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.
Elucidating protein function from its structure is central to the understanding of cellular mechanisms. This involves deciphering the dependence of local structural motifs on sequence. These structural motifs may be stabilized by direct or water‐mediated hydrogen bonding among the constituent residues. π‐Turns, defined by interactions between (i) and (i + 5) positions, are large enough to contain a central space that can embed a water molecule (or a protein moiety) to form a stable structure. This work is an analysis of such embedded π‐turns using a nonredundant dataset of protein structures. A total of 2965 embedded π‐turns have been identified, as also 281 embedded Schellman motif, a type of π‐turn which occurs at the C‐termini of α‐helices. Embedded π‐turns and Schellman motifs have been classified on the basis of the protein atoms of the terminal turn residues that are linked by the embedded moiety, conformation, residue composition, and compared with the turns that have terminal residues connected by direct hydrogen bonds. Geometrically, the turns have been fitted to a circle and the position of the linker relative to its center analyzed. The hydroxyl group of Ser and Thr, located at (i + 3) position, is the most prominent linker for the side‐chain mediated π‐turns. Consideration of residue conservation among homologous sequences indicates the terminal and the linker positions to be the most conserved. The embedded π‐turn as a binding site (for the linker) is discussed in the context of “nest,” a concave depression that is formed in protein structures with adjacent residues having enantiomeric main‐chain conformations. © 2013 Wiley Periodicals, Inc. Biopolymers 101: 441–453, 2014.  相似文献   

2.
The two-component alkanesulfonate monooxygenase system from Escherichia coli is comprised of an FMN reductase (SsuE) and a monooxygenase enzyme (SsuD) that together catalyze the oxidation of alkanesulfonate to the corresponding aldehyde and sulfite products. To determine the effects of protein interactions on catalysis, the steady-state kinetic parameters for SsuE were determined in single-enzyme assays and in the presence of the monooxygenase enzyme and alkanesulfonate substrate. In single-enzyme kinetic assays, SsuE followed an ordered sequential mechanism, with NADPH as the first substrate to bind and NADP+ as the last product to dissociate. However, in the presence of SsuD and octanesulfonate the kinetic mechanism of SsuE is altered to a rapid equilibrium ordered mechanism, and the Km value for FMN is increased 10-fold. These results suggest that both the SsuD enzyme and alkanesulfonate substrate are required to ensure that the FMN reductase reaction proceeds to form the ternary complex with the subsequent generation of reduced flavin transfer.  相似文献   

3.
Short‐chain dehydrogenase/reductase (SDR) is distributed in many organisms, from bacteria to humans, and has significant roles in metabolism of carbohydrates, lipids, amino acids, and other biomolecules. An important intermediate in acidic polysaccharide metabolism is 2‐keto‐3‐deoxy‐d ‐gluconate (KDG). Recently, two short and long loops in Sphingomonas KDG‐producing SDR enzymes (NADPH‐dependent A1‐R and NADH‐dependent A1‐R′) involved in alginate metabolism were shown to be crucial for NADPH or NADH coenzyme specificity. Two SDR family enzymes—KduD from Pectobacterium carotovorum (PcaKduD) and DhuD from Streptococcus pyogenes (SpyDhuD)—prefer NADH as coenzyme, although only PcaKduD can utilize both NADPH and NADH. Both enzymes reduce 2,5‐diketo‐3‐deoxy‐d ‐gluconate to produce KDG. Tertiary and quaternary structures of SpyDhuD and PcaKduD and its complex with NADH were determined at high resolution (approximately 1.6 Å) by X‐ray crystallography. Both PcaKduD and SpyDhuD consist of a three‐layered structure, α/β/α, with a coenzyme‐binding site in the Rossmann fold; similar to enzymes A1‐R and A1‐R′, both arrange the two short and long loops close to the coenzyme‐binding site. The primary structures of the two loops in PcaKduD and SpyDhuD were similar to those in A1‐R′ but not A1‐R. Charge neutrality and moderate space at the binding site of the nucleoside ribose 2′ coenzyme region were determined to be structurally crucial for dual‐coenzyme specificity in PcaKduD by structural comparison of the NADH‐ and NADPH‐specific SDR enzymes. The corresponding site in SpyDhuD was negatively charged and spatially shallow. This is the first reported study on structural determinants in SDR family KduD related to dual‐coenzyme specificity. Proteins 2016; 84:934–947. © 2016 Wiley Periodicals, Inc.  相似文献   

