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.
Creaser CS  Lill JR  Bonner PL  Hill SC  Rees RC 《The Analyst》2000,125(4):599-603
The formation of copper/peptide complex ions by nano-electrospray and microbore HPLC-electrospray mass spectrometry has been investigated for major histocompatibility complex (MHC) class I and class II restricted peptides. Post-column addition of copper(II) acetate following microbore HPLC-MS separation was carried out using a mixing T-piece or via the sheath flow inlet of the electrospray source. Optimal analytical conditions for copper complex ion formation were determined by variation of copper concentration, pH, nebulization gas supply and spray voltage. Tandem mass spectrometry of copper/peptide complex ions provides peptide sequence information and insight into the peptide chelation sites. Copper associated y fragment ions dominate the product ion spectrum for non-histidine containing peptides, but both b and y copper complex ions were observed for the histidine containing MHC class I associated peptide gp70.  相似文献   

2.
The gas-phase oxidation of methionine residues is demonstrated here using ion/ion reactions with periodate anions. Periodate anions are observed to attach in varying degrees to all polypeptide ions irrespective of amino acid composition. Direct proton transfer yielding a charge-reduced peptide ion is also observed. In the case of methionine and, to a much lesser degree, tryptophan-containing peptide ions, collisional activation of the complex ion generated by periodate attachment yields an oxidized peptide product (i.e., [M?+?H?+?O]+), in addition to periodic acid detachment. Detachment of periodic acid takes place exclusively for peptides that do not contain either a methionine or tryptophan side chain. In the case of methionine-containing peptides, the [M?+?H?+?O]+ product is observed at a much greater abundance than the proton transfer product (viz., [M?+?H]+). Collisional activation of oxidized Met-containing peptides yields a signature loss of 64 Da from the precursor and/or product ions. This unique loss corresponds to the ejection of methanesulfenic acid from the oxidized methionine side chain and is commonly used in solution-phase proteomics studies to determine the presence of oxidized methionine residues. The present work shows that periodate anions can be used to ‘label’ methionine residues in polypeptides in the gas phase. The selectivity of the periodate anion for the methionine side chain suggests several applications including identification and location of methionine residues in sequencing applications.
Figure
?  相似文献   

3.
Gas-phase complexes of cysteine-containing peptides and Fe2+ were produced by fast atom bombardment and studied by tandem mass spectrometry. Specific and strong interactions of the iron and sulfur from the thiol group of the cysteine side chain are preserved in the gas phase and are the basis for highly specific fragmentation to give abundant [an − 2H + Fe]+ ions, where n is position of the cysteine residue from the N-terminus of peptide. Metal/peptide complexes containing more than one Cys residue were also investigated; they display similar chemistry upon collisionally activated decompositions, indicating that the Fe2+ ion primarily binds at cysteine sites.  相似文献   

4.
The collision-activated dissociations (CAD) of gas phase salt complexes composed of chiral ions were studied in a quadrupole ion trap mass spectrometer. Because both partners in the salt are chiral, diastereomeric complexes can be formed (e.g., RR, RS). Two general types of complexes were investigated. In the first, the complex was composed of deprotonated binaphthol and a chiral bis-tetraalkylammonium dication. CAD of these complexes leads to the transfer of a proton or an alkyl cation to the binaphtholate leading to a singly-charged tetraalkylammonium cation. During CAD, diastereomeric complexes give significantly different product distributions indicating reasonable stereoselectivity in the process. In the second system, the complexes involved a peptide dianion and a chiral tetraalkylammonium cation. These systems may be viewed as very simple models for the interactions of peptides/proteins with small chiral molecules. Again, stereoselectivity was evident during CAD, but the extent was dependent on the nature of the peptide and not observable in some cases. To better understand the structural features needed to achieve stereoselectivity in gas phase salt complexes, representative transition states were modeled computationally. The results suggest that it is critical for the asymmetry of the nucleophile (i.e., anion) to be well represented in the vicinity of its reactive center.  相似文献   

