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 532 毫秒
1.
Using a murine hypodermic air pouch infection model designed to mimic the release of bacterial products at physiological levels, 3-hydroxy fatty acid (3-OH FA) and endotoxin unit levels from Burkholderia cenocepacia isolates were assessed. The B. cenocepacia environmental isolates (n = 35) survived in the hypodermic air pouch but did not invade across the peritoneal epithelial layer during a 72-h infection. For all 35 strains, when the molar ratio of C14:0 3-OH FA to C16:0 3-OH FA in the air pouch fluid wash samples was between 1.4 and 2.5, the concentrations of C14:0 3-OH FA were correlated with the endotoxin unit levels. However, both surrogate markers exhibited different correlations to the inflammatory response. The linear regression coefficient was 0.4234 for C14:0 3-OH FA concentrations vs. NO productions, 0.223 for endotoxin unit levels vs. NO productions, 0.5008 for C14:0 3-OH FA concentrations vs. TNF-alpha productions and 0.2869 for endotoxin unit levels vs. TNF-alpha productions. Therefore, C14:0 3-OH FA concentrations, rather than endotoxin unit levels, acted as an immunostimulatory indicator for LPS in the B. cenocepacia isolates.  相似文献   

2.
BackgroundWith dry eye, the ratio of cholesteryl ester (CE) to wax ester (WE) decreases substantially in meibum, but the functional and structural consequences of this change are speculative. The aim of this study is to confirm this finding and to bridge this gap in knowledge by investigating the effect of varying CE/WE ratios on lipid structure and thermodynamics.MethodsInfrared spectroscopy was use to quantify CE and WE in human meibum and to measure hydrocarbon chain conformation and thermodynamics in a cholesteryl behenate, stearyl stearate model system.ResultsThe CE/WE molar ratio was 36% lower for meibum from donors with dry eye due to meibomian gland dysfunction compared with meibum from donors without dry eye. CE (5 mol %) dramatically increased the phase transition temperature of pure WE from -0.12 °C to 63 °C in the mixture. Above 5 mol % CB, the phase transition temperature increased linearly, from 68.5 °C to 85 °C. In the ordered state, CE caused an increase in lipid order from about 72% trans rotamers to about 86% trans rotamers. Above 10% CE, the hydrocarbon chains were arranged in a monoclinic geometry.ConclusionsThe CE/WE is lower in meibum from donors with dry eye due to meibomian-gland dysfunction. Major conformational changes in the hydrocarbon chains of wax and cholesteryl ester mixtures begin to occur with just 5% CB and above.General significanceCE-WE interactions may be important for in understanding lipid layer structure and functional relationships on the surface of tears, skin and plants.  相似文献   

3.
The flowers of 23 species of grass and herb plants were collected from a mesotrophic grassland to assess natural variability in bulk, monosaccharide and fatty acid δ13C values from one plant community and were compared with previous analyses of leaves from the same species. The total mean bulk δ13C value of flower tissues was −28.1‰, and there was no significant difference between the mean δ13Cflower values for grass (−27.8‰) and herb (−28.2‰) species. On average bulk δ13Cflower values were 1.1‰ higher than bulk δ13Cleaf values, however, the δ13Cflower and δ13Cleaf values of grasses did not differ between organs suggesting that carbon isotope discrimination is different in grass and herb species. The abundance of different monosaccharides abundance varied between plant types, i.e. xylose concentrations in the grass flowers were as high as 40%, compared with up to 15% in the herb species, but the general relationship δ13Carabinose > δ13Cxylose > δ13Cglucose > δ13Cgalactose which had been observed in leaves was similar in flowers (total mean δ13C values = −25.9‰, −27.2‰, −28.8‰ and −28.1‰, respectively). However, the average 5.4‰ depletion in the δ13C values of the C16:0, C18:2 and C18:3 fatty acids in flowers compared to bulk tissue was significantly greater than observed for leaves. The trend C16:0 < C18:2 < C18:3 previously observed in leaves was also observed in grass flowers (δ13CC16:0 = −33.8‰; δ13CC18:2 = −33.1‰; δ13CC18:3 = −34.2‰) but not herb flowers (δ13CC16:0 = −34.1‰; δ13CC18:2 = −32.4‰; δ13CC18:3 = −34.5‰). We conclude: (i) that the biological processes influencing carbon isotope discrimination in grass flowers are different from herbs flowers; and, (ii) that a range of post-photosynthetic fractionation effects caused the observed differences between flower and leaf δ13C values, especially the significant 13C-depletion in flower fatty acid δ13C values.  相似文献   

