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.
Although mesenchymal stem cells (MSCs) are the natural source for bone regeneration, the exact mechanisms governing MSC crosstalk with collagen I have not yet been uncovered. Cell adhesion to collagen I is mostly mediated by three integrin receptors – α1β1, α2β1 and α11β1. Using human MSC (hMSC), we show that α11 subunit exhibited the highest basal expression levels but on osteogenic stimulation, both α2 and α11 integrins were significantly upregulated. To elucidate the possible roles of collagen-binding integrins, we applied short hairpin RNA (shRNA)-mediated knockdown in hMSC and found that α2 or α11 deficiency, but not α1, results in a tremendous reduction of hMSC numbers owing to mitochondrial leakage accompanied by Bcl-2-associated X protein upregulation. In order to clarify the signaling conveyed by the collagen-binding integrins in hMSC, we analyzed the activation of focal adhesion kinase, extracellular signal-regulated protein kinase and serine/threonine protein kinase B (PKB/Akt) kinases and detected significantly reduced Akt phosphorylation only in α2- and α11-shRNA hMSC. Finally, experiments with hMSC from osteoporotic patients revealed a significant downregulation of α2 integrin concomitant with an augmented mitochondrial permeability. In conclusion, our study describes for the first time that disturbance of α2β1- or α11β1-mediated interactions to collagen I results in the cell death of MSCs and urges for further investigations examining the impact of MSCs in bone conditions with abnormal collagen I.  相似文献   

2.
3.
We perform a systematical investigation on the geometry, thermodynamic/kinetic stability, and bonding nature of low-lying isomers of BnPt (n=1-6) at the CCSD(T)/[6-311+G(d)/LanL2DZ]//B3LYP/[6-311+G(d)/LanL2DZ] level. The most stable isomers of BnPt (n=1-6) adopt planar or quasi-planar structure. BnPt (n=2-5) clusters can be generated by capping a Pt atom on the B-B edge of pure boron clusters. However, For B6Pt with non-planar structure, a single doped Pt atom significantly affects the shape of the host boron cluster. The dopant of the Pt atom can improve the stability of pure boron clusters. The valence molecular orbital (VMO), electron localization function (ELF), and Mayer bond order (MBO) are applied to gain insight into the bonding nature of BnPt (n=2-6) isomers. The aromaticity for some isomers of BnPt (n=2-6) is analyzed and discussed in terms of VMO, ELF, adaptive natural density partitioning (AdNDP), and nucleus-independent chemical shift (NICS) analyses. Results obtained from the energy and cluster decomposition analyses demonstrate that B2Pt and B4Pt exhibits as highly stable. Importantly, some isomers of BnPt (n=2-5) are stable both thermodynamically and kinetically, which are observable in future experiment.  相似文献   

4.
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   

5.
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   

6.
The insertion reactions of the silylene H2Si with H2BXHn-1 (X?=?F, Cl, Br, O, N; n?=?1, 1, 1, 2, 3) have been studied by DFT and MP2 methods. The calculations show that the insertions occur in a concerted manner, forming H2Si(BH2)(XHn-1). The essences of H2Si insertions with H2BXHn-1 are the transfers of the σ electrons on the Si atom to the positive BH2 group and the electrons of X into the empty p orbital on the Si atom in H2Si. The order of reactivity in vacuum shows the barrier heights increase for the same-family element X from up to down and the same-row element X from right to left in the periodic table. The energies relating to the B-X bond in H2BXHn-1, and the bond energies of Si-X and Si-B in H2Si(BH2)(XHn-1) may determine the preference of insertions of H2Si into B-X bonds for the same-column element X or for the same-row element X. The insertion reactions in vacuum are similar to those in solvents, acetone, ether, and THF. The barriers in vacuum are lower than those in solvents and the larger polarities of solvents make the insertions more difficult to take place. Both in vacuum and in solvents, the silylene insertions are thermodynamically exothermic.
Graphical Abstract The insertion process of H2Si and H2BXHn-1(X?=?F, Cl, Br, O, and N; n?=?1, 1 , 1, 2, 3).
  相似文献   

