Download - Columbia University Medical Center

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Types of artificial neural networks wikipedia , lookup

Convolutional neural network wikipedia , lookup

Biological neuron model wikipedia , lookup

Biochemistry of Alzheimer's disease wikipedia , lookup

Electrophysiology wikipedia , lookup

Artificial general intelligence wikipedia , lookup

Bird vocalization wikipedia , lookup

Environmental enrichment wikipedia , lookup

Activity-dependent plasticity wikipedia , lookup

Rheobase wikipedia , lookup

Endocannabinoid system wikipedia , lookup

Neuromuscular junction wikipedia , lookup

Single-unit recording wikipedia , lookup

Neurogenomics wikipedia , lookup

Molecular neuroscience wikipedia , lookup

Metastability in the brain wikipedia , lookup

Stimulus (physiology) wikipedia , lookup

Axon wikipedia , lookup

Multielectrode array wikipedia , lookup

Axon guidance wikipedia , lookup

Synaptogenesis wikipedia , lookup

Neural oscillation wikipedia , lookup

Embodied language processing wikipedia , lookup

Clinical neurochemistry wikipedia , lookup

Neural coding wikipedia , lookup

Mirror neuron wikipedia , lookup

Caridoid escape reaction wikipedia , lookup

Development of the nervous system wikipedia , lookup

Neuropsychopharmacology wikipedia , lookup

Nervous system network models wikipedia , lookup

Neuroanatomy wikipedia , lookup

Circumventricular organs wikipedia , lookup

Central pattern generator wikipedia , lookup

Synaptic gating wikipedia , lookup

Optogenetics wikipedia , lookup

Feature detection (nervous system) wikipedia , lookup

Pre-Bötzinger complex wikipedia , lookup

Premovement neuronal activity wikipedia , lookup

Channelrhodopsin wikipedia , lookup

Transcript
Cell, Vol. 109, 205–216, April 19, 2002, Copyright 2002 by Cell Press
Regulation of Motor Neuron Pool Sorting
by Differential Expression of Type II Cadherins
Stephen R. Price,1 Natalia V. De Marco Garcia,1
Barbara Ranscht,2 and Thomas M. Jessell1,3
1
Howard Hughes Medical Institute
Department of Biochemistry
and Molecular Biophysics
Columbia University
701 West 168th Street
New York, New York 10032
2
Neurobiology Program
The Burnham Institute
10901 N. Torrey Pines Road
La Jolla, California 92037
Summary
During spinal cord development, motor neurons with
common targets and afferent inputs cluster into discrete nuclei, termed motor pools. Motor pools can be
delineated by transcription factor expression, but cell
surface proteins that distinguish motor pools in a systematic manner have not been identified. We show
that the developmentally regulated expression of type
II cadherins defines specific motor pools. Expression
of one type II cadherin, MN-cadherin, regulates the
segregation of motor pools that are normally distinguished by expression of this protein. Type II cadherins
are also expressed by proprioceptive sensory neurons, raising the possibility that cadherins regulate
additional steps in the development of sensory-motor
circuits.
Introduction
Many hundreds of neuronal cell types are generated
during the development of the vertebrate central nervous system (CNS)—a diversity essential to the formation of selective neuronal circuits. Soon after their generation, neurons of a particular class segregate from those
of other classes—a process that is linked to their later
patterns of connectivity. The organization of neurons in
the developing CNS proceeds along two major schemes
(Ramon y Cajal, 1894). In some regions of the CNS, such
as the cerebral cortex and dorsal spinal cord, neurons
are assembled into stratified layers, or laminae (Rexed,
1952; Rakic, 1972). However, the prevalent strategy of
neuronal organization is based on nuclear subdivisions
(Ramon y Cajal, 1911): neurons with similar functions
are clustered into discrete nuclei. Advances have been
made in defining the cellular and molecular basis of
neuronal lamination (Ross and Walsh, 2001), but little
is known of the developmental mechanisms used to
segregate CNS neurons into discrete nuclei.
The nuclear organization of neurons is prominent in
the ventral spinal cord. Functional subsets of motor neurons in the lateral motor column (LMC) are clustered
into small groups, termed motor pools. The existence
3
Correspondence: [email protected]
of motor pools has long been appreciated (Elliott, 1942;
Romanes, 1942), and detailed motor pool maps have
been compiled in the developing and adult spinal cord
(Romanes, 1964; Landmesser, 1978a, 1978b; Hollyday,
1980). Motor pools have been linked to three main features of motor organization. First, all neurons within a
motor pool project to a single muscle target in the limb
(Landmesser, 1978a, 1978b). Second, all receive monosynaptic input from proprioceptive sensory neurons that
supply the same muscle target (Frank et al., 1988). Third,
neurons within an individual motor pool are electrically
coupled (Brenowitz et al., 1983), a feature that is thought
to coordinate the patterns of firing of functionally related
motor neurons in response to afferent, segmental, and
supraspinal inputs (Chang and Balice-Gordon, 2000).
The segregation of LMC motor neurons into discrete
pools begins soon after their exit from the cell cycle and
appears to occur in two phases. In one phase, motor
neurons segregate by means of an “inside-out” migration, during which prospective lateral LMC neurons migrate through neurons of the medial LMC to reach their
final settling position (Hollyday and Hamburger, 1977;
Sockanathan and Jessell, 1998). Superimposed on this
program, motor neurons within both the medial and lateral subdivisions of the LMC segregate to form discrete
pools (Lin et al., 1998).
Many of the physiological features that characterize
motor pools arise during early spinal cord development
(Milner and Landmesser, 1999; Landmesser, 2001), implying molecular distinctions in motor pool identity. Different motor pools can be distinguished by their profile
of expression of ETS and LIM homeodomain transcription factors (Lin et al., 1998; Arber et al., 2000; J. Livet
et al., submitted), and the expression of two of these
ETS genes, Er81 and PEA3, is matched in functionally
interconnected proprioceptive sensory and motor neurons in the chick spinal cord (Lin et al., 1998). The mechanisms that direct the segregation of motor neurons into
specific pools, however, remain unknown. Several
classes of cell surface proteins, notably Trks, Ephrins,
Eph kinases, and Neuropilins are expressed by subsets
of spinal motor neurons (Eberhart et al., 2000; VarelaEchavarria et al., 1997; Iwamasa et al., 1999; Feng et
al., 2000), but these proteins have not been shown to
segregate with specific motor pools in any systematic
manner. Another family of vertebrate recognition molecules is the classic cadherins (Nollet et al., 2000), and
their potential involvement in defining motor pools is
suggested by two observations. First, the differential
expression of classic cadherins demarcates distinct regions of the developing brain (Suzuki et al., 1997; Arndt
et al., 1998; Yoon et al., 2000). More tellingly, a divergent
type I cadherin, T cadherin, is expressed by certain motor
pools in the developing chick spinal cord (Fredette and
Ranscht, 1994). Nevertheless, the contribution of cadherins to the functional organization of neurons within
the developing CNS has not been defined.
