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 248 毫秒
1.
Effect of diethylstilbestrol on ion fluxes in oat roots   总被引:10,自引:5,他引:5       下载免费PDF全文
Effects of diethylstilbestrol (DES) on ion fluxes in oat roots (Avena sativa L.) were investigated by measuring K+ and Cl absorption and K+ efflux. DES rapidly decreased the absorption of K+ (86Rb) and 36Cl by excised roots; 10−4 molar DES inhibited Cl absorption in 1 minute and K+ absorption in 1 to 2 minutes. With a 10-minute incubation period, K+ and Cl absorption were inhibited 50% by 1.1×10−5 molar and 8.4×10−6 molar DES, respectively. Treatment for 3 minutes with 10−4 molar DES caused irreversible inhibition of K+ absorption. Increasing concentrations of KCl in the absorption media decreased the DES inhibition. Experiments with the DES analogs, DES dipropionate, dienestrol and hexestrol, showed that the steric configuration and the hydroxyl group of the DES molecule are important in determining the inhibitory capacity of the compound.  相似文献   

2.
Salicylic and acetylsalicylic acids applied on excised leaves of Cassia fasciculata modify the dark-induced (scotonastic) and light-induced (photonastic) leaflet movements. They inhibit the scotonastic movements in a dose-dependent manner from 1 × 10−4 to 1 × 10−3 molar and they promote the photonastic movements at an optimum concentration of 5 × 10−4 molar. These results suggest that these phenolic compounds do not act specifically on the K+ uptake, which was shown to be inhibited by their action on other materials.  相似文献   

3.
Harper JR  Balke NE 《Plant physiology》1981,68(6):1349-1353
The phenolic compounds salicylic acid (o-hydroxybenzoic acid) and ferulic acid (4-hydroxy-3-methoxycinnamic acid) inhibited K+ (86Rb+) absorption in excised oat (Avena sativa L. cv. Goodfield) root tissue. Salicylic acid was the most inhibitory. The degree of inhibition was both concentration- and pH-dependent. With decreasing pH, the inhibitory effect of the phenolic increased. During the early stages of incubation, the time required to inhibit K+ absorption was also pH- and concentration-dependent. At pH 4.0, 5×10−4 molar salicylic acid inhibited K+ absorption about 60% within 1 minute; whereas, at pH 6.5, this concentration affected absorption only after 10 to 15 minutes. However, at 5 × 10−3 molar and pH 6.5, salicylic acid was inhibitory within 1 minute. The capacity of the tissue to recover following a 1-hour treatment in 5 × 10−4 molar salicylic acid ranged from no recovery at pH 4.5 to complete recovery at pH 7.5. The absorption of salicylic acid was pH-dependent, also. As pH decreased, more of the phenolic compound was absorbed by the tissue. The increased absorption of the compound at low pH most likely contributed to apparent tissue damage at pH 4.5 and might have accounted for the lack of recovery of K+ absorption as pH decreased.  相似文献   

4.
Pyrophosphorylytic kinetic constants (S0.5, Vmax) of partially purified UDP-glucose- and ADP-glucose pyrophosphorylases from potato tubers were determined in the presence of various intermediary metabolites. The S0.5 of UDP-glucose pyrophosphorylase for UDP-glucose (0.17 millimolar) or pyrophosphate (0.30 millimolar) and the Vmax were not influenced by high concentrations (2 millimolar) of these substances. The most efficient activator of ADP-glucose pyrophosphorylase was 3-P-glycerate (A0.5 = 4.5 × 10−6 molar). The S0.5 for ADP-glucose and pyrophosphate was increased 3.5-fold (0.83 to 0.24 millimolar) and 1.8-fold (0.18 to 0.10 millimolar), respectively, with 0.1 millimolar 3-P-glycerate while the Vmax was increased nearly 4-fold. The magnitude of 3-P-glycerate stimulation was dependent upon the integrity of key sulfhydryl groups (−SH) and pH. Oxidation or blockage of −SH groups resulted in a marked reduction of enzyme activity. Stimulations of 3.1-, 2.9-, 4.8-, and 9.5-fold were observed at pH 7.5, 8.0, 8.5, and 9.0, respectively, in the presence of 3-P-glycerate (2 millimolar). The most potent inhibitor of ADP-glucose pyrophosphorylase was orthophosphate (I0.5 = 8.8 × 10−5. molar). This inhibition was reversed with 3-P-glycerate (1.2 × 10−4 molar), resulting in an increased I0.5 value of 1.5 × 10−3 molar. Likewise, orthophosphate (7.5 × 10−4 molar) caused a decrease in the activation efficiency of 3-P-glycerate (A0.5 from 4.5 × 10−6 molar to 6.7 × 10−5 molar). The significance of 3-P-glycerate activation and orthophosphate inhibition in the regulation of α-glucan biosynthesis in Solanum tuberosum is discussed.  相似文献   