7.
Many agents that activate hematopoietic cells use phos pha tidyl ino si tol 3,4,5-trisphosphate (PtdIns 3,4,5-P3) to initiate signaling cascades. The SH2 domain-containing inositol 5′ phosphatase, SHIP1, regulates hematopoietic cell function by opposing the action of phos pha tidyl ino si tol 3-kinase and reducing the levels of PtdIns 3,4,5-P3. Activation of the cyclic AMP-de pend ent protein kinase (PKA) also opposes many of the pro-inflammatory responses of hematopoietic cells. We tested to see whether the activity of SHIP1 was regulated via phos pho ryl a tion with PKA. We prepared pure recombinant SHIP1 from HEK-293 cells and found it can be rapidly phos pho ryl a ted by PKA to a stoichiometry of 0.6 mol of PO4/mol of SHIP1. In 32P-labeled HEK-293 cells transfected with SHIP1, stimulation with Sp-adenosine 3′,5′-cyclic monophosphorothioate triethylammonium salt hydrate (Sp-cAMPS) or activation of the β-adrenergic receptor increased the phos pho ryl a tion state of SHIP1. Inhibition of protein phosphatase activity with okadaic acid also increased the phos pho ryl a tion of SHIP1. Phosphorylation of SHIP1 in vitro or in cells by PKA increased the 5′ phosphatase activity of SHIP1 by 2–3-fold. Elevation of Ca2+ in DT40 cells in response to B cell receptor cross-linking, an indicator of PtdIns 3,4,5-P3 levels, was markedly blunted by pretreatment with Sp-cAMPS. This effect was absent in SHIP−/− DT40 cells showing that the effect of Sp-cAMPS in DT40 cells is SHIP1-de pend ent. Sp-cAMPS also blunted the ability of the B cell receptor to increase the phos pho ryl a tion of Akt in DT40 and A20 cells. Overall, activation of G protein-coupled receptors that raise cyclic AMP cause SHIP1 to be phos pho ryl a ted and stimulate its inositol phosphatase activity. These results outline a novel mechanism of SHIP1 regulation.Activation of phosphatidylinositol 3-kinase (PtdIns 3-kinase)2 is central to regulation of multiple cell functions including cell shape changes, cell migration, cell activation, and proliferation (1). PtdIns 3-kinase phosphorylates phosphatidylinositol 4,5-bisphosphate in the inner leaflet of the plasma membrane to generate phosphatidylinositol 3,4,5-trisphosphate (PtdIns 3,4,5-P3) (2). PtdIns 3,4,5-P3 then activates downstream signaling pathways by interacting with pleckstrin homology domain-containing proteins, such as phosphoinositide-dependent kinase 1 and the serine-threonine kinase Akt (3). The finding of abnormal activation of the PtdIns 3-kinase pathway in cancer cells has led to interest in the development of inhibitors for PtdIns 3-kinase (4).The level of PtdIns 3,4,5-P3 is stimulated by multiple members of the PtdIns 3-kinase family (2) and is opposed by two phosphatidylinositol phosphatases: the Src homology 2 (SH2) domain-containing inositol 5′ phosphatase (SHIP) and the 3′ inositol phosphatase, phosphatase and tensin homolog (PTEN) (5). PTEN removes phosphate from the 3′ position in the inositol ring of PtdIns 3,4,5-P3 and converts it to phosphatidylinositol 4,5-bisphosphate (6). PTEN has a C2 domain, a PDZ-binding motif, and a N-terminal phosphatidylinositol 4,5-bisphosphate binding motif essential for translocation to the membrane and interaction with other regulatory proteins (7). There are serine and threonine residues in PTEN that have been found to be phosphorylated, but their role in regulating the activity of the enzyme is not clear (8). Mutations in the PTEN protein have been observed in many tumors, suggesting a role for this enzyme in cancer (9).In contrast, SHIP dephosphorylates the 5′ position on the inositol ring and produces phosphatidylinositol 3,4-bisphosphate (10). There are three isoforms of SHIP: the 145-kDa hematopoietic cell restricted SHIP (also known as SHIP1); the 104-kDa stem cell-restricted SHIP, sSHIP; and the more widely expressed 150-kDa SHIP2 (11). SHIP1 is the major inositol phosphatase regulating PtdIns 3,4,5-P3 in monocytes, macrophages, B cells, and T cells (11). SHIP1 has three known structural features: the N-terminal SH2 domain, the central inositol 5′ phosphatase domain, and two NPXY sequences in the C-terminal region. The currently accepted model for regulation of PtdIns 3,4,5-P3 levels by SHIP1 envisions translocation of SHIP1 from the cytosol to the membrane. Upon stimulation by growth factors, cytokine receptors, or immunoreceptors, SHIP1 is recruited via its N-terminal SH2 domain to phosphorylated tyrosine residues in receptor kinases and degrades the elevated levels of PtdIns 3,4,5-P3 near the activated receptor (12). During this translocation process, SHIP1 is not thought to change its 5′ phosphatase activity (13). Although it is known that SHIP1 can be phosphorylated on tyrosine residues by the lyn cytoplasmic kinase (12) or following the activation of the T cell receptor (14), neither event appears to influence the 5′ phosphatase activity. To date, direct regulation of SHIP1 activity by serine/threonine kinases has not been studied.Activation of G protein-coupled receptors that raise cAMP (i.e. β-adrenergic receptors or adenosine A2a receptors) is known to blunt the pro-inflammatory responses generated by receptors that raise the level of PtdIns 3,4,5-P3 (15). Therefore, we investigated the possibility that phosphorylation of SHIP1 by cyclic AMP-dependent protein kinase (PKA) might regulate the activity of SHIP1. We found that SHIP1 can be phosphorylated by PKA both in vitro and in cells leading to a stimulation of SHIP1 activity. Activation of PKA in DT40 and A20 cells blunted indicators of the PtdIns 3,4,5-P3 response to B cell receptor stimulation. These results indicate that SHIP1 activity can be regulated both in vitro and in cells by activation of the cyclic AMP-dependent protein kinase and highlight a new mode of SHIP regulation by G protein-coupled receptors.  相似文献   

