Anda di halaman 1dari 10

2710 Biophysical Journal Volume 106 June 2014 2710–2719

Drag of the Cytosol as a Transport Mechanism in Neurons

Matan Mussel,† Keren Zeevy,† Haim Diamant,‡ and Uri Nevo†§*



Department of Biomedical Engineering, Tel Aviv University, Tel Aviv, Israel; ‡School of Chemistry, Tel Aviv University, Tel Aviv, Israel; and
§
Sagol School of Neuroscience, Tel Aviv University, Tel Aviv, Israel

ABSTRACT Axonal transport is typically divided into two components, which can be distinguished by their mean velocity.
The fast component includes steady trafficking of different organelles and vesicles actively transported by motor proteins.
The slow component comprises nonmembranous materials that undergo infrequent bidirectional motion. The underlying mech-
anism of slow axonal transport has been under debate during the past three decades. We propose a simple displacement
mechanism that may be central for the distribution of molecules not carried by vesicles. It relies on the cytoplasmic drag
induced by organelle movement and readily accounts for key experimental observations pertaining to slow-component trans-
port. The induced cytoplasmic drag is predicted to depend mainly on the distribution of microtubules in the axon and the organ-
elle transport rate.

INTRODUCTION
Neurons are highly polarized cells with axons capable of Axonal transport plays a few major roles. 1), It is respon-
extending long distances. Individual axons are microscopic sible for bidirectional molecular signaling over long dis-
in diameter (typically of the order of 1 mm), whereas their tances. 2), It provides essential cellular organelles and
lengths may span as much as one meter in humans (e.g., materials from the cell body to the entire axon (anterograde
axons in the sciatic nerve and spinal cord) and even more transport by the kinesin motor protein). This supply of
in large animals such as giraffes or whales. Since axons newly synthesized proteins and lipids to the distal synapse
are cellular structures with a unique morphology, the maintains axonal activity. 3), It causes molecules destined
morphogenesis, integrity, and function of neurons are partic- for degradation to return from the periphery to the lyso-
ularly dependent on active intracellular transport, also called somes in the soma (retrograde transport by the dynein motor
axonal transport (1,2). protein) (8,9). The modes of transport through the axon are
There are three known major constituents of the neuronal typically divided into two categories: a fast component and
cytoskeleton: microtubules, microfilaments, and neurofila- a slow component (1). The components are distinct in their
ments (2). In the axon and dendrites, microtubules and mean velocity, as measured in studies of radioactively
neurofilaments are the major longitudinal cytoskeletal labeled proteins (10–15).
filaments. In the synaptic regions, such as the presynaptic The fast component is a directed movement of mem-
terminals and postsynaptic spines, microfilaments and branous cargoes. Small cargoes move continuously in one
microtubules form the major cytoskeletal architecture (3). direction and maintain almost uniform velocity that ranges
There are also microfilaments along the axon, which mainly between 0.4 and 3 mm/s, whereas larger cargoes exhibit
provide a scaffold beneath the axonal plasma membrane and irregular saltatory motion (16–21). The fast transport is
around microtubules (4,5). carried primarily by molecular motor proteins that bind
The lines of transport, communication, and control in the directly to the cargo surface (22–24). The directionality of
cell are provided by microtubules. These dynamic polymers the cargo depends on the motor type and the polarity of
create a network of tracks for motor proteins that carry the microtubule tracks (1).
various vesicles and organelles (referred to hereafter as The slow component (SC), which is the focus of this work,
cargoes), with diameters of the order of 10–100 nm (6). is usually divided into two subgroups with distinct overall
Microtubules appear as bundles of tubes dispersed across rates: SCa and SCb (1). SCa transports cytoskeletal proteins,
the axon and are aligned mostly along the axon axis to allow whereas SCb transports hundreds of different proteins that
long-range transport (see, e.g., Fig. 8-1 in Siegel (2)). In are important for synaptic and axonal maintenance,
comparison to the two other intracellular cytoskeletal including metabolic enzymes, chaperons, presynaptic pro-
filaments, microtubules are the most rigid, thus ensuring teins, and cytoskeletal proteins (see Lasek and colleagues
mechanical stability under the forces of large cargo (25,26), Baitinger and Willard (27), Nixon et al. (28), Yuan
movement (7). et al. (29), Bourke et al. (30), and Roy et al. (31) and refer-
ences therein). Key experimental observations concerning
both SCa and SCb are as follows (31–35). 1), Transported
Submitted January 7, 2014, and accepted for publication April 28, 2014.
soluble molecules spend most of their time pausing, and their
*Correspondence: nevouri@tau.ac.il
movement appears bidirectional. 2), They are not attached to
Matan Mussel and Keren Zeevy contributed equally to this work.
vesicles but homogeneously distributed throughout the entire
Editor: David Odde.
Ó 2014 by the Biophysical Society
0006-3495/14/06/2710/10 $2.00 http://dx.doi.org/10.1016/j.bpj.2014.04.037
Cytosol Drag in Neurons 2711

