研究者業績

林 謙介

ハヤシ ケンスケ  (Hayashi Kensuke)

基本情報

所属
上智大学 理工学部物質生命理工学科 教授
学位
理学博士(東京大学)

連絡先
kensuk-hsophia.ac.jp
研究者番号
50218567
J-GLOBAL ID
200901093609757481
researchmap会員ID
1000161071

1983-1988 東京大学大学院理学研究科生物学専攻 ニワトリ胚消化管の発生に関する研究
1988-1993 国立精神神経センターI神経研究所 ニワトリ胚骨格筋の発生に関する研究
1993-1998 群馬大学医学部 神経細胞のアクチン結合タンパクに関する研究
1998-2004 群馬大学生体調節研究所 神経細胞の細胞極性と細胞移動に関する研究
2004-現在 上智大学理工学部 神経細胞の微小管形成に関する研究


論文

 49
  • Akari Nakamura, Mami Ikeda, Seina Kusayanagi, Kensuke Hayashi
    IBRO Neuroscience Reports 13 264-273 2022年12月  査読有り最終著者責任著者
  • Kaho Miyata, Kensuke Hayashi
    Developmental Neuroscience 44 1-11 2022年9月  査読有り最終著者責任著者
    Neuronal migration and axon elongation in the developing brain are essential events for neural network formation. Leading processes of migrating neurons and elongating axons have growth cones at their tips. Cytoskeletal machinery for advance of growth cones of the two processes has been thought the same. In this study, we compared axonal-elongating growth cones and leading-process growth cones in the same conditions that manipulated filopodia, lamellipodia, and drebrin, the latter mediates actin filament-microtubule interaction. Cerebral cortex (CX) neurons and medial ganglionic eminence (MGE) neurons from embryonic mice were cultured on less-adhesive cover glasses. Inhibition of filopodia formation by triple knockdown of mammalian-enabled, Ena-VASP-like, and vasodilator-stimulated phosphoprotein or double knockdown of Daam1 and fascin affected axon formation of CX neurons but did not affect the morphology of leading process of MGE neurons. On the other hand, treatment with CK666, to inhibit lamellipodia formation, did not affect axons but destroyed the leading-process growth cones. When drebrin was knocked down, the morphology of CX neurons remained unchanged, but the leading processes of MGE neurons became shorter. In vivo assay of radial migration of CX neurons revealed that drebrin knockdown inhibited migration, while it did not affect axon elongation. These results showed that the filopodia-microtubule system is the main driving machinery in elongating growth cones, while the lamellipodia-drebrin-microtubule system is the main system in leading-process growth cones of migrating neurons.
  • Koyo Ide, Mika Muko, Kensuke Hayashi
    Histochemistry and Cell Biology 156 273-281 2021年6月10日  査読有り最終著者責任著者
  • Suzuki Y, Otake A, Ueno S, Hayashi K, Ishii H, Miyoshi N, Kuroiwa K, Tachikawa M, Fujimaki Y, Nishiyama K, Manabe K, Yamazaki R, Asai A
    ACS Med. Chem. Lett. 11(6) 1287-1291 2020年  査読有り
  • Mimori Yamada, Kensuke Hayashi
    Cytoskeleton 76 339-345 2019年  査読有り最終著者責任著者
  • Hatakeyama Eiko, Hayashi Kensuke
    Biochemical and Biophysical Research Communications 507 389-394 2018年11月  査読有り最終著者責任著者
  • Wataru Kojima, Kensuke Hayashi
    Histochemistry and Cell Biology 149(5) 529-536 2018年5月1日  査読有り最終著者責任著者
    Demyelination induced by cuprizone in mice has served a useful model system for the study of demyelinating diseases, such as multiple sclerosis. Severity of demyelination by cuprizone, however, varies across different regions of the central nervous system the corpus callosum is sensitive, while the optic nerves are resistant. Here, we investigated the effects of cuprizone on optic nerves, focusing on the axo-glial junctions. Immunostaining for sodium channels, contactin-associated protein, neurofascins, and potassium channels revealed that there were no massive changes in the density and morphology of the axo-glial junctions in cuprizone-treated optic nerves. However, when we counted the number of incomplete junctional complexes, we observed increased numbers of isolated paranodes. These isolated paranodes were immunopositive for both axonal and glial membrane proteins, indicating that they were the contact sites between axons and glia. These were not associated with sodium channels or potassium channels, suggesting the absence of physiological functions. When teased axons from cuprizone-treated optic nerves were immunostained, the isolated paranodes were found at the internode region of the myelin. From these observations, we conclude that cuprizone induces new contacts between axons and myelins at the internode region.
