ArticlePDF AvailableLiterature Review

Expression of two neuronal markers, growth-associated protein 43 and neuron-specific enolase, in rat glial cells

Authors:

Abstract and Figures

Recent studies have revealed that proteins such as growth-associated protein 43 (GAP-43) and neuron-specific enolase (NSE), believed for many years to be expressed exclusively in neurons, are also present in glial cells under some circumstances. Here we present an overview of these observations. GAP-43 is expressed both in vitro and in vivo transiently in immature rat oligodendroglial cells of the central nervous system, in Schwann cell precursors, and in non-myelin-forming Schwann cells of the peripheral nervous system. GAP-43 mRNA is also present in oligodendroglial cells and Schwann cells, indicating that GAP-43 is synthesized in these cells. GAP-43 is also expressed in type 2 astrocytes (stellate-shaped astrocytes) and in some reactive astrocytes but not in type 1 astrocytes (flat protoplasmic astrocytes). These results suggest that GAP-43 plays a more general role in neural plasticity during development of the central and peripheral nervous systems. NSE enzymatic activity and protein and mRNA have been detected in rat cultured oligodendrocytes at levels comparable to those of cultured neurons. NSE expression increases during the differentiation of oligodendrocyte precursors into oligodendrocytes. In vivo, NSE protein is expressed in differentiating oligodendrocytes and is repressed in fully mature adult cells. The upregulation of NSE in differentiating oligodendrocytes coincides with the formation of large amounts of membrane structures and of protoplasmic processes. Similarly, NSE becomes detectable in glial neoplasms and reactive glial cells at the time when these cells undergo morphological changes. The expression of the glycolytic isozyme NSE in these cells, which do not normally contain it, could reflect a response to higher energy demands. This expression may also be related to the neurotrophic and neuroprotective properties demonstrated for this enolase isoform. NSE activity and protein and mRNA have also been found in cultured rat type 1-like astrocytes but at much lower levels than in neurons and oligodendrocytes. Thus GAP-43 and NSE should be used with caution as neuron-specific markers in studies of normal and pathological neural development.
Content may be subject to copyright.
&p.1:Abstract Recent studies have revealed that proteins
such as growth-associated protein 43 (GAP-43) and neu-
ron-specific enolase (NSE), believed for many years to
be expressed exclusively in neurons, are also present in
glial cells under some circumstances. Here we present an
overview of these observations. GAP-43 is expressed
both in vitro and in vivo transiently in immature rat oli-
godendroglial cells of the central nervous system, in
Schwann cell precursors, and in non-myelin-forming
Schwann cells of the peripheral nervous system. GAP-43
mRNA is also present in oligodendroglial cells and
Schwann cells, indicating that GAP-43 is synthesized in
these cells. GAP-43 is also expressed in type 2 astrocytes
(stellate-shaped astrocytes) and in some reactive astro-
cytes but not in type 1 astrocytes (flat protoplasmic as-
trocytes). These results suggest that GAP-43 plays a
more general role in neural plasticity during develop-
ment of the central and peripheral nervous systems. NSE
enzymatic activity and protein and mRNA have been de-
tected in rat cultured oligodendrocytes at levels compara-
ble to those of cultured neurons. NSE expression in-
creases during the differentiation of oligodendrocyte pre-
cursors into oligodendrocytes. In vivo, NSE protein is
expressed in differentiating oligodendrocytes and is re-
pressed in fully mature adult cells. The upregulation of
NSE in differentiating oligodendrocytes coincides with
the formation of large amounts of membrane structures
and of protoplasmic processes. Similarly, NSE becomes
detectable in glial neoplasms and reactive glial cells at
the time when these cells undergo morphological chang-
es. The expression of the glycolytic isozyme NSE in
these cells, which do not normally contain it, could re-
flect a response to higher energy demands. This expres-
sion may also be related to the neurotrophic and neuro-
protective properties demonstrated for this enolase iso-
form. NSE activity and protein and mRNA have also
been found in cultured rat type 1-like astrocytes but at
much lower levels than in neurons and oligodendrocytes.
Thus GAP-43 and NSE should be used with caution as
neuron-specific markers in studies of normal and patho-
logical neural development.
&kwd:Key words Growth-associated protein 43 · Neuron-
specific enolase · Central nervous system · Peripheral
nervous system · Glia
Abbreviations bFGF Basic fibroblast growth factor ·
CNS Central nervous system · GAP-43 Growth-
associated protein 43 · GC Galactocerebroside ·
GFAP Glial fibrillary acidic protein · MAP microtubule-
associated protein · MBP Myelin basic protein ·
NSE Neuron-specific enolase · PNS Peripheral nervous
system&bdy:
Introduction
During the development of the nervous system changes
in the expression of neuron-specific markers are indica-
Monique Sensenbrenner (
) · Jean-Christophe Deloulme
Laboratoire de Neurobiologie Ontogénique, CNRS ERS 110,
Centre de Neurochimie, 5 rue Blaise Pascal,
F-67084 Strasbourg Cedex, France
Marguerite Lucas
Biochimie Cellulaire, CNRS UPR 9065, Collège de France,
Paris, France.
J Mol Med (1997) 75:653–663 © Springer-Verlag 1997
REVIEW
&roles:Monique Sensenbrenner · Marguerite Lucas
Jean-Christophe Deloulme
Expression of two neuronal markers, growth-associated protein 43
and neuron-specific enolase, in rat glial cells
&misc:Received: 27 January 1997 / Accepted: 9 May 1997
MONIQUE SENSENBRENNER
received the degree of Docteur
és Sciences from the Universi-
ty of Strasbourg, France. She
is presently Directeur de Re-
cherche at CNRS and Chief of
a CNRS Unit (Laboratoire de
Neurobiologie Ontogénique)
at the Centre de Neurochimie
du CNRS in Strasbourg. She
has been a pioneer in her stud-
ies on neural cell cultures, and
her major research interests
concern the effects and mech-
anisms of action of growth
factors on neural cell develop-
ment.&/fn-block:
tive of neuronal cell differentiation and maturation.
However, recent findings from several laboratories, in-
cluding our own, have demonstrated that neuron-specific
proteins such as the growth-associated protein 43 (GAP-
43) and the neuron-specific enolase (NSE) are not con-
fined to neurons but may occur also in glial cells.
GAP-43, a membrane-associated neuron-specific phos-
phoprotein, was first identified as a protein whose ex-
pression is associated with neuronal development and
neurite outgrowth in the nervous system. Proteins similar
to GAP-43 were later discovered in several laboratories,
and molecular cloning has confirmed that GAP-43 is
identical to neuromodulin, protein P-57, B-50, pp46, and
F1 [1, 2] Expression of GAP-43 has been described for
many years to occur exclusively in neurons both in vivo
[1] and in vitro [3, 4]. Moreover, it has been shown that
this protein is synthesized in the neuronal somata, rapid-
ly transported throughout axons, and accumulates in
growth cones and presynaptic terminals. It has also been
demonstrated that GAP-43 levels remain high during
neuronal development; its expression then declines
markedly following the establishment of final synapto-
genesis [1]. Some mature neurons, however, continue to
express high amounts of GAP-43 throughout adult life
[5]. Moreover, GAP-43 is reexpressed in adult neurons
following nervous system damage and during regenera-
tion processes [6]. GAP-43 was therefore thought to be
involved in the initiation of neurite outgrowth, axonal
elongation, axonal regeneration, synaptic plasticity and
somehow in neurotransmitter release [7, 8].
The notion of GAP-43 as a neuron-specific protein
has been questionable since recent findings from our lab-
oratory and from others have shown that GAP-43 expres-
sion also occurs in nonneuronal cells under some cir-
cumstances. The presence of GAP-43 has been demon-
strated in various glial cells in vitro and in vivo (see
“GAP-43 and glial cell development”). More recent find-
ings indicate that GAP-43 may even be expressed tran-
siently in nonneural cells, such as myoblasts [9, 10].
In 1972, it was demonstrated that the accumulation of
the 14-3-2 protein is a regionally selective process which
can be related to the differential maturation of the CNS
[11]. Later it was shown that the 14-3-2 protein is identi-
cal to the γ subunit of the enolase enzyme [12].
Enolase (2-phospho-
D-glycerate hydrolase, EC
4.2.1.11) is a glycolytic enzyme active as a dimer. In
higher vertebrates these dimers are composed of three
subunits α, β, and γ encoded by different genes. The αα
homodimer is found predominantly in immature cells
and remains expressed in most adult tissues, including
brain. During ontogenesis a switch towards specific iso-
forms occurs, from αα to αβ and ββ in the striated mus-
cle, and from αα to αγ and γγ in the nervous system,
specifically in the neuronal lineage [13]. The functional
significance of this switch towards specific isoforms in
two cell types with high energy demands is not clear.
Two nonexclusive hypothesis can be proposed. First, the
regulation of expression of each enolase subunit can be
specifically controlled at the transcriptional and/or post-
transcriptional levels and thereby adapted to the specific
energy demands of different cell types. Second, the pro-
teins themselves may have specific characteristics lead-
ing them to optimal functioning in their respective envi-
ronmental conditions. It has been shown, for example,
that γγ enolase is more resistant than αα enolase to dena-
turation by a high concentration of chloride ions as can
be found in active neurons. The appearance and expres-
sion of NSE (αγ + γγ enolases) are correlated directly
with structural and functional differentiation processes of
the central nervous system (CNS). Moreover, a net in-
crease in γ enolase gene product is related to the differ-
entiation and the synaptic activity of neuronal cells [14].
In adult tissues high levels of γ enolase are present
only in neurons and neuroendocrine cells. However, ex-
pression of γ enolase is less restricted than previously re-
ported since sensitive immunoassays have demonstrated
significant levels of γ enolase in differentiated tissues
other than nervous tissues [15, 16]. It has been shown
that early rat embryos contain a nonnegligible amount of
γ enolase before the formation of neural tissues [17], and
it has been suggested that γ enolase is rather common in
immature cells [18]. Numerous tumors of cells originat-
ing from various tissues express the γ enolase gene. Con-
cerning tumors of neural origin, glioblastoma, astrocyto-
ma, oligodendroglioma, and schwannoma contain NSE.
