ArticlePDF Available

Diffusion enhancement in bacterial cytoplasm through an active random force

Authors:

Abstract

Experiments have found that diffusion in metabolically active cells is much faster than in dormant cells, especially for large particles. However, the mechanism of this size-dependent diffusion enhancement in living cells is still unclear. In this Letter, we approximate the net effect of metabolic processes as a white-noise active force and simulate a model system of bacterial cytoplasm with a highly polydisperse particle size distribution. We find that diffusion enhancement in active cells relative to dormant cells can be more substantial for large particles. Our simulations agree quantitatively with the experimental data of Escherichia coli, suggesting an autocorrelation function of the active force proportional to the cube of the particle radius. We demonstrate that such a white-noise active force is equivalent to an active force of about 0.57 pN with random orientation. Our work unveils an emergent simplicity of random processes inside living cells.
PHYSICAL REVIEW RESEARCH 5, L032018 (2023)
Letter
Diffusion enhancement in bacterial cytoplasm through an active random force
Lingyu Meng ,1Yiteng Jin ,2Yichao Guan ,3Jiayi Xu,4and Jie Lin 1,2,*
1Peking-Tsinghua Center for Life Sciences, Academy for Advanced Interdisciplinary Studies, Peking University, Beijing 100871, China
2Center for Quantitative Biology, Academy for Advanced Interdisciplinary Studies, Peking University, Beijing 100871, China
3The Graduate Program in Biophysical Sciences, University of Chicago, Chicago, Illinois 60637, USA
4Yuanpei College, Peking University, Beijing 100871, China
(Received 5 September 2022; accepted 10 July 2023; published 7 August 2023)
Experiments have found that diffusion in metabolically active cells is much faster than in dormant cells,
especially for large particles. However, the mechanism of this size-dependent diffusion enhancement in living
cells is still unclear. In this Letter, we approximate the net effect of metabolic processes as a white-noise active
force and simulate a model system of bacterial cytoplasm with a highly polydisperse particle size distribution.
We find that diffusion enhancement in active cells relative to dormant cells can be more substantial for large
particles. Our simulations agree quantitatively with the experimental data of Escherichia coli, suggesting an
autocorrelation function of the active force proportional to the cube of the particle radius. We demonstrate that
such a white-noise active force is equivalent to an active force of about 0.57 pN with random orientation. Our
work unveils an emergent simplicity of random processes inside living cells.
DOI: 10.1103/PhysRevResearch.5.L032018
The efficient diffusion of cellular components is crucial
for various biological processes in bacteria since they do not
have active transport systems involving protein motors and
cytoskeletal filaments [14]. Meanwhile, bacterial cytoplasm
is highly crowded [58]. Particle diffusion inside the bacterial
cytoplasm is significantly suppressed compared with a dilute
solution [914]. Interestingly, the diffusion of large cellular
components, such as plasmids, protein filaments, and storage
granules, turns out to be much faster in metabolically active
cells than in dormant cells, i.e., cells depleted of ATP [1519].
Cellular metabolic activities appear to fluidize the cytoplasm
and allow large components to sample cytoplasmic space.
Another important feature of bacterial cytoplasm is its poly-
dispersity with constituent sizes spanning from subnanometer
to micrometers [2023]. Intriguingly, diffusion enhancement
of a metabolically active cell relative to a dormant cell is
size dependent as large components’ diffusion constants are
much more significantly increased while small molecules dif-
fuse with virtually the same diffusion constants in active and
dormant cells [16].
The physical mechanism underlying the size-dependent
diffusion behaviors in active cells is far from clear. Because
of the numerous ATP-consuming processes in vivo, finding
the dominant biological processes that speed up diffusion
may be difficult or even impossible. In a passive solution,
particles receive random kicks from neighboring molecules
*Corresponding author: linjie@pku.edu.cn
Published by the American Physical Society under the terms of the
Creative Commons Attribution 4.0 International license. Further
distribution of this work must maintain attribution to the author(s)
and the published article’s title, journal citation, and DOI.
due to thermal fluctuation. Therefore, the thermal noise’s am-
plitude is proportional to the temperature according to the
fluctuation-dissipation (FD) theorem. In contrast, in the cy-
toplasm of a metabolically active cell, particles also receive
random kicks from biomolecules such as ATPs, amino acids,
and other metabolites that do not follow a detailed balance
and are therefore out of equilibrium [2427]. As a result, the
net effect of their random collision with a particle is a random
force not constrained by the FD theorem.
In this Letter, we simulate a model system of bacterial cy-
toplasm and adopt a coarse-grained approach by introducing
an active random force as the net effect of multiple active
processes in cells. We model this active random force as white
noise with an amplitude independent of temperature. Our
system is highly polydisperse, and according to the Stokes-
Einstein relation, the autocorrelation function of the thermal
random force is proportional to the particle radius. Inspired by
that, we set the autocorrelation function of the active random
force proportional to a power-law function of the particle
radius. We find that in both passive systems without an active
force (corresponding to dormant cells) and active systems
(corresponding to metabolically active cells), the diffusion
constants are reduced under high density compared with the
dilute limit. This diffusion reduction is stronger for larger
particles in both passive and active systems.
