You are on page 1of 7

Loopy Lévy flights enhance tracer diffusion in active suspensions

Kiyoshi Kanazawa∗
Faculty of Engineering, Information and Systems,
University of Tsukuba, Tennodai, Tsukuba, Ibaraki 305-8577, Japan

Tomohiko G. Sano
Flexible Structures Laboratory, Institute of Mechanical Engineering,
École polytechnique fédérale de Lausanne, Lausanne, CH-1015, Switzerland.

Andrea Cairoli
arXiv:1906.00608v1 [cond-mat.stat-mech] 3 Jun 2019

Department of Bioengineering, Imperial College London, London SW7 2AZ, UK

Adrian Baule
School of Mathematical Sciences, Queen Mary University of London, London E1 4NS, UK
(Dated: June 4, 2019)

Brownian motion is widely used as a paradig- Lévy flight model [11] from a physical microscopic
matic model of diffusion in equilibrium media dynamics, but can also be applied to address im-
throughout the physical, chemical, and biologi- portant conceptual questions, such as the ther-
cal sciences. However, many real world systems, modynamics of active systems [13], and practical
particularly biological ones, are intrinsically out- ones regarding, e.g., the interaction of swimming
of-equilibrium due to the energy-dissipating ac- microorganisms with nutrients and other small
tive processes underlying their mechanical and particles like degraded plastic [14] and the design
dynamical features [1]. The diffusion process fol- of artificial nanoscale machines [15].
lowed by a passive tracer in prototypical active
media such as suspensions of active colloids or A passive tracer immersed in a fluid at equilibrium
swimming microorganisms [2] indeed differs sig- moves randomly due to its collisions with the surround-
nificantly from Brownian motion, manifest in a ing fluid molecules. Understanding how the observed
greatly enhanced diffusion coefficient [3–10], non- stochastic process followed by the tracer relates to the
Gaussian tails of the displacement statistics [6, 9, statistical mechanics of the surrounding fluid, as accom-
10], and crossover phenomena [9, 10] from non- plished in the seminal works by Einstein, Smoluchowski,
Gaussian to Gaussian scaling. While such char- and Langevin [16], has provided deep insight into the
acteristic features have been extensively observed connection between molecular transport and equilibrium
in experiments, there is so far no comprehensive thermodynamics , which has been widely exploited to
theory explaining how they emerge from the mi- describe soft matter and other complex physical systems
croscopic active dynamics. Here we present a the- [17]. However, when either artificial self-propelled col-
oretical framework of the enhanced tracer diffu- loids or biological swimming micro-organisms, such as
sion in an active medium from its microscopic dy- bacteria like Escherichia coli or algae like Volvox and
namics by coarse-graining the hydrodynamic in- Chlamydomonas reinhardtii [2], are also suspended, the
teractions between the tracer and the active par- diffusion of the tracer changes dramatically due to the
ticles as a stochastic process. The tracer is shown active stirring of the fluid exerted by the self-propelled
to follow a non-Markovian coloured Poisson pro- particles. Indeed, the active diffusion of the tracer ex-
cess that accounts quantitatively for all empirical perimentally exhibits the following unique features that
observations. The theory predicts in particular a can no longer be explained as a Brownian motion: (i) the
long-lived Lévy flight regime [11] of the tracer mo- tracer exhibits dancing loopy trajectories [4, 8, 9]; (ii) its
tion with a non-monotonic crossover between two mean square displacement (MSD) exhibits a crossover
different power-law exponents. The duration of between superdiffusion with ∼ tα (1 < α ≤ 2) for short
this regime can be tuned by the swimmer density, times and normal diffusion (α = 1) for long times, where
thus suggesting that the optimal foraging strat- the effective diffusion coefficient D is greatly enhanced
egy of swimming microorganisms might crucially compared with that of the equilibrium case, D0 , revealing
depend on the density in order to exploit the for both three and quasi-two dimensional systems a linear
Lévy flights of nutrients [12]. Our framework not dependence on the density of swimmers ρ: D = D0 + Bρ,
only provides the first validation of the celebrated with the coefficient B being system dependent [3–5, 7–
9]; (iii) the probability density function (PDF) P∆t of
position displacements in a given time interval ∆t ex-
hibits strong non-Gaussian features manifest as power-
∗ kiyoshi@sk.tsukuba.ac.jp law tails [6, 10]; (iv) P∆t eventually reverts to a Gaus-
2