This enzyme has also been found in reactive astrocytes.
Recently the presence of NSE has been described in nor-
mal glial cells in vitro and in vivo (see “NSE and glial
cell development”). Neurons in culture express αα, αγ,
and γγ enolases, and the neuron specific isoforms accu-
mulate as differentiation proceeds with time in culture
[19–22].
This review summarizes the recent findings that re-
port on regulation of the expression of GAP-43 and NSE
in various glial populations from the central (CNS) and
peripheral (PNS) nervous systems during in vitro and in
vivo development and discusses the functional signifi-
cance of these observations. The review is divided into
three sections. First, we provide some information on the
various types of vertebrate glial cells in vivo and in cul-
ture. We then present results concerning the expression
of GAP-43 in glial cells. Finally, we address the issue of
the presence of NSE in glial cells.
Glial cells in vivo and in culture
The two major macroglial cell types in the vertebrate
CNS are astrocytes and oligodendrocytes [23]. Fibrous
astrocytes have long, relatively straight processes with
few branches. They have numerous gliofilaments distrib-
uted throughout their cytoplasm. Protoplasmic astrocytes
have thick, extensively branched processes. Their cyto-
plasm contains relatively few gliofilaments. Fibrous as-
trocytes predominate among the myelinated axons in the
white matter, while protoplasmic astrocytes are com-
monly found among the neuronal cell bodies in the gray
matter.
654
During development of the rat and mouse brain astro-
glial precursor cells are generated primarily from the
subventricular zone in the early embryonic and postnatal
periods. Some astrocytes also develop from radial glial
cells, which arise very early in development from the
ventricular zone but are transitory and act as a guiding
substrate for the migration of neurons. Oligodendroglial
precursor cells also arise prenatally from the subventric-
ular zone and proliferate actively during the first 2 weeks
after birth. Both types of precursors then migrate into the
neighboring gray and white matter where they differenti-
ate into astrocytes and oligodendrocytes. In the white
matter oligodendrocytes are closely associated with mye-
linated fibers. Astroglial and oligodendroglial cells are
able to proliferate during the whole postnatal life and
even in aged animals.
Schwann cells are the predominant glial cell type of
the vertebrate PNS and derive from neural crest cells
during the embryonic period [23]. Two distinct subtypes
of Schwann cells have been characterized, myelinating
and nonmyelinating. However, the phenotype of these
cells depends on their association with neurites and un-
der certain experimental conditions can be changed; non-
myelin-forming Schwann cells can become myelin-form-
ing Schwann cells and vice versa.
Many functions have been attributed to astroglial cells
[23]. They are considered to be the supporting cells in
the nervous tissue. They interact with other neural cells
and respond to chemical signals from neurons by releas-
ing molecules such as neuroactive and neurotrophic sub-
stances. Astrocytes also play a role in neural develop-
ment and regeneration processes by promoting neuronal
migration and neurite outgrowth. An important role of
astroglial cells is contributing to the homeostasis of the
extracellular environment. Finally, astrocytes respond to
brain injury, become activated, change their shape, and
transform into reactive fibrous astrocytes. Such reactive
gliosis occurs also under several neuropathological con-
ditions.
For oligodendrocytes and myelinating Schwann cells
a functional role is clearly defined, consisting in myelin-
ation. During this process the cells produce high quanti-
ties of plasma membrane and insert into the lipid matrix
the myelin-specific proteins that participate in the orga-
nization and maintenance of the myelin sheath. Schwann
cells maintain the ability to dedifferentiate after nerve in-
jury and reenter the cell cycle, providing new Schwann
cells for the regeneration process [23].
In culture two populations of astrocytes, the epithelio-
id flat type 1 astrocyte and the stellate type 2 astrocyte,
can be distinguished on the basis of morphological, anti-
genic, and functional criteria [24]. These type 1 and type
2 astroglia have been suggested to reflect protoplasmic
and fibrous astrocytes of the brain, respectively. More-
over, in vitro observations indicate that bipotential oligo-
dendrocyte-type 2 astrocyte precursor cells, termed O-
2A progenitors (characterized by the monoclonal anti-
body A2B5 that binds to polysialogangliosides) differen-
tiate into oligodendrocytes when cultured in a serum-
free, chemically defined medium [24]. On the other
hand, these glial progenitor cells develop into type 2 as-
trocytes when maintained in the presence of serum. The
flat, polygonal-shaped type 1 astrocytes appear to derive
from a different lineage than the O-2A lineage and can
be cultured in either the presence or the absence of se-
rum [25, 26].
Astroglial and oligodendroglial cell cultures are ob-
tained from newborn rat cerebral hemispheres [25–27].
The astrocytes (type 1 and type 2 astrocytes) are charac-
terized by glial fibrillary acidic protein (GFAP) immuno-
reactivity. The progressive developmental stages of oli-
godendroglial cells can be followed by immunostaining
against specific surface markers, such as A2B5, a marker
of 0-2A precursors; O4, a marker of immature oligoden-
drocytes binding to surface sulfatide; and galactocere-
broside (GC) and myelin basic protein (MBP), markers
of mature oligodendrocytes.
Schwann cells, in contact with sensory neurons, are
present in cultures of dissociated adult rat dorsal root
ganglia [28]. Pure cultures of Schwann cells deprived of
axonal contact are obtained from 1- to 5-day-old rat pe-
ripheral sciatic nerves [29, 30]. The Schwann cells can
be characterized by S100 protein immunoreactivity.
GAP-43 and glial cell development
CNS macroglial cells and GAP-43
The expression of a protein with a similar molecular
weight and isoelectric point as GAP-43 in macroglial
cells was first mentioned by the group of Vitkovic in
1987 [31]. This protein was found to be associated with
the plasma membranes of both the somata and cellular
processes of cultured rat cortical type 1 astrocytes. Fur-
ther in vitro studies of this group confirmed the presence
of GAP-43 in cultured astrocytes [32, 33]. In addition,
these cells also express GAP-43 mRNA, indicating that
GAP-43 protein is synthesized rather than taken up by
the astrocytes [34]. These authors point out that GAP-43
expression decreases in type 1 astrocytes with time in
culture. After 2 weeks these glial cell type contains little
if any GAP-43 protein and mRNA.
We have also undertaken in vitro investigations to an-
alyze in more detail the presence and the regulation of
GAP-43 expression in both astrocytes and oligodendro-
glial cells derived from rat brain during their differentia-
tion in culture. Immunocytochemical observations indi-
cate that GAP-43 is present in oligodendroglial cells and
type 2 astrocytes but is absent in type 1 astrocytes [35].
The expression of GAP-43 protein and mRNA in oligo-
dendroglial cells and their absence in protoplasmic astro-
cytes have been confirmed by western and northern blot
analysis [35]. In addition, we have demonstrated that
principally the O-2A bipotential glial precursors (mor-
phologically immature A2B5 positive cells) express
GAP-43 protein, that GAP-43 is also expressed in O4
(sulfatide) positive cells, corresponding to a later devel-
655
opmental stage of the oligodendroglial precursor cells
(Fig. 1), and that cells which acquire a typical oligoden-
drocyte morphology and begin to express the GC antigen
become progressively GAP-43 negative. The precursor
glial cell cultures also show the presence of GAP-43
mRNA, indicating that this protein is synthesized early
in the oligodendrocyte lineage [35]. In the presence of
basic fibroblast growth factor (bFGF), a growth factor
known to enhance the multiplication of oligodendroglial
cells, the proliferation of the GAP-43 positive glial pre-
cursors is stimulated, and the levels of GAP-43 protein
and mRNA are increased [36].
The group of Vitkovic [37] have discussed a constitu-
tive expression of GAP-43 in cultured oligodendrocytes.
Other authors, however, have observed that GAP-43 is
expressed in cultured O-2A progenitor cells but seldom
in GC positive and never in MBP positive mature cul-
tured oligodendrocytes [38]. Moreover, an in vivo immu-
nohistochemical study performed by the latter group on
sections of rat brain demontrated a similar regulation of
GAP-43 in oligodendroglial cells during development.
Indeed, GAP-43 immunoreactivity was seen to be locat-
ed only in the membranes of immature oligodendroglial
cells, and no GAP-43 was found in mature oligodendro-
cytes [38].
Taken together these in vitro and in vivo data indicate
that GAP-43 is associated principally with glial precur-
sor cells, and that its expression characterizes an early
differentiation stage of the oligodendroglial cell lineage
and is lost when these cells differentiate into more ma-
ture oligodendrocytes. Such a transient expression of
GAP-43 has also been described for other nonneuronal
cells belonging to the muscle cell lineage [9].
Biochemical characterization of the GAP-43 has re-
vealed both similarities and differences between the
GAP-43 protein of cultured oligodendroglial and cul-
tured neuronal cells [35]. While three distinct protein
spots are resolved for neurons in two-dimensional gel
electrophoresis, the oligodendroglial GAP-43 migrates
as a single protein spot, with an isoelectric point identi-
cal to that of the more basic form of the neuronal pro-
tein. Moreover, the GAP-43 form of oligodendroglial
cells shows a lower electrophoretic mobility than the
neuronal GAP-43 forms, which may arise from varia-
tions of protein conformation and/or posttranslational
modifications.
Finally, we have also found that GAP-43 immunore-
activity remains in O-2A progenitor cells which differen-
tiate into type 2 astrocytes in serum-containing medium
(Fig. 2) [36], in agreement with the in vitro observations
of Vitkovic’s group [34, 37] and Wilkin’s group [38].
The identification of type 2 astrocytes in tissue sections
has been difficult and controversial. However, it has been
suggested that reactive astrocytes in the damaged hippo-
campus have the phenotypic features of type 2 astrocytes
and express GAP-43 [39]. By contrast, another group has
reported that normal and reactive astrocytes in brain sec-
tions never show GAP-43 immunoreactivity [38]. These
in vivo results should be taken with caution, and further
experiments are required to clarify such contradictory
data.
