ArticlePDF Available

Functional Requirement for Class I MHC in CNS Development and Plasticity

Authors:

Abstract and Figures

Class I major histocompatibility complex (class I MHC) molecules, known to be important for immune responses to antigen, are expressed also by neurons that undergo activity-dependent, long-term structural and synaptic modifications. Here, we show that in mice genetically deficient for cell surface class I MHC or for a class I MHC receptor component, CD3ζ, refinement of connections between retina and central targets during development is incomplete. In the hippocampus of adult mutants,N-methyl-d-aspartate receptor–dependent long-term potentiation (LTP) is enhanced, and long-term depression (LTD) is absent. Specific class I MHC messenger RNAs are expressed by distinct mosaics of neurons, reflecting a potential for diverse neuronal functions. These results demonstrate an important role for these molecules in the activity-dependent remodeling and plasticity of connections in the developing and mature mammalian central nervous system (CNS).
Abnormal retinogeniculate projections but normal dLGN ultrastructure in mice deficient in class I MHC signaling. At P12, one eye was injected with WGA-HRP ( 23 ); after 1 day, anterograde axonal transport results in labeling of the entire retinal projection to the LGN. Labeling pattern in the dLGN is shown in bright-field optics (label is black) or as dark-field composites [label is white; see ( 24 )]. ( A ) Representative projection from retina to dLGN contralateral (dashed lines; coronal section; dorsal is up; lateral is left) or ipsilateral to eye injected with WGA-HRP (asterisks indicate labeled area from ipsilateral eye: lateral is to right) in a P13 ␤ 2 M ϩ / ϩ wild-type mouse and a ␤ 2 M –/– mutant mouse. ( B and C ) Representative (B) and extreme (C) examples of the projection from the ipsilateral eye observed in ␤ 2 M Ϫ / Ϫ TAP1 Ϫ / Ϫ mice. ( D and E ) Representative (D) and extreme (E) examples of the projection in CD3 ␨ –/– mice. Arrowheads indicate ectopic projections, which appear extensive under the more sensitive dark-field optics. Scale bar, 200 ␮ m. ( F ) Graph of areas ( Ϯ SEM) occupied by the ipsilateral retinal projection to the LGN for ␤ 2 M ϩ / ϩ (wild-type), ␤ 2 M –/– , ␤ 2 M –/– TAP1 –/– , and CD3 ␨ Ϫ / Ϫ mice ( 24 ), normalized to total dLGN area. The ipsilateral projection area in ␤ 2 M ϩ / ϩ animals is set as 100% (horizontal dashed line). Asterisks indicate significant differences from ␤ 2 M ϩ / ϩ mice ( P Ͻ 0.05, Student’s two-tailed t test). ( G ), Electron micrograph of the dLGN from a ␤ 2 M –/– TAP1 –/– mouse (at P24), showing a typical R-type synaptic bouton (R) making contacts with a dendrite (d). A well-myelinated axon (ax) is also present in this field. Scale bar, 1 ␮ m.
… 
Content may be subject to copyright.
Functional Requirement for
Class I MHC in CNS
Development and Plasticity
Gene S. Huh,* Lisa M. Boulanger, Hongping Du, Patricio A. Riquelme,
Tilmann M. Brotz, Carla J. Shatz*
Class I major histocompatibility complex (class I MHC) molecules, known to be
important for immune responses to antigen, are expressed also by neurons that
undergo activity-dependent, long-term structural and synaptic modifications.
Here, we show that in mice genetically deficient for cell surface class I MHC or
for a class I MHC receptor component, CD3, refinement of connections be-
tween retina and central targets during development is incomplete. In the
hippocampus of adult mutants, N-methyl-
D-aspartate receptor–dependent
long-term potentiation (LTP) is enhanced, and long-term depression (LTD) is
absent. Specific class I MHC messenger RNAs are expressed by distinct mosaics
of neurons, reflecting a potential for diverse neuronal functions. These results
demonstrate an important role for these molecules in the activity-dependent
remodeling and plasticity of connections in the developing and mature mam-
malian central nervous system (CNS).
The development of precise connections in
the CNS is critically dependent on neural
activity, which drives the elimination of in-
appropriate connections and the stabilization
of appropriate ones. In the visual system of
higher mammals, the refinement of initially
imprecise axonal connections requires spon-
taneously generated activity early in develop-
ment and visually driven activity later (14).
Fine-tuning of neural connectivity is thought to
result from changes in synaptic strength, driven
by patterned impulse activity (1, 2, 5, 6).
To identify molecules critical for activity-
dependent structural remodeling, we previ-
ously conducted an unbiased screen for
mRNAs selectively regulated by blocking
spontaneously generated activity in the devel-
oping cat visual system. This manipulation
prevents the remodeling of retinal axons from
each eye into layers within the lateral genic-
ulate nucleus (LGN) (79). Although many
known neural genes were not detectably reg-
ulated by activity blockade, this screen re-
vealed to our surprise that members of the
class I MHC protein family are expressed by
neurons and are regulated by spontaneous
and evoked neural activity (10). Neuronal
class I MHC expression corresponds to well-
characterized times and regions of activity-
dependent development and plasticity of
CNS connections, including retina, LGN, and
hippocampus. Furthermore, the mRNA for
CD3[a class I MHC receptor subunit in the
immune system (11)] is also expressed by
neurons (10), consistent with its interaction
with class I MHC during activity-dependent
remodeling and plasticity. Although class I
MHC is primarily known for its function in
cell-mediated immune recognition, the above
findings from our differential screen suggest
that class I MHC molecules may play roles in
structural and synaptic remodeling in the de-
veloping and mature CNS.
To explore these possibilities by genetic
means, we first confirmed by in situ hybrid-
ization that class I MHC and CD3 were
expressed in the developing mouse CNS. Be-
cause numerous class I MHC genes exist in
the mouse genome (12), we used a pan-
specific cDNA probe expected to detect
many class I MHC molecules (13). This
probe detected elevated amounts of mRNAs
in the dorsal LGN (dLGN) during the first
two postnatal weeks, exactly when ganglion
cell axons sort into eye-specific layers in the
mouse (14); mRNA levels declined at later
ages (Fig. 1A, compare postnatal days P6 and
P40). Expression was also evident in the gan-
glion cell layer of the retina (Fig. 1A, P6 eye),
in neocortex (in layer 4 at early ages and in
deeper layers later; Fig. 1A), and in granule
and pyramidal cell layers of the hippocam-
pus (Fig. 1A, P40, and Fig. 2). CD3
mRNA, like that of class I MHC, was
expressed in the mouse dLGN during the
first two postnatal weeks (Fig. 1B); expres-
sion appeared higher medially. CD3
mRNA was also detected in small amounts
in P40 hippocampus (15). Therefore, as in
cat (10), class I MHC and CD3 transcripts
are present in the developing murine CNS at
locations and times consistent with a role for
these molecules in activity-dependent struc-
tural remodeling and synaptic plasticity.
Strikingly, different class I MHC genes
are expressed in unique subsets of neurons
throughout the mature CNS, as revealed by
using probes (13) that react more specifically
with each of two class Ia (H–2D, H–2K)or
two class Ib MHC genes (Qa-1, T22). For
example, within the somatosensory cortex,
H–2D probe signal was distributed through
many layers but was strongest in layer 4;
Qa-1 signal was specific to layer 6, and T22
signal was evident in both layers 5 and 6 (Fig.
2). H–2D and T22 signals were both strong in
the pyramidal layers of the hippocampus and
in the habenula; in contrast, that of Qa-1 was
weak in those locations (Fig. 2). Transcripts
detected by the T22 probe were particularly
abundant in the thalamic reticular nucleus,
globus pallidus, and substantia nigra [Fig. 2
and (15)]. H–2K signal paralleled that of
H–2D but was much lower throughout the
brain (16). The distinct expression patterns
detected by these probes extended prior in-
ferences from RNase protection experiments
in cat (10) and demonstrated conclusively
that several class I MHC mRNA subtypes are
differentially expressed by distinct subsets of
neurons in the CNS. These findings suggest a
potential for functional diversity among class
Ia and Ib genes within the CNS. Such heter-
ogeneity of function occurs among these
genes within the immune system (17).
