ArticlePDF AvailableLiterature Review

The origin and evolution of synapses

Authors:

Abstract

Understanding the evolutionary origins of behaviour is a central aim in the study of biology and may lead to insights into human disorders. Synaptic transmission is observed in a wide range of invertebrate and vertebrate organisms and underlies their behaviour. Proteomic studies of the molecular components of the highly complex mammalian postsynaptic machinery point to an ancestral molecular machinery in unicellular organisms--the protosynapse--that existed before the evolution of metazoans and neurons, and hence challenges existing views on the origins of the brain. The phylogeny of the molecular components of the synapse provides a new model for studying synapse diversity and complexity, and their implications for brain evolution.
The synapse is an intensely studied and highly special-
ized cellular site yet surprisingly little is known about its
origins and evolution. A century of electrophysiologi-
cal and pharmacological studies have demonstrated the
importance of synaptic function as a universal property
of neural circuits. During the past two decades, molecu-
lar studies have provided us with an understanding of
the fundamental mechanisms of synaptic transmission
and plasticity, and have shown that these are remarkably
conserved across a range of animal species1.
It remains unknown whether there are central behav-
ioural mechanisms shared by all organisms and whether
these share common ancestry. For example Hebbs learn-
ing model proposed in the 1940s2 applies only to animals
with neural circuits. Broader concepts of behaviour tra-
versing unicellular organisms to humans were stated in
the nineteenth century. In 1891, C. L. Morgan observed
that “the primary end and object of the receptions of
the influences (stimuli) of the external world or environ-
ment, is to enable the organisms to answer or respond
to these special modes of influence, or stimuli. In other
words, their purpose is to set agoing certain activities.
Now in the unicellular organisms, where both the recep-
tion and the response are effected by one and the same
cell, the activities are for the most part simple, though
even among these protozoa there are some which show
no little complexity of response3.
Scholarly attempts to investigate or draw theories on
the evolution of the nervous system did not until recently
address the question of the origin and evolution of the
synapse itself. This is not surprising, given the lack of
tools available to explore the synapse in an evolutionar-
ily meaningful way, as the traditional methods to study
synapse structure and function, which rely on micros-
copy and electrophysiology, do not serve this purpose.
To examine the synapse through an evolutionary frame-
work, experimental methods are required that allow the
identification of the specific adaptations that mediated
the diversification and natural selection of synapses
between species. Recent systematic approaches to study
synapse evolution are based on advancements in pro-
teomics and genomics and are coupled with molecular
phylogenetic approaches. Proteomics has provided the
key step in the understanding of this field by identifying
the protein constituents of synapses and their interac-
tions. The postsynaptic proteome of mice has ~1,500 pro-
teins and the presynaptic proteome and synaptic vesicles
also comprise hundreds of proteins4,5, demonstrating the
complexity of mammalian synapses.
The postsynaptic density (PSD) is composed of a set of
multiprotein complexes assembled from diverse protein
classes, such as ion channels, receptors, cell adhesion
and cytoskeletal proteins, kinases and phosphatases,
scaffolding proteins and signalling molecules4,6 (see
Supplementary information S1 (table)). Multiprotein
complexes that are associated with postsynaptic neuro-
transmitter receptors (such as glutamate, serotonin and
acetylcholine receptors) as well as potassium channels
have also been extensively characterised at the PSD
because of their key role in synaptic transmission5.
N-methyl--aspartate (NMDA) receptors, and their
interacting MAGUK (membrane-associated guanylate
kinase) proteins, are particularly well studied and dis-
cussed in detail in this Review. Interestingly, a large
proportion of the MAGUK associated signalling com-
plexes (MASC) proteins has been implicated in synaptic
plasticity and/or learning phenotypes in mutant mice, or
associated with psychiatric disorders in humans5–8. The
general concept of multiprotein signalling complexes
as ‘signalosomes’ has now been widely accepted, and is
*Wellcome Trust Sanger
Institute, Hinxton, Cambridge
CB10 1SA, UK.
Wolfson College, University
of Cambridge, Barton Road,
Cambridge, CB3 9BB, UK.
Correspondence to S.G.N.G.
e‑mail: sg3@sanger.ac.uk
doi:10.1038/nrn2717
Published online
9 September 2009
Postsynaptic proteome
The complete set of proteins
currently identified at the
postsynaptic side of the
synapse.
MAGUK
Membrane-associated
guanylate kinase (MAGUK)
proteins act as scaffolds for the
clustering of receptors, ion
channels and associated
signalling proteins at
postsynaptic sites.
The origin and evolution of synapses
Tomás J. Ryan*and Seth G. N. Grant*
Abstract | Understanding the evolutionary origins of behaviour is a central aim in
the study of biology and may lead to insights into human disorders. Synaptic
transmission is observed in a wide range of invertebrate and vertebrate
organisms and underlies their behaviour. Proteomic studies of the molecular
components of the highly complex mammalian postsynaptic machinery point to an ancestral
molecular machinery in unicellular organisms — the protosynapse — that existed before the
evolution of metazoans and neurons, and hence challenges existing views on the origins of
the brain. The phylogeny of the molecular components of the synapse provides a new model
for studying synapse diversity and complexity, and their implications for brain evolution.
REVIEWS
NATURE R EVIEWS
|
NEUROSCIENCE VOLUME 10
|
OCTOBER 2009
|
701
FOCUS ON CNS EVOLUTION
© 2009 Macmillan Publishers Limited. All rights reserved
Ursynapse
The last common ancestor
of all synapses. This was the
platform from which diversity
of synaptic proteins
between different organisms
and different synapse types
evolved.
Orthologues
Homologous genes that
separated due to a speciation
event.
Protosynapse
Those synaptic components
that were present before the
emergence of synapses and
most likely contributed to their
evolution.
Bilaterians
Animals belonging to the
phylum Bilateria. These are a
clade of animals with bilateral
symmetry that possess
complex nervous systems.
They are divided into
protostomes and
deuterostomes.
Outgroup
A group of organisms that
serves as a reference group for
determination of the
evolutionary relationship
between monophyletic groups
of organisms.
Choanoflagellates
Organisms belonging to the
phylum Choanoflagellata.
These are unicellular
eukaryotes that can exist in
both free-living and colonial
forms, and are multicellular
metazoans considered to be
the closest unicellular relative
of multicellular metazoans.
Porifera
Phylum of multicellular animals
(poriferans or sponges) that
lack a nervous system.
illustrated by examples in immunology (the mamma-
lian T cell receptor)9 and vision research (the Drosophila
melanogaster InaD complex)10.
Complementary to charting the proteome of syn-
apses, the flood of data from genome sequencing projects
allows the use of comparative genomics to examine how
and when synaptic proteins originated and diversified.
Comparative genomics has revolutionized evolutionary
biology, bolstering our understanding of the relation-
ships between species, and also into the mechanisms
of molecular evolution11. By combining proteomic and
genomic approaches, the postsynaptic proteome pro-
vides unprecedented opportunities for probing and map-
ping the evolution of the synapse across species. Here we
review such attempts and present the evidence accumu-
lated so far as a working model of synapse molecular
evolution, based primarily on the postsynaptic machin-
ery, which has been more extensively studied. We also
make suggestions as to how this topic may be addressed
in the future.
The origin of synapses
The question of origin is central to the study of evo-
lution12. When did synapses first form? What was the
identity of the organism that necessitated this evolution-
ary step? Were the first synapses capable of plasticity?
How are the origins of neurons and synapses related?
Knowing the proteomic composition of synaptic struc-
tures allows the use of comparative genomics to question
what the minimal, ancestral components of the synapse
might have been. If we can deduce the composition of
the last common ancestor of all synapses, the ursynapse,
then we should be in a position to address the question
of how and why the first synapse originated.
We approach the question of the composition of the
ursynapse by taking synaptic proteins identified in ver-
tebrate model organisms and searching for orthologues
in the genomes of two categories of organism. The first
category includes unicellular eukaryotes and multicel-
lular metazoans that lack a nervous system, and pro-
vides a means of identifying synaptic components that
were present before the evolution of the nervous system.
Such proteins, which make up the protosynapse, origi-
nated before the emergence of classical morphological
synapses and have been co-opted for synaptic roles. The
second category is composed of non-bilaterian multi-
cellular metazoans that have a nervous system, which
allows the identification of synaptic components present
in primitive synapses, giving us insight into the com-
position of the synapse at a relatively short time after it
originated (FIG. 1).
Protosynaptic organisms. The most evolutionary ancient
synaptic protein families are conserved in unicellular
eukaryotes such as the yeast Saccharomyces cerevisiae
and the amoeba Dictyostelium discoideum. The calcium
transporter PMCA (plasma membrane calcium ATPase)
and protein kinase C (PKC) are found in both these uni-
cellular species and also have fundamental synaptic
roles in animals1315. More broadly, when the mamma-
lian MASCs and larger PSD gene sets were compared
against the S. cerevisiae genome, over 21% of the MASC
genes and 25% of the PSD genes were found to have
direct orthologues in protosynaptic organisms, labelling
them as protosynaptic proteins of pre-metazoan ances-
try16. Whether protosynaptic components in yeast form
functional multiprotein complexes in a similar manner
to which they do in neurons is unknown, but loss-of-
function studies in yeast show that many of these genes
are involved in regulating the cell’s response to the envi-
ronment (TABLE 1)16, including vesicular trafficking and
cytoskeletal regulation in response to nutrients, ions
and pheromones from the environment16. A portion
(15%) of yeast PSD orthologues functions in signal trans-
duction pathways that mediate environmental response.
Yeast orthologues of the mammalian RAS regulator
neuro fibronin 1 (NF1; that is, yeast IRA2), the extracel-
lular signal-regulated kinase 2 (ERK2; also known as
MAPK1; that is, yeast FUS3), and the G-protein guanine
nucleotide binding protein 5 (GNB5; that is, yeast STE4)
regulate cell morphology, stress response and cell pro-
liferation induced by pheromone signalling. Strikingly,
quintessential players that mediate changes dependant on
neuronal activity during synaptic plasticity are also essen-
tial components of the yeast’s response to environmental
changes. The calcium binding protein calmodulin that
activates calcium/calmodulin-dependent protein kinase II
(CaMKII) and the protein phosphatase calcineurin
(which has been associated with Schizophrenia17)
both regulate calcium influx at the cell membrane and
are implicated in postsynaptic signalling pathways in
neurons and in environmental responses in yeast cells18.
Yeast is an excellent organism for the study of genet-
ics and genomics, but is a distant outgroup of metazoans
and is unlikely to hold a quasi-comprehensive set of
proto synaptic proteins. To understand the origin of the
synapse it is desirable to consider a unicellular organ-
ism that is closely related to multicellullar Metazoa.
Choanoflagellates fulfil this role. These unicellular eukary-
otes are the closest unicellular relatives of multicellular
metazoans on account of their cell body structure and
phylogenetic analyses of nuclear and mitochondrial
DNA19,20 (FIG. 1). The genomes of choanoflagellates have
assisted in reconstructing the genome of the last unicel-
lular ancestor of animals, and could also be useful for elu-
cidating the origin of the synapse21. Multicellular Porifera
(sponges), which lack neurons, are further candidates for
the identification of protosynaptic components22.
Bioinformatic analyses of choanoflagellate genomes
using comprehensive PSD gene sets as a starting point
have yet to be reported. However, pioneering studies of
candidate genes have demonstrated the presence of syn-
aptic molecules in choanoflagellates that are absent in
all other non-metazoans, including numerous tyrosine
kinases of the choanoflagellate Monosiga brevicollis23–26.
Rapid phosphorylation of tyrosine kinase substrates was
observed following treatment of starved M. brevicollis
cells with seawater and Enterobacter spp., which demon-
strates that the tyrosine kinase pathway has a role in the
environmental response24. Inhibition of tyrosine kinases
delayed progression of M. brevicollis cultures into log-
arithmic growth, wherease a SRC family tyrosine kinase
REVIEWS
702
|
OCTOBER 2009
|
VOLUME 10 www.nature.com/reviews/neuro
REVIEWS
© 2009 Macmillan Publishers Limited. All rights reserved
Deuterostomes
Bilaterians
Eumetazoans
Metazoans
Eukaryotes
Ursynapse
Protostomes
Greater signalling complexity,
MASC gene family expansion,
NMDA receptor duplication,
MAGUK duplication
Stargazin
LIMK
Excitatory glutamate receptors (NMDA and AMPA),
kainate receptors,
K+ channels, neuroligin, CASK, erbin
GABA receptors, metabotropic glutamate receptors,
CaMKII, KIR channel, NOS, SynGAP,
S-SCAM, Homer, GKAP, GRIP, CRIPT, agrin, MuSK, ankyrin, neurexin, NCAM
Cadherins, ephrin receptors, receptor and non-receptor tyrosine kinases,
Dlg (MAGUK), shank, calpain, spectrin, PDZ binding proteins
MASC downstream signaling components, PKC, PMCA, NF 1, calmodulin, calcineurin
Vertebrates Lophotrochozans Ecdysozoans Cnidarians Poriferans Choanoflagellates Fungi
Nature Reviews | Neuroscience
With synapses
Without synapses
(protosynaptic)
619–790 mya
Av. 790 mya
581–1,141 mya
Av. 910 mya
766–1,351 mya
Av. 1,036 mya
766–1,351 mya
Av. 1,237 mya
970–1,070 mya
Av. 1,020 mya
1,220–1,513 mya
Av. 1,368 mya
Demosponge
Organism belonging to the
primary class of Porifera.
