首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Synapses are asymmetric intercellular junctions that mediate neuronal communication. The number, type, and connectivity patterns of synapses determine the formation, maintenance, and function of neural circuitries. The complexity and specificity of synaptogenesis relies upon modulation of adhesive properties, which regulate contact initiation, synapse formation, maturation, and functional plasticity. Disruption of adhesion may result in structural and functional imbalance that may lead to neurodevelopmental diseases, such as autism, or neurodegeneration, such as Alzheimer''s disease. Therefore, understanding the roles of different adhesion protein families in synapse formation is crucial for unraveling the biology of neuronal circuit formation, as well as the pathogenesis of some brain disorders. The present review summarizes some of the knowledge that has been acquired in vertebrate and invertebrate genetic model organisms.Synapses are asymmetric, intercellular junctions that are the basic structural units of neuronal transmission. The correct development of synaptic specializations and the establishment of appropriate connectivity patterns are crucial for the assembly of functional neuronal circuits. Improper synapse formation and function may cause neurodevelopmental disorders, such as mental retardation (MsR) and autism spectrum disorders (ASD) (McAllister 2007; Sudhof 2008), and likely play a role in neurodegenerative disorders, such as Alzheimer''s disease (AD) (Haass and Selkoe 2007).At chemical synapses (reviewed in Sudhof 2004; Zhai and Bellen 2004; Waites et al. 2005; McAllister 2007; Jin and Garner 2008), the presynaptic compartment contains synaptic vesicles (SV), organized in functionally distinct subcellular pools. A subset of SVs docks to the presynaptic membrane around protein-dense release sites, named active zones (AZ). Upon the arrival of an action potential at the terminal, the docked and “primed” SVs fuse with the plasma membrane and release neurotransmitter molecules into the synaptic cleft. Depending on the type of synapse (i.e., excitatory vs. inhibitory synapses), neurotransmitters ultimately activate an appropriate set of postsynaptic receptors that are accurately apposed to the AZ.Synapse formation occurs in several steps (Fig. 1) (reviewed in Eaton and Davis 2003; Goda and Davis 2003; Waites et al. 2005; Garner et al. 2006; Gerrow and El-Husseini 2006; McAllister 2007). Spatiotemporal signals guide axons through heterogeneous cellular environments to contact appropriate postsynaptic targets. At their destination, axonal growth cones initiate synaptogenesis through adhesive interactions with target cells. In the mammalian central nervous system (CNS), immature postsynaptic dendritic spines initially protrude as thin, actin-rich filopodia on the surface of dendrites. Similarly, at the Drosophila neuromuscular junction (NMJ), myopodia develop from the muscles (Ritzenthaler et al. 2000). The stabilization of intercellular contacts and their elaboration into mature, functional synapses involves cytoskeletal arrangements and recruitment of pre- and postsynaptic components to contact sites in spines and boutons. Conversely, retraction of contacts results in synaptic elimination. Both stabilization and retraction sculpt a functional neuronal circuitry.Open in a separate windowFigure 1.(A–C) Different stages of synapse formation. (A) Target selection, (B) Synapse assembly, (C) Synapse maturation and stabilization. (D–F) The role of cell adhesion molecules in synapse formation is exemplified by the paradigm of N-cadherin and catenins in regulation of the morphology and strength of dendritic spine heads. (D) At an early stage the dendritic spines are elongated from motile structures “seeking” their synaptic partners. (E) The contacts between the presynaptic and postsynaptic compartments are stabilized by recruitment of additional cell adhesion molecules. Adhesional interactions activate downstream pathways that remodel the cytoskeleton and organize pre- and postsynaptic apparatuses. (F) Cell adhesion complexes, stabilized by increased synaptic activity, promote the expansion of the dendritic spine head and the maturation/ stabilization of the synapse. Retraction and expansion is dependent on synaptic plasticity.In addition to the plastic nature of synapse formation, the vast heterogeneity of synapses (in terms of target selection, morphology, and type of neurotransmitter released) greatly enhances the complexity of synaptogenesis (reviewed in Craig and Boudin 2001; Craig et al. 2006; Gerrow and El-Husseini 2006). The complexity and specificity of synaptogenesis relies upon the modulation of adhesion between the pre- and postsynaptic components (reviewed in Craig et al. 2006; Gerrow and El-Husseini 2006; Piechotta et al. 2006; Dalva et al. 2007; Shapiro et al. 2007; Yamada and Nelson 2007; Gottmann 2008). Cell adhesive interactions enable cell–cell recognition via extracellular domains and also mediate intracellular signaling cascades that affect synapse morphology and organize scaffolding complexes. Thus, cell adhesion molecules (CAMs) coordinate multiple synaptogenic steps.However, in vitro and in vivo studies of vertebrate CAMs are often at odds with each other. Indeed, there are no examples of mutants for synaptic CAMs that exhibit prominent defects in synapse formation. This apparent “resilience” of synapses is probably caused by functional redundancy or compensatory effects among different CAMs (Piechotta et al. 2006). Hence, studies using simpler organisms less riddled by redundancy, such as Caenorhabditis elegans and Drosophila, have aided in our understanding of the role that these molecules play in organizing synapses.In this survey, we discuss the roles of the best characterized CAM families of proteins involved in synaptogenesis. Our focus is to highlight the complex principles that govern the molecular basis of synapse formation and function from a comparative perspective. We will present results from cell culture studies as well as in vivo analyses in vertebrate systems and refer to invertebrate studies, mainly performed in Drosophila and C. elegans, when they have provided important insights into the role of particular CAM protein families. However, we do not discuss secreted factors, for which we refer the reader to numerous excellent reviews (as for example Washbourne et al. 2004; Salinas 2005; Piechotta et al. 2006; Shapiro et al. 2006; Dalva 2007; Yamada and Nelson 2007; Biederer and Stagi 2008; Salinas and Zou 2008).  相似文献   

