ArticlePDF Available

Small GTPases Rac and Rho in the Maintenance of Dendritic Spines and Branches in Hippocampal Pyramidal Neurons

Authors:

Abstract and Figures

The shape of dendritic trees and the density of dendritic spines can undergo significant changes during the life of a neuron. We report here the function of the small GTPases Rac and Rho in the maintenance of dendritic structures. Maturing pyramidal neurons in rat hippocampal slice culture were biolistically transfected with dominant GTPase mutants. We found that expression of dominant-negative Rac1 results in a progressive elimination of dendritic spines, whereas hyperactivation of RhoA causes a drastic simplification of dendritic branch patterns that is dependent on the activity of a downstream kinase ROCK. Our results suggest that Rac and Rho play distinct functions in regulating dendritic spines and branches and are vital for the maintenance and reorganization of dendritic structures in maturing neurons.
Dominant-negative Rac1 expression results in a progressive reduction of the dendritic spine density and mild changes in the dendritic-branching pattern. A, Transfected myc-tagged wild-type Rac1 protein is distributed along the dendrites and in dendritic spines ( green , anti-myc; red , anti- mCD8; composite confocal image using 40 ϫ objective; digital zoom factor of 3). B, Representative images are shown of apical dendrites that were transfected with the marker mCD8 alone ( top ) or with mCD8 and myc-tagged Rac1N17 ( bottom ) for 1, 2, or 3 d (composite confocal images of mCD8 staining using 100 ϫ objective). C, Apical dendrites from neighboring pyramidal neurons are shown 3 d after expressing mCD8 alone ( red ) or expressing both mCD8 ( red ) and myc-tagged Rac1N17 ( green ) and therefore appearing yellow (composite confocal image using a 40 ϫ objective with a digital zoom factor of 3). Although it appears that Rac1N17 dendrites have thicker den- drites in this image, quantification of the average diameter of dendrites within every image used to measure dendritic spine density does not reveal any significant difference (paired t test, apical, p ϭ 0.19; Rac1N17 ϭ 1.50 Ϯ 0.14 ␮ m; control ϭ 1.22 Ϯ 0.51 ␮ m; n ϭ 18, 9, respectively; basal, p ϭ 0.12; Rac1N17 ϭ 0.83 Ϯ 0.1 ␮ m; control ϭ 1.05 Ϯ 0.08 ␮ m; n ϭ 9, 8, respective- ly). D, Rac1N17 expression progressively reduces the number of spines on apical and basal dendrites (for 1, 2, 3 d, apical mCD8, n ϭ 16, 19, 10; apical Rac1N17, n ϭ 19, 11, 24; basal mCD8, n ϭ 15, 14, 10; basal Rac1N17, n ϭ 18, 11, 17, respec- tively). E, Quantification of dendritic branch segments after 3 d of Rac1N17 expression is shown ( n ϭ 17, 23, 18, 21 for apical mCD8, apical Rac1N17, basal mCD8, basal Rac1N17, respec- tively). F, Sholl profiles of the basal dendrites of Rac1N17- expressing and control CA1 pyramidal neurons (3 d after transfection) illustrate the slight change in dendritic-branching pattern with Rac1N17 expression ( n ϭ 14, 15 for basal mCD8, Rac1N17, respectively). Post hoc t tests reveal that the only significant difference occurs at 50 ␮ m (paired t test, ** p Ͻ 0.01). Scale bars: A, B , 5 ␮ m; C , 10 ␮ m.
… 
Function of ROCK in the maintenance of dendritic branches. A–E, Representative images of neurons 2 d after transfection and 100 ␮ M Y-27632 treatment are shown. Neurons were transfected with mCD8 alone ( A, B ) or cotransfected with RhoAV14 ( C, D ) or ROCK ⌬ 3 ( E ). E , Inset, The myc immunostaining for ROCK ⌬ 3 is shown. Additionally, neurons in B and D were treated with 100 ␮ M Y-27632 at the time of transfection (composite confocal images using 16 ϫ objective). F, Quantification of basal dendritic branch segment numbers of mCD8- and of mCD8 plus RhoAV14-expressing neurons with or without Y-27632 treatment is shown. Y-27632 treatment alone does not affect dendritic segment number ( p ϭ 0.31; control treatment, n ϭ 23; Y-27632 treatment, n ϭ 23). Y-27632 treatment blocks RhoAV14-associated dendritic segment reduction (*** p Ͻ 0.001; RhoAV14 ϩ control, n ϭ 17; RhoAV14 expression ϩ Y-27632 treatment, n ϭ 24). G, Y-27632 application does not alter the basal dendritic spine density of neurons expressing mCD8 alone ( p ϭ 0.65; n ϭ 29, 21 for control, Y-27632 treatment, respectively). Y-27632 treatment is capable of restoring the spine density of neurons expressing RhoAV14 close to control level ( p ϭ 0.08; n ϭ 16 for Y-27632 treatment and RhoA V14 expression). H, Activated ROCK ⌬ 3 expression results in significant reduction of dendritic segments compared with that in mCD8 alone ( p Ͻ 0.001 for both apical and basal; n ϭ 19, 20, respectively, for ROCK ⌬ 3). Scale bars, 50 ␮ m.
… 
Content may be subject to copyright.
Small GTPases Rac and Rho in the Maintenance of Dendritic Spines
and Branches in Hippocampal Pyramidal Neurons
Ann Y. Nakayama,
1,2
Matthew B. Harms,
1
and Liqun Luo
1,2
1
Department of Biological Sciences and
2
Neurosciences Program, Stanford University, Stanford, California 94305-5020
The shape of dendritic trees and the density of dendritic spines
can undergo significant changes during the life of a neuron. We
report here the function of the small GTPases Rac and Rho in the
maintenance of dendritic structures. Maturing pyramidal neurons
in rat hippocampal slice culture were biolistically transfected with
dominant GTPase mutants. We found that expression of
dominant-negative Rac1 results in a progressive elimination of
dendritic spines, whereas hyperactivation of RhoA causes a
drastic simplification of dendritic branch patterns that is depen-
dent on the activity of a downstream kinase ROCK. Our results
suggest that Rac and Rho play distinct functions in regulating
dendritic spines and branches and are vital for the maintenance
and reorganization of dendritic structures in maturing neurons.
Key words: Rac; Rho; dendritic spines; biolistic transfection;
pyramidal neurons; effector domain mutants; ROCK; Y-27632;
PSD-95
Perhaps the most striking features of many mammalian neurons are
their complex dendritic trees. Dendrites are the principal site
where neurons receive, process, and integrate inputs from their
multiple presynaptic partners. These functions are primarily influ-
enced by the branching pattern of the dendritic tree (Rall, 1964). In
addition to complex dendritic arbors, some mammalian neurons,
including cerebellar Purkinje cells and cortical and hippocampal
pyramidal neurons, have dendritic specializations called spines.
These spines, which protrude from the dendritic branches, are the
primary site of excitatory synapses and may function as the basic
unit of synaptic integration (for review, see Harris and Kater, 1994;
Yuste and Tank, 1996). Both the shape of dendritic trees and the
density and shape of their spines can undergo significant changes
during the development and life of a neuron. For instance, den-
dritic branches have been shown to undergo significant remodeling
in adult superior cervical ganglion neurons in vivo (Purves and
Hadley, 1985; Purves et al., 1986). Recent studies have also shown
that dendritic filopodia and spines undergo dynamic changes in
response to synaptic activity (Engert and Bonhoeffer, 1999;
Maletic-Savatic et al., 1999; Toni et al., 1999) and neurotrophin
overexpression (Horch et al., 1999).
Several extracellular molecules have been identified as potential
regulators of dendrite and spine development (Purves et al., 1988;
Snider, 1988; Lein et al., 1995; McAllister et al., 1995, 1997; Nedivi
et al., 1998; Guo et al., 1999; Horch et al., 1999). To induce
dendritic growth, branching, or retraction, as well as the formation,
elaboration, or elimination of dendritic spines, these extracellular
factors must eventually exert their effects by signaling to the cy-
toskeleton. Actin is one of the main components of the dendritic
cytoskeleton and is highly enriched in dendritic spines (see Fischer
et al., 1998). The Rho family of small GTPases, including Rho,
Rac, and Cdc42, regulates various aspects of the actin cytoskeleton
(for review, see Hall, 1994; Van Aelst and D’Souza-Schorey, 1997),
and these GTPases are therefore good candidates for mediating
these signals. In agreement with their differential effects in fibro-
blasts (for review, see Van Aelst and D’Souza-Schorey, 1997), both
in vivo and in vitro studies have shown that these GTPases appear
to have distinct effects on different aspects of neuronal morpho-
genesis (for review, see Luo et al., 1997; Mueller, 1999).
Although many studies illustrate the roles of Rho family
GTPases in the establishment of neuronal processes, it remains
unclear whether they also function in later stages of neuronal
development or in mature neurons to regulate dendritic reorgani-
zation and dynamic changes of dendritic spines. To address these
questions, we used biolistic transfection (Arnold et al., 1994; Lo et
al., 1994) to introduce dominant mutants of Rac1 and RhoA
acutely into organotypic hippocampal slices at a stage when pyra-
midal neurons have established dendritic arbors and possess den-
dritic spines yet still express the GTPases. We found that Rac1 is
required for the maintenance of dendritic spines, whereas the
elevation of RhoA activity leads to pronounced simplification of
dendritic trees. We also identified candidate downstream effector
pathways that mediate these morphological changes. These studies
demonstrate that signaling pathways used in the early development
of neuronal processes are also used in more mature neurons for the
maintenance and reorganization of dendritic structures.
MATERIALS AND METHODS
In situ hybridizations
Sense and antisense S
35
-riboprobes were generated to the C-terminal ends
of rat Rac1 and RhoA by the following steps: 5-Phosphorylated primers
were designed from published sequences to amplify base pairs 259–577 of
the mouse Rac1 ORF and base pairs 226 –551 of the rat RhoA ORF. These
regions were PCR amplified from an embryonic day 13 rat cDNA library,
and the bands were purified and ligated into pBLUESCRIPT II (SK
)
digested with EcoRV. Plasmids containing insertions of both orientations
were linearized with EcoRI and transcribed with T7 RNA polymerase to
generate the sense and antisense riboprobes.
Freshly dissected brains of postnatal day 8 (P8) Long–Evans rats were
submerged in ornithine carbamyltransferase and frozen in an ethanol and
dry ice bath and then stored at 80°C. T wenty micrometer cryostat
sections were prepared, fixed, washed in 1PBS, and serially dehydrated
before being returned to 80°C in a desiccated, sealed box. Sections were
hybridized and processed for autoradiography according to the protocol
used by Frantz et al. (1994).
Digoxigenin (DIG)-labeled in sit u hybridizations were performed ac-
cording to the protocol used by Wright and Snider (1995).
Received Feb. 8, 2000; revised April 19, 2000; accepted April 24, 2000.
McKnight, K lingenstein, and Sloan Fellowships to L.L. supported this study. A.Y.N.
is a Howard Hughes Medical Institute predoctoral fellow. We are gratef ul to Larry
Katz for introducing us to the biolistic transfection method and to Matt Scott for use
of his equipment. We thank members of the laboratories of Sue McConnell, Stephen
Smith, and Dan Madison and in particular Jim Weimann, Ami Okada, Aparna Desai,
Jack Waters, and Eric Schaible for sharing experimental expertise essential for our
study. We also thank Marc Symons, Rong-Guo Qiu, Li-Huei Tsai, Alan Hall, Jun-ichi
Miyazaki, Morgan Sheng, Haruhiko Bito, and Shuh Narumiya for plasmids and
Yoshitomi Pharmaceuticals for the Y-27632 compound. We thank Ben Barres, Ste-
phen Smith, Li-Huei Tsai, and members of the Luo laboratory for stimulating
discussions and their comments on thi s manuscript. We thank Haruhiko Bito and Shuh
Narumiya for communicating results before publication.
A.Y.N. and M.B.H. contributed equally to this work.
Correspondence should be addressed to Dr. Liqun Luo, Department of Biological
Sciences, Stanford University, 371 Serra Mall, Herrin Labs 144A, Stanford, CA
94305-5020. E-mail: lluo@stanford.edu.
Copyright © 2000 Society for Neuroscience 0270-6474/00/205329-10$15.00/0
The Journal of Neuroscience, July 15, 2000, 20(14):5329–5338
Molecular biology
p-chicken
actin–mouse CD8. pUC–mouse CD8 (mCD8) (Liaw et al.,
1986) was digested with X hoI and BamHI, and the end nucleotides were
filled in using K lenow to give a 900 bp fragment encoding mCD8. The
p-chicken
actin (pCA)–GAP– enhanced green fluorescent protein
(EGFP) vector (Okada et al., 1999) was digested with BamHI and NotI,
and the end nucleotides were filled to generate a 5.1 kb pCA backbone.