We have explored the role of cadherins in the organization of neuronal nuclei through an analysis of the development of motor pools in the chick spinal cord. We
Cell
206
find that the expression of classic type II cadherins defines specific motor pools and provide in vivo evidence
for their role in motor pool segregation. Type II cadherins
are also expressed by subsets of proprioceptive neurons, raising the possibility that these proteins have additional roles in the formation of sensory-motor connections.
Results
Expression of Type II Cadherins by Embryonic
Motor Neurons
To identify cadherin genes expressed by developing
spinal motor neurons, we performed a degenerate PCR
screen on cDNA isolated from HH (Hamburger and Hamilton, 1951) stage 35 chick spinal cord. We cloned 15
cadherin genes that were dispersed throughout the major cadherin subfamilies (Nollet et al., 2000; see Supplementary Figure S1 at http://www.cell.com/cgi/content/
full/109/2/205/DC1). Analysis of the expression of these
cadherins in HH stage 35 spinal cord, a stage when
motor pool segregation approximates the mature state
(Hollyday, 1980), showed that all 15 cadherins were expressed at detectable levels by spinal cord cells (Figure
1; data not shown). We focused on the type II cadherins,
since members of this subfamily were prominently expressed in subsets of motor neurons in the LMC, both
at forelimb and hindlimb levels (Figures 1E–1L; see Supplemental Figure S2; data not shown). Of the type I
cadherins analyzed, only the divergent member, T-cad,
was expressed by selected motor pools (Figure 1I; data
not shown; Fredette and Ranscht, 1994). To determine
whether type II cadherins are expressed by specific motor pools within the LMC, we examined lumbosacral
(LS) segments 1–3, levels of the spinal cord for which
detailed motor pool maps are available (Landmesser,
1978a, 1978b; Hollyday, 1980). In addition, we used the
motor pool-specific expression of Er81 and PEA3 (Lin
et al., 1998) as a molecular guide to pool identity (Figures
1A–1D and 1M).
The expression of each type II cadherin appeared to
conform to identified motor pools (Figures 1E–1L; data
not shown). The Adductor (A) pool expressed high levels
of MN-cad, T-cad, cad-6b, and cad-8 (Figures 1E and
1G–1J); the external Femorotibialis (eF) pool expressed
high levels of T-cad, cad-6b, and cad-8 (Figures 1H–1J);
the internal Femorotibialis (iF) pool selectively expressed cad-8 (Figure 1J); the Iliotrochanterici (ITR) pool
expressed cad-7 and cad-6b (Figures 1K and 1H) and
a low level of cad-8 (Figure 1J); the anterior Iliotibialis
(ITB) pool expressed cad-6b (Figure 1H) and a low level
of MN-cad (Figure 1G); the Hip-Retractor (HR) motor
pools expressed cad-8 and a low level of cad-12 (Figures
1J and 1L); and the Sartorius (S) motor pool selectively
expressed cad-10 (Figure 1F; data not shown). We also
performed dual color fluorescence in situ hybridization
with probes for selected cadherin and ETS genes and
found coincident expression of Er81 and MN-cad in essentially all A neurons (Supplemental Figures S2E–S2G
at http://w ww.cell.com/cgi /content/full/109/2 /205/
DC1). Similar findings were made with cad-6b and cad-8
in A and eF neurons (data not shown). Thus, all motor
neurons within a pool appear to express a given cadherin. We conclude that a combinatorial profile of classic
cadherin genes distinguishes individual motor pools
within the lumbar LMC. At rostral lumbar levels, individual motor pools are typically defined by the expression
of more than one cadherin gene, and no two motor
pools exhibit an identical profile of cadherin expression
(Figure 1N).
Developmental Regulation of Cadherin
Expression in Motor Neurons
Motor pool segregation occurs between HH stages 24
and 29, and we therefore addressed when the poolrestricted patterns of cadherin expression become evident. Two contrasting developmental profiles of cadherin expression were detected in rostral lumbar LMC
motor neurons.
For one set of cadherins, MN-cad, cad-12, and cad-8,
expression appeared to be initiated in most or all LMC
neurons at the time of their generation and was subsequently extinguished from subsets of neurons to produce the mature pool-restricted pattern (Figures 2A–2C;
data not shown). Thus, at HH stage 21, soon after the
generation of postmitotic motor neurons, MN-cad expression was evident in most or all LMC motor neurons
(Figure 2A). By HH stage 24, many LMC motor neurons
still expressed MN-cad, but laterally positioned motor
neurons had extinguished expression of the gene (Figure 2B), and by HH stage 27 only the more medially
located neurons within the LMC expressed detectable
levels of MN-cad (Figures 2C and 2J). The lateral, presumptive eF motor neurons expressed Er81 but no
longer expressed MN-cad (Figures 2C and 2I).
For a second set of cadherins, T-cad, cad-6b, and
cad-7, motor pool-specific gene expression was evident
well after the exit of motor neurons from the cell cycle
(Figures 2D–2F; data not shown). The onset of expression of T-cad, cad-6b, and cad-7 expression by LMC
pools occurred only around HH stages 26–27 and was
restricted to a subset of LMC neurons (Figures 2D–2F
and 2J; data not shown). The temporal profile of expression of these cadherins by LMC neurons appeared to
parallel the pool-specific profile of ETS gene expression
(Figures 2G–2J; data not shown; Lin et al., 1998). For
both programs, the restriction in cadherin expression to
subsets of motor neurons appeared to coincide with the
segregation of motor neurons into discrete pools (Figure 2J).
The coincidence in expression of cadherin and ETS
genes by motor pools led us to examine whether there
is a functional link between the expression of ETS and
cadherin genes. The onset of expression of ETS genes
in motor pools is dependent on the limb target (Lin et
al., 1998), and we therefore examined whether the poolspecific profile of classic cadherin expression is subject
to a similar regulation. To assess this, the expression
of Er81, MN-cad, and T-cad was examined in embryos
in which the hindlimb had been ablated unilaterally at
stage 18 (Lin et al., 1998).
Analysis of operated embryos at stage 29 revealed
that the absence of Er81 expression in motor neurons
on the side of the spinal cord lacking a limb bud (Figure
2K) was accompanied by the loss of a pool-specific
pattern of MN-cad and T-cad expression (Figures 2L
and 2M). In contrast, analysis of MN-cad expression at
stage 24, prior to the pool-specific extinction of MN-cad
Cadherins and Motor Neuron Sorting
207
Figure 1. Expression of Classic Cadherins in
HH Stage 35 Chick Spinal Cord
(A) ETS gene Er81 expression at LS2.
(B) Er81 and Isl1 expression defines three motor pools at LS2: A, Er81⫹ Isl1⫹; eF, Er81⫹
Isl1⫺; and HR, Er81⫺ Isl1⫹.
(C) PEA3 expression at LS2.
(D) Er81 and PEA3 expression A and eF, and
ITR/ITB pools.
(E) MN-cad expression at LS1.
(F) Cad-10 expression at LS1.
(G–L) Comparison of MN-cad, cad-6b T-cad,
cad-8, cad-7, and cad-12 expression at LS2.
Arrow in (K) indicates expression of cad-7 by
cells at the ventral root exit point.
(M) Expression of LIM-HD and ETS proteins
in motor pools at LS2 level.
(N) Cadherin expression in motor pools at
LS2. Cadherins indicated in brackets are expressed at low levels.
The scale bar equals 150 ␮m.
expression, revealed similar levels of MN-cad expression on the operated and control sides of the spinal
cord (Figure 2N). These findings indicate that the poolspecific profile of classic cadherin expression depends