  • Kensuke Hayashi
    Advances in Experimental Medicine and Biology 1006 83-101 2017年  査読有り招待有り筆頭著者
    Drebrin is localized in actin-rich regions of neuronal and non-neuronal cells. In mature neurons, its localization is strictly restricted to the postsynaptic sites. In order to understand the function of drebrin in cells, many studies have been performed to examine the effect of overexpression or knocking down of drebrin in various cell types, including neurons, myoblasts, kidney cells, and intestinal epithelial cells. In most cases alteration of cell shape and impairment or facilitation of actin-based activities of these cells were observed. Interestingly, overexpression of drebrin in matured neurons results in the alteration in dendritic spine morphology. Further studies have shown alteration in the localization of postsynaptic receptors and even changes in synaptic transmission caused by drebrin overexpression or depletion in neurons. These drebrin’s effects are thought to come from drebrin’s actin-cross-linking activity or competitive binding to actin against tropomyosin, fascin, and α-actinin. Furthermore, drebrin binds to various molecules, such as homer, EB3, and cell-cell junctional proteins, indicating that drebrin is a multifunctional cytoskeletal regulator.
  • Yuri Niimura, Yuichi Aminaka, Kensuke Hayashi
    NEUROREPORT 26(4) 179-185 2015年3月  査読有り最終著者責任著者
    Migrating neurons have leading processes that direct cell movement in response to guidance cues. We investigated the involvement of glycogen synthase kinase 3 (GSK3) in the formation of leading processes and migration of neurons in vitro. We used embryonic rat medial ganglionic eminence (MGE) neurons, which are precursors of inhibitory neurons that migrate into the cerebral cortex. When MGE neurons were placed on an astrocyte layer, they migrated freely with the highest speed among neurons from other parts of the embryonic forebrain. When they were cultured alone, they showed bipolar morphology and extended leading processes within 20 h. Their leading processes had large growth cones, but did not elongate during 3 days in culture, indicating that leading processes are distinct from short axons. Next, we examined the effect of GSK3 inhibitors on leading processes and the migratory behavior of MGE neurons. MGE neurons treated with GSK3 inhibitors showed multipolar morphology and altered process shapes. Moreover, migration of MGE neurons on the astrocyte layer was significantly decreased in the presence of GSK3 inhibitors. These data suggest that GSK3 is involved in the formation of leading processes and in the migration of MGE neurons. Copyright (C) 2015 Wolters Kluwer Health, Inc. All rights reserved.
  • Satoshi Yonezawa, Momoko Shigematsu, Kazuto Hirata, Kensuke Hayashi
    Acta Histochemica et Cytochemica 48(5) 145-152 2015年  査読有り最終著者責任著者
  • Makoto T. Fujiwara, Yasushi Yoshioka, Tomonari Hirano, Yusuke Kazama, Tomoko Abe, Kensuke Hayashi, Ryuuichi D. Itoh
    Plant Signaling and Behavior 7(1) 34-37 2012年  査読有り
    Organelle dynamics in the plant male gametophyte has received attention for its importance in pollen tube growth and cytoplasmic inheritance. We recently revealed the dynamic behaviors of plastids in living Arabidopsis pollen grains and tubes, using an inherent promoter-driven FtsZ1-green fluorescent protein (GFP) fusion. Here, we further monitored the movement of pollen tube plastids with an actin1 promoter-driven, stroma-targeted yellow fluorescent protein (YFP). In elongating pollen tubes, most plastids localized to the tube shank, where they displayed either retarded and unsteady motion, or fast, directional, and long-distance movement along the tube polarity. Efficient plastid tracking further revealed a population of tip-forwarding plastids that undergo a fluctuating motion(s) before traveling backward. The behavior of YFP-labeled plastids in pollen basically resembled that of FtsZ1-GFP-labeled plastids, thus validating the use of FtsZ1-GFP for simultaneous visualization of the stroma and the plastid-dividing FtsZ ring. © 2012 Landes Bioscience.
  • Wataru Kawabata, Satoru Yonezawa, Kensuke Hayashi
    DEVELOPMENTAL NEUROSCIENCE 34(1) 20-29 2012年  査読有り最終著者責任著者
    Axon elongation is usually performed by the migration of growth cones that leave axons. Axon microtubules are generated by enhanced polymerization of tubulin in the growth cones. Some kinds of neurons like cerebellar granule cells, however, generate axons as a result of migration of the cell body leaving axons at the rear. The mechanism to generate microtubules during such growth cone-independent elongation of axons is not well understood. To establish an experimental model to study this mechanism, we cultured neuroblastoma (Neuro-2a) cells on substrates that facilitate cell migration. When cultured on laminin-treated substrate, cells migrated actively and left processes at the rear. We investigated the role of the centrosome in this process formation. The centrosomes were always located at the base of the processes, i.e., at the rear side of the migrating cell body. Close observation of cytoskeletons revealed microtubules limited around the centrosomes, but concentrated at the periphery of the cells or within the processes. Microtubule re-growth experiments showed the ability of the centrosomes to nucleate microtubules. We thus examined the role of microtubule release from the centrosomes, by knocking down the expression of spastin, a microtubule-severing enzyme. Introducing siRNA for spastin into Neuro-2a cells reduced both the migration speed and the length of the processes. Taken together, Neuro-2a cells on laminin proved useful as a model to study the alternative type of axon elongation in which cell migration leaves axons at the rear. This model provided evidence for the involvement of microtubule release from centrosomes in the mechanisms for this type of process elongation. Copyright (C) 2012 S. Karger AG, Basel
  • Yusaku Ohama, Kensuke Hayashi
    HISTOCHEMISTRY AND CELL BIOLOGY 132(5) 515-524 2009年11月  査読有り最終著者責任著者
    Microtubules in typical cells form radial arrays with their plus-ends pointing toward the cell periphery. In contrast, microtubules in dendrites of neurons are free from centrosomes and have a unique arrangement in which about half have a polarity with a minus-end distal orientation. Mechanisms for generation and maintenance of the microtubule arrangement in dendrites are not well understood. Here, we examined dendritic localization of a centrosomal protein, ninein, which has microtubule-anchoring and stabilizing functions. Immunohistochemical analysis of developing mouse cerebral and cerebellar cortices showed that ninein is localized at the centrosome in undifferentiated neural precursors. In contrast, ninein was barely detected in migrating neurons, such as those in the intermediate layer of the cerebral cortex and the internal granular layer of the cerebellar cortex. High expression was observed in thick dendrite-bearing neurons such as pyramidal neurons of the cerebral cortex and Purkinje neurons in the cerebellar cortex. Ninein was not detected at the centrosome of these cells, but was diffusely present in cell soma and dendrites. In cultured cortical neurons, ninein formed granular structures in soma and dendrites, being not associated with gamma-tubulin. About 60% of these structures showed resistance to detergent and association with microtubules. Our observations suggest that the minus-ends of microtubules may be anchored and stabilized by centrosomal proteins localized in dendrites.