sian shape for large ∆t [9, 10]; (v) the associated non- Stokes equation. For force- and torque-free swimmers,
Gaussian parameter exhibits a scaling regime ∼ ∆t−1 for the leading-order term of this solution in a far-field ex-
large times [10]. pansion is a dipole force [26]. Therefore, F is specified
Developing a single theory that captures all features without loss of generality as the stresslet
(i)–(v) has been a major challenge, due to the multi-
(ni · ri )2
 
particle hydrodynamic, and thus long-range, interactions p ri
F (ri , ni ) ≈ 2 3 2 − 1 (2)
of the tracer with the swimmers underlying its transport. ri ri ri
While the loop-like motion (i) results from an individual
for ri > d, with the difference vector ri = xi − X, the
scattering event of the tracer in the dipolar flow field of
dipole strength p and the system specific cut-off d. The
a single swimmer [7, 18, 19], and the linear form of D
dipole strength parameter p specifies the universal fea-
(ii) has been explained phenomenologically based on the
tures of the far-flow hydrodynamic field [26]: p < 0 de-
active flux of the swimmers, which is defined as the prod-
notes pusher swimmers whose flow lines are oriented out-
uct of their number density and characteristic swimming
ward along the direction of its velocity vector and inward
speed [5, 7, 8, 19, 20], the statistical observations (iii)–(v)
laterally (e.g., E. coli [26]); p > 0 denotes instead puller
could so far not be explained consistently. The power-
swimmers whose flow lines are oriented in the opposite
law tails in P∆t and their convergence to Gaussian scal-
directions (e.g., Chlamydomonas [26]). The length pa-
ing for long observation times, which is expected based
rameter d separates the far-flow field from any near-flow
on the central limit theorem (CLT) arguments [21, 22],
field hydrodynamic contributions and hard-core interac-
have been reproduced in [10, 23, 24] assuming a static
tions, and thus determines when the approximation (2)
force distribution akin to the Holtsmark theory of gravi-
is valid.
p The model has one further length parameter
tating particles (see SI for a review) [25]. However, this
b∗ ≡ |p|/ΓvA , which can be related to the typical
approach neglects any dynamics of the swimmers and is
length scale of the swimmer [26]. All these parameters
thus not sufficient to capture the enhanced diffusion ob-
can be determined experimentally: for E. coli d ' 6 µm
served in experiments [3–10]. Here, we present a deriva-
and b∗ ' 2 µm [5, 27]; for Chlamydomonas d ' 35 µm
tion of the stochastic process underlying the enhanced
and b∗ ' 8 µm [10, 23, 24, 28], satisfying d ≥ b∗ .
diffusion of the tracer from microscopic dynamics that is
Conversely, for ri ≤ d the interaction force F is not
valid at all timescales. The resulting process captures all
universal but system-specific. Nevertheless, all swimmer-
characteristic features (i)–(v) and is in excellent quanti-
tracer interactions in this regime can be accurately cap-
tative agreement with simulation data.
tured by using arguments based on the CLT (see below),
We consider a three-dimensional system composed of
which do not require a detailed form of F . Therefore, in
m active particles and a passive tracer suspended in a vis-
our present implementation we set for simplicity F = 0
cous fluid in a cubic box of linear length L (Fig. 1a). The
in this regime, but we remark that our qualitative results
active particles (swimmers) are assumed self-propelled
are independent of this specific choice and that this ap-
and their motion is driven by unidirectional forces with
proximation has also been validated experimentally [10].
constant amplitude [26]. In general, the dynamics of such
Numerical simulations of Eqs. (1,2) in dilute conditions
multi-particle systems where interactions are mediated
reproduce all features (i)–(v) (see Fig. 1c and Fig. 3).
by the fluid environment in the form of hydrodynamic
Crucially, an inspection of time-series trajectories under
forces are complex and analytically intractable. How-
such dilute conditions shows that the dynamics of the
ever, suspensions of micro-organisms often considered in
tracer can be resolved as a sequence of individual scat-
experiments (see above) are generally characterized by
tering events, where only two-body tracer-swimmer inter-
(a) low Reynolds number swimming and (b) a low den-
actions are relevant (Fig. 2a. See also SI). In the kinetic
sity of swimmers (dilute conditions, see below). In par-
theory of gases the dilute condition is ensured by requir-
ticular the dilute condition (b) allows us to neglect the
ing ρrc3  1, where ρ is the number density and rc the
mutual hydrodynamic interactions of the swimmers [27],
range of interparticle interactions. In our system, how-
thus leading to the overdamped equations of motion:
ever, the hydrodynamic force is long-range such that an
m interaction range cannot be well defined. Nevertheless,
dxi dX X
we can identify rc with the maximum geometrical length
= vA ni , Γ = F (xi − X, ni ), (1)
dt dt i=1 parameter as rc ≡ max{d, b∗ } = d and define dilute con-
ditions accordingly. The motivation behind this defini-
where F is the force on the tracer generated by a sin- tion of rc is two-fold. Firstly, with this definition the
gle swimmer and Γ is a viscous coefficient for the passive condition ρd3  1 is indeed realized in experiments that
particle. In Eq. (1), xi (t) and X(t) denote the posi- exhibit the features (i)–(v) (e.g., ρd3 ∼ 0.15 in [10] and
tions of the i-th active particle and the passive particle, ρd3 ∼ 0.1 in [5]). Secondly, this parameter regime allows
respectively, and vA is the constant amplitude of the ac- for a self-consistent description of the tracer dynamics,
tive self-propulsion force with unit vector ni specifying in which the dipole interaction governs the displacement
the swimming direction. statistics on short and intermediate time scales, while
The low Reynolds number condition (a) further yields the statistics for longer times reverts to a Gaussian form
a closed form expression for F as the solution of the due to the CLT. To this end, we note that in a dilute
3

