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 15 毫秒
1.
A heterologous phosphoribulokinase (PRK) gene probe was used to analyze two recombinant plasmids isolated from a Rhodopseudomonas (Rhodobacter) sphaeroides gene library. These plasmids were previously shown to carry the genes for form I and form II ribulose 1,5-bisphosphate carboxylase/oxygenase (RuBPC/O). Southern blot hybridization analysis indicated that there were two PRK genes linked to the RuBPC/O coding sequences. Restriction mapping showed the arrangement of the duplicate sets of PRK and RuBPC/O to be distinct. Subcloning of the hybridizing PRK sequences downstream of the lac promoter of pUC8 allowed expression of the two PRK enzymes in Escherichia coli. Analysis of the purified proteins by sodium dodecyl sulfate-slab gel electrophoresis revealed polypeptides with molecular weights of 32,000 and 34,000 corresponding to the form I and form II PRKs, respectively. Preliminary experiments on sensitivity to NADH regulation suggested that the two PRK enzymes differ in catalytic properties.  相似文献   

2.
3.
We purified two isozymes of coniferyl alcohol dehydrogenase (CADH I and II) to homogeneity from cell-free extracts of Streptomyces sp. NL15-2K. The apparent molecular masses of CADH I and II were determined to be 143 kDa and 151 kDa respectively by gel filtration, whereas their subunit molecular masses were determined to be 35,782.2 Da and 37,597.7 Da respectively by matrix-assisted laser-desorption ionization time-of-flight mass spectrometry (MALDI-TOF-MS). Thus, it is probable that both isozymes are tetramers. The optimum pH and temperature for coniferyl alcohol dehydrogenase activity were pH 9.5 and 45 °C for CADH I and pH 8.5 and 40 °C for CADH II. CADH I oxidized various aromatic alcohols and allyl alcohol, and was most efficient on cinnamyl alcohol, whereas CADH II exhibited high substrate specificity for coniferyl alcohol, and showed no activity as to the other alcohols, except for cinnamyl alcohol and 3-(4-hydroxy-3-methoxyphenyl)-1-propanol. In the presence of NADH, CADH I and II reduced cinnamaldehyde and coniferyl aldehyde respectively to the corresponding alcohols.  相似文献   

4.
We purified two isozymes of coniferyl alcohol dehydrogenase (CADH I and II) to homogeneity from cell-free extracts of Streptomyces sp. NL15-2K. The apparent molecular masses of CADH I and II were determined to be 143 kDa and 151 kDa respectively by gel filtration, whereas their subunit molecular masses were determined to be 35,782.2 Da and 37,597.7 Da respectively by matrix-assisted laser-desorption ionization time-of-flight mass spectrometry (MALDI-TOF-MS). Thus, it is probable that both isozymes are tetramers. The optimum pH and temperature for coniferyl alcohol dehydrogenase activity were pH 9.5 and 45 °C for CADH I and pH 8.5 and 40 °C for CADH II. CADH I oxidized various aromatic alcohols and allyl alcohol, and was most efficient on cinnamyl alcohol, whereas CADH II exhibited high substrate specificity for coniferyl alcohol, and showed no activity as to the other alcohols, except for cinnamyl alcohol and 3-(4-hydroxy-3-methoxyphenyl)-1-propanol. In the presence of NADH, CADH I and II reduced cinnamaldehyde and coniferyl aldehyde respectively to the corresponding alcohols.  相似文献   

5.
Although the structures of mammalian cytosolic and mitochondrial ALDH have been determined, several differences, mainly functional, between these two 70% identical isozymes remain unexplained. A major difference is the differential effect of Mg(2+) ions that inhibits the cytosolic and activates the mitochondrial isozyme. Here, we have investigated the effect of Mg(2+) ions on each individual kinetic step of ALDH1 and ALDH2. The metal ions were found not to affect either acylation or hydride transfer for either isozyme. The lack of a Mg(2+) ion effect on hydride transfer was further demonstrated with an E399Q mutant of ALDH1 whose rate-limiting step had been changed from NADH dissociation to hydride transfer. The other steps, however, were affected by Mg(2+) ions for both isozymes. The metal ions inhibited NADH dissociation, the rate-limiting step for ALDH1, and enhanced deacylation, the rate-limiting step for ALDH2. Our results indicated that, with both isozymes, Mg(2+) ions tightened the binding of NADH, and by binding to the coenzyme, they increased the nucleophilicity of the nucleophile Cys302. The inhibition of ALDH1 and activation of ALDH2 at pH 7.4 are due to their different rate-limiting steps. Mg(2+) ions affected similarly the NADH activation of the esterase reaction for both isozymes. In contrast, the metal ions affected only the NAD(+) activation of ALDH1. This latter finding and other features described here can be rationalized on the basis of the known three-dimensional structures of the isozymes.  相似文献   

