1. Trang chủ >
  2. Khoa Học Tự Nhiên >
  3. Sinh học >

8 Conclusions, the Concept of a Synaptic Code and Future Directions

Bạn đang xem bản rút gọn của tài liệu. Xem và tải ngay bản đầy đủ của tài liệu tại đây (14.07 MB, 451 trang )


362



A.A. Chubykin



variations of different neurexin isoforms in different cell types and possibly even

individual cells.

Consequently, more detailed characterization of the various neurexin and

neuroligin isoforms and their expression patterns will be essential for our

understanding of the postulated neurexin–neuroligin-dependent synaptic

code. Understanding this neurexin–neuroligin-dependent code might help us

to learn how the brain is wired by specific connections between different cell

types and by individual neurexin–neuroligin pairs, which specify different types

of synapses. More modern, high-throughput, sensitive assays will shed more

light on synaptic complexity in vertebrate brains. One of the predictions would

be that the expression of neurexins–SS#4 is mostly in excitatory neurons, and

neurexins+SS#4 in inhibitory neurons. An individual neuron may express

either neurexin-SS#4 or neurexin+SS#4 isoform, but not both. The SS#4 splice

site could represent one of the potential markers of the neuronal type. Consequently, investigation of how the splicing is regulated becomes very important,

particularly, identification of the cis elements in the neurexin and neuroligin

genes, as well as the trans-acting splicing factors interacting with these elements.

A growing body of evidence suggests that the activity-dependent regulation

of neurexins and neuroligins at both excitatory and inhibitory synapses may be

involved in synaptic plasticity. However, we will need a considerably better and

more detailed knowledge of how synaptic activity is regulated by neurexins and

neuroligins at the molecular level, before we will understand the role of these

molecules in higher cognitive functions, such as memory formation and

behavior.

Acknowledgments The author thanks Thomas C. Suădhof and Dilja Krueger for their helpful

comments on the manuscript.



References

Alarcon M, Abrahams BS, Stone JL et al. (2008) Linkage, association, and gene-expression

analyses identify CNTNAP2 as an autism-susceptibility gene, Am J Hum Genet

82:150–159

Arac D, Boucard AA, Ozkan E et al. (2007) Structures of neuroligin-1 and the neuroligin-1/

neurexin-1 beta complex reveal specific protein-protein and protein-Ca2+ interactions,

Neuron 56:992–1003

Atlas D (2001) Functional and physical coupling of voltage-sensitive calcium channels with

exocytotic proteins: ramifications for the secretion mechanism, J Neurochem 77:972–985

Bakkaloglu B, O’Roak BJ, Louvi A et al. (2008) Molecular cytogenetic analysis and resequencing of contactin associated protein-like 2 in autism spectrum disorders, Am J Hum Genet

82:165–173

Barresi R and Campbell KP (2006) Dystroglycan: from biosynthesis to pathogenesis of

human disease, J Cell Sci 119:199–207

Baumgartner S, Littleton JT, Broadie K et al. (1996) A Drosophila neurexin is required for

septate junction and blood-nerve barrier formation and function, Cell 87:1059–1068



17



Neurexins and Neuroligins



363



Bear MF, Huber KM and Warren ST (2004) The mGluR theory of fragile X mental retardation, Trends Neurosci 27:370377

Biederer T and Suădhof TC (2001) CASK and protein 4.1 support F-actin nucleation on

neurexins, J Biol Chem 276:47869–47876

Blasi F, Bacchelli E, Pesaresi G et al. (2006) Absence of coding mutations in the X-linked

genes neuroligin 3 and neuroligin 4 in individuals with autism from the IMGSAC collection, Am J Med Genet B Neuropsychiatr Genet 141:220–221

Bolliger MF, Frei K, Winterhalter KH et al. (2001) Identification of a novel neuroligin in

humans which binds to PSD-95 and has a widespread expression, Biochem J 356:581–588