Schwann cells and GAP-43
Schwann cells, which are the functional counterparts of
oligodendrocytes in the PNS also express GAP-43 both
in vitro and in vivo. The presence of GAP-43 in
Schwann cells, however, has raised controversy. An in
vitro study reported that Schwann/satellite cells grown
together with sensory neurons in dorsal root ganglion
cell cultures express low levels of GAP-43 during the
first 8 days of culture [28]. In this mixed neuronal-glial
culture system the possibility could not be eliminated
that the glial cells might take up the protein from the
neurons. Further in vitro studies have demonstrated that
about 25–40% of Schwann cells cultured in the absence
of neurons can synthesize GAP-43, which is identical to
the neuronal protein [29, 30]. Moreover, cultured
656
Fig. 1a, b Double-immunofluorescence staining of oligodendro-
glial cells grown for 4 days under serum-free condition in second-
ary culture. (The same microscopic fields are shown in a and b.) a
GAP-43 TRITC immunofluorescence. b O4 FITC immunofluores-
cence. Note that all O4 positive oligodendroglial cells are also
GAP-43 positive. Arrows, two GAP-43 positive glial precursor
cells which are O4 negative. Bar=10 µm&/fig.c:
Schwann cells have also been shown to express GAP-43
mRNA [30, 40].
Several in vivo observations support the results of the
in vitro studies. GAP-43 is detectable in all Schwann cell
precursors in developing peripheral nerves [29]. More-
over, this protein is expressed in most mature non-my-
elin-forming Schwann cells [29, 30]. By contrast, adult
myelin-forming Schwann cells are not normally immu-
noreactive for GAP-43 but begin to synthesize GAP-43
only several weeks following peripheral nerve injury [29,
30]. Furthermore, expression of GAP-43 mRNA is high
in Schwann cells of degenerating peripheral nerves [30,
40]. Other authors have shown that non-myelin-forming
Schwann cells present on adult skeletal muscle motor
endplates are not GAP-43 immunoreactive [41]. These
cells, however, become positive for a GAP-43 mRNA in
situ hybridization signal and are GAP-43 immunoreac-
tive shortly after denervation of the muscles by either a
cut or crush lesion of the sciatic nerve [41]. A parallel
has been noted between GAP-43 expression and the
elaboration by denervated Schwann cells of long fine
processes and branches. After reinnervation occurs, the
Schwann cell processes disappear, and the cells loose
GAP-43 immunoreactivity. This indicates that GAP-43
expression in Schwann cells is state dependent, and in
this respect it resembles neuronal expression of the pro-
tein.
The developmental expression of GAP-43 in rat glial
cells of the CNS and PNS is summarized in Table 1.
657
Fig. 2a, b Double-immunoflu-
orescence staining of type 2 as-
trocytes grown for 1 week in
the presence of serum in sec-
ondary culture. (The same mi-
croscopic fields are shown in a
and b.) a GAP-43 FITC immu-
nofluorescence. b GFAP
TRITC-immunofluorescence.
Note that the GFAP-positive
astrocytes are also GAP-43
positive. Bar=10 µm&/fig.c:
Table 1 Rat glial cells: expres-
sion of GAP-43: comparison
of in vitro and in vivo results.
(M myelinating Schwann cells,
Non-M nonmyelinating
Schwann cells, R reactive
Schwann cells; n.d. not deter-
mined)&/tbl.c:&tbl.b:
Develop- Glial cell type
mental stage
Astrocytes Oligodendrocytes Schwann cells
Type 1 Type 2 (in vitro)
or reactive (in vivo)
Precursors Culture ± + + +
In vivo n.d. n.d. n.d. +
Immature Culture ± + + +
cells In vivo n.d. n.d. + +
Mature Culture + +
cells In vivo +/ M Non-M R
+/ +
&/tbl.b:
NSE and glial cell development
NSE in neoplastic glial cells and reactive astrocytes
Studies performed a number of years ago revealed that
many types of CNS and PNS tumors of glial origin such
as glioblastoma, astrocytoma, oligodendroglioma, and
schwannoma contain the γ subunit of the glycolytic en-
zyme enolase [42–45]. However, no relationship between
the presence of γ enolase and cellular anaplasia has been
found. The putative cell type from which these tumors
originate was found at that time not to contain demon-
strable NSE. All these findings which indicate that γ
enolase is detected not only in neuronal-derived tumors
but also in glial neoplasms demonstrate that this protein
can no longer be regarded as a neuron-specific marker in
neuropathology. Indeed, clinical interest has focused on
this enolase isoform over the years as a possible marker
antigen of some neuronal diseases. However, NSE im-
munohistochemistry is still a valuable method in neuro-
nal and glial tumor diagnosis and for evaluating tumor
metabolic activity.
In addition, nonneoplastic reactive astrocytes adjacent
to and within CNS tumors have been demonstrated often
to be NSE positive [42]. Moreover, reactive astrocytes in
the adult brain after injury also become NSE positive
[46]. It has been hypothesized that for cells normally
containing only ubiquitous enolase (αα enolase) the
change to a neoplastic or a reactive process necessitates
an increase in glycolytic activity, and therefore that NSE
must additionally be synthesized in the altered glial cell
to adapt to the new metabolic demands [47]. Further-
more, it has been suggested that glial cells in response to
injury may reexpress fetal characteristics of progenitors.
This response may represent a reversion to a common or-
igin of glial cells/neurons. This putative conversion to
early developmental stages may be favorable for repair
and regeneration processes through the neurotrophic ef-
fect of γ enolase (see “Discussion”).
NSE in normal glial cells
An in vivo study performed by Vinores et al. in 1984
[48] reported that fibrillary astrocytes in the white mat-
ter of rat and mouse cerebrum and cerebellum contain
NSE, while no other nonneuronal cells in the brain ex-
hibit positive immunostaining for NSE. Another investi-
gation later demonstrated that Müller cells in the retina
of Anura are immunolabeled with an antihuman NSE se-
rum [49].
We recently undertook studies to examine the pres-
ence of NSE in rat brain glial elements and to analyze
the regulation of its expression at various steps of their
differentiation/maturation stages [50, 51]. We showed
that NSE, both αγ- and γγ dimer forms, is expressed in
rat cultured oligodendrocytes. Both NSE enzymatic ac-
tivity and γ enolase protein and mRNA are detected in
these glial cells at levels similar to those of cultured neu-
rons (Figs. 3–5). We have also demonstrated that NSE is
expressed in oligodendroglial cells only at a certain stage
of differentiation both in vitro and in vivo [50]. Indeed,
NSE immunoreactivity is never seen in the cultured O-
2A precursors, which are A2B5 immunopositive. NSE is
expressed only in developing oligodendrocytes that pres-
ent GC and MBP immunostaining (Fig. 6). The immuno-
reactivity has been found to be localized mostly in the
cytoplasm, as expected for a glycolytic enzyme. The lev-
els of NSE protein and mRNA are high in oligodendro-
cytes but low in O-2A progenitor cells (Fig. 5).
We have also shown that in the presence of bFGF,
when the proliferation of O-2A precursor cells is stimu-
lated and their differentiation into oligodendrocytes re-
tarded [27], the amounts of γ enolase protein and mRNA
remain low in these oligodendroglial cell cultures. By
658
Fig. 3 Specific activities of NSE (γγ enolase) in cultured neurons,
astrocytes, and oligodendrocytes at various times of culture. d,
Days; w, weeks. Values are mean±SEM of 7–10 determinations&/fig.c:
Fig. 4 Western blot analysis of NSE in protein extracts from adult
rat brain (Br; 150 µg protein), 6-day-old oligodendrocyte cultures
(OL; 400 µg protein), 3-week-old astrocyte cultures (As; 400 µg
protein), and 4-day-old neuronal cell cultures (N; 80 µg protein).
Last lane, sample of purified γγ enolase (NSE; 0.5 µg protein).
The γ enolase subunit was detected by a rabbit anti-rat NSE poly-
clonal antibody (1:5000)&/fig.c:
contrast, treatment with bFGF has no effects on α eno-
lase protein or mRNA levels [50]. Thus bFGF prevents
specifically γ gene expression in cultured oligodendro-
glial cells, which is probably due to the fact that this fac-
tor delays the differentiation of these cells into oligoden-
drocytes. The cultures have been used only over a short
period of time (4 days), and therefore the possibility
cannot be excluded of a transient expression of NSE in
these oligodendroglial cells in vitro, as has been found
for oligodendroglial cells in vivo. Indeed, we found NSE
highly expressed in oligodendroglial cells of 8-day-old
rat brain, while it is no longer detected in adult rats
by day 60 [50]. All these data indicate that γ enolase
gene expression is associated with differentiation of the
oligodendrocytes and is repressed in fully mature adult
cells.
Biochemical analysis by two-dimensional gel electro-
phoresis has revealed a heterogeneity of γ enolase sub-
unit, consisting in the identification of six distinct pro-
tein spots, which is strictly the same in cultured oligo-
dendrocytes and in adult rat brain [50]. This indicates
that the same posttranscriptional modifications of γ sub-
unit occur in neurons in vivo and in oligodendroglial
cells in vitro.
We have also shown that NSE activity, γ enolase pro-
tein, and mRNA are present in cultured rat brain type
1 astrocytes [51]. However, NSE levels in this glial
cell type are much lower than those in cultured neurons
and oligodendrocytes (Figs. 4, 5). One of us (M.L. in
collaboration with Y. Berwald-Netter’s group) also ob-
served the presence of NSE mRNA in cultured mouse
type 1 astrocytes. On the other hand, type 2 astrocytes,
which develop in culture from the O-2A progenitor cells
in the presence of serum, have not been found to express
NSE.
