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.
Several CF3Se-substituted α-amino acid derivatives, such as (R)-2-amino-3-((trifluoromethyl)selanyl)propanoates ( 5 a / 6 a ), (S)-2-amino-4-((trifluoromethyl)selanyl)butanoates ( 5 b / 6 b ), (2R,3R)-2-amino-3-((trifluoromethyl)selanyl)butanoates ( 5 c / 6 c ), (R)-2-((S)-2-amino-3-phenylpropanamido)-3-((trifluoromethyl)selanyl)propanoates ( 11 a / 12 a ), and (R)-2-(2-aminoacetamido)-3-((trifluoromethyl)selanyl)propanoates ( 11 b / 12 b ), were readily synthesized from natural amino acids and [Me4N][SeCF3]. The primary in vitro cytotoxicity assays revealed that compounds 6 a , 11 a and 12 a were more effective cell growth inhibitors than the other tested CF3Se-substituted derivatives towards MCF-7, HCT116, and SK-OV-3 cells, with their IC50 values being less than 10 μM for MCF-7 and HCT116 cells. This study indicated the potentials of CF3Se moiety as a pharmaceutically relevant group in the design and synthesis of novel biologically active molecules.  相似文献   

2.
The performance of poly(carbon monofluoride) (PMF) (synonym: graphite fluoride) (CFx)n, ( 1 ) as oxidizer in a fuel rich Mg based pyrolant (ξ(Mg)=0.45) is investigated. The radiance [W sr−1] under both static and dynamic conditions, the spectral radiant exitance [W cm−2 μm−1], the linear burn rate [mm s−1] and the mass consumption rate [g s−1 cm−2] have been determined. Calculations for enthalpy of anaerobic and aerobic combustion [kJ g−1] and for the equilibrium composition of the combustion products as well as combustion temperature are given. A mechanism for the reaction is discussed. The combustion product from Mg/(CFx)n surprisingly reveals the formation of single walled carbon nanotubes (SWCNT) and “carbon nano carpet rolls” (CNCR). For part V see Ref. [1].  相似文献   

3.
Reactions of the fluorous primary phosphines Rfn(CH2)2PH2 [Rfn=(CF2)n−1CF3; n=6, 8, 10] and Rfn′CHCH2 [(n′=6, 8, 10) (1 : 1; THF, reflux) in the presence of AIBN give the title compounds [Rfn(CH2)2][Rfn′(CH2)2]PH [n/n′=6/6 ( 4 , 55%), 8/8 ( 5 , 58%), 10/10 ( 6 , 53%), 8/6 ( 7 , 52%), 10/8 ( 8 , 51%)] as low-melting white solids on up to 10-g scales. The chiral tertiary phosphine [Rf6(CH2)2][Rf8(CH2)2][Rf10(CH2)2]P ( 9 ) is similarly prepared from 7 and Rf10CHCH2 in the presence of VAZO (neat, 100 °C; 67%). The reaction of 5 and THF⋅BH3 yields the phosphine borane 5 ⋅BH3 (95%). Additions of triphosgene [(CCl3O)2CO] to 5 or Rf8(CH2)2PH2 give [Rf8(CH2)2]2PCl or Rf8(CH2)2PCl2, which are characterized in situ. The CF3C6F11/toluene partition coefficients of 4 – 9 increase with the number and lengths of the Rfn segments.  相似文献   

4.
Ruthenium complexes with the formulae Ru(CO)2(PR3)2(O2CPh)2 [ 6a – h ; R=n‐Bu, p‐MeO‐C6H4, p‐Me‐C6H4, Ph, p‐Cl‐C6H4, m‐Cl‐C6H4, p‐CF3‐C6H4, m,m′‐(CF3)2C6H3] were prepared by treatment of triruthenium dodecacarbonyl [Ru3(CO)12] with the respective phosphine and benzoic acid or by the conversion of Ru(CO)3(PR3)2 ( 8e – h ) with benzoic acid. During the preparation of 8 , ruthenium hydride complexes of type Ru(CO)(PR3)3(H)2 ( 9g , h ) could be isolated as side products. The molecular structures of the newly synthesized complexes in the solid state are discussed. Compounds 6a – h were found to be highly effective catalysts in the addition of carboxylic acids to propargylic alcohols to give valuable β‐oxo esters. The catalyst screening revealed a considerably influence of the phosphine′s electronic nature on the resulting activities. The best performances were obtained with complexes 6g and 6h , featuring electron‐withdrawing phosphine ligands. Additionally, catalyst 6g is very active in the conversion of sterically demanding substrates, leading to a broad substrate scope. The catalytic preparation of simple as well as challenging substrates succeeds with catalyst 6g in yields that often exceed those of established literature systems. Furthermore, the reactions can be carried out with catalyst loadings down to 0.1 mol% and reaction temperatures down to 50 °C.