Demosponges account for
~90% all sponge species.
inhibitor abolished cell proliferation24. Given the central
role of SRC family kinases in plasticity at excitatory syn-
apses, the establishment of this apparatus and mechanism
before the evolution of synapses is of significance27,28.
Molecules that have been implicated in synaptogene-
sis such as cadherins are also present in choanoflagellates,
and co-localize with actin at the cell’s apical collar mem-
brane, surrounding the singular flagellum29. Homophilic
cadherin interactions are essential for excitatory synapse
formation in mammals and in D. melanogaster, in which
cadherin interacts directly with actin at the postsynap-
tic terminal30. Cadherins may therefore be important
for cytoskeletal rearrangement in choanoflagellates and
may represent a precursor to synapse formation. It is
conceivable that the first protein–protein interaction
that led to synaptogenesis would be homophilic, as the
same molecules would be required by both cells, whereas
heterophilic transynaptic protein interactions probably
evolved later.
Comparisons between the PSDs of Bilateria and
sponges have allowed the identification of genes that
were present in a multicellular ancestor of Metazoa
that immediately preceded the emergence of organisms
with synaptic junctions13,31. Additional cell-signalling
and adhesion molecules that have roles in synaptogen-
esis have been identified in the sponge Oscarella carmela,
including ankyrin, neurexin and tyrosine kinase signal-
ling components 32. We postulate that these molecules
originated before the evolution of synapses, in a meta-
zoan/choanoflagellate common ancestor, and there-
fore were instrumental to the development of the first
synaptic junction. Transgenic experiments in unicellu-
lar organisms and sponges may prove useful in testing
this idea.
Many MASC components that have been implicated
in synaptic plasticity in mammals are also present in
the demosponge Amphimedon queenslandica (TABLE 1)13.
Various interaction domains of postsynaptic proteins,
Figure 1 | Phylogenetic tree depicting taxons of current relevance to synapse evolution. An extant model
organism of each clade is displayed at the top of the phylogeny (see Supplementary information S2 (Box) for additional
details on the phylogenetic placing of Porifera relative to Cnidaria). Nodes on the phylogeny represent the divergence
points of various clades and are presented by coloured circles. The red node represents Urbilateria (the last common
ancestor of all bilaterians). The small grey circle represents the ursynapse (last common ancestor of all synapses). Beside
each node the range of published estimations of the given divergence time are given in mya (millions of years ago), as well
as the average (av) estimated divergence time based on published studies (available through the public resource for
knowledge on the timescale and evolutionary history of life, Timetree43,104). Superimposed on the phylogeny are notable
proteins that are involved in synapse formation and/or function, showing at what intervals in evolutionary history various
synaptic components arose. See Supplementary information S3 (Box) for additional details on the possible origins of
GABA and metabotropic glutamate receptors. AMPA, α‑amino‑3‑hydroxy‑5‑methyl‑4‑isoxazolepropionic acid;
CaMKII, calcium/calmodulin‑dependent protein kinase II; CASK, calcium/calmodulin dependent serine protein kinase;
CRIPT, cysteine‑rich PDZ‑binding protein; Dlg, discs, large homolog; GABA, g‑aminobutyric acid; GKAP, guanylate
kinase associated protein; GRIP, glutamate receptor interacting protein; KIR channel, inwardly rectifying potassium
channel; LIMK, LIM domain kinase; MAGUK, membrane‑associated guanylate kinase; MASC, MAGUK associated
signalling complex; MuSK, muscle specific kinase; NCAM, neural cell adhesion molecule; NF1, neurofibromin 1;
NMDA, N‑methylase‑d‑aspartate; NOS, nitric oxide synthase; PKC, protein kinase C; PMCA, plasma membrane calcium
transporting ATPase; Shank, SH3 and multiple ankyrin repeat domains; S‑SCAM, membrane associated guanylate kinase,
WW and PDZ domain containing 2; SynGAP, synaptic Ras GTPase activating protein.
REVIEWS
NATURE R EVIEWS
|
NEUROSCIENCE VOLUME 10
|
OCTOBER 2009
|
703
FOCUS ON CNS EVOLUTION
© 2009 Macmillan Publishers Limited. All rights reserved
including PDZ binding domains, are almost completely
conserved between the sponge and the Bilateria, argu-
ing strongly in favour of the existence of an assembled
multiprotein structure. Five of the postsynaptic proteins
are present and co-expressed in epithelial flask cells of
A. queenslandica (discs large homologue (Dlg), guanylate
kinase-associated protein (GKAP), glutamate receptor
interacting protein (GRIP), Homer and cysteine-rich
PDZ-binding protein (CRIPT)), implying that they
may form a protosynaptic complex involved in sensing
environmental stimuli and that they could represent an
evolutionary precursor to synaptic sites33. Importantly, a
number of channels and receptors that have crucial roles
at synapses are also expressed in the sponge, including
inhibitory GABA receptors, the K+ channel KIR, and
metabotropic (G-protein coupled) glutamate receptors
Table 1 | The origin and ancestral functions of protosynaptic proteins
Molecule Clade of origin Organism reported Ancestral function
PKA Fungi S. cerevisiae16 Nutrient induced cell proliferation106
NF1 Fungi S. cerevisiae16 Stress response107
Calmodulin Fungi S. cerevisiae16 Ca2+ dependant stress response18
Calcineurin Fungi S. cerevisiae16 Ca2+ dependant stress response108
ERK2 Fungi S. cerevisiae16 Pheromone induced cell proliferation109
GNB5 Fungi S. cerevisiae16 Pheromone induced signalling110
SNAP‑25 Fungi S. cerevisiae111 Vacuolar morphogenesis and trafficking111
Syntaxin Fungi S. cerevisiae16 Vacuolar morphogenesis and trafficking111
TRP channels Fungi S. cerevisiae112 Osmolarity stress response112
Cadherin Choanoflagellata M. brevicollis24 Unknown, co‑localizes with actin filaments
at the apical collar29
SRC kinase Choanoflagellata M. brevicollis113 Regulation of cell proliferation24
RAF kinase Choanoflagellata M. brevicollis24 Unknown
Ephrin receptors Choanoflagellata M. brevicollis26 Unknown
Calpain Choanoflagellata M. brevicollis114 Unknown
Spectrin Choanoflagellata M. brevicollis115 Unknown
Dlg (MAGUK) Choanoflagellata M. brevicollis13,116 Unknown
Shank Choanoflagellata M. brevicollis13,116 Unknown
Agrin Porifera O. carmela31 Unknown
MuSK Porifera O. carmela31 Unknown
Ankyrin Porifera O. carmela31 Unknown
Neurexin Porifera O. carmela31 Unknown
NCAM Porifera O. carmela31 Unknown
GABA receptors Porifera A. queenslandica13 Unknown
mGluR receptors Porifera A. queenslandica13 Unknown, activity modulates Ca2+ influx35
KIR channels Porifera A. queenslandica13 Unknown
CaMKII Porifera A. queenslandica13 Unknown
NOS Porifera A. queenslandica13 Unknown
SynGAP Porifera A. queenslandica13 Unknown
S‑SCAM Porifera A. queenslandica13 Unknown, located in epithelial cells117
Homer Porifera A. queenslandica13 Unknown, located in epithelial cells13
GKAP Porifera A. queenslandica13 Unknown, located in epithelial cells13
GRIP Porifera A. queenslandica13 Unknown, located in epithelial cells13
CRIPT Porifera A. queenslandica13 Unknown, located in epithelial cells13
A. queenslandica, Amphimedon queenslandica; CaMKII, calcium/calmodulin‑dependent protein kinase II; CRIPT, cysteine‑rich
PDZ‑binding protein; Dlg, discs large homolog; ERK2, extracellular signal‑regulated kinase 2; GKAP, guanylate kinase associated
protein; GNB5, guanine nucleotide binding protein β5; GRIP, glutamate receptor interacting protein; KIR channel, inwardly rectifying
potassium channel; M. brevicollis, Monosiga brevicollis; MuSK, muscle specific kinase; NCAM, neural cell adhesion molecule; NF1,
neurofibromin 1; NOS, nitric oxide synthase; O. carmela, Oscarella carmela; PKA, protein kinase A; S. cerevisiae, Saccharomyces
cerevisiae; Shank, SH3 and multiple ankyrin repeat domains; SNAP‑25, synaptosome‑associated protein of 25,000 daltons; S‑SCAM,
membrane associated guanylate kinase, WW and PDZ domain containing 2; SynGAP, synaptic Ras GTPase activating protein; TRP
channel, transient receptor potential channel.
REVIEWS
704
|
OCTOBER 2009
|
VOLUME 10 www.nature.com/reviews/neuro
REVIEWS
© 2009 Macmillan Publishers Limited. All rights reserved
Cnidarian
Animal belonging to the
phylum Cnidaria. Cnidarians
are animals with radial
symmetry including jellyfish,
coral, hyrda and anemones.
Cnidarian nervous systems
consist of diffuse neuronal
net-like structures.
Clade
An evolutionary group
consisting of a given single
common ancestor and all of its
descendants.
Protostomes
Animals belonging to the
phylum Protostomia, an animal
clade that includes the
superphyla Ecdysozoa
(arthropods and nematodes)
and Lophotrochozoa.
Deuterostomes
Animals belonging to the
superphylum Deuterostomia
that includes the subphylum
Vertebrata.
(mGluRs)13,34. mGluR activity has been demonstrated in
cells cultured from the sponge Geodia cydonium, which
responds to the presence of glutamate by rising its intra-
cellular calcium concentration35. Indeed, application of
electrical or tactile stimuli to the sponge Rhabdocalyptus
dawsoni, causes it to propagate an electrical impulse that
controls water flow through the organism36. As no exci-
tatory, ionotropic glutamate receptors have been identi-
fied in sponges it seems that inhibitory GABA receptors
evolved in the common ancestor of Porifera and Bilateria,
and importantly, preceded the origin of excitatory
ionotropic glutamate receptors.
The above evidence not only demonstrates the exist-
ence of MASC components and electrically active ion
channels in sponges, but implies that there is a co-localized
protosynaptic complex expressed in an anatomical region
that has a role in environmental adaptation, and this
complex emerged around the same time, in evolution-
ary terms, as the first synaptic channels and receptors.
Of course, many of these proteins have also evolved non-
neuronal functions in animals and indeed their presence
in unicellular organisms and sponges demonstrates their
fundamental roles in intracellular signalling. For exam-
ple PI3K (phosphoinositide 3-kinase) is a protooncogene
that regulates cell division, and MAGUK family mem-
bers are known regulators of mammalian T cell activa-
tion9,37,38. Nevertheless, these pleiotropic proteins have
evolved discrete synaptic functions and it is compelling
to see that a substantial number of synaptic proteins
regulate environmental adaptations in unicellular organ-
isms such as yeast and choanoflagellates, and co-localize
in the sensory cells of sponges13,37,39. Protosynaptic pro-
teins represent therefore a pre-adaptation that later co-
opted for synaptic function and may have contributed to
the development of the first synapse.
Primitive metazoans with synapses. The complimen-
tary approach to investigate the origins of the synapse
is to search the genomes of primitive metazoans with
visible synapses, such as the cnidarian Nematostella vect-
ensis, which phylogenetically branched off ~200 million
years before the bilaterians (FIG. 1) and has a rudimentary
nervous system13. Common features of cnidarians and
bilaterians are representative of a common ancestor that
would have existed following the origin of the synapse
in early metazoans (TABLE 2). The most notable feature
of cnidarians is the emergence of postsynaptic iono-
tropic glutamate receptors including NMDA receptors
and AMPA receptors (α-amino-3-hydroxy-5-methyl-
4-isoxazolepropionic acid receptors), which mediate
synaptic plasticity in the vertebrate CNS40,41.