659
Fig. 5 NSE mRNA expression studied by northern blot analysis
of total RNA (15 µg per sample) in cultures of neurons (N; 4-day-
old), O-2A precursors (P; 2-day-old secondary culture), oligoden-
drocytes (O; 6-day-old secondary culture), and astrocytes (A; 3-
week-old). The same blot was rehybridized with an oligonucle-
otide probe for 18S ribosomal RNA, confirming that equal
amounts of total RNAs were used in all samples (not shown)&/fig.c:
Fig. 6a, b Double-immunoflu-
orescence staining of rat brain
oligodendrocytes grown under
serum-free conditions for 4
days in secondary culture. (The
same microscopic fields are
shown in a and b). a NSE
DTAF-immunofluorescence. b
MBP TRITC-immunofluores-
cence. Note that all MBP posi-
tive oligodendrocytes are also
NSE positive. Bar=10 µm.
(From [50])&/fig.c:
Discussion
For many years GAP-43 and NSE were thought to have a
restricted neuron-specific expression. However, this view
has now been questioned, particularly by in vitro studies
and by a few in vivo reports summarized in this review,
showing the expression of these proteins in various types
of central and peripheral glial cells, such as differentiat-
ing astrocytes, oligodendrocytes, and Schwann cells. We
have performed investigations preferentially on oligo-
dendroglial cells, and a transient expression of both
GAP-43 and NSE has been found during the in vitro and
in vivo development of these cells. Figure 7 presents
schematic representation of the expression of these two
proteins in developing oligodendrocytes. GAP-43 and
NSE gene expressions are regulated differentially. In-
deed GAP-43 is downregulated while NSE is upregulat-
ed during the differentiation of oligodendrocyte precur-
sors into oligodendrocytes. Moreover, it should be point-
ed out that the time courses of expression of each of
these proteins are similar in culture and in vivo.
Some investigators have mentioned other examples of
proteins, such as microtubule-associated protein 2 (MAP
2) and neurofilament (NF) protein M, normally found in
neurons and which are in fact also expressed in glial
cells under certain circumstances. Kelly et al. [52] have
reported that myelin-forming Schwann cells of the rat
sciatic nerve in culture express NF-M at a very early
stage of their development. Other authors have found de-
tectable levels of the high molecular weight isoforms of
MAP 2 in specific populations of normal rat optic nerve
astrocytes [53]. Furthermore, reactive astrocytes after
brain injury have been described to contain the low mo-
lecular weight isoform MAP 2c [54]. More recently a
transient expression of a MAP 2c isoform in cultured oli-
godendrocyte precursors has been described [55]. All
these data demonstrating that neuron-specific proteins
can also be expressed in vitro and in vivo by glial cells in
the CNS and PNS have raised new questions about the
role of these proteins.
In both neuronal and glial cells GAP-43 is preferen-
tially expressed in vitro and in vivo during early develop-
ment, and its level dramatically declines late in develop-
ment [1, 29, 36, 38]. Moreover, in glial cells this protein
is expressed at very restricted developmental stages [29,
36, 38]. It is not known, however, whether the upregula-
tion or the downregulation of the GAP-43 gene within
neuronal and glial cell types uses the same regulatory
mechanisms.
In neurons of the CNS as well as of the PNS, GAP-43
expression increases during periods of growth cone ac-
tivity and axonal outgrowth both in vivo and in vitro [1,
3, 4]. The highest amounts of expression occur in vitro
during axonal growing and in vivo when branching neu-
rites establish stable synaptic formations. Thereafter
GAP-43 expression decreases sharply in the cell body of
cultured neurons and in both the cell bodies and the neu-
rites of most neurons in vivo. However, in several brain
areas which retain the capacity to undergo synaptic reor-
ganization some mature neurons continue to express
GAP-43 at high levels throughout adult life. High ex-
pression of GAP-43 also has been associated with re-
growth of injured neurites as regeneration procedes [6].
Thus GAP-43 may play a key role in regulating mem-
brane dynamics during neurite growth and synaptic mod-
ulation and has been considered as a plasticity protein in-
volved in neuronal shaping and repair [6].
This hypothesis has been confirmed by recent find-
ings. An in vivo study has shown that overexpression of
GAP-43 in adult neurons of transgenic mice induces
both spontanous nerve sprouting and the formation of
new synapses and promotes massive sprouting after inju-
ry [56]. Another in vivo investigation observed that mice
which lack GAP-43 display abnormal axonal pathfinding
during early development [57]. Finally, in vitro experi-
ments show that cultured neurons depleted of GAP-43
by an antisense oligonucleotide approach extend only
thin, unramified neurites with very small growth cones
adhering poorly and displaying unstable lamellar exten-
sions [58].
Other functions have been ascribed to GAP-43 in neu-
ronal cells, such as to modulate neurotransmitter secre-
tion [8].
In oligodendrocytes and Schwann cells, as in neurons,
the expression of both GAP-43 and NF-M intermediate
filaments and MAP-2 coincides with developmental
stages of major morphological changes of these cells,
consisting in the formation of extensive branches and of
elevated motile activity [1, 29, 36, 52, 55]. GAP-43 is
also expressed in cultured type 2 astrocytes [36] and in
hippocampal reactive astrocytes, when large and long
processes are formed, and this protein may contribute to
an increase in the motility of these hypertrophic astro-
cytes [39]. By contrast, type 1 astrocytes in culture lack
long processes and express little if any GAP-43. All
these observations indicate that GAP-43 distribution in
glial cells is correlated with the formation of cytoplasmic
processes and with sites of new membrane synthesis.
660
Fig. 7 Schematic representation of GAP-43 and NSE expression
during the developmental stages of rat oligodendroglial cells in
culture and in vivo. A2B5, a marker of O-2A precursors; O4, a
marker of immature oligodendrocytes binding to surface sulfa-
tides; galactocerebroside (GC) and myelin basic myelin (MBP),
markers of mature oligodendrocytes&/fig.c:
Similarly, transfection experiments have shown that the
expression of GAP-43 in nonneural cells causes changes
in their shape, associated with the formation of long and
thin processes [59, 60]. Thus GAP-43 appears to be a
protein in glial cells involved in the regulation of their
shape, and which plays a role in their motility at specific
stages of development. This is consistent with the pro-
posed function of GAP-43 in neuronal morphological
changes.
The observations showing the presence of GAP-43 in
neurons and in glial cells of both the CNS and the PNS
and the findings indicating changes in the amounts of
GAP-43 expression during development and regenera-
tion suggest that GAP-43 is probably a multifunctional
protein. It has been proposed that some of its functions
are specific to neurons while some others are associated
with various types of morphological remodeling in both
neurons and glial cells.
It has been demonstrated that levels of NSE are very
low in embryos and at birth and rapidly increase con-
comitantly with morphological and functional matura-
tion of neurons in vivo [14, 61] and in vitro [19, 20, 22,
62, 63]. Similarly, the expression of NSE seems to be
linked to the process of glial cell, astroglia and oligoden-
droglia, differentiation [50, 51]. Moreover, relatively ma-
ture cultured neurons and oligodendrocytes which no
longer proliferate express high amounts of NSE. The ex-
pression of NSE is therefore indicative not only of neu-
ronal maturation but also of the relative degree of oligo-
dendroglial cell differentiation.
In vivo a transient expression of NSE has been found
in oligodendroglial cells at a period corresponding to the
beginning of myelination when these cells must to form
a large amount of membranous structures. It has been
suggested that higher levels of NSE reflect more intense
metabolic activity of neurons [64]. Therefore the expres-
sion of the glycolytic isozyme γ enolase in oligodendro-
cytes which do not normally contain it, could reflect a
response to higher energy demands occurring in these
cells. In culture, the upregulation of NSE coincides with
the formation of the protoplasmic processes, suggesting
again that increased glycolytic activity takes place as an
adjustement to new metabolic demands. The same has
been hypothesized for glial cells that change to neoplas-
tic glial cells or to reactive astrocytes and begin to ex-
press NSE [42, 46, 47].
Concerning the hypothetical physiological signifi-
cance of γ enolase expression in all these cells it is worth
noting that, as it is the case of many other glycolytic en-
zymes [65, 66], γγ enolase has another function in addi-
tion to its enzymatic activity. Indeed, γγ enolase but not
αα enolase has neurotrophic and neuroprotective effects
on a rather broad spectrum of neurons [67, 68]. Increas-
ing levels of γ enolase are found in cerebrospinal fluid
after cerebrovascular accidents, brain tumors, and Creuz-
feld-Jakob disease (see [68]), and this can be helpful in
promoting the survival of injured neurons. Furthermore,
NSE shows neuroprotective action on neurons cultivated
in a low-oxygen atmosphere [68]. The presence of this
protein in reactive astrocytes surrounding tumors or in-
farcts may also support survival of neurons. It is not
known whether γ enolase can be secreted, but it is of in-
terest to mention that it is present in nonnegligible levels
in the intact retinal extracellular matrix [69]. Thus NSE
may play an important role during brain development as
well as during repair of injured neurons in the adult
CNS.
Finally, unraveling how specific brain proteins are ex-
pressed and regulated in neural cells is of great interest
for understanding normal and disturbed brain develop-
ment and function. The data presented in this review un-
derline the difficulty of defining specific neural markers,
since they can be expressed transiently in different cell
types during development. GAP-43 and NSE are clearly
not exclusively neuron-specific proteins, but they proba-
bly play a fundamental role in neuronal development and
regeneration as well as a critical function in normal and
pathological glial development.
&p.2:Acknowledgements The work of the authors was supported by
grants from the French Centre National de la Recherche Scientifi-
que, Institut National de la Santé et de la Recherche Médicale,
Association pour la Recherche sur le Cancer, Fondation pour la
Recherche Médicale, and Association Française contre les Myo-
pathies. The authors acknowledge the collaboration of Drs. J.
Baudier, F. Eclancher, T. Janet, G. Labourdette, P. Laeng, and M.
Ledig; of Mrs. A. Helies and M.F. Knoetgen; and of Mr. P. Bouil-
lon and C. Gaber. We thank Dr. A. Keller for critical reading of
the manuscript, Mrs. C. Orphanides for excellent secretarial work,
and Mr. J.C. Barthe for skillful assistance with the photomicro-
graphs.