cytoplasm. However, their transport was found to depend on a b


motor proteins (35). 3), They move at broadly distributed
velocities, ranging from close to zero to the velocity of the
fast component, i.e., a few mm/s. 4), Slow-component pro-
teins were even observed to move side by side with fast-
component cargoes (34).
In contrast to the transport mechanism of the fast compo-
nent, that of the SC is still under debate (36). Blum and Reed
(37) suggested a set of interactions between soluble proteins
FIGURE 1 (a) Diagram highlighting key features in axonal cross section.
and motor proteins. When bound to the motor, the soluble Black dots represent microtubules going through the plane of the figure and
proteins move, and when not bound, they are stationary. distributed uniformly over the cross section. One cargo is attached to a
Additional stochastic models have also been suggested microtubule at the upper right corner of the axon. For more details, see
(38–41) in which soluble proteins bound to motors move Bray and Bunge (46). (b) Schematic diagram of the model. Spheres of
in the cytoplasm at different discrete velocity states and in radius a move at constant velocity along the axial direction of a cylindrical
tube of radius R and length L.
which the transition from motion to halt is a probabilistic
one. Scott et al. (35) proposed that soluble proteins move
in the cytoplasm through both passive diffusion and a direct We model the axon as a rigid, fluid-filled, cylindrical tube
interaction with hypothetical mobile units. In that model, of length L and radius R. The problem is treated using cylin-
the assembly and disassembly from the mobile units is prob- drical coordinates r ¼ (r, 4, z), where the tube axis lies
abilistic. Miller and Heidmann (36) offered a qualitative along the bz direction. The cargoes are modeled as rigid
description according to which soluble protein dynamics spheres of radius a moving inside and along the tube. Effects
is governed by several mechanisms, including diffusion related to shape changes of both the axon and the cargoes, as
from high to low concentrations, vesicles dispersing soluble well as the internal dynamics inside the cargoes, are thus
proteins as they travel, and displacement induced by the neglected. We assume that each cargo moves at its own con-
elongation of axons. stant velocity, Ui, parallel to the axis of the tube (21). The
In this work, we propose an alternative, simpler, mecha- tube is filled with a viscous Newtonian fluid at zero
nism for SC axonal transport. It relies on the indirect hydro- Reynolds number (47) having viscosity h. This may seem
dynamic interaction (flow-induced drag) that exists between a crude oversimplification, given the complexity of the cyto-
soluble particles and actively transported cargoes along the plasm. However, based on our results, the detailed response
axon. We show, without further conjecture, that this simple of the fluid should not have a strong effect on the long-range
mechanism is sufficient to account for the most relevant transport through the tube. We return to this issue in the
experimental observations mentioned above. An illustrative Discussion. The hydrodynamics of particles moving
metaphor for the suggested mechanism is that of log driving, through fluid-filled tubes have been widely studied (see,
i.e., the transportation of logs down a river. However, in the e.g., Brenner and colleagues (47–49), Liron and Shahar
case of intracellular drag, the river flows intermittently, (50), and Blake (51)). Here, we focus on the fluid displace-
being driven by anterogradely and retrogradely moving ment in the Lagrangian framework, i.e., the displacement of
vesicles, rather than by a steady external field (gravity). fluid particles resulting from the active motion of the
Earlier theoretical works addressing cytoplasmic flow cargoes. We consider the transport of the SC as a passive
along the elongated cells of Characean algae (42–45) are entrainment of soluble proteins by these active flows. This
worthy of mention here. Although those studies also consid- assumption relies on the fact that the proteins (a few nano-
ered flows induced by actively moving organelles, their flow meters in diameter) are much smaller than the cargoes (with
scenarios differ qualitatively in their geometry from that radius a on the order of 10–100 nm).
considered here. Flow problems in such confined geometries are highly
The outline of the article is as follows. In the next section, dependent on boundary conditions (52). For example, when
we describe the model and its assumptions. Technical a sphere moves an infinite distance in an unbounded quies-
methods used to analyze the model are explained in the cent fluid, the net displacement of any fluid particle is infinite
Methods section. In the Results section, we provide detailed (53). This idealized result arises from the slow 1/r decay of
data from the analysis in different scenarios. In the Discus- fluid velocity with distance r away from the moving sphere.
sion section, we discuss specific implications of the analysis The opposite example is of a sphere moving in a closed tube
for actual axonal transport, as well as further implications. with no-slip boundary conditions. The fluid velocity induced
by the sphere decays exponentially with distance (50), result-
ing in a finite displacement of fluid particles.
MODEL
A key question is the extent of solvent permeation
The system under consideration is schematically shown in through the boundaries along the axon. It is known that fluid
Fig. 1. diffuses through the membrane, e.g., via aquaporin channels

Biophysical Journal 106(12) 2710–2719


2712 Mussel et al.

(54). However, how appreciable such permeation is as a element software (COMSOL, version 4.2a). For a single sphere moving
result of cargo movement remains unclear. Hence, we treat at velocity U, the calculation is done in the reference frame of the
sphere, i.e., the sphere is at rest and the fluid flows past it with a far velocity
two simplified limits for the boundary conditions to be taken of U. Trajectories of fluid particles are subsequently calculated by
at the tube edges. 1), For the case of negligible permeation, integrating the velocity field and transforming back to the fluid reference
we model the axon as a closed tube of a given length, L. 2), frame. For several spheres, we employ a superposition approximation after
If permeation is significant, we replace the leaky axon with confirming its validity for a number of spheres moving at the same velocity
an open-ended impermeable tube of effective length L (see Results for details).
Equation 1 is Galilean-invariant and, therefore, the same in the fluid and
related to the membrane permeability, m. The correspon- sphere reference frames. The entire system, however, is not Galilean-
dence between these two systems is found by a simple anal- invariant, which is reflected in different boundary conditions. We begin
ysis of a pressure-driven flow through a permeable tube,
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi with the boundary conditions at the edge of the tube. For a closed
yielding a decay length of L ¼ R3 =ð8hmÞ. (One can tube the boundary conditions are zero fluid velocity, v(r, 4, 0) ¼
model the axon more accurately as an actual permeable v(r, 4, L) ¼ 0, which in the reference frame of the sphere turn into condi-
tions of constant velocity, vðr; 4; 0Þ ¼ vðr; 4; LÞ ¼ Ubz . For an open-
tube, yet we prefer to defer such a detailed analysis to a ended tube, the boundary conditions at the edges, in both reference frames,
future study.) We note that the flow patterns in these two are zero pressure difference, p(r, 4, 0) ¼ p(r, 4, L), such that the flow is
limits are quite different. When a sphere moves in a closed caused solely by the motion of the sphere.
tube in the axial direction, the fluid is quiescent far away Concerning the boundary conditions along the tube walls, r ¼ R, no-slip
from the sphere, whereas near the sphere, some fluid trajec- boundary conditions in the fluid frame, v(R, 4, z) ¼ 0, turn into
vðR; 4; zÞ ¼ Ubz in the sphere frame. Full-slip boundary conditions,
tories must go against the direction of motion due to mass vr(R, 4, z) ¼ vrv4(R, 4, z) ¼ vrvz(R, 4, z) ¼ 0, remain the same in both
conservation (50). By contrast, a similar sphere moving reference frames. Partial-slip boundary conditions in the fluid frame read,
through an open-ended tube creates a Poiseuille flow far vr(R, 4, z) ¼ 0 and v4(R, 4, z) þ ‘vrv4(R, 4, z) ¼ vz(R, 4, z) þ
from the sphere in the direction of motion (55). ‘vrvz(R, 4, z) ¼ 0, where ‘ is the slip length. In the sphere frame, the right
An additional property that may influence the fluid hand side of the last two conditions is changed from zero to Ubz . The actual
application of the partial-slip conditions requires further consideration due to
behavior is the friction at the tube walls. For example, obser- software limitations. We use the method presented in Wolff et al. (43): a thin
vations in the giant internodal cells of Characean algae fluid layer of width e ¼ 0.001R is added around the tube. The additional layer
showed that the fluid slides on the inner surface of the cell has a different viscosity, he < h. The boundary conditions are no-slip at the
(56). Accordingly, slip boundary conditions were used in external radius, R þ e, and full-slip at R. The relation between the slip length
modeling that system (43). The appropriate boundary condi- and these parameters is ‘ ¼ e(h/he 1) þ O(e2) (43).
tions applied by the axonal membrane and the thick protein
layer adjacent to it are unknown. Therefore, we consider the
entire range of possibilities, from the no-slip boundary con- RESULTS
dition through partial slip to full slip. Single sphere in a closed tube
Typical trajectories of fluid particles dragged by a moving
METHODS sphere along the central axis (r ¼ 0) of a closed tube
with no-slip boundary conditions are depicted in Fig. 2 a.
We numerically solve the Navier-Stokes equations for an incompressible
fluid in the limit of zero Reynolds number (Stokes equations), At a particular radial position, r ¼ r0, fluid particles
move outward and circle back to their initial position
Vp þ hV2 v ¼ 0
; (1) (Fig. 2 a, inset). Fluid particles located near the center
Vv ¼ 0 line (0 % r < r0) are dragged forward, whereas those at
for a fluid confined inside a tube. In Eq. 1, v(r, 4, z) and p(r, 4, z) are the the periphery (r0 < r < R) move backward. The value of
fluid velocity and pressure fields. The calculation is done using a finite- r0/R depends on a/R and L/R.

