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 31 毫秒
1.
Elemental mapping in the energy filtering transmission electron microscope (EFTEM) can be extended into three dimensions (3D) by acquiring a series of two‐dimensional (2D) core‐edge images from a specimen oriented over a range of tilt angles, and then reconstructing the volume using tomographic methods. EFTEM has been applied to imaging the distribution of biological molecules in 2D, e.g. nucleic acid and protein, in sections of plastic‐embedded cells, but no systematic study has been undertaken to assess the extent to which beam damage limits the available information in 3D. To address this question, 2D elemental maps of phosphorus and nitrogen were acquired from unstained sections of plastic‐embedded isolated mouse thymocytes. The variation in elemental composition, residual specimen mass and changes in the specimen morphology were measured as a function of electron dose. Whereas 40% of the total specimen mass was lost at doses above 106 e?/nm2, no significant loss of phosphorus or nitrogen was observed for doses as high as 108 e?/nm2. The oxygen content decreased from 25 ± 2 to 9 ± 2 atomic percent at an electron dose of 104 e?/nm2, which accounted for a major component of the total mass loss. The specimen thickness decreased by 50% after a dose of 108 e?/nm2, and a lateral shrinkage of 9.5 ± 2.0% occurred from 2 × 104 to 108 e?/nm2. At doses above 107 e?/nm2, damage could be observed in the bright field as well in the core edge images, which is attributed to further loss of oxygen and carbon atoms. Despite these artefacts, electron tomograms obtained from high‐pressure frozen and freeze‐substituted sections of C. elegans showed that it is feasible to obtain useful 3D phosphorus and nitrogen maps, and thus to reveal quantitative information about the subcellular distributions of nucleic acids and proteins.  相似文献   

2.
Colloidal particle size is an important characteristic to consider when choosing a radiopharmaceutical for diagnosis and therapeutic purposes in nuclear medicine. Photon correlation spectroscopy (PCS) and transmission electron microscopy (TEM) were used to determine the particle‐size distribution of 90Y‐ and 99mTc‐labelled antimony trisulfide (Sb2S3) and tin colloids (Sn‐colloid). 90Y‐Sb2S3 and 99mTc‐Sb2S3 were found to have a diameter of 28.92 ± 0.14 and 35.61 ± 0.11 nm, respectively, by PCS. By TEM, 90Y‐Sb2S3 particles were measured to be 14.33 ± 0.09 nm. 90Y‐labelled Sn colloid were found to exist with a dv(max1) of 805 nm and a dv(max2) of 2590 nm, by PCS, whereas 99mTc‐Sn colloid was shown to have more than 80% of radioactive particles of approximately 910 nm by PCS. For 90Y‐labelled Sb2S3 and Sn colloid, a comparison of TEM and PCS indicates that these techniques found significantly different mean diameters. TEM has an excellent resolution necessary for radiocolloid particle‐sizing analysis, and it is a desirable size‐measuring technique because it is more reliable than PCS.  相似文献   

3.
Mach–Zehnder interferometers have attracted attention due to their flexible structures and extensive applications. In this study, a simple and highly sensitive curvature sensor based on an asymmetrical Mach–Zehnder interferometer comprising an up-taper and a peanut-like section was theoretically and experimentally characterized. A simulation was conducted to investigate the influence of structural parameters on the interference spectrum. Experimental measurements show that a high curvature sensitivity of ?35.3?nm/m?1 from 1.35 to 2.20?m?1 was obtained and high mechanical strength was demonstrated. The temperature–curvature cross-sensitivity of 4.44?×?10?4?m?1/°C was lower than the curvature resolution. The high curvature sensitivity and low temperature cross-sensitivity make this sensor a candidate for practical applications.  相似文献   

4.
HgTe/Cd0.735Hg0.265Te nanostructures with HgTe quantum wells 16.2 and 21.0 nm thick are grown without additional doping on (013)CdTe/ZnTe/GaAs substrates by the method of molecular beam epitaxy. The compositions and thicknesses of the wide-gap layer and quantum well in the course of growth are performed by means of ellipsometry. The accuracy is Δx ? ±0.002 mole fractions of cadmium telluride in determining the composition and Δd ? 0.5 nm in determining the thickness of the wide-gap layer and quantum well. The central fragments of the wide-gap layers ≈ 10 nm thick are additionally doped by indium for a ~ 1015 cm?3 volume concentration of charge carriers to be reached. Galvanomagnetic research in a wide range of magnetic field intensities at liquid helium temperatures reveals dimensional quantization levels and the presence of a two-dimensional electron gas in grown nanostructures. High mobility of the two-dimensional electron gas μ e is obtained: 2 · 105 and 5 · 105 cm2/V · s for electron densities N s equal to 1.5 · 1011 and 3.5 · 1011 cm?2, respectively.  相似文献   