The mCD8 fragment and the pCA backbone were ligated together, gen-
erating the 6 kb pCA–mCD8.
pCA–hRac1V12. pBS–hRac1V12 (Luo et al., 1996) was digested with
Asp718, blunted with Klenow, and cut again with NotI to yield a 600 bp
fragment encoding hRac1V12. This fragment was ligated into the 5.1 kb
pCA backbone generated by digesting pCA–GAP–EGFP with SmaI and
NotI, yielding the 5.7 kb pCA–hRac1V12.
pCA–myc–hRac1W T. pUHD–myc–hRac1WT (Qiu et al., 1995a) was
digested with EcoRI, and the resulting fragment encoding myc–hRac1WT
(600 bp) was ligated into the compatible site of pNN0
3
.pNN0
3
–myc
hRac1WT was then digested with EcoRV and NotI, and the resulting
fragments were ligated into the pCA backbone as described above to yield
pCA–myc–hRac1WT.
pCA–myc–hRac1N17, pCA–myc–hRhoAV14, and pC A–myc–hR hoAN19.
pEXV–myc–hRac1N17, pEX V–myc –hRhoAN19, and pEXV–myc–
hRhoAV14 (Qiu et al., 1995a,b) were digested with EcoRI, and the
resulting fragments encoding myc –hRac1N17 (600 bp), myc –hRhoAN19
(1 kb), and myc–hRhoAV14 (1 kb) were ligated into the compatible site of
pBLUESCRIPT II (SK
). pBS–myc–hRac1N17, pBS –myc –hRhoAN19,
and pBS–myc–hRhoAV14 were digested with EcoRV and NotI, and the
resulting fragments were then ligated into the 5.1 kb pCA backbone
generated by digesting pCA–GAP–EGFP with SmaI and NotI. The result-
ing plasmid pCA–myc–hRac1N17 is 5.7 kb in length, whereas pCA–myc
hRhoAV14 and pCA–myc–hRhoAN19 are 6.1 kb.
pCA–myc–hRac1L61, pCA–myc–hRac1L61K40, and pCA–myc–
hRac1L61A37. pRK5–myc–hRac1L61, pRK5–myc–hRac1L61K40, and
pRK5–myc–hRac1L61A37 (Lamarche et al., 1996; Nikolic et al., 1998)
were digested with ClaI and EcoRI. The resulting 600 bp fragments
encoding myc–hRac1L61, myc–hRac1L61K40, and myc–hRac1L61A37
were ligated into the compatible sites of pBLUESCRIPT II (SK
). These
plasmids were then digested with X hoI, blunted with K lenow, and then
digested with NotI. The resulting 600 bp fragments representing the
myc-tagged Rac1 mutants were isolated and ligated into the 5.1 kb pCA
backbone generated by digesting pCA–GAP–EGFP with SmaI and NotI.
The resulting pCA–myc –hRac1L61, pCA–myc –hRac1L61K40, and pCA–
myc–hRac1L61A37 are 5.7 kb in length.
Preparation of DNA-coated gold particles
Qiagen-Midi prepped plasmids were precipitated onto 1.6
m gold beads
(Bio-Rad, Hercules, CA) at a concentration of 2
g of each plasmid per
milligram of gold beads, according to the manufacturer’s instructions.
Briefly, DNA was precipitated onto gold beads that were subsequently
dried to the sides of plastic tubing. This tubing was chopped into 0.5 inch
fragments known as “bullets.” When a gold bead was to carry two plasmids
simultaneously (cotransfection experiments), both plasmids (2
g/mg of
gold each) were first mixed and coprecipitated using a standard ethanol
precipitation protocol before precipitation onto gold beads. For “dual
gold” experiments, in which a single bullet contains beads carrying differ-
ent DNAs, the two populations of gold beads were prepared separately and
mixed just before precipitation onto the plastic tubing.
Preparation of rat hippocampal organotypic cultures
Hippocampal slices were prepared from P8 Long–Evans rats as described
previously (Stoppini et al., 1991). Briefly, the hippocampus was dissected in
ice-cold, sterile dissection medium: 1MEM (Hank’s salts, 25 mM
HEPES, without L-glutamine; from Life Technologies, Grand Island, NY)
and 100 units/ml each penicillinstreptomycin (Life Technologies). Hip-
pocampi were sliced transversely at a thickness of 400
m on a tissue
chopper (Stoelting, Wood Dale, IL) and separated from one another in
filtered and preincubated (37°C; 5% CO
2
) culture medium: 0.5MEM,
0.25HBSS (Life Technologies), 0.25horse serum (defined, heat-inac-
tivated; from HyClone, L ogan, UT), 100 units/ml each penicillin–strepto-
mycin, and 1 mML-glutamine (Life Technologies). Slices were immediately
plated onto Millicell C M membrane inserts (Millipore, Bedford, M A) in
Petri dishes containing 1 ml of preincubated culture media. Slices were
kept under 5% CO
2
at 37°C, with media changes at 1 d in vitro (DIV), 3
DIV, and every 3 d thereafter. In experiments using Y-27632 (gift from
Yoshitomi Pharmaceuticals, Tokyo, Japan) slices were transferred to Petri
dishes containing 100
MY-27632 in 1 ml of culture media just before
biolistic transfection. Control slices were similarly transferred, but to Petri
dishes containing media supplemented with the same volume of sterile
water instead of Y-27632. Animals were treated in accordance with the
animal safety protocols of the host institution.
Biolistic transfection
To identify optimal transfection and culture conditions, slices were pre-
pared from P8 rats, transfected at various DIV, and fixed for immunocy-
tochemistry 24 hr later. The number of healthy pyramidal neurons relative
to those that were dead or dying (characterized by fragmented processes or
blebbing processes, respectively) increased significantly when transfection
occurred at 2 DIV, compared with transfection at 0 or 1 DIV. When
transfection occurred at 2 DIV, the number of transfected glia increased
significantly, often obscuring the dendritic arbors we wished to quantify.
Thus, to maximize the number of healthy pyramidal cells transfected and
to minimize interfering glia, experiments throughout this study were
performed on a standard preparation: hippocampal slices prepared from
P8 rats and cultured for 2 d before transfection.
After 2 DIV, slice inserts were removed from the incubator briefly for
biolistic transfection using the Gene Gun (Bio-Rad). Gold beads contain-
ing expression plasmids were propelled from plastic tubing bullets into
slices with a rapid helium burst of 160 psi. The gold beads exited the gun
3.5 cm above the slices, which were plated toward the periphery of the
insert to avoid “ground zero” of the gold blast.
Immunocytochemistry
Slices were fixed according to the protocol used by McAllister et al. (1995)
on the insert membrane for 1.5 hr in 2.5% formaldehyde and 4% sucrose
in 1PBS and then soaked in 30% sucrose (w/v of water) for at least 2 hr.
After a quick freeze on dry ice, slices were thawed, rinsed in 1PBS for
5 min, and incubated overnight (O/N) at 4°C in blocking solution: 10%
normal goat serum and 0.25% Triton X-100 in 0.1 Mphosphate buffer.
Blocking solution was replaced with primary antibodies diluted in blocking
solution: 1:50 mouse anti-myc (Santa Cruz Biotechnology, Santa Cruz,
CA), 1:100 rat anti-mCD8 (Caltag, Burlingame, CA), or both before
another O/N incubation at 4°C. The slices were then washed three times
for 20 min each with fresh blocking solution and incubated O/N at 4°C in
secondary antibody, diluted into blocking solution: 1:200 FITC anti-mouse
(Jackson ImmunoResearch, West Grove, PA), 1:1000 indocarbocyanine
(Cy3) anti-rat (Jackson ImmunoResearch), or both (in which the low
cross-reactivity secondaries were used). Slices were subsequently washed
three times for 15 min each in 1PBS, with the first wash containing 1.5
g/ml 4,6-diamidino-2-phenylindole (DAPI), and mounted in SlowFade
according to the manufacturer’s specifications (Molecular Probes, Eu-
gene, OR).
Image analysis and quantification
Transfected pyramidal neurons were identified by their typical morpholog y
as well as their cell body locations in the CA1 and CA3 pyramidal layers as
indicated by DAPI staining (see Fig. 2A). Individual neurons were imaged
using a Zeiss microscope attached to a Bio-Rad MRC -1024 scanning laser
confocal microscope. For quantification of dendritic branch segments, a
stack of confocal images (Z steps of 1
m) taken with a 16objective and
comprising the entire cell were merged using NIH Image 1.62 (National
Institutes of Health, Bethesda, MD) and printed. From these printouts, the
number of dendritic segments was derived by counting the number of
dendrite branch points and dendrite terminal ends. To obtain the Sholl
profiles of dendritic arbors (Sholl, 1953), printouts were placed under a
clear sheet printed with concentric circles with increasing radii of 25
m.
To minimize the effect of cell body shape variation, we positioned the
center of the circles at the base of the apical dendrite for analysis of apical
dendrites. In addition, because CA1 and CA3 pyramidal neurons exhibited
different Sholl profiles (data not shown), all Sholl analysis was with only
CA1 neurons. The center of the circles was placed at the cell body edge,
opposite the apical dendrite when analyzing basal dendrites. The number
of dendrites crossing each concentric circle was then counted. If a branch
point fell on a line, it was counted as two crossings.
For spine quantification, images of apical dendrites were taken just distal
to the first apical branch point (25–100
m from the cell body). Basal
spines were imaged at the point of the first basal branch (25–75
m from
the cell body). Serial confocal images (Z steps of 0.5
m) were taken of
Cy3 fluorescence (detecting mCD8) with a 40objective with a digital
zoom factor of three. Sections were merged using N IH Image, and the
number of spines and filopodia was tallied on the screen. Spines were
defined as a headless dendritic protrusion 1–3
m long or a headed
protrusion of any length up to 3
m. Filopodia were defined as headless
protrusions 3
m but not long enough to be visible on the 16printout.
The length of all dendrites in the fields used to quantify spine density was
measured using NIH Image. The average dendritic diameter for each cell
was based on three measurements (at the start, middle, and end) of every
dendritic segment within the images used to analyze dendritic spine
density.
Identical procedures for acquiring images and measurements of den-
dritic length were also used to analyze GFP-tagged postsynaptic density-95
(PSD-95:GFP) clustering with Rac1N17 expression. To count PSD-95:
GFP clusters optimally in relation to spine profiles as defined above,
images were placed into channels in Photoshop 4.0 (Adobe Systems, San
Jose, CA). GFP clusters were counted independent of location, and the
spine profiles with GFP clusters either within the spine head or just
external to the head outline were also tallied. The dendrites and their
spines were outlined using the magic wand tool in the channel correspond-
ing to the myc label, which fills the entire dendritic tree including the
spines (see Fig. 3).
Graphs throughout represent averages and SEM. Statistical comparisons
of Sholl segments were done with StatView 5.0.1 (SAS Institute, Cary,
5330 J. Neurosci., July 15, 2000, 20(14):5329–5338 Nakayama et al. Rac and Rho Maintain Dendritic Spines and Branches
NC), and all post hoc two-tailed ttests assuming equal variances were done
using Microsoft Excel 98 (Microsoft, Seattle, WA).
RESULTS
Rac1 and RhoA are expressed in developing
hippocampal pyramidal neurons
Although Rac1 and RhoA are ubiquitously expressed (for review,
see Hall, 1994), the expression pattern of these GTPases in the
hippocampus has not been described. To identify Rac1- and RhoA-
expressing cells in the hippocampus, in situ hybridizations were
performed on coronal P8 rat brain sections, the same stage at which
our slice cultures were produced. Antisense S
35
-labeled riboprobes
were generated to the C-terminal portions of each mRNA, where
the nucleic acid sequence identities between Rac and Rho are
lowest (50% identical). As shown in Figure 1, Rac1 (Fig. 1 A) and
RhoA (Fig. 1B) antisense riboprobes labeled CA1 and CA3 pyra-
midal cell layers and the dentate granular cell layer, with the CA3
layer being the most intense. Rac and Rho transcripts are also
widely distributed in other brain areas (data not shown). The sense
riboprobes to both GTPases gave only diffuse background signal,
similar to that shown for Rac1 (Fig. 1C). In sit u hybridization using
DIG-labeled probes was performed on sections made from P10
hippocampi (equivalents to the day slices were transfected), with
similar results (data not shown). No obvious dendritic localization
of the two mRNAs was observed.
Pyramidal neurons maintain in vivo characteristics while
in slice culture
We have used particle-mediated gene transfer (biolistic transfec-
tion) (Arnold et al., 1994; Lo et al., 1994) of mCD8 cDNA, and
subsequent immunocytochemical detection of the protein, to visu-
alize the dendritic arbors and spines of pyramidal neurons in
hippocampal slice cultures. mCD8 is a cell-surface marker shown
previously to label the entire morphology and fine structures of
Drosophila neurons without toxic side effects (Lee and L uo, 1999).
