Network


Latest external collaboration on country level. Dive into details by clicking on the dots.

Hotspot


Dive into the research topics where Hideyo Inouye is active.

Publication


Featured researches published by Hideyo Inouye.


Biophysical Journal | 1993

Structure of beta-crystallite assemblies formed by Alzheimer beta-amyloid protein analogues: analysis by x-ray diffraction

Hideyo Inouye; P.E. Fraser; Daniel A. Kirschner

To elucidate the relation between amyloid fibril formation in Alzheimer disease and the primary structure of the beta/A4 protein, which is the major component of the amyloid, we have been investigating the ability of peptides sharing sequences with beta/A4 to form fibrils in vitro. In previous studies we focused on the macroscopic morphology of the assemblies formed by synthetic peptides corresponding in sequence to different regions of this protein. In the present study we analyze the x-ray diffraction patterns obtained from these assemblies. All specimens showed wide angle reflections that could be indexed by an orthogonal lattice of beta-crystallites having unit cell dimensions a = 9.4 A, b = 7 A, and c = 10 A, where a refers to hydrogen bonding direction, b to polypeptide chain direction, and c to intersheet direction. Given the amino acid sequence of beta/A4 as NH2-DAEFRHDSGYEVHHQKLVFFAEDVGSNKGAIIGLMVGGVVIAT-COOH, we found that, based on their orientation and assembly, the analogues could be classified into three groups: Group A, residues 19-28, 13-28, 12-28, 11-28, 9-28, 1-28, 1-38, 1-40, 6-25, 11-25 and 34-42; Group B, residues 18-28, 17-28, and 15-28; and Group C, residues 22-35 and 26-33. For Groups A and C, the sharpest reflections were (h00), indicating that the assemblies were fibrillar, i.e., elongated in a single direction. Lateral alignment of the crystallites in Group A account for its cross-beta pattern, in which the hydrogen bonding (H-bonding) direction is the fiber (rotation) axis. By comparison, the beta-crystallites of Group C had no preferential orientation, thus giving circular scattering. For Group B, the sharpest reflections were (h0l) on the meridian, indicating that the assemblies were plate-like, i.e., extended in two directions. A series of equatorial Bragg reflections having a 40 A period indicated regular stacking of the plates, and the rotation axis was normal to the surface of the plates. Of the Group A peptides, the analogues 11-28 and 6-25 showed intensity maxima on the equator as well as on higher layer lines, indicating that the beta-crystallites are highly ordered relative to one another in the axial, H-bonding direction. This sampling of the layer lines by a larger period (60 A) suggests that the beta-crystallites are arrayed either in cylindrical or small restricted crystalline lattices. Consistent with its electron microscopic images, we modeled the structure as a tube with five or six f,-crystallites constituting the wall and with the individual crystallite, which either rotates freely or is restricted, made of five or fewer beta-pleated sheets. For the Group B peptides, the electron density projection along the b-axis was calculated from the observed intensities using phase combinations from fl-keratin.Amino acid side-chain positions were apparent and, when refined as 4-A-diameter spheres, led to a substantial decrease in the R-factors.For peptide 18-28 the electron density peaks, which are thought to correspond to side chains, were centered 3.3 A from the peptide backbone, whereas for peptides 17-28 and 15-28, these peaks were centered 1 A or more further from the backbone. Peaks having high electron density faced peaks having lower density, suggesting a favorable stereochemical arrangement of the residues. Thus, our analysis of the fiber x-ray patterns from beta/A4 peptides shows the organization of the beta-crystallites that form the wall of the amyloid fibrils as well as possible side-chain interactions.


Biophysical Journal | 1998

Structural analysis of Alzheimer's beta(1-40) amyloid: protofilament assembly of tubular fibrils.