References
1. Skene JHP (1989) Axonal growth-associated proteins. Ann
Rev Neurosci 12:127–156
2. Coggins PJ, Zwiers H (1991) B-50 (GAP-43): biochemistry
and functional neurochemistry of a neuron-specific phospho-
protein. J Neurochem 56:1095–1106
3. Perrone-Bizzozero NI, Finklestein SP, Benowitz LI (1986)
Synthesis of a growth-associated protein by embryonic rat ce-
rebrocortical neurons in vitro. J Neurosci 6:3721–3730
4. Van Lookeren Campagne M, Dotti CG, Verkleij AJ, Gispen
WH, Oestreicher AB (1992) Redistribution of B-50/growth-as-
sociated protein 43 during differentiation and maturation of rat
hippocampal neurons in vitro. Neuroscience 51:601–619
5. Benowitz LI, Apostolides PJ, Perrone-Bizzozero N, Finkle-
stein SP, Zwiers H (1988) Anatomical distribution of the
growth-associated protein GAP-43/B-50 in the adult rat brain.
J Neurosci 8:339–352
6. Gispen WH, Boonstra J, De Graan PNE, Jennekens FGI, Oest-
reicher AB, Schotman P, Schrama LH, Verhaagen J, Margolis
FL (1990) B-50/GAP-43 in neuronal development and repair.
Restor Neurol Neurosci 1:237–244
7. Benowitz LI, Routenberg A (1987) A membrane phosphopro-
tein associated with neural development, axonal regeneration,
phospholipid metabolism, and synaptic plasticity. Trends Neu-
rosci 12:527–531
8. Dekker LV, De Graan PNE, Versteeg DHG, Oestreicher AB,
Gispen WH (1989) Phosphorylation of B-50 (GAP43) is cor-
related with neurotransmitter release in rat hippocampal slices.
J Neurochem 52:24–30
9. Stocker KM, Baizer L, Ciment G (1992) Transient expression
of GAP-43 in nonneuronal cells of the embryonic chicken
limb. Dev Biol 149:406–414
661
10. Moos T, Christensen LR (1993) GAP43 identifies developing
muscle cells in human embryos. Neuroreport 4:1299–1302
11. Cicero TJ, Ferrendelli JA, Suntzeff V, Moore BW (1972) Re-
gional changes in CNS levels of the S-100 and 14-3-2 proteins
during development and aging of the mouse. J Neurochem
19:2119–2125
12. Bock E, Fletcher L, Rider CC, Taylor CB (1978) The nature of
the two proteins of brain specific antigen 14-3-2. J Neurochem
30:181–185
13. Fletcher L, Rider CC, Taylor CB, Adamson ED, Luke BM,
Graham CF (1978) Enolase isozymes as markers of differenti-
ation in teratocarcinoma cells and normal tissues of mouse.
Dev Biol 65:462–475
14. Schmechel DE, Brightman MW, Marangos PJ (1980) Neurons
switch from nonneuronal enolase to neuron-specific enolase
during differentiation. Brain Res 190:195–214
15. Haimoto H, Takahashi T, Kohikawa H, Nagura H, Kato K
(1985) Immunohistochemical localization of γ-enolase in nor-
mal human tissues other than nervous and neuroendocrine tis-
sues. Laboratory Invest 52:257–263
16. Schmechel DE (1985) γ-Subunit of the glycolytic enzyme eno-
lase: nonspecific of neuron specific? Lab Invest 52:239–242
17. Kato K, Suzuki F, Watanabe R, Semba R, Keino H (1984) De-
velopmental profile of three enolase isozymes in rat brain de-
termination from one-cell embryo to adult brain. Neurochem
Int 6:51–54
18. Shinohara H, Semba R, Kato K, Kashiwamata S, Tanaka O
(1986) Immunohistochemical localization of γ-enolase in early
embryos. Brain Res 382:33–38
19. Schengrund C-L, Marangos PJ (1980) Neuron-specific enolase
levels in primary cultures of neurons. J Neurosci Res
5:305–311
20. Di Liegro I, Cestelli A, Barbieri G, Giallongo A (1991) Devel-
opmental changes in neuron-specific enolase mRNA in prima-
ry cultures of rat neurons. Cell Mol Neurobiol 11:289–294
21. Lamandé N, Lucas M, Koulakoff A, Cambier H, Berwald-Net-
ter Y, Legault-Demare L, Lazar M (1989) Expression of α and
γ enolase subunits and mRNAs in primary cultures of CNS
neurons. J Neurochem 52 [Suppl]:S174B
22. Schilling K, Blanco Barco E, Rhinehart D, Pilgrim C (1989)
Expression of synaptophysin and neuron-specific enolase dur-
ing neuronal differentiation in vitro: effects of dimethyl sul-
foxide. J Neurosci Res 24:347–354
23. Kettenmann H, Ransom BR (1995) Neuroglia. Oxford Univer-
sity Press. Oxford, pp 3–116
24. Raff MC, Miller RH, Noble M (1983) A glial progenitor cell
that develops in vitro into an astrocyte or an oligodendrocyte
depending on culture medium. Nature 303:390–396
25. Weibel M, Pettmann B, Daune G, Labourdette G, Sensenbren-
ner M (1984) Chemically defined medium for rat astroglial
cells in primary culture. Int J Devl Neurosci 2:355–366
26. Weibel M, Pettmann B, Labourdette G, Miehe M, Bock E,
Sensenbrenner M (1985) Morphological and biochemical mat-
uration of rat astroglial cells grown in chemically defined me-
dium: influence of an astroglial growth factor. Int J Devl Neu-
roscience 3:617–630
27. Besnard F, Perraud F, Sensenbrenner M, Labourdette G (1989)
Effects of acidic and basic fibroblast growth factors on prolif-
eration and maturation of cultured rat oligodendrocytes. Int J
Dev Neurosci 7:401–409
28. Woolf CJ, Reynolds ML, Molander C, O’Brien C, Lindsay
RM, Benowitz LI (1990) The growth-associated protein GAP-
43 appears in dorsal root ganglion cells and in the dorsal horn
of the rat spinal cord following peripheral nerve injury. Neuro-
science 34:465–478
29. Curtis R, Stewart HJS, Hall SM, Wilkin GP, Mirsky R, Jesser
MF (1992) GAP-43 is expressed by nonmyelin-forming
Schwann cells of the peripheral nervous system. J Cell Biol
116:1455–1464
30. Scherer SS, Xu Y-T, Roling D, Wrabetz L, Feltri ML, Kam-
holz J (1994) Expression of growth-associated protein-43 kD
in Schwann cells is regulated by axon-Schwann cell interac-
tions and cAMP. J Neurosci Res 38:575–589
31. Steisslinger HW, Aloyo VJ, Vitkovic L (1987) Characteriza-
tion of two plasma membrane proteins abundant in rat brain.
Brain Res 85:375–379
32. Vitkovic L, Steisslinger HW, Aloyo VJ, Mersel M (1988) The
43-kDa neuronal growth-associated protein (GAP-43) is pres-
ent in plasma membranes of rat astrocytes. Proc Natl Acad Sci
USA 85:8296–8300
33. Vitkovic L, Mersel M (1989) Growth-associated protein 43 is
down-regulated in cultured astrocytes. Metab Brain Dis
4:47–53
34. da Cunha A, Aloyo VJ, Vitkovic L (1991) Developmental reg-
ulation of GAP-43, glutamine synthetase and β-actin mRNA
in rat cortical astrocytes. Dev Brain Res 64:212–215
35. Deloulme J-C, Janet T, Au D, Storm DR, Sensenbrenner M,
Baudier J (1990) Neuromodulin (GAP43): a neuronal protein
kinase C substrate is also present in O-2A glial cell lineage.