5.
A new method of interferometry employing a Fabry–Perot etalon model was used to locate and size microscale features on the surface of the pine bark beetle. Oscillations in the reflected light spectrum, caused by self‐interference of light reflecting from surfaces of foreleg setae and spores on the elytrum, were recorded using white light hyperspectral microscopy. By making the assumption that pairs of reflecting surfaces produce an etalon effect, the distance between surfaces could be determined from the oscillation frequency. Low frequencies of less than 0.08 nm?1 were observed in the spectrum below 700 nm while higher frequencies generally occupied wavelengths from 600 to 850 nm. In many cases, two frequencies appeared separately or in combination across the spectrum. The etalon model gave a mean spore size of 3.04 ± 1.27 μm and a seta diameter of 5.44 ± 2.88 μm. The tapering near the setae tip was detected as a lowering of frequency. Spatial fringes were observed together with spectral oscillations from surfaces on the exoskeleton at higher magnification. These signals were consistent with embedded multi‐layer reflecting surfaces. Possible applications for hyperspectral interferometry include medical imaging, detection of spore loads in insects and other fungal carriers, wafer surface and subsurface inspection, nanoscale materials, biological surface analysis, and spectroscopy calibration. This is, to our knowledge, the first report of oscillations directly observed by microscopy in the reflected light spectra from Coleoptera, and the first demonstration of broadband hyperspectral interferometry using microscopy that does not employ an internal interferometer. Microsc. Res. Tech. 78:873–885, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

6.
We investigate the ultrastructural features and 3D electron tomography of chameleon (Chamaeleon calyptratus) which is a native of desert environments of Saudi Arabia. The corneas of the chameleon were fixed in 2.5% glutaraldehyde containing cuprolinic blue in sodium acetate buffer for electron microscopy and tomography, and observed under a JEOL 1400 transmission electron microscope. The thin cornea (21.92 μm) contained 28–30 collagen fibril lamellae. The middle stromal lamellae (from 13 to 19) contained keratocytes with a long cell process and filled with granular material. The CF diameter increased from lamella 1 (30.44 ± 1.03) to Lamella 5 (52.83 ± 2.00) then decreased towards the posterior stoma. The percentage of large CF diameters (55–65 nm) was very high in the lamellae L14 (38.8%) and L15 (85.7%). The mean PGs area of the posterior stroma (448.21 ± 24.84 nm2) was significantly larger than the mean PGs area of the anterior, (309.86 ± 8.2 nm2) and middle stroma 245.94 ± 8.28 nm2). 3D electron tomography showed the distribution of PGs around and over the CF. Variable diameters of CFs in the anterior stroma may provide compact lamellae which may restrict the low wavelength of light. Variable diameters of CFs in the anterior stroma may provide compact lamellae which may restrict the low wavelength of light. This accommodation function is achieved by bending of the cornea. During bending the anterior stroma was stretched and the posterior stroma was compressed due to the presence of small CFs. The middle stroma remained stiff due to the presence of large CFs. Large proteoglycans not only maintain hydration for a longer period of time, but also act as a lubricant to neutralise the shear forces in the anterior and posterior stroma during bending.  相似文献   

7.
Morphometric characterization of nanoparticles is crucial to determine their biological effects and to obtain a formulation pattern. Determining the best technique requires knowledge of the particles being analyzed, the intended application of the particles, and the limitations of the techniques being considered. The aim of this article was to present transmission (TEM) and scanning (SEM) electron microscopy protocols for the analysis of two different nanostructures, namely polymeric nanoemulsion and poly(lactic‐co‐glycolic acid) (PLGA) nanoparticles, and to compare these results with conventional dynamic light scattering (DLS) measurements. The mean hydrodynamic diameter, the polydispersity index, and zeta potential of the nanostructures of polymeric nanoemulsion were 370.5 ± 0.8 nm, 0.133 ± 0.01, and ?36.1 ± 0.15 mV, respectively, and for PLGA nanoparticles were 246.79 ± 5.03 nm, 0.096 ± 0.025, and ?4.94 ± 0.86 mV, respectively. TEM analysis of polymeric nanoemulsion revealed a mean diameter of 374 ± 117 nm. SEM analysis showed a mean diameter of 368 ± 69 nm prior to gold coating and 448 ± 70 nm after gold coating. PLGA nanoparticles had a diameter of 131 ± 41.18 nm in TEM and 193 ± 101 nm in SEM. Morphologically, in TEM analysis, the polymeric nanoemulsions were spherical, with variable electron density, very few showing an electron‐dense core and others an electron‐dense surface. PLGA nanoparticles were round, with an electron‐lucent core and electron‐dense surface. In SEM, polymeric nanoemulsions were also spherical with a rough surface, and PLGA nanoparticles were round with a smooth surface. The results show that the “gold standards” for morphometric characterization of polymeric nanoemulsion and PLGA nanoparticles were, respectively, SEM without gold coating and TEM with negative staining. Microsc. Res. Tech. 77:691–696, 2014. © 2014 Wiley Periodicals, Inc.  相似文献   

