ArticlePDF AvailableLiterature Review

New Insights into the Role of the Cortical Cytoskeleton in Exocytosis from Neuroendocrine Cells

Authors:

Abstract and Figures

The cortical cytoskeleton is a dense network of filamentous actin (F-actin) that participates in the events associated with secretion from neuroendocrine cells. This filamentous web traps secretory vesicles, acting as a retention system that blocks the access of vesicles to secretory sites during the resting state, and it mediates their active directional transport during stimulation. The changes in the cortical cytoskeleton that drive this functional transformation have been well documented, particularly in cultured chromaffin cells. At the biochemical level, alterations in F-actin are governed by the activity of molecular motors like myosins II and V and by other calcium-dependent proteins that influence the polymerization and cross-linking of F-actin structures. In addition to modulating vesicle transport, the F-actin cortical network and its associated motor proteins also influence the late phases of the secretory process, including membrane fusion and the release of active substances through the exocytotic fusion pore. Here, we discuss the potential interactions between the F-actin cortical web and proteins such as SNAREs during secretion. We also discuss the role of the cytoskeleton in organizing the molecular elements required to sustain regulated exocytosis, forming a molecular structure that foments the efficient release of neurotransmitters and hormones.
No caption available
… 
Content may be subject to copyright.
Provided for non-commercial research and educational use only.
Not for reproduction, distribution or commercial use.
This chapter was originally published in the book International Review of Cell and
Molecular Biology, Vol. 295, published by Elsevier, and the attached copy is provided
by Elsevier for the author's benefit and for the benefit of the author's institution, for
non-commercial research and educational use including without limitation use in
instruction at your institution, sending it to specific colleagues who know you, and
providing a copy to your institution’s administrator.
All other uses, reproduction and distribution, including without limitation commercial
reprints, selling or licensing copies or access, or posting on open internet sites, your
personal or institution’s website or repository, are prohibited. For exceptions,
permission may be sought for such use through Elsevier's permissions site at:
http://www.elsevier.com/locate/permissionusematerial
From: Luis M. Gutiérrez, New Insights into the Role of the Cortical Cytoskeleton in
Exocytosis from Neuroendocrine Cells. In Kwang W. Jeon, editor:
International Review of Cell and Molecular Biology, Vol. 295,
Burlington: Academic Press, 2012, pp. 109-137.
ISBN: 978-0-12-394306-4
© Copyright 2012 Elsevier Inc.
Academic Press
CHAPTER THREE
New Insights into the Role of the
Cortical Cytoskeleton in Exocytosis
from Neuroendocrine Cells
Luis M. Gutie
´rrez
Contents
1. Introduction 110
2. Organization of Cortical Cytoskeleton 110
3. Changes Associated with Exocytosis 111
4. Forces Driving Cytoskeletal Changes Related to Exocytosis 113
4.1. Molecular motors influencing F-actin dynamics 113
4.2. Proteins modifying F-actin polymerization, fragmentation,
and cross-linking 114
5. Cytoskeletal Control of Vesicle Transport 116
5.1. Contribution of F-actin and microtubules to the
transport systems 116
5.2. Myosins II and V as molecular motors of vesicle motion 118
6. Role of Cytoskeleton in Membrane Fusion 119
6.1. Role of myosin II in exocytotic fusion 120
6.2. Role of filamentous actin in membrane fusion 121
6.3. Interactions between cytoskeleton and secretory machinery 122
6.4. Association of F-actin cortex with ion channels involved in
exocytosis 124
7. Cortical Cytoskeleton as an Organizer of Molecular
Cytoarchitecture of Exocytosis 125
8. Conclusion 128
Acknowledgments 128
References 128
Abstract
The cortical cytoskeleton is a dense network of filamentous actin (F-actin) that
participates in the events associated with secretion from neuroendocrine cells.
This filamentous web traps secretory vesicles, acting as a retention system that
International Review of Cell and Molecular Biology, Volume 295 #2012 Elsevier Inc.
ISSN 1937-6448, DOI: 10.1016/B978-0-12-394306-4.00009-5 All rights reserved.
Instituto de Neurociencias, Centro Mixto Universidad Miguel Herna
´ndez-CSIC, Sant Joan d’Alacant,
Alicante, Spain
109
Author's personal copy
blocks the access of vesicles to secretory sites during the resting state, and it
mediates their active directional transport during stimulation. The changes in
the cortical cytoskeleton that drive this functional transformation have been
well documented, particularly in cultured chromaffin cells. At the biochemical
level, alterations in F-actin are governed by the activity of molecular motors like
myosins II and V and by other calcium-dependent proteins that influence the
polymerization and cross-linking of F-actin structures. In addition to modulating
vesicle transport, the F-actin cortical network and its associated motor proteins
also influence the late phases of the secretory process, including membrane
fusion and the release of active substances through the exocytotic fusion pore.
Here, we discuss the potential interactions between the F-actin cortical web and
proteins such as SNAREs during secretion. We also discuss the role of the
cytoskeleton in organizing the molecular elements required to sustain regu-
lated exocytosis, forming a molecular structure that foments the efficient
release of neurotransmitters and hormones.
Key Words: Cytoarchitecture, Neuroendocrine cells, Chromaffin vesicles,
Exocytosis, F-actin, Myosins, Phosphorylation, Vesicle transport,
SNAREs. ß2012 Elsevier Inc.
1. Introduction
Cytoskeletal structures play fundamental roles in many aspects of cell
biology, influencing both basic structure and functional specialization. At
the cell periphery in particular, the F-actin cortical cytoskeleton influences
the cell’s shape, topology, motility, and signaling. In endocrine and neuro-
endocrine cells, this dense network of actin filaments not only provides
architectural support but it also entraps secretory vesicles that contain
neurotransmitters and hormones, playing an important regulatory role in
the exocytotic process. In this review, we summarize our current under-
standing of the cortical cytoskeleton’s role in the physiology of neuroendo-
crine cells. Specifically, we focus on chromaffin cells, an important model
that has significantly influenced our current understanding of the role of the
cortical F-actin network in regulated exocytosis.
2. Organization of Cortical Cytoskeleton
In the early 1980s, a number of studies using immunofluorescence
microscopy techniques demonstrated that cytoskeletal proteins such as
actin, caldesmon, and myosin localize to the cortex of cultured chromaffin
cells, forming a peripheral ring (Aunis and Bader, 1988;Aunis et al., 1980;
110 Luis M. Gutie
´rrez
Author's personal copy
Burgoyne et al., 1986;Lee and Trifaro, 1981;Trifaro et al., 1985). This
distribution was consistent with earlier descriptions of a dense web of
F-actin beneath the plasma membrane in other endocrine systems, such as
pancreatic b-cells (Orci et al., 1972). Accordingly, it was inferred that this
dense cortical cytoskeleton was continuous throughout the cell periphery,
preventing secretory vesicles having free access to the plasma membrane
(Aunis and Bader, 1988;Burgoyne and Cheek, 1987;Trifaro et al., 1985).
The three-dimensional ultrastructural architecture of this cortical net-
work was elegantly demonstrated using quick-freeze, deep etch electronic
microscopy (Nakata and Hirokawa, 1992). This technique revealed that
F-actin filament bundles were heterogeneously distributed, coinciding with
the plasma membrane in some areas but being sparse in others. The contin-
uous nature of the cortex was a source of some controversy, even in
unstimulated cells, particularly given the potential of chemical fixation
and low temperature procedures that accompanied fluorescence and
electronic microscopy to affect the integrity of the “native” cortex.
The use of the transmitted light channel of confocal scanning microscopes
in conjunction with high numerical aperture objectives provided an alterna-
tive means of analyzing the cortical network, based on the differential refrac-
tion index presented by F-actin in the gel and sol states (Giner et al., 2005). In
living intact chromaffin cells, this technique revealed the cortex as an intricate
meshwork of cytoskeletal cages, which contained infrequent openings that
permit vesicles to access the membrane. Subsequentthree-dimensional recon-
structions of transmitted light images of the cortical area of unstimulated living
cells revealed numerous areas devoid of the cytoskeleton that resemble “cra-
ters” covering the cell’s surface (Fig. 3.1A; Giner et al., 2007).
3. Changes Associated with Exocytosis
Immunocytochemical analyses provided the first evidence that the
cortical cytoskeleton was a dynamic structure, demonstrating the redistri-
bution of fodrin, a spectrin-like protein, after cell stimulation (Perrin and
Aunis, 1985). Indeed, when F-actin was visualized with phalloidin-coupled
rhodamine, rapid fragmentation after cell stimulation with nicotine was
evident (Cheek and Burgoyne, 1986). The same authors later reported
that different secretagogues decrease the F-actin/G-actin ratio and they
suggested that this change is necessary prior to exocytosis (Burgoyne and
Cheek, 1987). The initial reports of cortical fragmentation were followed
by other studies using conventional epifluorescence (Marxen and Bigalke,
1991;Vitale et al., 1991) and confocal microscopy (Gil et al., 2000;Nakata
and Hirokawa, 1992). The hypothesis that F-actin disassembly was linked to
secretion was further investigated using a variety of natural chemicals that
Role of Cortical Cytoskeleton in Exocytosis 111
Author's personal copy
stabilize (e.g., jasplakinolide, phallotoxins) or disrupt polymerization (e.g.,
cytochalasins, latrunculins, clostridium C2 and iota toxins, snake venom
toxins) in a variety of cellular systems, including pancreatic b-cells (Orci
et al., 1972), melanotrophs (Chowdhury et al., 1999), mast cells (Pendleton
and Koffer, 2001), lactotropes (Carbajal and Vitale, 1997), acinar cells
(Jerdeva et al., 2005), chromaffin cells (Gasman et al., 2004;Gil et al.,
2000), and the related PC12 line (Matter et al., 1989).
Many of these studies demonstrated that the role of F-actin in exocytosis was
far more complex than that of a simple retention system that prevents granule
access and fusion. For example, by promoting F-actin depolymerization with
0 s
10 s
50 s
1–5 s
10–30 s
40–80 s
C
ABD
Figure 3.1Structure of the cortical cytoskeleton and its response to stimulation in
chromaffin cells. (A) Three-dimensional reconstruction of the F-actin cortex of a
cultured chromaffin cell based on transmitted light images from different planes. The
image shows the characteristic network and cages formed by the F-actin cortex in the
resting state. (B) Magnitude of the stimulation-induced changes. In the same cell
depicted in panel A, open spaces devoid of F-actin are generated by cell depolarization
for 30s. (C) Phases of the cytoskeletal changes. The scheme shows the different phases
of F-actin reorganization observed in a typical cell when stimulated. From the resting
state, cortical disruption occurs soon (1–5s) after stimulation, leading to the formation
of wider cytoskeletal depleted areas (10–30s) and the subsequent regeneration of the
cortical barrier. (D) Example of the cortical structures seen during stimulation. This
panel shows transmitted light images of the F-actin cortex in resting and stimulated
states. During stimulation, channel-like structures conduct vesicles to the secretory
sites. Scale bars¼10mm (panel A) and 1mm (panel D).
112 Luis M. Gutie
´rrez
Author's personal copy
the botulinum toxin C2, low levels of F-actin ADP-ribosylation were seen to
promote secretion, while very high levels significantly reduced exocytosis,
indicating an active role for F-actin during secretion (Matter et al., 1989).
Moreover, analyzing vesicle motion in the cortical area by evanescent light
microscopy revealed that, during stimulation, F-actin filaments mediate the
active transport of vesicles within 300nm of the cell’s limits (Lang et al., 2000;
Oheim and Stuhmer, 2000).
How does the F-actin cortex accomplish this dual role? This was
recently revealed in simultaneous transmitted light scanning and confocal
fluorescent microscopy studies of F-actin structural dynamics and granule
motion (Giner et al., 2005). This technique revealed major reversible
changes in the cortical cytoskeleton during stimulation, as illustrated in
the three-dimensional reconstructions in Fig. 3.1. Hence, the resting state
is characterized by a dense cytoskeletal web (Fig. 3.1A) that reorganizes to
reveal extensive areas devoid of F-actin (Fig. 3.1B). Dynamic changes occur
a few seconds after stimulation, with the formation of cortical disruptions
and channel-like structures (Fig. 3.1D), which allow vesicles access to
docking sites. If the stimulus persists, these alterations can continue for up
to 40s, with the formation of extended subplasmalemmal areas devoid of
cytoskeleton. The cortex finally reorganizes to impede vesicle from acces-
sing the plasma membrane (see Fig. 3.1C and D). In this cortical remodeling
scenario, F-actin transiently redistributes in a calcium-dependent manner,
acting as a structural barrier and a functional vesicle carrier during the
secretory cycle. The formation of cytoplasmic areas devoid of F-actin
during prolonged stimulation was previously reported in electron micros-
copy studies (Nakata and Hirokawa, 1992;Tchakarov et al., 1998).
4. Forces Driving Cytoskeletal Changes
Related to Exocytosis
The dynamic nature of the cortical network suggests that F-actin
plasticity must be exquisitely controlled to adapt to the dual effects of
stimulation. This control may be dependent on intracellular calcium levels
and accomplished by regulating actin activity via two different elements:
molecular motors such as specific myosin subtypes that control actin
dynamics and a variety of auxiliary proteins that regulate polymerization,
fragmentation, and F-actin cross-linking.
4.1. Molecular motors influencing F-actin dynamics
Many cellular processes depend upon the mechanical properties of the
cytoskeleton, such as cell adhesion and motility, in which cytoplasmic
reorganization requires the contractile force provided by actomyosin
Role of Cortical Cytoskeleton in Exocytosis 113
Author's personal copy
complexes (Bershadsky et al., 2006;Borisy and Svitkina, 2000). Cortical
reorganization appears to be dependent on contractile systems, and indeed,
the motion of cytosolic cytoskeletal structures under scanning transmitted
light microscopy is reduced by wortmannin and 2,3-butanodione mono-
xime (BDM), inhibitors of myosin II (Giner et al., 2005). Moreover,
chromaffin cells contain cortical myosin II (Trifaro et al., 1984), a non-
processive motor that binds to F-actin and undergoes a conformational
change after ATP hydrolysis. This protein self-associates to form minifila-
ments (Verkhovsky et al., 1999), which act as “disordered micromuscles”
after binding to F-actin bundles, producing local contractions in the cortical
meshwork. The expression of an inactive nonphosphorylatable form of the
regulatory light chain of myosin II inhibits cytoskeletal motion in the
chromaffin cell cytosol (Neco et al., 2004), suggesting that the
phosphorylation–dephosphorylation cycle sustains F-actin cytoskeletal
dynamics. Accordingly, calcium-dependent phosphorylation of the myosin
II light chain accompanies the secretory cycle (Cote et al., 1986;Gutierrez
et al., 1988, 1989), while inhibition of the myosin light chain kinase
(MLCK) partially blocks secretion from intact and permeabilized chromaf-
fin cells (Kumakura et al., 1994;Reig et al., 1993).
