the generalized hy-


√ √
pergeometric function 69 . The functions T ( 2bd, i/ b)
and 2 F2 ({1, 1}, {3/2, 2}, −d) can be evaluated in Math-
ematica using OwenT[Sqrt[2*b*d],I/Sqrt[b]] and
HypergeometricPFQ[{1,1},{3/2,2},-d], respectively.
D.2 Self-propulsion driven by self-advection
Finally a solvability condition can also be used to show that
Lw = 0, therefore one can substitute ξ = 2/L0 in Eq. (65).

D.3 Self-propulsion driven by capillarity and modulated by


We consider here the case of pure advective crawling (ω = 0, activity
ϕ = 0, A = 0, W 6= 0), so the ODE reduces to
g We consider here the case of pure capillary sliding (ω = 0, ϕ 6= 0,
3

′′′ ℓw A 6= 0, W = 0), so the governing ODE reduces to
ξ + ℓu gg +W [1 − exp (−g/ℓw )] = V (62)
3 g
 2 
g g 
Writing the perturbative solution as ξ3 + ℓu g g ′′′ − Aξ + ℓu g ′ = V (68a)
3 3
g = g0 + Wgw + O(W 2 ) (63a) with boundary conditions

L = L0 + WLw + O(W 2 ) ϕ
(63b) g(±1) = 0, ξg ′ (±1) = ∓φ + , (68b)
2
V = V0 + WVw + O(W 2 ), (63c)
If the contact angle are the same (ϕ = 0), the drop does not
we have the following problem at linear order move, therefore write the solution to Eq. (68) as a perturbation
ℓw power series of both ϕ and A:
Vw − [1 − exp (−g0 /ℓw )]
′′′ g0
gw = g
0
 (64a) g = g0 + ϕgϕ + Agα + Aϕgϕ,α + O(A2 , ϕ2 ) (69a)
ξ3 + ℓ u g0
3
L = L0 + ϕLϕ + ALα + AϕLϕ,α + O(A2 , ϕ2 ) (69b)

with gw (±1) = 0, gw (±1) = 0. (64b)
V = V0 + ϕVϕ + AVα + AϕVϕ,α + O(A2 , ϕ2 ). (69c)
The constant Vw can be determined from a solvability condition.
Using the test function t = 1 − y 2 we obtain The subscript ϕ denotes the correction at order ϕ due to asym-
√ metric contact angles for a passive drop (A = 0). It is solution of
ξℓw b
Vw = √ I (65)
φ arctanh(1/ b)
′′′ Vϕ
gϕ = g  (70a)
0
where I is the definite integral ξ3 + ℓ u g0
3
Z 1
{1 − exp[−d(1 − y 2 )]} ′
I= dy (66) with gϕ (±1) = 0, gϕ (±1) = 1/(2ξ) (70b)
−1 (1 − y 2 )(b − y 2 )

1–19 | 15
and we find after integration: Lϕ,α using a solvability condition and we obtain

1 βφ n
Lϕ,α = 0 (75a)
gϕ = y+ 2 2
(71a)
2ξ 2ξ(β − φ ) ln [(β + φ)/(β − φ)]
(
1 β+φ 1
(1 + y)2 ln(1 + y) − (1 − y)2 ln(1 − y) − 22 ln(2)y
 
Vϕ,α = 2 (75b)
12ξβ β − φ
 
β+φ
ln
1  β−φ
−(β + φy)2 ln(β + φy) + (β − φy)2 ln(β − φy)

+
βφ  
β+φ
− −2β 2 + (β − φ)2 dilog
 
2y o
β−φ
+ [(β + φ) ln(β + φ) + (β − φ) ln(β − φ)]
β  

− 2β 2 + (β − φ)2 dilog
 
Lϕ = 0 (71b) β+φ

βφ + β 2 − [ln(β + φ)]2 + [ln(β − φ)]2



Vϕ =   (71c)
β+φ
6 ln  
β−φ β+φ
+ 2β 2 + (β − φ)2 ln(2φ) ln
 
p β−φ
where we have introduced β = φ 1 + 6ℓu ξ/φ.  
β+φ
− (β − φ)2 [1 + ln(β + φ)] ln
β−φ
)
π2  2