8.
Adulteration is the root cause of producing not only a chemically and pharmacologically inferior but also in some instances hazardous or poisonous drug. Despite availability of several techniques, microscopy and physicochemical analyses are the most practical approaches for crude drug authentication. Hence, the present study aimed to evaluate morphological, microscopic, and physicochemical properties of root, bark, leaf, and fruit of Diospyros montana Morphological properties were determined by sensory organs, whereas microscopic features of cross‐sections and powders were determined by light and scanning electron microscopy. The proximate and fluorescence analyses were performed using the standard guidelines. The physical examination of fresh, shade‐dried, and powdered material showed no significant change in color. The identifying cellular structures included cuboidal cork, pitted tracheids, scalariform, reticulate and spiral xylary vessels, and rosettes, raphide, and cuboidal calcium oxalate crystals. The stomatal number, stomatal index, vein‐islet and vein‐termination number, and palisade ratio in the leaf were found to be 293.91 ± 32.68 mm?2, 64.18 ± 3.42%, 22.00 ± 3.81 mm?2 and 38.40 ± 5.81 mm?2, and 3.85 ± 0.60, respectively. Total ash, acid insoluble ash, water soluble ash and sulfated ash of leaf (9.00 ± 0.50%, 1.67 ± 0.23%, 2.00 ± 0.22% and 14.50 ± 0.99%, respectively), foaming index of bark and root (111.11 ± 2.11), and swelling index of fruit (19.00 ± 3.45) were higher than the other parts. The powder of different parts showed characteristic colors in the daylight and UV light upon treatment with various regents. The plant was found to be rich in saponins, fibers, and flavonoids. The results of the present study may serve as identifiers of different parts of Diospyros montana.  相似文献   

9.
The present article reviews the mass spectrometric fragmentation processes and fragmentation energetics of leucine enkephalin, a commonly used peptide, which has been studied in detail and has often been used as a standard or reference compound to test novel instrumentation, new methodologies, or to tune instruments. The main purpose of the article is to facilitate its use as a reference material; therefore, all available mass spectrometry‐related information on leucine enkephalin has been critically reviewed and summarized. The fragmentation mechanism of leucine enkephalin is typical for a small peptide; but is understood far better than that of most other compounds. Because ion ratios in the MS/MS spectra indicate the degree of excitation, leucine enkephalin is often used as a thermometer molecule in electrospray or matrix‐assisted laser desorption ionization (ESI or MALDI). Other parameters described for leucine enkephalin include collisional cross‐section and energy transfer; proton affinity and gas‐phase basicity; radiative cooling rate; and vibrational frequencies. The lowest‐energy fragmentation channel of leucine enkephalin is the MH+b4 process. All available data for this process have been re‐evaluated. It was found that, although the published Ea values were significantly different, the corresponding Gibbs free energy change showed good agreement (1.32 ± 0.07 eV) in various studies. Temperature‐ and energy‐dependent rate constants were re‐evaluated with an Arrhenius plot. The plot showed good linear correlation among all data (R2 = 0.97), spanned over a 9 orders of magnitude range in the rate constants and yielded 1.14 eV activation energy and 1011.0 sec?1 pre‐exponential factor. Accuracy (including random and systematic errors, with a 95% confidence interval) is ±0.05 eV and 10±0.5 sec?1, respectively. The activation entropy at 470 K that corresponds to this reaction is ?38.1 ± 9.6 J mol?1 K?1. We believe that these re‐evaluated values are by far the most accurate activation parameters available at present for a protonated peptide and can be considered as “consensus” values; results on other processes might be compared to this reference value. © 2010 Wiley Periodicals, Inc., Mass Spec Rev 30:298–320, 2011  相似文献   