Among other factors regulating contractile force that have been detected
in chromaffin granule preparations (Burgoyne et al., 1986), caldesmon is a
protein that stabilizes actin filaments against actin-severing proteins and
inhibits actomyosin ATPase activity. The inhibitory activity of caldesmon
is regulated by calcium and ATP through phosphorylation (Wang, 2008),
and when the calcium concentration is elevated to the levels reached during
the secretory process in chromaffin cells, caldesmon and tropomyosin
dissociate from F-actin gels (Cheek et al., 1986).
Molecular motors, such as myosin II, are abundant in the cell cortex
of endocrine and neuroendocrine cells, cross-linking F-actin in an ATP-
dependent manner. These proteins change the physical nature of the cyto-
solic sol–gel to that of an “active gels,” dynamic structures that induce
the instability of cortical actin ( Joanny and Prost, 2009), and they affect
organelle motion in a manner that will be discussed later.
4.2. Proteins modifying F-actin polymerization,
fragmentation, and cross-linking
Fodrin, a neural form of spectrin, has long been known to be a key factor in
cortical fragmentation (Perrin and Aunis, 1985). This protein cross-links
F-actin at low concentrations of intracellular calcium, but when cells are
stimulated and the intracellular calcium levels rise, fodrin dissociates from
actin and there is a local dissolution of the cortex. The role of this protein in
exocytosis was further demonstrated when calcium-induced secretion from
114 Luis M. Gutie
´rrez
Author's personal copy
digitonin-permeabilized cells was partially inhibited by purified antibodies
against this protein (Perrin et al., 1987).
The search for factors that regulate the disassembly of chromaffin cortex
upon stimulation led to the discovery of the calcium-dependent F-actin
severing protein, scinderin (Rodriguez del Castillo et al., 1990). Scinderin is
a protein localized in the subplasmalemmal region of chromaffin cells that is
found in a variety of secretory tissues including the brain, pituitary,and salivary
glands, as well as in platelets, and it alters the viscosity of F-actin gels. Interest-
ingly, scinderin is redistributed during chromaffin cell stimulation, following
the pattern of F-actin fragmentation, which is a feature not attributed to gelsolin
(Vitale et al., 1991). Scinderin contains 2 actin and 2 phosphatidylinositol 4,5
bisphosphate-binding domains (PIP2) and a relatively high degree of homol-
ogy to the gelsolin and villin genes (Marcu et al., 1994). The role of scinderin in
exocytosis was witnessed by the increased calcium-induced secretory response
of permeabilized chromaffin cells induced by recombinant scinderin, an effect
blocked by peptides competing with the actin PIP2-binding domains (Zhang
et al., 1996). These findings suggest that in resting conditions, PIP2 binding to
scinderin sequesters the protein in the plasma membrane, the correct cortical
location for this protein. Cell stimulation and the subsequent increase
in intracellular calcium activate actin-severing activity and induce F-actin
fragmentation in the region just below the plasma membrane.
The control of F-actin polymerization (Yin and Janmey, 2003) and the
regulation of the cytoskeletal–plasma membrane interactions (Meiri, 2004)
are associated with the production of phosphatidylinositol and its phos-
phorylated derivatives. In permeabilized chromaffin cells, manipulating
phosphoinositide levels with phospholipase C implicated phosphatidylino-
sitol polyphosphates in exocytosis (Eberhard et al., 1990). Two enzymes
that maintain the PIP2 levels, the phosphatidylinositol transfer protein and
phosphatidylinositol 4-kinase, are also required to sustain secretion in PC12
cells (Hay and Martin, 1993;Hay et al., 1995).
In terms of the signaling pathway that mediates PIP2 generation in
chromaffin and other secretory cells, a number of findings point to the
central role of the Rho GTPase family in this process. Specifically, RhoA is
associated with secretory granules and it is a downstream partner of vesicu-
lar-associated Go, a trimeric GTPase that regulates the priming steps of
exocytosis (Gasman et al., 1997;Vitale et al., 1996). Activation of RhoA via
the Go pathway (Gasman et al., 1997), or expression of a constitutively
activated GTP-RhoA mutant, impairs secretion and stabilizes the F-actin
cortical barrier in PC12 (Frantz et al., 2002) and chromaffin cells (Bader
et al., 2004). The phosphatidylinositol 4-kinase enzyme that associates with
chromaffin granules is thought to be the effector mediating RhoA inhibi-
tion of secretion (Gasman et al., 1998). Thus, the increase in PIP2 in the
granular membrane after RhoA activation may be involved in the nuclea-
tion of actin (Yin and Janmey, 2003), the regulation of cytoskeleton–plasma
Role of Cortical Cytoskeleton in Exocytosis 115
Author's personal copy
membrane interactions (Meiri, 2004), or even in the regulation of vesicle
motion in the vicinity of secretory sites (Holz and Axelrod, 2002).
Other GTPases may regulate actin dynamics by promoting F-actin
polymerization. For example, the constitutively active GTP-loaded mutants
of Rac1 (Rac1L61) and Cdc42 (Cdc42L61) facilitate secretion in chromaf-
fin and PC12 cells, respectively, suggesting that these GTPases play an active
role in exocytosis (Gasman et al., 2004;Li et al., 2003). The active Cdc42
L61 mutant also enhances F-actin polymerization in the subplasmalemmal
region (Gasman et al., 2004), indicating that Cdc42 triggers F-actin forma-
tion in the vicinity of active sites, thereby enhancing secretion efficiency.
The annexin family of proteins may be particularly important in exocy-
tosis. These proteins are characterized by their calcium-sensing and phos-
pholipid-binding domains (Geisow et al., 1987), and some subtypes interact
with actin, such as types II and VI (Hayes et al., 2004). Translocation of
these proteins from the cytosol to the plasmalemma may dynamically link
these two structures (Mangeat, 1988). Annexin II and annexin VI (synexin)
activity in chromaffin cells is essential to sustain regulated exocytosis
(Burgoyne et al., 1991;Creutz et al., 1982;Pollard et al., 1982;Roth
et al., 1993), suggesting that the translocation of tetramers of these proteins
to the plasma membrane is an essential step in the formation of the lipid
microdomains required for exocytosis (Chasserot-Golaz et al., 2005).
5. Cytoskeletal Control of Vesicle Transport
Early studies investigating the influence of antimitotic drugs on adre-
nal gland secretion suggested a key role for microtubules in the exocytotic
transport of vesicles (Douglas and Sorimachi, 1972;Poisner and Bernstein,
1971). Thus, the relative importance of the two major cytoskeletal-
associated systems for organelle transport is central to understanding vesicle
transport in neuroendocrine cells.
5.1. Contribution of F-actin and microtubules to the
transport systems
The classical view proposes that cytoskeletal elements play a central role in the
transport of organelles from the inner cytoplasm to the cell periphery. In
neuroendocrine cells, immature vesicles follow the exocytotic pathway
through the endoplasmic reticulum to mature in different compartments of
the Golgi apparatus (Winkler, 1977). In these early phases, F-actin structures
support the architecture of the cisternae and they cooperate with microtubules
in transporting vesicles (Caviston and Holzbaur, 2006;DePina and Langford,
1999). In chromaffin cells, vesicle movement visualized by dynamic confocal
116 Luis M. Gutie
´rrez
Author's personal copy
microscopy appears to be sensitive to chemicals that affect microtubules in
regions adjacent to the chromaffin cell nucleus (Neco et al., 2003). By contrast,
F-actin inhibitors affect granule motionboth in the interior (Giner et al., 2005;
Neco et al., 2003) and in the cortical region (Allersma et al., 2006;Giner et al.,
2007;Lang et al., 2000;Oheim and Stuhmer, 2000). This observation may
reflect the distribution and density of these cytoskeletal elements since micro-
tubules have a radial distribution in chromaffin cells, concentrating in internal
regions and contacting the cell periphery tangentially (Neco et al., 2003).
Conversely, F-actin is localized in the cortical zone, and consequently, it
influences granule transport in this area. Accordingly, chemicals that disrupt
F-actin affect both the fast and slow secretory phases, whereas microtubule
inhibitors have a minor impact on secretion and they only affect the slow
secretory components recruited by repetitive stimulation (Neco et al., 2003).
Indeed, the microtubule-depolymerizing and stabilizing agents, colchicine
and nocodazole, inhibit the late and sustained phase of granule secretion in
b-pancreatic cells (Farshori and Goode, 1994). It is noteworthy that a variety
of antimitotic agents have a clear inhibitory effect on secretion in cultured
chromaffin cells (Cooke and Poisner, 1979;McKay et al., 1991), probably
reflecting their effects on the cytoskeleton.
The fundamental role of F-actin in directing or limiting the movement
of vesicles in the cortical region was clearly established in the first total
internal reflection fluorescence microscopy studies. Vesicle motion within
the cortical region can be described in detail by TIRFM, as the evanescent
field spans a region of 100–300nm from the cell–coverslip interface. This
technique reveals that the movement of chromaffin granules becomes more
restricted when approaching the plasma membrane ( Johns et al., 2001;Lang
et al., 1997;Oheim and Stuhmer, 2000;Steyer and Almers, 1999), princi-
pally due to two events: the interaction of vesicles with docking or tethering
elements (Lopez et al., 2009;Toonen et al., 2006) and vesicle entrapment in
the cytoskeletal network (Oheim and Stuhmer, 2000;Steyer and Almers,
1999). Recently, the use of transmission light scanning microscopy has
provided compelling evidence that vesicle motion is influenced by the
organization of the F-actin cortex into small polygonal cages (Giner et al.,
2007). Vesicles move in association with F-actin structures, and thus their
movement is restricted to the space available within the cages.
The consensus from multiple studies performed in neuronal, neuroendo-
crine, and endocrine cellular systems is that microtubules and kinesins are
important for the transport of secretory vesicles and other organelles from the
cell interior to the cortical region, where they are transferred to F-actin trails
(Fig. 3.2), the dominant system that controls the movement of vesicles within
the cell cortex. The possible interaction between the molecular motors of the
microtubule (kinesin) and F-actin-based (myosin V) transport systems may
create a “transportome” complex, facilitating the distribution of their cargo
(DePina and Langford, 1999).
Role of Cortical Cytoskeleton in Exocytosis 117
Author's personal copy
5.2. Myosins II and V as molecular motors of vesicle motion
Members of two classes of the myosin family are especially abundant in
chromaffin cells, myosins II and V (Neco et al., 2002;Rose et al., 2002).
Myosin II associates with F-actin and concentrates in the cell cortex, whereas
myosin V colocalizes with vesicles and it is therefore more widely distributed
in the cell cytoplasm. Class V myosins are strong candidates to participate in
vesicle transport. They possess a tail region designed for interaction with
cargo (Mooseker and Cheney, 1995), they are regulated by calcium (Sellers
1. Actin-mediated granule tensional pressure
2. Actin-mediated pore dilation
3. Direct influence on fusion machinery
Fusion
F-actin
Myosin II
Myosin Va
Microtubules
Kinesin
Maturation
Vesicle transport
2
1
3
Figure 3.2Cytoskeletal proteins involved in vesicle transport and the late phases of
the secretory process. Microtubules and kinesin are key players in the initial stages of
vesicle transport from the cell interior to the cortical zone. F-actin mediated transport
dominates in the cortex, characterized by two processes: F-actin dynamic remodeling
by myosin II and processive motion along F-actin trails by myosin V motors. These
molecular motors may play active roles during the maturation associated with exocyto-
sis. Membrane fusion is also influenced by F-actin-myosin II, and three possibilities
mechanisms are depicted: (1) exertion of a tensional pressure over the entire vesicle,
driving the release of the vesicular content; (2) F-actin-myosin II control of the
expansion of the exocytotic pore, acting as a cytoskeletal scaffold; (3) F-actin or myosin
motors directly affect the secretory machinery (SNARE and associated proteins).
118 Luis M. Gutie
´rrez
Author's personal copy
and Knight, 2007), and they associate with chromaffin (Rose et al., 2003)
and PC12 vesicles (Rudolf et al., 2003). Moreover, expression of a domi-
nant-negative tail of myosin V alters the distribution of vesicles, which
appear to dissociate from the cell cortex (Rudolf et al., 2003). The small
GTPase Rab27 and the MyRIP proteins appear to mediate the interaction
between myosin V, vesicles, and actin filaments (Desnos et al., 2003).
Consequently, altering the function of these proteins affects the movement
of the vesicles in the subplasmalemmal area. Nevertheless, myosin V does not
appear to be simply associated with vesicle transport, as it appears to partici-
pate in the maturation process (Kogel et al., 2010), and in the docking of
dense vesicles (Desnos et al., 2007).
In addition to myosin V, there is clear evidence of a role for myosin II in
the transport of vesicles in neuroendocrine systems. This conventional
nonmuscle myosin colocalizes with cortical F-actin (Neco et al., 2002;
Rose et al., 2003), and its inhibitors alter the motion of vesicles in the
cortical region, such as BDM and blebbistatin (Berberian et al., 2009;Neco
et al., 2002). This effect can be observed by confocal and evanescent wave
fluorescent microscopy, and moreover, an inactive nonphosphorylatable
mutant of the myosin II regulatory light chain drastically reduces vesicle
movement within the cytoplasm of chromaffin cells (Neco et al., 2004).
How does conventional myosin II influence vesicle motion? As men-
tioned earlier, myosin II binds to actin filaments and forms cross-linked
structures, the so-called active gels ( Joanny and Prost, 2009). These struc-
tures are static in the absence of ATP, but in its presence, they display
dynamic fluctuations and induce the movement of small colloidal spheres
(Mizuno et al., 2007). Indeed, expression of a ATP-insensitive light chain of
myosin II causes F-actin structures to become static and prevents vesicular
motion in the chromaffin cell cytoplasm (Neco et al., 2004).
