Sie sind auf Seite 1von 12

Collective behavior of penetrable self-propelled rods in two dimensions

Masoud Abkenar, Kristian Marx, Thorsten Auth, and Gerhard Gompper


Theoretical Soft Matter and Biophysics, Institute of Complex Systems and Institute for Advanced Simulation,
Forschungszentrum Julich, D-52425 Julich, Germany

Collective behavior of self-propelled particles is observed on a microscale for swimmers such as


sperm and bacteria as well as for protein filaments in motility assays. The properties of such systems
depend both on their dimensionality and the interactions between their particles. We introduce a
model for self-propelled rods in two dimensions that interact via a separation-shifted Lennard-Jones
potential. Due to the finite potential barrier, the rods are able to cross. This model allows us to
efficiently simulate systems of self-propelled rods that effectively move in two dimensions but can
occasionally escape to the third dimension in order to pass each other. Our quasi-two-dimensional
self-propelled particles describe a class of active systems that encompasses microswimmers close to a
wall and filaments propelled on a substrate. Using Monte Carlo simulations, we first determine the
isotropic-nematic transition for passive rods. Using Brownian dynamics simulations, we characterize
cluster formation of self-propelled rods as a function of propulsion strength, noise, and energy barrier.
Contrary to rods with an infinite potential barrier, an increase of the propulsion strength does not
only favor alignment but also effectively decreases the potential barrier that prevents crossing of rods.
arXiv:1309.2829v2 [cond-mat.soft] 6 Jan 2014

We thus find a clustering window with a maximum cluster size at medium propulsion strengths.

PACS numbers: 82.70.y, 47.63.Gd, 87.18.Hf, 64.70.M

I. INTRODUCTION Motility assays with actin filaments or microtubules


are essentially two-dimensional systems, but with a fi-
nite probability for the filaments to cross each other
Collective behavior of active bodies is frequently found
[15, 47]. Because the filaments are not tightly bound to
in macroscopic systems such as bird flocks and fish
the surface, one of them might be slightly and temporar-
schools [1], but also is found in microscopic systems
ily pushed away from the surface when two filaments
such as sperm cells [2, 3], bacteria [47], and manmade
collide. In Ref. [15], microtubules have been found to
microswimmers that propel themselves forward using a
cross each other with a probability of 40 % if they ap-
chemical or physical mechanism [812]. Despite the dif-
proach perpendicularly. Two-dimensional models with
ferent natures of these systems, they all exhibit interac-
impenetrable swimmers thus do not adequately describe
tions that favor alignment of neighboring bodies, thus
these systems, while full three-dimensional calculations
leading to similar forms of collective behavior. Of partic-
are computationally expensive. In Ref. [14], a cellular
ular interest for us are experiments with elongated self-
automaton model with an imposed alignment rule that
propelled particles on the microscopic scale in two dimen-
allows two filaments to occupy the same site has been
sions, such as motility assays where actin filaments are
used to simulate actin motility assays.
propelled on a carpet of myosin motor proteins [13, 14],
microtubules propelled by surface-bound dyneins [15], In this paper, we propose a model for self-propelled
and microswimmers that are attracted to surfaces [16 rods (SPRs) in two dimensions that interact with a phys-
20]. ical interaction potential. We discretize each rod by a
In the pioneering work of Vicsek et al. [21], nonequi- number of beads to calculate rod-rod interactions. In
librium phase transitions were observed for systems with contrast to previous models with strict excluded-volume
self-propelled point particles that interact via an imposed interactions [38, 4042, 44], our interaction potential al-
alignment rule and thermal noise. This work led to nu- lows rods to cross. Our simulations thus combine the
merous analytical [2228] as well as computational [29 computational efficiency of two-dimensional simulations
36] studies for systems of self-propelled particles. Be- with a possibility to mimic an escape to the third dimen-
cause each particle consumes energy to generate motion, sion when two rods collide. Simulation snapshots of the
the systems are far from equilibrium and interesting new system which display disordered states, motile clusters,
dynamic properties emerge. For rods with strong short- lanes, etc. are shown in Fig. 1, and movies can be found
range repulsive interactions (volume exclusion) it has in the Supplemental Material [48].
been shown that self-propelled motion leads to alignment The paper is organized as follows. We introduce model,
of rods [3743]. Moreover, self-propulsion enhances ag- simulation methods, and numerical parameters in Sec. II.
gregation and cluster formation [36, 4042, 44, 45]. Near We calculate a phase diagram for passive (nonswimming)
the transition from a disordered to an ordered state, rods in Sec. III using Monte Carlo simulations, followed
the cluster size distribution obeys a power-law decay by a short discussion on the probability of crossing events
[6, 32, 40, 46]. In simulations at higher densities, longi- in Sec. IV. We focus on cluster formation in Sec. V, in-
tudinally moving bands [36] and lanes [38, 41] have been troducing gas density and cluster break-up in Sec. V A,
observed. cluster size analysis in Sec. V B, and autocorrelation func-
2

tions for rod orientations in Sec. V C. We summarize our and


main results in Sec. VI.

Nrod
1 X
rod,i (t) = Mij + r , (4)
r
j6=i

where ek and e are unit vectors parallel and perpendic-


II. MODEL AND SIMULATION TECHNIQUE
ular to the rod axis, respectively. Frod is the propulsion
force for each rod. The friction coefficients are given by
We simulate rods with and without an intrinsic propul- k = 0 Lrod , = 2k , and r = k L2rod /6, where Lrod is
sion force. Our systems consist of Nrod rods in a two- the rod length [40]. The random values k , , and r for
dimensional box of size Lx Ly with periodic bound- the forces in parallel and perpendicular direction and for
ary conditions; see Fig. 1. We use Brownian dynamics the torque are drawn from Gaussian distributions with
2 2 2
simulations for active systems and Monte Carlo simula- variances rod Lrod , 2rod Lrod , and rod L3rod /12, respec-
tions for passive systems. The rods are characterized by tively. We employ thermal noise, thus the variances are
their center-of-mass positions rrod,i , their orientation an- 2
calculated using rod = 2kB T /0 t [50]. Finally, Fij and
gles rod,i with respect to the x axis, their center-of-mass Mij are the force and torque from rod j to rod i, calcu-
velocities vrod,i , and their angular velocities rod,i ; see lated using Eq. (1). Hydrodynamic interactions between
Fig. 2. To calculate energy, force, and torque due to rod- the rods are largely screened because of the nearby wall
rod interactions, we discretize each rod into nb beads, and the high rod density [1720], and hence are neglected
separated from each other by a distance of Lrod /nb . in our simulations.
Beads from different rods interact by a separation-shifted We study systems with approximately 10 000 rods at
Lennard-Jones potential [49], scaled number densities ranging from L2rod = 2.5 to
10, where the number density of rods is defined as =