  • Daisuke Takahashi, Wenqian Yu, Peter W. Baas, Rika Kawai-Hirai, Kensuke Hayashi
    CELL MOTILITY AND THE CYTOSKELETON 64(5) 347-359 2007年5月  査読有り最終著者責任著者
    Axons and dendrites of neurons differ in the polarity orientation of their microtubules. Whereas the polarity orientation of microtubules in axons is uniform, with all plus ends distal, that in dendrites is nonuniform. The mechanisms responsible for establishment and maintenance of microtubule polarity orientation in neuronal processes remain unclear, however. We previously described a culture system in which dendrites of rat cortical neurons convert to axons. In the present study, we examined changes in microtubule polarity orientation in such dendrites. With the use of the hooking procedure and electron microscopy, we found that microtubule polarity orientation changed from nonuniform to uniform, with a plus end-distal arrangement, in dendrites that gave rise to axons during culture of neurons for 24 It. Microtubule polarity orientation remained nonuniform in dendrites that did not elongate. Axon regeneration at the dendritic tip thus triggered the disappearance of minus end-distal microtubules from dendrites. These minus end-distal microtubules also disappeared from dendrites during axon regeneration in the presence of inhibitors of actin polymerization, suggesting that actin-dependent transport of microtubules is not required for this process and implicating a previously unidentified mechanism in the establishment and maintenance of microtubule polarity orientation in neuronal processes.
  • SEKINO YUKO, TANAKA SATOSHI, HANAMURA KENJI, YAMAZAKI HIROYUKI, SASAGAWA YOSHIO, XUE YINHUAN, HAYASHI KENSUKE, SHIRAO TOMOAKI
    Molecular and Cellular Neuroscience 31(3) 493-504 2006年3月  査読有り
  • K Sato, E Malchinkhuu, T Muraki, K Ishikawa, K Hayashi, M Tosaka, A Mochiduki, K Inoue, H Tomura, C Mogi, H Nochi, K Tamoto, F Okajima
    JOURNAL OF NEUROCHEMISTRY 92(4) 904-914 2005年2月  査読有り
    Cerebrospinal fluid (CSF) induced neurite retraction of differentiated PC12 cells; the action was observed in 15 min (a rapid response) and the activity further increased until 6 h (a long-acting response) during exposure of CSF to the cells. The CSF action was sensitive to monoglyceride lipase and diminished by homologous desensitization with lysophosphatidic acid (LPA) and by pretreatment with an LPA receptor antagonist Ki16425. Although fresh CSF contains LPA to some extent, the LPA content in the medium was increased during culture of PC12 cells with CSF. The rapid response was mimicked by exogenous LPA, and a long-acting response was duplicated by a recombinant autotaxin, lysophospholipase D (lyso-PLD). Although the lyso-PLD substrate lysophosphatidylcholine (LPC) was not detected in CSF, lyso-PLD activity and an similar to120-kDa autotaxin protein were detected in CSF. On the other hand, LPC but not lyso-PLD activity was detected in the conditioned medium of a PC12 cell culture without CSF. Among neural cells examined, leptomeningeal cells expressed the highest lyso-PLD activity and autotaxin protein. These results suggest that leptomeningeal cells may work as one of the sources for autotaxin, which may play a critical role in LPA production and thereby regulate axonal and neurite morphological change.