Additionally, the postsynaptic transmembrane
protein neuroligin originated in a common ances-
tor of N. vectensis and bilaterians, and interacts tran-
synaptically with the presynaptic and protosynaptic
protein neurexin during synaptogenesis42. The neurexin–
neuroligin interaction has an essential and general role in
the formation of both excitatory and inhibitory synapses.
It is noteworthy that ectopic expression of neurexin
or neuroligin in non-neuronal cells co-cultured with
neurons is sufficient to induce synaptic differentiation
of the non-neuronal cells42. All of the ion channels that
emerged after the cnidarian–poriferan split interact
with protosynaptic intracellular scaffold proteins that
are expressed in sponges and which therefore evolved
earlier. In effect, upstream transmembrane receptors
plug into a pre-existing intracellular machinery and
this provides a framework for the emergence of a higher
complexity of signalling pathways (FIG. 2).
Comparing vertebrates and invertebrates
Owing to the pioneering studies of the molecular
mechanisms of synaptic plasticity in the lophotro-
chozoan Aplysia californica (sea slug) and the parallel
mechanisms subsequently discovered in mice1, one
might assume that an invertebrate synapse is essentially
identical to a vertebrate synapse. Recent evidence chal-
lenges this paradigm, highlighting the degree of diver-
gence that has occurred between the synaptic proteomes
of vertebrates and invertebrates16. Bilaterians comprise
two major clades: the protostomes and the deuterostomes.
Protostomes include among other phyla the invertebrate
Arthropoda (arthropods) and Nematoda (nematodes)
(FIG. 1), whereas deuterostomes include the subphylum
Vertebrata (vertebrates). The time of divergence of pro-
tostomes and deuterostomes is estimated at ~910 mil-
lion years ago43. The hypothetical last common ancestor
of all protostomes and deuterostomes is referred to as
Urbilateria, which gave rise to the vast majority of ani-
mal diversity that exists in nature today44. On the one
Table 2 | Synaptic proteins present in early organisms with a nervous system
Molecule Clade idenified Organism Synaptic function
NMDA receptors Cnidaria N. vectensis13 Induction of synaptic plasticity118
AMPA receptors Cnidaria N. vectensis13 Fast synaptic transmission and plasticity119
Kainate receptors Cnidaria N. vectensis13 Modulation of synaptic transmission and
plasticity120
Shaker channel Cnidaria N. vectensis13 Synaptic homeostasis121
Neuroligin Cnidaria N. vectensis13 Synapse formation42
Erbin Cnidaria N. vectensis13 Modulation of voltage dependant calcium
channels122
CASK Cnidaria N. vectensis13 Regulation of neurotransmitter release123
AMPA, α‑amino‑3‑hydroxy‑5‑methyl‑4‑isoxazolepropionic acid; CASK, calcium/calmodulin‑dependent serine protein kinase;
NMDA, N‑methyl‑d‑aspartate; N. vectensis, Nematostella vectenis.
REVIEWS
NATURE R EVIEWS
|
NEUROSCIENCE VOLUME 10
|
OCTOBER 2009
|
705
FOCUS ON CNS EVOLUTION
© 2009 Macmillan Publishers Limited. All rights reserved
Homologues
Set of genes or proteins that
are related by descent, that is,
they share a common ancestor.
Genome duplication
Duplication of an entire
genome that results in an
abundance of duplicated
genes, most of which are lost.
Two rounds of genome
duplication are believed to
have occurred at the base of
the chordate lineage.
Gene duplication
Duplication of a given gene
owing to replication errors and
resulting in two redundant
copies of the original gene.
Paralogues
Homologous genes that
separated because of a gene
duplication event.
hand, comparisons between protostomes and deuteros-
tomes allow inferences as to the features of the urbilat-
erian, because whatever is common to protostomes and
deuterostomes was present in the common ancestor. On
the other hand, comparisons between protostomes and
deuterostomes can identify features that have evolved
specifically within each clade16.
Few direct comparisons have been made for synapses
at the morphological level. One anecdotal difference is
that D. melanogaster dendrites can develop from the pri-
mary neurite (that normally give rise to axons) as well as
from the neuronal body, whereas in deuterostomes den-
drites have only been reported to arise independently
from axons45. At the molecular level, the urbilaterian
would have possessed many of the canonical neuro-
transmitter receptors, including acetylcholine, GABA,
glycine, dopamine and serotonin receptors, as these are
found in both protostomes and deuterostomes (FIG. 1)46,47.
The voltage-dependant calcium channel subunit star-
gazin (FIG. 2) is a canonical urbilaterian synaptic protein,
as it is present in all bilaterians and is absent in cnidar-
ians. When protostome genomes including D. mela-
nogaster, Caenorhabditis elegans and Apis mellifera
were searched using the mouse PSD and MASC gene
sets, 44.8% and 46.2%, respectively, had orthologues in
D. melanogaster16. This shows that the PSDs of proto-
stomes and deuterostomes share a core set of homologous
proteins that has expanded divergently since the urbilate-
rian ancestor, and that in vertebrates the PSD and MASC
complexes have more components. Genomic compari-
sons can identify which mouse PSD proteins are absent
from protostome genomes and are therefore specific
to vertebrate synapses. However, such analyses cannot
identify the actual components of the protostome PSD or
MASC complexes. Indeed, the D. melanogaster PSD
could include numerous protostome specific proteins.
Comparative proteomics can address this problem. Mass
spectrometric analysis of the D. melanogaster MASC
detected 220 proteins, showing that D. melanogaster
possesses a MASC similar in size to that of mouse (186
proteins). However, it is the content of the MASC that
differs between the two organisms, and this feature can-
not be observed at the level of genomic sequence com-
parison. When the isolated D. melanogaster MASC was
compared across species in the same manner as for the
mouse, 71% of the D. melanogaster MASC components
were found to have orthologues in yeast, whereas mice
MASC components only had 21.2% yeast orthologues,
showing that the majority of D. melanogaster MASC pro-
teins are ancestral, being common to all eukaryotes and
not of invertebrate or metazoan origin16. Therefore, the
mouse MASC is enriched for recently evolved proteins,
relative to that of invertebrates. Additionally, when the
components of the mouse MASC were classified accord-
ing to gene ontology categories, over 60% were found to
be ‘upstream’ signalling components, such as receptors,
scaffolding proteins and signal transduction molecules
(FIG. 3). Conversely, only 25% of the D. melanogaster
MASC falls into this category, with the majority pertain-
ing to the category of ‘downstream’ components, such
as metabolic enzymes, chaperones and mitochondrial
proteins. These numbers suggest that the deuterostome
MASC has evolved a significantly greater degree of
signalling complexity.
The paradigm of the NMDA receptor carboxy‑terminus
and its interactions. Besides the evolution of synaptic
complexity, there is evidence that the organization of
these proteins, mediated by their interaction domains,
has also differentially evolved. For example, GKAP
(guanylate kinase associated protein) and shank (SH3
and multiple ankyrin repeats domain) are both post-
synaptic scaffolding proteins that regulate synapse
assembly in animals and are present in sponges13,48,49.
However, the GKAP PDZ domain necessary for its
interaction with shank is deuterostome specific, and
is likely to influence synaptogenesis13,50. The signifi-
cance of domain evolution at the synapse is exempli-
fied by the intracellular carboxy-terminal domains of
the NMDA receptor, which interacts with proteins of the
MASC51,52. Comparison of the amino acid sequence of
various excitatory and metabotropic glutamate recep-
tors across species revealed a significant vertebrate–
invertebrate dichotomy in the protein size of NMDA
receptors51. Specifically, the vertebrate NMDA receptor
NR2 (NMDA receptor 2) subunits possess intracellu-
lar carboxy-terminal domains five times larger than all
known invertebrate, protostome NR2 protein sequences
(FIG. 4). This evolutionary dichotomy is unique to the
NR2 subunit as no differences of this magnitude were
observed for any other neurotransmitter receptor sub-
unit. Within the protostome and deuterostome clades
the NR2 carboxy-terminal domains retain a relatively
conserved size of ~100 amino acids in protostomes and
~600 amino acids in deuterostomes. The larger, deutero-
stome NR2 carboxy-terminus contains many protein
interaction sites and phosphorylation sites that are not
present in protostomes, arguing strongly in favour of a
higher complexity of NMDA receptor interactions in
vertebrates (FIG. 4). A high degree of sequence conserva-
tion of the carboxy-terminal domain is seen in the deu-
terostome clade, demonstrating functional significance
of this domain, which has been elegantly demonstrated
with respect to hippocampal synaptic plasticity and
memory formation by in vivo genetic studies53.
NR2 duplication and carboxy‑terminal divergence.
Two rounds of whole genome duplication have occurred
in a common ancestor of chordates in the deutero-
stome clade54. As a result many D. melanogaster genes
have up to four orthologues in mouse. Synaptic gene
families that have undergone chordate specific expan-
sion include sodium and potassium channels, calcium
channels, transient receptor potential (TRP) channels,
glutamate and GABA receptors, ephrin receptors, pro-
tein kinase C (PKC), plasma membrane calcium ATPase
(PMCA), cadherin, neuroligin, Dlg, nitric oxide syn-
thase (NOS), calcium/calmodulin-dependent protein
kinase II (CaMKII) and GKAP13,55. NMDA recep-
tors have evolved further in vertebrates due to gene
duplications of the NR2 subunit that have resulted in
four distinct paralogues (NR2A–NR2D), which have
REVIEWS
706
|
OCTOBER 2009
|
VOLUME 10 www.nature.com/reviews/neuro
REVIEWS
© 2009 Macmillan Publishers Limited. All rights reserved
Nature Reviews | Neuroscience
Ionotropic glutamate receptors
Metabotropic glutamate receptors
Calcium ion binding proteins
Protein kinase activity
Small GTPase signal transduction
Translational elongation
Protein biosynthesis
Fungi Protostomes Deuterostomes
NR2/MAGUK gene
family expansion
Stargazin
NMDA receptors
AMPA receptors
Neuroligins
mGluRs
SynGAP
CaMKII
Dlg (MAGUKS)
Calcineurin
Calmodulin
Fungi
Emergence of titular MASC components
Evolution of PSD and MASC components
Choanoflagellates Poriferans Cnidarians Protosomes Deuterostomes
Upstream
Downstream
Upstream
Downstream
Bilaterians
a
b
diverged a great extent with respect to their spatio-
temporal expression patterns56. The four NR2 subunits
have also diverged at the level of protein sequence, but
pri marily in their intracellular domains — with no
motif being conserved between the four paralogues
except the terminal PDZ binding domain that inter-
acts with the MAGUKs51. Therefore although the NR2
amino-terminal domains that contain the extracellular
Figure 2 | Evolution of postsynaptic components. a | Various protein types (left column) that constitute the
postsynaptic density (PSD) and membrane‑associated guanylate kinase (MAGUK) associated signalling complexes
(MASCs) in unicellular eukaryotes (fungi), protostomes and deuterostomes are ordered based on whether they have
‘upstream’ or ‘downstream’ signalling roles. Non‑coloured fields represent the absence of a given protein. Dark grey
rectangles represent presence of protein. Grey rectangles represent enrichment of a protein type in protostomes, or
protostomes and deuterostomes when compared with unicellular eukaryotes. Light grey rectangles represent enrichment
of a protein type in deuterostomes when compared with protostomes. The diagrams above each column represent the
molecular assembly of MASC, in which upstream proteins (blue circles) are connected to downstream proteins (yellow
triangles) through intermediate signalling proteins (red rectangle). The relative proportions of these proteins in
eukaryotes, protostomes and deuterostomes is therefore illustrated. b | The emergence of titular MASC components
across clades is illustrated. Proteins are ordered based on whether they are located ‘upstream’ or ‘downstream’ in synaptic
signal transduction pathways16. Non‑coloured fields represent the absence of a given protein, whereas dark grey
rectangles denote its presence. Diagrams of MASC structure are placed above each clade, along with an illustration of a
representative model organism. AMPA, α‑amino‑3‑hydroxy‑5‑methyl‑4‑isoxazolepropionic acid; CaMKII, calcium/cal
modulin‑dependent protein kinase II; Dlg, discs large homologue; mGluRs, metabotropic glutamate receptors; NMDA,
N‑methyl‑d‑aspartate; NR2, NMDA receptor 2; SynGAP, synaptic Ras GTPase activating protein. Diagrams in part a are
modified, with permission, from REF. 16 (2008) Macmillan Publishers Ltd. All rights reserved.