4 (2 + r2 )6 (2 + r2 )3 + 0 , r < rmin
  Nrod /Lx Ly . We measure lengths in units of rod length
(r) = Lrod , energies in units of kB T , and times in units of the
0, r rmin ,
(1) orientational diffusion time for a single rod, 0 = 1/Dr =
where r is the distance between two beads and  gives 0 L3rod /6 kB T . The system size is Lx = Ly = 36 Lrod , the
the interaction energy. The potential is shifted by 0 cutoff rmin = Lrod /nb = 0.056 Lrod , the rod aspect ratio
to avoid a discontinuity at r = rmin . The parameter Lrod /rmin = 18, the time interval t = 1.65 104 0 ,
characterizes the capping of the potential. For 6= 0, and unless mentioned otherwise, E = 1.5 kB T .
does not diverge at r = 0, hence allowing bead-bead There are three different energy scales in our sys-
overlap; for = 0, (r) becomes the truncated Lennard- tem; the thermal energy kB T , the propulsion strength
Jones potential. Frod Lrod , and the energy barrier E. Therefore, there
are two dimensionless ratios that characterize the impor-
E = (0) (rmin ) is the energy for two beads that tance of the different contributions: the Peclet number,
completely overlap and is used as independent parameter defined as [51]
in our simulations. Setting E to any value will dictate
 = 12 E/(12 46 + 4). The constant = (21/3 Lrod v0 Lrod Frod
2
Pe = = , (5)
rmin )1/2 is calculated by forcing (r) to be zero at r = Dk kB T
rmin . Considering the weak repulsion between rods, we
which is the ratio of propulsion strength to noise, and the
define r = rmin /2 as the effective radius for each bead,
penetrability coefficient, Q, defined as
which results in the effective rod thickness rmin and the
rod aspect ratio Lrod /rmin . The number of beads nb Lrod Frod
Q= , (6)
used for discretization is chosen such that the rod has a E
relatively smooth potential profile, so that no interlocking which is the ratio of propulsion strength to energy bar-
occurs when rods slide along each other; see Fig. 2. rier. Dk = kB T /k is the diffusion coefficient parallel to
For the Brownian dynamics simulations, we decompose the rod orientation.
the rod velocity into parallel and perpendicular compo- We simulate rods with Peclet (Pe) numbers in the
nents with respect to its axis, vrod,i = vrod,i,k + vrod,i, . range 0 Pe < 200 and penetrabilities in the range
In each simulation step, the velocities are calculated us- 0 Q < 200. We change Pe by changing Frod for fixed
2
ing rod and t, i. e., for fixed temperature. We change Q
by changing both Frod and E.

Nrod
1 X
vrod,i,k (t) = Fij,k + k ek + Frod ek , (2) III. ISOTROPIC-NEMATIC TRANSITION FOR
k
j6=i PASSIVE SYSTEMS

Nrod
1 X
vrod,i, (t) = Fij, + e , (3) Suspensions of passive rodlike particles in thermal
equilibrium are isotropic for low densities and nematic
j6=i
3

(a) Pe = 0, L2rod = 10.2 (b) Pe = 0, L2rod = 5.1 (c) Pe = 20, L2rod = 5.1 (d) Pe = 100, L2rod = 5.1

(e) Pe = 25, L2rod = 10.2 (f) close-up of (e) (g) Pe = 75, L2rod = 25.5 (h) color coding guide

FIG. 1. (Color online) Snapshots of self-propelled rod systems simulated using Brownian dynamics simulations. Each rod is
colored based on its orientation. (a) Nematic state at high scaled density L2rod and zero Peclet number Pe, (b) isotropic state
at low and zero Pe, (c) and (e) giant clusters at medium Pe, (f) close-up of a boundary of a giant cluster, (d) cluster break-up
at high Pe, (g) laning phase at high , and (h) color coding for rod orientation. For movies, see the Supplemental Material [48].

S < 0.11 0.11 < S < 0.8 S > 0.8


single bead
rod,i along rod axis
E 12.5
vrod,i Frod
nematic
rmin 10
(r)

rod,i
L2rod

7.5
rrod,i

5
0
isotropic
r 2.5

FIG. 2. (Color online) Left: schematic representation of the 0.2 0.5 1 2 3 5 10


model of a self-propelled rod and coordinates used in two E/kB T
dimensions. The rod is discretized into nb beads to calculate
the rod-rod interaction. Right: potential profile of a rod along
FIG. 3. (Color online) Phase diagram for passive rod systems
its long axis. Tics on the horizontal axis show the position of
with different densities () and energy barriers (E). In addi-
beads, separated from each other by rmin . In our simulations,
tion to color/symbol coding, the size of each triangle is pro-
we use nb = 18.
portional to the nematic order parameter S [Eq. (7)]. Bottom
left: isotropic phase with S < 0.11; top right: nematic phase
with S > 0.8; middle: nematic phase with 0.11 < S < 0.8.
for high densities [52]. For E and Lrod /rmin  1, The black arrow indicates Onsagers isotropic-nematic tran-
the transition density c L2rod = 3/2 has been pre- sition density, c L2rod = 3/2. The red (gray) line is given
dicted using Onsagers theory for infinitely thin hard rods by Eq. (8). Crosses () mark the parameters that have been
[52, 53]. For our systems with the capped potential given used for the Brownian dynamics simulations in Figs. 1(a) and
1(b).
by Eq. (1), not only the aspect ratio of the rod but also
4

the energy barrier E affects the density for the isotropic- 1


nematic transition. As E becomes smaller, the tendency Q Pe E/kB T
17 25 1.5
for rods to align becomes weaker because overlaps occur 0.8
12 25 2.0
10 25 2.5
more frequently. For E = 0, the rods do not interact 7 10 1.5
mutually and thus are in the isotropic phase for all den- 5 10 2.0
0.6 4 10 2.5
sities. experiment