  • 林謙介
    生命科学研究所紀要 23(23) 39-47 2004年  査読有り筆頭著者
  • K Hayashi, R Kawai-Hirai, A Harada, K Takata
    JOURNAL OF CELL SCIENCE 116(21) 4419-4428 2003年11月  査読有り筆頭著者責任著者
    Inhibitory and excitatory neurons exhibit distinct patterns of development in the mammalian cerebral cortex. The morphological development of inhibitory and excitatory neurons derived from fetal rat cerebral cortex has now been compared in vitro. Inhibitory neurons were identified by immunofluorescence staining with antibodies to gamma-aminobutyric acid, and axon formation was detected by staining with antibodies to phosphorylated neurofilaments. In chemically defined, glia-free and low-density cultures, excitatory neurons formed axons within three days of plating. By contrast, inhibitory neurons required more than six days to form axons. Time-lapse analysis over six days revealed that most inhibitory neurons were bipolar and that their two processes exhibited alternate growth and retraction without giving rise to axons. Movement of the cell body towards the growing process was apparent in about one-half of inhibitory neurons, whereas such movement was never seen in excitatory neurons. The migratory behavior of neurons was further investigated by culture on a glial cell monolayer. Inhibitory neurons migrated over substantially larger distances than did excitatory neurons. The centrosome of inhibitory neurons translocated to the base of the newly emerging leading process, suggesting the existence of a force that pulls intracellular organelles towards the leading process. Centrosome translocation was not detected in excitatory neurons. These observations suggest that the developmental programs of excitatory and inhibitory neurons differ. Inhibitory neurons thus possess a more effective cytoskeletal machinery for migration than excitatory neurons and they form axons later.
  • K Hayashi, R Kawai-Hirai, K Ishikawa, K Takata
    NEUROSCIENCE 110(1) 7-17 2002年  査読有り筆頭著者責任著者
    The mechanisms for the establishment and maintenance of cell polarity in neurons are not well understood. Axon regeneration from dendrites has been reported after axotomy near the cell body in vivo. We report here in vitro a reversal of neuronal polarity characterized by the conversion of dendrites into axons. We isolated neurons from the neonatal rat cerebral cortex. Neurons that exhibited an apical dendrite with a length of > 100 mum were monitored for 3 days in culture. In 66% of neurons examined, a new axon, as identified by reactivity with an antibody to dephosphorylated tau or by lack of reactivity with an antibody to the a and b isoforms of microtubule-associated protein 2, appeared to form from the tip of the original dendrite. Further analysis of such neurons revealed that the distal half of the original dendrite became positive for dephosphorylated tau or negative for microtubule-associated protein 2. Time-lapse video microscopy demonstrated the conversion of the original dendrite into an axon without dendritic retraction. Axon regeneration from dendritic tips required a significantly longer time than axon regeneration from minor processes. Our observations thus demonstrate in vitro a time-consuming reversal of neuronal polarity and the conversion of a dendritic cytoskeleton into an axonal one. (C) 2002 IBRO. Published by Elsevier Science Ltd. All rights reserved.
  • T Kanno, Y Kubohara, F Goto, K Takata, K Hayashi
    DEVELOPMENT GROWTH & DIFFERENTIATION 43(6) 709-716 2001年12月  査読有り最終著者責任著者
    Differentiation-inducing factor-1 (DIF-1) is a chlorinated alkylphenone (small lipophilic hormone) that induces stalk cell formation in the cellular slime mold Diclyostelium discoideum. Recent studies have revealed that DIF-1 inhibits growth and induces the differentiation of mammalian tumor cells. The present study examines the effects of DIF-1 on rat cortical neurons in primary culture. We found that DIF-1 induced rapid neuronal cell death, The release of lactate dehydrogenase (LDH), as an indicator of cell death, increased dose-dependently with DIF-1, The release of LDH was inhibited by the N-methyl-D-aspartate (NMDA) receptor antagonists MK801 and AP5, suggesting that the NMDA receptor is involved in the induction of cell death by DIF-1. However, glutamate cytotoxicity could not explain the entire action of DIF-1 on neurons because the estimated concentration of glutamate around DIF-1-treated neurons was below 50 pm and DIF-1 caused more severe cell death than 500 pm glutamate. We discovered that another portion of DIF-1 cytotoxicity is independent of the NMDA receptor; that is, coaddition of DIF-1 and MK801 induced dendritic beading and increased expression of the immediate early genes c-fos and zif/268. These results indicate that DIF-1 induces rapid cell death via both NMDA receptor-dependent and -independent pathways in rat cortical neurons.
  • 林謙介
    日本神経精神薬理学雑誌 21(3) 83-87 2001年  査読有り筆頭著者
  • XT Cheng, K Hayashi, T Shirao
    NEUROSCIENCE RESEARCH 36(2) 167-173 2000年2月  
    Dendritic spines ale extremely motile, providing a structural mechanism for synaptic plasticity. Actin-myosin interaction is thought to be responsible for the change in the shape of spine. We have already reported that drebrin, an actin-binding protein, inhibits actin-myosin interaction and is enriched in the dendritic spine of mature neurons. In this study, we prepared the actin cytoskeleton of dendritic spines as an immunoprecipitate with anti-drebrin antibody from adult guinea-pig brain, immunized mice with the cytoskeleton, and obtained a monoclonal antibody (MAb) called MAb G650. MAb G650 reacted with non-muscle myosin IIB, but it did not react with muscle myosin II or non-muscle myosin IIA. Immunoblotting with this antibody revealed that drebrin-binding cytoskeleton contains this myosin IIB-like immunreactivity. Immunohistochemistry using MAb G650 demonstrated that this myosin IIB-like immunreactivity can be detected in the neuronal cell bodies and their apical dendrites, where drebrin is hardly detected. These data demonstrate that a myosin subtype associated with drebrin-binding actin filaments in the dendritic spines is myosin IIB, although this myosin is widely distributed in somato-dendritic subdomains of neurons. Furthermore, it is indicated that the cytoskeletons in dendritic spine were uniquely characterized with actin-binding proteins such as drebrin, but not with myosins. (C) 2000 Elsevier Science Ireland Ltd. All rights reserved.
