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.
Corynebacterium glutamicum took up glutamine by a sodium-dependent secondary transport system. Both the membrane potential and the sodium gradient were driving forces. Glutamine uptake showed Michaelis-Menten kinetics, with aK m of 36 μM and aV max of 12.5 nmol min−1 (mg dry weight)−1 at pH 7. Despite a pH optimum in the alkaline range around pH 9, it was shown that uncharged glutamine is the transported species. The affinity for the cotransported sodium was relatively low; an apparentK m of 1.4 mM was determined. Among various substrates tested, only asparagine, when added in 50-fold excess, led to an inhibition of glutamine transport. It was concluded that glutamine uptake occurs via a specific transport system in symport with at least one sodium ion.  相似文献   

2.
The effect of trace metal ions (Co2+, Cu2+, Fe2+, Mn2+, Mo6+, Ni2+, Zn2+, SeO4 and WO4 ) on growth and ethanol production by an ethanologenic acetogen, Clostridium ragsdalei was investigated in CO:CO2-grown cells. A standard acetogen medium (ATCC medium no. 1754) was manipulated by varying the concentrations of trace metals in the media. Increasing the individual concentrations of Ni2+, Zn2+, SeO4 and WO4 from 0.84, 6.96, 1.06, and 0.68 μM in the standard trace metals solution to 8.4, 34.8, 5.3, and 6.8 μM, respectively, increased ethanol production from 35.73 mM under standard metals concentration to 176.5, 187.8, 54.4, and 72.3 mM, respectively. Nickel was necessary for growth of C. ragsdalei. Growth rate (μ) of C. ragsdalei improved from 0.34 to 0.49 (day−1), and carbon monoxide dehydrogenase (CODH) and hydrogenase (H2ase)-specific activities improved from 38.45 and 0.35 to 48.5 and 1.66 U/mg protein, respectively, at optimum concentration of Ni2+. At optimum concentrations of WO4 and SeO4 , formate dehydrogenase (FDH) activity improved from 32.3 to 42.6 and 45.4 U/mg protein, respectively. Ethanol production and the activity of FDH reduced from 35 mM and 32.3 U/mg protein to 1.14 mM and 8.79 U/mg protein, respectively, upon elimination of WO4 from the medium. Although increased concentration of Zn2+ enhanced growth and ethanol production, the activities of CODH, FDH, H2ase and alcohol dehydrogenase (ADH) were not affected by varying the Zn2+ concentration. Omitting Fe2+ from the medium decreased ethanol production from 35.7 to 6.30 mM and decreased activities of CODH, FDH, H2ase and ADH from 38.5, 32.3, 0.35, and 0.68 U/mg protein to 9.07, 7.01, 0.10, and 0.24 U/mg protein, respectively. Ethanol production improved from 35 to 54 mM when Cu2+ was removed from the medium. The optimization of trace metals concentration in the fermentation medium improved enzyme activities (CODH, FDH, and H2ase), growth and ethanol production by C. ragsdalei.  相似文献   

3.
The potential of alginate-immobilized Microcystis packed in a column for maximum removal of Cu2+ at different flow rates, biomass, and initial metal ion concentration was assessed in a continuous flow system. Although Cu2+ removal did occur at all the flow rates tested, it was maximum (54%) at 0.75-ml min−1 flow rate, 30 μg ml−1 initial metal ion concentration and 0.016 g biomass. Cu2+ removal was influenced by inlet metal ion concentration and biomass density. An increase in the biomass concentration from 0.016 to 0.128 g resulted in an apparent increase in percentage removal but the Cu2+ adsorbed per unit dry wt. declined. When the flow rate (0.75 ml min−1) and biomass density (0.064 g) were kept constant and the inlet metal ion concentration was varied from 10 to 150 μg ml−1, a 68% removal of Cu2+ was obtained at 50 μg ml−1 initial concentration in a time duration of 15 min. The metal-laden columns were efficiently desorbed and regenerated following elution with double distilled water (DDW) (pH 2) (89%). This was followed by 1 mm EDTA > 1 mm NTA > 0.1 mm EDTA > 1 mm HCl > 1 mm HNO3 > 5 mm CaCl2 > DDW (pH 7.0) > 1 mm NaHCO3 > 1 mm CaCl2. Of the total (2.83 mg) adsorbed Cu2+, 1.89 mg (67%) was desorbed by DDW (pH 2) within the first 20 min of elution time. Thereafter the desorption rate slowed down and only 22% (0.632 mg) desorption was obtained in the last 20 min. In contrast to water pH 2, the desorption of Cu2+ by 1 mm EDTA was very slow, the maximum being 8% after 40 min of elution. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

