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.
Cross-flow filtration (CFF) has been investigated as a method of separating filamentously growing fungal cells and purifying the polysaccharide produced. The effects of transmembrane pressure, module geometry (e.g. channel height or tube diameter), tangential feed velocity and cell as well as polysaccharide concentration are discussed. Apart from these experiments, influences by the recirculation pump used are shown.List of Symbols b f fouling index - b factor refering to the behaviour of the sublayer - C kg · m–3 concentration - C g kg · m–3 solute concentration at the membrane - C b kg · m–3 solute concentration in the bulk phase - D s-1 shear rate - k m · s–1 mass-transfer coefficient - K mPa · sn consistency index - n flow behaviour index - P w m3 · s–1 · m–2 rate of permeation - P w1 m3 · s–1 · m–2 rate of permeation at 1 minute - P w m3 · s–1 · m–2 rate of permeation at the beginning - p Pa pressure - Q m2 largest cross-section of a particle - q m2 smallest cross-section of a particle - Re Reynolds number - R f –1 fouling resistance - R m m–1 membrane resistance - t s time - w m · s–1 tangential feed velocity Greek Symbols friction factor - pTM Pa transmembrane pressure - mPa · s shear viscosity - sp specific viscosity (rel. increase of viscosity sp=rel-1) - [] m3· kg–1 intrinsic viscosity - w m2 · s–1 kinematic viscosity - kg · m–3 density Indices b bulk - cell cells - f fouling - g gelling - PS polysaccharide - rel relative - sp specific - w water  相似文献   

2.
N. O. Dungey  D. D. Davies 《Planta》1982,154(5):435-440
Protein turnover was examined, using tritiated water, in various 2-cm regions of 7-11-d-old, first leaves of barley (Hordeum vulgare). Differences were found between the regions in their protein turnover and their responses to stress. The rate constant for degradation for total protein was the same throughout the leaf and the average half-life (t1/2) of protein=approx. 220 h. Only in the older regions did a 24-h pulse of3H2O preferentially label protein with a t1/2 (90 h) considerably shorter than the t1/2 for total protein. Soluble protein was degraded faster than insoluble protein and contained an appreciable short-lived protein component observable by short-pulse labelling. The rate of protein synthesis was greatest in the cells of the youngest region and declined as each region aged. The mean rate of protein synthesis over the 4-d period was 4 and 7 nmol h-1 of amino-N with respect to the regions 1–3 and 7–9 cm from the leaf tip. Seedlings, stressed by adding polyethylene glycol (2.0 MPa) to the roots, showed a marked loss of protein from the older leaf regions with only small losses in the younger regions. Amino acids accumulated in the younger region continuously whereas in the older region little accumulation occurred until day 3 of stress when proline levels increased. Protein synthesis was decreased by between 30% and 50% in all leaf regions. In the region 1–3 cm from the leaf tip, the rate of protein degradation of total protein was enhanced and equalled the rate of degradation of 24-h-pulse-labelled protein which was not itself significantly affected by stress (t1/2=approx. 90 h). In the region 3–5 cm, the degradation of both 4-d and 24-h-labelled protein was enhanced by stress to rates similar to those found in the region 1–3 cm. This was largely through increases in the degradation of the insoluble protein, but the degradation of soluble protein was also raised. Protein degradation in the region 7–9 cm was not affected by stress.Abbreviations t1/2 average half-life - PEG polyethylene glycol  相似文献   

3.
Summary The mating behavior of a number of brewer's and distiller's yeasts was determined with a and haploid and aa and diploid tester strains. Mating frequencies were not high, ranging from one to (rarely) 2,000/108 cells in the mating mixture. Sporulating hybrids were obtained in most matings, though the percentage spore viability initially obtained was often low. Notable the spore viability obtained in hybrids with the haploid tester strains and the brewing strains DIB and DICH was much higher than from the a haploid tester strain, and higher in hybrids between these strains and the aa diploid tester than in those from the tester strain. With the brewing strain NBA, the spore viability in hybrids with the a haploid tester strain was higher than in the case of strains DIB and DICH, but the spore viability in the hybrid of NBA x the haploid strain was higher still. The data are consistent with the hypothesis that with the a and aa tester strains, most of the industrial yeasts tested mate as diploids, and with the and testers, they mate as haploids, an hypothesis which is supported by the segregation of adenine markers in the progeny of these hybrids.Presented in part at the 6th International Specialized Symposium on Yeasts, Montpellier, France, 2–8 July, 1978  相似文献   