Nevertheless, the enhancements of diffusion constants in
active cells relative to dormant cells can be more substantial
for larger particles given an appropriate size-dependent active
random force. Most importantly, the experimentally measured
ratios of diffusion constants between active and dormant E.
coli cells agree quantitatively with our simulations and sug-
gest an autocorrelation function of the active random force
proportional to the cube of the particle radius. We further
demonstrate that such a white-noise active force is equiva-
2643-1564/2023/5(3)/L032018(6) L032018-1 Published by the American Physical Society
MENG, JIN, GUAN, XU, AND LIN PHYSICAL REVIEW RESEARCH 5, L032018 (2023)
FIG. 1. A snapshot of three-dimensional simulations using poly-
disperse spheres as a model system of the bacterial cytoplasm. In
this example, the volume fraction φ=0.58. Particles are colored
according to their sizes.
lent to an active force with a constant magnitude undergoing
rotational diffusion. From the data of E. coli, we infer the
magnitude of this active force as 0.57 pN, consistent with
typical force magnitudes in the cytoplasm [28]. Our results
shed light on the mechanical nature of out-of-equilibrium
processes in the bacterial cytoplasm and unveil an emergent
simplicity in complex living systems.
Size-dependent active random force. We model the various
cellular components by spherical particles with heterogeneous
radii to mimic the polydisperse cytoplasm (Fig. 1) and their
equations of motion using the Langevin dynamics,
ηi
dri
dt =−U
ri
+ξi +κi .(1)
Here, i=1,2,...,Nwhere Nis the number of particles
and α=x,y,zthe directions in the Cartesian coordinate. ηi
is the friction coefficient of the ith particle and obeys the
Stokes’ law ηi=6πνai, where aiis the radius, and νis the
viscosity of the background solvent. Uis the pairwise interac-
tion between particles which we model as U=1
2i= jk(ai+
ajrij)2(ai+ajrij) where rij =|rirj|and is the
Heaviside step function: Particles repel each other only when
they overlap. ξi is the thermal noise, and its autocorrelation
function obeys the FD theorem [29]
ξi (t)ξj (t)=12πνaikBTδijδαβδ(tt).(2)
Here, kBis the Boltzmann constant, and Tis the temperature.
We introduce an active random force κi as the coarse-grained
outcome of active processes in a metabolically active cell, and
its autocorrelation function is independent of temperature,
κi (t)κj (t)=2Aaγ
iδijδαβδ(tt).(3)
Here, we assume that the noise amplitude is a power-law
function of the particle radius where Aand γare constants.
Later, we will show that this size dependence of active random
force is consistent with experimental measurements.
To nondimensionalize the model, we choose the aver-
age particle radius a0, the thermal energy kBT, and t0=
6πνa3
0/kBTas the length, energy, and time unit, respectively.
The dimensionless equation of motion becomes
d˜ri
d˜
t=−1
˜ai
˜
U
˜ri
+˜
ξi +˜κi,(4)
where ˜
ξi is the dimensionless thermal noise with its autocor-
relation function ˜
ξi (˜
t),˜
ξj (˜
t)=a1
iδijδαβδ(˜
t˜
t), and
˜κi is the dimensionless active noise with its autocorrelation
function ˜κi (˜
t),˜κj (˜
t)=2˜
A˜aγ2
iδijδαβδ(˜
t˜
t). Here, the
dimensionless active noise amplitude ˜
A=Aaγ1
0/(6πνkBT).
In the following, variables with a tilde above are dimension-
less. We simulate Nparticles in a three-dimensional cubic box
with a dimensionless side length ˜
Lunder the periodic bound-
ary condition. The Nparticles’ radii obey a shifted lognormal
distribution ˜a=˜amin +exp[μ+σN(0,1)], where N(0,1) is
a standard normal random number, and μand σare constants.
The minimum radius ˜amin is 0.1. We set σ=0.85 to mimic
the polydisperse environment and choose μby setting the di-
mensionless average radius ˜amean =˜amin +exp(μ+σ2/2) as
1. We set the physical value of the average radius as a0=10
nm, so the particles’ radii cover two orders of magnitude, from
1 to about 100 nm, consistent with real bacterial cytoplasm
[9,16]. We fix the volume fraction φin each simulation. We
also set T=300 K and choose a large dimensionless spring
constant ˜
k=1000 to mimic a hard-sphere system. The total
simulation time ˜
ttot is 10 000 with a time step d˜
t=2×104.
Active noise facilitates diffusion of large particles. We first
investigate the effects of volume fractions on the diffusion
constants relative to the dilute limit for both passive and active
systems. We compute the diffusion constant for each particle
from the time-averaged mean square displacement (MSD)
as D=˜
r2(˜
t)/6˜
twith ˜
t=1 where ˜
r(˜
t)isthe
displacement vector during a time interval ˜
t. We choose
multiple values of φ, including 0.58 and 0.64, the critical
volume fractions of the glass transition, and random close
packing of a monodisperse system in three dimensions [30].
In the dilute limit, collisions between particles are negli-
gible, and the diffusion constant of an active particle with a
dimensionless radius ˜abecomes Ddilute =1/˜a+˜
A˜aγ2.For
a passive particle with ˜
A=0, D0,dilute =1/˜a. In both pas-
sive and active systems, the reduction of diffusion constants
in high-volume fractions relative to the dilute limit is more
significant for larger particles [Figs. 2(a) and 2(b)]. To demon-
strate the effects of active random force, we compare the
diffusion constants of active and passive systems in the same
volume fraction, which is more biologically relevant. In the
dilute limit, the ratio of diffusion constants between active and
dormant cells is
Ddilute
D0,dilute
=1+˜
A˜aγ1.(5)
Our simulation results for φ=0.01 confirm Eq. (5)
[Fig. 2(c)].