5.
Effect of several parameters on inhibition of potato (Solanum tuberosum) invertase by its endogenous proteinaceous inhibitor was determined using homogeneous preparations of both proteins. The inhibitor and invertase formed an inactive complex with an observed association rate constant at pH 4.70 and 37°C of 8.82 × 102 per molar per second and a dissociation rate constant of 3.3 × 10−3 per minute. The inhibitor appeared to bind to invertase in more than one step. Initial interaction (measured by loss of invertase activity) was rapid, relatively weak, readily reversible (Ki of 2 × 10−6 molar) and noncompetitive with substrate at pH 4.70. Initial interaction was probably followed by isomerization to a tighter (Ki of 6.23 × 10−8 molar) complex, which dissociated slowly with a half-time of 3.5 hour. Interaction between enzyme and inhibitor appeared to be of ionic character and essentially pH independent between pH 3.5 and 7.4.  相似文献   

6.
We purified a secreted fungal laccase from filtrates of Gaeumannomyces graminis var. tritici cultures induced with copper and xylidine. The active protein had an apparent molecular mass of 190 kDa and yielded subunits with molecular masses of 60 kDa when denatured and deglycosylated. This laccase had a pI of 5.6 and an optimal pH of 4.5 with 2,6-dimethoxyphenol as its substrate. Like other, previously purified laccases, this one contained several copper atoms in each subunit, as determined by inductively coupled plasma spectroscopy. The active enzyme catalyzed the oxidation of 2,6-dimethoxyphenol (Km = 2.6 × 10−5 ± 7 × 10−6 M), catechol (Km = 2.5 × 10−4 ± 1 × 10−5 M), pyrogallol (Km = 3.1 × 10−4 ± 4 × 10−5 M), and guaiacol (Km = 5.1 × 10−4 ± 2 × 10−5 M). In addition, the laccase catalyzed the polymerization of 1,8-dihydroxynaphthalene, a natural fungal melanin precursor, into a high-molecular-weight melanin and catalyzed the oxidation, or decolorization, of the dye poly B-411, a lignin-like polymer. These findings indicate that this laccase may be involved in melanin polymerization in this phytopathogen’s hyphae and/or in lignin depolymerization in its infected plant host.  相似文献   

7.
Synthetic and natural polymers are often used as drug delivery systems in vitro and in vivo. Biodegradable chitosan of different sizes were used to encapsulate antitumor drug tamoxifen (Tam) and its metabolites 4-hydroxytamoxifen (4-Hydroxytam) and endoxifen (Endox). The interactions of tamoxifen and its metabolites with chitosan 15, 100 and 200 KD were investigated in aqueous solution, using FTIR, fluorescence spectroscopic methods and molecular modeling. The structural analysis showed that tamoxifen and its metabolites bind chitosan via both hydrophilic and hydrophobic contacts with overall binding constants of K tam-ch-15  = 8.7 (±0.5)×103 M−1, K tam-ch-100  = 5.9 (±0.4)×105 M−1, K tam-ch-200  = 2.4 (±0.4)×105 M−1 and K hydroxytam-ch-15  = 2.6(±0.3)×104 M−1, K hydroxytam – ch-100  = 5.2 (±0.7)×106 M−1 and K hydroxytam-ch-200  = 5.1 (±0.5)×105 M−1, K endox-ch-15  = 4.1 (±0.4)×103 M−1, K endox-ch-100  = 1.2 (±0.3)×106 M−1 and K endox-ch-200  = 4.7 (±0.5)×105 M−1 with the number of drug molecules bound per chitosan (n) 2.8 to 0.5. The order of binding is ch-100>200>15 KD with stronger complexes formed with 4-hydroxytamoxifen than tamoxifen and endoxifen. The molecular modeling showed the participation of polymer charged NH2 residues with drug OH and NH2 groups in the drug-polymer adducts. The free binding energies of −3.46 kcal/mol for tamoxifen, −3.54 kcal/mol for 4-hydroxytamoxifen and −3.47 kcal/mol for endoxifen were estimated for these drug-polymer complexes. The results show chitosan 100 KD is stronger carrier for drug delivery than chitosan-15 and chitosan-200 KD.  相似文献   