  • K Hayashi, R Ishikawa, R Kawai-Hirai, T Takagi, A Taketomi, T Shirao
    EXPERIMENTAL CELL RESEARCH 253(2) 673-680 1999年12月  
    Drebrin is an actin-binding protein which is expressed at highly levels in neurons. When introduced into fibroblasts, it has been known to bind to F-actin and to cause remodeling of F-actin. Here, we performed a domain analysis of the actin-binding and actin-remodeling activities of drebrin. Various fragments of drebrin cDNA were fused with green fluorescent protein cDNA and introduced into Chinese hamster ovary cells. Association of the fusion protein with F-actin and remodeling of the F-actin were examined. We found that the central 85-amino-acid sequence (residues 233-317) was sufficient for the binding to and remodeling of F-actin. The binding activity of this fragment was relatively low compared with that of full-length drebrin, but all the types of abnormalities of F-actin that are observed with full-length drebrin were also observed with this fragment. When this sequence was further fragmented, the actin-binding activity was greatly reduced and the actin-remodeling activity disappeared. The actin-binding activity of the central region of drebrin was confirmed by a cosedimentation assay of chymotryptic fragments of drebrin with purified actin. These data indicate that the actin-binding domain and actin-remodeling domain are identical and that this domain is located at the central region of drebrin. (C) 1999 Academic Press.
  • 林謙介
    生化学 71(7) 520-523 1999年7月  
  • K Hayashi, T Shirao
    JOURNAL OF NEUROSCIENCE 19(10) 3918-3925 1999年5月  
    Dendritic spines are known to be extremely motile, providing a structural mechanism for synaptic plasticity. Actin filaments are thought to be responsible for the changes in the shape of spines. We tested our hypothesis that drebrin, an actin-binding protein, is a regulator of spine shape. In high-density long-term primary cultures of rat cerebral cortex neurons, drebrin was colocalized with actin filaments at spines. We introduced drebrin tagged with green fluorescent protein (GFP) into these neurons to test the ability of exogenous drebrin to localize at spines and the effect of overexpression of drebrin on spine shape. We observed that exogenous drebrin indeed accumulated in spines. But when the actin-binding domain of drebrin was deleted, the protein was distributed in both spines and dendritic shafts, indicating that accumulation of drebrin in the spines required its actin-binding activity. Statistical analysis of the lengths of spines as determined from confocal laser microscopic images revealed that the spines were significantly longer in GFP-drebrin-expressing neurons than in GFP-expressing neurons. The longer spines labeled with GFP-drebrin were demonstrated to be postsynaptic by double labeling of the presynaptic terminals with antibody against synaptophysin. These results directly indicate that drebrin binds to actin filaments at dendritic spines and alters spine shape.
  • T Shirao, R Yong, K Hayashi, Y Sekino
    NEURAL DEVELOPMENT-BOOK 2 381-385 1999年  査読有り
    The cytoskeleton has been suggested to be one of the important endogenous factors that control the elaborate morphology of neuronal processes. In a previous paper, we showed that overexpression of neuronal actin-binding protein, drebrin A, in fibroblasts results in the appearance of thick, curving bundles of actin filaments. In this chapter, we have analyzed, using a time-lapse recording microscopy system, how drebrin-expressing fibroblasts formed highly branching cell processes. Because drebrin is a cytoskeletal protein, drebrin in living cells could not be immunostained. Therefore, L cells were cotransfected with cDNAs of both drebrin A and green fluorescent protein (GFP), and drebrin-expressing cells were identified as GFP-positive cells. Their cell shape was at first thick and round, seldom bearing any filopodia. Then, outgrowth of filopodia-like cell processes suddenly increased, apparently because of accumulation of drebrin A. The cell processes of drebrin-expressing cells were more stable than normal filopodia and seldom retract. Finally, these cell processes changed into long cell processes.
  • K Hayashi, T Shirao
    NEURAL DEVELOPMENT-BOOK 2 399-403 1999年  査読有り
    We have previously reported that an actin-binding protein, drebrin, accumulates at spines and modulates actin-myosin interaction within spines. Drebrin may be one of the regulators of actin filament reorganization during the shape change of spines. We constructed chimeric cDNAs carrying a green fluorescent protein (GFP) cDNA and various fragments of drebrin cDNA and introduced them into CHO cells to test their actin-binding activity. We found that GFP-drebrin bound to actin filaments in CHO cells and caused remodeling of actin filaments and cell shape. A central 28-amino-acid sequence was identified as the core for actin binding. Next, we introduced the GFP-drebrin fusion cDNA into primary cultured cortical neurons. Three weeks after plating, spines branching off the neurons were visible under a fluorescence microscope. The fusion protein was localized at the head of spines. Furthermore, spines of neurons transfected with a GFP-drebrin cDNA were longer than those of GFP-transfected neurons. These results support our hypothesis that drebrin regulates the morphology of spines.