8.
Activated macrophages are essential effectors of immunity and a rich source of matrix metalloproteinase-9 (MMP-9; gelatinase B). To search for cellular substrates of the enzyme, we subjected wild-type macrophages and macrophages expressing an autoactivating form of pro-MMP-9 (M9A macrophages) to proteomics analysis. Two-dimensional liquid chromatography together with tandem mass spectrometry identified 467 proteins in medium conditioned by M9A and/or wild-type macrophages. Subtractive proteomics identified 18 candidate MMP-9 substrates. Biochemical studies confirmed that two transmembrane proteins, β2 integrin subunit (CD18) and amyloid protein precursor (APP), were enriched in the medium of M9A macrophages. To identify potential cleavage sites, we synthesized an overlapping library of peptides that spanned 60 residues of the ectodomain and transmembrane domain of β2 integrin. Active MMP-9 cleaved a single peptide, ECVKGPNVAAIVGGT, at residues corresponding to Ala705 and Ile706 of the β2 integrin. Peptides corresponding to this cleavage site were detected by tandem mass spectrometric analysis only in medium from M9A macrophages, strongly supporting the proposal that β2 integrin is shed by autoactivating MMP-9. Our observations indicate that subtractive proteomics in concert with peptide substrate mapping is a powerful approach for identifying proteolytic substrates and suggest that MMP-9 plays previously unsuspected roles in the regulation and shedding of β2 integrin.Matrix metalloproteinases (MMPs),1 a subfamily of metazincins, are a structurally related group of zinc-dependent proteases (1). They are synthesized in latent form as pro-MMPs, and their prodomain must be removed or modified before they are proteolytically active. Some MMPs are secreted, whereas others are anchored to the cell surface, but their proteolytic activity is thought to be confined locally within the secretory pathway at the cell surface and nearby extracellular space (13). Individual MMPs have distinct substrate specificities and act on diverse extracellular and membrane proteins, such as chemokines, cell surface adhesion proteins, and extracellular matrix components. Proteolysis by MMPs plays an important role in a wide variety of normal and pathological processes, such as host defense, inflammation, and tumor progression (19).High levels of MMP-9 (gelatinase B) are expressed by activated macrophages (10), which are key effector cells of both innate and acquired immunity. In addition to having homeostatic functions, MMP-9 secreted by macrophages has been implicated in aneurysm formation, tumor progression, and disruption of atherosclerotic plaques (8, 9, 11, 12). Although the pathogenesis of those processes is generally thought to involve inappropriate degradation of extracellular matrix proteins, it has become increasingly clear that MMPs cleave a number of diverse substrates to mediate their varied functions (3, 13). Because MMP-9 can accumulate on the cell surface (14), it is likely to act on membrane proteins.To understand the specific roles of individual MMPs in inflammatory and immune responses, it is critical to identify their physiological substrates (3, 1517). Most studies have focused on identifying substrates by their ability to be cleaved in defined in vitro reactions (18, 19), but this approach is biased in two ways. First, the candidate substrate must be selected a priori. Second, in vitro reactions fail to account for the complexity of the pericellular environment. Another method is to identify sequences in synthetic peptides that MMPs can cleave (20, 21). However, individual MMPs cleave different proteins at a variety of sites rather than at a consensus site. Moreover MMPs often interact with substrates through domains remote from the active site (exosites) (22), and exosites of MMP-2 have been used in a yeast two-hybrid system to trap candidate substrates (23). However, some substrates may bind weakly or not at all to exosites, limiting the utility of this approach for global substrate screening.An emerging strategy for finding MMP substrates is to conduct an unbiased, global search by coupling gel electrophoresis or liquid chromatography with MS-based protein identification. For example, two-dimensional (2D) gel electrophoresis (24) and derivatization of cysteine-containing peptides with an isotope affinity tag (25) have identified candidate substrates for membrane type-1 MMP (MT1-MMP) in plasma and cultured cells. Quantitative approaches using 2D difference gel electrophoresis have identified potential substrates of MMP-2 and MMP-9 in bronchoalveolar lavage fluid (26) and of MMP-9 and the related metalloproteinases ADAM-10 and ADAM-17 in cancer cells (27, 28). Lectin affinity chromatography detected glycosylated proteins that were selectively enriched in medium from a monocyte cell line expressing ADAM-17 and in phorbol ester-stimulated monocytes (16). Recently iTRAQ (isobaric tags for relative and absolute quantitation) labeling was used to identify substrates of MMP-2 (29). It is important to note, however, that proteases can affect protein abundance by pathways not involving proteolysis. Thus, an important limitation of many of these studies is that they fail to provide evidence that proteins with altered abundance in cells expressing a protease are direct substrates for proteolytic cleavage.In the current studies, we used subtractive proteomics to identify proteins enriched in the medium of a macrophage cell line. Subtractive proteomics compares two or more proteomes to identify proteins that are specifically enriched or depleted under certain conditions (30, 31). Our biochemical studies confirmed that two integral membrane proteins, amyloid precursor protein (APP) and the β2 integrin subunit (CD18), were shed by macrophages expressing autoactivating MMP-9. We next used a peptide substrate mapping strategy to identify potential MMP-9 cleavage sites in β2 integrin subunit. Targeted MS/MS analysis demonstrated that β2 integrin subunit peptides with the same cleavage site were detected only in the medium of macrophages expressing autoactivating MMP-9, providing strong evidence that β2 integrin is a direct substrate for proteolysis. Our observations indicate that subtractive proteomics in concert with peptide substrate mapping is a robust, high throughput technique for identifying cellular substrates that are proteolytically shed from macrophages.  相似文献   