As shown in Figure 2, delivery of a single gold particle carrying
mCD8 under the control of the chicken
-actin promoter resulted
in the intense labeling of not only the axon and the entire dendritic
arbor (Fig. 2B) but individual spines as well (Fig. 2C). In addition
to pyramidal neurons, many other hippocampal cell types were
readily transfected, including glia, the granule cells of the dentate
gyrus, and interneurons (Fig. 2A).
In our standard preparation, hippocampal slices were obtained
from postnatal day 8 rat pups and cultured for 2 d before transfec-
tion (see Materials and Methods). To determine the developmental
state of pyramidal neurons at the time of transfection, 1 d in vitro
cultures were transfected with mCD8 and fixed 24 hr later (2 DIV).
By this time, pyramidal neurons had already acquired their char-
acteristic dendritic-branching pattern, including well differentiated
apical and basal dendrites (Fig. 2B). Counting the number of
dendritic branch segments (see Materials and Methods) gave us a
quantitative measure of dendritic complexity and revealed that
over the first few days after transfection, there is a gradual increase
in the number of dendritic segments (Fig. 2 E). At each time point
examined, we found no statistically significant difference between
the number of dendritic segments for apical and basal dendrites or
between CA1 and CA3 neurons (data not shown).
These 2 DIV neurons also displayed numerous protrusions from
their apical and basal dendrites. Although a minority of these
protrusions were filopodial in shape (Fig. 2C,arrowheads), many
more had the well defined neck and head structure characteristic of
mature spines found in adults (Fig. 2C,arrows). To verify indepen-
dently the maturity of these dendritic protrusions in our culture, we
cotransfected GFP-tagged PSD-95 as a marker for postsynaptic
density (Arnold and Clapham, 1999). We found that PSD-95:GFP
labeling was concentrated in most of the heads of the protrusions
but was absent from longer filopodia (Fig. 2D). Using the morpho-
logical criterion described above (see Materials and Methods for
details), we quantified the density of spine-like protrusions and
filopodial protrusions. We found that spine density increased with
the time spent in culture, whereas the number of filopodial protru-
sions gradually diminished (Fig. 2 F). The spine densities found in
our cultures are similar to those reported previously for similarly
aged cultures [P12 equivalent 2 spines/10
m by EM (Boyer et
al., 1998)]. In accordance with Drakew et al. (1996), we did not find
a significant difference in spine density between the apical and
basal dendrites at any time point examined (data not shown).
Additionally, we did not observe a significant difference in spine
density between CA1 and CA3 pyramidal dendrites (data not
shown). Because CA1 and CA3 neurons did not differ in either
their branch segment number or their spine density, neurons from
both areas have been combined in our quantification of branch
segments and spine densities.
Expression of dominant-negative Rac1 results in a
progressive reduction of spine density
To study the function of Rho family GTPases in the maintenance
of dendritic structures, we cotransfected pyramidal neurons in our
standard preparation with both mCD8 and dominant mutants of
the GTPases. Cotransfection was accomplished by the introduction
of gold beads coated with plasmids for both genes, each under the
control of the chicken
-actin promoter. Although analogous ex-
periments with two marker genes (mCD8 and GAP–EGFP)
showed a high cotransfection rate (85%), quantitative variations
in the relative expression levels of the two constructs were ob-
served. Therefore, we independently monitored the expression of
the dominant mutants by using myc epitope-tagged GTPases, an-
alyzing only those cells in which both plasmids were strongly
expressed. As an internal control and to minimize variation among
hippocampal slices, most experiments were performed using dual
gold preparations. In these experiments, slices were transfected
with a mixed population of gold beads, some carrying the mCD8
and GTPase plasmids and some the mCD8 plasmid alone. This
enabled us to visualize both control (mCD8 alone) and experimen-
tal (mCD8 and GTPase) neurons from the same slice (see Fig. 3C).
We found that transfected wild-type Rac1, as well as all mutant
proteins, was distributed throughout the entire dendritic tree, in-
cluding dendritic spines and other fine processes (see Figs. 3A,C,
4B–D,F–H).
To test whether endogenous Rac1 is necessary for the mainte-
nance of dendritic spine and branch morphology, we cotransfected
pyramidal neurons with gold beads carrying mCD8 and myc-tagged
dominant-negative Rac1 (Rac1N17). This allele is thought to exert
its dominant-negative effect by sequestering rate-limiting GDP–
GTP nucleotide exchange factors necessary for activation of en-
dogenous Rac1 (Ridley et al., 1992). Expression of Rac1N17 re-
sulted in a significant time-dependent loss of pyramidal dendritic
Figure 1. Rac1 and RhoA are expressed in developing hippocampus. A,
Dark-field image of a coronally sectioned P8 rat hippocampus hybridized
with an antisense S
35
-riboprobe against Rac1, showing distribution of Rac1
mRNA in the dentate gyrus and CA1 and CA3 pyramidal cell layers. B,
Dark-field image of antisense riboprobe showing distribution of RhoA
mRNA. C, Dark-field image of the Rac1 sense riboprobe showing back-
ground staining. D, Bright-field image of the same hippocampus shown in
Cwith the CA1, CA3, and dentate gyrus (DG) labeled. Scale bar, 2 mm.
Nakayama et al. Rac and Rho Maintain Dendritic Spines and Branches J. Neurosci., July 15, 2000, 20(14):5329–5338 5331
Figure 2. Development of hippocampal pyramidal neurons in
cultured slices. A–C, Representative images of transfected hip-
pocampal neurons show the developmental stage at the onset of
experiments in our standard preparation. For these images
only, transfection was performed at 1 DIV, and slices were
fixed 24 hr later. A, Low magnification is shown of a hippocam-
pal slice biolistically transfected with mCD8 (red), with brack-
ets defining the DAPI-labeled (blue) pyramidal cell layers of
CA1 and CA3. Pink labeling (from overlapping red and blue
signals) represents transfected cells, including a granule cell in
the dentate gyrus (arrowhead) and a pyramidal neuron over-
lapped by glia in the CA1 cell layer (ar row). B, A CA3 pyra-
midal neuron (composite confocal image using a 16objec-
tive) is shown. The immunocytochemical detection of mCD8
reveals the structure of both apical and basal dendrites, as well
as the axon (arrow). C, The spines on the apical dendrites of the
neuron pictured in B(composite confocal image using a 100
objective) are shown. Arrows point to spines with characteristic
head and neck morphology, whereas arrowheads show filopo-
dial protrusions. D, Double labeling of mCD8 and PSD-95:
GFP proteins in apical dendrites (composite confocal images
using a 40objective with a digital zoom factor of 3) is shown.
mCD8 labeling is in red (D,D⬘⬘), and PSD-95:GFP labeling
(PSD-95:GFP)ising reen (D,D⬘⬘). Arrows indicate spines with
a head (PSD-95:GFP positive), whereas arrowheads indicate
filopodial protrusions (PSD-95:GFP negative). E, Dendritic
branch segments gradually increase on both apical and basal
dendrites with successive days in culture (apical, n10, 13, 8,
10; basal, n10, 14, 7, 9 for 2, 4, 6, 9 DIV, respectively). F,
Spine density increases over successive days in culture, whereas
filopodial protrusions decrease. Spines and filopodia were
counted from stacked confocal images collected at 40with a
zoom of three, from a stereotyped region of the dendrite as
defined in Materials and Methods (apical, n10, 19, 17, 12;
basal, n11, 14, 17, 11 for 2, 4, 6, 9 DIV, respectively). Scale
bars: A, 1 mm; B,50
m; C, D,10
m.
Figure 3. Dominant-negative Rac1 expression results in a
progressive reduction of the dendritic spine density and mild
changes in the dendritic-branching pattern. A, Transfected
myc-tagged wild-type Rac1 protein is distributed along the
dendrites and in dendritic spines ( green, anti-myc; red, anti-
mCD8; composite confocal image using 40objective; digital
zoom factor of 3). B, Representative images are shown of
apical dendrites that were transfected with the marker mCD8
alone (top) or with mCD8 and myc-tagged Rac1N17 (bottom)
for 1, 2, or 3 d (composite confocal images of mCD8 staining
using 100objective). C, Apical dendrites from neighboring
pyramidal neurons are shown3dafter expressing mCD8 alone
(red) or expressing both mCD8 (red) and myc-tagged Rac1N17
(green) and therefore appearing yellow (composite confocal
image using a 40objective with a digital zoom factor of 3).
Although it appears that Rac1N17 dendrites have thicker den-
drites in this image, quantification of the average diameter of
dendrites within every image used to measure dendritic spine
density does not reveal any significant difference (paired ttest,
apical, p0.19; Rac1N17 1.50 0.14
m; control 1.22
0.51
m; n18, 9, respectively; basal, p0.12; Rac1N17
0.83 0.1
m; control 1.05 0.08
m; n9, 8, respective-
ly). D, Rac1N17 expression progressively reduces the number
of spines on apical and basal dendrites (for 1, 2, 3 d, apical
mCD8, n16, 19, 10; apical Rac1N17, n19, 11, 24; basal
mCD8, n15, 14, 10; basal Rac1N17, n18, 11, 17, respec-
tively). E, Quantification of dendritic branch segments after 3 d
of Rac1N17 expression is shown (n17, 23, 18, 21 for apical
mCD8, apical Rac1N17, basal mCD8, basal Rac1N17, respec-
tively). F, Sholl profiles of the basal dendrites of Rac1N17-
expressing and control CA1 pyramidal neurons (3 d after transfection) illustrate the slight change in dendritic-branching pattern with Rac1N17 expression
(n14, 15 for basal mCD8, Rac1N17, respectively). Post hoc t tests reveal that the only significant difference occurs at 50
m (paired ttest, **p0.01).
Scale bars: A, B,5
m; C,10
m.
5332 J. Neurosci., July 15, 2000, 20(14):5329–5338 Nakayama et al. Rac and Rho Maintain Dendritic Spines and Branches
spines from both apical and basal dendrites. Whereas neurons
expressing mCD8 alone displayed an increasing trend of spine
density with successive days after transfection, those also express-
ing Rac1N17 possessed fewer and fewer spines on both their apical
and basal dendrites (Fig. 3B—D; see also Fig. 2F).Only1dafter
transfection, the basal dendritic spine density of Rac1N17 neurons
was significantly reduced compared with that of neurons expressing
mCD8 alone (68% fewer; p0.05), whereas the apical spine
density was significantly lower than that of controls2dafter
transfection (54% fewer; p0.05). An example of a dual gold
experiment is shown in Figure 3C.After3dofexpression, mCD8-
transfected pyramidal dendrites (red) displayed adult-like spines,
whereas the dendrites from Rac1N17-expressing cells ( yellow be-
cause of red labeling of mCD8 and green labeling of myc-tagged
Rac1N17) were nearly devoid of spines.
To address whether the corresponding synapse number is de-
creased with the loss of the dendritic spine profile in Rac1N17-
transfected neurons, we cotransfected PSD-95:GFP with Rac1N17.
In agreement with the loss of dendritic spine density, we observed
a significant decrease in PSD-95:GFP clusters with3dofRac1N17
expression ( p0.05; Rac1N17 2.14 0.35; mCD8 3.84
0.52; units are PSD-95:GFP clusters per 10
m; n9, 8,
respectively).
To determine whether Rac1 may play a role in the maintenance
of dendritic growth and branching, we quantified the number of
dendritic segments after3dofRac1N17 expression. As shown in
Figure 3E, there is no statistically significant difference between
Rac1N17 and control cells (apical, p0.38; basal, p0.48). As
another measure of dendritic complexity, a standard Sholl analysis
(Sholl, 1953), which counts the number of dendritic crossings at 25
m concentric circles, was performed. Although the Sholl profiles
for the apical dendrites of Rac1N17-expressing neurons and con-
trol neurons are similar (repeated measures ANOVA, p0.40),
their basal profiles are slightly different (Fig. 3F; repeated measures
ANOVA, p0.05). There are more branches concentrated in the
proximal regions in control neurons compared with Rac1N17 neu-
rons, which appear to have more distally distributed branches (Fig.
3F). This mild effect of Rac1N17 on the dendritic-branching pat-
tern of pyramidal neurons is in contrast to its pronounced effect on
spine density, suggesting a preferential role for endogenous Rac1
in the maintenance of dendritic spines.