(a) (b) (c)


Scattering

Collision

FIG. 1. (a) Schematic of the microscopic model: active particles (green ellipsoids) and a passive tracer (filled violet sphere)
are suspended in a cubic box of linear length L. The direction of the active swimmers ni ≡ (sin θi cos φi , sin θi sin φi , cos θi ) is
randomized upon hitting the box boundary. We finally take the large system-size limit keeping the number density of active
particles ρ ≡ N/L3 constant (see SI for protocol details). The bead is dragged by the hydrodynamic flow fields generated by
the swimming of active particles. (b) Exemplary two-body scattering event with impact parameter b > d and injection angles
θ, φ, and φ0 , unit vector eθ , and collision event with b ≤ d. The force shape function during scattering is thus characterized
by the parameter set b ≡ (b, θ, φ, φ0 ). An active particle travels the distance vA ∆t in a time interval ∆t. This distance is equal
to the characteristic lengthscale b∗ at the time τH such that vA τH = b∗ ⇐⇒ τH ≡ b∗ /vA . Since b∗ is related to the typical
length scale of the active swimmers, for ∆t  τH their motion can be effectively neglected (see SI). Conversely, for a given set
of injection angles active particles can collide against the tracer in the time interval ∆t if they are contained in a cylinder with
cross section area πd2 and linear length vA ∆t surrounding the passive particle. The mean free time of the tracer τC is then
estimated as ρπd2 vA τC = 1 ⇐⇒ τC ≡ 1/ρπd2 vA . For ∆t  τC such collision events do not contribute to the tracer dynamics.
(c) Typical trajectory of the passive tracer observed in simulations, exhibiting its dancing Lévy motion in the case of pushers
(p < 0 and d = b∗ ). Once scattered by an active swimmer, the tracer exhibits a triangular-shaped but non-closed trajectory in
the plane specified by the injection angles of the active swimmer θ, φ, and φ0 (dashed lines). We choose b∗ as unit for lengths.