5.
To enable the development of a tandem mass spectrometry (MS/MS) based methodology for selective protein identification and differential quantitative analysis, a novel derivatization strategy is proposed, based on the formation of a "fixed-charge" sulfonium ion on the side-chain of a methionine amino acid residue contained within a protein or peptide of interest. The gas-phase fragmentation behavior of these side chain fixed charge sulfonium ion containing peptides is observed to result in exclusive loss of the derivatized side chain and the formation of a single characteristic product ion, independently of charge state or amino acid composition. Thus, fixed charge containing peptide ions may be selectively identified from complex mixtures, for example, by selective neutral loss scan mode MS/MS methods. Further structural interrogation of identified peptide ions may be achieved by subjecting the characteristic MS/MS product ion to multistage MS/MS (MS3) in a quadrupole ion trap mass spectrometer, or by energy resolved "pseudo" MS3 in a triple quadrupole mass spectrometer. The general principles underlying this fixed charge derivatization approach are demonstrated here by MS/MS, MS3 and "pseudo" MS3 analysis of side chain fixed-charge sulfonium ion derivatives of peptides containing methionine formed by reaction with phenacylbromide. Incorporation of "light" and "heavy" isotopically encoded labels into the fixed-charge derivatives facilitates the application of this method to the quantitative analysis of differential protein expression, via measurement of the relative abundances of the neutral loss product ions generated by dissociation of the light and heavy labeled peptide ions. This approach, termed "selective extraction of labeled entities by charge derivatization and tandem mass spectrometry" (SELECT), thereby offers the potential for significantly improved sensitivity and selectivity for the identification and quantitative analysis of peptides or proteins containing selected structural features, without requirement for extensive fractionation or otherwise enrichment from a complex mixture prior to analysis.  相似文献   

6.
Alkali metal ions and anionic peptides can be desorbed into the gas phase to give metal-bound peptides and bis(peptide) complexes bearing a ? 1 charge. Although amide nitrogens of peptide bonds are deprotonated in the gas phase by alkali metal ions, this reacion does not occur in solution. Metal-bound dipeptide anions exist as a single structure, whereas those of tripeptide complexes have three structures as revealed by tandem mass spectrometric studies. Ions of bis(peptide) complexes of alkali metals decompose upon collisional activation principally to form deprotonated peptides, in contrast to bis(peptide) complexes of alkaline earth metal ions, which undergo elimination of a neutral peptide.  相似文献   

7.
Stimulated by the interest in developing gold compounds for treating cancer, gold ion–angiotensin peptide interactions are investigated by mass spectrometry. Under the experimental conditions used, the majority of gold ion–angiotensin peptide complexes contain gold in the oxidation states I and III. Both ESI-MS and MALDI-TOF MS detect singly/multiply charged ions for mononuclear/multinuclear gold-attached peptides, which are represented as [peptide + a Au(I) + b Au(III) + (e - a -3b) H]e+, where a,b ≥ 0 and e is charge. ESI-MS data shows singly/multiply charged ions of Au(I)-peptide and Au(III)-peptide complexes. This study reveals that MALDI-TOF MS mainly detects singly charged Au(I)-peptide complexes, presumably due to the ionization process. The electrons in the MALDI plume seem to efficiently reduce Au(III) to Au(I). MALDI also tends to enhance the higher polymeric forms of gold-peptide complexes regardless of the laser power used. Collision-induced dissociation experiments of the mononuclear and dinuclear gold-attached peptide ions for angiotensin peptides show that the gold ion (a soft acid) binding sites are in the vicinity of Cys (a soft ligand), His (a major anchor of peptide for metal ion chelation), and the basic residue Arg. Data also suggests that the abundance of gold-attached peptides increases with higher gold concentration until saturation, after which an increase in gold ion concentration leads to the aggregation and/or precipitation of gold-bound peptides.  相似文献   

8.
Covalent modification of primary amine groups in multiply protonated or deprotonated polypeptides in the gas phase via ion/ion reactions is demonstrated using N-hydroxysuccinimide (NHS) esters as the modifying reagents. During the ion/ion reaction, the peptide analyte ions and the NHS or sulfo-NHS based reagent form a long-lived complex, which is a prerequisite for the covalent modification chemistry to occur. Ion activation of the peptide-reagent complex results in a neutral NHS or sulfo-NHS molecule loss, which is a characteristic signature of covalent modification. As the NHS or sulfo-NHS group leaves, an amide bond is formed between a free, unprotonated, primary amine group of a lysine side chain in the peptide and the carboxyl group in the reagent. Subsequent activation of the NHS or sulfo-NHS loss product ions results in sequence informative fragment ions containing the modification. The N-terminus primary amine group does not make a significant contribution to the modification process; this behavior has also been observed in solution phase reactions. The ability to covalently modify primary amine groups in the gas phase with N-hydroxysuccinimide reagents opens up the possibility of attaching a wide range of chemical groups to gaseous peptides and proteins and also for selectively modifying other analytes containing free primary amine groups.  相似文献   

