Download as pdf or txt
Download as pdf or txt
You are on page 1of 47

Physical Biology

TOPICAL REVIEW • OPEN ACCESS You may also like


- Advances in nanowire bioelectronics
Getting around the cell: physical transport in the Wei Zhou, Xiaochuan Dai and Charles M
Lieber
intracellular world - Quasi-Two-Dimensional Diffusion in
Adherent Cells Revealed by Three-
Dimensional Single Quantum Dot Tracking
To cite this article: Saurabh S Mogre et al 2020 Phys. Biol. 17 061003 Chao Jiang, , Bo Li et al.

- Intracellular FRET-based probes: a review


Clare E Rowland, Carl W Brown, Igor L
Medintz et al.
View the article online for updates and enhancements.

Recent citations
- Fluid flow and interface motion in gels: A
finite-strain theory and its application to a
channel flow problem
Fernando P. Duda et al

- Diffusive search and trajectories on tubular


networks: a propagator approach
Zubenelgenubi C. Scott et al

- Saurabh Mogre et al

This content was downloaded by cesarnc from IP address 190.239.215.40 on 24/11/2021 at 20:58
Phys. Biol. 17 (2020) 061003 https://doi.org/10.1088/1478-3975/aba5e5

O P E N AC C E S S TOPICAL REVIEW

Getting around the cell: physical transport in the intracellular


R E C E IVE D
world
31 March 2020
R E VISE D
11 June 2020
Saurabh S Mogre , Aidan I Brown and Elena F Koslover 1
Department of Physics, University of California, San Diego, San Diego, California 92093, United States of America
AC C E PTE D FOR PUBL IC ATION
1
14 July 2020 Author to whom any correspondence should be addressed
PUBL ISHE D E-mail: ekoslover@ucsd.edu
9 October 2020
Keywords: intracellular transport, organelle dynamics, cytoplasmic flow, diffusion, molecular motors, cellular biophysics

Original content from


this work may be used
under the terms of the Abstract
Creative Commons
Attribution 4.0 licence. Eukaryotic cells face the challenging task of transporting a variety of particles through the complex
Any further distribution intracellular milieu in order to deliver, distribute, and mix the many components that support cell
of this work must
maintain attribution to function. In this review, we explore the biological objectives and physical mechanisms of
the author(s) and the intracellular transport. Our focus is on cytoplasmic and intra-organelle transport at the whole-cell
title of the work, journal
citation and DOI. scale. We outline several key biological functions that depend on physically transporting
components across the cell, including the delivery of secreted proteins, support of cell growth and
repair, propagation of intracellular signals, establishment of organelle contacts, and spatial
organization of metabolic gradients. We then review the three primary physical modes of transport
in eukaryotic cells: diffusive motion, motor-driven transport, and advection by cytoplasmic flow.
For each mechanism, we identify the main factors that determine speed and directionality. We also
highlight the efficiency of each transport mode in fulfilling various key objectives of transport, such
as particle mixing, directed delivery, and rapid target search. Taken together, the interplay of
diffusion, molecular motors, and flows supports the intracellular transport needs that underlie a
broad variety of biological phenomena.

1. Introduction of localized regions with high metabolic demand.


In addition, key functional roles are attributed to
The movement of intracellular components, rang- physical contacts between multiple organelles, and
ing from ions and small metabolites to proteins and the formation of these contacts, as well as delivery
micron-scale organelles, underlies the vast majority of macromolecules to the contact zones, requires the
of cellular functions. Cellular transport needs vary regulated transport of cellular components.
from the nanoscale mixing that supports biomolec- In order to accomplish this diverse array of trans-
ular reaction kinetics, to delivery and sorting of car- port tasks, eukaryotic cells utilize several distinct
gos across whole-cell scales that can reach up to a physical mechanisms of transport (figure 1). For
meter in length. Newly synthesized proteins or mes- short distances and small (nanoscale) components,
senger RNA (mRNA) molecules must be transported stochastic ‘Brownian’ motion allows for mixing and
from their site of synthesis in perinuclear regions to rapid particle encounters. For longer distances and
distant peripheral locations for secretion or inser- larger particles, the cell harnesses the directed motion
tion into the plasma membrane. Cellular growth and of molecular motors along cytoskeletal filaments to
injury response, in particular, require a robust flux deliver vesicular organelles and RNA–protein com-
of components toward the newly synthesized regions plexes. The active transport machinery is controlled
of the cell. Conversely, external signals received at by a broad variety of regulatory factors that allow
the cell membrane often require the transport of for controlled sorting and distribution of cellular
activated protein molecules toward the nucleus in components. In addition, many cell types utilize
order to initiate a transcriptional response. Cellular advective flows of cytosolic fluids to rapidly drive par-
metabolism necessitates the efficient distribution of ticles through the cytoplasm. Each of these transport
ATP and metabolites to all subcellular regions. In modes is embedded in a highly complex, crowded,
large cells such as neurons, the spatial organization of and actively fluctuating intracellular environment.
metabolism is key to supporting the energetic needs Consequently, understanding the movement of

© 2020 The Author(s). Published by IOP Publishing Ltd


Phys. Biol. 17 (2020) 061003 Topical Review

Figure 1. Overview of common physical mechanisms for intracellular transport. (a) Diffusive motion. Bottom: diffusion of
membrane protein (Sec61) in peripheral tubular network of the ER, in COS7 cell (data from [1]) (b) motor-driven transport
along cytoskeletal highways. Bottom: lysosome trajectory (red) and microtubules (green) in monkey kidney cell (image from [2]).
(c) Advection in a flowing cytoplasm. Bottom: trajectories of acidified organelles in migrating HL60 cell (data from [3]).

cellular components requires expanding the classic There are several existing reviews in the literature on
models of physical transport processes to incorporate the molecular components and biochemical regula-
the unique milieu inside a living cell. In this sense, tion of motor-driven transport [18–20]. Our focus
cell biology can serve as a source of inspiration for here is on the quantitative exploration of transport
new fundamental questions in fields such as soft at the whole-cell scale, including diffusive, motor-
matter, non-equilibrium statistical mechanics and driven, and advective motion. Given the broad diver-
stochastic processes. sity of tranport systems in different cell types, we
A number of studies have explored the con- focus on animal cells where possible, touching upon
nection between defects in intracellular transport other eukaryotic cell systems when needed to illus-
and human pathologies (reviewed in [4, 5]). Owing trate specific physical effects. Furthermore, we restrict
to their spatially extended structure, human neu- the discussion primarily to transport in the cytoplasm
rons are especially susceptible to diseases linked with and within cytoplasmic organelles. The movement
transport defects. Neurodegenerative disorders such of particles across semipermeable membranes and
as multiple sclerosis, amyotrophic lateral sclerosis, within the nucleus is not considered.
Parkinson’s disease, and Alzheimer’s, among oth- We begin with a brief overview of the general
ers, are attributed to disruption of axonal transport properties of transport, in section 2. We then address
by mutations or other abnormalities [6–9]. Primary the broad biological question: why do cells require
cilia in mammalian cells [10] provide a non-neuronal transport? In section 3 we summarize several key
example of cellular structures that rely on functional functional roles played by transport processes in
transport processes for their formation. Defects in the cell, noting the relevant length and time scales.
motor proteins result in abnormal ciliary structures In section 4 we proceed to discuss the fundamen-
which are linked to developmental defects, lung dis- tal mechanisms of intracellular transport: diffusion,
ease, and hearing loss [11]. A number of pathogenic motors, and advection. In each case, we outline key
viruses are also known to hijack the intracellular physical parameters that govern transport efficiency
transport machinery to deliver them to different cel- and organization, as well as identifying the cellu-
lular regions and aid in uncoating, replication, and lar mechanisms that modulate those parameters. In
packaging [12–14]. For example, calciviruses rely on section 5 we highlight some outstanding physical
acidification within the endocytic pathway for their questions regarding intracellular transport. The over-
replication, a process dependent on vesicular trans- arching aim of this review is to provide a broad
port [15]. The Ebola virus and some coronaviruses overview of the physics of transport in animal cells,
have also been shown to depend upon intracellular highlighting those aspects that support biological
trafficking to late endosomes and lysosomal vesicles function.
prior to release into the cytoplasm [16, 17]. A quan-
titative understanding of the limitations and con- 2. Fundamentals of transport
sequences of intracellular transport is thus critical
to unraveling the mechanistic basis for a variety of Physical mass transport (as distinct from heat or
human pathologies. information transfer), is defined by the move-
In this review, we explore the biological objectives ment of particles between different spatial regions.
and physical mechanisms of intracellular transport. Transport behavior is generally characterized by the

2
Phys. Biol. 17 (2020) 061003 Topical Review

relationship between the length scale explored by undergoing diffusion superimposed on an underlying
the particles and the transport time. The nature of cytoplasmic flow [27–29].
this relationship is itself determined by the transport The relative contribution of directed versus effec-
mechanism (the forces that drive particle motion), as tively diffusive transport is characterized by the
well as the properties of the environment in which dimensionless Péclet number [30–32]:
transport occurs.
Intracellular transport, in particular, takes place Pe(L) = vL/D, (2)
within a dense aqueous medium where the motion This number gives the ratio of time required to tra-
of any particle necessitates flow or rearrangement of verse a region of length L by diffusion (t ∼ L2 /D) and
the surrounding fluid. Consequently, the response of by directed motion (t ∼ L/v). High values (Pe  1)
a particle to applied force is determined in part by the indicate that processive transport is dominant.
hydrodynamic properties of the cytoplasmic medium. Because diffusivity increases inversely with particle
The importance of viscous versus inertial forces in a size [23, 24, 33] while motor-driven and flow-driven
fluid is governed by the Reynolds number: transport tend to be size-independent [34], the
vLρ Péclet number is particularly high for large particles
Re = , (1) transported over long distances.
η
Even slower scaling of distance explored ver-
where v is the flow speed, L the characteristic linear sus time (γ < 0.5) arises when particles undergo
size scale of the object in motion, ρ the fluid den- so-called subdiffusive motion [35]. This form of
sity, and η the fluid viscosity [21, 22]. Within a typical transport (discussed further in section 4.1.2) is char-
animal cell, the relevant length scale is generally L < acterized by negative correlations in particle veloci-
100 μm and transport velocities are v < 100 μm s−1 . ties during consecutive time-steps [36]. Such an effect
Even assuming a density and viscosity of pure water, can arise, for example, for particles that must push
the intracellular world has Re < 0.01, and is thus through a viscoelastic medium such as a polymer
well in the regime of low Reynold’s number hydrody- gel [33, 37, 38]. The cytoplasmic transport of large
namics. Consequently, inertial forces inside a cell are protein complexes and organelle-sized particles is
negligible relative to viscous forces, and the instanta- generally observed to exhibit subdiffusive behavior
neous velocity rather than the acceleration of a par- [39–41].
ticle is determined by the applied force. For example, The relevant lengths and times for intracellular
if we consider a vesicle of size 1 μm moving at speed transport vary broadly depending on the cell size, the
1 μm s−1 in water, when the force pushing that particle type, and the functional role of the trans-
organelle is removed it will coast a distance of less than port process. The need to transport material between
10−4 nm before coming to a stop [22, 23]. the cell surface and the bulk has been suggested as
The relationship between length scale covered a fundamental physical limitation on cell shape and
and transport time can often be expressed as a size [32, 42–45]. At one extreme is the transport of
power-law L ∼ tγ . In the intracellular world, given the small metabolites (∼1 nm in size, D ≈ 200 μm2 s−1
dominance of viscous versus inertial forces, the scal- [46, 47]) between the cell periphery and metabolic
ing exponent is generally in the range of 0 < γ  1. organelles in globular cells such as fibroblasts, over
For directed motion, driven by a constant force, we length scales on the order of ∼ 2 μm. Diffusive trans-
have γ = 1 and the particle moves at a constant port is sufficient in this case to allow delivery in about
velocity (L = vt). This type of motion is seen 20 ms. At the other extreme is the transport of vesi-
for the transport of cellular particles attached to cles (∼ 100 nm in size, D ≈ 0.01 μm2 s−1 [26, 27,
active molecular motors (section 4.2) or for those 34]) over the meter-long length of neuronal axons in
driven by large-scale flows of the intracellular fluid the human peripheral nervous system. For this pur-
(section 4.3). The velocity v, of course, can be both pose, diffusive transport would require over a million
position- and time-dependent. However, so long as it years and is clearly impractical. Even motor-driven
has a finite average value, the long-time transport will transport (at a typical rate of 1 μm s−1 [48]) requires
obey this scaling behavior. about 10 days to deliver particles from the cell body
By contrast, a different scaling of length versus to the tips of these long cellular projections. Time-
time [L ∼ (Dt)0.5 ] is expected for particles whose scales that may be considered physiologically relevant
transport behavior resembles a random walk. This for a given transport process also vary by many orders
includes diffusive particles (with diffusivity D) in a of magnitude. A turnover time of a week to deliver
viscous fluid, whose steps are uncorrelated over all new mitochondria to distal regions of an axon seems
time-scales. It also includes the long-time behavior of to be sufficient to maintain a homeostatic population
particles that switch the direction of transport many of these energy-producing organelles [48]. On the
times, without retaining a memory of their previ- other hand, the most rapid intracellular enzymes can
ous motion [24]. Many cellular components engage catalyze reactions with microsecond turnover [49],
in multiple forms of transport, switching between necessitating the delivery of reactants over these very
diffusive and motor-driven states [19, 25, 26], or rapid time-scales.

3
Phys. Biol. 17 (2020) 061003 Topical Review

The many functional roles of intracellular trans- compartments’ (ERGICs) [63]. In mammalian cells,
port (section 3) span across the broad range of rele- the ERGIC are thought to mediate transit to the Golgi
vant length and time scales. In addition, each comes by generating membrane-bound compartments of
with its own limitations in terms of the amount of varying size that are transported along microtubule
material that must be transported and the necessity highways [64–67]. Microtubules in mammalian cells
for precise control over where, when, and which intra- tend to be polarized with their minus ends anchored
cellular components are transported. Cells thus rely at a microtubule organizing center (MTOC) proximal
on several complementary physical transport mecha- to the Golgi, enabling rapid delivery of cargo-carrying
nisms (section 4) to address their functional transport compartments to the Golgi by minus-end directed
needs. dynein/dynactin motor complexes [65]. Mobile struc-
tures, generated from the ERGIC and carrying
3. Functional roles for transport ER-derived proteins, have been observed to move
processively on curvilinear trajectories toward the
A fundamental question underlies, explicitly or indi- centrally located Golgi, over distances of 6–20 μm
rectly, all studies of intracellular transport—what are [65, 67].
the functional objectives or consequences of any given During the early secretory pathway, transport
transport system? In this section, we outline several of membrane-bound vesicular organelles serves not
key categories of biological functions that rely on only to physically move proteins between compart-
intracellular transport processes. The broad diversity ments located in different areas of the cell, but also to
of these functions suggests a variety of metrics for the mediate quality control and protein sorting [55, 63,
utility of a transport process. While some transport 68, 69]. Packaging of proteins for ER exit relies on a
systems need to be optimized for rapid delivery of combination of factors, including specific binding of
components to a specific target within the cell, others secretion tags to ERES scaffolding proteins [70–72],
require efficient mixing and uniform distribution of bulk transport of small proteins captured within
particles throughout a cellular region. In some cases nascent vesicles [62, 73], and clustering of membrane
a stable transport infrastructure is sufficient to meet proteins with similar transmembrane domain lengths
cellular needs over long time periods, whereas other that are poorly matched to the thickness of the ER
systems require the ability to respond quickly to varia- membrane [74, 75]. Together, these factors combine
tions in the desired flux or target location of delivered to prevent many misfolded proteins or ER-resident
particles. proteins from being transported out of the ER. Once
at the Golgi, a recycling pathway relies on the motor-
3.1. Delivery of secreted and plasma membrane driven transport of vesicles coated with coat protein
proteins I (COPI) to shuttle transport receptors and leaked
One major functional role for intracellular transport ER-resident proteins back into the ER, maintaining
is to drive the secretory pathway (figure 2). Proteins proteostasis within the organelles [76].
destined for extracellular secretion or insertion into In the Golgi, proteins are further processed
the plasma membrane are manufactured by ribo- and decorated with post-translational modifications,
somes attached to the rough endoplasmic reticulum while passing from the perinuclear cis-Golgi region
(ER), generally located adjacent to the cell nucleus to the trans-Golgi side. The mechanism of transport
[53, 54]. Such proteins are inserted co-translationally within Golgi compartments remains under debate
into the ER lumen or membrane, wherein they are [77] and may include vesicular transport [78], pro-
folded and processed before moving into an ER exit gression and maturation of transient cisternae [79],
site (ERES) [55, 56], as illustrated in figure 2(a). While or rapid partitioning between phase-separated lipid
transport within the ER is generally assumed to be domains [80]. Recent theoretical work indicates that
diffusive in nature [57, 58], recent evidence from sin- multiple mechanisms can be encompassed by a
gle particle tracking studies implies the existence of kinetic model that relies on tuning of vesicle fusion
short-range processive movements that push proteins and budding rates to achieve optimal sorting [81].
rapidly from node to node within the tubular ER From the trans-Golgi cisternae, secretory proteins are
network [59]. Exit sites appear as distinct long-lived sorted into a network of membranous tubules that are
puncta [60] scattered throughout both the perinu- extruded by the action of kinesin motors pulling along
clear and peripheral ER network (figure 2(b)). There microtubule tracks [82]. The tubes are then cleaved
are on the order of 200 ERES [56, 61] in a typical- to create pleiomorphic membranous carriers that are
sized mammalian cell (40 μm in diameter), implying transported to the plasma membrane for secretion
that proteins must explore over a spatial distance of [83]. After fission from the trans-Golgi network, car-
roughly 3 μm to encounter a site for ER exit. riers are transported by kinesin motors across dis-
At the ERES, proteins are packaged into vesicles tances on the order of 10 μm in a typical mammalian
coated with coat protein II (COPII) [62], which bud cell [84].
from the ER, shed their coats, and fuse into vesicular- Measurements of secretory pathway kinetics, via
tubular clusters termed the ‘ER-Golgi intermediate a retention and synchronized release system [85],

4
Phys. Biol. 17 (2020) 061003 Topical Review

Figure 2. Transport processes in the early secretory pathway. (a) Newly-synthesized proteins are inserted into the ER lumen or
membrane. After folding, these proteins are trafficked to the Golgi in vesicular bodies that form with the aid of COPII coat
proteins. Retrograde trafficking of COPI-coated vesicles from the Golgi to the ER allows for homeostasis of ER-resident proteins.
(b) Proteins must find punctate ERES to leave the ER and proceed along the secretory pathway. Green signal shows COPII
proteins at ERES, while red signal shows ER structure via an ER-resident luminal marker protein [50]. (c) Schematic of mRNA
transport and local translation in neuronal axons, involving motor-driven transport from the soma along the cytoskeleton to
provide mRNA for translation at axonal terminals (adapted from [51]). (d) Dendritic Golgi outposts in rat hippocampal neuron
indicate sites of local secretory processing (from [52], Copyright 2013, Society for Neuroscience).

indicate that newly released proteins are exported compartments, and processive transport of vesicles is
from the ER within 2–3 min, reach the Golgi within then required to enable sufficiently rapid delivery to
10 min, and are secreted at the plasma membrane the cell periphery.
within 20 min. Given a directed transport rate Highly extended cell types such as neurons face a
for motor-driven exocytic vesicles of approximately particularly challenging transport problem to deliver
1 μm s−1 , and a typical distance of 10 μm from the components manufactured near the nucleus to dis-
nucleus to the cell periphery, the transport of pro- tant secretion regions that can be up to a meter
teins across the cell does not appear to be rate-limiting away. Neuronal axons are capable of rapidly releas-
in the secretory pathway, at least in globular ani- ing large quantities of secreted neurotransmitter
mal cells. Notably, however, diffusion coefficients of proteins at the presynaptic terminals located on their
vesicular organelles in cytoplasm tend to be in the distal tips. Rapid variation in the complement of neu-
range of 0.002–0.08 μm2 s−1 [27, 34, 86], implying a rotransmitter receptors expressed on the dendritic
time-scale of several hours to traverse the cell by post-synaptic membrane plays an important role in
diffusion alone. Thus, motor-driven transport is a synaptic plasticity and adaptation [87]. The criti-
fundamental necessity for maintaining the complex cal need to control secreted and membrane protein
secretion processes of eukaryotic cells. Vesicular pack- availability at the distant tips of axons and dendrites
aging of proteins provides a functional benefit in raises the question of how the proteins themselves
allowing regulated protein sorting between different or the components needed for their manufacture are

5
Phys. Biol. 17 (2020) 061003 Topical Review

transported across such long distances from the cell which consists of sporadic bidirectional motion with
nucleus. Many synaptic proteins are manufactured at average rates on the order of 0.05 μm s−1 [104]. The
the cell body via the canonical secretory pathway [88]. origin of this transport mechanism remains under
They are then sorted into post-Golgi vesicles bound debate [105], but it has been suggested to arise from
toward either axonal or dendritic compartments and transient interactions with molecular motors [106] or
delivered to their eventual destinations by long-range entrainment in cytoplasm dragged by passing motor-
motor-driven transport along microtubule highways driven organelles [107].
[87, 89]. Even with rapid unidirectional motor-driven Motor-driven transport of tubulin also plays an
motion, a delivery time on the order of 10 days is important role in the extension of cellular projec-
required to transport somatically synthesized proteins tions such as flagella and primary cilia [108, 109]. The
to the end of a meter-long axon. dynamics of the intraflagellar transport (IFT) trains
More efficient response to changing protein responsible for tubulin delivery are crucial to regulat-
requirements at axonal and dendritic terminals can be ing the length distribution of these organelles [110,
achieved by local protein translation (figure 2(c)). The 111].
existence of rough ER, ERGIC, and Golgi outposts at In addition to cytoskeletal components, an axonal
distal dendritic regions (figure 2(d)) allows secretory growth cone also requires the continuous incorpo-
protein synthesis and modification to proceed with- ration of new proteins and lipids. While a number
out the need for delivery to and from the cell body of proteins are locally translated at the growth cone
[90]. Emerging evidence indicates that local transla- [112], many others are delivered by Golgi-derived
tion at axonal terminals is prevalent, particularly in vesicles that also serve as a source of membrane upon
the context of development, regeneration, and repair eventual fusion with the growth cone tip [96]. Such
[51, 91–93]. Local translation bypasses the problem vesicles have been shown to accumulate at the plus
of long-range protein delivery but does require trans- ends of microtubules in newly formed growth cones
port of mRNA, which is usually bound by RNA- of regenerating axons [113]. In a growing axonal tip,
binding proteins that couple directly to molecular the accumulation of these vesicles can out-pace their
motors [94]. This transport system allows for a con- incorporation into the growth cone, leaving behind
stant, relatively slow, turnover of mRNA molecules organelle-filled varicosities (figure 3(a), right) that
at distal translation outposts, while enabling rapid then serve as nascent pre-synaptic structures [95].
variation in protein manufacture and secretion in The transport of protein-filled vesicles from the soma
response to local signals. to the axonal tip, balanced against the rate of deliv-
ery and incorporation of structural growth cone com-
3.2. Distribution of components for growth, ponents, thus plays an important role in both axon
injury repair, and cell division growth and the placement of pre-synaptic terminals.
Growing and regenerating cells require the delivery of In several cell types, rapid vesicle transport has
a broad array of structural components to supply the the additional function of plugging holes in the
necessary material for growth and repair in specific plasma membrane generated by cellular injury. A
regions. In addition, cell division and the separation severed axon seals its plasma membrane, over a time-
of a syncytium into distinct cellular regions (as in fun- scale of minutes to hours, with the aid of multi-
gal hyphae and animal embryonic development [98, vesicular structures derived from endocytosis along
99]) necessitates the maintenance of a controlled dis- the axon membrane followed by transport of the
tribution of proteins and organelles to ensure appro- resulting vesicles to the cut end [96, 114] (figure 3(b)).
priate partitioning into the newly formed cells. All of Certain fungi form extensive multi-cellular hyphae,
these processes require intracellular transport of com- where individual cells are separated by perforated
ponents, often along substantial cellular distances. septa that allow for free passage of cytoplasmic con-
Neurons again present an important example tents. In case of injury, peroxisome-derived organelles
where long-range transport is required for growth. called Woronin bodies are rapidly delivered, primar-
Axonal growth during development and regenera- ily through bulk cytoplasmic flow, to plug up septal
tion is mediated by a distal growth cone structure pores and prevent large scale loss of cytoplasm [115].
that contains both the cytoskeletal components that A further critical role for organelle transport in
drive growth and an abundance of regulatory factors growth and development is to maintain a spatially
that determine growth rate and direction [100]. Axon well-mixed distribution of organelles, allowing for
protrusion is dependent on the delivery of micro- equitable partitioning during cell division or cellular-
tubule components to the tip of the growth cone ization. In mammalian cells, motor-driven transport
[101, 102] (figure 3(a)) where their incorporation of mitochondria is required for maintaining their dis-
into the axonal shaft both directly drives extension tribution throughout the soma [116], and in yeast
and contributes to mechanical forces that stretch the cells an active transport mechanism is used to parti-
axonal axis [103]. Tubulin monomers are translated tion and sort mitochondria between the mother cell
in the cell body and delivered to the growth cone via and the bud [97] (figure 3(c)). Furthermore, motor-
the so-called ‘slow component’ of axonal transport, driven transport enhances the fission and fusion of

6
Phys. Biol. 17 (2020) 061003 Topical Review

Figure 3. Transport processes in cell growth, division, and healing. (a) Delivery to axonal growth cone. Left: motor-driven
transport of tubulin, growth factors, and vesicles supports and directs growth cone protrusion. Right: varicosity in an advancing
growth cone due to accumulation of vesicles (image from [95]). (b) Role of cytoplasmic transport in axonal injury response. The
membrane at the injured end of an axon is sealed by accumulation of anterograde-moving motor-driven vesicles. Image adapted
from [96]. (c) Schematic of mitochondrial rearrangement during cell division in budding yeast. Image adapted from [97].

mitochondria [117], which can switch between glob- (figure 4(a)). A well-known example is the JAK/STAT
ular and extensively networked structures to facili- pathway, where an activated transmembrane
tate homogenization of mitochondrial contents [118, receptor JAK (Janus kinase) phosphorylates latent
119]. Other membrane-bound organelles such as transcription factors STATs (signal transducer and
peroxisomes also rely on microtubule-based trans- activator of transcription proteins) that reside in the
port mechanisms for controlling segregation between cytoplasm [124]. These factors diffuse throughout
dividing cells. In mammalian cells, peroxisomes con- the cell until they encounter the nucleus, where
gregate at spindle poles to ensure equitable parti- their phosphorylated nuclear localization sequence
tioning, in yeast they are delivered directly to the enables nuclear import, triggering subsequent
nascent bud, and in fungal hyphae they hitchhike cellular response through the regulation of gene
on other motile organelles to allow rapid equilibra- expression.
tion throughout the growing hypha [120]. An effi- A related approach is exemplified by several recep-
cient transport process to either deliver the organelles tor tyrosine kinase signaling pathways, including
to specific cellular regions or to maintain a uniform Notch and insulin signaling, where receptor acti-
distribution of organelles throughout the cell is thus vation at the plasma membrane triggers cleavage
necessary for homeostasis of organelle content in of a soluble intracellular domain that binds to a
growing and dividing cells. cytoplasmic transcription factor and escorts it dif-
fusively toward nuclear import sites [125]. A simi-
3.3. Intracellular signal propagation lar strategy is employed by a branch of the unfolded
Given the complex spatial organization of eukary- protein response pathway, in which accumulation of
otic cells, signals from the extracellular environment misfolded proteins in the ER triggers the transport
received at the cell periphery must be propagated over of the ATF6 (activating transcription factor 6) trans-
substantial distances to reach the nucleus or other membrane protein from the ER to the Golgi. In the
distant cellular regions. In certain specific cases, such Golgi, ATF6 is cleaved to release a cytoplasmic domain
as the action potential in neurons or mitotic signaling that diffuses to the nucleus and serves as a transcrip-
in oocytes, these signals can propagate very rapidly tion factor to upregulate the expression of chaperones
by a ‘trigger wave’ mechanism, that involves local promoting protein folding [126] (figure 4(b)).
diffusion of activating factors that trigger a switch- The speed and efficiency of signal propagation to
like self-propagating response [121–123]. Many the nucleus using these diffusive mechanisms is lim-
signaling pathways, however, rely on the physical ited by both the mobility of proteins in the cytoplasm
transport of specific proteins from the cell periphery and the timescale of deactivation and turnover of the
to the nucleus, where they can activate a response signaling proteins. The typical diffusivity of globu-
through transcriptional regulation. lar proteins in mammalian cytoplasm is in the range
A simple approach to transporting a signal of 3–30 μm2 s−1 [127–129], so that a signal from the
across relatively small cellular distances relies on plasma membrane would take on the order of 10 s to
the diffusion of an activated protein to the nucleus reach the nucleus in a modestly-sized cell of radius

7
Phys. Biol. 17 (2020) 061003 Topical Review

Figure 4. Common pathways of signal propagation through physical transport of molecular components. (a) Signals activate
cytoplasmic transcription factors which diffuse toward the nucleus. Schematic of the JAK/STAT signaling pathway is shown.
(b) Signals result in translocation of receptor to another organelle and/or cleavage of activated region which diffuses to the
nucleus and serves as a transcription factor. The ATF6 branch of the unfolded protein response pathway uses both mechanisms.
(c) Activated receptors are encapsulated in vesicles and carried via motor-driven transport toward the nucleus or to contact other
organelles which trigger deactivation, degradation, or recycling. The EGFR pathway is an example of this propagation
mechanism. (d) IFT machinery controls receptor localization and turnover in primary cilia. Motor-driven motion along a central
microtubule bundle allows for anterograde transport of inactive receptors to cilia tips and retrograde transport of activated
receptors to the cell body.