Taken together, these findings indicate that the movement of secretory
vesicles in the cortex of neuroendocrine systems depends on the activity of
two molecular systems: nonprocessive motion mediated by the F-actin-
myosin II network and processive movement generated by the activity of
myosin V along F-actin trails (concepts summarized in Fig. 3.2, which also
illustrates the cooperation of F-actin structures and microtubules in the
interior of the cytosol to transport chromaffin granules from the inner
region, adjacent to the nucleus, to the peripheral cortex).
6. Role of Cytoskeleton in Membrane Fusion
While F-actin and myosin motors clearly participate in vesicle trans-
port, the role of cytoskeletal elements in exocytosis is far more complex than
solely ensuring the access of vesicles to docking sites. This first became
Role of Cortical Cytoskeleton in Exocytosis 119
Author's personal copy
evident through the effects of botulinum toxin C2, an F-actin ribosylating
agent, on secretion in PC12 cells (Matter et al., 1989), and subsequently, it
was shown that inhibitors of the MLCK affect the fast phases of exocytosis
associated with fusion of primed vesicles (Kumakura et al., 1994). These
findings were supported by studies on endocrine cells such as insulin-
secreting lines (Iida et al., 1997).
6.1. Role of myosin II in exocytotic fusion
The direct involvement of a member of the myosin family in vesicle fusion
during the final stages of secretion was demonstrated by analyzing single
fusion kinetics in bovine chromaffin cells, expressing a nonphosphorylatable
form of the regulatory light chain of myosin II (Neco et al., 2004). In these
cells, the kinetics of amperometric events was slowed by the expression of
this mutated myosin, which also reduced the rate of fusion of the ready
releasable pool of vesicles. Patch amperometry, probably the most efficient
technique to analyze membrane fusion kinetics (Albillos et al., 1997), was
later used to demonstrate a direct link between myosin II phosphorylation
and the rate of expansion of the fusion pore, the structure that initially
connects the intravesicular space with the cell exterior (Neco et al., 2008).
The role of the F-actin-myosin II contractile system in controlling single
vesicle fusion kinetics was further studied using specific inhibitors. This
study suggested that both the rate of expansion of the fusion pore and the
increase in its lifespan enhanced the degree of catecholamine dissociation
from the vesicular matrix (Berberian et al., 2009).
The complex role of the cortical cytoskeleton in controlling the final
phases of membrane fusion was elegantly demonstrated when kiss and run
and full collapse events were discriminated by altering the frequency of
electrical stimulation in mouse chromaffin cells (Doreian et al., 2008).
Accordingly, disrupting the F-actin cortical cytoskeleton through myosin
II phosphorylation favors the full collapse of chromaffin granules, whereas
kiss-and-run events are stabilized by F-actin structures. These conclusions
are supported by data from PC12 cells expressing vesicular proteins of
variable size, and hence with different diffusion rates through the fusion
pore (Aoki et al., 2010). Using TIRFM, myosin II was clearly seen to
control the duration of fusion pore opening. Taken together, these findings
indicate that F-actin-myosin II interactions govern vesicle transport and that
they are essential to ensure “normal” vesicular fusion kinetics, ending in full
collapse. The regulation of this contractile system appears to determine the
type of release: from incomplete kiss-and-run events to full fusion (the
possible molecular mechanisms underlying this regulatory activity are
described in Fig. 3.2). F-actin-myosin II may influence fusion by exerting
mechanical tension over the entire vesicle or alternatively (mechanism 1);
this tension may only affect a hypothetical cytoskeletal scaffold required for
120 Luis M. Gutie
´rrez
Author's personal copy
pore dilation (mechanism 2). By contrast, F-actin and myosin II (or other
myosin classes) molecules may directly affect the proteins responsible for the
formation and expansion of the fusion pore (probably SNARE proteins,
mechanism 3).
In addition to myosin II, several observations suggest that other classes of
myosin, such as myosins V, VI, and 1c/1e, fulfill similar roles in the final
stages of secretion (Bond et al., 2011). Thus, it should be stressed that a
variety of motor enzymes may play essential roles in fusion pore opening or
in driving the release of active compounds in diverse secretory systems.
6.2. Role of filamentous actin in membrane fusion
Exocytosis is inhibited when more than 60% of F-actin is sequestered by
ADP-ribosylation suggesting that a threshold of F-actin is required for
exocytosis (Matter et al., 1989). However, the source of this F-actin remains
unclear, yet it seems likely that after F-actin reorganization, sufficient
F-actin remains in the proximity of secretory sites to fulfill this role.
F-actin remodeling results in the redistribution but not the net destruction
of F-actin structures, and vesicles are associated with F-actin structures for
the duration of the secretory cycle (Giner et al., 2005). Further, although
exocytosis requires contact with the walls of cytoskeletal structures
(Torregrosa-Hetland et al., 2011), this does not preclude the possibility
that de novo F-actin polymerization occurs during secretion. In fact, in
pancreatic acinar cells, vesicles appear to be coated with new F-actin during
exocytosis (Valentijn et al., 2000), stabilizing the structures required sus-
taining exocytosis (Nemoto et al., 2004). However, a subsequent study
proposed that F-actin coating occurs following exocytosis, as active coating
only occurred in granules undergoing exocytotic pore formation (Turvey
and Thorn, 2004).
The possible existence of a molecular cascade leading to de novo F-actin
synthesis has been investigated in different secretory cells. The Rho GTPase
family may be involved in this cascade as Cdc42 is a positive regulator of
mast cell degranulation and it controls exocytosis in pancreatic b-cells
(Abdel-Latif et al., 2004;Hong-Geller and Cerione, 2000). In chromaffin
cells, constitutive activation of Cdc42 triggers the formation of F-actin
filaments in the subplasmalemmal area (Gasman et al., 2004), a process
accompanied by an increased secretory activity. This recruitment of
F-actin appears to be mediated by N-WASP and the Arp2/3 complex,
two factors that govern actin nucleation during the propulsion of secretory
and endocytic vesicles (Malacombe et al., 2006;Taunton et al., 2000).
Overexpression of N-WASP stimulates secretion in a manner that is depen-
dent on its F-actin polymerizing activity (Gasman et al., 2004).
There is as yet no consensus regarding the requirement of de novo F-actin
recruitment for exocytosis in secretory systems. Active coating has only
Role of Cortical Cytoskeleton in Exocytosis 121
Author's personal copy
been demonstrated in slow secretory systems characterized by large granules
or insoluble vesicles, a form of exocytosis termed “kiss and coat” (Sokac and
Bement, 2006). Further, it remains unclear whether these structures are
involved in generating extrusion forces or in the subsequent recovery of the
membrane by endocytosis. It has been suggested that F-actin coating may
simultaneously drive closure of the fusion pore and compression of the
secretory compartment (Sokac et al., 2003). Indeed, F-actin local assembly
was proposed to stabilize the secretory compartment during docking and
drive the posterior compensative endocytosis, triggered by a molecular
mechanism that detects the mixing of vesicular and plasma membranes
(Yu and Bement, 2007). Following the incorporation of diacylglycerol
from the plasmalemma, the bisoform of protein kinase C would be
recruited into the granular membrane, which would otherwise phosphory-
late Cdc 42. Finally, the Arp2/3 complex controls the exocyst in the final
phases of F-actin-mediated exocytosis (Zuo et al., 2006). The exocyst is a
multiprotein evolutionary conserved complex that, along with other possi-
ble protein complexes, mediates the tethering of vesicles to the plasma
membrane (Lipschutz and Mostov, 2002).
6.3. Interactions between cytoskeleton and
secretory machinery
Protein–protein interactions probably provide the specificity required for
granule docking, as well as contributing the essential machinery necessary
for vesicular fusion. These proteins include the plasma membrane proteins
(t-SNAREs) and their counterparts in the vesicular membrane (v-SNAREs:
Sollner et al., 1993a,b; Weber et al., 1998). Indeed, a ternary complex
consisting of syntaxin 1 (a t-SNARE), SNAP-25, and the vesicular synap-
tobrevin II constitutes the neuronal secretory machinery (Sollner et al.,
1993a,b) and has being described in chromaffin cells (Hodel et al., 1994;
Roth and Burgoyne, 1994) as well as in other endocrine and neuroendo-
crine secretory systems.
There is evidence that SNARE proteins interact with the F-actin
cytoskeleton, and, for example, the binding of syntaxin 4 to F-actin is
disrupted by stimulation in insulin-secreting cells (Thurmond et al.,
2003). F-actin readily binds to the first two coiled-coil domains of this
t-SNARE, apparently inducing granule accumulation prior to fusion
(Jewell et al., 2008). This mechanism is implicated in the targeting of
insulin-containing granules to secretory sites, and it involves the association
of cytoskeletal and SNARE proteins, and the small GTPase participating in
F-actin remodeling, Cdc42 (Daniel et al., 2002). Similarly, Cdc42 associates
with synaptobrevin II during vesicle transport, and a trimeric complex of
Cdc42, synaptobrevin II, and syntaxin 1 may arise (Nevins and Thurmond,
122 Luis M. Gutie
´rrez
Author's personal copy
2005). Associations between cytoskeletal elements and SNARE proteins
during transport are further demonstrated by the calcium-dependent bind-
ing of myosin V to syntaxin 1A, a process implicated in vesicle tethering
(Watanabe et al., 2005). In addition, other proteins that interact with either
individual SNAREs or the SNARE complex may be involved in the nexus
between the secretory machinery and the cortical cytoskeleton. The associ-
ation of Munc-18A with syntaxin-1 (Han et al., 2010;Hata et al., 1993;
Haynes et al., 1999;Jahn, 2000;Katagiri et al., 1995) has been implicated in
the morphological docking of dense vesicles controlled by the F-actin
cortical cytoskeleton (de Wit, 2010a,b). Interestingly, overexpression of
this protein strongly influences the density of the submembrane F-actin
layer (Toonen et al., 2006), indicating that Munc-18 is a protein that
influences the interaction between the F-actin cortical cytoskeleton and
SNARE proteins.
These findings suggest that the secretory machinery interacts with
F-actin and cytoskeleton-related proteins, probably via SNAREs and
other factors. But are these interactions essential for fusion? To answer
this question, we must consider studies of the porosome, a structural
element that aggregates the molecular factors required for the formation
of the fusion pore during secretion. During exocytosis, the intravesicular
and extracellular spaces communicate through a pore structure that permits
the entrance of water molecules, generating an electrical conductance that
can be measured by electrophysiology (Alvarez de and Fernandez, 1988;
Breckenridge and Almers, 1987;Chow et al., 1992;Spruce et al., 1990).
Studies in mast and chromaffin cells, and in other cellular models, proposed
the existence of a scaffold formed by a group of proteins that bring together
membranes implicated in vesicle fusion (Monck and Fernandez, 1992;
Monck et al., 1995). This concept assumed the transient formation of
temporal pores with a conductance similar to that of an ion channel,
which induced either kiss and run (Ales et al., 1999;Artalejo et al., 1998;
Schneider, 2001;Stevens and Williams, 2000) or full collapse of vesicles.
However, permanent fusion pore structures, where secretory vesicles dock
and fuse to release their contents, have also been described using atomic
force and electron microscopy in a variety of secretory lines, including
exocrine and endocrine pancreatic cells, chromaffin and mast cells, and
neurons (Cho et al., 2002, 2009;Jena, 2005;Jena and Cho, 2002;Jeremic
et al., 2003;Schneider et al., 1997). These porosomes have a cup-shaped
morphology and they are of variable size, ranging from 10–15nm in
neurons to 150–200nm in pancreatic exocrine cells. The porosome has
been isolated to determine its composition (Cho et al., 2004;Jena et al.,
2003) and reconstituted in artificial membranes (Cho et al., 2004;Jeremic
et al., 2003). These supramolecular structures appear to be composed of
SNAREs (such as SNAP-25 and syntaxins, synaptotagmin), the ATPase
Role of Cortical Cytoskeleton in Exocytosis 123
Author's personal copy
NSF, calcium channel subunits, and cytoskeletal proteins, such as actin,
a-fodrin, and vimentin. Actin is thought to associate with the neck of the
porosome (Schneider et al., 1997), indicating that fusion pore opening and
dilation may be directly regulated by cytoskeletal elements, as described for
myosin II (Berberian et al., 2009;Neco et al., 2004, 2008).
The association of F-actin with the secretory machinery was further
demonstrated in chromaffin cells overexpressing SNAREs (Villanueva
et al., 2010) or the native protein (Torregrosa-Hetland et al., 2011). In
these conditions, dynamic clusters form in the plasma membrane (Lopez
et al., 2009) that associate with the borders of cytoskeletal F-actin cages in
chromaffin cells. These clusters of SNARE proteins also appear to be
associated with L and P/Q-type calcium channels, though this association
varies between intact tissue and isolated cell cultures (Lopez et al., 2007).
6.4. Association of F-actin cortex with ion channels involved
in exocytosis
A number of studies suggest that actin filaments may modulate the activity
of a range of ion channels. For example, cytochalasin D alters the ion flux
associated with calcium channels ( Johnson and Byerly, 1993) and voltage-
dependent sodium channels (Cantiello et al., 1991). Moreover, calcium
channels are endocytosed when the F-actin cytoskeleton is disrupted
(Cristofanilli et al., 2007). This probably reflects the alterations of the
microdomains in the plasma membrane, in the form of cholesterol-rich
lipid rafts that are anchored to cytoskeletal structures, which help organize
the secretory machinery (Churchward and Coorssen, 2009). Indeed,
alpha-7 nicotinic acetylcholine receptors were recently shown to tether to
the postsynaptic membrane in cholesterol-rich domains via PDZ scaffolds
and actin filaments (Fernandes et al., 2010). Interestingly, the microdomains
of L-type voltage-dependent calcium channels exhibit a dynamic behavior
during the fusion of individual vesicles at the ribbon synapse (Mercer et al.,
2011). These domains move within a confined region that is modulated by
actin and cholesterol. After vesicular fusion, the domains spread to occupy a
larger region, although they quickly return to their original position within
a molecular scaffold in order to sustain efficient neurotransmission.
In endocrine cells, cytoplasmic loops II and III of L-type calcium
channels are essential for channels to localize to lipid rafts ( Jacobo et al.,
2009). Moreover, the potentiation of exocytosis by repetitive stimulation of
neuroendocrine chromaffin cells is dependent upon the clustering of L-type
calcium channels in lipid rafts (Park and Kim, 2009). The capacity of F-actin
to agglutinate other relevant molecules, forming clusters in a lipid raft
environment, suggests that the cytoskeleton organizes the cytoarchitecture
associated with secretion in different cellular systems.