P ()
We performed Brownian dynamics simulations for sys-
tems with Pe = 0 at various densities. At low densities, 0.4
2.5 L2rod 5.1, the systems are in an isotropic state
as shown in Fig. 1(b). For high densities, 7.7 L2rod 0.2
10.2, nematic states are found that are composed of large
interlocked groups of rods with similar orientations; see
0
Fig. 1(a). Because the simulation of passive rods with 10 30 50 70 90 110 130 150 170
Brownian dynamics is computationally very expensive,
we used Monte Carlo simulations to systematically study
the state of the system for several values of and E. We
characterize the state using the nematic order parameter
[42],
* N
+
X 1
S= cos[2(i j )] , (7)
N (N 1)
i6=j
FIG. 4. (Color online) Crossing probability for two rods as a
where the average is over cells of side length 4.5 Lrod . function of their crossing angle (as defined in the schematic).
S = 0 and S = 1 correspond to perfectly isotropic and For each angle, 1000 simulations have been performed. The
nematic states, respectively. simulations are divided into 10 groups and the error bars are
Figure 3 shows a phase diagram of the system with calculated as the standard deviation of the mean (m ) for
varying density and energy barrier. According to ana- these groups. The experimental data are taken from Ref. [15].
lytical theory [52], for E = the transition from the
isotropic to the nematic state occurs at c L2rod = 3/2,
as indicated by the black arrow in Fig. 3. This den-
sity corresponds to S = 0.11, which we thus define as least 0.2Lrod away from the ends of each rod. We thus do
threshold value to calculate the transition density for fi- not count events when one rod only touches the other
nite values of the energy barrier; see Appendix A. We rod, which frequently happens due to the weak repulsion
have also calculated the density for the isotropic-nematic between the rods.
transition,
As shown in Fig. 4, P () is low near ' 0 and
3 1 ' 180 and has a peak near ' 90 . There is a small
c L2rod = , (8)
2 [1 exp(E/kB T )] asymmetry in the peak with an enhancement for direc-
tions > 90 , which may be attributed to the increased
by generalizing Onsagers approach for finite energy bar- relative velocity between two rods for > 90 and the
riers, as described in Appendix A. We find very good fact that the rods are not perfectly smooth. Comparison
agreement between the analytical theory shown by the between P () for different penetrabilities shows that an
red (gray) line in Fig. 3 and our Monte Carlo simulations. increased Q generally increases the probability for rod
The phase diagram is also consistent with our Brownian crossing. In addition, for small Pe, noise also plays an
dynamics simulations for Pe = 0 and E = 1.5 kB T ; see important role to enhance rod crossing. For example,
snapshots in Figs. 1(a) and 1(b). the curves for Q = 10 and Q = 7 in Fig. 4 have approx-
imately the same height, and this could be explained by
the fact that the effect of noise is higher for the case
IV. CROSSING PROBABILITY FOR ROD-ROD Q = 7 that has a smaller Pe.
COLLISIONS
The results are qualitatively similar to the crossing
To find the probability of crossing events P (), we per- probability measured in experiments with microtubules
formed simulations for two rods that initially touch each propelled on surfaces. In Fig. 3(d) in Ref. [15], the maxi-
other in a tip-center arrangement with crossing angle ; mum crossing probability for two microtubules in a motil-
see Fig. 4. We measure P () for several penetrabilities ity assay is 40 % and corresponds to Q = 5 and Pe = 10
and Peclet numbers using Brownian dynamics simula- in our simulations. However, the same crossing probabil-
tions. We count a crossing event when two rods intersect ity may be achieved by reducing Q and increasing Pe at
significantly, i. e., such that the intersection point is at the same time.
5

Pe = 0, i L2rod = 10.2
25 Pe = 0, i L2rod = 5.1
nematic
Pe = 20, i L2rod = 5.1
20
Pe = 100, i L2rod = 5.1
lanes

