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 850 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.
8.
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   

9.
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   

10.
11.
12.

Background

Several felids are endangered and threatened by the illegal wildlife trade. Establishing geographic origin of tissues of endangered species is thus crucial for wildlife crime investigations and effective conservation strategies. As shown in other species, stable isotope analysis of hydrogen and oxygen in hair (δDh, δ18Oh) can be used as a tool for provenance determination. However, reliably predicting the spatial distribution of δDh and δ18Oh requires confirmation from animal tissues of known origin and a detailed understanding of the isotopic routing of dietary nutrients into felid hair.

Methodology/Findings

We used coupled δDh and δ18Oh measurements from the North American bobcat (Lynx rufus) and puma (Puma concolor) with precipitation-based assignment isoscapes to test the feasibility of isotopic geo-location of felidae. Hairs of felid and rabbit museum specimens from 75 sites across the United States and Canada were analyzed. Bobcat and puma lacked a significant correlation between H/O isotopes in hair and local waters, and also exhibited an isotopic decoupling of δ18Oh and δDh. Conversely, strong δD and δ18O coupling was found for key prey, eastern cottontail rabbit (Sylvilagus floridanus; hair) and white-tailed deer (Odocoileus virginianus; collagen, bone phosphate).

Conclusions/Significance

Puma and bobcat hairs do not adhere to expected pattern of H and O isotopic variation predicted by precipitation isoscapes for North America. Thus, using bulk hair, felids cannot be placed on δ18O and δD isoscapes for use in forensic investigations. The effective application of isotopes to trace the provenance of feline carnivores is likely compromised by major controls of their diet, physiology and metabolism on hair δ18O and δD related to body water budgets. Controlled feeding experiments, combined with single amino acid isotope analysis of diets and hair, are needed to reveal mechanisms and physiological traits explaining why felid hair does not follow isotopic patterns demonstrated in many other taxa.  相似文献   

13.
We previously found that pigeon IgG possesses unique N-glycan structures that contain the Galα1–4Galβ1–4Galβ1–4GlcNAc sequence at their nonreducing termini. This sequence is most likely produced by putative α1,4- and β1,4-galactosyltransferases (GalTs), which are responsible for the biosynthesis of the Galα1–4Gal and Galβ1–4Gal sequences on the N-glycans, respectively. Because no such glycan structures have been found in mammalian glycoproteins, the biosynthetic enzymes that produce these glycans are likely to have distinct substrate specificities from the known mammalian GalTs. To study these enzymes, we cloned the pigeon liver cDNAs encoding α4GalT and β4GalT by expression cloning and characterized these enzymes using the recombinant proteins. The deduced amino acid sequence of pigeon α4GalT has 58.2% identity to human α4GalT and 68.0 and 66.6% identity to putative α4GalTs from chicken and zebra finch, respectively. Unlike human and putative chicken α4GalTs, which possess globotriosylceramide synthase activity, pigeon α4GalT preferred to catalyze formation of the Galα1–4Gal sequence on glycoproteins. In contrast, the sequence of pigeon β4GalT revealed a type II transmembrane protein consisting of 438 amino acid residues, with no significant homology to the glycosyltransferases so far identified from mammals and chicken. However, hypothetical proteins from zebra finch (78.8% identity), frogs (58.9–60.4%), zebrafish (37.1–43.0%), and spotted green pufferfish (43.3%) were similar to pigeon β4GalT, suggesting that the pigeon β4GalT gene was inherited from the common ancestors of these vertebrates. The sequence analysis revealed that pigeon β4GalT and its homologs form a new family of glycosyltransferases.  相似文献   