124 Luis M. Gutie
´rrez
Author's personal copy
7. Cortical Cytoskeleton as an Organizer of
Molecular Cytoarchitecture of Exocytosis
Fast exocytosis in neurons is very precisely controlled in spatial and
temporal terms, occurring within a specialized region known as the active
zone (Couteaux and Pecot-Dechavassine, 1970;Landis et al., 1988). This is
an electron-dense structure composed of a cytoskeletal matrix and a variety
of associated molecular complexes, and it is involved in the tethering,
docking, and release of neuronal light secretory vesicles (for active zone
function and organization, see Rosenmund et al., 2003 and Schoch and
Gundelfinger, 2006). From the structural viewpoint, the active zone cyto-
matrix is formed by regular arrays of cone-shaped electron-dense particles
that extend 50nm into the neuronal cytosol (Bloom and Aghajanian, 1968;
Gray, 1963). These arrays are interconnected by a meshwork of cytoskeletal
proteins and long filamentous strands of other proteins, forming a scaffold
structure that holds together the constituents of the exo- and endocytotic
neuronal machinery. SNARE proteins are among the proteins that form
these organized structures (Munc-18), as well as scaffolding proteins (e.g.,
CASK, Mint, Bassoon and Piccolo), cell adhesion molecules (including
neurexins, cadherins, integrins), voltage-dependent calcium channels,
and cytoskeletal proteins (such as actin, tubulin, myosins, spectrin, and
b-catenin: Schoch and Gundelfinger, 2006).
Given this level of organization and the presence of specialized structures
that prevent the free diffusion of essential components of the secretory
machinery, there is no doubt that one of the major roles of submembranous
F-actin in neurons is its scaffold function (Sankaranarayanan et al., 2003).
Thus, it is pertinent question to ask whether the cytoskeleton plays a similar
role in endocrine and neuroendocrine cells, thereby organizing the essential
elements of the secretory machinery.
The description of “hot spots” in neuroendocrine cells provided the first
indication of organization of the secretory machinery (Robinson et al.,
1995). These spatially restricted calcium microdomains coincide with indi-
vidual exocytotic events that can be measured with thin amperometric
carbon electrodes in chromaffin cells. However, the sensitivity of secretion
to slow calcium buffers and the latency of secretion suggest that calcium
signals originate at distances within 300–600nm in most chromaffin gran-
ules (Klingauf and Neher, 1997). Interestingly, a fraction of granules (10%)
experienced much higher calcium signals, suggesting they colocalize with
calcium channels. Recently, simultaneous determination of calcium signals
and exocytosis using dual-color TIRFM revealed that most released vesicles
are located within 300nm of calcium microdomains (Becherer et al., 2003).
In chromaffin cells, this distance was reduced in an activity-dependent
manner, indicating that vesicles approach calcium microdomains in
Role of Cortical Cytoskeleton in Exocytosis 125
Author's personal copy
response to continuous stimulation, increasing the efficiency of exocytosis.
Direct analysis of the colocalization of L and P/Q calcium channels with
SNARE microdomains in cultured chromaffin cells (Lopez et al., 2007)
revealed that 30% of SNARE patches containing SNAP-25 and syntaxin-1
colocalize with calcium channel patches. However, the majority of these
SNARE microdomains are within 300nm of calcium channel clusters.
Further, the colocalization of calcium channel and SNARE microdomains
is elevated in intact tissue (adrenomedullary slices), suggesting that the
secretory machinery is more efficiently coupled in neuroendocrine cells of
the adrenal gland, and that cell dissociation and culturing may affect the
organization of the secretory machinery. This would explain the accelerated
secretory kinetics observed in adrenomedullary slices when compared with
dissociated cells (Moser and Neher, 1997).
Does the cortical F-actin cytoskeleton mediate the interaction between
calcium channel clusters and SNARE microdomains? The Cav 2.2 and Cav
2.2 subunits that form the P/Q and N-subtypes of voltage-dependent
calcium channels can directly interact with SNAREs via the synaptic
protein interaction motif (synprint: Sheng et al., 1994, 1997, 1998). These
interactions involve SNARE proteins, such as syntaxin-1A and SNAP-25,
as well as synaptotagmin-1, thought to sense calcium in exocytosis. Thus,
calcium channels appear to contain structural elements that integrate the
molecular machinery required for vesicle docking, priming, and fusion.
Interestingly, recent evidence indicates that the clusters formed by syn-
taxin-1/SNAP-25 dimers are associated with the borders of cytoskeletal
cages in chromaffin cells, as demonstrated by the immunocytochemical
detection of native clusters (Torregrosa-Hetland et al., 2011) and by study-
ing the expression of exogenous SNAP-25 (Villanueva et al., 2010). As
these t-SNARE microdomains colocalize with patches formed by voltage-
dependent calcium channels (Lopez et al., 2007), it is not surprising that
calcium channel microdomains are also found at the border of cytoskeletal
cortical structures (Torregrosa-Hetland et al., 2011). Therefore, despite the
absence of a proper active zone in neuroendocrine cells, it appears that the
cytoskeleton serves as a structural element that recruits calcium channels and
the secretory machinery microdomains to specific membrane regions
(Torregrosa-Hetland et al., 2010).
An important question arising from these observations relates to the
benefits of associating the components of the secretory machinery with a
network of cytoskeletal cavities? To answer this question, we proposed
mathematical models of secretory behavior based on Monte Carlo algo-
rithms (Segura et al., 2000), including the active zone as a domain of
cylindrical geometry where calcium channel clusters and in which secretory
sites are located either in the walls or the center. Of the two geometries, the
most robust initial rise [Ca
2þ
]
i
in the proximity of the active site occurs
when this site is close to the border of the simulated spherical cage. The
126 Luis M. Gutie
´rrez
Author's personal copy
predicted [Ca
2þ
]
i
distributions resulted in faster exocytotic responses when
the secretory machinery (SNAREs) and calcium channels form clusters
associated with the borders of cortical cytoskeletal cages. Therefore, as
seen in neurons, the association of components of the secretory machinery
with the F-actin cortical cytoskeleton, including calcium channels and
SNARE proteins, appears to define a special cytoarchitecture that shapes
secretory kinetics (see Fig. 3.3). This configuration would favor vesicle
docking at active sites in neuroendocrine cells where SNARE clusters
colocalize with VDCC microdomains, leading to fast release of a “ready
Vesicles
Snare clusters
F-actin structures
VDCC clusters
Figure 3.3Cytoarchitecture of the secretory machinery in neuroendocrine chromaf-
fin cells. As mentioned in Fig. 3.1, the cortical cytoskeleton of chromaffin cells forms an
intricate network characterized by the presence of polygonal cages. Recent findings
revealed the association of SNARE microdomains and voltage-dependent calcium
channels (VDCC) clustered along the borders of these cytoskeletal structures
(Torregrosa-Hetland et al., 2011;Villanueva et al., 2010). Other observations indicate
that vesicle fusion occurs in the precise sites occupied by SNARE clusters, and that
VDCC microdomains can colocalize with SNAREs (as observed for the vesicle on the
left). In other cases, VDCC microdomains localize within 300nm of the secretory
machinery (vesicle in the right: Lopez et al., 2007). The F-actin cytoskeleton organizes
the key elements of the secretory machinery to ensure rapid intracellular calcium rises in
the proximity of some active sites (ready releasable vesicle pool). In other populations
of docked vesicles, slower calcium elevations are induced by the diffusion of this second
messenger from nearby VDCC microdomains.
Role of Cortical Cytoskeleton in Exocytosis 127
Author's personal copy
releasable” secretory component (Gillis et al., 1996;Heinemann et al.,
1994;Horrigan and Bookman, 1994). Alternatively, vesicles may dock at
secretory sites located at a distance from the closest calcium channel cluster,
which would involve a delay due to calcium diffusion, resulting in the
release of a slower “docked pool” of vesicles (Becherer et al., 2003;Chow
et al., 1992;Stevens et al., 2011).
8. Conclusion
The cortical F-actin cytoskeleton of neuroendocrine cells is a dynamic
structure that fulfills a variety of fundamental roles required for exocytosis.
This structure reorganizes during secretion in a calcium-dependent manner,
acting as a barrier to vesicles in the resting state and as a transport system to
facilitate vesicle access to exocytotic sites when cells are stimulated by
secretagogues. It is now evident that during the initial phases of vesicle
transport, the F-actin cortical network and molecular motors (such as
myosins II and V) play an essential role in both the dynamics of cytoskeletal
structures and in the processive motion of the vesicles along F-actin trails. In
addition to its expected role in vesicle transport, the cortical network and its
associated proteins also shape the kinetics of neurotransmitter release by
acting in the late membrane fusion steps of exocytosis. In addition, this
network organizes the molecular elements of the secretory machinery by
the forming of specific structures that promote the efficient coupling of
these elements.
ACKNOWLEDGMENTS
The author wishes to thank his mentors and especially his postgraduate students for their
dedication in the field of exocytosis. This work was supported by grants from the Spanish
Ministry of Science and Innovation (MICINN, BFU2008-00731) and the Generalitat
Valenciana (ACOMP2011/090).
REFERENCES
Abdel-Latif, D., Steward, M., Macdonald, D.L., Francis, G.A., Dinauer, M.C., Lacy, P.,
2004. Rac2 is critical for neutrophil primary granule exocytosis. Blood 104, 832–839.
Albillos, A., Dernick, G., Horstmann, H., Almers, W., varez de, T.G., Lindau, M., 1997.
The exocytotic event in chromaffin cells revealed by patch amperometry. Nature 389,
509–512.
Ales, E., Tabares, L., Poyato, J.M., Valero, V., Lindau, M., Alvarez de, T.G., 1999. High
calcium concentrations shift the mode of exocytosis to the kiss-and-run mechanism. Nat.
Cell Biol. 1, 40–44.
128 Luis M. Gutie
´rrez
Author's personal copy
Allersma, M.W., Bittner, M.A., Axelrod, D., Holz, R.W., 2006. Motion matters: secretory
granule motion adjacent to the plasma membrane and exocytosis. Mol. Biol. Cell 17,
2424–2438.
Alvarez de, T.G., Fernandez, J.M., 1988. The events leading to secretory granule fusion.
Soc. Gen. Physiol. Ser. 43, 333–344.
Aoki, R., Kitaguchi, T., Oya, M., Yanagihara, Y., Sato, M., Miyawaki, A., et al., 2010.
Duration of fusion pore opening and the amount of hormone released are regulated by
myosin II during kiss-and-run exocytosis. Biochem. J. 429, 497–504.
Artalejo, C.R., Elhamdani, A., Palfrey, H.C., 1998. Secretion: dense-core vesicles can kiss-
and-run too. Curr. Biol. 8, R62–R65.
Aunis, D., Bader, M.F., 1988. The cytoskeleton as a barrier to exocytosis in secretory cells.
J. Exp. Biol. 139, 253–266.
Aunis, D., Guerold, B., Bader, M.F., Cieselski-Treska, J., 1980. Immunocytochemical and
biochemical demonstration of contractile proteins in chromaffin cells in culture. Neuro-
science 5, 2261–2277.
Bader, M.F., Doussau, F., Chasserot-Golaz, S., Vitale, N., Gasman, S., 2004. Coupling actin
and membrane dynamics during calcium-regulated exocytosis: a role for Rho and ARF
GTPases. Biochim. Biophys. Acta 1742, 37–49.
Becherer, U., Moser, T., Stuhmer, W., Oheim, M., 2003. Calcium regulates exocytosis at
the level of single vesicles. Nat. Neurosci. 6, 846–853.
Berberian, K., Torres, A.J., Fang, Q., Kisler, K., Lindau, M., 2009. F-actin and myosin II
accelerate catecholamine release from chromaffin granules. J. Neurosci. 29, 863–870.
Bershadsky, A.D., Ballestrem, C., Carramusa, L., Zilberman, Y., Gilquin, B., Khochbin, S.,
et al., 2006. Assembly and mechanosensory function of focal adhesions: experiments and
models. Eur. J. Cell Biol. 85, 165–173.
Bloom, F.E., Aghajanian, G.K., 1968. Fine structural and cytochemical analysis of the
staining of synaptic junctions with phosphotungstic acid. J. Ultrastruct. Res. 22,
361–375.
Bond, L.M., Brandstaetter, H., Sellers, J.R., Kendrick-Jones, J., Buss, F., 2011. Myosin
motor proteins are involved in the final stages of the secretory pathways. Biochem. Soc.
Trans. 39, 1115–1119.
Borisy, G.G., Svitkina, T.M., 2000. Actin machinery: pushing the envelope. Curr. Opin.
Cell Biol. 12, 104–112.
Breckenridge, L.J., Almers, W., 1987. Currents through the fusion pore that forms during
exocytosis of a secretory vesicle. Nature 328, 814–817.
Burgoyne, R.D., Cheek, T.R., 1987. Reorganisation of peripheral actin filaments as a
prelude to exocytosis. Biosci. Rep. 7, 281–288.
Burgoyne, R.D., Cheek, T.R., Norman, K.M., 1986. Identification of a secretory granule-
binding protein as caldesmon. Nature 319, 68–70.
Burgoyne, R.D., Handel, S.E., Morgan, A., Rennison, M.E., Turner, M.D., Wilde, C.J.,
1991. Calcium, the cytoskeleton and calpactin (annexin II) in exocytotic secretion
from adrenal chromaffin and mammary epithelial cells. Biochem. Soc. Trans. 19,
1085–1090.
Cantiello, H.F., Stow, J.L., Prat, A.G., Ausiello, D.A., 1991. Actin filaments regulate
epithelial Naþchannel activity. Am. J. Physiol. 261, C882–C888.
Carbajal, M.E., Vitale, M.L., 1997. The cortical actin cytoskeleton of lactotropes as an
intracellular target for the control of prolactin secretion. Endocrinology 138, 5374–5384.
Caviston, J.P., Holzbaur, E.L., 2006. Microtubule motors at the intersection of trafficking
and transport. Trends Cell Biol. 16, 530–537.
Chasserot-Golaz, S., Vitale, N., Umbrecht-Jenck, E., Knight, D., Gerke, V., Bader, M.F.,
2005. Annexin 2 promotes the formation of lipid microdomains required for calcium-
regulated exocytosis of dense-core vesicles. Mol. Biol. Cell 16, 1108–1119.