Bolliger MF, Pei J, Maxeiner S et al. (2008) Unusually rapid evolution of Neuroligin-4 in

mice, Proc Natl Acad Sci USA 105:6421–6426

Boucard AA, Chubykin AA, Comoletti D et al. (2005) A splice code for trans-synaptic cell

adhesion mediated by binding of neuroligin 1 to alpha- and beta-neurexins, Neuron

48:229–236

Budreck EC and Scheiffele P (2007) Neuroligin-3 is a neuronal adhesion protein at GABAergic and glutamatergic synapses, Eur J Neurosci 26:1738–1748

Butz S, Okamoto M and Suădhof TC (1998) A tripartite protein complex with the potential to

couple synaptic vesicle exocytosis to cell adhesion in brain, Cell 94:773–782

Chen X, Liu H, Shim AH et al. (2008) Structural basis for synaptic adhesion mediated by

neuroligin–neurexin interactions, Nat Struct Mol Biol 15:50–56

Chih B, Afridi SK, Clark L et al. (2004) Disorder-associated mutations lead to functional

inactivation of neuroligins, Hum Mol Genet 13:1471–1477

Chih B, Gollan L and Scheiffele P (2006) Alternative splicing controls selective trans-synaptic

interactions of the neuroligin–neurexin complex, Neuron 51:171–178

Chubykin AA, Atasoy D, Etherton MR et al. (2007) Activity-dependent validation of

excitatory versus inhibitory synapses by neuroligin-1 versus neuroligin-2. Neuron

54:919–931

Chubykin AA, Liu X, Comoletti D et al. (2005) Dissection of synapse induction by

neuroligins: effect of a neuroligin mutation associated with autism, J Biol Chem

280:22365–22374

Comoletti D, De Jaco A, Jennings LL et al. (2004) The Arg451Cys-neuroligin-3 mutation

associated with autism reveals a defect in protein processing, J Neurosci 24:4889–4893

Comoletti D, Flynn R, Jennings LL et al. (2003) Characterization of the interaction of a

recombinant soluble neuroligin-1 with neurexin-1beta, J Biol Chem 278:50497–50505

Comoletti D, Flynn RE, Boucard AA et al. (2006) Gene selection, alternative splicing, and

post-translational processing regulate neuroligin selectivity for beta-neurexins, Biochemistry 45:12816–12827

Dean C, Scholl FG, Choih J et al. (2003) Neurexin mediates the assembly of presynaptic

terminals, Nat Neurosci 6:708–716

Durand CM, Betancur C, Boeckers TM et al. (2007) Mutations in the gene encoding the

synaptic scaffolding protein SHANK3 are associated with autism spectrum disorders, Nat

Genet 39:25–27

Fabrichny IP, Leone P, Sulzenbacher G et al. (2007) Structural analysis of the synaptic

protein neuroligin and its beta-neurexin complex: determinants for folding and cell adhesion, Neuron 56:979–991

Fu Z, Washbourne P, Ortinski P et al. (2003) Functional excitatory synapses in HEK293 cells

expressing neuroligin and glutamate receptors, J Neurophysiol 90:3950–3957

Futai K, Kim MJ, Hashikawa T et al. (2007) Retrograde modulation of presynaptic release

probability through signaling mediated by PSD-95-neuroligin, Nat Neurosci 10:186–195

Graf ER, Kang Y, Hauner AM et al. (2006) Structure function and splice site analysis of the

synaptogenic activity of the neurexin-1 beta LNS domain, J Neurosci 26:4256–4265

Graf ER, Zhang X, Jin SX et al. (2004) Neurexins induce differentiation of GABA and

glutamate postsynaptic specializations via neuroligins, Cell 119:1013–1026



364



A.A. Chubykin



Hata Y, Butz S and Suădhof TC (1996) CASK: a novel dlg/PSD95 homolog with an Nterminal calmodulin-dependent protein kinase domain identified by interaction with

neurexins, J Neurosci 16:2488–2494

Hines RM, Wu L, Hines DJ et al. (2008) Synaptic imbalance, stereotypies, and impaired social

interactions in mice with altered neuroligin 2 expression, J Neurosci 28:6055–6067