For systems with high volume fractions, neither Dwith
˜
A>0 nor D0with ˜
A=0 is equal to the dilute limit prediction
[Figs. 2(a) and 2(b)]. Intriguingly, we find that the simulation
results of D/D01 with high volume fractions still agree
reasonably well with the theoretical prediction from the dilute
limit, particularly for small particles with ˜a<1 [Fig. 2(d)].
L032018-2
DIFFUSION ENHANCEMENT IN BACTERIAL CYTOPLASM PHYSICAL REVIEW RESEARCH 5, L032018 (2023)
(a) (b)
(c) (d)
FIG. 2. Diffusion constants of the simulated polydisperse sys-
tems. (a) The diffusion constants of passive systems relative to the
dilute limit. ˜ais the dimensionless radius. (b) The same analysis
as (a) but for an active system. Here, ˜
A=1, γ=3. The diffusion
constants in the dilute limit differ in (a) and (b). (c) D/D01for
different γs. Dis the diffusion constant of the active system, and
D0is the diffusion constant of the passive system. Here, φ=0.01,
˜
A=1. The lines with corresponding colors are the predictions in
the dilute limit. (d) The same analysis as (c) but for systems with
φ=0.58. In all panels, N=1000.
The above observation is nontrivial because all particles’
diffusion constants deviate from the dilute limit regardless of
size [Figs. 2(a) and 2(b)]. Deviations are observed for large
particles, and we define an effective γeff for large particles with
˜a>1.2 such that D/D01˜aγeff 1(see Supplemental Ma-
terial [31], Fig. S1). Experimentally, the diffusion constants of
small particles are close in active and dormant cells; however,
the diffusions of large particles are much faster in active cells
than in dormant cells [16]. Our simulations in the regime of
γ>1 agree with experiments [Fig. 2(d)].
We also compute the diffusion constant D=
˜
r2(˜
t)/6˜
tusing a longer ˜
twhere ˜
r2(˜
t)= ˜
L2/32
but still short enough to ensure that the finite system size
does not confine the particles’ MSDs, and our conclusions
are equally valid under this definition [Figs. S2(a), S2(b), and
S3]. We also compute the diffusion constants by fitting the
MSDs using ˜
r2(˜
t)=6D˜
tand obtain similar results
[Figs. S2(c) and S2(d)].
We track the trajectories of single particles (Fig. 3). We find
that particles with small radii can equally explore the system
in active and passive systems; however, particles with large
radii can only explore space in active systems and remain
localized in passive systems, consistent with experimental
observations [16].
Comparison with experiments. In Ref. [16], the authors
measured the MSDs of exogenous particles of multiple sizes
in active and dormant cells, including GFP-μNS particles,
which are GFP-labeled self-assembling avian reovirus pro-
teins with changeable sizes, and mini-RK2 plasmid, which is
an engineered low-copy-number plasmid in E. coli. We find
FIG. 3. Single particles’ trajectories projected to two dimen-
sions. (a) The trajectory of a large particle with ˜a=6.75 in a passive
system. The color represents the time elapsed from the beginning of
the trajectory. (b) The trajectory of the same particle in (a) but in an
active system. (c) The trajectory of a small particle with ˜a=2.44 in
a passive system. (d) The trajectory of the same particle in (c) but in
an active system. In (b) and (d), ˜
A=0.42 and γ=3. In all panels,
φ=0.58, N=1000.
that the measured MSDs are subject to large noises, making it
difficult to compute the diffusion constants accurately. To cir-
cumvent this problem, we compute the ratios of MSDs in ac-
tive and dormant cells at every moment and calculate the ratios
of diffusion constants between active and dormant cells D/D0
as the averaged MSD ratios over time. The experimental parti-
cle radius is converted to a dimensionless number using a0=
10 nm.
To compare with experiments, we simulate several dif-
ferent volume fractions since the actual volume fraction of
bacterial cytoplasm is unknown. Intriguingly, the simulated
D/D01 with γ=3 nicely matches the experimental data
[Fig. 4(a)], and we will explain the physical mechanism
of γ=3 later. We find that the simulated diffusion en-
hancements D/D01 are insensitive to the volume fraction,
although the diffusion constants Dand D0by themselves
change significantly with the volume fraction [Figs. 2(a) and
2(b)]. To confirm the robustness of our results, we also sim-
ulate a narrower distribution of particle radii that is more
Gaussian-like. The agreement between simulations and ex-
periments is equally valid (Fig. S4). Our results are also
independent of the length unit as we choose a different length
unit and obtain the same results (Fig. S5). Using the alter-
native definition of diffusion constants does not affect our
conclusions (Fig. S6).
We note that for a dilute active system with γ=3, the
absolute diffusion constant is a nonmonotonic function of
particle size (Ddilute =1/˜a+˜
A˜a). Nevertheless, we find that
the absolute diffusion constant continuously decreases with
particle size in systems with large volume fractions (Fig. S7),
consistent with the experimental observations [32].
L032018-3
MENG, JIN, GUAN, XU, AND LIN PHYSICAL REVIEW RESEARCH 5, L032018 (2023)
(a) (b)
FIG. 4. Comparison between simulations and experiments. (a) The relative enhancements of diffusion constants (D/D01) from the
experimental data match the simulation results of the white-noise active force model. Here, γ=3, ˜
A=0.42. (b) The same analysis as (a) but
for simulations of the self-propelled model with the longer ˜
t. Here, ˜
F=1.38. The results are binned over particles with a bin interval of 0.02
in (a) and of 0.05 in (b) in the log10 scale. In both panels, N=4000.