Role of Cortical Cytoskeleton in Exocytosis 129
Author's personal copy
Cheek, T.R., Burgoyne, R.D., 1986. Nicotine-evoked disassembly of cortical actin fila-
ments in adrenal chromaffin cells. FEBS Lett. 207, 110–114.
Cheek, T.R., Hesketh, J.E., Richards, R.C., Burgoyne, R.D., 1986. Assembly and char-
acterisation of a multi-component cytoskeletal gel from adrenal medulla. Biochim.
Biophys. Acta 887, 164–172.
Cho, S.J., Wakade, A., Pappas, G.D., Jena, B.P., 2002. New structure involved in transient
membrane fusion and exocytosis. Ann. N. Y. Acad. Sci. 971, 254–256.
Cho, W.J., Jeremic, A., Rognlien, K.T., Zhvania, M.G., Lazrishvili, I., Tamar, B., et al.,
2004. Structure, isolation, composition and reconstitution of the neuronal fusion pore.
Cell Biol. Int. 28, 699–708.
Cho, W.J., Shin, L., Ren, G., Jena, B.P., 2009. Structure of membrane-associated neuronal
SNARE complex: implication in neurotransmitter release. J. Cell. Mol. Med. 13,
4161–4165.
Chow, R.H., von, R.L., Neher, E., 1992. Delay in vesicle fusion revealed by electro-
chemical monitoring of single secretory events in adrenal chromaffin cells. Nature 356,
60–63.
Chowdhury, H.H., Popoff, M.R., Zorec, R., 1999. Actin cytoskeleton depolymerization
with clostridium spiroforme toxin enhances the secretory activity of rat melanotrophs.
J. Physiol. 521 (Pt 2), 389–395.
Churchward, M.A., Coorssen, J.R., 2009. Cholesterol, regulated exocytosis and the physi-
ological fusion machine. Biochem. J. 423, 1–14.
Cooke, P.H., Poisner, A.M., 1979. The role of cytoskeleton in adreno-medullary secretion.
Methods Achiev. Exp. Pathol. 9, 137–146.
Cote, A., Doucet, J.P., Trifaro, J.M., 1986. Phosphorylation and dephosphorylation of
chromaffin cell proteins in response to stimulation. Neuroscience 19, 629–645.
Couteaux, R., Pecot-Dechavassine, M., 1970. Synaptic vesicles and pouches at the level of
“active zones” of the neuromuscular junction. C. R. Acad. Sci. Hebd. Seances Acad. Sci.
D 271, 2346–2349.
Creutz, C.E., Scott, J.H., Pazoles, C.J., Pollard, H.B., 1982. Further characterization of the
aggregation and fusion of chromaffin granules by synexin as a model for compound
exocytosis. J. Cell. Biochem. 18, 87–97.
Cristofanilli, M., Mizuno, F., Akopian, A., 2007. Disruption of actin cytoskeleton causes
internalization of Ca(v)1.3 (alpha 1D) L-type calcium channels in salamander retinal
neurons. Mol. Vis. 13, 1496–1507.
Daniel, S., Noda, M., Cerione, R.A., Sharp, G.W., 2002. A link between Cdc42 and
syntaxin is involved in mastoparan-stimulated insulin release. Biochemistry 41,
9663–9671.
De Wit, W.H., 2010a. Molecular mechanism of secretory vesicle docking. Biochem. Soc.
Trans. 38, 192–198.
De Wit, W.H., 2010b. Morphological docking of secretory vesicles. Histochem. Cell Biol.
134, 103–113.
DePina, A.S., Langford, G.M., 1999. Vesicle transport: the role of actin filaments and
myosin motors. Microsc. Res. Tech. 47, 93–106.
Desnos, C., Schonn, J.S., Huet, S., Tran, V.S., El-Amraoui, A., Raposo, G., et al., 2003.
Rab27A and its effector MyRIP link secretory granules to F-actin and control their
motion towards release sites. J. Cell Biol. 163, 559–570.
Desnos, C., Huet, S., Fanget, I., Chapuis, C., Bottiger, C., Racine, V., et al., 2007. Myosin
va mediates docking of secretory granules at the plasma membrane. J. Neurosci. 27,
10636–10645.
Doreian, B.W., Fulop, T.G., Smith, C.B., 2008. Myosin II activation and actin reorganiza-
tion regulate the mode of quantal exocytosis in mouse adrenal chromaffin cells.
J. Neurosci. 28, 4470–4478.
130 Luis M. Gutie
´rrez
Author's personal copy
Douglas, W.W., Sorimachi, M., 1972. Colchicine inhibits adrenal medullary secretion
evoked by acetylcholine without affecting that evoked by potassium. Br. J. Pharmacol.
45, 129–132.
Eberhard, D.A., Cooper, C.L., Low, M.G., Holz, R.W., 1990. Evidence that the inositol
phospholipids are necessary for exocytosis. Loss of inositol phospholipids and inhibition
of secretion in permeabilized cells caused by a bacterial phospholipase C and removal of
ATP. Biochem. J. 268, 15–25.
Farshori, P.Q., Goode, D., 1994. Effects of the microtubule depolymerizing and stabilizing
agents Nocodazole and taxol on glucose-induced insulin secretion from hamster islet
tumor (HIT) cells. J. Submicrosc. Cytol. Pathol. 26, 137–146.
Fernandes, C.C., Berg, D.K., Gomez-Varela, D., 2010. Lateral mobility of nicotinic acetyl-
choline receptors on neurons is determined by receptor composition, local domain, and
cell type. J. Neurosci. 30, 8841–8851.
Frantz, C., Coppola, T., Regazzi, R., 2002. Involvement of Rho GTPases and their
effectors in the secretory process of PC12 cells. Exp. Cell Res. 273, 119–126.
Gasman, S., Chasserot-Golaz, S., Popoff, M.R., Aunis, D., Bader, M.F., 1997. Trimeric G
proteins control exocytosis in chromaffin cells. Go regulates the peripheral actin network
and catecholamine secretion by a mechanism involving the small GTP-binding protein
Rho. J. Biol. Chem. 272, 20564–20571.
Gasman, S., Chasserot-Golaz, S., Hubert, P., Aunis, D., Bader, M.F., 1998. Identification of
a potential effector pathway for the trimeric Go protein associated with secretory
granules. Go stimulates a granule-bound phosphatidylinositol 4-kinase by activating
RhoA in chromaffin cells. J. Biol. Chem. 273, 16913–16920.
Gasman, S., Chasserot-Golaz, S., Malacombe, M., Way, M., Bader, M.F., 2004. Regulated
exocytosis in neuroendocrine cells: a role for subplasmalemmal Cdc42/N-WASP-
induced actin filaments. Mol. Biol. Cell 15, 520–531.
Geisow, M.J., Walker, J.H., Boustead, C., Taylor, W., 1987. Annexins—new family of
Ca2þ-regulated-phospholipid binding protein. Biosci. Rep. 7, 289–298.
Gil, A., Rueda, J., Viniegra, S., Gutierrez, L.M., 2000. The F-actin cytoskeleton modulates
slow secretory components rather than readily releasable vesicle pools in bovine
chromaffin cells. Neuroscience 98, 605–614.
Gillis, K.D., Mossner, R., Neher, E., 1996. Protein kinase C enhances exocytosis from
chromaffin cells by increasing the size of the readily releasable pool of secretory granules.
Neuron 16, 1209–1220.
Giner, D., Neco, P., Frances, M.M., Lopez, I., Viniegra, S., Gutierrez, L.M., 2005. Real-
time dynamics of the F-actin cytoskeleton during secretion from chromaffin cells. J. Cell
Sci. 118, 2871–2880.
Giner, D., Lopez, I., Villanueva, J., Torres, V., Viniegra, S., Gutierrez, L.M., 2007. Vesicle
movements are governed by the size and dynamics of F-actin cytoskeletal structures in
bovine chromaffin cells. Neuroscience 146, 659–669.
Gray, E.G., 1963. Electron microscopy of presynaptic organelles of the spinal cord. J. Anat.
97, 101–106.
Gutierrez, L.M., Ballesta, J.J., Hidalgo, M.J., Gandia, L., Garcia, A.G., Reig, J.A., 1988. A
two-dimensional electrophoresis study of phosphorylation and dephosphorylation of chro-
maffin cell proteins in response to a secretory stimulus. J. Neurochem. 51, 1023–1030.
Gutierrez, L.M., Hidalgo, M.J., Palmero, M., Ballesta, J.J., Reig, J.A., Garcia, A.G., et al.,
1989. Phosphorylation of myosin light chain from adrenomedullary chromaffin cells in
culture. Biochem. J. 264, 589–596.
Han, G.A., Malintan, N.T., Collins, B.M., Meunier, F.A., Sugita, S., 2010. Munc18-1 as a
key regulator of neurosecretion. J. Neurochem. 115, 1–10.
Hata, Y., Slaughter, C.A., Sudhof, T.C., 1993. Synaptic vesicle fusion complex contains
unc-18 homologue bound to syntaxin. Nature 366, 347–351.
Role of Cortical Cytoskeleton in Exocytosis 131
Author's personal copy
Hay, J.C., Martin, T.F., 1993. Phosphatidylinositol transfer protein required for ATP-
dependent priming of Ca(2þ)-activated secretion. Nature 366, 572–575.
Hay, J.C., Fisette, P.L., Jenkins, G.H., Fukami, K., Takenawa, T., Anderson, R.A., et al.,
1995. ATP-dependent inositide phosphorylation required for Ca(2þ)-activated secre-
tion. Nature 374, 173–177.
Hayes, M.J., Rescher, U., Gerke, V., Moss, S.E., 2004. Annexin-actin interactions. Traffic
5, 571–576.
Haynes, L.P., Morgan, A., Burgoyne, R.D., 1999. nSec-1 (munc-18) interacts with both
primed and unprimed syntaxin 1A and associates in a dimeric complex on adrenal
chromaffin granules. Biochem. J. 342 (Pt 3), 707–714.
Heinemann, C., Chow, R.H., Neher, E., Zucker, R.S., 1994. Kinetics of the secretory
response in bovine chromaffin cells following flash photolysis of caged Ca
2þ
. Biophys. J.
67, 2546–2557.
Hodel, A., Schafer, T., Gerosa, D., Burger, M.M., 1994. In chromaffin cells, the mammalian
Sec1p homologue is a syntaxin 1A-binding protein associated with chromaffin granules.
J. Biol. Chem. 269, 8623–8626.
Holz, R.W., Axelrod, D., 2002. Localization of phosphatidylinositol 4,5-P(2) important in
exocytosis and a quantitative analysis of chromaffin granule motion adjacent to the plasma
membrane. Ann. N. Y. Acad. Sci. 971, 232–243.
Hong-Geller, E., Cerione, R.A., 2000. Cdc42 and Rac stimulate exocytosis of secretory
granules by activating the IP(3)/calcium pathway in RBL-2H3 mast cells. J. Cell Biol.
148, 481–494.
Horrigan, F.T., Bookman, R.J., 1994. Releasable pools and the kinetics of exocytosis in
adrenal chromaffin cells. Neuron 13, 1119–1129.
Iida, Y., Senda, T., Matsukawa, Y., Onoda, K., Miyazaki, J.I., Sakaguchi, H., et al., 1997.
Myosin light-chain phosphorylation controls insulin secretion at a proximal step in the
secretory cascade. Am. J. Physiol. 273, E782–E789.
Jacobo, S.M., Guerra, M.L., Jarrard, R.E., Przybyla, J.A., Liu, G., Watts, V.J., et al., 2009.
The intracellular II-III loops of Cav1.2 and Cav1.3 uncouple L-type voltage-gated Ca
2þ
channels from glucagon-like peptide-1 potentiation of insulin secretion in INS-1 cells via
displacement from lipid rafts. J. Pharmacol. Exp. Ther. 330, 283–293.
Jahn, R., 2000. Sec1/Munc18 proteins: mediators of membrane fusion moving to center
stage. Neuron 27, 201–204.
Jena, B.P., 2005. Molecular machinery and mechanism of cell secretion. Exp. Biol. Med.
(Maywood) 230, 307–319.
Jena, B.P., Cho, S.J., 2002. The atomic force microscope in the study of membrane fusion
and exocytosis. Methods Cell Biol. 68, 33–50.
Jena, B.P., Cho, S.J., Jeremic, A., Stromer, M.H., bu-Hamdah, R., 2003. Structure and
composition of the fusion pore. Biophys. J. 84, 1337–1343.
Jerdeva, G.V., Wu, K., Yarber, F.A., Rhodes, C.J., Kalman, D., Schechter, J.E., et al., 2005.
Actin and non-muscle myosin II facilitate apical exocytosis of tear proteins in rabbit
lacrimal acinar epithelial cells. J. Cell Sci. 118, 4797–4812.
Jeremic, A., Kelly, M., Cho, S.J., Stromer, M.H., Jena, B.P., 2003. Reconstituted fusion
pore. Biophys. J. 85, 2035–2043.
Jewell, J.L., Luo, W., Oh, E., Wang, Z., Thurmond, D.C., 2008. Filamentous actin
regulates insulin exocytosis through direct interaction with Syntaxin 4. J. Biol. Chem.
283, 10716–10726.
Joanny, J.F., Prost, J., 2009. Active gels as a description of the actin-myosin cytoskeleton.
HFSP J. 3, 94–104.
Johns, L.M., Levitan, E.S., Shelden, E.A., Holz, R.W., Axelrod, D., 2001. Restriction of
secretory granule motion near the plasma membrane of chromaffin cells. J. Cell Biol. 153,
177–190.
132 Luis M. Gutie
´rrez
Author's personal copy
Johnson, B.D., Byerly, L., 1993. A cytoskeletal mechanism for Ca
2þ
channel metabolic
dependence and inactivation by intracellular Ca
2þ
. Neuron 10, 797–804.
Katagiri, H., Terasaki, J., Murata, T., Ishihara, H., Ogihara, T., Inukai, K., et al., 1995. A
novel isoform of syntaxin-binding protein homologous to yeast Sec1 expressed ubiqui-
tously in mammalian cells. J. Biol. Chem. 270, 4963–4966.
Klingauf, J., Neher, E., 1997. Modeling buffered Ca
2þ
diffusion near the membrane:
implications for secretion in neuroendocrine cells. Biophys. J. 72, 674–690.
Kogel, T., Bittins, C.M., Rudolf, R., Gerdes, H.H., 2010. Versatile roles for myosin Va in
dense core vesicle biogenesis and function. Biochem. Soc. Trans. 38, 199–204.