β−φ
The subscript α refers to the correction at order A for a sym- 2β + (β − φ)2 − 2βφ

+ .
metric active drop (ϕ = 0, A 6= 0). It is the solution of 6 β+φ
g
Note that since A affects L, it also affects indirectly Vϕ (L). We

0
Vα + ξ + ℓu g0′
′′′
gα = g 3  (72a) can expand
3 0
ξ + ℓ u g0
3 dVϕ

2 2
Vϕ (L) = Vϕ (L0 ) + ALα + O(A , ϕ ), (76)
′ dL L0
with gα (±1) = 0, gα (±1) = 0 (72b)
therefore in the main text we write
and we find
1  V = ϕ (Vϕ,L0 + AVϕ,Lα ) + ϕAVϕ,α + O(A2 , ϕ̃2 ) (77)
gα = 2 (1 + y)2 ln(1 + y) + (1 − y)2 ln(1 − y)
2ξ dVϕ

where Vϕ,L0 = Vϕ (L0 ) and Vϕ,Lα = Lα .
dL L0
− (2 ln 2 + 1)y 2 − 2 ln 2 + 1

(73a)
E Including nematic stresses

Lα = (73b) n
2φ2 We now retain nematic stresses σij = −K∂i nk ∂j nk in the mo-
mentum balance at leading order. The additional relevant dimen-
Vα = 0 (73c) sionless parameter is N = K/(ηU L). Note that in the following,
m = m(h̃) is the regularizing function defined by Eq. (11), and
m′ = dm/dh̃.
The expressions of the shear stress [Eq. (14)] and of the ve-
locity [Eq. (15)] now contain nematic contributions. Those are
Finally gϕ,α is the correction at order Aϕ for an asymmetric
given by
active drop which satisfies
N π2 ω2 2
  
  ′ ′ h̃
2g0 (m − h̃mm )h̃ z̃ − if ω 6= 0,


Vϕ,α − ξ 3 + ℓu (gα gϕ ′′′ ′′′
+ gϕ g α ) 2

 2h̃3

′′′ 3 n
σ̃xz =
gϕ,α = g 
0
ξ3 2N ǫ2 h 2 i 
+ ℓ u g0 h̃


3 (m − h̃mm′ )h̃′3 − m2 h̃h̃′ h̃′′ z̃ − if ω = 0,



(74a) h̃3 2
1 g
0

(78)
ξgϕ g0′ + ξ + ℓ u gϕ′

+3 g 3 
N π2 ω2 2
0
  2 
ξ3 + ℓ u g0 ′ ′ z
3 (m − h̃mm )h̃ − h̃z̃ − ℓ̃u h̃ if ω 6= 0,


2

 4h̃3

n
ũx =
N ǫ2 h 2 i  z2 


with gϕ,α (±1) = 0, gϕ,α (±1) = 0 (74b) (m − h̃mm′ )h̃′3 − m2 h̃h̃′ h̃′′

− h̃z̃ − ℓ̃u h̃ if ω = 0.


2

h̃3
Here we do not solve Eq. (74), but determine instead Vϕ,α and (79)

16 | 1–19
(a) Self-propulsion driven by active stresses, Ã = (αL)/(2πωγǫ2 ), Ñ = (Kω 2 π 2 )/(4γLǫ3 )

(b) Self-propulsion driven by self-advection, W̃ = (ηw0 )/(γǫ3 ), Ñ = K/(γLǫ)

(c) Self-propulsion driven by capillarity, Ã = (αL)/(γǫ), Ñ = K/(γLǫ)

Fig. 6 Effect of nematic stresses on the drop shape and velocity (shaded drops are those obtained for Ñ = 0 in the main text), for each mode of
motion. Colored symbols/lines mark corresponding state points between left and right panels.

1–19 | 17
Averaging the velocity over the drop height gives: 3 E. K. Paluch, I. M. Aspalter and M. Sixt, Annu. Rev. Cell Dev.
Biol., 2016, 32, 469–490.
1 h̃ n
Z
ũx dz̃ = 4 K. I. Anderson, Y. L. Wang and J. V. Small, J. Cell Biol., 1996,
h̃ 0 134, 1209–18.