We also calculate the radius of gyration, the root-mean-
square distance from the center of the trajectory, for both
passive and active systems [Fig. S8(a)]. The radii of gyration
from simulations decrease linearly with the particle radius in
both passive and active systems. The ratio between the passive
and active systems also has a linear relationship with the par-
ticle radius [Fig. S8(b)]. These results agree with experiments
[16], further supporting the validity of our simulations.
In Ref. [16], the authors observed much stronger glassylike
properties in dormant cells than in living cells. We find similar
hallmarks of a glass transition in our simulations, includ-
ing dynamic heterogeneity and non-Gaussian displacements
[3335]. The MSDs of particles with similar radii in the same
passive system vary over two orders of magnitude [Fig. 5(a)],
showing significant dynamic heterogeneity. Meanwhile, this
dynamic heterogeneity is much weaker in the corresponding
active system. Furthermore, the displacement distributions
have a non-Gaussian tail in the passive systems while the devi-
ation from Gaussian distribution is much less significant in the
(a) (b)
FIG. 5. Activity fluidizes the glassy polydisperse system under
a high volume fraction. (a) Violin plot of MSDs of particles whose
˜a4 for passive and active systems under φ=0.75. The MSDs
within a dimensionless time of 1000 are shown in the log scale
and normalized by the average value. (b) The non-Gaussian pa-
rameter α2of the displacement distribution as a function of the
particle radius for passive and active systems under φ=0.75. α2=
˜
r(˜
t)4/[3˜
r(˜
t)22] and it equals 1 if the displacements obey
the Gaussian distribution. The results are averaged with 50 particles
in each bin. The maximum α2is shown from ˜
t=50 to 2500. For
the active systems, ˜
A=0.42 and γ=3. In both panels, N=4000.
active systems (Fig. S9). Indeed, the non-Gaussian degree of
displacement distributions for large particles is much weaker
in the active system than in the passive system [Fig. 5(b)].
Our results show that activity fluidizes the glassy polydisperse
system, in agreement with the experimental observations [16].
We find that a higher volume fraction is needed to observe
the hallmarks of a glass transition in polydisperse systems
(Fig. S10) than in monodisperse systems (Fig. S11), consis-
tent with observations that polydispersity can smear out a
glass transition [36].
Mechanism of the cubic scaling γ=3.In the follow-
ing, we explain the cubic scaling of the white-noise active
force with the particle radius. We consider a self-propelled
model in which an active force with a constant magnitude
is exerted on each particle [3740]. The orientation of this
active force is random due to the rotational diffusion of the
particle. The equation of motion for the ith active particle
becomes
ηi
dri
dt =−U
ri
+ξi +Fn
i ,(6)
where Fis the magnitude of the active force and ni is the
orientation vector of the active force in the direction α.The
orientation vector niobeys dni/dt =Ti×ni, where Tiis
the thermal random torque. Its autocorrelation function satis-
fies the FD theorem, Ti(t),Tj (t)=2DR,iδijδαβδ(tt).
Here, DR,i=kBT/8πνa3
iis the rotational diffusion constant
for a spherical particle with radius ai.
At long times, the additional diffusion constant due to
activity Dactive =(F/6πνa)2×(1/2DR)/3, where F/6πνais
the speed of the active particle and 1/2DRis the time for
the active force to change its orientation in three-dimensional
space. Therefore, the diffusion enhancement of an active par-
ticle relative to a passive particle in the dilute limit becomes
Ddilute
D0,dilute
=1+2˜
F2
9˜a2.(7)
Here, ˜
Fis the dimensionless active force with unit kBT/a0.
Comparing Eqs. (5) and (7), we find that the two models are
equivalent in terms of diffusion enhancement when γ=3.
L032018-4
DIFFUSION ENHANCEMENT IN BACTERIAL CYTOPLASM PHYSICAL REVIEW RESEARCH 5, L032018 (2023)
This conclusion applies to the dilute limit, and we hypothesize
that the equivalence of the two models is still valid under
high-volume fractions.
To test our hypothesis, we simulate the self-propelled
model with ˜
Fsatisfying ˜
A=2˜
F2/9 so that the two models
lead to the same diffusion enhancement in the dilute limit.
The agreement between simulations and experimental data
also holds for the self-propelled model [Fig. 4(b)]. We use
the longer ˜
tand confirm that the ˜
tcalculated in this
way is longer than the rotational relaxation time. We find
that the magnitude of the active force F=0.57 pN in the
physical unit, consistent with typical force magnitudes in the
cytoplasm [28]. We remark that the constant magnitude of
the active force is crucial to obtain the correct scaling of
diffusion enhancement. In an alternative model with a con-
stant active speed, Ddilute/D0,dilute 1˜a4, inconsistent with
experiments. In a more biologically plausible scenario, the
magnitude of the active force can be random among differ-
ent particles [41]. Therefore, we also simulate a modified
model in which the active force obeys a size-independent nor-
mal distribution and obtain similar results (Fig. S12), further
strengthening our proposed mechanism.
Discussion. In this Letter, we introduce a white-noise ac-
tive force to a highly polydisperse system to mimic bacterial
cytoplasm. While prior works have investigated the effect of
activities on particle mobility [4244], our work simultane-
ously incorporates crowding by different particle sizes and
active forces. Due to its out-of-equilibrium nature, the FD the-
orem does not constrain the active random force. Surprisingly,
a white-noise active force reproduces the experimentally mea-
sured ratios of diffusion constants between living and dormant
bacteria with its autocorrelation function proportional to the
cube of the particle radius. We note that an active random
force generally generates an additional friction coefficient that
is inversely proportional to an active temperature [27,4547].