(a) (b) (c) 0.4×10-3


-4
3×10 0
0 0 τ = 26.35
-0.4 -0.4×10-3 b = 27.20
θ = 2.14
-3×10-4 5 -0.8 φ = 0.02
4 -0.8×10-3 φ' = 0.93
-4
3×10 3
2
1 10 5 0 -5 -10 -1.2×10-3
11200 11600 12000
0
0.3×10-4 τ = -150.21
-3×10-4 0.2×10-4 b = 152.23
θ = 0.68
3×10-4 φ = 3.36
0.2 0.1×10-4 φ' = 3.48
0 0 0
-0.2
-3×10-4 5 -0.1×10-4
4 -0.4
0 4
3×10 6×10 9×104 4 3 -10
2 0 -5 74000 76000 78000
1 10 5

FIG. 2. We choose b∗ , τH as units for length and time respectively. (a) Typical time series of force exerted on the tracer by the
active swimmers in dilute conditions under the parameters p < 0 (pushers) and d = b∗ . (b) 3D plots of the force shape function
for the stresslet hydrodynamic force in Eq. (2), obtained by perturbatively solving the two-body scattering problem (SI). We
set ΓvA = 1, and {θ, φ, φ0 } = {π/2, 0, π/2} (top panel) or {θ, φ, φ0 } = {0, 0, 0} (bottom panel). The general formula is obtained
by linearly combining these two solutions weighted by coefficients that depend only on the angular variables. (c) Exemplary fits
(black solid line) of the force shape function fb (t) to simulation data of force exerted on the tracer in the direction ex during
different scattering events arbitrarily extracted from force time-series. Fit parameters are the set b and the scattering time τ .
The fit is obtained by using non-linear least squares. The agreement is excellent.
4

system, at every instant in time, swimmers on average tional techniques (see SI), we derive the displacement
have b  d, where b is introduced as the impact parame- PDF P∆t (|∆X|) and obtain its scaling behaviours in the
ter of a binary swimmer-tracer interaction (see Fig. 1b). regimes (1.)–(3.) as
The tracer statistics will then be governed by three dis- 
tinct dynamical regimes: (1.) For short times ∆t  τH , −α
|∆X| H
 (∆t  τH )
the tracer experiences the long-range forces of effectively P∆t (|∆X|) ∝ |∆X| S −α
(τH  ∆t  τC ) (4)
static swimmers as in the Holtzmark theory (“Holtzmark e−∆X 2 /2σ2