Nω π 2 2


 5 A. B. Verkhovsky, T. M. Svitkina and G. G. Borisy, Curr. Biol.,
− + ℓ̃u (m2 − h̃mm′ )h̃′ if ω 6= 0,

1999, 9, 11–20.

3


 4h̃2
6 P. T. Yam, C. A. Wilson, L. Ji, B. Hebert, E. L. Barnhart, N. A.
N ǫ2 h̃
  h
Dye, P. W. Wiseman, G. Danuser and J. A. Theriot, J. Cell Biol.,
 i
2 ′ ′3 2 ′ ′′
− + ℓ̃u (m − h̃mm )h̃ − m h̃h̃ h̃ if ω = 0.


h̃2 3 2007, 178, 1207–21.
(80) 7 K. Keren, Z. Pincus, G. M. Allen, E. L. Barnhart, G. Marriott,
A. Mogilner and J. A. Theriot, Nature, 2008, 453, 475–480.
Note that we must have N ∼ 1 for ω 6= 0 and N ∼ ǫ−2 for ω = 0
8 K. Kruse, J. F. Joanny, F. Jülicher and J. Prost, Phys. Biol.,
such that nematic stresses play a role at leading order.
2006, 3, 130–137.
The thin drop equation (21a) becomes
9 D. Shao, H. Levine and W.-J. Rappel, Proc. Natl. Acad. Sci.
 
h̃ USA, 2012, 109, 6851–6.
+ ℓ̃u h̃h̃ + Ãf˜ (h̃) + W̃ f˜ (h̃) + Ñ f˜ (h̃) = Ṽ
′′′ α w n
3 10 C. Blanch-Mercader and J. Casademunt, Phys. Rev. Lett., 2013,
   110, 078102.
1 h̃ 2 ′ ′
− + ℓ̃u (m − h̃mm )h̃ if ω 6= 0, 11 A. C. Callan-Jones and R. Voituriez, New J. Phys., 2013, 15,


 h̃2 3


˜n 025022.
f (h̃) =

 1 h̃
 h i 12 P. Recho, T. Putelat and L. Truskinovsky, Phys. Rev. Lett., 2013,
− + ℓ̃u (m2 − h̃mm′ )h̃′3 − m2 h̃h̃′ h̃′′ if ω = 0,


h̃ 2 3 111, 108102.
(81) 13 E. Tjhung, A. Tiribocchi, D. Marenduzzo and M. E. Cates, Nat.
Commun., 2015, 6, 5420.
with 14 D. Khoromskaia and G. P. Alexander, Phys. Rev. E, 2015, 92,
Kω 2 π 2



 4γLǫ
 3
if ω 6
= 0, 062311.
Ñ = (82) 15 T. Sanchez, D. T. N. Chen, S. J. DeCamp, M. Heymann and
 K Z. Dogic, Nature, 2012, 491, 431–434.


if ω = 0.