Sergey B. Malinchik; Hideyo Inouye; Karen E. Szumowski; Daniel A. Kirschner

Detailed structural studies of amyloid fibrils can elucidate the way in which their constituent polypeptides are folded and self-assemble, and exert their neurotoxic effects in Alzheimers disease (AD). We have previously reported that when aqueous solutions of the N-terminal hydrophilic peptides of AD beta-amyloid (A beta) are gradually dried in a 2-Tesla magnetic field, they form highly oriented fibrils that are well suited to x-ray fiber diffraction. The longer, more physiologically relevant sequences such as A beta(1-40) have not been amenable to such analysis, owing to their strong propensity to polymerize and aggregate before orientation is achieved. In seeking an efficient and inexpensive method for rapid screening of conditions that could lead to improved orientation of fibrils assembled from the longer peptides, we report here that the birefringence of a small drop of peptide solution can supply information related to the cooperative packing of amyloid fibers and their capacity for magnetic orientation. The samples were examined by electron microscopy (negative and positive staining) and x-ray diffraction. Negative staining showed a mixture of straight and twisted fibers. The average width of both types was approximately 70 A, and the helical pitch of the latter was approximately 460 A. Cross sections of plastic-embedded samples showed a approximately 60-A-wide tubular structure. X-ray diffraction from these samples indicated a cross-beta fiber pattern, characterized by a strong meridional reflection at 4.74 A and a broad equatorial reflection at 8.9 A. Modeling studies suggested that tilted arrays of beta-strands constitute tubular, 30-A-diameter protofilaments, and that three to five of these protofilaments constitute the A beta fiber. This type of structure--a multimeric array of protofilaments organized as a tubular fibril--resembles that formed by the shorter A beta fragments (e.g., A beta(6-25), A beta(11-25), A beta(1-28)), suggesting a common structural motif in AD amyloid fibril organization.


Proteins | 2005

Polyglutamine homopolymers having 8–45 residues form slablike β‐crystallite assemblies

Deepak Sharma; Leonid M. Shinchuk; Hideyo Inouye; Ronald Wetzel; Daniel A. Kirschner

At least nine inherited neurodegenerative diseases, including Huntingtons, are caused by poly(L‐glutamine) (polyGln, polyQ) expansions > 35–40 repeats in widely or ubiquitously expressed proteins. Except for their expansions, these proteins have no sequence homologies, and their functions mostly remain unknown. Although each disease is characterized by a distinct pathology specific to a subset of neuronal cells, the formation of neuronal intranuclear aggregates containing protein with an expanded polyQ is the hallmark and common feature to most polyQ disorders. The neurodegeneration is thought to be caused by a toxic gain of function that occurs at the protein level and depends on the length of the expansion: Longer repeats cause earlier age of onset and more severe symptoms. To address whether there is a structural difference between polyQ having < 40 versus > 40 residues, we undertook an X‐ray fiber diffraction study of synthetic polyQ peptides having varying numbers of residues: Ac‐Q8‐NH2, D2Q15K2, K2Q28K2, and K2Q45K2. These particular lengths bracket both the range of normalcy (9–36 repeats) and the pathological (45 repeats), and therefore could be indicative of the structural changes expected in expanded polyQ domains. Contrary to expectations of different length‐dependent morphologies, we accounted for all the X‐ray patterns by slablike, β‐sheet structures, ∼20 Å thick in the β‐chain direction, all having similar monoclinic lattices. Moreover, the slab thickness indicates that K2Q45K2, rather than forming a water‐filled nanotube, must form multiple reverse turns. Proteins 2005.


Biophysical Journal | 1988

Membrane interactions in nerve myelin: II. Determination of surface charge from biochemical data

Hideyo Inouye; Daniel A. Kirschner

In our accompanying paper (Inouye and Kirschner, 1988) we calculated the surface charge density at the extracellular surfaces in peripheral and central nervous system (PNS; CNS) myelins from observations on the dependency of the width of the extracellular space on pH and ionic strength. Here, we have determined the surface charge density of the membrane surfaces in myelin from its chemical composition and the localization of some of its molecular components. We then analyzed the attractive and repulsive forces between the apposed surfaces and calculated equilibrium periods for comparison with the measured values. The biochemical model accounts for the observed isoelectric range of the myelin period and, with the surface charge reduced (possibly by divalent cation binding or a space charge approximation), the model also accounts for the dependency of period on pH above the isoelectric range. At the extracellular (and cytoplasmic) surfaces the contribution of lipid (with pI approximately 2) to the net surface charge is about the same in both PNS and CNS myelin, whereas the contribution of protein depends on which ones are exposed at the two surfaces. The protein conformation and localization modulate the surface charge of the lipid, resulting in positively-charged cytoplasmic surfaces (pI approximately 9) and negatively-charged extracellular surfaces (pI approximately 2-4). The net negative charge at the extracellular surface is due in CNS myelin to lipid, and in PNS myelin to both lipid and (PO) glycoprotein. The net positive charge at the cytoplasmic surface is due in CNS myelin mostly to basic protein, and in PNS myelin to PO glycoprotein and basic protein. The invariance of the cytoplasmic packing may be due to specific short-range interactions. Our models demonstrate how the particular myelin proteins and their localization and conformation can account for the differences in inter-membrane interactions in CNS and PNS myelins.