10.
The aim of this study was to evaluate the toxicity of titanium dioxide nanoparticles (TiO2NPs) by short‐term toxicity tests in Apis mellifera, considered an excellent bioindicator organism mainly due to its sensitivity. Bees have been exposed to several concentrations of TiO2NPs (1 × 10?3, 1 × 10?4, 1 × 10?5, 1 × 10?6 mg/10 ml) for 10 days. Morphostructural and histological assays were done on gut and honey sac. The research of exposure biomarkers like metallothioneins 1 (MT1) and Heat Shock Protein 70 (HSP70) was performed to verify if a detoxification mechanism has been activated in the exposed animals. No histological alteration on the epithelium of the gut and honey sac were observed in exposed samples. A significant positivity for anti‐MT1 antibody was observed only in the honey sac cells. A weak positivity for HSP70 was observed in both structures analyzed. In several studies have shown the non‐toxicity of TiO2NPs on other model organisms, in our study, titanium dioxide nanoparticles was proven to be highly toxic at the highest concentration tested (100% of lethality to 1 × 10?3 mg/10 ml) and moderately toxic at lower concentrations. Honey bees proved to be excellent models for study of NPs toxicity and for monitoring environment.  相似文献   

11.
AIMS: To evaluate measurements of collagen fibril spacing using different shaped regions of interest (ROI) on transmission electron micrograph (TEM) images of rabbit corneal stroma. METHODS: Following glutaraldehyde fixation and phosphotungstic acid staining, TEM images of collagen fibrils in cross section were projected at a final magnification close to 250,000 × to obtain overlays. Interfibril distances (IFDs; center‐to‐center spacing) were measured within different ROIs of the same nominal area (0.25 μm2) but different shape (with the length to width, L:W, ratio from 1:1 to 6:1). The IFD distribution was analyzed, and the 2D organization assessed using a radial distribution analysis. RESULTS: The fibrils had an average diameter of 35.3 ± 3.8 (SD) nm, packing density of 393 ± 4 fibrils / μm2 and a fibril volume fraction of 0.39 ± 0.02. IFDs ranged from 29 to 1400 nm depending on the shape of the ROI, with average values ranging from 263 to 443 nm. By artificially selecting IFD data only to a radial distance of 250 nm, the average IFDs were just 145–157 nm. The radial distributions, to 250 nm, all showed a nearest neighbors first peak which shifted slightly from predominantly at 45–54 nm with more rectangular ROIs. The radial distribution profiles could be shown to be statistically different if the ROI L:W ratio was 2:1 or greater. CONCLUSION: Selection of an ROI for assessment of packing density and interfibril distances should be standardized for comparative assessments of TEMs of collagen fibrils. Microsc. Res. Tech., 2011. © 2011 Wiley‐Liss, Inc.  相似文献   

12.
In a previous study, the dislocations in Fe30Ni20Mn25Al25 (at. %), which consist of 50 nm wide alternating b.c.c. and B2 phases, were shown to have a/2<111> Burgers vectors after room temperature deformation. The dislocations were found to glide in pairs on both {110} and {112} slip planes and were relatively widely separated in the b.c.c. phase, where the dislocations were uncoupled, and closely spaced in the B2 phase, where the dislocations were connected by an anti‐phase boundary. In this article, we analyze the dislocations in the two ~5 nm‐wide B2 phases in a related two‐phase alloy Fe30Ni20Mn20Al30, with compositions Fe‐23Ni‐21Mn‐24Al and Fe‐39Ni‐12Mn‐34Al, compressed to ~3% strain at a strain rate 5 × 10?4 s?1 at 873 K (the lowest temperature at which substantial plastic flow was observed). It is shown that slip occursby the glide of a<100> dislocations. A review of the literature suggests that the differences in the observed slip vector between these B2 phases could be due to the differences in composition, differences in deformation temperature, or possibly both. Microsc. Res. Tech. 76:263–267, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

13.
A gold film of about half a monolayer in thickness was vacuum-deposited onto air-cleaved MoS2 surface at room temperature at a pressure of 10?5 Pa. Quasi-two-dimensionally grown gold islands made of several atomic layers were observed by scanning an area of 40×30 nm2. The shape of them was oblate and about 20 nm in diameter. Atomic corrugations which reflect the Au(111) surface were observed. The islands were found to move and changed their shape during the microscopy.  相似文献   