Hirao K, Hata Y, Ide N et al. (1998) A novel multiple PDZ domain-containing molecule

interacting with N-methyl-D-aspartate receptors and neuronal cell adhesion proteins, J

Biol Chem 273:21105–21110

Huber KM, Gallagher SM, Warren ST et al. (2002) Altered synaptic plasticity in a mouse

model of fragile X mental retardation, Proc Natl Acad Sci USA 99:7746–7750

Hung AY, Futai K, Sala C et al. (2008) Smaller dendritic spines, weaker synaptic transmission, but enhanced spatial learning in mice lacking Shank1. J Neurosci 28:1697–1708

Ichtchenko K, Hata Y, Nguyen T et al. (1995) Neuroligin 1: a splice site-specific ligand for

beta-neurexins, Cell 81:435–443

Ichtchenko K, Nguyen T and Suădhof TC (1996) Structures, alternative splicing, and neurexin

binding of multiple neuroligins, J Biol Chem 271:2676–2682

Iida J, Hirabayashi S, Sato Y et al. (2004) Synaptic scaffolding molecule is involved in the

synaptic clustering of neuroligin, Mol Cell Neurosci 27:497–508

Irie M, Hata Y, Takeuchi M et al. (1997) Binding of neuroligins to PSD-95. Science 277:1511–1515

Jamain S, Quach H, Betancur C et al. (2003) Mutations of the X-linked genes encoding

neuroligins NLGN3 and NLGN4 are associated with autism, Nat Genet 34:27–29

Jamain S, Radyushkin K, Hammerschmidt K et al. (2008) Reduced social interaction and

ultrasonic communication in a mouse model of monogenic heritable autism, Proc Natl

Acad Sci USA 105:1710–1715

Kim HG, Kishikawa S, Higgins AW et al. (2008a) Disruption of neurexin 1 associated with

autism spectrum disorder, Am J Hum Genet 82:199–207

Kim J, Jung SY, Lee YK et al. (2008b) Neuroligin-1 is required for normal expression of LTP

and associative fear memory in the amygdala of adult animals, Proc Natl Acad Sci USA

Koehnke J, Jin X, Budreck EC et al. (2008) Crystal structure of the extracellular cholinesterase-like domain from neuroligin-2. Proc Natl Acad Sci USA 105:1873–1878

Kootz JP, Marinelli B and Cohen DJ (1981) Sensory receptor sensitivity in autistic children:

response times to proximal and distal stimulation, Arch Gen Psychiatry 38:271–273

Laumonnier F, Bonnet-Brilhault F, Gomot M et al. (2004) X-linked mental retardation and

autism are associated with a mutation in the NLGN4 gene, a member of the neuroligin

family, Am J Hum Genet 74:552–557

Leonoudakis D, Conti LR, Radeke CM et al. (2004) A multiprotein trafficking complex

composed of SAP97, CASK, Veli, and Mint1 is associated with inward rectifier Kir2

potassium channels, J Biol Chem 279:19051–19063

Marble DD, Hegle AP, Snyder ED, 2nd et al. (2005) Camguk/CASK enhances Ether-a-go-go

potassium current by a phosphorylation-dependent mechanism, J Neurosci 25:48984907

Maximov A, Suădhof TC and Bezprozvanny I (1999) Association of neuronal calcium channels with modular adaptor proteins, J Biol Chem 274:24453–24456

Meyer G, Varoqueaux F, Neeb A et al. (2004) The complexity of PDZ domain-mediated

interactions at glutamatergic synapses: a case study on neuroligin, Neuropharmacology