6.
ATP7B, the Wilson disease-associated Cu(I)-transporter, and ZntA from Escherichia coli are soft metal P1-type ATPases with mutually exclusive metal ion substrates. P1-type ATPases have a distinctive amino-terminal domain containing the conserved metal-binding motif GXXCXXC. ZntA has one copy of this motif while ATP7B has six copies. The effect of interchanging the amino-terminal domains of ATP7B and ZntA was investigated. Chimeric proteins were constructed in which either the entire amino-terminal domain of ATP7B or only its sixth metal-binding motif replaced the amino-terminal domain of ZntA. Both chimeras conferred resistance to lead, zinc, and cadmium salts but not to copper salts. The purified chimeras displayed activity with lead, cadmium, zinc, and mercury, which are substrates of ZntA. There was no activity with copper or silver, which are substrates of ATP7B. The chimeras were 2-3-fold less active than ZntA. Thus, the amino-terminal domain of P1-type ATPases cannot alter the metal specificity determined by the transmembrane segment. Also, these results suggest that this domain interacts with the rest of the transporter in a metal ion-specific manner; the amino-terminal domain of ATP7B cannot replace that of ZntA in restoring full catalytic activity.  相似文献   

7.
Lactate dehydrogenase [L-lactate: NAD oxidoreductase, EC 1.1.1.27] was isolated from sweet potato root tissues. Two species of the enzyme (isozymes I and II) were separated by DE-52 cellulose column chromatography from healthy, cut, and black-rot diseased tissues. Isozymes I and II were purified from healthy and diseased tissues, respectively. Reduction of pyruvate by NADH with either isozyme I or II was inhibited by pyruvate at high concentrations, by NAD+ and by several mononucleotides. Isozyme I was inhibited by a lower concentration of adenine nucleotide than isozyme II, and Km for pyruvate was increased markedly at acidic pH in the case of isozyme I, but only slightly in the case of isozyme II. The molecular weights of both isozymes were determined to be 150,000 and they were found to be charge isomers by polyacrylamide gel electrophoresis. The enzyme activity increased in response to infection by black-rot fungus but decreased in response to cutting.  相似文献   

8.
Detailed in vitro comparisons of the biochemical characteristics of three protein kinase C isozymes were performed. As an alternative to earlier uncertain separation methods and expression schemes, highly purified and genetically distinct protein kinase C enzymes were produced using the baculovirus expression system. The baculovirus expression system yielded approximately 200-300 micrograms of the purified isozyme from 3 x 10(8) (100 ml of culture medium) baculovirus-infected insect cells. Biochemical characterization of the expressed isozymes indicated that the three isozymes had virtually indistinguishable Ca2+, Mg2+, and ATP dependencies. However, in certain critical functional characteristics such as phosphatidylserine dependencies, phospholipid and substrate preferences, and arachidonic acid activation, the gamma isozyme exhibited distinctive properties when compared with both the alpha and beta II subtypes. In addition, the activity of the beta II subtype was more dependent upon diacylglycerol or phorbol esters for activation than either the alpha or gamma isoforms. The alpha isozyme, unlike the beta II and gamma forms, was totally dependent on Ca2+ for activation in the presence of free arachidonic acid. These studies provide definitive characterizations of the pure isoforms; many of the findings were consistent with earlier enzymatic observations using hydroxyapatite-purified isoforms. Thus, the distinctive biochemical properties of the protein kinase C isozymes are consistent with the hypothesis that each isoform may have distinct roles in signal transduction processes.  相似文献   

