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 640 毫秒
1.
2.
Microglia engage in the clearance of dead cells or dangerous debris. When neighboring cells are injured, the cells release or leak ATP into extracellular space and microglia rapidly move toward or extend a process to the nucleotides as chemotaxis through P2Y12 receptors. In the meanwhile, microglia express the metabotropic P2Y6 receptors, the activation of which by uridine 5′-diphosphate (UDP) triggers microglial phagocytosis in a concentration-dependent fashion. UDP/UTP was leaked when hippocampal neurons were damaged by kainic acid in vivo and in vitro. Systemic administration of kainic acid in rats resulted in neuronal cell death in the hippocampal CA1 and CA3 regions, where increases in mRNA for P2Y6 receptors in activated microglia. Thus, the P2Y6 receptor is upregulated when neurons are damaged, and would function as a sensor for phagocytosis by sensing diffusible UDP signals.Key Words: microglia, phagocytosis, P2Y6 receptors, UDPAccumulating findings indicate that nucleotides play an important role in neuron to glia communication through P2 purinoceptors, even though ATP is recognized primarily to be a source of free energy and nucleotides are key molecules in cells. P2 purinoceptors are divided into two families, ionotropic receptors (P2X) and metabotropic receptors (P2Y) (Fig. 1). P2X receptors (seven types; P2X1-P2X7) contain intrinsic pores that open by binding with ATP. P2Y (eight types; P2Y1,2,4,6 and 11–14) are activated by nucleotides and couple to intracellular second-messenger systems through heteromeric G-proteins.1 Microglia express P2X4, P2X7, P2Y2, P2Y6 and P2Y121 and are known as resident macrophages in CNS, accounting for 5–10% of the total population of glia.2,3 When neurons are injured or dead, microglia are activated, resulting in their interaction with immune cells, active migration to the site of injury, release of pro-inflammatory substances and the phagocytosis of damaged cells or debris. For such activation of microglial motilities, extracellular nucleotides have a central role. Extracellular ATP functions as a chemoattractant. Microglial chemotaxis by ATP via P2Y12 receptors was originally found by Honda et al.,4 and has recently been confirmed in vivo in P2Y12 receptor knockout animals.5 Neuronal injury results in the release or leakage of ATP that appears to be a “find-me” signal from damaged neurons to microglia to cause chemotaxis. In addition to microglial migration by ATP, another nucleotide, UDP, an endogenous agonist of the P2Y6 receptor, greatly activates the motility of microglia and orders microglia to engulf damaged neurons.6Open in a separate windowFigure 1P2 purinergic receptors (ATP receptors).Phagocytosis is a specialized form of endocytosis taking relatively large particles (> 1.0 µm) into vacuoles and has a central role in tissue remodeling, inflammation and the defense against infectious agents.7 Phagocytosis is initiated by the activation of cell-surface phagocytosis receptors, including Fc receptors, complement receptors, integrins, endotoxin receptors (CD18, CD14), mannose receptors and scavenger receptors8 which are activated by corresponding extracellular ligands called as “eat-me” signals. Since recognition is the most important step for phagocytosis, extensive studies on phagocytosis receptors have been reported. With regard to apoptotic cells, it is well known that dying cells express so called “eat-me” signals such as phosphatidylserine (PS) on their surface membrane,8 by which microglia recognize the apoptotic cells in order to catch and remove them.8 As for amyloid β protein (Aβ), a key molecule that mediates Alzheimer''s disease, microglia remove Aβ presumably via Fc receptor-dependent phagocytosis.9,10 It, however, is unclear how phagocytotic cells come to the target cells or debris. Our findings suggest that nucleotides might be the molecules to guide phagocytotic cells to the targets.We found that exogenously applied UDP caused microglial phagocytosis through P2Y6 in a concentration-dependent manner, and that neuronal injury caused by kainic acid (KA) upregulated P2Y6 receptors in microglia, the KA evoked neuronal injury resulted in an increase in extracellular UTP, which was immediately metabolized into UDP in vivo and in vitro. We also found that UDP leaked from injured neurons caused P2Y6 receptor-dependent phagocytosis in vivo and in vitro. Thus, UDP could be a diffusible molecule that signals the crisis of damaged neurons to microglia, triggering phagocytosis. Nucleotides seem to have the ability to act as “eat-us” signals for necrotic cells suffering traumatic or ischemic injury because such necrotic cells cause swelling, followed by shrinkage, leading to the leakage of cytoplasmic molecules including a large amount of ATP and UTP and extracellular nucleotides are immediately degraded by ecto-nucleotideases, suggesting that leaked nucleotides could be transient and localized signals that alert to the crisis created by the presence of the necrotic cells. These findings suggest that microglia might be attracted by ATP/ADP4,5,11,12 and subsequently recognize UDP, starting to recognize “eat-me” signals attached to the targets and engulf them (Fig. 2). It is interesting that ATP/ADP is not able to efficiently activate P2Y6 receptors, nor can UDP act on P2Y12 receptors. Thus, adenine and uridine nucleotides would regulate microglial motilities, i.e. chemotaxis and phagocytosis, in a coordinated fashion.Open in a separate windowFigure 2Illustration of nucleotide-activated microglial chemotaxix and phagocytosis. Activated microglia might be attracted by ATP/ADP is not able to efficiently activate P2Y6 receptors, nor ca UDP act on P2Y12 receptors.  相似文献   

3.
As a second messenger, H2O2 generation and signal transduction is subtly controlled and involves various signal elements, among which are the members of MAP kinase family. The increasing evidences indicate that both MEK1/2 and p38-like MAP protein kinase mediate ABA-induced H2O2 signaling in plant cells. Here we analyze the mechanisms of similarity and difference between MEK1/2 and p38-like MAP protein kinase in mediating ABA-induced H2O2 generation, inhibition of inward K+ currents, and stomatal closure. These data suggest that activation of MEK1/2 is prior to p38-like protein kinase in Vicia guard cells.Key words: H2O2 signaling, ABA, p38-like MAP kinase, MEK1/2, guard cellAn increasing number of literatures elucidate that reactive oxygen species (ROS), especially H2O2, is essential to plant growth and development in response to stresses,14 and involves activation of various signaling events, among which are the MAP kinase cascades.13,5 Typically, activation of MEK1/2 mediates NADPH oxidase-dependent ROS generation in response to stresses,4,68 and the facts that MEK1/2 inhibits the expression and activation of antioxidant enzymes reveal how PD98059, the specific inhibitor of MEK1/2, abolishes abscisic acid (ABA)-induced H2O2 generation.6,8,9 It has been indicated that PD98059 does not to intervene on salicylic acid (SA)-stimulated H2O2 signaling regardless of SA mimicking ABA in regulating stomatal closure.2,6,8,10 Generally, activation of MEK1/2 promotes ABA-induced stomatal closure by elevating H2O2 generation in conjunction with inactivating anti-oxidases.Moreover, activation of plant p38-like protein kinase, the putative counterpart of yeast or mammalian p38 MAP kinase, has been reported to participate in various stress responses and ROS signaling. It has been well documented that p38 MAP kinase is involved in stress-triggered ROS signaling in yeast or mammalian cells.1113 Similar to those of yeast and mammals, many studies showed the activation of p38-like protein kinase in response to stresses in various plants, including Arabidopsis thaliana,1416 Pisum sativum,17 Medicago sativa18 and tobacco.19 The specific p38 kinase inhibitor SB203580 was found to modulate physiological processes in plant tissues or cells, such as wheat root cells,20 tobacco tissue21 and suspension-cultured Oryza sativa cells.22 Recently, we investigate how activation of p38-like MAP kinase is involved in ABA-induced H2O2 signaling in guard cells. Our results show that SB203580 blocks ABA-induced stomatal closure by inhibiting ABA-induced H2O2 generation and decreasing K+ influx across the plasma membrane of Vicia guard cells, contrasting greatly with its analog SB202474, which has no effect on these events.23,24 This suggests that ABA integrate activation of p38-like MAP kinase and H2O2 signaling to regulate stomatal behavior. In conjunction with SB203580 mimicking PD98059 not to mediate SA-induced H2O2 signaling,23,24 these results generally reveal that the activation of p38-like MAP kinase and MEK1/2 is similar in guard cells.On the other hand, activation of p38-like MAP kinase23,24 is not always identical to that of MEK1/28,25 in ABA-induced H2O2 signaling of Vicia guard cells. For example, H2O2- and ABA-induced stomatal closure was partially reversed by SB203580. The maximum inhibition of both regent-induced stomatal closure were observed at 2 h after treatment with SB203580, under which conditions the stomatal apertures were 89% and 70% of the control values, respectively. By contrast, when PD98059 was applied together with ABA or H2O2, the effects of both ABA- and H2O2-induced stomatal closure were completely abolished (Fig. 1). These data imply that the two members of MAP kinase family are efficient in H2O2-stimulated stomatal closure, but p38-like MAP kinase is less susceptive than MEK1/2 to ABA stimuli.Open in a separate windowFigure 1Effects of SB203580 and PD98059 on ABA- and H2O2-induced stomatal closure. The experimental procedure and data analysis are according to the previous publication.8,23,24It has been reported that ABA or NaCl activate p38 MAP kinase in the chloronema cells of the moss Funaria hygrometrica in 2∼10 min.26 Similar to this, SB203580 improves H2O2-inhibited inward K+ currents after 4 min and leads it to the control level (100%) during the following 8 min (Fig. 2). However, the activation of p38-like MAP kinase in response to ABA need more time, and only recovered to 75% of the control at 8 min of treatment (Fig. 2). These results suggest that control of H2O2 signaling is required for the various protein kinases including p38-like MAP kinase and MEK1/2 in guard cells,1,2,8,23,24 and the ABA and H2O2 pathways diverge further downstream in their actions on the K+ channels and, thus, on stomatal control. Other differences in action between ABA and H2O2 are known. For example, Köhler et al. (2001) reported that H2O2 inhibited the K+ outward rectifier in guard cells shows that H2O2 does not mimic ABA action on guard cell ion channels as it acts on the K+ outward rectifier in a manner entirely contrary to that of ABA.27Open in a separate windowFigure 2Effect of SB203580 on ABA- and H2O2-inhibited inward K+ currents. The experimental procedure and data analysis are according to the previous publication.24 SB203580 directs ABA- and H2O2-inactivated inward K+ currents across plasma membrane of Vicia guard cells. Here the inward K+ currents value is stimulated by −190 mV voltage.Based on the similarity and difference between PD98059 and SB203580 in interceding ABA and H2O2 signaling, we speculate the possible mechanism is that the member of MAP kinase family specially regulate signal event in ABA-triggered ROS signaling network,14 and the signaling model as follows (Fig. 3).Open in a separate windowFigure 3Schematic illustration of MAP kinase-mediated H2O2 signaling of guard cells. The arrows indicate activation. The line indicates enhancement and the bar denotes inhibition.  相似文献   