8.
A procedure was developed for the enzymic isolation of large quantities of protoplasts from the cortex of Zea mays L. WF9 × MO 17 roots. Cortex was separated from the primary root, sectioned, and the cell walls digested for 3.5 hours in 2% (w/v) Cellulysin, 0.1% Pectolyase Y-23, 1 millimolar CaCl2, 0.05% bovine serum albumin, 0.5 millimolar dithiothreitol in 0.6 molar mannitol (pH 5.6). Cortical cell protoplasts were collected by centrifugation and purified by flotation in a Ficoll step gradient. The yield of protoplasts was approximately 650 × 103/gram fresh tissue. To obtain maximum yield it was essential to include an effective pectinase (Pectolyase Y-23) and protectants (bovine serum albumin and dithiothreitol) in the digestion medium.

Cortical cell protoplasts exhibited energy-dependent uptake of K+ (86Rb), H232PO4, and 36Cl as well as net H+ extrusion. Ion fluxes were sustained for at least 3 hours. Influx of K+ was highest between pH 7.5 and 8.0, whereas the influx of H2PO4 was greatest between pH 4.0 and 5.0. K+ and H2PO4 influx and net H+ efflux were inhibited by respiratory poisons such as cyanide (0.1 millimolar) and oligomycin (5 micrograms per milliliter), and by inhibitors of plasma membrane ATPase such as diethylstilbestrol (50 micromolar). Calculated flux for Cl was low, but not greatly different from that observed for other plant cells. K+ flux was somewhat high, probably because the K+ concentration in the cortical cells was below steady-state. The results indicate that isolated cortical cell protoplasts retain transport properties which are similar to those of root tissue.

  相似文献   

9.
Studies in human populations and mouse models of disease have linked the common leptin receptor Q223R mutation to obesity, multiple forms of cancer, adverse drug reactions, and susceptibility to enteric and respiratory infections. Contradictory results cast doubt on the phenotypic consequences of this variant. We set out to determine whether the Q223R substitution affects leptin binding kinetics using surface plasmon resonance (SPR), a technique that allows sensitive real-time monitoring of protein-protein interactions. We measured the binding and dissociation rate constants for leptin to the extracellular domain of WT and Q223R murine leptin receptors expressed as Fc-fusion proteins and found that the mutant receptor does not significantly differ in kinetics of leptin binding from the WT leptin receptor. (WT: ka 1.76×106±0.193×106 M−1 s−1, kd 1.21×10−4±0.707×10−4 s−1, KD 6.47×10−11±3.30×10−11 M; Q223R: ka 1.75×106±0.0245×106 M−1 s−1, kd 1.47×10−4±0.0505×10−4 s−1, KD 8.43×10−11±0.407×10−11 M). Our results support earlier findings that differences in affinity and kinetics of leptin binding are unlikely to explain mechanistically the phenotypes that have been linked to this common genetic variant. Future studies will seek to elucidate the mechanism by which this mutation influences susceptibility to metabolic, infectious, and malignant pathologies.  相似文献   

10.
1. Purified rabbit-muscle and -liver glucose phosphate isomerase, free of contaminating enzyme activities that could interfere with the assay procedures, were tested for inhibition by fructose, fructose 1-phosphate and fructose 1,6-diphosphate. 2. Fructose 1-phosphate and fructose 1,6-diphosphate are both competitive with fructose 6-phosphate in the enzymic reaction, the apparent Ki values being 1·37×10−3−1·67×10−3m for fructose 1-phosphate and 7·2×10−3−7·9×10−3m for fructose 1,6-diphosphate; fructose and inorganic phosphate were without effect. 3. The apparent Km values for both liver and muscle enzymes at pH7·4 and 30° were 1·11×10−4−1·29×10−4m for fructose 6-phosphate, determined under the conditions in this paper. 4. In the reverse reaction, fructose, fructose 1-phosphate and fructose 1,6-diphosphate did not significantly inhibit the conversion of glucose 6-phosphate into fructose 6-phosphate. 5. The apparent Km values for glucose 6-phosphate were in the range 5·6×10−4−8·5×10−4m. 6. The competitive inhibition of hepatic glucose phosphate isomerase by fructose 1-phosphate is discussed in relation to the mechanism of fructose-induced hypoglycaemia in hereditary fructose intolerance.  相似文献   