P ()
15
L2rod

giant cluster
10

5
small clusters
isotropic
0 0 5 10 15 20 25 30
0 25 50 75 100
L2rod
Pe

FIG. 6. (Color online) Density distributions for the sys-


FIG. 5. (Color online) Phase diagram for self-propelled rods tems shown in snapshots of Fig. 1. i is the average density.
with different densities () and Peclet numbers (Pe). The The distributions are not normalized and only the position of
energy barrier is E = 1.5 kB T ; the gray lines are guides to the peaks can be compared.
eye. Note that the region Pe < 0 has no physical meaning and
that the nematic state is found for passive rods with Pe = 0.
a laning phase that is composed of streams of rods that
move in opposite directions; see Fig. 1(g). The laning
V. CLUSTER FORMATION FOR ACTIVE phase is nematic, similar to the nematic lanes that have
SYSTEMS been observed for the Vicsek model in simulations [36]
and analytical calculations [55].
To characterize the collective behavior, we have per- Our phase diagram in Fig. 5 may be compared with the
formed simulations with large numbers of rods. After phase diagram in Ref. [41] for self-propelled rods that in-
initiating the rods with random positions and orienta- teract segment-wise via a Yukawa potential. Since our
tions in the two-dimensional (2D) plane, the rods move model incorporates noise and has a capped repulsive in-
by their propulsion force and are affected by interactions teraction potential, we can only compare both models in
with other rods and thermal noise. Snapshots of the sys- the medium Pe regime, where the noise does not dom-
tem are shown in Fig. 1. More snapshots and movies inate (Pe  1) and where the rods are not completely
can be found in the Supplemental Material [48]. A phase penetrable (Pe . 75). For aspect ratio 18 used in our
diagram of self-propelled rods with varying density and simulation, we see qualitatively similar behavior with in-
Peclet number is shown in Fig. 5. creasing density, namely the transition from the isotropic
For 1 Pe . 80, we find giant clusters that span phase to the swarming (clustering) phase and then to the
the entire simulation box and form as a result of the laning phase.
alignment interaction due to the rod-rod repulsion, as A comparison of our phase diagram in Fig. 5 with that
explained qualitatively in Refs. [36, 54]; see Figs. 1(c) of Ref. [40] shows that we do not observe jammed giant
and 1(e). At the cluster perimeter, the clusters steadily clusters as reported in Ref. [40], because we employ a
lose rods due to the rotational diffusion and at the same smoother potential profile along the rod; see Fig. 2.
time acquire new rods that collide and align. The clus-
ters are polar and almost all rods within a giant cluster
move in the same direction. However, we expect that the A. Rod densities
system is essentially in an isotropic phase, and that for a
sufficiently large system size the clusters can randomly We measure densities of rods in cells of side length
change direction. The polar order of our giant clus- 2 Lrod and construct a distribution of monomer densities
ters which span the simulation box is due to symmetry- for each system; see Fig. 6. For a homogeneous system
breaking collisions because of the roughness of the rods. of rods, the distribution has a single narrow peak at the
In the early stage of the formation of giant clusters, some average density of the system, i . This can be seen for
of the eventually polar clusters are composed of streams example in the histograms for Pe = 0 that correspond to
of rods that move in opposite directions. the systems where no cluster formation is observed; see
Upon further increase of Pe the clusters start to break; Figs. 1(a) and 1(b). For systems with self-propelled rods,
see Fig. 1(d). Smaller clusters are observed until they the density distribution can change from a binomial to
become as small as about five rods per cluster for Pe & a more complicated distribution that shows phase sep-
100. For very high densities, 15.1 L2rod 25.5, when aration between dilute and dense regions of rods. For
the dense region spans the entire simulation box, we find Pe = 20 the distribution has a large peak at low density
6

the volume of the liquid phase.


5 L2rod = 10.2 The dependence of gas on Pe and in the low Pe range
10 7.7
5.1 can be quantitatively explained by a rate equation [29].
2
3.8 In the stationary state, the rate of rods joining a cluster
8 theory
equals the rate of rods leaving a cluster. Assuming an
1 isotropic distribution of rods in the gas, the number of
gas L2rod

6 rods joining the cluster from an infinitesimally small box


0.5 of side length dx and dy is d3 N = dx dy dt(1 cos )/2,
5 10 15 20 25
4 where = cos1 (dx/(vdt)) is the half angle of a cone
inside which rods reach the wall in a given time dt, and
2 x is the distance to the cluster wall. Integrating d3 N
over x from 0 to vdt gives the attachment rate
(a) 0
0 25 50 75 100 d2 N v Dk Pe
Jatt = = = , (9)
(a) Pe dtdy 4 4Lrod
P e1
where we have used the definition of Peclet number in
Eq. (5).
The detachment rate Jdet is determined by the rota-
4 E/kB T = 1.0 tional diffusion of the rods; the typical time a rod needs
1.5 to diffuse by an angle is
gas L2rod

2.0
2.5
3.0 2
10.0 = . (10)
theory 2Dr
2
Assuming that a complete detachment from the cluster
requires = /2 and that rods are placed regularly along
the border of a cluster, the detachment rate is found to
(b) 0
be
0 25 50 75 100 125 150 175
1 8Dr
(b) Pe Jdet = = 2 . (11)
Lrod Lrod
FIG. 7. (Color online) Gas density as a function of the Peclet By equating Jatt and Jdet , we find gas as a function
number for different average rod densities and energy barri- of Pe,
ers compared with the estimate in Eq. (12). (a) Gas densities
for E = 1.5 kB T and several rod densities. Inset: double- 32 Dr 1 192 1
logarithmic plot of gas for 5 < Pe < 25. (b) Gas densities gas = 2
= 2 2 , (12)
Dk Pe Lrod Pe
for average rod density L2rod = 5.1 and several energy bar-
riers. The errors are given by the peak width for the density where we have used Dr /Dk = k /r = 6/L2rod . Note
histograms, L2rod ' 0.5.
that the gas density in Eq. (12) only depends on Lrod
and Pe and is independent of the average system density
, which is consistent with the simulation results. This
and a very broad peak at higher densities. The noise
implies that the giant cluster grows until the density of
in the distribution is due to the poor statistics in the
the dilute region reaches gas .
intermediate density regime. The system consists of a
Note that this estimate includes several approxima-
(high-density) cluster in a gas of rods; the density of
tions, in particular using free rotational diffusion for rods
this cluster-free region corresponds to the position of the
at the border of the cluster and assuming that complete
first peak in the density distribution. In the following,
detachment requires the rods to diffuse by = /2. As
we denote the density of this cluster-free region as gas .
shown in Fig. 7, the analytical estimate in Eq. (12) agrees
In Fig. 7, gas is plotted as a function of Pe for several well with the simulation results in the small-Pe range
values of i and E. We define gas as the position of the without any adjustable parameters. Assuming a two-
first local maximum in the density distribution, which is dimensional gas for the dilute rod phase, we can thus
at least as high as 80% of the absolute maximum. We estimate an effective binding energy per rod for the rods
find gas Pe1 for small Pe and an increase of gas inside the giant cluster,
with increasing Pe for high Pe. The gas density is to a
large extent independent of the average rod density of the Eb = kB T ln(gas Lrod rmin )
entire system;see Fig. 7(a). This behavior is analogous = kB T ln 192/ 2 ln(Pe Lrod /rmin ) , (13)
  
to the vapor density for liquid-gas phase coexistence in
conventional liquids, where the density of the gas phase as explained in Appendix B. The effective binding
only depends on the temperature and is independent of strength increases logarithmically with the product of
7

50 104
= 2.1 Pe = 0
Pe = 100
103 Pe = 25
40

102
30
= 1.6
Q

101


20
= 3.7
100

10 theory
101
simulation
0
102
1 1.5 2 2.5 3
100 101 102 103 104
E/kB T n

FIG. 8. Critical penetrability coefficient Q at which clusters FIG. 9. Cluster size distributions (n) for systems shown
start to break vs the energy barrier E, compared with the in the snapshots of Fig. 1. Average rod density is L2rod =
analytical estimate given by Eq. (14). Average rod density is 10.2. For small n, the distributions can be fit by a power law,
L2rod = 5.1. The points from the simulations are the Peclet (n) n . The distributions have been averaged over 200
numbers at which gas has a minimum for each energy barrier; frames in the last 40 000 time steps.
compare Fig. 7.