47:724–733

Missler M, Zhang W, Rohlmann A et al. (2003) Alpha-neurexins couple Ca2+ channels to

synaptic vesicle exocytosis, Nature 423:939–948

Moretti P, Levenson JM, Battaglia F et al. (2006) Learning and memory and synaptic

plasticity are impaired in a mouse model of Rett syndrome, J Neurosci 26:319–327

Mukherjee K, Sharma M, Urlaub H et al. (2008) CASK Functions as a Mg2+-independent

neurexin kinase, Cell 133:328–339

Nam CI and Chen L (2005) Postsynaptic assembly induced by neurexin-neuroligin interaction

and neurotransmitter, Proc Natl Acad Sci USA 102:6137–6142



17



Neurexins and Neuroligins



365



Oertel WH and Mugnaini E (1984) Immunocytochemical studies of GABAergic neurons in

rat basal ganglia and their relations to other neuronal systems, Neurosci Lett 47:233–238

Petrenko AG, Ullrich B, Missler M et al. (1996) Structure and evolution of neurexophilin, J

Neurosci 16:4360–4369

Piluso G, Carella M, D’Avanzo M et al. (2003) Genetic heterogeneity of FG syndrome: a

fourth locus (FGS4) maps to Xp11.4-p11.3 in an Italian family, Hum Genet 112:124–130

Prange O, Wong TP, Gerrow K et al. (2004) A balance between excitatory and inhibitory

synapses is controlled by PSD-95 and neuroligin, Proc Natl Acad Sci USA 101:13915–13920

Rowen L, Young J, Birditt B et al. (2002) Analysis of the human neurexin genes: alternative

splicing and the generation of protein diversity, Genomics 79:587–597

Rudenko G, Hohenester E and Muller YA (2001) LG/LNS domains: multiple functions – one

business end? Trends Biochem Sci 26:363–368

Rudenko G, Nguyen T, Chelliah Y et al. (1999) The structure of the ligand-binding domain of

neurexin Ibeta: regulation of LNS domain function by alternative splicing, Cell 99:93–101

Sara Y, Biederer T, Atasoy D et al. (2005) Selective capability of SynCAM and neuroligin for

functional synapse assembly, J Neurosci 25:260–270

Scheiffele P, Fan J, Choih J et al. (2000) Neuroligin expressed in nonneuronal cells triggers

presynaptic development in contacting axons, Cell 101:657–669

Sebat J, Lakshmi B, Malhotra D et al. (2007) Strong association of de novo copy number

mutations with autism, Science 316:445–449

Serajee FJ, Zhong H, Nabi R et al. (2003) The metabotropic glutamate receptor 8 gene at

7q31: partial duplication and possible association with autism, J Med Genet 40:e42

Shen KC, Kuczynska DA, Wu IJ et al. (2008) Regulation of neurexin 1beta tertiary structure

and ligand binding through alternative splicing, Structure 16:422–431

Sheng M and Kim E (2000) The Shank family of scaffold proteins, J Cell Sci 113 (Pt 11):18511856

Song JY, Ichtchenko K, Suădhof TC et al. (1999) Neuroligin 1 is a postsynaptic cell-adhesion

molecule of excitatory synapses, Proc Natl Acad Sci USA 96:1100–1105

Squire LR and Zola SM (1996) Structure and function of declarative and nondeclarative

memory systems, Proc Natl Acad Sci USA 93:13515–13522

Steffenburg S, Gillberg C, Hellgren L et al. (1989) A twin study of autism in Denmark,

Finland, Iceland, Norway and Sweden, J Child Psychol Psychiatry 30:405–416

Sugita S, Saito F, Tang J et al. (2001) A stoichiometric complex of neurexins and dystroglycan

in brain, J Cell Biol 154:435–445

Szatmari P, Paterson AD, Zwaigenbaum L et al. (2007) Mapping autism risk loci using

genetic linkage and chromosomal rearrangements, Nat Genet 39:319–328

Tabuchi K, Blundell J, Etherton MR et al. (2007) A neuroligin-3 mutation implicated in

autism increases inhibitory synaptic transmission in mice, Science 318:71–76

Tabuchi K and Suădhof TC (2002) Structure and evolution of neurexin genes: insight into the

mechanism of alternative splicing, Genomics 79:849–859

Ullrich B, Ushkaryov YA and Suădhof TC (1995) Cartography of neurexins: more than 1000

isoforms generated by alternative splicing and expressed in distinct subsets of neurons,