4.
Long-term clinical outcomes are dependent on whether carcinoma cells leave the primary tumor site and invade through adjacent tissue. Recent evidence links tissue rigidity to alterations in cancer cell phenotype and tumor progression. We found that rigid extracellular matrix (ECM) substrates promote invasiveness of tumor cells via increased activity of invadopodia, subcellular protrusions with associated ECM-degrading proteinases. Although the subcellular mechanism by which substrate rigidity promotes invadopodia function remains to be determined, force sensing does appear to occur through myosin-based contractility and the mechanosensing proteins FAK and p130Cas. In addition to rigidity, a number of ECM characteristics may regulate the ability of cells to invade through tissues, including matrix density and crosslinking. 3-D biological hydrogels based on type I collagen and reconstituted basement membrane are commonly used to study invasive behavior; however, these models lack some of the tissue-specific properties found in vivo. Thus, new in vitro organotypic and synthetic polymer ECM substrate models will be useful to either mimic the properties of specific ECM microenvironments encountered by invading cancer cells or to manipulate ECM substrate properties and independently test the role of rigidity, integrin ligands, pore size and proteolytic activity in cancer invasion of various tissues.Key words: cancer, invasion, invadopodia, rigidity, mechanotransduction, microenvironmentIn multicellular organisms, cells must sense and respond to multiple cues for proper functioning within tissues. Although most experimental research has focused on the regulation of cellular processes by external chemical signals, there is increasing recognition that mechanical forces also regulate critical cellular functions. Indeed, rigidity of the extracellular environment has been shown to regulate such diverse processes as muscle cell differentiation, stem cell lineage fate, breast epithelial signaling and phenotype, and fibroblast motility.15In breast cancer, accumulating evidence suggests a role for tissue rigidity in promoting both the formation and invasiveness of tumors. Mammographic density of breast tissue has been correlated with increased cancer risk and included in models to predict the likelihood of in situ and invasive breast cancers.6 Histologically, dense breast tissue has increased stromal collagen content and in vitro analyses have shown that cancerous breast tissue is much stiffer than normal tissue (as represented by values for the elastic or Young''s moduli).3,7 In addition, experimentally increased expression of collagen fibrils in a mouse mammary model of spontaneous breast cancer was recently shown to promote tumor formation, invasion and metastasis.8 Therefore, both clinical and animal data suggest a correlation between tissue density and cancer aggressiveness, and mechanical factors appear likely to play a role in this process.9A well-established mechanism by which extracellular matrix (ECM) rigidity signals can drive phenotypic transformations is through mechano-signal transduction (mechanotransduction) pathways in which external forces are transmitted via integrin receptors at linear focal adhesion structures to cytoskeletal and signaling proteins inside the cell. Actomyosin contractility leads to stretching and activation of proteins such as talin, p130Cas and potentially focal adhesion kinase (FAK).1012 For example, stem cell lineage was found to be dependent on formation of cellular focal adhesions and actomyosin contractility in response to substrate tensile properties.2 Mammary epithelial cells grown on compliant matrices will differentiate and polarize to form lactating 3-D structures that resemble in vivo acini but fail to do so on stiff matrices due to increased cytoskeletal contractility.3 Activation of mechanotransduction molecules, such as FAK, Rho and ROCK, are required for the rigidity-induced phenotype changes.3,5 Using polyacrylamide (PA) gel systems, Yu-li Wang''s group found that rigid substrates induce fibroblast and epithelial cells to migrate away from each other instead of aggregating to form tissue-like structures.13 This transformation in phenotype is characteristic of the epithelial to mesenchymal transition and thought to be crucial for tumor cell migration.14A critical feature of tumor aggressiveness is the ability to invade across tissue boundaries, through degradation of ECM. The subcellular structures responsible for this invasive activity are thought to be invadopodia: actin-rich, finger-like cellular protrusions that proteolytically degrade local ECM. These structures are characteristic of invasive cells and have been implicated in tumor cell metastasis due to their association with ECM degradation.15 Similar structures, podosomes, are formed in src-transformed cells, as well as normal cells such as osteoclasts and dendritic cells that need to degrade matrix and/or cross tissue boundaries.16 In addition to mediating ECM degradation, podosomes have been postulated to function as adhesion structures, since well-characterized adhesion proteins localize to podosomes and many podosome-expressing cells no longer express focal adhesions.17 Furthermore, podosomes have been shown to be essential for chemotactic motility and transendothelial migration, although not for chemokinetic motility.18,19We recently found that ECM rigidity increases both the number and activity of invadopodia, and this effect was dependent on the cellular contractile machinery (Fig. 1A).20 Consistent with a role for mechanotransduction in this process, we found localization of the active, phosphorylated forms of the mechanosensing proteins FAK and p130Cas in actively degrading invadopodia and an increase in invadopodia-associated degradation in breast cancer cells overexpressing FAK and p130Cas. These results suggest that in breast cancer, increases in tissue rigidity may directly lead to increased cellular invasiveness and tumor progression.Open in a separate windowFigure 1Potential rigidity sensing mechanisms by invadopodia. (A) Invadopodia are typically identified by colocalization of fluorescent antibodies for actin and cortactin at puncta that correspond to areas of ECM degradation visualized as dark regions in FITC-labeled fibronectin (Fn) overlaying gelatin. In this case, ECM was layered on top of either soft (storage modulus = 360 Pa) or hard (storage modulus = 3,300 Pa) polyacrylamide gels (PA) to determine if invadopodia activity was regulated by differences in mechanical properties. On hard PA, invasive MCF10ACA1d breast carcinoma cells produced more invadopodia and degraded more ECM than on soft PA. Yellow arrows indicate examples of invadopodia. (B) The localization of rings of the contractile protein myosin IIA (myoIIA) surrounding invadopodia (actin puncta) suggests a role for these structures in mechanosensing by potentially linking invadopodia with the contractile apparatus to detect differences in substrate rigidity. An example ring structure is indicated with a yellow arrow and shown in the zoomed portion of the myosin IIA image, and an example of no or weak localization of myosin IIA with an invadopodium is indicated with the red arrow. (C) Activated forms of FAK and p130Cas localize to invadopodia and depend on cytoskeletal contractility.20 Rings of myosin IIA also frequently surround invadopodia. These results suggest that invadopodia may act as mechanosensing organelles, either directly through localized mechanoresponsiveness at the invadopodia or through longer-range connections to neighboring or even distant focal adhesions. In either case, traction forces may be generated as a result of changes in cytoskeletal tension in response to ECM properties. Alternatively, invadopodia function could be regulated in the absence of local traction forces, secondary to distant intracellular signaling that leads to alterations in whole cell phenotypic changes. (A and B) are reprinted from Current Biology, Volume 18, Nelson R. Alexander, Kevin M. Branch, Aron Parekh, Emily S. Clark, Izuchukwu C. Iwueke, Scott A. Guelcher and Alissa M. Weaver, Extracellular Matrix Rigidity Promotes Invadopodia Activity, pp. 1295–9, 2008; with permission from Elsevier.The localization of phosphorylated FAK and p130Cas at invadopodia and the requirement for actomyosin contractility in our study suggests that invadopodia have the potential to act as mechanosensing organelles. This concept is supported by our finding that ∼40% of breast cancer cells cultured on rigid substrates had rings of myosin IIA surrounding invadopodia (Fig. 1B)20 and the recent finding that similar podosome structures can exert local traction forces.21 In addition, a few studies have implicated integrin activity in invadopodia function as well as localized β1 and β3 integrins to invadopodia.2225 However, whether invadopodia can serve as tension-generating adhesion structures is controversial, in part because of the presence of both focal adhesions and invadopodia in many cancer cells (Fig. 1C).Regulation of invadopodia and podosome function is also not straightforward. Although our data,20 along with results from Collin et al.,21 suggests that mechanical tension promotes invadopodia and podosome activity, in some systems podosome formation is promoted by a loss rather than a gain of cytoskeletal tension. That is, local cytoskeletal relaxation has been shown to promote podosome formation coincident with focal adhesion dissolution in both vascular smooth muscle cells treated with phorbol ester26 and neuroblastoma cells.27 A yin-yang activity between focal adhesions and podosomes has been known for many years, whereby activation of src kinase leads to both disassembly of focal adhesions28 and formation of podosomes.29 However, the role of tension in this process is unclear, particularly since activation of src kinase occurs downstream of mechanical stimuli30 and should promote podosome/invadopodia activity, yet loss of tension apparently induces biological activities dependent on src kinase (focal adhesion disassembly and podosome formation).26,27 For invadopodia, the role of tension is even less clear. Basic characterization studies need to be performed to establish molecular and structural differences between invadopodia and focal adhesions and to measure force profiles at the two structures. Since invadopodia have much smaller diameters compared to podosomes (50–100 nm vs ∼1 µm, respectively),15,16 the latter task of determining traction forces may be difficult due to resolution limitations in measuring potentially tiny substrate displacements. The standard identification of invadopodia, by association of actin-rich puncta with sites of degradation of fluorescent ECM, adds another technical limitation since the thickness and fluorescence of the ECM matrix used to identify proteolytic activity may hinder visualization of embedded fluorescent beads in the underlying PA gel (displacement of beads is typically used to calculate traction forces).31 Thus, an important future direction should be the development of new in vitro experimental systems that have manipulable substrate properties and allow simultaneous identification of subcellular forces and proteolytic activity.The cellular response to rigidity is often characterized using PA gels with tunable stiffness in the range spanning that of normal and cancerous breast tissue (elastic moduli = 100–10,000 Pa).3,7 PA gels will likely continue to be invaluable tools for understanding cellular responses to rigidity. However, this system is inherently simple and cannot fully replicate cellular events occurring in a complex in vivo ECM microenvironment. Given that invading breast cancer cells are likely to experience different microenvironments as they cross through the basement membrane (BM) and into neighboring collagenous stromal tissue (Fig. 2), biological hydrogels such as reconstituted basement membrane (Matrigel) and type I collagen gels are often utilized to mimic these ECM substrates. However, both of these models lack many of the chemical, physical, and mechanical characteristics of tissues found in vivo and have been recently questioned as suitable models for studying cancer cell invasion.32 Type I collagen gels have a fibrillar architecture but a low density and high porosity33 and frequently lack crosslinking sites.34 Although Matrigel contains many of the biochemical components of the BM, it is tumor-derived35 and the major component is laminin-1, which is only abundant in fetal tissues.36 By contrast, the major component of normal BM is type IV collagen. In addition, Matrigel is a solubilized preparation that lacks crosslinks37 and a fibrillar component.38 Both sparse collagen gels and Matrigel are quite compliant with Young''s moduli of ∼1,000 and ∼200 Pa, respectively;3 therefore, without further manipulation these substrates lack the rigidity required to mimic tumor-associated ECM.Open in a separate windowFigure 2Navigation of basement membranes and stromal collagen by invading cancer cells. Invasive cancer cells are thought to navigate different tissue microenvironments in the process of invasion. In order for invasion to occur, tumor cells must first breach the basement membrane, a thin and highly crosslinked specialized ECM that requires proteolytic degradation for subsequent transmigration. Once past this barrier, cells must proceed through the neighboring stroma composed of collagenous connective tissue. The meshwork in the stroma is looser and may facilitate diverse migration modes dependent on local microenvironmental conditions and cellular cohesiveness. These modes of migration include a single cell, proteinase-independent amoeboidal phenotype (left) and single cell (middle) and collective (right) proteinase-dependent mesenchymal phenotypes that locally degrade matrix at enzymatically active invadopodia. Note the absence of collagen stroma surrounding and along the migration track of proteolytically active cells. New physiologically relevant models that mimic these interactions in vitro will be useful to elucidate mechanisms of cancer cell migration and invasion in various tissues.In order to invade neighboring stromal tissue, carcinoma cells must first breach the BM, a complex, interwoven meshwork composed of type IV collagen, laminin, nidogen/entactin, and various proteoglycans and glycoproteins.32 The highly ordered and crosslinked type IV collagen network is regarded as the limiting barrier to cancer cell invasion since it forms pores on the order of 100 nm that are too small for passage of cells without proteolytic degradation of the BM.32 In addition to degradation, decreased BM synthesis may contribute to the initial steps of cancer invasion by altering the balance between BM formation and remodeling.39 Once cancer cells cross the BM, they encounter stromal collagen tissue. In tumors, this desmoplastic stroma is frequently fibrotic due to increased ECM deposition and crosslinking by carcinoma-associated fibroblasts.9 Although controversial, cancer cells are thought to use a nonproteolytic, amoeboid mode to traverse this connective tissue;40 therefore, different modes of migration may be necessary to traverse BM or stromal collagenous matrices (Fig. 2). However, the amoeboid phenotype has been described using either sparse collagen gels without crosslinks41 or Matrigel.42 In vivo, the process of invading through tumor-associated stromal collagen is likely to depend on the pore size, the crosslinking status, and whether cells are migrating collectively or individually.34,43In light of these concerns and many others, there has been a push for more physiologically relevant in vitro models that represent closer approximations of BM or stromal collagen tissue. Successful models, whether natural or synthetic, must be able to mimic the composition, architecture and mechanical properties of the in vivo environment as well as support cell culture in ex vivo conditions. Natural substrates can be produced by cultured cells, such as the epithelial basement membranes synthesized by MDCK cells.37 Alternatively, organotypic models derived from biological specimens have recently been utilized to study invasion. These materials can be based on processed biological tissue, such as detergent-extracted mouse embryo sections,44 homogenized involution matrix,38 and decellularized human dermis,45 or on native tissue such as chick chorioallantoic membrane46 and explanted peritoneal or mammary tissue.34,37 In addition, the field of tissue engineering has already provided novel hybrid scaffolds and advanced tissue culturing methods that can be utilized for cancer research.47 Biological materials developed for clinical use in tissue reconstruction and regeneration, such as small intestinal submucosa and urinary bladder matrix, are attractive candidates as new in vitro models since they maintain their tissue-like properties and have been extensively characterized.48,49 These tissue-derived scaffolds are composed of well-defined structural and functional proteins, originally produced by cells in vivo, and maintain their complex 3-D architecture. Thus, such materials can provide an environment that recapitulates the chemical, physical and mechanical properties found in vivo.48 In addition, synthetic materials, such as poly(ethylene glycol)-based hydrogels, will likely play a large role in cancer research since they can be designed with defined chemistries to obtain appropriate physical and mechanical properties as well as specific spatial arrangements of biologically relevant moieties on relevant length scales.33,50 Similarly, engineered adhesive microenvironments created with microfabrication techniques can also be utilized to probe molecular and cellular phenomena.51 Due to this flexibility in fabrication, these materials are good candidates for novel in vitro models to probe the effects of specific mechanical, topographical and chemical factors on cellular migration and invasion.In summary, the physical microenvironment is increasingly recognized as a major influence on cellular phenotype. Recent data emphasizes the importance of mechanical factors in tumor progression, including cellular invasiveness. Exciting future directions include understanding how stromal and BM environments affect cellular invasiveness at multiple scales, including subcellular and molecular regulation of ECM degradation in response to ECM rigidity and the role of proteinases in crossing diverse tissue barriers. The development of novel model systems with appropriate biological and physical properties will facilitate all of these goals.  相似文献   