on peripheral signals. To examine whether cadherin expression in motor neurons is regulated by ETS genes,
we ectopically expressed Er81 by electroporation in
stage 16 chick spinal cord and analyzed the resulting
pattern of MN-cad expression at stage 29. Misexpression of Er81 resulted in ectopic expression of MN-cad
within the ventral spinal cord, both at hindlimb and thoracic levels (see Supplementary Figures S3A–S3C at
http://www.cell.com/cgi/content/full/109/2/205/DC1).
Together, these findings support the idea that cadherin
expression in motor pools is regulated by ETS proteins.
MN-cad Expression Regulates Motor
Pool Segregation
To test the possibility that the differential expression of
type II cadherins is involved in the segregation of LMC
neurons into pools, we focused on the eF and A motor
pools because at rostral lumbar levels these are the
only two motor pools that can be distinguished by the
expression of a single cadherin: MN-cad is expressed
by A but not by eF neurons (Figures 1E, 1G, and 1N). If
the differential expression of MN-cad contributes to the
segregation of eF and A motor neurons, we reasoned
that the expression of MN-cad in eF neurons and/or the
elimination of MN-cad activity from A neurons might be
expected to perturb the normal segregation of these two
motor pools. MN-cad expression becomes restricted to
subsets of LMC neurons after stage 24, and thus, ectopic expression of MN-cad is likely to perturb the normal pattern of MN-cad expression only from this stage
onward. Moreover, since the eF pool resides within the
lateral LMC and the A pool within the medial LMC, a
perturbation in motor pool segregation by ectopic expression of MN-cad is likely to reflect interactions between eF and A neurons during the inside-out phase of
neuronal migration.
We used in ovo electroporation to express three
cadherins—MN-cad, E-cad, and Cad-6b—in mosaic
Cell
208
Figure 2. Developmental
Regulation
Cadherin Expression in LMC Neurons
of
(A–C) MN-cad expression at LS2.
(D–F) T-cad expression at LS2.
(G–I) Er81 expression at LS2.
(J) MN-cad and T-cad expression in the eF
motor pool.
(K–M) Er81, MN-cad, and T-cad expression
at LS1 in stage 29 embryos, after unilateral
(left side) ablation of limb bud at HH stage
18. Similar findings were obtained in five embryos.
(N) Motor neuron expression of MN-cad at
HH stage 24 is not affected by unilateral (left
side) ablation of the limb bud at HH stage 18.
The scale bar equals 75 ␮m in (A) and (G),
100 ␮m in (B), (D), (H), and (N), and 150 ␮m
in (C), (F), (I), and (K–M)
fashion in motor neurons located at rostral lumbar levels
of the spinal cord, and we examined the impact on
motor pool segregation (Figure 3A). Chick embryos were
electroporated unilaterally at HH stages 15–18 and analyzed at HH stages 29–30, at which time motor pool
segregation is essentially complete (Figure 2J; Lin et al.,
1998). Expression was analyzed with antibodies directed against native cadherin proteins or through the
use of cad-GFP fusions (Figures 3B–3E). Electroporation
resulted in high levels of expression of MN-cad, E-cad,
and Cad-6b on the cell bodies, dendrites, and axons
of motor neurons (Figures 3B–3E; data not shown). In
functional experiments, we expressed cadherin cDNAs
in a bicistronic vector, from which nuclear-targeted LacZ
was also expressed, permitting the identification of neurons that expressed cadherins. Our analysis was restricted to embryos in which over 30% of LMC neurons
expressed cadherins (mean incidence of ectopic MN-cad
expression in LMC neurons was 46%; assessed by LacZ
expression; n ⫽ 18 embryos).
The consequence of cadherin expression on eF and
A motor pool segregation was assessed by analysis of
Isl1 and Er81 expression, and in all instances the identity
of individual neurons expressing ectopic cadherins was
determined by LacZ expression, using triple-label immunocytochemistry (Figures 4A–4C). We first examined
whether ectopic expression of MN-cad altered the specification of motor pool identity, as assessed by the number of Er81⫹, Isl1⫹ A and Er81⫹, Isl1⫺ eF neurons present
at rostral lumbar levels of HH stage 30 spinal cord. Ectopic expression of MN-cad did not change the total
number of A or eF motor neurons (control side, 292 A
neurons and 307 eF neurons at LS1/LS2 levels; electroporated side, 314 A neurons and 304 eF neurons at LS1/
LS2 levels; mean values from five embryos ␹2 : p ⫽ 0.5,
1 df). Thus, the profile of ETS protein expression in these
two motor pools appears not to be affected by MN-cad
expression. This finding permitted us to examine the
influence of MN-cad misexpression on the segregation
of A and eF motor neurons through an analysis of Er81
and Isl1 expression.
Ectopic expression of MN-cad did not alter the position or integrity of the LMC itself but did result in a
marked increase in the extent of intermixing of eF and
A neurons within the LMC (Figures 4A and 4B). After
MN-cad misexpression, A neurons were found in ectopic lateral positions, and conversely eF neurons were
found in ectopic medial positions (Figure 4B). To quantify this effect, we devised an index of neuronal mixing.
Individual motor neurons were assigned to specific motor pools on the basis of Er81 and Isl1 expression, and
the pool identity of each of the five motor neurons that
were located adjacent to the motor neuron under study
was scored. This index permits an analysis of the spatial
relationship between two motor neuron subtypes that
is independent of the medio-lateral position of these
neurons, and thus provides a quantitative measure of
the segregation of the two neuronal populations.
In control spinal cord analyzed at stage 29, eF neurons
were surrounded almost exclusively by other eF neurons
(Figure 4A), resulting in a neuronal mixing index that is
dominated by zero scores (Figure 4E). After MN-cad
expression, eF neurons were sometimes surrounded by
two to three A neurons (Figures 4B, 4C, and 4E), a situation rarely, if ever, encountered on the control side of
the spinal cord (Figure 4E), and the mixing of eF and A
neurons was highly significant (␹2: p ⬍ 0.001; Figure 4E;
data not shown). In addition, the degree of intermixing
of A and eF neurons was strictly dependent on the status
of MN-cad expression by eF neurons. All eF neurons
that intermixed with A neurons expressed MN-cad,
whereas eF neurons that lacked MN-cad expression
showed no increase in intermixing with A neurons over
controls (␹2: p ⬎ 0.05; Figure 4F). Thus, expression of
MN-cad in eF neurons promotes their intermixing with
A neurons in a cell-autonomous manner.
Since electroporation directed mosaic MN-cad expression in all LMC motor pools, two populations of A
neurons were generated in these experiments: nonelectroporated A neurons that exhibited wild-type levels of
MN-cad and electroporated A neurons that presumably
had elevated levels of MN-cad expression (Figure 4G).
Cadherins and Motor Neuron Sorting
209
Figure 3. Mosaic Expression of Cadherin Transgenes in LMC
Neurons
(A) Strategy for expression of cadherins by in ovo electroporation.
(B–E) Ectopic expression of an MN-cad-GFP fusion (B and C), E-cad
(D), and cad-6b (E) results in labeling of axons, somata, and dendrites of spinal cord neurons. The level of ectopic cad-6b expression
(E) is ⵑ5- to 10-fold that of endogenous cad-6b expression in motor
neurons.
Scale bar equals 150 ␮m.
We examined whether the intermixing of A and eF neurons varied according to the status of MN-cad expression in A neurons. As predicted, this analysis revealed