4.
Artemisinin production by hairy roots of Artemisia annua L. was increased 6-fold to 1.8 μg mg−1 dry wt over 6 days by adding 150 mg chitosan l−1. The increase was dose-dependent. Similar treatment of hairy roots with methyl jasmonate (0.2 mM) or yeast extract (2 mg ml−1) increased artemisinin production to 1.5 and 0.9 μg mg−1 dry wt, respectively.  相似文献   

5.
The effects of macronutrients (NO3 , NH4 + and PO4 3−) on cell growth and triterpenoids production inCentella asiatica cell suspension cultures were analyzed using the Box-Behnken response surface model experimental design. In screening and optimization experiments, PO4 3− as a single factor significantly influenced cell growth where increasing the phosphate level from 0.1 to 2.4 or 2.6 mM, elevated cell growth from 3.9 to 14–16 g/L. The optimum values predicted from the response surface model are 5.05 mM NH4 +, 15.0 mM NO3 and 2.6 mM PO4 3−, yielding 16.0 g/L cell dry weight with 99% fitness to the experimental data. While the NH4 +-NO3 interaction influenced cell growth positively in the optimization experiment, NH4 + and NO3 as single factors; and interactions of NO3 -PO4 3−, NH4 +-PO4 3− and NH4 +-NO3 were all negative in the screening experiment. Cell growth and the final pH level were positively affected by PO4 3−, but negatively affected by NH4 + and NH4 +-PO4 3− interactions. The different effects of factors and their interactions on cell growth and final pH are influenced by a broad or narrow range of macronutrient concentrations. The productions of triterpenoids however were lower than 4 mg/g cell dry weight.  相似文献   

6.
The growth, biofiltering efficiency and uptake rates of Ulva clathrata were studied in a series of outdoor tanks, receiving waste water directly from a shrimp (Litopenaeus vannamei) aquaculture pond, under constant aeration and two different water regimes: (1) continuous flow, with 1 volume exchange a day (VE day-1) and (2) static regime, with 1 VE after 4 days. Water temperature, salinity, pH, dissolved inorganic nitrogen (DIN), phosphate (PO4), chlorophyll-a (chl-a), total suspended solids (TSS), macroalgal biomass (fresh weight) and tissue nutrient assimilation were monitored over 12 days. Ulva clathrata was highly efficient in removing the main inorganic nutrients from effluent water, stripping 70–82% of the total ammonium nitrogen (TAN) and 50% PO4 within 15 h. Reductions in control tanks were much lower (Tukey HSD, P < 0.05). After 3 days, the mean uptake rates by the seaweed biomass under continuous flow were 3.09 mg DIN g DW day−1 (383 mg DIN m−2 day−1) and 0.13 mg PO4 g DW day−1 (99 mg PO4 m−2 day−1), being significantly higher than in the static regime (Tukey HSD, P < 0.05). The chl-a decreased in seaweed tanks, suggesting that U. clathrata inhibited phytoplankton growth. Correlations between the cumulative values of DIN removed from the water and total nitrogen assimilated into the seaweed biomass (r = 0.7 and 0.8, P < 0.05), suggest that nutrient removal by U. clathrata dominated over other processes such as phytoplankton and bacterial assimilation, ammonia volatilization and nutrient precipitation.  相似文献   