9.
A combined density functional and ab initio quantum chemical study of the insertion reactions of the germylenoid H2GeLiF with SiH3X (X?=?F, Cl, Br) was carried out. The geometries of all the stationary points of the reactions were optimized using the DFT B3LYP method and then the QCISD method was used to calculate the single-point energies. The theoretical calculations indicated that along the potential energy surface, there were one precursor complex (Q), one transition state (TS), and one intermediate (IM) which connected the reactants and the products. The calculated barrier heights relative to the respective precursors are 102.26 (X?=?F), 95.28 (X?=?Cl), and 84.42 (X?=?Br) kJ mol-1 for the three different insertion reactions, respectively, indicating the insertion reactions should occur easily according to the following order: SiH3-Br?>?SiH3-Cl?>?SiH3-F under the same situation. The solvent effects on the insertion reactions were also calculated and it was found that the larger the dielectric constant, the easier the insertion reactions. The elucidations of the mechanism of these insertion reactions provided a new reaction model of germanium-silicon bond formation.  相似文献   

10.
最常见的人类X染色体的结构异常是长臂等臂X染色体,以及X染色体与常染色体之间的易位。关于idic(X)染色体结构异常的病例较为罕见。国外Chapella等曾有研究,国内周氏等首次报道了有关idic(X)染色体结构异常的病例。  相似文献   