9.
10.
Oxidative damage of biopharmaceuticals during manufacturing and storage is a key concern throughout pharmaceutical development. However, few simple and robust analytical methods are available for the determination of oxidation sites. Here, the potential of affinity capillary electrophoresis (ACE) in the separation of proteins with oxidized methionine (Met) residues is shown. Silver(I) and gold(I) ions have the attribute to selectively form complexes with thioethers over sulfoxides. The addition of these ions to the BGE leads to a selective complexation of Met residues and, thus, to a change of charge allowing separation of species according to the different oxidation states of Met. The mechanisms of these interactions are discussed and binding constants for peptides containing Met with silver(I) are calculated. Additionally, the proposed method can be used as an indicator of oxidative stress in large proteins. The presented technique is easily accessible, economical, and has rapid analysis times, adding new approaches to the analytical toolbox of Met sulfoxide detection.  相似文献   

11.
Protonated angiotensin II and protonated leucine enkephalin‐based peptides, which included YGGFL, YGGFLF, YGGFLH, YGGFLK and YGGFLR, were subjected to ion/ion reactions with the doubly deprotonated reagents 4‐formyl‐1,3‐benzenedisulfonic acid (FBDSA) and 1,3‐benzenedisulfonic acid (BDSA). The major product of the ion/ion reaction is a negatively charged complex of the peptide and reagent. Following dehydration of [M + FBDSA‐H]? via collisional‐induced dissociation (CID), angiotensin II (DRVYIHPF) showed evidence for two product populations, one in which a covalent modification has taken place and one in which an electrostatic modification has occurred (i.e. no covalent bond formation). A series of studies with model systems confirmed that strong non‐covalent binding of the FBDSA reagent can occur with subsequent ion trap CID resulting in dehydration unrelated to the adduct. Ion trap CID of the dehydration product can result in cleavage of amide bonds in competition with loss of the FBDSA adduct. This scenario is most likely for electrostatically bound complexes in which the peptide contains both an arginine residue and one or more carboxyl groups. Otherwise, loss of the reagent species from the complex, either as an anion or as a neutral species, is the dominant process for electrostatically bound complexes. The results reported here shed new light on the nature of non‐covalent interactions in gas phase complexes of peptide ions that can be used in the rationale design of reagent ions for specific ion/ion reaction applications. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

12.
Polypeptide ions comprising different cationizing agents show distinct fragmentation behavior in the gas phase. Thus, it is desirable to be able to form ions with different cationizing agents such as protons and metal ions. Usually, metal-cationized peptide/protein ions are introduced to the mass spectrometer by electrospraying solutions containing a mixture of the peptide/protein of interest and a metal salt. A new technique for generating metal-containing polypeptide ions that involves gas-phase ion/ion reactions is described. In this strategy, solutions of metal-containing ions and solutions of proteins are each electrosprayed into separate ion sources. The approach allows for independent maximization of ion signal and selection of ions prior to gas-phase reactions. Selected ions are stored in a quadrupole ion trap where reactions of ions of opposite polarity form metal-cationized peptides and proteins in the gas phase by cation switching. This approach affords a high degree of flexibility in forming metal-containing peptide and protein ions via the ability to mass-select reactant ions. The ability to form a variety of peptide/protein ions with various cationizing reagents in the gas phase is attractive both for the study of intrinsic interactions of metal ions with polypeptides and for maximizing the structural information available from tandem mass spectrometry of peptides and proteins.  相似文献   