9.
A series of chimeric receptors was generated between the epidermal growth factor (EGF) receptor, ErbB-1, and its homologue, ErbB-4, to investigate the roles of the extracellular domains (I-IV) in the ligand specificities. As compared with ErbB-1 and the chimeras with both domains I and III of ErbB-1, the chimeras with only one of these domains exhibited reduced binding of 125I-labeled EGF. Particularly, the contribution of domain III was appreciably larger than that of domain I of ErbB-1 in 125I-labeled EGF binding. Nevertheless, the chimeras with domain III of ErbB-1 and domain I of ErbB-4 were prevented from binding to 125I-labeled EGF competitively by the ErbB-4 ligand, neuregulin (NRG). On the other hand, NRG did not compete with 125I-labeled EGF for binding to the chimeras with the ErbB-1 domain I and the ErbB-4 domain III. Therefore, NRG binding to ErbB-4 depends much more on domain I than on domain III. With respect to autophosphorylation and subsequent ERK activation, EGF activated the chimeras with either domain I or III of ErbB-1. In contrast, NRG activated the chimeras with the ErbB-4 domain I and the ErbB-1 domain III, but not those with the ErbB-1 domain I and the ErbB-4 domain III. Therefore, the relative contributions between domains I and III of ErbB-4 in the NRG signaling are different from those of ErbB-1 in the EGF signaling.  相似文献   

10.
Two isozymes of carbonic anhydrase (CA I and CA II) were quantified by a radio-immunoassay in 10 different tissues of the pig-tailed macaque. There were clearly differences in relative amounts of the two isozymes, indicating a differential regulation of these two different gene products. An inherited deficiency variant reduced red cell CA I and CA II 5000-fold and 2.7-fold, respectively. In nine other tissues, CA I was reduced from approximately twofold to 110-fold, and CA II was essentially unchanged. The CA I in deficient red cells was immunochemically and electrophoretically identical to common electrophoretic variants of CA I in the pig-tailed macaque and was enzymatically active.This work was part of a doctoral dissertation submitted in partial fulfillment of the Doctor of Philosophy degree in the Horace H. Rackham School of Graduate Studies at The University of Michigan. Supported by NIH training grant 5-T01-GM-71-11 and NIH research grant GM-15419.  相似文献   

11.
Citrate synthase (citrate-oxaloacetate lyase (CoA acetylating), EC 4.1.3.7) has been purified to electrophoretic homogeneity from a marine Pseudomonas. The enzyme was made up of identical subunits, with a molecular wieght of about 53 000, as determined by sodium dodecyl sulphate - polyacrylamide gel electrophoresis. The native enzyme (citrate synthase II, CS II) could be dissociated by dialysis against 20 mM phosphate (Pi), pH 7; the enzyme thus obtained (citrate synthase I, CS I) was still active, but presented different molecular weight and kinetic and regulatory properties. CS II was activated by adenosine monophosphate (AMP), Pi, and KCl, and inhibited by reduced nicotinamide adenine dinucleotide (NADH), being apparently insensitive to adenosine triphosphate (ATP) and adenosine diphosphate (ADP). The inhibition by NADH was completely counteracted by 0.1 mM AMP, but not by 50 mM Pi or 0.1 M KCl. The activation by KCl and Pi, or by KCl and AMP was nearly additive, whereas that by AMP and Pi was not. The activators acted essentially by increasing Vmax, although they also caused a decrease in the Km values. CS I was inhibited by ATP, ADP, AMP, and KCl, and was insensitive to NADH. CS I could be reassociated after elimination of Pi by dialysis, regaining the higher molecular weight and the activation by AMP characteristic of CS II.  相似文献   

12.
Protein carboxymethylase (EC 2.1.1.24) from cytosol of bovine brain was found to exist as two apparent isozymes that could be separated by chromatography on DEAE-cellulose at pH 8.O. Rechromatography of the two forms, designated PCM I and PCM II, indicated that they are not interconvertible. Both enzymes have a molecular weight of 24,300 by sodium dodecyl sulfate-polyacryl-amide gel electrophoresis. PCM I consists mainly of one isoelectric form, pI 6.5, whereas PCM II resolves into two forms of pI 5.6 and 5.7. The relative amounts of PCM I and PCM II show a marked tissue dependence. Brain has approximately twice as much PCM I as II, whereas liver contains only the type II enzyme. The two enzymes were found to have similar substrate specificities when tested with five different methyl-accepting proteins. Synapsin I, a basic protein associated with synaptic vesicles, was found to be an excellent methyl-accepting protein with regard to its Km (1.2 μM), but it exhibited a low stoichiome-try of methyl incorporation.  相似文献   