Neuron 14:497–507

Ushkaryov YA, Petrenko AG, Geppert M et al. (1992) Neurexins: synaptic cell surface

proteins related to the alpha-latrotoxin receptor and laminin, Science 257:50–56

Varoqueaux F, Aramuni G, Rawson RL et al. (2006) Neuroligins determine synapse maturation and function, Neuron 51:741–754

Varoqueaux F, Jamain S and Brose N (2004) Neuroligin 2 is exclusively localized to inhibitory

synapses, Eur J Cell Biol 83:449–456

Zhao X, Leotta A, Kustanovich V et al. (2007) A unified genetic theory for sporadic and

inherited autism, Proc Natl Acad Sci USA 104:12831–12836

Zoghbi HY (2003) Postnatal neurodevelopmental disorders: meeting at the synapse? Science

302:826–830



Chapter 18



Synaptic Adhesion-Like Molecules (SALMs)

Philip Y. Wang and Robert J. Wenthold



Abstract The synaptic adhesion-like molecules (SALMs) are a newly discovered family of cell adhesion molecules that have a variety of functions

in neuronal development, including aspects of neurite outgrowth and synapse

formation (Ko et al. Neuron 50:233–245, 2006, Morimura et al. Gene

380:72–83, 2006, Wang et al. J Neurosci 26:2174–2183, 2006, Seabold et al.

J Biol Chem 283:8395–8405, 2008, Wang et al. Mol Cell Neurosci, 39:83–94,

2008). Also known as Lrfn (leucine-rich and fibronectin III domain-containing), five family members have been identified thus far: SALM1/Lrfn2,

SALM2/Lrfn1, SALM3/Lrfn4, SALM4/Lrfn3, and SALM5/Lrfn5. The

SALMs have been shown to interact with NMDA receptors and the PSD-95

family of MAGUK proteins. Recent studies also indicate that the individual

SALMs, while similar in structure, play distinct roles in heteromeric and

homomeric protein interactions and neurite outgrowth (Seabold et al. J Biol

Chem 283:8395–8405, 2008, Wang et al. Mol Cell Neurosci, 39:83–94, 2008).

Neurite outgrowth and synapse formation are fundamental mechanisms in the

development of the nervous system. While a considerable amount of information is known about both phenomena, the mechanism connecting the two is still

enigmatic. SALMs join a growing mosaic of synaptic proteins that contribute

to both neurite outgrowth and synapse formation during the course of development. Investigating SALMs and related proteins is essential for addressing

fundamental questions of neuronal development.

Keywords Synapse Á Dendrite Á Axon Á PDZ proteins Á LRR Á Growth cone Á

Heteromeric Á Homomeric Á Neurite outgrowth



R.J. Wenthold (*)

Laboratory of Neurochemistry, National Institute on Deafness and Other

Communication Disorders, National Institutes of Health, Bethesda, MD 20892, USA

e-mail: wenthold@nidcd.nih.gov



M. Hortsch, H. Umemori (eds.), The Sticky Synapse,

DOI 10.1007/978-0-387-92708-4_18, Ó Springer ScienceỵBusiness Media, LLC 2009



367



368



P.Y. Wang and R.J. Wenthold



18.1 SALM Family Structure and Expression

The synaptic adhesion-like molecules (SALMs) form a family of five adhesion

molecules. In the mouse, they range from 626 to 788 amino acid residues in

length. All SALMs contain a characteristic domain structure including six

leucine-rich repeat (LRR) regions, an immunoglobulin C2-like domain

(IgC2), a fibronectin type 3 domain (FN3), and a transmembrane region

(TM) (Fig. 18.1). Additionally, SALMs 1–3 contain a PDZ-binding domain

(PDZ-BD) at their distal C-termini, while SALMs 4–5 do not. The SALMs

share a considerable amount of sequence similarity, though there are regions of

high variability in both the N- and C-termini that could lend distinct characteristics to their individual functions (Fig. 18.2). Studies on SALMs and their