15 μm. Given that dephosphorylation times for acti- membrane or to multivesicular bodies and late endo-
vated proteins tend to be on the order of 1 s, such somes [135]. Fusion of these organelles with lyso-
signals would be attenuated to non-detectable lev- somes carrying proteolytic enzymes eventually results
els before they ever reached the nucleus [130]. In in cargo degradation, leading to attenuation of the
small cells, the signal can be propagated over suffi- signal. Other pathways, such as EGFR (epidermal
cient distances by cascades of sequential phosphory- growth factor receptor) signaling, rely on phos-
lation of multiple cytoplasmic proteins, as occurs in phatases localized to the perinuclear ER to dephos-
the mitogen-activated protein kinase (MAPK) path- phorylate and shut off active receptors [137, 138].
way [131]. For larger animal cells, however, diffusive Thus, the transport processes that shuttle endosomes
transport of activated proteins is too slow to be of to different cellular regions and facilitate organelle
practical use in signaling. For example, an activated interactions play an important role in regulating the
peripheral protein would require several hours to dif- duration and time-course of signaling events [139,
fuse to the nucleus in a 1 mm frog egg, and several 140].
months to diffuse from the distal tip of a centimeter- An additional transport process crucial to intra-
long axon to the cell body. cellular signaling is the IFT that moves proteins within
Many signaling pathways intertwine with the primary cilia (figure 4(d)). Primary cilia are nar-
endocytic pathway, leveraging vesicular encapsula- row cellular projections, roughly 5–10 μm long and
tion and motor-driven transport to deliver activated 0.3 μm in width that serve as a signaling nexus in
components to regions near the nucleus. A canon- many mammalian cell types and play an important
ical example of signaling via retrograde transport role in development, vision, and olfaction [10, 141].
is the neurotrophic signaling pathway that regulates Signaling receptors are concentrated on the ciliary
neuronal survival, axon and dendrite growth, and membrane in a highly regulated manner that relies on
synapse formation [132]. Neurotrophin growth fac- their transport into, out of, and throughout the cil-
tors bind to receptors on the distal tips of axonal ium by coupling to trains of molecular motors that
projections, which are packaged into endosomes and move them along the central bundle of microtubules
carried to the cell body by dynein motors walking [142, 143]. A particularly well-characterized example
along microtubule highways [133, 134]. For a meter- is hedgehog signaling, which plays a key role in tissue
long axon, this process takes approximately 10 days, development and homeostasis. The hedgehog ligand
putting a substantial limit on the ability of the neuron receptor, patched, accumulates in primary cilia in the
to respond to distal growth signals. absence of signaling, and is exported from the cilium
In general, a broad variety of signaling cascades is upon activation, thereby allowing the ciliary entry
known to involve packaging and activation of compo- and accumulation of other receptors such as Smo and
nents within endosomes [135] (figure 4(c)). Motor- Gli proteins [144]. The latter, in turn, are activated
driven transport of the endosomes can rapidly deliver within the cilium, transported to the ciliary base,
activated signals to the nucleus, as in the case of Smad and from there relocate to the nucleus where they
proteins activated in the transforming growth fac- act as a transcription factor regulating gene expres-
tor beta signaling pathway [136]. Alternatively, early sion [145, 146]. Mutations in adaptor proteins that
endosome-encapsulated receptors can be trafficked form the complex connecting signaling receptors to
to a recycling compartment for return to the cell IFT motors result in failure of signaling receptors to

8
Phys. Biol. 17 (2020) 061003 Topical Review

Figure 5. Interactions between various organelles. (a) Matrix representation of absolute number of contacts between lysosomes
(Lyso), the Golgi body, the ER, mitochondria (Mito), peroxisomes (Perox), and lipid droplets (LDs) within a single cell. Color
denotes number of contacts between each pair of organelles. (b) Network representation of interaction frequency. The length of
an edge represents the inverse number of contacts between organelles at each end. Figures adapted from [149].

localize to cilia and/or abnormal accumulation of In addition to lipid transfer and signal prop-
activated receptors within the cilium [147, 148]. IFT agation (discussed in the previous section),
is thus critical for regulating the spatial organiza- inter-organelle contacts can themselves facilitate the
tion of ciliary receptors as well as downstream signal transport and morphological dynamics of the par-
propagation. ticipating organelles. For example, peroxisomes and
lipid droplets have both been shown to hitchhike
on early endosomes [167, 168], allowing them to
3.4. Organelle interaction and exchange
move rapidly through the cell by attaching to mobile
Membrane-bound organelles are topologically dis-
carrier organelles. Contacts between ER tubules
tinct compartments within eukaryotic cells that serve
and mitochondria are known to be required for
to spatially organize a broad array of intracellu-
fission of mitochondrial networks into globular
lar reactions. Recent measurements have highlighted
structures, which can be redistributed by transport
the plethora of direct physical interactions between
processes throughout the cell [169]. Furthermore,
different organelle structures [149], and the biolog-
motor-driven transport along the cytoskeleton allows
ical role of these inter-organelle contacts is increas-
for the formation of mitochondrial networks through
ingly appreciated [150–152]. Lipid droplets (LDs),
fusion, allowing for the mixing of mitochondrial
mitochondria, peroxisomes, lysosomes, endosomes,
contents on a cellular scale [170].
the ER, and the Golgi complex all form an exten-
Experimental evidence suggests that disrupting
sive dynamic network of interacting organelles that
the cytoskeleton affects many features of the organelle
coordinate and colocalize with each other (figure 5).
interactome [149]. Transport processes thus play an
The establishment and turnover of contact sites relies
important role in modulating organelle interactions
on intracellular transport to place regions of different
that are crucial for cellular function.
organelles in spatial proximity.
One well-established role for organelle contacts
is lipid homeostasis and metabolism. Lipids are 3.5. Control of nutrient and metabolite gradients
synthesized in the ER, stored and transported in Several studies have pointed toward the existence
lipid droplets, metabolized in mitochondria and of substantial intracellular gradients in nutri-
peroxisomes, and recycled in lysosomes [153–157]. ents, metabolites, and ATP [174–176], prompting
Colocalization of lipid droplets with mitochon- increased interest in unraveling the spatial hetero-
dria and lysosomes, in particular, is essential for geneity of metabolism [177, 178]. Although small
fatty acid metabolism and starvation response [157, metabolites diffuse rapidly through the cytoplasm
158]. The organelles involved in lipid turnover are (with diffusivity of around 200 μm2 s−1 for glucose
generally distributed throughout the cell, allow- and ATP [46, 47]) such gradients can arise as a result
ing for frequent transient contacts that permit of locally enhanced metabolism in the vicinity of
signaling and delivery of components [159]. Main- mitochondria or rapid ATP consumption in localized
taining the relatively uniform distribution of per- cellular regions. An additional source of metabolite
oxisomes, lipid droplets, and ER tubules requires gradients is extracellular spatial heterogeneity in
bidirectional motor-driven transport along micro- nutrient levels [171], or spatial variation in the
tubule highways [25, 26, 160–162]. The vesicular density of transporter proteins allowing nutrient
nature of lipid droplets, in particular, makes them import into the cell [172, 179] (figures 6(a) and (b)).
well-suited for targeted transport of lipids to specific Intracellular transport and positioning of mitochon-
cellular regions with distinct metabolic requirements dria, glucose transporters, and a variety of metabolic
[163, 164]. Mitochondria and lysosomes also move in enzymes thus have a key role to play in maintaining
a regulated fashion along microtubules to enable the the spatial organization of metabolism, particularly
spatial organization of metabolism and lipid recycling in large cells such as oocytes, neurons, and plant
[165, 166]. cells. The mitigation and control of metabolite

9
Phys. Biol. 17 (2020) 061003 Topical Review

Figure 6. Spatially heterogeneous distribution of metabolic components in neurons. (a) Glucose sensor distribution in rat
hippocampal neuron, from [171]. (b) Colocalization of presynaptic marker synaptophysin (green) and punctae of glucose
transporter GLUT3 (red), from [172]. (c) Localized mitochondrion (red) and ATP sensor (green) at presynaptic boutons in rat
hippocampal neuron, from [173]. (d) Mitochondria (blue) localized in region surrounding a node of Ranvier (membrane in
yellow), from [48]. (e) Mitochondria (blue) in a zebrafish sensory neuron (membrane in yellow), from [48].

gradients in plant cells has long been proposed to The task of mitochondrial localization poses a
rely on convective transport in a flowing cytoplasm number of challenges to the intracellular trans-
[180], while motor-driven towing of mitochondria is port machinery. It must be able to robustly con-
thought to contribute to metabolic organization in trol mitochondrial position in response to shallow
neurons [165]. gradients in long cellular projections [185]. Efficient
Neuronal cells tend to exhibit a high degree of spa- redistribution of mitochondria must be achieved in
tial and temporal heterogeneity in metabolic activ- response to growth, injury, or changing activity pat-
ity. Rapid ATP turnover is required for vesicle release terns [186–188]. In addition, because mitochondrial
in presynaptic boutons [181], with metabolic needs biogenesis and the synthesis of many mitochondrial
peaking during neuronal firing and activity [172, proteins is believed to occur largely (though not
179]. In myelinated neurons of the peripheral nervous entirely) in the soma [189], maintenance of localized
system, saltatory signal conduction relies on ion chan- mitochondrial health requires either periodic replace-
nels localized near narrow nodes of Ranvier, which ment by younger mitochondria or transient fusion
and protein exchange with a motile mitochondrial
can be separated by hundreds of micrometers. The
population [48].
energetic demands of ion pumping to restore resting
In very large cell types, active transport of small
potential are then spatially peaked in the vicinity of
nutrient molecules themselves may be of functional
these nodes [182, 183]. Neurons are known to reg-
benefit to the cell. An extreme example is the long-
ulate mitochondrial localization (figures 6(c)–(e)),
distance delivery of resources within the mycelial
concentrating them specifically in regions of high
networks of filamentous fungi, which can stretch
demand (including presynpatic boutons and areas
to many meters in extent, and whose multinucle-
near the nodes of Ranvier in electrically active neu- ated and septated structures blur the line between
rons) to enable rapid local generation of ATP [165]. cells and tissues [98]. Given the enormous size of
Such mitochondrial positioning is governed by a these syncytia in the uncontrolled environment of
number of mechanisms for halting motor-driven the forest floor, the extracellular nutrient levels can
transport in response to high calcium concentrations vary widely, necessitating long-range transport of
[183, 184] or high glucose [171]. These transport- resources through a combination of vesicle movement
regulation mechanisms enable mitochondria to accu- and flow of the cytoplasmic fluid [190, 191]. Several
mulate in regions with both high activity levels and studies have shown that the slime-mold Physarum
high fuel supply. It should be noted that mitochon- polycephalum reconfigures its own filamentous net-
dria also act as calcium buffers for the cytoplasm, work morphology to connect multiple food sources
and their controlled localization helps to regulate cal- in a manner reminiscent of man-made transporta-
cium gradients crucial to neuronal signaling as well as tion networks [192, 193], optimizing the transport of
gradients in ATP. nutrients and signaling molecules through peristaltic

10
Phys. Biol. 17 (2020) 061003 Topical Review

‘shuttle-streaming’ flows [28, 194, 195]. In large mean squared displacement (MSD) in each dimen-
algal cells, which can grow up to a millimeter sion scales linearly with time according to
in width and several centimeters long, cytoplasmic
  kB T
streaming flows are responsible for the long-distance MSD = x2 = 2Dt, D= (3)
delivery of nutrients from regions of uptake to sites μ
of active growth [29, 196]. In each of these cases, where μ is the friction coefficient of the particle, kB
with their broadly different cell types and morpholo- is Boltzmann’s constant, and T is the temperature of
gies, the necessity for nutrient dispersion over long the medium. The friction coefficient μ depends on
length-scales requires the introduction of flow-based the size and shape of the particle [198, 199], as well
active transport mechanisms that vastly outpace as the viscosity of the medium [200]. For a sphere,
diffusion. μ = 6πηa, where η is the medium viscosity and a the
radius of the particle [23, 33].
4. Physical mechanisms of transport The simple Stokes–Einstein relation
(equation (3)) rests on several major assumptions:
In order to fulfill the varied functional objectives of the particle must be embedded in a continuous,
intracellular transport, eukaryotic cells rely on trans- purely viscous, three-dimensional (3D) fluid of
port mechanisms that can be categorized into three infinite extent, with no external sources of energy.
classes: (1) diffusion-like random motion of small Below we discuss how the breakdown of each of these
particles down their concentration gradient, driven assumptions affects intracellular particle diffusion.
by broadly distributed fluctuations in the intracel-
4.1.1. Lateral diffusion on membranes
lular medium; (2) processive movements associated
with the ATP hydrolysis-driven stepping of motor Many biologically important proteins are embed-
proteins along cytoskeletal highways; and (3) advec- ded in cellular lipid membranes, including both the
tive motion arising from fluid flows in the cytoplasmic plasma membrane surrounding the cell itself and
medium. the much more extensive membranes of eukaryotic
Each of these mechanisms has its advantages organelles [202]. Lateral diffusivities of membrane
and disadvantages for different cellular tasks. For proteins have ranges of 6–10 μm2 s−1 in plasma
instance, diffusion through the cytoplasm requires no membranes [203] and 0.2–0.5 μm2 s−1 in the ER
additional energy input beyond the ongoing active membrane [204]. Confinement of a particle to a two-
processes that drive cytoplasmic fluctuations. This dimensional fluid membrane fundamentally alters its
mechanism can be very efficient at spreading small diffusivity in a manner dependent on the thickness
molecules over relatively short distances (e.g.: pro- and curvature of the membrane. A critical feature of
teins require only a few seconds to diffuse across a a purely two-dimensional fluid is that hydrodynamic
typical 20 μm animal cell). correlations do not decay but rather extend over the
Motor-driven transport, with its typical processive entire domain, leading to the famous Stokes para-
rates of ∼1 μm s−1 , requires burning ATP for every dox [205]. As a consequence, the size of the domain
step taken by a motor, but can allow much more rapid can be an important length-scale for determining
delivery of cargo over long distances. This form of the diffusivity even of very small particles far from
active transport also has the advantage of enabling the boundary. The classic Saffman–Delbrück model
the cell to control which cargo gets delivered to which [206] derives the lateral diffusivity of a particle of
cellular region through selective packaging into vesi- radius a in a thin membrane of thickness h and vis-
cles, regulation of the motor complement attached to cosity μm , embedded within a bulk fluid with lower
each organelle, and modification of the cytoskeletal viscosity μs , as
tracks.  
kB T Rcorr
Advective flow can enable faster motions still (up DSD = ln −γ , (4)
4πμm h a
to 1 mm s−1 in the shuttle flows of Physarum [197]),
driving broad populations of intracellular particles, where γ  0.6 is the Euler constant and Rcorr gives an
but with less control over the precise delivery of spe- effective length-scale limiting planar hydrodynamic
cific components. Below, we review the main physical correlations. In contrast to free diffusion in a 3D sol-
factors that underlie each of these transport mecha- vent, this expression implies that lateral diffusivity
nisms, their inherent limitations, and their coupling on a membrane is only weakly dependent on par-
and control in cellular systems. ticle size and is inversely proportional to the mem-
brane thickness. The Saffman–Delbrück model is
4.1. Diffusive transport supported by in vitro experimental measurements
The canonical diffusion of particles in a fluid arises [207], but requires significant alterations when the
from Brownian motion—spatially and temporally membrane is near a solid substrate [208], or when the
uncorrelated movements due to thermally driven protein radius is comparable to the membrane thick-
fluctuations in the medium. Diffusing particles in ness [209]. The latter case, in particular, is relevant
a viscous medium execute random walks whose in the intracellular world, where typical membrane

11
Phys. Biol. 17 (2020) 061003 Topical Review

Figure 7. Lateral diffusion of transmembrane proteins. (a) Schematic showing a membrane protein (green) diffusing in the plane
of the membrane, with relevant parameters determining its diffusivity (equation (4)) labeled. (b) A thickness mismatch between a
protein’s transmembrane region and the membrane itself results in local membrane deformation that can bias diffusion and lead
to interactions between neighboring proteins. (Image from [201]).

thicknesses (∼4 nm [210]) are comparable to protein preference can also attract proteins toward specific
dimensions. cellular regions. In particular, an energetic preference
The correlation length scale Rcorr is defined by for membrane thickness has been implicated as a pro-
an interplay of several physical effects. In the case tein sorting mechanism in the secretory pathway [216,
of a very large flat membrane domain, it is given by 221, 222], including capture at ERESs [55] and par-
the Saffman–Delbruck screening length [206]: LSD = titioning to secretion-bound lipid rafts in the Golgi
hμm /μs , beyond which planar hydrodynamics are [223, 224]. Similarly, curvature preference is believed
screened out by flows in the bulk fluid [208]. Alter- to facilitate protein sorting into membrane tubules
nately, it can be given by the overall extent of the [225], the necks of budding vesicles [226], and the
membrane domain itself (Rmem ; figure 7(a)), when curved regions of dividing bacterial cells [227].
this is smaller than the screening length [211]. The
domain size Rmem is not well-defined for many bio- 4.1.2. Medium rheology
logical systems. It may correspond to the size of mem- For particles diffusing within the bulk of the cell,
brane compartments with fixed boundaries defined a key assumption of the Stokes–Einstein relation
by interaction with cytoskeletal filaments [212, 213]. (equation (3)) is that the cytoplasmic environ-
In the specific case of particles diffusing laterally along ment behaves as a purely viscous medium. This
a tubule-shaped membrane, it can be approximated as assumption has been challenged by a variety of stud-
the radius of the tubule [214]. As a consequence, the ies that actively probe the rheological properties of
lateral diffusivity of particles is expected to decrease the cytoplasm [232, 233], or else leverage ‘passive
with decreasing tubule radius, accounting for the microrheology’—visualizing and tracking the appar-
experimentally observed slowing of diffusive spread ently passive trajectories of individual particles in live
on narrow reconstituted tubules [211]. cells [36, 39, 228, 231, 234]. These studies are sum-
Mechanical properties of the membrane can also marized in several excellent reviews on intracellular
have an important impact on the lateral diffusivity of rheology [200, 235].
embedded proteins. For example, important physical Passive particle-tracking microrheology enables
effects arise when there is a mismatch between the explicit calculation of the MSD as a function of time
preferred curvature of the embedded protein and the (figures 8(a)–(c)), for comparison with the expected
surrounding membrane curvature [215]. Alternately, diffusive behavior described by equation (3). In some
many proteins show a mismatch between the length cases, injected beads or endogenous vesicles exhibit
of the transmembrane region and the preferred thick- linear scaling of the MSD with time, as would be
ness of the membrane [216] (figure 7(b)). In both expected for a diffusing particle [26, 234, 236, 237].
cases, the mismatch engenders an elastic deformation More commonly, however, particle motion in cyto-
field in the surrounding membrane [217, 218]. When plasm is characterized as subdiffusive, with a sublin-
multiple proteins come sufficiently close together for ear scaling MSD ∼ tα , where α < 1 [39–41].
the deformation fields to overlap, they can experience Subdiffusive scaling is expected when motion
attractive or repulsive forces mediated by the mem- is driven by thermal fluctuations in a power-law
brane elasticity [219]. Such interactions have a range fluid—a material with complex rheology, whose vis-
of 1–2 nm for thickness deformations and 5–500 nm cous and elastic moduli vary as a characteristic power
for curvature deformations [218]. law of the probing frequency [235]. For example, sub-
As a result of these effects, membrane proteins diffusive motion with α ≈ 0.75 is both theoretically
diffuse across an effective potential energy landscape expected and observed for particles embedded in gels
that can guide and modulate their motion. On a of semielastic polymer filaments, such as F-actin [39,
thermally fluctuating membrane, protein curvature 238].
preference has been postulated to enhance lateral The usual physical model for passive particle
diffusion by up to a factor of two, due to the attrac- movement in a viscoelastic fluid is termed ‘fractional
tion of the protein toward transient regions of match- Brownian motion’ [37, 200, 239]. This model derives
ing curvature [220]. Both curvature and thickness from an overdamped generalized Langevin equation

12
Phys. Biol. 17 (2020) 061003 Topical Review

Figure 8. Characterizing diffusive transport by analysis of single particle trajectories. (a)–(c) MSD. (a) Expected behavior is
shown for processive motion with speed 1 μm s−1 (dotted, purple), diffusion with D = 5 μm2 s−1 (dashed, cyan), and subdiffusion
with α = 0.6 (solid, orange). Inset shows example trajectories for each type of motion. (b) MSD for peroxisomes in COS-7 cells,
with black showing linear scaling for untreated cells, red showing subdiffusive scaling for ATP-depleted cells, from [26]. Blue and
green curves are for cells treated with nocodazole (Noc) and latrunculin A (LatA) to hinder polymerizaton of microtubules and
actin filaments, respectively. (c) Subdiffusive MSD for 100 nm nanoparticles in cytoplasm of 3T3 fibroblasts, adapted from [228].
Red line shows ensemble average. (d)–(f) Rescaled VACFs. (d) Expected VACF for a diffusing particle (dashed) and particle
undergoing fractional Brownian motion with α = 0.6 (solid). Inset shows local displacements on a sample trajectory used to
calculate the VACF at Δt = δ = 4. (e) VACF for RNA–protein particles in Escherichia coli cytoplasm, from [229]. (f) VACF for
quantum dots in HeLa cells, adapted from [230]. (g)–(i) Distribution of step sizes. (g) Predictions for a diffusing-diffusivity
model, with correlation time τ , showing transition from exponential scaling at short times (blue) to Gaussian scaling a long times
(red). Solid and dashed black lines show the t  τ and t  τ limits, respectively. (h) Displacement distribution for colloidal √
beads in F-actin suspensions, over different time intervals, with inset showing universal behavior when distance is rescaled by t,
from [231]. (i) Displacement distributions for RNA–protein particles in yeast cytoplasm, using rescaled distance, from [229].

[240] featuring a power-law memory kernel K(t) distributed interaction timescales [253] can all give
which is convolved with the past time-course of par- rise to MSDs with sublinear scaling. Other metrics
ticle velocities to give the drag force: have thus been developed to quantify the behavior of
 t particles undergoing subdiffusive motion. One com-
d r(t  ) (B)
μ dt  K(t − t  ) = F (t) mon metric is the velocity autocorrelation function
0 dt (VACF), which tracks how velocities (defined by steps
  (5)
Fi(B) (t)Fj(B) (t  ) = μδij kB TK(t − t  ) over different timescales δ) are correlated across a
time-lag Δt. Namely, the velocity autocorrelation is
where F(B) is a Brownian force satisfying the fluc- given by
tuation–dissipation relation and hence exhibiting
1
the medium-dependent time correlations indicated Cvδ (Δt) = [ r(Δt + δ) − r(Δt)] · [ r(δ) − r(0)]
.
δ2
above [241–244]. When the memory kernel is (7)
replaced by a delta-function, corresponding to an Unlike classical diffusion, where velocities are fully
instantaneous relation between force and velocity as uncorrelated for all Δt > δ, fractional Brownian
in a purely viscous fluid, the model reduces to clas- motion gives rise to negative velocity correlations
sical Brownian motion. In a power-law fluid, the that are self-similar across time-scales (figure 8(d)).
memory kernel is K(t) ∼ t−α , effectively replacing the Because several microrheology studies have shown
medium viscosity η with a frequency-dependent vis- similar behavior for the velocity autocorrelation of
cosity η(ω) ∼ ωα−1 [200, 245]. Fractional Brown- intracellular particles (figures 8(e) and (f)) [36,
ian motion gives rise to a sublinear MSD of passive 229, 230, 246], the cytoplasm is often treated
particles [37]: as a power-law fluid whose viscoelastic proper-
 2 kB T sin(απ) ties lead to fractional Brownian motion of passive
x fBM = tα (6) components.
μ π(1 − α/2)(1 − α)α
This model has been used to explain the observed sub- 4.1.3. Active diffusion
diffusion of a variety of intracellular particles, includ- Brownian or fractional Brownian motion in a pas-
ing genomic loci [246, 247], mRNA molecules [248], sive medium is driven by equilibrium thermal fluctu-
and RNA–protein complexes [229, 247]. ations. Thermally generated fluctuating forces must
However, the MSD by itself cannot distin- have a specific time-dependent correlation func-
guish between several different models for subdif- tion determined by the rheology of the medium
fusive motion [249]. For example, localization error (equation (5)). The interior of a living cell, however, is
[250] in tracking particle positions, crowding [251], an environment that is manifestly outside the equilib-
confinement [252], or binding events with broadly rium regime, with fluctuations driven by a wide array

13
Phys. Biol. 17 (2020) 061003 Topical Review

of active energy-consuming processes with different stochastic particle movements with negligible proces-
underlying temporal correlations. sivity as apparently diffusive in the remainder of this
A plethora of recent experimental evidence has manuscript.
shown that even apparently diffusive particle dynam-
ics rely on active cellular processes and are not 4.1.4. Crowding and heterogeneity
driven primarily by thermal fluctuations [254]. Active An assumption of the Stokes–Einstein relationship
microrheology measurements can be used to probe for diffusing particles (equation (3)) is that the par-
the force-response dynamics of the cytoplasm by ticles are embedded in a continuous medium. The
directly controlling the forces applied to beads caught interior of a eukaryotic cell is inherently very crowded,
in optical and magnetic traps. Such measurements with proteins constituting over 20% by mass of mam-
tend to indicate that the cytoplasm responds to force malian cell cytoplasm [264]. In addition, organelle
as a largely elastic material, in direct contrast with structures ranging from vesicles to reticulated tubules
the apparently diffusive motion of passive particles and cytoskeletal networks are interspersed through-
[232, 255]. Attenuation of active cellular processes out the cell, occupying 40%–50% of cell volume
(e.g.: by ATP depletion) results in severe reduction [210]. In most models of particle motion within the
in the mobility of cytoplasmic particles [26, 232, cytoplasm, these crowding agents are averaged out
256]. Furthermore, the temperature dependence of to yield an effective viscous or viscoelastic medium.
apparent particle diffusivity inside the cell is non- However, this approximation can lead to inaccurate
linear, in contrast to expected behavior for generalized predictions for transport behavior in the cytoplasm.
diffusive motion (equation (5)). Instead, the tempera- For instance, the dependence of the diffusion coeffi-
ture dependence is Arrhenius-like, with mobility scal- cient on particle size (D ∼ R in a continuum fluid)
ing according to D ∼ exp(−Ea /kB T) as expected for is highly non-linear, with nanoscale proteins typically
reaction rates of activated processes [257]. experiencing an effective viscosity that is orders of
A number of different active processes are believed magnitude lower than that measured with micron-
to play a role in the apparent particle diffusivity sized beads or vesicles [265]. Furthermore, protein
inside the cell. Myosin motor activity been shown complexes sized in the tens of nanometers tend to
to contribute substantially to overall particle mobil- exhibit purely diffusive motion [266], rather than
ity in mammalian cytoplasm [232, 256, 258, 259]. the subdiffusive behavior observed with vesicle-sized
Inhibition of directed motor-driven transport is also probes that are an order of magnitude larger [40, 41,
known to reduce active diffusivity of apparently pas- 228]. This strong dependence of medium properties
sive organelles [26]. Recent evidence indicates the dif- on probe size is generally found in gels, where par-
fusivity of individual active enzyme molecules can ticles much smaller than the pore size move freely
be significantly enhanced in the presence of their through the gel while larger particles rely on rare
substrates, through mechanisms that are currently jump events or large-scale rearrangements to move
unclear [260–262]. between pores [267, 268]. For proteins embedded
The behavior of particles driven by active fluctu- in the plasma membrane, the actin cortex has also
ations is determined by the spatiotemporal correla- been shown to form a meshwork of obstacles that
tions of forces acting on the particles [ F (a) ] and the reduces effective diffusivity, particularly for larger
memory kernel (K) describing medium  response.  If probes [269].
the active forces have correlation F (a) (t)F (a) (t  ) ∼ In addition to individual crowders of all shapes
|t − t  |−β and the memory kernel scales as K(t − t ) ∼ and sizes, broadly-distributed spatial heterogene-
|t − t |−α , then the MSD of the particle is given by ity has been shown to play an important role in
[263] governing Brownian motion of cytoplasmic compo-
MSD ∼ t 2α−β . (8) nents. Quantification of individual step size distri-
butions for RNA–protein particles [229], colloidal
For a particle pushed by a purely processive force tracers [249, 270], and membrane-bound receptors
β = 0, while forces with delta-function correlations [271] indicates that they do not follow a Gaussian
correspond to the limit β → 1. Linear scaling of the distribution as would be expected for thermally
MSD arises for particles undergoing thermal diffu- diffusing particles in a uniform viscous or a vis-
sion in a purely viscous medium (α = β = 1). Alter- coelastic medium (figures 8(g)–(i)). Instead, the step
nately, it can also arise for particles in a purely elastic sizes have a Laplace distribution, with probability
medium (α = 0) pushed by random processive forces density P(Δx) ∼ exp[−Δx/λ(t)]. This scaling is
that themselves accumulate as a random walk over indicative of a breakdown of spatiotemporal homo-
time (β = −1). The latter model has been proposed geneity in the particle motion, which would imply by
for movement driven by an accumulation of acto- the central limit theorem that each time-step should
myosin contraction events, over timescales shorter involve the sum of many uncorrelated displacements
than the processivity time of an individual myosin and should thus follow a Gaussian distribution. Simi-
motor [232]. In the interest of brevity, regardless lar long-tailed distributions of step-sizes are observed
of the underlying physical cause, we will refer to for the dynamics of tracers in a suspension of active