REVIEWS
NATURE R EVIEWS
|
NEUROSCIENCE VOLUME 10
|
OCTOBER 2009
|
707
FOCUS ON CNS EVOLUTION
© 2009 Macmillan Publishers Limited. All rights reserved
Nature Reviews | Neuroscience
Signalling and/or structural
components
Channels and
receptors
Cytoskeletal and
cell adhesion
G proteins and
modulators
Kinases
MAGUKs; adaptors;
scaffolders
Protein phosphatases
Signaling molecule
and enzymes
Synaptic vesicles;
protein transport
Transcription and
translation
Uncharacterized
dMASC
mMASC
Immunological synapse
A region that can form
between two cells of the
immune system in close
contact. The immunolgical
synapse originally reffered to
the interaction between a T cell
and an antigen-presenting cell.
Positive selection
Positive selection is said to
occur when a given genetic
variant rises to prevalence in a
population by increasing the
reproductive fitness of the
organism in a given
environment. Positive selection
at the level of amino acid
sequence is identified by the
dN/dS ratio.
Non-synonymous nucleotide
substitution
A nucleotide substitution in the
coding sequence of a gene that
alters the amino acid sequence
of the protein.
Synonymous nucleotide
substitution
A nucleotide substitution in the
coding sequence of a gene that
does not alter the amino acid
sequence of the protein.
ligand-binding and transmembrane domains, are largely
conserved throughout vertebrate evolution, selection
has acted primarily on the intracellular domains of the
NR2 proteins. The evolutionary divergence of the NR2
carboxy-terminal domains has resulted in different
MASC sets being recruited by different NMDA recep-
tors depending on the particular NR2 subunit present.
Experimental evidence has shown that a number of
NMDA receptor interacting proteins, such as SynGAP
(synaptic Ras GTPase activating protein)57, adaptor pro-
tein 2 (AP2)58, calcineurin59, cyclin dependant kinase 5
(CDK5)60, PSD95 and SAP102 (REF. 61) interact specifically
or preferentially with particular NR2 subunits.
NMDA‑receptor–MAGUK interactions. MAGUKs are
intracellular scaffolds that anchor and traffic NMDA
receptors and AMPA receptors to the synapse39,62. They
are composed of PDZ domains, a SRC homology 3 (SH3)
domain and a guanylate kinase domain. The MAGUK
family has 22 members classified into 7 subfamilies
(MAG1, CASK, MPP, ZO, DLG, CCNB and CARMA)52,
all of which originated in metazoans: 4 subfamilies are of
ancient origin and are present in Porifera (MAGI, MPP,
DLG, DLG5); 2 subfamilies (ZO, CACNB) emerged
later in cnidarians and bilaterians; and the CARMA
(CARD–MAGUK) subfamily, which functions in signal
transduction at immunological synapses63, is deuterostome
specific. Although most MAGUK subfamilies are com-
mon to both protostomes and deuterostomes, five show
deuterostome specific gene expansion thereby bolster-
ing the number of MAGUKs available. For example, Dlg
has only been duplicated in vertebrates, giving rise to
four paralogues: DLG1–DLG4 (also known as SAP97,
PSD93, SAP102, and PSD95, respectively), and muta-
tions in DLG1–DLG4 in mice result in distinct behav-
ioural phenotypes64–66. Deuterostomes have single NR2
and Dlg genes whereas chordates have four paralogues
of each, thus the number of potential chordate NR2–
Dlg interactions has expanded to be 12-fold greater67
(FIG. 4b). This particular interaction serves as an exam-
ple for the combinatorial complexity at the level of
protein interactions that has evolved for chordate
synaptic proteins.
GABA and metabotropic glutamate receptor evolution.
Inhibitory GABA receptors also display deuterostome
specific expansion, with two primary clades of genomi-
cally clustered subunits. The first contains α, γ and ε
subunits, the second includes ρ, β, σ, θ and π subunits68.
The two groups diverged within the deuterostome clade,
in a common ancestor of chordates and urochordates68.
Protostomes by contrast possess a small compliment of
GABA-receptor-like subunits69. GABA receptor subunits
have also undergone mammalian specific gene dupli-
cation and loss. The GABA receptor ε and τ subunits
evolved by duplication of the γ and β subunits, respec-
tively, and the β4 and γ4 subunits were eliminated dur-
ing evolution68,70. Positive selection has since acted on the
protein coding region of the θ subunit, whereas the ε has
displayed evidence of relaxation of constraint (neutral
evolution)68. The functional significance of these evolu-
tionary events is unknown, but may lie in ligand binding
affinity or receptor sensitivity.
Metabotropic glutamate receptors, also have an
ancient origin at the base of Metazoans, and show
substantial expansion in vertebrates13,71. mGluRs are
classified into three groups (I–III) based on sequence
homology. Whereas vertebrates have two to three
members in each group, invertebrates have one71,72.
Presynaptic proteins. Comparative genomics has made
progress in examining the evolution of presynaptic pro-
teins across bilaterians. One study addressed the conser-
vation of 120 presynaptic proteins between vertebrates
and insects73. This protein set showed strong but variable
conservation, and interestingly the degree of conserva-
tion for exocytotic proteins (such as synaptobrevin 2
(VAMP2)) was found to correlate with the number of
interacting partners. This is consistent with reports dem-
onstrating that protein connectivity generally correlates
with protein conservation, that is, the more interactions a
given protein has, the more likely it is to have its sequence
conserved throughout evolution74. However, for endocytic
proteins (such as synaptojanin) the degree of conservation
did not correlate with the number of protein interactions.
The reason for this dichotomy is unknown but may lie in
group specific protein size and domain structure.
Another study selected 150 primarily presynaptic
genes and studied them with respect to conservation of
their coding sequences and adjacent non-coding genomic
elements across 8 vertebrate species including human and
mouse75. The vast majority of the gene set was found to
be under strong purifying selection, with non-synonymous
nucleotide substitution to synonymous nucleotide substitution
ratios (dN/dS ratios) over fivefold lower than the genomic
average. This outcome is consistent with reports of genes
expressed in the brain being under more conservative
selective pressure than genes expressed in other tissues76.
Figure 3 | Comparative proteomics of mouse and Drosophila melanogaster MASC.
Pie charts showing the proportion of various functional classes of proteins in the isolated
Drosophila melanogaster membrane‑associated guanylate kinase (MAGUK) associated
signalling complex (dMASC) and mouse MASC (mMASC). The mMASC is enriched for
‘upstream’ signalling components (blue circles in right panels). Figure is modified, with
permission from REF. 16 (2008) Macmillan Publishers Ltd. All rights reserved.
REVIEWS
708
|
OCTOBER 2009
|
VOLUME 10 www.nature.com/reviews/neuro
REVIEWS
© 2009 Macmillan Publishers Limited. All rights reserved
Nature Reviews | Neuroscience
NR2ANR2 NR2B NR2C NR2D
E
S
S
V
L
D
V
E
S
D
V
E
S
D
V
E
S
D
V
PSD95 PSD93 SAP102
Dlg
b Mouse NMDA receptor
a Drosophila NMDA
receptor
Membrane
PSD95
CaMKII
NMDA receptor
complex/MASC
Synapse diversity
Variability of synaptic protein composition is visible not
only between different organisms. There is also a paral-
lel between the evolutionary origin of synaptic genes
and their expression pattern in the mammalian brain.
In general terms, the evolution of synaptic genes at the
eukaryote–metazoan and metazoan–chordate bounda-
ries preceded their expression in different populations
of neurons and synapses and thereby allowed diversity of
function in nervous systems that have emerged later, in
evolutionary terms, and that are generally larger16. For
example, MASC mRNA expression has been compared
between 22 regions of the mouse brain, and this has shown
that the genes of most variable expression originated since
the deuterostome common ancestor, whereas genes of low
expression variability are of ancient, pre-metazoan origin
(FIG. 5). It is likely that the relative expression of MASC
components results in different combinatorial versions of
MASC components in separate brain regions, contributing
to synaptic and neuronal diversity.
Innovative transgenic methods have contributed to our
knowledge of synapse diversity, showing that expression
pattern of synaptic proteins is an identifier of neuronal
subsets. One study conducted whole genome expression
analysis on mRNA isolated from 12 discrete green fluo-
rescent protein (GFP)-tagged GABAergic or glutamater-
gic neuronal populations of the mouse brain, to produce
a phylogeny of neuronal subtypes77. Using this method
it was shown that among genes that exhibited variable
expression between brain regions, synaptic and axonal
components were significantly over-represented and
that there was more expression diversity in GABAergic
neuronal subpopulations than in glutamatergic ones.
Importantly, genes expressed in the brain with hetero-
geneous expression patterns are enriched for paralogues,
which lends credence to the hypothesis that gene dupli-
cation leads to the subfunctionalisation of duplicates by
divergence of expression patterns77,78. Indeed, the excita-
tory glutamate receptors and MAGUKs fall into the highly
variable region and have undergone chordate specific gene
duplication16. The resolution of such profiling studies will
increase in the future as larger sets of genes are studied in
more discrete anatomical regions79. The development of
sophisticated transgenic technologies, such as those using
tagged ribosomal proteins expressed in discrete neuronal
populations, will facilitate such studies80,81.
Human synapse evolution
What makes us human is one of the oldest and most
tantalising questions in biology82–84. The most striking
feature of humans is our increased cognitive capacity.
Little is known about the genetic events that led to the
emergence of human specific behaviour but it is widely
assumed that the enlarged and convoluted neocortex is
the basis for our increased mental abilities (see article by
Pasko Rakic in this issue)85. Here, we posit that the evolu-
tion of the synapse complement might have enabled the
increase of neuronal cell types and, in turn, the neuronal
network complexity of the human brain. Due to difficul-
ties in obtaining appropriate tissue samples for proteomic
work the investigation of human synaptic complexes is
Figure 4 | NMDA receptor carboxy-terminal evolution. Schematic comparing the
Drosophila melanogaster NR2 containing NMDA receptor (a) and the mouse NR2
containing NMDA (N‑methyl‑d‑aspartate) receptors (b). The PDZ binding domains at the
carboxy‑terminus of both D. melanogaster NMDA receptor 2 (NR2; SVL) and mouse NR2
(ESDV) are indicated. Note that the mouse NR2 intracellular domain is five times larger
than that of D. melanogaster. The diagrams compare the evolution of NMDA receptor
signalling complexity in D. melanogaster and mouse showing protein–protein
interactions at the NR2 carboxy‑termini. The only established protein interaction site on
the D. melanogaster NR2 carboxy‑terminus is the interaction of Dlg (discs large
homologue; SAP97 orthologue) (Bayes A. and S.G.N.G., unpublished data) through the
PDZ binding domain. The vertebrate NR2B carboxy‑terminus has numerous established
primary and secondary interacting proteins (zoom out) and therefore a greater degree of
NMDA receptor signalling complexity. Furthermore, the number of potential interactions
of the NMDA receptor with MAGUK (membrane‑associated guanylate kinase)
components differ between protostome and deuterostome synapses. In the case of
protostomes only one such interaction can occur, between NR2 and the protostome
MAGUK, Dlg. Because of gene family expansion in chordates there are four available NR2
subunits (NR2A–NR2D) and four MAGUKs (PSD95, SAP102, PSD93 and SAP97). Three of
the MAGUK paralogues can interact with any of the four NR2 subunits, making twelve
potential deuterostome NR2–MAGUK interactions. As NMDA receptors are considered
to be tetramers that contain two NR2 subunits, the existence of tri‑heteromeric NMDA
receptor channel further increases this combinatorial complexity105.
REVIEWS
NATURE R EVIEWS
|
NEUROSCIENCE VOLUME 10
|
OCTOBER 2009
|
709
FOCUS ON CNS EVOLUTION
© 2009 Macmillan Publishers Limited. All rights reserved
Nature Reviews | Neuroscience
60
50
40
30
20
10
0
Yeast
Protostome
Deuterostome
Percentage of (high/low)
variability genes
a
b
High variability
Low variability
dN/dS ratio
The ratio of non-synonymous
nucleotide substitutions to
synonymous nucleotide
substitution for a given
protein-coding gene. A dN/dS
ratio of <1 implies purifying
selection or conservative
evolution, ~0 implies
relaxation of constraint or
neutral evolution, >1 implies
positive selection or adaptive
evolution. This measure is
based on Kimura’s theory of
molecular evolution, which
argues that the vast majority of
nucleotide sequence changes
are functionally neutral.
not straightforward. In the mean time, there are examples
of positive selection (accelerated evolution) of particular
synaptic, primate and human proteins, including recep-
tors for acetylcholine, oxytocin, kainate, dopamine and
glutamate receptor subunits86. Of particular interest is the
NR2A subunit, which shows evidence of positive selec-
tion in primates and happens to be the NR2A paralogue
that is expressed postnatally in mammals.