13.
Interactions among the three structural domains of Bacillus thuringiensis Cry1 toxins were investigated by functional analysis of chimeric proteins. Hybrid genes were prepared by exchanging the regions coding for either domain I or domain III among Cry1Ab, Cry1Ac, Cry1C, and Cry1E. The activity of the purified trypsin-activated chimeric toxins was evaluated by testing their effects on the viability and plasma membrane permeability of Sf9 cells. Among the parental toxins, only Cry1C was active against these cells and only chimeras possessing domain II from Cry1C were functional. Combination of domain I from Cry1E with domains II and III from Cry1C, however, resulted in an inactive toxin, indicating that domain II from an active toxin is necessary, but not sufficient, for activity. Pores formed by chimeric toxins in which domain I was from Cry1Ab or Cry1Ac were slightly smaller than those formed by toxins in which domain I was from Cry1C. The properties of the pores formed by the chimeras are therefore likely to result from an interaction between domain I and domain II or III. Domain III appears to modulate the activity of the chimeric toxins: combination of domain III from Cry1Ab with domains I and II of Cry1C gave a protein which was more strongly active than Cry1C.  相似文献   

14.
We examined the metabolic responses of the hypoxia-tolerant killifish (Fundulus heteroclitus) to 15 h of severe hypoxia and recovery with emphasis on muscle substrate usage and the regulation of the mitochondrial protein pyruvate dehydrogenase (PDH), which controls carbohydrate oxidation. Hypoxia survival involved a transient activation of substrate-level phosphorylation in muscle (decreases in [creatine phospate] and increases in [lactate]) during which time mechanisms to reduce overall ATP consumption were initiated. This metabolic transition did not affect total cellular [ATP], but had an impact on cellular energy status as indicated by large decreases in [ATP]/[ADP(free)] and [ATP]/[AMP(free)] and a significant loss of phosphorylation potential and Gibbs free energy of ATP hydrolysis (DeltafG'). The activity of PDH was rapidly (within 3 h) decreased by approximately 50% upon hypoxia exposure and remained depressed relative to normoxic samples throughout. Inactivation of PDH was primarily mediated via posttranslational modification following the accumulation of acetyl-CoA and subsequent activation of pyruvate dehydrogenase kinase (PDK). Estimated changes in cytoplasmic and mitochondrial [NAD(+)]/[NADH] did not parallel one another, suggesting the mitochondrial NADH shuttles do not function during hypoxia exposure. Large increases in the expression of PDK (PDK isoform 2) were consistent with decreased PDH activity; however, these changes in mRNA were not associated with changes in total PDK-2 protein content assessed using mammalian antibodies. No other changes in the expression of other known hypoxia-responsive genes (e.g., lactate dehydrogenase-A or -B) were observed in either muscle or liver.  相似文献   

15.
We have studied two aspects of calcium channel activation. First, we investigated the molecular regions that are important in determining differences in activation between low- and high-voltage activated channels. For this, we made chimeras between the low-voltage activating CaV3.1 channel and the high-voltage activating CaV1.2 channel. Chimeras were expressed in oocytes, and calcium channel currents recorded by voltage clamp. For domain I, we found that the molecular region that is important in determining the voltage dependence of activation comprises the pore regions S5-P as well as P-S6, but surprisingly not the voltage sensor S1–S4 region, which might have been expected to play a major part. By contrast, the smaller, but still significant, modulating effects of domain II on activation properties were due to effects involving both S1–S4 and S5–S6 but not the I/II linker. Second, during channel activation we studied movement of the S4 segment in domain I of one of the chimeras, using cysteine-scanning mutagenesis. The reagent parachloromercuribenzensulfonate inhibited currents for mutants V263, A265, L266 and A268, but not for F269 and V271, and voltage dependence of inhibition for residue V263 indicated S4 movement, which occurred before channel opening. The data indicate movement outwards upon depolarisation so as to expose amino acids up to residue 268 in S4.Junying Li and Louisa Stevens contributed equally to this work.  相似文献   