that A neurons were intermixed with eF neurons, but
there was no significant difference in the extent to which
electroporated and nonelectroporated A neurons mixed
with eF neurons (␹2: p ⬎ 0.05; comparison of data in
Figures 4H and 4I). We also observed that after MN-cad
misexpression, Isl1⫹ Er81⫺ HR neurons (green neurons
in Figure 4B) were interspersed within A neurons, a finding that we return to in more detail below.
Expression of a Truncated MN-cad Isoform in A
Neurons Promotes Intermixing with eF Neurons
Since ectopic MN-cad expression in eF neurons promotes their intermixing with A neurons, we considered
whether a reduction in MN-cad function in A neurons
might similarly promote their intermixing. Truncated
forms of classical cadherins that lack carboxy-terminal
regions of their cytoplasmic domain have been shown
to function in vivo in a dominant-negative manner, selectively blocking the function of their cognate cadherins
(Levine et al., 1994; Ozawa and Kemler, 1998). A truncated version of MN-cad, termed ⌬MN-cad, which lacks
114 residues at its carboxyl terminus but retains the
membrane-proximal cytoplasmic domain, was expressed in the developing spinal cord. Since MN-cad
appears initially to be expressed by all postmitotic motor
neurons (Figures 2A and 2J), the expression of ⌬MN-cad
could, in principle, perturb the normal profile of MN-cad
expression in A neurons as soon as they have left the
cell cycle.
The expression of ⌬MN-cad in A neurons promoted
their mixing with eF neurons (Figures 5A–5E). In contrast,
A neurons that lacked ⌬MN-cad expression did not mix
with eF neurons (Figure 5F). The degree of intermixing
of ⌬MN-cad-electroporated A neurons with eF neurons
was comparable to that obtained by electroporation of
MN-cad within eF neurons (compare Figures 4E and
5E). As predicted, analysis of the position of eF neurons
after ⌬MN-cad expression revealed a significant mixing
with A neurons, but the intermixed eF neurons were
comprised of those that expressed and lacked ⌬MN-cad
at equal incidence (data not shown). Together, these
findings provide evidence that the mixing of neurons
in the A and eF motor pools can be elicited either by
expression of MN-cad in eF neurons or by expression
of ⌬MN-cad in A neurons.
As a control for the specificity of action of ⌬MN-cad,
we expressed a version of MN-cad that lacks the entire
extracellular domain but retains the transmembrane and
cytoplasmic domains (⌬EC-MN-cad). With other classic
cadherins, deletion of the extracellular domain has been
shown to generate proteins that inhibit the function of
multiple cadherins (Kintner, 1992; Fujimori and Takeichi,
1993). Expression of ⌬EC-MN-cad did not result in the
intermixing of neurons in the A and eF pools (see Supplemental Figures S4C and S4D at http://www.cell.com/
cgi/content/full/109/2/205/DC1) but did result in an apparent disruption of the segregation of motor columns.
Motor neurons with a median motor column (MMC) identity were found interspersed with LMC neurons (Supplemental Figure S4E). A similar intermixing of MMC and
LMC neurons was not observed after expression of
MN-cad or ⌬MN-cad (data not shown). These findings
provide evidence for the specificity of ⌬MN-cad function
and also raise the possibility that certain classic cadherins function in the segregation of motor neurons into
columns as well as pools.
Selectivity of Motor Pool Mixing
after MN-cad Expression
As a further test of the specificity of MN-cad in eliciting
A and eF pool mixing, we examined the actions of two
other classic cadherins, E-cad and cad-6b. Between
HH stages 24 and 30, E-cad is not expressed by LMC
neurons (data not shown), and cad-6b comes to be
expressed in both A and eF neurons (Figures 1H and 1N).
Thus, ectopic expression of either of these cadherins is
Cell
210
Figure 4. Misexpression of MN-cad in eF Neurons Promotes Mixing with A Neurons
(A) eF and A neurons are segregated on the control side of the spinal cord.
(B) After misexpression of MN-cad, eF neurons are found in ectopic medial positions, and A neurons are found in ectopic lateral positions.
(C) Neuronal MN-cad expression, assessed by LacZ expression.
(D) Diagram indicating that expression of MN-cad generates a mosaic of electroporated and wild-type eF neurons.
(E) eF neurons that express MN-cad mix with A neurons (n ⬎ 450 neurons in each class from nine embryos; ␹2 , p ⬍ 0.001). The histogram
plots the incidence with which eF neurons are surrounded by A neurons, indicated as neuronal mixing index.
(F) eF neurons that do not express MN-cad remain laterally positioned and segregated from A neurons (n ⬎ 500 neurons in each class from
nine embryos; ␹2 , p ⬎ 0.05). Data in (E) and (F) indicate that eF neurons that express MN-cad are segregated from nonelectroporated eF
neurons.
(G) Expression of MN-cad generates a mosaic of electroporated and wild-type A neurons.
(H and I) Both electroporated and wild-type A neurons mix with electroporated eF neurons (nine embryos, comparison of [H] and [I]; ␹2 , p ⬎
0.05).
Scale bar equals 60 ␮m.
not predicted to equate the profile of cadherin expression of eF and A neurons (Figure 1N). Misexpression of
E-cad did not promote intermixing of A and eF neurons
(␹2: p ⬎ 0.05; Figures 6A–6C). Similarly, ectopic expression of cad-6b did not influence the segregation of A
and eF motor pools (␹2: p ⬎ 0.05; Figures 6D–6F). Thus,
the regulation of A and eF motor pool mixing by MN-cad
is not mimicked by two other classic cadherins, despite
their high-level expression by motor neurons (Figures
3D and 3E).
Specificity of Motor Pool Desegregation after
MN-cad and ⌬MN-cad Misexpression
We also examined the influence of MN-cad and
⌬MN-cad on the segregation of eF or A neurons from
neurons in other motor pools present at the LS1-LS3
level—pools that would not be predicted to acquire
equivalent cadherin profiles by manipulation of expression of MN-cad alone. We first assayed the effect of
MN-cad misexpression on the segregation of eF neurons from neurons in the ITR/ITB motor pools, two laterally positioned pools adjacent to the eF pool that express PEA3 (Figures 7A and 7B; Lin et al. 1998). Ectopic
expression of MN-cad or ⌬MN-cad in eF and ITR/ITB
neurons did not perturb the segregation of these two
motor pools (Figures 7C–7E; data not shown). Thus,
there is selectivity in the desegregation of LMC motor
pools elicited by MN-cad misexpression.
As noted above, we also examined the segregation
of A neurons from Er81⫺, Isl1⫹ neurons in the HR pool,
a motor pool complex that normally forms ventral to A
neurons and expresses cad-8 and cad-12 (Figure 1N).
After MN-cad expression, a significant intermixing of A
and HR neurons was detected (Figures 4B and 7H). The
intermixing of HR and A neurons, however, occurred
independent of the status of MN-cad expression by HR
neurons. Thus, nonelectroporated as well as electroporated HR neurons showed increased mixing with A neurons (Figures 7I and 7J). In contrast, expression of
Cadherins and Motor Neuron Sorting
211
Figure 5. Expression of ⌬MN-cad Impairs the Segregation of A and eF Neurons
(A) Segregation of A and eF neurons on the control side of the spinal cord.
(B) After misexpression of ⌬MN-cad, eF neurons are found in ectopic medial positions, and A neurons are found in ectopic lateral positions.
(C) LacZ expression indicates ⌬MN-cad electroporated neurons.