  相似文献   


5.
Polymer surfaces have been modified by chemical reaction or thin film deposition using a capacitatively coupled plasma RF-discharge system. This has allowed the production of a graded, reproducible series of surfaces with surface energies ranging from low to high. Three gases, hexafluoroethane (C2F6), perfluoropropane (C3F8), and hexafluoropropene (CF3CF?CF2), were used to prepare the low-energy surfaces. Ethylene oxide (C2H4O) was used to prepare the higher energy surfaces. Intermediate values of surface energy were obtained by using gas mixtures of perfluoropropane and ethylene oxide. The plasma-deposited films were characterized by ESCA and contact-angle studies. In the low-energy range, the critical surface tension (γc) values of the films prepared were found to be lower than those for Teflon. For the higher energy surfaces, values up to about 45 dyn/cm were obtained. Addition of oxygen to the plasma permitted the production of films with still higher γc values. ESCA studies indicate that, under the reaction conditions used, C3F8 and CF3CF?CF2 form a fluoropolymer deposit on glass that is at least 100 Å thick. For the same reaction parameters, C2F6 resulted in more etching and deposition of a much thinner film.  相似文献   

6.
Chemical modification based on incorporation of flame retardants (FR) into the polymer backbone was used in order to reduce polystyrene flammability. 3‐(trifluoromethyl)styrene (StCF3) and 3,5‐bis(trifluoromethyl)styrene (St(CF3)2) were applied as reactive FR. Copolymers were synthesized with different feed ratios and it gave series of copolymers with various amounts of StCF3 and St(CF3)2 (5–50% mol/mol of St). Glass transition temperature (Tg) and thermal stability of obtained (co)polymers were determined from differential scanning calorimetry (DSC) and thermogravimetric analysis (TGA), respectively. Kinetic parameters such as the thermal decomposition activation energy (E) and frequency factor (A) were estimated by Ozawa and Kissinger models. Pyrolysis combustion flow calorimeter (PCFC) was applied as a tool for assessing the flammability of the synthesized (co)polymers. Relative reactivity ratios were determined by applying the conventional linearization Jaacks method (rSt = 1.34, rStCF3 = 0.54), (rSt = 0.47, rSt(CF3)2 = 0.13). The results suggest that incorporation of fluorinated styrenes into PSt enchance flame retardance. © 2015 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2015 , 132, 42839.  相似文献   

7.
The reaction of [RhCl(PiPr3)2] ( 1 ) with 1,4-C6H4(C≡CH)2 at 0°C leads almost quantitatively to the formation of the bis(alkyne) complex [(PiPr3)2ClRh-(HC≡C6H4-C≡CH)RhCl(PiPr3)2] (2). At elevated temperatures (THF, 60°C) it rearranges to give the isomeric bis(vinylidene) complex [(PiPr3)2ClRh-(=C=CH-C6H4-CH=C=)RhCl(PiPr3)2] (3). A one-pot synthesis of 3 is also described. Treatment of either 2 or 3 with pyridine affords the bis(alkynyl)dihydrido compound [(PiPr3)2(py)Cl(H)Rh(-C≡C-C6H4-C≡C-)Rh(H)Cl(py)(PiPr3)2] ( 4 ) in which both metal centers are octahedrally coordinated. Whereas the reaction of 2 with NaC5H5 produces the complex CsH5(PiPr3)Rh(HC≡C-C6H4-C≡CH)Rh(PiPr3)C5H5 ( 7 ), the bis(vinyl-idene) isomer C5H5(PiPr3)Rh(=C=CH-C6H4-CH=C=)Rh(PiPr3)C5Hs ( 8 ) is obtained from 4 and NaC5H5. Electrophiles preferably attack the Rh=C bonds of 8 and thus on protonation with CF3CO2H the bis(vinyl) complex C5H5(PiPr3)(CF3CO2)-Rh(Z,Z-CH=CH-C6H4-CH=CH)Rh(O2CCF3)(PiPr3)C5Hs ( Z-9 ) is formed. In acetone solution, it rearranges to give the E isomer. Reaction of 8 with sulfur affords the bis(thioketene) complex C5H5(PiPr3)Rh(≡2-C,S; η2-C,S-S=C=CH-C6H4-CH=C=S)-Rh(PiPr3)C5H5 ( 12 ), for which only one diastereomer is observed. All attempts to prepare mononuclear rhodium compounds containing the diyne HC≡C-C6H4-G≡CH or the isomeric vinylidene: C=CH-C6H4-G≡CH as ligand failed.  相似文献   