Activated Rac1 disrupts normal spine morphology
To elucidate how Rac1 activity affects spine morphogenesis, we
transfected pyramidal cells with a constitutively activated form of
Rac1 that is deficient in GTP hydrolysis (hereafter referred to as
activated Rac1). Previous transgenic expression of the activated
allele Rac1V12 in Purkinje cells results in a slight simplification of
dendritic complexity, a loss of normal dendritic spines, and the
formation of supernumerary smaller protrusions that were only
detectable with EM (Luo et al., 1996). We found that pyramidal
neurons expressing Rac1L61, another activated Rac1 allele, for 1,
2, or 3 d also showed similar defects in spine development (Fig.
4B). With the exception of some distal dendrites, apical and basal
dendrites lacked normal spines. In contrast to our control neurons
(Fig. 4A), which have even dendritic branches that are regularly
punctuated by spines, Rac1L61 dendrites were irregular in thick-
ness because of the presence of regions of numerous overlapping
bumps and ruffle-like structures (Fig. 4B,F, arrowheads). Addition-
ally, numerous long and fine processes were found on the cell soma
and proximal dendritic shafts (Fig. 4, compare F,arrows,E).
Although these fine processes prevented quantification of dendritic
segments, 70% of Rac1L61-expressing neurons had an otherwise
normal dendritic-branching pattern despite their perturbed den-
dritic spines. These findings suggest that dendritic spines are more
sensitive to the hyperactivation of Rac1 than are the dendritic
branches themselves. Expression of another constitutively active
form of Rac1, Rac1V12, in hippocampal neurons gave indistin-
guishable results compared with Rac1L61 expression (data not
shown). Although normal dendritic spines are lost with the expres-
sion of either dominant-negative or constitutively active Rac1,
Rac1L61 expression resulted in a net increase in dendritic protru-
sion exemplified by an increase in filopodial-like structures and
membrane ruffling, which were not seen with Rac1N17 expression
(Fig. 3B,C).
The F37A effector domain mutant abolishes the
activated Rac1 phenotype
Rac1 binds to different effectors to activate distinct downstream
signal transduction pathways. Previous in vitro experiments have
characterized several effector domain mutants that separate Rac1’s
role in lamellipodia formation from its role in activation of the Jun
kinase cascade and nuclear signaling (Joneson et al., 1996;
Lamarche et al., 1996). The Y40K mutation abolishes the binding
of several CRIB-containing proteins [e.g., p21-activated kinase
(Pak)] and when combined with an activating Rac1 mutation blocks
the stimulation of the Jun kinase (JNK) pathway without affecting
membrane ruffling and lamellipodia formation. Conversely F37A
mutation does not bind to Rho-associated kinase (ROCK) and
prevents the induction of membrane ruffling and lamellipodia by an
activating mutation of Rac (Lamarche et al., 1996), yet it still binds
to Pak and activates the JNK pathway. These data suggest that
effectors dependent on tyrosine residue 40 for binding to activated
Rac (e.g., Pak and other CRIB-containing proteins) are required
for the activation of Jun kinase, whereas effectors dependent on
phenylalanine residue 37 (e.g., ROCK) for binding are candidate
mediators for Rac1’s regulation of lamellipodia formation.
We attempted to identify the pathway responsible for the acti-
vated Rac1 phenotype in hippocampal pyramidal neurons using
these same effector domain mutants. Rac1L61K40 expression re-
sulted in irregular, bumpy dendrites containing numerous thin
processes (Fig. 4C,G) that were indistinguishable from Rac1L61
dendrites (Fig. 4B,F ), suggesting that effectors dependent on ty-
rosine 40 for binding (e.g., Pak) are not necessary to mediate the
effects of activated Rac1. In contrast, the dendritic morphology and
spine density of pyramidal neurons expressing Rac1L61A37 (Fig.
Figure 4. Effects of activated Rac1 expression on dendritic
morphology. A–H, Representative distal ( A–D) and proximal
(E–H, just above cell bodies) apical dendrites and spines of
hippocampal pyramidal neurons, 24 hr after being transfected
with mCD8 only (A, E), mCD8 and Rac1L61 (B, F ), mCD8 and
Rac1L61K40 (C, G), or mCD8 and Rac1L61A37 (D, H ) (com-
posite confocal images using a 40objective with a digital
zoom factor of 3). Red staining in all images represents anti-
mCD8 immunoreactivity, and green staining in B–D and F–H
represents anti-myc immunoreactivity. All mutant Rac proteins
were expressed at comparable levels and were distributed
throughout the entire neuron on the basis of their myc staining
(see yellow labeling in B–D,F–H because of overlapping green
signal for myc and red signal for mCD8). Asterisk s represent
mature spines with heads, ar rowheads indicate ruffle-like struc-
tures, and arrows point to long and thin filopodial-like protru-
sions. Scale bars, 10
m.
Nakayama et al. Rac and Rho Maintain Dendritic Spines and Branches J. Neurosci., July 15, 2000, 20(14):5329–5338 5333
4D,H) resembled those of control neurons (Fig. 4A,E), implying a
requirement for effectors dependent on phenylalanine 37 to medi-
ate the activated Rac1 phenotype. ROCK is one such protein
(Joneson et al., 1996; Lamarche et al., 1996). We attempted to
mimic the Rac1L61A37 rescue by treating Rac1L61-transfected
slices with 100
MY-27632, a compound that specifically inhibits
ROCK activity without affecting other similar kinases (see Uehata
et al., 1997; Madaule et al., 1998). Although this treatment effec-
tively blocked the activated RhoA phenotype (see below), the
Rac1L61 effect persisted, suggesting that ROCK is not responsible
for mediating the effect of activated Rac1.
Expression of activated RhoA results in marked
simplification of the dendritic tree
Although expression of activated Rac1 had a mild effect on the
overall dendritic branch complexity (discounting the filopodia-like
thin processes), expression of activated RhoA (RhoAV14, analo-
gous mutation to Rac1V12) caused a drastic simplification of the
dendritic tree compared with those of neurons expressing mCD8
alone (Fig. 5A,B). This effect was evident1dafter transfection and
persisted for 2 and 3 d after transfection. Branch segment quanti-
fication (Fig. 5E) indicated a highly penetrant and significant sim-
plification of both the apical and basal dendrites as early as1dafter
transfection of RhoAV14 (ttest, p0.001 for both apical and
basal). The Sholl profile indicates that both the length and branch-
ing pattern of RhoAV14-expressing neurons are affected (Fig. 5G;
repeated measures ANOVA, p0.001). Interestingly, there ap-
peared to be a recovery of dendritic simplification after3dof
RhoAV14 transfection, as seen by quantification of both apical and
basal dendritic segment numbers (Fig. 5E), despite the constant
level of RhoAV14 expression as judged by myc staining (data not
shown).
Strikingly, many neurons expressing activated RhoA (80%) had
unusual structures at the end of their reduced dendrites: threads of
extremely thin processes trailing behind the shortened dendrites
(Fig. 5D) resembling certain aspects of retracting axon terminals
(for review, see Bernstein and Lichtman, 1999). These shortened
dendrites often exhibited a fattened portion that appeared to have
an increased aggregation of cytoplasm and membranous structures
on the basis of the intensity of the cytoplasmic-myc and plasma
membrane-localized CD8 labeling.
Perturbation of endogenous RhoA activity with transfection of
the dominant-negative RhoA construct (RhoAN19, analogous mu-
tation as Rac1N17) did not affect the dendritic morphology of our
pyramidal neurons (Fig. 5C,F,H ) despite their high level of expres-
sion (see Fig. 5C, inset). In addition, expression of RhoAN19 did
not significantly affect the spine density when compared with con-
trol neurons2dafter transfection ( p0.56; n15 and 13 for
RhoAN19 and mCD8 alone, respectively), supporting the specific-
ity of the spine reduction effect seen with dominant-negative
Rac1N17 expression.
The ROCK inhibitor blocks RhoAV14-induced
dendritic simplification
RhoA acts via many different effector molecules to regulate diverse
cellular functions, including organization of the actin cytoskeleton.
To investigate the pathway by which RhoA regulates dendritic
complexity, we tested the involvement of ROCK, because ROCK
inhibition has been shown to block Rho-induced cell rounding and
process retraction in cultured neuroblastoma cells (Hirose et al.,
1998). We applied the ROCK inhibitor Y-27632 (100
M) (Uehata
et al., 1997) in culture media at the time of transfection. We limited
our quantification to the basal dendrites of neurons2dafter
transfection because the dendritic simplification was most robust at
this time, and both apical and basal dendrites were equally affected
by RhoAV14 expression (Fig. 5E).
The treatment of slices with Y-27632 alone did not result in a
significant change in dendritic complexity (Fig. 6 B,F;p0.30).
Remarkably, Y-27632 treatment completely blocked the dendritic
simplification associated with RhoAV14 expression (Fig. 6, com-
pare D, C). As is quantified in Figure 6 F, although RhoAV14
expression caused a fivefold decrease in dendritic segments, appli-
Figure 5. Expression of activated RhoA results in den-
dritic simplification. A–C, Representative images of pyra-
midal neurons that have expressed the marker mCD8
alone (A), mCD8 and myc-tagged RhoAV14 (B), or
mCD8 and myc-tagged RhoAN19 ( C) for 1 d (composite
confocal images using 16objective). Insets, Myc immu-
nostaining for RhoAV14 (B) and RhoAN19 ( C). D, Tip
of an apical dendrite from a pyramidal neuron expressing
RhoAV14 for 2 d (composite confocal image using 40
objective, with a digital zoom factor of 3). The soma is
toward the bottom of the image. E, Quantification of
dendritic branch segments after 1, 2, and3dofRhoAV14
expression. Both apical and basal dendrites exhibit a re-
duced number of dendritic segments (see Materials and
Methods) with RhoAV14 expression compared with that
of neurons expressing mCD8 alone (***p0.001; **p
0.01; for 1, 2, 3 d, apical mCD8, n14, 13, 17, and
RhoAV14, n25, 19, 23; basal mCD8, n11, 14, 18, and
RhoAV14, n25, 18, 22). F, Quantification of dendritic
branch segments after 1, 2, and 3 d of RhoAN19 expres-
sion. Neurons expressing RhoAN19 do not show a change
in dendritic branch segment number (for 1, 2, 3 d, apical,
n9, 10, 7; basal, n7, 11, 7, respectively). G, H, Sholl
profiles for basal dendrites of CA1 neurons2dafter
expressing mCD8 alone (n11) or expressing RhoAV14
(G;n16) or RhoAN19 (H;n8). Scale bars: A–C and
insets,50
m; D,10
m.
5334 J. Neurosci., July 15, 2000, 20(14):5329–5338 Nakayama et al. Rac and Rho Maintain Dendritic Spines and Branches
cation of Y-27632 restored the number of dendritic segments to the
level of our control neurons. This result suggests that ROCK is a
mediator of RhoAV14-induced dendritic simplification.
The significant dendritic simplification induced with RhoAV14
expression impeded our ability to quantify the density of dendritic
spines. On a qualitative level it appeared that expression of acti-
vated RhoA resulted in a loss of adult-like dendritic spines. Inter-
estingly, although Y-27632 alone does not affect spine density, it
also restores spine density of RhoAV14-expressing neurons close to
the control level (Fig. 6G;p0.08).
ROCK activation is sufficient to induce
dendritic simplification
We next tested whether activation of ROCK is sufficient to induce
dendritic simplification. Pyramidal neurons were transfected with a
truncation mutant, ROCK3, which eliminates the negative regu-
latory domain in the C terminal of the ROCK protein and has been
shown to behave as a constitutively active version of ROCK in
cultured cells (Ishizaki et al., 1997). We found that ROCK3
expression resulted in the reduction of dendritic complexity to an
extent similar to that of RhoAV14 expression (Fig. 6 E,H), suggest-
ing that ROCK activation is not only necessary but also sufficient in
inducing the pruning of dendritic branches.
DISCUSSION
Using biolistic transfection to introduce dominant mutants of Rac1
and RhoA acutely in hippocampal pyramidal neurons with estab-
lished dendritic arbors and some adult-like dendritic spines, we
have shown that these GTPases play important roles in the main-
tenance and reorganization of dendritic structures. Rac1 seems to
have preferential roles in regulating spine morphogenesis, whereas
RhoA is implicated in limiting the growth of dendritic branches.
The fact that expression of analogous dominant-negative or acti-
vated mutants of these similar GTPases gave qualitatively different
phenotypes argues for the relative specificity of these perturbation
experiments.
Generation and maintenance of dendritic spines
Although dendritic spines are important sites of synaptic input and
plasticity (Harris and Kater, 1994; Yuste and Tank, 1996), little is
known about the molecular mechanisms that regulate their mor-
phogenesis (Harris, 1999). We have shown previously that trans-
genic expression of activated Rac1 in mouse Purkinje cells results
in drastic changes in dendritic spine morphology (Luo et al., 1996).