To test directly our hypothesis that class I
MHC is required for activity-driven structural
remodeling and synaptic plasticity, mice de-
ficient either for cell surface class I MHC
expression or for CD3 were analyzed. Be-
cause numerous class I MHC genes may be
expressed by specific subsets of neurons (Fig.
2), we examined mice lacking two molecules
required for the stable cell-surface expression
of nearly all fully assembled class I MHC
molecules:
2
-microglobulin [
2
M, a class I
MHC cosubunit (18)], and TAP1 [a compo-
nent of the transporter that supplies peptides
to class I MHC enroute to the cell surface (19,
20)].
2
-M is expressed by neurons in LGN,
cortex, and hippocampus (10) and, as in non-
neuronal cells, induction of class I MHC on
the cell surface of neurons requires expres-
sion of
2
M and TAP1 mRNAs (21). In
addition, to examine whether CD3-contain-
ing receptors were involved in class I MHC–
mediated signaling in the CNS, we analyzed
mice lacking CD3 (22). When raised in a
germ-free facility, all mutant mice are out-
wardly normal and are not obviously differ-
ent from wild-type mice in weight, body
length, appearance, or behavior.
We hypothesized that mice deficient in
class I MHC–mediated signaling might have
abnormal patterns of retinogeniculate projec-
tions because blockade of neural activity si-
multaneously prevents the segregation of ret-
inal ganglion cell axons into eye-specific lay-
Department of Neurobiology, Harvard Medical
School, 220 Longwood Avenue, Boston, MA 02115,
USA.
*To whom correspondence may be addressed. E-mail:
gshuh@alum.mit.edu or carla_shatz@hms.harvard.edu
Present address: Experimental Immunology Branch,
National Cancer Institute, National Institutes of
Health, Building 10, Room 4B36, Bethesda, MD
20892, USA.
R EPORTS
www.sciencemag.org SCIENCE VOL 290 15 DECEMBER 2000 2155
ers and reduces class I MHC expression in
the LGN (710). The normal adult mouse
dLGN has a small layer that receives inputs
from ganglion cells in the ipsilateral eye;
inputs from the contralateral eye occupy the
remainder of the dLGN (Fig. 3A). The refine-
ment of these eye-specific connections in the
mouse occurs between postnatal day 4 (P4)
and P8 (14). We therefore examined the dis-
tribution of retinal inputs at P13, 5 days after
segregation was complete, using the antero-
grade transport of horseradish peroxidase–
conjugated wheat germ agglutinin (WGA-
HRP) injected into one eye (23). Compared
with wild-type animals (Fig. 3, A and F,
2
M
/
), the pattern of the retinogeniculate
projection was significantly altered in all
three mutant genotypes tested. This point is
best appreciated by inspecting the size and
shape of the ipsilateral retinal projection to
the dLGN (Figs. 3, A to E). Although all
mutants still form an ipsilateral patch located
approximately normally in the mediodorsal
dLGN, the area of this patch was significantly
larger in mutant mice and, in extreme cases,
was accompanied by multiple ectopic clusters
of inputs that were never observed in wild-
type mice (Fig. 3, C and E, arrowheads).
These ectopic clusters appeared in medial
areas of the dLGN, where the highest levels
of CD3 mRNA are normally present (com-
pare Fig. 3, C and E, with Fig. 1A). In these
extreme cases, ectopic clusters were also ob-
served in the ipsilateral superior colliculus,
another retinorecipient target that expresses
low-to-moderate levels of class I MHC
mRNA in mouse (15).
To assess quantitatively the altered retino-
geniculate projection in mutant mice, com-
puterized image analysis was used to measure
the fraction of dLGN area occupied by the
ipsilateral projection. All image analyses
were carried out by an observer blind to
genotype (24 ). In all mutant genotypes, there
was a significant increase in area occupied by
the ipsilateral projection over that of wild-
type controls [Fig. 3F:
2
M
–/–
, 130.3 7.3%
(n 10);
2
M
–/–
TAP1
–/–
, 133.3 5.7%
(n 13); CD3
–/–
, 122.7 4.2% (n 13);
wild-type
2
M
/
, 100.0 9.1% (n 12);
P 0.05, Student’s two-tailed t-test]. These
observations support the hypothesis that class
I MHC function is required for the develop-
mental refinement of the retinal projections
and the formation of precise eye-specific re-
gions in the LGN.
Although the refinement of retinogenicu-
late axons was abnormal in mutant mice,
many other aspects of LGN development ap-
pear to proceed normally. The histological
appearance, size, shape, and location of
the dLGN and thalamus, as viewed in
Nissl-stained sections, were indistinguishable
among all experimental groups (15). The
bulk of the ipsilateral projection was posi-
tioned, as expected, in the binocular region of
the dLGN. At the ultrastructural level, the
synaptic organization of the LGN in
2
M
–/–
TAP1
–/–
mice appeared qualitatively indistin-
Fig. 1. Class I MHC expression in mouse CNS. (A) Expression of class I MHC transcripts in coronal
sections of the mouse CNS at P6 and P40 and in a cross section of P6 eye (13). Left column,
adjacent Nissl-stained section; middle column, hybridization with antisense riboprobe under
dark-field optics; right column, hybridization with control sense probe. D, dorsal; L, lateral; hc,
hippocampus; ctx, neocortex; gcl, ganglion cell layer. Arrowheads and dashed lines indicate dLGN.
Scale bar for P6 and P40 brains, 0.5 mm; scale bar for P6 eye, 250 m. (B) Expression of CD3 in
the dLGN during eye-specific layer formation. Upper panel, adjacent Nissl-stained coronal section
of P6 mouse brain (arrowhead, dLGN). Middle panel, hybridization with CD3 antisense probe
(dashed lines, dLGN); hybridization is also present in the ventroposterior nucleus of thalamus (down and
to right of dLGN). Lower panel (cptr), excess of unlabeled competitor probe. Scale bar, 200 m.
Fig. 2. Expression of multiple class I MHC sub-
classes in distinct regions of the mature CNS.
Coronal sections of P40 mouse brain analyzed
by in situ hybridization, using subclass-specific
probes indicated at top of each panel (13). S1,
somatosensory cortex; hb, habenula; hc, hip-
pocampus; rs, retrosplenial cortex; tr, thalamic
reticular nucleus; gp, globus pallidus. Numerals
(4, 6, 56) indicate neocortical layers. Scale
bar, 1 mm.
R EPORTS
15 DECEMBER 2000 VOL 290 SCIENCE www.sciencemag.org2156
guishable from that of wild type (23). Reti-
nogeniculate axons were well-myelinated,
and glomeruli and R-type synaptic boutons
[hallmarks of retinogeniculate synapses; (25
27)] were present, indicating that normal ret-
inal synapses do form in the LGN (Fig. 3G).
These observations suggest that many activ-
ity-independent processes (1, 2, 28, 29) are
not perturbed in mice with abnormal class I
MHC function.
Because similar abnormalities in the ipsi-
lateral projection result from blockade of
spontaneous activity at comparable ages in
the cat or ferret visual system (79), we tested
whether the mutant retinogeniculate pheno-
types were secondary to abnormal retinal ac-
tivity. Calcium imaging of mutant retinas
revealed spontaneous retinal waves with spa-
tiotemporal properties indistinguishable from
those of normal mice (30). Thus, we ascribe
abnormalities in the mutant retinogeniculate
projection directly to a loss of class I MHC
signaling downstream of neural activity.
Because activity-dependent structural re-
organizations during development are thought
to arise from cellular mechanisms of synaptic
plasticity (1, 2, 6), we next asked whether
synaptic plasticity is altered in mutant mice.
Because little is known about such mecha-
nisms in the developing LGN, we used a
well-characterized model system for studying
long-lasting changes in the strength of syn-
aptic transmission: the Schaffer collateral-
CA1 synapse of the hippocampus (31, 32).
Class I MHC and CD3were both expressed
in the adult hippocampus (Fig. 1) (10, 15).
Furthermore, class I MHC immunoreactivity
can be detected in synaptosome preparations,
suggesting that some class I molecules are
synaptically associated (33). We therefore as-
sessed hippocampal synaptic plasticity in
wild-type and mutant mice. Data collection
was performed by an observer blind to geno-
type (34 ).