1.5
L2rod = 10.2
Peclet number and the rod aspect ratio. For aspect ratio 7.7
18 and Pe 25 used in our simulations, we find effective 5.1
2 3.8
binding energies of Eb 0.1 kB T , which are compara-
ble to binding energies for the gas-liquid critical point for
colloidal systems [56]. 2.5

1.5
For E = 1.5kB T , clusters break up when Pe & 80,
2
which implies Q & 50. We observe that in the regime of 3
cluster break-up, individual rods and even small clusters 2.5

can pass through each other. In our simulations Pe is pro- 3


3.5
portional to the propulsion force, and a high propulsion 3.5
force thus facilitates crossing of rods. As a result, fewer 4
0 5 10 15 20 25
rods aggregate in a large cluster and the rod density in 4
0 25 50 75 100 125 150 175
the dilute region gas increases; see Fig. 7. Cluster break- Pe
up starts when the propulsion force, Frod = Q E/Lrod ,
is comparable with the maximum force for bead-bead in- FIG. 10. (Color online) The exponent of the power law
teraction, Fint = max(d/dr). Equating Fint to Frod for cluster size distributions as a function of Pe for systems
gives the critical value of the penetrability coefficient for with E = 1.5 kB T and several average rod densities . The
cluster break-up, exponents have very weak dependence on . Inset: magnified
view for 0 < Pe < 25.
Fint Lrod
Q = = 28 , (14)
E
B. Cluster size distributions
where Fint = d/dr|r=r0 and r0 = 0.192 is found by
numerically solving d2 (r)/dr2 = 0 for the potential in We define two rods to be in the same cluster if the
Eq. (1). In Fig. 8, Q is plotted for various energy bar- nearest distance between them is less than 2rmin and the
riers. Although the angular dependence for crossing of difference in their orientation angles is less than /6. In
rods (Fig. 4) is neglected in the estimate in Eq. (14), Fig. 9, sample cluster size distributions (n) are pre-
we find reasonable agreement with the simulation results sented. For small cluster size n, (n) decreases with a
without any adjustable parameters. However, there is power law, (n) n with < 0; for large n, (n)
less agreement for small energy barriers, corresponding decreases exponentially [40, 46]. For systems with giant
to small Pe. The deviations may be accounted for by the clusters, such as the system with Pe = 25, there is a
noise that for small Pe is comparable with the propul- gap in the distribution because they consist of one giant
sion force (but that is not considered in the analytical cluster (n > 10 000) and small clusters (n < 30) that
estimate). mostly form near the boundary of the giant cluster. In
8

500 1
20 20
500

400
15 0.9
400
300
15 10
200 0.8
5 100 300
N

C(t)
10 0 0 0.7
0 5 10 15 20 25
L2rod N N 200
10.2 0.6 Pe = 0, L2rod = 10.2
7.7
5 Pe = 25, L2rod = 10.2
5.1 100
3.8 0.5 Pe = 20, L2rod = 5.1
Pe = 0, L2rod = 5.1
0 0 0.4
0 25 50 75 100 125 150 175 0 0.5 1 1.5
Pe (a) t/0

1
FIG. 11. (Color online) Cluster size average N and spread
N as function of Pe for several average system densities. The 0.9
number of rods in systems with L2rod = 10.2 is 13107. Inset: 0.8
magnified view for 0 Pe 25. The cluster sizes have been
0.7
averaged over 200 frames in the last 40 000 time steps.
0.6

a, X
0.5
such systems, the exponent is calculated only based on 0.4
the distribution of small clusters. L2rod a X
0.3
The power-law exponent for the cluster size distribu- 10.2
tion first decreases with increasing Peclet number, has a 0.2 7.7
5.1
minimum for Pe 25, and then increases for increasing 0.1
3.8
the values of Pe, see Fig. 10. We find the exponent to 0
be in the range 1.5 3.5, which agrees with 0 5 10 15 20 25
the range 2 3.6 found in Ref. [40] for rods (b) Pe
with different aspect ratio and a different interaction po-
4
tential than in our simulations. A recent experimental L2rod = 10.2
7.7
study found = 1.88 0.07 for clusters of M. xanthus 5.1
bacteria [6]. As shown in Fig. 11, the average size of the 3.8
3
clusters, N , increases with increasing Peclet number for
Pe . 25 and decreases if Pe is further increased. The
spread of the cluster size, N , shows the same qualita-
2
/0

tive behavior but decays faster at high Pe values, which


shows that the system becomes more homogeneous.
1

C. Polar autocorrelation functions


0
The clustering dynamics in the systems can be char- 0 5 10 15 20 25
acterized by autocorrelation functions for the rod orien- Pe
(c)
tation
C(t) = hni (t0 ) ni (t0 + t)i , (15) FIG. 12. (Color online) (a) Autocorrelation of rod orientation
with lag time t for the systems shown in Fig. 1. Thick lines
for lag time t, where ni (t0 ) is the orientation vector of are simulation results; thin horizontal lines are autocorrela-
rod i at time t0 , and the average is over all rods and over tion base values, calculated by fitting the data with Eq. (16).
(b) Comparison of the autocorrelation base value a and the
all times t0 . Figure 12(a) shows C(t) for systems shown
fraction of rods in the largest cluster X. (c) Autocorrelation
in Fig. 1. The autocorrelation function C(t) can be fit time of the rod orientation as function of Pe for several av-
using a shifted exponential function erage rod densities. 0 = 1/Dr is the time unit; see Sec. II.
The observables have been calculated based on 200 frames in
A(t) = (1 a)et/ + a , (16) the last 40 000 time steps.
where is the autocorrelation time and a is an autocor-
relation base value. A finite value of a is the ratio of
9