11.
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   

12.
13.
Transforming growth factor-β1 (TGF-β1) is a multifunctional cytokine that signals through the interaction of type I (TβRI) and type II (TβRII) receptors to activate distinct intracellular pathways. TAK1 is a serine/threonine kinase that is rapidly activated by TGF-β1. However, the molecular mechanism of TAK1 activation is incompletely understood. Here, we propose a mechanism whereby TAK1 is activated by TGF-β1 in primary mouse mesangial cells. Under unstimulated conditions, endogenous TAK1 is stably associated with TβRI. TGF-β1 stimulation causes rapid dissociation from the receptor and induces TAK1 phosphorylation. Deletion mutant analysis indicates that the juxtamembrane region including the GS domain of TβRI is crucial for its interaction with TAK1. Both TβRI-mediated TAK1 phosphorylation and TGF-β1-induced TAK1 phosphorylation do not require kinase activity of TβRI. Moreover, TβRI-mediated TAK1 phosphorylation correlates with the degree of its association with TβRI and requires kinase activity of TAK1. TAB1 does not interact with TGF-β receptors, but TAB1 is indispensable for TGF-β1-induced TAK1 activation. We also show that TRAF6 and TAB2 are required for the interaction of TAK1 with TβRI and TGF-β1-induced TAK1 activation in mouse mesangial cells. Taken together, our data indicate that TGF-β1-induced interaction of TβRI and TβRII triggers dissociation of TAK1 from TβRI, and subsequently TAK1 is phosphorylated through TAB1-mediated autophosphorylation and not by the receptor kinase activity of TβRI.Members of the transforming growth factor-β (TGF-β)3 superfamily are key regulators of various biological processes such as cellular differentiation, proliferation, apoptosis, and wound healing (1, 2). TGF-β1, the prototype of TGF-β family, is a potent inducer of extracellular matrix synthesis and is well established as a central mediator in the final common pathway of fibrosis associated with progressive kidney diseases (3, 4). Upon ligand stimulation, TGF-β type I (TβRI) and type II (TβRII) receptors form heterotetrameric complexes, by which TβRI is phosphorylated in the GS domain and activated. Smad signaling pathway is well established as a canonical pathway induced by TGF-β1 (5, 6). Receptor-regulated Smads (Smad2 and Smad3) are recruited and activated by the activated TβRI. The phosphorylation in the GS domain (7) and L45 loop (8) of TβRI are crucial for its interaction with receptor-regulated Smads. After phosphorylation, receptor-regulated Smads are rapidly dissociated from TβRI and interact with common Smad (Smad4) followed by nuclear translocation. In addition to the Smad pathway, a recently emerging body of evidence has demonstrated that TGF-β1 also induces various Smad-independent signaling pathways (917) by which mitogen-activated protein kinases (MAPKs), c-Jun N-terminal kinase (JNK) (18, 19), p38 MAPK (2022), and extracellular signal-regulated kinase 1/2 (23, 24) can be activated by TGF-β1.TAK1, initially identified as a MAPK kinase kinase 7 (MKKK7 or MAP3K7) in the TGF-β signaling pathway (11, 12), also can be activated by environmental stress (25), proinflammatory cytokines such as IL-1 and TNF-α (26, 27) and lipopolysaccharide (28). For TAK1 activation, phosphorylation at Thr-187 and Ser-192 in the activation loop of TAK1 is essentially required (2931). TAK1 can transduce signals to several downstream signaling cascades, including the MAPK kinase (MKK) 4/7-JNK cascade, MKK3/6-p38 MAPK cascade, and nuclear factor κB (NF-κB)-inducing kinase-IκB kinase cascade (2628). A recent report has shown that TAK1 is also activated by agonists of AMP-activated kinase (AMPK) and ischemia, which in turn activates the LKB1/AMPK pathway, a pivotal energy-sensor pathway (32). TAK1 is also involved in Wnt signaling (33). We and others have previously demonstrated that TAK1 is a major mediator of TGF-β1-induced type I collagen and fibronectin expression through activation of the MKK3-p38 MAPK and MKK4-JNK signaling cascades, respectively (3437). Furthermore, increased expression and activation of TAK1 enhance p38 phosphorylation and promote interstitial fibrosis in the myocardium from 9-day-old TAK1 transgenic mice (37). These data implicate a crucial role of TAK1 in extracellular matrix production and tissue fibrosis. TAK1 is also implicated in regulation of cell cycle (38), cell apoptosis (3941), and the Smad signaling pathway (4244). Thus, TAK1 may function as an important regulator and mediator of TGF-β1-induced Smad-dependent and Smad-independent signaling pathways.It has been demonstrated that TAK1 can be activated by the interaction with TAK1-binding protein 1 (TAB1) by in vitro binding assays and in overexpression studies (2931); however, it is not clear whether TAB1 plays a crucial role in ligand-induced TAK1 activation. In embryonic fibroblasts from TAB1 null mice, IL-1 and TNF-α could induce TAK1-mediated NF-κB and JNK activation (45). TAK1 activation induced by TNF-α, IL-1, and T-cell receptor requires TAB2 or its homologous protein TAB3 (4650). Although many questions still remain, much progress has been made in understanding the activation mechanism of TAK1 by inflammatory cytokines (46, 47, 5153). Ligand binding of IL-1 receptor (IL-1R) results in recruitment of MyD88, which serves as an adaptor for IL-1 receptor-associated kinase (IRAK) 1 and 4. Subsequently IRAK1 is hyperphosphorylated and induces interaction with TNF-α receptor-associated factor 6 (TRAF6), resulting in TRAF6 oligomerization. After oligomerization of TRAF6, IRAK1-TRAF6 complex is dissociated from the receptor and associated with TAK1, which is mediated by TAB2 (or TAB3). In this process polyubiquitination of TRAF6 by Ubc13/Uev1A is thought to be critical for the association with TAB2 (or TAB3), which links TAK1 activation (46, 54, 55). In the case of TNF-α stimulation, TNF-α receptors form trimers and recruit adaptor proteins, TRAF2/5, and receptor-interacting protein 1 on the membrane. Ubc13/Uev1A- and TRAF2-dependent polyubiquitination of receptor-interacting protein 1 induce association of TAB2 (or TAB3), which then activates TAK1. Thus, TAB2 is required for ubiquitin-dependent activation of TAK1 by TRAFs. On the other hand, it has been demonstrated that hematopoietic progenitor kinase 1 plays a role as an upstream mediator of TGF-β-induced TAK1 activation, which in turn activates the MKK4-JNK signaling cascade in 293T cells (56, 57). Besides hematopoietic progenitor kinase 1, it has been also suggested that X-linked inhibitor of apoptosis (XIAP) might link TAK1 to TGF-β/BMP receptors through the capability of XIAP to interact with TGF-β/BMP receptors and TAB1 (58). Thus, although various molecules participate in the activation of TAK1, the precise mechanism by which TGF-β1 induces TAK1 activation is incompletely understood. Here, we provide evidence that the association of TAK1 with TGF-β receptors is important for TGF-β1-induced activation of TAK1 in mouse mesangial cells. TGF-β1 stimulation induces interaction of TβRI and TβRII, triggering dissociation of TAK1 from TβRI, and subsequently TAK1 is phosphorylated through TAB1-mediated autophosphorylation, independent of receptor kinase activity of TβRI.  相似文献   

