Network


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

Hotspot


Dive into the research topics where Natsuhiko Sugimura is active.

Publication


Featured researches published by Natsuhiko Sugimura.


Journal of Mass Spectrometry | 2016

Differentiation of AB-FUBINACA positional isomers by the abundance of product ions using electron ionization-triple quadrupole mass spectrometry.

Takaya Murakami; Yoshiaki Iwamuro; Reiko Ishimaru; Satoshi Chinaka; Natsuhiko Sugimura; Nariaki Takayama

Mass spectrometric differentiation of structural isomers is important for the analysis of forensic samples. Presently, there is no mass spectrometric method for differentiating halogen positional isomers of cannabimimetic compounds. We describe here a novel and practical method for differentiating one of these compounds, N-(1-amino-3-methyl-1-oxobutan-2-yl)-1-(4-fluorobenzyl)-1H-indazole-3-carboxamide (AB-FUBINACA (para)), and its fluoro positional (ortho and meta) isomers in the phenyl ring by electron ionization-triple quadrupole mass spectrometry. It was found that the three isomers differed in the relative abundance of the ion at m/z 109 and 253 in the product ion spectra, while the detected product ions were identical. The logarithmic values of the abundance ratio of the ions at m/z 109 to 253 (ln(A109 /A253 )) were in the order meta < ortho < para and increased linearly with collision energy. The differences in abundances were attributed to differences in the dissociation reactivity between the indazole moiety and the fluorobenzyl group because of the halogen-positional effect on the phenyl ring. Our methodology, which is based on the abundance of the product ions in mass spectra, should be applicable to determination of the structures of other newly encountered designer drugs. Copyright


Journal of Physical Chemistry B | 2015

Photochromic Solid Materials Based on Poly(decylviologen) Complexed with Alginate and Poly(sodium 4-styrenesulfonate)

Luis Sanhueza; Joaquín Castro; Esteban Urzúa; Lorena Barrientos; Felipe Oyarzun-Ampuero; Héctor Pesenti; Toshimichi Shibue; Natsuhiko Sugimura; Wataru Tomita; Hiroyuki Nishide; Ignacio Moreno-Villoslada

Photochromic solid materials based on the cationic polymer poly(decylviologen) are reported. The solids were obtained by freeze-drying colloidal suspensions of nanocomplexes obtained by mixing aqueous solutions of the polycation with different solutions of polyanions such as poly(sodium 4-styrenesulfonate) or sodium alginate, at a cationic/anionic polymeric charge ratio of 0.7. The photochromic responses of the solid materials fabricated with alginate as complementary charged polyelectrolyte of the cationic polyviologen are faster than those of the solid materials fabricated with poly(sodium 4-styrenesulfonate), achieving coloration kinetics in the order of minutes, and discoloration kinetics in the order of hours for the former. Aromatic-aromatic interactions between the latter polyanion and the polyviologen may stabilize the dicationic form of the viologen derivative, increasing the necessary energy to undergo photoreduction, thus decreasing the reduction kinetics.


European Journal of Mass Spectrometry | 2015

Comparison of the applicability of mass spectrometer ion sources using a polarity-molecular weight scattergram with a 600 sample in-house chemical library

Natsuhiko Sugimura; Asami Furuya; Takahiro Yatsu; Toshimichi Shibue

To provide a practical guideline for the selection of a mass spectrometer ion source, we compared the applicability of three types of ion source: direct analysis in real time (DART), electrospray ionization (ESI) and fast atom bombardment (FAB), using an in-house high-resolution mass spectrometry sample library consisting of approximately 600 compounds. The great majority of the compounds (92%), whose molecular weights (MWs) were broadly distributed between 150 and 1000, were detected using all the ion sources. Nevertheless, some compounds were not detected using specific ion sources. The use of FAB resulted in the highest sample detection rate (>98%), whereas the detection rates obtained using DART and ESI were slightly lower (>96%). A scattergram constructed using MW and topological polar surface area (tPSA) as a substitute for molecular polarity showed that the performance of ESI was weak in the low-MW (<400), low-polarity (tPSA < 60) area, whereas the performance of DART was weak in the high-MW (>800) area. These results might provide guidelines for the selection of ion sources for inexperienced mass spectrometry users.


Journal of Physical Chemistry B | 2017

Water-Induced Phase Transition in Cyclohexane/n-Hexanol/Triton X-100 Mixtures at a Molar Composition of 1/16/74 Studied by NMR

Mario E. Flores; Francisco Martínez; Andrés F. Olea; Toshimichi Shibue; Natsuhiko Sugimura; Hiroyuki Nishide; Ignacio Moreno-Villoslada