5.
Growth factors regulate a diverse array of cellular functions including proliferation, survival and movement, and the ability to do this often involves interactions with the extracellular matrix (ECM) and particularly heparan sulfate proteoglycans (HSPGs). HSPGs have been shown to sequester growth factors and to act as growth factor co-receptors or receptors themselves. Recent studies, however, have revealed a new role for HSPGs in mediating the interactions of growth factors with the ECM. Specifically, heparan sulfate has been shown to modulate fibronectin structure to reveal previously masked growth factor binding sites. In vivo, this mechanism appears to control the guidance of migrating cells during embryonic development as HSPG-modification of fibronectin enables direct platelet derived growth factor-fibronectin interactions necessary for this process. A model based on this observation is discussed here as well as the possibility that other growth factors/morphogens utilize similar mechanisms involving fibronectin or additional ECM proteins.Key words: heparin, heparan sulfate, fibronectin, cell movement, cell motility, mesoderm, mesendoderm, embryonic development, embryo, XenopusThe ability of cells to move is critical to normal processes, such as embryonic development and wound healing, and is also a feature of pathological conditions such as cancer metastasis. In normal circumstances, cell movement is tightly regulated and leads to precise cell reorganizations. This is particularly evident during embryonic development where cells need to be in the right place at the right time to give and receive signals, to form tissues and organs and ultimately, to become a functional organism. To achieve this, cells must be able to move in a directed way. This requires that cells sense and move in response to extracellular signals, while coordinating changes in shape with the creation of traction and force and balancing attachment and detachment to the extracellular matrix (ECM) and neighboring cells.Evidence from a number of systems suggests that growth factors play a significant role in the spatial and temporal control of cell movements. In particular, members of the platelet derived growth factor (PDGF)/vascular endothelial growth factor (VEGF) family act as guidance cues for cells in a variety of processes including the development of both invertebrate and vertebrate embryos (reviewed in ref. 1). How such cues are laid down is an area of intensive investigation; however, a new study from our group suggests that the microenvironment within the ECM, principally the composition of heparan sulfate, modulates direct interactions between PDGF and fibronectin necessary for directed migration of embryonic cells during Xenopus development.2During the development of all vertebrates, cell movements begin at gastrulation, when changes in the shape, adhesion and motility of cells drive the complex series of tissue rearrangements necessary to establish the basic body plan of the embryo. In Xenopus embryos, some of the first cells to move during gastrulation are the anterior mesendoderm cells. These cells migrate in a characteristic pattern that provides a tractable model to identify the molecular control of directed cell movement. Anterior mesendoderm cells move away from the blastopore lip and towards the animal pole across a fibrillar, fibronectin-rich ECM that covers the ectoderm (see Fig. 1A; reviewed in ref. 3). They migrate as a sheet and are polarized with lamelliform protrusions extending in the direction of movement and the trailing edges underlapping one another like shingles on a roof (reviewed in ref. 3). This mesendoderm movement is guided by PDGF-A signaling.4 Immediately prior to and during Xenopus gastrulation, PDGF-A is expressed in the ectoderm while its receptor, PDGFRα, is expressed in the migrating mesendoderm cells (reviewed in ref. 5). Disruption of PDGF-A signaling by interference with either the receptor or ligand in vivo or in an ex vivo assay disorients the cells and the direction of migration is randomized, although the cells are still able to physically move.4Open in a separate windowFigure 1Heparan sulfate modification of fibronectin enables PDGF-AA-fibronectin interactions necessary for directed cell migration during Xenopus gastrulation. (A) A schematic of a Xenopus embryo immediately after the onset of gastrulation (stage 10+). The prospective mesoderm (red), ectoderm (blue) and endoderm (yellow) are indicated. The black circle indicates migrating anterior mesendoderm cells and the region of tissue shown in (B and C). (B) In vivo the inner surface of the ectoderm (blue blocks) is lined with a fibronectin-rich ECM. This fibronectin is proposed to be in its open conformation (blue dumbbells) due to the presence of heparan sulfate proteoglycans on the surface of these cells (indicated in red). PDGF (green triangle) secreted by the ectoderm cells binds to the C-terminus of the fibronectin. This PDGF is then available for PDGFRα (green fork) binding and activation. The PDGFRα is expressed on the migrating mesendoderm cells. This ligand receptor interaction guides the directed movement of these cells. (C) Heparinase III treatment of these tissues during matrix deposition digests the heparan sulfate side chains leaving the proteoglycan core (red line), which does not alter fibronectin structure leaving it in the closed conformation. PDGF secreted by the ectoderm cells can then diffuse from the site of secretion leaving it available to bind PDGFRαs over the entire surface of the mesendoderm cells and thus, disrupting directed movement. (D and E) A schematic of a “higher magnification” view of one cell in (B and C). A key of the symbols used in (B–D) is shown.PDGF-A must be associated with the ECM for directed mesendoderm migration. In an ex vivo assay, overexpression of the long form of PDGF-A, which associates with the ECM via a positively charged carboxy-terminal retention motif (reviewed in ref. 6) in the ectoderm cells, disrupts the normal guidance cues causing randomization of mesendoderm movement.4 In contrast, overexpression of the short form of PDGF-A, which diffuses from its site of secretion,6 does not.4 Our recent work now suggests that this growth factor-ECM association represents a direct interaction between PDGF-A and fibronectin.2 Importantly, this interaction requires that the fibronectin is pre-exposed to heparan sulfate proteoglycan (HSPG) before PDGF-A can bind.2 A similar interaction between the related growth factor VEGF and fibronectin is also enhanced by pre-treatment of the fibronectin with heparin or heparan sulfate.79 Heparin acts by modifying the structure of fibronectin,8 mediating the transition from the globular form to a stable, more extended conformation,7,8,1012 which reveals growth factor binding sites.7,8 These data are significant because in an ex vivo directed migration assay, treatment of the native ECM with heparinase III, which degrades heparan sulfate chains, abolishes the guidance cues and results in randomized mesendoderm migration.2 Here we propose a model in which prior to and during Xenopus gastrulation, fibronectin is modified to an extended conformation by endogenous HSPGs as it is being laid down (Fig. 1B and D). PDGF-AA secreted from the ectoderm cells binds to this modified fibronectin and is thus, retained in the ECM (Fig. 1B and D). PDGF-AA then guides the mesendoderm cells by local activation of their cell surface PDGFRαs (Fig. 1B and D). In the absence of HSPG, the fibronectin remains in a more globular form and PDGF-AA binding sites are shielded (Fig. 1C and E). In this case, it is proposed that PDGF-AA diffuses away from the site of secretion resulting in a less localized activation of PDGFRαs and a loss of persistent cell migration (Fig. 1C and E). This idea is supported by data showing that flooding the native ECM with exogenous PDGF-AA prior to performing the ex vivo migration assay, also causes a randomization of directed movement.4 In this case, it seems likely that any gradient or uneven distribution of PDGF-AA is overwhelmed and the guidance cues consequently lost.In cultured cells, PDGF can also associate with HSPGs via its retention motif.13 However, in the Xenopus ex vivo directed migration assay, heparinase treatment of the ectoderm ECM after its deposition does not alter the direction of mesendoderm migration.2,14 This suggests that in this system, PDGF-AA interacts directly with fibronectin and not with heparan sulfate chains, since heparinase III would have degraded them and caused the release of HS-bound PDGF-AA.13The type of proteoglycan as well as the specific structural composition of their heparan sulfate chains likely contributes to the modulation of PDGF-fibronectin binding. Both chain length and the chemical composition of heparan sulfate play crucial roles in the binding of VEGF to fibronectin, in which long (>22 saccharides) heparin chains with sulfation on the 6-O and N positions of glucosamine units are required for activity.8 This suggests that a localized control of heparan sulfate biosynthesis through the regulation of the array of enzymes involved might be a critical component to directed cell migration. Interestingly, an analysis of heparan sulfate expression during mouse lung development noted dynamic and discrete localization within the mesenchyme at sites of FGF10-mediated epithelial budding,15 indicating that rapid alterations of heparan sulfate structure may provide a general mechanism for positional control of growth factor activity.Several HSPGs are expressed in tissue specific patterns during Xenopus gastrulation including Syndecans-1 and -2 (expressed in the ectoderm);16 and Syndecan-4, Glypican-4 and Biglycan (expressed in the ectoderm and mesoderm).1619 Although the role(s) of these HSPGs in mesendoderm migration are not known, they do affect embryological processes that involve cell rearrangements. For example, Syndecan-2 has recently been shown to regulate the migration of organ primordia and fibrillogenesis in zebrafish embryos.20 During gastrulation, knock-down of either Syndecan-1 or -2 disrupts the native ECM for mesendoderm migration and results in gaps in fibril deposition.21 In addition, Syndecan-4 and Glypican-4 are both essential for convergent extension, another important form of cell movement during gastrulation that involves the coordinated rearrangement of cells.17,18 Similarly, Syndecan-4 is necessary for the directional migration of neural crest cells.22 In these cases, Syndecan-4 and Glypican-4 regulate the planar cell polarity (PCP) pathway, which is also involved in polarized matrix deposition during convergent extension.23 Compellingly, embryos develop with anterior defects when either Syndecan-4 or Glypican-4 is knocked-down, indicating that mesendoderm cell migration is also likely to be affected.17,18The Slb/Wnt-11 mediated-PCP pathway has been shown in zebrafish embryos to regulate the polarity and directional movement of ingressing mesendoderm cells.24 In Slb/Wnt-11 mutants, however, although these processes are disrupted they are not completely abolished and evidence suggests that PDGF signaling through phosphatidyl inositol 3-kinase (PI3K) is responsible for establishing cell polarity and controlling the velocity of these cells.25 Rho GTPases act downstream in both PCP and PDGF signaling pathways. Recent evidence suggests that RhoA and Rac1 are important for the polarity and protrusive activity of migrating mesendoderm cells and play a role in the guidance of neural crest cells,22,26 but the mechanisms by which these pathways are integrated with signals from other factors that also modulate these processes are still to be determined.How retention of PDGF-AA in the ECM translates into directional migration still needs to be determined. For example, PDGF-AA may be present in a gradient such that receptors at the front of each cell are exposed to a different level of PDGF than those at the rear. Alternatively, control might involve selective expression of molecules that locally release PDGF from matrix binding sites. Either process might provide a means for uneven activation of PDGFRs and a polarized distribution of downstream signaling components at the leading and lagging edges. Such signaling mechanisms have been identified in other cell types and organisms including leukocytes, fibroblasts and Dictyostelium in which the initial response to a chemoattractant creates cell polarity that is further amplified by feedback (reviewed in ref. 27). For example, in Dictyostelium PI3K and PTEN (phosphatase and tensin homolog) are regulated reciprocally to control the spatial and temporal levels of phosphatidyl inositol-3,4,5-triphosphate (PI(3,4,5)P3). PI(3,4,5)P3 is localized to the leading edge of the cell in response to an extracellular signal, whereas PTEN is downregulated in these regions. Recruitment of proteins that bind preferentially to PI(3,4,5)P3, many of which affect remodeling of the actin cytoskeleton, further enhances this polarity in the direction of the chemoattractant.27 In Xenopus, PDGF signaling does appear to regulate cell polarity since overexpression of PDGF-A in the ectoderm disorients the protrusive activity and abolishes the normal shingle arrangement of the receptor expressing, migrating mesendoderm cells.4 Similarly, in the zebrafish mesendoderm, PDGF signaling induces cell protrusions and polarization.25 In this case, upon PDGF treatment, protein kinase B (PKB)/Akt become localized to the leading edge of the cells in a PI3K dependent manner.25 Such mechanisms require the activation of PDGFR at the leading edge of the cell. In mammalian cells, however, it has been demonstrated that activation of epidermal growth factor receptor (EGFR) can spread laterally in the cell membrane.28 If the same is true for PDGFR, even cells exposed to a gradient or localized source of PDGF-AA may lose spatial information. Recent evidence in Drosophila, however, suggests an alternate mechanism involving receptor endocytosis that ensures that a graded extracellular signal is maintained intracellularly.29The Drosophila member of the PDGF/VEGF family, PVF1, along with epidermal growth factor, directs the migration of border cells towards the oocyte during oogenesis.3032 The patterns of expression of ligand and receptor are similar to those during Xenopus gastrulation in that a stationary cell, the oocyte, expresses the ligand, PVF1, while the motile border cells express its receptor, PVR.30 PVR activation is concentrated at the leading edge, and specifically maintained in a localized region of the cell by endocytosis.29 PDGFRβ endocytosis also regulates PDGF-dependent migration of NIH3T3 cells in culture.33 This process involves recruitment of a ternary complex of DOCK4-Grb2-Dynamin2 to the leading edge of the migrating cell, which results in rapid receptor endocytosis, where it is maintained in an active phosphorylation state and leads to polarized signaling. Disruption of the complex, for example by overexpression of DOCK4 ΔC, which cannot bind Grb2, blocks chemotaxis in response to PDGF. Whether a similar mechanism creates spatial signaling during Xenopus gastrulation requires further investigation, but this tantalizing evidence suggests that it is a possibility.  相似文献   