11.
The effects of the photosystem II herbicides diuron (3-(3,4-dichlorophenyl)-1,1-dimethylurea) and atrazine (2-chloro-4-ethylamino-6-isopropylamino-s-triazine) on the photosynthetic membranes of a cyanobacterium, Aphanocapsa 6308, were compared to the effects on a higher plant, Spinacia oleracea. The inhibition of photosystem II electron transport by these herbicides was investigated by measuring the photoreduction of the dye 2,6-dichlorophenol-indophenol spectrophotometrically using isolated membranes. The concentration of herbicide that caused 50% inhibition of electron transport (I50 value) in Aphanocapsa membranes for diuron was 6.8 × 10−9 molar and the I50 value for atrazine was 8.8 × 10−8 molar. 14C-labeled diuron and atrazine were used to investigate herbicide binding with calculated binding constants (K) being 8.2 × 10−8 molar for atrazine and 1.7 × 10−7 molar for diuron. Competitive binding studies carried out on Aphanocapsa membranes using radiolabeled [14C]atrazine and unlabeled diuron revealed that diuron competed with atrazine for the herbicide-binding site. Experiments involving the photoaffinity label [14C]azidoatrazine (2-azido-4-ethylamino-6-isopropylamino-2-triazine) and autoradiography of polyacrylamide gels indicated that the herbicide atrazine binds to a 32-kilodalton protein in Aphanocapsa 6308 cell extracts.  相似文献   

12.
We have examined the activity of the thiamin phosphate pyrophosphorylase in Arabidopsis thaliana wild type and in a mutant (th-1) which requires exogenous thiamin for growth. Mutant and wild-type plants grown in 1 × 10−7 molar thiamin were used for the examination of the production of thiamin and thiamin monophosphate (TMP) using 4-methyl-5-hydroxyethylthiazole phosphate and 2-methyl-4-amino-5-hydroxymethylpyrimidine pyrophosphate as substrates. While the wild-type strain formed both thiamin and TMP, the th-1 mutant did not. When TMP was added to the extracts, the th-1 mutant, as well as wild type, produced thiamin. Accordingly, it was concluded that the th-1 mutant was defective in the activity of TMP pyrophosphorylase. Some of the characteristics of the enzyme from the wild-type plant were examined. The optimum temperature for the reaction is 45°C, and the Km values for the substrates are 2.7 × 10−6 molar for 4-methyl-5-hydroxyethylthiazole phosphate and 1.8 × 10−6 molar for 2-methyl-4-amino-5-hydroxymethylpyrimidine pyrophosphate.  相似文献   

13.
Phanerochaete chrysosporium produces intracellular soluble and particulate β-glucosidases and an extracellular β-glucosidase. The extracellular enzyme is induced by cellulose but repressed in the presence of glucose. The molecular weight of this enzyme is 90,000. The Km for p-nitrophenyl-β-glucoside is 1.6 × 10−4 M; the Ki for glucose, a competitive inhibitor, is 5.0 × 10−4 M. The Km for cellobiose is 5.3 × 10−4 M. The intracellular soluble enzyme is induced by cellobiose; this induction is prevented by cycloheximide. The presence of 300 mM glucose in the medium, however, had no effect on induction. The Km for p-nitrophenyl-β-glucoside is 1.1 × 10−4 M. The molecular weight of this enzyme is ~410,000. Both enzymes have an optimal temperature of 45°C and an Eact of 9.15 kcal (ca. 3.83 × 104 J). The pH optima, however, were ~7.0 and 5.5 for the intracellular and extracellular enzymes, respectively.  相似文献   