rods that do not lose their orientation for the time scale ity for two microtubules in a motility assay is 40 % and
of the measurement. Rods that are inside clusters are corresponds to Q = 5 and Pe = 10 in our simulations
less likely to lose their orientation, which corresponds to [58].
a high value of a, while free rods in the gas change ori-
entation more frequently because of rotational diffusion. Self-propelled rods align due to their soft repulsive in-
In Fig. 12(b), we compare a to the averaged fraction of teraction [54]. For high rod densities, we find a laning
rods X that are part of the largest cluster in the system phase. For intermediate rod densities and Peclet num-
for several densities and Peclet numbers. In general, we bers, we observe the formation of giant clusters that span
find good agreement between a and X [57]. The autocor- the entire simulation box, which we denote as clustering
relation time , shown in Fig. 12(c), does not change sub- window. Clusters break if the propulsion force is strong
stantially for different values of Pe and . The correlation enough to overcome the repulsive force due to rod-rod
time obtained from the fit with Eq. (16) is very simi- interaction. We find a critical value Q = 28 for cluster
lar to the autocorrelation time 0 for a single rod, which break-up. We characterize our systems by cluster size
shows that the rotational diffusion is only weakly affected distributions that can be fit by power-laws (n) n
by occasional collisions of the rods. Therefore, the giant with 1.5 3.5, which is consistent with previous
cluster moves persistently within simulation time. experimental and simulation results [6, 40]. By analyzing
the autocorrelation function for rod orientation, we can
separate the contributions from rods in a cluster from the
VI. SUMMARY AND CONCLUSIONS contributions from free rods. We find that the free rods
show almost the same orientational correlation as single
We have studied collective behavior for self-propelled rods.
rigid rods in two dimensions constructed by single beads
that interact with a separation-shifted Lennard Jones po- We can analytically estimate the density of free rods
tential. The finite potential strength mimics the ability of in systems with giant clusters, which we denote as gas
microswimmers close to a wall and of filaments in motil- density, gas = 192/( 2 L2rod Pe). The gas density is inde-
ity assays to temporarily escape to the third dimension pendent of the average rod density in the system, which
and cross each other. For a high potential barrier, we re- is analogous to the molecule density in the gas phase for
cover the limit of impenetrable rods studied, for example, liquid-gas coexistence that does not depend on the vol-
in Refs. [40, 41]. For most simulations, we have used an ume of the liquid phase but only on temperature. Using
interaction energy E = 1.5 kB T that for complete overlap Eb = kB T ln[192rmin /( 2 Lrod Pe)], we calculate effective
of two beads is of the order of the thermal energy; cross- binding energies for rods in the cluster. For aspect ratio
ing of rods therefore occurs with a high probability; see 18 used in our simulations and Pe 25, we find effective
Fig. 4. However, the interaction energy is much larger binding energies of about 0.1 kB T , which is comparable
than the bead-bead interaction energy if rods cross at a to binding energies for the gas-liquid critical point for
small angle or if a single rod approaches a cluster. For colloidal systems [56].
our system with nb = 18 beads per rod, the energy for
complete overlap is nb E = 27 kB T and thus the proba- Phase separation into high-density and low-density re-
bility for such events is very low. gions is an intrinsic property of self-propelled particle sys-
We have calculated a phase diagram for rod density tems and has also been observed for nonaligning spherical
and energy barrier to characterize the isotropic-nematic particles [29, 5961]. As for the rods, the gas density of
transition for passive rods. The isotropic-nematic tran- the spheres is inversely proportional to the propulsion
sition is shifted to higher densities for reduced overlap velocity [29]. However, the nature of cluster formation is
energy, because of the reduced rod-rod interaction. We different in the two models: While we observe motile clus-
find significant deviations from the transition density cal- ters as a result of particle alignment, systems with non-
culated for hard rods [52, 53] if the bead-bead interaction aligning spheres exhibit jammed nonmotile clusters as a
energy is below 2 kB T . For E = 1.5 kB T , the isotropic- result of steric trapping. Moreover, the internal struc-
nematic transition occurs for = 1.3 c , where c is the ture of clusters is nematic in our model, contrary to the
transition density for hard rods [52]. Our results using a isotropic structure for nonaligning spheres. Of course,
modified Onsager theory show excellent agreement with also laning phases are only possible for anisotropic par-
our Monte Carlo simulations. ticles.
Using Brownian dynamics simulations, we have deter-
mined the crossing probability for two colliding rods as We have introduced and characterized a model of self-
function of their relative angles for several values of pen- propelled rods that interact with a physical interaction
etrability coefficient and Peclet number. The crossing that allows for crossing events. The model can now be
probability is highest for almost perpendicular collisions, used to interpret experiments for almost two-dimensional
which is qualitatively similar to the crossing probability systems with good computational efficiency and allows
measured in experiments with microtubules propelled on predictions beyond those based on models using point
surfaces. In Ref. [15], the maximum crossing probabil- particles with phenomenological alignment rules.
10