4.
5.
The concentration dependence of the influx ofl-lysine in excised roots ofArabidopsis thaliana seedlings was analyzed for the wild-type (WT) and two mutants,rlt11 andraec1, which had been selected as resistant to lysine plus threonine, and to S-2-aminoethyl-l-cysteine, respectively. In the WT three components were resolved: (i) a high-affinity, low-capacity component [K m = 2.2 M;V max = 23 nmol·(g FW)–1·h–1]; (ii) a low-affinity, high-capacity component [K m = 159 M;V max = 742 nmol·(g FW)–1·h–1]; (iii) a component which is proportional to the external concentration, with a constant of proportionalityk = 104 nmol·(g FW)–1 h–1];·mM–1. The influx ofl-lysine in the mutants was lower than in the WT, notably in the concentration range 0.1–0.4 mM, where it was only 7% of that in the WT. In both mutants the reduced influx could be fully attributed to the absence of the low-affinity (high-K m ) component. This component most likely represents the activity of a specific basic-amino-acid transporter, since it was inhibited by several other basic amino acids (arginine, ornithine, hydroxylysine, aminoethylcysteine) but not byl-valine. The high-affinity uptake ofl-lysine may be due to the activity of at least two general amino acid transporters, as it was inhibitable byl-valine, and could be further dissected into two components with a high affinity (K i = 1–5 M; and a low affinity (K i = 0.5–1mM) forl-valine, respectively. Therlt11 andraecl mutant have the same phenotype and the corresponding loci were mapped on chromosome 1, but it is not yet clear whether they are allelic.Abbreviations AEC S-2-aminoethyl-l-cysteine - K i equilibrium constant - WT wild-type  相似文献   

6.
The photooxidation of the primary electron donor in several Photosystem I-related organisms (Synechocystis sp. PCC 6803, Heliobacillus mobilis, and Chlorobium limicola f. sp. thiosulphatophilum) has been studied by light-induced FTIR difference spectroscopy at 100 K in the 4000 to 1200 cm–1 spectral range. The data are compared to the well-characterized FTIR difference spectra of the photooxidation of the primary donor P in Rhodobacter sphaeroides (both wild type and the heterodimer mutant HL M202) in order to get information on the charge localization and the extent of coupling within the (bacterio)chlorophylls constituting the oxidized primary donors. In Rb. sphaeroides RC, four marker bands mostly related to the dimeric nature of the oxidized primary donor have been previously observed at 2600, 1550, 1480, and 1295 cm–1. The high-frequency band has been shown to correspond to an electronic transition (Breton et al. (1992) Biochemistry 31: 7503–7510) while the three other marker bands have been described as phase-phonon bands (Reimers and Hush (1995) Chem Phys 197: 323–332). The absence of these bands in PS I as well as in the heterodimer HL M202 demonstrates that in P700+ the charge is essentially localized on a single chlorophyll molecule. For both H. mobilis and C. limicola, the presence of a high-frequency band at 2050 and 2450 cm–1, respectively, and of phase-phonon bands (at 1535 and 1300 cm–1 in H. mobilis, at 1465 and 1280 cm-1 in C. limicola) indicate that the positive charge in the photooxidized primary donor is shared between two coupled BChls. The structure of P840+ in C. limicola, in terms of the resonance interactions between the two BChl a molecules constituting the oxidized primary donor, is close to that of P+ in purple bacteria reaction centers while for H. mobilis the FTIR data are interpreted in terms of a weaker coupling of the two bacteriochlorophylls.Abbreviations (B)Chl (bacterio)chlorophyll - BPhe bacteriopheophytin - C. Chlorobium - FTIR Fourier transform infrared - H. Heliobacillus - PS I, PS II Photosystem I, Photosystem II - P primary electron donor - RC reaction center - Rb. Rhodobacter - Rp. Rhodopseudomonas - QA primary quinone acceptor - Wt wild type  相似文献   