14
Phys. Biol. 17 (2020) 061003 Topical Review

swimmers [272], in glassy systems [273], and in poly- Confinement within stable tubular geometries is
mer solutions (figure 8(h)) [231, 274]. found in mitochondrial [303], ER [290], and peroxi-
The origin of such distributions has been some [304, 305] networks, as well as bead-on-a-string
attributed to broadly distributed diffusivities of structures formed by nuclei in certain cell types such
individual particles caught in different regions of a as human leukocytes [306]. Transient tubules are also
heterogeneous environment [229, 275, 276], with observed during vesicle budding and organelle fis-
exponential distributions of the diffusion constant sion [307], ER-to-Golgi transport [63], and peroxi-
giving rise to the observed Laplace distribution some division [308, 309]. Tubule radii can range from
in stepping times. Indeed, diffusion coefficients ∼10 nm for dynamin-constricted regions [310] to
extracted from individual trajectories of intracellular ∼300 nm for mitochondrial network tubules [303].
particles generally exhibit very broad distributions For membrane proteins, models of diffusion on
that are not strongly peaked around a preferred value curved surfaces have shown that confinement to
[27, 247, 277]. In some cases, this observation has increasingly narrow tubules leads to slower spread-
been attributed directly to local variations in the ing over the surface even when the diffusion constant
density of obstacles formed by organelle structures and membrane surface area are kept constant [311].
such as the ER [278]. This purely geometric effect is thought to arise from
More sophisticated models of ‘diffusing- the local curvature and global topology of tubular
diffusivity’ incorporate time correlations as the membranes. Crowding of proteins on tubular mem-
particle moves through the heterogeneous environ- branes can lead to additional effects, including effec-
ment, with D(t) itself treated as a time-dependent tively anisotropic diffusion in the lateral versus cir-
random variable [276, 279, 280]. Beyond a charac- cumferential directions [312].
teristic correlation time, such a particle samples over For proteins in the lumen of a tubule, variation in
many diffusivities and its step-size distribution again tube radius can give rise to an entropic effect wherein
begins to look Gaussian (figure 8(g)), as has been locally narrower regions serve as effective diffusion
observed in some experimental measurements [231, barriers [313]. The local axial diffusivity in a tubule of
280]. heterogeneous radius R(x) is given by the Fick–Jacobs
Overall, the broad non-Gaussian distributions of equation:
step sizes over commonly measured time-scales high- D0
D(x) = , (9)
light the heterogeneity of the intracellular medium [1 + R (x)2 ]α
and the difficulty of making general conclusions based where D0 is the diffusivity in free space, α = 1/3 for
on ‘typical’ particle diffusivities. two dimensions and α = 1/2 for three dimensions
[313, 314]. An extreme case of entropic traps can
4.1.5. Confinement and geometry be seen in geometries with narrow-necked regions
In addition to macromolecular crowding, the dif- branching away from a main tubule, as in dendritic
fusion of many intracellular particles is limited by spines (figure 9(d)). These traps serve as effective
confinement in subcellular regions of complex geom- obstacles to diffusive motion along the tubule, result-
etry. Subcellular morphology is diverse, including ing in a reduced diffusivity at long times (when many
shapes resembling spheres, tubes, sheets, labyrinths, traps have been sampled) and anomalous diffusion at
beads on a string, and networks. Tubes and sheets intermediate times [243, 315].
are particularly common, and effectively confine Confinement within complex organelle geome-
diffusion to one or two dimensions, respectively. tries gives rise to a discrepancy between the actual
Here we outline the effect of these morphologies on domain explored by a particle and its apparent
the diffusive spreading of proteins confined within motion in the 3D space where the organelle is embed-
organelles. ded. For instance, particles on a curved membrane
The ER and mitochondria are ubiquitous intra- surface generally traverse longer lengths than the
cellular structures that exemplify several of these Euclidean distance between a start and end point
morphologies. Both of these organelles can form (figure 9(e)), leading to underestimation of diffusivity
extensive tubular networks (figures 9(a) and (b), left) when 3D spreading is analyzed [287]. Particles con-
[288–290] or dynamically break up into globular fined to a reticulated network of tubules are restricted
structures [291, 292] (restricted to certain stress or to move along one dimension between each neigh-
perturbative conditions in the case of the ER [293, boring node. This effect decreases long-range diffu-
294]). The ER also forms stacks of flat membranous sivity by a factor of 2 or 3 for a fully connected regular
sheets in the perinuclear region (figure 9(a), bottom planar or 3D lattice, respectively (figure 9(f)). Com-
right). The morphology of these organelles is altered parison of fluorescence recovery after photobleaching
in different cell types [54, 295], growth conditions experiments with simulations on extracted ER struc-
[296, 297], cell cycle stages [298, 299], or states of tures suggests that diffusive recovery times in the ER
stress [300, 301]. The complex geometries of both ER are 1.8–4.2× longer than would be expected from
and mitochondria are believed to be linked to their local diffusivity measurements [316]. Furthermore,
functional roles in the cell [290, 302]. the convoluted geometry of this organelle seems to

15
Phys. Biol. 17 (2020) 061003 Topical Review

Figure 9. Diffusive transport is modulated by confinement in intracellular structures of complex morphology. (a) Structure of
the mammalian ER. Left: peripheral ER network in COS7 cell, from [281]. Right, top: cross-section of individual ER tubules,
from [282]. Right, bottom: 3D reconstruction of helicoidal ramps connecting ER sheets in mouse salivary gland cell, from [283].
(b) Structure of mitochondrial network (left) in pancreatic β-cells, and the inner membrane cristae that form occlusions within a
mitochondrion (right), from [284]. (c) Schematic of murine rod photoreceptor cell (left) and 3D reconstruction of membranous
discs in the rod cell outer segment, adapted from [285]. (d) Dendritic spines that serve as diffusive traps in mouse pyramidal
neurons, from [286]. (e) Confinement to curved surfaces results in reduced apparent diffusivity when measured with 2D or 3D
Euclidean distance metrics (from [287]). Employing a geodesic distance over the surface corrects this effect. (f) Confinement
within planar networks. Blue curve shows MSD of simulated particles on a fully connected (complete) honeycomb network, with
effective diffusivity reduced by a factor of 2 compared to a free particle (black solid line). Red curve is for simulated particles on a
honeycomb network with 29% of edges removed while maintaining a single connected component (decimated network). Both
networks are confined in a circle of radius 20 μm, and particle diffusivity is set to D = 5 μm2 s−1 , to give relevant units for
proteins diffusing in an animal cell.

have a greater effect on membrane than on luminal example of sheet geometry in the form of stacks of
proteins. Simulations on realistic tubular ER geome- flat parallel cisternae bounded by membrane sheets
tries indicate that membrane proteins explore ER [290]. The cisternae are interconnected with ramp-
regions up to 4× slower than luminal ones, even when like spiral dislocations reminiscent of a parking garage
the diffusion coefficient is identical for both [317]. [283] (figure 9(a), bottom right). The morphology
The effect of complex confining geometry or of this structure is thought to directly modulate
occluding barriers on long-range particle diffusion diffusion, with the ramps substantially enhancing dif-
has been extensively explored in the context of trans- fusive transport between stacked cisternae when com-
port through porous media and over spatial networks pared to individual holes in the membrane [323]. In
[318]. These effects are often described via an emer- particular, the unique connection geometry allows
gent quantity called ‘tortuosity’—which is conceptu- diffusing particles to transition between sheets by spi-
ally defined as the ratio between the typical length raling around dislocations rather than searching for
traversed by a diffusing particle and the Euclidean dis- small holes serving as localized connections between
tance between its start and end points (figure 9(e)) flat sheets. Even in the presence of these spiral struc-
[319]. A common simplifying model for environ- tures, the limited connectivity of stacked ER sheets
ments with high tortuosity is that of percolation on results in an effective perpendicular diffusion that is
a lattice [320]. The medium is represented as a lat-
roughly 10-fold slower than local diffusivity [323].
tice network with randomly removed edges. Such sys-
An example of labyrinthine structures that limit
tems exhibit a phase transition when the fraction of
connectivity within organelle compartments are the
remaining edges reaches a critical value pc , below
mitochondrial cristae—convoluted folds of inner
which the network becomes disconnected and parti-
mitochondrial membrane that occlude much of the
cles can no longer penetrate throughout the domain.
mitochondrial matrix space (figure 9(b), right) [324].
Percolation systems exhibit a number of universal
There are about 6–8 cristae per μm of mitochon-
scaling behaviors, including a slow-down in effective
drial length, each of which serves as an impene-
diffusivity according to
trable barrier to the diffusion of molecular species
D ∼ (p − pc )μ (10) [325]. Early studies indicated that these protrusions
as the fraction of remaining edges approaches the crit- must stretch across nearly the entire mitochondrial
ical value [321]. The scaling exponent is μ = 1.3 for cross-section in order to have a substantial impact
two-dimensional and μ = 2.0 for 3D lattices [322]. on diffusivity [326]. Simulations of diffusive spread-
Within cellular organelles, the connectivity of ing in the presence of multiple such overlapping
the space available for diffusion is determined by barriers show that cristae are expected to slow long-
the overall organelle geometry, as well as the pres- range axial diffusion of matrix proteins by a factor of
ence of intra-organelle substructures that serve as 5–6 [325]. Effective axial diffusion of proteins embed-
obstacles for mobility. The perinuclear ER presents an ded in the inner mitochondrial membrane may also

16
Phys. Biol. 17 (2020) 061003 Topical Review

Figure 10. Diffusive target search processes. (a) Illustration of non-compact (top) versus compact (bottom) search processes.
Left: effect of medium dimensionality, with simulated diffusion trajectory in 3D (top) and on a network of 1D edges (bottom);
color progression represents time. Right: transition from non-compact to compact search with increased compartmentalization
of the domain (figure from [330]). (b) Dependence of search rate (inverse of mean first-passage time) on density of targets in a
network structure extracted from fluorescent images of peripheral ER in COS7 cell, showing transition from an effectively 2D to
effectively 1D search process. Yellow region indicates physiologically relevant concentrations. (c) Sample geometries for
intracellular narrow escape processes: (i) transition from ER sheets to adjacent tubules (figure from [331]), (ii) protein
accumulation at ERESs, (iii) transport of cytoplasmic transcription factors into nucleus and mRNA out of nucleus through
nuclear pores [332], (iv) diffusion of ions and proteins into narrow-necked dendritic spines (figure from [286]).

be slowed by up to an order of magnitude by the their target search times strongly depend on their
convoluted morphology of these structures [327]. starting position [337]. By contrast, a non-compact
In the outer segment of mammalian photorecep- search sparsely samples subregions of the domain, will
tor cells, flat lammellar disc membranes form simi- generally reach the domain boundary before finding
lar occlusions, leading to a high tortuosity for axial the target, and has search times largely independent
transport (figure 9(c)) [328, 329]. Axial diffusivity of starting position.
in this compartment has been measured as roughly For random walks on self-similar (i.e. fractal)
50-fold slower than the nearby inner segment com- geometries, the behavior of the search process is deter-
partment, with a factor of 20–40× accounted for by mined by two key dimensions. The dimensionality
the increased tortuosity due to membrane occlusions of the random walk itself (dw ) can be defined by the
[329]. scaling of MSD with respect to time (in the absence
of confinement): MSD ∼ t 2/dw [318]. Equivalently,
4.1.6. Diffusive target search
dw describes the scaling between the time to exit a
In the preceding discussion we addressed the impact sphere and the sphere size R: texit ∼ Rdw [338]. The
of various physical factors on diffusive particle fractal dimension (df ) describes the dimensionality
motion. Here, we consider the interplay of diffu- of the medium within which the walk is embed-
sion and morphology in limiting the kinetics of ded, relating the number of sites (N) with the spa-
intracellular encounters and reactions. For freely tial extent of a region (R) according to N ∼ Rdf
diffusing particles in a 3D continuum, maximal [318]. Compact search corresponds to the regime
reaction rates are proportional to particle concentra- where dw > df , such as canonical diffusion (dw = 2)
tions. The steady-state current of particles to a per- on a one-dimensional line (df = 1). The opposite
fectly absorbing spherical target of radius a is given regime (dw < df ) is termed non-compact search and
by includes canonical diffusion in three dimensions
J = 4πDac0 , (11) (df = 3).
where D is the particle diffusivity, c0 the bulk parti- The mean time T
for a randomly moving parti-
cle concentration [333], and J represents the rate of cle to find a target site in the fractal medium is then
particles arriving at the target. However, this linear given by the following scaling laws with respect to the
relationship between concentration and reaction rate domain volume N and initial distance from the target
does not necessarily hold when particles are confined r [338]:

to complex geometries or embedded in domains of ⎪ d −d
⎪N(A − Br w f ), for dw < df (non-compact)

reduced dimensionality [318, 334].
T
∼ N(A + B ln r), for dw = df .
Target search processes involving randomly mov- ⎪


ing particles fall into two broad categories: compact N(A + Brdw −df ), for dw > df (compact)
and non-compact [318, 335, 336] (see figure 10(a)). (12)
In a compact search process, a particle will cover most In the case of non-compact search, the dependence
of the sites within each subregion it visits. Such par- on starting position disappears for sufficiently large
ticles generally find the target after comprehensively r, and the search time is simply proportional to the
exploring a finite subsection of their domain, and system volume, as expected for classical 3D kinet-

17
Phys. Biol. 17 (2020) 061003 Topical Review

ics (equation (11)). The distribution of search times nuclear pores [343] (see figure 10(c)). It also includes
in this case exhibits an exponential drop-off with a reactions with a fixed target on the membrane of an
single characteristic time-scale corresponding to the organelle within which the searcher is confined. The
average search time [335]. By contrast, the compact mean first-passage time for a diffuser to reach a nar-
case results in ‘geometry-controlled’ kinetics, with a row target whose area covers a small fraction ( ) of the
search time that depends strongly on starting posi- boundary can be approximated as:
tion, even for initially distant particles. In this situ-

ation, the distribution of search times exhibits decay R2 1
over a range of different time-scales whose breadth τMFP  ln + O(1) , (2D)
D
depends on the dimensions df and dw [335]. The 
V 1
mean search time, averaged over all starting positions, τMFP  1+ ln + O( ) , (3D)
scales as T
∼ N dw /df , indicating that the slowing of 4 D π
(14)
kinetics with increased volume is super-linear [335].
for a circular or a spherical domain, respectively
It should be noted, however, that the broadly dis-
[344–346]. The case of a particle trapped in a short
tributed reaction times in compact systems are not
cylinder lies intermediate between the two regimes,
well-described by this single mean first-passage time
transitioning from two- to three-dimensional as the
[339]. For particles undergoing unhindered canon-
height of the cylinder increases [347]. This geometry
ical diffusion (dw = 2), the overall reaction rate for
can be particularly relevant for target search by parti-
a particle to find any stationary target (defined by
cles trapped between flat sheets, as in the ER cisternae
k := 1/ T
) is expected to scale as follows depending
or lamellar discs of photoreceptor cells.
on the dimensionality of the confining domain [337,
A common model for diffusion in the presence
340]:
⎧ of obstacles or in reticulated or porous structures
⎪ 2
⎪c , (1D) is to treat the process as a series of hops between

compartments that are themselves rapidly equili-
k ∼ c log c, (2D) , (13)

⎪ brated [330, 348]. Such geometries can result in a sub-

c, (3D) stantial reduction in long-range diffusivity without a
concomitant decrease in the reaction rate [330]. Inter-
where c is the target concentration (or the inverse of estingly, the connectivity of compartments can be
the volume per target). tuned in such a way that diffusive particles propagate
The impact of confinement geometry on tar- in a wave-like manner, with transient concentration
get search is particularly relevant for molecules peaks appearing in different containers [348].
that must find sparsely scattered binding partners The nature of a target-search process in compart-
within an organelle. This includes, for example, ment networks is determined by the dimensionless
newly-translated secretory proteins searching for an parameter x = DL/D0 a, where D0 is the diffusivity
exit site within the ER network [341], or mito- within a compartment, D the long-range effective dif-
chondrial matrix proteins searching for nucleoids fusivity, L the compartment size, and a the particle
[342]. While realistic cellular structures are not true reaction radius. The reaction rate exhibits one of two
fractals, similar considerations of compact versus possible behaviors [330]:
non-compact search processes can be applied to
understand the effect of organelle morphology on
k = 4π(1 − Pr )DL, x  1
kinetics. For example, calculation of diffusive first-  
passage times to find one of many point-like targets a aD (15)
k = 4πD0 a 1 − + , x  1,
on planar ER networks extracted from mammalian L LD0
cell images indicate that the search domain transi-
tions from effectively 2D to effective 1D with increas- where Pr is the probability of returning to an
ing concentrations of the target sites (figure 10(b)). already-sampled compartment. When x < 1, the pro-
Due to the compact nature of this process, the rate cess is compact and each compartment is fully
at which proteins find punctate exit sites in the explored as the particle moves through the medium
ER is expected to scale super-linearly with exit site (figure 10(a)). By contrast, for x > 1, the search pro-
density. cess is sparse and the particle typically encounters
One important class of target-search processes, the target only after multiple visits to the compart-
known as ‘narrow escape’ problems, consists of par- ment containing the target. In this regime, when the
ticles that must find their way to a very small region target size is much smaller than the compartment,
on the boundary of their confining domain. This class the long-range diffusivity may be greatly reduced
of problems encompasses molecules that need to exit (D  D0 ) without significantly changing kinetic
specific cellular regions, such as ER proteins moving rates. For enzyme diffusion in the cytoplasm, esti-
from cisternae to peripheral tubules [290] or reach- mated pore sizes are roughly 10 times bigger than the
ing an exit site for export [55], signaling factors leav- protein size [264, 349], implying that the sparse search
ing dendritic spines [286], or mRNA encountering regime is relevant for cytoplasmic kinetics.

18
Phys. Biol. 17 (2020) 061003 Topical Review

The effect of macromolecular crowding on reac- sis as an energy source to walk in a directed manner
tion rates can be approximated in an analogous along cytoskeletal highways [354]. A variety of cargos
manner by treating reactants as moving between including vesicles [355–357], mitochondria [165],
crowder-free cavities [350]. It should be noted that ribonucleoprotein particles [358], protein complexes
non-specific binding to reactants, and finite local [359], and ER tubules [160], among others, navigate
reaction rate upon encounter can further slow the the cytoplasm using motor-based transport.
overall reactive flux in the presence of crowding A key advantage of this transport mechanism is
[350]. Once interacting molecules are coincident its ability to move cargo processively over very long
in space, they must also find the correct relative length scales (up to a meter in neuronal axons). The
rotational orientation for binding or activity [351]. relative efficiency of motor-driven versus diffusive
Molecules coming together will typically experience transport over a given length scale can be quantified
many ‘microcollisions’, allowing time for reorien- by the dimensionless Péclet number (equation (2)).
tation through random chance or intermolecular Typical velocities for motor-driven cargos in ani-
interactions that favor alignment [351]. Effective con- mal cells fall in the range of 0.3–2 μm s−1 [2, 48, 167,
finement from crowding cavities provides further 355], with individual vesicle velocities reported up to
opportunity for sites to align and a reaction or bind- 10 μm s−1 [355, 357]. Speeds of motor-driven cargo
ing event to occur. tend to be independent of particle size [34], allowing
For particles diffusing on a network, the connec- this transport mechanism to vastly outpace diffusion
tivity of compartments (or nodes) plays an important for long lengths and large cargos. For RNA–protein
role in regulating target search times, as well as large- complexes and vesicular organelles, with typical cyto-
scale diffusivity [352]. For reticulated structures sim- plasmic diffusivities of D ≈ 0.01–0.1 μm2 s−1 [26,
ilar to those of the peripheral ER or mitochondrial 27, 34, 360], motor-driven motion tends to dom-
networks, target search times were recently shown inate (i.e.: Pe > 1) on length scales above a few
to be determined largely by the total network edge microns.
length and the loop (or cyclomatic) number [281, An additional advantage to motor-driven motion
289]. Loop number is a global metric of connectivity, is the ability to regulate and control transport behav-
defined by Γ = Ne − Nn + 1 where Ne is the num- ior. The mechanochemical properties of individual
ber of edges and Nn is the number of nodes. The motors can be tuned to optimize their speed or pro-
parameter corresponds to the number of independent cessivity under varying loads [361, 362]. Selective
cycles in the network structure [353]. Increasing loop recruitment of different motor proteins and biochem-
number decreases search times, while increasing edge ical modification of key molecular components in the
length increases them, with a scaling relationship that transport machinery can also tune cargo distribution
can be derived from the slowed diffusivity on a per- and dynamics [20, 363]. Furthermore, the cellular-
colation lattice (equation (10)) [281]. A recent study scale organization of cytoskeletal transport highways
on yeast mitochondrial networks demonstrated that enables sorting of cargo to different destinations in
network connectivity can be altered by mutations in the cell [364]. The plethora of molecular components
specific proteins responsible for mitochondrial fusion involved in motor-driven transport thus allows for a
and fission [289]. Simulations of diffusive search over broad variety of control mechanisms to regulate cargo
these network structures indicate that the reduced delivery.
connectivity in mutant networks is expected to slow 4.2.1. Components of the motor transport
encounter times by almost two-fold for particles at machinery
low concentrations [289]. The basic components of motor-driven transport
Diffusive transport inside cells is modulated by include the cargo itself, the motor proteins, a vari-
the mechanics of intracellular media, by active non- ety of adaptor proteins and linkers that attach motors
thermal fluctuations, and by the presence of obstacles to the cargo, and the cytoskeletal filaments that serve
and complex subcellular geometries. These physical as a substrate for walking motors (figure 11). Both
factors control both the overall dispersion and the actin filaments and microtubules can serve as high-
rates of encounter between particles. Diffusive trans- ways for motor-driven transport. Both are polarized,
port thus provides a physical link between the mor- with distinct ‘+’ and ‘−’ ends, governing the direc-
phology and dynamics of cellular structures and the tion of motor movement. In plant cells, the motion
kinetics of biomolecular reactions that underlie cell of a variety of myosin motors along polarized actin
function. filaments is responsible for long-range cargo delivery,
4.2. Motor-driven transport as well as the establishment of persistent cytoplasmic
For transport tasks where diffusive motion is too flows [365]. In animal cells, the myosin-V motor has
poorly controlled or too slow, eukaryotic cells have been shown to contribute to local organelle position-
evolved an extensive system of motor-driven trans- ing in actin-dense cortical regions [366–368]. How-
port. This system relies on the attachment of cellular ever, long-distance transport in animal cells primarily
cargo to motor proteins, which employ ATP hydroly- occurs along microtubule highways.

19
Phys. Biol. 17 (2020) 061003 Topical Review

of individual molecular motors has been extensively


explored at the single-molecule level in vitro [379,
380]. In a living cell, many motors can attach to each
cargo, and their cooperative behavior determines the
speed, processivity, and direction of cargo motion
[19, 381–383].
Specialized adaptor proteins control the comple-
ment of motor molecules recruited to a particular
cargo [20, 375]. These adaptors make it possible for
a wide range of cargos to be transported by a limited
variety of motor proteins, as well as controlling the
direction and processivity of motion [376, 384–386].
In general, adaptor proteins are bound directly by
receptors on the cargo surface, by both kinesin and
dynein motor complexes, and by a variety of signal-
ing proteins that serve to activate or repress transport
Figure 11. Schematic of components involved in [20].
motor-driven transport on microtubules. Kinesin and As an alternative to the direct recruitment of
dynein motors attach the cargo (in green) to the
microtubule (blue and white) via motor adaptors (dark motors via an adaptor protein, some cargos have
blue). A hitchhiking cargo (pink) can attach to a been found to engage transiently with other motile
motor-driven carrier via a linker protein (red). Microtubule organelles, moving by a non-canonical form of
associated proteins (MAPs) and post-translational
modifications to the microtubule help regulate motor-driven transport termed ‘hitchhiking’ [9, 26,
motor-driven transport. Figure is not to scale [332]. 167, 387–392]. In place of an adaptor protein, a linker
protein attaches the hitchhiking cargo to a carrier
organelle, which connects through an adaptor pro-
Microtubules (MTs) form long hollow tubes, con- tein to the motor. Specific linker proteins have been
sisting of 13 parallel protofilaments, with motor pro- identified for several hitchhiking cargos [9, 168, 392],
teins attaching to the outside of the tube. Interestingly, and the density, length, and stiffness of these linker
diffusive transport in the hollow interior of a micro- proteins can serve to modulate the efficiency of the
tubule has also been shown to play an important hitchhiking interaction [393]. Both linker proteins
role in the spread of several microtubule-modifying and adaptors share the common feature of enabling
proteins [369, 370]. Microtubules are quite stiff, with specific control of transport for a particular cargo,
effective in vivo persistence lengths on the order of without affecting the movement of other cellular
30 μm, enabling them to fluctuate around relatively components.
straight configurations on typical cellular scales [371].
In many animal cell types, they are organized with 4.2.2. Direction and processivity along a
their minus ends anchored near the nucleus and microtubule
their plus ends extending toward the cell periphery. The direction and processivity of transport along
Microtubules are highly dynamic, undergoing cycles a single microtubule can vary widely for different
of growth and depolymerization that allow for rapid cellular systems. Some cargos, such as post-Golgi
remodeling of the transport highway network [372], synaptic precursor vesicles in proximal regions of
as well as bending and sliding events that contribute neuronal axons, move primarily in the anterograde
to cargo motion [373]. direction toward the cell periphery [397–399]. Oth-
Two families of motor proteins execute trans- ers, such as endocytic vesicles carrying growth factor
port along microtubules. The kinesin superfamily signals [400] and neuronal autophagosomes [166],
[374] is generally responsible for anterograde trans- move primarily in the retrograde direction, toward
port: movement toward microtubule plus ends, which the cell nucleus. Many cargos are bidirectional,
often corresponds to the direction away from the exhibiting both types of motion with varying run-
nucleus. Dynein motors drive retrograde motion lengths prior to switching directions [19, 165, 401,
toward microtubule minus ends [375, 376]. Both 402]. At one extreme of highly processive bidirec-
types of motors form protein complexes with two tional motion lie mitochondria in neuronal axons,
ATP-burning motor domains that bind to the micro- which move for many tens of microns in either
tubule, linked to a long tail that attaches to cargo, anterograde or retrograde directions, undergoing
often via an adaptor complex [18]. The motors pauses of varying duration, but rarely reversing their
walk in a hand-over-hand fashion, with some (e.g.: direction after pausing (figure 12(a)) [394]. By way of
kinesin-1) following individual protofilaments while contrast, lipid droplets in Drosophila embryos [403],
others (including kinesin-2 and dynein) undergo lysosomes in neurons (figure 12(b)) [382], as well
frequent side-stepping to neighboring protofila- as endosomes and hitchhiking peroxisomes in fun-
ments [377, 378]. The mechanochemical behavior gal hyphae (figures 12(c) and (d)) [26, 392] all switch

20
Phys. Biol. 17 (2020) 061003 Topical Review

directions frequently, with typical run-lengths of alternate mechanism relies on inactive motors enter-
about 0.3–10 μm. Cytosolic proteins engaged in slow ing a weakly-bound diffusive state wherein they func-
axonal transport have been observed to exhibit even tion as tethers that prevent cargo dissociation from
shorter processive runs of about 0.1 μm, thought to the microtubule and hence increase processive run-
arise from transient interactions with passing cargos lengths driven by the dominant active motor [19].
[106]. Such an effect may account for the increased proces-
The direction and run-length for a motor-driven sivity of kinesin-carried cargos along microtubules in
cargo moving along a single microtubule is thought the presence of myosin-V motors, and the reciprocal
to be determined by the complement of associated increase in myosin-V processivity on actin filaments
motors, as well as regulatory modifications to motors, in the presence of kinesin [418, 419].
adaptor proteins, and the microtubules themselves Cooperation between multiple motors pulling in
(figure 13). Cargos that exhibit bidirectional motion the same direction has also been proposed to enhance
are generally attached to both kinesin and dynein the speed and processivity of transport. In vitro
motors simultaneously [382, 404–406]. Even axonal measurements on reconstituted systems show that
mitochondria and autophagosomes, with their very the presence of multiple kinesin motors allows for
long processive run-lengths, have been shown to carry longer run lengths and larger stall forces [420, 421],
both kinesin and dynein motors regardless of whether with similar cooperative effects observed for multi-
they are stationary or moving in the anterograde or ple dyneins [422]. Furthermore, coupling of many
retrograde direction [166, 407, 408]. The question kinesins bound to a fluid lipid membrane has been
of how multiple motors coordinate to determine the shown to increase cargo transport velocity without
direction, speed, and processivity of cargo has been altering the behavior of individual motors [423]. The-
the topic of much theoretical and experimental work oretical studies of load-sharing between motors help
over the past two decades. clarify the importance of key mechanical parameters
in determining motor cooperativity, as well as high-
The classic model for opposing motor interac-
lighting the limitations of purely mechanical mod-
tions is a ‘tug-of-war’ between multiple motors that
els and the need to incorporate biochemical coupling
come on and off the microtubule stochastically and
effects [383, 424–426].
pull in their characteristic direction when engaged
In addition to interactions between the comple-
(figure 13(a)), with the overall direction of move-
ment of motors attached to a cargo, processive motion
ment dictated by the net generated force [381, 409].
along a microtubule can also be regulated by exter-
When coupled with experimental measurements of
nal signals targeting motors and adaptor proteins
the number of motors on a cargo and the force-
[427] (figure 13(b)). These signals often take the
response parameters of individual motors, the tug-of-
form of a biochemical modification through a sig-
war model can quantitatively recapitulate aspects of in
naling pathway that responds to the local intracellu-
vivo bidirectional motion for vesicles in mammalian
lar environment or the state of the cargo itself. For
neurons [382] and endosomes in Dictyostelium slime
example, calcium ion binding to the mitochondrial
molds [410], as well as multi-motor assemblies in adaptor complex consisting of Miro and Milton pro-
vitro [411]. teins results in transient halting by dissociation of
However, this simple model fails to account for kinesin motors from the microtubule [165, 407]. Sim-
a number of puzzling observations indicating coop- ilarly, a byproduct of glucose metabolism serves as a
erative rather than competitive behavior between substrate for modifying the Milton adaptor protein,
kinesin and dynein motors on the same cargo [19, inhibiting mitochondrial motility [171]. By coupling
383]. Qualitatively, the presence of both kinesin and transport behavior to the local biochemical environ-
dynein motors has been found to be necessary to ment, these pathways can result in targeted localiza-
activate motion in both anterograde and retrograde tion of mitochondria to regions with high metabolic
directions [412–414], raising the so-called ‘paradox demand [48, 184, 428] or high glucose supply [171,
of co-dependence’ [19]. Quantitatively, a thorough 185] within extended neuronal projections.
parameter scan for the tug-of-war model has shown A permanent cessation of mitochondrial trans-
that no variant of the model can simultaneously port can also be triggered through the PINK1/Parkin
reproduce the in vivo distribution of run-lengths pathway, which is activated when the mitochondrial
and pausing behavior of bidirectionally motile lipid membrane potential (a marker for mitochondrial
droplets [415, 416]. health) drops too low and results in the degradation
A number of mechanisms for positive cooperativ- of the Miro adaptor protein [165]. A cell can thus pre-
ity between opposing motors have been proposed as cisely control the positioning of its mitochondria in
an alternative to the antagonistic tug-of-war model response to local cytoplasmic conditions and mito-
[19]. One possibility is the existence of direct bio- chondrial health. Another example of organelle state
chemical and mechanical interactions wherein one modulating transport behavior can be seen in neu-
motor type serves to activate the other or to push ronal autophagosomes, whose biochemical matura-
it out of an auto-inhibited state [406, 414, 417]. An tion is coupled to their transition from bidirectional

21
Phys. Biol. 17 (2020) 061003 Topical Review

Figure 12. Bidirectional processive motion of cargo with varying run-lengths. (a) Long-range transport of mitochondria in the
anterograde (top) and retrograde (bottom) direction (adapted from [394]). (b) Kymograph showing directional reversal of
LysoTracker-labeled vesicles in primary neurons (image from [382]). (c) Kymograph showing bidirectional motility of early
endosomes (green) in Ustilago maydis fungal hypha (image from [395]). (d) Kymograph of bidirectional processive transport for
a peroxisome in Aspergillus nidulans fungal hypha (from [396]).