ACKNOWLEDGMENTS For finite energy barrier E, we generalize the approach


presented in Ref. [52] based on bifurcation theory to ob-
We thank Adam Wysocki and Jens Elgeti for stimu- tain the critical density for the isotropic-nematic transi-
lating discussions. M.A. and K.M. acknowledge support tion. The distribution function for the rod orientation is
by the International Helmholtz Research School of Bio- given by f (), which satisfies
physics and Soft Matter (IHRS BioSoft). CPU time al- Z 2
lowance from the Julich Supercomputing Centre (JSC) is ln[2f ()] = C + F (, 0 )f (0 )d0 , (A1)
gratefully acknowledged. =0

where the constant C is determined by the normalization


of f (),
Appendix A: Phase transition of passive rods Z 2
f ()d = 1 . (A2)
The nematic order parameter is plotted in Fig. 13 for =0
various cuts through the phase diagram in Fig. 3. It
We define h() as
has been suggested that the isotropic-nematic transition
of rods is continuous in two dimensions [52]. We have h() = 2f () 1 , (A3)
chosen a threshold value for the isotropic-nematic phase
transition, St = 0.11, such that for an infinite interaction such that h() = 0 corresponds to an isotropic distribu-
energy the value predicted by the Onsager theory is re- tion. Using Eqs. (A1)(A3), we can write
covered. Our threshold value is similar to the threshold
exp( F (, 0 )h(0 )d0 /2)
R
value S = 0.2 that has been chosen in Ref. [42]. h() = 1 + R R .
(1/2) exp( F (, 0 )h(0 )d0 /2)d
(A4)
1
E/kB T = We assume that the interaction energy of two rods is ei-
7.5 ther 0 or E, depending on whether they cross each other.
0.8 1.25
0.63 This approximation is justified if the rods are very thin
0.19 and the complete overlap of two rodswhich is energeti-
0.6 cally very unfavorableis excluded. In the regime where
Lrod  rmin , the parameter and the kernel F are given
S

by
0.4
1
= L2rod , (A5)
0.2 4
(1/2)F (, 0 ) = (2/ 2 ) sin() (A6)
0 [1 exp(E/kB T )] .
4 6 8 10 12 Substituting Eqs. (A5) and (A6) in Eq. (A4) gives
(a) L2rod
exp[Kh()]
1 h() = 1 + R , (A7)
(1/) 0 exp[Kh()]d

0.8 where the operator K is defined as


Kh() = (L2rod /)[1 exp(E/kB T )] (A8)
0.6 Z
sin(0 )h( 0 )d0 .
S

0
0.4 L2rod = 12.7
10.2 For h() to have bifurcation point in ,
7.7
5.1
0.2 2.6 w() = Kw() (A9)
has to have an eigenfunction with two maxima at w(0) =
0
w() and no further maxima. The corresponding eigen-
0 0.5 1 1.5 2 2.5 value determines the density at which bifurcation occurs.
(b) E/kB T The desired eigenfunction is cos 2 with the eigenvalue
3/2; thus the bifurcation density that corresponds to
FIG. 13. (Color online) Nematic order parameter S, used to the isotropic-nematic transition is
determine the phase transition for passive rods, as a function
of (a) the rod density for several energy barriers E and (b) 3 1
c = . (A10)
the energy barrier E for several rod densities . 2L2rod [1 exp(E/kB T )]
11

Appendix B: Effective binding energy for rod ergy. The free energy for the rods in the gas is thus
adsorption to the cluster  
N
Fgas = N kB T ln (B1)
A
= N kB T ln (gas Lrod rmin ) ,
The independence of gas density gas from the average where N is the number of rods in the gas, is the area
density of the system is analogous to a vapor density of each rod, and A is the area accessible for the rods in
for rods. Here we follow this analogy to obtain an effec- the gas.
tive binding energy gain Eb for rods that are part of the In the cluster, each rod gains a binding energy Eb ,
cluster.
Fcluster = N Eb , (B2)

We use an ideal-gas model in two dimensions to rep- where N here is the number of rods in the cluster.
resent the rods in the gas phase. The activity and the In equilibrium, the chemical potential = F/N in
anisotropy of the rods are intentionally not taken into the gas and in the cluster should be equal. This gives
account explicitly and enter via the effective binding en- Eq. (13).