functions thus far have focused on the rat and mouse representatives of this

gene family. However, SALMs are present in a variety of mammalian species, as

well as in fish and amphibians (Morimura et al. 2006). Interestingly, additional

sequence analysis using the Ensembl database (www.ensembl.org) indicates

that the protein structure of the SALMs is quite conserved across these species.

As depicted in Fig. 18.3, the LRRs, IgC2, FN3, and PDZ-BD regions are all

conserved across various SALM1 sequences, including those of human, mouse,

platypus, chicken, and medaka (Japanese killifish). Analysis of the Drosophila

genome indicates that SALMs share the closest sequence identity with a family

of LRR and Ig-like domain containing transmembrane proteins called Kekkon

(Kekkon 1–5), which function in the developmental regulation of EGF receptor

activity during oogenesis (Ghiglione et al. 1999, 2003). Phylogenic analysis

reveals that SALMs are closely related to a variety of leucine-rich molecules

including the AMIGO, LINGO, NGL, PAL, and FLRT family of proteins (for

review of these proteins, see Chen et al. 2006). As we describe later in this

chapter, SALMs and these highly related proteins have one major characteristic

in common: their role in regulating neurite outgrowth.

Northern blot analysis indicates that all SALM transcripts are found in the

mouse and the rat brain, and transcripts for SALMs 2, 3, and 4 are seen to some

extent in testis (Ko et al. 2006, Morimura et al. 2006). Additionally, SALM4



Fig. 18.1 SALM protein domain structure. Schematic diagram illustrating the domain structure of the SALM family of proteins. The SALMs contain an N-terminal signal peptide, six

extracellular LRR regions flanked by N- and C-terminal LRR regions, an IgC2 domain, an

FN3 domain, a single TM region, and a PDZ-BD (present in SALMs 1–3)



18



Synaptic Adhesion-Like Molecules (SALMs)



369



Fig. 18.2 SALM family sequence comparison. (A) Sequence analysis reveals that mouse

SALMs contain six LRR regions (flanked by N- and C-terminal LRR sequences), an IgC2



370



P.Y. Wang and R.J. Wenthold



transcripts are found in the gastrointenstinal tract and kidney (Morimura et al.

2006). Temporal expression profile blots indicate that transcript levels for

SALMs 2–4 show an incremental increase starting from E10.5, while SALM1

and SALM5 increase from around E11.5–E12.5 (Morimura et al. 2006). In situ

hybridization reveals that SALM transcripts are distinctly expressed in a variety

of brain regions, including the cerebral cortex, hippocampus, dentate gyrus,

and olfactory bulb (Ko et al. 2006, Morimura et al. 2006).

Western blot and subcellular fractionation experiments showed that

SALM proteins are highly expressed in the rat brain, and enriched in synaptosomal and postsynaptic density fractions (Ko et al. 2006, Wang et al. 2006).

SALM protein expression levels exhibit some differentiation among family

members. Protein levels for SALM1 are detectable from E18, display high

expression from P1 to P21, and then decrease at P28 (Wang et al. 2006).