13.
Arginine forms a stable noncovalent anionic salt bridge complex with DP (a crown ether which contains two endocyclic dialkylhydrogenphosphate esters). Abundant adduct formation with DP is observed for complexes with arginine, YAKR, HPPGFSPFR, AAKRKAA, RR, RPPGFSPFR, RYLGYL, RGDS, and YGGFMRGL in electrospray ionization mass spectrometry (ESI-MS) experiments. DFT calculations predict a hydrogen bonded salt bridge structure with a protonated guanidinium flanked by two deprotonated phosphates to be the lowest energy structure. Dissociation of DP/peptide adducts reveals that, in general, the relative gas phase acidity of a peptide is dependent on peptide length, with longer peptides being more acidic. In particular, peptides that are six residues or more in length can stabilize the deprotonated C-terminus by extensive hydrogen bonding with the peptide backbone. Dissociation of DP/peptide complexes often yields the deprotonated peptide, allowing for the facile formation of anionic peptides that otherwise would be difficult to generate in high abundance. Although DP has a preference for binding to arginine residues in peptides, DP is also observed to form less abundant complexes with peptides containing multiple lysines. Lys-Xxx-Lys and Lys-Lys sequences form low abundance anionic adducts with DP. For example, KKKK exclusively forms a double adduct with one net negative charge on the complex.  相似文献   

14.
气相条件下金属离子/肽复合物电喷雾串联质谱研究   总被引:3,自引:1,他引:2  
研究蛋白质与金属离子之间的相互作用的本质一直是生物学家感兴趣的课题。但蛋白结构的复杂性使得二者之间相互作用的研究难度很大,常选用氨基酸或小肽作为模型化合物进行研究^[1-4]。目前该方法的研究甩缺乏系统性。  相似文献   

15.
As a means of generating fixed-charge peptide radicals in the gas phase we have examined the collision-induced dissociation (CID) chemistry of ternary [Cu(II)(terpy)(TMPP-M)]2+ complexes, where terpy = 2,2':6'2'-terpyridine and TMPP-M represents a peptide (M) modified by conversion of the N-terminal amine to a [tris(2,4,6-trimethoxyphenyl)phosphonium]acetamide (TMPP-) fixed-charge derivative. The following modified peptides were examined: oligoglycines, (Gly)n (n = 1-5), alanylglycine, glycylalanine, dialanine, trialanine and leucine-enkephaline (YGGFL). The [Cu(II)(terpy)(TMPP-M)]2+ complexes are readily formed upon electrospray ionization (ESI) of a mixture of derivatized peptide and [Cu(II)(terpy)(NO3)2] and generally fragment to form transient peptide radical cations, TMPP-M+*, which undergo rapid decarboxylation for the simple aliphatic peptides. This is contrasted with the complexes containing the unmodified peptides, which predominantly undergo fragmentation of the coordinated peptide. These differences demonstrate the importance of proton mobility in directing fragmentation of ternary copper(II) peptide complexes. In the case of leucine-enkephaline, a sufficient yield of the radical cation was obtained to allow further CID. The TMPP-YGGFL+* ion showed a rich fragmentation chemistry, including CO2 loss, side-chain losses of an isopropyl radical, 2-methylpropene and p-quinomethide, and *a1 and *a4 sequence ion formation. In contrast, the even-electron TMPP-YGGFL+ ion fragments to form *a(n) and *b(n) sequence ions as well as the [*b4 + H2O]+ rearrangement ion.  相似文献   