8.
Cotton fabrics were treated by radio‐frequency plasma with tetrafluoromethane (CF4) and hexafluoropropene (C3F6) gases under different exposure times, pressures, and power levels. The hydrophobicity and water repellency were analyzed with measurements of the cosine of the contact angle (cos θ) and wet‐out time. The hydrophobicity was enhanced with treatments of both gases. X‐ray photoelectron spectroscopy (XPS) revealed increases in the surface fluorine content of 1–2% for CF4 plasma and of 2.3–7.8% for C3F6 plasma. The relative chemical composition of the C1s spectra after CF4 and C3F6 plasma treatments showed increases in the relative amounts of ? C? O? C? and fluorocarbon groups (? CF, ? CF2, and ? CF3), whereas peak areas for ? COH and ? COOH decreased. The hydrophobicity was enhanced by the increase in the fluorine content and fluorocarbon groups. C3F6 plasma treatment resulted in higher hydrophobicity than CF4 plasma treatment according to not only cos θ and wet‐out measurements but also XPS analysis. © 2003 Wiley Periodicals, Inc. J Appl Polym Sci 88: 2038–2047, 2003  相似文献   

9.
Summary [2,5-Bis(trifluoromethyl)phenyl]acetylene [BTFPA; HCCC6H3-2, 5-(CF3)2]polymerized with W, Mo, and Nb catalysts to produce methanol-insoluble polymers in high yields. The poly(BTFPA) produced by the W(CO)6-based catalyst at 30 °C was soluble in p-(CF3)2C6H4, and had relatively high molecular weight ([]=0.352 dL/g in p-(CF3)2C6H4). The main chain of the polymer was composed of alternating double bonds, and the polymer was a dark brown solid. The temperature at which the weight loss of the polymer started was higher than 300 °C. The polymerization behavior and polymer properties for BTFPA are compared with those for phenylacetylene and [o-(trifluoromethyl)phenyl]acetylene.  相似文献   

10.
The effect of the nature of the anion on the performance of ionic rhodium catalysts has received little attention. Herein it is shown that the use of highly fluorous tetraphenylborate anions can enhance catalyst activity in both conventional and fluorous media. For hydrogenation catalysts of the type [Rh(COD)(dppb)][X] {COD=1,5‐cis,cis‐cyclooctadiene; dppb=1,4‐bis(diphenylphosphino)butane; X=BF4 ( 1a ), [BPh4] ( 1b ), [B{C6H4(SiMe3)‐4}4] ( 1c ), [B{C6H3(CF3)2‐3,5}4] ( 1d ), [B{C6H4(SiMe2CH2CH2C6F13)‐4}4] ( 1e ), [B{C6H4(C6F13)‐4}4] ( 1f ) and [B{C6H3(C6F13)2‐3,5}4] ( 1 g )} the activity towards the hydrogenation of 1‐octene in acetone increased in the order 1c < 1b < 1e < 1a < 1d ~ 1f < 1g with 1g being twice as active as the commonly applied 1a . Despite the fluorophilic character introduced by the substituted tetraarylborate anions, the presence of some perfluoroalkyl‐substituents in the cation was still required for achieving high partition coefficients. Therefore, [Rh(COD)(Ar2PCH2CH2PAr2)][X] {Ar=C6H4(SiMe2CH2CH2C6F13)‐4, X=[B{C6H3(C6F13)2‐3,5}4] ( 3f ); Ar=C6H4(SiMe(CH2CH2C6F13)2)‐4 and X=[B{C6H4(C6F13)‐4}4] ( 2g )} were prepared, which were active in the hydrogenation of 1‐octene, 2g even more so than 3f . Both these highly fluorous catalysts could be recycled with 99% efficiency through fluorous biphasic separation, whereas the corresponding BF4 complex of 2g ( 2a ) did not show any affinity for the fluorous phase.  相似文献   