γLǫ 16 P.-G. de Gennes, F. Brochard-Wyart and D. Quéré, Capillar-
We show in Fig. 6 how Ñ affects the solutions presented in ity and Wetting Phenomena: Drops, Bubbles, Pearls, Waves,
the main text. Overall, increasing Ñ causes the drop to flatten. Springer-Verlag, New York, 2004.
For self-propulsion driven by active stresses [Fig. 6(a)], the trac- 17 D. Bonn, J. Eggers, J. Indekeu, J. Meunier and E. Rolley, Rev.
tionless flat solution at high à is essentially independent of Ñ . Mod. Phys., 2009, 81, 739–805.
For self-propulsion driven by self-advection [Fig. 6(b)], Ñ favors 18 F. Brochard, Langmuir, 1989, 5, 432–438.
the growth of the protrusion and simply renormalizes the tran- 19 S. Ramaswamy, Annu. Rev. Condens. Matter Phys., 2010, 1,
sition to a protruded shape (W̃c decreases upon increasing Ñ ). 323–345.
For self-propulsion driven by capillarity [Fig. 6(c)], increasing Ñ 20 M. C. Marchetti, J. F. Joanny, S. Ramaswamy, T. B. Liverpool,
dramatically reduces the drop speed. J. Prost, M. Rao and R. A. Simha, Rev. Mod. Phys., 2013, 85,
1143–1189.
Conflicts of interest
21 D. Saintillan and M. J. Shelley, C. R. Phys., 2013, 14, 497–517.
There are no conflicts to declare.
22 J. Prost, F. Jülicher and J.-F. Joanny, Nat. Phys., 2015, 11,
Acknowledgements 111–117.
23 F. Jülicher, S. W. Grill and G. Salbreux, Rep. Prog. Phys., 2018,
A.L. thanks Dmitry Zhdanov and Francisco Gonzalez Montoya for
81, 076601.
deriving (67). Part of this work was funded by a Leverhulme Trust
Research Project Grant RPG-2016-147. T.B.L. acknowledges sup- 24 R. J. Hawkins, R. Poincloux, O. Bénichou, M. Piel, P. Chavrier
port of BrisSynBio, a BBSRC/EPSRC Advanced Synthetic Biology and R. Voituriez, Biophys. J., 2011, 101, 1041–1045.
Research Centre (grant number BB/L01386X/1). 25 E. Tjhung, D. Marenduzzo and M. E. Cates, Proc. Natl. Acad.
Sci. USA, 2012, 109, 12381–6.
Notes and references 26 L. Giomi and A. DeSimone, Phys. Rev. Lett., 2014, 112,
1 T. Lämmermann and M. Sixt, Curr. Opin. Cell Biol., 2009, 21, 147802.
636–644. 27 C. A. Whitfield and R. J. Hawkins, New J. Phys., 2016, 18,
2 Y.-J. Liu, M. Le Berre, F. Lautenschlaeger, P. Maiuri, A. Callan- 123016.
Jones, M. Heuzé, T. Takaki, R. Voituriez and M. Piel, Cell, 28 F. Ziebert, S. Swaminathan and I. S. Aranson, J. Royal Soc.
2015, 160, 659–672. Interface, 2012, 9, 1084–1092.