Kumakura, K., Sasaki, K., Sakurai, T., Ohara-Imaizumi, M., Misonou, H., Nakamura, S.,
et al., 1994. Essential role of myosin light chain kinase in the mechanism for
MgATP-dependent priming of exocytosis in adrenal chromaffin cells. J. Neurosci. 14,
7695–7703.
Landis, D.M., Hall, A.K., Weinstein, L.A., Reese, T.S., 1988. The organization of cyto-
plasm at the presynaptic active zone of a central nervous system synapse. Neuron 1,
201–209.
Lang, T., Wacker, I., Steyer, J., Kaether, C., Wunderlich, I., Soldati, T., et al., 1997. Ca2þ-
triggered peptide secretion in single cells imaged with green fluorescent protein and
evanescent-wave microscopy. Neuron 18, 857–863.
Lang, T., Wacker, I., Wunderlich, I., Rohrbach, A., Giese, G., Soldati, T., et al., 2000. Role
of actin cortex in the subplasmalemmal transport of secretory granules in PC-12 cells.
Biophys. J. 78, 2863–2877.
Lee, R.W., Trifaro, J.M., 1981. Characterization of anti-actin antibodies and their use in
immunocytochemical studies on the localization of actin in adrenal chromaffin cells in
culture. Neuroscience 6, 2087–2108.
Li, Q., Ho, C.S., Marinescu, V., Bhatti, H., Bokoch, G.M., Ernst, S.A., et al., 2003.
Facilitation of Ca(2þ)-dependent exocytosis by Rac1-GTPase in bovine chromaffin
cells. J. Physiol. 550, 431–445.
Lipschutz, J.H., Mostov, K.E., 2002. Exocytosis: the many masters of the exocyst. Curr.
Biol. 12, R212–R214.
Lopez, I., Giner, D., Ruiz-Nuno, A., Fuentealba, J., Viniegra, S., Garcia, A.G., et al., 2007.
Tight coupling of the t-SNARE and calcium channel microdomains in adrenomedullary
slices and not in cultured chromaffin cells. Cell Calcium 41, 547–558.
Lopez, I., Ortiz, J.A., Villanueva, J., Torres, V., Torregrosa-Hetland, C.J., del Mar, F.M.,
et al., 2009. Vesicle motion and fusion are altered in chromaffin cells with increased
SNARE cluster dynamics. Traffic 10, 172–185.
Malacombe, M., Bader, M.F., Gasman, S., 2006. Exocytosis in neuroendocrine cells: new
tasks for actin. Biochim. Biophys. Acta 1763, 1175–1183.
Mangeat, P.H., 1988. Interaction of biological membranes with the cytoskeletal framework
of living cells. Biol. Cell 64, 261–281.
Marcu, M.G., Rodriguez Del Castillo, A., Vitale, M.L., Trifaro, J.M., 1994. Molecular
cloning and functional expression of chromaffin cell scinderin indicates that it belongs to
the family of Ca(2þ)-dependent F-actin severing proteins. Mol. Cell. Biochem. 141,
153–165.
Marxen, P., Bigalke, H., 1991. Tetanus and botulinum A toxins inhibit stimulated F-actin
rearrangement in chromaffin cells. Neuroreport 2, 33–36.
Matter, K., Dreyer, F., Aktories, K., 1989. Actin involvement in exocytosis from PC12 cells:
studies on the influence of botulinum C2 toxin on stimulated noradrenaline release.
J. Neurochem. 52, 370–376.
McKay, D.B., Lopez, I., Sanchez, P.A., English, J.L., Wallace, L.J., 1991. Characterization
of muscarinic receptors of bovine adrenal chromaffin cells: binding, secretion and anti-
microtubule drug effects. Gen. Pharmacol. 22, 1185–1189.
Role of Cortical Cytoskeleton in Exocytosis 133
Author's personal copy
Meiri, K.F., 2004. Membrane/cytoskeleton communication. Subcell. Biochem. 37, 247–282.
Mercer, A.J., Chen, M., Thoreson, W.B., 2011. Lateral mobility of presynaptic L-type
calcium channels at photoreceptor ribbon synapses. J. Neurosci. 31, 4397–4406.
Mizuno, D., Tardin, C., Schmidt, C.F., Mackintosh, F.C., 2007. Nonequilibrium mechan-
ics of active cytoskeletal networks. Science 315, 370–373.
Monck, J.R., Fernandez, J.M., 1992. The exocytotic fusion pore. J. Cell Biol. 119,
1395–1404.
Monck, J.R., Oberhauser, A.F., Fernandez, J.M., 1995. The exocytotic fusion pore inter-
face: a model of the site of neurotransmitter release. Mol. Membr. Biol. 12, 151–156.
Mooseker, M.S., Cheney, R.E., 1995. Unconventional myosins. Annu. Rev. Cell Dev.
Biol. 11, 633–675.
Moser, T., Neher, E., 1997. Rapid exocytosis in single chromaffin cells recorded from
mouse adrenal slices. J. Neurosci. 17, 2314–2323.
Nakata, T., Hirokawa, N., 1992. Organization of cortical cytoskeleton of cultured chro-
maffin cells and involvement in secretion as revealed by quick-freeze, deep-etching, and
double-label immunoelectron microscopy. J. Neurosci. 12, 2186–2197.
Neco, P., Gil, A., del Mar, F.M., Viniegra, S., Gutierrez, L.M., 2002. The role of myosin in
vesicle transport during bovine chromaffin cell secretion. Biochem. J. 368, 405–413.
Neco, P., Giner, D., del Mar, F.M., Viniegra, S., Gutierrez, L.M., 2003. Differential
participation of actin- and tubulin-based vesicle transport systems during secretion in
bovine chromaffin cells. Eur. J. Neurosci. 18, 733–742.
Neco, P., Giner, D., Viniegra, S., Borges, R., Villarroel, A., Gutierrez, L.M., 2004. New
roles of myosin II during vesicle transport and fusion in chromaffin cells. J. Biol. Chem.
279, 27450–27457.
Neco, P., Fernandez-Peruchena, C., Navas, S., Gutierrez, L.M., de Toledo, G.A., Ales, E.,
2008. Myosin II contributes to fusion pore expansion during exocytosis. J. Biol. Chem.
283, 10949–10957.
Nemoto, T., Kojima, T., Oshima, A., Bito, H., Kasai, H., 2004. Stabilization of exocytosis
by dynamic F-actin coating of zymogen granules in pancreatic acini. J. Biol. Chem. 279,
37544–37550.
Nevins, A.K., Thurmond, D.C., 2005. A direct interaction between Cdc42 and vesicle-
associated membrane protein 2 regulates SNARE-dependent insulin exocytosis. J. Biol.
Chem. 280, 1944–1952.
Oheim, M., Stuhmer, W., 2000. Tracking chromaffin granules on their way through the
actin cortex. Eur. Biophys. J. 29, 67–89.
Orci, L., Gabbay, K.H., Malaisse, W.J., 1972. Pancreatic beta-cell web: its possible role in
insulin secretion. Science 175, 1128–1130.
Park, Y., Kim, K.T., 2009. Dominant role of lipid rafts L-type calcium channel in activity-
dependent potentiation of large dense-core vesicle exocytosis. J. Neurochem. 110,
520–529.
Pendleton, A., Koffer, A., 2001. Effects of latrunculin reveal requirements for the actin
cytoskeleton during secretion from mast cells. Cell Motil. Cytoskeleton 48, 37–51.
Perrin, D., Aunis, D., 1985. Reorganization of alpha-fodrin induced by stimulation in
secretory cells. Nature 315, 589–592.
Perrin, D., Langley, O.K., Aunis, D., 1987. Anti-alpha-fodrin inhibits secretion from
permeabilized chromaffin cells. Nature 326, 498–501.
Poisner, A.M., Bernstein, J., 1971. A possible role of microtubules in catecholamine release
from the adrenal medulla: effect of colchicine, vinca alkaloids and deuterium oxide.
J. Pharmacol. Exp. Ther. 177, 102–108.
Pollard, H.B., Creutz, C.E., Fowler, V., Scott, J., Pazoles, C.J., 1982. Calcium-dependent
regulation of chromaffin granule movement, membrane contact, and fusion during
exocytosis. Cold Spring Harb. Symp. Quant. Biol. 46 (Pt 2), 819–834.
134 Luis M. Gutie
´rrez
Author's personal copy
Reig, J.A., Viniegra, S., Ballesta, J.J., Palmero, M., Guitierrez, L.M., 1993. Naphthalene-
sulfonamide derivatives ML9 and W7 inhibit catecholamine secretion in intact and
permeabilized chromaffin cells. Neurochem. Res. 18, 317–323.
Robinson, I.M., Finnegan, J.M., Monck, J.R., Wightman, R.M., Fernandez, J.M., 1995.
Colocalization of calcium entry and exocytotic release sites in adrenal chromaffin cells.
Proc. Natl. Acad. Sci. USA 92, 2474–2478.
Rodriguez Del, C.A., Lemaire, S., Tchakarov, L., Jeyapragasan, M., Doucet, J.P., Vitale, M.
L., et al., 1990. Chromaffin cell scinderin, a novel calcium-dependent actin filament-
severing protein. EMBO J. 9, 43–52.
Rose, S.D., Lejen, T., Casaletti, L., Larson, R.E., Pene, T.D., Trifaro, J.M., 2002. Molecu-
lar motors involved in chromaffin cell secretion. Ann. N. Y. Acad. Sci. 971, 222–231.
Rose, S.D., Lejen, T., Casaletti, L., Larson, R.E., Pene, T.D., Trifaro, J.M., 2003. Myosins
II and V in chromaffin cells: myosin V is a chromaffin vesicle molecular motor involved
in secretion. J. Neurochem. 85, 287–298.
Rosenmund, C., Rettig, J., Brose, N., 2003. Molecular mechanisms of active zone function.
Curr. Opin. Neurobiol. 13, 509–519.
Roth, D., Burgoyne, R.D., 1994. SNAP-25 is present in a SNARE complex in adrenal
chromaffin cells. FEBS Lett. 351, 207–210.
Roth, D., Morgan, A., Burgoyne, R.D., 1993. Identification of a key domain in annexin
and 14-3-3 proteins that stimulate calcium-dependent exocytosis in permeabilized adre-
nal chromaffin cells. FEBS Lett. 320, 207–210.
Rudolf, R., Kogel, T., Kuznetsov, S.A., Salm, T., Schlicker, O., Hellwig, A., et al., 2003.
Myosin Va facilitates the distribution of secretory granules in the F-actin rich cortex of
PC12 cells. J. Cell Sci. 116, 1339–1348.
Sankaranarayanan, S., Atluri, P.P., Ryan, T.A., 2003. Actin has a molecular scaffolding, not
propulsive, role in presynaptic function. Nat. Neurosci. 6, 127–135.
Schneider, S.W., 2001. Kiss and run mechanism in exocytosis. J. Membr. Biol. 181, 67–76.
Schneider, S.W., Sritharan, K.C., Geibel, J.P., Oberleithner, H., Jena, B.P., 1997. Surface
dynamics in living acinar cells imaged by atomic force microscopy: identification of
plasma membrane structures involved in exocytosis. Proc. Natl. Acad. Sci. USA 94,
316–321.
Schoch, S., Gundelfinger, E.D., 2006. Molecular organization of the presynaptic active
zone. Cell Tissue Res. 326, 379–391.
Segura, J., Gil, A., Soria, B., 2000. Modeling study of exocytosis in neuroendocrine cells:
influence of the geometrical parameters. Biophys. J. 79, 1771–1786.
Sellers, J.R., Knight, P.J., 2007. Folding and regulation in myosins II and V. J. Muscle Res.
Cell Motil. 28, 363–370.
Sheng, Z.H., Rettig, J., Takahashi, M., Catterall, W.A., 1994. Identification of a syntaxin-
binding site on N-type calcium channels. Neuron 13, 1303–1313.
Sheng, Z.H., Yokoyama, C.T., Catterall, W.A., 1997. Interaction of the synprint site of N-
type Ca2þchannels with the C2B domain of synaptotagmin I. Proc. Natl. Acad. Sci.
USA 94, 5405–5410.
Sheng, Z.H., Westenbroek, R.E., Catterall, W.A., 1998. Physical link and functional
coupling of presynaptic calcium channels and the synaptic vesicle docking/fusion
machinery. J. Bioenerg. Biomembr. 30, 335–345.
Sokac, A.M., Bement, W.M., 2006. Kiss-and-coat and compartment mixing: coupling
exocytosis to signal generation and local actin assembly. Mol. Biol. Cell 17, 1495–1502.
Sokac, A.M., Co, C., Taunton, J., Bement, W., 2003. Cdc42-dependent actin polymeriza-
tion during compensatory endocytosis in Xenopus eggs. Nat. Cell Biol. 5, 727–732.
Sollner, T., Bennett, M.K., Whiteheart, S.W., Scheller, R.H., Rothman, J.E., 1993a.
A protein assembly-disassembly pathway in vitro that may correspond to sequential
steps of synaptic vesicle docking, activation, and fusion. Cell 75, 409–418.
Role of Cortical Cytoskeleton in Exocytosis 135
Author's personal copy
Sollner, T., Whiteheart, S.W., Brunner, M., Erdjument-Bromage, H., Geromanos, S.,
Tempst, P., et al., 1993b. SNAP receptors implicated in vesicle targeting and fusion.
Nature 362, 318–324.
Spruce, A.E., Breckenridge, L.J., Lee, A.K., Almers, W., 1990. Properties of the fusion pore
that forms during exocytosis of a mast cell secretory vesicle. Neuron 4, 643–654.
Stevens, C.F., Williams, J.H., 2000. “Kiss and run” exocytosis at hippocampal synapses.
Proc. Natl. Acad. Sci. USA 97, 12828–12833.
Stevens, D.R., Schirra, C., Becherer, U., Rettig, J., 2011. Vesicle pools: lessons from adrenal
chromaffin cells. Front Synaptic Neurosci. 3, 2.
Steyer, J.A., Almers, W., 1999. Tracking single secretory granules in live chromaffin cells by
evanescent-field fluorescence microscopy. Biophys. J. 76, 2262–2271.
Taunton, J., Rowning, B.A., Coughlin, M.L., Wu, M., Moon, R.T., Mitchison, T.J., et al.,
2000. Actin-dependent propulsion of endosomes and lysosomes by recruitment of N-
WASP. J. Cell Biol. 148, 519–530.