16.
Interactions among the three structural domains of Bacillus thuringiensis Cry1 toxins were investigated by functional analysis of chimeric proteins. Hybrid genes were prepared by exchanging the regions coding for either domain I or domain III among Cry1Ab, Cry1Ac, Cry1C, and Cry1E. The activity of the purified trypsin-activated chimeric toxins was evaluated by testing their effects on the viability and plasma membrane permeability of Sf9 cells. Among the parental toxins, only Cry1C was active against these cells and only chimeras possessing domain II from Cry1C were functional. Combination of domain I from Cry1E with domains II and III from Cry1C, however, resulted in an inactive toxin, indicating that domain II from an active toxin is necessary, but not sufficient, for activity. Pores formed by chimeric toxins in which domain I was from Cry1Ab or Cry1Ac were slightly smaller than those formed by toxins in which domain I was from Cry1C. The properties of the pores formed by the chimeras are therefore likely to result from an interaction between domain I and domain II or III. Domain III appears to modulate the activity of the chimeric toxins: combination of domain III from Cry1Ab with domains I and II of Cry1C gave a protein which was more strongly active than Cry1C.  相似文献   

17.
Two isozymes of horse liver aldehyde dehydrogenase (aldehyde, NAD oxidoreductase (EC 1.2.1.3)), F1 and F2, have been purified to homogeneity using salt fractionation followed by ion exchange and gel filtration chromatography. The specific activities of the two isozymes in a pH 9.0 system with propionaldehyde as substrate were approximately 0.35 and 1.0 mumol of NADH/min/mg of protein for the F1 and F2 isozymes, respectively. The multiporosity polyacrylamide gel electrophoresis molecular weights of the F1 and F2 isozymes were approximately 230,000 and 240,000 respectively. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis gave subunit molecular weight estimates of 52,000 and 53,000 for the F1 and F2 isozymes, respectively. The amino acid compositions of the two isozymes were found to be similar; the ionizable amino acid contents being consistent with the electrophoretic and chromatographic behavior of the two isozymes. Both isozymes exhibited a broad aldehyde specificity, oxidizing a wide variety of aliphatic and aromatic aldehydes and utilized NAD as coenzyme, but at approximately 300-fold higher coenzyme concentration could use NADP. The F1 isozyme exhibited a very low Km for NAD (3 muM) and a higher Km for acetaldehyde (70 muM), while the F2 isozyme was found to have a higher Km for NAD (30 muM) and a low Km for acetaldehyde (0.2 muM). The two isozymes showed similar chloral hydrate and p-chloromercuribenzoate inhibition characteristics, but the F1 isozyme was found to be several orders of magnittude more sensitive to disulfiram, a physiological inhibitor of acetaldehyde oxidation. Based on its disulfiram inhibition characteristics, it has been suggested that the F1 isozyme may be the primary enzyme for oxidizing the acetyldehyde produced during ethanol oxidation in vivo.  相似文献   

18.
Marri L  Trost P  Pupillo P  Sparla F 《Plant physiology》2005,139(3):1433-1443
Calvin cycle enzymes glyceraldehyde-3-phosphate dehydrogenase (GAPDH) and phosphoribulokinase (PRK) form together with the regulatory peptide CP12 a supramolecular complex in Arabidopsis (Arabidopsis thaliana) that could be reconstituted in vitro using purified recombinant proteins. Both enzyme activities were strongly influenced by complex formation, providing an effective means for regulation of the Calvin cycle in vivo. PRK and CP12, but not GapA (A(4) isoform of GAPDH), are redox-sensitive proteins. PRK was reversibly inhibited by oxidation. CP12 has no enzymatic activity, but it changed conformation depending on redox conditions. GapA, a bispecific NAD(P)-dependent dehydrogenase, specifically formed a binary complex with oxidized CP12 when bound to NAD. PRK did not interact with either GapA or CP12 singly, but oxidized PRK could form with GapA/CP12 a stable ternary complex of about 640 kD (GapA/CP12/PRK). Exchanging NADP for NAD, reducing CP12, or reducing PRK were all conditions that prevented formation of the complex. Although GapA activity was little affected by CP12 alone, the NADPH-dependent activity of GapA embedded in the GapA/CP12/PRK complex was 80% inhibited in respect to the free enzyme. The NADH activity was unaffected. Upon binding to GapA/CP12, the activity of oxidized PRK dropped from 25% down to 2% the activity of the free reduced enzyme. The supramolecular complex was dissociated by reduced thioredoxins, NADP, 1,3-bisphosphoglycerate (BPGA), or ATP. The activity of GapA was only partially recovered after complex dissociation by thioredoxins, NADP, or ATP, and full GapA activation required BPGA. NADP, ATP, or BPGA partially activated PRK, but full recovery of PRK activity required thioredoxins. The reversible formation of the GapA/CP12/PRK supramolecular complex provides novel possibilities to finely regulate GapA ("non-regulatory" GAPDH isozyme) and PRK (thioredoxin sensitive) in a coordinated manner.  相似文献   