4.
The alkanesulfonate monooxygenase system from Escherichia coli is involved in scavenging sulfur from alkanesulfonates under sulfur starvation. An FMN reductase (SsuE) catalyzes the reduction of FMN by NADPH, and the reduced flavin is transferred to the monooxygenase (SsuD). Rapid reaction kinetic analyses were performed to define the microscopic steps involved in SsuE catalyzed flavin reduction. Results from single-wavelength analyses at 450 and 550 nm showed that reduction of FMN occurs in three distinct phases. Following a possible rapid equilibrium binding of FMN and NADPH to SsuE (MC-1) that occurs before the first detectable step, an initial fast phase (241 s(-1)) corresponds to the interaction of NADPH with FMN (CT-1). The second phase is a slow conversion (11 s(-1)) to form a charge-transfer complex of reduced FMNH(2) with NADP(+) (CT-2), and represents electron transfer from the pyridine nucleotide to the flavin. The third step (19 s(-1)) is the decay of the charge-transfer complex to SsuE with bound products (MC-2) or product release from the CT-2 complex. Results from isotope studies with [(4R)-(2)H]NADPH demonstrates a rate-limiting step in electron transfer from NADPH to FMN, and may imply a partial rate-limiting step from CT-2 to MC-2 or the direct release of products from CT-2. While the utilization of flavin as a substrate by the alkanesulfonate monooxygenase system is novel, the mechanism for flavin reduction follows an analogous reaction path as standard flavoproteins.  相似文献   

5.
Flavodoxins in combination with the flavin mononucleotide (FMN) cofactor play important roles for electron transport in prokaryotes. Here, novel insights into the FMN‐binding mechanism to flavodoxins‐4 were obtained from the NMR structures of the apo‐protein from Lactobacillus acidophilus (YP_193882.1) and comparison of its complex with FMN. Extensive reversible conformational changes were observed upon FMN binding and release. The NMR structure of the FMN complex is in agreement with the crystal structure (PDB ID: 3EDO ) and exhibits the characteristic flavodoxin fold, with a central five‐stranded parallel β–sheet and five α‐helices forming an α/β‐sandwich architecture. The structure differs from other flavoproteins in that helix α2 is oriented perpendicular to the β‐sheet and covers the FMN‐binding site. This helix reversibly unfolds upon removal of the FMN ligand, which represents a unique structural rearrangement among flavodoxins.  相似文献   

6.
Bilirubin is a potent antioxidant that is produced from the reduction of the heme degradation product biliverdin. In mammalian cells and Cyanobacteria, NADH/NADPH‐dependent biliverdin reductases (BVRs) of the Rossmann‐fold have been shown to catalyze this reaction. Here, we describe the characterization of Rv2074 from Mycobacterium tuberculosis, which belongs to a structurally and mechanistically distinct family of F420H2‐dependent BVRs (F‐BVRs) that are exclusively found in Actinobacteria. We have solved the crystal structure of Rv2074 bound to its cofactor, F420, and used this alongside molecular dynamics simulations, site‐directed mutagenesis and NMR spectroscopy to elucidate its catalytic mechanism. The production of bilirubin by Rv2074 could exploit the anti‐oxidative properties of bilirubin and contribute to the range of immuno‐evasive mechanisms that have evolved in M. tuberculosis to allow persistent infection.  相似文献   