6.
Recently we have studied the secretion pattern of a pectin methylesterase inhibitor protein (PMEI1) and a polygalacturonase inhibitor protein (PGIP2) in tobacco protoplast using the protein fusions, secGFP-PMEI1 and PGIP2-GFP. Both chimeras reach the cell wall by passing through the endomembrane system but using distinct mechanisms and through a pathway distinguishable from the default sorting of a secreted GFP. After reaching the apoplast, sec-GFP-PMEI1 is stably accumulated in the cell wall, while PGIP2-GFP undergoes endocytic trafficking. Here we describe the final localization of PGIP2-GFP in the vacuole, evidenced by co-localization with the marker Aleu-RFP, and show a graphic elaboration of its sorting pattern. A working model taking into consideration the presence of a regulated apoplast-targeted secretion pathway is proposed.Key words: cell wall trafficking, endocytosis, GPI-anchor, PGIP2, PMEI1, secretion pathway, vacuole fluorescent markerCell wall biogenesis, growth, differentiation and remodeling, as well as wall-related signaling and defense responses depend on the functionality of the secretory pathway. Matrix polysaccharides, synthesized in the Golgi stacks, and cell wall proteins, synthesized in the ER, are packaged into secretory vesicles that fuse with the plasma membrane (PM) releasing their cargo into the cell wall. Also the synthesis and deposition of cellulose itself are driven by the endomembrane system which controls the assembly, within the Golgi, and the export to the plasma membrane of rosette complexes of cellulose synthase.1 Secretion to the cell wall has always been considered a default pathway2 but recent studies have evidenced a complex regulation of wall component trafficking that does not seem to follow the default secretion model. Recent evidence that several cell wall proteins are retained in the Golgi stacks until specific signals at the N-terminal domain are proteolitically removed is a case in point.35 Moreover, it has previously been reported that secretion of exogenous marker proteins (secGFP and secRGUS) and cell wall polysaccharides reach the PM through different pathways.6 More recently, we have reported that cell wall protein trafficking also occurs through mechanisms distinguishable from that of a secreted GFP suggesting that more complex events than the mechanisms of bulk flow control cell wall growth and differentiation.7 To follow cell wall protein trafficking we used a Phaseolus vulgaris polygalacturonase inhibitor protein (PGIP2) and an Arabidopsis pectin methylesterase inhibitor protein (PMEI1) fused to GFP (PGIP2-GFP and secGFP-PMEI1). Both apoplastic proteins are involved in the remodeling of pectin network with different mechanisms. PGIP2 specifically inhibits exogenous fungal polygalacturonases (PGs) and is involved in the plant defense mechanisms against pathogenic fungi.8,9 PMEI1 counteracts endogenous PME and takes part in the physiological synthesis and remodeling of the cell wall during growth and differentiation.10,11 The specific functions of the two apoplastic proteins seem to be strictly related to the distinct mechanisms that control their secretion and stability in the cell wall. In fact, while secGFP-PMEI1 moves through ER and Golgi stacks linked to a glycosyl phosphatidylinositol (GPI)-anchor, PGIP2-GFP moves as a cargo soluble protein. Furthermore, secGFP-PMEI1 is stably accumulated in the cell wall, while PGIP2-GFP, over the time, is internalized into endosomes and targeted to vacuole, likely for degradation. After reaching the cell wall, the different fate of the two proteins seems to be strictly related to the presence/absence of their physiological counteractors. PMEI regulates the demethylesterification of homogalacturonan by inhibiting pectin methyl esterase (PME) activity through the formation of a reversible 1:1 complex which is stable in the acidic cell wall environment.12 Stable wall localization of PMEI1 is likely related to its interaction with endogenous PME, always present in the wall. Unlike PMEs, fungal polygalacturonases (PGs), the physiological interactors of PGIP2, are present in the cell wall only during a pathogen attack. The absence of PGs may determine PGIP2 internalization. Internalization events have been already reported for PM proteins,1316 while cell wall protein internalization is surely a less well-known event. To date, only internalization of an Arabidopsis pollen-specific PME4,5,17 and PGIP2 7 has been reported.To further confirm the internalization of PGIP2-GFP and its final localization into the vacuole, we constructed a red fluorescent variant (RFP) of the green fluorescent marker protein that accumulates in lytic or acidic vacuole because of the barley aleurain sorting determinants (Aleu-RFP).18 The localization of PGIP2-GFP was compared to that of Aleu-RFP by confocal microscopy in tobacco protoplasts transiently expressing both fusions. Sixty hours after transformation, PGIP2-GFP labeled the central vacuole as indicated by complete co-localization with the vacuolar marker (Fig. 1A–D). Instead, at the same time point, secGFP-PMEI1 still labeled the cell wall (Fig. 1E–H) and never reached the vacuolar compartment. To summarize PGIP2-GFP secretion pattern, a graphic elaboration of confocal images is reported describing the sorting of PGIP2GFP in tobacco protoplast (Fig. 1I). The protein transits through the endomembrane system (green) and reaches the cell wall which is rapidly regenerating as evidenced by immunostaining with the red monoclonal antibody JIM7 that binds to methylesterified pectins.19 PGIP2-GFP is then internalized in endosomes, labeled in yellow because of the co-localization with the styryl dye FM4-64, a red marker of the endocytic pathway.Open in a separate windowFigure 1PGIP2-GFP, but not secGFP-PMEI1, is internalized and reaches the vacuole in tobacco leaf protoplasts. (A) Approximately 60 h after transformation, PGIP2-GFP labeled the central vacuole as indicated by co-localization with the vacuole marker Aleu-RFP (B). (C) Merged image of (A and B). (D) Differential interference contrast (DIC) image of (A–C). On the contrary, secGFP-PMEI1 still labeled cell wall (E). (F) No co-localization is present in the vacuole labeled by Aleu-RFP. (G) Merged image of (E and F). (H) DIC image of (E–G). (I) Graphic elaboration of confocal images describing the sorting of PGIP2. The protein is sorted by the endomembrane system (green) to the cell wall (red) that is regenerated by the protoplast. Lacking the specific ligand, it is then internalized in endosome (yellow). Details are reported in the text.In Figure 2 we propose a model of the mechanism of secGFP-PMEI1 and PGIP2-GFP secretion derived from the different lines of evidence previously reported in reference 7. SecGFPPMEI1 (Fig. 2-1), but not PGIP2-GFP (Fig. 2-2), carries a GPI-anchor, required for its secretion to the cell wall. When the anchorage of GPI is inhibited by mannosamine (Fig. 2-a) or by the fusion of GFP to the C-terminus of PMEI1 (Fig. 2-b), the two non-anchored proteins accumulate in the Golgi stacks. Evidence of retention in Golgi stacks has already been reported for other two cell wall proteins.35 Unlike secGFP-PMEI1, PGIP2-GFP is not stably accumulated in the cell wall and undergoes endocytic trafficking (Fig. 2-3). PGIP2-GFP internalization, likely due to the absence of PGs, might also be related with its ability to interact with homogalacturonan and oligogalacturonides,20 which have been reported to internalize21,22 (Fig. 2-4). Since SYP 121, a Qa-SNARE, is involved in the default secretion of secGFP,23 but not in secretion of PGIP2-GFP and secGFP-PMEI1, trafficking mechanisms underlying secretion into the apoplast are likely different from those underlying the default route (Figs. 2-5). Taken as a whole, evidence suggests the existence of currently undefined signals that control apoplast-targeted secretion.Open in a separate windowFigure 2Schematic illustration for secGFP-PMEI1 and PGIP2-GFP trafficking. See text for details.  相似文献   

7.
Myofibers with an abnormal branching cytoarchitecture are commonly found in various neuromuscular diseases as well as after severe muscle injury. These aberrant myofibers are fragile and muscles containing a high percentage of these myofibers are weaker and more prone to injury. To date the mechanisms and molecules regulating myofiber branching have been obscure. Recent work analyzing the role of mouse odorant receptor 23 (MOR23) in muscle regeneration revealed that MOR23 is necessary for proper skeletal muscle regeneration in mice as loss of MOR23 leads to increased myofiber branching. Further studies demonstrated that MOR23 expression is induced when muscle cells were extensively fusing and plays an important role in controlling cell migration and adhesion. These data demonstrate a novel role for an odorant receptor in tissue repair and identify the first molecule with a functional role in myofiber branching.Key words: muscle regeneration, odorant receptor, olfactory receptor, MOR23, myofiber splitting, myofiber branching, myoblast, fusion, myotube, olfr16Skeletal muscle is characterized by an extensive ability to regenerate after injury due to trauma or disease. Muscle regeneration results from a finely orchestrated series of steps that are spatially and temporally regulated, many of which are not understood. Elucidation of the mechanisms that regulate muscle regeneration may be beneficial for enhancing the rate or extent of muscle regeneration in injury, disease or age.Skeletal muscle is composed of myofibers, which are long cylindrical cells containing hundreds of myonuclei in a common cytoplasm (Fig. 1). Each myofiber is surrounded by a basal lamina sheath; between the myofiber and the basal lamina lie myogenic stem cells called satellite cells. When muscle is injured, myofibers degenerate and satellite cells proliferate to give rise to progeny myoblasts. Myoblasts differentiate and undergo migration, adhesion and fusion to form regenerated myofibers and normal tissue architecture is restored. In many neuromuscular diseases muscle regeneration is aberrant and various abnormalities, such as variation in myofiber size, decreased myofiber number, fibrosis and branched myofibers, are observed. In the clinical literature, branched myofibers are more commonly referred to as “split myofibers.”Open in a separate windowFigure 1Myofiber growth during normal muscle regeneration. (A) Myofibers contain many myonuclei within a common cytoplasm and are surrounded by a basal lamina sheath. Underneath the basal lamina lie satellite cells, myogenic stem cells responsible for muscle regeneration. (B) Myofiber degeneration leads to activation of quiescent satellite cells and their reentry into the cell cycle. Their progeny myoblasts proliferate to yield a pool of progenitor cells. (C) Myoblasts differentiate and undergo migration, adhesion and fusion to form nascent myofibers within the original basal lamina sheath. Additional myoblasts fuse with these newly formed myofibers and the myofiber will continue to grow in size. (D) At later time points regenerated myofibers are similar in size to undamaged myofibers but contain centrally located myonuclei, a hallmark of a regenerated myofiber.Branched myofibers are malformed cells which, instead of having a normal cylindrical shape, contain one or more offshoots of small daughter myotubes contiguous with the parent myofiber (Fig. 2). Branched myofibers can be simple with only one branch (Fig. 2A) or complex with many anastomosing branches resembling a gnarled tree root (Fig. 2B).1 In myofibers with complex cytoarchitecture, individual branches can persist up to hundreds of microns and then eventually recombine with the parent myofiber. Each daughter branch is enclosed in its own basal lamina, which is contiguous with the basal lamina of the parent myofiber.2 The frequency of branched myofibers in rodent muscle under normal conditions is low, on the order of 0.003%.3 However, the frequency in both rodent and human muscle is increased in response to hypertrophy4,5 as well as regeneration due to induced injury,68 muscle transplantation,9,10 or muscular dystrophy.1,1116 In mdx mice, a model of Duchenne Muscular Dystrophy, up to 65–90% of myofibers by 7 months and older display this abnormal branched morphology compared to 6–17% of myofibers at 1–3 months of age.1214 Branched myofibers display functional abnormalities such as alterations in myofiber calcium signaling.13 Additionally, isolated branched myofibers are more prone to rupture at branches during stimulation17 and mdx muscles containing a high percentage of branched myofibers are more vulnerable to contraction-induced injury.12,14 Thus, decreasing the number of branched myofibers would likely be beneficial to muscle function.Open in a separate windowFigure 2Myofiber branching during aberrant muscle regeneration. (A) Phase contrast microscopy of a normal (left) and a branched (right) myofiber. The branched myofiber contains one branch at the end of the myofiber. (B) Schematic diagrams of myofibers with more complex patterns of branching than depicted in (A).Although branched myofibers have been reported in literature for over 100 years, the mechanism by which they arise is unknown and no causative molecules have been identified. The aberrant cytoarchitecture of branched myofibers likely arises from incomplete fusion of small myotubes during muscle regeneration8 though direct proof is lacking. Evidence in favor of these malformed myofibers arising from abortive regenerative processes includes the expression of neonatal myosin, a marker of early muscle regeneration, in the small branches.13 That these branches are newly formed is further suggested by the presence of centrally located nuclei,13,18 a hallmark of regenerated myofibers. The observation that during muscle regeneration multiple small myotubes can form within the old basal lamina sheath8,16,19 leads credence to the idea that aberrant fusion of such small myotubes underlies generation of branched myofibers. Indeed, electron microscopic studies support the ability of myotubes to fuse with one another in vivo;20 myotubes readily fuse with one another in vitro.21 Recent studies of odorant receptor function during muscle regeneration in mice18 have identified the first molecule with a functional role in controlling myofiber branching and suggest that defects in muscle cell migration and/or adhesion may underlie the genesis of these branches.  相似文献   

8.
Organelle movement in plants is dependent on actin filaments with most of the organelles being transported along the actin cables by class XI myosins. Although chloroplast movement is also actin filament-dependent, a potential role of myosin motors in this process is poorly understood. Interestingly, chloroplasts can move in any direction and change the direction within short time periods, suggesting that chloroplasts use the newly formed actin filaments rather than preexisting actin cables. Furthermore, the data on myosin gene knockouts and knockdowns in Arabidopsis and tobacco do not support myosins'' XI role in chloroplast movement. Our recent studies revealed that chloroplast movement and positioning are mediated by the short actin filaments localized at chloroplast periphery (cp-actin filaments) rather than cytoplasmic actin cables. The accumulation of cp-actin filaments depends on kinesin-like proteins, KAC1 and KAC2, as well as on a chloroplast outer membrane protein CHUP1. We propose that plants evolved a myosin XI-independent mechanism of the actin-based chloroplast movement that is distinct from the mechanism used by other organelles.Key words: actin, Arabidopsis, blue light, kinesin, myosin, organelle movement, phototropinOrganelle movement and positioning are pivotal aspects of the intracellular dynamics in most eukaryotes. Although plants are sessile organisms, their organelles are quickly repositioned in response to fluctuating environmental conditions and certain endogenous signals. By and large, plant organelle movements and positioning are dependent on actin filaments, although microtubules play certain accessory roles in organelle dynamics.1,2 Actin inhibitors effectively retard the movements of mitochondria,36 peroxisomes,5,711 Golgi stacks,12,13 endoplasmic reticulum (ER),14,15 and nuclei.1618 These organelles are co-aligned and associated with actin filaments.5,7,8,1012,15,18 Recent progress in this field started to reveal the molecular motility system responsible for the organelle transport in plants.19Chloroplast movement is among the most fascinating models of organelle movement in plants because it is precisely controlled by ambient light conditions.20,21 Weak light induces chloroplast accumulation response so that chloroplasts can capture photosynthetic light efficiently (Fig. 1A). Strong light induces chloroplast avoidance response to escape from photodamage (Fig. 1B).22 The blue light-induced chloroplast movement is mediated by the blue light receptor phototropin (phot). In some cryptogam plants, the red light-induced chloroplast movement is regulated by a chimeric phytochrome/phototropin photoreceptor neochrome.2325 In a model plant Arabidopsis, phot1 and phot2 function redundantly to regulate the accumulation response,26 whereas phot2 alone is essential for the avoidance response.27,28 Several additional factors regulating chloroplast movement were identified by analyses of Arabidopsis mutants deficient in chloroplast photorelocation.2932 In particular, identification of CHUP1 (chloroplast unusual positioning 1) revealed the connection between chloroplasts and actin filaments at the molecular level.29 CHUP1 is a chloroplast outer membrane protein capable of interacting with F-actin, G-actin and profilin in vitro.29,33,34 The chup1 mutant plants are defective in both the chloroplast movement and chloroplast anchorage to the plasma membrane,22,29,33 suggesting that CHUP1 plays an important role in linking chloroplasts to the plasma membrane through the actin filaments. However, how chloroplasts move using the actin filaments and whether chloroplast movement utilizes the actin-based motility system similar to other organelle movements remained to be determined.Open in a separate windowFigure 1Schematic distribution patterns of chloroplasts in a palisade cell under different light conditions, weak (A) and strong (B) lights. Shown as a side view of mid-part of the cell and a top view with three different levels (i.e., top, middle and bottom of the cell). The cell was irradiated from the leaf surface shown as arrows. Weak light induces chloroplast accumulation response (A) and strong light induces the avoidance response (B).Here, we review the recent findings pointing to existence of a novel actin-based mechanisms for chloroplast movement and discuss the differences between the mechanism responsible for movement of chloroplasts and other organelles.  相似文献   