The human genes of GABA(A) receptor subunits
display a disproportionate number of mutations in
microRNA target recognition sites, indicating alterations
in human GABA receptor gene expression87. MAOA
(monoamine oxidase A), an enzyme that catabolises
neurotransmitters, has been shown to be under posi-
tive selection in humans88. Within populations, genetic
variations of MAOA have also been implicated in anti-
social behaviour in humans89,90. Additionally, dopamine
receptor subunit D4 (DRD4) and GABA(A) receptor β2
subunit (GABRB2) variants are under positive selection
in human populations, and are associated with attention
deficit hyperactivity disorder, and schizophrenia, respec-
tively9196 . It is possible that positive selection on synaptic
proteins has contributed to the evolution of human behav-
iour at the cost of increased vulnerability to psychiatric
conditions97–99.
Comparison of protein expression levels in the
brains of human and non-human primates has led to
the identification of numerous proteins, including syn-
aptic proteins such as CaMKII, that are upregulated
specifically in the human brain100,101. Currently, we can
only speculate about the function of the human specific
variants of synaptic proteins and the reasons why their
selection sometimes correlates with ‘disease’ phenotypes.
Cell culture studies allied with transgenic ‘humaniza-
tion’ of genes in model organisms may help to elucidate
the role of these genes in primate and/or human brain
evolution.
Conclusions
In this Review we discuss how the synapse, a fundamental
structural and functional unit of the nervous system, is
a fertile platform with which to explore the evolution of
the brain. Many mammalian synaptic components existed
before the appearance of synapses, and some of these may
have been pivotal to the origin of the ursynapse. It is con-
ceivable that synapses would have evolved before axons
and dendrites, as an organism with a small number of
closely associated neurons would not need axonal pro-
jections for neuron to neuron communication. By the
same logic, without the synaptic specifications, axons or
dendrites would have no relevance. Similarly, synapse for-
mation would have evolved before other stages in neural
development including neuronal migration. Therefore we
posit that synapse formation is one of the exceptions to the
trend that ontogeny recapitulates phylogeny as, although
it is the final stage of neural development, it necessarily
evolved before the earlier developmental stages. We con-
sider the emergence of the synapse to be a crucial step
in the origin of the nervous system. Indeed, one could
argue that the evolution of synapses led to the evolution
of neurons, as a neuron is defined by synaptic connec-
tions. We refer to this model of the origins of the brain
as the ‘synapse first’ model. It may prove useful to profile
the expression of PSD and MASC genes during synaptic
development with respect to phylogenetic origin, in order
to test the ‘ontogeny recapitulates phylogeny’ evo–devo
paradigm in the context of synaptogenesis102,103.
Studies focusing on synapse evolution are at early stages,
but it is now an intellectually and technologically oppor-
tune time to launch into this topic. The focus has been on
excitatory glutamate receptor associated postsynaptic pro-
teins, as this family of proteins is well known and of cen-
tral importance to synaptic function. These studies should
be expanded to include all presynaptic and post synaptic
proteins. Most of our knowledge on the topic of synapse
evolution is based on data derived from comparative
genomics and proteomics. It will be necessary to expand
on these studies with thoughtful functional experiments.
Comparing the physiological and behavioural importance
of ancient and mammalian specific synaptic genes through
loss-of-function experiments in mice may aid in identi-
fying the functional roles for which they were selected.
Transgenic gain-of-function studies in organisms without
synapses might establish what combinations of synaptic
proteins resulted in the origin of the synapse. Finally, meth-
ods already applied to mice should be applied to humans
to see to what extent the evolution of the synapse has
contributed to the evolution of the human brain.
Figure 5 | MASC signalling diversity within the brain. a | Model depicting expression
variation of MASC (membrane‑associated guanylate kinase (MAGUK) associated
signalling complex) components throughout the deuterostome brain where different
combinations of MASC proteins are found in different synapses. Four different MASC
complexes comprised of various combinations of discrete upstream (coloured circles),
intermediate (red rectangles) and downstream (yellow triangles) proteins are expressed
in different regions of the brain (arrows). b | MASC components of ancient pre‑metazoan
origin (present in yeast), protostome origin and deuterostome origin are displayed in a
histogram that shows their percentage composition of genes of high expression
variability and low expression variability. MASC genes of deuterostome origin are
enriched for genes of high expression variability in the mouse brain: ancient genes are
uniformly expressed and recent genes are most variable. Figure is modified, with
permission from REF. 16 (2008) Macmillan Publishers Ltd. All rights reserved.
REVIEWS
710
|
OCTOBER 2009
|
VOLUME 10 www.nature.com/reviews/neuro
REVIEWS
© 2009 Macmillan Publishers Limited. All rights reserved
1. Kandel, E. R. The molecular biology of memory
storage: a dialogue between genes and synapses.
Science 294, 1030–1038 (2001).
2. Hebb, D. The Organization of Behavior; a
Neuropsychological Theory. (New York, Wiley, ,
1949).
3. Morgan, C. L. Animal Life and Intelligence (E. Arnold,
London, 1891).
4. Collins, M. O. et al. Molecular characterization and
comparison of the components and multiprotein
complexes in the postsynaptic proteome.
J. Neurochem. 97 (Suppl. 1), 16–23 (2006).
A comprehensive collation of postsynaptic protein
complexes
5. Bayes, A. & Grant, S. G. Neuroproteomics:
understanding the molecular organization and
complexity of the brain. Nature Rev. Neurosci. 10,
635–646 (2009).
A detailed review on the state of the art of
proteomics in neuroscience
6. Fernandez, E. et al. Targeted tandem affinity
purification of PSD-95 recovers core postsynaptic
complexes and schizophrenia susceptibility proteins.
Mol. Syst. Biol. 5, 269 (2009).
7. Pocklington, A. J., Cumiskey, M., Armstrong, J. D. &
Grant, S. G. The proteomes of neurotransmitter
receptor complexes form modular networks with
distributed functionality underlying plasticity and
behaviour. Mol. Syst. Biol. 2, 2006 0023 (2006).
8. Grant, S. G., Marshall, M. C., Page, K. L., Cumiskey,
M. A. & Armstrong, J. D. Synapse proteomics of
multiprotein complexes: en route from genes to
nervous system diseases. Hum. Mol. Genet. 14 Spec.
No. 2, R225–234 (2005).
9. Round, J. L. et al. Scaffold protein Dlgh1 coordinates
alternative p38 kinase activation, directing T cell
receptor signals toward NFAT but not NF-κB
transcription factors. Nature Immunol. 8, 154–161
(2007).
10. Tsunoda, S. & Zuker, C. S. The organization of INAD-
signaling complexes by a multivalent PDZ domain
protein in Drosophila photoreceptor cells ensures
sensitivity and speed of signaling. Cell Calcium 26,
165–171 (1999).
11. Wolfe, K. H. & Li, W. H. Molecular evolution meets the
genomics revolution. Nature Genet. 33 , S255–S265
(2003).
12. Darwin, C. On the Origin of Species by Means of
Natural Selection (John Murray, Albemarle Street,
London, 1859).
13. Sakarya, O. et al. A post-synaptic scaffold at the origin
of the animal kingdom. PLoS ONE 2, e506 (2007).
A detailed and careful study of 36 postsynaptic
gene families in poriferans and cnidarians
14. Garcia, M. L. & Strehler, E. E. Plasma membrane
calcium ATPases as critical regulators of calcium
homeostasis during neuronal cell function. Front.
Biosci. 4, D869–D882 (1999).
15. Muller, D., Buchs, P. A., Stoppini, L. & Boddeke, H.
Long-term potentiation, protein kinase C, and
glutamate receptors. Mol. Neurobiol. 5, 277–288
(1991).
16. Emes, R. D. et al. Evolutionary expansion and
anatomical specialization of synapse proteome
complexity. Nature Neurosci. 11, 799–806 (2008).
The first study to combine comparative genomics
and proteomics to address the origin and evolution
of the PSD and MASC complexes
17. Miyakawa, T. et al. Conditional calcineurin knockout
mice exhibit multiple abnormal behaviors related to
schizophrenia. Proc. Natl Acad. Sci. USA 100,
8987–8992 (2003).
18. Cyert, M. S. Genetic analysis of calmodulin and its
targets in Saccharomyces cerevisiae. Annu. Rev.
Genet. 35, 647–672 (2001).
19. King, N. Choanoflagellates. Curr. Biol. 15,
R113–R114 (2005).
20. Ruiz-Trillo, I. et al. The origins of multicellularity: a
multi-taxon genome initiative. Trends Genet. 23,
113–118 (2007).
21. King, N. The unicellular ancestry of animal
development. Dev. Cell 7, 313–325 (2004).
An informative review on the importance of
choanoflagellate research to our knowledge of the
evolution of multicellularity.
22. Leys, S. P., Rohksar, D. S. & Degnan, B. M.
Sponges. Curr. Biol. 15, R114–R115 (2005).
23. King, N. & Carroll, S. B. A receptor tyrosine kinase
from choanoflagellates: molecular insights into early
animal evolution. Proc. Natl Acad. Sci. USA 98,
15032–15037 (2001).
24. King, N., Hittinger, C. T. & Carroll, S. B. Evolution of
key cell signaling and adhesion protein families
predates animal origins. Science 3 01, 361–363
(2003).
25. Pincus, D., Letunic, I., Bork, P. & Lim, W. A. Evolution
of the phospho-tyrosine signaling machinery in
premetazoan lineages. Proc. Natl Acad. Sci. USA 10 5,
9680–9684 (2008).
26. Manning, G., Young, S. L., Miller, W. T. & Zhai, Y.
The protist, Monosiga brevicollis, has a tyrosine
kinase signaling network more elaborate and diverse
than found in any known metazoan. Proc. Natl Acad.
Sci. USA 105, 9674–9679 (2008).
27. Grant, S. G. et al. Impaired long-term potentiation,
spatial learning, and hippocampal development
in fyn mutant mice. Science 258, 1903–1910
(1992).
28. Salter, M. W. & Kalia, L. V. Src kinases: a hub for
NMDA receptor regulation. Nature Rev. Neurosci. 5,
317–328 (2004).
29. Abedin, M. & King, N. The premetazoan ancestry of
cadherins. Science 319, 946–948 (2008).
30. Arikkath, J. & Reichardt, L. F. Cadherins and
catenins at synapses: roles in synaptogenesis and
synaptic plasticity. Trends Neurosci. 31, 487–494
(2008).
31. Nichols, S. A., Dirks, W., Pearse, J. S. & King, N.
Early evolution of animal cell signaling and adhesion
genes. Proc. Natl Acad. Sci. USA 103, 12451–12456
(2006).
32. Jessell, T. M. & Sanes, J. R. Development. The decade
of the developing brain. Curr. Opin. Neurobiol. 10,
599–611 (2000).
33. Leys, S. P. & Degnan, B. M. Cytological basis of
photoresponsive behavior in a sponge larva. Biol. Bull.
201, 323–338 (2001).
34. Muller, W. E. Review: How was metazoan threshold
crossed? The hypothetical Urmetazoa. Comp. Biochem.
Physiol. A Mol. Integr. Physiol. 129, 433–460
(2001).
35. Perovic, S., Krasko, A., Prokic, I., Muller, I. M. &
Muller, W. E. Origin of neuronal-like receptors in
Metazoa: cloning of a metabotropic glutamate/GABA-
like receptor from the marine sponge Geodia
cydonium. Cell Tissue Res. 296, 395–404 (1999).
36. Leys, S. P., Mackie, G. O. & Meech, R. W. Impulse
conduction in a sponge. J. Exp. Biol. 202, 1139–1150
(1999).
37. Lin, C. H. et al. A role for the PI-3 kinase signaling
pathway in fear conditioning and synaptic plasticity in
the amygdala. Neuron 31, 841–851 (2001).
38. Sweatt, J. D. Protooncogenes subserve memory
formation in the adult CNS. Neuron 31, 671–674
(2001).
39. Elias, G. M. & Nicoll, R. A. Synaptic trafficking of
glutamate receptors by MAGUK scaffolding proteins.
Trends Cell Biol. 17, 343–352 (2007).
40. Kessels, H. W. & Malinow, R. Synaptic AMPA receptor
plasticity and behavior. Neuron 61, 340–350
(2009).
41. Nakazawa, K., McHugh, T. J., Wilson, M. A. &
Tonegawa, S. NMDA receptors, place cells and
hippocampal spatial memory. Nature Rev. Neurosci.
5, 361–372 (2004).
42. Craig, A. M. & Kang, Y. Neurexin-neuroligin signaling
in synapse development. Curr. Opin. Neurobiol. 17,
43–52 (2007).