7.
Three new thieno[3,2‐b][1]benzothiophene ( TBT )‐based donor–π–acceptor (D–π–A) sensitizers, coded as SGT ‐ 121 , SGT ‐ 129 , and SGT ‐ 130 , have been designed and synthesized for dye‐sensitized solar cells (DSSCs), for the first time. The TBT , prepared by fusing thiophene unit with the phenyl unit of triphenylamine donor, is utilized as the π‐bridge for all sensitizers with good planarity. They have been molecularly engineered to regulate the highest occupied molecular orbital (HOMO)‐lowest unoccupied molecular orbital (LUMO) energy levels and extend absorption range as well as to control the electron‐transfer process that can ensure efficient dye regeneration and prevent undesired electron recombination. The photovoltaic performance of SGT‐sensitizer‐based DSSCs employing Co(bpy)32+/3+ (bpy = 2,2′‐bipyridine) redox couple is systematically evaluated in a thorough comparison with Y123 as a reference sensitizer. Among them, SGT ‐ 130 with benzothiadiazole‐phenyl ( BTD ‐ P ) unit as an auxiliary acceptor exhibits the highest power‐conversion efficiency (PCE) of 10.47% with Jsc = 16.77 mA cm?2, Voc = 851 mV, and FF = 73.34%, whose PCE is much higher than that of Y123 (9.5%). It is demonstrated that the molecular combination of each fragment in D–π–A organic sensitizers can be a pivotal factor for achieving the higher PCEs and an innovative strategy for strengthening the drawbacks of the π‐bridge.  相似文献   

8.
The role aromatic amino acids play in the formation of amyloid is a subject of controversy. In an effort to clarify the contribution of aromaticity to the self‐assembly of human islet amyloid polypeptide (hIAPP)22‐29, peptide analogs containing electron donating groups (EDGs) or electron withdrawing groups (EWGs) as substituents on the aromatic ring of Phe‐23 at the para position have been synthesized and characterized using turbidity measurements in conjunction with Raman and fluorescence spectroscopy. Results indicate the incorporation of EDGs on the aromatic ring of Phe‐23 virtually abolish the ability of hIAPP22‐29 to form amyloid. Peptides containing EWGs were still capable of forming aggregates. These aggregates were found to be rich in β‐sheet secondary structure. Transmission electron microscopy images of the aggregates confirm the presence of amyloid fibrils. The observed difference in amyloidogenic propensity between peptides containing EDGs and those with EWGs appears not to be based on differences in peptide hydrophobicity. Fluorescence and Raman spectroscopic investigations reveal that the environment surrounding the aromatic ring becomes more hydrophobic and ordered upon aggregation. Furthermore, Raman measurements of peptide analogs containing EWGs, conclusively demonstrate a distinct downshift in the ? C?C? ring mode (ca. 1600 cm?1) upon aggregation that has previously been shown to be indicative of π‐stacking. While previous work has demonstrated that π‐stacking is not an absolute requirement for fibrillization, our findings indicate that Phe‐23 also contributes to fibril formation through π‐stacking interactions and that it is not only the hydrophobic nature of this residue that is relevant in the self‐assembly of hIAPP22‐29. © Proteins 2013. © 2012 Wiley Periodicals, Inc.  相似文献   

9.
The NADPH‐cytochrome P450 oxidoreductase (CYPOR) enzyme is a membrane‐bound protein and contains both FAD and FMN cofactors. The enzyme transfers two electrons, one at a time, from NADPH to cytochrome P450 enzymes to function in the enzymatic reactions. We previously expressed in Escherichia coli the membrane‐bound CYPOR (flAnCYPOR) from Anopheles minimus mosquito. We demonstrated the ability of flAnCYPOR to support the An. minimus CYP6AA3 enzyme activity in deltamethrin degradation in vitro. The present study revealed that the flAnCYPOR purified enzyme, analyzed by a fluorometric method, readily lost its flavin cofactors. When supplemented with exogenous flavin cofactors, the activity of flAnCYPOR‐mediated cytochrome c reduction was increased. Mutant enzymes containing phenylalanine substitutions at leucine residues 86 and 219 were constructed and found to increase retention of FMN cofactor in the flAnCYPOR enzymes. Kinetic study by measuring cytochrome c–reducing activity indicated that the wild‐type and mutant flAnCYPORs followed a non‐classical two‐site Ping‐Pong mechanism, similar to rat CYPOR. The single mutant (L86F or L219F) and double mutant (L86F/L219F) flAnCYPOR enzymes, upon reconstitution with the An. minimus cytochrome P450 CYP6AA3 and a NADPH‐regenerating system, increased CYP6AA3‐mediated deltamethrin degradation compared to the wild‐type flAnCYPOR enzyme. The increased enzyme activity could illustrate a more efficient electron transfer of AnCYPOR to CYP6AA3 cytochrome P450 enzyme. Addition of extra flavin cofactors could increase CYP6AA3‐mediated activity supported by wild‐type and mutant flAnCYPOR enzymes. Thus, both leucine to phenylalanine substitutions are essential for flAnCYPOR enzyme in supporting CYP6AA3‐mediated metabolism. © 2010 Wiley Periodicals, Inc.  相似文献   