In those experiments, transgenes were expressed before the onset
of dendritic growth and remained active thereafter. Although we
concluded that hyperactivation of Rac affects the development of
dendritic spines, we could not determine whether regulation of Rac
activity is important for the dynamic changes of dendritic spines in
more mature neurons, which may contribute to the morphological
plasticity of neurons underlying learning and memory (Harris and
Kater, 1994).
In this study, we have extended the previous findings in several
new directions. First, because expression of activated Rac1 mutants
in hippocampal pyramidal neurons results in a spine perturbation
phenotype analogous to transgenic expression of activated Rac1 in
Purkinje cells, it seems that the effect of increased Rac1 activity in
dendritic spine morphogenesis is not specific to Purkinje cells. A
notable difference between the two experimental conditions, how-
ever, is the presence of prof use long filopodial-like extensions in
this study (Fig. 4) that were not observed in the previous transgenic
studies (Luo et al., 1996). This could be caused by the differences
in cell types, transgene expression level, and the relative timing of
transfection versus neuronal differentiation. For instance, in the
transgenic experiments, neurons may have adapted to the constant
high level of Rac1 activity over a long period of time, such that the
filopodial response is diminished. A second new insight gained
from the current study is that activated Rac1 expression perturbs
the maintenance of spines, because at the time of transfection these
pyramidal neurons already possess dendritic spines.
Third and more importantly, our current study finds that expres-
sion of dominant-negative Rac1 resulted in a progressive reduction
Figure 6. Function of ROCK in the maintenance of dendritic branches. A–E, Representative images of neurons2dafter transfection and 100
MY-27632
treatment are shown. Neurons were transfected with mCD8 alone (A, B) or cotransfected with RhoAV14 (C, D) or ROCK3(E). E,Inset, The myc
immunostaining for ROCK3 is shown. Additionally, neurons in Band Dwere treated with 100
MY-27632 at the time of transfection (composite confocal
images using 16objective). F, Quantification of basal dendritic branch segment numbers of mCD8- and of mCD8 plus RhoAV14-expressing neurons with
or without Y-27632 treatment is shown. Y-27632 treatment alone does not affect dendritic segment number ( p0.31; control treatment, n23; Y-27632
treatment, n23). Y-27632 treatment blocks RhoAV14-associated dendritic segment reduction (***p0.001; RhoAV14 control, n17; RhoAV14
expression Y-27632 treatment, n24). G, Y-27632 application does not alter the basal dendritic spine density of neurons expressing mCD8 alone ( p
0.65; n29, 21 for control, Y-27632 treatment, respectively). Y-27632 treatment is capable of restoring the spine density of neurons expressing RhoAV14
close to control level ( p0.08; n16 for Y-27632 treatment and RhoA V14 expression). H, Activated ROCK3 expression results in significant reduction
of dendritic segments compared with that in mCD8 alone ( p0.001 for both apical and basal; n19, 20, respectively, for ROCK3). Scale bars, 50
m.
Nakayama et al. Rac and Rho Maintain Dendritic Spines and Branches J. Neurosci., July 15, 2000, 20(14):5329–5338 5335
of spine number in both apical and basal dendrites. Taken together
with the expression of Rac1 mRNA in the hippocampal pyramidal
neurons at the time of our experiments, these results suggest that
endogenous Rac1 is used for the maintenance of spine density in
maturing neurons. Because of the slow but significant turnover of
dendritic spines at analogous developmental stages reported from
several recent live-imaging studies (Engert and Bonhoeffer, 1999;
Horch et al., 1999), we cannot distinguish between the following
two possibilities. (1) Rac1 is required for the generation of new
spines. The net spine loss over time is a result of the failure of new
spine formation while existing spines naturally turn over. (2) Rac1
is required for the maintenance of existing spine structure. Reduc-
tion in Rac activity speeds up the turnover rate of existing spines.
Future time-lapse observation of dominant-negative Rac1-
transfected neurons may help to distinguish between these two
possibilities.
Interestingly, the number of PSD-95:GFP clusters, markers for
postsynaptic density, decreases with Rac1N17 expression com-
pared with controls. Although it is not possible to conclude whether
loss of Rac1 activity causes a loss of PSD-95 clustering before spine
structure loss or vice versa, this result strongly suggests that syn-
apses are lost with concomitant loss of dendritic spines with
Rac1N17 expression. It will be interesting to determine whether
Rac1 activity directly regulates the development and maintenance
of the postsynaptic density or whether it acts on the overall integ-
rity of the dendritic spine itself.
Rac has been shown to regulate a number of biological processes
including lamellipodia formation, transcription regulation, and cell
cycle progression via activation of distinct effectors and down-
stream signaling pathways (Van Aelst and D’Souza-Schorey, 1997).
We attempted to define the downstream signal transduction path-
ways by using an activated Rac1 with specific effector domain
mutations shown previously to bind to a subset of effectors and
have a subset of biological effects (Joneson et al., 1996; Lamarche
et al., 1996). We found that the effector domain mutant shown
previously to block the ability of Rac1L61 to induce lamellipodial
formation in fibroblasts (Rac1L61A37) eliminates Rac1L61-
induced morphological changes in pyramidal neurons. The similar
properties of Rac1L61A37 on spine morphogenesis and lamellipo-
dial formation, along with the fact that Rac proteins are found
throughout the dendritic tree (Fig. 3A), suggest that Rac’s role in
spine morphogenesis is more likely via local regulation of the actin
cytoskeleton.
Generation and maintenance of dendritic branches
Previous studies have implied that the Rho GTPases have different
effects on the growth of axons and neuronal processes from neu-
ronal cell lines and primary neuronal cultures. It has been generally
thought that although Rac and Cdc42 have positive effects on
process extension, Rho has a negative effect on process outgrowth
or a positive effect on process retraction. Thus Rho activation
opposes the effect of Rac (see Kozma et al., 1997; Van Leeuwen et
al., 1997). An exception to this generalization was reported by
Threadgill et al. (1997), who showed similar effects of Rho and that
of Cdc42 and Rac on dendritic as well as axonal growth in disso-
ciated mammalian cortical neurons in culture.
We investigated the effects of Rac1 and RhoA in the regulation
of dendritic branch dynamics in neurons that have a relatively
mature and established dendritic tree (Fig. 2 B). Expression of
dominant-negative and constitutively active Rac1, although pro-
foundly affecting dendritic spines, had a relatively mild effect on the
overall dendritic tree complexity (Figs. 3, 4). In contrast, expression
of activated RhoA resulted in a drastic reduction of dendritic
branches (Fig. 5B,E,G).
Several lines of evidence suggest that the dendritic simplification
associated with RhoAV14 expression is caused by the retraction of
existing dendritic branches. First, at the time of transfection the
dendrites are more complex than after 24 hr of RhoAV14 expres-
sion (compare Figs. 2B, 5B). The dynamics of dendrite addition
and elimination (Dailey and Smith, 1996) is too slow for the simple
blockage of new dendrite formation to explain the degree of
simplification observed after1dofRhoAV14 expression. Second,
if the elimination of existing dendrites were caused by local degen-
eration of dendrites, one would expect to catch remnants of such
events. Although we occasionally observed this in all transfection
conditions when the slice culture was not healthy, we observed no
increase of such remnants above background associated with
RhoAV14-transfected neurons. Finally, in most of the RhoAV14-
expressing neurons, we observed a unique enlargement at the end
of their dendrites. It appeared that the dendrites have an aggrega-
tion of cytoplasm and plasma membrane at the tip of the simplified
dendrites. The detailed process involved in the reduction of the
dendritic branches and the identity of the enlarged end structure
remain to be characterized by live imaging and electron micros-
copy, respectively.
Several different effectors have been identified for RhoA signal-
ing to the actin cytoskeleton (for review, see Narumiya, 1996; Van
Aelst and D’Souza-Schorey, 1997). In particular, a multidomain
serine/threonine kinase ROCK has been shown to mediate RhoA’s
effect on cell rounding and process retraction from cultured neu-
roblastoma cells (Hirose et al., 1998). This RhoA-mediated retrac-
tion is likely to result from RhoA regulation, via ROCK, of myosin
light chain phosphorylation and actomyosin contractility (Jalink et
al., 1994; Kimura et al., 1996; Hirose et al., 1998). We have
extended these cell culture studies into hippocampal pyramidal
neurons in slice cultures that retain a relatively native environ-
ment. We found that treatment with the ROCK inhibitor Y-27632
(Uehata et al., 1997) effectively blocked the dendritic simplification
caused by RhoAV14 expression, indicating that ROCK is necessary
for Rho-mediated dendritic branch reduction. The fact that expres-
sion of an activated ROCK mutant mimics the effect of RhoAV14
expression further supports the importance of the RhoA–ROCK
pathway in mediating dendritic branch elimination.
Bito et al. (2000) have demonstrated recently that the RhoA–
ROCK pathway is vital for axonogenesis in cerebellar granule cells.
Despite the expression of RhoA mRNA and the highly penetrant
dendritic simplification phenotype with either activated RhoA or
ROCK, we did not detect any significant effect of inhibiting the
activity of RhoA (by expressing dominant-negative RhoAN19) or
ROCK (by application of Y-27632) on dendritic growth and
branching complexity. One explanation is that RhoAN19 is not
sufficient to block endogenous Rho activity, although similar treat-
ment has been shown to be effective in blocking axonogenesis (Bito
et al., 2000). We favor the hypothesis that the RhoA–ROCK
pathway, although intact in these neurons, is primarily inactive
during normal physiological conditions, thus allowing for the main-
tenance of the dendritic tree (and their spines). Inhibition of a
primarily inactive pathway would not result in detectable conse-
quences. The fact that inhibition of ROCK activity by itself does
not result in any significant change of dendritic complexity, but can
potently inhibit the effect of RhoA activation, strongly supports this
possibility. Such an intact yet dormant signaling pathway may be
useful in reorganizing local dendritic branches in response to local
activation of RhoA.
In a separate study, we showed that removal of RhoA activity by
null mutations of RhoA in Drosophila mushroom body neurons in
mosaic animals resulted in an overextension of dendrites. Con-
versely, expression of activated RhoA in mushroom body neurons
resulted in a significant reduction of dendritic field volume and
complexity (Lee et al., 2000). Taken together, these studies indicate
that RhoA plays an evolutionarily conserved role in limiting den-
dritic growth and complexity.
Functional significance of dendritic structural changes
in normal physiology and in mental retardation
One implication from our study is that molecules and mechanisms
used for initial dendritic morphogenesis may be reused in maturing
neurons for the continuous reorganization of neuronal cytoarchi-
tecture, possibly mediating plasticity in response to environmental
changes. We found that similar to their effect in development, Rac1
5336 J. Neurosci., July 15, 2000, 20(14):5329–5338 Nakayama et al. Rac and Rho Maintain Dendritic Spines and Branches
had a positive effect on dendritic outgrowth (at the level of den-
dritic spines and filopodia-like extensions), although hyperactiva-
tion of Rac could lead to disruption of normal spine morphology.
On the other hand, RhoA hyperactivation had a negative effect on
the maintenance of dendritic branches. Recent findings demon-
strate that dendritic spines can undergo dynamic changes in re-
sponse to synaptic stimulation under conditions that generate long-
term potentiation (Engert and Bonhoeffer, 1999; Toni et al., 1999)
or to local increases in the neurotrophin BDN F (Horch et al.,
1999). Estrogen treatment also increases dendritic spine density in
vivo (Woolley and McEwen, 1993). Regulation of Rac GTPase
activity may be one way these extracellular factors and synaptic
activity are able to exert their effects on dendritic spines. It is
noteworthy that links between RhoA and neurotrophin receptor
p75
NTR
and NMDA have been suggested by two recent studies
(Yamashita et al., 1999; Li et al., 2000).
One consequence of how the misregulation of the Rho family of
GTPases may adversely affect the physiology of neurons comes
from the recent identification of genes responsible for nonspecific
X-linked mental retardation (for review, see Chelly, 1999). T wo of
the first three identified genes encode components of the Rho
GTPase-signaling pathways, oligophrenin (Billuart et al., 1998) and
Pak3 (Allen et al., 1998). Oligophrenin possesses a GAP domain
for Rho GTPases and has GAP activity in vitro for Rho, Rac, and
Cdc42 (Billuart et al., 1998). Loss of GAP activity would increase
the activity of Rho GTPases and would lead to disruption of spine
morphology or pruning of dendritic branches according to our
present study. Pak3 belongs to a family of p21-activated kinases
that was identified as effectors of Rac and Cdc42 (Manser et al.,
1994). Although our effector domain mutant analysis did not im-
plicate Pak as directly responsible for mediating the effect of
activated Rac1 on the disruption of spine morphology, Pak has also
been shown recently to act upstream of Rac (Obermeier et al.,
1998). Pak can also act downstream of Rac1 indirectly via associ-
ation with the p35/Cdk5 complex, whose binding to Rac1L61 is
abolished by both F37A and Y40K mutations (Nikolic et al., 1998).