Molecular aggregation in a mixture of Triton X-100/n-hexanol/cyclohexane at a molar ratio of 1/16/74 is studied upon addition of small amounts of water. The composition of organic components has been chosen at a ratio n-hexanol/cyclohexane where a well-formed hydrogen bond network has been described. The ratio Triton X-100/n-hexanol has been chosen to afford a stoichiometry of ethylene oxide (EO) residues/n-hexanol of 1/2. At these conditions the addition of water consecutively produces the appearance of three defined phases: a clear solution, a lamellar phase, and a microemulsion. The two corresponding transitions occur at water/EO/n-hexanol molar ratios of 2/1/2 (clear to lamella), and 3/1/2 (lamella to microemulsion), while phase separation occurs at a molar ratio of 4/1/2, highlighting the important role of stoichiometry. Molecular dynamics measured by 1H NMR techniques, such as DOSY, and calculations of T1 and T2 relaxation times allow distinguishing the transition between the different phases and justifying their structure. Molecular assembly in the three phases is organized around hydrogen bond networks in which the hydroxyl groups of both TX-100 and n-hexanol, ethylene oxide groups of TX-100, and water participate. 1D 1H NMR spectral changes correlate with the different characteristics of the different phases. As the main characteristics of the lamellar phase we find a very restricted mobility of the molecules involved, and water chemical shifts in 1D 1H NMR spectra of around 5.0 ppm, higher than that of bulk water appearing at 4.7 ppm.


Journal of Physical Chemistry B | 2015

Stability of Water/Poly(ethylene oxide)43-b-poly(ε-caprolactone)14/Cyclohexanone Emulsions Involves Water Exchange between the Core and the Bulk

Mario E. Flores; Francisco Martínez; Andrés F. Olea; Toshimichi Shibue; Natsuhiko Sugimura; Hiroyuki Nishide; Ignacio Moreno-Villoslada

The formation of emulsions upon reverse self-association of the monodisperse amphiphilic block copolymer poly(ethylene oxide)43-b-poly(ε-caprolactone)14 in cyclohexanone is reported. Such emulsions are not formed in toluene, chloroform, or dichloromethane. We demonstrate by magnetic resonance spectroscopy the active role of the solvent on the stabilization of the emulsions. Cyclohexanone shows high affinity for both blocks, as predicted by the Hansen solubility parameters, so that the copolymer chains are fully dissolved as monomeric chains. In addition, the solvent is able to produce hydrogen bonding with water molecules. Water undergoes molecular exchange between water molecules associated with the polymer and water molecules associated with the solvent, dynamics of major importance for the stabilization of the emulsions. Association of polymeric chains forming reverse aggregates is induced by water over a concentration threshold of 5 wt %. Reverse copolymer aggregates show submicron average hydrodynamic diameters, as seen by dynamic light scattering, depending on the polymer and water concentration.


European Journal of Mass Spectrometry | 2015

Prediction of adducts on positive mode electrospray ionization mass spectrometry: proton/sodium selectivity in methanol solutions.

Natsuhiko Sugimura; Asami Furuya; Takahiro Yatsu; Toshimichi Shibue

We used positive mode electrospray ionization (ESI) mass spectrometry to examine 540 in-house high-resolution mass spectrometry (HRMS) samples that formed an adducted positive ion. Of the 540 samples, the sodium adduct ([M + Na]+) was detected in 480 samples, and the protonated molecule ([M + H]+) was detected in 92 samples; both [M + Na]+ and [M + H]+ were detected in 32 samples. No other adduct ions were predominant. The selectivities of these adducts were evaluated by a two-dimensional plot using topological polar surface area (tPSA) and molecular weight. Two predominant trends were observed: [M + H]+ converged around tPSA (Å2) = 20 and molecular weight = 250, and the selectivity for [M + Na]+ correlated with the tPSA value. These observations were found to be related to the elemental composition of the sample compounds. From the results obtained by positive mode ESI mass spectroscopy under our experimental conditions, predominant trends were observed with respect to adduct selectivity: compounds containing oxygen atom(s) form [M + Na]+, and compounds containing nitrogen but not oxygen atom(s) form [M + H]+. Based on these trends, we developed the “Nitrogen–Oxygen rule” (NO rule) to predict the adduct formed by a given compound on positive mode ESI. This NO rule provides a guideline to estimate elemental composition using ESI-HRMS with methanol as mobile phase.


Artificial Cells Nanomedicine and Biotechnology | 2013

Swine hemoglobin as a potential source of artificial oxygen carriers, hemoglobin-vesicles

Hiromi Sakai; Kiayi Ng; Bing Li; Natsuhiko Sugimura

Abstract Hemoglobin-based oxygen carriers (HBOCs) have been developed as a transfusion alternative and oxygen therapy. The Hb source is usually outdated donated human blood or cow blood obtained from cattle industries because of its abundance. This study examined the feasibility of using swine Hb (SHb) for preparation of cellular-type HBOCs, hemoglobin-vesicles (HbV). Purification of SHb from fresh swine whole blood was conducted with processes including carbonylation (SHbO2 ◊ SHbCO), pasteurization (60 °C, 15 hours) and tangential flow ultrafiltration, with yield of 90%. Actually, differential scanning calorimetric analysis showed a denaturation temperature of SHbCO at 83 °C and assures its stability during pasteurization. Concentrated SHbCO together with pyridoxal 5’-phosphate (PLP) as an allosteric effector was encapsulated in phospholipid vesicles to prepare SHbV. After decarbonylation (SHbCO ◊ SHbO2), the oxygen affinity (P50) of SHb changes mainly by PLP, and the influence of Cl‐ was small, in a manner similar to that of human Hb (HHb). However, after encapsulation, vesicles of SHbV showed much lower oxygen affinity (higher P50) than HHbV did. Autoxidation of SHbV was slightly faster than HHbV. Although some differences are apparent in oxygen affinity and autoxidation rates, results clarified that SHb is useful as a starting material for HbV production.