16.
The mechanism of the catalysis of the reversible (propargyl ester)/(allenyl ester) rearrangement ( 10 ? 11 ) by silver(I) ions was investigated, using optically active and diastereoisomeric esters as well as 14C- and 18O-labelling. In order to work with crystalline materials, mainly p-nitrobenzoates ( 10 , 11 : R4 = p? O2N? C6H4) were used. In some cases the rearrangement 10 ? 11 was studied using acetates (R4 = CH3). The alkyl substituents R1, R2, R3, were widely varied (cf. Tables 1, 2). The solvents in which the rearrangements were performed were in most cases dry chlorobenzene and 96% aqueous dioxane. Silver tetrafluoroborate, the benzene complex of the latter, and silver trifluoroacetate (in chlorobenzene) as well as silver nitrate (in aqueous dioxane) served as catalysts. The amounts of the silver catalysts used varied between 0,5 and 10 mol-%; reaction temperatures applied were in the range 35–95°, The results obtained are as follows:
  • 1 The rate-determining step of the (propargyl ester)/(allenyl ester) rearrangement ( 10 ? 11 ) occurs in a silver(I) complex with the substrates ( 10 , 11 ), which is formed in a pre-equilibrium. This follows from kinetic experiments (cf. Fig. 6, 7, 8, 10) and the fact that the rate of rearrangement (of 10a ) is strongly decreased when cyclohexene is added (cf. Fig. 9). In solvents which are known to form complexes with silver(I) ions the rate of rearrangement (of 10a )is much slower than in solvents with similar dielectric constants but with small capacity for complex formation with silver(I) ions (cf. Table 4). Taking into account what is known about silver(I)-alkene and -alkyne complexes (cf. [18]), it is obvious that the (propargyl ester)/(allenyl ester) rearrangement occurs in a π-complex of the silver(I) ion with the triple bond in the propargyl ester or one of the two C,C double bonds in the allenyl ester, respectively.
  • 2 The shift of the carboxyl moiety in the reversible rearrangement 10 ? 11 occurs intramolecularly. p-Nitrobenzoic acid-[carboxyl-14C] is not incorporated during the rearrangement, neither in the reactant 10 nor in the product 11 and vice versa. A crossing experiment gave no mixed products (cf. Scheme 2, p. 882).
  • 3 An internal ion pair can be excluded for the rearrangement 10 ? 11 because the 18O-carbonyl label in the reactant is found exclusively in the alkoxy part of the product (cf. Scheme 3, p. 886, and Table 9). Thus, the rearrangement 10 ? 11 occurs with inversion of the carboxyl moiety.
  • 4 The rearrangement of optically active propargyl esters ( 10g , 10i ) leads to completely racemic allenyl esters ( 11g , 11i ). However, rearrangement of erythro- and threo- 10j -[carbonyl-18O] (Scheme 3) shows that the stereospecifically formed allenyl esters erythro- and threo- 11j -[18O]-epimerize rapidly in the presence of silver(I) ions. This epimerization is twice and forty times, respectively, as fast as the rearrangement of the corresponding propargyl esters (cf. Fig. 1–5). During epimerization or racemization the 18O-label is not randomized (cf. also Scheme 4, p. 898).
  • 5 The equilibrium of the rearrangement 10 ? 11 depends on the bulkiness of the substituents R1, R2, R3 and of the carboxyl moiety (cf. Table 2).
Taking into account these facts (points 1–5), the reversible (propargyl ester)/(allenyl ester) rearrangement promoted by silver(I) ions can be described as a [3s, 3s]-sigmatropic reaction occurring in a silver(I)-π-complex with the C,C triple bond in 10 and a C,C double bond in 11 . It is suggested that complex formation in 10 and 11 occurs with the π-bond which is not involved in the quasicyclic (containing six orbitals and six electrons) transition state of the rearrangement (Fig. 11). Thus, the rearrangement is of a type which has recently been called a charge-induced sigmatropic reaction (cf. [26]). Therefore, in our case, the catalysis by silver(I) ions is of a different type from that of transformations of strained cyclic molecules promoted by silver(I) ions (cf. [14] [16] [27]–[31]). Side reactions. Whereas the rearrangement of propargyl esters 10 in presence of silver tetra- fluoroborate in chlorobenzene or silver nitrate in aqueous dioxane leads to the corresponding allenyl esters 11 , the rearrangement of 10 with silver trifluoroacetate, especially in the presence of trifluoroacetic acid, results in the formation of the dienol esters 12 and 13 , which clearly are derived from 11 (see Scheme 1, p. 881). As shown by the rearrangement of 11 in the presence of p-nitrobenzoic acid-[carboxyl-14C], 12 and 13 arise in part from a not isolated di-p-nitrobenzoate (cf. Scheme 6, p. 905), since radioactivity is found in 12 and 13 .  相似文献   