43. Hedges, S. B., Dudley, J. & Kumar, S. TimeTree: a
public knowledge-base of divergence times among
organisms. Bioinformatics 22, 2971–2972 (2006).
44. De Robertis, E. M. Evo-devo: variations on ancestral
themes. Cell 132, 185–195 (2008).
45. Sanchez-Soriano, N. et al. Are dendrites in
Drosophila homologous to vertebrate dendrites? Dev.
Biol. 288, 126–138 (2005).
46. Walker, R. J., Brooks, H. L. & Holden-Dye, L.
Evolution and overview of classical transmitter
molecules and their receptors. Parasitology 113 ,
S3–S33 (1996).
47. Littleton, J. T. & Ganetzky, B. Ion channels and
synaptic organization: analysis of the Drosophila
genome. Neuron 26, 35–43 (2000).
48. Naisbitt, S. et al. Shank, a novel family of postsynaptic
density proteins that binds to the NMDA receptor/
PSD-95/GKAP complex and cortactin. Neuron 23,
569–582 (1999).
49. Sheng, M. & Kim, E. The Shank family of scaffold
proteins. J. Cell Sci. 11 3 , 1851–1856 (2000).
50. Gerrow, K. et al. A preformed complex of postsynaptic
proteins is involved in excitatory synapse
development. Neuron 49, 547–562 (2006).
51. Ryan, T. J., Emes, R. D., Grant, S. G. & Komiyama,
N. H. Evolution of NMDA receptor cytoplasmic
interaction domains: implications for organisation of
synaptic signalling complexes. BMC Neurosci. 9, 6
(2008).
This paper identifies the deuterostome specific
elongated NR2 intracellular domain and discusses
it implications for NMDA receptor evolution
52. te Velthuis, A. J., Admiraal, J. F. & Bagowski, C. P.
Molecular evolution of the MAGUK family in metazoan
genomes. BMC Evol. Biol. 7, 129 (2007).
53. Sprengel, R. et al. Importance of the intracellular
domain of NR2 subunits for NMDA receptor function
in vivo. Cell 92, 279–289 (1998).
This study establishes the indispensable role of
the vertebrate NR2 intracellular carboxy‑terminal
domain in NMDA receptor function
54. McLysaght, A., Hokamp, K. & Wolfe, K. H. Extensive
genomic duplication during early chordate evolution.
Nature Genet. 31, 200–204 (2002).
55. Okamura, Y. et al. Comprehensive analysis of the
ascidian genome reveals novel insights into the
molecular evolution of ion channel genes. Physiol.
Genomics 22, 269–282 (2005).
56. Cull-Candy, S., Brickley, S. & Farrant, M. NMDA
receptor subunits: diversity, development and
disease. Curr. Opin. Neurobiol. 11, 327–335
(2001).
57. Kim, M. J., Dunah, A. W., Wang, Y. T. & Sheng, M.
Differential roles of NR2A- and NR2B-containing
NMDA receptors in Ras-ERK signaling and AMPA
receptor trafficking. Neuron 46, 745–760 (2005).
58. Prybylowski, K. et al. The synaptic localization of
NR2B-containing NMDA receptors is controlled by
interactions with PDZ proteins and AP-2. Neuron 47,
845–857 (2005).
59. Townsend, M., Liu, Y. & Constantine-Paton, M.
Retina-driven dephosphorylation of the NR2A subunit
correlates with faster NMDA receptor kinetics at
developing retinocollicular synapses. J. Neurosci. 24,
11098–11107 (2004).
60. Li, B. S. et al. Regulation of NMDA receptors by
cyclin-dependent kinase-5. Proc. Natl Acad. Sci. USA
98, 12742–12747 (2001).
61. Sans, N. et al. A developmental change in NMDA
receptor-associated proteins at hippocampal
synapses. J. Neurosci. 20, 1260–1271 (2000).
62. Funke, L., Dakoji, S. & Bredt, D. S. Membrane-
associated guanylate kinases regulate adhesion and
plasticity at cell junctions. Annu. Rev. Biochem. 74,
219–245 (2005).
63. Wang, D. et al. CD3/CD28 costimulation-induced
NF-κB activation is mediated by recruitment of protein
kinase C-τ, Bcl10, and IκB kinase β to the
immunological synapse through CARMA1. Mol. Cell
Biol. 24, 164–171 (2004).
64. Migaud, M. et al. Enhanced long-term potentiation
and impaired learning in mice with mutant
postsynaptic density-95 protein. Nature 396,
433–439 (1998).
65. Cuthbert, P. C. et al. Synapse-associated protein
102/dlgh3 couples the NMDA receptor to specific
plasticity pathways and learning strategies.
J. Neurosci. 27, 2673–2682 (2007).
66. Carlisle, H. J., Fink, A. E., Grant, S. G. & O’Dell, T. J.
Opposing effects of PSD-93 and PSD-95 on long-term
potentiation and spike timing-dependent plasticity.
J. Physiol. 586, 5885–5900 (2008).
67. Xia, S. et al. NMDA receptors mediate olfactory
learning and memory in Drosophila. Curr. Biol. 15,
603–615 (2005).
The first study to clearly demonstrate that NMDA
receptors mediate learning in an invertebrate
organism.
68. Martyniuk, C. J., Aris-Brosou, S., Drouin, G., Cahn, J.
& Trudeau, V. L. Early evolution of ionotropic GABA
receptors and selective regimes acting on the
mammalian-specific τ and ε subunits. PLoS ONE 2,
e894 (2007).
69. Tsang, S. Y., Ng, S. K., Xu, Z. & Xue, H. The evolution
of GABAA receptor-like genes. Mol. Biol. Evol. 24,
599–610 (2007).
70. Simon, J., Wakimoto, H., Fujita, N., Lalande, M. &
Barnard, E. A. Analysis of the set of GABA(A) receptor
genes in the human genome. J. Biol. Chem. 279,
41422–41435 (2004).
71. Mitri, C., Parmentier, M. L., Pin, J. P., Bockaert, J. &
Grau, Y. Divergent evolution in metabotropic
glutamate receptors. A new receptor activated by an
endogenous ligand different from glutamate in insects.
J. Biol. Chem. 279, 9313–9320 (2004).
REVIEWS
NATURE R EVIEWS
|
NEUROSCIENCE VOLUME 10
|
OCTOBER 2009
|
711
FOCUS ON CNS EVOLUTION
© 2009 Macmillan Publishers Limited. All rights reserved
72. Dillon, J., Hopper, N. A., Holden-Dye, L. & O’Connor, V.
Molecular characterization of the metabotropic
glutamate receptor family in Caenorhabditis elegans.
Biochem. Soc. Trans. 34, 942–948 (2006).
73. Yanay, C., Morpurgo, N. & Linial, M. Evolution of
insect proteomes: insights into synapse organization
and synaptic vesicle life cycle. Genome Biol. 9, R27
(2008).
74. Fraser, H. B., Hirsh, A. E., Steinmetz, L. M., Scharfe, C.
& Feldman, M. W. Evolutionary rate in the protein
interaction network. Science 296, 750–752
(2002).
75. Hadley, D. et al. Patterns of sequence conservation in
presynaptic neural genes. Genome Biol. 7, R105
(2006).
76. Winter, E. E., Goodstadt, L. & Ponting, C. P. Elevated
rates of protein secretion, evolution, and disease
among tissue-specific genes. Genome Res. 14, 54–61
(2004).
77. Sugino, K. et al. Molecular taxonomy of major
neuronal classes in the adult mouse forebrain.
Nature Neurosci. 9, 99–107 (2006).
78. Ohno, S. Evolution by gene duplication (Allen &
Unwin, 1970).
79. Thompson, C. L. et al. Genomic anatomy of the
hippocampus. Neuron 60, 1010–1021 (2008).
80. Heiman, M. et al. A translational profiling approach
for the molecular characterization of CNS cell types.
Cell 135, 738–748 (2008).
81. Doyle, J. P. et al. Application of a translational
profiling approach for the comparative analysis of CNS
cell types. Cell 135, 749–762 (2008).
82. Varki, A., Geschwind, D. H. & Eichler, E. E. Explaining
human uniqueness: genome interactions with
environment, behaviour and culture. Nature Rev.
Genet. 9, 749–763 (2008).
83. Mekel-Bobrov, N. & Lahn, B. T. What makes us human:
revisiting an age-old question in the genomic era.
J. Biomed. Discov. Collab. 1, 18 (2006).
84. Sikela, J. M. The jewels of our genome: the search for
the genomic changes underlying the evolutionarily
unique capacities of the human brain. PLoS Genet. 2,
e80 (2006).
85. Gilbert, S. L., Dobyns, W. B. & Lahn, B. T. Genetic links
between brain development and brain evolution.
Nature Rev. Genet. 6, 581–590 (2005).
86. Dorus, S. et al. Accelerated evolution of nervous
system genes in the origin of Homo sapiens. Cell 11 9 ,
1027–1040 (2004).
87. Gardner, P. P. & Vinther, J. Mutation of miRNA target
sequences during human evolution. Trends Genet. 24,
262–265 (2008).
88. Andres, A. M. et al. Positive selection in MAOA gene is
human exclusive: determination of the putative amino
acid change selected in the human lineage.
Hum. Genet. 115 , 377–386 (2004).
89. Caspi, A. et al. Role of genotype in the cycle of
violence in maltreated children. Science 297,
851–854 (2002).
90. Buckholtz, J. W. & Meyer-Lindenberg, A. MAOA and
the neurogenetic architecture of human aggression.
Trends Neurosci. 31, 120–129 (2008).
91. Ding, Y. C. et al. Evidence of positive selection acting
at the human dopamine receptor D4 gene locus.
Proc. Natl Acad. Sci. USA 99, 309–314 (2002).
92. Swanson, J. et al. Attention deficit/hyperactivity
disorder children with a 7-repeat allele of the
dopamine receptor D4 gene have extreme behavior
but normal performance on critical neuropsychological
tests of attention. Proc. Natl Acad. Sci. USA 97,
4754–4759 (2000).
93. Wang, E. et al. The genetic architecture of selection at
the human dopamine receptor D4 (DRD4) gene locus.
Am. J. Hum. Genet. 74, 931–944 (2004).
94. Lo, W. S. et al. Association of SNPs and haplotypes
in GABAA receptor β2 gene with schizophrenia.
Mol. Psychiatry 9, 603–608 (2004).
95. Lo, W. S. et al. Positive selection within the
Schizophrenia-associated GABA(A) receptor β2 gene.
PLoS ONE 2, e462 (2007).
96. Zhao, C. et al. Two isoforms of GABA(A) receptor β2
subunit with different electrophysiological properties:
Differential expression and genotypical correlations in
schizophrenia. Mol. Psychiatry 11 , 1092–1105
(2006).
97. Crow, T. J. The ‘big bang’ theory of the origin of
psychosis and the faculty of language. Schizophr Res.
102, 31–52 (2008).
98. Pearlson, G. D. & Folley, B. S. Schizophrenia,
psychiatric genetics, and Darwinian psychiatry: an
evolutionary framework. Schizophr Bull. 34, 722–733
(2008).
99. Nieoullon, A. & Coquerel, A. Dopamine: a key
regulator to adapt action, emotion, motivation and
cognition. Curr. Opin. Neurol. 16 (Suppl 2), 3–9
(2003).
100. Caceres, M. et al. Elevated gene expression levels
distinguish human from non-human primate brains.
Proc. Natl Acad. Sci. USA 100, 13030–13035
(2003).
101. Enard, W. et al. Intra- and interspecific variation in
primate gene expression patterns. Science 296,
340–343 (2002).
102. Valor, L. M., Charlesworth, P., Humphreys, L.,
Anderson, C. N. & Grant, S. G. Network activity-
independent coordinated gene expression program
for synapse assembly. Proc. Natl Acad. Sci. USA 104,
4658–4663 (2007).
103. Mattson, M. P. & Bruce-Keller, A. J.
Compartmentalization of signaling in neurons:
evolution and deployment. J. Neurosci. Res. 58,
2–9 (1999).
104. Hedges, S. B. & Kumar, S. The Timetree of Life
(Oxford University Press, Oxford, 2009).
105. Kohr, G. NMDA receptor function: subunit
composition versus spatial distribution. Cell Tissue
Res. 326, 439–446 (2006).
106. Dechant, R. & Peter M. Nutrient signals driving cell
growth. Curr. Opin. Cell Biol. 20, 678–687 (2008).
107. Park, J. I., Grant, C. M. & Dawes, I. W. The high-affinity
cAMP phosphodiesterase of Saccharomyces
cerevisiae is the major determinant of cAMP levels in
stationary phase: involvement of different branches of
the Ras-cyclic AMP pathway in stress responses.