In wild-type mice (C57BL/6), tetanic
stimulation (4 100 Hz) resulted in a sus-
tained increase in the slope of the field exci-
tatory postsynaptic potential (f EPSP) (167
13% of pretetanus baseline; n 15; Fig. 4, A
and C). In contrast, in CD3
–/–
mutant ani-
mals, LTP in response to the same tetanus
was significantly enhanced relative to that in
wild-type mice (248 29% of baseline; n
8; P 0.05; Fig. 4, A and C). A similar
enhancement of LTP was observed in
2
M
–/–
TAP1
–/–
mutant mice (227 22% of base-
line; n 10; P 0.05; Fig. 4C). Basal
synaptic transmission is not significantly dif-
ferent among all experimental groups (35).
Enhanced LTP in gene knockout animals was
not due to changes in inhibition, because
GABA
A
-mediated transmission was blocked
with 100 M picrotoxin in all experiments.
Nor was the enhanced LTP due to induction
of an N-methyl-
D-aspartate (NMDA) recep-
tor-independent form of LTP, because LTP
was completely abolished in all genotypes in
the presence of the NMDA antagonist 2-ami-
no-5-phosphonovalerate [50 M D-APV;
Fig. 4B and (36)].
It is conceivable that enhancement of LTP
seen in these genotypes is due to some non-
specific effect of immune compromise on the
CNS. Thus we also examined LTP in a more
severely immunodeficient strain of mice that
lacks recombination activating gene-1
(RAG1). RAG1 is required for production of
B and T cells and is also transcribed by
neurons in the CNS (37, 38). LTP in RAG1
–/–
mice was indistinguishable from that of wild
type [153 13% of baseline (n 10),
compared with 167 13% in wild type; P
0.48; Fig. 4C], indicating that the LTP abnor-
malities seen in
2
M
–/–
TAP1
–/–
or CD3
–/–
mice are specific to their genotypes rather
than to immune status.
Synaptic plasticity in the hippocampus is
dependent on stimulation frequency, with
high frequencies producing LTP and low fre-
quencies producing LTD (31, 3941). We
therefore examined the effect of other stimu-
lation frequencies on synaptic plasticity in
animals deficient for class I MHC signaling.
In adult wild-type slices, the delivery of 900
pulses at 0.5 Hz induced significant LTD
(82 6% of baseline; n 8; P 0.05; Fig.
4D). In adult slices from both mutant geno-
types, however, there was no significant
change in f EPSP slope upon 0.5 Hz stimula-
tion [CD3
–/–
, 107 7% of baseline (n 5,
P 0.29);
2
M
–/–
TAP1
–/–
,99 5% of
baseline (n 8, P 0.78); Fig. 4D]. Fur-
thermore, after 900 pulses at 1 Hz, transmis-
sion was significantly enhanced over baseline
in both CD3
–/–
(141 14% of baseline, n
5, P 0.05) and
2
M
–/–
TAP1
–/–
slices
(128 9%, n 6, P 0.05) but was
unchanged in wild-type slices (94 5%, n
14, P 0.41; Fig. 4D). Thus, in mutant mice,
LTD could not be detected, and the frequen-
cy-response curve of hippocampal synaptic
plasticity was consistently shifted across a
broad range of stimulation frequencies.
These results indicate that class I MHC/
CD3 signaling is important for mediating
Fig. 3. Abnormal retinogeniculate projections but normal dLGN ultrastructure in mice deficient in
class I MHC signaling. At P12, one eye was injected with WGA-HRP (23); after 1 day, anterograde
axonal transport results in labeling of the entire retinal projection to the LGN. Labeling pattern in
the dLGN is shown in bright-field optics (label is black) or as dark-field composites [label is white;
see (24)]. (A) Representative projection from retina to dLGN contralateral (dashed lines; coronal
section; dorsal is up; lateral is left) or ipsilateral to eye injected with WGA-HRP (asterisks indicate
labeled area from ipsilateral eye: lateral is to right) in a P13
2
M
/
wild-type mouse and a
2
M
–/–
mutant mouse. (B and C) Representative (B) and extreme (C) examples of the projection from the
ipsilateral eye observed in
2
M
/
TAP1
/
mice. (D and E) Representative (D) and extreme (E)
examples of the projection in CD3
–/–
mice. Arrowheads indicate ectopic projections, which appear
extensive under the more sensitive dark-field optics. Scale bar, 200 m. (F) Graph of areas (SEM)
occupied by the ipsilateral retinal projection to the LGN for
2
M
/
(wild-type),
2
M
–/–
,
2
M
–/–
TAP1
–/–
, and CD3
/
mice (24), normalized to total dLGN area. The ipsilateral projection area in
2
M
/
animals is set as 100% (horizontal dashed line). Asterisks indicate significant differences
from
2
M
/
mice (P 0.05, Student’s two-tailed t test). (G), Electron micrograph of the dLGN
from a
2
M
–/–
TAP1
–/–
mouse (at P24), showing a typical R-type synaptic bouton (R) making
contacts with a dendrite (d). A well-myelinated axon (ax) is also present in this field. Scale bar, 1
m.
R EPORTS
www.sciencemag.org SCIENCE VOL 290 15 DECEMBER 2000 2157
activity-dependent synaptic depression, be-
cause, in mutants, there is a shift in the
bidirectional regulation of synaptic strength
[i.e., the frequency response function (39
41)] that favors potentiation. In the absence
of class I MHC or CD3, patterns of neural
activity that normally have no effect on syn-
aptic strength or that lead to synaptic depres-
sion result, instead, in abnormal synaptic
strengthening. Likewise, in the dLGN, en-
hanced LTP and lack of LTD at the develop-
ing retinogeniculate synapse could account
for the structural phenotype observed: a per-
sistence of inappropriate connections that
would be normally be removed via an activ-
ity-dependent process of synaptic weakening
during eye-specific segregation (14, 4244 ).
Class I MHC and CD3 are expressed in the
CNS by specific sets of neurons that undergo
activity-dependent changes (10). Here, we
show that mice lacking these molecules exhibit
abnormalities in connections between these
neurons, suggesting a direct neuronal function
for class I signaling. In addition, both mutants
have strikingly similar phenotypes, implying
that class I MHC signaling in the brain is
transduced via a CD3-containing receptor, ei-
ther an unknown CNS-specific or a known
immune receptor. The expression patterns of
class I MHC and CD3in the CNS are consis-
tent with signaling via a number of possible
receptor-ligand configurations. For example,
both class I MHC and CD3are expressed by
neurons in the hippocampus; in addition, class I
MHC mRNA is also expressed by retinal gan-
glion cells when CD3 is detected in the dLGN
[Fig. 1A and (10)]. Detailed information con-
cerning the ultrastructural localization of these
molecules will be needed to resolve this issue.
Whatever the case, the evidence to date
supports a model in which class I MHC func-
tions in the CNS by engaging CD3-containing
receptors to signal activity-dependent changes
in synaptic strength that ultimately lead to the
establishment of appropriate synapses. Class I
MHC may act directly at the synapse to pro-
mote the elimination of inappropriate connec-
tions, by using signaling mechanisms already
characterized in immune cells (11), possibly via
phosphorylation of CD3by fyn [a kinase pre-
viously implicated in hippocampal plasticity
(45)]. Because different class I MHC mem-
bers are expressed by different subsets of
CNS neurons, additional signaling specificity
may be furnished by the particular repertoire
of MHC molecules present in any given neu-
ron. In the immune system, recognition of
class I MHC by T cell receptors can result in
functional elimination of inappropriate self-
reactive T cell populations (46, 47). Our
results demonstrate that class I MHC is also
required for normal regressive events in the
developing and adult CNS, including activi-
ty-dependent synaptic weakening and struc-
tural refinement.
References and Notes
1. C. S. Goodman, C. J. Shatz, Cell 72, 77 (1993).
2. L. C. Katz, C. J. Shatz, Science 274, 1133 (1996).
3. M. C. Crair, Curr. Opin. Neurobiol. 9, 88 (1999).
4. R. O. Wong, Annu. Rev. Neurosci. 22, 29 (1999).
5. C. J. Shatz, Neuron 5, 745 (1990).
6. D. E. Feldman, R. A. Nicoll, R. C. Malenka, J. Neurobiol.
41, 92 (1999).