(τC  ∆t)
regime”). Accordingly, τH ≡ b∗ /vA . (2.) For times
τH  ∆t  τC , the tracer is scattered by the moving
with power-law exponents αH = 5/2, αS = 5/3 and pos-
swimmers in a sequence of binary interactions that we
itive constant σ 2 that depends on d (see SI). For finite
call “scatterings” (“scattering regime”). Since the sys-
d a truncation appears in the power-law tails, which is
tem is dilute, the swimmer-tracer interaction is governed
realistic because any physical system must accommodate
by the far-flow field as in Eq. (2) in this regime. The
finite cutoffs [30].
time scale τC ≡ 1/ρvA πd2 (see Fig. 1b) thus estimates
The Holtsmark regime (1.) yields the same scaling
the time necessary for a swimmer to arrive close enough
behaviour as found in other purely static approaches
to the tracer to interact via hard-core and near-field hy-
|∆X|−αH [10, 23, 24]. For ∆t & τH , these theories are
drodynamic interactions. (3.) For ∆t  τC , these inter-
not applicable because the rearrangements of the active
action events thus represent “collisions” and the tracer is
swimmers are no longer negligible. In fact, the force
displaced by an accumulation of such collisions with the
induced during binary scatterings has to be fully taken
swimmers such that the CLT applies (“CLT regime”).
into account, and the stochastic process X is then non-
Remarkably, these three regimes are captured by a
Markovian (see SI). In the scattering regime (2.), the
coarse-grained description of the tracer dynamics in
description nevertheless becomes effectively Markovian
terms of the Langevin equation
as the FSF can be approximated effectively as a Dirac
N (t) δ-function (see SI). In this regime the coloured Poisson
dX X
model (3) is equivalent to a compound Poisson process
Γ = F (t), F (t) ≡ fbi (t − τi ), (3)
dt i=1
with jump-length distribution prescribed as a truncated
power-law with scaling behaviour |∆X|−αS . Our results
where N (t) counts the number of scattering events up thus validate the well known Lévy flight model [11] as an
to a fixed time t and fb (t) is the force shape function approximate description of the tracer dynamics at this
(FSF) describing the force exerted on the tracer during timescale. A related jump-diffusion process underlying
each scattering. The transition from the fully determinis- the enhanced diffusion has previously been proposed in
tic dynamics of Eqs. (1, 2) to a stochastic description by [9] on purely phenomenological grounds. In the collision
Eq. (3) is realised by assuming N (t) to be a Poisson pro- regime (3.), collisions become dominant over scatterings.
cess with intensity λ(b). The FSF fb (t) is characterized Since collisional impact has finite cutoff, accumulation
by the set of impact parameter and injection angles (de- of a sufficient number of collisions leads to the Gaussian
noted by b in Fig. 2b) and is obtained in analytical form tail as a consequence of the central limit theorem. We
by solving the binary swimmer–tracer scattering prob- note that the detailed form of F for ri ≤ d is renormal-
lem. Remarkably, for b  b∗ (valid on average in dilute ized into the variance σ 2 , though it is irrelevant to the
conditions and in the Holtzmark and scattering regimes, power-law tail (e.g., αH and αS ). Overall, the striking
see above) an analytical approximation of the FSF can non-monotonic behaviour of the scaling exponents of the
be obtained using a Picard iteration up to 2nd order and tracer displacement statistics as predicted by Eq. (4) is
subsequent Taylor expansion (see SI), which is in excel- in excellent agreement with simulation results (Fig. 3a).
lent agreement with numerics (Fig. 2c). The FSF is cen- Using the exact expression for P∆t (SI), the tracer
tred at the scattering time point τ , which is set by the MSD for different concentrations of active swimmers can
condition n·r = 0. The intensity can be shown to satisfy be shown to collapse onto the same universal curve upon
λ(b) ∝ b directly from its microscopic dynamics (see SI), rescaling by 1/ρb∗5 , which reveals a crossover from bal-
and the total intensity
R diverges in the large system size listic motion ∼ (∆t/τH )2 for short timescales to normal
limit as limL→∞ dbλ(b) = ∞. Consequently, F is a diffusion ∼ ∆t/τH for longer ones (see Fig. 3b). This
2
coloured Poisson noise with infinite intensity, also known yields in the ballistic regime: MSD(∆t) ∝ ρvA ∆t2 . Like-
as generalized Campbell’s process [29]. This infinite in- wise, in the diffusive regime we find: MSD(∆t) ∝ ρvA ∆t.
tensity is a typical singular character of a Lévy process, Eq. (3) thus predicts the enhancement of the tracer dif-
and is a physical consequence of the long-range hydro- fusive motion at all time scales. While in the diffusive
dynamic interactions, which cause an infinite number of regime it captures the linear dependence of the diffusion
small scatterings at possibly infinite impact parameter. coefficient D on the active flux ρvA that has been vali-
We now calculate the statistics resulting from Eq. (3) dated experimentally [3–5, 7–9], in the ballistic regime it
2
for the tracer displacement in the x-direction ∆X ≡ predicts a dependence of the MSD on ρvA .
[X(t + ∆t) − X(t)] · ex in agreement with the experi- Likewise, we can also calculate the non-Gaussian
mental protocols typically used [4, 6, 9, 10]. Using func- parameter for the tracer displacement statistics
5

(a) (b) (c)