9.
The migration of neuronal growth cones, driving axon extension, is a fascinating process which has been subject of intense investigation over several decades. Many of the key underlying molecules, in particular adhesion proteins at the cell membrane which allow for target recognition and binding, and cytoskeleton filaments and motors which power locomotion have been identified. However, the precise mechanisms by which growth cones coordinate, in time and space, the transmission of forces generated by the cytoskeleton to the turnover of adhesion proteins are still partly unresolved. To get a better grasp at these processes, we put here in relation the turnover rate of ligand/receptor adhesions and the degree of mechanical coupling between cell adhesion receptors and the actin rearward flow. These parameters were obtained recently for N-cadherin and IgCAM based adhesions using ligand-coated microspheres in combination with optical tweezers and photo-bleaching experiments. We show that the speed of growth cone migration requires both a fairly rapid adhesion dynamics and a strong physical connection between adhesive sites and the cytoskeleton.Key words: actin retrograde flow, molecular clutch, myosin, N-cadherin, IgCAMGrowth cones are motile structures at the distal extremity of axons responsible for pathfinding and neurite extension during nervous system development and repair (Fig. 1A). Growth cone advance relies on two coupled processes. First, an internal dynamics of the cytoskeletal network, with actin polymerization occurring at the leading edge, depolymerization in the central region, and myosin activity pulling on lamellipodial actin filaments.1 These integrated mechanisms altogether result in a continuous retrograde flow of actin (Fig. 1B). This flow provides the mechanical tension that drives axonal extension, through a connection to the dynamic array of microtubules that fills the axon and invades the growth cone central domain.2 Second, there is repeated formation and dissociation of transient contacts between growth cones and the extracellular matrix or adjacent cells. These contacts are mediated by trans-membrane cell adhesion molecules (CAMs), e.g. integrins,3 immunoglobulin CAMs (IgCAMs)4,5 and cadherins,6,7 which form specific ligand/receptor bonds with variable lifetimes. A still open question is how these two processes, i.e. actin flow and adhesion dynamic, are coordinated at the growth cone level and contribute to set migration speed. A thorough understanding of these mechanisms is important both from a fundamental perspective and for the design of new compounds to foster axon regeneration after injury.Open in a separate windowFigure 1Growth cone advance and actin flow. (A) Growth cone from a 2 DIV rat hippocampal neuron plated on N-cadherin coated glass. This growth cone moved forward at a speed of about 1 µm/min (B) Raw fluorescence image of transfected actin-GFP. (C) Sequential actin-GFP images were subtracted, giving rise to intensity variations that display the movement of newly assembled actin (black). Note the rapid retrograde movement of actin spots (arrowheads), at a velocity of several µm/min.The coupling between actin-based motility and substrate adhesion has been shown for certain adhesion molecules such as NCAM and N-cadherin to involve a “molecular clutch” (Fig. 2). This mechanism implies a direct transmission of traction forces from the cytoskeleton to the substrate through a strong physical connection between the actin flow and ligand-bound adhesion receptors.8,9 The connection is likely provided by adaptor proteins that can make transient bridges between actin filaments and the cytoplasmic domain of adhesion molecules, i.e. α-and β-catenin in the case of N-cadherin,10 ankyrin and ezrin in the case of IgCAMs such as L1.1114 These purely mechanical connections can also be accompanied by signalling events such as Rac-1 activation by N-cadherin liganding15 and phosphorylation of the L1 intracellular tail that regulates binding to ankyrin.11,12 When only few molecular bonds are formed, e.g. at low ligand density, coupling to the actin flow is not strong enough, resulting in “slippage.” In this process, transient bonds can be formed and broken repeatedly between ligand-occupied adhesion receptors and the actin network. This is how the speed of growth cone translocation usually reaches at most 1 µm/min, whereas the internal actin flow rate proceeds at a rate of several µm/min (Fig. 1A and B). Such slippage is best demonstrated by the use of optical tweezers to impose low forces on ligand-coated microspheres presented to the growth cone dorsal surface (Fig. 3A). Beads tend to move rearward as they couple to the actin flow, and then suddenly snap back into the trap center, when receptor-cytoskeleton bonds break16 (the force of optical tweezers is usually not enough to rupture ligand-receptor bonds, which remain intact at the cell surface). Thus, a step in which a nucleating cluster of adhesion receptors recruits a minimal number of intracellular partners allowing coupling to the actin flow, can be a rate-limiting factor in growth cone progression.Open in a separate windowFigure 2Molecular components involved in growth cone migration. (A) Top view diagram showing filopodia which sense the environment, a flat lamellipodium which is the site of actin dynamics and the thicker central domain and axon which contain dynamic microtubules. The plus signs are sites of actin polymerization and the minus signs indicate actin depolymerization. (B) Side view showing the life cycle of ligand/receptor adhesions.Open in a separate windowFigure 3Optical tweezers and FRAP experiments to measure ligand/receptor and receptor/cytoskeleton dynamics. (A) Optical tweezers experiments performed on ligand-coated beads placed on the growth cone dorsal surface.16 (B) The distance traveled rearward with respect to the trap center is measured. A 2 min trajectory is indicated in red. A pooled parameter called coupling index taking into account the latency for bead escape, as well as the mean velocity and lateral diffusion of the bead, measures the strength of receptor/cytoskeleton interactions. (C) FRAP experiments on membrane GFP-tagged molecules accumulated at ligand-coated microspheres having sedimented on growth cones. The fluorescence intensity is normalized to represent the receptor enrichment level at the bead contact. (D) The recovered intensity is fit by a diffusion/reaction model, which yields a collective equilibrium turnover rate of ligand/receptor bonds. In red is the average of a series of individual curves (grey).In contrast, when strong connection is formed and if the substrate is resistant enough, then the molecular clutch engages and the cell reacts. In growth cones from Aplysia bag cells, forces were imposed on microspheres coated with ApCAM (the homolog of vertebrate NCAM) using a microneedle to locally block the retrograde actin flow. This was systematically followed by a protrusion of the microtubule-rich central domain towards those stiff contacts and forward expansion of the actin-rich lamellipodium.9 These phenomena were later shown to be controlled by a src protein kinase.17 In the case of rat hippocampal neurons, a dramatic accumulation of actin at N-cadherin coated microspheres is observed when the latter are restrained from moving rearward by a microneedle.16 This phenomenon is mediated by a connection between N-cadherin and α-catenin, likely triggering local actin polymerization. By careful analysis of the bead trajectories at varying ligand densities and computation of the latency for bead escape when the optical trap is applied continuously, one can extract a quantitative index of receptor-cytoskeleton coupling (Fig. 3B). Overall, a strong correlation was observed between such coupling index and the velocity of growth cone migration on N-cadherin substrates, both by varying N-cadherin ligand density and by expressing mutated N-cadherin molecules, supporting the clutch concept.16 This mechanism is consistent with in vivo experiments showing that overexpression of the N-cadherin intracellular tail in retinal ganglion cells results in severely impaired axon outgrowth.18 As a negative example of the clutch model, beads coated with fibronectin (our unpublished data) or anti-α1 integrin antibodies3 couple weakly to the actin flow in growth cones while, in parallel, the migration of growth cones on fibronectin- or collagen-coated substrates is rather limited.6,19Molecular mechanisms parallel to the “clutch” can also be involved in growth cone migration. For example, IgCAM adhesions can not only couple to the rearward actin flow but also to static components of the cytoskeleton. Indeed, a 30% fraction of TAG-1 or anti-L1 coated beads can stay immobile on the growth cone surface.12,20 These contrasting behaviors are likely mediated by interactions between the IgCAM intracellular domain and different binding partners (ankyrin vs. ERM),13 and may be responsible for the pauses which alternate with phases of growth cone advance. Also, homophilic adhesions between molecules of cadherin-11 couple very weakly to the actin flow, but promote substantial growth cone migration when cadherin-11 is presented as a substrate. This effect seems to be mediated by an independent interaction with the FGF receptor, which triggers actin dynamic through a signaling cascade.21,22As growth cones migrate, adhesion sites must be recycled at a rate that somehow matches the speed of migration. Adhesion turnover can be schematically decomposed in several sequential phases (Fig. 2B). (1) Initiation of a first single ligand/receptor bond powered by membrane diffusion23 and followed by trapping through a key/lock interaction; (2) Formation of small adhesion clusters through the recruitment of more ligand/receptor pairs, and possibly stabilized by cis-oligomerization (cadherins through the same interface as the trans-dimer, IgCAMs through FnIII domains). These clusters might form very transiently and serve as sites of actin recruitment, as demonstrated for N-cadherin;16 (3) contact maturation and possible reinforcement by connection to the cytoskeleton (as demonstrated for integrins in fibroblasts24); (4) Adhesion rupture, which can proceed through ligand/receptor dissociation triggered by cytoskeleton tension. Indeed, the intrinsic lifetime of ligand/receptor bonds such as cadherins, is sensitive to the mechanical force applied on them.25 Furthermore, the loosening of receptor/cytoskeketon connections can cause inside-out rupture of ligand/receptor bonds. This was demonstrated for fibronectin/integrin interactions by the fact that when fibronectin coated-beads reach the base of a fibroblast lamellipodium, they spontaneously detach from the cell surface.26 In the case of very sticky ligand/receptor interactions such as SynCAM homophilic adhesions,27 this process can actually be a limiting step that slows down growth cone advance. Indeed, SynCAM couples very well to the actin flow, but is unable to support growth cone migration.16 Finally, adhesion rupture might also proceed through membrane rupture, the adhesion receptors being extracted from the cell membrane and left behind on the substrate (demonstrated for integrins at the tail of fibroblasts28).These basic processes can be accompanied by more complex and active phenomena, e.g. involving forward surface transport as shown for NCAM29 or internal trafficking in the case of L1.30 By interacting with the clathrin adaptor AP-2 through a specific RSLE motif in its intracellular tail, L1 can undergo endocytosis in the central domain and exocytosis at the periphery of the growth cone.30 This mechanism generates a density gradient of L1 molecules which accelerates the formation of bonds with a variety of ligands, including L1 itself. The use of an L1-GFP construct in which the N-terminal GFP could be rapidly cleaved off by thrombin, together with L1-Fc microspheres manipulated by optical tweezers showed that local exocytosis of L1-rich vesicles at the growth cone periphery indeed participates in enhancing the formation of L1 homophilic contacts.31 We did not observe such internal traffic for N-cadherin within the growth cone, partly because of a difficulty to introduce a fluorescent protein tag in the ectodomain, which otherwise perturbs the adhesive function. However, the use of an N-cadherin molecule with triple mutation in the juxta-membrane domain that abolishes binding to p120 catenin, involved in the export of N-cadherin to the cell surface, suggested that recycling events might also play a role.32By measuring the fluorescence recovery after photobleaching (FRAP) of GFP-tagged receptors transiently trapped at ligand-coated microspheres and analyzing the curves using a diffusion/reaction model, we were able to compute the equilibrium turnover rates of ligand/receptor pairs in controlled adhesive contacts involving many simultaneous bonds (Fig. 3C and D). We found that mature L1 homophilic adhesions recycle fast compared to other IgCAMs such as TAG-1/NrCAM adhesions,20 likely owing to the specific internalization motif present in L1. Indeed, the recycling rate was reduced by a factor of 3 after truncation of the L1 intracellular tail, which prevented endocytosis.31 N-cadherin homophilic adhesions have an intermediate turnover rate, which is sensitive to the binding to catenin partners.32 Using these measurements as well as data from the literature, we plotted the impact of both receptor-cytoskeleton coupling and adhesion turnover rate on neurite outgrowth (Fig. 4A and B), which is strongly proportional to growth cone velocity.16 The graphs show that a strong coupling between ligand-occupied receptors and the actin flow is necessary, but not sufficient for neurite extension (Fig. 4C). Another requirement is that the turnover of ligand/receptor adhesions lies in an optimal range: not too high, otherwise bonds detach before coupling can occur, and not too slow either, since sticky bonds which do not rupture paralyze growth cone progression (Fig. 4D). A similar bell-shape curve between the strength of cell-substrate adhesion and cell migration speed was demonstrated for fibroblasts33 and keratocytes,34 indicating that these coupled mechanisms are fundamental to cell migration. To fully understand the quantitative relationship between adhesion turnover and the clutch process, it would be helpful to add data to this preliminary graph. For example, the extracellular matrix molecule laminin is known to support axon growth very efficiently but, to our knowledge, neither the coupling to the actin flow in growth cones or the adhesive turnover rate of integrins has been evaluated yet. Conversely, NCAM was shown to couple well to the actin flow35 and induce neurite outgrowth,4,36 but measurements of the lifetime of NCAM homophilic adhesions within growth cones are still lacking.Open in a separate windowFigure 4Impact of ligand/receptor turnover rate and receptor/cytoskeletal coupling on neurite outgrowth. (A and B) Example of 2 DIV rat hippocampal neurons plated on N-cadherin-Fc coated substrate and transfected at 1 DIV with N-cadherin-GFP. (A) DIC image. (B) Fluorescence image. The longest neurite, most likely the axon, is outlined by arrowheads. (C and D) In both graphs, the y-axis represents the longest neurite length after two days plating on ligand-coated glass. (Red) Rat hippocampal neurons transfected with either wild type or mutated N-cadherin molecules, interacting with purified N-cadherin ligands.16,32 The scale in red intensity represents from dark to light: wild type N-cadherin, N-cadherin deleted of the whole ectodomain, N-cadherin truncated in the C-terminal region binding to β-catenin, N-cadherin with triple mutation in the juxta-membrane domain interacting with p120, and wild type N-cadherin in the presence of cytochalasin D. (Blue) Neurons transfected with either wild type L1 (dark) or L1 truncated in the intracellular tail (light), interacting with purified L1.31 Neurite growth on L1 was estimated from references.7,14,30 (Grey) Interaction between endogenous SynCAM1 molecules expressed on growth cones and SynCAM-Fc ligands coated on microspheres or flat glass.16 The turnover of SynCAM homophilic interactions was estimated from SynCAM-coated Quantum dots detaching from neurons transfected with SynCAM1.42 (Green) Neuroblastoma cells expressing NrCAM-GFP in contact with TAG-1 coated microspheres.20 DRG neurite growth on TAG-1 was taken from reference.43 (Orange) The coupling index was taken from optical tweezers experiments using anti-β1 integrin coated beads interacting with DRG neurons,3 the turnover rate was inferred from FRAP experiments on fibroblast focal contacts44 and neurite growth on fibronectin was estimated from reference.19 We omitted statistics for clarity. The SEM are usually in the order of 5–15% of the mean, for sample sizes of typically 20–30 beads (coupling index and turnover rate) and 40–100 transfected cells (neurite length).One important question is how these observations obtained from simplified in vitro systems using stiff substrates of well-defined geometry coated with specific purified proteins at controlled density, translate to the in vivo situation. There, the 3D substrate is comprised of extracellular matrix and multiple cell types, co-expressing many different CAMs that can bind simultaneously in various stoechiometries and also generating local gradients of chemo-attractant and chemo-repulsive signals.37 Substrate flexibility is probably an important factor, since axons grow more slowly when neurons are plated on a layer of fibroblasts expressing CAMs21 than when molecules are immobilized on a substrate.16,30 This preference for cells to move on stiff substrates, called durotaxis, has been well described for fibroblasts.38 Another specific feature of the in vivo situation is the existence of decision points, often correlated with the presence of guidepost cells where growth cones make a pause and often change shape and reorient before turning to another direction.39 This type of behavior has been successfully mimicked in vitro using artificial guideposts made of fibronectin or laminin coated microspheres.40 Whereas growth cones display a fairly continuous displacement on a homogeneous substrate,16 the presence of these guideposts make growth cones either slow down, pause or even collapse, or conversely accelerate, depending on the CAM grafted on the bead.40,41 Finally, the shape itself of the growth cone can be an indicator of its motile state:39 this is also true in vitro where small growth cones are often the most rapid whereas large and flat growth cones stay rather immobile. Thus, although the in vivo situation seems at first sight awfully complex, some general trends can be explained given a small number of interacting molecular species and rather simple bio-chemical and mechanical models.In conclusion, the dynamic regulation of growth cone advance can take place at several levels: (1) the actin-associated proteins controlling actin dynamics (nucleation, polymerization, sequestering, branching); (2) the activity of motors pulling on the actin network, generating the retrograde flow; (3) the intracellular adaptor proteins that link actin to the CAMs; (4) the membrane delivery and retrieval of CAMs; (5) the ligand/receptor interaction properties themselves; and (6) the processes regulating microtubule assembly and microtubule/actin interactions at the base of the growth cone. The orchestration in time and space of all these processes generates the movement and reactivity of growth cones necessary to lead axons to their target cells.  相似文献   