7. C. J. Shatz, M. P. Stryker, Science 242, 87 (1988).
8. D. W. Sretavan, C. J. Shatz, M. P. Stryker, Nature 336,
468 (1988).
9. A. A. Penn, P. A. Riquelme, M. B. Feller, C. J. Shatz,
Science 279, 2108 (1998).
10. R. A. Corriveau, G. S. Huh, C. J. Shatz, Neuron 21, 505
(1998).
11. J. E. van Leeuwen, L. E. Samelson, Curr. Opin. Immu-
nol. 11, 242 (1999).
12. C. Amadou et al., Immunol. Rev. 167, 211 (1999).
13. The pan-specific class I MHC probe was obtained by
reverse transcription–polymerase chain reaction (RT-
PCR) of adult rat spleen total RNA, using primers
targeting the 3 domain of the rat class I MHC
molecule rat RT1.Aa [nucleotides (nts) 673 to 859 of
GenBank accession M31018; primers were 5-GATGT
SACCC TGAGG TGCTG-3 and 5-GGCAT GTGTA
MYTCT GCTCC-3]. The resultant clone RATMHC1
exhibited greater than 95% homology with all mouse
class Ia, as well as many class Ib, MHC sequences.
Subclass-specific class I MHC probes were cloned
from mouse CNS by RT-PCR of C57BL/6 mouse hip-
pocampal RNA; primers targeted a segment that
varies considerably among class Ia and Ib subfamily
members [nts 143 to 463 of GenBank accession
U47325 (H–2D
b
); primers were 5-NNGTN GGCTA
YGTKG ACRAC-3 and 5-KYRGG TYYTC RTTCA
GGG-3]. Clones corresponding to H–2D
b
, H–2K
b
(12), Qa-1
b
, and T22
b
[J. L. Lalanne et al., Cell 41, 469
(1985); L. Van Kaer et al., Immunol. Rev. 120,89
(1991)] were identified via BLAST database compar-
ison. Cloning of CD3from mouse spleen RNA and in
situ hybridization analysis were carried out as de-
scribed (10).
14. P. Godement, J. Salaeun, M. Imbert, J. Comp. Neurol.
230, 552 (1984).
15. G. S. Huh et al., data not shown.
16. Cross-hybridization between D, Qa-1, and T22 probes
was minimal, as assessed by cross-competition stud-
ies: cohybridization of each labeled riboprobe with a
10,000-fold mass excess of homologous unlabeled
transcript abolished all signal, whereas cohybridiza-
tion with an excess of the other two unlabeled tran-
scripts resulted in little if any alteration in the hy-
bridization pattern.
17. S. M. Shawar, J. M. Vyas, J. R. Rodgers, R. R. Rich,
Annu. Rev. Immunol. 12, 839 (1994).
18. M. Zijlstra et al., Nature 344, 742 (1990).
19. J. R. Dorfman, J. Zerrahn, M. C. Coles, D. H. Raulet,
J. Immunol. 159, 5219 (1997).
20. M. R. Jackson, P. A. Peterson, Annu. Rev. Cell Biol. 9,
207 (1993).
21. H. Neumann, H. Schmidt, A. Cavalie, D. Jenne, H.
Wekerle, J. Exp. Med. 185, 305 (1997).
22. P. E. Love et al., Science 261, 918 (1993).
23. All surgeries on postnatal mice were performed accord-
ing to institutional guidelines and approved protocols.
2
M
–/–
(5 backcrossed to C57BL/6) and
2
M
–/–
TAP1
–/–
double mutant mice (5 backcrossed to
C57BL/6) were obtained from D. Raulet (University of
California at Berkeley) (18, 19). CD3
–/–
mice (22)(8
backcrossed to C57BL/6) and RAG1
/
mice (38) (10
backcrossed to C57BL/6) were obtained from Jackson
Laboratories (Bar Harbor, ME). As part of the blind
study,
2
M
/–
heterozygotes (from
2
M
–/–
C57BL/6
crosses) were intercrossed;
2
M
/
and
2
M
–/–
pups
were not revealed until after image analysis was com-
plete. P12 mouse pups were anesthetized with isoflu-
rane, and one eye was injected with 1 to 2 l WGA-HRP
(4 to 10% in saline; L7017 from Sigma, St. Louis, MO, or
PL-1026 from Vector Laboratories, Burlingame, CA).
After 22 to 26 hours, 50-m brain sections were pre-
pared for histology essentially as described (9); the
nitroprusside solution used to stabilize the reaction
product was ice-cold and included 10 mM sodium ac-
etate (pH 3.3). For electron microscopy of
2
M
–/–
TAP1
–/–
mice, P24 animals were perfused first with
buffer (0.1 M sodium cacodylate pH 7.35, 5 U/ml hep-
arin) and then 1% paraformaldehyde, 2% glutaralde-
hyde, 0.2% acrolein, and 4 mM CaCl
2
in buffer. The
thalamus was fixed overnight at 4°C; dLGN was isolated
from 150-m Vibratome sections and processed for
electron microscopy.
24. The following series of steps was carried out on all
slide sets by an observer blind to genotype. Only sets
exhibiting equivalent degrees of anterograde labeling
Fig. 4. Enhanced hippocam-
pal LTP in mice deficient ei-
ther for cell surface class I
MHC expression or for
CD3.(A) Field EPSP
(fEPSP) slopes in wild-type
versus CD3
–/–
-deficient
mice. Tetanus was applied
at time 0. (Insets) Superim-
posed sample fEPSPs re-
corded 10 min before or
180 min after tetanic stim-
ulation from individual
wild-type (left) and
CD3
–/–
(right) slices. Scale
bar, 10 msec/0.25 mV. (B)
NMDA receptor depen-
dence of LTP in CD3-defi-
cient mice. Tetanus was ap-
plied at time 0 either in the
absence [filled circles; from
(A)] or presence (hollow cir-
cles) of 50 M D-APV. All points in (A) and (B) are averages of four consecutive fEPSPs (means
SEM, normalized to 15-min baseline) recorded from CA1. (C) Graphs summarizing degree of
potentiation in wild-type,
2
M
–/–
TAP1
–/–
, CD3
–/–
, or RAG1
–/–
mice after 100-Hz tetanus. Data are
shown for mice with histologically normal brains (48). Asterisks indicate significant differences
from wild type (one-way ANOVA, P 0.05). (D) Relation (logarithmic plot) between synaptic
enhancement and stimulation frequency. Points at 0.033 Hz (test pulse frequency) indicate
baseline values (horizontal dashed line). Points at 100 Hz are taken from (C). Values in (C) and (D)
are mean fEPSP slopes for each genotype over the 1-hour period following tetanus. See text and
(34) for methods.
R EPORTS
15 DECEMBER 2000 VOL 290 SCIENCE www.sciencemag.org2158
were selected for analysis. Eight-bit TIFF images con-
taining the dLGN were acquired on a Macintosh-
linked charge-coupled device camera (MTI VE1000)
attached to a Nikon Microphot FXA. Using NIH Image
(v1.62b7), images of the dLGN ipsilateral and con-
tralateral to the injected eye were cropped to exclude
ventral LGN, intrageniculate leaf and extrageniculate
optic tract; images of ipsilateral dLGN were also
modified to eliminate the optic tract running above
the dLGN. NIH Image macros were used to eliminate
background blood vessel-derived staining (very
heavily stained blood vessels were removed by hand)
and to calculate areas occupied by retinal projections
(9). For each brain, an internally controlled measure
of the area occupied by the ipsilateral projection was
obtained by dividing the average of the four largest
ipsilateral areas (corresponding to the middle third of
the dLGN) by the average of the four largest total
dLGN areas (assessed by the outer boundaries of the
contralateral projection zones). Sections in Fig. 3, A
to E, were photographed in both bright-field and
dark-field optics. Although dark-field optics are more
sensitive and reveal lightly labeled regions as white,
very heavily labeled regions become saturated and
appear black. Therefore, for accuracy, composites of
bright-field and dark-field images of the same section
were constructed to ensure that heavily labeled re-
gions appeared white, while detailed information
about lightly-labeled regions revealed in dark-field
was preserved.