11.
The iron complex generated in situ from Fe(CF3SO3)2 and the new tridentate N ligand mpzbpy (mpzbpy stands for 6-(3,5-dimethylpyrazol-1-ylmethyl)-6′-methyl-2,2′-bipyridine) mediates the room temperature oxidation of cyclohexane to cyclohexanol and cyclohexanone, and the epoxidation of cyclooctene to cyclooctene oxide. X-ray diffraction studies of single crystals obtained by reaction of iron(II) triflate with mpzbpy in methanol in the presence of ascorbic acid reveals the formation of a dinuclear complex, namely [Fe2(mpzbpy)2(C2O4)(CF3SO3)2] (1), where two iron(II) ions are bridged by an oxalato dianion originating from the oxidative cleavage of ascorbic acid. Magnetic susceptibility measurements of 1 show that the two metallic centres are weakly antiferromagnetically coupled, with a J value of −1.74(1) cm−1.  相似文献   

12.
Four new fully triphenylamine-based polyamides coded as polyamide (CF3,CF3), polyamide (CF3,CH3), polyamide (CH3,CF3), and polyamide (CH3,CH3) were synthesized by the phosphorylation polyamidation reaction from various combinations of 3,5-bis(trifluoromethyl)-4′,4″-dicarboxytriphenylamine, 3,5-dimethyl-4′,4″-dicarboxytriphenylamine, 3,5-bis(trifluoromethyl)-4′,4″-diaminotriphenylamine, and 3,5-dimethyl-4′,4″-diaminotriphenylamine. All the polyamides were amorphous and readily soluble in many common organic solvents and could be solution-cast into transparent, flexible, and strong films with good mechanical properties. They had useful levels of thermal stability associated with high glass-transition temperatures of 268–287°C and 10?wt%-loss temperatures in excess of 500°C. Cyclic voltammograms of the film of polyamide (CH3,CH3) on the indium-tin oxide-coated glass substrates exhibited two oxidation redox couples with E1/2 around 0.82 and 1.29?V vs. Ag/AgCl in tetrabutylammonium perchlorate/acetonitrile solution, accompanied by a color change from pale yellow neutral state to dark green oxidized state. The CF3-substituted polyamides displayed a higher oxidation potential and less electrochemical stability as compared to the CH3-substituted analogues.  相似文献   

13.
The electrochemical behaviour of the series of ten [Rh(RCOCHCOR′)(P(OPh)3)2] complexes with R, R′ = CF3, CF3 (1), CF3, CH3 (2), CF3, Ph (C6H5) (3), CF3, Fc (ferrocenyl = (C5H5)Fe(C5H4)) (4), CH3, Ph (5), CH3, CH3 (6), Ph, Ph (7), Fc, CH3 (8), Fc, Ph (9) and Fc, Fc (10) were studied in acetonitrile containing 0.100 mol dm−3 tetra-n-butylammonium hexafluorophosphate as supporting electrolyte utilizing a glassy carbon working electrode. Results are consistent with Rh(I) being first oxidized in an electrochemically irreversible two-electron transfer process at peak anodic potentials ranging Epa(Rh) = 0.124–0.881 V vs. Fc/Fc+. For the ferrocene-containing complexes (4), and (8)–(10) the rhodium oxidation was followed by the electrochemically reversible oxidation of the ferrocenyl group in a one-electron transfer process at a slightly more positive potential. Relationships were established between the electrochemical quantity Epa(Rh) and kinetic parameter log k2 as well the sum of experimental group electronegativities (Gordy Scale) of the R and R′ groups (χR + χR′), the Hammett σ values (σR + σR′) and the Lever ligand parameter EL for the [Rh(RCOCHCOR′)(P(OPh)3)2] complexes: Epa(Rh) (vs. Fc/Fc+/V) = 0.31 (χR + χR′)–1.09 = 0.56 (σR + σR′) + 0.28 = SMEL) + (IM − 0.66 V) = −0.23 log k2 − 0.03 (k2 = second order rate constant for the oxidative addition of methyl iodide to rhodium). A profound shift of Epa(Rh) to a more positive potential was observed for Rh(I) substrates containing β-diketonato ligands with increasing electronegative substituents R and R′. An exponential dependence of Epa(Rh) on the pKa of the β-diketone was obtained.  相似文献   