a b c

FIGURE 2 (a) Typical trajectories of fluid particles due to the motion of a single cargo along the axis of a cylindrical tube. Backward particle trajectories
are marked in red. (Inset) Enlargement of particle trajectories with positive, zero, and negative net displacement. (b) The displacement profile. (c) Depen-
dence of the fluid displacement on a/R for four radial positions. Physical parameters are a/R ¼ 0.1 (in a and b), L/R ¼ 10. To see this figure in color, go online.

Biophysical Journal 106(12) 2710–2719


Cytosol Drag in Neurons 2713

We define a net fluid displacement, d(r), as the total to a times a function that diverges as r / 0. For a(r(r0,
distance traversed by a fluid particle that started at the displacement is proportional to R, arising mainly from
position r, for given initial and final positions of the moving the stages outside the near field. Finally, for r0 (r<R, the
sphere, zi and zf. The inset of Fig. 2 a shows three trajec- (negative) displacement is governed by the opposite flow
tories with positive, zero, and negative net displacements. due to mass conservation.
The displacement profile corresponding to the trajectories We now consider the effect of partial slip at the tube
of Fig. 2 a is shown in Fig. 2 b. The displacement is walls. Fig. 3 a depicts the displacement profiles for several
zero at r ¼ r0, and for a no-slip tube, it becomes zero again values of the slip length ‘, compared against the profile of a
at r ¼ R. no-slip (‘ ¼ 0) and full-slip (‘ / N) tube.
We now discuss additional general properties of the net An important observation is that the effect of wall slip on
displacement profile. 1), It is independent of U and h. the displacement profile is minor. In addition, we find a
This can be shown by dimensional arguments and is also special radial position, r ¼ r‘, at which the fluid displace-
verified numerically. 2), If the initial and final positions of ment is a constant independent of ‘ (see Fig. 3 a, inset).
the sphere, and the initial position of the fluid particle, are Closer to the axis, r < r‘, the net displacement increases
all well separated, zi  z  zf, with jzf  zij [ R, the as ‘ is increased, whereas closer to the boundary, r > r‘,
displacement profile becomes independent of axial position the net displacement becomes more negative with the slip.
d(r) ¼ d(r,4). This results from the fact that the fluid We note that the insensitivity to ‘ is restricted to the total
velocity field decays exponentially with z/R away from the displacement. The detailed trajectories of fluid particles
moving sphere (50). In addition, in the case of azimuthal significantly depend on slip, as demonstrated in Fig. 3 b.
symmetry (the sphere moving along the central axis), the Another way to examine fluid drag is through the drift
profile depends only on the radial position of the fluid par- volume (57), defined as the volume enclosed between the
ticle, d(r) ¼ d(r). 3), For the same reason, unless the fluid initial and final positions of a marked fluid surface—here
particle is arbitrarily close to the center line, the net taken as the cross section of the tube. In the case of a closed
displacement is achieved during a finite period of time. tube, described here, the drift volume vanishes due to mass
This local response in space and time, resulting from the conservation, as explained above. We define the positive and
confinement of the fluid, allows us to restrict the discussion negative contributions to the drift volume, Vpos and Vneg,
hereafter to spheres that move along the entire tube (zi ¼ 0, which are the integrals of the displacement profile, d(r),
zf ¼ L [ R). In this limit, d/R becomes a function of r/R for 0 % r % r0 and r0 % r % R, respectively, such that
and a/R only. 4), For sufficiently small values of a/R and a Vpos ¼ jVnegj. Fig. 4 shows the change in these partial
finite value of r/R, we recover the displacement profile volumes as the slip length, ‘, is varied. Although they
in an unbounded fluid (53), with d proportional to a, increase monotonically with ‘, the increase is not very
as shown in Fig. 2 c. For small values of both r/R and significant, reflecting again the insensitivity of the net fluid
a/R, the displacement profile has the form d ¼ c1a2/r motion to the boundary conditions at the tube wall.
c2aln(r/R) (53). 5), In a closed tube, due to mass con-
servation, the integral of the displacement profile vanishes,
Many spheres in a closed tube
! d (r, 4) rdrd4 ¼ 0.
The shape of d(r) (Fig. 2 b) may be divided into three For a system with N simultaneously moving spheres in a
regions. For r(a, the displacement is governed by the closed tube, we introduce an approximate solution and
near field—the stage of motion when the sphere and fluid verify it as valid if the separation between the spheres is
particle were close together. This behavior is similar to larger than the tube radius R. Due to hydrodynamic
that in an unbounded fluid, with a displacement proportional screening, the fluid trajectories and the displacement profile