25. N. Aggelopoulos, J. G. Parnevelas, S. Edmunds, Anat.
Embryol. (Berl.) 180, 243 (1989).
26. J. A. Rafols, F. Valverde, J. Comp. Neurol. 150, 303
(1973).
27. A. R. Lieberman, K. E. Webster, J. Neurocytol. 3, 677
(1974).
28. M. Tessier-Lavigne, Cell 82, 345 (1995).
29. D. A. Feldheim et al., Neuron 21, 1303 (1998).
30. D. Stellwagen, G. S. Huh, C. J. Shatz, data not shown.
31. R. C. Malenka, R. A. Nicoll, Science 285, 1870 (1999).
32. T. V. Bliss, G. L. Collingridge, Nature 361, 31 (1993).
33. Supplementary data are available on Science Online
at www.sciencemag.org/cgi/content/full/290/5499/
2155/DC1
34. Slices of mouse brain 400 m thick (from 8- to
17-week-old animals, killed with halothane) were
maintained at 25°C in a submerged recording cham-
ber (perfused at 2 to 3 ml/min) with artificial cere-
brospinal fluid (ACSF: 126 mM NaCl, 2.5 mM KCl,
1.25 mM NaH
2
PO
4
, 1.3 mM MgSO
4
, 2.5 mM CaCl
2
,
26 mM NaHCO
3
, and 10 mM glucose). Connections
to the CA3 region of the hippocampus were cut, and
100 M picrotoxin (Sigma) was added to the bath
ACSF. Stainless-steel bipolar electrodes were used to
stimulate Schaffer collateral/commissural fibers;
glass microelectrodes filled with ACSF (2 to 6 M)
were inserted into the stratum radiatum to record
currents from populations of CA1 pyramidal cells.
Test pulses (0.033 Hz) were applied at a stimulation
intensity required to produce an fEPSP that was 30%
(for 100 Hz stimulation) or 50% (for 0.5 and 1 Hz
stimulation) of the maximal response for each re-
cording. High-frequency stimulation (tetanus) con-
sisted of four trains of 100 pulses at 100 Hz (inter-
train interval 15 s), applied at time 0. All values are
reported as means SEM, n is the number of slices
(one slice per mouse). Data collection was performed
by an observer blind to genotype. Before the blind
was dropped, recordings were omitted from analysis
if the extracellular resistance changed significantly
(3/94) or if the stimulating electrode had visibly
drifted over the course of the recording (4/94). LTP
was calculated as the average of responses between
0 and 60 min after tetanus, normalized to a 15-min
pretetanus control period. Stimulus intensity was
relatively high because of the use of electrodes with
uninsulated tips to maximize the number of fibers
stimulated. Stimulus artifacts were clearly complete
well before fEPSP onset and so were easily excluded
from analysis. In experiments using D-APV, drug was
added at least 30 min before tetanic stimulation and
was present throughout the entire recording. Statis-
tical significance was assessed by two-tailed one-way
ANOVA or Student’s t-test.
35. Pretetanus test pulse fEPSP slopes for mice with
normal ventricles (millivolts per millisecond) were as
follows: wild type, 0.091 0.007 (n 14); CD3
–/–
,
0.091 0.009 (n 8, P 0.95 compared with wild
type);
2
M
–/–
TAP1
–/–
, 0.089 0.010 (n 9, P
0.85). Stimulation intensities required to evoke an
fEPSP at 30% of the maximal response (in microam-
peres) were as follows: wild type, 136 27; CD3
–/–
,
134 24 (P 0.75 compared with wild type);
2
M
–/–
TAP1
–/–
, 128 17 (P 0.59).
36. Posttetanus fEPSP slopes, averaged over 180 min, did
not differ significantly from baseline in the presence
of 50 M D-APV: wild type, 107 14% (n 3; P
0.64); CD3
–/–
,97 9% (n 3; P 0.93); and
2
M
–/–
TAP1
–/–
, 106 10% (n 5; P 0.56, Stu-
dent’s t-test).
37. J. J. Chun, D. G. Schatz, M. A. Oettinger, R. Jaenisch, D.
Baltimore, Cell 64, 189 (1991).
38. P. Mombaerts et al., Cell 68, 869 (1992).
39. M. F. Bear, Neuron 15, 1 (1995).
40. K. Deisseroth, H. Bito, H. Schulman, R. W. Tsien, Curr.
Biol. 5, 1334 (1995).
41. M. Migaud et al., Nature 396, 433 (1998).
42. K. F. So, G. Campbell, A. R. Lieberman, J. Exp. Biol.
153, 85 (1990).
43. G. Campbell, C. J. Shatz, J. Neurosci. 12, 1847 (1992).
44. D. W. Sretavan, C. J. Shatz, J. Neurosci. 6, 234 (1986).
45. S. G. Grant et al., Science 258, 1903 (1992).
46. D. Amsen, A. M. Kruisbeek, Immunol. Rev. 165, 209
(1998).
47. E. Sebzda et al., Annu. Rev. Immunol. 17, 829 (1999).
48. When brains of otherwise normal-appearing animals
at age P13 were examined, 52% (16/31) of
2
M
–/–
TAP1
–/–
mice and 22% of CD3
–/–
mice (10/45) had
enlarged lateral ventricles. This phenotype is unlikely
to be due to immunocompromise because severely
immunodeficient RAG1
–/–
mice, when cohoused in
our facility, do not exhibit this phenotype (0/18). This
phenotype also occurs in 57% (12/21) of
2
M
–/–
TAP1
–/–
and 20% (2/10) of CD3
–/–
adult mice. Al-
though ventricular enlargement does not affect the
appearance of the dLGN and thalamus (assessed by
Nissl stains), the size, placement, and appearance of
extrathalamic structures such as the hippocampus
can be altered. In the LTP analysis, animals with
enlarged ventricles were treated separately because
in these animals, LTP measurements could be con-
founded by abnormal hippocampal architecture and
the known reduction of LTP by hydrocephalus [T.
Tsubokawa, Y. Katayama, T. Kawamata, Brain Inj. 2,
19 (1988)]. Consistent with the latter idea, LTP at
100 Hz in
2
M
–/–
TAP1
–/–
mice with dilated ventri-
cles, while still present, is significantly lower than
that of
2
M
–/–
TAP1
–/–
mice with normal-appearing
brains (168 15% relative to 227 22%; P 0.05).
CD3
–/–
mice with dilated ventricles also displayed
diminished LTP (data not shown).
49. We thank S. Wiese, C. Cowdrey, and H. Aaron for
expert technical assistance; D. Stellwagen for exam-
ination of retinal waves in mutant mice; A. Toroian-
Raymond for assistance with electron microscopy; M.
Bennett and E. Choi for advice and assistance with
synaptosome preparations; D. Raulet for generously
providing
2
M
–/–
and
2
M
–/–
TAP1
–/–
mice. Support-
ed in part by grant NIH MH48108 and an Alcon
Research Institute Award to C.J.S. G.S.H. and L.M.B.
were Howard Hughes Associates. L.M.B. was support-
ed by NRSA 1F32EY07016. H.D. was supported by
NRSA EY06912.
31 July 2000; accepted 14 November 2000
R EPORTS
www.sciencemag.org SCIENCE VOL 290 15 DECEMBER 2000 2159
... Cells 2024, 13, 1006 2 of 22 spleen tyrosine kinase (Syk), are expressed by RGCs [14,[19][20][21][22][23]. The genetic mutation of MHCI or CD3 ζ-chain (CD3ζ) resultes in significant defects in the development of RGC structure and function [14,19]. In addition, activation of MHCI promotes locomotor abilities after spinal cord injury [13,17,18,24] while inactivation of MHCI lessens brain injury after stroke [25]. ...