Figure 13. Postulated mechanisms for pauses and reversals in microtubule-based transport. (a) ‘Tug-of-war’ between opposing
motors. Pauses are resolved when force by engaged motors in one direction dominates. (b) Biochemical regulation of motors and
adaptors. External signals can trigger pausing by dissociation of motors or cargo, or by tethering to the cytoskeleton. (c)
Roadblocks in the form of microtubule-associated proteins and post-translational tubulin modifications can result in motor
dissociation. (d) Cytoplasmic obstacles and intersections lead to pausing or directional changes. (e) Traffic jams of free motors
reduce speed and processivity.

motion near sites of synthesis at distal axonal tips Even in the case where a cargo follows a single
to robust retrograde motility toward the cell body microtubule or polarized bundle, its direction and
[166]. run-length are thus a complicated function of the
Microtubules themselves can serve as a substrate complement of attached motors, the decoration of the
for post-translational modifications and other signals microtubule track, and the spatial profile of signaling
molecules that inhibit transport.
that regulate transport processivity [427, 429, 430]
(figure 13(c)). MAPs bind to the external surface of
microtubules and differentially regulate motor pro- 4.2.3. Obstacles and traffic jams
tein behavior. For example, tau proteins tend to cause The processive motion of a motor-driven cargo
kinesin detachment at low concentrations with lit- along a microtubule is inherently limited by the
tle effect on dynein [431]. Gradients of tau proteins crowded environment within a living cell. Crowding
by filamentous macromolecules gives rise to a vis-
(which have been observed in neuronal axons [432]),
coelastic rheology of the cytoplasm (section 4.1.2)
can thus be used to tune the anterograde or retrograde
which results in size-dependent and time-dependent
bias, as well as processivity, of cargo transport [421,
drag forces experienced by the moving cargo. As a
433–435]. Other MAPs differentiate the microtubule- consequence, in vivo movements of cargo tend to
binding affinity of separate types of kinesin motors, be ‘bursty’, with speed fluctuations consistent with
allowing kinesin-3-bearing cargos to be sorted into a slow build-up and rapid release of mechanical
dendritic projections while those carrying kinesin-1 stresses [437, 438]. Models of motor-driven motion
are relegated to the axons [436]. which incorporate complex fluid rheology predict the

22
Phys. Biol. 17 (2020) 061003 Topical Review

emergence of an anomalous transport regime with value [453]. Because traffic jams depend both on total
superdiffusive yet sub-ballistic scaling of the mean- motor density and accumulation at microtubule ends,
squared displacement (MSD ∼ tα with 1 < α < 2) the moderate processivity and high end detachment
[439, 440]. In reconstituted in vitro systems with a rates of kinesin have been hypothesized to be advan-
viscoelastic medium, increased densities of filamen- tageous for overall cellular transport [454]. Inter-
tous crowders have been shown to drastically reduce estingly, for cargo that can bind motors reversibly,
the transport velocity of cargos carried by teams of increased free motor density can actually give rise
kinesin motors [441]. to longer run-lengths [455], possibly due to the
In addition to altering the rheology of the cyto- cargo’s ability to associate with more motors to bypass
plasmic medium, crowded conditions within the cell localized traffic jams or effectively surf along densely
imply the ubiquitous presence of obstacles, both packed neighboring motors [442].
directly bound to the microtubule track and in The motion of a motor-driven cargo along a sin-
the cytoplasm at large [442]. Individual molecu- gle microtubule is determined by a complex inter-
lar motors vary in their ability to bypass MAPs action between the complement and regulation of
that serve as roadblocks along the transport high- motors attached to the cargo, the distribution of road-
way (figure 13(c)). Single kinesin-1 motors generally blocks and traffic jams along the microtubule, and the
dissociate when encountering a road-block, though presence of cytoplasmic obstacles encountered by the
teams of such motors can effectively bypass the obsta- cargo. We next proceed to consider how cargo dis-
cle [377]. Individual dynein motors, on the other tribution on a cellular scale is governed by a combi-
hand, are much more capable of side-stepping to nation of limited-processivity runs interspersed with
neighboring protofilaments, allowing them to suc- passive periods.
cessfully bypass microtubule-bound obstacles [377,
443]. The increased ability to maneuver around obsta-
cles afforded by the presence of different motor types 4.2.4. Run and pause: intermittent transport
has been proposed as a key evolutionary advantage to Cellular cargos engaged in long-range transport often
bidirectional motion [444]. undergo periods of processive runs interspersed with
When encountering large obstacles, such as other pauses of varying duration [19, 26, 382, 393–395].
vesicles attached to the same track or intersecting The pauses can be very long, as is the case for axonal
microtubules (figure 13(d)), 3D motion of the cargo mitochondria that have been observed to switch from
around its track is required for maneuvering around a motile to a long-lived stationary state [165]. They
the obstacle [445]. In vivo tracking of anisotropic par- can also be transient, associated with maneuvering
ticles indicates that 3D rotation of the cargo occurs around an obstacle [2, 446], tug-of-war between
during long pauses that result in directional rever- opposing molecular motors [382, 456], or dissocia-
sals on the same microtubule or a nearby parallel tion from the microtubule or hitchhiking carrier [26,
track [446]. These pauses were postulated to arise 457]. During such pauses the cargo can remain sta-
from obstacle encounters, with release and engage- tionary, tethered to the microtubule itself or to nearby
ment of alternate motors allowing the cargo to bypass filaments of the actin cytoskeleton [165, 458, 459].
the obstacle. When encountering a microtubule inter- Alternatively, the cargo can be free to diffuse within
section, cargo can also switch to the intersecting the cytoplasm until the next run of processive motion
microtubule, reverse, or pass by it, in a manner [26, 457].
dependent on the geometry of the intersection [447] A simple mathematical model for transport con-
and the complement of attached motors [356, 448]. sisting of interspersed periods of diffusive and proces-
The extent to which bypassing of an intersection in sive motion is the one-dimensional ‘halting creeper’
vivo involves side-stepping of individual motors ver- (figure 14, inset) [25]. This model comprises one-
sus switching or tug-of-war behavior between multi- dimensional particle motion, switching at fixed rate
ple motors remains largely unknown [445]. kstop from processive motion with velocity ±v to
An additional source of transport obstacles comes pauses with diffusivity D and vice versa with rate
from traffic jams formed by individual molecular kstart . Such a particle has a run length  = vkstop
motors bound to and moving along microtubules and is processive a fraction f = kstart /(kstart + kstop )
(figure 13(e)). These traffic jams can be described of its time. The 1D model is particularly relevant
by the classic physical model of a ‘totally asymmet- for particles within cellular regions that form highly
ric simple exclusion process’ [449], which consists of extended tubules, such as fungal hyphae and neuronal
non-intersecting particles moving along a line and axons.
predicts the onset of jamming as a phase transition The transport range (length of domain explored)
[450–452]. Such models are quantitatively consis- for a halting creeper particle transitions from a
tent with in vitro observations of the steep drop in diffusion-dominated regime at short times, to a ballis-
both velocity and run-length when the density of tic intermediate motion above a characteristic length
kinesin-1 motors on a microtubule reaches a critical scale x∗ which can be estimated by setting f Pe(x∗ ) > 1

23
Phys. Biol. 17 (2020) 061003 Topical Review

Figure 14. Dispersion of particles via multimodal transport. Inset: schematic of the halting creeper model for a particle switching
between ballistic and diffusive motion. Plot shows range explored by a halting creeper versus time. Two transitions in behavior are
evident: at t∗ , x∗ processive motion begins to dominate; at t∗∗ , x∗∗ a sufficient number of reversals have occurred that particle
motion begins to look effectively diffusive. Parameters used apply to peroxisome transport in fungal hyphae
(D = 0.015 μm2 s−1 , v = 2 μm s−1 ,  = 6 μm, f = 0.05). Adapted from [25], copyrighted by the American Physical Society.

[25]. At much longer length and time scales, when the that both diffusion and active transport contribute
particle has had the opportunity to sample repeatedly substantially to target search processes [25].
between the different modes, it again exhibits effec- While the velocity v of processive motion is fairly
tively diffusive transport (figure 14). Similar transi- constant (of order 1 μm s−1 ), cells can regulate both
tions, albeit on different time-scales, are also observed the typical run length  and the pause time tpause =
for the MSD of a particle engaged in multi-modal 1/kstart for transported particles. The pause time, in
transport [409, 460]. particular, can be reduced by tethering the particle
The relative importance of diffusive versus pro- to the microtubule track and thereby increasing the
cessive motion thus depends on both the length scale rate at which it can re-engage with the machinery
of interest and the overall objective of transport. for motor-driven transport. Such tethering is partic-
For instance, the uniform dispersion of an initially ularly effective when the microtubules themselves are
concentrated bolus of particles is optimized at inter- sparsely distributed and diffusion toward a micro-
mediate values of the run length  and of the active tubule becomes rate-limiting for initiating transport
fraction f [25]. For particles that are only able to [25, 393]. Recent mechanical modeling of hitchhik-
carry out their function in the passive state (e.g.: pro- ing transport for fungal peroxisomes indicates that
teins that must be released from a vesicle), reaction tethering to microtubules could enhance the rate
kinetics are fastest at intermediate fractions of time in of starting a hitchhiking run by up to an order of
active motion [461]. Even for constantly active par- magnitude [393]. For directly motor-driven cargo,
ticles, when the domain is sufficiently long and f is tethering and preventing dissociation from the micro-
sufficiently high, the search time for a single particle tubule has been proposed as a cooperativity mech-
to hit a target is also optimized at intermediate run anism for motors with opposing polarity [19]. In
lengths, which preclude very long excursions in the terms of transport efficiency, the enhanced starting
wrong direction [462]. rate for active motion due to tethering is balanced by
When the transport objective comprises efficient reduced diffusive exploration during the paused state.
encounter of a target by the first in a uniform popu- For organelles that spend a small fraction of time
lation of particles, the relevant length scale becomes engaged in processive motion, tethering is beneficial
the inverse of the particle spatial density, which tends for transport only on length scales beyond Lcrit ≈
to be on the order of 0.1–10 μm. For densities higher x∗ /(1 − â2 )2 , where â is the ratio between the capture
than 1/x∗ , target search is dominated by diffusive radius around a microtubule and the characteristic
transport, whereas for lower densities motor-driven separation between parallel microtubules [25].
motion predominates. Interestingly, many organelles By tuning pause rates and durations, as well as the
capable of motor-driven transport have been found to mobility state of a particle while paused, cells can thus
spend only a small fraction of time actually engaged regulate overall particle dispersion through an inter-
in processive motion [26, 382, 392, 457]. These par- play of passive and processively moving transport
ticles can have sufficiently high values of x∗ such modes.

24
Phys. Biol. 17 (2020) 061003 Topical Review

Figure 15. Patterns of cytoskeletal organization in animal cells. Sketches of microtubule organization show nuclei in gray,
microtubules in red, minus ends in cyan (adapted from [463]). (a) Polarized microtubules in a neuronal axon organize in
bundles. Right: confocal migrograph of growing neurons show microtubules in purple, actin in green, and nuclei in blue (from
[464]) (b) radially polarized microtubules in mouse fibroblast cell. Right: fluorescent image of actin (purple) and microtubules
(yellow) (from [465]). (c) and (c ) Microtubule organization in Drosophila oocytes changes depending on the stage of
development. Microtubules are visualized via immunofluoresence in early (c) and late (c ) stages of embryo development
(micrographs from [466]). (d) A reverse polarized organization is seen in mouse keratinocytes. Middle: apical confocal slice of
fixed primary cultured keratinocytes with fluorescently labeled microtubule binding protein (esconsin). Right: for the same cell
type, nucleus labeled in blue and microtubule-anchoring protein ninein shown in green. Micrographs from [467].
(e) Unidirectionally polarized microtubules in intestinal epithelial cells. Right: Super-resolution image showing minus-end
binding protein CAMSAP3 (red, inset) and microtubules (green). Micrograph from [468].

4.2.5. Organization of cytoskeletal tracks The two types of cytoskeletal filaments serving
The intracellular distribution of cargos and their as transport highways exhibit very different orga-
efficiency at reaching cellular regions can be con- nizations within the cell. Actin filaments tend to
trolled at several levels. As discussed in section 4.2.2, form branched networks of varying densities. In
biochemical modification or binding of signaling mammalian cells, dense actin networks are usually
restricted to a cortical layer (∼100 nm thick) beneath
molecules to motor-proteins, adaptors, linkers, and
the cell membrane [472]. Away from the leading edge
cytoskeletal tracks can regulate the processivity and
of migrating cells, these cortical actin filaments tend
directional bias of cargo moving along a single micro-
to be isotropic, without a defined polarity [473].
tubule. However, models of transport that rely on Consequently, transport within the actin network
uniform constant-rate processes at the single-cargo tends to appear characteristically diffusive, even when
level tend to be insufficient to reproduce the com- driven by motor proteins [366, 461]. The effective
plex behavior of motor-driven cargos in vivo [19, 415, diffusivity of particles moving within the actin net-
416]. Some of this complexity may be due to spa- work is thought to be regulated in different cellular
tially or temporally heterogeneous regulation, with states by altering the switching probability at each fil-
gradients in signaling molecules responsible for mod- ament intersection, thereby controlling the processive
ulating transport parameters in different regions of run-length of the cargo [366].
the cell [185, 188, 435]. However, an additional key By contrast, the microtubule cytoskeleton can
source of spatial heterogeneity is the organization of form a variety of structures with different degrees
of polarity and spatial organization. Microtubules
the cytoskeletal highways themselves. This organiza-
nucleate at discrete sites termed MTOCs, which
tion both determines and is set by cell shape and
anchor their minus ends while allowing plus ends
polarity, allowing for a close coupling between cel-
to grow outward. The best-studied MTOC in ani-
lular function, morphology, and transport logistics mal cells is the centrosome, which is located near
[364]. In some systems, incorporating the explicit dis- the nucleus, and nucleates an aster-like structure of
tribution of cytoskeletal filaments has been shown to microtubules extending their plus ends toward the
be sufficient to explain observed transport behavior cell periphery [463] (figure 15(b)). At the periph-
while maintainining spatially uniform cargo unbind- ery, microtubules can penetrate the cortical actin
ing rates [469–471]. network, allowing cargos to switch from long-range

25
Phys. Biol. 17 (2020) 061003 Topical Review

transport on microtubules to short range motion the dendrite tip but more efficient establishment of
on actin filaments [474], in a manner dependent a uniform distribution of cargos within the dendrites
on the complement of attached motors [475]. A [486]. Modeling of early endosome transport in fun-
number of non-centrosomal microtubule-organizing gal hyphae demonstrated that spatially uniform rates
structures have also been identified, allowing for of motor switching and microtubule nucleation are
anchoring of minus ends in many different cellular sufficient to reproduce experimentally observed accu-
regions, and giving rise to microtubule networks with mulation of endosomes in different hyphal regions in
varying polarity and orientational alignment [463] response to dynein and kinesin-3 motor mutations
(figure 15). Some MTOCs are associated with the [469].
Golgi body and its outposts, allowing for direct deliv- Additional effects beyond a purely one-
ery of dynein-driven vesicles carrying secretory cargo dimensional system arise when considering the
from the ER to the Golgi [476]. radial spacing of microtubules within a cylin-
In certain cell types, including Drosophila [477] drical cellular projection. Because motor-driven
and Xenopus [478] oocytes as well as epithelial cells organelle transport can only be initiated when the
[479], microtubule nucleation is localized at the organelle passes close to a microtubule track, the
cell cortex (figures 15(c)–(e)). While fully polarized cross-sectional movement of organelles can play
epithelial cells can establish unidirectional micro- an important role in their dispersion. Modeling of
tubule structures (figure 15(e)), oocytes tend to 3D particle dynamics has shown, for instance, that
exhibit largely disordered cytoskeletal organization tethering of hitchhiking peroxisomes to microtubule
[477]. Nevertheless, a statistical bias in microtubule tracks is expected to greatly increase their overall
orientation can be sufficient to enable robust local- rate of transport, particularly when there are very
ization of cellular components [358, 470, 471, 480]. few parallel microtubules in the cellular region
For Drosophila oocytes in particular, a gradient of [393]. Cylindrical models with explicit microtubule
microtubule nucleation densities at the cell periphery
arrangements form a natural transition from one-
was shown to be sufficient to establish a structured
dimensional models to local regions of fully 3D
velocity field for motor-driven motion throughout
systems that are lacking in microtubule intersections.
the ooplasm, when averaged over many realizations
For example, the asymmetric densities of parallel
of a cytoskeleton that turns over on minute time-
microtubules observed in Drosophila cell spindles
scales [470]. The resulting orientational bias allows
can be incorporated into a 1D transport model
kinesin-driven mRNA molecules to accumulate at the
that explains the uneven distribution of endosomes
posterior pole despite executing many rapid runs in
between daughter cells [487]. Other modeling efforts
all directions [481]. Simulation studies incorporat-
have shown that random spacing of locally parallel
ing the biased orientation field accurately reproduce
microtubules leads to a higher long-range effective
both this posterior localization and the more complex
diffusivity of motor-driven particles than does purely
splitting behavior of dynein-driven mRNAs, whose
uniform spacing [471].
ultimate localization depends on the point of injec-
tion [470, 482]. In many animal cell types, microtubules form 3D
Elongated cellular regions, such as neuronal pro- networks with frequent intersections between indi-
jections or fungal hyphae, generally exhibit arrays of vidual filaments [2, 356]. These intersections serve
parallel microtubules, arranged into polarized bun- as both obstacles for cargo moving along a micro-
dles [483]. Microtubules in neuronal axons are uni- tubule (section 4.2.3) and as an opportunity to alter
formly oriented, with their plus ends pointing to the the direction of motion. The probability of switch-
distal end of the projection [484]. In dendrites, the ing tracks at a microtubule intersection is dependent
orientation can be uniform with minus end outwards on the 3D spacing and orientation of the intersecting
(in Drosophila and Caenorhabditis elegans neurons) microtubules [2, 447], as well as the cargo size [445]
[484] or mixed with plus ends in both directions and complement of attached motors [448]. Live-cell
(in vertebrate neurons) [485]. The ability of cargos tracking studies indicate that most cargos tend to pre-
to be transported selectively to dendrites or axons is serve the anterograde or retrograde polarity of their
thought to rely on varying recruitment of motor sub- motion upon passing microtubule intersections [2],
types [486] together with post-translational modifi- an effect which may arise from the radially polarized
cations of the microtubule tracks [483]. organization of the microtubule network.
The parallel architecture of microtubules in these As with one-dimensional models, the motion of
cellular projections is conducive to modeling studies motor-driven particles over cytoskeletal networks is
that treat the system as essentially one-dimensional, generally assumed to consist of stochastic switching
representing the density of microtubules, cargos, and between processive runs along filaments and slow
motors as mean-field distributions along the axis of passive phases [461]. While the passive phases are
the projection. For example, a model of dynein-driven generally treated as diffusive, they may also involve
dendritic transport showed that microtubule arrays of tethering to stationary structures [19, 458, 459].
mixed polarity resulted in slower delivery of cargo to Recent work in which the passive mode is treated as

26
Phys. Biol. 17 (2020) 061003 Topical Review

a continuous-time random walk with broadly dis- plethora of control parameters tuned for different car-
tributed step times indicates that such intermittent gos, cell types, and cellular states. The factors sub-
motion would give rise to a characteristic distribu- ject to cellular control include cytoskeletal organi-
tion of first passage times to the cell periphery [488]. zation, motor recruitment, processivity of individual
Namely, a peak of particles arriving at short times is motors, and cooperative interactions between motor
expected, followed by a sustained long tail of sporadic teams. However, motor transport is limited in its
particle arrivals—a biphasic pattern which has been maximum speed, has a high metabolic cost in ATP
observed for the exocytic release of insulin granules consumption, and requires additional complexity in
[489]. the packaging of molecular components into vesi-
The density, spatial distribution, and polarity of cles or motor-driven complexes. An alternate mode of
cytoskeletal filaments in a 2D or 3D cellular region directed intracellular transport, the movement of par-
plays an important role in determining the over- ticles by cytoplasmic flow, offers cells the opportunity
all transport of cargo. Denser networks of filaments to circumvent some of these challenges.
allow cargos to spend more time in the actively mov-
ing phase. However, more dense networks also imply 4.3. Advective transport: intracellular flows
more frequent filament intersections and thus shorter In addition to directed motor-driven motion along
processive runs. Simulations on randomly oriented cytoskeletal highways, active transport in the cell can
2D networks indicate that the mean first-passage time be achieved through advection, with particles car-
from a central nucleus to the cell periphery is largely ried along by the flow of intracellular fluids. This
determined by the total mass of cytoskeletal tracks, phenomenon was first discovered in plant cells [493,
with faster transport at higher total filament content 494], but has since been observed in a variety of pro-
[490]. For the same total network densities, struc- tist [194, 495, 496], fungal [26, 497], and animal [3,
tures with a few long filaments tended to exhibit 466, 498] cell types. In plant cells, particularly, cyto-
much greater variation in transit times than those plasmic flow has long been thought to play a crucial
with many short filaments, an effect arising from the role in distributing molecular components through-
presence of ‘traps’ where processively moving cargo is out the cell: replenishing depleted regions, control-
directed into a localized region of the network [490, ling delivery rates of metabolic reactants, and (with
491]. The polarity of randomly scattered filaments the aid of diffusion) smoothing intracellular gradients
plays an important role in determining transition [180, 499].
times across the network, and reversing the polarity The processivity, speed, and spatial correlations
of a single filament can alter the first-passage times for transport by fluid flow can vary widely among
several-fold [491]. different cellular systems. At one extreme are highly
Spatially inhomogeneous network structures can coordinated and extensive flows in macroscopic cells,
also help optimize transport of intermittently motor- such as cytoplasmic streaming in the internodal cells
driven cargos. Regions of randomly oriented short of characean algae (persistent spiral flows over cen-
filaments serve to locally enhance the effective par- timeter scales at speeds of 100 μm s−1 ) [29, 32] or
ticle diffusivity. Continuum models show that when peristaltic shuttle flows in the hyphae of the giant
such a region is placed closer to the center of a cir- slime mold P. polycephalum (reaching speeds up to
cular domain, the mean first-passage time of particles 1 mm s−1 ) [194, 197]. At the other extreme are short-
from the center to the domain boundary can be sig- range perturbations due to hydrodynamic entrain-
nificantly decreased [490]. By contrast, when the goal ment by passing motor-driven cargo, which have been
of a transport system involves locating a specific nar- hypothesized to contribute to ‘active diffusion’ of
row target on the periphery, then optimal search rates axonal vesicles [107] and fungal peroxisomes [26].
can be obtained by an ordered radial arrangement For simplicity, many studies of intracellular fluid
of polarized filaments in the cell bulk, coupled with flow represent the cytoplasm as a linearly viscous
a thin shell of random filaments near the periphery (i.e.: Newtonian) fluid, subject to various bound-
[474, 492]. In this case, cargo is delivered in a directed ary conditions and perturbed by stresses that can be
fashion to the peripheral layer, followed by effectively generated both at the cellular boundary and within
diffusive exploration of the boundary. Such a mor- the bulk [3, 28, 29, 107, 500, 501]. More complex
phology is indeed observed in many cell types which mechanical models have also been developed, treat-
maintain a radially polarized microtubule cytoskele- ing the cytoplasm as a poroelastic material consist-
ton originating at the centrosome near the nucleus ing of a fluid phase intercalated with and rubbing
and a thin largely disordered cortex of actin filaments against an elastic solid phase [502, 503]. Such poroe-
that may contribute to localized cargo transport in lastic models can more accurately reproduce the flow
peripheral or distal regions [368]. patterns arising in response to specific cellular forces
Motor-driven transport is a ubiquitous feature involved in blebbing, motility, and indentation [349,
of eukaryotic cells. Its unique advantage lies in its 496, 504], as well as propagating waves that arise
ability to deliver and disperse cargo in an efficient from mechanochemical coupling between cytoplas-
and regulated manner that can be modified via a mic activators and cytoskeletal contractions [505].

27
Phys. Biol. 17 (2020) 061003 Topical Review

Here, we focus primarily on the role of flow patterns Waves of actomyosin contraction are responsible for
in particle transport, and we restrict our discussion to the peristaltic shuttle flows in slime mold hyphae
models of the cytoplasm as a simple fluid. [197, 519], as well as flows that drive spindle posi-
tioning in mammalian oocytes [520] and nuclei dis-
4.3.1. Fundamentals of advective transport persion in Drosophila embryos [508]. Myosin-driven
As discussed in section 2, flows of intracellular flu- contraction at the cell rear also drives flow toward
ids generally lie in the regime of very low Reynold’s the leading edge in migrating keratocytes [521] and
numbers, where viscous forces dominate over iner- neutrophil cells [522]. These flows can be regulated
tia. In this ‘Stokes flow’ regime, fluid flows are lam- by gradients in the distribution of myosin motors
inar, particle velocities are proportional to applied or of signaling molecules that trigger myosin activa-
forces, and flow patterns are established nearly instan- tion. When the molecules regulating contraction are
taneously throughout the domain for any given pat- driven by the flow itself, precise patterning of flows
tern of applied stresses [21]. Such systems are subject and molecular distributions can be established across
to an effect which has been whimsically referred to the entire cell [28, 508, 523, 524]. Example flow pat-
as the ‘scallop theorem’, where time-reversing flows terns generated by actomyosin contraction are shown
result in no net movement of the advective particles in figures 16(a)–(c).
[22]. In essence, particles that are mixed by stirring Large-scale contraction of the actomyosin net-
in a low Reynold’s number fluid can be un-mixed by work is often associated with deformation of the
repeating the same stirring motions in reverse [499, cell shape during migration [496, 522, 525], division
510, 511]. As a result, simple oscillatory back-and- [526, 527], and development [498, 528]. In many
forth flows cannot, in and of themselves, result in cases, however, cell shape dynamics are driven pri-
particle transport. However, long-range transport can marily by leading edge extension through directed
be achieved by the establishment of steady, persis- polymerization of the actin cytoskeleton [529–531],
tent flow patterns (as for cytoplasmic streaming in as in the migrating neutrophil-like cell in figure 16(d).
plant cells [32, 512]) or by coordinated oscillations Growing cells, such as fungal hyphae, may also har-
that propel material via peristalsis (as in the shuttle ness gradients in osmotic or turgor pressure to drive
flows of slime molds [194, 197]). flow toward extending tips [497, 532] (figure 16(e)).
The spatiotemporal distribution c(x, t) of particles Regardless of its origin, deformation of the cell
subject to both diffusive motion and flow is described boundary gives rise to cytoplasmic flows that can con-
by the advection-diffusion-reaction equation [513]: tribute to intracellular mixing [3] or overall transla-
tion of the cytoplasm [514].
d c
=∇
· (D∇c)
−∇ · ( v c) + R(x, t), (16) An additional major source of flow is hydrody-
dt namic entrainment by motor-driven cargo. Long-
where D is the diffusivity, v the fluid flow field (which range, persistent flows are particularly prominent
can vary over space and time), and R is a reaction in plant cells (figures 16(f) and (g)), where myosin
term that describes sources or sinks that may arise motors carry a variety of organelles along bundled
from chemical reactions. This general equation can actin filaments organized around the cell periphery
be leveraged to describe pattern formation and signal [29, 499, 533, 534]. The motion of these organelles
propagation in a variety of cellular systems with cyto- entrains a thick layer of cytoplasmic fluid, result-
plasmic flow [28, 506]. The importance of flow versus ing in streaming flows that can reach 100 μm s−1
diffusion over a length scale L is characterized by the [534]. In animal cells, entrainment-driven flows tend
Péclet number Pe(L) [32], which is defined generally to be slower and more spatially heterogeneous. In
for directed transport processes (equation (2)). Drosophila oocytes, for instance, kinesin-bound car-
A large Péclet number (Pe  1) indicates gos are responsible for slow, apparently random
advection-dominated transport. For non-stationary flows (25 nm s−1 ) and rapid, coordinated stream-
flows, the length scale can be replaced by L = vτ , ing (300 nm s−1 ) during different stages of oogenesis
where τ is the persistence time of the flow pat- [466] (figures 16(h) and (i)). Seemingly random flow
tern. Cellular transport systems where advection patterns in the early oocyte tend to be spatially cor-
is believed to play an important biological role related on the few-micron scale (figure 16(h)), likely
have Péclet numbers in the range Pe ≈ 2–1000, as due to the underlying organization of the microtubule
summarized in table 1. cytoskeleton [86, 507, 538] (see figures 15(c) and
(c )).
4.3.2. Generating cytoplasmic flow patterns In other systems, where cellular-scale flows are
Several distinct mechanisms are capable of gener- not directly evident, the bidirectional motion of
ating intracellular flows. The first mechanism relies motor-driven cargos may nevertheless give rise to very
on the contraction of actin filament networks by short-range entrainment events for nearby tracer par-
myosin motors. Large-scale flow patterns have been ticles [107]. When the cargo motion is slightly biased
observed in reconstituted in vitro active gel systems toward one direction, an overall slow flow of pas-
with actin turnover and myosin activity [517, 518]. sive cytoplasmic contents will arise. For example, a

28
Phys. Biol. 17 (2020) 061003 Topical Review

Table 1. Péclet numbers for example cellular systems where flows have been shown
to play a role in cytoplasmic transport.