Tchakarov, L.E., Zhang, L., Rose, S.D., Tang, R., Trifaro, J.M., 1998. Light and electron
microscopic study of changes in the organization of the cortical actin cytoskeleton during
chromaffin cell secretion. J. Histochem. Cytochem. 46, 193–203.
Thurmond, D.C., Gonelle-Gispert, C., Furukawa, M., Halban, P.A., Pessin, J.E., 2003.
Glucose-stimulated insulin secretion is coupled to the interaction of actin with the
t-SNARE (target membrane soluble N-ethylmaleimide-sensitive factor attachment
protein receptor protein) complex. Mol. Endocrinol. 17, 732–742.
Toonen, R.F., Kochubey, O., de, W.H., Gulyas-Kovacs, A., Konijnenburg, B., Sorensen, J.B.,
et al., 2006. Dissecting docking and tethering of secretory vesicles at the target membrane.
EMBO J. 25, 3725–3737.
Torregrosa-Hetland, C.J., Villanueva, J., Lopez-Font, I., Garcia-Martinez, V., Gil, A.,
Gonzalez-Velez, V., et al., 2010. Association of SNAREs and calcium channels with
the borders of cytoskeletal cages organizes the secretory machinery in chromaffin cells.
Cell. Mol. Neurobiol. 30, 1315–1319.
Torregrosa-Hetland, C.J., Villanueva, J., Giner, D., Lopez-Font, I., Nadal, A., Quesada, I.,
et al., 2011. The F-actin cortical network is a major factor influencing the organization of
the secretory machinery in chromaffin cells. J. Cell Sci. 124, 727–734.
Trifaro, J.M., Kenigsberg, R.L., Cote, A., Lee, R.W., Hikita, T., 1984. Adrenal paraneur-
one contractile proteins and stimulus-secretion coupling. Can. J. Physiol. Pharmacol. 62,
493–501.
Trifaro, J.M., Bader, M.F., Doucet, J.P., 1985. Chromaffin cell cytoskeleton: its possible role
in secretion. Can. J. Biochem. Cell Biol. 63, 661–679.
Turvey, M.R., Thorn, P., 2004. Lysine-fixable dye tracing of exocytosis shows F-actin
coating is a step that follows granule fusion in pancreatic acinar cells. Pflugers Arch. 448,
552–555.
Valentijn, J.A., Valentijn, K., Pastore, L.M., Jamieson, J.D., 2000. Actin coating of secretory
granules during regulated exocytosis correlates with the release of rab3D. Proc. Natl.
Acad. Sci. USA 97, 1091–1095.
Verkhovsky, A.B., Svitkina, T.M., Borisy, G.G., 1999. Network contraction model for cell
translocation and retrograde flow. Biochem. Soc. Symp. 65, 207–222.
Villanueva, J., Torregrosa-Hetland, C.J., Gil, A., Gonzalez-Velez, V., Segura, J.,
Viniegra, S., et al., 2010. The organization of the secretory machinery in chromaffin
cells as a major factor in modeling exocytosis. HFSP J. 4, 85–92.
Vitale, M.L., Rodriguez Del, C.A., Tchakarov, L., Trifaro, J.M., 1991. Cortical filamentous
actin disassembly and scinderin redistribution during chromaffin cell stimulation precede
exocytosis, a phenomenon not exhibited by gelsolin. J. Cell Biol. 113, 1057–1067.
Vitale, N., Gensse, M., Chasserot-Golaz, S., Aunis, D., Bader, M.F., 1996. Trimeric G
proteins control regulated exocytosis in bovine chromaffin cells: sequential involvement
136 Luis M. Gutie
´rrez
Author's personal copy
of Go associated with secretory granules and Gi3 bound to the plasma membrane. Eur. J.
Neurosci. 8, 1275–1285.
Wang, C.L., 2008. Caldesmon and the regulation of cytoskeletal functions. Adv. Exp. Med.
Biol. 644, 250–272.
Watanabe, M., Nomura, K., Ohyama, A., Ishikawa, R., Komiya, Y., Hosaka, K., et al.,
2005. Myosin-Va regulates exocytosis through the submicromolar Ca2þ-dependent
binding of syntaxin-1A. Mol. Biol. Cell 16, 4519–4530.
Weber, T., Zemelman, B.V., McNew, J.A., Westermann, B., Gmachl, M., Parlati, F., et al.,
1998. SNAREpins: minimal machinery for membrane fusion. Cell 92, 759–772.
Winkler, H., 1977. The biogenesis of adrenal chromaffin granules. Neuroscience 2,
657–683.
Yin, H.L., Janmey, P.A., 2003. Phosphoinositide regulation of the actin cytoskeleton. Annu.
Rev. Physiol. 65, 761–789.
Yu, H.Y., Bement, W.M., 2007. Control of local actin assembly by membrane fusion-
dependent compartment mixing. Nat. Cell Biol. 9, 149–159.
Zhang, L., Marcu, M.G., Nau-Staudt, K., Trifaro, J.M., 1996. Recombinant scinderin
enhances exocytosis, an effect blocked by two scinderin-derived actin-binding peptides
and PIP2. Neuron 17, 287–296.
Zuo, X., Zhang, J., Zhang, Y., Hsu, S.C., Zhou, D., Guo, W., 2006. Exo70 interacts with
the Arp2/3 complex and regulates cell migration. Nat. Cell Biol. 8, 1383–1388.
Role of Cortical Cytoskeleton in Exocytosis 137
Author's personal copy
... According to the data shown in the present study, α3β4 nAChR clusters were located very close to active secretory sites formed by t-SNARE proteins and, therefore, near the cortical structures of VDCC associated through F-actin, which are located below the plasma membrane in these neuroendocrine cells. Thus, this cytoskeletal structural element would be responsible for redirecting to specific membrane regions for voltage-dependent calcium channels and microdomains of the secretory machinery [28,29], as well as nAChRs. ...
Article
Full-text available
The heteromeric assembly of α3 and β4 subunits of acetylcholine nicotinic receptors (nAChRs) seems to mediate the secretory response in bovine chromaffin cells. However, there is no information about the localization of these nAChRs in relationship with the secretory active zones in this cellular model. The present work presents the first evidence that, in fact, a population of these receptors is associated through the F-actin cytoskeleton with exocytotic machinery components, as detected by SNAP-25 labeling. Furthermore, we also prove that, upon stimulation, the probability to find α3β4 nAChRs very close to exocytotic events increases with randomized distributions, thus substantiating the clear dynamic behavior of these receptors during the secretory process. Modeling on secretory dynamics and secretory component distributions supports the idea that α3β4 nAChR cluster mobility could help with improving the efficiency of the secretory response of chromaffin cells. Our study is limited by the use of conventional confocal microscopy; in this sense, a strengthening to our conclusions could come from the use of super-resolution microscopy techniques in the near future.
... The fusion of vesicles with other lipid bilayers is essential for intracellular trafficking and release of neurotransmitters and hormones [1][2][3][4]. Release of neurotransmitters and hormones involves the active transport of the vesicles using cytoskeletal elements such as F-actin and microtubules [5,6], tethering and docking of the vesicles with the target membrane [7,8], and finally calcium-induced fusion of the membranes, resulting in the release of vesicular contents to the extracellular media via exocytosis [9,10]. ...
Article
Full-text available
The fusion of membranes is a central part of the physiological processes involving the intracellular transport and maturation of vesicles and the final release of their contents, such as neurotransmitters and hormones, by exocytosis. Traditionally, in this process, proteins, such SNAREs have been considered the essential components of the fusion molecular machinery, while lipids have been seen as merely structural elements. Nevertheless, sphingosine, an intracellular signalling lipid, greatly increases the release of neurotransmitters in neuronal and neuroendocrine cells, affecting the exocytotic fusion mode through the direct interaction with SNAREs. Moreover, recent studies suggest that FTY-720 (Fingolimod), a sphingosine structural analogue used in the treatment of multiple sclerosis, simulates sphingosine in the promotion of exocytosis. Furthermore, this drug also induces the intracellular fusion of organelles such as dense vesicles and mitochondria causing cell death in neuroendocrine cells. Therefore, the effect of sphingosine and synthetic derivatives on the heterologous and homologous fusion of organelles can be considered as a new mechanism of action of sphingolipids influencing important physiological processes, which could underlie therapeutic uses of sphingosine derived lipids in the treatment of neurodegenerative disorders and cancers of neuronal origin such neuroblastoma.
... The copyright holder for this preprint this version posted October 8, 2021. ; https://doi.org/10.1101/2021.10.06.463274 doi: bioRxiv preprint 'prevent' or 'repair' membrane injury via expression of complement regulatory factors or endocytosis/exocytosis, respectively [37][38][39], also remains to be seen. Ultimately, use of the rhesus monkey model of early AMD to uncover the mechanobiological basis for complementmediated choroidal EC loss may help identify new classes of therapeutic targets for effective AMD management in the future. ...
Preprint
Full-text available
Age-related macular degeneration (AMD) is the leading cause of blindness in the aging population. Yet, no therapies exist for ∼85% of all AMD patients who have the dry form that is marked by degeneration of the retinal pigment epithelium (RPE) and underlying choroidal vasculature. As the choroidal vessels are crucial for RPE development and maintenance, understanding how they degenerate may lead to effective therapies for dry AMD. One likely causative factor for choroidal vascular loss is the cytolytic membrane attack complex (MAC) of the complement pathway that is abundant on choroidal vessels of humans with early dry AMD. To examine this possibility, we studied the effect of complement activation on choroidal endothelial cells (ECs) isolated from a rhesus monkey model of early AMD that, we report, exhibits MAC deposition and choriocapillaris endothelial loss similar to that seen in human early AMD. Treatment of choroidal ECs from AMD eyes with complement-competent normal human serum caused extensive actin cytoskeletal injury that was significantly less pronounced in choroidal ECs from young normal monkey eyes. We further show that ECs from AMD eyes are significantly stiffer than their younger counterparts and exhibit peripheral actin organization that is distinct from the longitudinal stress fibers in young ECs. Finally, these differences in complement susceptibility and mechanostructural properties were found to be regulated by the differential activity of small GTPases Rac and Rho because Rac inhibition in AMD cells led to simultaneous reduction in stiffness and complement susceptibility while Rho inhibition in young cells exacerbated complement injury. Thus, by identifying cell stiffness and cytoskeletal regulators Rac and Rho as important determinants of complement susceptibility, the current findings offer a new mechanistic insight into choroidal vascular loss in early AMD that warrants further investigation for assessment of translational potential.
... A similar domain enriched in positive charges that could potentially bind to anionic PA has been described in vesicular SNAREs and reported to be required for membrane destabilization and fusion (Rathore et al., 2019). The actin cytoskeleton, well known to influence several steps leading to exocytosis in neuroendocrine cells (Gutié rrez, 2012;Li et al., 2018;Malacombe et al., 2006), is also a major potential target of PA. First viewed as a physical barrier preventing granule recruitment, cortical actin also acts as an active network that has been proposed to stabilize granule docking sites, control fusion pore lifetime, and/or directly expel granule secretory content. ...
Article
Full-text available
Specific forms of fatty acids are well known to have beneficial health effects, but their precise mechanism of action remains elusive. Phosphatidic acid (PA) produced by phospholipase D1 (PLD1) regulates the sequential stages underlying secretory granule exocytosis in neuroendocrine chromaffin cells, as revealed by pharmacological approaches and genetic mouse models. Lipidomic analysis shows that secretory granule and plasma membranes display distinct and specific composition in PA. Secretagogue-evoked stimulation triggers the selective production of several PA species at the plasma membrane near the sites of active exocytosis. Rescue experiments in cells depleted of PLD1 activity reveal that mono-unsaturated PA restores the number of exocytotic events, possibly by contributing to granule docking, whereas poly-unsaturated PA regulates fusion pore stability and expansion. Altogether, this work provides insight into the roles that subspecies of the same phospholipid may play based on their fatty acyl chain composition.
... Due to its density, cortical actin was long regarded as a diffusion barrier that prevents access of granules to secretory sites and traps secretory vesicles in the cortical network [41,42]. In non-neuronal secretory cells, the role of cortical actin for exocytosis was mostly investigated by chemical disruption of actin cytoskeleton. ...
Article
Full-text available
Cellular secretion depends on exocytosis of secretory vesicles and discharge of vesicle contents. Actin and myosin are essential for pre-fusion and post-fusion stages of exocytosis. Secretory vesicles depend on actin for transport to and attachment at the cell cortex during the pre-fusion phase. Actin coats on fused vesicles contribute to stabilization of large vesicles, active vesicle contraction and/or retrieval of excess membrane during the post-fusion phase. Myosin molecular motors complement the role of actin. Myosin V is required for vesicle trafficking and attachment to cortical actin. Myosin I and II members engage in local remodeling of cortical actin to allow vesicles to get access to the plasma membrane for membrane fusion. Myosins stabilize open fusion pores and contribute to anchoring and contraction of actin coats to facilitate vesicle content release. Actin and myosin function in secretion is regulated by a plethora of interacting regulatory lipids and proteins. Some of these processes have been first described in non-neuronal cells and reflect adaptations to exocytosis of large secretory vesicles and/or secretion of bulky vesicle cargoes. Here we collate the current knowledge and highlight the role of actomyosin during distinct phases of exocytosis in an attempt to identify unifying molecular mechanisms in non-neuronal secretory cells.
Article
Vesicular fusion plays a pivotal role in cellular processes, involving stages like vesicle trafficking, fusion pore formation, content release, and membrane integration or separation. This dynamic process is regulated by a complex interplay of protein assemblies, osmotic forces, and membrane tension, which together maintain a mechanical equilibrium within the cell. Changes in cellular mechanics or external pressures prompt adjustments in this equilibrium, highlighting the system’s adaptability. This review delves into the synergy between intracellular proteins, structural components, and external forces in facilitating vesicular fusion and release. It also explores how cells respond to mechanical stress, maintaining equilibrium and offering insights into vesicle fusion mechanisms and the development of neurological disorders.