  • K Hayashi, K Suzuki, T Shirao
    DEVELOPMENTAL BRAIN RESEARCH 111(1) 137-141 1998年11月  
    Easier methods to evaluate synapse formation in cultured neurons are desirable to investigate the regulatory mechanisms of synaptic development. We focused on drebrin, which changes from embryonic-type to adult-type isoform during postnatal development. The adult-type isoform of drebrin was detected by Western blotting after seventh day in primary cultured cortical neurons and the expression was coincidental with that of synaptophysin. Reverse transcription-polymerase chain reaction could demonstrate reversal of the predominant drebrin mRNA isoform from embryonic type to adult type between 8 and 13 days of culture. This is an easy and quick method to evaluate synapse formation in vitro. (C) 1998 Elsevier Science B.V. All rights reserved.
  • 任永, 平井里香, 薛音歓, 林謙介, 白尾智明
    北関東医学 48(5) 343-350 1998年  
  • 林謙介
    北関東医学 47(6) 461-463 1997年  
  • K Hayashi, R Ishikawa, LH Ye, XL He, K Takata, K Kohama, T Shirao
    JOURNAL OF NEUROSCIENCE 16(22) 7161-7170 1996年11月  
    Morphological changes in the dendritic spines have been postulated to participate in the expression of synaptic plasticity. The cytoskeleton is likely to play a key role in regulating spine structure. Here we examine the molecular mechanisms responsible for the changes in spine morphology, focusing on drebrin, an actin-binding protein that is known to change the properties of actin filaments. We found that adult-type drebrin is localized in the dendritic spines of rat forebrain neurons, where it binds to the cytoskeleton. To identify the cytoskeletal proteins that associated with drebrin, we isolated drebrin-containing cytoskeletons using immunoprecipitation with a drebrin antibody. Drebrin, actin, myosin, and gelsolin were co-precipitated. We next examined the effect of drebrin on actomyosin interaction. In vitro, drebrin reduced the sliding velocity of actin filaments on immobilized myosin and inhibited the actin-activated ATPase activity of myosin. These results suggest that drebrin may modulate the actomyosin interaction within spines and may play a role in the structure-based plasticity of synapses.
  • S Kobayashi, K Isa, K Hayashi, HK Inoue, K Uyemura, T Shirao
    DEVELOPMENTAL BRAIN RESEARCH 90(1-2) 122-128 1995年12月  
    In order to elucidate the cellular mechanisms of migrating neurons, we developed an assay system in vitro, using an aggregation culture of developing granule cells from the rat cerebellum. This assay system allowed us to eliminate the effects of various factors other than neurons and to examine the direct effects of individual molecules on neuronal migration. In this assay system, we examined the effects of several protein kinase inhibitors on cerebellar granule cell migration, and revealed that K252a, an inhibitor of protein kinases and of the actions of neurotrophins, inhibited the migration. Within 5 min after the addition of K252a to the culture medium, most of the migrating spindle-shaped cells changed into non-migrating large and polygonal cells, which had many microspikes. Staining with rhodamine-phalloidin revealed the appearance of actin bundles that resembled stress fibers within these large cells. On the other hand, extension of neurites was not severely inhibited by the addition of K252a. These results suggest that the migration is regulated by a different mechanism from that of neurite growth.
  • K HAYASHI, E OZAWA
    DEVELOPMENT 121(3) 661-669 1995年3月  
    It is known that myogenic cells in limb buds are derived from somites. In order to examine the potential of the limb primordium (presumptive limb somatopleure) to induce myogenic cell migration, we transplanted chick presumptive limb somatopleure to the flank region of an embryo, a region that does not normally contribute myogenic cells to the limb. Semitic cell migration was examined using a vital labeling technique. When the presumptive limb somatopleure was transplanted and was in contact with the host flank somite, semitic-cell migration toward the graft was observed. The labeled semitic cells within the graft were identified as myogenic cells in two ways: first, we found that N-cadherin-expressing cells appeared in the graft. Second, after 3 further days of incubation, the semitic cells formed dorsal and ventral masses and expressed sarcomeric myosin heavy chain within the graft. Cell migration occurred only when the somite was in contact with the medial region of the presumptive limb somatopleure. When the somite was not in contact with the limb somatopleure, or when the somite was in contact with the lateral region of the limb somatopleure, migration did not occur. These observations indicate that the potential to induce myogenic cell migration is restricted to the medial region of the presumptive limb somatopleure and that tissue contact is required.