14.
Conditions are described for isolating functional phycobilisome-thylakoid vesicles from the red alga Porphyridium cruentum. Phycobilisome-thylakoid vesicles were prepared by brief sonication and centrifugation in a medium containing 0.5 molar sucrose, 0.5 molar potassium phosphate, and 0.3 molar sodium citrate (pH 7.0). They required ferricyanide as an oxidant and had O2 evolution rates (about 450 micromoles O2 per hour per milligram chlorophyll) higher than whole cells (about 250 micromoles O2 per hour per milligram chlorophyll). Energy transfer to photosystem II chlorophyll was evident from a high F695 nanometer (−196 C) emission peak. Preparations could be stored for over 24 hours and were considerably more stable than those from the cyanobacterium Anabaena variabilis (Katoh T, E Gantt 1979 Biochim Biophys Acta 546: 383-393). In electron micrographs of negatively stained material, the active thylakoid vesicles were found covered by closely spaced phycobilisomes on their external surface. The phycobilisome number in negatively stained vesicles was 450 per square micrometer, which was in the same range as the 400 per square micrometer observed in surface sections. A cell containing 1.5 × 10−6 micrograms phycoerythrin and 1.3 × 10−6 micrograms chlorophyll was found to contain 5 to 7 × 105 phycobilisomes on a thylakoid area of 1.1 to 1.6 × 103 square micrometers.  相似文献   

15.
1. An improved method is given for preparation of pyruvate,phosphate dikinase from Bacteroides symbiosus. 2. The bacterial enzyme is stable, free from interfering enzyme activities, and does not require thiol compounds to maintain stability during storage or assay. 3. New direct assays of enzyme activity are based on acid evolution or consumption as measured at constant pH in a pH-stat. 4. The optimum rate of reaction in the direction of pyruvate formation occurs at about pH6.4; in the direction of phosphoenolpyruvate formation, it is at pH7.2–7.8. 5. Newly determined substrate Km values for the enzyme are: AMP, 3.5×10−6m; ATP, 1×10−4m; pyruvate, 8×10−5m; Pi, 6×10−4m. 6. K+ may substitute for NH4+ in activating the reaction catalysed by the B. symbiosus enzyme. 7. In the direction of pyruvate formation the bivalent metal ion requirement of the enzyme is fulfilled by salts of nickel, manganese, magnesium and cobalt. In the other direction only magnesium salts were effective. 8. The nucleotide specificity of the enzyme is strictly limited to the adenine nucleotides. CTP and ITP strongly inhibit the reaction in the direction of phosphoenolpyruvate formation.  相似文献   

16.
Changes in the concentrations of NH4+ and amides during the growth of suspension cultures of rose (Rosa cv. Paul's Scarlet) cells were examined. When cells were grown in medium possessing only NO3 as a nitrogen source, the concentrations of NH4+ and amides increased to 4.0 × 10−1 and 5.9 micromoles per gram fresh weight, respectively. The amounts of both constituents declined during the later stages of growth. When a trace amount of NH4+ was added to the NO3 base starting medium, the concentration of NH4+ in the cells was increased to 7.0 × 10−1 micromoles per gram fresh weight.  相似文献   

17.
A 2-gram fresh weight inoculum of bromegrass (Bromus inermis Leyss. culture BG970) cell suspension culture treated with 7.5 × 10−5 molar abscisic acid (ABA) for 7 days at 25°C survived slow cooling to −60°C. Over 80% of the cells in ABA treated cultures survived immersion in liquid N2 after slow cooling to −40 or −60°C. In contrast, a 6-gram fresh weight inoculum only attained a hardiness level of −28°C after 5 days of ABA treatment. Ethanol (2 × 10−2 molar) added to the culture medium at the time of ABA addition, inhibited the freezing tolerance of bromegrass cells by 25°C. A 6-gram inoculum of both control and ABA treated bromegrass cells altered the pH of the medium more than a 2-gram inoculum. ABA inhibited the increase in fresh weight of bromegrass by 20% after 4 days. Both control and ABA (10−4 molar) treated alfalfa cells (Medicago sativa L.) grown at 25°C hardened from an initial LT50 of −5°C to an LT50 of −23°C by the third to fifth day after subculture. Thereafter, the cells dehardened but the ABA treated cells did not deharden to the same level as the control cells. ABA inhibited the increase in fresh weight of alfalfa by 50% after 5 days.  相似文献   