Particle Cell type Péclet number References

PAR proteins C. elegans zygote 3 [506]


Acidified vesicles Human neutrophil-like (HL-60) 11 [3]
mRNA Drosophila oocytes 10–100 [507]
Bicoid morphogens Drosophila embryos 80 [508, 509]
Small molecules Characean algae internodal cells 100–1000 [500]

Figure 16. Spatial patterns of cytoplasmic flow. (a) Peristaltic shuttle flow in P. polycephalum plasmodium fragment (from
[514]). (b) Elongational flow and unidirectional flow in contracting starfish oocytes (from [498]). (c) Bidirectional fountain flow
in Drosophila embryo, at cell cycle 6. Red arrows show cytoplasmic flow and blue arrows show nuclear trajectories (adapted from
[508]). (d) Flow in migrating neutrophil-like HL60 cell associated with deformation of cell boundary (pink arrows). Red arrows
show velocity of acidified organelles, green arrows show the computed flow pattern based on boundary deformation (adapted
from [3]). (e) Eddies formed near pore constriction for hyphal flow in Neurospora crassa fungi (adapted from [515]). (f) Reverse
fountain flow in lily pollen tube, with organelle velocities shown (from [516]). (g) Spiral streaming in characean algae, with
indifference zone marking boundary between axial flow directions (from [32]). (h) Disordered yet spatially correlated flows in
stage 9 Drosophila oocyte (from [507]). (i) Circulating flow in stage 11 Drosophila oocyte (from [466]).

bias toward anterograde cargo motion in growing teins [506, 535]. Directional advective transport also
fungal tips has been suggested to give rise to a very contributes to delivering cytoplasmic contents that
slow directed polar drift (0.5 nm s−1 ) that leads to drive cellular growth in a variety of systems, including
organelle accumulation when other active transport the developing axon [536], slime mold plasmodium
mechanisms are removed [26]. [194], fungal hypha [497], and elongated plant and
The variety of spatiotemporal flow patterns gen- algal cells [32, 512].
erated by different cellular mechanisms contributes The simplest model for localization and gradient-
to the distribution and dispersion of intracellular formation by advection consists of a one-dimensional
particles ranging from small nutrient molecules to domain of length L with reflecting boundary con-
proteins and organelles. Unlike diffusion, flows can ditions and a steady unidirectional flow of velocity
drive the motion of even very large particles. Unlike v. The steady-state distribution of a particle with
motor-driven active transport, they affect all particles diffusivity D is then given by solution of equation
passing a particular region, without the level of regu- (16) as
lation derived from specific adaptors coupling motors c Pe ePe·(x/L)
c(x) = (17)
to cargos. We proceed to consider the functional con- ePe − 1
sequences of various flow patterns on both directed
localization of cellular components and overall mix- where c is the average density and Pe is the Péclet
ing of cell contents. number over the domain. High Péclet numbers lead
to sharp accumulation of density at the domain
4.3.3. Directed transport and localization by flow boundary, while lower values result in a more uni-
Stable, persistent cytoplasmic flow provides a mech- form distribution (figure 17). Because diffusivity gen-
anism for directed transport of cellular components, erally scales with particle size, larger particles develop
allowing the establishment of intracellular gradients sharper gradients under a given flow—an effect that
and the localized positioning of organelles. In mam- has been used to estimate flow velocities in the leading
malian oocytes, cytoplasmic flow drives the place- edge of crawling keratocytes [521].
ment of the meiotic spindle near the cortical cap Gradients can be further enhanced by a polarized
[520]. In C. elegans zygotes, flows with Pe ≈ 3 con- distribution of molecules capable of binding the par-
tribute to the anterior accumulation of PAR pro- ticle of interest (figure 17). Weak binding, along with

29
Phys. Biol. 17 (2020) 061003 Topical Review

through the pore leads to the formation of circular


eddies on the upstream side of the septum, which can
serve as a subcellular compartment. These compart-
ments locally entrap nuclei that proceed to differenti-
ate to a transcriptional program which differs from
other nuclei in the same cytoplasm [515]. Further-
more, the flow-driven accumulation of vesicles at the
septa has been hypothesized to contribute to hyphal
branch formation [191].
The entrainment of cytoplasm by motor-driven
cargo also raises the problem of fluid cycling when the
cargo approaches the end of a cellular region. Model-
ing studies indicate that the recirculatory flow engen-
dered by this entrainment may counteract directed
transport, washing unbound cargo and other passive
Figure 17. Particle localization through the interplay of particles out of the target zone [538]. The resultant
advection, diffusion, and binding. Solid lines show
steady-state profile of particles at average concentration c in coupling between motor-driven motion and advec-
a 1D domain under the influence of flow and diffusion tive flow implies that disordered, weakly directional
(equation (17)). Color corresponds to different Péclet
numbers. Dashed lines show the distribution profiles in the
cytoskeletal networks may in fact lead to more opti-
presence of binding sites at concentration 10c and mal local accumulation of particles [538].
dissociation constant KD = c, located within the last 10% Many cellular advective transport systems rely on
of the domain (green region). Inset illustrates example
particle distributions corresponding to these profiles. relatively stationary flow patterns that persist over suf-
ficient time periods to enable particle delivery across
the cell. However, important counterexamples exist,
directed flow, can combine to segregate a molecule where large-scale directed movement of cytoplas-
into a specific cellular region, while allowing for mic contents is achieved through coordinated time-
rapid equilibration within that region. An analogous varying flows. A particularly well-studied example is
mechanism has recently been shown to underlie the peristaltic shuttle flow observed in slime molds,
the accumulation of proteins in the outer seg- both in their migrating ameboid [496] and their
ment of mammalian photoreceptor cells [537]. It hyphal network [194] state. These flows are gener-
should be noted that the distributions described by ated by directionally propagating contraction fronts
equation (17) and its generalizations do not require that are thought to be self-organizing via a signal-
that v represent fluid flow specifically. Any kind of ing molecule that both amplifies contractions and is
directed transport process that moves all relevant advected by the flow itself [28]. In general, peristaltic
particles passing a particular point in space with flows require an organized spatial gradient of contrac-
the same velocity can supply the advective drift v. tion phases, allowing for overall directed transport of
This could include, for instance, the IFT trains that fluid contents [539]. In tubular network structures,
transport proteins into primary cilia [142, 143]. By advective transport is optimized when the wavelength
interacting only with certain specific proteins, such of the peristaltic wave is comparable to the network
forms of directed transport allow for more precise size, consistent with the observed phase correlation
control over the patterning and accumulation of patterns in P. polycephalum hyphae [194]. An alter-
intracellular particles. nate example of cytoplasmic transport by oscillatory
Conservation of mass implies that when advec- flows has recently been observed in multinucleate
tive flow delivers cytoplasmic contents to specific cel- Drosophila embryos, where vortex-like flow patterns
lular regions, the fluid itself must either recirculate (figure 16(c)) oscillate in coordination with the cell
or deform the cell contour. In growing or migrat- cycle. These flows are able to drive the separation
ing cells, expansion of protrusions provides a reser- of nuclei originally clustered near the embryo center
voir for newly arriving cytoplasm (figures 16(a) and to well-spaced positions along the anterior–posterior
(d)). Other transport systems rely on fountain flow axis [508].
patterns (figures 16(c) and (f)) that cycle the incom-
ing fluid with peripheral flow in the reverse direction
from flow along the central axis. In these flow pat- 4.3.4. Enhancing mixing through flow
terns, local binding or rapid removal via metabolism In addition to targeted delivery and patterning, cyto-
or exocytosis is needed to prevent newly delivered plasmic flows can also drive more efficient mixing
molecules from being flushed back by the recirculat- of cellular components. Mixing in the world of low
ing flow [516]. Yet another pattern of advective deliv- Reynold’s number fluids relies on two distinct phys-
ery is seen in some fungal hyphae, where flows pass ical effects: Taylor dispersion (the smearing out of
between cellular regions separated by septa with a nar- concentration gradients by diffusion) [541, 542] and
row central pore (figure 16(e)). The focusing of flow Lagrangian stirring (the chaotic motion of particles

30
Phys. Biol. 17 (2020) 061003 Topical Review

Figure 18. Enhancement of mixing by cytoplasmic flow. (a) Taylor dispersion of simulated particles in a Poiseuille flow (red
arrows). Initial bolus of particles is shown spreading by flow only (Pe = inf), diffusion only (Pe = 0), and both processes
(Pe = 100). (b) Computed distribution of material in the cross-section of a cylindrical algal cell undergoing spiral streaming.
Flow generates steep asymmetric gradients, increasing flux of diffusive material into cell (from [29]). (c)–(e) Lagrangian stirring
by unsteady flows. (c) Stretching and folding of initial particle distribution due to flows in a 2D circular domain with boundary
dynamics representing deformation of migrating HL60 cell (adapted from [3]). (d) Simulated spreading of initially clustered
nuclei along the cell axis, based on measured fluid flows in Drosophila embryos during cell cycles 4–7 (from [508]). (e) Computed
trajectories of a tracer sphere entrained in the flow field generated by an active sphere moving processively along the central axis
of a no-slip cylinder. Spheres represent organelles of radius 100 nm; cylinder represents a hypha of radius 1 μm. Each trajectory
corresponds to a different starting position. Inset shows entrained sphere velocity at each position. Flow field computed as
described in [540].

driven by a spatially heterogeneous, unsteady, non- to increase flow speeds in a few large central tubules,
reversing flow) [543, 544]. Taylor dispersion arises increasing particle dispersion [195].
from spatially varying rates of flow, which give rise In addition to Taylor dispersion, Lagrangian stir-
to gradients in particle densities, resulting in an effec- ring resulting from unsteady fluid flow patterns also
tively higher diffusivity of particles across the stream- contributes to mixing in cellular systems, even for
lines (figure 18(a)). For steady Poiseuille flow in a tube particles whose diffusion is negligible. Stirring is
[545], the effective diffusivity along the cross-section often described by quantifying the extent to which
of the tube is given by a given region of fluid stretches and folds under
the flow (figure 18(c)), increasing the length of its
Deff = D(1 + Pe2 /48), (18) boundary with the surrounding fluid [544]. These
boundaries mark regions of high gradients, which
can then be smoothed by diffusion. Stirring thus
where Pe refers to the Péclet number (equation (2))
acts together with Taylor dispersion to mix the sys-
computed for the average velocity in the tube over
tem across different length scales. Lagrangian stirring
the length-scale of the tube radius. For the rapid
arises from the fact that even a relatively simple lami-
contractile flows in P. polycephalum (velocity ≈
nar flow pattern for a low Reynold’s number fluid can
0.1 mm s−1 , radius ≈ 50 μm, Pe ≈ 50), the effective
nevertheless lead to highly complex (chaotic) trajec-
dispersion of small molecules (defined as 1/Deff
−1 )
is increased by up to 7-fold [195]. tories of individual particles or fluid elements [543,
Several cellular systems with more complicated 546, 547]. Chaotic trajectories are characterized by
flow patterns have also been hypothesized to enhance positive Lyapunov exponents [544], which quantify
diffusive transport through the flow-induced for- the exponential divergence of paths for two initially
mation of steep gradients. In late-stage Drosophila close particles carried by the fluid. Although a steady
oocytes, streaming flows exhibit faster velocities flow field can yield such diverging trajectories in three
toward the cortex (figure 16(i)), leading to cytoplas- dimensions, time-varying flow patterns are needed
mic shear gradients that may contribute to mixing to generate chaotic stirring in 2D fluids [544, 547].
[466]. In the long cylindrical cells of characean algae, An example of diverging particle trajectories due to
high shear rates result from rapid spiral cytoplas- unsteady flow is seen in the spreading of nuclei along
mic streaming [29, 32, 500]. These flows are expected the anterior–posterior axis of Drosophila embryos
to give rise to radial concentration gradients that (figure 18(d)) [508].
augment diffusive entry of nutrients into the cell In practice, extensive and efficient stirring
(figure 18(b)). Interestingly, the geometry of flow pat- can be achieved by unsteady flows that are not
terns can be used to tune the gradient steepness and time-reversing. In such systems, the Péclet number
hence the rate of mixing or diffusive uptake. Intern- associated with instantaneous flow velocities does
odal cells of the algae Nitella axillaris alter the wave- not adequately describe the overall stirring behavior,
length of their spiral flows as the cell grows, with a since partial flow reversals tend to drive particles
maximum in both diffusive uptake and growth rate back toward their starting points. Instead, one can
arising at a specific cell length [29]. Foraging P. poly- characterize the effect of flow on mixing by defining
cephalum slime molds prune their network structure an ‘effective Péclet number’ on any given time-scale,

31
Phys. Biol. 17 (2020) 061003 Topical Review

as the overall displacement of a tracer particle driven their importance in cellular transport is still lacking.
by flow alone versus the diffusive displacement over In a tubular system, each entrainment event from the
the same time period [3]. Starfish oocytes serve an passage of a single organelle should yield a finite short
example of a cellular system with rapid back and displacement () of a tracer particle of length com-
forth flows (figure 16(b)) but no significant overall parable to the organelle size [107] (figure 18(e)). If
displacement of large cytoplasmic particles [498]. organelles pass near the tracer at a frequency kpass ,
More complex dynamically evolving flow patterns the effective diffusion coefficient for the tracer is then
are observed in the cytoplasm of crawling cells given by
executing amoeboid-like deformations [3, 496, 548].
Deff = D + 2 kpass (19)
Numerical simulations indicate that the flows arising
from deformation of neutrophil-like migrating where D is the tracer diffusivity in the absence of
cells (figure 16(d)) are sufficient to substantially active motion. In fungal hyphae, knocking out an
enhance the mixing of lysosome-like organelles in endosomal motor adaptor results in a decrease in the
the cytoplasm (Peeff ≈ 11 over 30 s timescales) [3]. diffusivity of passive peroxisome organelles by ΔD ≈
In late-stage Drosophila oocytes, dynamic buck- 0.01 μm2 [26]. This effect is comparable to the pre-
ling of microtubule tracks due to drag forces on dicted contribution due to entrainment by passing
motor-driven cargo is thought to give rise to local endosomes, at a frequency of kpass ≈ 1/s [392], in
time-variation in the overall flow pattern [466, 549]. accordance with equation (19).
The resulting unsteady flows (figures 16(h) and (i)) Flow of cytoplasmic fluids thus constitutes a ver-
have been shown to homogenize the distribution satile mechanism for transport across a broad range of
of initially concentrated yolk granules within the length scales. Coordinated patterns of flow can result
cytoplasm [466, 550]. in the directed delivery of bulk cytoplasmic contents
Locally oscillating flows can efficiently drive dis- at speeds far higher than those reached by motor-
persion when the particles are confined in a domain driven transport. Furthermore, flows contribute to
of complex geometry and the overall spatial pat- mixing of cytoplasmic contents through the forma-
tern of flows is stochastic. A biologically relevant tion of gradients smoothed by Taylor dispersion,
example is the luminal flow generated by random through Lagrangian stirring, and potentially through
contractions in a tubular network, as in slime-mold the generation of effectively diffusive active motion
hyphae [195]. In order for such flows to contribute via stochastic local entrainment events.
substantially to mixing, they must be rapid enough
and persistent enough to enable individual particles 5. Perspectives
to transition between nodes before the contraction re-
opens, reversing the flow. Once a particle reaches a Over the past decades, many of the molecular compo-
network junction, flow splitting and small time delays nents driving transport within eukaryotic cells have
in flow reversal at adjacent edges ensure that the parti- been characterized in great detail. Studies of in vitro
cle does not get restored to its initial position, thereby systems have allowed for a quantitative understand-
promoting mixing through Lagrangian stirring [551]. ing of the mechanochemical behavior of molecular
On a smaller scale, flows arising from unco- motors, both individually [379, 380] and in cooper-
ordinated tubular contractions have recently been ating or competing groups [19, 383, 403, 423]. More
hypothesized to drive node-to-node transport of pro- recently a plethora of adaptors and regulatory fac-
teins in the mammalian ER network [59]. Processive tors modifying either the motor-cargo complex or
particle velocities on the order of 20 μm s−1 , over time the cytoskeletal tracks have been identified [20, 427,
scales of 30 ms, have been measured for individual 430]. In the context of non-directed transport, the
proteins tracked in ER tubules, which exhibit a lumi- effects of crowding [265], filamentous networks [267,
nal diffusivity in the nodes of about 0.5 μm2 s−1 [59]. 268], and actively contracting gels [517, 518] on par-
These estimates yield a Péclet number of Pe ≈ 20 and ticle motion have also been extensively explored in
allow for individual processive trajectories to cover a reconstituted systems. However, the behavior of this
distance comparable to the typical edge length in an formidable array of molecular players in the com-
ER network (∼ 1.5 μm). plex and dynamic intracellular environment remains
An additional role for stochastic intracellular in many ways mysterious. We summarize below some
flows in driving cytoplasmic mixing is through unco- of the main outstanding questions associated with
ordinated entrainment by motor-driven cargos. Such each of the physical transport modes employed by
entrainment events result in short-range runs, lead- eukaryotic cells.
ing to an enhanced ‘active diffusion’ driven by Perhaps one of the largest outstanding ques-
bidirectionally moving cargos. Localized entrain- tions pertaining to the stochastic ‘Brownian’ motion
ment has been hypothesized to account for the of intracellular particles is the nature of the non-
‘slow component’ of axonal transport [107] and the thermal active forces that drive their movements.
kinesin-dependent ‘active diffusion’ of peroxisomes How much of the apparently diffusive particle motion
in fungal hyphae [26], although direct evidence of can be attributed to active contraction of cytoskeletal

32
Phys. Biol. 17 (2020) 061003 Topical Review

networks [232, 259], to localized hydrodynamic ponents. The roles of confinement in complex mor-
entrainment [26, 107], or to non-specific microscopic phologies, as well as crowding and medium vis-
agitation of the medium associated with conforma- coelasticity, in modulating diffusion-limited reac-
tional changes of ATP-burning enzymes [260, 262]? tion rates have been explored theoretically [330, 335,
To what extent can decreased mobilities associated 338, 350]. However the importance of these effects
with ATP depletion or myosin inhibition be treated in specific intracellular reaction systems remains
as a rigidification of the medium [552, 553] or a unclear. Similarly, the interplay of motor trans-
reduction in the ‘effective temperature’ [554] within port, flow, and diffusion [474, 492, 538], as well
the cell? Recent studies have begun to tease apart the as the role of cytoskeletal track arrangements [30,
nature of these delocalized driving forces, separating 31] in particle delivery and sorting is still an area
them out from the continuum rheological proper- of active exploration. The contribution of transport
ties of the intracellular medium [232, 255]. However, limitations to the behavior of complex biochem-
the consequences of this breakdown in the fluctua- ical reaction networks in eukaryotic cells remains
tion–dissipation relationship on the overall cellular- poorly understood, although theoretical studies hint
scale transport of molecules and organelles remain at their qualitative as well as quantitative importance
unclear. [555, 556].
With regards to motor-driven transport, our Ultimately, unraveling the biological conse-
understanding of what controls processive run- quences of transport, its defects, and its regulation,
lengths, pausing, directionality, and track selection in will require synthesizing our understanding of
vivo remains incomplete. One of the key unanswered multiple physical transport mechanisms with newly
questions is the extent to which cargo sorting and emerging data on patterns of motion within living
distribution by motor-driven transport is locally self- cells.
organized [469] versus guided by external signals such
as pre-existing spatial heterogeneity in, e.g., adaptor-
binding signaling factors or microtubule-associated Acknowledgments
proteins [188, 435]. Recent live-cell measurements
have begun to identify the role of microtubule inter- We thank Matthias Weiss, Laura Westrate, Christo-
sections in pauses, reversals, and directional switches pher Obara, and Jenna Christensen for sharing data
of moving cargos [2, 445]. However, the contribu- prior to publication. This work was supported in
tion of other factors in regulating processivity in vivo part by funding from the NSF CAREER grant PHY-
remains unclear. Furthermore, the factors that control 1848057, the Hellman Fellows Fund, and the Alfred
the particular set of motors recruited to a given cargo, P Sloan Foundation, as well as a predoctoral fellow-
the interaction of those motors under in vivo condi- ship to SSM from the Visible Molecular Cell Consor-
tions, and the consequences of motor interactions on tium / Center for Trans-scale Structural Biology and
cellular-scale cargo delivery remain topics of ongoing Biophysics.
research.
The role of fluid flows in driving intracellular ORCID iDs
transport and mixing is beginning to be appreciated
in a widening variety of cell types. While rapid, exten- Saurabh S Mogre https://orcid.org/0000-0002-
sive flows in plant cells, fungi, and slime molds have 3781-5161
been the target of extensive study, the contribution Aidan I Brown https://orcid.org/0000-0002-6600-
of more modest flows in animal cells is now begin- 8289
ning to be unraveled. Cytoplasmic flows help drive Elena F Koslover https://orcid.org/0000-0003-
the segregation of subcellular components in devel- 4139-9209
opment [508], generate gradients that establish cell
polarity [506], and may enhance the mixing and dis-
persion of molecules and vesicular organelles [3, 107]. References
A potential role for flow in driving mixing within
[1] Data from personal communication with Christopher
reticulated organelles has also been recently pro- Obara, Janelia Research Campus.
posed [59]. Primary outstanding questions include [2] Bálint Š, Vilanova I V, Álvarez Á S and Lakadamyali M
the extent to which cells can control flow patterns to 2013 Correlative live-cell and superresolution microscopy
regulate advective transport and the importance of reveals cargo transport dynamics at microtubule
intersections Proc. Natl Acad. Sci. 110 3375–80
flow relative to other mechanisms for specific trans- [3] Koslover E F, Chan C K and Theriot J A 2017 Cytoplasmic
port systems. flow and mixing due to deformation of motile cells
An overarching question of key biological impor- Biophys. J. 113 2077–87
tance is how to draw a quantitative connection [4] Aridor M and Hannan L A 2000 Traffic jam: a
compendium of human diseases that affect intracellular
between our understanding of transport (i.e.: speed, transport processes Traffic 1 836–51
directional bias, processivity of particle movement) [5] Aridor M and Hannan L A 2002 Traffic jams ii: an update
and the kinetics of reactions between cellular com- of diseases of intracellular transport Traffic 3 781–90

33
Phys. Biol. 17 (2020) 061003 Topical Review

[6] Liu X-A, Rizzo V and Puthanveettil S 2012 Pathologies of [30] Godec A and Metzler R 2015 Signal focusing through
axonal transport in neurodegenerative diseases J. Transl. active transport Phys. Rev. E 92 10701
Neurosci. 3 355–72 [31] Godec A and Metzler R 2016 Active transport improves the
[7] Millecamps S and Julien J-P 2013 Axonal transport deficits precision of linear long distance molecular signalling J.
and neurodegenerative diseases Nat. Rev. Neurosci. 14 161 Phys. A: Math. Theor. 49 364001
[8] Smith B N et al 2017 Mutations in the vesicular trafficking [32] Goldstein R E and van de Meent J-W 2015 A physical
protein annexin a11 are associated with amyotrophic perspective on cytoplasmic streaming Interface Focus 5
lateral sclerosis Sci. Trans. Med. 9 eaad9157 20150030
[9] Liao Y-C et al 2019 Rna granules hitchhike on lysosomes [33] Doi M and Edwards S F 1988 The Theory of Polymer
for long-distance transport, using annexin a11 as a Dynamics vol 73 (Oxford: Oxford University Press)
molecular tether Cell 179 147–64 [34] Bandyopadhyay D, Cyphersmith A, Zapata J A, Joseph
[10] Goetz S C and Anderson K V 2010 The primary cilium: a Kim Y and Payne C K 2014 Lysosome Transport as a
signalling centre during vertebrate development Nat. Rev. Function of Lysosome Diameter PloS One 9 e86847
Genet. 11 331 [35] Metzler R and Klafter J 2000 The random walk’s guide to
[11] Pennarun G et al 1999 Loss-of-function mutations in a anomalous diffusion: a fractional dynamics approach Phys.
human gene related to chlamydomonas reinhardtii dynein Rep. 339 1–77
ic78 result in primary ciliary dyskinesia Am. J. Hum. Genet. [36] Weber S C, Thompson M A, Moerner W E, Spakowitz A J
65 1508–19 and Theriot J A 2012 Analytical tools to distinguish the
[12] Radtke K, Döhner K and Sodeik B 2006 Viral interactions effects of localization error, confinement, and medium
with the cytoskeleton: a hitchhiker’s guide to the cell Cell. elasticity on the velocity autocorrelation function Biophys.
Microbiol. 8 387–400 J. 102 2443–50
[13] Blondot M-L, Bruss V and Kann M 2016 Intracellular [37] Weber S C, Theriot J A and Spakowitz A J 2010
transport and egress of hepatitis b virus J. Hepatol. 64 Subdiffusive motion of a polymer composed of
S49–59 subdiffusive monomers Phys. Rev. E 82 011913
[14] Carnes S K and Aiken C 2019 Host proteins involved in [38] Valentine M T, Kaplan P D, Thota D, Crocker J C, Gisler T,
microtubule-dependent hiv-1 intracellular transport and Prud’homme R K, Beck M and Weitz D A 2001
uncoating Future Virol. 14 361–74 Investigating the microenvironments of inhomogeneous
[15] Shivanna V, Kim Y and Chang K-O 2014 Endosomal soft materials with multiple particle tracking Phys. Rev. E
acidification and cathepsin l activity is required for 64 061506
calicivirus replication Virol. 464 287–95 [39] Yamada S, Wirtz D and Kuo S C 2000 Mechanics of living
[16] Mingo R M, Simmons J A, Shoemaker C J, Nelson E A, cells measured by laser tracking microrheology Biophys. J.
Schornberg K L, D’Souza R S, Casanova J E and White J M 78 1736–47
2015 Ebola virus and severe acute respiratory syndrome [40] Tolić-Nørrelykke I M, Munteanu E-L, Thon G,
coronavirus display late cell entry kinetics: evidence that Oddershede L and Berg-Sørensen K 2004 Anomalous
transport to NPC1+endolysosomes is a rate-defining step diffusion in living yeast cells Phys. Rev. Lett. 93 078102
J. Virol. 89 2931–43 [41] Hoffman B D, Massiera G, Van Citters K M and Crocker J
[17] Wang H, Yang P, Liu K, Guo F, Zhang Y, Zhang G and C 2006 The consensus mechanics of cultured mammalian
Jiang C 2008 SARS coronavirus entry into host cells cells Proc. Natl Acad. Sci. 103 10259–64
through a novel clathrin- and caveolae-independent [42] Thompson D 1992 On Growth and Form ed. Bonner J
endocytic pathway Cell Res. 18 290–301 (Cambridge: Cambridge University Press)
[18] Vale R D 2003 The molecular motor toolbox for [43] Sanderson Haldane J B 1926 On Being the Right Size (New
intracellular transport Cell 112 467–80 York City, NY: Harper & Brothers)
[19] Hancock W O 2014 Bidirectional cargo transport: moving [44] Soh S, Banaszak M, Kandere-Grzybowska K and
beyond tug of war Nat. Rev. Mol. Cell Biol. 15 615 Grzybowski B A 2013 Why cells are microscopic: a
[20] Fu M-m and Holzbaur E L F 2014 Integrated regulation of transport-time perspective J. Phys. Chem. Lett. 4 861–5
motor-driven organelle transport by scaffolding proteins [45] Levin P A and Angert E R 2015 Small but mighty: cell size
Trends Cell Biol. 24 564–74 and bacteria Cold Spring Harbor Perspect. Biol. 7 a019216
[21] Happel J and Brenner H 1983 Low Reynolds Number [46] Riley M R, Muzzio F J and Reyes S C 1999 Experimental
Hydrodynamics: With Special Applications to Particulate and modeling studies of diffusion in immobilized cell
Media vol 1 (Berlin: Springer) systems: a review of recent literature and patents Appl.
[22] Purcell E M 1977 Life at low Reynolds number Am. J. Phys. Biochem. Biotechnol. 80 151–88
45 3–11 [47] Vendelin M, Kongas O and Saks V 2000 Regulation of
[23] Phillips R, Kondev J, Theriot J and Garcia H 2012 Physical mitochondrial respiration in heart cells analyzed by
Biology of the Cell (New York: Garland Science) reaction-diffusion model of energy transfer Am. J. Physiol.:
[24] Berg H C 1993 Random Walks in Biology (Princeton, NJ: Cell Physiol. 278 C747–64
Princeton University Press) [48] Misgeld T and Schwarz T L 2017 Mitostasis in neurons:
[25] Mogre S S and Koslover E F 2018 Multimodal transport maintaining mitochondria in an extended cellular
and dispersion of organelles in narrow tubular cells Phys. architecture Neuron 96 651–66
Rev. E 97 042402 [49] Bar-Even A, Noor E, Savir Y, Liebermeister W, Davidi D,
[26] Lin C, Schuster M, Guimaraes S C, Ashwin P, Schrader M, Tawfik D S and Milo R 2011 The moderately efficient
Metz J, Hacker C, Jane Gurr S and Steinberg G 2016 Active enzyme: evolutionary and physicochemical trends shaping
diffusion and microtubule-based transport oppose myosin enzyme parameters Biochemistry 50 4402–10
forces to position organelles in cells Nat. Commun. 7 11814 [50] Image from Personal Communication with Laura
[27] Koslover E F, Chan C K and Theriot J A 2016 Westrate, Calvin University.
Disentangling random motion and flow in a complex [51] Spaulding E L and Burgess R W 2017 Accumulating
medium Biophys. J. 110 700–9 evidence for axonal translation in neuronal homeostasis
[28] Alim K, Andrew N, Pringle A and Brenner M P 2017 Front. Neurosci. 11 312
Mechanism of signal propagation in physarum [52] Horton A C and Ehlers M D 2003 Dual modes of
polycephalum Proc. Natl Acad. Sci. 114 5136–41 endoplasmic reticulum-to-golgi transport in dendrites
[29] Goldstein R E, Tuval I and van de Meent J-W 2008 revealed by live-cell imaging J. Neurosci. 23 6188–99
Microfluidics of cytoplasmic streaming and its [53] Alberts B, Johnson A, Lewis J, Raff M, Roberts K and
implications for intracellular transport Proc. Natl Acad. Sci. Walter P 2007 Molecular Biology of the Cell (New York:
USA 105 3663–7 Garland Science)