17.
We report here the generation of gas-phase complexes containing Pd(II), a ligand (deprotonated alanine, A-), and/or N-terminus derivatized peptides containing histidine as one of the amino acids. The species were produced by electrospray ionization, and their gas-phase reactions were investigated using ion-trap tandem mass spectrometry. Pd(II) forms a stable diaqua complex in the gas phase of the formula, [Pd(A-) (H(2)O)(2)]+, (where A- = deprotonated alanine) along with ternary complexes containing A- and peptide. The collision-induced dissociation (CID) patterns of the binary and ternary complexes were investigated, and the dissociation patterns for the ternary complexes suggest that: (a) the imidazole ring of the histidine side group may be the intrinsic binding site of the metal ion, and (b) the peptides fragment primarily by cleavage of the amide bond to the C-terminal side of the histidine residues. These observations are in accord with previous solution-state studies in which Pd(II) was shown to cause hydrolysis of an amide bond of a peptide at the same position.  相似文献   

18.
A peptide containing a single disulfide bond was sequenced using high-energy collision-induced dissociation (HE-CID) in conjunction with a high mass resolution time-of-flight tandem mass spectrometer equipped with a matrix-assisted laser desorption/ionization source. This mass spectrometer, which has spiral ion trajectory, allowed both high mass resolution and high precursor ion selectivity. It is difficult to obtain sufficient product ions from peptides containing disulfide bonds using HE-CID due to the single collision in the gas phase. To compensate for insufficient dissociation, the disulfide bond was cleaved via an in-source reduction process using 1,5-diaminonaphthalene, a reducing matrix. After applying the reduction in the ionization, subsequent sequencing using HE-CID provided the detailed structural information of the peptide containing the single disulfide bond.  相似文献   

19.
Various peptide modifications have been explored recently to facilitate the acquisition of sequence information. N-terminal sulfonation is an interesting modification because it allows unambiguous de novo sequencing of peptides, especially in conjunction with MALDI-PSD-TOF analysis; such modified peptide ions undergo fragmentation at energies lower than those required conventionally for unmodified peptide ions. In this study, we systematically investigated the fragmentation mechanisms of N-terminal sulfonated peptide ions prepared using two different N-terminal sulfonation reagents: 4-sulfophenyl isothiocyanate (SPITC) and 4-chlorosulfophenyl isocyanate (SPC). Collision-induced dissociation (CID) of the SPC-modified peptide ions produced a set of y-series ions that were more evenly distributed relative to those observed for the SPITC-modified peptides; y(n-1) ion peaks were consistently and significantly larger than the signals of the other y-ions. We experimentally investigated the differences between the dissociation energies of the SPITC- and SPC-modified peptide ions by comparing the MS/MS spectra of the complexes formed between the crown ether 18-crown-6 (CE) and the modified peptides. Upon CID, the complexes formed between 18-crown-6 ether and the protonated amino groups of C-terminal lysine residues underwent either peptide backbone fragmentation or complex dissociation. Although the crown ether complexes of the unmodified ([M + CE + 2H]2+) and SPC-modified ([M* + CE + 2H]2+) peptides underwent predominantly noncovalent complex dissociation upon CID, the low-energy dissociations of the crown ether complexes of the SPITC-modified peptides ([M' + CE + 2H]2+) unexpectedly resulted in peptide backbone fragmentations, along with a degree of complex dissociation. We performed quantum mechanical calculations to address the energetics of fragmentations observed for the modified peptides.  相似文献   

20.
A series of ternary copper(II) complexes of the type [Cu(II)(L)(M)](2+), where M represents the hexapeptides GGGFLR, YGGFLR and WGGFLR and L a set of 12 nitrogen donor ligands have been evaluated for their ability to form cationic peptide radicals, M(+)*, in the gas phase. Although the fragmentation chemistry of these ions is complex, two main conclusions emerge: (i) Complexes containing a tri- or tetra-dentate ligand were found to be more effective at producing the peptide radical because in these instances competitive loss of the ligand from the complex is inhibited; (ii) The ligands ought not possess any acidic protons in order to prevent competitive loss of the protonated peptide, [M + H](+). There is significant interaction of the N-terminal aromatic residues in YGGFLR and WGGLFR with the copper(ii) ion in several of the complexes as revealed by the formation of [Cu(I)(L)(p-quinomethide)](+) and [Cu(I)(L)(3-methyleneindoline)](+) fragment ions. Following its dissociation from the ternary complex, CID of the YGGFLR(+)* radical cation shows a dependence on the ligand in the complex from which it was formed. This 'memory effect' most likely reflects differences in the coordinated peptide structure induced by the ligand in the precursor complex which are maintained following dissociation.  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