Journal of Physical Chemistry Letters | 2017

Structural Insights into a Hemoglobin–Albumin Cluster in Aqueous Medium

Ryuichi Shinohara; Taiga Yamada; Boris Schade; Christoph Böttcher; Takaaki Sato; Natsuhiko Sugimura; Toshimichi Shibue; Teruyuki Komatsu

A hemoglobin (Hb) wrapped covalently by three human serum albumins (HSAs) is a triangular protein cluster designed as an artificial O2-carrier and red blood cell substitute. We report the structural insights into this Hb-HSA3 cluster in aqueous medium revealed by 3D reconstruction based on cryogenic transmission electron microscopy (cryo-TEM) data and small-angle X-ray scattering (SAXS) measurements. Cryo-TEM observations showed individual particles with approximately 15 nm diameter in the vitrified ice layer. Subsequent image processing and 3D reconstruction proved the expected spatial arrangements of an Hb in the center and three HSAs at the periphery. SAXS measurements demonstrated the monodispersity of the Hb-HSA3 cluster having a molecular mass of 270 kDa. The pair-distance distribution function suggested the existence of oblate-like particles with a maximum dimeter of ∼17 nm. The supramolecular 3D structure reconstructed from the SAXS intensity using an ab initio procedure was similar to that obtained from cryo-TEM data.


Journal of Clinical Biochemistry and Nutrition | 2017

Stabilizers of edaravone aqueous solution and their action mechanisms. 1. Sodium bisulfite

Masahiko Tanaka; Natsuhiko Sugimura; Akio Fujisawa; Yorihiro Yamamoto

Edaravone (3-methyl-1-phenyl-2-pyrazolin-5-one) has been used as a free radical scavenging drug for the treatment of acute ischemic stroke in Japan since 2001. Edaravone is given to patients intravenously; therefore, it is distributed in the form of an aqueous solution. However, aqueous solutions of edaravone are very unstable because it is present as edaravone anion, which is capable of transferring an electron to free radicals including oxygen, and becomes edaravone radical. We observed the formation of hydrogen peroxide and edaravone trimer when aqueous edaravone solution was kept at 60°C for 4 weeks. We proposed the mechanism of edaravone trimer formation from edaravone radicals. Lowering the pH and deoxygenation can effectively increase the stability of aqueous edaravone solution, since the former reduces edaravone anion concentration and the latter inhibits edaravone radical formation. Addition of sodium bisulfite partially stabilized aqueous edaravone solutions and partially inhibited the formation of edaravone trimer. Formation of bisulfite adduct was suggested by 13C NMR and HPLC studies. Therefore, the stabilizing effect of sodium bisulfite is ascribed to the formation of a bisulfite adduct of edaravone and, consequently, reduction in the concentration of edaravone anion.


European Journal of Mass Spectrometry | 2017

Observed adducts on positive mode direct analysis in real time mass spectrometry – Proton/ammonium adduct selectivities of 600-sample in-house chemical library:

Natsuhiko Sugimura; Asami Furuya; Takahiro Yatsu; Yoko Igarashi; Reiko Aoyama; Chisato Izutani; Yorihiro Yamamoto; Toshimichi Shibue

In this study, direct analysis in real time adduct selectivities of a 558 in-house high-resolution mass spectrometry sample library was evaluated. The protonated molecular ion ([M + H]+) was detected in 462 samples. The ammonium adduct ion ([M + NH4]+) was also detected in 262 samples. [M + H]+ and [M + NH4]+ molecular ions were observed simultaneously in 166 samples. These adduct selectivities were related to the elemental compositions of the sample compounds. [M + NH4]+ selectivity correlated with the number of oxygen atom(s), whereas [M + H]+ selectivity correlated with the number of nitrogen atom(s) in the elemental compositions. For compounds including a nitrogen atom and an oxygen atom [M + H]+ was detected; [M + NH4]+ was detected for compounds including an oxygen atom only. Density functional theory calculations were performed for selected library samples and model compounds. Energy differences were observed between compounds detected as [M + H]+ and [M + NH4]+, and between compounds including a nitrogen atom and an oxygen atom in their elemental compositions. The results suggested that the presence of oxygen atoms stabilizes [M + NH4]+, but not every oxygen atom has enough energy for detection of [M + NH4]+. It was concluded that the nitrogen atom(s) and oxygen atom(s) in the elemental compositions play important roles in the adduct formation in direct analysis in real time mass spectrometry.

Collaboration


Dive into the Natsuhiko Sugimura's collaboration.

Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Mario E. Flores

Austral University of Chile

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Researchain Logo
Decentralizing Knowledge