2.
The free energy of transfer of nonpolar solutes from water to lipid bilayers is often dominated by a large negative enthalpy rather than the large positive entropy expected from the hydrophobic effect. This common observation has led to the idea that membrane partitioning is driven by the "nonclassical" hydrophobic effect. We examined this phenomenon by characterizing the partitioning of the well-studied peptide melittin using isothermal titration calorimetry (ITC) and circular dichroism (CD). We studied the temperature dependence of the entropic (-TΔS) and enthalpic (ΔH) components of free energy (ΔG) of partitioning of melittin into lipid membranes made of various mixtures of zwitterionic and anionic lipids. We found significant variations of the entropic and enthalpic components with temperature, lipid composition and vesicle size but only small changes in ΔG (entropy-enthalpy compensation). The heat capacity associated with partitioning had a large negative value of about -0.5 kcal mol(-1) K(-1). This hallmark of the hydrophobic effect was found to be independent of lipid composition. The measured heat capacity values were used to calculate the hydrophobic-effect free energy ΔG (hΦ), which we found to dominate melittin partitioning regardless of lipid composition. In the case of anionic membranes, additional free energy comes from coulombic attraction, which is characterized by a small effective peptide charge due to the lack of additivity of hydrophobic and electrostatic interactions in membrane interfaces [Ladokhin and White J Mol Biol 309:543-552, 2001]. Our results suggest that there is no need for a special effect-the nonclassical hydrophobic effect-to describe partitioning into lipid bilayers.  相似文献   

3.
Alcadeins (Alcs) constitute a family of neuronal type I membrane proteins, designated Alcα, Alcβ, and Alcγ. The Alcs express in neurons dominantly and largely colocalize with the Alzheimer amyloid precursor protein (APP) in the brain. Alcs and APP show an identical function as a cargo receptor of kinesin-1. Moreover, proteolytic processing of Alc proteins appears highly similar to that of APP. We found that APP α-secretases ADAM 10 and ADAM 17 primarily cleave Alc proteins and trigger the subsequent secondary intramembranous cleavage of Alc C-terminal fragments by a presenilin-dependent γ-secretase complex, thereby generating “APP p3-like” and non-aggregative Alc peptides (p3-Alcs). We determined the complete amino acid sequence of p3-Alcα, p3-Alcβ, and p3-Alcγ, whose major species comprise 35, 37, and 31 amino acids, respectively, in human cerebrospinal fluid. We demonstrate here that variant p3-Alc C termini are modulated by FAD-linked presenilin 1 mutations increasing minor β-amyloid species Aβ42, and these mutations alter the level of minor p3-Alc species. However, the magnitudes of C-terminal alteration of p3-Alcα, p3-Alcβ, and p3-Alcγ were not equivalent, suggesting that one type of γ-secretase dysfunction does not appear in the phenotype equivalently in the cleavage of type I membrane proteins. Because these C-terminal alterations are detectable in human cerebrospinal fluid, the use of a substrate panel, including Alcs and APP, may be effective to detect γ-secretase dysfunction in the prepathogenic state of Alzheimer disease subjects.  相似文献   