18.
The objective of this study was to improve the growth of in vitro shoot cultures of Brugmansia × candida ‘Creamsickle’. Several mineral nutrient experiments were conducted to determine the effect of NH4+, NO3, K+, FeSO4/EDTA, ZnSO4, MnSO4, and CuSO4 on quality, leaf width and length, size and weight of shoot mass, and shoot number. The experiment to determine the levels of NH4+, NO3, and K+, was conducted as a 2-component NH4+: K+ mixture crossed by [NO3] and resulted in an experimental design free of ion confounding and capable of separating the effects of proportion and concentration. The results of the NH4+-K+-NO3 experiment revealed a region in the design space where growth was significantly improved; the region generally had lower total nitrogen and lower NH4+:K+ ratios than MS medium. The experiments to determine the appropriate levels of Fe, Zn, Mn, and Cu were conducted at six log levels ranging from 0 to 1 mM. Of the four metal salts tested, MnSO4 had the least effect on in vitro shoot growth and its concentration was reduced from 0.1 mM (MS level) to 0.001 mM. CuSO4 had large effects on in vitro shoot growth and was increased from 0.0001 mM to 0.001 mM. A 2-level factorial of NH4+-K+-NO3, FeSO4/EDTA, and ZnSO4 was conducted and several formulations identified for their improvements of quality and growth. In addition to the changes to MnSO4 and CuSO4, these formulations were characterized by lower levels of NH4+, K+, NO3 and Zn, and higher levels of FeSO4/EDTA. Overall, several nutrient formulations were identified as superior to MS medium for growth of in vitro shoot cultures of B. ‘Creamsickle’.  相似文献   

19.
Complexes of cationic liposomes with DNA are promising tools to deliver genetic information into cells for gene therapy and vaccines. Electrostatic interaction is thought to be the major force in lipid–DNA interaction, while lipid-base binding and the stability of cationic lipid–DNA complexes have been the subject of more debate in recent years. The aim of this study was to examine the complexation of calf-thymus DNA with cholesterol (Chol), 1,2-dioleoyl-3-trimethylammonium-propane (DOTAP), dioctadecyldimethylammoniumbromide (DDAB) and dioleoylphosphatidylethanolamine (DOPE), at physiological condition, using constant DNA concentration and various lipid contents. Fourier transform infrared (FTIR), UV-visible, circular dichroism spectroscopic methods and atomic force microscopy were used to analyse lipid-binding site, the binding constant and the effects of lipid interaction on DNA stability and conformation. Structural analysis showed a strong lipid–DNA interaction via major and minor grooves and the backbone phosphate group with overall binding constants of KChol = 1.4 (±0.5) × 104 M−1, KDDAB = 2.4 (±0.80) × 104 M−1, KDOTAP = 3.1 (±0.90) × 104 M−1 and KDOPE = 1.45 (± 0.60) × 104 M−1. The order of stability of lipid–DNA complexation is DOTAP>DDAB>DOPE>Chol. Hydrophobic interactions between lipid aliphatic tails and DNA were observed. Chol and DOPE induced a partial B to A-DNA conformational transition, while a partial B to C-DNA alteration occurred for DDAB and DOTAP at high lipid concentrations. DNA aggregation was observed at high lipid content.  相似文献   

20.
The following equilibrium constants (given as logK in units of m−1) were determined for the substitution of co-ordinated H2O in aquocobalamin by glycine (bound through N) 5.8, cysteine (bound through S) 6.0 or 8.3, depending on the value chosen for the pK of the thiol group, and phenolate 2.9. The spectrum of the phenolate cobalamin shows an additional intense absorption band at 468nm with a molar extinction coefficient of 1.1×104, which is assigned to a charge transfer from the phenolate to the cobalt ion. Equilibrium constants have also been determined for the equilibria between adenylcobamide cyanide and CN, HO and H+, which show that the adenine is more easily displaced by CN and HO than is 5,6-dimethylbenziminazole in vitamin B12, but can be protonated by acid while still remaining co-ordinated to the cobalt. It is shown that in the binding of corrinoids to proteins and polypeptides the formation of hydrogen bonds is far more important than co-ordination by the metal.  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