a b

FIGURE 3 (a) Displacement profile for no-slip (black), full-slip (red), and partial-slip tubes with several ‘ values (blue). (b) Trajectory of a fluid particle,
initially located at r‘, for no-slip (black), full-slip (red), and several finite values of slip length, ‘. Parameters: a/R ¼ 0.3, L/R ¼ 30. To see this figure in color,
go online.

Biophysical Journal 106(12) 2710–2719


2714 Mussel et al.

Open-ended tube
In an open-ended tube with zero pressure difference
between the tube ends, the sphere movement generates in
the entire tube a z- and 4-independent global flow, vðrÞ, in
addition to a fast-decaying local flow. The latter screened
contribution is very similar to the flow produced by the
sphere in a closed tube, as described in the previous section.
This is demonstrated in Fig. 7.
Consequently, the superposition approximation for N
spheres in an open-ended tube takes the form
FIGURE 4 Change in the partial volumes, Vpos ¼ jVnegj, as ‘ is varied. X
N
Parameters: a/R ¼ 0.01, L/R ¼ 6. dN ðx; yÞy ½d1 ðx  xi ; y  yi Þ þ vi ðrÞti ; (3)
i¼1
can be calculated by superimposing the motions induced by
each sphere separately. Furthermore, for trajectories and where d1 is the single-sphere displacement profile discussed
displacements occurring sufficiently far away from the in the previous section, and ti ¼ L/Ui is the total time of
tube walls, one gets good estimates by considering the motion of sphere i through the tube. Since vi is proportional
much simpler situation, where each sphere moves along to Ui/L (58), the contributions of the global flows, vi ti , add
the symmetry axis of the tube. One subsequently shifts the to the net displacement a constant, dðrÞ, independent of
resulting fluid motion back to the actual cross-sectional Ui and L.
position of the sphere before superimposing the effects of The contribution of the global flow to the net displace-
all the spheres. This approximation does not satisfy the ment in an open-ended tube, d, makes it always larger
boundary conditions at the tube walls, yet the error is found than the net displacement in the corresponding closed
to be small. For example, the total fluid displacement profile tube. In addition, it makes the displacement sensitive to
is calculated according to the boundary conditions on the tube wall. Both effects can
be seen in Fig. 8. As the slip length is increased, the net
X
N
displacement increases because of the reduced friction at
dN ðx; yÞy d1 ðx  xi ; y  yi Þ; (2)
i¼1
the walls. Its profile flattens with increasing ‘, as the global
flow changes from parabolic to plug-shaped. For full-slip
where (xi, yi) denotes the cross-sectional position of the ith boundary conditions, the net displacement becomes propor-
sphere, and d1 is the displacement induced by a single tional to the system size, L. Without the effect of the global
sphere moving along the central axis as obtained, for flow, the curves shown in Fig. 8 become similar to those in
example, numerically. Fig. 3 a.
We demonstrate the superposition approximation for 11
spheres arranged as shown in Fig. 5, and Fig. 6 compares
the displacement profiles, d(r), as obtained from the full DISCUSSION
numerical solution and from the superposition approxima- Consequences for cytoplasmic drag in axons
tion in Fig. 5. The approximation error is observable but
small, on the order of a few percent. Transport due to the motion of a single cargo

Let us estimate the mean displacement of a soluble molecule


due to the motion of a single cargo. We assume a uniform
distribution of ~100 microtubules over an axonal cross sec-
tion of ~1 mm2 (46,59). Each soluble molecule is therefore
within a 50 nm distance of a microtubule. We consider the
displacement of the molecule due to the motion of a cargo
on the nearest microtubule only. Including cargoes moving
farther away will only increase the mean displacement,
and the estimate given below, therefore, is a lower bound.
The mean displacement induced by a cargo moving along
the nearest microtubule (i.e., averaged over 0 < r <
50 nm) is proportional to the cargo radius (see Results).
FIGURE 5 Geometry of 11 spheres arranged at five different initial Because of the slow radial decay, the proportionality factor
radial positions. (a) Side view. (b) Cross-sectional view. Parameters: is large (of order 10), as demonstrated in Fig. 9. The conse-
a/R ¼ 0.1, L/R ¼ 12. quence is that a single cargo induces a ballistic motion of the

Biophysical Journal 106(12) 2710–2719


Cytosol Drag in Neurons 2715

FIGURE 6 (a) Displacement profile of the


system described in Fig. 5 as obtained numerically.
(b) Approximate displacement profile as obtained
from superposition. (c) Difference between the
two displacement profiles. To see this figure in
color, go online.

fluid near the microtubule to a distance an order of magni- tic motion, the soluble molecule is expected to undergo
tude larger than the cargo itself, which is of order 1 mm. smaller drags in both directions due to cargoes on more
Since the cargo moves at a typical velocity of 1 mm/s, this distant microtubules.
motion should last for a few seconds. In summary, our lower-bound estimate yields an occa-
The rate of cargoes traveling along an axon was sional micron-scale ballistic motion (stop-and-go motion)
measured to be ~1–100 cargoes/min in axons with a cross of a soluble molecule, lasting a few seconds. These
section of 1 mm2 (60,61). This is a lower-bound estima- results are consistent with those from experiments by
tion, since not all cargoes are visible in such experiments. Roy et al. (31,34). (Note that in those experiments, the
We therefore consider a cargo rate of the order of analysis was deliberately restricted to proteins moving
100 cargoes/min. Assuming that the cargoes are uniformly along large distances of >10 mm.) For illustration,
distributed among the microtubules, the rate of cargoes on Fig. 10 shows two such intermittent trajectories of a
a single microtubule, g, is ~1 cargo/min. Thus, on average, soluble molecule due to cargo motion that bear qualita-
a soluble molecule is dragged ballistically for a few tive resemblance to trajectories reported in the Roy
seconds every minute. In addition to the significant ballis- et al. studies (31,34).