  • S TANAKA, K HAYASHI, K ISA, T SHIRAO
    JOURNAL OF NEUROCHEMISTRY 65 S91-S91 1995年  
  • R ISHIKAWA, K HAYASHI, T SHIRAO, YH XUE, T TAKAGI, Y SASAKI, K KOHAMA
    JOURNAL OF BIOLOGICAL CHEMISTRY 269(47) 29928-29933 1994年11月  
    Drebrin is a development-associated neuroprotein whose cDNA into fibroblasts causes the formation of dendrite-like structures (Shirao, T., Kojima, N., and Obata, K. (1992) Neuroreport 3, 109-112). To explore molecular functions of drebrin during brain development, we purified drebrin from brains of rat embryos. Drebrin bound to actin filaments at a stoichiometry of 1:5 with a dissociation constant (K-d) of 1.2 x 10(-7) M. It strongly inhibited the actin binding activity of tropomyosin. Excess amounts of tropomyosin also inhibited the drebrin binding to actin filaments, suggesting that drebrin and tropomyosin competitively bind to actin filaments. Further, drebrin inhibited not only the actin binding activity of alpha-actinin but also the actin cross-linking activity of alpha-actinin. Gene transfection experiments revealed that tropomyosin was dissociated from actin filaments in drebrin-overexpressing fibroblasts. Thus we hypothesize that drebrin may destabilize actin filaments by dissociating tropomyosin and alpha-actinin from actin filaments, resulting in the formation of axon and dendrites during neuronal development.
  • T SHIRAO, K HAYASHI, R ISHIKAWA, K ISA, H ASADA, K IKEDA, K UYEMURA
    EXPERIMENTAL CELL RESEARCH 215(1) 145-153 1994年11月  
    Drebrin A is a neuron-specific protein, the expression of which is regulated during development. Upon transfection of fibroblasts with drebrin A cDNA, the protein is expressed at high levels in fibroblasts and the outgrowth of highly branched, neurite-like cell processes is induced. In this report, we describe a biochemical examination of the binding of drebrin A to actin filaments. We also demonstrate by an immunocytochemical method that, when drebrin A is expressed in transfected cells, it binds to actin filaments and is concentrated in cell processes. Furthermore, we provide evidence that thick, curving bundles of actin together with drebrin are formed in some of the transfected cells. Our results suggest that the actin filaments that bind drebrin might be a novel class of actin filaments and might play a role in neuronal morphogenesis. (C) 1994 Academic Press, Inc.
  • A SUZUKI, M YOSHIDA, K HAYASHI, Y MIZUNO, Y HAGIWARA, E OZAWA
    EUROPEAN JOURNAL OF BIOCHEMISTRY 220(2) 283-292 1994年3月  
    Direct interaction between the C-terminal portion of dystrophin and dystrophin-associated proteins was investigated. The binding of dystrophin to each protein was reconstituted by overlaying bacterially expressed dystrophin fusion proteins onto the blot membranes to which dystrophin-associated proteins were transferred after separation by SDS/PAGE with the following results. (a) Among the components of the glycoprotein complex which links dystrophin to the sarcolemma, a 43-kDa dystrophin-associated glycoprotein binds directly to dystrophin. Although at least one of the binding sites of this protein resides within the cysteine-rich domain of dystrophin, a contribution of additional amino acid residues within the first half of the C-terminal domain was also suggested for more secure binding. (b) Two other proteins also directly bind to dystrophin. Their binding sites are suggested to reside within the last half of the C-terminal domain which is alternatively spliced depending on the tissue type. Previously, based on the enzyme digestion experiments, we showed that the binding site for the glycoprotein complex on dystrophin is present within the cysteine-rich domain and the first half of the C-terminal domain [Suzuki, A., Yoshida, M., Yamamoto, H. and Ozawa, E. (1992) FEBS Lett. 308, 154-160]. Here, we have extended this work and found that the region which is involved in interaction with the complex is widely extended to the entire length of this part of the molecule. On the basis of the present results, we propose a model of molecular architecture at the binding site for the complex on dystrophin.
  • H YAMAMOTO, Y MIZUNO, K HAYASHI, NONAKA, I, M YOSHIDA, E OZAWA
    JOURNAL OF BIOCHEMISTRY 115(1) 162-167 1994年1月  
    A monoclonal antibody MA4-2 against a dystrophin-associated protein 35DAG (A4) was established and applied to examine the distribution of 35DAG in monkey tissue and its expression in DMD patients. In immunoblotting after two-dimensional gel electrophoresis of crude skeletal muscle extracts, MA4-2 reacted exclusively with an apparent single spot located in a similar position to rabbit 35DAG in each animal examined. 35DAG was detected only in striated muscles (quadriceps femoris and cardiac muscles), but not in other tissues examined, including smooth muscle (aorta, uterus), brain, nerve, lung, and liver. This distribution pattern is the same as that of 50DAG but different from that of 43DAG (A3a) [Mizuno et al. (1993) J. Biochem. 114,936-941]. In Duchenne muscular dystrophy muscles, 35DAG was distinctly present, but greatly reduced in amount. This is also the case with 50DAG. On the basis of these results, it is concluded that 35DAG and 50DAG are striated muscle-specific and may be important in the pathogenesis of Duchenne muscular dystrophy.