We show that the additional friction coefficient does not affect
our conclusions because the active temperature is typically
much higher than the thermal temperature (see Supplemental
Material [31]).
We further demonstrate that the white-noise active force
model with γ=3 is equivalent to the self-propelled model
with a constant-magnitude active force regarding the diffu-
sion enhancement. Our results suggest an emergent simplicity
when many active processes are averaged simultaneously.
Importantly, we identify the magnitude of the active force,
F=0.57 pN. While the origin of this active force is still
unclear [16], we hypothesize that it may come from the colli-
sion of small molecules with proteins, e.g., amino acids and
ions, which are out of equilibrium [45]. We note that the
active force can be estimated as the thermal energy divided
by typical protein sizes, F=kBT/a0.4 pN, where ais the
typical size of proteins around 10 nm. The active force applied
to proteins is just enough to overcome thermal fluctuation.
This particular magnitude of active force allows proteins to
move or change their configurations according to some spe-
cific intracellular signaling. On the other hand, it lets proteins
quickly change their dynamics when the signaling changes.
Therefore, the magnitude of the active force around the pN
range may be evolutionarily selected.
We note that the time-averaged MSDs of particles of the
same size differ among independent simulations. Neverthe-
less, the MSD averaged over the time-averaged MSDs of inde-
pendent simulations are close to the ensemble-averaged MSD
(Fig. S13), suggesting a weak nonergodicity effect, presum-
ably because polydispersity smears out the glass transition
(Fig. S11). Finally, we remark that while a hydrodynamic
interaction has been shown to reduce the diffusion coefficient,
its effect may be negligible for particles with a radius above
25 nm that we use to compare with experimental data [22].
Acknowledgments. We thank Yiyang Ye, Hua Tong, Sheng
Mao, and Ming Han for helpful discussions related to this
work. The research was funded by National Key R&D Pro-
gram of China (2021YFF1200500) and supported by grants
from Peking-Tsinghua Center for Life Sciences.
[1] P. H. von Hippel and O. G. Berg, Facilitated target location in
biological systems, J. Biol. Chem. 264, 675 (1989).
[2]M.B.Elowitz,M.G.Surette,P.-E.Wolf,J.B.Stock,andS.
Leibler, Protein mobility in the cytoplasm of escherichia coli, J.
Bacteriol. 181, 197 (1999).
[3] A. Chari and U. Fischer, Cellular strategies for the assembly of
molecular machines, Trends Biochem. Sci. 35, 676 (2010).
[4] J. T. Mika and B. Poolman, Macromolecule diffusion and con-
finement in prokaryotic cells, Curr. Opin. Biotechnol. 22, 117
(2011).
[5] S. Cayley, Jr., B. A. Lewis, H. J. Guttman, and M. T. Record, Jr.,
Characterization of the cytoplasm of Escherichia coli K-12 as a
function of external osmolarity: Implications for protein-DNA
interactions in vivo,J. Mol. Biol. 222, 281 (1991).
[6] S. B. Zimmerman and S. O. Trach, Estimation of macro-
molecule concentrations and excluded volume effects for the
cytoplasm of Escherichia coli,J. Mol. Biol. 222, 599 (1991).
[7] R. Swaminathan, C. P. Hoang, and A. Verkman, Photobleach-
ing recovery and anisotropy decay of green fluorescent protein
GFP-S65T in solution and cells: Cytoplasmic viscosity probed
by green fluorescent protein translational and rotational diffu-
sion, Biophys. J. 72, 1900 (1997).
[8] R. J. Ellis, Macromolecular crowding: obvious but underappre-
ciated, Trends Biochem. Sci. 26, 597 (2001).
[9] R. Milo and R. Phillips, CellBiologybytheNumbers(Garland
Science, New York, 2015).
[10] J. A. Dix and A. Verkman, Crowding effects on diffusion in
solutions and cells, Annu. Rev. Biophys. 37, 247 (2008).
[11] H.-X. Zhou, G. Rivas, and A. P. Minton, Macromolecu-
lar crowding and confinement: Biochemical, biophysical, and
potential physiological consequences, Annu. Rev. Biophys. 37,
375 (2008).
[12] I. Golding and E. C. Cox, Physical Nature of Bacterial Cyto-
plasm, Phys. Rev. Lett. 96, 098102 (2006).
[13] A. Nenninger, G. Mastroianni, and C. W. Mullineaux, Size
dependence of protein diffusion in the cytoplasm of Escherichia
coli,J. Bacteriol. 192, 4535 (2010).
[14] K. A. Dill, K. Ghosh, and J. D. Schmit, Physical limits of
cells and proteomes, Proc. Natl. Acad. Sci. USA 108, 17876
(2011).
L032018-5
MENG, JIN, GUAN, XU, AND LIN PHYSICAL REVIEW RESEARCH 5, L032018 (2023)
[15] S. C. Weber, A. J. Spakowitz, and J. A. Theriot, Nonthermal
ATP-dependent fluctuations contribute to the in vivo motion
of chromosomal loci, Proc. Natl. Acad. Sci. USA 109, 7338
(2012).
[16] B. Parry, I. Surovtsev, M. Cabeen, C. OHern, E. Dufresne,
and C. Jacobs-Wagner, The bacterial cytoplasm has glass-like
properties and is fluidized by metabolic activity, Cell 156, 183
(2014).