7.
Effects of nitrate,(NO3) chloride (Cl), sulfate (SO42-, and acetate (Ac) on Cu2+ adsorption and affinity of the adsorbed Cu2+ were evaluated in two Fe and Al enriched variable charge soils from Southern China. The maximum adsorption of Cu2+ (M, a parameter from the Langmuir isotherm model) in the presence of different anions decreased in the order Cl > Ac > NO3 > SO42- for both soils. The clayey loamy soil (mixed siliceous thermic Typic Dystrochrept, TTD), developed on the Arenaceous rock, adsorbed less Cu2+ than the clayey soil (kaolinitic thermic Plinthudults, KTP), derived from the Quaternary red earths, regardless of anion type present in the medium. The affinity of adsorbed Cu2+ to both soils could be characterized by the Kd (distribution coefficient) values and successive extraction of the adsorbed Cu2+ with 1-mol NH4Ac L−1. The log10Kd value was smaller for the TTD soil than for the KTP soil and decreased in the order of Cl > NO3 > SO42- > Ac at low initial Cu2+ concentrations (≤40 mg Cu2+L−1), whereas at 80 mg Cu2+L−1, the log10Kd value was similar for NO3, SO42-, and Ac, but was slightly higher for Cl. Complete extraction of Cu2+ adsorbed in the presence of Ac was achieved. Influence of NO3 and SO42- on the affinity of adsorbed Cu2+ was similar, but the effects of Cl depended on the initial Cu2+ concentrations. The extracted percentage of the adsorbed Cu2+ in the presence of NO3 or SO42- increased with increasing Cu2+ adsorption saturation. The presence of Cl, NO3, or SO42- markedly decreased the equilibrium solution pH for both soils with increasing initial Cu2+ concentrations, and the delta pH values at the highest Cu2+ level were 0.5, 0.63, and 0.55 U for the TTD soil and 0.79, 0.84, and 0.93 U for the KTP soil, respectively for the three anions. The presence of Ac had a minimal influence on the equilibrium solution pH because of the buffering nature of the NaAc/HAc medium which buffered the released protons. The effects of anions on Cu2+ adsorption and affinity of the adsorbed Cu2+ were dependent on anion types and were apparently related to the altered surface properties caused by anion adsorption and/or the formation of anion– Cu2+ complexes.  相似文献   

8.
The role played by a bacterial community composed ofPseudomonas putida, strain 21;Pseudomonas stutzeri, strain 18; andPseudomonas sp., strain 5, and by physical and chemical factors in the degradation of CN and SCN was studied. It was shown that the degradation of CN is determined both by the action of bacteria and by abiotic physical and chemical factors (pH, O2, temperature, the medium agitation rate, etc.). The contribution of chemical degradation was found to increase drastically at pH below 9.0; when air was blown through the medium (irrespective of the pH value); under active agitation of the medium; and when the medium surface interfacing air was increased. Even at elevated pH values (9.0-9.2), suboptimal for bacterial growth, the microbial degradation could account for at most 20–25 mg/1 of CN, regardless of its initial concentration. When CN and SCN were concurrently present in the medium, the former compound was the first to be degraded by microorganisms. The rate of bacterial degradation of SCN under continuous cultivation in a chain of reactors was found to depend on its concentration, the medium flow rate, agitation rate, and the pattern of carbon source supply and could exceed 1 g/(l day). CN and SCN are utilized by bacteria solely as nitrogen sources. The mechanism of CN and SCN degradation by the microbial community is discussed. Deceased.  相似文献   

9.
In our study, we isolated the isolate Trichoderma SP2F1 from sediment samples from the Penchala River, heavily contaminated with effluents from nearby industrial areas. Qualitative and quantitative screening using plate and broth assay, respectively, supplemented with various concentrations of Cu(II) showed the isolate was able to tolerate 6 mM CuSO4, although growth was also detected in broths with 10 mM CuSO4. Trichoderma spp. was able to remove Cu(II) in aqueous solutions in both viable and non-viable cell forms. Bioaccumulation capacity of viable SP2F1 cells removed 19.60 mg g−1 of Cu(II) after 168 h incubation, while the maximum Cu(II) biosorption capacity for non-viable SP2F1 cells was 28.75 mg g−1 of Cu(II). Results here showed that Trichoderma spp isolate SP2F1 has good potential for application in Cu(II) removal, can be used to treat sewage waste by applying either in viable or non-viable cell forms.  相似文献   