7.
The primary act of charge separation was studied in P+BA and P+HA states (P, primary electron donor; BA and HA, primary and secondary electron acceptor) of native reaction centers (RCs) of Rhodobacter sphaeroides R-26 using femtosecond absorption spectroscopy at low (90 K) and room temperature. Coherent oscillations were studied in the kinetics of the stimulated emission band of P* (935 nm), of absorption band of BA (1020 nm) and of absorption band of HA (760 nm). It was found that in native RCs kept in heavy water (D2O) buffer the isotopic decreasing of basic oscillation frequency 32 cm –1 and its overtones takes place by the same factor 1.3 in the 935, 1020, and 760 nm bands in comparison with the samples in ordinary water H2O. This suggests that the femtosecond oscillations in RC kinetics with 32 cm –1 frequency may be caused by rotation of hydrogen-containing groups, in particular the water molecule which may be placed between primary electron donor PB and primary electron acceptor BA. This rotation may appear also as high harmonics up to sixth in the stimulated emission of P*. The rotation of the water molecule may modulate electron transfer from P* to BA. The results allow for tracing of the possible pathway of electron transfer from P* to BA along a chain consisting of polar atoms according to the Brookhaven Protein Data Bank (1PRC): Mg(PB)-N-C-N(His M200)-HOH-O = BA. We assume that the role of 32-cm –1 modulation in electron transfer along this chain consists of a fixation of electron density at BA during a reversible electron transfer, when populations of P* and P+BA states are approximately equal.  相似文献   

8.
The formation of metarhodopsin II in various bovine rhodopsin preparations (rod outer segment (ROS) suspensions and rhodopsin-detergent solutions) was measured by means of flash spectrophotometry. The half-lifetime and formation of metarhodopsin II in ROS did not depend on the calcium concentration in the range of less than 10–9 M (using EGTA or EDTA) to 15×10–3 M calcium at pH values of 5.0, 7.1, and 9.0 (Table 1).The regeneration of rhodopsin from opsin by adding 11-cis retinal to ROS-suspensions and rhodopsin digitonin solutions was measured spectrophotometrically. It was not substantially different in either saline, one containing less than 10–7 M calcium (by adding EGTA), the other containing 10–3 M calcium (Table 2).Abbreviations A absorption - A absorption change - CTAB N-Cetyl-N,N,N-trimethylammoniumbromide - E700 extinction at =700 nm - EDTA ethylenediamine-NNNN-tetraacetic acid - EGTA 2,2-ethylenedioxybis [ethyliminodi (acetic acid)] - MI metarhodopsin I - MII metarhodopsin II - Rh rhodopsin - ROS rod outer segment This work is based upon a Ph. D. dissertation (Nöll, 1974) and was presented in part at the Jahrestagung der Deutschen Gesellschaft für Biophysik, Freiburg, Germany, October 1974  相似文献   

9.
We have recently shown that addition of human erythrocyte glycosphingolipids (GSL) to non-human CD4+ or GSL-depleted human CD4+ cells rendered those cells susceptible to gp120-gp41-mediated cell fusion (Puri et al., BBRC, 1998). One GSL fraction (Fraction 3) isolated from human erythrocyte GSL mixture exhibited the highest recovery of fusion following incorporation into CD4+ non-human and GSL-depleted HeLa-CD4 cells (HeLa-CD4/GSL). Structural analysis of Fraction 3 showed that this GSL had identical head group as the known GSL, Gal(14)Gal(1 4)Glc-Ceramide (Gb3) (Puri et al., PNAS, 1998). Here we report that presence of Gb3 in CD4+/CXCR4+ cells but not CD4+/CXCR4 cells allows fusion with HIV-1Lai-envelope glycoprotein expressing cells (TF228). Therefore, Gb3 functions in conjunction with HIV-1 co-receptor, CXCR4 to promote fusion. We propose that Gb3 functions by recruiting CD4 and/or CXCR4 at the fusion site through structurally specific interactions.  相似文献   