Characterization of neuromodulin in secondary cultures of oli-
godendrocytes and comparison with the neuronal antigen. J
Cell Biol 111:1559–1569
36. Deloulme J-C, Laeng P, Janet T, Sensenbrenner M, Baudier J
(1993) Expression of neuromodulin (GAP-43) and its regula-
tion by basic fibroblast growth factor during the differentiation
of O-2A progenitor cells. J Neurosci Res 36:147–162
37. Da Cunha A, Vitkovic L (1990) Regulation of immunoreactive
GAP-43 expression in rat cortical macroglia is cell type spe-
cific. J Cell Biol 111:209–215
38. Curtis R, Hardy R, Reynolds R, Spruce BA, Wilkin GP (1991)
Down-regulation of GAP-43 during oligodendrocyte develop-
ment and lack of expression by astrocytes in vivo: implications
for macroglial differentiation. Eur J Neurosci 3:876–886
39. Represa A, Niquet J, Charriaut-Marlangue C, Ben-Ari Y
(1993) Reactive astrocytes in the kainic acid-damaged hippo-
campus have the phenotypic features of type-2 astrocytes. J
Neurocytol 22:299–310
40. Plantinga LC, Verhaagen J, Edwards PM, Hol EM, Bär PR,
Gispen WH (1993) The expression of B-50/GAP-43 in
Schwann cells is upregulated in degenerating peripheral nerve
stumps following nerve injury. Brain Res 602:69–76
41. Woolf CJ, Reynolds ML, Chong MS, Emson P, Irwin N, Beno-
witz LI (1992) Denervation of the motor endplate results in the
rapid expression by terminal Schwann cells of the growth-as-
sociated protein GAP-43. J Neurosci 12:3999–4010
42. Vinores SA, Bonnin JM, Rubinstein LJ, Marangos PJ (1984)
Immunohistochemical demonstration of neuron-specific eno-
lase in neoplasms of the CNS and other tissues. Arch Pathol
Lab Med 108:536–540
43. Vinores SA, Herman MM, Rubinstein LJ (1987) Localization
of neuron-specific (γγ) enolase in proliferating (supportive and
neoplastic) Schwann cells. An immunohisto- and electron-im-
munocytochemical study of ganglioneuroblastoma and
schwannomas. Histochem J 19:439–448
44. Royds JA, Ironside JW, Taylor CB, Graham DI, Timperley
WR (1986) An immunohistochemical study of glial and neuro-
nal markers in primary neoplasms of the central nervous
system. Acta Neuropathol (Berl) 70:320–326
45. Cras P, Martin JJ, Gheuens J (1988) γ-Enolase and glial fibril-
lary acidic protein in nervous system tumors. An immunohis-
tochemical study using specific monoclonal antibodies. Acta
Neuropathol (Berl) 75:377–384
46. Lin RCS, Matesic DF (1994) Immunohistochemical demon-
stration of neuron-sepcific enolase and microtubule-associated
protein 2 in reactive astrocytes after injury in the adult fore-
brain. Neuroscience 60:11–16
47. Vinores SA, Herman MM, Rubinstein LJ (1986) Electron-im-
munocytochemical localization of neuron-specific enolase in
cytoplasm and on membranes of primary and metastatic cere-
bral tumours on glial filaments of glioma cells. Histopathol
10:891–908
48. Vinores SA, Herman MM, Rubinstein LJ, Marangos PJ (1984)
Electron microscopic localization of neuron-specific enolase
in rat and mouse brain. J Histochem Cytochem 32:1295–
1302
662
49. Wilhelm M, Straznicky C, Gabriel R (1992) Neuron-specific
enolase-like immunoreactivity in the vertebrate retina: selec-
tive labelling of Müller cells in Anura. Histochemistry
98:243–252
50. Deloulme JC, Lucas M, Gaber C, Bouillon P, Keller A,
Eclancher F, Sensenbrenner M (1996) Expression of the neu-
ron-specific enolase gene by rat oligodendroglial cells during
their differentiation. J Neurochem 66:936–945
51. Deloulme JC, Helies A, Ledig M, Lucas M, Sensenbrenner M
(1997) A comparative study of the distribution of α- and γ-
enolase subunits in cultured rat neural cells and fibroblasts. Int
J Devl Neurosci 15:183–194
52. Kelly BM, Gillespie CS, Sherman DL, Brophy PJ (1992)
Schwann cells of the myelin-forming phenotype express neu-
rofilament protein NF-M. J Cell Biol 118:397–410
53. Papasozomenos SC, Binder LI (1986) Microtubule-associated
protein 2 (MAP2) is present in astrocytes of the optic nerve
but absent from astrocytes of the optic tract. J Neurosci
6:1748–1756
54. Geisert EE Jr, Johnson HG, Binder LI (1990) Expression of
microtubule-associated protein 2 by reactive astrocytes. Proc
Natl Acad Sci USA 87:3967–3971
55. Vouyiouklis DA, Brophy PJ (1995) Microtubule-associated
proteins in developing oligodendrocytes: transient expression
of a MAP2c isoform in oligodendrocyte precursors. J Neurosci
Res 42:803–817
56. Aigner L, Arber S, Kapfhammer JP, Laux T, Schneider C, Bot-
teri F, Brenner HR, Caroni P (1995) Overexpression of the
neural growth-associated protein GAP-43 induces nerve
sprouting in the adult nervous system of transgenic mice. Cell
83:269–278
57. Strittmatter S, Fankhauser C, Huang PL, Mashimo H, Fishman
MC (1995) Neuronal pathfinding is abnormal in mice lacking
the neuronal growth cone protein GAP-43. Cell 80:445–452
58. Aigner L, Caroni P (1995) Absence of persistent spreading,
branching, and adhesion in GAP-43-depleted growth cones. J
Cell Biol 128:647–660
59. Zuber MX, Goodman DW, Karns LR, Fishman MC (1989)
The neuronal growth-associated protein GAP-43 induces filo-
podia in non-neuronal cells. Science 244:1193–1195
60. Widmer F, Caroni P (1993) Phosphorylation-site mutagenesis
of the growth-associated protein GAP-43 modulates its effects
on cell spreading and morphology. J Cell Biol 120:503–512
61. Marangos PJ, Schmechel DE, Parma AM, Goodwin FK (1980)
Developmental profile of neuron-specific (NSE) and nonneu-
ronal (NNE) enolase. Brain Res 190:185–193
62. Trapp BD, Marangos PJ, Webster HDeF (1981) Immunocyto-
chemical localization and developmental profile of neuron-
specific enolase (NSE) and non-neuronal enolase (NNE) in ag-
gregating cell cultures of fetal rat brain. Brain Res
220:121–130
63. Reisert I, Jirikowski G, Pilgrim C, Oertel W, Marangos PJ
(1982) The development of immunoreactivity for neuron-spe-
cific enolase of preoptic and septal neurons in dissociated cul-
tures. Neuroscience 7:1317–1322
64. Silverman WF (1992) Neuron-specific enolase reflects meta-
bolic activity in mesencephalic neurons of the rat. Brain Res
577:276–284
65. Meyer-Siegler K, Mauro DJ, Seal G, Wurzer J, de Riel JK,
Sirover MA (1991) A human nuclear uracil DNA glycosylase
is the 37-kDa subunit of glyceraldehyde-3-phosphate deshy-
drogenase. Proc Natl Acad Sci USA 88:8460–8464
66. Kato H, Fukuda T, Parkinson C, McPhie P, Cheng S (1989)
Cytosolic thyroid hormone-binding protein is a monomer of
pyruvate kinase. Proc Natl Acad Sci USA 86:7861–7865
67. Takei N, Kondo J, Nagaike K, Ohsawa K, Kato K, Kohsaha S
(1991) Neuronal survival factor from bovine brain is identical
to neuron-specific enolase. J Neurochem 57:1178–1184
68. Hattori T, Takei N, Mizuno Y, Kato K, Kohsaka S (1995) Neu-
rotrophic and neuroprotective effects of neuron-specific eno-
lase on cultured neurons from embryonic rat brain. Neurosci
Res 21:191–198
69. Li A, Lane WS, Johnson LV, Chader GJ, Tombran-Tink J
(1995) Neuron-specific enolase: a neuronal survival factor in
the retinal extracellular matrix? J Neurosci 15:385–393
663
... Besides, NSE is also detectable in glial neoplasms and reactive glial cells while undergoing morphological changes [49]. In astrocytes, NSE is expressed but at a lower level than those in cultured neurons [48][49][50]. Furthermore, it has been suggested that glial cells in response to injury may reexpress fetal characteristics of progenitors. ...
... Furthermore, it has been suggested that glial cells in response to injury may reexpress fetal characteristics of progenitors. This response may represent a reversion to a common origin of glial cells/neurons [50]. Here, we also found that NSE is slightly present in the cultured RBA ( Figure 6A), consistent with the study in rat brain type 1 astrocytes [50]. ...
... This response may represent a reversion to a common origin of glial cells/neurons [50]. Here, we also found that NSE is slightly present in the cultured RBA ( Figure 6A), consistent with the study in rat brain type 1 astrocytes [50]. Moreover, BK can induce a time-dependently increase of NSE protein in RBA by Western blot analysis ( Figure 6A), and the same result was also observed by immunofluorescent staining of NSE ( Figure 6B). ...
Article
Full-text available
Kinins are endogenous, biologically active peptides released into the plasma and tissues via the kallikrein-kinin system in several pathophysiological events. Among kinins, bradykinin (BK) is widely distributed in the periphery and brain. Several studies on the neuro-modulatory actions of BK by the B2BK receptor (B2BKR) indicate that this neuropeptide also functions during neural fate determination. Previously, BK has been shown to induce differentiation of nerve-related stem cells into neuron cells, but the response in mature brain astrocytes is unknown. Herein, we used rat brain astrocyte (RBA) to investigate the effect of BK on cell transdifferentiation into a neuron-like cell morphology. Moreover, the signaling mechanisms were explored by zymographic, RT-PCR, Western blot, and immunofluorescence staining analyses. We first observed that BK induced RBA transdifferentiation into neuron-like cells. Subsequently, we demonstrated that BK-induced RBA transdifferentiation is mediated through B2BKR, PKC-δ, ERK1/2, and MMP-9. Finally, we found that BK downregulated the astrocytic marker glial fibrillary acidic protein (GFAP) and upregulated the neuronal marker neuron-specific enolase (NSE) via the B2BKR/PKC-δ/ERK pathway in the event. Therefore, BK may be a reprogramming factor promoting brain astrocytic transdifferentiation into a neuron-like cell, including downregulation of GFAP and upregulation of NSE and MMP-9 via the B2BKR/PKC-δ/ERK cascade. Here, we also confirmed the transdifferentiative event by observing the upregulated neuronal nuclear protein (NeuN). However, the electrophysiological properties of the cells after BK treatment should be investigated in the future to confirm their phenotype.
... As differentiation progresses from oligodendrocyte precursors to mature oligodendrocytes, γ-enolase expression significantly increases, indicating its role in this transformation. Interestingly, in fully mature adult cells, γ-enolase expression is repressed, implying that its primary function is intricately tied to differentiation and maturation [133]. However, γ-enolase expression in oligodendrocytes is lower than that in neurons [133,134]. ...
... Interestingly, in fully mature adult cells, γ-enolase expression is repressed, implying that its primary function is intricately tied to differentiation and maturation [133]. However, γ-enolase expression in oligodendrocytes is lower than that in neurons [133,134]. According to previous studies using enzymatic and immunological techniques, the γ-subunit of enolase is present in cultured rat astrocytes, meningeal fibroblasts, and oligodendrocytes. ...
Article
Full-text available
Enolase, a multifunctional protein with diverse isoforms, has generally been recognized for its primary roles in glycolysis and gluconeogenesis. The shift in isoform expression from α-enolase to neuron-specific γ-enolase extends beyond its enzymatic role. Enolase is essential for neuronal survival, differentiation, and the maturation of neurons and glial cells in the central nervous system. Neuron-specific γ-enolase is a critical biomarker for neurodegenerative pathologies and neurological conditions, not only indicating disease but also participating in nerve cell formation and neuroprotection and exhibiting neurotrophic-like properties. These properties are precisely regulated by cysteine peptidase cathepsin X and scaffold protein γ 1 -syntrophin. Our findings suggest that γ-enolase, specifically its C-terminal part, may offer neuroprotective benefits against neurotoxicity seen in Alzheimer's and Parkinson's disease. Furthermore, although the therapeutic potential of γ-enolase seems promising, the effectiveness of enolase inhibitors is under debate. This paper reviews the research on the roles of γ-enolase in the central nervous system, especially in pathophysiological events and the regulation of neurodegenerative diseases.