10.
Li GQ  Li SS  Qu SW  Liu QK  Ma T  Zhu L  Liang FL  Liu RL 《Biotechnology letters》2008,30(10):1759-1764
Substituted benzothiophenes (BTs) and dibenzothiophenes (DBTs) remain in diesel oil following conventional desulfurization by hydrodesulfurization. A mixture of washed cells (13.6 g dry cell wt l−1) of Rhodococcus erythropolis DS-3 and Gordonia sp. C-6 were employed to desulfurize hydrodesulfurized diesel oil; its sulfur content was reduced from 1.26 g l−1 to 180 mg l−1, approx 86% (w/w) of the total sulfur was removed from diesel oil after three cycles of biodesulfurization. The average desulfurization rate was 0.22 mg sulfur (g dry cell wt)−1 h−1. A bacterial mixture is therefore efficient for the practical biodesulfurization of diesel oil.  相似文献   

11.
Formate oxidase was found in cell-free extracts of Debaryomyces vanrijiae MH201, a soil isolate. After purification by column chromatography, the preparation showed a protein band corresponding to a molecular mass (MM) of 64 kDa on sodium dodecyl sulfate–polyacrylamide gel electrophoresis. The MM, estimated by a gel filtration, was 99 kDa. The preparation showed two and three bands on isoelectric focusing under denaturing and native conditions, respectively. These results suggest that the preparation contained three isoforms, each of which might be composed of αα, αβ, and ββ subunits with apparently similar MM. The preparation acted on formate with K m and V max values of 11.7 mM and 262 μmol min−1 mg−1, respectively, at pH 4.5 and 25°C, but showed no evidence of activity on the other compounds tested. The optimum pH and temperature were pH 4.0 and 35°C, respectively. The preparation showed activities of 85% of the initial activity after storage at pH 6.0 and 4°C for 8 weeks. When 10 mM formaldehyde was reacted with 2.0 U ml−1 of the enzyme preparation at pH 5.5 and room temperature in the presence of 2.0 U ml−1 of a microbial aldehyde oxidase and 100 U ml−1 of catalase for 180 min, neither of formate nor formaldehyde was detected, suggesting that the reaction involved the quantitative conversion of formaldehyde to carbon dioxide.  相似文献   

12.
In this work, we characterized an ecto-ATPase activity in intact mycelial forms of Fonsecaea pedrosoi, the primary causative agent of chromoblastomycosis. In the presence of 1 mM EDTA, fungal cells hydrolyzed adenosine-5′-triphosphate (ATP) at a rate of 84.6 ± 11.3 nmol Pi h−1 mg−1 mycelial dry weight. The ecto-ATPase activity was increased at about five times (498.3 ± 27.6 nmol Pi h−1 mg−1) in the presence of 5 mM MgCl2, with values of V max and apparent K m for Mg-ATP2−corresponding to 541.9 ± 48.6 nmol Pi h−1 mg−1 cellular dry weight and 1.9 ± 0.2 mM, respectively. The Mg2+-stimulated ecto-ATPase activity was insensitive to inhibitors of intracellular ATPases such as vanadate (P-ATPases), bafilomycin A1 (V-ATPases), and oligomycin (F-ATPases). Inhibitors of acid phosphatases (molybdate, vanadate, and fluoride) or alkaline phosphatases (levamizole) had no effect on the ecto-ATPase activity. The surface of the Mg2+-stimulated ATPase in F. pedrosoi was confirmed by assays in which 4,4′-diisothiocyanostylbene-2,2′-disulfonic acid (DIDS), a membrane impermeant inhibitor, and suramin, an inhibitor of ecto-ATPase and antagonist of P2 purinoreceptors. Based on the differential expression of ecto-ATPases in the different morphological stages of F. pedrosoi, the putative role of this enzyme in fungal biology is discussed.  相似文献   

13.
Tong P  Hong Y  Xiao Y  Zhang M  Tu X  Cui T 《Biotechnology letters》2007,29(2):295-301
A new basidiomycete, Trametes sp. 420, produced laccase at 6,810 U l−1 (268 mg, 25.4 U mg−1 protein for guaiacol) in glucose medium and 7,870 U l−1 (310 mg) in cellobiose medium with induction by 0.5 mM Cu2+ and 6 mM o-toluidine. Laccase isozyme E (LacE) was the sole laccase in the fermentation products. It was stable at pH 5–9 and below 70°C over 30 min. The K m values of LacE for four substrates (guaiacol ABTS, 2,6-dimethoxyphenol and syringaldazine) varied from 5 to 245 μM. The activity of LacE was strongly inhibited by NaN3 but not by EDTA or dimethylsulfoxide. LacE at 0.5 U l−1 could decolorize industrial dyes. The open reading frame of the lacE gene was 2,130 bp and was interrupted by 10 introns. It displayed a high homology to laccases from other fungi. Pingui Tong and Yuzhi Hong contributed equally to the study  相似文献   