[17] M. Guo, A. Ehrlicher, M. Jensen, M. Renz, J. Moore, R.
Goldman, J. Lippincott-Schwartz, F. Mackintosh, and D. Weitz,
Probing the stochastic, motor-driven properties of the cy-
toplasm using force spectrum microscopy, Cell 158, 822
(2014).
[18] M. C. Munder, D. Midtvedt, T. Franzmann, E. Nuske, O.
Otto, M. Herbig, E. Ulbricht, P. Müller, A. Taubenberger, S.
Maharana et al., A pH-driven transition of the cytoplasm from
a fluid-to a solid-like state promotes entry into dormancy, eLife
5, e09347 (2016).
[19] R. P. Joyner, J. H. Tang, J. Helenius, E. Dultz, C. Brune, L. J.
Holt, S. Huet, D. J. Müller, and K. Weis, A glucose-starvation
response regulates the diffusion of macromolecules, eLife 5,
e09376 (2016).
[20] N. Chebotareva, B. Kurganov, and N. Livanova, Biochemical
effects of molecular crowding, Biochemistry (Moscow) 69,
1239 (2004).
[21] D. J. Bicout and M. J. Field, Stochastic dynamics simulations
of macromolecular diffusion in a model of the cytoplasm of
Escherichia coli,J. Phys. Chem. 100, 2489 (1996).
[22] T. Ando and J. Skolnick, Crowding and hydrodynamic inter-
actions likely dominate in vivo macromolecular motion, Proc.
Natl. Acad. Sci. USA 107, 18457 (2010).
[23] S. R. McGuffee and A. H. Elcock, Diffusion, crowding &
protein stability in a dynamic molecular model of the bacterial
cytoplasm, PLoS Comput. Biol. 6, e1000694 (2010).
[24] C. Wilhelm, Out-of-Equilibrium Microrheology inside Living
Cells, Phys.Rev.Lett.101, 028101 (2008).
[25] F. S. Gnesotto, F. Mura, J. Gladrow, and C. P. Broedersz, Bro-
ken detailed balance and non-equilibrium dynamics in living
systems: a review, Rep. Prog. Phys. 81, 066601 (2018).
[26] A. P. Solon, J. Stenhammar, R. Wittkowski, M. Kardar, Y. Kafri,
M. E. Cates, and J. Tailleur, Pressure and Phase Equilibria
in Interacting Active Brownian Spheres, Phys. Rev. Lett. 114,
198301 (2015).
[27] A. P. Solon, Y. Fily, A. Baskaran, M. E. Cates, Y. Kafri, M.
Kardar, and J. Tailleur, Pressure is not a state function for
generic active fluids, Nat. Phys. 11, 673 (2015).
[28] P. Roca-Cusachs, V. Conte, and X. Trepat, Quantifying forces
in cell biology, Nat. Cell Biol. 19, 742 (2017).
[29] M. Doi, Soft Matter Physics (Oxford University Press, Oxford,
UK, 2013).
[30] G. L. Hunter and E. R. Weeks, The physics of the colloidal glass
transition, Rep. Prog. Phys. 75, 066501 (2012).
[31] See Supplemental Material at http://link.aps.org/supplemental/
10.1103/PhysRevResearch.5.L032018 for additional discus-
sions and figures.
[32] M. Kumar, M. S. Mommer, and V. Sourjik, Mobility of cyto-
plasmic, membrane, and dna-binding proteins in Escherichia
coli,Biophys. J. 98, 552 (2010).
[33] A. H. Marcus, J. Schofield, and S. A. Rice, Experimental ob-
servations of non-Gaussian behavior and stringlike cooperative
dynamics in concentrated quasi-two-dimensional colloidal liq-
uids, Phys. Rev. E 60, 5725 (1999).
[34] E. R. Weeks, J. C. Crocker, A. C. Levitt, A. Schofield, and
D. A. Weitz, Three-dimensional direct imaging of structural
relaxation near the colloidal glass transition, Science 287, 627
(2000).
[35] W. K. Kegel and A. van Blaaderen, Direct observation of dy-
namical heterogeneities in colloidal hard-sphere suspensions,
Science 287, 290 (2000).
[36] E. Zaccarelli, S. M. Liddle, and W. C. Poon, On polydispersity
and the hard sphere glass transition, Soft Matter 11, 324 (2015).
[37] J. R. Howse, R. A. L. Jones, A. J. Ryan, T. Gough, R.
Vafabakhsh, and R. Golestanian, Self-Motile Colloidal Parti-
cles: From Directed Propulsion to Random Walk, Phys. Rev.
Lett. 99, 048102 (2007).
[38] Y. Fily and M. C. Marchetti, Athermal Phase Separation of Self-
Propelled Particles with No Alignment, Phys. Rev. Lett. 108,
235702 (2012).
[39] J. Bialké, T. Speck, and H. Löwen, Crystallization in a Dense
Suspension of Self-Propelled Particles, Phys. Rev. Lett. 108,
168301 (2012).
[40] G. S. Redner, M. F. Hagan, and A. Baskaran, Structure and
Dynamics of a Phase-Separating Active Colloidal Fluid, Phys.
Rev. Lett. 110, 055701 (2013).
[41] M. Xu, J. L. Ross, L. Valdez, and A. Sen, Direct Single
Molecule Imaging of Enhanced Enzyme Diffusion, Phys. Rev.
Lett. 123, 128101 (2019).
[42] R. Mandal, P. J. Bhuyan, M. Rao, and C. Dasgupta, Active flu-
idization in dense glassy systems, Soft Matter 12, 6268 (2016).