(D) Analysis of A-eF neuronal mixing after ⌬MN-cad misexpression.
(E) A neurons that express ⌬MN-cad are impaired in their ability to segregate from eF neurons (n ⬎ 200 neurons in each class from five
embryos; ␹2 , p ⬍ 0.001).
(F) A neurons that do not express ⌬MN-cad segregate normally from eF neurons. (n ⬎ 200 neurons in each class from five embryos; ␹2, p ⬎
0.05).
Scale bar equals 60 ␮m.
⌬MN-cad did not result in a similar increase in the mixing
of neurons in the HR and A pools (Figure 5B; data not
shown). Furthermore, expression of cad-6b did not promote the intermixing of HR and A neurons (Figure 6E;
data not shown). Thus, the influence of MN-cad on HR
and A mixing is not simply a general response to ectopic
cadherin expression. Taken together, these findings reveal a largely selective perturbation of motor pool sorting after manipulation of MN-cad expression.
Expression of Classic Cadherins
by Proprioceptive Sensory Neurons
We observed that type II cadherins are also expressed
by proprioceptive sensory neurons in the embryonic
dorsal root ganglion (DRG). At HH stage 36, a time at
which monosynaptic sensory-motor connections have
just been established (Lee et al., 1988; Davis et al., 1989),
subsets of DRG neurons expressed cad-8, cad-12, cad6b, and MN-cad and, in addition, T-cad (Figures 8C–8H).
At this developmental stage, proprioceptive sensory
neurons are located in a lateral domain of the DRG and
can be identified by expression of Er81 and PEA3 (Figures 8A and 8B; Lin et al., 1998). We found that DRG
neurons that expressed cad-8, T-cad, cad-12, cad-6b,
and MN-cad were concentrated in the ventrolateral region of the DRG, coincident with the domain of PEA3
and Er81 expression (Figures 8A–8G; data not shown;
Lin et al., 1998). Of the total population of rostral lumbar
Figure 6. E-cad and cad-6b Do Not Promote
the Mixing of A-eF Neurons
(A) Ectopic E-cad expression in LMC
neurons.
(B) A and eF neurons segregate normally after
E-cad misexpression.
(C) Quantitation of A and eF neuronal segregation after E-cad misexpression (␹2, p ⬎
0.05; n ⬎ 150 neurons in each class from four
embryos).
(D) Ectopic cad-6b expression. Electroporated neurons are marked by LacZ expression.
(E) A and eF neurons segregate normally after
cad-6b misexpression.
(F) Quantitation of A and eF neuron segregation after cad-6b misexpression (␹2, p ⬎ 0.05;
n ⬎ 200 neurons in each class from five embryos).
Scale bar equals 60 ␮m.
Cell
212
Figure 7. Selectivity of Motor Pool Sorting
after MN-cad Expression
(A–E) The lateral ITR/ITB pool (labeled I) segregates from eF neurons after MN-cad expression.
(A) Analysis of eF-I neuron mixing after
MN-cad expression.
(B) PEA3⫹ I neurons are found lateral to eF
neurons on the control side of the spinal cord.
(C) I neurons are in a normal position after
MN-cad expression.
(D) MN-cad expression in I neurons, detected
by Lac-Z expression.
(E) Quantitation of eF-I mixing after MN-cad
expression (n ⬎ 70 neurons in each class from
three embryos; ␹2 , p ⫽ 0.9; in order to detect
a significant difference, ⬎4000 neurons from
each class would have to be counted).
(F) Analysis of HR-A neuron sorting after
MN-cad expression.
(G) Normal segregation of A and HR neurons.
(H) Mixing of A and HR neurons after MN-cad
expression.
(I and J) Both electroporated and wild-type
HR neurons mix with A neurons after MN-cad
expression. (n ⬎ 200 neurons in each class
from five embryos; ␹2, p ⬍ 0.001).
Scale bar equals 60 ␮m.
level proprioceptive neurons, defined by the sum of Er81
and PEA3 expression (see Lin et al., 1998), 22% expressed cad-8, 21% expressed cad-12, 18% expressed
cad-6b, 15% expressed T-cad, and 13% expressed
MN-cad (Figure 8).
The expression of Er81 and PEA3 is linked in proprioceptive sensory and motor neurons that supply the same
muscle (Lin et al., 1998), prompting us to examine
whether there might be a similar linkage in cadherin
expression. To test this, we injected HRP into identified
muscles in the hindlimb of HH stage 35 or 36 chick
embryos and examined cadherin mRNA or protein expression in HRP-labeled proprioceptive neurons. This
analysis focused on the profile of MN-cad and T-cad
expression in DRG neurons that innervated A, F, and S
muscles, since the motor neurons that supply these
muscles differ in their profile of expression of these two
cadherins (Figure 1N). We detected a significant, albeit
imperfect, correlation in cadherin expression by proprioceptive sensory and motor neurons that supplied the
same muscle target. A motor neurons express both
MN-cad and T-cad, and we found that 62% of A DRG
neurons expressed MN-cad and 93% expressed T-cad
(Figures 8K, 8L, and 8O). eF motor neurons express
T-cad but not MN-cad, and we found that 62% of the
total population of F DRG neurons expressed T-cad,
whereas only 29% expressed MN-cad (Figures 8M–8O).
S motor neurons express neither MN-cad nor T-cad, and
we found that only 19% of S DRG neurons expressed
MN-cad and only 29% expressed T-cad (Figures 8I, 8J,
and 8O). A ␹2 analysis of pair-wise muscle combinations
for each cadherin indicates that the sensory expression
of cadherins is significantly correlated with motor neuron expression of the same cadherin (p ⬍ 0.01; correlation coefficient R ⫽ 0.91).
Discussion
The segregation of spinal motor neurons into discrete
motor pools typifies the nuclear organization of neuronal
cell groups evident throughout the vertebrate CNS. Type
II cadherins are differentially expressed by neurons in
specific motor pools over the period of motor pool sorting. Perturbation of the expression of one type II cadherin, MN-cad, impairs the normal program of segregation of motor pools that differ selectively in the
expression of this gene. We consider the implication of
these findings to the contribution of cadherin-mediated
recognition to nuclear organization within the developing CNS.
A Role for Cadherins in Neuronal Segregation
Classic cadherins are a large class of vertebrate cell
surface proteins and have been implicated in cell adhe-
Cadherins and Motor Neuron Sorting
213
Figure 8. Expression of Classic Cadherins in Proprioceptive Sensory Neurons
(A and B) Expression of ER81 and PEA3 in DRG at LS1-LS3 in HH
stage 36 embryos.
(C–H) Cadherin expression in DRG at LS1-LS3 in HH stage 36 embryos.
(I–O) Expression of MN-cad and T-cad in proprioceptive DRG neurons, defined by retrograde transport of HRP from specific muscles.
The cadherin expression status of corresponding motor neurons is
indicated under each panel.
(I) MN-cad is largely excluded from sensory neurons that project to
the F muscle.
(J) T-cad is expressed in eF sensory neurons.
(K and L) MN-cad (K) and T-cad (L) label A proprioceptive sensory
neurons.
(M and N) MN-cad (M) and T-cad (N) expression is largely excluded
from S sensory neurons.
(O) Quantitation of proprioceptive sensory neuron expression of
MN-cad and T-cad. ␹2 analysis indicates that sensory neuron expression of MN-cad and T-cad is positively correlated with motor
neuron expression. (For each muscle, ⬎100 HRP-labeled neurons
were counted in ⬎5 embryos, p ⬍ 0.01; correlation coefficient of
regression analysis R ⫽ 0.91.)
Scale bar equals 150 ␮m in (A–H) and 200 ␮m in (I–N).
sion and recognition in many tissue types (Shapiro and
Colman, 1998; Luo et al., 2001). Cadherins exhibit regional patterns of expression in the developing brain,
notably in the cerebral cortex, cerebellum, and thalamus
(Suzuki et al., 1997). Studies on cad-6b and cad-7 have