Journal of Neurochemistry | 1989

Myelin membrane structure and composition correlated: a phylogenetic study.

Daniel A. Kirschner; Hideyo Inouye; Allen L. Ganser; Vicki Mann

Abstract: We have correlated myelin membrane structure with biochemical composition in the CNS and PNS of a phylogenetic series of animals, including elasmobranchs, teleosts, amphibians, and mammals. X‐ray diffraction patterns were recorded from freshly dissected, unfixed tissue and used to determine the thicknesses of the lipid bilayer and the widths of the spaces between membranes at their cytoplasmic and extracellular appositions. The lipid and protein compositions of myelinated tissue from selected animals were determined by TLC and sodium dodecyl sulfate‐polyacrylamide gel elec‐trophoresis/immunoblotting, respectively. We found that(l) there were considerable differences in lipid (particularly gly‐colipid) composition, but no apparent phylogenetic trends; (2) the lipid composition did not seem to affect either the bilayer thickness, which was relatively constant, or the membrane separation; (3) the CNS of elasmobranch and teleost and the PNS of all four classes contained polypeptides that were recognized by antibodies against myelin Po glycoprotein; (4) antibodies against proteolipid protein (PLP) were recognized only by amphibian and mammalian CNS; (5) wide extracellular spaces (ranging from 36 to 48 Å) always correlated with the presence of Po‐immunoreactive protein; (6) the narrowest extracellular spaces (∼31 Å) were observed only in PLP‐containing myelin; (7) the cytoplasmic space in PLP‐containing myelin (∼31 Å) averaged ∼5Å less than that in Po‐containing myelin; (8) even narrower cytoplasmic spaces (∼24Å) were measured when both Po and 11–13‐kilodalton basic protein were detected; (9) proteins immu‐noreactive to antibodies against myelin P2 basic protein were present in elasmobranch and teleost CNS and/or PNS, and in mammalian PNS, but not in amphibian tissues; and (10) among mammalian PNS myelins, the major difference in structure was a variation in membrane separation at the cytoplasmic apposition. These findings demonstrate which features of myelin structure have remained constant and which have become specifically altered as myelin composition changed during evolutionary development.


Proceedings of the National Academy of Sciences of the United States of America | 2009

SCAP is required for timely and proper myelin membrane synthesis

Mark H. G. Verheijen; Nutabi Camargo; Karim Nadra; Anne-Sophie de Preux Charles; Jean-Jacques Médard; Adrienne M. Luoma; Michelle Crowther; Hideyo Inouye; Hitoshi Shimano; Su Chen; Jos F. Brouwers; J. Bernd Helms; M. Laura Feltri; Lawrence Wrabetz; Daniel A. Kirschner; Roman Chrast; August B. Smit

Myelination requires a massive increase in glial cell membrane synthesis. Here, we demonstrate that the acute phase of myelin lipid synthesis is regulated by sterol regulatory element-binding protein (SREBP) cleavage activation protein (SCAP), an activator of SREBPs. Deletion of SCAP in Schwann cells led to a loss of SREBP-mediated gene expression involving cholesterol and fatty acid synthesis. Schwann cell SCAP mutant mice show congenital hypomyelination and abnormal gait. Interestingly, aging SCAP mutant mice showed partial regain of function; they exhibited improved gait and produced small amounts of myelin indicating a slow SCAP-independent uptake of external lipids. Accordingly, extracellular lipoproteins partially rescued myelination by SCAP mutant Schwann cells. However, SCAP mutant myelin never reached normal thickness and had biophysical abnormalities concordant with abnormal lipid composition. These data demonstrate that SCAP-mediated regulation of glial lipogenesis is key to the proper synthesis of myelin membrane, and provide insight into abnormal Schwann cell function under conditions affecting lipid metabolism.


Biophysical Journal | 1988

Membrane interactions in nerve myelin. I. Determination of surface charge from effects of pH and ionic strength on period.