18 | 1–19
29 S. Trinschek, F. Stegemerten, K. John and U. Thiele, arXiv e- 52 T. J. Pedley and J. O. Kessler, Annu. Rev. Fluid Mech., 1992,
prints, 2019, arXiv:1911.08258. 24, 313–358.
30 C. A. Whitfield, D. Marenduzzo, R. Voituriez and R. J. 53 T. Lämmermann, B. L. Bader, S. J. Monkley, T. Worbs,
Hawkins, Eur. Phys. J. E, 2014, 37, 8. R. Wedlich-Söldner, K. Hirsch, M. Keller, R. Förster, D. R.
31 W. Marth, S. Praetorius and A. Voigt, J. Royal Soc. Interface, Critchley, R. Fässler and M. Sixt, Nature, 2008, 453, 51–55.
2015, 12, 20150161. 54 R. Poincloux, O. Collin, F. Lizárraga, M. Romao, M. Debray,
32 A. Oron, S. H. Davis and S. G. Bankoff, Rev. Mod. Phys., 1997, M. Piel and P. Chavrier, Proc. Natl. Acad. Sci. USA, 2011, 108,
69, 931–980. 1943–8.
33 R. V. Craster and O. K. Matar, Rev. Mod. Phys., 2009, 81, 55 W.-C. Hung, S.-H. Chen, C. D. Paul, K. M. Stroka, Y.-C. Lo,
1131–1198. J. T. Yang and K. Konstantopoulos, J Cell Biol., 2013, 202,
34 M. Ben Amar and L. J. Cummings, Phys. Fluids, 2001, 13, 807–24.
1160–1162. 56 H. M. López, J. Gachelin, C. Douarche, H. Auradou and E. Clé-
35 L. J. Cummings, Eur. J. Appl. Math., 2004, 15, 651–677. ment, Phys. Rev. Lett., 2015, 115, 028301.
36 L. J. Cummings, T.-S. Lin and L. Kondic, Phys. Fluids, 2011, 57 S. Guo, D. Samanta, Y. Peng, X. Xu and X. Cheng, Proc. Natl.
23, 043102. Acad. Sci. USA, 2018, 115, 7212–7217.
37 T.-S. Lin, L. Kondic, U. Thiele and L. J. Cummings, J. Fluid 58 M. Dembo, T. Oliver, A. Ishihara and K. Jacobson, Biophys. J.,
Mech., 2013, 729, 214–230. 1996, 70, 2008–22.
38 T.-S. Lin, L. J. Cummings, A. J. Archer, L. Kondic and 59 J. L. Tan, J. Tien, D. M. Pirone, D. S. Gray, K. Bhadriraju and
U. Thiele, Phys. Fluids, 2013, 25, 082102. C. S. Chen, Proc. Natl. Acad. Sci. USA, 2003, 100, 1484–1489.
39 M. Crespo, A. Majumdar, A. M. Ramos and I. M. Griffiths, 60 O. du Roure, A. Saez, A. Buguin, R. H. Austin, P. Chavrier,
Physica D: Nonlinear Phenomena, 2017, 351-352, 1–13. P. Siberzan and B. Ladoux, Proc. Natl. Acad. Sci. USA, 2005,
102, 2390 –2395.
40 S. Sankararaman and S. Ramaswamy, Phys. Rev. Lett., 2009,
102, 118107. 61 N. Koumakis, C. Maggi and R. Di Leonardo, Soft Matter, 2014,
10, 5695–5701.
41 J.-F. Joanny and S. Ramaswamy, J. Fluid Mech., 2012, 705,
46–57. 62 O. Sipos, K. Nagy, R. Di Leonardo and P. Galajda, Phys. Rev.
Lett., 2015, 114, 258104.
42 G. Kitavtsev, A. Münch and B. Wagner, Proc. R. Soc. A, 2018,
474, 20170828. 63 Y. Yuan and Y. Zhang, Nanomedicine: Nanotechnology, Biology
and Medicine, 2017, 13, 2199–2207.
43 A. Loisy, J. Eggers and T. B. Liverpool, Phys. Rev. Lett., 2019,
123, 248006. 64 A. Muñoz-Bonilla, R. Cuervo-Rodríguez, F. López-Fabal, J. L.
Gómez-Garcés and M. Fernández-García, Materials, 2018, 11,
44 D. Shao, W.-J. Rappel and H. Levine, Phys. Rev. Lett., 2010,
1266.
105, 108104.
65 J. Hasan, R. J. Crawford and E. P. Ivanova, Trends Biotechnol.,
45 F. Ziebert and I. S. Aranson, PLoS ONE, 2013, 8, e64511.
2013, 31, 295–304.
46 P.-G. de Gennes, Physica A Stat. Mech. Appl., 1998, 249, 196–
66 R. Voituriez, J. F. Joanny and J. Prost, Europhys. Lett., 2005,
205.
70, 404–410.
47 M. K. Chaudhury and G. M. Whitesides, Science, 1992, 256,
67 T. P. Loisel, R. Boujemaa, D. Pantaloni and M.-F. Carlier, Na-
1539–1541.
ture, 1999, 401, 613–616.
48 C. D. Bain, G. D. Burnett-Hall and R. R. Montgomerie, Nature,
68 D. B. Owen, Ann. Math. Stat., 1956, 27, 1075–1090.
1994, 372, 414–415.
69 NIST Handbook of Mathematical Functions, ed. F. W. J. Olver,
49 F. D. Dos Santos and T. Ondarçuhu, Phys. Rev. Lett., 1995, 75,
D. W. Lozier, R. F. Boisvert and C. W. Clark, Cambridge Uni-
2972–2975.
versity Press, 2010.
50 N. J. Cira, A. Benusiglio and M. Prakash, Nature, 2015, 519,
70 J. Eggers and M. A. Fontelos, Singularities: Formation, Struc-
446–450.
ture, and Propagation, Cambridge University Press, Cam-
51 P.-G. de Gennes and J. Prost, The Physics of Liquid Crystals,
bridge, 2015.
Oxford University Press, Oxford, 2nd edn, 1993.
71 L. M. Pismen and J. Eggers, Phys. Rev. E, 2008, 78, 056304.

1–19 | 19

You might also like