... Отмечали увеличение размеров трансплантированных нейронов, и к 3-му и 6-му месяцам в мозолистом теле обнаруживали длинные NSE-позитивные отростки, направленные латерально, по ходу волокон мозолистого тела. Хотя NSE служит маркером развития нейронов трансплантата, следует учесть, что в отдельных работах показана её экспрессия [22] в олигодендроглии и глиальных новообразованиях, что требует сочетания с другими маркерами. Созревание нейронов сопровождалось увеличением экспрессии SYP, белка синаптических везикул, связанного с формированием синапсов и синаптической активностью нейронов [23]. ...
Article
Full-text available
Introduction. Development of cell therapy for Parkinson's disease (PD) requires protocols based on transplantation of neurons derived from human induced pluripotent stem cells (hiPSCs) into the damaged area of the brain. Objective: to characterize neurons transplanted into a rat brain and evaluate neural transplantation efficacy using a PD animal model. Materials and methods. Neurons derived from hiPSCs (IPSRG4S line) were transplanted into the striatum of rats after intranigral injection of 6-hydroxydopamine (6-OHDA). Immunostaining was performed to identify expression of glial and neuronal markers in the transplanted cells within 224 weeks posttransplant. Results. 4 weeks posttransplant we observed increased expression of mature neuron markers, decreased expression of neural progenitor markers, and primary pro-inflammatory response of glial cells in the graft. Differentiation and maturation of neuronal cells in the graft lasted over 3 months. At 3 and 6 months we detected 2 graft zones: one mainly contained the transplanted neurons and the other human astrocytes. We detected human neurites in the corpus callosum and surrounding striatal tissue and large human tyrosine hydroxylase-expressing neurons in the graft. Conclusion. With graft's morphological characteristics identified at different periods we can better understand pathophysiology and temporal patterns of new dopaminergic neurons integration and striatal reinnervation in a rat PD model in the long-term postoperative period.
... The nervous system has a limited self-regenerating capacity due to its complex physiology, where the main functional units are neurons [1]. Peripheral nerve repair (PNR) happens due to the numerous growth factors (GFs) that Schwann cells (SCs) secrete and upregulate, such as neurotrophin-3 or neurotrophin-4, the nerve growth factor, glial cell line-derived neurotrophic factor, ciliary neurotrophic factor, and brainderived neurotrophic factor [2,3]. Peripheral nerve injuries (PNIs) can happen as a consequence to a trauma and often is related to limb injuries. ...
Article
Full-text available
Peripheral nerve injuries represent a major clinical challenge, if nerve ends retract, there is no spontaneous regeneration, and grafts are required to proximate the nerve ends and give continuity to the nerve. The nerve guidance conduits (NGCs) presented in this work are silk fibroin (SF)-based, which is biocompatible and very versatile. The formation of conduits is obtained by forming a covalently cross-linked hydrogel in two concentric moulds, and the inner longitudinally aligned pattern of the SF NGCs is obtained through the use of a patterned inner mould. SF NGCs with two wall thicknesses of ~ 200 to ~ 400 μm are synthesized. Their physicochemical and mechanical characteristics have shown improved properties when the wall thickness is thicker such as resistance to kinking, which is of special importance as conduits might also be used to substitute nerves in flexible body parts. The Young modulus is higher for conduits with inner pattern, and none of the conduits has shown any salt deposition in presence of simulated body fluid, meaning they do not calcify; thus, the regeneration does not get impaired when conduits have contact with body fluids. In vitro studies demonstrated the biocompatibility of the SF NGCs; proliferation is enhanced when iSCs are cultured on top of conduits with longitudinally aligned pattern. BJ fibroblasts cannot infiltrate through the SF wall, avoiding scar tissue formation on the lumen of the graft when used in vivo. These conduits have been demonstrated to be very versatile and fulfil with the requirements for their use in PNR.
... NSE is an isoenzyme of the glycolytic enzyme enolase, which is mainly expressed in the mature neurons and the late differentiation of oligodendrocytes. It's a marker of mature neurons and neuroendocrine cells [43]. Similar to S100β, it can be used as a biomarker reflecting acute brain injury for brain trauma, stroke, hypoxic-ischemic encephalopathy, sepsis-associated encephalopathy, etc. [32]. ...
Article
Full-text available
Sepsis-associated encephalopathy (SAE) is a common complication of sepsis, raise the mortality rate with an incidence of up to 71%. Pathological neuroinflammation after sepsis leads to acute brain dysfunction, survivors may remain long-term cognitive impairment. At present, the evaluation of SAE severity and prognosis mainly depends on clinical manifestations and imaging features, but lack of effectiveness and timeliness. Biomarkers of nerve injuries nowadays, have shown good application value and perspectives in the diagnosis and evaluation of SAE. This article will review the current biomarkers for accurate diagnosis and evaluation, basing on the possible pathophysiological mechanism of different stages of SAE.
... Peripheral nerve repair happens due to the numerous growth factors (GFs) that Schwann cells (SCs) secrete and up-regulate [1], such as neurotrophin-3 or neurotrophin-4, the nerve growth factor, glial cell line-derived neurotrophic factor (GDNF), ciliary neurotrophic factor, and brain-derived neurotrophic factor [4,5]. Several works are directed to the exogenous administration of these GFs [6][7][8], as they are incorporated into artificial NGCs in order to promote peripheral nerve regeneration (PNR) to overcome long nerve gaps [9,10]. ...
Article
Full-text available
(1) Background: Peripheral nerve injuries represent a major clinical challenge. If nerve ends retract, there is no spontaneous regeneration and grafts are required to proximate the nerve ends and give continuity to the nerve. (2) Methods: GDNF-loaded NPs were characterized physicochemically. For that, NPs stability at different pH’s was assessed, and GDNF release was studied through ELISA. In vitro studies are performed with Schwann cells, and the NPs are labeled with fluorescein-5(6)-isothiocyanate for uptake experiments with SH-SY5Y neural cells. (3) Results: GDNF-loaded NPs are stable in physiological conditions, releasing GDNF in a two-step profile, which is beneficial for nerve repair. Cell viability is improved after 1 day of culture, and the uptake is near 99.97% after 3 days of incubation. (4) Conclusions: The present work shows the efficiency of using CMCht/PAMAM NPs as a GDNF-release system to act on peripheral nerve regeneration.
... Staining for NSE, however, was very weak and usually similar in intensity to the PO staining. Although not a common marker of glial cells, NSE-immunoreactivity has been demonstrated for differentiating oligodendrocytes, reactive astrocytes and reactive and neoplastic SCs (Filie et al., 1996;Sensenbrenner et al., 1997). The presence of NSE, a glycolytic enzyme, in these cells is suggested to be a product of the high energy demands of these cell types. ...
Thesis
The factors underlying the failure of axon regeneration in the CNS are thought to comprise of both the lack of supportive factors as well as the presence of inhibitory ones. Transplantation work has shown that the PNS is able to provide some of the necessary criteria and with it an increased capacity to regenerate. Studies have further shown that the crucial ingredient to such peripheral grafts is the presence of Schwann cells (SCs), the major glial cell of the PNS. I have used an extrusion transplantation system, recently developed in this laboratory, to study the effects of a SC column placed into the origin of the septo-hippocampal cholinergic projection. These SC columns integrate with host glia with minimal tissue damage, form a tight and ordered column with aligned cellular processes, and are able to recruit modest numbers of axons. Immunostaining with a cholinergic axon marker suggests that these axons arise from the septal nuclei. Given the limited availability and yield of primary SCs that current preparation protocols offer, I have engineered neonatal SC lines by transfecting the SV40 large T antigen into a population of primary neonatal rat SCs. Characterisation of these cell lines, with the use of immunocytochemistry, Western blotting and RT-PCR, shows that they retain the immunophenotype of primary SCs in vitro, although in vivo studies have posed more difficult with the lack of a suitable marker. In addition, I have set up a retroviral transfection system with the use of a bicistronic vector containing the Green Fluorescent Protein. This would provide a means of unique and efficient labelling prior to transplantation, and moreover offer the potential for further transfections of an additional gene of interest within the same vector.