10.
The distribution of 12 acid hydrolase and two polysaccharide depolymerase enzymes in the rumen entodiniomorphid ciliatePolyplastron multivesiculatum, isolated from the ovine rumen 2 h after feeding, was examined by differential and density-gradient centrifugation. Approximately 60%–70% of the recovered activity was sedimentable in fractions prepared by centrifugation at 103 g for 10 min (F1) and 104 g for 10 min (F2) with 25%–35% of the acid hydrolases and 15%–20% of acid phosphatase and the polysaccharidases remaining nonsedimentable (in fraction F5) after centrifugation at 105 g for 60 min. Approximately 60% of the sedimentable activity was located in fraction F1. Latency of the hydrolase activity was demonstrated. After isopycnic centrifugation in sucrose density gradients, the hydrolytic enzymes cosedimented in acid phosphatase-containing, membrane-bound, pleomorphic lysosomelike vesicles 0.1–1.0 m in size, with a mean equilibrium density of 1.17 (1.15–1.19) g/ml.  相似文献   

11.
The biogeochemistry of nitrogen (N)was evaluated for three forest ecosystems[Woods Lake (WL), Pancake-Hall Creek (PHC) andHuntington Forest (HF)] in the Adirondackregion of New York, U.S.A. to evaluate theresponse of a range of N atmospheric inputsand experimental N additions. Bulk Ndeposition was higher at sites in the westthan those in the central and easternAdirondacks. These higher atmospheric N inputswere reflected in higher bulk throughfallfluxes of N (WL and PHC, 10.1 and 12.0 kg Nha–1 yr–1, respectively) in thewestern Adirondacks than at HF (4.6 kg Nha–1 yr–1) in the centralAdirondacks. Nitrogen was added to plots as(NH4)2SO4 at 14 and 28 kg Nha–1 yr–1 or as HNO3 at 14 kg Nha–1 yr–1. Litter decompositionrates of Fagus grandifolia and Acerrubrum were substantially higher at WL andPHC compared to HF but were not affected byexperimental N additions. Results usingmineral soil bags showed no effects of Naddition on N and C concentrations in soilorganic matter, but C and N concentrationincreases were less at WL and PHC compared toHF. Soil solution nitrate (NO3 )concentrations at 15-cm depth in the referenceplots were higher at PHC than at WL and HFwhile at 50-cm concentrations were higher atPHC and WL than at HF. The reference plots atthe two sites (WL and PHC) with the highestatmospheric inputs of N exhibited lower Nretention (53 and 33%, respectively) than HF(68%) in reference plots. The greatestincrease in NO3 loss in response tothe experimental treatments occurred at HFwhere the HNO3 additions resulted in thehighest NO3 concentrations andlowest N retentions. In contrast, at WL andPHC increases in soil water NO3 were not evident in response to experimental Nadditions. The results suggest that the twosites (WL and PHC) in the western Adirondacksdid not respond to additional N inputsalthough they have experienced elevatedatmospheric N inputs and higher N drainagelosses in reference plots than the HF site inthe central Adirondacks. Some of thesedifferences in site response may have alsobeen a function of stand age of WL and PHCthat were younger (24 and 33 years,respectively) than the HF (age 70).Highest NO3 fluxes in thereference plots across the sites correspondedto higher 15N values in soil andplants. An experimental addition experimentat PHC found that the forest floor and themineral soil were the largest sinks forexperimentally added N.  相似文献   

12.
An efficient protocol has been developed for the in vitro propagation of Bambusa tulda through shoot proliferation. Shoots from 3-week-old aseptically grown seedlings were used to initiate cultures. Multiple shoots were obtained on liquid Murashige and Skoog (MS) medium supplemented with 6-benzylaminopurine (8×10–6M) and kinetin (4×10–6M). Continuous shoot proliferation at a rate of 4–5 fold every three weeks was achieved through forced axillary branching. More than 90% of the shoots could be rooted on a modified MS medium containing indoleacetic acid (1×10–5M) and coumarin (6.8×10–5M). Following simple hardening procedures, the in vitro raised plants were transferred to the soil with more than 80% success.Abbreviations BAP 6-benzylaminopurine - 2-ip 6-,-dimethylallylaminopurine - Kn kinetin - IAA 3-indoleacetic acid - IBA 3-indolebutyric acid - NAA 1-naphthaleneacetic acid  相似文献   