Hideyo Inouye; Daniel A. Kirschner

We have used x-ray diffraction to study the interactions between myelin membranes in the sciatic nerve (PNS) and optic nerve (CNS) as a function of pH (2-10) and ionic strength (0-0.18). The period of myelin was found to change in a systematic manner with pH and ionic strength. PNS periods ranged from 165 to 250 A or more, while CNS periods ranged from 150 to 230 A. The native periods were observed only near physiological ionic strength at neutral or alkaline pH. The smallest periods were observed in the pH range 2.5-4 for PNS myelin and pH 2.5-5 for CNS myelin. The minimum period was also observed for PNS myelin after prolonged incubation in distilled water. At pH 4, within these acidic pH ranges, myelin period increased slightly with ionic strength; however, above these ranges, the period increased with pH and decreased with ionic strength. Electron density profiles calculated at different pH and ionic strength showed that the major structural alteration underlying the changes in period was in the width of the aqueous space at the extracellular apposition of membranes; the width of the cytoplasmic space was virtually constant. Assuming that the equilibrium myelin periods are determined by a balance of nonspecific forces/i.e., the electrostatic repulsion force and the van der Walls attractive force, as well as the short-range repulsion force (hydration force, or steric stabilization), then values in the period-dependency curve can be used to define the isoelectric pH and exclusion length of the membrane. The exclusion length, which is related to the minimum period at isoelectric pH, was used to calculate the electrostatic repulsion force given the other forces. The electrostatic repulsion was then used to calculate the surface potential, which in turn was used to calculate the surface charge density (at different pH and ionic strength). We found the negative surface charge increases with pH at constant ionic strength and with ionic strength at constant pH. We suggest that the former is due to deprotonation of the ionizable groups on the surface while the latter is due to ion binding. Interpretation of our data in terms of the chemical composition of myelin is given in the accompanying paper (Inouye and Kirschner, 1988). We also calculated the total potential energy functions for the different equilibrium periods and found that the energy minima became shallower and broader with increasing membrane separation. Finally, it was difficult to account directly for certain structural transitions from a balance of nonspecific forces. Such transitions included the abrupt appearance of the native period at alkaline pH and physiological ionic strength and the discontinuous compaction after prolonged treatment in distilled water. Possibly, in PNS myelin conformational modification of PO glycoprotein occurs under these conditions. The invariance of the cytoplasmic space suggests the presence of specific short-range interactions between surfaces at this apposition.


Proteins | 2005

Poly‐(L‐alanine) expansions form core β‐sheets that nucleate amyloid assembly

Leonid M. Shinchuk; Deepak Sharma; Sylvie E. Blondelle; Natàlia Reixach; Hideyo Inouye; Daniel A. Kirschner

Expansion to a total of 11–17 sequential alanine residues from the normal number of 10 in the polyadenine‐binding protein nuclear‐1 (PABPN1) results in formation of intranuclear, fibrillar inclusions in skeletal muscle and hypothalamic neurons in adult‐onset, dominantly inherited oculopharyngeal muscular dystrophy (OPMD). To understand the role that homopolymeric length may play in the protein misfolding that leads to the inclusions, we analyzed the self‐assembly of synthetic poly‐(L‐alanine) peptides having 3–20 residues. We found that the conformational transition and structure of polyalanine (polyAla) assemblies in solution are not only length‐dependent but also are determined by concentration, temperature, and incubation time. No β‐sheet complex was detected for those peptides characterized by n < 8, where n is number of alanine residues. A second group of peptides with 7 < n < 15 showed varying levels of complex formation, while for those peptides having n > 15, the interconversion process from the monomeric to the β‐sheet complex was complete under any of the tested experimental conditions. Unlike the typical tinctorial properties of amyloid fibrils, polyalanine fibrils did not show fluorescence with thioflavin T or apple‐green birefringence with Congo red; however, like amyloid, X‐ray diffraction showed that the peptide chains in these fibrils were oriented normal to the fibril axis (i.e., in the cross‐β arrangement). Neighboring β‐sheets are quarter‐staggered in the hydrogen‐bonding direction such that the alanine side‐chains were closely packed in the intersheet space. Strong van der Waals contacts between side‐chains in this arrangement likely account for the high stability of the macromolecular fibrillar complex in solution over a wide range of temperature (5–85°C), and pH (2–10.5), and its resistance to denaturant (< 8 M urea) and to proteases (protease K, trypsin). We postulate that a similar stabilization of an expanded polyalanine stretch could form a core β‐sheet structure that mediates the intermolecular association of mutant proteins into fibrillar inclusions in human pathologies. Proteins 2005.