invoked a role for these type II cadherins in the delamination of neural crest cells from the dorsal neural tube and
in the establishment of compartment boundaries in the
early forebrain (Inoue et al., 2001; Nakagawa and Takeichi, 1998). In the Drosophila and vertebrate visual systems, the type I cadherin, N-cad, has been implicated
in the targeting of retinal axons (Lee et al., 2001; Inoue
and Sanes, 1997). In addition, type I cadherins have been
localized to synaptic sites in the mature CNS (Uchida et
al., 1996).
A role for type II cadherins in the segregation of CNS
neurons into discrete nuclei is supported by two sets
of observations. First, the combinatorial profile of type
II cadherin expression delineates neurons within a functional motor pool in the developing spinal cord. Second,
for motor pools that differ by expression of a single
cadherin, changing their profile of cadherin expression
to one in which both pools are predicted to express
the same combination markedly impairs their normal
program of segregation (see Supplemental Figure S5 at
http://www.cell.com/cgi/content/full/109/2/205/DC1).
The activity of cadherins revealed by these findings
raises three specific issues. First, what is the contribution of cadherin-based recognition to the normal segregation of neurons during motor pool development?
Second, how selective are type II cadherin-mediated
interactions between developing motor neurons? Third,
what steps in motor pool formation are regulated by
differential type II cadherin expression?
The Contribution of MN-cad to eF
and A Pool Mixing
The known profile of type II cadherin expression by A
and eF motor neurons differs solely in the expression
of MN-cad. Expression of MN-cad in eF neurons, or
⌬MN-cad in A neurons, results in an intermixing of the
neurons that populate these two motor pools (see Supplemental Figure S5 at http://www.cell.com/cgi/content/
full/109/2/205/DC1). Yet the mixing of eF and A neurons
is incomplete after both of these manipulations—a finding that has several possible explanations. First, fewer
than half of the motor neurons within a pool ectopically
express transgenic cadherins, and thus some electroporated eF neurons may simply not come into proximity
with A neurons. Second, the levels of cell surface
MN-cad expression achieved after misexpression of
MN-cad or ⌬MN-cad may vary between neurons within
an individual motor pool, with the consequence that a
significant change in the level of functional MN-cad may
be achieved in only a fraction of electroporated neurons.
Third, the subtype diversity of motor neurons known to
exist within an individual pool—␣ and ␥ (Eccles and
Sherrington, 1930) and fast and slow (Rafuse et al.,
1996)—may limit the actions of ectopic MN-cad to a
subset of neurons within a pool. Finally, additional cadherin genes could distinguish eF and A neurons, with the
consequence that manipulation of MN-cad expression
does not completely equate their profile of cadherin
expression.
The Selectivity of Cadherin-Mediated Pool Mixing
The actions of MN-cad on eF and A mixing are not
mimicked by two other classic cadherins, E-cad and
cad-6b. Thus, the desegregation of eF and A motor
neurons cannot simply be accounted for by a nonspecific elevation of cadherin expression on motor neurons.
Cell
214
Moreover, manipulation of MN-cad expression does not
promote the intermixing of eF neurons with other motor
pools that exhibit additional distinctions in cadherin expression profile, notably the ITR/ITB pools (see Supplemental Figure S5 at http://www.cell.com/cgi/content/
full/109/2/205/DC1). Nevertheless, predictions based on
the known profile of cadherin expression by LMC neurons do not readily explain the intermixing of HR and A
neurons observed after MN-cad expression (see Supplemental Figure S5). The mixing of HR and A neurons
is unlikely to be a secondary consequence of A and eF
mixing, since HR and A neuron mixing does not occur
after a similar disruption of eF-A neuron segregation by
⌬MN-cad expression. The aberrant mixing of electroporated HR and A neurons could mean that MN-cad expression in HR neurons results in a cell surface cadherin
profile that, although distinct from A neurons, still promotes interactions between these two neuronal classes.
Since heterophilic interactions have been demonstrated
in vitro between classic cadherins (Shimoyama et al.,
2000; Shan et al., 2000), MN-cad expressed on HR neurons could interact with other type II cadherins expressed on A neurons, thus promoting their intermixing.
However, this scenario fails to explain the mixing of
nonelectroporated HR neurons with A neurons. An alternative, and perhaps more likely, possibility is that HR
neurons that express MN-cad acquire the ability to mix
with both A neurons and nonelectroporated HR neurons
such that the resultant dispersal of electroporated HR
neurons leads to a secondary mixing of A and nonelectroporated HR neurons.
More generally, our findings emphasize the fact that
the molecular basis of cadherin-mediated recognition
has traditionally been considered under conditions of
interaction between single cadherins (Steinberg and Takeichi, 1993). Our results reveal that developing CNS
neurons typically express more than one, and in some
cases as many as four, classic cadherins. The rules of
cadherin-based interaction under these more complex,
but perhaps more developmentally relevant, conditions
remain to be delineated.
The Cellular Basis of Cadherin Action during
Motor Pool Formation
How does the differential expression of MN-cad influence the segregation of eF and A neurons? The downregulation of MN-cad expression by eF neurons is evident from stage 24 onward. At this stage, eF neurons,
as other lateral LMC neurons, are beginning to migrate
past A neurons to reach their lateral setting position
(Tsuchida et al., 1994; Sockanathan and Jessell, 1998).
Thus, the early downregulation of MN-cad expression
by eF neurons may ensure that they do not interact
with A neurons during the inside-out phase of neuronal
migration. In this view, a reduction in MN-cad activity
in A neurons or expression of MN-cad by eF neurons
would result in an aberrant interaction between these
two sets of motor neurons, perturbing their segregation.
We note that the degree of intermixing of eF and A
neurons achieved by perturbing the cadherin profile of
A neurons at early stages (through ⌬MN-cad expression) and of eF neurons at later stages (through MN-cad
expression) appear similar. This finding implies that
the distinction in profile of cadherin expression that
emerges after stage 24 is relevant to the segregation of
neurons in these two motor pools.
The formation of motor pools depends not only on
the early segregation of the lateral and medial subdivisions of the LMC, but also on the subsequent segregation of neurons within each of these subdivisions (Lin
et al., 1998). Our studies have focused on the function
of MN-cad, a protein that appears to perturb the segregation of motor neurons during the inside-out phase of
migration, but the differential expression of other type
II cadherins clearly distinguishes each of the motor
pools that form within the medial and lateral subdivisions
of the LMC. Cadherins are therefore plausible mediators
of the secondary phase of motor pool segregation.
Correlated Expression of Type II Cadherins
by Proprioceptive Sensory and Motor Neurons
Several findings suggest that the expression of type II