34
Phys. Biol. 17 (2020) 061003 Topical Review

[54] Schwarz D S and Blower M D 2016 The endoplasmic [77] Glick B S and Luini A 2011 Models for golgi traffic: a
reticulum: structure, function and response to cellular critical assessment Cold Spring Harbor Perspect. Biol. 3
signaling Cell. Mol. Life Sci. 73 79–94 a005215
[55] Borgese N 2016 Getting membrane proteins on and off the [78] Orci L, Ravazzola M, Volchuk A, Engel T, Gmachl M,
shuttle bus between the endoplasmic reticulum and the Amherdt M, Perrelet A, Söllner T H and Rothman J E 2000
golgi complex J. Cell Sci. 129 1537–45 Anterograde flow of cargo across the Golgi stack
[56] Hammond A T and Glick B S 2000 Dynamics of potentially mediated via bidirectional “percolating” COPI
transitional endoplasmic reticulum sites in vertebrate cells vesicles Proc. Natl Acad. Sci. 97 10400–5
Mol. Biol. Cell 11 3013–30 original ERES characterization [79] Rizzo R, Parashuraman S, Mirabelli P, Puri C, Lucocq J
[57] Dayel M J, Hom E F Y and Verkman A S 1999 Diffusion of and Luini A 2013 The dynamics of engineered resident
green fluorescent protein in the aqueous-phase lumen of proteins in the mammalian golgi complex relies on
endoplasmic reticulum Biophys. J. 76 2843–51 cisternal maturation J. Cell Biol. 201 1027–36
[58] Runions J, Brach T, Kühner S and Hawes C 2005 [80] Patterson G H, Hirschberg K, Polishchuk R S, Gerlich D,
Photoactivation of gfp reveals protein dynamics within the Phair R D and Lippincott-Schwartz J 2008 Transport
endoplasmic reticulum membrane J. Exp. Bot. 57 through the golgi apparatus by rapid partitioning within a
43–50 two-phase membrane system Cell 133 1055–67
[59] Holcman D, Parutto P, Chambers J E, Fantham M, Young [81] Vagne Q and Sens P 2018 Stochastic model of vesicular
L J, Marciniak S J, Kaminski C F, Ron D and Avezov E 2018 sorting in cellular organelles Phys. Rev. Lett. 120 058102
Single particle trajectories reveal active endoplasmic [82] Antonietta De Matteis M and Luini A 2008 Exiting the
reticulum luminal flow Nat. Cell Biol. 20 1118 golgi complex Nat. Rev. Mol. Cell Biol. 9 273
[60] Watson P, Townley A K, Koka P, Palmer K J and Stephens [83] Polishchuk E V, Di Pentima A, Luini A and Polishchuk R S
D J 2006 Sec16 defines endoplasmic reticulum exit sites 2003 Mechanism of constitutive export from the golgi:
and is required for secretory cargo export in mammalian bulk flow via the formation, protrusion, and en bloc
cells Traffic 7 1678–87 cleavage of large trans-golgi network tubular domains Mol.
[61] Stephens D J 2003 De novo formation, fusion and fission Biol. Cell 14 4470–85
of mammalian COPII-coated endoplasmic reticulum exit [84] Polishchuk R S, Polishchuk E V, Marra P, Alberti S,
sites EMBO Rep. 4 210–7 Buccione R, Luini A and Mironov A A 2000 Correlative
[62] Barlowe C and Helenius A 2016 Cargo capture and bulk light-electron microscopy reveals the tubular-saccular
flow in the early secretory pathway Annu. Rev. Cell Dev. ultrastructure of carriers operating between golgi
Biol. 32 197–222 apparatus and plasma membrane J. Cell Biol. 148 45–58
[63] Appenzeller-Herzog C and Hauri H-P 2006 The er-golgi [85] Boncompain G et al 2012 Synchronization of secretory
intermediate compartment (ergic): in search of its identity protein traffic in populations of cells Nat. Methods 9 493
and function J. Cell Sci. 119 2173–83 [86] Drechsler M, Giavazzi F, Cerbino R and Palacios I M 2017
[64] Brandizzi F and Barlowe C 2013 Organization of the Active diffusion and advection in drosophila oocytes result
ER-Golgi interface for membrane traffic control Nat. Rev. from the interplay of actin and microtubules Nat.
Mol. Cell Biol. 14 382 Commun. 8 1520
[65] Presley J F, Cole N B, Schroer T A, Hirschberg K, Zaal K J [87] Kennedy M J and Ehlers M D 2006 Organelles and
M and Lippincott-Schwartz J 1997 Er-to-golgi transport trafficking machinery for postsynaptic plasticity Annu.
visualized in living cells Nature 389 81 Rev. Neurosci. 29 325–62
[66] Stephens D J and Pepperkok R 2001 Illuminating the [88] Akins M R, Berk-Rauch H E and Fallon J 2009 Presynaptic
secretory pathway: when do we need vesicles? J. Cell Sci. translation: stepping out of the postsynaptic shadow Front.
114 1053–9 Neural Circuits 3 17
[67] Ben-Tekaya H, Miura K, Pepperkok R and Hauri H-P 2005 [89] Black M M 2016 Axonal transport: the orderly motion of
Live imaging of bidirectional traffic from the ergic J. Cell axonal structures Methods in Cell Biology vol 131
Sci. 118 357–67 (Amsterdam: Elsevier) pp 1–19
[68] Rothman J E and Wieland F T 1996 Protein sorting by [90] Hanus C and Ehlers M D 2008 Secretory outposts for the
transport vesicles Science 272 227–34 local processing of membrane cargo in neuronal dendrites
[69] Dancourt J and Barlowe C 2010 Protein sorting receptors Traffic 9 1437–45
in the early secretory pathway Annu. Rev. Biochem. 79 [91] Twiss J L, Kalinski A L, Sachdeva R and Houle J D 2016
777–802 Intra-axonal protein synthesis–a new target for neural
[70] Malkus P, Jiang F and Schekman R 2002 Concentrative repair? Neural Regener. Res. 11 1365
sorting of secretory cargo proteins into copii-coated [92] Costa C J and Willis D E 2018 To the end of the line: axonal
vesicles J. Cell Biol. 159 915–21 mrna transport and local translation in health and
[71] Barlowe C 2003 Signals for COPII-dependent export from neurodegenerative disease Dev. Neurobiol. 78 209–20
the ER: what’s the ticket out? Trends Cell Biol. 13 [93] González C, Cornejo V H and Couve A 2018 Golgi bypass
295–300 for local delivery of axonal proteins, fact or fiction? Curr.
[72] Pagant S, Wu A, Edwards S, Diehl F and Miller E A 2015 Opin. Cell Biol. 53 9–14
Sec24 is a coincidence detector that simultaneously binds [94] Donnelly C J, Fainzilber M and Twiss J L 2010 Subcellular
two signals to drive er export Curr. Biol. 25 403–12 communication through rna transport and localized
[73] Thor F, Gautschi M, Geiger R and Helenius A 2009 Bulk protein synthesis Traffic 11 1498–505
flow revisited: transport of a soluble protein in the [95] Malkinson G and Spira M E 2010 Clustering of excess
secretory pathway Traffic 10 1819–30 growth resources within leading growth cones underlies
[74] Dukhovny A, Yaffe Y, Shepshelovitch J and Hirschberg K the recurrent “deposition” of varicosities along developing
2009 The length of cargo-protein transmembrane neurites Exp. Neurol. 225 140–53
segments drives secretory transport by facilitating cargo [96] Bradke F, Fawcett J W and Spira M E 2012 Assembly of a
concentration in export domains J. Cell Sci. 122 1759–67 new growth cone after axotomy: the precursor to axon
[75] Ronchi P, Colombo S, Francolini M and Borgese N 2008 regeneration Nat. Rev. Neurosci. 13 183
Transmembrane domain-dependent partitioning of [97] Vevea J D, Swayne T C, Boldogh I R and Pon L A 2014
membrane proteins within the endoplasmic reticulum J. Inheritance of the fittest mitochondria in yeast Trends Cell
Cell Biol. 181 105–18 Biol. 24 53–60
[76] Spang A 2013 Retrograde traffic from the golgi to the [98] Riquelme M et al 2018 Fungal morphogenesis, from the
endoplasmic reticulum Cold Spring Harbor Perspect. Biol. 5 polarized growth of hyphae to complex reproduction and
a013391 infection structures Microbiol. Mol. Biol. Rev. 82 e00068 17

35
Phys. Biol. 17 (2020) 061003 Topical Review

[99] Haglund K, Nezis I P and Stenmark H 2011 Structure and [123] Ross W N 2012 Understanding calcium waves and sparks
functions of stable intercellular bridges formed by in central neurons Nat. Rev. Neurosci. 13 157
incomplete cytokinesis during development Commun. [124] Rawlings J S, Rosler K M and Harrison D A 2004 The
Integr. Biol. 4 1–9 jak/stat signaling pathway J. Cell Sci. 117 1281–3
[100] Lowery L A and Van Vactor D 2009 The trip of the tip: [125] Carpenter G and Liao H-J 2013 Receptor tyrosine kinases
understanding the growth cone machinery Nat. Rev. Mol. in the nucleus Cold Spring Harbor Perspect. Biol. 5 a008979
Cell Biol. 10 332 [126] Walter P and Ron D 2011 The unfolded protein response:
[101] Dent E W and Gertler F B 2003 Cytoskeletal dynamics and from stress pathway to homeostatic regulation Science 334
transport in growth cone motility and axon guidance 1081–6
Neuron 40 209–27 [127] Kiuchi T, Nagai T, Ohashi K and Mizuno K 2011
[102] Lewis T L, Courchet J and Polleux F 2013 Cellular and Measurements of spatiotemporal changes in g-actin
molecular mechanisms underlying axon formation, concentration reveal its effect on stimulus-induced actin
growth, and branching J. Cell Biol. 202 837–48 assembly and lamellipodium extension J. Cell Biol. 193
[103] Suter D M and Miller K E 2011 The emerging role of forces 365–80
in axonal elongation Prog. Neurobiol. 94 91–101 [128] Sadovsky R G, Brielle S, Kaganovich D and England J L
[104] Brown A 2000 Slow axonal transport: stop and go traffic in 2017 Measurement of rapid protein diffusion in the
the axon Nat. Rev. Mol. Cell Biol. 1 153 cytoplasm by photo-converted intensity profile expansion
[105] Miller K E and Heidemann S R 2008 What is slow axonal Cell Rep. 18 2795–806
transport? Exp. Cell Res. 314 1981–90 [129] Verkman A S 2002 Solute and macromolecule diffusion in
[106] Scott D A, Das U, Tang Y and Roy S 2011 Mechanistic logic cellular aqueous compartments Trends Biochem. Sci. 27
underlying the axonal transport of cytosolic proteins 27–33
Neuron 70 441–54 [130] Kholodenko B N 2002 Map kinase cascade signaling and
[107] Mussel M, Zeevy K, Diamant H and Nevo U 2014 Drag of endocytic trafficking: a marriage of convenience? Trends
the cytosol as a transport mechanism in neurons Biophys. Cell Biol. 12 173–7
J. 106 2710–9 [131] Kholodenko B N 2006 Cell-signalling dynamics in time
[108] Avasthi P and Marshall W F 2012 Stages of ciliogenesis and and space Nat. Rev. Mol. Cell Biol. 7 165
regulation of ciliary length Differentiation 83 S30–42 [132] Harrington A W and Ginty D D 2013 Long-distance
[109] Lechtreck K F 2015 IFT-Cargo interactions and protein retrograde neurotrophic factor signalling in neurons Nat.
transport in cilia Trends Biochem. Sci. 40 765–78 Rev. Neurosci. 14 177
[110] Marshall W F and Rosenbaum J L 2001 Intraflagellar [133] Ibáñez C F 2007 Message in a bottle: long-range retrograde
transport balances continuous turnover of outer doublet signaling in the nervous system Trends Cell Biol. 17 519–28
microtubules J. Cell Biol. 155 405–14 [134] Chowdary P D, Che D L and Cui B 2012 Neurotrophin
[111] Engel B D, Ludington W B and Marshall W F 2009 signaling via long-distance axonal transport Annu. Rev.
Intraflagellar transport particle size scales inversely with Phys. Chem. 63 571–94
flagellar length: revisiting the balance-point length control [135] Sorkin A and Von Zastrow M 2009 Endocytosis and
model J. Cell Biol. 187 81–9 signalling: intertwining molecular networks Nat. Rev. Mol.
[112] Lin A C and Holt C E 2008 Function and regulation of Cell Biol. 10 609
local axonal translation Curr. Opin. Neurobiol. 18 [136] Hata A and Chen Y-G 2016 TGF-β signaling from
60–8 receptors to smads Cold Spring Harbor Perspect. Biol. 8
[113] Erez H, Malkinson G, Prager-Khoutorsky M, De Zeeuw C a022061
I, Hoogenraad C C and Spira M E 2007 Formation of [137] Bakker J, Spits M, Neefjes J and Berlin I 2017 The EGFR
microtubule-based traps controls the sorting and odyssey - from activation to destruction in space and time
concentration of vesicles to restricted sites of regenerating J. Cell Sci. 130 4087–96
neurons after axotomy J. Cell Biol. 176 497–507 [138] York H, Kaur A, Patil A, Bhowmik A, Moorthi U K, Hyde
[114] Fishman H M and Bittner G D 2003 Vesicle-mediated G J, Gandhi H, Gaus K and Arumugam S 2018 Rapid
restoration of a plasmalemmal barrier in severed axons whole cell imaging reveals an appl1-dynein nexus that
Physiology 18 115–8 regulates stimulated egfr trafficking bioRxiv:481796
[115] Steinberg G, Harmer N J, Schuster M and Kilaru S 2017 [139] Miaczynska M, Pelkmans L and Zerial M 2004 Not just a
Woronin body-based sealing of septal pores Fungal Genet. sink: endosomes in control of signal transduction Curr.
Biol. 109 53–5 Opin. Cell Biol. 16 400–6
[116] Mishra P and Chan D C 2014 Mitochondrial dynamics and [140] Neefjes J, Jongsma M M L and Berlin I 2017 Stop or go?
inheritance during cell division, development and disease endosome positioning in the establishment of
Nat. Rev. Mol. Cell Biol. 15 634 compartment architecture, dynamics, and function Trends
[117] Cagalinec M, Safiulina D, Liiv M, Liiv J, Choubey V, Cell Biol. 27 580–94
Wareski P, Veksler V and Kaasik A 2013 Principles of the [141] Nachury M V 2014 How do cilia organize signalling
mitochondrial fusion and fission cycle in neurons J. Cell cascades? Phil. Trans. R. Soc. B 369 20130465
Sci. 126 2187–97 [142] Mourao A, Christensen S T and Lorentzen E 2016 The
[118] Chen H, Chomyn A and Chan D C 2005 Disruption of intraflagellar transport machinery in ciliary signaling Curr.
fusion results in mitochondrial heterogeneity and Opin. Struct. Biol. 41 98–108
dysfunction J. Biol. Chem. 280 26185–92 [143] Nachury M V 2018 The molecular machines that traffic
[119] Chen H, Vermulst M, Wang Y E, Chomyn A, Prolla T A, signaling receptors into and out of cilia Curr. Opin. Cell
McCaffery J M and Chan D C 2010 Mitochondrial fusion Biol. 51 124–31
is required for mtdna stability in skeletal muscle and [144] Eguether T, Cordelieres F P and Pazour G J 2018
tolerance of mtdna mutations Cell 141 280–9 Intraflagellar transport is deeply integrated in hedgehog
[120] Knoblach B and Rachubinski R A 2016 How peroxisomes signaling Mol. Biol. Cell 29 1178–89
partition between cells. a story of yeast, mammals and [145] Kim J, Kato M and Beachy P A 2009 Gli2 trafficking links
filamentous fungi Curr. Opin. Cell Biol. 41 73–80 hedgehog-dependent activation of smoothened in the
[121] Chang J B and Ferrell J E Jr 2013 Mitotic trigger waves and primary cilium to transcriptional activation in the nucleus
the spatial coordination of the xenopus cell cycle Nature Proc. Natl Acad. Sci. 106 21666–71
500 603 [146] Mukhopadhyay S and Rohatgi R 2014 G-protein-coupled
[122] Gelens L, Anderson G A and Ferrell J E Jr 2014 Spatial receptors, hedgehog signaling and primary cilia Seminars
trigger waves: positive feedback gets you a long way Mol. in Cell & Developmental Biology vol 33 (Amsterdam:
Biol. Cell 25 3486–93 Elsevier) pp 63–72

36
Phys. Biol. 17 (2020) 061003 Topical Review

[147] Eguether T et al 2014 Ift27 links the bbsome to ift for [169] Abrisch R G, Gumbin S C, Wisniewski B T, Lackner L L
maintenance of the ciliary signaling compartment Dev. and Voeltz G K 2020 Fission and fusion machineries
Cell 31 279–90 converge at er contact sites to regulate mitochondrial
[148] Datta P, Allamargot C, Hudson J S, Andersen E K, morphology J. Cell Biol. 219 e201911122
Bhattarai S, Drack A V, Sheffield V C and Seo S 2015 [170] Moore A S and Holzbaur E L F 2018
Accumulation of non-outer segment proteins in the outer Mitochondrial-cytoskeletal interactions: dynamic
segment underlies photoreceptor degeneration in associations that facilitate network function and
Bardet-Biedl syndrome Proc. Natl Acad. Sci. 112 E4400–9 remodeling Curr. Opinion Physiol. 3 94–100
[149] Valm A M et al 2017 Applying systems-level spectral [171] Pekkurnaz G, Trinidad J C, Wang X, Kong D and Schwarz
imaging and analysis to reveal the organelle interactome T L 2014 Glucose regulates mitochondrial motility via
Nature 546 162–7 milton modification by o-glcnac transferase Cell 158
[150] Cohen S, Valm A M and Lippincott-Schwartz J 2018 54–68
Interacting organelles Curr. Opin. Cell Biol. 53 84–91 [172] Ferreira J M, Burnett A L and Rameau G A 2011
[151] Gatta A T and Levine T P 2017 Piecing together the Activity-dependent regulation of surface glucose
patchwork of contact sites Trends Cell Biol. 27 214–29 transporter-3 J. Neurosci. 31 1991–9
[152] Prinz W A, Toulmay A and Balla T 2019 The functional [173] Pathak D et al 2015 The role of mitochondrially derived atp
universe of membrane contact sites Nat. Rev. Mol. Cell in synaptic vesicle recycling J. Biol. Chem. 290 22325–36
Biol. 21 7–24 [174] Aw T Y and Jones D P 1985 Atp concentration gradients in
[153] Phillips M J and Voeltz G K 2016 Structure and function of cytosol of liver cells during hypoxia Am. J. Physiol.: Cell
er membrane contact sites with other organelles Nat. Rev. Physiol. 249 C385–92
Mol. Cell Biol. 17 69 [175] Jones D P t 1986 Intracellular diffusion gradients of o2 and
[154] Daniel Barbosa A and Siniossoglou S 2017 Function of atp Am. J. Physiol.: Cell Physiol. 250 C663–75
lipid droplet-organelle interactions in lipid homeostasis [176] Niethammer P, Kueh H Y and Mitchison T J 2008 Spatial
Biochim. Biophys. Acta, Mol. Cell Res. 1864 1459–68 patterning of metabolism by mitochondria, oxygen, and
[155] Settembre C, Fraldi A, Medina D L and Ballabio A 2013 energy sinks in a model cytoplasm Curr. Biol. 18 586–91
Signals from the lysosome: a control centre for cellular [177] Zecchin A, Stapor P C, Goveia J and Carmeliet P 2015
clearance and energy metabolism Nat. Rev. Mol. Cell Biol. Metabolic pathway compartmentalization: an
14 283–96 underappreciated opportunity? Curr. Opin. Biotechnol. 34
[156] Shai N, Schuldiner M and Zalckvar E 2016 No peroxisome 73–81
is an island - Peroxisome contact sites Biochim. Biophys. [178] Agapakis C M, Boyle P M and Silver P A 2012 Natural
Acta, Mol. Cell Res. 1863 1061–9 strategies for the spatial optimization of metabolism in
[157] Pu J, Ha C W, Zhang S, Jung J P, Huh W-K and Liu P 2011 synthetic biology Nat. Chem. Biol. 8 527
Interactomic study on interaction between lipid droplets [179] Weisová P, Concannon C G, Devocelle M, Prehn J H M
and mitochondria Protein Cell 2 487–96 and Ward M W 2009 Regulation of glucose transporter 3
[158] Rambold A S, Cohen S and Lippincott-Schwartz J 2015 surface expression by the amp-activated protein kinase
Fatty acid trafficking in starved cells: regulation by lipid mediates tolerance to glutamate excitation in neurons J.
droplet lipolysis, autophagy, and mitochondrial fusion Neurosci. 29 2997–3008
dynamics Dev. Cell 32 678–92 [180] Hochachka P W 1999 The metabolic implications of
[159] Islinger M, Felipe Godinho L, Costello J and Schrader M intracellular circulation Proc. Natl Acad. Sci. 96
2015 The different facets of organelle interplay-an 12233–9
overview of organelle interactions Frontiers in Cell and [181] Rangaraju V, Calloway N and Ryan T A 2014
Developmental Biology 3 56 Activity-driven local atp synthesis is required for synaptic
[160] Woźniak M J, Bola B, Brownhill K, Yang Y-C, Levakova V function Cell 156 825–35
and Allan V J 2009 Role of kinesin-1 and cytoplasmic [182] Harris J J and Attwell D 2012 The energetics of cns white
dynein in endoplasmic reticulum movement in VERO cells matter J. Neurosci. 32 356–71
J. Cell Sci. 122 1979–89 [183] Ohno N, Kidd G J, Mahad D, Kiryu-Seo S, Avishai A,
[161] Friedman J R, Webster B M, Mastronarde D N, Verhey K J Komuro H and Trapp B D 2011 Myelination and axonal
and Voeltz G K 2010 ER sliding dynamics and electrical activity modulate the distribution and motility of
ER-mitochondrial contacts occur on acetylated mitochondria at cns nodes of ranvier J. Neurosci. 31
microtubules J. Cell Biol. 190 363–75 7249–58
[162] Herms A et al 2015 Ampk activation promotes lipid [184] Zhang C L, Ho P L, Kintner D B, Sun D and Chiu S Y 2010
droplet dispersion on detyrosinated microtubules to Activity-dependent regulation of mitochondrial motility
increase mitochondrial fatty acid oxidation Nat. Commun. by calcium and na/k-atpase at nodes of ranvier of
6 1–14 myelinated nerves J. Neurosci. 30 3555–66
[163] Zehmer J K, Huang Y, Peng G, Pu J, Anderson R G W and [185] Agrawal A, Pekkurnaz G and Koslover E F 2018 Spatial
Liu P 2009 A role for lipid droplets in inter-membrane control of neuronal metabolism through glucose-mediated
lipid traffic Proteomics 9 914–21 mitochondrial transport regulation eLife 7 e40986
[164] Schuldiner M and Bohnert M 2017 A different kind of love [186] Chang D T W and Reynolds I J 2006 Mitochondrial
- lipid droplet contact sites Biochim. Biophys. Acta, Mol. trafficking and morphology in healthy and injured
Cell Biol. Lipids 1862 1188–96 neurons Prog. Neurobiol. 80 241–68
[165] Schwarz T L 2013 Mitochondrial trafficking in neurons [187] Sheng Z-H 2014 Mitochondrial trafficking and anchoring
Cold Spring Harbor Perspect. Biol. 5 a011304 in neurons: new insight and implications J. Cell Biol. 204
[166] Maday S and Holzbaur E L F 2012 Autophagosome 1087–98
assembly and cargo capture in the distal axon Autophagy 8 [188] MacAskill A F and Kittler J T 2010 Control of
858–60 mitochondrial transport and localization in neurons
[167] Guimaraes S C, Schuster M, Bielska E, Dagdas G, Kilaru S, Trends Cell Biol. 20 102–12
Meadows B R A, Schrader M and Steinberg G 2015 [189] Amiri M and Hollenbeck P J 2008 Mitochondrial
Peroxisomes, lipid droplets, and endoplasmic reticulum biogenesis in the axons of vertebrate peripheral neurons
“hitchhike” on motile early endosomes J. Cell Biol. 211 Dev. Neurobiol. 68 1348–61
945–54 [190] Fricker M D, Lee J A, Bebber D P, Tlalka M, Hynes J,
[168] Salogiannis J and Reck-Peterson S L 2017 Hitchhiking: a Darrah P R, Watkinson S C and Boddy L 2008 Imaging
non-canonical mode of microtubule-based transport complex nutrient dynamics in mycelial networks J.
Trends Cell Biol. 27 141–50 Microsc. 231 317–31

37
Phys. Biol. 17 (2020) 061003 Topical Review

[191] Fricker M D, Heaton L L M, Jones N S and Boddy L 2017 high-speed single-molecule tracking of membrane
The mycelium as a network Microbiol. Spectrum 5 3 molecules Annu. Rev. Biophys. Biomol. Struct. 34 351–78
[192] Tero A, Takagi S, Saigusa T, Ito K, Bebber D P, Fricker M [213] Krapf D 2018 Compartmentalization of the plasma
D, Yumiki K, Kobayashi R and Nakagaki T 2010 Rules for membrane Curr. Opin. Cell Biol. 53 15–21
biologically inspired adaptive network design Science 327 [214] Daniels D R and Turner M S 2007 Diffusion on membrane
439–42 tubes: a highly discriminatory test of the
[193] Fessel A, Oettmeier C, Bernitt E, Gauthier N C and Saffman−Delbruck theory Langmuir 23 6667–70
Döbereiner H-G 2012 Physarum polycephalum [215] Antonny B 2011 Mechanisms of membrane curvature
percolation as a paradigm for topological phase transitions sensing Annu. Rev. Biochem. 80 101–23
in transportation networks Phys. Rev. Lett. 109 [216] Dumas F, Lebrun M C and Tocanne J-F 1999 Is the
078103 protein/lipid hydrophobic matching principle relevant to
[194] Alim K, Amselem G, Peaudecerf F, Brenner M P and membrane organization and functions? FEBS Lett. 458
Pringle A 2013 Random network peristalsis in physarum 271–7
polycephalum organizes fluid flows across an individual [217] Nielsen C, Goulian M and Andersen O S 1998 Energetics
Proc. Natl Acad. Sci. 110 13306–11 of inclusion-induced bilayer deformations Biophys. J. 74
[195] Marbach S, Alim K, Andrew N, Pringle A and Brenner M P 1966–83
2016 Pruning to increase taylor dispersion in physarum [218] Phillips R, Ursell T, Wiggins P and Sens P 2009 Emerging
polycephalum networks Phys. Rev. Lett. 117 178103 roles for lipids in shaping membrane-protein function
[196] Raven J A 2003 Long-distance transport in non-vascular Nature 459 379–85
plants Plant, Cell Environ. 26 73–85 [219] Haselwandter C A and Phillips R 2013 Directional
[197] Alim K 2018 Fluid flows shaping organism morphology interactions and cooperativity between mechanosensitive
Phil. Trans. R. Soc. B 373 20170112 membrane proteins Europhys. Lett. 101 68002
[198] Nenninger A, Mastroianni G and Mullineaux C W 2010 [220] Reister E and Seifert U 2005 Lateral diffusion of a protein
Size dependence of protein diffusion in the cytoplasm of on a fluctuating membrane Europhys. Lett. 71 859
escherichia coli J. Bacteriol. 192 4535–40 [221] Lippincott-Schwartz J and Phair R D 2010 Lipids and
[199] Lukacs G L, Haggie P, Seksek O, Lechardeur D, cholesterol as regulators of traffic in the endomembrane
Freedman N and Verkman A S 2000 Size-dependent dna system Annu. Rev. Biophys. 39 559–78
mobility in cytoplasm and nucleus J. Biol. Chem. 275 [222] Brandizzi F, Frangne N, Marc-Martin S, Hawes C, Neuhaus
1625–9 J-M and Paris N 2002 The destination for single-pass
[200] Wirtz D 2009 Particle-tracking microrheology of living membrane proteins is influenced markedly by the length of
cells: principles and applications Annu. Rev. Biophys. 38 the hydrophobic domain Plant Cell 14 1077–92
301–26 [223] Bretscher M S and Munro S 1993 Cholesterol and the golgi
[201] Itel F, Najer A, Palivan C G and Meier W 2015 Dynamics of apparatus Science 261 1280–1
membrane proteins within synthetic polymer membranes [224] Lundbaek J A, Andersen O S, Werge T and Nielsen C 2003
with large hydrophobic mismatch Nano Lett. 15 3871–8 Cholesterol-induced protein sorting: an analysis of
[202] Tan S, Tan H T and Chung M C M 2008 Membrane energetic feasibility Biophys. J. 84 2080–9
proteins and membrane proteomics Proteomics 8 3924–32 [225] Aimon S, Callan-Jones A, Berthaud A, Pinot M, Toombes
[203] Fujiwara T K et al 2016 Confined diffusion of G E S and Bassereau P 2014 Membrane shape modulates
transmembrane proteins and lipids induced by the same transmembrane protein distribution Dev. Cell 28 212–8
actin meshwork lining the plasma membrane Mol. Biol. [226] Simunovic M, Voth G A, Callan-Jones A and Bassereau P
Cell 27 1101–19 2015 When physics takes over: Bar proteins and membrane
[204] Lippincott-Schwartz J 2001 The secretory membrane curvature Trends Cell Biol. 25 780–92
system studied in real-time Histochem. Cell Biol. 116 [227] Huang K C and Ramamurthi K S 2010 Macromolecules
97–107 that prefer their membranes curvy Mol. Microbiol. 76
[205] Marušić-Paloka E 2001 On the stokes paradox for 822–32
power-law fluids ZAMM-Journal of Applied Mathematics [228] Lee J S H, Panorchan P, Hale C M, Khatau S B, Kole T P,
and Mechanics/Zeitschrift für Angewandte Mathematik und Tseng Y and Wirtz D 2006 Ballistic intracellular
Mechanik: Applied Mathematics and Mechanics 81 31–6 nanorheology reveals rock-hard cytoplasmic stiffening
[206] Saffman P G and Delbrück M 1975 Brownian motion in response to fluid flow J. Cell Sci. 119 1760–8
biological membranes Proc. Natl Acad. Sci. 72 3111–3 [229] Lampo T J, Stylianidou S, Backlund M P, Wiggins P A and
[207] Weiß K, Neef A, Van Q, Kramer S, Gregor I and Enderlein J Spakowitz A J 2017 Cytoplasmic rna-protein particles
2013 Quantifying the diffusion of membrane proteins and exhibit non-gaussian subdiffusive behavior Biophys. J. 112
peptides in black lipid membranes with 2-focus 532–42
fluorescence correlation spectroscopy Biophys. J. 105 [230] Sabri A, Xu X, Krapf D and Weiss M 2020 Elucidating the
455–62 origin of heterogeneous anomalous diffusion in the
[208] Block S 2018 Brownian motion at lipid membranes: a cytoplasm of mammalian cells Phys. Rev. Lett. 125 058101
comparison of hydrodynamic models describing and [231] Wang B, Anthony S M, Bae S C and Granick S 2009
experiments quantifying diffusion within lipid bilayers Anomalous yet brownian Proc. Natl Acad. Sci. 106 15160–4
Biomolecules 8 30 [232] Guo M, Ehrlicher A J, Jensen M H, Renz M, Moore J R,
[209] Gambin Y, Lopez-Esparza R, Reffay M, Sierecki E, Gov N Goldman R D, Lippincott-Schwartz J, Mackintosh F C and
S, Genest M, Hodges R S and Urbach W 2006 Lateral Weitz D A 2014 Probing the stochastic, motor-driven
mobility of proteins in liquid membranes revisited Proc. properties of the cytoplasm using force spectrum
Natl Acad. Sci. 103 2098–102 microscopy Cell 158 822–32
[210] Milo R and Phillips R 2015 Cell Biology by the Numbers [233] Wilhelm C 2008 Out-of-equilibrium microrheology inside
(New York: Garland Science) living cells Phys. Rev. Lett. 101 028101
[211] Domanov Y A, Aimon S, Toombes G E S, Renner M, [234] Weihs D, Mason T G and Teitell M A 2006
Quemeneur F, Triller A, Turner M S and Bassereau P 2011 Bio-microrheology: a frontier in microrheology Biophys. J.
Mobility in geometrically confined membranes Proc. Natl 91 4296–305
Acad. Sci. 108 12605–10 [235] Kollmannsberger P and Fabry B 2011 Linear and nonlinear
[212] Kusumi A, Nakada C, Ritchie K, Murase K, Suzuki K, rheology of living cells Annu. Rev. Mater. Res. 41 75–97
Murakoshi H, Kasai R S, Kondo J and Fujiwara T 2005 [236] Arcizet D, Meier B, Sackmann E, Rädler J O and Heinrich
Paradigm shift of the plasma membrane concept from the D 2008 Temporal analysis of active and passive transport in
two-dimensional continuum fluid to the partitioned fluid: living cells Phys. Rev. Lett. 101 248103