14.
Li W  Wang MD  Chen H  Chen JM  Shi Y 《Biotechnology letters》2006,28(15):1175-1179
A new isolate, identified as Gordonia sp. ZD-7 by 16S rDNA sequence analysis, grew in n-hexadecane containing dibenzothiophene (DBT) which was degraded from 2.8 mM to 0.2 mM within 48 h. Biodesulfurization could be repeatedly performed for more than 190 h, with average desulfurization rates of 5 mmol DBT kg cells (dry wt)−1 h−1.  相似文献   

15.
Detailed knowledge about the fate and transport of tungsten in soils is critical to understanding and effectively addressing tungsten behavior in the environment. Recent studies have shown that tungsten anions may polymerize (depending upon concentration, pH, and aquatic geochemistry) in aquatic and soil systems. However, to date, of all soluble tungstate species only monotungstates have been scrutinized to a fair extent in adsorption studies. There is a lack of information evaluating adsorption mechanisms of mono- and polytungstates onto clay minerals. The objective of this work is to investigate the adsorption behavior of monotungstates (sodium tungstate, Na2WO4) and polytungstates (sodium metatungstate, 3Na2WO4·9WO3) onto different types of clay minerals (montmorillonite, kaolinite, illite) and an organic adsorbent (Pahokee peat). Batch equilibrium experiments as a function of concentration (adsorption isotherms) and pH (adsorption envelopes) were performed to provide information about mono- and polytungstate adsorption onto clays and Pahokee peat. Adsorption equilibrium data for mono- and polytungstates onto different types of clay minerals and Pahokee peat were modeled with Freundlich and Langmuir isotherms. The adsorption affinity of clays and Pahokee peat for monotungstates follows the order: Pahokee peat>kaolinite>montmorillonite>illite; for polytungstates, the order is as follows: kaolinite>Pahokee peat>montmorillonite>illite. Results of this study suggest that the charges of the clay mineral surface, tungsten species, and solution pH are the main factors controlling tungsten adsorption. Moreover, polymeric tungsten species (i.e., metatungstate) appear to be more mobile in the environment than monomeric tungstate.  相似文献   

16.
This work was aimed at producing inulinase by solid-state fermentation of sugarcane bagasse, using factorial design to identify the effect of corn steep liquor (CSL) and soybean bran concentration, particle size of bagasse and size of inoculum. Maximum inulinase activity achieved was 250 U per g of dry substrate (gds) at 20% (w/w) of CSL, 5% (w/w) of soybean bran, 1 × 1010 cells mL−1 and particle size of bagasse in the range 9/32 mesh. The use of soybean bran decreased the time to reach maximum activity from 96 to 24 h and the maximum productivity achieved was 8.87 U gds−1 h−1. The maximum activity was obtained at pH 5.0 and 55.0°C. Within the investigated range, the enzyme extract was more thermostable at 50.0°C, showing a D-value of 123.1 h and deactivation energy of 343.9 kJ gmol−1. The extract showed highest stability from pH 4.5 to 4.8. Apparent K m and V max are 7.1 mM and 17.79 M min−1, respectively.  相似文献   

17.
This study examined whether ‘Risnod2’ and ‘Risnod27’ non-nodulating mutants of pea (Pisum sativum L.) provided with increasing concentrations of nitrate could achieve a growth and nitrogen accumulation comparable to their parental N2-fixing cv. Finale. In the cv. Finale, nodule number, nodule dry mass accumulation, total C2H2-reducing activity of nodulated roots (TAR) and estimated N2 fixation were considerably inhibited at 5.0 and 10.0 mM root medium NO3 concentrations. In contrast a 0.63 mM level stimulated both the nodule dry mass and TAR. The cv. Finale N2-fixing plants grown on 0 to 2.5 mM NO3 levels had higher shoot N concentrations than the Nod mutants, but within the 5.0 to 10.0 mM levels the Nod mutants approached or even overtopped the N concentration of the cv. Finale plants. Compared with a high positive correlation found in the Nod mutants, shoot N concentration in the cv. Finale was negatively correlated with the root medium NO3 concentration. The pattern of nitrogen content in shoot dry mass was very similar to that seen in the shoot dry mass accumulation. The Nod mutants grown on the 5.0 and/or 10.0 mM NO3 level had plant dry mass, shoot nitrogen concentration, shoot nitrogen content, and root/shoot dry mass ratio comparable with those of the nodulating cv. Finale grown on the same nitrate levels.  相似文献   