19.
Voltage-dependent calcium channels are classified into low voltage-activated and high voltage-activated channels. We have investigated the molecular basis for this difference in voltage dependence of activation by constructing chimeras between a low voltage-activated channel (Ca(V)3.1) and a high voltage-activated channel (Ca(V)1.2), focusing on steady-state activation properties. Wild type and chimeras were expressed in oocytes, and two-electrode voltage clamp recordings were made of calcium channel currents. Replacement of domains I, III, or IV of the Ca 3.1 channel with the corresponding domain of Ca(V)1.2 led (V)to high voltage-activated channels; for these constructs the current/voltage (I/V) curves were similar to those for Ca(V)1.2 wild type. However, replacement of domain II gave only a small shift to the right of the I/V curve and modulation of the activation kinetics but did not lead to a high voltage-activating channel with an I/V curve like Ca 1.2. We also investigated the role of the voltage sensor (V)S4 by replacing the S4 segment of Ca(V)3.1 with that of Ca 1.2. For domain I, there was no shift in the I/V curve (V)as compared with Ca(V)3.1, and there were relatively small shifts to the right for domains III and IV. Taken together, these results suggest that domains I, III, and IV (rather than domain II) are apparently critical for channel opening and, therefore, contribute strongly to the difference in voltage dependence of activation between Ca 3.1 and Ca(V)1.2. However, the S4 segments in domains I, (V)III, and IV did not account for this difference in voltage dependence.  相似文献   

20.
Two distinct ferredoxins, Fd I and Fd II, were isolated and purified to homogeneity from photoautotrophically grown Chlorobium tepidum, a moderately thermophilic green sulfur bacterium that assimilates carbon dioxide by the reductive tricarboxylic acid cycle. Both ferredoxins serve a crucial role as electron donors for reductive carboxylation, catalyzed by a key enzyme of this pathway, pyruvate synthase/pyruvate ferredoxin oxidoreductase. The reduction potentials of Fd I and Fd II were determined by cyclic voltammetry to be -514 and -584 mV, respectively, which are more electronegative than any previously studied Fds in which two [4Fe-4S] clusters display a single transition. Further spectroscopic studies indicated that the CD spectrum of oxidized Fd I closely resembled that of Fd II; however, both spectra appeared to be unique relative to ferredoxins studied previously. Double integration of the EPR signal of the two Fds yielded approximately approximately 2.0 spins per molecule, compatible with the idea that C. tepidum Fd I and Fd II accept 2 electrons upon reduction. These results suggest that the C. tepidum Fd I and Fd II polypeptides each contain two bound [4Fe-4S] clusters. C. tepidum Fd I and Fd II are novel 2[4Fe-4S] Fds, which were shown previously to function as biological electron donors or acceptors for C. tepidum pyruvate synthase/pyruvate ferredoxin oxidoreductase (Yoon, K.-S., Hille, R., Hemann, C. F., and Tabita, F. R. (1999) J. Biol. Chem. 274, 29772-29778). Kinetic measurements indicated that Fd I had approximately 2.3-fold higher affinity than Fd II. The results of amino acid sequence alignments, molecular modeling, oxidation-reduction potentials, and spectral properties strongly indicate that the C. tepidum Fds are chimeras of both clostridial-type and chromatium-type Fds, suggesting that the two Fds are likely intermediates in the evolutional development of 2[4Fe-4S] clusters compared with the well described clostridial and chromatium types.  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