Protein levels for SALM2 increase from P21 and remain high at 6 weeks

(Ko et al. 2006). SALM2 protein is widely distributed in brain, and for

example, has been detected in cortical pyramidal neurons, hippocampal

CA3 and CA1 neurons, and cerebellar Purkinje cells (Ko et al. 2006). At the

subcellular level, SALM2 proteins localize to cell bodies, neurites, and punctate structures that co-localize with the presynaptic protein, synapsin I (Ko

et al. 2006). Ultrastructural analysis using immunogold electron microscopy

shows that native SALM4 is present at a variety of presynaptic, postsynaptic,

and extrasynaptic sites in hippocampus, olfactory bulb, and cerebellar cortex

(Seabold et al. 2008). Overexpressed SALMs are localized throughout the cell

in the soma, axons, dendrites, and growth cones in young (
vitro 7) neuronal cultures, both on the cell surface and intracellularly (Wang

et al. 2006, 2008). In older (>DIV14) neuronal cultures, overexpressed

SALMs are localized throughout the cell, on the cell surface, and at synapses

(Ko et al. 2006, Wang et al. 2006, and unpublished observations). Additionally, Ko et al. (2006) demonstrated that SALM2 is localized to excitatory, but

not inhibitory synapses, and that perturbation of SALM2 expression leads to

aberrations in excitatory synaptic formation.



Fig 18.2 (continued) domain, an FN3 domain, a TM region, and a PDZ-BD at the distal Ctermini (present in SALMs 1–3). Illustrated domain locations based on the SALM1 sequence.

Similar amino acid residues are highlighted by the gray background, and the consensus

sequence/strength is shown above the alignment. Consensus strength correlates with the

length of the individual bars. Protein sequences for SALMs 1–5 (accession numbers

NM_027452, NM_030562, NM_153388, NM_175478, and NM_178714, respectively) were

aligned by the Clustal V method using MegAlign computer software. (B) Phylogenic tree

comparing the mouse SALMs was constructed with MegAlign based on the alignment

produced in (A)



18



Synaptic Adhesion-Like Molecules (SALMs)



371



Fig. 18.3 SALM1 species comparison. SALM1 protein sequences from a variety of species

were acquired using the Ensembl database (www.ensembl.org). (A) The SALM1 sequence

structure is highly conserved among a variety of mammalian, avian, and fish species including



372



P.Y. Wang and R.J. Wenthold



18.2 SALM-Associated Proteins and Functional Significance

The known binding partners for the SALMs include the PSD-95 family of

membrane-associated guanylate kinase (MAGUK) proteins, the NR1 subunit

of the N-methyl-D-aspartate receptor (NMDAR), and the SALMs themselves –

through homomeric and heteromeric interactions (Ko et al. 2006, Morimura

et al. 2006, Wang et al. 2006, Seabold et al. 2008). The SALMs were independently identified in two different laboratories through yeast two-hybrid

screens using the PDZ domains of MAGUKs as bait (SAP97 in Wang et al.

2006, and PSD-95 in Ko et al. 2006). Classically described as scaffolding

proteins that assist in the tethering of receptors and associated proteins at

the postsynaptic density (PSD), MAGUKs have been linked to a variety of

functions in the CNS, including the trafficking of NMDARs to the synapse

and neurite outgrowth (Kim and Sheng 2004, Charych et al. 2006). Overexpression of PSD-95 decreases dendritic branching in immature neurons,

while knocking down PSD-95 increases it (Charych et al. 2006). Overexpression of SALMs in young neurons promotes neurite outgrowth, while overexpression of SALM 1–3 constructs lacking the PDZ-BD do not (Wang et al.

2008). This suggests that there may be a direct link in the process of neurite

outgrowth and SALM–MAGUK associations, at least for SALMs 1–3. Interestingly, SALMs 4, 5 lack a PDZ-BD, but still promote neurite outgrowth.

This may indicate that they act via a different mechanism or through heteromeric associations with other SALMs to induce MAGUK-associated neurite

outgrowth. In mature neurons, deletion of the PDZ-BD has effects on synapse

formation and morphology (Wang et al. 2006), further emphasizing the

importance of associated PDZ proteins in SALM function.