38
Phys. Biol. 17 (2020) 061003 Topical Review

[237] Guo M, Ehrlicher A J, Mahammad S, Fabich H, Jensen M High-resolution mapping of intracellular fluctuations
H, Moore J R, Fredberg J J, Goldman R D and Weitz D A using carbon nanotubes Science 344 1031–5
2013 The role of vimentin intermediate filaments in [260] Jee A-Y, Dutta S, Cho Y-K, Tlusty T and Granick S 2018
cortical and cytoplasmic mechanics Biophys. J. 105 Enzyme leaps fuel antichemotaxis Proc. Natl Acad. Sci. 115
1562–8 14–8
[238] Gardel M L, Shin J H, MacKintosh F C, Mahadevan L, [261] Illien P, Zhao X, Dey K K, Butler P J, Sen A and
Matsudaira P A and Weitz D A 2004 Scaling of f-actin Golestanian R 2017 Exothermicity is not a necessary
network rheology to probe single filament elasticity and condition for enhanced diffusion of enzymes Nano Lett. 17
dynamics Phys. Rev. Lett. 93 188102 4415–20
[239] Kou S C and Xie X S 2004 Generalized langevin equation [262] Xu M, Ross J L, Valdez L and Sen A 2019 Direct single
with fractional gaussian noise: subdiffusion within a single molecule imaging of enhanced enzyme diffusion Phys. Rev.
protein molecule Phys. Rev. Lett. 93 180603 Lett. 123 128101
[240] Zwanzig R 2001 Nonequilibrium Statistical Mechanics [263] Gallet F, Arcizet D, Bohec P and Richert A 2009 Power
(Oxford: Oxford University Press) spectrum of out-of-equilibrium forces in living cells:
[241] Lutz E 2001 Fractional langevin equation Phys. Rev. E 64 amplitude and frequency dependence Soft Matter 5
051106 2947–53
[242] Min W, Luo G, Cherayil B J, Kou S C and Xie X S 2005 [264] Luby-Phelps K 1999 Cytoarchitecture and physical
Observation of a power-law memory kernel for properties of cytoplasm: volume, viscosity, diffusion,
fluctuations within a single protein molecule Phys. Rev. intracellular surface area International Review of Cytology
Lett. 94 198302 vol 192 (Amsterdam: Elsevier) pp 189–221
[243] Bressloff P C and Newby J M 2013 Stochastic models of [265] Dix J A and Verkman A S 2008 Crowding effects on
intracellular transport Rev. Mod. Phys. 85 135–96 diffusion in solutions and cells Annu. Rev. Biophys. 37
[244] Deng W and Barkai E 2009 Ergodic properties of fractional 247–63
brownian-langevin motion Phys. Rev. E 79 011112 [266] Etoc F, Balloul E, Vicario C, Normanno D, Liße D,
[245] Mason T G and Weitz D A 1995 Optical measurements of Sittner A, Jacob P, Dahan M and Coppey M 2018
frequency-dependent linear viscoelastic moduli of Non-specific interactions govern cytosolic diffusion of
complex fluids Phys. Rev. Lett. 74 1250 nanosized objects in mammalian cells Nat. Mater. 17 740
[246] Lucas J S, Zhang Y, Dudko O K and Murre C 2014 3d [267] Fatin-Rouge N, Starchev K and Buffle J 2004 Size effects on
trajectories adopted by coding and regulatory dna diffusion processes within agarose gels Biophys. J. 86
elements: first-passage times for genomic interactions Cell 2710–9
158 339–52 [268] Cai L-H, Panyukov S and Rubinstein M 2015 Hopping
[247] Weber S C, Spakowitz A J and Theriot J A 2010 Bacterial diffusion of nanoparticles in polymer matrices
chromosomal loci move subdiffusively through a Macromolecules 48 847–62
viscoelastic cytoplasm Phys. Rev. Lett. 104 238102 [269] Heinemann F, Vogel S K and Schwille P 2013 Lateral
[248] Magdziarz M, Weron A, Burnecki K and Klafter J 2009 membrane diffusion modulated by a minimal actin cortex
Fractional brownian motion versus the continuous-time Biophys. J. 104 1465–75
random walk: a simple test for subdiffusive dynamics Phys. [270] Fodor É, Guo M, Gov N S, Visco P, Weitz D A and van
Rev. Lett. 103 180602 Wijland F 2015 Activity-driven fluctuations in living cells
[249] Gal N, Lechtman-Goldstein D and Weihs D 2013 Particle Europhys. Lett. 110 48005
tracking in living cells: a review of the mean square [271] He W, Song H, Su Y, Geng L, Ackerson B J, Peng H B and
displacement method and beyond Rheol. Acta 52 Tong P 2016 Dynamic heterogeneity and non-gaussian
425–43 statistics for acetylcholine receptors on live cell membrane
[250] Martin D S, Forstner M B and Käs J A 2002 Apparent Nat. Commun. 7 11701
subdiffusion inherent to single particle tracking Biophys. J. [272] Leptos K C, Guasto J S, Gollub J P, Pesci A I and Goldstein
83 2109–17 R E 2009 Dynamics of enhanced tracer diffusion in
[251] Weiss M 2013 Single-particle tracking data reveal suspensions of swimming eukaryotic microorganisms
anticorrelated fractional brownian motion in crowded Phys. Rev. Lett. 103 198103
fluids Phys. Rev. E 88 010101 [273] Scott Shell M, Debenedetti P G and Stillinger F H 2005
[252] Daumas F, Destainville N, Millot C, Lopez A, Dean D and Dynamic heterogeneity and non-gaussian behaviour in a
Salomé L 2003 Confined diffusion without fences of a model supercooled liquid J. Phys.: Condens. Matter. 17
g-protein-coupled receptor as revealed by single particle S4035
tracking Biophys. J. 84 356–66 [274] Xue C, Zheng X, Chen K, Tian Y and Hu G 2016 Probing
[253] Saxton M J 1996 Anomalous diffusion due to binding: a non-gaussianity in confined diffusion of nanoparticles J.
monte carlo study Biophys. J. 70 1250–62 Phys. Chem. Lett. 7 514–9
[254] Brangwynne C P, Koenderink G H, MacKintosh F C and [275] Wang B, Kuo J, Bae S C and Granick S 2012 When
Weitz D A 2009 Intracellular transport by active diffusion brownian diffusion is not gaussian Nat. Mater. 11 481
Trends Cell Biol. 19 423–7 [276] Chechkin A V, Seno F, Metzler R and Sokolov I M 2017
[255] Gupta S K and Guo M 2017 Equilibrium and Brownian yet non-gaussian diffusion: from superstatistics
out-of-equilibrium mechanics of living mammalian to subordination of diffusing diffusivities Phys. Rev. X 7
cytoplasm J. Mech. Phys. Solids 107 284–93 021002
[256] Smelser A M, Macosko J C, O’Dell A P, Smyre S, Bonin K [277] Duits M H G, Li Y, Vanapalli S A and Mugele F 2009
and Holzwarth G 2015 Mechanical properties of normal Mapping of spatiotemporal heterogeneous particle
versus cancerous breast cells Biomech. Model. Mechanobiol. dynamics in living cells Phys. Rev. E 79 051910
14 1335–47 [278] Li H, Dou S-X, Liu Y-R, Li W, Xie P, Wang W-C and Wang
[257] Weber S C, Spakowitz A J and Theriot J A 2012 P-Y 2015 Mapping intracellular diffusion distribution
Nonthermal atp-dependent fluctuations contribute to the using single quantum dot tracking: compartmentalized
in vivo motion of chromosomal loci Proc. Natl Acad. Sci. diffusion defined by endoplasmic reticulum J. Am. Chem.
109 7338–43 Soc. 137 436–44
[258] Posey D, Blaisdell-Pijuan P, Knoll S K, Saif T A and Ahmed [279] Chubynsky M V and Slater G W 2014 Diffusing diffusivity:
W W 2018 Small-scale displacement fluctuations of a model for anomalous, yet brownian, diffusion Phys. Rev.
vesicles in fibroblasts Sci. Rep. 8 1–9 Lett. 113 098302
[259] Fakhri N, Wessel A D, Willms C, Pasquali M, Klopfenstein [280] Metzler R 2020 Superstatistics and non-gaussian diffusion
D R, MacKintosh F C and Schmidt C F 2014 Eur. Phys. J.: Spec. Top. 229 711–28

39
Phys. Biol. 17 (2020) 061003 Topical Review

[281] Brown A I, Westrate L M and Koslover E F 2020 Impact of stress independently of the unfolded protein response J.
global structure on diffusive exploration of organelle Cell Biol. 187 525–36
networks Sci. Rep. 10 1–13 [302] McCarron J G, Wilson C, Sandison M E, Olson M L,
[282] Schroeder L K, Barentine A E S, Merta H, Schweighofer S, Girkin J M, Saunter C and Chalmers S 2013 From
Zhang Y, Baddeley D, Bewersdorf J and Bahmanyar S 2019 structure to function: mitochondrial morphology, motion
Dynamic nanoscale morphology of the er surveyed by sted and shaping in vascular smooth muscle J. Vasc. Res. 50
microscopy J. Cell Biol. 218 83–96 357–71
[283] Terasaki M et al 2013 Stacked endoplasmic reticulum [303] Long Q, Zhao D, Fan W, Yang L, Zhou Y, Qi J, Wang X
sheets are connected by helicoidal membrane motifs Cell and Liu X 2015 Modeling of mitochondrial donut
154 285–96 formation Biophys. J. 109 892–9
[284] Ježek P and Dlasková A 2019 Dynamics of mitochondrial [304] Schrader M, King S J, Stroh T A and Trina A Schroer 2000
network, cristae, and mitochondrial nucleoids in Real time imaging reveals a peroxisomal reticulum in
pancreatic β-cells Mitochondrion 49 245–58 living cells J. Cell Sci. 113 3663–71
[285] Nickell S, Park P S-H, Baumeister W and Palczewski K [305] Barton K A, Mathur N and Mathur J 2013 Simultaneous
2007 Three-dimensional architecture of murine rod outer live-imaging of peroxisomes and the er in plant cells
segments determined by cryoelectron tomography J. Cell suggests contiguity but no luminal continuity between the
Biol. 177 917–25 two organelles Frontiers in Physiology 4 196
[286] Tønnesen J and Valentin Nägerl U 2016 Dendritic spines [306] Chan Y-K, Tsai M-H, Huang D-C, Zheng Z-H and Hung
as tunable regulators of synaptic signals Frontiers K-D 2010 Leukocyte nucleus segmentation and nucleus
Psychiatry 7 101 lobe counting BMC Bioinf. 11 558
[287] Adler J, Sintorn I-M, Strand R and Parmryd I 2019 [307] Praefcke G J K and McMahon H T 2004 The dynamin
Conventional analysis of movement on non-flat surfaces superfamily: universal membrane tubulation and fission
like the plasma membrane makes brownian motion appear molecules? Nat. Rev. Mol. Cell Biol. 5 133–47
anomalous Commun. Biol. 2 1–10 [308] Schrader M, Bonekamp N A and Islinger M 2012 Fission
[288] Zamponi N, Zamponi E, Cannas S A, Billoni O V, and proliferation of peroxisomes Biochim. Biophys. Acta,
Helguera P R and Chialvo D R 2018 Mitochondrial Mol. Basis Dis. 1822 1343–57
network complexity emerges from fission/fusion dynamics [309] Rodrı́guez-Serrano M, Romero-Puertas M C,
Sci. Rep. 8 1–10 Sanz-Fernández M, Hu J and Sandalio L M 2016
[289] Viana M P, Brown A I, Mueller I A, Goul C, Koslover E F Peroxisomes extend peroxules in a fast response to stress
and Rafelski S M 2020 Mitochondrial fission and fusion via a reactive oxygen species-mediated induction of the
dynamics generate efficient, robust, and evenly distributed peroxin pex11a Plant Physiol. 171 1665–74
network topologies in budding yeast cells Cell Systems 10 [310] Sundborger A C, Fang S, Heymann J A, Ray P, Chappie J S
287–97 and Hinshaw J E 2014 A dynamin mutant defines a
[290] Westrate L M, Lee J E, Prinz W A and Voeltz G K 2015 superconstricted prefission state Cell Rep. 8
Form follows function: the importance of endoplasmic 734–42
reticulum shape Annu. Rev. Biochem. 84 791–811 [311] Klaus C J S, Raghunathan K, DiBenedetto E and
[291] Peng J-Y et al 2011 Automatic morphological subtyping Kenworthy A K 2016 Analysis of diffusion in curved
reveals new roles of caspases in mitochondrial dynamics surfaces and its application to tubular membranes Mol.
Plos Comput. Biol. 7 e1002212 Biol. Cell 27 3937–46
[292] Wang S, Tukachinsky H, Romano F B and Rapoport T A [312] Kusters R, Paquay S and Storm C 2015 Confinement
2016 Cooperation of the er-shaping proteins atlastin, without boundaries: anisotropic diffusion on the surface of
lunapark, and reticulons to generate a tubular membrane a cylinder Soft Matter 11 1054–7
network eLife 5 e18605 [313] Reguera D and Rubi J M 2001 Kinetic equations for
[293] Dickens J A et al 2016 The endoplasmic reticulum remains diffusion in the presence of entropic barriers Phys. Rev. E
functionally connected by vesicular transport after its 64 061106
fragmentation in cells expressing Z-α 1 -antitrypsin [314] Zwanzig R 1992 Diffusion past an entropy barrier J. Phys.
FASEB J. 30 4083–97 Chem. 96 3926–30
[294] Espadas J et al 2019 Dynamic constriction and fission of [315] Santamaria F, Wils S, De Schutter E and Augustine G J
endoplasmic reticulum membranes by reticulon Nat. 2006 Anomalous diffusion in purkinje cell dendrites
Commun. 10 1–11 caused by spines Neuron 52 635–48
[295] Kuznetsov A V, Hermann M, Saks V, Hengster P and [316] Sbalzarini I F, Mezzacasa A, Helenius A and Koumoutsakos
Margreiter R 2009 The cell-type specificity of P 2005 Effects of organelle shape on fluorescence recovery
mitochondrial dynamics Int. J. Biochem. Cell Biol. 41 after photobleaching Biophys. J. 89 1482–92
1928–39 [317] Sbalzarini I F, Hayer A, Helenius A and Koumoutsakos P
[296] Benard G and Rossignol R 2008 Ultrastructure of the 2006 Simulations of (an)isotropic diffusion on curved
mitochondrion and its bearing on function and biological surfaces Biophys. J. 90 878–85
bioenergetics Antioxid. Redox Signaling 10 1313–42 [318] Ben-Avraham D and Havlin S 2000 Diffusion and Reactions
[297] Patil N, Bonneau S, Joubert F, Bitbol A-F and in Fractals and Disordered Systems (Cambridge: Cambridge
Berthoumieux H 2019 Mitochondrial cristae modeled as University Press)
an out-of-equilibrium membrane driven by a proton field [319] Shen L and Chen Z 2007 Critical review of the impact of
(arXiv:1912.06373) tortuosity on diffusion Chem. Eng. Sci. 62 3748–55
[298] Martı́nez-Diez M, Santamarı́a G, Ortega Á D and Cuezva J [320] Havlin S and Ben-Avraham D 1987 Diffusion in
M 2006 Biogenesis and dynamics of mitochondria during disordered media Adv. Phys. 36 695–798
the cell cycle: significance of 3útrs PloS One 1 e107 [321] Havlin S, Ben-Avraham D and Sompolinsky H 1983
[299] Puhka M, Vihinen H, Joensuu M and Jokitalo E 2007 Scaling behavior of diffusion on percolation clusters Phys.
Endoplasmic reticulum remains continuous and Rev. A 27 1730
undergoes sheet-to-tubule transformation during cell [322] Stauffer D and Aharony A 2018 Introduction to Percolation
division in mammalian cells J. Cell Biol. 179 895–909 Theory (Boca Raton, FL: CRC Press)
[300] Koopman W J H, Willems P H G M and Jan A M S 2012 [323] Huber G and Wilkinson M 2019 Terasaki spiral ramps and
Monogenic mitochondrial disorders New Engl. J. Med. 366 intracellular diffusion Phys. Biol. 16 065002
1132–41 [324] Cogliati S, Enriquez J A and Scorrano L 2016
[301] Schuck S, Prinz W A, Thorn K S, Voss C and Walter P 2009 Mitochondrial cristae: where beauty meets functionality
Membrane expansion alleviates endoplasmic reticulum Trends Biochem. Sci. 41 261–73

40
Phys. Biol. 17 (2020) 061003 Topical Review

[325] Dieteren C E J, Gielen S C A M, Nijtmans L G J, Smeitink J [348] Lizana L and Konkoli Z 2005 Diffusive transport in
A M, Swarts H G, Brock R, Willems P H G M and networks built of containers and tubes Phys. Rev. E 72
Koopman W J H 2011 Solute diffusion is hindered in the 026305
mitochondrial matrix Proc. Natl Acad. Sci. 108 8657–62 [349] Moeendarbary E, Valon L, Fritzsche M, Harris A R,
[326] Ölveczky B P and Verkman A S 1998 Monte carlo analysis Moulding D A, Thrasher A J, Stride E, Mahadevan L and
of obstructed diffusion in three dimensions: application to Charras G T 2013 The cytoplasm of living cells behaves as
molecular diffusion in organelles Biophys. J. 74 2722–30 a poroelastic material Nat. Mater. 12 253–61
[327] Sukhorukov V M and Bereiter-Hahn J 2009 Anomalous [350] Berezhkovskii A M and Szabo A 2016 Theory of crowding
diffusion induced by cristae geometry in the inner effects on bimolecular reaction rates J. Phys. Chem. B 120
mitochondrial membrane PloS One 4 e4604 5998–6002
[328] Najafi M, Maza N A and Calvert P D 2012 Steric volume [351] Berg O G and von Hippel P H 1985 Diffusion-controlled
exclusion sets soluble protein concentrations in macromolecular interactions Annu. Rev. Biophys. Bioeng.
photoreceptor sensory cilia Proc. Natl Acad. Sci. 109 203–8 14 131–58
[329] Calvert P D, Schiesser W E and Pugh E N 2010 Diffusion [352] Masuda N, Porter M A and Lambiotte R 2017 Random
of a soluble protein, photoactivatable gfp, through a walks and diffusion on networks Phys. Rep. 716 1–58
sensory cilium J. Gen. Physiol. 135 173–96 [353] Barthélemy M 2011 Spatial networks Phys. Rep. 499 1–101
[330] Li R, Fowler J A and Todd B A 2014 Calculated rates of [354] Brown A I and Sivak D A 2019 Theory of nonequilibrium
diffusion-limited reactions in a three-dimensional network free energy transduction by molecular machines Chem.
of connected compartments: application to porous Rev. 120 434–59
catalysts and biological systems Phys. Rev. Lett. 113 028303 [355] Nan X, Sims P A and Xie X S 2008 Organelle tracking in a
[331] Park S H and Blackstone C 2010 Further assembly living cell with microsecond time resolution and
required: construction and dynamics of the endoplasmic nanometer spatial precision ChemPhysChem 9 707–12
reticulum network EMBO Rep. 11 515–21 [356] Ross J L, Ali M Y and Warshaw D M 2008 Cargo transport:
[332] Figure partially adapted from servier medical art, licensed molecular motors navigate a complex cytoskeleton Curr.
under a creative common attribution 3.0 generic license. Opin. Cell Biol. 20 41–7
http://smart.servier.com/ [357] Kikushima K, Kita S and Higuchi H 2013 A non-invasive
[333] Berg H C and Purcell E M 1977 Physics of chemoreception imaging for the in vivo tracking of high-speed vesicle
Biophys. J. 20 193–219 transport in mouse neutrophils Sci. Rep. 3 1913
[334] Li R and Todd B A 2015 Diffusion-limited encounter rate [358] Gagnon J A and Mowry K L 2011 Molecular motors:
in a three-dimensional lattice of connected compartments directing traffic during rna localization Crit. Rev. Biochem.
studied by brownian-dynamics simulations Phys. Rev. E 91 Mol. 46 229–39
032801 [359] Hirokawa N, Niwa S and Tanaka Y 2010 Molecular motors
[335] Bénichou O, Chevalier C, Klafter J, Meyer B and Voituriez in neurons: transport mechanisms and roles in brain
R 2010 Geometry-controlled kinetics Nat. Chem. 2 472 function, development, and disease Neuron 68 610–38
[336] De Gennes P G 1982 Kinetics of diffusion-controlled [360] Katz Z B, English B P, Lionnet T, Yoon Y J, Monnier N,
processes in dense polymer systems. I. Nonentangled Ovryn B, Bathe M and Singer R H 2016 Mapping
regimes J. Chem. Phys. 76 3316–21 translation ’hot-spots’ in live cells by tracking single
[337] Bénichou O and Voituriez R 2014 From first-passage times molecules of mRNA and ribosomes eLife 5 e10415
of random walks in confinement to geometry-controlled [361] Brown A I and Sivak D A 2017 Allocating dissipation
kinetics Phys. Rep. 539 225–84 across a molecular machine cycle to maximize flux Proc.
[338] Condamin S, Bénichou O, Tejedor V, Voituriez R and Natl Acad. Sci. 114 11057–62
Klafter J 2007 First-passage times in complex [362] Brown A I and Sivak D A 2018 Allocating and splitting free
scale-invariant media Nature 450 77–80 energy to maximize molecular machine flux J. Phys. Chem.
[339] Grebenkov D S, Metzler R and Oshanin G 2018 Strong B 122 1387–93
defocusing of molecular reaction times results from an [363] Gibbs K L, Greensmith L and Schiavo G 2015 Regulation
interplay of geometry and reaction control Commun. of axonal transport by protein kinases Trends Biochem. Sci.
Chem. 1 1–12 40 597–610
[340] Montroll E W and Weiss G H 1965 Random walks on [364] Burute M and Kapitein L C 2019 Cellular logistics:
lattices. ii J. Math. Phys. 6 167–81 unraveling the interplay between microtubule organization
[341] Hughes H et al 2009 Organisation of human ER-exit sites: and intracellular transport Annu. Rev. Cell Dev. Biol. 35
requirements for the localisation of sec16 to transitional er 29–54
J. Cell Sci. 122 2924–34 [365] Geitmann A and Nebenführ A 2015 Navigating the plant
[342] Ruhanen H, Borrie S, Szabadkai G, Tnismaa H, Jones A W cell: intracellular transport logistics in the green kingdom
E, Kang D, Taanman J-W and Yasukawa T 2010 Mol. Biol. Cell 26 3373–8
Mitochondrial single-stranded dna binding protein is [366] Snider J, Lin F, Zahedi N, Rodionov V, Yu C C and Gross S
required for maintenance of mitochondrial dna and 7s dna P 2004 Intracellular actin-based transport: How far you go
but is not required for mitochondrial nucleoid depends on how often you switch Proc. Natl Acad. Sci. USA
organisation Biochim. Biophys. Acta, Mol. Cell Res. 1803 101 13204–9
931–9 [367] Kapitein L C, van Bergeijk P, Lipka J, Keijzer N, Wulf P S,
[343] Oeffinger M and Zenklusen D 2012 To the pore and Katrukha E A, Akhmanova A and Hoogenraad C C 2013
through the pore: a story of mRNA export kinetics Myosin-v opposes microtubule-based cargo transport and
Biochim. Biophys. Acta, Gene Regul. Mech. 1819 494–506 drives directional motility on cortical actin Curr. Biol. 23
[344] Holcman D and Schuss Z 2004 Escape through a small 828–34
opening: receptor trafficking in a synaptic membrane J. [368] Hammer J A and Sellers J R 2012 Walking to work: roles
Stat. Phys. 117 975–1014 for class v myosins as cargo transporters Nat. Rev. Mol. Cell
[345] Schuss Z, Singer A and Holcman D 2007 The narrow Biol. 13 13–26
escape problem for diffusion in cellular microdomains [369] Szyk A, Deaconescu A M, Spector J, Goodman B,
Proc. Natl Acad. Sci. 104 16098–103 Valenstein M L, Ziolkowska N E, Kormendi V, Grigorieff N
[346] Singer A, Schuss Z, Holcman D and Eisenberg R S 2006 and Roll-Mecak A 2014 Molecular basis for age-dependent
Narrow escape, part i J. Stat. Phys. 122 437–63 microtubule acetylation by tubulin acetyltransferase Cell
[347] Grebenkov D S and Krapf D 2018 Steady-state reaction 157 1405–15
rate of diffusion-controlled reactions in sheets J. Chem. [370] Ross J L and Fygenson D K 2003 Mobility of taxol in
Phys. 149 064117 microtubule bundles Biophys. J. 84 3959–67

41
Phys. Biol. 17 (2020) 061003 Topical Review

[371] Brangwynne C P, MacKintosh F C and Weitz D A 2007 [392] Salogiannis J, Egan M J and Reck-Peterson S L 2016
Force fluctuations and polymerization dynamics of Peroxisomes move by hitchhiking on early endosomes
intracellular microtubules Proc. Natl Acad. Sci. 104 using the novel linker protein PxdA J. Cell Biol. 212
16128–33 201512020
[372] de Forges H, Bouissou A and Perez F 2012 Interplay [393] Mogre S S, Christensen J R, Niman C S, Reck-Peterson S L
between microtubule dynamics and intracellular and Koslover E F 2020 Hitching a ride: Mechanics of
organization Int. J. Biochem. Cell Biol. 44 266–74 transport initiation through linker-mediated hitchhiking
[373] Kulić I M, Brown A E X, Kim H, Kural C, Blehm B, Selvin Biophys. J. 118 1357–1369
P R, Nelson P C and Gelfand V I 2008 The role of [394] Wehnekamp F, Plucińska G, Thong R, Misgeld T and
microtubule movement in bidirectional organelle Lamb D C 2019 Nanoresolution real-time 3d orbital
transport Proc. Natl Acad. Sci. 105 10011–6 tracking for studying mitochondrial trafficking in
[374] Hirokawa N, Noda Y, Tanaka Y and Niwa S 2009 Kinesin vertebrate axons in vivo eLife 8 e46059
superfamily motor proteins and intracellular transport [395] Schuster M, Lipowsky R, Assmann M-A, Lenz P and
Nat. Rev. Mol. Cell Biol. 10 682–96 Steinberg G 2011 Transient binding of dynein controls
[375] Cianfrocco M A, DeSantis M E, Leschziner A E and bidirectional long-range motility of early endosomes Proc.
Reck-Peterson S L 2015 Mechanism and regulation of Natl Acad. Sci. 108 3618–23
cytoplasmic dynein Annu. Rev. Cell Dev. Biol. 31 [396] Unpublished lattice light sheet imaging data collected by S
83–108 Mogre, Jenna R Christensen (Reck-Peterson lab, University
[376] Reck-Peterson S L, Redwine W B, Vale R D and Carter A P of California San Diego), and Hiroyuki Hakozaki
2018 The cytoplasmic dynein transport machinery and its (National Center for Microscopy and Imaging Research,
many cargoes Nat. Rev. Mol. Cell Biol. 19 382 University of California San Diego). Cell culture and
[377] Ferro L S, Can S, Turner M A, ElShenawy M M and Yildiz imaging methods as described in [393].
A 2019 Kinesin and dynein use distinct mechanisms to [397] Maday S, Twelvetrees A E, Moughamian A J and Holzbaur
bypass obstacles eLife 8 e48629 E L F 2014 Axonal transport: cargo-specific mechanisms of
[378] DeWitt M A, Chang A Y, Combs P A and Yildiz A 2012 motility and regulation Neuron 84 292–309
Cytoplasmic dynein moves through uncoordinated [398] Nakazawa H, Sada T, Toriyama M, Tago K, Sugiura T,
stepping of the aaa+ ring domains Science 335 221–5 Fukuda M and Inagaki N 2012 Rab33a mediates
[379] Schnitzer M J, Visscher K and Block S M 2000 Force anterograde vesicular transport for membrane exocytosis
production by single kinesin motors Nat. Cell Biol. 2 and axon outgrowth J. Neurosci. 32 12712–25
718–23 [399] Maeder C I, San-Miguel A, Wu E Y, Lu H and Shen K 2014
[380] Gennerich A, Carter A P, Reck-Peterson S L and Vale R D In vivo neuron-wide analysis of synaptic vesicle precursor
2007 Force-induced bidirectional stepping of cytoplasmic trafficking Traffic 15 273–91
dynein Cell 131 952–65 [400] Cui B, Wu C, Chen L, Ramirez A, Bearer E L, Li W-P,
[381] Müller M J I, Klumpp S and Lipowsky R 2008 Tug-of-war Mobley W C and Chu S 2007 One at a time, live tracking
as a cooperative mechanism for bidirectional cargo of ngf axonal transport using quantum dots Proc. Natl
transport by molecular motors Proc. Natl Acad. Sci. 105 Acad. Sci. 104 13666–71
4609–14 [401] Welte M A 2004 Bidirectional transport along
[382] Hendricks A G, Perlson E, Ross J L, Schroeder H W, Tokito microtubules Curr. Biol. 14 R525–37
M and Holzbaur E L F 2010 Motor coordination via a [402] Wong M Y, Zhou C, Shakiryanova D, Lloyd T E, Deitcher
tug-of-war mechanism drives bidirectional vesicle D L and Levitan E S 2012 Neuropeptide delivery to
transport Curr. Biol. 20 697–702 synapses by long-range vesicle circulation and sporadic
[383] McLaughlin R T, Diehl M R and Kolomeisky A B 2016 capture Cell 148 1029–38
Collective dynamics of processive cytoskeletal motors Soft [403] Shubeita G T, Tran S L, Xu J, Vershinin M, Cermelli S,
Matter 12 14–21 Cotton S L, Welte M A and Gross S P 2008 Consequences
[384] Elshenawy M M, Canty J T, Oster L, Ferro L S, Zhou Z, of motor copy number on the intracellular transport of
Blanchard S C and Yildiz A 2019 Cargo adaptors regulate kinesin-1-driven lipid droplets Cell 135 1098–107
stepping and force generation of mammalian [404] Rogers S L, Tint I S, Fanapour P C and Gelfand V I 1997
dynein-dynactin Nat. Chem. Biol. 15 1093–101 Regulated bidirectional motility of melanophore pigment
[385] Akhmanova A and Hammer J A III 2010 Linking granules along microtubules in vitro Proc. Natl Acad. Sci.
molecular motors to membrane cargo Curr. Opin. Cell 94 3720–5
Biol. 22 479–87 [405] Ligon L A, Tokito M, Finklestein J M, Grossman F E and
[386] Barlan K, Rossow M J and Gelfand V I 2013 The journey of Holzbaur E L F 2004 A direct interaction between
the organelle: teamwork and regulation in intracellular cytoplasmic dynein and kinesin i may coordinate motor
transport Curr. Opin. Cell Biol. 25 483–8 activity J. Biol. Chem. 279 19201–8
[387] Higuchi Y, Ashwin P, Roger Y and Steinberg G 2014 Early [406] Encalada S E, Szpankowski L, Xia C-h and Goldstein L S B
endosome motility spatially organizes polysome 2011 Stable kinesin and dynein assemblies drive the axonal
distribution J. Cell Biol. 204 343–57 transport of mammalian prion protein vesicles Cell 144
[388] Baumann S, König J, Koepke J and Feldbrügge M 2014 551–65
Endosomal transport of septin mrna and protein indicates [407] Wang X and Schwarz T L 2009 The mechanism of
local translation on endosomes and is required for correct ca2+-dependent regulation of kinesin-mediated
septin filamentation EMBO Rep. 15 94–102 mitochondrial motility Cell 136 163–74
[389] Baumann S, Pohlmann T, Jungbluth M, Brachmann A and [408] Russo G J, Louie K, Wellington A, Macleod G T, Hu F,
Feldbrügge M 2012 Kinesin-3 and dynein mediate Panchumarthi S and Zinsmaier K E 2009 Drosophila miro
microtubule-dependent co-transport of mrnps and is required for both anterograde and retrograde axonal
endosomes J. Cell Sci. 125 2740–52 mitochondrial transport J. Neurosci. 29 5443–55
[390] Pohlmann T, Baumann S, Haag C, Albrecht M and [409] Müller M J I, Klumpp S and Lipowsky R 2010 Bidirectional
Feldbrügge M 2015 A fyve zinc finger domain protein transport by molecular motors: enhanced processivity and
specifically links mrna transport to endosome trafficking response to external forces Biophys. J. 98 2610–8
eLife 4 e06041 [410] Soppina V, Rai A K, Ramaiya A J, Barak P and Mallik R
[391] Schmid M, Jaedicke A, Du T-G and Jansen R-P 2006 2009 Tug-of-war between dissimilar teams of microtubule
Coordination of endoplasmic reticulum and mrna motors regulates transport and fission of endosomes Proc.
localization to the yeast bud Curr. Biol. 16 1538–43 Natl Acad. Sci. 106 19381–6