14.
T.A. Centeno  F. Stoeckli 《Carbon》2008,46(7):1025-1030
Based on more than 80 carbons, the paper shows that immersion calorimetry into benzene, water and carbon tetrachloride can be used to assess with a good accuracy the limiting capacitance Co at low current densities in both acidic (2 M H2SO4) and aprotic (1 M tetraethyl ammonium tetrafluoroborate in acetonitrile) electrolytic solutions. The enthalpies of immersion ΔiH(C6H6) and ΔiH(H2O) provide information on Co-acidic, where both the surface area and the oxygen content play a role. On the other hand, in the case of the organic electrolyte the oxygen content has only a small influence and Co-aprotic is directly related to ΔiH(C6H6) and ΔiH(CCl4). Carbon tetrachloride has a critical dimension (0.65 nm), which is close to the size of the (C2H5)4N+ ion (0.68 nm) and therefore ΔiH(CCl4) provides better information in the case of carbons with small micropores. The advantage of this approach lies in the fact that immersion calorimetry, in itself a useful tool for the structural and the chemical characterization of carbons, can also be used to evaluate directly the gravimetric capacitances of these solids at low current densities.  相似文献   

15.
In order to increase the formation ratio of perfluorotrimethylamine, (CF3)3N, to overall anode gas in electrolytic production using Ni anode, mixed melts of (CH3)3mHF + CsF·2.3HF were used as electrolytes at room temperature. The ionic conductivity of the mixed melts decreased with an increase in the CsF concentration, whereas the viscosity of the mixed melts increased with increasing the CsF concentration. AC impedance and XRD analysis revealed that the presence of CsNi2F6 in the oxidized layer formed on the Ni anode after electrolysis. The gas evolved at the Ni anode was composed of (CF3)3N, (CF3)2CHF2N, CF3(CHF2)2N, (CHF2)3N, CF4, NF3, CHF3, C2HF5, and C2F6. The best ratio of (CF3)3N to the overall anode gas (52.11%) was obtained in the electrolyte of (CH3)3N·5.0HF + 50 wt% CsF·2.3HF mixed melt at 20 mA cm−2.  相似文献   

16.
Reactions of (CO)5Re(Br), (η5‐C5H5)Ru(Cl)(PPh3)2, and [Pt(μ‐Cl)(C6F5)(S(CH2CH2‐)2)]2 with the alkyne‐containing phosphine Ph2P(CH2)6C≡CCH3 give the bis(phosphine) complexes fac‐(CO)3Re(Br)(Ph2P(CH2)6C≡CCH3)2 ( 5 ), (η5‐C5H5)Ru(Cl)(Ph2P(CH2)6C≡CCH3)2 ( 6 ), and trans‐(Cl)(C6F5)Pt(Ph2P(CH2)6C≡CCH3)2 ( 7 ). Alkyne metatheses with the catalyst (t‐BuO)3W(≡C‐t‐Bu) (10–15 mol %, chlorobenzene, 80 °C) give the seventeen‐membered metallamacrocycles fac‐(CO)3Re(Br)(Ph2P(CH2)6CC(CH2)6P Ph2) ( 8 ), (η5‐C5H5)Ru(Cl)(Ph2P(CH2)6CC(CH2)6P Ph2) ( 9 ), and trans‐(Cl)(C6F5)Pt(PPh2(CH2)6CC(CH2)6P Ph2) ( 10 ). 31P NMR analyses show 90–75% conversions to 8 – 10 (59–47% isolated after chromatography). The identity of 8 was confirmed by a crystal structure, and 10 was hydrogenated over Pd/C to fac‐(CO)3Re(Br)(Ph2P(CH2)6CC(CH2)6P Ph2) ( 12 , 87%), which was crystallographically characterized earlier. A catalyst derived from Mo(CO)6/4‐chlorophenol effects a slower conversion of 7 to 10 at 140 °C. In the case of 5 , a mer, trans isomer of 8 is isolated ( 11 , 44%), as established by NMR and IR data. In 10 – 12 , the diphosphines span trans positions. These results, together with previous examples involving group VIII metallocenes, establish the wide viability of the title reaction.  相似文献   