10.
Kim SH  Hisano T  Iwasaki W  Ebihara A  Miki K 《Proteins》2008,70(3):718-730
The two-component enzyme, 4-hydroxyphenylacetate 3-monooxygenase, catalyzes the conversion of 4-hydroxyphenylacetate to 3,4-dihydroxyphenylacetate. In the overall reaction, the oxygenase component (HpaB) introduces a hydroxyl group into the benzene ring of 4-hydroxyphenylacetate using molecular oxygen and reduced flavin, while the reductase component (HpaC) provides free reduced flavins for HpaB. The crystal structures of HpaC from Thermus thermophilus HB8 in the ligand-free form, the FAD-containing form, and the ternary complex with FAD and NAD(+) were determined. In the ligand-free form, two large grooves are present at the dimer interface, and are occupied by water molecules. A structural analysis of HpaC containing FAD revealed that FAD has a low occupancy, indicating that it is not tightly bound to HpaC. This was further confirmed in flavin dissociation experiments, showing that FAD can be released from HpaC. The structure of the ternary complex revealed that FAD and NAD(+) are bound in the groove in the extended and folded conformation, respectively. The nicotinamide ring of NAD(+) is sandwiched between the adenine ring of NAD(+) and the isoalloxazine ring of FAD. The distance between N5 of the isoalloxazine ring and C4 of the nicotinamide ring is about 3.3 A, sufficient to permit hydride transfer. The structures of these three states are essentially identical, however, the side chains of several residues show small conformational changes, indicating an induced fit upon binding of NADH. Inactivity with respect to NADPH can be explained as instability of the binding of NADPH with the negatively charged 2'-phosphate group buried inside the complex, as well as a possible repulsive effect by the dipole of helix alpha1. A comparison of the binding mode of FAD with that in PheA2 from Bacillus thermoglucosidasius A7, which contains FAD as a prosthetic group, reveals remarkable conformational differences in a less conserved loop region (Gly83-Gly94) involved in the binding of the AMP moiety of FAD. These data suggest that variations in the affinities for FAD in the reductases of the two-component flavin-diffusible monooxygenase family may be attributed to difference in the interaction between the AMP moiety of FAD and the less conserved loop region which possibly shows structural divergence.  相似文献   

11.
The three‐dimensional structure of Rv2607, a putative pyridoxine 5′‐phosphate oxidase (PNPOx) from Mycobacterium tuberculosis, has been determined by X‐ray crystallography to 2.5 Å resolution. Rv2607 has a core domain similar to known PNPOx structures with a flavin mononucleotide (FMN) cofactor. Electron density for two FMN at the dimer interface is weak despite the bright yellow color of the protein solution and crystal. The shape and size of the putative binding pocket is markedly different from that of members of the PNPOx family, which may indicate some significant changes in the FMN binding mode of this protein relative to members of the family. Proteins 2006. © 2005 Wiley‐Liss, Inc.  相似文献   

12.
A cell-free protein synthesis system is a powerful tool with which unnatural amino acids can be introduced into polypeptide chains. Here, the authors describe unnatural amino acid probing in a wheat germ cell-free translation system as a method for detecting the structural changes that occur in a cofactor binding protein on a conversion of the protein from an apo-form to a holo-form. The authors selected the FMN-binding protein from Desulfovibrio vulgaris as a model protein. The apo-form of the protein was synthesized efficiently in the absence of FMN. The purified apo-form could be correctly converted to the holo-form. Thus, the system could synthesize the active apo-form. Gel filtration chromatography, analytical ultracentrifugation, and circular dichroism-spectra studies suggested that the FMN-binding site of the apo-form is open as compared with the holo-form. To confirm this idea, the unnatural amino acid probing was performed by incorporating 3-azido-L-tyrosine at the Tyr35 residue in the FMN-binding site. The authors optimized three steps in their system. The introduced 3-azido-L-tyrosine residue was subjected to specific chemical modification by a fluorescein-triarylphosphine derivative. The initial velocity of the apo-form reaction was 20 fold faster than that of the holo-form, demonstrating that the Tyr35 residue in the apo-form is open to solvent.  相似文献   