13.
The light-induced Q A /QA FTIR difference spectra of Rb. sphaeroides and Rp. viridis show very broad positive bands of small amplitude peaking around 2750 cm–1. Upon 1H/2H exchange these bands shift to about 2150 cm–1. Similarly, the Q B /QB spectra exhibit broad continuum bands at 2600 and 2800 cm–1 shifting to 2100 and 2200 cm–1 in 2H2O for Rb. sphaeroides and Rp. viridis, respectively. These continuum bands are tentatively interpreted in terms of highly polarizable hydrogen bonds in a large web of polar bonds involving cofactors, amino acid residues, and structured water molecules. As a working hypothesis, we propose that the protons participating in this web redistribute upon quinone reduction, increasing their concentration around the newly formed charged species, and leading to net proton uptake. Assuming that the precise localization of the mobile protons is dependent on the local electrostatic, this model can explain the apparent discrepancies between some results of FTIR experiments and of electrostatic calculations. Notably, it could help rationalize the observation that mobile protons tend to localize on Glu L212 upon QB reduction in Rb. sphaeroides, while for QB reduction in Rp. viridis and for QA reduction in both Rb. sphaeroides and Rp. viridis, proton uptake by a small number of carboxylic residues is not supported by the FTIR data.  相似文献   

14.
The effect of growth retardants on anthocyanin production was studied in wild carrot (Daucus carota) cell suspension cultures. Paclobutrazol [(2RS,3RS) — 1 — (4-chlorophenyl) — 4,4 —dimethyl-2-(1,2,4-triazol-1-yl) pentan-3-ol], uniconazole [(E)-1-(4-chlorophenyl-4,4 —) dimethyl-2-(1,2,4-triazol-1-yl)-1-penten-3-ol], tetcyclacis [5-(4-chloro-phenyl) -3,4,5,9,10-pentaaza-tetracyclo-5, 4, 102,6, O8,11 — dodeca-3, 9-diene], ancymidol [-cyclopropyl — 4 — methoxy-(pyrimidine-5-yl)benzyl alcohol] and CCC (2-chloro-ethyltrimethylammonium chloride) increased anthocyanin accumulation. AMO-1618 [(2-isopropyl-5-methyl-4-trimethyl-ammonium-chloride)-phenyl-1-piperidinium carboxylate] did not increase anthocyanin accumulation in the first passage but did increase it during the second passage on medium for improved anthocyanin accumulation. Prohexadione (3,5-dioxo-4-propionylcyclohexane carboxylic acid) decreased anthocyanin accumulation by 10%–12.5%.The inhibitory effect of gibberellin on anthocyanin accumulation was reversed by paclobutrazol. Paclobutrazol together with 10–6M GA3 increased anthocyanin level from 33% of control in GA3 treated cell suspension to 76%. These results are consistent growth retardants increasing anthocyanin accumulation in carrot cell suspension cultures by inhibiting gibberellin biosynthesis.  相似文献   

15.
We have measured the rate constant for the formation of the oxidized chlorophyll a electron donor (P680+) and the reduced electron acceptor pheophytin a (Pheo a ) following excitation of isolated Photosystem II reaction centers (PS II RC) at 15 K. This PS II RC complex consists of D1, D2, and cytochrome b-559 proteins and was prepared by a procedure which stabilizes the protein complex. Transient absorption difference spectra were measured from 450–840 nm as a function of time with 500fs resolution following 610 nm laser excitation. The formation of P680+-Pheo a is indicated by the appearance of a band due to P680+ at 820 nm and corresponding absorbance changes at 490, 515 and 546 nm due to the formation of Pheo a . The appearance of the 490 nm and 820 nm bands is monoexponenital with =1.4±0.2 ps. Treatment of the PS II RC with sodium dithionite and methyl viologen followed by exposure to laser excitation results in accumulation of Pheo a . Laser excitation of these prereduced RCs at 15 K results in formation of a transient absorption spectrum assigned to 1*P680. We observe wavelength-dependent kinetics for the recovery of the transient bleach of the Qy absorption bands of the pigments in both untreated and pre-reduced PS II RCs at 15K. This result is attributed to an energy transfer process within the PS II RC at low temperature that is not connected with charge separation.Abbreviations PS I Photosystem I - PS II Photosystem II - RC reaction center - P680 primary electron donor in Photosystem II - Chl a chlorophyll a - Pheo a pheophytin a  相似文献   