Dominant-negative C dk5 has been shown to block the effects of
Rac1 in axonogenesis (Ruchhoeft et al., 1999). Taken together,
these findings suggest that the regulation of Rho GTPase-signaling
pathways is important for the generation and maintenance of
neuronal morphology that may eventually contribute to proper
mental function.
REFERENCES
Allen K M, Gleeson JG, Bagrodia S, Partington MW, MacMillan JC,
Cerione RA, Mulley JC, Walsh CA (1998) PAK3 mutation in nonsyn-
dromic X-linked mental retardation. Nat Genet 20:25–30.
Arnold D, Feng L, Kim J, Heintz N (1994) A strategy for the analysis of
gene expression during neural development. Proc Natl Acad Sci USA
91:9970–9974.
Arnold DB, Clapham DE (1999) Molecular determinants for subcellular
localization of PSD-95 with an interacting K
channel. Neuron
23:149–157.
Bernstein M, Lichtman JW (1999) Axonal atrophy: the retraction reaction.
Curr Opin Neurobiol 9:364–370.
Billuart P, Bienvenu T, Ronce N, des Portes V, Vinet MC, Z emni R,
Crollius HR, C arrie A, Fauchereau F, Cherry M, Briault S, Hamel B,
Fryns JP, Beldjord C, Kahn A, Moraine C, Chelly J (1998)
Oligophrenin-1 encodes a rhoGAP protein involved in X-linked mental
retardation. Nature 39:923–926.
Bito H, Furuyashiki T, Ishihara H, Shibasaki Y, Ohashi K, Mizuno K,
Maekawa M, Ishizaki T, Narumiya S (2000) A critical role for a Rho-
associated kinase p160ROCK in determining axon outgrowth in mam-
malian CNS neurons. Neuron 26:431–441.
Boyer C, Schikorski T, Stevens CF (1998) Comparison of hippocampal
dendritic spines in culture and in brain. J Neurosci 18:5294–5300.
Chelly J (1999) Breakthroughs in the molecular and cellular mechanisms
underlying X-linked mental retardation. Hum Mol Genet 8:1833–1838.
Dailey ME, Smith SJ (1996) The dynamics of dendritic structure in devel-
oping hippocampal slices. J Neurosci 16:2983–2994.
Drakew A, Muller M, Gahwiler BH, Thompson SM, Frotscher M (1996)
Spine loss in experimental epilepsy: quantitative light and electron mi-
croscopic analysis of intracellularly stained CA3 pyramidal cells in hip-
pocampal slice cultures. Neuroscience 70:31– 45.
Engert F, Bonhoeffer T (1999) Dendritic spine changes associated with
hippocampal long-term synaptic plasticity. Nature 399:66–70.
Fischer M, Kaech S, Knutti D, Matus A (1998) Rapid actin-based plastic-
ity in dendritic spines. Neuron 20:847–854.
Frantz GD, Bohner AP, Akers RM, McConnell SK (1994) Regulation of
the POU domain gene SCIP during cerebral cortical development.
J Neurosci 14:472–485.
Guo X, Chandrasekaran V, Lein P, Kaplan PL, Higgins D (1999) Leuke-
mia inhibitory factor and ciliary neurotrophic factor cause dendritic
retraction in cultured rat sympathetic neurons. J Neurosci 19:2113–2121.
Hall A (1994) Small GTP-binding proteins and the regulation of the actin
cytoskeleton. Annu Rev Cell Biol 10:31–54.
Harris KM (1999) Structure, development, and plasticity of dendritic
spines. Curr Opin Neurobiol 9:343–348.
Harris KM, Kater SB (1994) Dendritic spines: cellular specializations
imparting both stability and flexibility to synaptic function. Annu Rev
Neurosci 17:341–371.
Hirose M, Ishizaki T, Watanabe N, Uehata M, Kranenburg O, Moolenaar
WH, Matsumura F, Maekawa M, Bito H, Narumiya S (1998) Molecular
dissection of the Rho-associated protein kinase (p160ROCK)-regulated
neurite remodeling in neuroblastoma N1E-115 cells. J Cell Biol
141:1625–1636.
Horch HW, Kruttgen A, Protbury SD, Katz LC (1999) Destabilization of
cortical dendrites and spines by BDNF Neuron 23:353–364.
Ishizaki T, Naito M, Fujisawa K , Maekawa M, Watanabe N, Saito Y,
Narumiya S (1997) p160ROCK, a Rho-associated coiled-coil forming
protein kinase, works downstream of Rho and induces focal adhesions.
FEBS Lett 404:118 –124.
Jalink K, van Corven EJ, Hengeveld T, Morii N, Narumiya S, Moolenaar
WH (1994) Inhibition of lysophosphatidate- and thrombin-induced neu-
rite retraction and neuronal cell rounding by ADP ribosylation of the
small GTP-binding protein Rho. J Cell Biol 126:801–810.
Joneson T, McDonough M, Bar-Sagi D, Van Aelst L (1996) RAC regula-
tion of actin polymerization and proliferation by a pathway distinct from
Jun kinase. Science 274:1374 –1376.
Kimura K, Ito M, Amano M, Chihara K, Fukata Y, Nakafuku M,
Yamamori B, Feng J, Nakano T, Okawa K, Iwamatsu A, Kaibuchi K
(1996) Regulation of myosin phosphatase by Rho and Rho-associated
kinase (Rho-kinase). Science 273:245–248.
Kozma R, Sarner S, Ahmed S, Lim L (1997) Rho family GTPases and
neuronal growth cone remodeling: relationships between increased com-
plexity induced by Cdc42Hs, Rac1, and acetylcholine and collapse in-
duced by RhoA and lysophosphatidic acid. Mol Cell Biol 17:1201–1211.
Lamarche N, Tapon N, Stowers L, Burbelo PD, Aspenstrom P, Bridges T,
Chant J, Hall A (1996) Rac and Cdc42 induce actin polymerization and
G1 cell cycle progression independently of p65
PAK
and the JNK/SAPK
MAP kinase cascade. C ell 87:519 –529.
Lee T, Luo L (1999) Mosaic analysis with a repressible cell marker for
studies of gene function in neuronal morphogenesis. Neuron 22:451– 461.
Lee T, Winter C, Marticke SS, Lee A, L uo L (2000) Essential roles of
Drosophila RhoA in the regulation of neuroblast proliferation and den-
dritic but not axonal morphogenesis. Neuron 25:307–316.
Lein P, Johnson M, Guo X, Rueger D, Higgins D (1995) Osteogenic
protein-1 induces dendritic growth in rat sympathetic neurons. Neuron
15:597–605.
Li Z, Van Aelst LV, Cline HT (2000) Rho GTPases regulate distinct
aspects of dendritic arbor growth in Xenopus central neurons in vivo. Nat
Neurosci 3:217–225.
Liaw CW, Zamoyska R, Parnes JR (1986) Structure, sequence, and poly-
morphism of the Lyt-2 T cell differentiation antigen gene. J Immunol
137:1037–1043.
Lo DC, McAllister AK, Katz LC (1994) Neuronal transfection in brain
slices using particle-mediated gene transfer. Neuron 13:1263–1268.
Luo L, Hensch TK, Ackerman L, Barbel S, Jan LY, Jan YN (1996)
Differential effects of the Rac GTPase on Purkinje cell axons and den-
dritic trunks and spines. Nature 379:837– 840.
Luo L, Jan LY, Jan YN (1997) Rho family small GTP-binding proteins in
growth cone signalling. Curr Opin Neurobiol 7:81–86.
Madaule P, Eda M, Watanabe N, Fujisawa K , Matsuoka T, Bito H, Ishizaki
T, Narumiya S (1998) Role of citron kinase as a target of the small
GTPase Rho in cytokinesis. Nature 394:491– 494.
Maletic-Savatic M, Malinow R, Svoboda K (1999) Rapid dendritic mor-
phogenesis in CA1 hippocampal dendrites induced by synaptic activity.
Science 283:923–927.
Manser E, Leung T, Salihuddin H, Z hao Z S, Lim L (1994) A brain
serine/threonine protein kinase activated by C dc42 and Rac1. Nature
367:4046.
McAllister AK, Lo DC, Katz LC (1995) Neurotrophins regulate dendritic
growth in developing visual cortex. Neuron 15:791–803.
McAllister AK, Katz LC, Lo DC (1997) Opposing roles for endogenous
BDNF and NT-3 in regulating cortical dendritic growth. Neuron
18:767–778.
Mueller BK (1999) Growth cone guidance: first steps towards a deeper
understanding. Annu Rev Neurosci 22:351–388.
Narumiya S (1996) The small GTPase Rho: cellular functions and signal
transductions. J Biochem 120:215–228.
Nakayama et al. Rac and Rho Maintain Dendritic Spines and Branches J. Neurosci., July 15, 2000, 20(14):5329–5338 5337
Nedivi E, Wu GY, Cline HT (1998) Promotion of dendritic growth by
CPG15, an activity-induced signaling molecule. Science 281:1863–1866.
Nikolic M, Chou MM, Lu W, Mayer BJ, Tsai L H (1998) The p35/Cdk5
kinase is a neuron-specific Rac effector that inhibits Pak1 activity. Nature
395:194–198.
Obermeier A, Ahmed S, Manser E, Yen SC, Hall C, Lim L (1998) PAK
promotes morphological changes by acting upstream of Rac. EMBO J
17:43284339.
Okada A, Weimann JM, Fraser SE, McConnell SK (1999) Imaging cells in
the developing nervous system with retrovirus expressing modified green
fluorescent protein. Exp Neurol 156:394 –406.
Purves D, Hadley RD (1985) Changes in the dendritic branching of adult
mammalian neurones revealed by repeated imaging in situ. Nature
315:404406.
Purves D, Hadley RD, Voyvodic JT (1986) Dynamic changes in the den-
dritic geometry of individual neurons visualized over periods of up to
three months in the superior cervical ganglion of living mice. J Neurosci
6:1051–1060.
Purves D, Snider WD, Voyvodic J T (1988) Trophic regulation of nerve
cell morphology and innervation in the autonomic nervous system. Na-
ture 336:123–128.
Qiu RG, Chen J, Kirn D, McCormick F, Symons M (1995a) An essential
role for Rac in Ras transformation. Nature 374:457–459.
Qiu RG, Chen J, Kirn D, McCormick F, Symons M (1995b) A role for
Rho in Ras transformation. Proc Natl Acad Sci USA 92:11781–11785.
Rall W (1964) Theoretical significance of dendritic trees for neuronal
input-output relations. In: Neural theory and modeling (Reiss RF, ed),
pp 63–97. Palo Alto, CA: Stanford UP.
Ridley AJ, Paterson HF, Johnston CL, Diekmann D, Hall A (1992) The
small GTP-binding protein rac regulates growth factor-induced mem-
brane ruffling. Cell 70:401– 410.
Ruchhoeft ML, Ohnuma S, McNeill L, Holt CE, Harris WA (1999) The
neuronal architecture of Xenopus retinal ganglion cells is sculpted by
Rho-family GTPases in vivo. J Neurosci 19:8454 8463.
Sholl DA (1953) Dendritic organization in the neurons of the visual and
motor cortices of the cat. J Anat 87:387– 406.
Snider WD (1988) Nerve growth factor enhances dendritic arborizations
of sympathetic ganglion cells in developing mammals. J Neurosci
8:2628–2634.
Stoppini L, Buchs DA, Muller D (1991) A simple method for organotypic
cultures of nervous tissue. J Neurosci Methods 37:173–182.
Threadgill R, Bobb K , Ghosh A (1997) Regulation of dendritic growth
and remodeling by Rho, Rac, and Cdc42. Neuron 19:625–634.
Toni N, Buchs PA, Nikonenko I, Bron CR, Muller D (1999) LTP promotes
formation of multiple spine synapses between a single axon terminal and
a dendrite. Nature 402:421–424.
Uehata M, Ishizaki T, Satoh H, Ono T, Kawahara T, Morishita T, Tama-
kawa H, Yamagami K, Inui J, Maekawa M, Narumiya S (1997) C alcium
sensitization of smooth muscle mediated by a Rho-associated protein
kinase in hypertension. Nature 389:990–994.
Van Aelst L, D’Souza-Schorey C (1997) Rho GTPases and signaling net-
works. Genes Dev 11:2295–2322.
Van Leeuwen FN, Kain HET, van der Kammen RA, Nichiels F, Kranen-
burg OW, Collard JG (1997) The guanine nucleotide exchange factor
Tiam1 affects neuronal morphology; opposing roles for the small GT-
Pases Rac and Rho. J Cell Biol 139:797–807.
Woolley CS, McEwen BS (1993) Roles of estradiol and progesterone in
regulation of hippocampal dendritic spine density during the estrous
cycle in the rat. J Comp Neurol 336:293–306.