Biochem. Biophys. Res. Commun. 327, 311–319
(2005).
108. Kraus, P. R. & Heitman J. Coping with stress:
calmodulin and calcineurin in model and pathogenic
fungi. Biochem. Biophys. Res. Commun. 311 ,
1151–1157 (2003).
109. Zeitlinger, J. et al. Program-specific distribution of a
transcription factor dependent on partner
transcription factor and MAPK signaling. Cell 11 3,
395–404 (2003).
110. Slessareva, J. E. & Dohlman H. G. G protein signaling
in yeast: new components, new connections, new
compartments. Science 314, 1412–1413 (2006).
111. Sato, T. K. Vam7p, a SNAP-25-like molecule, and
Vam3p, a syntaxin homolog, function together in yeast
vacuolar protein trafficking. Mol. Cell Biol. 18, 5308–
5319 (1998).
112. Denis, V. & Cyert M. S. Internal Ca2+ release in yeast is
triggered by hypertonic shock and mediated by a TRP
channel homologue. J. Cell Biol. 156, 29–34 (2002).
113. Li, W. et al. Signaling properties of a non-metazoan Src
kinase and the evolutionary history of Src negative
regulation. J. Biol. Chem. 283, 15491–15501 (2008).
114. Cai, X. Unicellular Ca2+ signaling ‘toolkit’ at the origin
of metazoa. Mol. Biol. Evol. 25, 1357–1361 (2008).
115. Baines, A. J. Evolution of spectrin function in
cytoskeletal and membrane networks. Biochem. Soc.
Trans. 37, 796–803 (2009).
116. Kosik, K. S. Exploring the early origins of the synapse
by comparative genomics. Biol. Lett (2008).
117. Adell, T. et al. Evolution of metazoan cell junction
proteins: the scaffold protein MAGI and the
transmembrane receptor tetraspanin in the
demosponge Suberites domuncula. J. Mol. Evol. 59,
41–50 (2004).
118. Nakazawa, K. et al. NMDA receptors, place cells and
hippocampal spatial memory. Nature Re v. Neurosci.
5, 361–372 (2004).
119. Kessels, H. W. & Malinow, R. Synaptic AMPA receptor
plasticity and behavior. Neuron 61, 340–350
(2009).
120. Jane, D. E. Kainate receptors: pharmacology, function
and therapeutic potential. Neuropharmacology 56,
90–113 (2009).
121. Lee, J. et al. Pre- and post-synaptic mechanisms of
synaptic strength homeostasis revealed by slowpoke
and shaker K+ channel mutations in Drosophila.
Neuroscience 154, 1283–1296 (2008).
122. Calin-Jageman, I. et al. Erbin enhances voltage-
dependent facilitation of Cav1.3 Ca2+ channels
through relief of an autoinhibitory domain in the
Cav1.3 α1 subunit. J. Neurosci. 27, 1374–1385
(2007).
123. Atasoy, D. et al. Deletion of CASK in mice is lethal and
impairs synaptic function. Proc. Natl Acad. Sci. USA
104, 2525–2530 (2007).
Acknowledgements
We thank members of the Genes to Cognition Programme for
useful discussions. T.J.R. was supported by a Wellcome Trust
Ph.D. Studentship at time of writing.
DATABASES
UniProtKB: http://www.uniprot.org
CRIPT | Dlg | ERK2 | GKAP | GNB5 | GRIP | Homer | NF1 | PKC |
PMCA
FURTHER INFORMATION
Seth G. N. Grant’s homepage:http://www.sanger.ac.uk/
Teams/faculty/grant/
Timetree: http://www.timetree.org/
SUPPLEMENTARY INFORMATION
See online article: S1 (table) | S2 (box) | S3 (box)
ALL LINKS ARE ACTIVE IN THE ONLINE PDF
REVIEWS
712
|
OCTOBER 2009
|
VOLUME 10 www.nature.com/reviews/neuro
REVIEWS
© 2009 Macmillan Publishers Limited. All rights reserved
Biographies
Tomás Ryan received a B.A. in human genetics from the University of
Dublin, Trinity College (Ireland). He was awarded a Wellcome Trust
Ph.D. Studentship to study in the research group of Seth Grant at the
Wellcome Trust Sanger Institute and Darwin College, University of
Cambridge (UK), where he was a final year graduate student at time
of writing. His thesis research focused on the generation and analysis
of transgenic mouse models of NMDA (N-methyl--aspartate) recep-
tor molecular evolution. He is currently a Junior Research Fellow at
Wolfson College, University of Cambridge.
Seth Grant is a neuroscientist best known for his work using mouse
genetics and synapse proteomics to study plasticity and learning.
This work has uncovered unexpectedly high molecular complexity
in synapse protein complexes and the postsynaptic density providing
avenues to study synapse evolution and brain disease. He received
degrees in physiology, medicine and surgery from the University of
Sydney, Australia, and postdoctoral training at Cold Spring Harbor
Laboratory, USA, and with Eric Kandel at Columbia University,
USA. He currently heads the Genes to Cognition Programme
at the Wellcome Trust Sanger Institute in Cambridge UK and is
Professor of Molecular Neuroscience at Edinburgh and Cambridge
Universities, UK.
Online summary
• The molecular composition of the synapse has recently been
proved to be useful for studying the evolution of the brain.
• Synapse proteomics data sets, such as those of the postsynaptic
density (PSD) and associated protein complexes when combined
with comparative genomics have provided unprecedented insights
into the evolution of synapses.
• The PSD that is found in organisms with nervous systems has
evolved from an ancient protosynaptic core that exists in uni-
cellular organisms and multicellular organisms without nervous
systems.
• Comparisons of vertebrate PSD and synaptogenesis genes with
orthologues from sponges and cnidarians open an avenue for
speculating as to what may have contributed to the origin of the
first synapse.
• Comparative proteomics has shown that vertebrate excitatory
synapses have evolved to be significantly more complex than
invertebrates.
TOC
000 The origin and evolution of synapses
Tomás J. Ryan and Seth G. N. Grant
Tracing the phylogeny of the molecular components
of synapses, Ryan and Grant speculate on the core
components of the last common ancestor of all
synapses and posit that the diversification of upstream
signalling components contributed to increased
signalling complexity later in evolution.
ONLINE ONLY
© 2009 Macmillan Publishers Limited. All rights reserved
... gap junctional) contacts between them. The appearance and evolution of the synapse is another extremely interesting topic (see, for example, Ryan & Grant, 2009). Importantly, the main molecules needed for synapse formation had already evolved in single cell organisms. ...
... Indeed, the receptors for neurotransmitters appeared previously in bacteria (pentameric receptors and glutamate receptors) and in protozoa (purinoceptors, which are present in amoeba). Similarly, ion channels and ion pumps and many molecules of the post-synaptic density appeared very early in evolution in prokaryotes and early unicellular eukaryotes (yeast and amoeba; see Case et al., 2007;Ryan & Grant, 2009). The epithelial cells that give rise to ancestral neurones are also endowed with exocytotic machinery underlying vesicular release of proto-neurotransmitters. ...
... ▪ One could equally (perhaps necessarily) argue for an empirical study of the consciousness of artificial intelligence utilizing the MoM consciousness theory. o Additionally, if one can conceptualize gradual descent or gradient descent (i.e., continuous complexification from optimization to a changing environment) as integral to and the very underpinnings of the evolutionary-derived and evolutionary-directed process of inclusive fitness, a brain or likely even a nervous system, which, ergo, is not required for complex behaviors to be associated with consciousness (Ryan and Grant, 2009). ...
Article
Full-text available
In simple terms, consciousness is constituted by multiple goals for action and the continuous adjudication of such goals to implement action, which is referred to as the maps of meaning (MoM) consciousness theory. The MoM theory triangulates through three parallel corollaries: action (behavior), mechanism (morphology/pathophysiology), and goals (teleology). (1) An organism’s consciousness contains fluid, nested goals. These goals are not intentionality, but intersectionality, via the Darwinian byproduct of embodiment meeting the world, i.e., Darwinian inclusive fitness or randomization and then survival of the fittest. (2) These goals are formed via a gradual descent under inclusive fitness and are the abstraction of a “match” between the evolutionary environment and the organism. (3) Human consciousness implements the brain efficiency hypothesis, genetics, epigenetics, and experience-crystallized efficiencies, not necessitating best or objective but fitness, i.e., perceived efficiency based on one’s adaptive environment. These efficiencies are objectively arbitrary but determine the operation and level of one’s consciousness, termed as extreme thrownness. (4) Since inclusive fitness drives efficiencies in the physiologic mechanism, morphology, and behavior (action) and originates one’s goals, embodiment is necessarily entangled to human consciousness as it is at the intersection of mechanism or action (both necessitating embodiment) occurring in the world that determines fitness. (5) Perception is the operant process of consciousness and is the de facto goal adjudication process of consciousness. Goal operationalization is fundamentally efficiency-based via one’s unique neuronal mapping as a byproduct of genetics, epigenetics, and experience. (6) Perception involves information intake and information discrimination, equally underpinned by efficiencies of inclusive fitness via extreme thrownness. Perception is not a ‘frame rate’ but Bayesian priors of efficiency based on one’s extreme thrownness. (7) Consciousness and human consciousness are modular (i.e., a scalar level of richness, which builds up like building blocks) and dimensionalized (i.e., cognitive abilities become possibilities as the emergent phenomena at various modularities such as the stratified factors in factor analysis). (8) The meta dimensions of human consciousness seemingly include intelligence quotient, personality (five-factor model), richness of perception intake, and richness of perception discrimination, among other potentialities. (9) Future consciousness research should utilize factor analysis to parse modularities and dimensions of human consciousness and animal models.
... Fast synaptic transmission in the nervous system is mediated by electrical and chemical synapses, with each type imparting distinct modes of communication [1][2][3]. Electrical synapses directly couple neurons via interneuronal gap junction (GJ) channels, whereas chemical synapses utilize neurotransmitter release/reception for signaling. Despite unique structures and functions, both synaptic types are supported by locally-assembled, cytoplasmic, associated proteins that regulate the development and function of transmission. ...
Article
Full-text available
Electrical synapses are neuronal gap junction (GJ) channels associated with a macromolecular complex called the electrical synapse density (ESD), which regulates development and dynamically modifies electrical transmission. However, the proteomic makeup and molecular mechanisms utilized by the ESD that direct electrical synapse formation are not well understood. Using the Mauthner cell of zebrafish as a model, we previously found that the intracellular scaffolding protein ZO1b is a member of the ESD, localizing postsynaptically, where it is required for GJ channel localization, electrical communication, neural network function, and behavior. Here, we show that the complexity of the ESD is further diversified by the genomic structure of the ZO1b gene locus. The ZO1b gene is alternatively initiated at three transcriptional start sites resulting in isoforms with unique N-termini that we call ZO1b-Alpha, -Beta, and -Gamma. We demonstrate that ZO1b-Beta and ZO1b-Gamma are broadly expressed throughout the nervous system and localize to electrical synapses. By contrast, ZO1b-Alpha is expressed mainly non-neuronally and is not found at synapses. We generate mutants in all individual isoforms, as well as double mutant combinations in cis on individual chromosomes, and find that ZO1b-Beta is necessary and sufficient for robust GJ channel localization. ZO1b-Gamma, despite its localization to the synapse, plays an auxiliary role in channel localization. This study expands the notion of molecular complexity at the ESD, revealing that an individual genomic locus can contribute distinct isoforms to the macromolecular complex at electrical synapses. Further, independent scaffold isoforms have differential contributions to developmental assembly of the interneuronal GJ channels. We propose that ESD molecular complexity arises both from the diversity of unique genes and from distinct isoforms encoded by single genes. Overall, ESD proteomic diversity is expected to have critical impacts on the development, structure, function, and plasticity of electrical transmission.
... Many molecular components of synapses originated long before neurons (T#15-16). Specifically, a substantial fraction of all the molecules typically associated with synapse formation and function have direct homologs (orthologs) in unicellular organisms such as yeast and choanoflagellates, as well as sponges (Ryan and Grant, 2009). This discovery implies that the evolution of chemical synapses involved "deploying existing genes and modules in new cellular contexts" (Colgren and Burkhardt, 2022, 787). ...