14.
A new technique of analytical transmission electron microscopy called ConceptEM has been developed for determining highly accurately small amounts of solute or dopant atoms incorporated into well‐defined planar defects such as stacking faults, grain boundaries or interfaces. The method is based on recording series of analytical spectra taken with different electron beam diameters on the same position centred above a defect that is orientated either edge‐on or slightly inclined with respect to the electron beam. It can be applied to energy‐dispersive X‐ray spectroscopy or electron energy‐loss spectroscopy and necessitates only a nano‐probe modus but no scanning unit. Reliability and accuracy have been tested numerically under various conditions using simulations for a specific geometry, as a function of specimen thickness, material, acceleration voltage, collection angle, random beam displacements and solid solubility. The accuracy has been found to be substantially better (by factors of 5–10) than that of any other current standard technique based on single measurements. Our calculations suggest an accuracy in the determination of the Gibbsian solute excess at a special grain boundary down to ±1% of a monolayer, i.e. around ±0.1 atoms nm?2 under typical experimental conditions, with a maximum error about twice as large. The parameter limiting a straightforward analysis is found to be the solid solubility, which itself, however, can be measured accurately by the technique so that it can be taken into account quantitatively and the above‐stated precision is retained.  相似文献   

15.
The combination of scanning transmission electron microscopy (STEM) and parallel-detection energy-loss spectroscopy (EELS) was used to detect specific bound elements within macromolecules and macromolecular assemblies prepared by direct freezing. After cryotransferring and freeze-drying in situ, samples were re-cooled to liquid nitrogen temperature and low-dose (about 103 e/nm2) digital dark-field images were obtained with single-electron sensitivity using a beam energy of approximately 100 keV and a probe current of approximately 5 pA. These maps provided a means of characterizing the molecular weights of the structures at low dose. The probe current was subsequently increased to about 5 nA in order to perform elemental analysis. The 320 copper atoms in a keyhole limpet haemocyanin molecule (mol.wt = 8 MDa) were detected with a sensitivity of ± 30 atoms in an acquisition time of 200 s. Phosphorus was detected in an approximately 10-nm length of single-stranded RNA contained in a tobacco mosaic virus particle (mol.wt = 130 kDa/nm) with a sensitivity of ± 25 atoms. Near single-atom sensitivity was achieved for the detection of iron in one haemoglobin molecule (mol.wt = 65 kDa, containing four Fe atoms). Such detection limits are only feasible if special processing methods are employed, as is demonstrated by the use of the second-difference acquisition technique and multiple least-squares fitting of reference spectra. Moreover, an extremely high electron dose (about 1010 e/nm2) is required resulting in mass loss that may be attributable to ‘knock-on’ radiation damage.  相似文献   

16.
Since the recent boost in the usage of electron microscopy in life‐science research, there is a great need for new methods. Recently minimal resin embedding methods have been successfully introduced in the sample preparation for focused‐ion beam scanning electron microscopy (FIB‐SEM). In these methods several possibilities are given to remove as much resin as possible from the surface of cultured cells or multicellular organisms. Here we introduce an alternative way in the minimal resin embedding method to remove excess of resin from two widely different cell types by the use of Mascotte filter paper. Our goal in correlative light and electron microscopic studies of immunogold‐labelled breast cancer SKBR3 cells was to visualise gold‐labelled HER2 plasma membrane proteins as well as the intracellular structures of flat and round cells. We found a significant difference (p < 0.001) in the number of gold particles of selected cells per 0.6 m2 cell surface: on average a flat cell contained 2.46 ± 1.98 gold particles, and a round cell 5.66 ± 2.92 gold particles. Moreover, there was a clear difference in the subcellular organisation of these two cells. The round SKBR3 cell contained many organelles, such as mitochondria, Golgi and endoplasmic reticulum, when compared with flat SKBR3 cells. Our next goal was to visualise crosswall associated organelles, septal pore caps, of Rhizoctonia solani fungal cells by the combined use of a heavy metal staining and our extremely thin layer plastification (ETLP) method. At low magnifications this resulted into easily finding septa which appeared as bright crosswalls in the back‐scattered electron mode in the scanning electron microscope. Then, a septum was selected for FIB‐SEM. Cross‐sectioned views clearly revealed the perforate septal pore cap of R. solani next to other structures, such as mitochondria, endoplasmic reticulum, lipid bodies, dolipore septum, and the pore channel. As the ETLP method was applied on two widely different cell types, the use of the ETLP method will be beneficial to correlative studies of other cell model systems and multicellular organisms.  相似文献   