FIG. 3. (a) The numerical distribution P∆t (|∆X|) for the displacement of the tracer ∆X, showing power-law tails. The tails
particularly exhibit a non-monotonic behaviour, becoming fatter from the Holtsmark regime (∆t  τH ), where P∆t ∼ |∆X|−αH ,
to the scattering regime (τH  ∆t  τC ), where instead P∆t ∼ |∆X|−αS . This striking behaviour of the enhanced diffusion
process is predicted by our theory (see Eq. (4)) and should be observable experimentally. (b) The mean-squared displacement
is plotted for various swimmer densities ρ, and shown to collapse onto a single master curve upon rescaling by 1/ρb∗5 . One
observes a clear crossover between ballistic and normal diffusive motion [3, 4, 6, 7, 10]. (c) The non-Gaussian parameter rescaled
by ρb∗3 is plotted for various densities, thus showing the universal power-law decay ∆t−1 consistently with the experimental
observations in [10].

NGP(∆t) ≡ h∆X 4 i/3h∆X 2 i2 − 1. Upon rescaling of the hydrodynamic interaction force, αS depends cru-
by ρb∗3 , simulation data for different concentrations of cially on its detailed form (i.e., the existence of loops).
active swimmers indeed collapse onto the same curve,
that scales as (∆t/τH )−1 for long observation times (see The enhanced diffusion process as determined by
Fig. 3c). This scaling prediction is confirmed by the (i)–(v) is a linear-in-time diffusion process with non-
model that yields NGP(∆t) ∝ (ρvA )−1 ∆t−1 (see SI). Gaussian displacement statistics. Such processes rep-
resent a ubiquitous feature of non-equilibrium diffusion
The tracer is also subjected to thermal fluctuations
phenomena [31], but yet they pose considerable chal-
that can be included in the model (3) as Brownian motion
lenges already for purely phenomenological modelling
(with variance 2kB ΓT , the Boltzmann constant kB and T
approaches [32]. The fact that the stochastic process
the temperature of the bath without swimmers). While
Eq. (3) captures all its characteristic features is thus re-
the scaling predictions (4) for the tracer displacement
markable, even more so since it is derived through an
distribution and that for the NGP are preserved, thermal
exact coarse-graining of microscopic dynamics. In con-
noise breaks the data collapse for both the MSD and
trast to many models that solve this “Brownian yet non-
the NGP and shifts the MSD by a term 2D0 ∆t with
Gaussian” diffusion conundrum, Eq. (3) is easily inter-
D0 ≡ kB T /Γ (see SI for details).
pretable physically due to the well-defined force pulse
Our approach can be extended straightforwardly to nature. Models of this type might thus be applicable
any dimensional systems. Further, more general hydro- much more generally. A striking prediction of our theory
dynamic force fields can be readily incorporated than is the intermediate fatter tail of the tracer displacement
Eq. (2), such as higher-order terms in the far-field ex- distribution in the scattering regime compared with the
pansion (e.g., the quadrupole). Let us assume the scal- Holtsmark regime, which is counter-intuitive since the
ing relations for the far-flow hydrodynamic interaction tails are typically expected to become thinner monoton-
force F ∝ 1/rinH (or equivalently, fb ∝ 1/bnH ) and for ically for longer times due to the CLT. Such signature
the
R ∞ net tracer displacement during a scattering event should be experimentally detectable. Moreover, our re-
nS
−∞
f b (t)dt ∝ 1/b (see SI). As a general recipe, our sults make predictions on the possible foraging behaviour
theory predicts power-law crossover with αH = D/nH of real swimming micro-organisms like Chlamydomonas.
and αS = (D − 1)/nS , considering that λ(b)|Jb | ∝ bD−2 On the one hand, τC diverges for ρ → 0, which implies
for D-dimensional active flows. This suggests that, while that the power-law tail persists for longer timescales the
αH is universally determined by the power-law exponent more dilute the system is. On the other hand, Lévy
6

flights have been shown to increase encounter probabil- gies [35] more efficient for the forager. On the contrary,
ities in the stochastic search for sparsely and randomly at low densities such displacements are power-law dis-
distributed revisitable targets [33, 34]. These results thus tributed, such that it might be advantageous for the for-
suggest that for swimming micro-organisms an optimal ager to simply wait for a nutrient to come close dragged
foraging strategy should significantly depend on the den- by the other swimmers [12]. Finally, we remark that su-
sity. For large densities the displacements of nutrients are perimposing additional force fields might lead to novel
primarily Gaussian (i.e. spatially localized), which make mechanisms to control and exploit enhanced diffusion in
an active search like intermittent Lévy-Brownian strate- artificial devices.