Biophysical Journal | 1989

Membrane structure in isolated and intact myelins

Hideyo Inouye; J. Karthigasan; Daniel A. Kirschner

The biochemical composition of myelin and the topology of its constituent lipids and proteins are typically studied using membranes that have been isolated from whole, intact tissue using procedures involving hypotonic shock and sucrose density gradient centrifugation. To what extent, however, are the structure and intermembrane interactions of isolated myelin similar to those of intact myelin? We have previously reported that intact and isolated myelins do not always show identical myelin periods, indicating a difference in membrane-membrane interactions. The present study addresses the possibility that this is due to altered membrane structure. Because x-ray scattering from isolated myelin sometimes consists of overlapping Bragg reflections or is continuous, we developed nonlinear least squares procedures for analyzing the total intensity distribution after film scaling, background subtraction, and Lorentz correction. We calculated electron density profiles of isolated myelin for comparison with membrane profiles from intact myelin. The change in the width of the extracellular space and the relative invariance of the cytoplasmic space as a function of pH and ionic strength that we previously found for intact nerve was largely paralleled by isolated myelin. There were two exceptions: isolated CNS myelin was resistant to swelling under all conditions, and isolated PNS myelin in hypotonic saline showed indefinite swelling at the extracellular apposition. However, electron density profiles of isolated myelins, calculated to 30 A resolution, did not show any major change in structure compared with intact myelin that could account for the differences in interactions.


Journal of Neuropathology and Experimental Neurology | 2005

Structure and Stability of Internodal Myelin in Mouse Models of Hereditary Neuropathy

Robin L. Avila; Hideyo Inouye; Rena C. Baek; Xinghua Yin; Bruce D. Trapp; M. Laura Feltri; Lawrence Wrabetz; Daniel A. Kirschner

Peripheral neuropathies often result in abnormalities in the structure of internodal myelin, including changes in period and membrane packing, as observed by electron microscopy (EM). Mutations in the gene that encodes the major adhesive structural protein of internodal myelin in the peripheral nervous system of humans and mice-P0 glycoprotein-correlate with these defects. The mechanisms by which P0 mutations interfere with myelin packing and stability are not well understood and cannot be provided by EM studies that give static and qualitative information on fixed material. To gain insights into the pathogenesis of mutant P0, we used x-ray diffraction, which can detect more subtle and dynamic changes in native myelin, to investigate myelin structure in sciatic nerves from murine models of hereditary neuropathies. We used mice with disruption of one or both copies of the P0 gene (models of Charcot-Marie-Tooth-like neuropathy [CMT1B] or Dejerine-Sottas-like neuropathy) and mice with a CMT1B resulting from a transgene encoding P0 with an amino terminal myc-tag. To directly test the structural role of P0, we also examined a mouse that expresses P0 instead of proteolipid protein in central nervous system myelin. To link our findings on unfixed nerves with EM results, we analyzed x-ray patterns from unembedded, aldehyde-fixed nerves and from plastic-embedded nerves. From the x-ray patterns recorded from whole nerves, we assessed the amount of myelin and its quality (i.e. relative thickness and regularity). Among sciatic nerves having different levels of P0, we found that unfixed nerves and, to a lesser extent, fixed but unembedded nerves gave diffraction patterns of sufficient quality to distinguish periods, sometimes differing by a few Å. Certain packing abnormalities were preserved qualitatively by aldehyde fixation, and the relative amount and structural integrity of myelin among nerves could be distinguished. Measurements from the same nerve over time showed that the amount of P0 affected myelins stability against swelling, thus directly supporting the hypothesis that packing defects underlie instability in “live” or intact myelin. Our findings demonstrate that diffraction can provide a quantitative basis for understanding, at a molecular level, the membrane packing defects that occur in internodal myelin in demyelinating peripheral neuropathies.

Collaboration


Dive into the Hideyo Inouye's collaboration.

Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Lee Makowski

Northeastern University

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Dennis J. Selkoe

Brigham and Women's Hospital

View shared research outputs
Top Co-Authors

Avatar

Lawrence K. Duffy

University of Alaska Fairbanks

View shared research outputs
Top Co-Authors

Avatar
Researchain Logo
Decentralizing Knowledge