... Many reports have shown that major histocompatibility complex I (MHCI) molecules and their receptors play critical roles in the pathogenesis of the CNS and retina. Mice with defective MHCI or its receptors have abnormal RGC axonal projections in the brain, abnormal synaptic connections in the visual cortex, abnormal motor neuron function, excessive loss of synaptic connections, and axonal regeneration after injury [13][14][15][16][17][18]. In the retina, MHCI, the key components of the T-cell receptor (TCR) complex, such as the CD3 complex (cluster of differentiation 3), and the major downstream cascades of the TCR complex, such as Src (proto-oncogene tyrosine-protein kinase) family kinases (SFK) and To determine whether CD3ζ regulates RGC death through SFK and Syk family kinases and which SFK and Syk family members are involved in the singling, four SFK inhibitors (PP2, A419259, SU6656, and Saracatinib) [31][32][33][34][35][36][37][38] and one Syk/Zap70 inhibitor (Piceatannol) [39,40] were used in the study. ...
... It is well demonstrated that both the TCR-mediated adaptive immune system and complement-mediated innate immune system regulate neuronal development and pathogenesis in the CNS [1,[13][14][15][16][17][18][19]25,55,56]. We have previously shown that CD3ζ is expressed by RGCs and displaced amacrine cells in the mouse retina [19], and mutation of CD3ζ impairs dendritic development of RGCs and ACs [19,53]. ...
Article
Full-text available
Excessive levels of glutamate activity could potentially damage and kill neurons. Glutamate excitotoxicity is thought to play a critical role in many CNS and retinal diseases. Accordingly, glutamate excitotoxicity has been used as a model to study neuronal diseases. Immune proteins, such as major histocompatibility complex (MHC) class I molecules and their receptors, play important roles in many neuronal diseases, while T-cell receptors (TCR) are the primary receptors of MHCI. We previously showed that a critical component of TCR, CD3ζ, is expressed by mouse retinal ganglion cells (RGCs). The mutation of CD3ζ or MHCI molecules compromises the development of RGC structure and function. In this study, we investigated whether CD3ζ-mediated molecular signaling regulates RGC death in glutamate excitotoxicity. We show that mutation of CD3ζ significantly increased RGC survival in NMDA-induced excitotoxicity. In addition, we found that several downstream molecules of TCR, including Src (proto-oncogene tyrosine-protein kinase) family kinases (SFKs) and spleen tyrosine kinase (Syk), are expressed by RGCs. Selective inhibition of an SFK member, Hck, or Syk members, Syk or Zap70, significantly increased RGC survival in NMDA-induced excitotoxicity. These results provide direct evidence to reveal the underlying molecular mechanisms that control RGC death under disease conditions.
... For example, proteins of the complement cascade (including C1q, C3, and C4) and members of MHCI (including H2-K and H2-D) contribute to synapse elimination in the developing visual system. [20][21][22][23][24][25] H2-K and H2-D also negatively regulate dendritic complexity and the frequency and amplitude of miniature excitatory postsynaptic currents (mEPSCs) in hippocampal and cortical neurons and limit the levels of Ca 2+ -permeable a-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) receptors in the lateral geniculate nucleus. 21,22,[26][27][28][29] The central roles of these and other immune proteins in neuronal homeostasis raises the possibility that an immune response to AAV-mediated gene delivery could disrupt neuronal structure and/or function, even in the absence of classical signs of inflammation. ...
... [20][21][22][23][24][25] H2-K and H2-D also negatively regulate dendritic complexity and the frequency and amplitude of miniature excitatory postsynaptic currents (mEPSCs) in hippocampal and cortical neurons and limit the levels of Ca 2+ -permeable a-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) receptors in the lateral geniculate nucleus. 21,22,[26][27][28][29] The central roles of these and other immune proteins in neuronal homeostasis raises the possibility that an immune response to AAV-mediated gene delivery could disrupt neuronal structure and/or function, even in the absence of classical signs of inflammation. ...
... Several immune proteins regulate neuronal morphology and synaptic transmission, including the complement protein C3, the classical MHCI proteins H2-K and H2-D, and the tyrosine kinase Fyn. [20][21][22][23][24][25][26][27][28][29][45][46][47][48] To determine whether any of these known pleiotropic immune molecules are rapidly upregulated in response to AAV8-hSyn-mCherry (1.2 Â 10 11 vg), mRNA levels for H2-K, H2-D, C3, and Fyn were measured in samples of somatosensory cortex via RT-qPCR. Samples were collected at 4 dpi, when dendritic loss is ongoing ( Figures 2D-2F). ...
... MHCI molecules are optimal candidates since they are activity-regulated, expressed in neurons at synapses 44,72,73 , and negatively regulate synapse density selectively during the period of the initial establishment of connections when glutamate rapidly alters synapse number 43,45 . Importantly, little is known about how MHCI causes synapse loss 17,74 . ...
... With few MHCI neuronal interactors known [106][107][108][109][110] , this is the first demonstration that MHCI negatively regulates NL1 levels to control synapse density. Future studies will be needed to determine how MHCI alters NL1 and whether that interaction also plays a role in homeostatic plasticity of synaptic strength at later ages 15,43,72,73,105,111 , and in synapse loss downstream of peripheral immune activation 15,45,74,[112][113][114] and during neurodegeneration 109,115 . ...
Preprint
Although neurons release neurotransmitter before contact, the role for this release in synapse formation remains unclear. Cortical synapses do not require synaptic vesicle release for formation, yet glutamate clearly regulates glutamate receptor trafficking and induces spine formation. Using a culture system to dissect molecular mechanisms, we found that glutamate rapidly decreases synapse density specifically in young cortical neurons in a local and calcium-dependent manner through decreasing NMDAR transport and surface expression as well as co-transport with neuroligin (NL1). Adhesion between NL1 and neurexin 1 protects against this glutamate-induced synapse loss. Major histocompatibility I (MHCI) molecules are required for the effects of glutamate in causing synapse loss through negatively regulating NL1 levels. Thus, like acetylcholine at the NMJ, glutamate acts as a dispersal signal for NMDARs and causes rapid synapse loss unless opposed by NL1-mediated trans-synaptic adhesion. Together, glutamate, MHCI and NL1 mediate a novel form of homeostatic plasticity in young neurons that induces rapid changes in NMDARs to regulate when and where nascent glutamatergic synapses are formed.
... Previous evidence has found that MHC-I molecules, present on the surface of neuronal cells during CNS development and maturation, play a critical in regulating brain development, synaptic plasticity, and social behavior in animals, independently of their typical immune function (Boulanger and Shatz 2004). However, it should be noted that these studies only utilized B2M-null mice or B2M-knockout neuron models to examine changes in the physiological function of neurons lacking MHC-I molecules (Loconto et al. 2003;Huh et al. 2000;Corriveau et al. 1998;Goddard et al. 2007). Caution is needed in interpreting these findings, as these models not only altered the amount of MHC-I on the surface of neurons, but also reduced free B2M levels in CNS. ...
Article
Full-text available
Central nervous system (CNS) disorders represent the leading cause of disability and the second leading cause of death worldwide, and impose a substantial economic burden on society. In recent years, emerging evidence has found that beta2 -microglobulin (B2M), a subunit of major histocompatibility complex class I (MHC-I) molecules, plays a crucial role in the development and progression in certain CNS diseases. On the one hand, intracellular B2M was abnormally upregulated in brain tumors and regulated tumor microenvironments and progression. On the other hand, soluble B2M was also elevated and involved in pathological stages in CNS diseases. Targeted B2M therapy has shown promising outcomes in specific CNS diseases. In this review, we provide a comprehensive summary and discussion of recent advances in understanding the pathological processes involving B2M in CNS diseases (e.g., Alzheimer's disease, aging, stroke, HIV-related dementia, glioma, and primary central nervous system lymphoma). Supplementary Information The online version contains supplementary material available at 10.1007/s10571-024-01481-6.
... Several MHC genes associated with VDM in this investigation (including the marginally replicated HLA-DRA) have been associated with AD (HLA-A, HLA-B, HLA-DRA [61][62][63]) or showed increased hippocampal (HLA-DMA, HLA-DMB, HLA-DPA1, HLA-DRA [60]) or pre-frontal cortex (HLA-A, HLA-C, HLA-E, HLA-F, HLA-G, HLA-DPB1 [60]) expression in mild AD dementia cases compared to non-demented controls. MHC genes may influence memory through their effects on synaptic plasticity, development, morphology, and function [57,[64][65][66]. ...