[1] Needleman, D. & Dogic, Z. Active matter at the interface [16] Gardiner, C. Stochastic methods. Springer Series in
between materials science and cell biology. Nat. Rev. Synergetics (Springer-Verlag, Berlin, 2009) (1985).
Mat. 2, 17048 (2017). [17] Coffey, W. T. & Kalmykov, Y. P. The Langevin equation:
[2] Koch, D. L. & Subramanian, G. Collective hydrodynam- with applications to stochastic problems in physics, chem-
ics of swimming microorganisms: Living fluids. Annu. istry and electrical engineering (World Scientific, 2004).
Rev. Fluid Mech. 43, 637–659 (2011). [18] Dunkel, J., Putz, V. B., Zaid, I. M. & Yeomans, J. M.
[3] Wu, X.-L. & Libchaber, A. Particle diffusion in a quasi- Swimmer-tracer scattering at low Reynolds number. Soft
two-dimensional bacterial bath. Phys. Rev. Lett. 84, 3017 Matter 6, 4268–4276 (2010).
(2000). [19] Lin, Z., Thiffeault, J.-L. & Childress, S. Stirring by
[4] Leptos, K. C., Guasto, J. S., Gollub, J. P., Pesci, A. I. & squirmers. J. Fluid Mech. 669, 167–177 (2011).
Goldstein, R. E. Dynamics of enhanced tracer diffusion [20] Morozov, A. & Marenduzzo, D. Enhanced diffusion of
in suspensions of swimming eukaryotic microorganisms. tracer particles in dilute bacterial suspensions. Soft Mat-
Phys. Rev. Lett. 103, 198103 (2009). ter 10, 2748–2758 (2014).
[5] Miño, G. et al. Enhanced diffusion due to active swim- [21] Pushkin, D. O. & Yeomans, J. M. Fluid mixing by curved
mers at a solid surface. Phys. Rev. Lett. 106, 048102 trajectories of microswimmers. Phys. Rev. Lett. 111,
(2011). 188101 (2013).
[6] Kurtuldu, H., Guasto, J. S., Johnson, K. A. & Gollub, [22] Thiffeault, J.-L. Distribution of particle displacements
J. P. Enhancement of biomixing by swimming algal cells due to swimming microorganisms. Phys. Rev. E 92,
in two-dimensional films. Proc. Natl. Acad. Sci. 108, 023023 (2015).
10391–10395 (2011). [23] Zaid, I. M., Dunkel, J. & Yeomans, J. M. Lévy fluc-
[7] Mino, G. L., Dunstan, J., Rousselet, A., Clément, E. tuations and mixing in dilute suspensions of algae and
& Soto, R. Induced diffusion of tracers in a bacterial bacteria. J. Royal Soc. Interface 8, 1314–1331 (2011).
suspension: theory and experiments. J. Fluid Mech. 729, [24] Zaid, I. & Mizuno, D. Analytical limit distributions from
423–444 (2013). random power-law interactions. Phys. Rev. Lett. 117,
[8] Jepson, A., Martinez, V. A., Schwarz-Linek, J., Morozov, 030602 (2016).
A. & Poon, W. C. K. Enhanced diffusion of nonswimmers [25] Holtsmark, J. Über die verbreiterung von Spektrallinien.
in a three-dimensional bath of motile bacteria. Phys. Rev. Ann. Phys. (Berl.) 363, 577–630 (1919).
E 88, 041002 (2013). [26] Lauga, E. & Powers, T. R. The hydrodynamics of
[9] Jeanneret, R., Pushkin, D. O., Kantsler, V. & Polin, swimming microorganisms. Rep. Prog. Phys. 72, 096601
M. Entrainment dominates the interaction of microal- (2009).
gae with micron-sized objects. Nat. Commun. 7, 12518 [27] Drescher, K., Dunkel, J., Cisneros, L. H., Ganguly, S. &
(2016). Goldstein, R. E. Fluid dynamics and noise in bacterial
[10] Kurihara, T., Aridome, M., Ayade, H., Zaid, I. & cell–cell and cell–surface scattering. Proc. Natl. Acad.
Mizuno, D. Non-Gaussian limit fluctuations in active Sci. 108, 10940–10945 (2011).
swimmer suspensions. Phys. Rev. E 95, 030601 (2017). [28] Drescher, K., Goldstein, R. E., Michel, N., Polin, M. &
[11] Hughes, B. D., Shlesinger, M. F. & Montroll, E. W. Ran- Tuval, I. Direct measurement of the flow field around
dom walks with self-similar clusters. Proc. Natl. Acad. swimming microorganisms. Phys. Rev. Lett. 105, 168101
Sci. 78, 3287–3291 (1981). (2010).
[12] Bartumeus, F., Catalan, J., Fulco, U., Lyra, M. & [29] Campbell, N. The study of discontinuous phenomena. In
Viswanathan, G. Optimizing the encounter rate in bi- Proc. Camb. Philos. Soc., vol. 15, 117–136 (1909).
ological interactions: Lévy versus Brownian strategies. [30] Mantegna, R. N. & Stanley, H. E. Stochastic process
Phys. Rev. Lett. 88, 097901 (2002). with ultraslow convergence to a Gaussian: the truncated
[13] Kanazawa, K., Sano, T. G., Sagawa, T. & Hayakawa, H. Lévy flight. Phys. Rev. Lett. 73, 2946 (1994).
Minimal model of stochastic athermal systems: Origin of [31] Wang, B., Kuo, J., Bae, S. C. & Granick, S. When
non-Gaussian noise. Phys. Rev. Lett. 114, 090601 (2015). Brownian diffusion is not Gaussian. Nature Mater. 11,
[14] Goldstein, R. E. Green algae as model organisms for 481–485 (2012).
biological fluid dynamics. Annu. Rev. Fluid Mech. 47, [32] Chechkin, A. V., Seno, F., Metzler, R. & Sokolov, I. M.
343–375 (2015). Brownian yet non-Gaussian diffusion: From superstatis-
[15] Bechinger, C. et al. Active particles in complex and tics to subordination of diffusing diffusivities. Phys. Rev.
crowded environments. Rev. Mod. Phys. 88, 045006 X 7, 021002 (2017).
(2016).
7