cadherins and ETS genes in motor pools is linked. First,
both ETS and cadherin expression in motor pools is
dependent of limb-derived signals. Second, ectopic expression of Er81 results in the deregulation of MN-cad
expression in the chick spinal cord. Third, the analysis
of PEA3 mutant mice indicates that the expression of
type II cadherins in specific motor pools within the LMC
is regulated by this ETS gene (J. Livet et al., submitted).
Nevertheless, the functional link between ETS and cadherin gene expression in motor neurons is not simple.
The multiple type II cadherin genes expressed by motor
neurons define many more motor pools than are revealed by PEA3 and Er81 (Lin et al., 1998). Secondly,
the early and widespread expression of MN-cad, cad-8,
and cad-12 by developing LMC neurons indicates that
the known ETS genes are not involved in the initial onset
of expression of these cadherins. Physiological studies
have suggested that motor neurons in the LMC acquire
an initial pool identity prior to the segregated expression
of ETS or cadherin genes (Milner and Landmesser,
1999). Defining the molecular basis of this early phase
of motor pool specification may clarify the link between
ETS and cadherin gene expression in developing LMC
neurons.
The expression of Er81 and PEA3 is linked in interconnected proprioceptive sensory and motor neurons (Lin
et al., 1998). The present studies reveal a significant
correlation in type II cadherin expression in proprioceptive sensory and motor neurons that project to the same
target muscle. A matching in cadherin expression on
the axons of proprioceptive sensory neurons and on the
dendrites of motor neurons could provide a basis for
the selectivity with which monosynaptic connections
are formed in this neuronal circuit (Frank et al., 1988).
Motor Pool Formation and the Nuclear
Organization of CNS Neurons
The organization of neuronal subtypes in the CNS is
based in large part on the segregation of neurons into
discrete nuclei. Since the formation of motor pools captures many of the essential features of neuronal nuclear
organization, the role of cadherin function in motor pool
segregation demonstrated in this study suggests a more
widespread role for cadherins in the organization of neuronal nuclei in the developing CNS. The patterns of
Cadherins and Motor Neuron Sorting
215
cadherin expression in many areas of the developing
brain (Inoue et al., 2001; Suzuki et al., 1997) are consistent with an involvement in neuronal clustering.
The functional rationale for organizing motor neurons
into specific pools remains enigmatic, however. Motor
pools have been correlated with electrical coupling, a
feature that has been suggested to coordinate the phasic firing of functionally related motor neurons (Chang
and Balice-Gordon, 2000). Physiological studies have
also suggested distinctions in local interneuronal connections with motor pools (Ritter et al., 1999), and such
specificity in local circuitry could depend on the clustered organization of neurons into motor pools. Motor
pool clustering might also be involved in late steps in
the establishment of motor axon projections (J. Livet et
al., submitted). The role of cadherins in motor pool sorting revealed in these studies should permit a dissociation between motor neuron position and identity. In turn,
defining how the organization of motor neurons into
discrete nuclei influences the function of motor circuits
may provide insights pertinent to neuronal organization
in other regions of the CNS.
Health grant HD25938. T.M.J. is supported by a grant from the
Harold and Leila Mathers Foundation and is a Howard Hughes Medical Institute Investigator.
Experimental Procedures
Eberhart, J., Swartz, M., Koblar, S.A., Pasquale, E.B., Tanaka, H.,
and Krull, C.E. (2000). Expression of EphA4, ephrin-A2 and ephrinA5 during axon outgrowth to the hindlimb indicates potential roles
in pathfinding. Dev. Neurosci. 22, 237–250.
Chick Embryo Preparation
Chick eggs (Spafas, Truslow Farms) were incubated and staged as
in Hamburger and Hamilton (1951).
Cloning of Cadherin Genes
Details of the cloning of chicken cadherin genes are available as
Supplemental Data at http://www.cell.com/cgi/content/full/109/2/
205/DC1.
In Situ Hybridization Histochemistry
Digoxigenin (DIG)-labeled cRNA probes were used for in situ hybridization histochemistry as in Schaeren-Wiemers and Gerfin-Moser
(1993). Dual color fluorescence in situ hybridization histochemistry
was performed as described in the Supplemental Data.
Immunohistochemistry
Antibodies used in this study are included in the Supplemental Data
at http://www.cell.com/cgi/content/full/109/2/205/DC1.
Retrograde Labeling of DRG Neurons
HRP labeling of DRG neurons was performed essentially as described in Lin et al. (1998).
Generation of Cadherin Expression Constructs
Details of cloning and expression of cadherin cDNAs are provided
as Supplemental Data.
In Ovo Electroporation
Expression of cDNAs was achieved by in ovo electroporation using
an ECM830 electro-squareporator (BTX Inc.). Pulses were of 50 ms
duration, 5 times at 30V. Embryos were electroporated at HH stages
15–18 and analyzed at HH stages 28–30.
Modeling Neuronal Mixing
Details of quantitation of neural mixing are provided in the Supplemental Data at http://www.cell.com/cgi/content/full/109/2/205/DC1.
Acknowledgments
We are grateful to M. Takeichi for providing reagents. We thank S.
Arber, R. Axel, K. Baldwin, J. de Nooij, C. Henderson, J. Kaltschmidt,
P. Scheiffele, and L. Shapiro for critical comments on the manuscript
and K. MacArthur and I. Schieren for help in preparing the text and
figures. S.R.P. is indebted to the Helen Hay Whitney Foundation, of
which he is a fellow. B.R. is supported by National Institutes of
Received: December 14, 2001
Revised: February 20, 2002
References
Arber, S., Ladle, D.R., Lin, J.H., Frank, E., and Jessell, T.M. (2000).
ETS gene Er81 controls the formation of functional connections
between group Ia sensory afferents and motor neurons. Cell 101,
485–498.
Arndt, K., Nakagawa, S., Takeichi, M., and Redies, C. (1998). Cadherins define segments and parasagital cell ribbons in the developing
chick cerebellum. Mol. Cell. Neurosci. 10, 211–228.
Brenowitz, G.L., Collins, W.F.D., and Erulkar, S.D. (1983). Dye and
electrical coupling between frog motor neurons. Brain Res. 274,
371–375.
Chang, Q., and Balice-Gordon, R.J. (2000). Gap junctional communication among developing and injured motor neurons. Brain Res.
Brain Res. Rev. 32, 242–249.
Davis, B.M., Frank, E., Johnson, F.A., and Scott, S.A. (1989). Development of central projections of lumbosacral sensory neurons in
the chick. J. Comp. Neurol. 279, 556–566.
Eccles, J.C., and Sherrington, C. (1930). Numbers and contraction
values of individual motor-units examined in some muscles of the
limb. Proc. R. Soc. Lond. B Biol. Sci. 106, 326–357.
Elliott, H.C. (1942). Studies on the motor cells of the spinal cord. I.
Distribution in the normal human cord. Am. J. Anat. 70, 95–117.
Feng, G., Laskowski, M.B., Feldheim, D.A., Wang, H., Lewis, R.,
Frisen, J., Flanagan, J.G., and Sanes, J.R. (2000). Roles for ephrins
in positionally selective synaptogenesis between motor neurons and
muscle fibers. Neuron 25, 295–306.
Frank, E., Smith, C., and Mendelson, B. (1988). Strategies for selective synapse formation between muscle sensory and motor neurons
in the spinal cord. In From Message to Mind, S.S. Easter, K.F. Barald,