Article
Full-text available
Background Uncovering the functional relevance underlying verbal declarative memory (VDM) genome-wide association study (GWAS) results may facilitate the development of interventions to reduce age-related memory decline and dementia. Methods We performed multi-omics and pathway enrichment analyses of paragraph (PAR-dr) and word list (WL-dr) delayed recall GWAS from 29,076 older non-demented individuals of European descent. We assessed the relationship between single-variant associations and expression quantitative trait loci (eQTLs) in 44 tissues and methylation quantitative trait loci (meQTLs) in the hippocampus. We determined the relationship between gene associations and transcript levels in 53 tissues, annotation as immune genes, and regulation by transcription factors (TFs) and microRNAs. To identify significant pathways, gene set enrichment was tested in each cohort and meta-analyzed across cohorts. Analyses of differential expression in brain tissues were conducted for pathway component genes. Results The single-variant associations of VDM showed significant linkage disequilibrium (LD) with eQTLs across all tissues and meQTLs within the hippocampus. Stronger WL-dr gene associations correlated with reduced expression in four brain tissues, including the hippocampus. More robust PAR-dr and/or WL-dr gene associations were intricately linked with immunity and were influenced by 31 TFs and 2 microRNAs. Six pathways, including type I diabetes, exhibited significant associations with both PAR-dr and WL-dr. These pathways included fifteen MHC genes intricately linked to VDM performance, showing diverse expression patterns based on cognitive status in brain tissues. Conclusions VDM genetic associations influence expression regulation via eQTLs and meQTLs. The involvement of TFs, microRNAs, MHC genes, and immune-related pathways contributes to VDM performance in older individuals.
Article
Mounting evidence indicates that a physiological function of amyloid-β (Aβ) is to mediate neural activity-dependent homeostatic and competitive synaptic plasticity in the brain. I have previously summarized the lines of evidence supporting this hypothesis and highlighted the similarities between Aβ and anti-microbial peptides in mediating cell/synapse competition. In cell competition, anti-microbial peptides deploy a multitude of mechanisms to ensure both self-protection and competitor elimination. Here I review recent studies showing that similar mechanisms are at play in Aβ-mediated synapse competition and perturbations in these mechanisms underpin Alzheimer’s disease (AD). Specifically, I discuss evidence that Aβ and ApoE, two crucial players in AD, co-operate in the regulation of synapse competition. Glial ApoE promotes self-protection by increasing the production of trophic monomeric Aβ and inhibiting its assembly into toxic oligomers. Conversely, Aβ oligomers, once assembled, promote the elimination of competitor synapses via direct toxic activity and amplification of “eat-me” signals promoting the elimination of weak synapses. I further summarize evidence that neuronal ApoE may be part of a gene regulatory network that normally promotes competitive plasticity, explaining the selective vulnerability of ApoE expressing neurons in AD brains. Lastly, I discuss evidence that sleep may be key to Aβ-orchestrated plasticity, in which sleep is not only induced by Aβ but is also required for Aβ-mediated plasticity, underlining the link between sleep and AD. Together, these results strongly argue that AD is a disease of competitive synaptic plasticity gone awry, a novel perspective that may promote AD research.
Article
Full-text available
Learning and memory require activity-induced changes in dendritic translation, but which mRNAs are involved and how they are regulated are unclear. In this study, to monitor how depolarization impacts local dendritic biology, we employed a dendritically targeted proximity labeling approach followed by crosslinking immunoprecipitation, ribosome profiling and mass spectrometry. Depolarization of primary cortical neurons with KCl or the glutamate agonist DHPG caused rapid reprogramming of dendritic protein expression, where changes in dendritic mRNAs and proteins are weakly correlated. For a subset of pre-localized messages, depolarization increased the translation of upstream open reading frames (uORFs) and their downstream coding sequences, enabling localized production of proteins involved in long-term potentiation, cell signaling and energy metabolism. This activity-dependent translation was accompanied by the phosphorylation and recruitment of the non-canonical translation initiation factor eIF4G2, and the translated uORFs were sufficient to confer depolarization-induced, eIF4G2-dependent translational control. These studies uncovered an unanticipated mechanism by which activity-dependent uORF translational control by eIF4G2 couples activity to local dendritic remodeling.
Article
Full-text available
Aging brings about a myriad of degenerative processes throughout the body. A decrease in cognitive abilities is one of the hallmark phenotypes of aging, underpinned by neuroinflammation and neurodegeneration occurring in the brain. This review focuses on the role of different immune receptors expressed in cells of the central and peripheral nervous systems. We will discuss how immune receptors in the brain act as sentinels and effectors of the age-dependent shift in ligand composition. Within this 'old-age-ligand soup,' some immune receptors contribute directly to excessive synaptic weakening from within the neuronal compartment, while others amplify the damaging inflammatory environment in the brain. Ultimately, chronic inflammation sets up a positive feedback loop that increases the impact of immune ligand-receptor interactions in the brain, leading to permanent synaptic and neuronal loss.
Chapter
There are several entrance ways for pathogens to enter the brain despite the blood–brain barrier. They can breach the tight junctions, cross through endothelial cells, or use the Trojan horse route by hiding in leukocytes. Less used are the nasal pathway and the meninges and subarachnoid space. All cells of the brain have intrinsic mechanisms to battle infections by viruses like other body cells. The difference is that infected neurons cannot easily be removed by immune cells as most of them have irreplaceable roles in circuits. Microglial cells have a key function in the defense. They are the first line of defense and one of their main priorities is to protect neurons that are infected. They even instruct invading T cells to avoid attacks on infected neurons. The downside is that latent virus infections of neurons is a common problem. Despite these protective mechanisms for neurons, direct damage by viruses is almost as common as neuronal damage by overreacting immune cells, for example, by a cytokine storm. In extreme cases, this can lead to encephalitis, an out-of-control inflammatory reaction due to infections.
Article
Full-text available
Mice homozygous for a beta2-microglobulin gene disruption do not express any detectable beta2-ITL protein. They express little if any functional major histocompatibility complex (MHC) class I antigen on the cell surface yet are fertile and apparently healthy. They show a normal distribution of gammadelta, CD4+8+ and CD4+8- cells, but have no mature CD4-8+ cells and are defective in CD4-8+ cell-mediated cytotoxicity. Our results strongly support earlier evidence that MHC class I molecules are crucial for positive selection of T cell antigen receptor alphabeta+ CD4-8+ T cells in the thymus and call into question the non-immune functions that have been ascribed to MHC class I molecules.
Article
Full-text available
Cells from mice with mutations in the genes for beta2-microglobulin (beta2m) or for TAP-1 express only low levels of MHC class I proteins on their surfaces, and are thus sensitive to attack by normal NK cells. Although NK cells are present in beta2m- mice and TAP-1(-) mice, they are completely self-tolerant. The underlying mechanism for this tolerance is unknown. It has been proposed that education processes render NK cells from these mice hypersensitive to class I-mediated inhibition, so that they can be inhibited even by the low levels of class I expressed on autologous cells. In this study, we present evidence against this hypothesis, by demonstrating that NK cells from beta2m- mice and TAP-1(-) mice fail to attack beta2m(-)TAP-1(-) double-mutant cells in both in vitro and in vivo assays. The latter cells express substantially lower levels of class I than single-mutant cells, based on serologic tests, as well as a significantly diminished sensitivity to attack by class I-specific CTL. Furthermore, the Ly-49 repertoire on NK cells derived from beta2m(-)TAP-1(-) mice is highly similar to that of either single mutant, indicating that the developmental processes that shape the Ly-49 repertoire cannot respond to the differences in class I levels among these mice. We propose that self-tolerance of NK cells in beta2m- mice and TAP-1(-) mice is likely to result from hyporesponsiveness of the cells to activating signals, or alternatively, to induction of inhibitory signaling through receptors specific for non-class I MHC ligands.