0.05
1.5

0
1
-0.05
0.4 0.5 0.6
0.5
ρ0/R

-0.50 0.2 0.4 0.6 0.8

FIGURE 7 Fluid trajectories due to the motion of a single sphere in


no-slip closed and open-ended tubes. Red trajectories are for an open-ended
tube after subtraction of the global (Poiseuille) flow, and black trajectories FIGURE 8 Displacement profiles induced by a single sphere in an open-
are for a closed tube. (Inset) Enlargement of trajectories near r0. ended tube for several slip lengths. The case of a no-slip closed tube is
Parameters: a/R ¼ 0.1, L/R ¼ 10. To see this figure in color, go online. shown for reference. Parameters: a/R ¼ 0.1, L/R ¼ 10.

Biophysical Journal 106(12) 2710–2719


2716 Mussel et al.

complete unidirectional transport (e.g. for Macropinosomes


0 < ρ/R ≤ 0.05 or mitochondria in specific regions) (64). As an example, a
4
0 < ρ/R ≤ 0.1
10% bias would give a mean fluid displacement of
0.1hdi/min, which is of the order of 0.1–1 mm/day for a
3
closed tube. This is a conservative estimate; describing the
axon as an open-ended tube with partial-slip boundary
2 conditions would yield a larger net flow (note the effect in
Fig. 8). These values are not inconsistent with the
1 experimentally observed overall rate of SC particles: 0.1–
1 mm/day for SCa and 1–10 mm/day for SCb (1).
0
0 0.05 0.1 0.15 0.2
Further implications

FIGURE 9 Mean fluid displacement within a given distance from the In this article, we propose that drag induced by active cargo
microtubule (rmax/R ¼ 0.05 or 0.1) as a function of a/R in a closed tube. motion may be a plausible mechanism contributing to the
Parameter: L/R ¼ 15. motion of SC proteins. The implications for axonal transport
are discussed above. The essential conclusion is that fluid
particles and soluble molecules should be dragged in a
Overall transport
stop-and-go bidirectional motion along micron-scale dis-
We now turn to the overall transport of soluble molecules tances during times on the order of seconds. This conclusion
due to the motion of many cargoes. This transport is is in line with the known experimental facts concerning SC
crucially dependent on the existence, or absence, of bias transport. Although additional mechanisms influencing the
in the direction of cargo motion. dynamics of SC proteins are not ruled out, the mechanism
Let us begin by assuming no such bias. In such a case, proposed here relies on simple physical effects without the
fluid particles are transported randomly in both directions. need to add extra elements such as binding of the molecules
Over a very long time ([g1, i.e., many minutes), their to motor proteins (35).
motion will become diffusive. The effective diffusion coef- Our analysis is based on two main assumptions. The first
ficient is Deff ~ ghdi2 ~ 1010  109 cm2/s. This is much is that transported molecules are assumed to be well dis-
smaller than the diffusion coefficient of water molecules in- solved in the medium and therefore carried along with it
side cells (106  105 cm2/s (62)), and still smaller than as it flows. The theory does not hold for molecules that do
that of proteins (~108 cm2/s (63)) as measured in nerve not conform to this assumption, e.g., proteins that bind to
cells. Thus, in the absence of bias, the cargo-induced motion intracellular structures.
considered here does not contribute appreciably to the over- The second major assumption is that the axonal medium
all transport over very long times. However, as discussed in is fluid, i.e., it should flow in response to the active motion of
the preceding section, it can occasionally make molecules the cargo. Such flows were observed in response to the
move ballistically over micron-scale distances—a type of forced motion of magnetic particles (65). Evidently, the
transport impossible to achieve through thermal diffusion. actual axoplasm is crowded, viscoelastic, and anisotropic
On the other hand, a bias in the direction of cargo motion (65). We have focused in this work on the simple question
along the microtubules will create a net drift resulting in the of whether the passive drag due to cargo motion is at all rele-
accumulation of net displacement in the direction of the vant or negligible as regards the SC of axonal transport.
bias. Experimentally, a previous report indicates significant Within this modest goal, the orders of magnitude presented
biases ranging from 10–30% for some cargoes, and up to above should be valid for actual axons, based on the

a 0.35 b 0.7
0.6
0.3 #2
0.5
FIGURE 10 Two demonstrative trajectories of
0.25
#3 0.4 fluid particles along the axonal direction, as
0.3 induced by the motions of three cargoes. Parame-
z [μm]

0.2
z [μm]

#1 #3
0.2 ters: R ¼ 1 mm, a/R ¼ 0.1; cargo coordinates:
0.15
0.1 (r1, 41) ¼ (0.5R, 0), (r2, 42) ¼ (0.1R, p) moving
0.1 0
#1 in the opposite direction, and (r3, 43) ¼ (0.4R,
-0.1 p); fluid particle coordinates: (ra, 4a) ¼ (0.345R,
0.05 #2
-0.2 0), (rb, 4b) ¼ (0.375R, p). To see this figure in
0 -0.3 color, go online.
0 2 4 6 8 10 12 0 2 4 6 8 10 12
time [s] time [s]