17.
One-dimensional structure of Zinc(II) metal organic framework ([Zn2(C10H8N2)3(NO3)4]) has been synthesized at room temperature and structurally characterized by elemental analysis, thermogravimetric analysis, and single-crystal X-ray diffraction. The five-coordinated Zn(II) complex exhibits trigonal bipyramidal coordination geometry. The Zn(II) ion is stabilized by three-connected nodes to form 1D ladder structure. The 1D chain is further connected to each other via hydrogen bonding to form 2D structure. Disordered CH2Cl2 solvent molecules were trapped in the pores. The luminescent and thermal properties have been also investigated. In addition, the activation thermodynamic parameters, ΔE*, ΔH*, ΔS* and ΔG* are calculated from the DTA curves using Coats–Redfern method.  相似文献   

18.
Surface modification of poly(tetrafluoroethylene-co-perfluoropropyl vinyl ether) (PFA) with vacuum UV (VUV) photo-oxidation using radiation from excited Ar atoms downstream from an Ar microwave (MW) plasma shows: (1) an improvement in wettability as observed by water contactangle measurements; (2) surface roughening; (3) incorporation of oxygen as C=O, CF—O—CF2 and CF2—O—CF2 moieties and (4) enhancement of the CF—O—C n F2n+1 concentration. Adhesion measurements of Cu sputter coated onto the photo-oxidized PFA surface results in failure within the PFA (cohesive failure) and not at the Cu–PFA interface.  相似文献   

19.
The synthesis and characterization of two tungsten carbonyl dimers containing bridging alkoxide or aryloxide ligands are described. The crystal and molecular structures of [PPN]2[W2(CO)8(OCH2CF3)2], 1, and [Et4N]3[W2(CO)6-(OPh)3]-CH3CN, 2 , are reported and compared with the structures of tetranuclear tungsten derivatives previously described. The dimer 1 crystalizes in the triclinic space group P 1 with unit cell parameters a = 13.460(11) Å, b = 12.318(5) Å, c = 13.842(10) Å, α = 82.73(5)°, β = 59.11(5)°, γ= 80.09(5)°, V = 1938(2) Å3, and Z = 1. The complex 2 crystalizes in the monoclinic space group P21/n with unit cell parameters a = 11.954(2) Å, b = 19.359(4) Å, c = 26.462(5) Å, β = 102.50(16)°, V = 5979(2) Å3, Z = 4. Molecular modeling software was utilized to construct a tetranuclear derivative from 1 similar to the structurally characterized [W(CO)3OH]4−4 tetramer. The two tetramers were found to possess similar molecular parameters. This supports the contention that dimers of type 1 are the precursors of the tetramers. Comparisons of the tungsten alkoxides and aryloxides with the behavior of W(CO)6 on γ-alumina are provided.  相似文献   

20.
In this research, we describe the application of the complexes o‐C6H4(NSiMe3)2ZrCl2 ( 1 ), o‐C6H4(NSiMe3)2TiBr2 ( 2 ), o‐C6H4(NSiMe3)2TiCl2 ( 3 ), C2H4(NSiMe3)2ZrCl2 ( 4 ), in the ethylene polymerization with different Al/M ratios and temperatures. These complexes presented significant catalytic activities in the presence of methyaluminoxane (MAO) as cocatalyst and toluene as solvent, producing high molecular weight linear polyethylenes. Zirconium complexes were more active at 60°C and titanium complexes at 40°C. Zirconium complex ( 1 ) showed the best values of activity (347 kg PE/mol Zr h atm) for Al/Zr ratio of 340 and 60°C of temperature. In ethylene‐1‐hexene copolymerization, the best result was also reached with catalyst 1 , at the same conditions. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2008  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