42
Phys. Biol. 17 (2020) 061003 Topical Review

[411] Derr N D, Goodman B S, Jungmann R, Leschziner A E, [432] Black M M, Slaughter T, Moshiach S, Obrocka M and
Shih W M and Reck-Peterson S L 2012 Tug-of-war in Fischer I 1996 Tau is enriched on dynamic microtubules in
motor protein ensembles revealed with a programmable the distal region of growing axons J. Neurosci. 16 3601–19
dna origami scaffold Science 338 662–5 [433] Vershinin M, Xu J, Razafsky D S, King S J and Gross S P
[412] Martin M A, Iyadurai S J, Gassman A, Gindhart J G Jr, 2008 Tuning microtubule-based transport through
Hays T S and Saxton W M 1999 Cytoplasmic dynein, the filamentous maps: the problem of dynein Traffic 9 882–92
dynactin complex, and kinesin are interdependent and [434] Chaudhary A R, Berger F, Berger C L and Hendricks A G
essential for fast axonal transport Mol. Biol. Cell 10 2017 Tau directs intracellular trafficking by regulating the
3717–28 forces exerted by kinesin and dynein teams Traffic 19
[413] Barkus R V, Klyachko O, Horiuchi D, Dickson B J and 111–21
Saxton W M 2008 Identification of an axonal kinesin-3 [435] Kuznetsov I A and Kuznetsov A V 2014 What tau
motor for fast anterograde vesicle transport that facilitates distribution maximizes fast axonal transport toward the
retrograde transport of neuropeptides Mol. Biol. Cell 19 axonal synapse? Math. Biosci. 253 19–24
274–83 [436] Karasmanis E P, Phan C-T, Angelis D, Kesisova I A,
[414] Ally S, Larson A G, Barlan K, Rice S E and Gelfand V I Hoogenraad C C, McKenney R J and Spiliotis E T 2018
2009 Opposite-polarity motors activate one another to Polarity of neuronal membrane traffic requires sorting of
trigger cargo transport in live cells J. Cell Biol. 187 kinesin motor cargo during entry into dendrites by a
1071–82 microtubule-associated septin Dev. Cell 46 204–18
[415] Kunwar A et al 2011 Mechanical stochastic tug-of-war [437] Wang B, Kuo J and Granick S 2013 Bursts of active
models cannot explain bidirectional lipid-droplet transport in living cells Phys. Rev. Lett. 111 1–5
transport Proc. Natl Acad. Sci. 108 18960–5 [438] Nam W and Epureanu B I 2012 The effects of viscoelastic
[416] Leidel C, Longoria R A, Gutierrez F M and Shubeita G T fluid on kinesin transport J. Phys.: Condens. Matter. 24
2012 Measuring molecular motor forces in vivo: 375103
implications for tug-of-war models of bidirectional [439] Goychuk I, Kharchenko V O and Metzler R 2014 How
transport Biophys. J. 103 492–500 molecular motors work in the crowded environment of
[417] Torisawa T, Ichikawa M, Furuta A, Saito K, Oiwa K, living cells: coexistence and efficiency of normal and
Kojima H, Toyoshima Y Y and Furuta K 2014 anomalous transport PloS One 9 e91700
Autoinhibition and cooperative activation mechanisms of [440] Bouzat S 2014 Influence of molecular motors on the
cytoplasmic dynein Nat. Cell Biol. 16 1118–24 motion of particles in viscoelastic media Phys. Rev. E 89
[418] Ali M Y, Lu H, Bookwalter C S, Warshaw D M and Trybus 062707
K M 2008 Myosin V and kinesin act as tethers to enhance [441] Gagliano J, Walb M, Blaker B, Macosko J C and Holzwarth
each others’ processivity Proc. Natl Acad. Sci. 105 G 2010 Kinesin velocity increases with the number of
4691–6 motors pulling against viscoelastic drag Eur. Biophys. J. 39
[419] Berger F, Müller M J I and Lipowsky R 2009 Enhancement 801–13
of the processivity of kinesin-transported cargo by myosin [442] Lakadamyali M 2014 Navigating the cell: how motors
v Europhys. Lett. 87 28002 overcome roadblocks and traffic jams to efficiently
[420] Kunwar A, Vershinin M, Xu J and Gross S P 2008 Stepping, transport cargo Phys. Chem. Chem. Phys. 16 5907–16
strain gating, and an unexpected force-velocity curve for [443] Bertalan Z, Budrikis Z, La Porta C A M and Zapperi S 2015
multiple-motor-based transport Curr. Biol. 18 1173–83 Navigation strategies of motor proteins on decorated
[421] Vershinin M, Carter B C, Razafsky D S, King S J and Gross tracks PloS One 10 e0136945
S P 2007 Multiple-motor based transport and its [444] Jolly A L and Gelfand V I 2011 Bidirectional intracellular
regulation by tau Proc. Natl Acad. Sci. 104 87–92 transport: utility and mechanism Biochem. Soc. Trans. 39
[422] Rai A K, Rai A, Ramaiya A J, Jha R and Mallik R 2013 1126–30
Molecular adaptations allow dynein to generate large [445] Verdeny-Vilanova I, Wehnekamp F, Mohan N, Álvarez Á S,
collective forces inside cells Cell 152 172–82 Borbely J S, Otterstrom J J, Lamb D C and Lakadamyali M
[423] Li Q, Tseng K-F, King S J, Qiu W and Xu J 2018 A fluid 2017 3d motion of vesicles along microtubules helps them
membrane enhances the velocity of cargo transport by to circumvent obstacles in cells J. Cell Sci. 130 1904–16
small teams of kinesin-1 J. Chem. Phys. 148 123318 [446] Gu Y, Sun W, Wang G, Jeftinija K, Jeftinija S and Fang N
[424] Kunwar A and Mogilner A 2010 Robust transport by 2012 Rotational dynamics of cargos at pauses during
multiple motors with nonlinear force-velocity relations axonal transport Nat. Commun. 3 1–8
and stochastic load sharing Phys. Biol. 7 016012 [447] Bergman J P, Bovyn M J, Dovald F F, Sharmad A,
[425] Klumpp S, Keller C, Berger F and Lipowsky R 2015 Gudhetie M V, Gross S P, Allard J F and Vershinin M D
Molecular motors: cooperative phenomena of multiple 2018 Cargo navigation across 3D microtubule
molecular motors Multiscale Modeling in Biomechanics and intersections Proc. Natl Acad. Sci. USA 115 537–42
Mechanobiology (Berlin: Springer) pp 27–61 [448] Ross J L, Shuman H, Holzbaur E L F and Goldman Y E
[426] Uçar M C and Lipowsky R 2019 Force sharing and force 2008 Kinesin and dynein-dynactin at intersecting
generation by two teams of elastically coupled molecular microtubules: motor density affects dynein function
motors Sci. Rep. 9 1–13 Biophys. J. 94 3115–25
[427] Barlan K and Gelfand V I 2017 Microtubule-based [449] Parmeggiani A, Franosch T and Frey E 2003 Phase
transport and the distribution, tethering, and organization coexistence in driven one-dimensional transport Phys. Rev.
of organelles Cold Spring Harbor Perspect. Biol. 9 a025817 Lett. 90 086601
[428] Chiu S Y 2011 Matching mitochondria to metabolic needs [450] Klumpp S and Lipowsky R 2004 Phase transitions in
at nodes of ranvier Neuroscientist 17 343–50 systems with two species of molecular motors Europhys.
[429] Janke C and Bulinski J C 2011 Post-translational regulation Lett. 66 90
of the microtubule cytoskeleton: mechanisms and [451] Müller M J I, Klumpp S and Lipowsky R 2005 Molecular
functions Nat. Rev. Mol. Cell Biol. 12 773 motor traffic in a half-open tube J. Phys.: Condens. Matter.
[430] Balabanian L, Chaudhary A R and Hendricks A G 2018 17 S3839
Traffic control inside the cell: microtubule-based [452] Ashwin P, Lin C and Steinberg G 2010 Queueing induced
regulation of cargo transport Biochemist 40 14–7 by bidirectional motor motion near the end of a
[431] Dixit R, Ross J L, Goldman Y E and Holzbaur E L F 2008 microtubule Phys. Rev. E 82 051907
Differential regulation of dynein and kinesin motor [453] Miedema D M, Kushwaha V S, Denisov D V, Acar S,
proteins by tau Science 319 1086–9 Nienhuis B, Peterman E J G and Schall P 2017 Correlation

43
Phys. Biol. 17 (2020) 061003 Topical Review

imaging reveals specific crowding dynamics of kinesin reconstruction of the membrane skeleton at the plasma
motor proteins Phys. Rev. X 7 041037 membrane interface by electron tomography J. Cell Biol.
[454] Leduc C, Padberg-Gehle K, Varga V, Helbing D, Diez S and 174 851–62
Howard J 2012 Molecular crowding creates traffic jams of [474] Hafner A E and Rieger H 2018 Spatial cytoskeleton
kinesin motors on microtubules Proc. Natl Acad. Sci. USA organization supports targeted intracellular transport
109 6100–015 Biophys. J. 114 1420–32
[455] Conway L, Wood D, Tüzel E and Ross J L 2012 Motor [475] Schroeder H W, Mitchell C, Shuman H, Holzbaur E L F
transport of self-assembled cargos in crowded and Goldman Y E 2010 Motor number controls cargo
environments Proc. Natl Acad. Sci. 109 20814–9 switching at actin-microtubule intersections in vitro Curr.
[456] Osunbayo O, Butterfield J, Bergman J, Mershon L, Biol. 20 687–96
Rodionov V and Vershinin M 2015 Cargo transport at [476] Zhu X and Kaverina I 2013 Golgi as an mtoc: making
microtubule crossings: evidence for prolonged tug-of-war microtubules for its own good Histochem. Cell Biol. 140
between kinesin motors Biophys. J. 108 1480–3 361–7
[457] Zajac A L, Goldman Y E, Holzbaur E L F and Ostap E M [477] Parton R M, Hamilton R S, Ball G, Yang L, Cullen C F, Lu
2013 Local cytoskeletal and organelle interactions impact W, Ohkura H and Davis I 2011 A PAR-1-dependent
molecular-motor-driven early endosomal trafficking Curr. orientation gradient of dynamic microtubules directs
Biol. 23 1173–80 posterior cargo transport in the Drosophila oocyte J. Cell
[458] Bharti P et al 2011 Pex14 is required for microtubule-based Biol. 194 121–35
peroxisome motility in human cells J. Cell Sci. 124 [478] Pfeiffer D C and Gard D L 1999 Microtubules in xenopus
1759–68 oocytes are oriented with their minus-ends towards the
[459] Kang J-S, Tian J-H, Pan P-Y, Zald P, Li C, Deng C and cortex Cell Motil. Cytoskeleton 44 34–43
Sheng Z-H 2008 Docking of axonal mitochondria by [479] Sugioka K and Sawa H 2012 Formation and functions of
syntaphilin controls their mobility and affects short-term asymmetric microtubule organization in polarized cells
facilitation Cell 132 137–48 Curr. Opin. Cell Biol. 24 517–25
[460] Klumpp S and Lipowsky R 2005 Active diffusion of motor [480] Gagnon J A, Kreiling J A, Powrie E A, Wood T R and
particles Phys. Rev. Lett. 95 268102 Mowry K L 2013 Directional transport is mediated by a
[461] Loverdo C, Bénichou O, Moreau M and Voituriez R 2008 dynein-dependent step in an rna localization pathway
Enhanced reaction kinetics in biological cells Nat. Phys. 4 PLoS Biol. 11 e1001551
134–7 [481] Zimyanin V L, Belaya K, Pecreaux J, Gilchrist M J, Clark A,
[462] Campos D, Abad E, Méndez V, Yuste S B and Lindenberg Davis I and Johnston D S 2008 In vivo imaging of oskar
K 2015 Optimal search strategies of space-time coupled mrna transport reveals the mechanism of posterior
random walkers with finite lifetimes Phys. Rev. E 91 localization Cell 134 843–53
052115 [482] Cha B-J, Koppetsch B S and Theurkauf W E 2001 In vivo
[463] Sanchez A D and Feldman J L 2017 analysis of drosophila bicoid mrna localization reveals a
Microtubule-organizing centers: from the centrosome to novel microtubule-dependent axis specification pathway
non-centrosomal sites Curr. Opin. Cell Biol. 44 93–101 Cell 106 35–46
[464] Wittmann T 2019 Neuron growth Nikon Small World 2019 [483] Tas R P, Chazeau A l, Cloin B M C, Lambers M L A,
Photomicrography Competition https://www. Hoogenraad C C and Kapitein L C 2017 Differentiation
nikonsmallworld.com/galleries/2019-photomicrography- between oppositely oriented microtubules controls
competition/neuron-growth-actin-in-green- polarized neuronal transport Neuron 96 1264–71
microtubules-in-purple-nuclei-in-blue [484] Kapitein L C and Hoogenraad C C 2015 Building the
[465] Wittmann T 2003 Filamentous actin and microtubules in neuronal microtubule cytoskeleton Neuron 87
mouse fibroblasts Nikon Small World 2003 492–506
Photomicrography Competition https://www. [485] Yau K W, Schätzle P, Tortosa E, Pagès S, Holtmaat A,
nikonsmallworld.com/galleries/2003-photomicrography- Kapitein L C and Hoogenraad C C 2016 Dendrites in vitro
competition/filamentous-actin-and-microtubules- and in vivo contain microtubules of opposite polarity and
structural-proteins-in-mouse-fibroblasts axon formation correlates with uniform plus-end-out
[466] Quinlan M E 2016 Cytoplasmic Streaming in microtubule orientation J. Neurosci. 36 1071–85
theDrosophilaOocyte Annu. Rev. Cell Dev. Biol. 32 173–95 [486] Kapitein L C, Schlager M A, Kuijpers M, Wulf P S, van
[467] Lechler T and Fuchs E 2007 Desmoplakin: an unexpected Spronsen M, MacKintosh F C and Hoogenraad C C 2010
regulator of microtubule organization in the epidermis J. Mixed microtubules steer dynein-driven cargo transport
Cell Biol. 176 147–54 into dendrites Curr. Biol. 20 290–9
[468] Toya M, Kobayashi S, Kawasaki M, Shioi G, Kaneko M, [487] Derivery E, Seum C, Daeden A, Loubéry S, Holtzer L,
Ishiuchi T, Misaki K, Meng W and Takeichi M 2016 Jülicher F and Gonzalez-Gaitan M 2015 Polarized
Camsap3 orients the apical-to-basal polarity of endosome dynamics by spindle asymmetry during
microtubule arrays in epithelial cells Proc. Natl Acad. Sci. asymmetric cell division Nature 528 280
113 332–7 [488] Maelfeyt B, Ali Tabei S M and Gopinathan A 2019
[469] Gou J, Edelstein-Keshet L and Allard J 2014 Mathematical Anomalous intracellular transport phases depend on
model with spatially uniform regulation explains cytoskeletal network features Phys. Rev. E 99
long-range bidirectional transport of early endosomes in 062404
fungal hyphae Mol. Biol. Cell 25 2408–15 [489] Wang Z and Thurmond D C 2009 Mechanisms of biphasic
[470] Trong P K, Doerflinger H, Dunkel J, Johnston D S and insulin-granule exocytosis - roles of the cytoskeleton, small
Goldstein R E 2015 Cortical microtubule nucleation can GTPases and SNARE proteins J. Cell Sci. 122 893–903
organise the cytoskeleton of drosophila oocytes to define [490] Ando D, Korabel N, Huang K C and Gopinathan A 2015
the anteroposterior axis eLife 4 e06088 Cytoskeletal network morphology regulates intracellular
[471] Ciocanel M-V, Sandstede B, Jeschonek S P and Mowry K L transport dynamics Biophys. J. 109 1574–82
2018 Modeling microtubule-based transport and [491] Mlynarczyk P J and Abel S M 2019 First passage of
anchoring of mrna SIAM J. Appl. Dyn. Syst. 17 2855–81 molecular motors on networks of cytoskeletal filaments
[472] Salbreux G, Charras G and Paluch E 2012 Actin cortex Phys. Rev. E 99 022406
mechanics and cellular morphogenesis Trends Cell Biol. 22 [492] Hafner A E and Rieger H 2016 Spatial organization of the
536–45 cytoskeleton enhances cargo delivery to specific target
[473] Morone N, Fujiwara T, Murase K, Kasai R S, Ike H, Yuasa areas on the plasma membrane of spherical cells Phys. Biol.
S, Usukura J and Kusumi A 2006 Three-dimensional 13 066003

44
Phys. Biol. 17 (2020) 061003 Topical Review

[493] Corti B 1774 Osservazioni Microscopiche Sulla Tremella: E [517] Tan T H, Malik-Garbi M, Abu-Shah E, Li J, Sharma A,
Sulla Circolazione del Fluido in Una Pianta Acquajuola MacKintosh F C, Keren K, Schmidt C F and Fakhri N 2018
(Lucca: Apresso G. Rocchi) Self-organized stress patterns drive state transitions in
[494] Allen N S and Allen R D 1978 Cytoplasmic streaming in actin cortices Sci. Adv. 4 eaar2847
green plants Annu. Rev. Biophys. Bioeng. 7 497–526 [518] Prost J, Jülicher F and Joanny J-F 2015 Active gel physics
[495] Kamiya N 1981 Physical and chemical basis of cytoplasmic Nat. Phys. 11 111–7
streaming Annu. Rev. Plant Physiol. 32 205–36 [519] Teplov V A 2017 Role of mechanics in the appearance of
[496] Lewis O L, Zhang S, Guy R D and Alamo J C D 2015 oscillatory instability and standing waves of the
Coordination of contractility, adhesion and flow in mechanochemical activity in thePhysarum
migrating physarum amoebae J. R. Soc., Interface 12 polycephalumplasmodium J. Phys. D: Appl. Phys. 50
20141359 213002
[497] Lew R R 2011 How does a hypha grow? the biophysics of [520] Yi K, Unruh J R, Deng M, Slaughter B D, Rubinstein B and
pressurized growth in fungi Nat. Rev. Microbiol. 9 509–18 Li R 2011 Dynamic maintenance of asymmetric meiotic
[498] Klughammer N, Bischof J, Schnellbächer N D, Callegari A, spindle position through arp2/3-complex-driven
Lénárt P and Schwarz U S 2018 Cytoplasmic flows in cytoplasmic streaming in mouse oocytes Nat. Cell Biol. 13
starfish oocytes are fully determined by cortical 1252–8
contractions PLoS Comput. Biol. 14 e1006588 [521] Keren K, Yam P T, Kinkhabwala A, Mogilner A and
[499] Pickard W F 2003 The role of cytoplasmic streaming in Theriot J A 2009 Intracellular fluid flow in rapidly moving
symplastic transport Plant, Cell Environ. 26 1–15 cells Nat. Cell Biol. 11 1219
[500] van de Meent J-W, Tuval I and Goldstein R E 2008 Nature’s [522] Tsai T Y-C et al 2019 Efficient front-rear coupling in
microfluidic transporter: rotational cytoplasmic streaming neutrophil chemotaxis by dynamic myosin ii localization
at high péclet numbers Phys. Rev. Lett. 101 178102 Dev. Cell 49 189–205
[501] Woodhouse F G and Goldstein R E 2013 Cytoplasmic [523] Gross P, Kumar K V and Grill S W 2017 How active
streaming in plant cells emerges naturally by microfilament mechanics and regulatory biochemistry combine to form
self-organization Proc. Natl Acad. Sci. 110 14132–7 patterns in development Annu. Rev. Biophys. 46 337–56
[502] Mitchison T J, Charras G T and Mahadevan L 2008 [524] Bois J S, Jülicher F and Grill S W 2011 Pattern formation in
Implications of a poroelastic cytoplasm for the dynamics active fluids Phys. Rev. Lett. 106 028103
of animal cell shape Seminars in Cell & Developmental [525] Paluch E K and Raz E 2013 The role and regulation of
Biology vol 19 (Amsterdam: Elsevier) pp 215–23 blebs in cell migration Curr. Opin. Cell Biol. 25 582
[503] Mogilner A and Manhart A 2018 Intracellular fluid –90
mechanics: coupling cytoplasmic flow with active [526] Li Y, He L, Gonzalez N A P, Graham J, Wolgemuth C,
cytoskeletal gel Annu. Rev. Fluid Mech. 50 347–70 Wirtz D and Sun S X 2017 Going with the flow: Water flux
[504] Charras G T, Coughlin M, Mitchison T J and and cell shape during cytokinesis Biophys. J. 113 2487
Mahadevan L 2008 Life and times of a cellular bleb –95
Biophys. J. 94 1836–53 [527] Paluch E, Sykes C, Prost J and Bornens M 2006 Dynamic
[505] Radszuweit M, Alonso S, Engel H and Bär M 2013 modes of the cortical actomyosin gel during cell
Intracellular mechanochemical waves in an active locomotion and division Trends Cell Biol. 16 5–10
poroelastic model Phys. Rev. Lett. 110 138102 [528] Roh-Johnson M et al 2012 Triggering a cell shape change
[506] Goehring N W, Trong P K, Bois J S, Chowdhury D, Nicola by exploiting preexisting actomyosin contractions Science
E M, Hyman A A and Grill S W 2011 Polarization of par 335 1232–5
proteins by advective triggering of a pattern-forming [529] Lämmermann T and Sixt M 2009 Mechanical modes of
system Science 334 1137–41 ‘amoeboid’ cell migration Curr. Opin. Cell Biol. 21 636–44
[507] Ganguly S, Williams L S, Palacios I M and Goldstein R E [530] Caswell P T and Zech T 2018 Actin-based cell protrusion
2012 Cytoplasmic streaming in drosophila oocytes varies in a 3d matrix Trends Cell Biol. 28 823–34
with kinesin activity and correlates with the microtubule [531] Bodor D L, Pönisch W, Endres R G and Paluch E K 2020
cytoskeleton architecture Proc. Natl Acad. Sci. 109 Of cell shapes and motion: the physical basis of animal cell
15109–14 migration Dev. Cell 52 550–62
[508] Deneke V E, Puliafito A, Krueger D, Narla A V, De Simone [532] Abadeh A and Lew R R 2013 Mass flow and velocity
A, Primo L, Vergassola M, De Renzis S and Di Talia S 2019 profiles in neurospora hyphae: partial plug flow dominates
Self-organized nuclear positioning synchronizes the cell intra-hyphal transport Microbiol. 159 2386–94
cycle in drosophila embryos Cell 177 925–41 [533] Tominaga M and Ito K 2015 The molecular mechanism
[509] Castle B T, Howard S A and Odde D J 2011 Assessment of and physiological role of cytoplasmic streaming Curr.
transport mechanisms underlying the bicoid morphogen Opin. Plant Biol. 27 104–10
gradient Cell. Mol. Bioeng. 4 116–21 [534] Shimmen T and Yokota E 2004 Cytoplasmic streaming in
[510] Taylor G I 1967 Low Reynolds Number Flows (Chicago: plants Curr. Opin. Cell Biol. 16 68–72
Encyclopaedia Britannica Educational Corp) Video [535] Mittasch M et al 2018 Non-invasive perturbations of
[511] Heller J P 1960 An unmixing demonstration Am. J. Phys. intracellular flow reveal physical principles of cell
28 348–53 organization Nat. Cell Biol. 20 344–51
[512] Verchot-Lubicz J and Goldstein R E 2010 Cytoplasmic [536] Bradke F and Dotti C G 1997 Neuronal polarity: vectorial
streaming enables the distribution of molecules and cytoplasmic flow precedes axon formation Neuron 19
vesicles in large plant cells Protoplasma 240 99–107 1175–86
[513] Hundsdorfer W and Verwer J G 2013 Numerical Solution of [537] Maza N A, Schiesser W E and Calvert P D 2019 An
Time-dependent Advection-Diffusion-Reaction Equations intrinsic compartmentalization code for peripheral
vol 33 (Berlin: Springer) membrane proteins in photoreceptor neurons J. Cell Biol.
[514] Matsumoto K, Takagi S and Nakagaki T 2008 Locomotive 218 3753–72
mechanism of physarum plasmodia based on [538] Trong P K, Guck J and Goldstein R E 2012 Coupling of
spatiotemporal analysis of protoplasmic streaming active motion and advection shapes intracellular cargo
Biophys. J. 94 2492–504 transport Phys. Rev. Lett. 109 028104
[515] Pieuchot L, Lai J, Loh R A, Leong F Y, Chiam K-H, Stajich [539] Shapiro A H, Jaffrin M Y and Weinberg S L 1969 Peristaltic
J and Jedd G 2015 Cellular subcompartments through pumping with long wavelengths at low Reynolds number J.
cytoplasmic streaming Dev. Cell 34 410–20 Fluid Mech. 37 799–825
[516] Chebli Y, Kroeger J and Geitmann A 2013 Transport [540] Liron N and Shahar R 1978 Stokes flow due to a stokeslet
logistics in pollen tubes Mol. Plant 6 1037–52 in a pipe J. Fluid Mech. 86 727–44

45
Phys. Biol. 17 (2020) 061003 Topical Review

[541] Taylor G I 1953 Dispersion of soluble matter in solvent microtubules and fast fluid flows Biophys. J. 110
flowing slowly through a tube P Roy Soc A-Math Phy 219 2053–65
186–203 [550] Serbus L R, Cha B-J, Theurkauf W E and
[542] Aris R 1956 On the dispersion of a solute in a fluid flowing Saxton W M 2005 Dynein and the actin cytoskeleton
through a tube Proc. - R. Soc. Edinburgh, Sect. A: Math. control kinesin-driven cytoplasmic streaming in
Phys. Sci. 235 67–77 drosophila oocytes Development 132 3743–52
[543] Aref H and Balachandar S 1986 Chaotic advection in a [551] Haupt M and Hauser M J B 2020 Effective mixing due to
stokes flow Phys. Fluids 29 3515–21 oscillatory laminar flow in tubular networks of plasmodial
[544] Ottino J M and Ottino J M 1989 The Kinematics of Mixing: slime moulds New J. Phys. 22 053007
Stretching, Chaos, and Transport vol 3 (Cambridge: [552] Parry B R, Surovtsev I V, Cabeen M T, O’Hern C S,
Cambridge University Press) Dufresne E R and Jacobs-Wagner C 2014 The
[545] Batchelor C K and Batchelor G K 2000 An Introduction to bacterial cytoplasm has glass-like properties and is
Fluid Dynamics (Cambridge: Cambridge University fluidized by metabolic activity Cell 156 183–94
Press) [553] Humphrey D, Duggan C, Saha D, Smith D and Käs J 2002
[546] Chaiken J, Chevray R, Tabor M and Tan Q M 1986 Active fluidization of polymer networks through
Experimental study of lagrangian turbulence in a stokes molecular motors Nature 416 413–6
flow Proc. R. Soc. London, Ser. A 408 165–74 [554] Fodor É, Ahmed W W, Almonacid M, Bussonnier M, Gov
[547] Aref H and Jones S W 1987 Chaotic advection: efficient N S, Verlhac M-H, Betz T, Visco P and van Wijland F 2016
stirring of viscous liquids Fifth Symposium on Energy Nonequilibrium dissipation in living oocytes Europhys.
Engineering Sciences p 209 Lett. 116 30008
[548] Reverey J F, Jeon J-H, Bao H, Leippe M, Metzler R and [555] Takahashi K, Tǎnase-Nicola S and Wolde P R T 2010
Selhuber-Unkel C 2015 Superdiffusion dominates Spatio-temporal correlations can drastically change the
intracellular particle motion in the supercrowded response of a mapk pathway Proc. Natl Acad. Sci. 107
cytoplasm of pathogenic acanthamoeba castellanii Sci. Rep. 2473–8
5 1–14 [556] Abel S M, Roose J P, Groves J T, Weiss A and Chakraborty
[549] Monteith C E, Brunner M E, Djagaeva I, Bielecki A M, A K 2012 The membrane environment can promote or
Deutsch J M and Saxton W M 2016 A mechanism for suppress bistability in cell signaling networks J. Phys.
cytoplasmic streaming: kinesin-driven alignment of Chem. B 116 3630–40

46

You might also like