Biophysical Journal 106(12) 2710–2719


Cytosol Drag in Neurons 2717

following arguments. 1), The complex forces exerted on the contribution of active mechanisms that may induce
the cargoes by molecular motors, and their resulting veloc- microstreaming. Experiments that involve fixation of cells
ities—the well-documented fast component of axonal trans- or blockage of various possible active processes (35) can
port—are not addressed theoretically but are rather treated provide the needed discrimination between diffusion and
as external biological constraints imposed on our calcula- microstreaming. Only with additional experiments of this
tions of the drag of the much smaller dissolved proteins. sort will we be able to quantify the relative contribution of
2), It is well established that confinement in a long tube thermal diffusion and active mechanisms in defining the
must screen out hydrodynamic interactions over long dis- SC and intracellular displacement of soluble proteins.
tances—an effect that is directly related to the lack of The implications of water displacement in neurons extend
momentum conservation in the confined fluid and remains beyond the basic understanding of transport and axonal
valid regardless of the complexity of the fluid. 3), The envi- physiology. Diffusion tensor imaging (DTI) uses water
ronmental response to an object moving through the axon displacement in the brain as a unique tool for diagnosis
depends on size. For objects of molecular size, such as pro- and research (70). Water displacement in neuronal projec-
teins, the medium’s response was found to be close to that of tions is anisotropic (with preferred displacement in the lon-
a simple solvent (66). 4), Concerning the long-distance flow gitudinal direction), and significantly drops during events of
induced by cargo motion, viscoelastic effects should set in metabolic or oxidative stress, as occurs, for example,
at sufficiently large length- and time-scales. The only mea- in cerebral artery occlusion (71). DTI measures water
surement of axoplasm viscoelasticity of which we are aware displacement over typical timescales of 1–100 ms. Our
(65) gave comparable values for the storage and loss moduli calculation implies that, contrary to a previous suggestion
over timescales of order 103 s. Measurements in the cyto- (72), within these timescales, the displacement of water
plasm of other cells (macrophages) (67) yielded a storage molecules, caused by drag due to transported cargoes, is
modulus of the same order as, or one order of magnitude too minor to be detected by DTI. On a more general level,
larger than, the loss modulus for timescales of order 10 s. it should be noted that some of the displacement of water
Related materials such as F-Actin networks are known to molecules, as well as the drop in displacement detected after
have a storage modulus that increases with time, where tissue insults, may be due to blockage of intracellular micro-
the crossover from viscous to elastic behavior takes place streaming. The existence of such active microstreaming, as
at ~1–10 s (68). Thus, the relevant timescale of cargo mo- well as its biophysical basis, should be further investigated.
tion, ~1 s, may be around that crossover; we expect visco- Transport mechanisms similar to those proposed here
elastic effects to be present but not very large. Clearly, might be relevant to the mixing and distribution of solutes
more accurate experimental data for the viscoelasticity in other scenarios. A particular example is that of cyto-
of the axoplasm is required. We further note that in the plasmic streaming in plant cells (56,73–75). Organelle-
noninertial (zero Reynolds number) limit considered here induced flow is observed in many plant cells in which
for given cargo velocities, the results for the displace- different flow patterns with variable velocity distributions
ments (rather than stresses) of the dragged molecules are have been identified. Nonetheless, organelle-induced flow
independent of the viscosity. Qualitatively, in a viscoelastic is expected to be most prominent in cells of anisotropic
medium, they should be similarly insensitive to the fre- cytoskeleton, where there are displacements over long
quency-dependent moduli. distances. These conditions are met in neurons.
Another issue, addressed in detail in the Results section,
is the effect of boundary conditions along the axonal mem- We thank S. Bhattacharya, A. Leshansky and E. Perlson for helpful
discussions, and A. Liberman for his assistance.
brane, namely, whether or not there is significant slip at the
boundary. We find that slip may be either unimportant or H.D. acknowledges support from the Israel Science Foundation (grant no.
8/10). U.N. acknowledges support from a European Union Marie Curie
important for axonal transport, depending on whether the Institutional Research grant (MMDTIAN) and an Israel Science Foundation
axon should be described as effectively closed or open- grant (1156/12).
ended, respectively (compare Figs. 3 a and 8). This issue,
in turn, depends on the amount of water exchange across
the axonal membrane. Thus, the issues of boundary slip REFERENCES
and membrane permeability are intertwined and should be
1. Brown, A. 2003. Axonal transport of membranous and nonmembra-
clarified in future experiments to obtain a more quantitative nous cargoes: a unified perspective. J. Cell Biol. 160:817–821.
description of the transport studied here.
2. Siegel, G. 2006. Basic Neurochemistry: Molecular, Cellular and
An important implication of this study relates to the Medical Aspects. Elsevier Academic Press, Amsterdam.
interpretation of nonlocal measurements of intracellular
3. Hirokawa, N., S. Niwa, and Y. Tanaka. 2010. Molecular motors in
diffusion of various proteins, such as by FRAP or diffu- neurons: transport mechanisms and roles in brain function, develop-
sion-weighted NMR (69). It should be taken into account ment, and disease. Neuron. 68:610–638.
that intracellular diffusion coefficients extracted from such 4. Bearer, E. L., and T. S. Reese. 1999. Association of actin filaments with
measurements in vital cells may be overestimated due to axonal microtubule tracts. J. Neurocytol. 28:85–98.