14.
Mcm10 is an essential eukaryotic protein required for the initiation and elongation phases of chromosomal replication. Specifically, Mcm10 is required for the association of several replication proteins, including DNA polymerase α (pol α), with chromatin. We showed previously that the internal (ID) and C-terminal (CTD) domains of Mcm10 physically interact with both single-stranded (ss) DNA and the catalytic p180 subunit of pol α. However, the mechanism by which Mcm10 interacts with pol α on and off DNA is unclear. As a first step toward understanding the structural details for these critical intermolecular interactions, x-ray crystallography and NMR spectroscopy were used to map the binary interfaces between Mcm10-ID, ssDNA, and p180. The crystal structure of an Mcm10-ID·ssDNA complex confirmed and extended our previous evidence that ssDNA binds within the oligonucleotide/oligosaccharide binding-fold cleft of Mcm10-ID. We show using NMR chemical shift perturbation and fluorescence spectroscopy that p180 also binds to the OB-fold and that ssDNA and p180 compete for binding to this motif. In addition, we map a minimal Mcm10 binding site on p180 to a small region within the p180 N-terminal domain (residues 286–310). These findings, together with data for DNA and p180 binding to an Mcm10 construct that contains both the ID and CTD, provide the first mechanistic insight into how Mcm10 might use a handoff mechanism to load and stabilize pol α within the replication fork.To maintain their genomic integrity, cells must ensure complete and accurate DNA replication once per cell cycle. Consequently, DNA replication is a highly regulated and orchestrated series of molecular events. Multiprotein complexes assembled at origins of replication lead to assembly of additional proteins that unwind chromosomal DNA and synthesize nascent strands. The first event is the formation of a pre-replicative complex, which is composed of the origin recognition complex, Cdc6, Cdt1, and Mcm2–7 (for review, see Ref. 1). Initiation of replication at the onset of S-phase involves the activity of cyclin- and Dbf4-dependent kinases concurrent with recruitment of key factors to the origin. Among these, Mcm10 (2, 3) is recruited in early S-phase and is required for loading of Cdc45 (4). Mcm2–7, Cdc45, and the GINS complex form the replicative helicase (58). Origin unwinding is followed by loading of RPA,3 And-1/Ctf4, and pol α onto ssDNA (912). In addition, recruitment of Sld2, Sld3, and Dpb11/TopBP1 are essential for replication initiation (13, 14), and association of topoisomerase I, proliferating cellular nuclear antigen (PCNA), replication factor C, and the replicative DNA polymerases δ and ϵ completes the replisome (for review, see Ref. 15).Mcm10 is exclusive to eukaryotes and is essential to both initiation and elongation phases of chromosomal DNA replication (6, 8, 16). Mutations in Mcm10 in yeast result in stalled replication, cell cycle arrest, and cell death (2, 3, 1719). These defects can be explained by the number of genetic and physical interactions between Mcm10 and many essential replication proteins, including origin recognition complex, Mcm2–7, and PCNA (3, 12, 2024). In addition, Mcm10 has been shown to stimulate the phosphorylation of Mcm2–7 by Dbf4-dependent kinase in vitro (25). Thus, Mcm10 is an integral component of the replication machinery.Importantly, Mcm10 physically interacts with and stabilizes pol α and helps to maintain its association with chromatin (16, 26, 27). This is a critical interaction during replication because pol α is the only enzyme in eukaryotic cells that is capable of initiating DNA synthesis de novo. Indeed, Mcm10 stimulates the polymerase activity of pol α in vitro (28), and interestingly, the fission yeast Mcm10, but not Xenopus Mcm10, has been shown to exhibit primase activity (29, 30). Mcm10 is composed of three domains, the N-terminal (NTD), internal (ID), and C-terminal (CTD) domains (29). The NTD is presumably an oligomerization domain, whereas the ID and CTD both interact with DNA and pol α (29). The CTD is not found in yeast, whereas the ID is highly conserved among all eukaryotes. The crystal structure of Mcm10-ID showed that this domain is composed of an oligonucleotide/oligosaccharide binding (OB)-fold and a zinc finger motif, which form a unified DNA binding platform (31). An Hsp10-like motif important for the interaction with pol α has been identified in the sequence of Saccharomyces cerevisiae Mcm10-ID (16, 26).DNA pol α-primase is composed of four subunits: p180, p68, p58, and p48. The p180 subunit possesses the catalytic DNA polymerase activity, and disruption of this gene is lethal (32, 33). p58 and p48 form the DNA-dependent RNA polymerase (primase) activity (34, 35), whereas the p68 subunit has no known catalytic activity but serves a regulatory role (36, 37). Pol α plays an essential role in lagging strand synthesis by first creating short (7–12 nucleotide) RNA primers followed by DNA extension. At the critical length of ∼30 nucleotides, replication factor C binds to the nascent strand to displace pol α and loads PCNA with pols δ and ϵ (for review, see Ref. 38).The interaction between Mcm10 and pol α has led to the suggestion that Mcm10 may help recruit the polymerase to the emerging replisome. However, the molecular details of this interaction and the mechanism by which Mcm10 may recruit and stabilize the pol α complex on DNA has not been investigated. Presented here is the high resolution structure of the conserved Mcm10-ID bound to ssDNA together with NMR chemical shift perturbation competition data for pol α binding in the presence of ssDNA. Collectively, these data demonstrate a shared binding site for DNA and pol α in the OB-fold cleft of Mcm10-ID, with a preference for ssDNA over pol α. In addition, we have mapped the Mcm10-ID binding site on pol α to a 24-residue segment of the N-terminal domain of p180. Based on these results, we propose Mcm10 helps to recruit pol α to origins of replication by a molecular hand-off mechanism.  相似文献   