Additional evidence suggests that SALMs may interact, indirectly or

directly, with various other postsynaptic proteins. For example, beadinduced aggregation experiments revealed co-clustering of AMPARs and

GKAP with SALM2 (Ko et al. 2006). SALM1 has also been shown to

interact directly with NR1 in heterologous cells, and co-immunoprecipitates

with NR1 and NR2 subunits of NMDARs from brain (Wang et al. 2006).

The various protein domains that SALMs possess offer potential binding

sites for other proteins.



Fig 18.3 (continued) human (Homo sapiens), mouse (Mus musculus), platypus (Ornithorhynchus anatinus), chicken (Gallus gallus), and Japanese killifish/medaka (Oryzias latipes).

(B) Phylogenic tree comparing the SALM1 sequences of the various species. SALM1 proteins

sequences were aligned by the Clustal V method using MegAlign. Ensembl gene IDs used for

SALM1 human, mouse, platypus, chicken, and killifish were ENSG00000156564,

ENSOANG00000014625, ENSGALG00000010050, and ENSORLG00000018222, respectively. Accession number NM_027452 was used for the SALM1 mouse sequence



18



Synaptic Adhesion-Like Molecules (SALMs)



373



18.3 Homomeric and Heteromeric SALM Interactions

Cell adhesion molecules (CAMs) are usually transmembrane proteins, and

participate in the formation of cellular junctions through interactions between

their extracellular protein domains. These CAM interactions can be cis, within

the same membrane, or trans, across the cell–cell junctions of adjacent cells.

These cis or trans interactions can be heteromeric (forming interactions with

other distinct proteins), or homomeric (forming dimers or higher order multimers with a single protein type). For example, nectins form homomeric cis

interactions (Takai and Nakanishi 2003), while L1-type CAMs and axonin-1

form heteromeric cis interactions (Buchstaller et al. 1996, Stoeckli et al. 1996).

Cadherins form homomeric trans interactions (Tepass et al. 2000), while neuroligin and neurexin form heteromeric trans interactions (Ichtchenko et al.

1995, Scheiffele et al. 2000).

Seabold et al. (2008) demonstrated that SALMs are able to form heteromeric

and homomeric interactions with each other, though with some distinctions

among the individual SALM family members. From brain extracts, SALMs

1–3 co-immunoprecipitate with each other, indicating the formation of heteromeric interactions, while SALMs 4 and 5 do not. When individually expressed

and co-plated in heterologous cells, only SALMs 4 and 5 are able to form

homomeric trans interactions. These interactions are due to the extracellular

N-terminus, as demonstrated through the use of chimera constructs made of the

N- and C-termini of SALMs 2 and 4. When transfected into heterologous cells,

a construct containing the N-terminus of SALM4 and the C-terminus of

SALM2 (SALM4/2) forms homomeric trans interactions, while the reverse

chimera construct (SALM2/4) does not. Application of antibodies directed to

the extracellular LRR of SALMs blocks this interaction, indicating a function

of the LRR region in these trans associations. Furthermore, when HeLa cells

and primary hippocampal neurons overexpressing SALM4 are co-cultured,

SALM4 is recruited to points of contact between the two different cell types.

These SALM4 accumulations are seen at both axonal and dendritic adhesions

between neurons and HeLa cells.

Interestingly, when co-transfected into heterologous cells in pairs, all five

SALMs appear to form both heteromeric and homomeric complexes, indicating that the potential for such complexes may exist in vivo under certain

conditions. The mechanism of SALM trafficking and the dynamics of their

early interactions in the secretory pathway are not yet known. A hypothetical

model of SALM interactions is depicted in Fig. 18.4. Early in the ER, SALMs

may form combinations of dimers or higher order multimers with each other or

other proteins. The constitution of such SALM multimers could contribute to

the distinctiveness of SALM function, including their cis/trans interactions. The

potential interplay between cis and trans interactions is highly complex, and

may contribute to the variability among the SALM function. Formation of cis

complexes may regulate the formation of trans complexes. For example,



Xem Thêm
Tải bản đầy đủ (.pdf) (451 trang)

×