  • K HAYASHI, Y HAGIWARA, E OZAWA
    DEVELOPMENT GROWTH & DIFFERENTIATION 35(3) 301-309 1993年6月  
  • K IKEYA, K SAITO, K HAYASHI, H TANAKA, Y HAGIWARA, M YOSHIDA, A YAMAUCHI, Y FUKUYAMA, T ISHIGURO, C EGUCHI, E OZAWA
    AMERICAN JOURNAL OF MEDICAL GENETICS 43(3) 580-587 1992年6月  
    We examined the nucleotide sequence of deleted part of dystrophin mRNA and its translational product with immunoblot and immunohistochemical methods in a 6-year-old boy with a deleted DMD/BMD gene. On Southern blot analysis of his genomic DNA, we found a deletion of exons 10 to 37 in the DMD/BMD gene, which was expected to preserve the translational open reading frame (ORF). Dystrophin mRNA from his biopsy sample was amplified by polymerase chain reaction (PCR) and sequenced. The mRNA lacked the sequence corresponding to the gene from exons 10-37, and the translational ORF was preserved. The transcript was expected to code a 260 kDa protein. Dystrophin expressed in this patient was investigated with immunological methods. A 260 kDa protein was detected by immunoblot analysis with antidystrophin antiserum against nondeleted regions. These observations confirmed the preservation of the reading frame and the 260 kDa protein was produced as a mutant dystrophin. All these are compatible with the diagnosis of BMD. However, the immunohistochemical pattern of his muscle cells was peculiar. With deleted-region-directed antiserum, the membrane was not stained at all as in DMD patients. In contrast, with nondeleted-region-directed antiserum, all the muscle cell membrane was stained continuously as in non-DMD/BMD individuals. These are quite different from the staining pattern in most BMD patients where muscles are stained patchily or discontinuously.
  • H TANAKA, K HAYASHI, E OZAWA
    PROCEEDINGS OF THE JAPAN ACADEMY SERIES B-PHYSICAL AND BIOLOGICAL SCIENCES 67(8) 148-152 1991年10月  
    The mdx mouse has X-chromosome linked myopathy due to dystrophin deficiency, and is therefore used as a model of human Duchenne muscular dystrophy. As the mdx mouse bears a point mutation resulting in a stop codon in the middle of the dystrophin gene, most of its muscle fibers are not stained in immunohistochemistry with anti-dystrophin antisera. However, a clear immunostaining occurs on the cell membrane in a small number of fibers. We characterized these rare positive fibers in the present study with five different antibodies raised against various regions of dystrophin. They were stained with four of them including those raised against the N- and C-terminals of dystrophin. However, they were not stained with antibody Dy4/6D3 raised against the middle region of dystrophin close to the site corresponding to the mdx mutation. The staining could not be explained by the expression of the dystrophin related protein, an autosomal homologue of dystrophin. We tentatively concluded that the positive staining is due to truncated dystrophin lacking the region corresponding to the mdx mutation.
  • K HAYASHI, E OZAWA
    ROUXS ARCHIVES OF DEVELOPMENTAL BIOLOGY 200(4) 188-192 1991年  
    In order to understand how myogenic cells migrate in the limb bud, it is indispensable to distinguish undifferentiated myogenic cells from other mesenchymal cells. Thus, a suitable method for this purpose has been sought. A method to exchange the somites of a chicken and a quail microsurgically has widely been used, since the nuclei of the two species are morphologically distinguishable. However, microsurgery is accompanied by disturbances at the operated locus, and introducing cells of different species might induce unexpected effects. We report a new method for labelling chicken myogenic cells without transplantational operations, and describe their migration pattern in limb buds. Injection of a fluorescent carbocyanine dye into the somite lumen intensely labelled the somitic cells. Myogenic cells derived from the somite were clearly detected in limb buds. Before stage 20, the labelled cells were diffusely distributed in the proximal region of the limb bud. At about stage 21 in both wing and leg buds, labelled cells began to form dorsal and ventral masses. The label was followed until the cells differentiated and expressed myosin. This vital labelling method has advantages over the somite transplantation method: it does not include surgical operations that may disturb the normal development, and the cells are labelled intensely enough to be detected in a whole mount preparation.
  • K HAYASHI, S YASUGI, T MIZUNO
    BIOCHEMICAL AND BIOPHYSICAL RESEARCH COMMUNICATIONS 152(2) 776-782 1988年4月  
  • K HAYASHI, K AGATA, M MOCHII, S YASUGI, G EGUCHI, T MIZUNO
    JOURNAL OF BIOCHEMISTRY 103(2) 290-296 1988年2月  
  • K HAYASHI
    DEVELOPMENT GROWTH & DIFFERENTIATION 29(3) 285-295 1987年6月  
  • S YASUGI, K HAYASHI, K TAKIGUCHI, T MIZUNO, M MOCHII, R KODAMA, K AGATA, G EGUCHI
    DEVELOPMENT GROWTH & DIFFERENTIATION 29(1) 85-91 1987年2月  
  • HAYASHI Kensuke, MIZUNO T
    Comptes Rendus des Seances de la Societe de Biologie et de Ses Filiales 178(5) 584-587 1984年  

MISC

 12

共同研究・競争的資金等の研究課題

 23