13.
Neuronal nitric-oxide synthase (nNOS) contains a unique autoinhibitory insert (AI) in its FMN subdomain that represses nNOS reductase activities and controls the calcium sensitivity of calmodulin (CaM) binding to nNOS. How the AI does this is unclear. A conserved charged residue (Lys842) lies within a putative CaM binding helix in the middle of the AI. We investigated its role by substituting residues that neutralize (Ala) or reverse (Glu) the charge at Lys842. Compared with wild type nNOS, the mutant enzymes had greater cytochrome c reductase and NADPH oxidase activities in the CaM-free state, were able to bind CaM at lower calcium concentration, and had lower rates of heme reduction and NO synthesis in one case (K842A). Moreover, stopped-flow spectrophotometric experiments with the nNOS reductase domain indicate that the CaM-free mutants had faster flavin reduction kinetics and had less shielding of their FMN subdomains compared with wild type and no longer increased their level of FMN shielding in response to NADPH binding. Thus, Lys842 is critical for the known functions of the AI and also enables two additional functions of the AI as newly identified here: suppression of electron transfer to FMN and control of the conformational equilibrium of the nNOS reductase domain. Its effect on the conformational equilibrium probably explains suppression of catalysis by the AI.  相似文献   

14.
Tear lipocalin and β‐lactoglobulin are members of the lipocalin superfamily. They have similar tertiary structures but unusually low overall sequence similarity. Non‐native helical structures are formed during the early stage of β‐lactoglobulin folding. To address whether the non‐native helix formation is found in the folding of other lipocalin superfamily proteins, the folding kinetics of a tear lipocalin variant were investigated by stopped‐flow methods measuring the time‐dependent changes in circular dichroism (CD) spectrum and small‐angle X‐ray scattering (SAXS). CD spectrum showed that extensive secondary structures are not formed during a burst‐phase (within a measurement dead time). The SAXS data showed that the radius of gyration becomes much smaller than in the unfolded state during the burst‐phase, indicating that the molecule is collapsed during an early stage of folding. Therefore, non‐native helix formation is not general for folding of all lipocalin family members. The non‐native helix content in the burst‐phase folding appears to depend on helical propensities of the amino acid sequence. Proteins 2009. © 2008 Wiley‐Liss, Inc.  相似文献   

15.
New carbo[6]helicene derivatives grafted with π‐conjugated cyano‐phenyl arms were synthesized in enantiopure forms and their π‐conjugation examined by UV–vis spectroscopy. The influence of the π‐conjugation on the circular dichroism spectra and molar rotations is discussed based on comparing experimental data with results from quantum‐chemical calculations. The results highlight the fact that increasing the spatial extension of the π‐system in a helicene molecule is an efficient way of increasing its molar rotation. Chirality 25:455–465, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

16.
The enzymatic dearomatization of aromatic ring systems by reduction represents a highly challenging redox reaction in biology and plays a key role in the degradation of aromatic compounds under anoxic conditions. In anaerobic bacteria, most monocyclic aromatic growth substrates are converted to benzoyl‐coenzyme A (CoA), which is then dearomatized to a conjugated dienoyl‐CoA by ATP‐dependent or ‐independent benzoyl‐CoA reductases. It was unresolved whether or not related enzymes are involved in the anaerobic degradation of environmentally relevant polycyclic aromatic hydrocarbons (PAHs). In this work, a previously unknown dearomatizing 2‐naphthoyl‐CoA reductase was purified from extracts of the naphthalene‐degrading, sulphidogenic enrichment culture N47. The oxygen‐tolerant enzyme dearomatized the non‐activated ring of 2‐naphthoyl‐CoA by a four‐electron reduction to 5,6,7,8‐tetrahydro‐2‐naphthoyl‐CoA. The dimeric 150 kDa enzyme complex was composed of a 72 kDa subunit showing sequence similarity to members of the flavin‐containing ‘old yellow enzyme’ family. NCR contained FAD, FMN, and an iron‐sulphur cluster as cofactors. Extracts of Escherichia coli expressing the encoding gene catalysed 2‐naphthoyl‐CoA reduction. The identified NCR is a prototypical enzyme of a previously unknown class of dearomatizing arylcarboxyl‐CoA reductases that are involved in anaerobic PAH degradation; it fundamentally differs from known benzoyl‐CoA reductases.  相似文献   