10.
Determination of the subcellular localization of an unknown protein is a major step towards the elucidation of its function. Lately, the expression of proteins fused to fluorescent markers has been very popular and many approaches have been proposed to express these proteins. Stable transformation using Agrobacterium tumefaciens generates stable lines for downstream experiments, but is time-consuming. If only colocalization is required, transient techniques save time and effort. Several methods for transient assays have been described including protoplast transfection, biolistic bombardment, Agrobacterium tumefaciens cocultivation and infiltration. In general colocalizations are preferentially performed in intact tissues of the same species, resembling the native situation. High transformation rates were described for cotyledons of Arabidopsis, but never for roots. Here we report that it is possible to transform Arabidopsis root epidermal cells with an efficiency that is sufficient for colocalization purposes.Key words: Arabidopsis, GFP-fusions, protein localization, root, transient transformationSince the release of the Arabidopsis thaliana genome sequence plant biologists set the goal to elucidate the functions of all coded genes. Apart from the spatio-temporal expression patterns of genes, the subcellular localization of gene products can play an essential role in deciphering their function. Classical immunological approaches to localize proteins can be hindered by cross-reactivity, time-consuming generation of antibodies and the low temporal resolution. Expression of tagged proteins forms a suitable alternative. Lately, fusions with fluorescent proteins in combination with confocal (CLSM)1 or spinning disc microscopy2 allow real time protein localization and even subcellular trafficking at high resolution. An overview of fluorescent tagging approaches can be found elsewhere.3Currently several techniques to introduce the coding region for a tagged protein in a plant are available. The generation of stable lines transformed by Agrobacterium tumefaciens offers a continuous source of plant material, but it is time-consuming especially when only colocalization experiments are required. Transient assays, on the other hand, offer the advantage of being fast and amenable to high throughput strategies. Each of these techniques, however, has some limitations and drawbacks. Particle bombardment (biolistics) 46 for example circumvents the host specificity of Agrobacterium strains, but requires expensive equipment. Moreover, it is rather disruptive and imposes a significant stress upon the plants, possibly influencing the results. Protoplasts lack a cell wall and protoplast transformation7,8 is therefore not suitable for certain experiments related to cell wall proteins or when interactions between cells on tissue level might be important.9 Moreover, protoplasts have lost their identity which might be critical for the correct functioning of certain transgenic constructs. Agrobacterium infiltration of tobacco leaves10 is regularly used and represents an efficient, fast and relatively easy transformation technique. However, tobacco leaves easily show autofluoresence due to tissue damage as a result of experimental manipulations. As it has been reported that some protein fusions expressed in an heterologous system localize to different subcellular localizations11 it is advisable not to use tobacco when localizing Arabidopsis proteins. Leaf infiltrations have been performed in Arabidopsis,12 but apparently their leaves are much more prone to mechanical damage and the leaf developmental stage is critical, complicating this technique. Cocultivation of Agrobacterium with seedlings offers a rapid and efficient approach applicable to many mono and dicot species. It was reported to work efficiently in Arabidopsis cotyledons, but not in roots.9 As an alternative method, Agrobacterium infiltration of Arabidopsis seedlings11 seems an efficient technique for transient expression. However, expression in root cells could not be obtained. Colocalizations are required in the native cells or tissue for the correct localization of an unknown protein or proteins that need interaction partners. As a consequence this technique can not be reliably used when root expressed gene products are studied. Here we show evidence that it is possible to use the described technique11 to induce transient expression in Arabidopsis roots.We used the Agrobacterium infiltration of Arabidopsis seedlings technique11 to colocalize several C-terminal (S65T)-sGFP fusions generated in the plant binary vector pGWB6.13 Each construct was transformed into Agrobacterium tumefaciens (C59C1RifR) containing the helper plasmid pMP90. Subsequently different stable marker lines, wild type Arabidopsis (Col-0) bearing mCherry fusion constructs,14 were transiently transformed.11 After 2 or 3 days seedlings were studied using CLSM. Besides being expressed in cotyledons fusion proteins were clearly observed in root epidermis and root cap cells (Fig. 1A and B). As reported11 the transformation efficiency in cotyledons was considerably higher than in root cells. However, in each experiment we obtained a considerable amount of transformed root epidermal cells which was more than sufficient for colocalization studies (Fig. 2). It was remarkable that transformation was repeatedly successful in groups of cells, adjacent or close to each other.Open in a separate windowFigure 1Transient transformation of Arabidopsis root cells. Expression of the protein-GFP fusion product can be seen in the epidermal (A) and root cap cells (B) on fluorescence/transmission merged images. As seen in (A) high efficiencies of root transformation can be reached.Open in a separate windowFigure 2Colocalization of mCherry and GFP constructs. Confocal image of the mCherry fluorescence (A), the GFP signal (B) and the merged image (C).In contrast to what was reported earlier we show here that the Agrobacterium infiltration technique11 is perfectly capable of transiently transforming Arabidopsis root epidermal cells. It allows the transient production and study of proteins in their native environment, considerably increasing the reliability of such experiments. Additionaly the use of RFP marker constructs in colocalisation studies in the root is free of interference by the red background autofluorescence of chlorophyll.  相似文献   

11.
12.
The prion hypothesis13 states that the prion and non-prion form of a protein differ only in their 3D conformation and that different strains of a prion differ by their 3D structure.4,5 Recent technical developments have enabled solid-state NMR to address the atomic-resolution structures of full-length prions, and a first comparative study of two of them, HET-s and Ure2p, in fibrillar form, has recently appeared as a pair of companion papers.6,7 Interestingly, the two structures are rather different: HET-s features an exceedingly well-ordered prion domain and a partially disordered globular domain. Ure2p in contrast features a very well ordered globular domain with a conserved fold, and—most probably—a partially ordered prion domain.6 For HET-s, the structure of the prion domain is characterized at atomic-resolution. For Ure2p, structure determination is under way, but the highly resolved spectra clearly show that information at atomic resolution should be achievable.Key words: prion, NMR, solid-state NMR, MAS, structure, Ure2p, HET-sDespite the large interest in the basic mechanisms of fibril formation and prion propagation, little is known about the molecular structure of prions at atomic resolution and the mechanism of propagation. Prions with related properties to the ones responsible for mammalian diseases were also discovered in yeast and funghi8,9 which provide convenient model system for their studies. Prion proteins described include the mammalian prion protein PrP, Ure2p,10 Rnq1p,11 Sup35,12 Swi1,13 and Cyc8,14 from bakers yeast (S. cervisiae) and HET-s from the filamentous fungus P. anserina. The soluble non-prion form of the proteins characterized in vitro is a globular protein with an unfolded, dynamically disordered N- or C-terminal tail.1518 In the prion form, the proteins form fibrillar aggregates, in which the tail adopts a different conformation and is thought to be the dominant structural element for fibril formation.Fibrills are difficult to structurally characterize at atomic resolution, as X-ray diffraction and liquid-state NMR cannot be applied because of the non-crystallinity and the mass of the fibrils. Solid-state NMR, in contrast, is nowadays well suited for this purpose. The size of the monomer, between 230 and 685 amino-acid residues for the prions of Figure 1, and therefore the number of resonances in the spectrum—that used to be large for structure determination—is now becoming tractable by this method.Open in a separate windowFigure 1Prions identified today and characterized as consisting of a prion domain (blue) and a globular domain (red).Prion proteins characterized so far were found to be usually constituted of two domains, namely the prion domain and the globular domain (see Fig. 1). This architecture suggests a divide-and-conquer approach to structure determination, in which the globular and prion domain are investigated separately. In isolation, the latter, or fragments thereof, were found to form β-sheet rich structures (e.g., Ure2p(1-89),6,19 Rnq1p(153-405)20 and HET-s(218-289)21). The same conclusion was reached by investigating Sup35(1-254).22 All these fragements have been characterized as amyloids, which we define in the sense that a significant part of the protein is involved in a cross-beta motif.23 An atomic resolution structure however is available presently only for the HET-s prion domain, and was obtained from solid-state NMR24 (vide infra). It contains mainly β-sheets, which form a triangular hydrophobic core. While this cross-beta structure can be classified as an amyloid, its triangular shape does deviate significantly from amyloid-like structures of smaller peptides.23Regarding the globular domains, structures have been determined by x-ray crystallography (Ure2p25,26 and HET-s27), as well as NMR (mammal prions15,2830). All reveal a protein fold rich in α-helices, and dimeric structures for the Ure2 and HET-s proteins. The Ure2p fold resembles that of the β-class glutathione S-transferases (GST), but lacks GST activity.25It is a central question for the structural biology of prions if the divide-and-conquer approach imposed by limitations in current structural approaches is valid. Or in other words: can the assembly of full-length prions simply be derived from the sum of the two folds observed for the isolated domains?  相似文献   