Thesis
p>The aim of this thesis is to investigate the control of dentate neurogenesis, especially after seizure, and to identify the principle cells involved in these processes. I have used a combination of in vivo and in vitro techniques to investigate the mechanisms and control of seizure induced denatate neurogenesis, chemoconvulsant kainate was used to induce seizures in all seizure models, maximizing the potential for comparison between experiments. Initial in vitro work established conditions in which cell death and proliferation could be accurately and reproducibly quantified in organotypic hippocampal slice cultures. Subsequent experiments established that in immature tissue kainate induced cell death, followed by increased neurogenesis. In vivo experiments in adult rats using a ‘clonal; BrdU labelling technique, where a cohort of cells labelled prior to seizure induction were followed in recovery, found that the prelabelled cohort contributed less to seizure induced cell proliferation than the cohort of cells that were not dividing prior to seizures implying the recruitment of an additional dividing cell population by seizures. Astrocytes with radial glial like morphology are putative stem cells for dentate neurogenesis. To test our recruitment hypothesis, I used transgenic mice expressing enhanced green fluorescent protein under the hGFAP promoter to readily identify a subset of the radial glial like cells, and found their proliferation was selectively increased (10 fold vs. 2.5 fold overall increase in proliferation) in response to seizures. The data obtained suggest that seizures result in either death or inactivation of a progenitor cell population with a consequent recruitment of either quiescent or slowly dividing stem cells, which divide to replenish the progenitor cell population and restore neurogenesis. This work also identifies GFAP expressing cells with a radial glial morphology as a quiescent stem cell population selectively recruited to divide by brain injury due to seizures, and is the first report of recruitment of an identified stem cell population in the dentate after brain injury.</p
Article
Peripheral nerve regeneration is associated with pain in several preclinical models of neuropathic pain. Some neuropathic pain conditions and preclinical neuropathic pain behaviors are improved by sympathetic blockade. In this study we examined the effect of a localized "microsympathectomy," i.e., cutting the gray rami containing sympathetic postganglionic axons where they enter the L4 and L5 spinal nerves, which is more analogous to clinically used sympathetic blockade compared to chemical or surgical sympathectomy. We also examined manipulations of CCL2 (monocyte chemoattractant protein 1; MCP-1), a key player in both regeneration and pain. We used rat tibial nerve crush as a neuropathic pain model in which peripheral nerve regeneration can occur successfully. CCL2 in the sensory ganglia was increased by tibial nerve crush and reduced by microsympathectomy. Microsympathectomy and localized siRNA-mediated knockdown of CCL2 in the lumbar DRG had very similar effects: partial improvement of mechanical hypersensitivity and guarding behavior; reduction of regeneration markers growth-associated protein 43 (GAP43) and activating transcription factor 3 (ATF3); and reduction of macrophage density in the sensory ganglia and regenerating nerve. Microsympathectomy reduced functional regeneration as measured by myelinated action potential propagation through the injury site and denervation-induced atrophy of the tibial-innervated gastrocnemius muscle at day 10. Microsympathectomy plus CCL2 knockdown had behavioral effects similar to microsympathectomy alone. The results show that local sympathetic effects on neuropathic pain may be mediated in large part by the effects on expression of CCL2, which in turn regulates the regeneration process.
Article
Full-text available
The growth-associated protein GAP-43 is a major protein kinase C substrate of growth cones and developing nerve terminals. In the growth cone, it accumulates near the plasma membrane, where it associates with the cortical cytoskeleton and membranes. The role of GAP-43 in neurite outgrowth is not yet clear, but recent findings suggest that it may be a crucial competence factor in this process. To define the role of GAP-43 in growth cone activity , we have analyzed neurite outgrowth and growth cone activity in primary sensory neurons depleted of GAP-43 by a specific antisense oligonucleotide procedure. Under optimal culture conditions, but in the absence of GAP-43, growth cones adhered poorly, displayed highly dynamic but unstable lamellar extensions, and were strikingly devoid of local f-actin concentrations. Upon stimulation, they failed to produce NGF-induced spreading or insulin-like growth factor-l-induced branching, whereas growth factor-induced phosphotyrosine immunoreactivity and acceleration of neurite elongation were not impaired. Unlike their GAP-43-expressing counterparts, they readily retracted when exposed to inhibitory central nervous system myelin-derived liposomes. Frequency and extent of induced retraction were attenuated by NGE Our results indicate that GAP-43 can promote f-actin accumulation, evoked morphogenic activity, and resistance to retraction of the growth cone, suggesting that it may promote regulated neurite outgrowth during development and regeneration.
Article
Full-text available
Recently it has been demonstrated that the growth-associated protein GAP-43 is not confined to neurons but is also expressed by certain central nervous system glial cells in tissue culture and in vivo. This study has extended these observations to the major class of glial cells in the peripheral nervous system, Schwann cells. Using immunohistochemical techniques, we show that GAP-43 immunoreactivity is present in Schwann cell precursors and in mature non-myelin-forming Schwann cells both in vitro and in vivo. This immunoreactivity is shown by Western blotting to be a membrane-associated protein that comigrates with purified central nervous system GAP-43. Furthermore, metabolic labeling experiments demonstrate definitively that Schwann cells in culture can synthesize GAP-43. Mature myelin-forming Schwann cells do not express GAP-43 but when Schwann cells are removed from axonal contact in vivo by nerve transection GAP-43 expression is upregulated in nearly all Schwann cells of the distal stump by 4 wk after denervation. In contrast, in cultured Schwann cells GAP-43 is not rapidly upregulated in cells that have been making myelin in vivo. Thus the regulation of GAP-43 appears to be complex and different from that of other proteins associated with nonmyelin-forming Schwann cells such as N-CAM, glial fibrillary acidic protein, A5E3, and nerve growth factor receptor, which are rapidly upregulated in myelin-forming cells after loss of axonal contact. These observations suggest that GAP-43 may play a more general role in the nervous system than previously supposed.
Article
Immature Schwann cells of the rat sciatic nerve can differentiate into myelin-forming or non-myelin-forming cells. The factors that influence this divergent development are unknown but certain markers such as galactocerebroside distinguish the two cell populations at an early stage of Schwann cell differentiation. Because myelination requires extensive changes in cell morphology, we have investigated the composition and structure of the Schwann cell cytoskeleton at a time when these cells become committed to myelination. Here we show that Schwann cells express a cytoskeletal protein of M(r) 145 before diverging into the myelin-forming path, i.e., before they acquire cell-surface galactocerobroside. The p145 protein has the characteristics of an intermediate filament (IF) protein and immunoelectron microscopy shows that it colocalizes with vimentin, which suggests that these two proteins can coassemble into IFs. Elevated intracellular cAMP levels, which can mimic some of the early effects of axons on Schwann cell differentiation, induced p145 synthesis, therefore, we conclude that myelin-forming Schwann cells express this protein at a very early stage in their development. Immunological comparisons with other IF proteins revealed a close similarity between p145 and the neurofilament protein NF-M; the identification of p145 as NF-M was confirmed by isolating and sequencing a full-length clone from a Schwann cell cDNA library. These data demonstrate that Schwann cells remodel their IFs by expressing NF-M before acquiring the myelin-forming phenotype and that IF proteins of the neurofilament-type are not restricted to neurons in the vertebrate nervous system.
Article
: The biochemistry and functional neurochemistry of the synaptosomal plasma membrane phosphoprotein B-50 (GAP-43) are reviewed. The protein is putatively involved in seemingly diverse functions within the nervous system, including neuronal development and regeneration, synaptic plasticity, and formation of memory and other higher cognitive behaviors. There is a considerable amount of information concerning the spatial and temporal localization of B-50 (GAP-43) in adult, fetal, and regenerating nervous tissue but far less is known about the physical chemistry and biochemistry of the protein. Still less information is available about posttranslational modifications of B-50 (GAP-43) that may be the basis of neurochemical mechanisms that could subsequently permit a variety of physiological functions. Hence, consideration is given to several plausible roles for B-50 (GAP-43) in vivo, which are discussed in the context of the cellular localization of the protein, significant posttranslational enzymes, and regulatory proteins, including protein kinases, phosphoinositides, calmodulin, and proteases.
Article
The three isoenzymes of rat brain enolase (2-phospho d-glycerate hydrolase EC 4.2.1.11.) χχ, χγ and γγ were separated by ion-exchange chromatography and were tested for reaction with an antiserum against brain specific antigen 14-3-2. This monospecific antiserum affects the enolase activity of only the χγ and γγ isoenzymes. Immunoelectrophoretic experiments show that the two proteins which react as 14-3-2 both contain γ enolase subunits, and one of these also contains χ enolase subunits. It is concluded that the 14-3-2 antigen and the γ enolase subunit are identical, and that the two proteins which react immunologically as 14-3-2 are the χγ and γγ enolase isoenzymes.
Article
We have developed a serum-free defined medium that supports the growth in primary culture of rat astroglial cells. Cells dissociated from cerebral hemispheres of newborn rats were maintained for 4 days in a basal medium (Waymouth's medium) containing 10% fetal calf serum, which was substituted by a serum-free medium. The basal medium was supplemented with insulin (5 μg/ml) and fatty acid free bovine serum albumin (0.5 mg/ml). Under these conditions the cells proliferate as estimated by cell counts and DNA content; however, growth was less than in Waymouth's medium supplemented with 10% fetal calf serum. In contrast, a very similar morphology was observed between cultures grown in the serum-free or serum-containing media. The serum-free medium allows some maturation of the astroglial cells as shown by the presence of glial fibrillary acidic (GFA) protein, S-100 protein and glutamine synthetase (GS) activity. The astroglial cells can survive and grow in this chemically defined medium for up to 5 weeks. The ability to culture astroglial cells in such a minimal defined medium should facilitate investigations concerning the effects of growth factors on their proliferation and maturation.
Article
Astroglial cells from cerebral hemispheres of newborn rats were cultured for 5 days in Waymouth's MD 705/1 medium containing 10% fetal calf serum. Thereafter, cells were grown in a chemically defined medium consisting of basal Waymouth's medium supplemented with insulin (5 μg/ml) and fatty acid free bovine serum albumin (0.5 mg/ml). The cells underwent morphological and biochemical development over a period of 28 days. The changes in the amount of glial fibrillary acidic protein indicated a development of gliofilaments. The level of S100 protein increased during the entire culture period, while glutamine synthetase activity remained low and relatively constant. The addition of an astroglial growth factor, partially purified from bovine brain soluble extract, stimulated the morphological maturation of the astroglial cells. The cells extended cytoplasmic processes and resembled mature astrocytes. At the ultrastructural level an increase in free ribosomes was observed and the intermediate filaments became organized into large bundles. The amount of glial fibrillary acidic protein was not significantly increased, but the level of S100 protein and the glutamine synthetase activity were greatly enhanced. Our results indicate that astroglial cells undergo limited maturation in the chemically defined medium and that this process is positively affected by the astroglial growth factor.
Article
It has now been shown that a membrane phosphoprotein initially studied independently by several laboratories and given different designations, is in fact the same protein. Thus, the growth-associated proteins, GAP-43 and GAP-48, a phosphoprotein of the growth cone, pp46, a C-kinase substrate related to phosphatidylinositol turnover in synaptic membranes, B-50, and a C-kinase substrate associated with hippocampal long-term potentiation, F1, are identical. This protein may thus play a general role in the formation of synaptic relationships during development or regeneration, and a continuing role in the functional modulation of certain synapses throughout life.