Article
Full-text available
Mice with mutations in four nonreceptor tyrosine kinase genes, fyn, src, yes, and abl, were used to study the role of these kinases in long-term potentiation (LTP) and in the relation of LTP to spatial learning and memory. All four kinases were expressed in the hippocampus. Mutations in src, yes, and abl did not interfere with either the induction or the maintenance of LTP. However, in fyn mutants, LTP was blunted even though synaptic transmission and two short-term forms of synaptic plasticity, paired-pulse facilitation and post-tetanic potentiation, were normal. In parallel with the blunting of LTP, fyn mutants showed impaired spatial learning, consistent with a functional link between LTP and learning. Although fyn is expressed at mature synapses, its lack of expression during development resulted in an increased number of granule cells in the dentate gyrus and of pyramidal cells in the CA3 region. Thus, a common tyrosine kinase pathway may regulate the growth of neurons in the developing hippocampus and the strength of synaptic plasticity in the mature hippocampus.
Article
Full-text available
The synaptic organization of identified retinogeniculate axons was studied during the prenatal development of eye-specific layers in the LGN of the cat. During this period, retinogeniculate axons undergo stereotyped morphological changes. Retinogeniculate axons originating from one eye and passing through LGN territory destined to be solely innervated by the other eye (inappropriate territory) initially give rise to many side branches. As the eye-specific layers emerge, these axons elaborate extensive terminal arbors within territory appropriate to their eye of origin and concurrently retract their side branches from inappropriate territory (Sretavan and Shatz, 1986). These transient side branches may therefore represent a morphological substrate for the observed functional convergence of inputs from the two eyes onto common LGN neurons during prenatal development (Shatz and Kirkwood, 1984). This possibility was investigated by examining whether identified axons and their side branches form synapses in inappropriate territory. Three retinogeniculate axons from two fetuses aged embryonic day 53 (E53) and E57 were filled with HRP in an in vitro preparation, prior to being processed for electron microscopy (EM). The HRP-filled axons, originating from the contralateral eye, were first reconstructed at the light microscope level. The portion of axon passing through the center of ipsilaterally innervated layer A1 was then serially sectioned and reconstructed by EM. Two sets of 450 serial EM sections revealed that all three contralateral axons established synaptic contacts in ipsilateral territory. Many of these synapses were made by side branches and a few were even formed by the main axon trunks. Both side branches and trunks formed mainly en passant asymmetrical contacts that were associated with spherical synaptic vesicles and that were apposed to immature dendritic elements and dendritic shafts. For comparison, a portion of the same E53 axon within the future contralateral layer A was also serially sectioned and reconstructed for EM. Within this contralateral zone, the E53 axon formed synaptic contacts similar to those established in the ipsilateral region, except that in the appropriate zone they contained significantly more synaptic vesicles. These results demonstrate that axons from the contralateral eye can establish synapses in territory simultaneously innervated by the ipsilateral eye, both via side branches and by means of contacts along the main axon trunk. Thus, the development of eye-specific layers is accompanied by the formation and subsequent elimination of synapses that almost certainly represent a morphological substrate for the known transient functional convergence of inputs from the two eyes.
Article
Specific patterns of neuronal firing induce changes in synaptic strength that may contribute to learning and memory. If the postsynaptic NMDA (N-methyl-D-aspartate) receptors are blocked, long-term potentiation (LTP) and long-term depression (LTD) of synaptic transmission and the learning of spatial information are prevented. The NMDA receptor can bind a protein known as postsynaptic density-95 (PSD-95), which may regulate the localization of and/or signalling by the receptor. In mutant mice lacking PSD-95, the frequency function of NMDA-dependent LTP and LTD is shifted to produce strikingly enhanced LTP at different frequencies of synaptic stimulation. In keeping with neural-network models that incorporate bidirectional learning rules, this frequency shift is accompanied by severely impaired spatial learning. Synaptic NMDA-receptor currents, subunit expression, localization and synaptic morphology are all unaffected in the mutant mice. PSD-95 thus appears to be important in coupling the NMDA receptor to pathways that control bidirectional synaptic plasticity and learning.
Article
When contacts are first forming in the developing nervous system, many neurons generate spontaneous activity that has been hypothesized to shape appropriately patterned connections. In Mustela putorius furo, monocular intraocular blockade of spontaneous retinal waves of action potentials by cholinergic agents altered the subsequent eye-specific lamination pattern of the lateral geniculate nucleus (LGN). The projection from the active retina was greatly expanded into territory normally belonging to the other eye, and the projection from the inactive retina was substantially reduced. Thus, interocular competition driven by endogenous retinal activity determines the pattern of eye-specific connections from retina to LGN, demonstrating that spontaneous activity can produce highly stereotyped patterns of connections before the onset of visual experience.
Article
The V(D)J recombination activation gene RAG-1 was isolated on the basis of its ability to activate V(D)J recombination on an artificial substrate in fibroblasts. This property and the expression pattern in tissues and cell lines indicate that RAG-1 either activates or catalyzes the V(D)J recombination reaction of immunoglobulin and T cell receptor genes. We here describe the introduction of a mutation in RAG-1 into the germline of mice via gene targeting in embryonic stem cells. RAG-1-deficient mice have small lymphoid organs that do not contain mature B and T lymphocytes. The arrest of B and T cell differentiation occurs at an early stage and correlates with the inability to perform V(D)J recombination. The immune system of the RAG-1 mutant mice can be described as that of nonleaky scid mice. Although RAG-1 expression has been reported in the central nervous system of the mouse, no obvious neuroanatomical or behavioral abnormalities have been found in the RAG-1-deficient mice.
Article
We have studied the ligand specificity of a gamma delta T-cell receptor (TCR) derived from a mouse T-cell hybridoma (KN6). KN6 cells reacted with syngeneic (C57BL/6) cells from various origins (splenocytes, thymocytes, peritoneal exudate cells, etc.) and cells from many different mouse strains. KN6 reactivity against cells from a panel of congenic and recombinant mouse strains demonstrated that the ligand recognized by KN6 is controlled by an MHC-linked gene that most probably maps in the TL region. We cloned this gene and formally proved that it does map in the TL region. This gene turned out to be a novel class I gene (designated T22b) belonging to a hitherto unidentified cluster of TL region genes in strain C57BL/6. This gene was expressed in many different tissues and cell types. We also examined the tissue expression of several other TL genes. One of these, the structural gene (T3b) encoding the thymus leukemia (TL) antigen from C57BL/6 mice, was specifically expressed in the epithelium of the small intestine. Since the intestinal epithelium of the mouse is known to be the homing site for a subset of gamma delta T cells (i-IEL) bearing diverse TCR with V7 rearranged gamma chains, we propose that the T3b gene product is part of the ligand recognized by some of the i-IEL. Our data support the idea that gamma delta T cells might be specific for non-classical class I or class I-like molecules and suggest that gamma delta TCR and non-classical MHC co-evolved for the recognition of a conserved set of endogenous or foreign peptides.
Article
The recombination activating genes, RAG-1 and RAG-2, are likely to encode components of the V(D)J site-specific recombination machinery. We report here the detection of low levels of the RAG-1 transcript in the murine central nervous system by polymerase chain reaction, in situ hybridization, and Northern blot analyses. In contrast, an authentic RAG-2 transcript could not be detected reproducibly in the central nervous system. The RAG-1 transcript was found to be widespread in embryonic and postnatal neurons, with transcription being most apparent in regions of the postnatal brain with a high neuronal cell density (the cerebellum and the hippocampal formation). The results suggest that RAG-1 functions in neurons, where its role might be to recombine elements of the neuronal genome site-specifically, or to prevent detrimental alterations of the genome in these long-lived cells.
Article
Guidelines for submitting commentsPolicy: Comments that contribute to the discussion of the article will be posted within approximately three business days. We do not accept anonymous comments. Please include your email address; the address will not be displayed in the posted comment. Cell Press Editors will screen the comments to ensure that they are relevant and appropriate but comments will not be edited. The ultimate decision on publication of an online comment is at the Editors' discretion. Formatting: Please include a title for the comment and your affiliation. Note that symbols (e.g. Greek letters) may not transmit properly in this form due to potential software compatibility issues. Please spell out the words in place of the symbols (e.g. replace “α” with “alpha”). Comments should be no more than 8,000 characters (including spaces ) in length. References may be included when necessary but should be kept to a minimum. Be careful if copying and pasting from a Word document. Smart quotes can cause problems in the form. If you experience difficulties, please convert to a plain text file and then copy and paste into the form.