[33] Viswanathan, G. M. et al. Optimizing the success of Theoretical Physics in Kyoto University.
random searches. Nature 401, 911 (1999).
[34] Humphries, N. E., Weimerskirch, H., Queiroz, N.,
Southall, E. J. & Sims, D. W. Foraging success of bi-
AUTHOR CONTRIBUTIONS
ological Lévy flights recorded in situ. Proc. Natl. Acad.
Sci. 109, 7169–7174 (2012).
[35] Bénichou, O., Loverdo, C., Moreau, M. & Voituriez, R. KK conceived the original idea. KK, TGS, AC and
Intermittent search strategies. Rev. Mod. Phys. 83, 81 AB designed research. KK and AC performed analyti-
(2011). cal calculations. TGS performed numerical simulations.
KK, AC, and AB wrote the paper.

ACKNOWLEDGMENTS
COMPETING INTERESTS
We appreciate D. Mizuno, H. Takayasu, M. Takayasu,
H. Hayakawa, and F. van Wijland for fruitful discus-
The authors declare no competing financial interests.
sions. This work was supported by Grant-in-Aid for
JSPS Fellows (Grant No. 16J05315), JSPS KAKENHI
(Grant Nos. 16K16016 and 18K13519), the Research Fel-
lowship granted by the Royal Commission for the Exhibi- DATA AVAILABILITY
tion of 1851, and Atoms program granted by the Yukawa
Institute for Theoretical Physics. The numerical calcula- We will provide all the numerical codes and plot files
tions were carried out on XC40 at Yukawa Institute for upon requests.

You might also like