Biophysical Journal 106(12) 2710–2719


2718 Mussel et al.

5. Fath, K. R., and R. J. Lasek. 1988. Two classes of actin microfilaments protein MAP 1A in mouse retinal ganglion cells: tubulin and MAP 1A
are associated with the inner cytoskeleton of axons. J. Cell Biol. display distinct transport kinetics. J. Cell Biol. 110:437–448.
107:613–621. 29. Yuan, A., R. G. Mills, ., J. J. Bray. 1999. Cotransport of glyceralde-
6. Bray, D. 2001. Cell Movements: From Molecules to Motility, 2nd ed. hyde-3-phosphate dehydrogenase and actin in axons of chicken moto-
Garland, New York. neurons. Cell. Mol. Neurobiol. 19:733–744.
7. Hawkins, T., M. Mirigian, ., J. L. Ross. 2010. Mechanics of 30. Bourke, G. J., W. El Alami, ., M. J. Carden. 2002. Slow axonal trans-
microtubules. J. Biomech. 43:23–30. port of the cytosolic chaperonin CCT with Hsc73 and actin in motor
8. Sabry, J., T. P. O’Connor, and M. W. Kirschner. 1995. Axonal neurons. J. Neurosci. Res. 68:29–35.
transport of tubulin in Ti1 pioneer neurons in situ. Neuron. 14:1247– 31. Roy, S., M. J. Winton, ., V. M. Y. Lee. 2008. Cytoskeletal re-
1256. quirements in axonal transport of slow component-b. J. Neurosci.
9. Chevalier-Larsen, E., and E. L. F. Holzbaur. 2006. Axonal transport 28:5248–5256.
and neurodegenerative disease. Biochim. Biophys. Acta. 1762:1094– 32. Wang, L., C. L. Ho, ., A. Brown. 2000. Rapid movement of axonal
1108. neurofilaments interrupted by prolonged pauses. Nat. Cell Biol.
10. Lasek, R. J. 1980. Axonal transport: a dynamic view of neuronal struc- 2:137–141.
tures. Trends Neurosci. 3:87–91. 33. Roy, S., P. Coffee, ., M. M. Black. 2000. Neurofilaments are trans-
11. Sjöstrand, J. 1970. Fast and slow components of axoplasmic ported rapidly but intermittently in axons: implications for slow axonal
transport in the hypoglossal and vagus nerves of the rabbit. Brain transport. J. Neurosci. 20:6849–6861.
Res. 18:461–467. 34. Roy, S., M. J. Winton, ., V. M. Lee. 2007. Rapid and inter-
12. Willard, M., W. M. Cowan, and P. R. Vagelos. 1974. The polypeptide mittent cotransport of slow component-b proteins. J. Neurosci.
composition of intra-axonally transported proteins: evidence for four 27:3131–3138.
transport velocities. Proc. Natl. Acad. Sci. USA. 71:2183–2187. 35. Scott, D. A., U. Das, ., S. Roy. 2011. Mechanistic logic underlying the
axonal transport of cytosolic proteins. Neuron. 70:441–454.
13. Willard, M. B., and K. L. Hulebak. 1977. The intra-axonal transport of
polypeptide H: evidence for a fifth (very slow) group of transported 36. Miller, K. E., and S. R. Heidemann. 2008. What is slow axonal trans-
proteins in the retinal ganglion cells of the rabbit. Brain Res. port? Exp. Cell Res. 314:1981–1990.
136:289–306. 37. Blum, J. J., and M. C. Reed. 1989. A model for slow axonal transport
14. Black, M. M., and R. J. Lasek. 1980. Slow components of axonal trans- and its application to neurofilamentous neuropathies. Cell Motil.
port: two cytoskeletal networks. J. Cell Biol. 86:616–623. Cytoskeleton. 12:53–65.
15. Lasek, R. J., J. A. Garner, and S. T. Brady. 1984. Axonal transport of 38. Brown, A., L. Wang, and P. Jung. 2005. Stochastic simulation of neuro-
the cytoplasmic matrix. J. Cell Biol. 99:212s–221s. filament transport in axons: the ‘‘stop-and-go’’ hypothesis. Mol. Biol.
Cell. 16:4243–4255.
16. Allen, R. D., J. Metuzals, ., S. P. Gilbert. 1982. Fast axonal transport
in squid giant axon. Science. 218:1127–1129. 39. Craciun, G., A. Brown, and A. Friedman. 2005. A dynamical system
model of neurofilament transport in axons. J. Theor. Biol. 237:316–322.
17. Vale, R. D., B. J. Schnapp, ., M. P. Sheetz. 1985. Movement of organ-
elles along filaments dissociated from the axoplasm of the squid giant 40. Jung, P., and A. Brown. 2009. Modeling the slowing of neurofilament
axon. Cell. 40:449–454. transport along the mouse sciatic nerve. Phys. Biol. 6:046002.
18. Schnapp, B. J., R. D. Vale, ., T. S. Reese. 1986. Microtubules and the 41. Kuznetsov, A. V. 2010. Analytical solution of equations describing
mechanism of directed organelle movement. Ann. N. Y. Acad. Sci. slow axonal transport based on the stop-and-go hypothesis. Cent.
466:909–918. Eur. J. Phys. 9:662–673.
19. Lalli, G., and G. Schiavo. 2002. Analysis of retrograde transport in 42. Nothnagel, E. A., and W. W. Webb. 1982. Hydrodynamic models of
motor neurons reveals common endocytic carriers for tetanus toxin viscous coupling between motile myosin and endoplasm in characean
and neurotrophin receptor p75NTR. J. Cell Biol. 156:233–239. algae. J. Cell Biol. 94:444–454.
20. Hill, D. B., M. J. Plaza, ., G. Holzwarth. 2004. Fast vesicle 43. Wolff, K., D. Marenduzzo, and M. E. Cates. 2012. Cytoplasmic
transport in PC12 neurites: velocities and forces. Eur. Biophys. J. streaming in plant cells: the role of wall slip. J. R. Soc. Interface.
33:623–632. 9:1398–1408.
21. Cui, B., C. Wu, ., S. Chu. 2007. One at a time, live tracking of NGF 44. Goldstein, R. E., I. Tuval, and J. W. van de Meent. 2008. Microfluidics
axonal transport using quantum dots. Proc. Natl. Acad. Sci. USA. of cytoplasmic streaming and its implications for intracellular trans-
104:13666–13671. port. Proc. Natl. Acad. Sci. USA. 105:3663–3667.
22. Howard, J. 2001. Mechanics of Motor Proteins and the Cytoskeleton. 45. van de Meent, J. W., I. Tuval, and R. E. Goldstein. 2008. Nature’s
Sinauer Associates, Sunderland, MA. microfluidic transporter: rotational cytoplasmic streaming at high
Péclet numbers. Phys. Rev. Lett. 101:178102.
23. Karcher, R. L., S. W. Deacon, and V. I. Gelfand. 2002. Motor-cargo
interactions: the key to transport specificity. Trends Cell Biol. 46. Bray, D., and M. B. Bunge. 1981. Serial analysis of microtubules in
12:21–27. cultured rat sensory axons. J. Neurocytol. 10:589–605.
24. Hirokawa, N., and R. Takemura. 2005. Molecular motors and 47. Happel, J. R., and H. Brenner. 1983. Low Reynolds Number
mechanisms of directional transport in neurons. Nat. Rev. Neurosci. Hydrodynamics: With Special Applications to Particulate Media.
6:201–214. Springer, New York.
25. Black, M. M., and R. J. Lasek. 1979. Axonal transport of actin: slow 48. Bungay, P. M., and H. Brenner. 1973. The motion of a closely-fitting
component b is the principal source of actin for the axon. Brain Res. sphere in a fluid-filled tube. Int. J. Multiphase Flow. 1:25–56.
171:401–413. 49. Brenner, H., and L. J. Gaydos. 1977. The constrained brownian move-
26. Brady, S. T., M. Tytell, ., R. J. Lasek. 1981. Axonal transport of ment of spherical particles in cylindrical pores of comparable radius.
calmodulin: a physiologic approach to identification of long-term asso- J. Colloid Interface Sci. 58:312–356.
ciations between proteins. J. Cell Biol. 89:607–614. 50. Liron, N., and R. Shahar. 1978. Stokes flow due to a Stokeslet in a pipe.
27. Baitinger, C., and M. Willard. 1987. Axonal transport of synapsin J. Fluid Mech. 86:727–744.
I-like proteins in rabbit retinal ganglion cells. J. Neurosci. 7:3723– 51. Blake, J. R. 1979. On the generation of viscous toroidal eddies in
3735. a cylinder. J. Fluid Mech. 95:209–222.
28. Nixon, R. A., I. Fischer, and S. E. Lewis. 1990. Synthesis, axonal trans- 52. Diamant, H. 2009. Hydrodynamic interaction in confined geometries.
port, and turnover of the high molecular weight microtubule-associated J. Phys. Soc. Jpn. 78:041002.