15.
16.
C. Tran 《PSN》2011,9(1):10-18
For the general public, Buddhist meditation has been a very fashionable subject for several years. It is presented either as a spiritual way to personal fulfilment or as a quick and effective relaxation technique. Otherwise, we have to keep in mind that Buddhist meditation is a lifestyle, born 2,500 years ago, integrated into a complex philosophic, religious, spiritual, and cultural system of thoughts. Moreover, it directly inspired the third cognitive wave of the cognitive and behavioural therapy, from which Mindfulness-Based Cognitive therapy (MBCT) was derived. The MBCT program is intended to prevent relapse/recurrence in major depression and has been the object of clinical trials with the aim of evaluating it using a scientific basis, but other indications are developed.  相似文献   

17.
18.
Family X DNA polymerases (PolXs) are involved in DNA repair. Their binding to gapped DNAs relies on two conserved helix-hairpin-helix motifs, one located at the 8-kDa domain and the other at the fingers subdomain. Bacterial/archaeal PolXs have a specifically conserved third helix-hairpin-helix motif (GFGxK) at the fingers subdomain whose putative role in DNA binding had not been established. Here, mutagenesis at the corresponding residues of Bacillus subtilis PolX (PolXBs), Gly130, Gly132 and Lys134 produced enzymes with altered DNA binding properties affecting the three enzymatic activities of the protein: polymerization, located at the PolX core, 3′-5′ exonucleolysis and apurinic/apyrimidinic (AP)-endonucleolysis, placed at the so-called polymerase and histidinol phosphatase domain. Furthermore, we have changed Lys192 of PolXBs, a residue moderately conserved in the palm subdomain of bacterial PolXs and immediately preceding two catalytic aspartates of the polymerization reaction. The results point to a function of residue Lys192 in guaranteeing the right orientation of the DNA substrates at the polymerization and histidinol phosphatase active sites. The results presented here and the recently solved structures of other bacterial PolX ternary complexes lead us to propose a structural model to account for the appropriate coordination of the different catalytic activities of bacterial PolXs.  相似文献   

19.
S. Iceta  F. Haesebaert  E. Poulet 《PSN》2011,9(2):78-85

Objective

Over the past 15 years, the therapeutic effects of repetitive transcranial magnetic stimulation (rTMS) have been well studied in psychiatric disorders. We propose to conduct a qualitative review of the literature to allow a focus on the value of this technique, including its main indications.

Results

This technique, by the induction of a change in cortical excitability, can provide symptomatic improvement in various psychiatric disorders. We first focused on its effectiveness in mood disorders (mainly in uni or bipolar depressive episodes), then in schizophrenia (auditory hallucinations and negative symptoms), and lately in other psychiatric disorders. We are also considering how to implement this technique in clinical practice and future development.

Conclusion

Results reported in literature suggest that rTMS might fit into the strategies of caremainly for unipolar depression and auditory hallucinations. Recommendations should be published in this regard during 2011. Optimization of this technique and the extension to other psychiatric indications are still under study.  相似文献   

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