[1] T. Vicsek and A. Zafeiris, Phys. Rep. 517, 71 (2012). [22] J. Toner and Y. Tu, Phys. Rev. Lett. 75, 4326 (1995).
[2] I. H. Riedel, K. Kruse, and J. Howard, Sci. Signal. 309, [23] R. A. Simha and S. Ramaswamy, Phys. Rev. Lett. 89,
300 (2005). 058101 (2002).
[3] Y. Yang, J. Elgeti, and G. Gompper, Phys. Rev. E 78, [24] S. Ramaswamy, R. A. Simha, and J. Toner, Europhys.
061903 (2008). Lett. 62, 196 (2003).
[4] E. Ben-Jacob, I. Cohen, and H. Levine, Adv. Phys. 49, [25] J. Toner, Y. Tu, and S. Ramaswamy, Ann. Phys. 318,
395 (2000). 170 (2005).
[5] A. Sokolov, I. S. Aranson, J. O. Kessler, and R. E. Gold- [26] F. Peruani, A. Deutsch, and M. Bar, Eur. Phys. J. Spe-
stein, Phys. Rev. Lett. 98, 158102 (2007). cial Topics 157, 111 (2008).
[6] F. Peruani, J. Starru, V. Jakovljevic, L. Sgaard- [27] E. Bertin, M. Droz, and G. Gregoire, J. Phys. A: Math.
Andersen, A. Deutsch, and M. Bar, Phys. Rev. Lett. Theor. 42, 445001 (2009).
108, 098102 (2012). [28] R. Golestanian, Phys. Rev. Lett. 102, 188305 (2009).
[7] J. Gachelin, G. Mino, H. Berthet, A. Lindner, A. Rous- [29] G. S. Redner, M. F. Hagan, and A. Baskaran, Phys. Rev.
selet, and E. Clement, Phys. Rev. Lett. 110, 268103 Lett. 110, 055701 (2013).
(2013). [30] B. Szabo, G. J. Szollosi, B. Gonci, Z. Juranyi,
[8] W. F. Paxton, K. C. Kistler, C. C. Olmeda, A. Sen, S. K. D. Selmeczi, and T. Vicsek, Phys. Rev. E 74, 061908
St. Angelo, Y. Cao, T. E. Mallouk, P. E. Lammert, and (2006).
V. H. Crespi, J. Am. Chem. Soc. 126, 13424 (2004). [31] G. Gregoire and H. Chate, Phys. Rev. Lett. 92, 025702
[9] G. Volpe, I. Buttinoni, D. Vogt, H.-J. Kummerer, and (2004).
C. Bechinger, Soft Matter 7, 8810 (2011). [32] C. Huepe and M. Aldana, Phys. Rev. Lett. 92, 168701
[10] Y. Mei, A. A. Solovev, S. Sanchez, and O. G. Schmidt, (2004).
Chem. Soc. Rev. 40, 2109 (2011). [33] M. R. DOrsogna, Y. L. Chuang, A. L. Bertozzi, and
[11] S. Ebbens, M.-H. Tu, J. R. Howse, and R. Golestanian, L. S. Chayes, Phys. Rev. Lett. 96, 104302 (2006).
Phys. Rev. E 85, 020401 (2012). [34] M. Aldana, V. Dossetti, C. Huepe, V. M. Kenkre, and
[12] M. Yang and M. Ripoll, Phys. Rev. E 84, 061401 (2011). H. Larralde, Phys. Rev. Lett. 98, 095702 (2007).
[13] Y. Harada, A. Noguchi, A. Kishino, and T. Yanagida, [35] H. Chate, F. Ginelli, G. Gregoire, and F. Raynaud, Phys.
Nature (London) 326, 805 (1987). Rev. E 77, 046113 (2008).
[14] V. Schaller, C. Weber, C. Semmrich, E. Frey, and A. R. [36] F. Ginelli, F. Peruani, M. Bar, and H. Chate, Phys. Rev.
Bausch, Nature (London) 467, 73 (2010). Lett. 104, 184502 (2010).
[15] Y. Sumino, K. H. Nagai, Y. Shitaka, D. Tanaka, [37] A. Baskaran and M. C. Marchetti, Phys. Rev. E 77,
K. Yoshikawa, H. Chate, and K. Oiwa, Nature (Lon- 011920 (2008).
don) 483, 448 (2012). [38] S. R. McCandlish, A. Baskaran, and M. F. Hagan, Soft
[16] E. Lauga, W. R. DiLuzio, G. M. Whitesides, and H. A. Matter 8, 2527 (2012).
Stone, Biophys. J. 90, 400 (2006). [39] A. Baskaran and M. C. Marchetti, Phys. Rev. Lett. 101,
[17] A. P. Berke, L. Turner, H. C. Berg, and E. Lauga, Phys. 268101 (2008).
Rev. Lett. 101, 038102 (2008). [40] Y. Yang, V. Marceau, and G. Gompper, Phys. Rev. E
[18] J. Elgeti and G. Gompper, EPL 85, 38002 (2009). 82, 031904 (2010).
[19] J. Elgeti, U. B. Kaupp, and G. Gompper, Biophys. J. [41] H. H. Wensink, J. Dunkel, S. Heidenreich, K. Drescher,
99, 1018 (2010). R. E. Goldstein, H. Lowen, and J. M. Yeomans, Proc.
[20] J. Elgeti and G. Gompper, EPL 101, 48003 (2013). Natl. Acad. Sci. U.S.A. 109, 14308 (2012).
[21] T. Vicsek, A. Czirok, E. Ben-Jacob, I. Cohen, and [42] P. Kraikivski, R. Lipowsky, and J. Kierfeld, Phys. Rev.
O. Shochet, Phys. Rev. Lett. 75, 1226 (1995). Lett. 96, 258103 (2006).
12

[43] A. Costanzo, R. Di Leonardo, G. Ruocco, and L. Ange- (1978).


lani, J. Phys. Condens. Matter 24, 065101 (2012). [53] L. Onsager, Phys. Rev. 65, 117 (1944).
[44] F. Peruani, A. Deutsch, and M. Bar, Phys. Rev. E 74, [54] A. Baskaran and M. C. Marchetti, Eur. Phys. J. E 35, 1
030904 (2006). (2012).
[45] F. Peruani, L. Schimansky-Geier, and M. Bar, Eur. [55] A. Peshkov, I. S. Aranson, E. Bertin, H. Chate, and
Phys. J. Special Topics 191, 173 (2010). F. Ginelli, Phys. Rev. Lett. 109, 268701 (2012).
[46] C. Huepe and M. Aldana, Phys. A (Amsterdam, Neth.) [56] G. Vliegenthart and H. N. Lekkerkerker, J. Chem. Phys.
387, 2809 (2008). 112, 5364 (2000).
[47] F. Ruhnow, D. Zwicker, and S. Diez, Biophys. J. 100, [57] The values of a and X do not agree for Pe = 0, where rods
2820 (2011). are either freely moving due to noise (isotropic regime) or
[48] See Supplemental Material at [URL] for movies of self- stuck in small nonmotile clusters (nematic regime). In the
propelled rod systems. former case, there is hardly any orientation preservation
[49] M. E. Fisher and D. Ruelle, J. Math. Phys. 7, 260 (1966). (a  1) and there is no large cluster (X  1). In the
[50] In biological and synthetic self-propelled systems, the latter case, rods hardly change their orientation (a 1)
noise arises from the environmental noise, for exam- but are distributed over small clusters (X  1).
ple, from density fluctuations of signaling molecules for [58] The same crossing probability may be achieved by in-
chemotactic swimmers or from motor activity. In this creasing E and Pe at the same time.
case, the noise is not proportional to kB T and also the [59] J. Tailleur and M. E. Cates, Phys. Rev. Lett. 100, 218103
coupling between translational and rotational noise may (2008).
be different. [60] Y. Fily and M. C. Marchetti, Phys. Rev. Lett. 108,
[51] The Peclet number can be alternatively defined as Per = 235702 (2012).
v0 /Dr Lrod with the rotational diffusion constant Dr . In [61] A. Wysocki, R. G. Winkler, and G. Gompper,
such case, Pe = 6 Per . arXiv:1308.6423 [cond-mat.soft].
[52] R. F. Kayser and H. J. Raveche, Phys. Rev. A 17, 2067

Das könnte Ihnen auch gefallen