Article
Full-text available
Neuroscience courses can be enriched by including an evolutionary perspective. To that end, this essay identifies several concepts critical to understanding nervous system evolution and offers numerous examples that can be used to illustrate those concepts. One critical concept is that the distribution of features among today’s species can be used to reconstruct a feature’s evolutionary history, which then makes it possible to distinguish cases of homology from convergent evolution. Another key insight is that evolution did not simply add new features to old nervous systems, leaving the old features unchanged. Instead, both new and old features have changed, and they generally did so along divergent trajectories in different lineages, not in a linear sequence. Some changes in nervous system organization can be linked to selective pressures (i.e, adaptation), especially if they occurred convergently in different lineages. However, nervous system evolution has also been subject to various constraints, which is why many neural features are, in a sense, suboptimal. An overarching theme is that evolution has brought forth tremendous diversity across all levels of the nervous system and at all levels of organization, from molecules to neural circuits and behavior. This diversity provides excellent research opportunities, but it can also complicate the extrapolation of research findings across species.
... For example, the number of variant forms of particular neurotransmitter receptors may differ in the two groups. The NMDA receptor has two major forms found in both vertebrates and invertebrates, but vertebrates have a component of the receptor (the NR2B subunit) that is not found in insects (Ryan et al. 2008;Emes et al. 2008;Ryan and Grant 2009). ...
Chapter
An examination of how widely distributed and specialized activities of the brain are flexibly and effectively coordinated. A fundamental shift is occurring in neuroscience and related disciplines. In the past, researchers focused on functional specialization of the brain, discovering complex processing strategies based on convergence and divergence in slowly adapting anatomical architectures. Yet for the brain to cope with ever-changing and unpredictable circumstances, it needs strategies with richer interactive short-term dynamics. Recent research has revealed ways in which the brain effectively coordinates widely distributed and specialized activities to meet the needs of the moment. This book explores these findings, examining the functions, mechanisms, and manifestations of distributed dynamical coordination in the brain and mind across different species and levels of organization. The book identifies three basic functions of dynamic coordination: contextual disambiguation, dynamic grouping, and dynamic routing. It considers the role of dynamic coordination in temporally structured activity and explores these issues at different levels, from synaptic and local circuit mechanisms to macroscopic system dynamics, emphasizing their importance for cognition, behavior, and psychopathology. ContributorsEvan Balaban, György Buzsáki, Nicola S. Clayton, Maurizio Corbetta, Robert Desimone, Kamran Diba, Shimon Edelman, Andreas K. Engel, Yves Fregnac, Pascal Fries, Karl Friston, Ann Graybiel, Sten Grillner, Uri Grodzinski, John-Dylan Haynes, Laurent Itti, Erich D. Jarvis, Jon H. Kaas, J.A. Scott Kelso, Peter König, Nancy J. Kopell, Ilona Kovács, Andreas Kreiter, Anders Lansner, Gilles Laurent, Jörg Lücke, Mikael Lundqvist, Angus MacDonald, Kevan Martin, Mayank Mehta, Lucia Melloni, Earl K. Miller, Bita Moghaddam, Hannah Monyer, Edvard I. Moser, May-Britt Moser, Danko Nikolic, William A. Phillips, Gordon Pipa, Constantin Rothkopf, Terrence J. Sejnowski, Steven M. Silverstein, Wolf Singer, Catherine Tallon-Baudry, Roger D. Traub, Jochen Triesch, Peter Uhlhaas, Christoph von der Malsburg, Thomas Weisswange, Miles Whittington, Matthew Wilson
... 6 In this way, the main element responsible for the prokaryote's and mainly the bacteria's learning and memory abilities might be defined as the postsynaptic terminal only as depicted by various researchers in analogy to realization of human synapses with two-dimensional materials, even though there are no synapses in the bacteria. [7][8][9] Bacteria's learning and memory abilities are modeled as a sensory membrane, which is surrounding the bacteria's cell. Many bacteria are small in size and has large surface areas. ...
Article
We mimic bacterial learning and memory abilities in tungsten based two-sided single layers of WSeO, WSeS, WSeSe, and WSeTe, where the thickness of the material represents the growth in time. We aim to create a quantum memristor like system to show learning and memory abilities of bacteria with time while it grows. Its governing equation is derived, and it was found to be similar to the gene regulatory response model of the bacteria. Polarization is calculated from the Berry phase theory to plot its relation with the degauss parameter in time, which leads to bow-tie like memory switches similar to phase-change memories. We attribute this behavior to a specific bacteria, that is, Geobacter metallireducens. Mimicking bacteria’s learning and memory abilities will open a way to merge physical intelligence with quantum computing computationally.
Article
How nervous systems evolved is a central question in biology. A diversity of synaptic proteins is thought to play a central role in the formation of specific synapses leading to nervous system complexity. The largest animal genes, often spanning hundreds of thousands of base pairs, are known to be enriched for expression in neurons at synapses and are frequently mutated or misregulated in neurological disorders and diseases. Although many of these genes have been studied independently in the context of nervous system evolution and disease, general principles underlying their parallel evolution remain unknown. To investigate this, we directly compared orthologous gene sizes across eukaryotes. By comparing relative gene sizes within organisms, we identified a distinct class of large genes with origins predating the diversification of animals and, in many cases, the emergence of neurons as dedicated cell types. We traced this class of ancient large genes through evolution and found orthologs of the large synaptic genes potentially driving the immense complexity of metazoan nervous systems, including in humans and cephalopods. Moreover, we found that while these genes are evolving under strong purifying selection, as demonstrated by low dN/dS ratios, they have simultaneously grown larger and gained the most isoforms in animals. This work provides a new lens through which to view this distinctive class of large and multi-isoform genes and demonstrates how intrinsic genomic properties, such as gene length, can provide flexibility in molecular evolution and allow groups of genes and their host organisms to evolve toward complexity.
Article
We offer our opinion on the benefits of integration of insights from active matter physics with principles of regulatory interactions and control to develop a field we term "smart active matter". This field can provide insight into important principles in living systems as well as aid engineering of responsive, robust and functional collectives.
Chapter
Neurons are excitable cells. They use ions and electrical signaling to talk to each to other, and when they talk to each other, neurons control behavior. The Oxford Handbook of Neuronal Ion Channels is an accessible reference describing the nature and properties of ion channels in neurons. The book explains how ion channels open and close, how they can be selective for specific ions, and how they give rise to action potentials. There are in-depth chapters discussing specific classes of ion channels: potassium channels, sodium channels, neurotransmitter-gated ion channels, and other specialized channels. Throughout the handbook, important insight is provided on the contribution ion channels make to neuronal excitability and to synaptic transmission. The handbook goes further to discuss a group of human diseases such as epilepsy, pain, and migraines that can be caused by ion channel dysfunction called channelopathies. For neuroscientists, biophysicists, and neuropharmacologists, this handbook is a valuable reference of ion channel biology and function.
Article
Full-text available
The membrane-associated guanylate kinases [Chapsyn-110/postsynaptic density-93 (PSD-93), synapse-associated protein-90 (SAP-90)/PSD-95, and SAP-102] are believed to cluster and anchor NMDA receptors at the synapse and to play a role in signal transduction. We have investigated the developmental changes in expression of these proteins in rat hippocampus using biochemical analyses and quantitative immunogold electron microscopy. At postnatal day 2 (P2), SAP-102 was highly expressed, whereas PSD-93 and PSD-95 were low. SAP-102 expression increased during the first week, stayed stable through P35, and showed a reduced expression at 6 months. From P2 through 6 months, PSD-93 and PSD-95 increased. For PSD-95, the percent of labeled synapses increased almost threefold with age, whereas the number of gold particles per labeled synapse did not change significantly, suggesting that the increase in PSD-95 is attributable primarily to an increase in the number of synapses containing PSD-95. In contrast, for SAP-102, both percent labeled synapses and the number of gold particles per labeled synapse decreased during this time. From Western blots of hippocampus and immunogold analysis of CA1 synapses, the high expression of NR2B at P2 coincides with the high level of SAP-102 at synapses, whereas the later expression of NR2A coincides with that of PSD-93 and PSD-95. To determine whether the changes in PSD-93/95 and SAP-102 reflect preferred associations with NR2A and NR2B, respectively, we measured co-immunoprecipitation in the adult hippocampus. These studies suggest that there is a preference for complexes of NR2A/PSD-93/95 and NR2B/SAP-102. These results indicate that individual receptor-associated proteins may have specific functions that are critical to synapse development.
Article
The evolutionary history of life includes two primary components: phylogeny and timescale. Phylogeny refers to the branching order (relationships) of species or other taxa within a group and is crucial for understanding the inheritance of traits and for erecting classifications. However, a timescale is equally important because it provides a way to compare phylogeny directly with the evolution of other organisms and with planetary history such as geology, climate, extraterrestrial impacts, and other features. The Timetree of Life is the first reference book to synthesize the wealth of information relating to the temporal component of phylogenetic trees. In the past, biologists have relied exclusively upon the fossil record to infer an evolutionary timescale. However, recent revolutionary advances in molecular biology have made it possible to not only estimate the relationships of many groups of organisms, but also to estimate their times of divergence with molecular clocks. The routine estimation and utilization of these so-called ‘time-trees’ could add exciting new dimensions to biology including enhanced opportunities to integrate large molecular data sets with fossil and biogeographic evidence (and thereby foster greater communication between molecular and traditional systematists). They could help estimate not only ancestral character states but also evolutionary rates in numerous categories of organismal phenotype; establish more reliable associations between causal historical processes and biological outcomes; develop a universally standardized scheme for biological classifications; and generally promote novel avenues of thought in many arenas of comparative evolutionary biology. This authoritative reference work brings together, for the first time, experts on all major groups of organisms to assemble a timetree of life. The result is a comprehensive resource on evolutionary history which will be an indispensable reference for scientists, educators, and students in the life sciences, earth sciences, and molecular biology. For each major group of organism, a representative is illustrated and a timetree of families and higher taxonomic groups is shown. Basic aspects of the evolutionary history of the group, the fossil record, and competing hypotheses of relationships are discussed. Details of the divergence times are presented for each node in the timetree, and primary literature references are included.
Article
Dopamine is expressed in restricted brain areas involved in numerous integrative functions contributing to automated behaviours that are highly adaptive. During ontogenesis, dopamine can have a trophic action, which influences cortical specification directly, especially in prefrontal areas. Such a role is attested by the close relationships existing between the development of dopamine cortical innervation and cognitive abilities. Interestingly, such a proposal is reinforced by phylogenetic data. During the last stages of evolution in mammals, the characteristic extension of dopamine cortical innervation is also correlated with the development of cognitive capacities. More generally, the contribution of dopamine will be to increase the processing of cortical information through basal ganglia, either during the course of evolution or development. In this respect, dysmaturation suspected in schizophrenia and attention deficit-hyperactivity disorder in children can emphasize such a defect in basal ganglia processing. Remarkably, these diseases are improved by dopamine antagonists and agonists, respectively. In this line of evidence, experimental studies showed that selective lesions of the dopamine neurones in rats or primates can actually provide motor, limbic and cognitive deficits, in later cases especially when the mesocorticolimbic dopamine pathways are altered. Data resulting from lesion studies also showed significant alteration in attentional processes, thus raising the question of the direct involvement of dopamine in regulating attention. Dopamine can act as a powerful regulator and integrator of different aspects of brain functions. For example, in Parkinson's disease, besides motor impairment, dopamine degeneration is also expressed by alterations of both limbic, executive and cognitive functions, both improved by dopamine receptor agonists and dopa therapy. Dopamine has thus to be considered as a key regulator that contributes to behavioural adaptation and to the anticipatory processes necessary for preparing voluntary action consequent upon intention. In this respect, the alteration of dopamine transmission with age could contribute to cognitive impairment. Therefore, to normalize dopamine transmission pharmacologically could actually improve the cognitive and limbic deficits during normal aging, as it does in psychiatric and neurological disorders.
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
Characterization of the composition of the postsynaptic proteome (PSP) provides a framework for understanding the overall organization and function of the synapse in normal and pathological conditions. We have identified 698 proteins from the postsynaptic terminal of mouse CNS synapses using a series of purification strategies and analysis by liquid chromatography tandem mass spectrometry and large-scale immunoblotting. Some 620 proteins were found in purified postsynaptic densities (PSDs), nine in AMPA-receptor immuno-purifications, 100 in isolates using an antibody against the NMDA receptor subunit NR1, and 170 by peptide-affinity purification of complexes with the C-terminus of NR2B. Together, the NR1 and NR2B complexes contain 186 proteins, collectively referred to as membrane-associated guanylate kinase-associated signalling complexes. We extracted data from six other synapse proteome experiments and combined these with our data to provide a consensus on the composition of the PSP. In total, 1124 proteins are present in the PSP, of which 466 were validated by their detection in two or more studies, forming what we have designated the Consensus PSD. These synapse proteome data sets offer a basis for future research in synaptic biology and will provide useful information in brain disease and mental disorder studies.