14.
Vasopressin controls renal water excretion largely through actions to regulate the water channel aquaporin-2 in collecting duct principal cells. Our knowledge of the mechanisms involved has increased markedly in recent years with the advent of methods for large-scale systems-level profiling such as protein mass spectrometry, yeast two-hybrid analysis, and oligonucleotide microarrays. Here we review this progress.Regulation of water excretion by the kidney is one of the most visible aspects of everyday physiology. An outdoor tennis game on a hot summer day can result in substantial water losses by sweating, and the kidneys respond by reducing water excretion. In contrast, excessive intake of water, a frequent occurrence in everyday life, results in excretion of copious amounts of clear urine. These responses serve to exact tight control on the tonicity of body fluids, maintaining serum osmolality in the range of 290–294 mosmol/kg of H2O through the regulated return of water from the pro-urine in the renal collecting ducts to the bloodstream.The importance of this process is highlighted when the regulation fails. For example, polyuria (rapid uncontrolled excretion of water) is a sometimes devastating consequence of lithium therapy for bipolar disorder. On the other side of the coin are water balance disorders that result from excessive renal water retention causing systemic hypo-osmolality or hyponatremia. Hyponatremia due to excessive water retention can be seen with severe congestive heart failure, hepatic cirrhosis, and the syndrome of inappropriate antidiuresis.The chief regulator of water excretion is the peptide hormone AVP,2 whereas the chief molecular target for regulation is the water channel AQP2. In this minireview, we describe new progress in the understanding of the molecular mechanisms involved in regulation of AQP2 by AVP in collecting duct cells, with emphasis on new information derived from “systems-level” approaches involving large-scale profiling and screening techniques such as oligonucleotide arrays, protein mass spectrometry, and yeast two-hybrid analysis. Most of the progress with these techniques is in the identification of individual molecules involved in AVP signaling and binding interactions with AQP2. Additional related issues are addressed in several recent reviews (14).  相似文献   

15.
16.
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