17.
Tubulins are an ancient family of eukaryotic proteins characterized by an amino‐terminal globular domain and disordered carboxyl terminus. These carboxyl termini play important roles in modulating the behavior of microtubules in living cells. However, the atomic‐level basis of their function is not well understood. These regions contain multiple acidic residues and their overall charges are modulated in vivo by post‐translational modifications, for example, phosphorylation. In this study, we describe an application of NMR and computer Monte Carlo simulations to investigate how the modification of local charge alters the conformational sampling of the γ‐tubulin carboxyl terminus. We compared the dynamics of two 39‐residue polypeptides corresponding to the carboxyl‐terminus of yeast γ‐tubulin. One polypeptide comprised the wild‐type amino acid sequence while the second contained a Y > D mutation at Y11 in the polypeptide (Y445 in the full protein). This mutation introduces additional negative charge at a site that is phosphorylated in vivo and produces a phenotype with perturbed microtubule function. NMR relaxation measurements show that the Y11D mutation produces dramatic changes in the millisecond‐timescale motions of the entire polypeptide. This observation is supported by Monte Carlo simulations that—similar to NMR—predict the WT γ‐CT is largely unstructured and that the substitution of Tyr 11 with Asp causes the sampling of extended conformations that are unique to the Y11D polypeptide.  相似文献   

18.
Diederik J. Opperman 《Proteins》2017,85(12):2252-2257
Ene‐reductases (ERs), or Old Yellow Enzymes, catalyze the asymmetric reduction of various activated alkenes. This class of biocatalysts is considered an attractive alternative to current chemical technologies for hydrogenation due to their high selectivity and specificity. Here the X‐ray crystal structure of RmER, a “thermophilic”‐like ER from Ralstonia (Cupriavidus) metallidurans, is reported. Unlike other members of this class of ERs, RmER is monomeric in solution which we previously related to its atypical elongated C‐terminus. A typical dimer interface was however observed in our crystal structure, with the conserved Arg‐“finger” forming part of the adjacent monomer's active site and the elongated C‐terminus extending into the active site through contacting the “capping” domain. This dimerization also resulted in the loss of one FMN cofactor from each dimer pair. This potential transient dimerization and dissociation of FMN could conceivably explain the rapid rates previously observed when an FMN light‐driven cofactor regeneration system was used during catalysis with RmER.  相似文献   

19.
Maleylacetate reductase plays a crucial role in catabolism of resorcinol by catalyzing the NAD(P)H‐dependent reduction of maleylacetate, at a carbon–carbon double bond, to 3‐oxoadipate. The crystal structure of maleylacetate reductase from Rhizobium sp. strain MTP‐10005, GraC, has been elucidated by the X‐ray diffraction method at 1.5 Å resolution. GraC is a homodimer, and each subunit consists of two domains: an N‐terminal NADH‐binding domain adopting an α/β structure and a C‐terminal functional domain adopting an α‐helical structure. Such structural features show similarity to those of the two existing families of enzymes in dehydroquinate synthase‐like superfamily. However, GraC is distinct in dimer formation and activity expression mechanism from the families of enzymes. Two subunits in GraC have different structures from each other in the present crystal. One subunit has several ligands mimicking NADH and the substrate in the cleft and adopts a closed domain arrangement. In contrast, the other subunit does not contain any ligand causing structural changes and adopts an open domain arrangement. The structure of GraC reveals those of maleylacetate reductase both in the coenzyme, substrate‐binding state and in the ligand‐free state. The comparison of both subunit structures reveals a conformational change of the Tyr326 loop for interaction with His243 on ligand binding. Structures of related enzymes suggest that His243 is likely a catalytic residue of GraC. Mutational analyses of His243 and Tyr326 support the catalytic roles proposed from structural information. The crystal structure of GraC characterizes the maleylacetate reductase family as a third family in the dehydroquinate synthase‐like superfamily. Proteins 2016; 84:1029–1042. © 2016 Wiley Periodicals, Inc.  相似文献   

20.
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