Article
Age‐related macular degeneration (AMD) is the leading cause of blindness in the aging population. Yet, no therapies exist for approximately 85% of all AMD patients who have the dry form that is marked by degeneration of the retinal pigmented epithelium (RPE) and underlying choroidal vasculature. As the choroidal vessels are crucial for RPE development and maintenance, understanding how they degenerate may lead to effective therapies for dry AMD. One likely causative factor for choroidal vascular loss is the cytolytic membrane attack complex (MAC) of the complement pathway that is abundant on choroidal vessels of humans with early dry AMD. To examine this possibility, we studied the effect of complement activation on choroidal endothelial cells (ECs) isolated from a rhesus monkey model of early AMD that, we report, exhibits MAC deposition and choriocapillaris endothelial loss similar to that seen in human early AMD. Treatment of choroidal ECs from AMD eyes with complement‐competent normal human serum caused extensive actin cytoskeletal injury that was significantly less pronounced in choroidal ECs from young normal monkey eyes. We further show that ECs from AMD eyes are significantly stiffer than their younger counterparts and exhibit peripheral actin organization that is distinct from the longitudinal stress fibers in young ECs. Finally, these differences in complement susceptibility and mechanostructural properties were found to be regulated by the differential activity of the small GTPases Rac and Rho, because Rac inhibition in AMD cells led to simultaneous reduction in stiffness and complement susceptibility while Rho inhibition in young cells exacerbated complement injury. Thus, by identifying cell stiffness and cytoskeletal regulators Rac and Rho as important determinants of complement susceptibility, the current findings offer a new mechanistic insight into choroidal vascular loss in early AMD that warrants further investigation for assessment of translational potential. This article is protected by copyright. All rights reserved.
Chapter
Full-text available
Models for how diatoms move were devised as early as 1753 and up to the present. Most of them have not been pursued to the point of proof or disproof. Elements of some of the oldest models, fanciful and disputed, still help in thinking about this unsolved problem, which has been tackled by a wide variety of scientists, amateur and professional. The current models are full of holes, and are not based on modern understandings of secretory mechanisms, cytoplasmic streaming, and physical chemistry. A testable working model is presented, which is an amalgam of old and new. In this model, the major component of the raphe fluid is a polysaccharide designated as “raphan” which is synthesized by a membrane protein “raphan synthase,” either in Golgi vesicles called crystalloid bodies or in the cell membrane. The raphan synthase is transported to each raphe via cytoplasmic streaming engendered by its pair of adjacent microfilament bundles and attached myosin motor molecules, hydrodynamically inducing motion of the whole fluid cell membrane. The raphan is initially hydrophobic and fills the hydrophobic raphe, which is a capillary nanochannel. Proteins in the raphe fluid trigger hydration of the raphan on contact with a substrate. The hydrated raphan can no longer wet the raphe and exits, producing the diatom trail. The capillary force generated is sufficient to explain the force that a moving diatom can exert, whereas cytoplasmic streaming is 10,000 times weaker. The cytoplasmic streaming controls the direction of the diatom, whereas capillarity provides the force. Capillary motion is sustained by hydration of the raphan. By analogy with an automobile, the steering wheel requires a small force, which controls an engine that produces a much larger force. What turns the steering wheel or determines the direction of the cytoplasmic streaming is a higher order problem of behavior.
Article
Full-text available
Previously we reported that adrenal chromaffin cells exposed to a 5 ns, 5 MV/m pulse release the catecholamines norepinephrine (NE) and epinephrine (EPI) in a Ca²⁺-dependent manner. Here we determined that NE and EPI release increased with pulse number (one versus five and ten pulses at 1 Hz), established that release occurs by exocytosis, and characterized the exocytotic response in real-time. Evidence of an exocytotic mechanism was the appearance of dopamine-β-hydroxylase on the plasma membrane, and the demonstration by total internal reflection fluorescence microscopy studies that a train of five or ten pulses at 1 Hz triggered the release of the fluorescent dye acridine orange from secretory granules. Release events were Ca²⁺-dependent, longer-lived relative to those evoked by nicotinic receptor stimulation, and occurred with a delay of several seconds despite an immediate rise in Ca²⁺. In complementary studies, cells labeled with the plasma membrane fluorescent dye FM 1-43 and exposed to a train of ten pulses at 1 Hz underwent Ca²⁺-dependent increases in FM 1-43 fluorescence indicative of granule fusion with the plasma membrane due to exocytosis. These results demonstrate the effectiveness of ultrashort electric pulses for stimulating catecholamine release, signifying their promise as a novel electrostimulation modality for neurosecretion.
Chapter
To study the formation and the architecture of exocytotic site, we generated plasma membrane (PM) sheets on electron microscopy grids to visualize the membrane organization and quantitatively analyze distributions of specific proteins and lipids. This technique allows observing the cytoplasmic face of the plasma membrane by transmission electron microscope. The principle of this approach relies on application of mechanical forces to break open cells. The exposed inner membrane surface can then be visualized with different electron-dense colorations, and specific proteins or lipids can be detected with gold-conjugated probes. Moreover, the membrane sheets are sufficiently resistant to support automated acquisition of multiple-tilt projections, and thus electron tomography allows to obtain three-dimensional (3D) ultrastructural images of secretory granule docked to the plasma membrane.
Chapter
The catecholamine-storing organlles of adrenal medulla, the chromaffin granules, have been analyzed in great detail. It is a characteristic feature of these membrane-limited vesicles that their content, which is destined for secretion, consists of both macromolecular components, that is, glycoproteins and mucopolysaccharides, and small molecules, including catecholamines, nucleotides, and calcium. This chapter discusses the formation of chromaffin granules to provide a conceptual basis for studies on the biogenesis of other organelles storing small molecules. It discusses the synthesis of the macromolecular components and their assembly into newly formed chromaffin granules. The chapter also describes various uptake mechanisms that endow these particles with the ability to accumulate a high concentration of small molecules. The macromolecular components of these granules destined for secretion are glycoproteins and mucopolysaccharides. In accordance with other secretory tissues, these components are synthesized in the rough endoplasmic reticulum with further additions on their way to the Golgi region, where the secretory products become packaged into new chromaffin granules.
Article
Neurons transmit signals to their target cells at specialized contact sites called synapses. At chemical synapses this signal propagation is mediated by the fusion of neurotransmitter-filled synaptic vesicles with the presynaptic plasma membrane and the subsequent release of transmitter into the synaptic cleft. In contrast to other fusion events in the cell the fusion of synaptic vesicles is extremely fast, highly regulated and spatially restricted but also dynamically modulated. Fusion occurs only at a specialized region of the presynaptic plasma membrane, called the active zone. Ultrastructural studies have shown that the active zone is precisely aligned with the postsynaptic reception apparatus and that the plasma membrane on both sides of the synaptic cleft is marked by an electron-dense structure. In the presynaptic terminal this electron-dense cytoskeletal matrix is referred to as the cytomatrix at the active zone (CAZ). So far five protein families whose members are highly enriched at active zones have been identified: Munc13s, RIMs, ELKS, Bassoon/Piccolo und Liprin-α In recent years studies using genetic, biochemical, structural and electrophysiological approaches have begun to elucidate how these proteins are involved in the regulation of synaptic vesicle exocytosis, in mediating use-dependent changes during different forms of plasticity and in the structural organization of the active zone.
Article
Initial integrin-mediated cell-matrix adhesions (focal complexes) appear underneath the lamellipodia, in the regions of the ''fast'' centripetal flow driven by actin polymerization. Once formed, these adhesions convert the flow behind them into a ''slow'', myosin II-driven mode. Some focal complexes then turn into elongated focal adhesions (FAs) associated with contractile actomyosin bundles (stress fibers). Myosin II inhibition does not suppress formation of focal complexes but blocks their conversion into mature FAs and further FA growth. Application of external pulling force promotes FA growth even under conditions when myosin II activity is blocked. Thus, individual FAs behave as mechanosensors responding to the application of force by directional assembly. We proposed a thermodynamic model for the mechanosensitivity of FAs, taking into account that an elastic molecular aggregate subject to pulling forces tends to grow in the direction of force application by incorporating additional subunits. This simple model can explain a variety of processes typical of FA behavior. Assembly of FAs is triggered by the small G-protein Rho via activation of two major targets, Rho-associated kinase (ROCK) and the formin homology protein, Dia1. ROCK controls creation of myosin II-driven forces, while Dia1 is involved in the response of FAs to these forces. Expression of the active form of Dia1, allows the external force-induced assembly of mature FAs, even in conditions when Rho is inhibited. Conversely, downregulation of Dia1 by siRNA prevents FA maturation even if Rho is activated. Dia1 and other formins cap barbed (fast growing) ends of actin filaments, allowing insertion of the new actin monomers. We suggested a novel mechanism of such ''leaky'' capping based on an assumption of elasticity of the formin/barbed end complex. Our model predicts that formin-mediated actin polymerization should be greatly enhanced by application of external pulling force. Thus, the formin-actin complex might represent an elementary mechanosensing device responding to force by enhancement of actin assembly. In addition to its role in actin polymerization, Dia1 seems to be involved in formation of links between actin filaments and microtubules affecting microtubule dynamics. Alpha-tubulin deacetylase HDAC6 cooperates with Dia1 in formation of such links. Since microtubules are known to (A.D. Bershadsky). promote FA disassembly, the Dia1-mediated effect on microtubule dynamics may possibly play a role in the negative feedback loop controlling size and turnover of FAs.
Article
Immunofluorescence and cytochemical studies have demonstrated that filamentous actin is mainly localized in the cortical surface of the chromaffin cell. It has been suggested that these actin filament networks act as a barrier to the secretory granules, impeding their contact with the plasma membrane. Stimulation of chromaffin cells produces a disassembly of actin filament networks, implying the removal of the barrier. The presence of gelsolin and scinderin, two Ca(2+)-dependent actin filament severing proteins, in the cortical surface of the chromaffin cells, suggests the possibility that cell stimulation brings about activation of one or more actin filament severing proteins with the consequent disruption of actin networks. Therefore, biochemical studies and fluorescence microscopy experiments with scinderin and gelsolin antibodies and rhodamine-phalloidin, a probe for filamentous actin, were performed in cultured chromaffin cells to study the distribution of scinderin, gelsolin, and filamentous actin during cell stimulation and to correlate the possible changes with catecholamine secretion. Here we report that during nicotinic stimulation or K(+)-evoked depolarization, subcortical scinderin but not gelsolin is redistributed and that this redistribution precedes catecholamine secretion. The rearrangement of scinderin in patches is mediated by nicotinic receptors. Cell stimulation produces similar patterns of distribution of scinderin and filamentous actin. However, after the removal of the stimulus, the recovery of scinderin cortical pattern of distribution is faster than F-actin reassembly, suggesting that scinderin is bound in the cortical region of the cell to a component other than F-actin. We also demonstrate that peripheral actin filament disassembly and subplasmalemmal scinderin redistribution are calcium-dependent events. Moreover, experiments with an antibody against dopamine-beta-hydroxylase suggest that exocytosis sites are preferentially localized to areas of F-actin disassembly.
Article
ELUCIDATION of the reactions responsible for the calcium-regulated fusion of secretory granules with the plasma membrane in secretory cells would be facilitated by the identification of participant proteins having known biochemical activities. The successful characterization of cytosolic1-3 and vesicle4,5 proteins that may function in calcium-regulated secretion has not yet revealed the molecular events underlying this process. Regulated secretion consists of sequential priming and triggering steps which depend on ATP and Ca2+, respectively, and require distinct cytosolic proteins6. Characterization of priming-specific factors (PEP proteins) should enable the ATP-requiring reactions to be identified. Here we show that one of the mammalian priming factors (PEP3) is identical to phosphatidylinositol transfer protein (PITP)7. The physiological role of PITP was previously unknown. We also find that SEC14p, the yeast phosphatidylinositol transfer protein which is essential for constitutive secretion8-10, can substitute for PEP3/PITP in priming. Our results indicate that a role for phospholipid transfer proteins is conserved in the constitutive and regulated secretory pathways.
Article
Botulinum C2 toxin is known to ADP-ribosylate actin. The toxin effect was studied on [3H]noradrenaline secretion of PC12 cells. [3H]Noradrenaline release was stimulated five- to 15-fold by carbachol (100 μM) or K+ (50 mM) and 10–30-fold by the ionophore A23187 (5 μM). Pretreatment of PC12 cells with botulinum C2 toxin for 4–8 h at 20°C, increased carbachol-, K+-, and A23187-induced, but not basal, [3H]noradrenaline release maximally 1.5- to threefold, whereas ≥75% of the cellular actin pool was ADP-ribosylated. Treatment of PC12 cells with botulinum C2 toxin for up to 1 h at 37°C also increased stimulated [3H]noradrenaline secretion, whereas toxin treatment for > 1 h decreased the enhanced [3H]noradrenaline release stimulated by carbachol and K+ but not by A23187. Concomitantly with toxin-induced stimulation of secretion, 20–50% of the cellular actin was ADP-ribosylated, whereas >60% of actin was modified when exocytosis was attenuated. The data indicate that ADP-ribosylation of actin by botulinum C2 toxin largely modulates stimulation of [3H]noradrenaline release. Moreover, the biphasic toxin effects suggest that distinct mechanisms are involved in the role of actin in secretion.
Article
Regulated secretion requires both calcium and MgATP. Studies in diverse secretory systems indicate that ATP is required to prime the exocytotic apparatus whereas Ca2+ triggers the final ATP-independent fusion event. In this paper, we examine the possible role of trimeric G proteins in these two steps of exocytosis in chromaffin cells. We show that in the presence of low concentrations of Mg2+, mastoparan selectively stimulates G proteins associated with purified chromaffin granule membranes. Under similar conditions in permeabilized chromaffin cells, mastoparan inhibits ATP-dependent secretion but is unable to trigger ATP-independent release. This inhibitory effect of mastoparan on secretion was specifically reversed by anti-Go antibodies and a synthetic peptide corresponding to the carboxyl terminus of Go. In contrast, mastoparan required millimolar Mg2+ for the activation of plasma membrane-bound G proteins and stimulation of ATP-independent secretion in permeabilized chromaffin cells. The latter effect was completely inhibited by anti-Gi3 antibodies and a synthetic peptide corresponding to the carboxyl terminus of Gi3. By confocal immunofluorescence and immunoreplica analysis, we provide evidence that in chromaffin cells Go is preferentially associated with secretory granules, while Gi3 is essentially present on the plasma membrane. Our findings suggest that these two trimeric G proteins act in series in the exocytotic pathway in chromaffin cells: a secretory granule-associated Go protein controls the ATP-dependent priming reaction, whereas a plasma membrane-bound Gi3 protein is involved in the late calcium-dependent fusion step, which does not require ATP.