17.
We introduce a label‐free spectroscopic method to classify subtypes of quinolone‐nonsusceptible Escherichia coli (E. coli ) isolates obtained from human blood cultures. Raman spectroscopy with a 30‐nm gold‐deposited, surface‐enhanced Raman scattering (SERS) substrate was used to evaluate three multilocus sequencing typing (MLST)‐predefined groups including E . coli ATCC25922, E . coli ST131:O75, and E . coli ST1193:O25b. Although there was a coffee‐ring effect, the ring zone was selected at the ideal position to screen E. coli isolates. Strong Raman peaks were present at 1001–1004 cm?1 (C? C aromatic ring breathing stretching vibrational mode of phenylalanine), 1447–1448 cm?1 (C? H2 scissoring deformation vibrational mode), and 1667 cm?1 (amide I α‐helix). Although the three MLST‐predefined E . coli isolates had similar Raman spectral patterns, a support vector machine (SVM) learning algorithm‐assisted principal component analysis (PCA) analysis had superior performance in detecting the presence of quinolone‐nonsusceptible E. coli isolates as well as classifying similar microbes, such as quinolone‐nonsusceptible E . coli ST131:O75 and E . coli ST1193:O25b isolates. Therefore, this label‐free and nondestructive technique is likely to be useful for clinically diagnosing quinolone‐nonsusceptible E. coli isolates with the MLST method.  相似文献   

18.
The aim of the this study was to quantitatively evaluate in three‐dimensional (3D), the porosity degree of three improved silicate‐based endodontic repair cements (iRoot BP Plus®, Biodentine®, and Ceramicrete) compared to a gold‐standard calcium silicate bioactive cement (Pro Root® MTA). From each tested cement, four samples were prepared by a single operator following the manufacturer's instructions in terms of proportion, time, and mixing method, using cylindrical plastic split‐ring moulds. The moulds were lubricated and the mixed cements were inserted with the aid of a cement spatula. The samples were scanned using a compact micro‐CT device (Skyscan 1174, Bruker micro‐CT, Kontich, Belgium) and the projection images were reconstructed into cross‐sectional slices (NRecon v.1.6.9, Bruker micro‐CT). From the stack of images, 3D models were rendered and the porosity parameters of each tested material were obtained after threshold definition by comparison with standard porosity values of Biodentine®. No statistically significant differences in the porosity parameters among the different materials were seen. Regarding total porosity, iRoot BP Plus® showed a higher percentage of total porosity (9.58%), followed by Biodentine® (7.09%), Pro Root® MTA (6.63%), and Ceramicrete (5.91%). Regarding closed porosity, Biodentine® presented a slight increase in these numbers compared to the other sealers. No significant difference in porosity between iRoot BP Plus®, Biodentine®, and Ceramicrete were seen. In addition, no significant difference in porosity between the new calcium silicate‐containing repair cements and the gold‐standard MTA were found. Microsc. Res. Tech., 76:1093–1098, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

19.
Abstract

A simple, rapid, highly sensitive, and selective method for detecting protein in biological samples has been developed, which is based on the interaction between protein and m‐nitrophenylfluorone‐Mo(VI) complex as a spectral probe. The optimum condition for the reaction is investigated. Bovine serum albumin (BSA) obeys Beer's law up to 10 µg · mL?1, having a molar absorption coefficient of 8.51×106 L · mol?1 · cm?1 at 535 nm. Many amino acids and metal ions do not interfere. The results of determination for biological samples are comparable to those obtained by the Bradford method. Meanwhile, the binding number is also determined.  相似文献   

20.
Electron tomography is applied to photocatalytic gold/titanium oxide and gold/silver/titanium oxide samples. In order to obtain a tilt series for the electron tomography measurement, high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) is used under cryogenic conditions. Dedicated programs have been developed for measuring volume, surface area, thickness distribution and nearest-neighbour distance of metallic nanoparticles on samples. Using these quantification programs, the 3D morphology of gold and silver nanoparticles is accurately characterized. We paid particular attention to the quantitative measurement of surface area. The measurement error of the method and appropriate magnification are defined using spherical nanoparticle models. We measured the 3D morphology of gold nanoparticles supported on titanium oxide (total volume=6.5×105 [nm3], surface area=1.4×105 [nm2], and average nearest-neighbour distance=40 [nm]).  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