13.
Polar auxin transport (PAT), which is controlled precisely by both auxin efflux and influx facilitators and mediated by the cell trafficking system, modulates organogenesis, development and root gravitropism. ADP-ribosylation factor (ARF)-GTPase protein is catalyzed to switch to the GTP-bound type by a guanine nucleotide exchange factor (GEF) and promoted for hybridization to the GDP-bound type by a GTPase-activating protein (GAP). Previous studies showed that auxin efflux facilitators such as PIN1 are regulated by GNOM, an ARF-GEF, in Arabidopsis. In the November issue of The Plant Journal, we reported that the auxin influx facilitator AUX1 was regulated by ARF-GAP via the vesicle trafficking system.1 In this addendum, we report that overexpression of OsAGAP leads to enhanced root gravitropism and propose a new model of PAT regulation: a loop mechanism between ARF-GAP and GEF mediated by vesicle trafficking to regulate PAT at influx and efflux facilitators, thus controlling root development in plants.Key Words: ADP-ribosylation factor (ARF), ARF-GAP, ARF-GEF, auxin, GNOM, polar transport of auxinPolar auxin transport (PAT) is a unique process in plants. It results in alteration of auxin level, which controls organogenesis and development and a series of physiological processes, such as vascular differentiation, apical dominance, and tropic growth.2 Genetic and physiological studies identified that PAT depends on efflux facilitators such as PIN family proteins and influx facilitators such as AUX1 in Arabidopsis.Eight PIN family proteins, AtPIN1 to AtPIN8, exist in Arabidopsis. AtPIN1 is located at the basal side of the plasma membrane in vascular tissues but is weak in cortical tissues, which supports the hypothesis of chemical pervasion.3 AtPIN2 is localized at the apical side of epidermal cells and basally in cortical cells.1,4 GNOM, an ARF GEF, modulates the localization of PIN1 and vesicle trafficking and affects root development.5,6 The PIN auxin-efflux facilitator network controls root growth and patterning in Arabidopsis.4 As well, asymmetric localization of AUX1 occurs in the root cells of Arabidopsis plants,7 and overexpression of OsAGAP interferes with localization of AUX1.1 Our data support that ARF-GAP mediates auxin influx and auxin-dependent root growth and patterning, which involves vesicle trafficking.1 Here we show that OsAGAP overexpression leads to enhanced gravitropic response in transgenic rice plants. We propose a model whereby ARF GTPase is a molecular switch to control PAT and root growth and development.Overexpression of OsAGAP led to reduced growth in primary or adventitious roots of rice as compared with wild-type rice.1 Gravitropism assay revealed transgenic rice overxpressing OsAGAP with a faster response to gravity than the wild type during 24-h treatment. However, 1-naphthyl acetic acid (NAA) treatment promoted the gravitropic response of the wild type, with no difference in response between the OsAGAP transgenic plants and the wild type plants (Fig. 1). The phenotype of enhanced gravitropic response in the transgenic plants was similar to that in the mutants atmdr1-100 and atmdr1-100/atpgp1-100 related to Arabidopsis ABC (ATP-binding cassette) transporter and defective in PAT.8 The physiological data, as well as data on localization of auxin transport facilitators, support ARF-GAP modulating PAT via regulating the location of the auxin influx facilitator AUX1.1 So the alteration in gravitropic response in the OsAGAP transgenic plants was explained by a defect in PAT.Open in a separate windowFigure 1Gravitropism of OsAGAP overexpressing transgenic rice roots and response to 1-naphthyl acetic acid (NAA). (A) Gravitropism phenotype of wild type (WT) and OsAGAP overexpressing roots at 6 hr gravi-stimulation (top panel) and 0 hr as a treatment control (bottom panel). (B) Time course of gravitropic response in transgenic roots. (C and D) results correspond to those in (A and B), except for treatment with NAA (5 × 10−7 M).The polarity of auxin transport is controlled by the asymmetric distribution of auxin transport proteins, efflux facilitators and influx carriers. ARF GTPase is a key member in vesicle trafficking system and modulates cell polarity and PAT in plants. Thus, ARF-GDP or GTP bound with GEF or GAP determines the ARF function on auxin efflux facilitators (such as PIN1) or influx ones (such as AUX1).ARF1, targeting ROP2 and PIN2, affects epidermal cell polarity.9 GNOM is involved in the regulation of PIN1 asymmetric localization in cells and its related function in organogenesis and development.6 Although VAN3, an ARF-GAP in Arabidopsis, is located in a subpopulation of the trans-Golgi transport network (TGN), which is involved in leaf vascular network formation, it does not affect PAT.10 OsAGAP possesses an ARF GTPase-activating function in rice.11 Specifically, our evidence supports that ARF-GAP bound with ARF-GTP modulates PAT and gravitropism via AUX1, mediated by vesicle trafficking, including the Golgi stack.1Therefore, we propose a loop mechanism between ARF-GAP and GEF mediated by the vascular trafficking system in regulating PAT at influx and efflux facilitators, which controls root development and gravitropism in plants (Fig. 2). Here we emphasize that ARF-GEF catalyzes a conversion of ARF-bound GDP to GTP, which is necessary for the efficient delivery of the vesicle to the target membrane.12 An opposite process of ARF-bound GDP to GTP is promoted by ARF-GTPase-activating protein via binding. A loop status of ARF-GTP and ARF-GDP bound with their appurtenances controls different auxin facilitators and regulates root development and gravitropism.Open in a separate windowFigure 2Model for ARF GTPase as a molecular switch for the polar auxin transport mediated by the vesicle traffic system.  相似文献   

14.
Understanding plant response to wind is complicated as this factor entails not only mechanical stress, but also affects leaf microclimate. In a recent study, we found that plant responses to mechanical stress (MS) may be different and even in the opposite direction to those of wind. MS-treated Plantago major plants produced thinner more elongated leaves while those in wind did the opposite. The latter can be associated with the drying effect of wind as is further supported by data on petiole anatomy presented here. These results indicate that plant responses to wind will depend on the extent of water stress. It should also be recognized that the responses to wind may differ between different parts of a plant and between plant species. Physiological research on wind responses should thus focus on the signal sensing and transduction of both the mechanical and drought signals associated with wind, and consider both plant size and architecture.Key words: biomechanics, leaf anatomy, phenotypic plasticity, plant architecture, signal transduction thigmomorphogenesis, windWind is one of the most ubiquitous environmental stresses, and can strongly affect development, growth and reproductive yield in terrestrial plants.13 In spite of more than two centuries of research,4 plant responses to wind and their underlying mechanisms remain poorly understood. This is because plant responses to mechanical movement themselves are complicated and also because wind entails not only mechanical effects, but also changes in leaf gas and heat exchange.57 Much research on wind has focused primarily on its mechanical effect. Notably, several studies that determine plant responses to mechanical treatments such as flexing, implicitly extrapolate their results to wind effects.810 Our recent study11 showed that this may lead to errors as responses to wind and mechanical stimuli (in our case brushing) can be different and even in the opposite direction. In this paper, we first separately discuss plant responses to mechanical stimuli, and other wind-associated effects, and then discuss future challenges for the understanding of plant responses to wind.It is often believed that responses to mechanical stress (thigmomorphogenesis) entail the production of thicker and stronger plant structures that resist larger forces. This may be true for continuous unidirectional forces such as gravity, however for variable external forces (such as wind loading or periodic flooding) avoiding such mechanical stress by flexible and easily reconfigurable structures can be an alternative strategy.1214 How plants adapt or acclimate to such variable external forces depends on the intensity and frequency of stress and also on plant structures. Reduced height growth is the most common response to mechanical stimuli.15,16 This is partly because such short stature increases the ability of plants to both resist forces (e.g., real-locating biomass for radial growth rather than elongation growth), and because small plants experience smaller drag forces (Fig. 1). Some plant species show a resistance strategy in response to mechanical stress by increasing stem thickness1,10 and tissue strength.7 But other species show an avoidance strategy by a reduction in stem or petiole thickness and flexural rigidity in response to MS.11,1518 These different strategies might be associated with plant size and structure. Stems of larger plants such as trees and tall herbs are restricted in the ability to bend as they carry heavy loads7,10,19 (Fig. 1). Conversely short plants are less restricted in this respect and may also be prone to trampling for which stress-avoidance would be the only viable strategy.18,20 Systematic understanding of these various responses to mechanical stress remains to be achieved.Open in a separate windowFigure 1A graphical representation of how wind effects can be considered to entail both a drying and a mechanical effect. Adaptation or acclimation to the latter can be through a force resistance strategy or a force avoidance strategy, the benefit of which may depend on the size and architecture of plants as well as the location of a given structure within a plant.Wind often enhances water stress by reducing leaf boundary layers and reduces plant temperature by transpiration cooling. The latter effect may be minor,11 but the former could significantly affect plant development. Anten et al. (2010) compared phenotypic traits and growth of Plantago major that was grown under mechanical stimuli by brushing (MS) and wind in the factorial design. Both MS and wind treatments reduced growth and influenced allocation in a similar manner. MS plants, however, had more slender petioles and narrower leaf blades while wind exposed plants exhibited the opposite response having shorter and relatively thicker petioles and more round-shaped leaf blades. MS plants appeared to exhibit stress avoidance strategy while such responses could be compensated or overridden by water stress in wind exposure.11 A further analysis of leaf petiole anatomy (Fig. 2) supports this view. The vascular fraction in the petiole cross-section was increased by wind but not by MS, suggesting that higher water transport was required under wind. Our results suggest that drying effect of wind can at least to some extent override its mechanical effect.Open in a separate windowFigure 2Representative images of petiole cross-sections of Plantago major grown in 45 days in continuous wind and/or mechanical stimuli (A–D). Petiole cross-section area (E) and vascular bundle fraction in the cross-section of petiole (F). mean + SD (n = 12) are shown. Significance levels of ANOVA; ***p < 0.001, **p < 0.01, *p < 0.05, ns p > 0.05.Physiological knowledge on plant mechanoreception and signal transduction has been greatly increased during the last decades. Plants sense mechanical stimuli through membrane strain with stretch activated channels21 and/or through some linker molecules connecting the cell wall, plasma membrane and cytoskeleton.4,22,23 This leads to a ubiquitous increase in intracellular Ca2+ concentration. The increased Ca2+ concentration is sensed by touch induced genes (TCHs),24,25 which activates downstream transduction machineries including a range of signaling molecules and phytohormones, consequently altering physiological and developmental processes.26 Extending this knowledge to understand plant phenotypic responses to wind however remains a challenge. As responses to wind have been found to differ among parts of a plant (e.g., terminal vs. basal stem) and also across species, physiological studies should be extended to the whole-plant as integrated system rather than focusing on specific tissue level. Furthermore to understand the general mechanism across species, it is required to study different species from different environmental conditions. Advances in bioinformatics, molecular and physiological research will facilitate cross-disciplinary studies to disentangle the complicated responses of plants to wind.  相似文献   

15.
16.
Co-localization of mitochondria with chloroplasts in plant cells has long been noticed as beneficial interactions of the organelles to active photosynthesis. Recently, we have found that mitochondria in mesophyll cells of Arabidopsis thaliana expressing mitochondrion-targeted green fluorescent protein (GFP) change their distribution in a light-dependent manner. Mitochondria occupy the periclinal and anticlinal regions of palisade cells under weak and strong blue light, respectively. Redistributed mitochondria seem to be rendered static through co-localization with chloroplasts. Here we further demonstrated that distribution patterns of mitochondria, together with chloroplasts, returned back to those of dark-adapted state during dark incubation after blue-light illumination. Reversible association of the two organelles may underlie flexible adaptation of plants to environmental fluctuations.Key words: Arabidopsis thaliana, blue light, chloroplast, green fluorescent protein, mesophyll cell, mitochondrion, organelle positioningHighly dynamic cell organelles, mitochondria, are responsible not only for energy production, but also for cellular metabolism, cell growth and survival as well as gene regulations.1,2 Appropriate intracellular positioning and distribution of mitochondria contribute to proper organelle functions and are essential for cell signaling.3,4 In plant cells operating photosynthesis, the co-localization of mitochondria with chloroplasts has been a well known phenomenon for a long period of time.5,6,7 Physical contact of mitochondria with chloroplasts may provide a means to transfer genetic information from the organelle genome,8 as well as to exchange metabolite components; a process required for the maintenance of efficient photosynthesis.9,10,11Using Arabidopsis thaliana stably expressing mitochondrion-targeted GFP,12 we have recently examined a different aspect of mitochondria positioning. Although mitochondria in leaf mesophyll cells are highly motile under dark condition, mitochondria change their intracellular positions in response to light illumination.13 The pattern of light-dependent positioning of mitochondria seems to be essentially identical to that of chloroplasts.14 Mitochondria occupy the periclinal regions under weak blue light (wBL; 470 nm, 4 µmol m−2s−1) and the anticlinal regions under strong blue light (sBL; 100 µmol m−2s−1), respectively. A gradual increase in the number of static mitochondria located in the vicinity of chloroplasts in the periclinal regions with time period of wBL illumination clearly demonstrates that the co-localization of these two organelles is a light-induced phenomenon.13In the present study, to ask whether the light-dependent positioning of mitochondria is reversible or not, a time course of mitochondria redistribution was examined transferring the sample leaves from light to dark conditions. The representative results (Fig. 1) clearly show that mitochondria re-changed their positions within several hours of dark treatment. Immediately after dark adaptation, mitochondria in the palisade mesophyll cells were distributed randomly throughout the cytoplasm (Fig. 1A and ref. 13). Chloroplasts were distributed along the inner periclinal walls and the lower half of the anticlinal walls. On the contrary, mitochondria accumulated along the outer (Fig. 1B) and inner periclinal walls when illuminated with wBL. Chloroplast position was also along the outer and inner periclinal walls. Many of the mitochondria located near the chloroplasts lost their motility. When wBL-illuminated leaves were transferred back to dark condition, the numbers of mitochondria and chloroplasts present on the periclinal regions began to decrease within several hours (Fig. 1C). After 10 h dark treatment, distribution patterns of mitochondria as well as chloroplasts almost recovered to those of dark-adapted cells (Fig. 1D).Open in a separate windowFigure 1Distribution of mitochondria and chloroplasts on the outer periclinal regions of palisade mesophyll cells of A. thaliana under different light conditions. Mitochondria (green; GFP) and chloroplasts (red; chlorophyll autofluorescence) were visualized with confocal microscopy after dark adaptation (A), immediately after wBL (470 nm, 4 µmol m−2s−1) illumination for 4 h (B), after dark treatment for 6 h (C) and 10 h (D) following the 4-h wBL illumination, respectively. Bar = 50 µm.To our knowledge, this may be the first report that directly demonstrates that wBL regulates mitochondria and chloroplast positioning in a reversible manner, though the nuclei in A. thaliana leaf cells were also found to reverse their positions when transferred from sBL to dark conditions.15 Reversible regulation of organelle positioning in leaf cells should play critical roles in adaptation of plants to highly fluctuating light conditions in the nature. Since distribution patterns of mitochondria under wBL and sBL are identical to those of chloroplasts, we can assume that phototropins, the BL receptors for chloroplast photo-relocation movement,16 may have some role in the redistribution of mitochondria. On the other hand, we also found that red light exhibited a significant effect on mitochondria positioning (Islam et al. 2009), suggesting an involvement of photosynthesis. These possibilities are now under investigation.  相似文献   