16.
Substrate properties of several dTTP analogues bearing a photoreactive 2-nitro-5-azidobenzoyl (NAB) group attached at position 5 of uracil through linkers of various lengths, dTTP–NAB-x-dUTP (where x = 2, 4, 7–13 is the number of atoms in the linker), were studied. All the analogues are substrates for thermostable Thermus thermophilus B35 DNA polymerase in the elongation reaction of the 5-32P-labeled primer–template complex. The kinetic parameters of some of the analogues were determined and compared with those of natural dTTP. It was shown that an increase in the linker length results in a higher efficiency of the analogue. The incorporation of NAB-x-dUP residues into the 3 primer end did not impede further elongation of the chain in the presence of natural dNTP.  相似文献   

17.
Summary Mercury concentration in intraoral air and urine of seven females with dental amalgam was measured before and after intake of one hard-boiled egg. A considerable decrease in mercury concentration in intraoral air was found. Twenty women with about equal dental amalgam status, with or without subjective symptoms related to dental amalgam, were also studied. Mercury concentrations in intraoral air and urine were measured. For all the 27 women the basal intraoral air concentration of mercury ranged over 0.6–10.4 g/m3 (median value 4.3 g/m3). This corresponds to a release of 0.02–0.38 ng/s (median value 0.16 ng/s). In urine, the mercury concentration varied from < 0.8–6.9 g/g creatinine (median value 1.9 g/g creatinine). Data from both parameters were significantly correlated to the total number of teeth areas with dental amalgam. Protein values in urine indicated no renal damage. Maximum concentrations of mercury vapour in intraoral air for the 27 women who had chewed chewing gum for 5 min varied between 2–60 g Hg/m3 (median value 19 g Hg/m3). This corresponds to 0.07–2.20 ng Hg/s and a median value of 0.70 ng Hg/s.  相似文献   

18.
The present study was conducted to characterize the possible interaction of Al3+ and Fe2+ with synthetic melanin in the potentiation of lipid peroxidation in liposomes and rat caudate-putamen homogenates. Al3+ stimulated melanin-initiated lipid peroxidation as measured by the production of 2-thiobarbituric acid-reactive substances (TBARS) and conjugated dienes. The effect of Al3+ was dependent on melanin (10–100 g/ml) and Al3+ (2.5–250 M) concentrations and no synergism between Fe2+ and Al3+ was observed. The prooxidant effect of Al3+ was partially inhibited by superoxide dismutase indicating the involvement of O 2 - . Ga3+ and Be2+ which can increase NADH oxidation in the presence of O 2 - , also were shown to stimulate melanin-initiated TBARS production. Based on the effect of Al3+ and other non redox metals, we suggest that Al3+ does not act through either the induction of melanin free radicals, or the induction of changes in membrane physical properties. Results show that Al3+ enhances melanin-initiated lipid peroxidation in part through an interaction with O 2 - generated from the autoxidation of melanin. We speculate that Al3+ contributes to neuromelanin-mediated oxidative damage in dopaminergic neurons and subsequent neuronal degeneration and death in Parkinson's disease.  相似文献   

19.
The proton/hydroxide (H+/OH) permeability of phospholipid bilayer membranes at neutral pH is at least five orders of magnitude higher than the alkali or halide ion permeability, but the mechanism(s) of H+/OH transport are unknown. This review describes the characteristics of H+/OH permeability and conductance through several types of planar phospholipid bilayer membranes. At pH7, the H+/OH conductances (G H/OH) range from 2–6 nS cm–2, corresponding to net H+/OH permeabilities of (0.4–1.7)×10–5 cm sec–1. Inhibitors ofG H/OH include serum albumin, phloretin, glycerol, and low pH. Enhancers ofG H/OH include chlorodecane, fatty acids, gramicidin, and voltages >80 mV. Water permeability andG H/OH are not correlated. The characteristics ofG H/OH in fatty acid (weak acid) containing membranes are qualitatively similar to the controls in at least eight different respects. The characteristics ofG H/OH in gramicidin (water wire) containing membranes are qualitatively different from the controls in at least four different respects. Thus, the simplest explanation for the data is thatG H/OH in unmodified bilayers is due primarily to weakly acidic contaminants which act as proton carriers at physiological pH. However, at low pH or in the presence of inhibitors, a residualG H/OH remains which may be due to water wires, hydrated defects, or other mechanisms.  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