Wright DE, Snider WD (1995) Neurotrophin receptor mRNA expression
defines distinct populations of neurons in rat dorsal root ganglia. J Comp
Neurol 351:329–338.
Yamashita T, Tucker KL, Barde YA (1999) Neurotrophin binding to the
p75 receptor modulates Rho activity and axonal outgrowth. Neuron
24:585–593.
Yuste R, Tank DW (1996) Dendritic integration in mammalian neurons, a
century after Cajal. Neuron 16:701–716.
5338 J. Neurosci., July 15, 2000, 20(14):5329–5338 Nakayama et al. Rac and Rho Maintain Dendritic Spines and Branches
... In most studies published thus far, increased activity of RhoA and inhibition of Rac1 or Cdc42 resulted in significant simplification of the dendritic trees in many neuron types (Threadgill et al. 1997, Nakayama et al. 2000, Hayashi et al. 2002. In contrast, activation of Rac1 or Cdc42 caused an increase in the number of dendrite branches (Threadgill et al. 1997, Nakayama et al. 2000, Hayashi et al. 2002. ...
... In most studies published thus far, increased activity of RhoA and inhibition of Rac1 or Cdc42 resulted in significant simplification of the dendritic trees in many neuron types (Threadgill et al. 1997, Nakayama et al. 2000, Hayashi et al. 2002. In contrast, activation of Rac1 or Cdc42 caused an increase in the number of dendrite branches (Threadgill et al. 1997, Nakayama et al. 2000, Hayashi et al. 2002. ...
... Activation of ROCK has complex cellular consequences, including a myosin contraction of actin filaments and an activation of LIMK. Overexpression of ROCK-CA (ROCK∆3), in rat hippocampal neurons in organotypic culture, caused simplification of dendritic trees (Nakayama et al. 2000). Consequently, inhibition of ROCK by a specific inhibitor, Y-27632, prevented dendritic arbor simplification caused by RhoA-CA overexpression or by suppression of Abl kinase which leads to RhoA activation (Nakayama et al. 2000, Jones et al. 2004). ...
Article
Full-text available
The pattern of dendritic branching along with the receptor and channel composition and density of synapses regulate the electrical properties of neurons. Abnormalities in dendritic tree development lead to serious dysfunction of neuronal circuits and, consequently, the whole nervous system. Not surprisingly, the complicated and multi-step process of dendritic arbor development is highly regulated and controlled at every stage by both extrinsic signals and intrinsic molecular mechanisms. In this review, we analyze the molecular mechanisms that contribute to cellular processes that are crucial for the proper formation and stability of dendritic arbors, in such distant organisms as insects (e.g. Drosophila melanogaster), amphibians (Xenopus laevis), and mammals.
... La identificación de los animales Control y Cdh1 cKO se realizó mediante la técnica de la Reacción en Cadena de la Polimerasa (PCR), amplificando secuencias concretas de varios alelos de Cdh1 y del transgén de Cre. Para ello, se empleó una biopsia de la cola de los ratones, que es el procedimiento de obtención de tejido que se realiza de manera mayoritaria en roedores que están identificados previamente (Naranjo & Pintado, 2017). ...
... Se sabe que Rock modifica el número, la morfología y la dendritas en varios tipos de neuronas. La sobreactivación de la señalización de Rock estabiliza el citoesqueleto de la actina y, con ello, previene la plasticidad sináptica, mientras que la inhibición de Rock promueve el ensamblaje de microtúbulos y restaura la complejidad y plasticidad de las dendritas(Chen & Firestein, 2007;Nakayama et al., 1995).La ausencia de Cdh1 altera la arborización de las dendritas y provoca la pérdida de espinas dendríticas y sinapsis en la corteza y el hipocampo, que, a su vez, produce deterioro cognitivo y neurodegeneración en ratones adultos. Rock2 es un sustrato de APC/C-Cdh1, y ante la ausencia del cofactor, la actividad de Rock2 aumenta en la corteza y el hipocampo. ...
Thesis
The Anaphase Promoting Complex/Cyclosome (APC/C), together with its coactivator Cdh1, plays an essential role in brain development, where it coordinates the exit of neural progenitors from the cell cycle and the initiation of neurogenesis and, with it, the size of the brain and organization of the cerebral cortex. Recently, we have identified a novel Cdh1 mutation (Asp187Gly) that causes loss of APC/C-Cdh1 activity and causes microcephaly, psychomotor retardation and epilepsy in humans. These disorders are associated with alterations in neurogenesis, oligodendrogenesis and myelination. Moreover, a preliminary imaging examination of the brain of the patient carrying the Cdh1 mutation showed decreased white matter and dysgenesis of the corpus callosum. Therefore, APC/C-C-Cdh1 could play a key role in oligodendrogenesis and myelination. In order to study the role of the APC/C-Cdh1 complex during postnatal development, an in vivo experimental model was established with the absence of Cdh1 conditionally to the Nestin promoter, thus generating Cdh1-deficient knockout mice (Cdh1 cKO) at late stages of embryonic development. Animals lacking Cdh1 showed brain morphological alterations, including hydrocephalus and severe ventriculomegaly, promoting the displacement of structures important for cognitive processes, such as the cortex, hippocampus and corpus callosum. The absence of Cdh1 delayed myelination in both the spinal cord and brain, demonstrating that APC/C activation is required for myelin formation during postnatal development. Moreover, analysis of myelin ultrastructure revealed a higher percentage of unmyelinated axons and increased g-ratio in Cdh1 cKO mice. In addition, myelin from Cdh1-deficient animals showed alterations and breaks characteristic of decompaction. These results confirm that APC/C-Cdh1 is essential not only for myelin formation, but also for its proper packaging. Untargeted metabolomic studies demonstrated that deletion of Cdh1 negatively regulates brain lipid composition, especially in the main components of the myelin sheath, such as galactosylceramides and sphingolipids. In conclusion, the APC/C-Cdh1 complex plays a key role in the coordination of myelination and oligodendrogenesis during postnatal development, which places Cdh1 in the etiopathogenesis of neurodevelopmental disorders involving hypomyelination and dysgenesis of the corpus callosum.
... dynamics (26), leading to dendrite elaboration (27) and dendritic spine elongation (28) in rat hippocampal neurons through inactivation of the RhoA-Rho-associated protein kinase (ROCK) pathway (26,(28)(29)(30)(31)(32)(33)(34)(35)(36). Previously, Rnd1 has been characterized to associate with the H/RBD region of plexin-A1, which contains the conserved amino acid motif lysine-valine-serine (LVS) (23,37,38), and plexin-B1 (19,37,39); its association with plexin-A1 induces axonal repulsion (15). ...
... Rnd1 promotes activity-dependent dendrite elaboration (27) and dendritic spine elongation in rat hippocampal neurons (28) by decreasing RhoA activity and its downstream effector, ROCK (26,(28)(29)(30)(31)(32)(33)(34)(35)(36). Therefore, we examined whether inactivation of RhoA/ROCK signaling by Rnd1 is required for dendritic elaboration in response to Sema3A. ...
Article
Full-text available
The precise development of neuronal morphologies is crucial to the establishment of synaptic circuits and, ultimately, proper brain function. Signaling by the axon guidance cue semaphorin 3A (Sema3A) and its receptor complex of neuropilin-1 and plexin-A4 has multifunctional outcomes in neuronal morphogenesis. Downstream activation of the RhoGEF FARP2 through interaction with the lysine-arginine-lysine motif of plexin-A4 and consequent activation of the small GTPase Rac1 promotes dendrite arborization, but this pathway is dispensable for axon repulsion. Here, we investigated the interplay of small GTPase signaling mechanisms underlying Sema3A-mediated dendritic elaboration in mouse layer V cortical neurons in vitro and in vivo. Sema3A promoted the binding of the small GTPase Rnd1 to the amino acid motif lysine-valine-serine (LVS) in the cytoplasmic domain of plexin-A4. Rnd1 inhibited the activity of the small GTPase RhoA and the kinase ROCK, thus supporting the activity of the GTPase Rac1, which permitted the growth and branching of dendrites. Overexpression of a dominant-negative RhoA, a constitutively active Rac1, or the pharmacological inhibition of ROCK activity rescued defects in dendritic elaboration in neurons expressing a plexin-A4 mutant lacking the LVS motif. Our findings provide insights into the previously unappreciated balancing act between Rho and Rac signaling downstream of specific motifs in plexin-A4 to mediate Sema3A-dependent dendritic elaboration in mammalian cortical neuron development.
... TRIO encodes a large cytoskeletal regulatory protein with two guanine nucleotide exchange factor (GEF) domains for Rho family GTPases -GEF1 activates Rac1 and RhoG, while GEF2 activates RhoA (35-37). As a dual RhoGEF, TRIO integrates signals downstream of cell surface receptors and acts on GTPases to coordinate cytoskeletal rearrangements critical for proper neurodevelopment (16,28,(38)(39)(40)(41)(42)(43)(44). Interestingly, nonsense variants spread throughout TRIO are enriched in individuals with SCZ (34, 45), whereas pathogenic missense TRIO variants in or surrounding the GEF1 domain are associated with ASD/ID (9, 32, 33). ...
... Based on the effects of Rac1 and RhoA signaling on the neuronal actin cytoskeleton(13,(39)(40)(41)(42), we hypothesized that Trio K1431M, K1918X, and M2145T variants would significantly impact dendritic arbor branching and length. Instead, the Trio variants had relatively subtle effects on cortical L5 PN dendrites. ...
Preprint
Full-text available
Heterozygosity for rare genetic variants in TRIO is associated with neurodevelopmental disorders (NDDs) including schizophrenia (SCZ), autism spectrum disorder (ASD) and intellectual disability. TRIO uses its two guanine nucleotide exchange factor (GEF) domains to activate GTPases (GEF1: Rac1 and RhoG; GEF2: RhoA) that control neuronal migration, synapse development and function. It remains unclear whether and how discrete TRIO variants differentially impact these neurodevelopmental events. Here, we elucidate how heterozygosity for NDD-associated Trio variants – +/K1431M (ASD), +/K1918X (SCZ), and +/M2145T (bipolar disorder, BPD) – impact mouse behavior, brain development, and synapse structure and function. Heterozygosity for different Trio variants impacts motor, social, and cognitive behaviors in distinct ways that align with clinical phenotypes in humans. ASD- and SCZ-linked Trio variants differentially impact head and brain size with corresponding changes in dendritic arbors of motor cortex layer 5 pyramidal neurons (M1 L5 PNs). Although dendritic spine density and synaptic ultrastructure were only modestly altered in the Trio variant heterozygotes, we observe significant changes in synaptic function and plasticity including excitatory/inhibitory imbalance and long-term potentiation defects. We also identify distinct changes in glutamate synaptic release in +/K1431M and +/M2145T cortico-cortical synapses, associated with deficiencies in crucial presynaptic release regulators. While TRIO K1431M has impaired ability to promote GTP exchange on Rac1, +/K1431M mice exhibit increased Rac1 activity, suggesting possible compensation by other GEFs. Our work reveals that discrete disease-associated Trio variants yield overlapping but distinct NDD-associated phenotypes in mice and demonstrates, for the first time, an essential role for Trio in presynaptic glutamate release, underscoring the importance of studying the impact of variant heterozygosity in vivo.
... Surprisingly, among the 73 causative variants of ARID identified by Najmabadi et al 31 ., 10 are direct targets of PRMT9-regulated splicing, and additional 28 have paralogs that are regulated by PRMT9 (Supplementary Fig. 6f). Although ARID is highly heterogeneous, we found that 27 of the 38 protein components of the Ras/Rho/PSD95 network 31,43 are splicing targets of PRMT9 (Supplementary Fig. 6g) and that PRMT9-regulated alternatively spliced genes are more enriched in postsynapse components ( Supplementary Fig. 6h), suggesting that PRMT9 might regulate synaptic functions through Ras and Rho GTPase signaling pathway, which includes critical factors controlling dendritic spine morphogenesis and neural circuit connectivity [44][45][46][47] . ...
Article
Full-text available
Protein arginine methyltransferase 9 (PRMT9) is a recently identified member of the PRMT family, yet its biological function remains largely unknown. Here, by characterizing an intellectual disability associated PRMT9 mutation (G189R) and establishing a Prmt9 conditional knockout (cKO) mouse model, we uncover an important function of PRMT9 in neuronal development. The G189R mutation abolishes PRMT9 methyltransferase activity and reduces its protein stability. Knockout of Prmt9 in hippocampal neurons causes alternative splicing of ~1900 genes, which likely accounts for the aberrant synapse development and impaired learning and memory in the Prmt9 cKO mice. Mechanistically, we discover a methylation-sensitive protein–RNA interaction between the arginine 508 (R508) of the splicing factor 3B subunit 2 (SF3B2), the site that is exclusively methylated by PRMT9, and the pre-mRNA anchoring site, a cis-regulatory element that is critical for RNA splicing. Additionally, using human and mouse cell lines, as well as an SF3B2 arginine methylation-deficient mouse model, we provide strong evidence that SF3B2 is the primary methylation substrate of PRMT9, thus highlighting the conserved function of the PRMT9/SF3B2 axis in regulating pre-mRNA splicing.