and B.M. Carlson, eds. (Sunderland, MA: Sinauer Associates Inc.),
pp. 180–202.
Fredette, B.J., and Ranscht, B. (1994). T-cadherin expression delineates specific regions of the developing motor axon-hindlimb projection pathway. J. Neurosci. 14, 7331–7346.
Fujimori, T., and Takeichi, M. (1993). Disruption of epithelial cell-cell
adhesion by exogenous expression of a mutated non-functional
N-cadherin. Mol. Biol. Cell 4, 37–47.
Hamburger, V., and Hamilton, H. (1951). A series of normal stages
in the development of the chick embryo. J. Morphol. 160, 535–546.
Hollyday, M. (1980). Organization of motor pools in the chick lumbar
lateral motor column. J. Comp. Neurol. 194, 143–170.
Hollyday, M., and Hamburger, V. (1977). An autoradiographic study
of the formation of the lateral motor column in the chick embryo.
Brain Res. 132, 197–208.
Inoue, A., and Sanes, J.R. (1997). Lamina-specific connectivity in
the brain: regulation by N-cadherin, neurotrophins, and glycoconjugates. Science 276, 1428–1431.
Inoue, T., Tanaka, T., Takeichi, M., Chisaka, O., Nakamura, S., and
Osumi, N. (2001). Role of cadherins in maintaining the compartment
boundary between the cortex and striatum during development.
Development 128, 561–569.
Iwamasa, H., Ohta, K., Yamada, T., Ushijima, K., Terasaki, H., and
Tanaka, H. (1999). Expression of Eph receptor tyrosine kinases and
their ligands in chick embryonic motor neurons and hindlimb muscles. Dev. Growth Differ. 41, 685–698.
Kintner, C. (1992). Regulation of embryonic cell adhesion by the
cadherin cytoplasmic domain. Cell 69, 225–236.
Cell
216
Landmesser, L. (1978a). The distribution of motoneurones supplying
chick hind limb muscles. J. Physiol. 284, 371–389.
man, D.R., and Shapiro, L. (2000). Functional cis-heterodimers of
N- and R-cadherins. J. Cell Biol. 148, 579–590.
Landmesser, L. (1978b). The development of motor projection patterns in the chick hind limb. J. Physiol. 284, 391–414.
Shapiro, L., and Colman, D.R. (1998). Structural biology of cadherins
in the nervous system. Curr. Opin. Neurobiol. 8, 593–599.
Landmesser, L.T. (2001). The acquisition of motoneuron subtype
identity and motor circuit formation. Int. J. Dev. Neurosci. 19,
175–182.
Shimoyama, Y., Tsujimoto, G., Kitajima, M., and Natori, M. (2000).
Identification of three human Type-II classic cadherins and frequent
heterophilic interactions between different subclasses of Type-II
classic cadherins. Biochem. J. 349, 159–167.
Lee, M.T., Koebbe, M.J., and O’Donovan, M.J. (1988). The development of sensorimotor synaptic connections in the lumbosacral spinal cord of the chick embryo. J. Neurosci. 8, 2530–2543.
Lee, C.H., Herman, T., Clandinin, T.R., Lee, R., and Zipursky, S.L.
(2001). N-cadherin regulates target specificity in the Drosophila visual system. Neuron 30, 437–450.
Levine, E., Lee, C.H., Kintner, C., and Gumbiner, B.M. (1994). Selective disruption of E-cadherin function in early Xenopus embryos by
a dominant negative mutant. Development 120, 901–909.
Lin, J.H., Saito, T., Anderson, D.J., Lance-Jones, C., Jessell, T.M.,
and Arber, S. (1998). Functionally related motor neuron pool and
muscle sensory afferent subtypes defined by coordinate ETS gene
expression. Cell 95, 383–407.
Luo, Y., Ferreira-Cornwell, M.C., Baldwin, H.S., Kostetski, I., Lenox,
J.M., Lieberman, M., and Radice, G.L. (2001). Rescuing the
N-cadherin knockout by cardiac-specific expression of N- or
E-cadherin. Development 128, 459–469.
Milner, L.D., and Landmesser, L.T. (1999). Cholinergic and GABAergic inputs drive patterned spontaneous motoneuron activity before
target contact. J. Neurosci. 19, 3007–3022.
Nakagawa, S., and Takeichi, M. (1998). Neural crest emigration from
the neural tube depends on regulated cadherin expression. Development 125, 2963–2971.
Nollet, F., Kools, P., and van Roy, F. (2000). Phylogenetic analysis
of the cadherin superfamily allows identification of six major subfamilies besides several solitary members. J. Mol. Biol. 299, 551–572.
Ozawa, M., and Kemler, R. (1998). The membrane proximal region
of the E-cadherin cytoplasmic domain prevents dimerization and
negatively regulates adhesion activity. J. Cell Biol. 142, 1605–1613.
Rafuse, V.F., Milner, L.D., and Landmesser, L.T. (1996). Selective
innervation of fast and slow muscle regions during early chick neuromuscular development. J. Neurosci. 16, 6864–6877.
Rakic, P. (1972). Mode of cell migration to the superficial layers of
fetal monkey neocortex. J. Comp. Neurol. 145, 61–84.
Ramon y Cajal, S. (1894). La fine structure des centres nerveux.
Proc. R. Soc. Lond 55, 444–468.
Ramon y Cajal, S. (1911). Histologie du system nerveux de l’home
et des vertebras, Volumes 1 and 2. (Paris: A Maloine). Reprinted by
Consejo Superior de Investigaciones Cientificas, Inst. Ramon y Cajal, Madrid, 1955.
Rexed, B. (1952). The cytoarchitectonic organization of the spinal
cord in the cat. J. Comp. Neurol. 96, 415–495.
Ritter, A., Wenner, P., Ho, S., Whelan, P.J., and O’Donovan, M.J.
(1999). Activity patterns and synaptic organisation of ventrally located interneurons in the embryonic chick spinal cord. J. Neurosci.
19, 3457–3471.
Romanes, G.J. (1942). The development and significance of the cell
columns in the ventral horn of the cervical and upper thoracic spinal
cord of the rabbit. J. Anat. Lond. 76, 112–130.
Romanes, G.J. (1964). The motor pools of the spinal cord. In Organization of the Spinal Cord, Volume 11, J.C. Eccles, and J.P. Schade,
eds. (Amsterdam: Elsevier Publishing), pp. 93–119.
Ross, M.E., and Walsh, C.A. (2001). Human brain malformations
and their lessons for neuronal migration. Annu. Rev. Neurosci. 24,
1041–1070.
Schaeren-Wiemers, N., and Gerfin-Moser, A. (1993). A single protocol to detect transcripts of various types and expression levels in
neural tissue and cultured cells: in situ hybridisation using digoxygenin labelled cRNA probes. Histochemistry 100, 431–440.
Shan, W.S., Tanaka, H., Phillips, G.R., Arndt, K., Yoshida, M., Col-
Sockanathan, S., and Jessell, T.M. (1998). Motor neuron-derived
retinoid signals determine the number and subtype identity of motor
neurons in the developing spinal cord. Cell 94, 503–514.
Steinberg, M.S., and Takeichi, M. (1993). Experimental specification
of cell sorting, tissue spreading and specific patterning by quantitative differences in cadherin expression. Proc. Natl. Acad. Sci. USA
91, 206–209.
Suzuki, S.C., Inoue, T., Kimura, Y., Tanaka, T., and Takeichi, M.
(1997). Neuronal circuits are subdivided by differential expression
of Type-II classic cadherins in postnatal mouse brains. Mol. Cell.
Neurosci. 9, 433–447.
Tsuchida, T., Ensini, M., Morton, S.B., Baldassare, M., Edlund, T.,
Jessell, T.M., and Pfaff, S.L. (1994). Topographic organisation of
embryonic motor neurons defined by expression of LIM homeobox
genes. Cell 79, 957–970.
Uchida, N., Honjo, Y., Johnson, K.R., Wheelock, M.J., and Takeichi,
M. (1996). The catenin/cadherin adhesion system is localized in
synaptic junctions bordering transmitter release zones. J. Cell Biol.
135, 767–779.
Varela-Echavarria, A., Tucker, A., Puschel, A.W., and Guthrie, S.
(1997). Motor axon subpopulations respond differentially to the
chemorepellents netrin-1 and semaphorin D. Neuron 18, 193–207.
Yoon, M.S., Puelles, L., and Redies, C. (2000). Formation of cadherinexpressing brain nuclei in diencephalic alar plate divisions. J. Comp.
Neurol. 427, 461–480.
Accession Numbers
The sequence of MN-cad has been submitted to GenBank with the
accession number AF459439.