[43] C. Yuan, A. Chen, B. Zhang, and N. Zhao, Activity–crowding
coupling effect on the diffusion dynamics of a self-propelled
particle in polymer solutions, Phys. Chem. Chem. Phys. 21,
24112 (2019).
[44] L. Abbaspour and S. Klumpp, Enhanced diffusion of a tracer
particle in a lattice model of a crowded active system, Phys.
Rev. E 103, 052601 (2021).
[45] A. Solon and J. M. Horowitz, On the Einstein relation between
mobility and diffusion coefficient in an active bath, J. Phys. A:
Math. Theor. 55, 184002 (2022).
[46] A. Shakerpoor, E. Flenner, and G. Szamel, The Einstein effec-
tive temperature can predict the tagged active particle density,
J. Chem. Phys. 154, 184901 (2021).
[47] O. Granek, Y. Kafri, and J. Tailleur, Anomalous Transport of
Tracers in Active Baths, Phys. Rev. Lett. 129, 038001 (2022).
L032018-6
Preprint
Full-text available
The mechanical properties of cytoplasm are crucial for cellular functions. While active processes significantly alter cytoplasmic viscoelasticity, the physical mechanisms remain elusive. Here, we model the cytoplasm as a colloidal suspension subject to passive and active noise, coarse-grained as an effective temperature. We show that a jammed cytoplasm transitions from a solid to a liquid phase at a critical effective temperature. Intriguingly, the simulated complex shear modulus at the critical state exhibits a 1/2 power-law scaling with frequency and quantitatively matches the experimental data of live cells without fitting parameters, given that the cytoplasmic volume fraction is slightly above the jamming transition. We further reveal the biological significance for the cytoplasm to be slightly above jamming: it is a regime in which the viscosity is ultrasensitive to changes in the effective temperature. Our results suggest that cells actively regulate their volume fractions in the sensitive regime in which they can tune cytoplasmic viscosity efficiently via active processes.
Article
Full-text available
An active bath, made of self-propelling units, is a nonequilibrium medium in which the Einstein relation D = μk B T between the mobility μ and the diffusivity D of a tracer particle cannot be expected to hold a priori . We consider here heavy tracers for which these coefficients can be related to correlation functions which we estimate. We show that, to a good approximation, an Einstein relation does hold in an active bath upon using a different temperature which is defined mechanically, through the pressure exerted on the tracer.
Article
Full-text available
We derive the long-time dynamics of a tracer immersed in a one-dimensional active bath. In contrast to previous studies, we find that the damping and noise correlations possess long-time tails with exponents that depend on the tracer symmetry. For generic tracers, shape asymmetry induces ratchet effects that alter fluctuations and lead to superdiffusion and friction that grows with time when the tracer is dragged at a constant speed. In the singular limit of a completely symmetric tracer, we recover normal diffusion and finite friction. Furthermore, for small symmetric tracers, the active contribution to the friction becomes negative: active particles enhance motion rather than oppose it. These results show that, in low-dimensional systems, the motion of a passive tracer in an active bath cannot be modeled as a persistent random walker with a finite correlation time.
Article
Full-text available
Living systems at the subcellular, cellular, and multicellular levels are often crowded systems that contain active particles. The active motion of these particles can also propel passive particles, which typically results in enhanced effective diffusion of the passive particles. Here we study the diffusion of a passive tracer particle in such a dense system of active crowders using a minimal lattice model incorporating particles pushing each other. We show that the model exhibits several regimes of motility and quantify the enhanced diffusion as a function of density and activity of the active crowders. Moreover, we demonstrate an interplay of tracer diffusion and clustering of active particles, which suppresses the enhanced diffusion. Simulations of mixtures of passive and active crowders show that a rather small fraction of active particles is sufficient for the observation of enhanced diffusion.
Article
Full-text available
Living systems operate far from thermodynamic equilibrium. Enzymatic activity can induce broken detailed balance at the molecular scale. This molecular scale breaking of detailed balance is crucial to achieve biological functions such as high-fidelity transcription and translation, sensing, adaptation, biochemical patterning, and force generation. While biological systems such as motor enzymes violate detailed balance at the molecular scale, it remains unclear how non-equilibrium dynamics manifests at the mesoscale in systems that are driven through the collective activity of many motors. Indeed, in several cellular systems the presence of non-equilibrium dynamics is not always evident at large scales. For example, in the cytoskeleton or in chromosomes one can observe stationary stochastic processes that appear at first glance thermally driven. This raises the question how non-equilibrium fluctuations can be discerned from thermal noise. We discuss approaches that have recently been developed to address this question, including methods based on measuring the extent to which the system violates the fluctuation-dissipation theorem. We also review applications of this approach to reconstituted cytoskeletal networks, the cytoplasm of living cells, and cell membranes. Furthermore, we discuss a more recent approach to detect actively driven dynamics, which is based on inferring broken detailed balance. This constitutes a non-invasive method that uses time-lapse microscopy data, and can be applied to a broad range of systems in cells and tissue. We discuss the ideas underlying this method and its application to several examples including flagella, primary cilia, and cytoskeletal networks. Finally, we briefly discuss recent developments in stochastic thermodynamics and non-equilibrium statistical mechanics, which offer new perspectives to understand the physics of living systems.&#13.