4.
5.
6.
7.
A Gram-negative bacterium that formed cream-colored colonies designated strain LF7 was isolated from soil collected in the Tambopata National Reserve in Madre de Dios, Peru. 16S rRNA sequence comparisons indicate that LF7 is a novel Enterobacter sp. closely related to E. asburiae JCM 6051(T) [AB004744] and E. aerogenes JCM 1235(T) [AB004750] based on their sequence homologies (p-distance: 1.06 and 1.19%, respectively). DNA G + C content was 52.8 mol% which is within the range reported for E. asburiae (55-57 mol%). The major cellular fatty acids present in the LF7 strain were C(16:0) (27.3%), C(16:1) ω7c and/or C(16:1) ω6c (16.3%), C(18:1) ω7c (16.1%), C(17:0) cyclo (12.4%), C(14:0) 3-OH and/or C(16:1) iso-I (8.9%), C(14:0) (7.6%), C(12:0) (3.9%), C(17:0) (2.4%), C(13:0) 3-OH and/or C(15:1) iso-H (1.7%), C(13:0) (1.1%), and C(18:2) ω6,9c and/or C(18:0) ante (0.5%). The cellular fatty acid profile, G + C content, phenotypic and biochemical characteristics were consistent with its placement in the genus Enterobacter. The name Enterobacter soli is proposed for this bacterium.  相似文献   

8.
According to Charles Hauss, “[i]n the last few years, reconciliation has become one of the ‘hottest’ topics in the increasingly ‘hot’ field of conflict resolution” (2003, ?1). However, despite the apparent interest in this “hot” academic topic (which is becoming increasingly warm in Canada as our own Truth and Reconciliation Commission commences), reconciliation studies have been dominated by Truth-based approaches. The restrictions of these approaches, which emphasize objectivity and rationality, often elide the body and the primacy of emotions in the reparative process. This essay begins a conversation on the role of the body and emotion in the study of reconciliation by engaging the work being done in the social sciences with contemporary trends in critical theory and literature. I argue that by looking at the fundamental role the body plays on the “road to reconciliation” we can devise a more vital approach to conflict resolution and the various processes that make it up.  相似文献   