Biophysical Journal 106(12) 2710–2719


Cytosol Drag in Neurons 2719

53. Eames, I., D. Gobby, and S. B. Dalziel. 2003. Fluid displacement by 65. Sato, M., T. Z. Wong, ., R. D. Allen. 1984. Rheological properties
Stokes flow past a spherical droplet. J. Fluid Mech. 485:67–85. of living cytoplasm: a preliminary investigation of squid axoplasm
54. Baslow, M. H. 1999. The existence of molecular water pumps in the (Loligo pealei). Cell Motil. 4:7–23.
nervous system: a review of the evidence. Neurochem. Int. 34:77–90. 66. Haak, R. A., F. W. Kleinhans, and S. Ochs. 1976. The viscosity of
55. Navardi, S., and S. Bhattacharya. 2010. Axial pressure-difference be- mammalian nerve axoplasm measured by electron spin resonance.
tween far-fields across a sphere in viscous flow bounded by a cylinder. J. Physiol. 263:115–137.
Phys. Fluids. 22:103305.
67. Bausch, A. R., W. Möller, and E. Sackmann. 1999. Measurement of
56. Shimmen, T. 2007. The sliding theory of cytoplasmic streaming: fifty local viscoelasticity and forces in living cells by magnetic tweezers.
years of progress. J. Plant Res. 120:31–43. Biophys. J. 76:573–579.
57. Darwin, C. 1953. Note on hydrodynamics. Math. Proc. Cambridge 68. Gardel, M. L., M. T. Valentine, ., D. A. Weitz. 2003. Microrheology
Philos. Soc. 49:342–354. of entangled F-actin solutions. Phys. Rev. Lett. 91:158302.
58. Navardi, S., and S. Bhattacharya. 2010. A new lubrication theory to
derive far-field axial pressure difference due to force singularities in cy- 69. Norris, D. G. 2001. The effects of microscopic tissue parameters on the
lindrical or annular vessels. J. Math. Phys. 51:043102. diffusion weighted magnetic resonance imaging experiment. NMR
Biomed. 14:77–93.
59. Nadelhaft, I. 1974. Microtubule densities and total numbers in selected
axons of the crayfish abdominal nerve cord. J. Neurocytol. 3:73–86. 70. Basser, P. J., J. Mattiello, and D. LeBihan. 1994. Estimation of the
effective self-diffusion tensor from the NMR spin echo. J. Magn.
60. Barkus, R. V., O. Klyachko, ., W. M. Saxton. 2008. Identification
Reson. B. 103:247–254.
of an axonal kinesin-3 motor for fast anterograde vesicle transport
that facilitates retrograde transport of neuropeptides. Mol. Biol. Cell. 71. Moseley, M. E., Y. Cohen, ., P. R. Weinstein. 1990. Early detection of
19:274–283. regional cerebral ischemia in cats: comparison of diffusion- and T2-
61. Moughamian, A. J., G. E. Osborn, ., E. L. F. Holzbaur. 2013. Ordered weighted MRI and spectroscopy. Magn. Reson. Med. 14:330–346.
recruitment of dynactin to the microtubule plus-end is required for effi- 72. Nevo, U., E. Ozarslan, ., P. J. Basser. 2010. A system and mathemat-
cient initiation of retrograde axonal transport. J. Neurosci. 33:13190– ical framework to model shear flow effects in biomedical DW-imaging
13203. and spectroscopy. NMR Biomed. 23:734–744.
62. Beaulieu, C., and P. S. Allen. 1994. Water diffusion in the giant axon of
73. Allen, N. S., and R. D. Allen. 1978. Cytoplasmic streaming in green
the squid: implications for diffusion-weighted MRI of the nervous
plants. Annu. Rev. Biophys. Bioeng. 7:497–526.
system. Magn. Reson. Med. 32:579–583.
63. Tsuriel, S., R. Geva, ., N. E. Ziv. 2006. Local sharing as a predomi- 74. Kachar, B., and T. S. Reese. 1988. The mechanism of cytoplasmic
nant determinant of synaptic matrix molecular dynamics. PLoS Biol. streaming in characean algal cells: sliding of endoplasmic reticulum
4:e271. along actin filaments. J. Cell Biol. 106:1545–1552.
64. Overly, C. C., H. I. Rieff, and P. J. Hollenbeck. 1996. Organelle motility 75. Sugi, H., and S. Chaen. 2003. Force-velocity relationships in actin-
and metabolism in axons vs dendrites of cultured hippocampal neu- myosin interactions causing cytoplasmic streaming in algal cells.
rons. J. Cell Sci. 109:971–980. J. Exp. Biol. 206:1971–1976.

Biophysical Journal 106(12) 2710–2719

Anda mungkin juga menyukai