Article
Full-text available
ELife digest Most organisms live in unpredictable environments, which can often lead to nutrient shortages and other conditions that limit their ability to grow. To survive in these harsh conditions, many organisms adopt a dormant state in which their metabolism slows down to conserve vital energy. When the environmental conditions improve, the organisms can return to their normal state and continue to grow. The interior of cells is known as the cytoplasm. It is very crowded and contains many molecules and compartments that carry out a variety of vital processes. The cytoplasm has long been considered to be fluid-like in nature, but recent evidence suggests that in bacterial cells it can solidify to resemble a glass-like material under certain conditions. When cells experience stress they stop dividing and alter their metabolism. However, it was not clear whether cells also alter their physical properties in response to changes in the environment. Now, Joyner et al. starve yeast cells of sugar and track the movements of two large molecules called mRNPs and chromatin. Chromatin is found in a cell compartment known as the nucleus, while mRNPs are found in the cytoplasm. The experiments show that during starvation, both molecules are less able to move around in their respective areas of the cell. This appears to be due to water loss from the cells, which causes the cells to become smaller and leads to the interior of the cell becoming more crowded. Joyner et al. also observed a similar response in bacteria. Furthermore, Joyner et al. suggest that the changes in physical properties are critical for cells to survive the stress caused by starvation. A separate study by Munder et al. found that when cells become dormant the cytoplasm becomes more acidic, which causes many proteins to bind to each other and form large clumps. Together, the findings of the studies suggest that the interior of cells can undergo a transition from a fluid-like to a more solid-like state to protect the cells from damage when energy is in short supply. The next challenge is to understand the molecular mechanisms that cause the physical properties of the cells to change under different conditions. DOI: http://dx.doi.org/10.7554/eLife.09376.002
Article
Cell cytoplasm contains high concentrations of high-molecular-weight components that occupy a substantial part of the volume of the medium (crowding conditions). The effect of crowding on biochemical processes proceeding in the cell (conformational transitions of biomacromolecules, assembling of macromolecular structures, protein folding, protein aggregation, etc.) is discussed in this review. The excluded volume concept, which allows the effects of crowding on biochemical reactions to be quantitatively described, is considered. Experimental data demonstrating the biochemical effects of crowding imitated by both low-molecular-weight and high-molecular-weight crowding agents are summarized.
Article
We derive a distribution function for the position of a tagged active particle in a slowly varying in space external potential, in a system of interacting active particles. The tagged particle distribution has the form of the Boltzmann distribution but with an effective temperature that replaces the temperature of the heat bath. We show that the effective temperature that enters the tagged particle distribution is the same as the effective temperature defined through the Einstein relation, i.e., it is equal to the ratio of the self-diffusion and tagged particle mobility coefficients. This result shows that this effective temperature, which is defined through a fluctuation–dissipation ratio, is relevant beyond the linear response regime. We verify our theoretical findings through computer simulations. Our theory fails when an additional large length scale appears in our active system. In the system we simulated, this length scale is associated with long-wavelength density fluctuations that emerge upon approaching motility-induced phase separation.
Article
The anomalous diffusion dynamics of an active particle in polymer solutions is studied based on a Langevin Brownian dynamics simulation. Firstly, the mean-square displacement (MSD) is investigated under various system parameters of active force FaF_a, probe size σa\sigma_a, polymer volume fraction ϕ\phi and polymer chain length N. A very novel transition between superdiffusion and subdiffusion is observed with varying FaF_a and ϕ\phi, owing to the activity and crowding competition effect. Diagram of the two anomalous diffusive regimes are identified in the parameter space. The increment of MSD under activity is examined in the intermediate time scales, which manifests a power law relation with the particle's dynamical persistence length L, i.e., \Delta \mbox{MSD}=2L^m where the exponent m decreases with ϕ\phi. Secondly, we explicitly evaluate the long-time diffusion coefficients Da0D_a^0 in pure solvent and DaD_a in polymer solutions. The dependence of relative diffusivity Da/Da0D_a/D_a^0 on volume fraction ϕ\phi reproduces the well-known Phillies' equation exp(κϕμ)\exp(-\kappa \phi^{\mu}). The fitting parameters show μ1\mu\simeq 1, but κ\kappa apparently increases with activity. More importantly, our simulation justifies a multi-length scaling relation in a very similar form to that for passive probes, depending on simple structural parameters of the probe-polymer system. By the aid of activation energy model, we find out a counterintuitive activity-crowding coupling effect: activity enhances effective viscosity experienced by the probe and thus strengthens the crowding-induced retardation to diffusion.
Article
Recent experimental results have shown that enzymes can diffuse faster when they are in the presence of their reactants (substrate). This faster diffusion has been termed enhanced diffusion. Fluorescence correlation spectroscopy (FCS), which has been employed as the only method to make these measurements, relies on analyzing the fluctuations in fluorescence intensity to measure the diffusion coefficient of particles. Recently, artifacts in FCS measurements due to its sensitivity to environmental conditions have been evaluated, calling prior enhanced diffusion results into question. It behooves us to adopt complementary and direct methods to measure the mobility of enzymes. Herein, we use a technique of direct single molecule imaging to observe the diffusion of individual enzymes in solution. This technique is less sensitive to intensity fluctuations and deduces the diffusion coefficient directly based on the trajectory of the enzyme. Our measurements recapitulate that enzyme diffusion is enhanced in the presence of its substrate and find that the relative increase in diffusion of a single enzyme is even higher than those previously reported using FCS. We also use this complementary method to test if the total enzyme concentration affects the relative increase in diffusion and if the enzyme oligomerization state changes during its catalytic turnover. We find that the diffusion increase is independent of the total concentration of enzymes and the presence of substrate does not change the oligomerization state of enzymes.