18.
A Bacillus sp. RE was resistant to chromium and reduced Cr(VI) without accumulating chromium inside the cell. When Cr(VI) was 10 and 40 μg ml−1, >95% of the total Cr(VI) was reduced in 24 and 72 h of growth, respectively, whereas at 80 μg Cr(VI) ml−1 only 50% of Cr(VI) was reduced. However growth was not affected; the cell mass was 0.7–0.8 mg ml−1 in all cases. The cell-free extract showed Cr(VI) reducing enzyme activity which was enhanced (>5 fold) by NADH and NADPH. Like whole cells the enzyme also reduced Cr(VI) with decreasing efficiency on increasing Cr(VI) concentration. The enzyme activity was optimal at pH 6.0 and 30 °C. The enzyme was stable up to 30 °C and from pH 5.5 to 8, but from pH 4 to 5 the enzyme was severely destabilized. Its Km and Vmax were 14 μm and 3.8 nmol min−1 mg−1 respectively. The enzyme activity was enhanced by Cu2+ and Ni2+ and inhibited by Hg2+. Received 21 September 2005; Revisions requested 5 October 2005; Revisions received 16 November 2005; Accepted 16 November 2005  相似文献   

19.
Yang CF  Lee CM 《Biodegradation》2008,19(3):329-336
The objectives of this research were to monitor the variations of species in mixed cultures during the enrichment period, isolate species and identify and characterize the pure 4-chlorophenol (4-CP) degrading strains from enriched mixed cultures. Strain Rhizobium sp. 4-CP-20 was isolated from the acclimated mixed culture. The DGGE result indicated that strain Rhizobium sp. 4-CP-20 was undetectable at the beginning but detectable after 2 weeks of enrichment. The optimum growth temperatures for Rhizobium sp. 4-CP-20 were both 36°C using 350 mg l−1 glucose or sodium acetate as the substrate. The optimum pH range for degrading 100 mg l−1 4-CP was between 6.89 and 8.20. Strain Rhizobium sp. 4-CP-20 could degrade 4-CP completely within 3.95 days, as the initial 4-CP concentration was 100 mg l−1. If the initial 4-CP concentration was higher than 240 mg l−1, the growth of bacterial cells and the activity of degrading 4-CP were both inhibited.  相似文献   

20.
Composition and accumulation of secondary carotenoids in Chlorococcum sp.   总被引:4,自引:0,他引:4  
A locally isolated Chlorococcum sp. could accumulate astaxanthin and its esters as secondary carotenoids. The secondary carotenoids could reach a concentration of 5.2 mg g−1 d. wt, and were located in the cytoplasm and chloroplast as globules. Cells grew best at pH 8.0 and 30 °C, at which the growth rate was about 0.066 h−1. Acidic condition (pH 5.5 and 6.5) and slightly elevated temperature (35 °C) enhanced the cellular accumulation of astaxanthin. Outdoor studies indicated that Chlorococcum sp. grew well in a tubular photobioreactor. In medium containing 2 mM and 10 mM NH4CI, the cellular contents of total secondary carotenoids and astaxanthin reached similar levels (5.0 mg g−1 d. wt and 2.0 mg g−1 d. wt, respectively) in the 15 days of cultivation, while the yield of total secondary carotenoids and astaxanthin in 10 mM NH4CI were higher (45 mg L−1 and 18 mg L−1, respectively). The advantages of tolerance to high temperature and extreme pH values, relative fast growth rate and ease of cultivation in outdoor system suggest that Chlorococcum sp. could be a potential candidate for mass production of secondary carotenoids in particular astaxanthin. This revised version was published online in September 2006 with corrections to the Cover Date.  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