9.
10.
11.
12.
Mammalian defensins are cationic antimicrobial peptides that play a central role in host innate immunity and as regulators of acquired immunity. In animals, three structural defensin subfamilies, designated as α, β, and θ, have been characterized, each possessing a distinctive tridisulfide motif. Mature α- and β-defensins are produced by simple proteolytic processing of their prepropeptide precursors. In contrast, the macrocyclic θ-defensins are formed by the head-to-tail splicing of nonapeptides excised from a pair of prepropeptide precursors. Thus, elucidation of the θ-defensin biosynthetic pathway provides an opportunity to identify novel factors involved in this unique process. We incorporated the θ-defensin precursor, proRTD1a, into a bait construct for a yeast two-hybrid screen that identified rhesus macaque stromal cell-derived factor 2-like protein 1 (SDF2L1), as an interactor. SDF2L1 is a component of the endoplasmic reticulum (ER) chaperone complex, which we found to also interact with α- and β-defensins. However, analysis of the SDF2L1 domain requirements for binding of representative α-, β-, and θ-defensins revealed that α- and β-defensins bind SDF2L1 similarly, but differently from the interactions that mediate binding of SDF2L1 to pro-θ-defensins. Thus, SDF2L1 is a factor involved in processing and/or sorting of all three defensin subfamilies.Mammalian defensins are tridisulfide-containing antimicrobial peptides that contribute to innate immunity in all species studied to date. Defensins are comprised of three structural subfamilies: the α-, β-, and θ-defensins (1). α- and β-Defensins are peptides of about 29–45-amino acid residues with similar three-dimensional structures. Despite their similar tertiary conformations, the disulfide motifs of α- and β-defensins differ. Expression of human α-defensins is tissue-specific. Four myeloid α-defensins (HNP1–4) are expressed predominantly by neutrophils and monocytes wherein they are packaged in granules, while two enteric α-defensins (HD-5 and HD-6) are expressed at high levels in Paneth cells of the small intestine. Myeloid α-defensins constitute about 5% of the protein mass of human neutrophils. HNPs are discharged into the phagosome during phagocytic ingestion of microbial particles. HD-5 and HD-6 are produced and stored as propeptides in Paneth cell granules and are processed extracellularly by intestinal trypsin (2). β-Defensins are produced primarily by various epithelia (e.g. skin, urogenital tract, airway) and are secreted by the producing cells in their mature forms. In contrast to pro-α-defensins, which contain a conserved prosegment of ∼40 amino acids, the prosegments in β-defensins vary in length and sequence. θ-Defensins are found only in Old World monkeys and orangutans and are the only known circular peptides in animals. These 18-residue macrocyclic peptides are formed by ligation of two nonamer sequences excised from two precursor polypeptides, which are truncated versions of ancestral α-defensins. Like myeloid α-defensins, θ-defensins are stored primarily in neutrophil and monocyte granules (3).Numerous laboratories have demonstrated that the antimicrobial properties of defensins derive from their ability to bind and disrupt target cell membranes (4), and studies have shown defensins to be active against Gram-positive and -negative bacteria (5), viruses (69), fungi (10, 11), and parasites such as Giardia lamblia (12). Defensins also play a regulatory role in acquired immunity as they are known to chemoattract T lymphocytes, monocytes, and immature dendritic cells (13, 14), act as adjuvants, stimulate B cell responses, and up-regulate proliferation and cytokine production by spleen cells and T helper cells (15, 16).Defensins are produced as pre-propeptides and undergo post-translational processing to form the mature peptides. While much has been learned about regulation of defensin expression, little is known about the factors involved in their biosynthesis. Valore and Ganz (17) investigated the processing of defensins in cultured cells and demonstrated that maturation of HNPs occurs through two proteolytic steps that lead to formation of mature α-defensins, but the proteases involved have yet to be identified. Moreover, there are virtually no published data regarding endoplasmic reticulum (ER)2 factors that are responsible for the folding, processing, and sorting steps necessary for defensin maturation and secretion or trafficking to the proper subcellular compartment. It is likely that several chaperones, proteases, and protein-disulfide isomerase (PDI) family proteins are involved. Consistent with this possibility, Gruber et al. (18) recently demonstrated the role of a PDI in biosynthesis of cyclotides, small ∼30-residue macrocyclic peptides produced by plants.The primary structures of α- and θ-defensin precursors are closely related. We therefore undertook studies to identify proteins that interact with representative propeptides of each defensin subfamily with the goal of determining common and unique processes that regulate biosynthesis of α- and θ-defensins. We used two-hybrid analysis to first identify interactors of the θ-defensin precursor, proRTD1a. As described, we identified SDF2L1, a component of the ER-chaperone complex as an interactor, and showed that it also specifically interacts with α- and β-defensins. This suggests that SDF2L1 is involved in the maturation/trafficking of defensins at a step common to all three subfamilies of mammalian defensins.  相似文献   

13.
14.
15.
16.
Recently, a few fish proteins have been described with a high homology to B-type lectins of monocotyledonous plants. Because of their mannose binding activity, they have been ascribed a role in innate immunity. By screening various fish venoms for their integrin inhibitory activity, we isolated a homologous protein from the fin stings and skin mucus of the scorpionfish (Scorpaena plumieri). This protein inhibits α1β1 integrin binding to basement membrane collagen IV. By protein chemical and spectroscopic means, we demonstrated that this fish protein, called plumieribetin, is a homotetramer and contains a high content of anti-parallel β strands, similar to the mannose-binding monocot B-lectins. It lacks both N-linked glycoconjugates and common O-glycan motifs. Despite its B-lectin-like structure, plumieribetin binds to α1β1 integrin irrespective of N-glycosylation, suggesting a direct protein-protein interaction. This interaction is independent of divalent cations. On the cellular level, plumieribetin failed to completely detach hepatocarcinoma HepG2 cells and primary arterial smooth muscle cells from the collagen IV fragment CB3. However, plumieribetin weakened the cell-collagen contacts, reduced cell spreading, and altered the actin cytoskeleton, after the compensating α2β1 integrin was blocked. The integrin inhibiting effect of plumieribetin adds a new function to the B-lectin family, which is known for pathogen defense.  相似文献   

17.
18.
19.
20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号