17.
Intracellular components in methyl jasmonate (MeJA) signaling remain largely unknown, to compare those in well-understood abscisic acid (ABA) signaling. We have reported that nitric oxide (NO) is a signaling component in MeJA-induced stomatal closure, as well as ABA-induced stomatal closure in the previous study. To gain further information about the role of NO in the guard cell signaling, NO production was examined in an ABA- and MeJA-insensitive Arabidopsis mutant, rcn1. Neither MeJA nor ABA induced NO production in rcn1 guard cells. Our data suggest that NO functions downstream of the branch point of MeJA and ABA signaling in Arabidopsis guard cells.Key words: abscisic acid, Arabidopsis thaliana, guard cells, methyl jasmonate, nitric oxideStomatal pores that are formed by pairs of guard cells respond to various environmental stimuli including plant hormones. Some signal components commonly function in MeJA- and ABA-induced stomatal closing signals,1 such as cytosolic alkalization, ROS generation and cytosolic free calcium ion elevation. Recently, we demonstrated that NO functions in MeJA signaling, as well as ABA signaling in guard cells.2NO production by nitric oxide synthase (NOS) and nitrate reductase (NR) plays important roles in physiological processes in plants.3,4 It has been shown that NO functions downstream of ROS production in ABA signaling in guard cells.5 NO mediates elevation of cytosolic free Ca2+ concentration ([Ca2+]cyt), inactivation of inward-rectifying K+ channels and activation of S-type anion channels,6 which are known to be key factors in MeJA- and ABA-induced stomatal closure.2,79It has been reported that ROS was not induced by MeJA and ABA in the MeJA- and ABA-insensitive mutant, rcn1 in which the regulatory subunit A of protein phosphatase 2A, RCN1, is impaired.7,10 We examined NO production induced by MeJA and ABA in rcn1 guard cells (Fig. 1). NO production by MeJA and ABA was impaired in rcn1 mutant (p = 0.87 and 0.25 for MeJA and ABA, respectively) in contrast to wild type. On the other hand, the NO donor, SNP induced stomatal closure both in wild type and rcn1 mutant (data not shown). These results are consistent with our previous results, i.e., NO is involved in both MeJA- and ABA-induced stomatal closure and functions downstream of the branching point of MeJA and ABA signaling in Arabidopsis guard cells.7 Our finding implies that protein phosphatase 2A might positively regulate NO levels in guard cells (Fig. 2).Open in a separate windowFigure 1Impairment of MeJA- and ABA-induced NO production in rcn1 guard cells. (A) Effects of MeJA (n = 10) and ABA (n = 9) on NO production in wild-type guard cells. (B) Effects of MeJA (n = 7) and ABA (n = 7) on NO production in rcn1 guard cells. The vertical scale represents the percentage of diaminofluorescein-2 diacetate (DAF-2 DA) fluorescent levels when fluorescent intensities of MeJA- or ABA-treated cells are normalized to control value taken as 100% for each experiment. Each datum was obtained from at least 30 guard cells. Error bars represent standard errors. Significance of differences between data sets was assessed by Student''s t-test analysis in this paper. We regarded differences at the level of p < 0.05 as significant.Open in a separate windowFigure 2A model of signal interaction in MeJA-induced and ABA-induced stomatal closure. Neither MeJA nor ABA induces ROS production, NO production, IKin and stomatal closure in rcn1 mutant. These results suggest that NO functions downstream of the branch point of MeJA signaling and ABA signaling in Arabidopsis guard cells.  相似文献   

18.
During development, dorsal root ganglion (DRG) neurons extend their axons toward the dorsolateral part of the spinal cord and enter the spinal cord through the dorsal root entry zone (DREZ). After entering the spinal cord, these axons project into the dorsal mantle layer after a “waiting period” of a few days. We revealed that the diffusible axonal guidance molecule netrin-1 is a chemorepellent for developing DRG axons. When DRG axons orient themselves toward the DREZ, netrin-1 proteins derived from the ventral spinal cord prevent DRG axons from projecting aberrantly toward the ventral spinal cord and help them to project correctly toward the DREZ. In addition to the ventrally derived netrin-1, the dorsal spinal cord cells adjacent to the DREZ transiently express netrin-1 proteins during the waiting period. This dorsally derived netrin-1 contributes to the correct guidance of DRG axons to prevent them from invading the dorsal spinal cord. In general, there is a complete lack of sensory axonal regeneration after a spinal cord injury, because the dorsal column lesion exerts inhibitory activities toward regenerating axons. Netrin-1 is a novel candidate for a major inhibitor of sensory axonal regeneration in the spinal cord; because its expression level stays unchanged in the lesion site following injury, and adult DRG neurons respond to netrin-1-induced axon repulsion. Although further studies are required to show the involvement of netrin-1 in preventing the regeneration of sensory axons in CNS injury, the manipulation of netrin-1-induced repulsion in the CNS lesion site may be a potent approach for the treatment of human spinal injuries.Key words: netrin-1, dorsal root ganglion, axon guidance, chemorepellent, Unc5, spinal cord, axon regenerationDeveloping axons navigate to their targets by responding to attractive and repulsive guidance cues working in a contact-dependent or diffusible fashion in their environment (reviewed in ref. 1). During early development of the primary sensory system, centrally projecting sensory axons from dorsal root ganglion (DRG) neurons extend toward the dorsolateral region of the spinal cord (Fig. 1A and C), where they enter the spinal cord exclusively through the dorsal root entry zone (DREZ), and never orient themselves toward the notochord or the ventral spinal cord (Fig. 1A; reviewed in ref. 2). We previously showed that the notochord but not the ventral spinal cord secretes semaphorin 3A (Sema3A), which is known to be a chemorepellent for DRG axons at early developmental stages (Fig. 1A).3 This is the reason why DRG axons never project toward the notochord. Along the same line, it is highly possible that the ventral spinal cord may secrete some chemorepulsive cue other than Sema3A for DRG axons.Open in a separate windowFigure 1Netrin-1 plays a critical role in sensory axonal guidance as an axon chemorepellent. (A) A schematic diagram of a thoracic transverse section of an E10 mouse embryo, summarizing the possible mechanism of netrin-1 action in early DRG axonal guidance. When DRG axons project toward the DREZ in the dorsal spinal cord (dSC), ventrally derived netrin-1 chemorepels DRG axons to prevent them from orienting aberrantly toward the ventral spinal cord (vSC) (upper). NC; notochord. In netrin-1-deficient embryos, some DRG axons misorient themselves toward the ventral spinal cord, because of the absence of netrin-1 proteins in the ventral spinal cord (lower). (B) At E12.5 when DRG axons grow to the marginal zone of the spinal cord longitudinally (arrows) to form the dorsal funiculus (DF), netrin-1 proteins are transiently expressed in a subpopulation of dorsal spinal cord cells adjacent to the dorsal funiculus (upper). In netrin-1-deficient embryos, the dorsal funiculus is disorganized because DRG axons are no longer waiting for invading the dorsal mantle layer (lower). (C) Gain-of-function experiments by electroporation confirm the repulsive activity of netrin-1 toward DRG axons. When netrin-1 is misexpressed in the dorsal spinal cord, the number of DRG axons that enter the DREZ is significantly reduced compared with the control, because some DRG axons fail to project toward the DREZ and turn in the wrong direction.After entering the spinal cord, DRG axons grow to the marginal zone of the spinal cord longitudinally to form the dorsal funiculus without projecting to the dorsal mantle layer for a few days (this delay of the axonal projection to the mantle layer is referred to as the ‘waiting period;’ Fig. 1B). A few days later, proprioceptive afferents of DRGs begin to send collaterals into the dorsal layers, and cutaneous afferents project ventrally through the dorsal layers.4 This evidence raises the possibility that some repulsive cues transiently prevent the collaterals of DRGs from penetrating the dorsal spinal cord during this waiting period.Netrins are a family of secreted proteins that play a key role in axonal guidance, cell migration, morphogenesis and angiogenesis.5 Netrin-1 is a bifunctional axonal guidance cue, attracting some axons including commissural axons via the Deleted in Colorectal Cancer (DCC) receptor and repelling others via Unc5 receptors (reviewed in ref. 6). However, it has not been clear whether netrin-1 plays a role in sensory axonal guidance during development.Several observations strongly suggest a role for netrin-1 in DRG axonal guidance as a repulsive guidance cue during development.7,8 First, in the mouse embryo at embryonic day (E) 10–11.5 when many DRG axons orient themselves to reach the DREZ, netrin-1 is strongly expressed in the floor plate of the ventral spinal cord but not in the dorsal spinal cord (Fig. 1A). Second, at E12.5 when DRG neurons extend their axons longitudinally along the dorsolateral margin of the spinal cord, netrin-1 is expressed in the dorsolateral region adjacent to the DREZ (Fig. 1B), but its expression is down-regulated in the dorsal spinal cord at E13.5 when many collaterals have entered the mantle layer. Third, repulsive netrin-1 receptor Unc5c is expressed in the DRG neurons during development.These observations motivated us to explore whether netrin-1/Unc5c signaling contributes to DRG axonal guidance. We used cell and tissue cultures combined with tissues from netrin-1-deficient mice. We clearly showed that netrin-1 exerts a chemorepulsive activity toward developing DRG axons and that the ventral spinal cord-derived repulsive activity depends on netrin-1 in vitro.8 Additional evidence for a chemorepulsive role of netrin-1 came from the observation of DRG axonal trajectories in netrin-1-deficient mice.7,8 In netrin-1-deficient embryos at E10, we showed that some DRG axons became misoriented toward the ventral spinal cord, probably because of the absence of netrin-1 proteins in the ventral spinal cord (Fig. 1A). In addition, at E12.5 when DRG axons grow to the marginal zone of the spinal cord longitudinally to form the dorsal funiculus, the dorsal funiculus is disorganized in netrin-1-deficient embryos, because in the absence of netrin-1 DRG axons are not waiting for invading the dorsal mantle layer adjacent to the dorsal funiculus (Fig. 1B). Gain-of-function experiments further confirmed the repulsive activity of netrin-1 toward DRG axons (Fig. 1C). These lines of evidence lead us to the conclusion that dorsally derived netrin-1 plays an important role in providing the ‘waiting period’ for extension of collaterals from sensory afferents and that ventrally derived netrin-1 prevents sensory axons from misorienting themselves toward the ventral spinal cord.At later developmental stages (E13.5), DRG axons still possess a weak responsiveness to the chemorepulsive activity of netrin-1 in vitro.8 In addition, both postnatal and adult DRG neurons respond to netrin-1-induced axon inhibition.9 Consistent with these results, DRG neurons at not only later developmental stages (E13.5) but also postnatal stages express the repulsion-mediating netrin-1 receptor Unc5c.8,9Generally, lesioning of the dorsal column projection of sensory axons results in a complete lack of regeneration. The possible explanation for the complete lack of regeneration is that the environment, the lesion site itself and/or oligodendrocytes adjacent to the lesion, may be non-permissive for regenerating axons.10 Sema3A and chondroitin sulfate proteoglycans (CSPGs) are candidates as major inhibitors of sensory axonal regeneration in the spinal cord, because they are expressed in the lesion site and can inhibit DRG axonal growth in vitro.3,1114 Recently, Kaneko et al. showed that a selective inhibitor of Sema3A also enhances axonal regeneration and functional recovery in a subpopulation of sensory neurons after lesioning of the dorsal column.12 More recently, McMahon''s group clearly demonstrated that enzymatic degradation of CSPGs on the dorsal column lesion of the spinal cord promotes sensory axonal regeneration and functional recovery.13,14 Although these treatments greatly improved functional recovery, complete sensory axonal growth and functional recovery have not been yet achieved after the spinal cord injury. To promote further recovery of sensory axonal regeneration in the CNS, we should focus on other candidate inhibitors of CNS injury sites.Following spinal cord injury, the expression of the attraction- mediating netrin-1 receptor DCC decreases, while the expression level of the repulsive receptor Unc5c returns to normal.15 Levels of netrin-1 expression also stay unchanged in neurons and oligodendrocytes adjacent to the lesion site. Together with the in vitro evidence described above, these data strongly suggest a possible role for netrin-1 as a novel inhibitor of CNS myelin for regenerating DRG axons in the dorsal column-lesioned spinal cord. Further studies will be required to show directly the functional recovery of sensory axons in the spinal cord by perturbation of netrin-1 in and around the lesion site after spinal cord injury.  相似文献   

19.
20.
Characterization of aggregation profiles of monoclonal antibodies (mAb) is gaining importance because an increasing number of mAb-based therapeutics are entering clinical studies and gaining marketing approval. To develop a successful formulation, it is imperative to identify the critical biochemical properties of each potential mAb drug candidate. We investigated the conformational change and aggregation of a human IgG1 using external dye-binding experiments with fluorescence spectroscopy and compared the aggregation profiles obtained to the results of size-exclusion chromatography. We show that using an appropriate dye at selected mAb concentration, unfolding or aggregation can be studied. In addition, dye-binding experiments may be used as conventional assays to study therapeutic mAb stability.Key words: therapeutic monoclonal antibody, protein aggregation, conformational change, stability and shelf-life prediction, accelerated studiesMonoclonal antibodies (mAbs) have emerged as a novel class of protein drugs and are utilized for a variety of mostly incurable and debilitating diseases such as cancer and rheumatoid arthritis.14 For treatment of chronic diseases, it is desirable for these drugs to be administered subcutaneously, in which case high protein concentrations (>100 mg/mL) are generally needed.5,6 Protein-based drugs containing mAbs must contain minimum amounts of aggregation and fragmentation and conserve their structural integrity during storage because degraded or aggregated protein may induce immunogenicity or reduce efficacy. Currently, size-exclusion chromatography-high performance liquid chromatography (SEC-HPLC) is the most commonly used method to characterize mAb aggregation profiles;7 however it is time consuming, expensive and requires expertise. SEC-HPLC cannot be used to obtain accurate biophysical profiles of mAbs at high concentrations because dilution during the experiment might lead to reversible aggregation. Furthermore, the potential interaction of aggregates with surfaces, e.g., needle, tubing, column, will lead to the loss of sample and thus an inaccurate analysis.8,9 Additional drawbacks of the technique are that different conformations such as partially unfolded monomers also cannot be distinguished by SEC-HPLC and large aggregates may be totally excluded during the injection into the column.External dye binding assays have been used to characterize protein stability and aggregation,1012 and studies involving biopharmaceuticals have been reported recently, e.g., for thermostability screening10 and detection of aggregation.1114 These methods are not limited by protein quantity and are more sensitive because they are fluorescence-based. We studied the accelerated unfolding of an IgG1 mAb with the hydrophobic dye 1-anilino-8-naphthale-nesulfonate (ANS), and its accelerated aggregation with aggregate specific Thioflavin T (ThT). We have also conducted accelerated aggregation studies with SEC-HPLC7 and compared the findings to the ThT binding results. We hypothesize that key structures formed during mAb aggregation can be probed selectively by the appropriate dyes (Fig. 1) with specific mAb concentrations.Open in a separate windowFigure 1Key structures of the mAb probed by fluorescent dyes. N and U are native and unfolded monomers, respectively. “n” reactive monomers form aggregates.  相似文献   

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