... Sholl analysis of primary rat hippocampal neurons Sholl analysis was performed employing the modified method as described previously (Nakayama et al., 2000). Individual neurons were imaged using a 200× objective, employing confocal microscopy, and the captured images were printed. ...
Article
Full-text available
ARL6IP1 is implicated in hereditary spastic paraplegia (HSP), but the specific pathogenic mechanism leading to neurodegeneration has not been elucidated. Here, we clarified the molecular mechanism of ARL6IP1 in HSP using in vitro and in vivo models. The Arl6ip1 knockout (KO) mouse model was generated to represent the clinically involved frameshift mutations and mimicked the HSP phenotypes. Notably, in vivo brain histopathological analysis revealed demyelination of the axon and neuroinflammation in the white matter, including the corticospinal tract. In in vitro experiments, ARL6IP1 silencing caused cell death during neuronal differentiation and mitochondrial dysfunction by dysregulated autophagy. ARL6IP1 localized on mitochondria-associated membranes (MAMs) to maintain endoplasmic reticulum and mitochondrial homeostasis via direct interaction with LC3B and BCl2L13. ARL6IP1 played a crucial role in connecting the endoplasmic reticulum and mitochondria as a member of MAMs. ARL6IP1 gene therapy reduced HSP phenotypes and restored pathophysiological changes in the Arl6ip1 KO model. Our results established that ARL6IP1 could be a potential target for HSP gene therapy.
Article
Full-text available
From fly to man, the Wingless (Wg)/Wnt signaling molecule is essential for both the stability and plasticity of the nervous system. The Drosophila neuromuscular junction (NMJ) has proven to be a useful system for deciphering the role of Wg in directing activity-dependent synaptic plasticity (ADSP), which, in the motoneuron, has been shown to be dependent on both the canonical and the noncanonical calcium Wg pathways. Here we show that the noncanonical planar cell polarity (PCP) pathway is an essential component of the Wg signaling system controlling plasticity at the motoneuron synapse. We present evidence that disturbing the PCP pathway leads to a perturbation in ADSP. We first show that a PCP-specific allele of disheveled (dsh) affects the de novo synaptic structures produced during ADSP. We then show that the Rho GTPases downstream of Dsh in the PCP pathway are also involved in regulating the morphological changes that take place after repeated stimulation. Finally, we show that Jun kinase is essential for this phenomenon, whereas we found no indication of the involvement of the transcription factor complex AP1 (Jun/Fos). This work shows the involvement of the neuronal PCP signaling pathway in supporting ADSP. Because we find that AP1 mutants can perform ADSP adequately, we hypothesize that, upon Wg activation, the Rho GTPases and Jun kinase are involved locally at the synapse, in instructing cytoskeletal dynamics responsible for the appearance of the morphological changes occurring during ADSP.
Article
The postsynaptic density (PSD) is a collection of specialized proteins assembled beneath the postsynaptic membrane of dendritic spines. The PSD proteome comprises ~1000 proteins, including neurotransmitter receptors, scaffolding proteins and signalling enzymes. Many of these proteins have essential roles in synaptic function and plasticity. During brain development, changes are observed in synapse density and in the stability and shape of spines, reflecting the underlying molecular maturation of synapses. Synaptic protein composition changes in terms of protein abundance and the assembly of protein complexes, supercomplexes and the physical organization of the PSD. Here, we summarize the developmental alterations of postsynaptic protein composition during synapse maturation. We describe major PSD proteins involved in postsynaptic signalling that regulates synaptic plasticity and discuss the effect of altered expression of these proteins during development. We consider the abnormality of synaptic profiles and synaptic protein composition in the brain in neurodevelopmental disorders such as autism spectrum disorders. We also explain differences in synapse development between rodents and primates in terms of synaptic profiles and protein composition. Finally, we introduce recent findings related to synaptic diversity and nanoarchitecture and discuss their impact on future research. Synaptic protein composition can be considered a major determinant and marker of synapse maturation in normality and disease.
Preprint
Full-text available
One of the most extensively studied members of the Ras superfamily of small GTPases, Rac1 is an intracellular signal transducer that remodels actin and phosphorylation signaling networks. Previous studies have shown that Rac1-mediated signaling is associated with hippocampal-dependent working memory and longer-term forms of learning and memory and that Rac1 can modulate forms of both pre- and postsynaptic plasticity. How these different cognitive functions and forms of plasticity mediated by Rac1 are linked, however, is unclear. Here, we show that spatial working memory is selectively impaired following the expression of a genetically encoded Rac1-inhibitor at presynaptic terminals, while longer-term cognitive processes are affected by Rac1 inhibition at postsynaptic sites. To investigate the regulatory mechanisms of this presynaptic process, we leveraged new advances in mass spectrometry to identify the proteomic and post-translational landscape of presynaptic Rac1 signaling. We identified serine/threonine kinases and phosphorylated cytoskeletal signaling and synaptic vesicle proteins enriched with active Rac1. The phosphorylated sites in these proteins are at positions likely to have regulatory effects on synaptic vesicles. Consistent with this, we also report changes in the distribution and morphology of synaptic vesicles and in postsynaptic ultrastructure following presynaptic Rac1 inhibition. Overall, this study reveals a previously unrecognized presynaptic role of Rac1 signaling in cognitive processes and provides insights into its potential regulatory mechanisms.
Article
Single nucleotide variants in the general population are common genomic alterations, where the majority are presumed to be silent polymorphisms without known clinical significance. Using human induced pluripotent stem cell (hiPSC) cerebral organoid modeling of the 1.4 megabase Neurofibromatosis type 1 (NF1) deletion syndrome, we previously discovered that the cytokine receptor-like factor-3 (CRLF3) gene, which is co-deleted with the NF1 gene, functions as a major regulator of neuronal maturation. Moreover, children with NF1 and the CRLF3L389P variant have greater autism burden, suggesting that this gene might be important for neurologic function. To explore the functional consequences of this variant, we generated CRLF3L389P-mutant hiPSC lines and Crlf3L389P-mutant genetically engineered mice. While this variant does not impair protein expression, brain structure, or mouse behavior, CRLF3L389P-mutant human cerebral organoids and mouse brains exhibit impaired neuronal maturation and dendrite formation. In addition, Crlf3L389P-mutant mouse neurons have reduced dendrite lengths and branching, without any axonal deficits. Moreover, Crlf3L389P-mutant mouse hippocampal neurons have decreased firing rates and synaptic current amplitudes relative to wild type controls. Taken together, these findings establish the CRLF3L389P variant as functionally deleterious and suggest that it may be a neurodevelopmental disease modifier.
Article
Full-text available
Recent observations have suggested that the dendritic arbors of sympathetic ganglion cells may be regulated by interactions with their peripheral targets (Voyvodic, 1987a; Yawo, 1987). In order to assess a potential mechanism for such interactions, I have investigated the effects of the target-derived trophic molecule for sympathetic ganglion cells on the development of dendrites in the rat superior cervical ganglion. Systemic treatment of neonatal animals with NGF for 1 or 2 weeks results in a striking expansion of ganglion cell dendritic arbors, as revealed by intracellular staining with HRP. During this period, neurons in treated animals extend new primary dendrites, and the length and branching of existing dendrites are increased compared to age-matched controls. These results support the idea that targets may regulate ganglion cell arbors via elaboration of NGF, and suggest an explanation for the correlation between animal size and dendritic complexity noted in several recent studies (Purves and Lichtman, 1985a; Snider, 1987; Voyvodic, 1987a).
Article
Full-text available
We describe a means of visualizing the same neuron in the superior cervical ganglion of young adult mice over intervals of up to 3 months. The dendrites of these neurons change during this interval; some branches retract, others elongate, and still others appear to form de novo. Thus, neuronal dendrites in this part of the nervous system are subject to continual change beyond what is usually considered the developmental period. The remodeling of postsynaptic processes further implies that the synaptic connections made onto these cells undergo substantial rearrangement well into adulthood.
Article
Activity shapes the structure of neurons and their circuits. Two-photon imaging of CA1 neurons expressing enhanced green fluorescent protein in developing hippocampal slices from rat brains was used to characterize dendritic morphogenesis in response to synaptic activity. High-frequency focal synaptic stimulation induced a period (longer than 30 minutes) of enhanced growth of small filopodia-like protrusions (typically less than 5 micrometers long). Synaptically evoked growth was long-lasting and localized to dendritic regions close (less than 50 micrometers) to the stimulating electrode and was prevented by blockade of N-methyl-D-aspartate receptors. Thus, synaptic activation can produce rapid input-specific changes in dendritic structure. Such persistent structural changes could contribute to the development of neural circuitry.
Article
After getting top billing at a recent astronomy meeting, speculative proposals that dark matter might lurk in unseen "ghost galaxies" or dim stars made headlines. But the media splash had little to do with the likely scientific impact of the two modest but interesting studies; the few astronomers who are familiar with the work suggest that the teams' broader conclusions were overstated.
Article
The function of rac, a ras-related GTP-binding protein, was investigated in fibroblasts by microinjection. In confluent serum-starved Swiss 3T3 cells, rac1 rapidly stimulated actin filament accumulation at the plasma membrane, forming membrane ruffles. Several growth factors and activated H-ras also induced membrane ruffling, and this response was prevented by a dominant inhibitory mutant rac protein, N17rac1. This suggests that endogenous rac proteins are required for growth factor-induced membrane ruffling. In addition to membrane ruffling, a later response to both rac1 microinjection and some growth factors was the formation of actin stress fibers, a process requiring endogenous rho proteins. Using N17rac1 we have shown that these growth factors act through rac to stimulate this rho-dependent response. We propose that rac and rho are essential components of signal transduction pathways linking growth factors to the organization of polymerized actin.
Article
Hippocampal slices prepared from 2-23-day-old neonates were maintained in culture at the interface between air and a culture medium. They were placed on a sterile, transparent and porous membrane and kept in petri dishes in an incubator. No plasma clot or roller drum were used. This method yields thin slices which remain 1-4 cell layers thick and are characterized by a well preserved organotypic organization. Pyramidal neurons labelled by extra- and intracellular application of horse radish peroxidase resemble by the organization and complexity of their dendritic processes those observed in situ at a comparable developmental stage. Excitatory and inhibitory synaptic potentials can easily be analysed using extra- or intracellular recording techniques. After a few days in culture, long-term potentiation of synaptic responses can reproducibly be induced. Evidence for a sprouting response during the first days in culture or following sections is illustrated. This technique may represent an interesting alternative to roller tube cultures for studies of the developmental changes occurring during the first days or weeks in culture.
Article
A remarkable feature of nerve cells is the complex and variable pattern of their axonal and dendritic branches. Quantitative studies of a simple part of the nervous system in mammals provide evidence that neuronal geometry and innervation are regulated by long-term trophic interactions between neurons and their targets. This trophic linkage may explain how nerve cells adjust their function to the needs of bodies that vary markedly in size and form.
Article
We have determined the nucleotide sequence and genomic organization of the mouse Lyt-2 T lymphocyte differentiation antigen gene. This gene consists of five exons and four introns, and the organization roughly parallels the protein domains. Alternative splicing to include or exclude exon IV (encoding part of the cytoplasmic tail) results in two forms of mRNA and accounts for the difference in size between the alpha- and alpha'-chains of Lyt-2. The gene structure provides further evidence for the evolutionary relationship between Lyt-2 and immunoglobulin genes. Comparison of the nucleotide sequence of the Lyt-2.1 and Lyt-2.2 alleles shows a high degree of conservation, but indicates that a single nucleotide change and consequent amino acid substitution in the variable region-like domain accounts for the serologic difference between these two alleles.
Article
A major obstacle to understanding the mechanism of long-term change in the vertebrate nervous system has been the inability to observe the same nerve cell at different times during the life of an animal. The possibility that changes in neural connectivity underlie the remarkable flexibility of the nervous systems of mammals has therefore not been tested by direct observation. Here, we report studies in which we have visualized the same neurone in the superior cervical ganglion of young adult mice at intervals of up to 33 days. This collection of nerve cells is particularly accessible and therefore well suited to our approach. We find that the dendritic branches of the neurones examined change appreciably over intervals of 2 weeks or more; some branches retract, others elongate and others seem to form de novo. The apparent remodelling of these postsynaptic elements implies that the synaptic connections of these cells normally undergo significant rearrangement beyond what is usually considered to be the developmental period.