4.
Leaves of 26 grass, herb, shrub and tree species were collected from mesotrophic grasslands to assess natural variability in bulk, fatty acid and monosaccharide δ13C values under different grazing management (cattle- or deer-grazed) on three sample dates (May, July and October) such that interspecific and spatiotemporal variations in whole leaf tissues and compound-specific δ13C values could be determined. The total mean leaf bulk δ13C value for plants was −28.9‰ with a range of values spanning 7.5‰. Significant interspecific variation between bulk leaf δ13C values was only determined in October (P = <0.001) when δ13C values of the leaf tissues from both sites was on average 1.5‰ depleted compared to during July and May. Samples from May were significantly different between fields (P = 0.03) indicating an effect from deer- or cattle-grazing in young leaves. The average individual monosaccharide δ13C value was 0.8‰ higher compared with whole leaf tissues. Monosaccharides were the most abundant components of leaf biomass, i.e. arabinose, xylose, mannose, galactose and glucose, and therefore, fluctuations in their individual δ13C values had a major influence on bulk δ13C values. An average depletion of ca. 1‰ in the bulk δ13C values of leaves from the deer-grazed field compared to the cattle-grazed field could be explained by a general depletion of 1.1‰ in glucose δ13C values, as glucose constituted >50% total leaf monosaccharides. In October, δ13C values of all monosaccharides varied between species, with significant variation in δ13C values of mannose and glucose in July, and mannose in May. This provided an explanation for the noted variability in the tissue bulk δ13C values observed in October 1999. The fatty acids C16:0, C18:2 and C18:3 were highly abundant in all plant species. Fatty acid δ13C values were lower than those of bulk leaf tissues; average values of −37.4‰ (C16:0), −37.0‰ (C18:2) and −36.5‰ (C18:3) were determined. There was significant interspecific variation in the δ13C values of all individual fatty acids during October and July, but only for C18:2 in May (P = <0.05). This indicated that seasonal trends observed in the δ13C values of individual fatty acids were inherited from the isotopic composition of primary photosynthate. However, although wide diversity in δ13C values of grassland plants ascribed to grazing management, interspecific and spatiotemporal influences was revealed, significant trends (P = <0.0001) for fatty acid and monosaccharide δ13C values: δ13C16:0 < δ13C18:2 < δ13C18:3 and δ13Carabinose > δ13Cxylose > δ13Cglucose > δ13Cgalactose, respectively, previously described, appear consistent across a wide range of species at different times of the year in fields under different grazing regimes.  相似文献   

5.
Electrochemistry of cytochrome c (cyt c) immobilized on a cardiolipin (CL)/phosphatidylcholine (PC) film supported on a glassy carbon electrode was investigated using variable-frequency AC voltammetry. At low ionic strength, we observed two redox-active subpopulations characterized by distinct values of potential (E1/2) and electron transfer rate constant (kET). At high ionic strength, only one subpopulation was detected, consistent with the existence of very stable cyt c–CL adducts, most probably formed by hydrophobic interactions between the protein and the fatty acid (FA) chains carried by CL. This subpopulation exhibits a comparatively high kET value (> 300 s− 1) apparently changing with the structure of the FA chains of CL, i.e. 18:2(n − 6) or 14:0. Our study suggests that electrochemistry can be a useful technique for probing protein–lipid interactions, and more particularly the role played by the specific structure of the FA chains of CL on cyt c binding.  相似文献   

6.
Novel ionic liquid (IL) sol-gel materials development, for enzyme immobilization, was the goal of this work. The deglycosylation of natural glycosides were performed with α-l-rhamnosidase and β-d-glucosidase activities expressed by naringinase. To attain that goal ILs with different structures were incorporated in TMOS/Glycerol sol-gel matrices and used on naringinase immobilization.The most striking feature of ILs incorporation on TMOS/Glycerol matrices was the positive impact on the enzyme activity and stability, which were evaluated in fifty consecutive runs. The efficiency of α-rhamnosidase expressed by naringinase TMOS/Glycerol@ILs matrices increased with cation hydrophobicity as follows: [OMIM] > [BMIM] > [EMIM] > [C2OHMIM] > [BIM] and [OMIM] ≈ [E2-MPy] ? [E3-MPy]. Regarding the imidazolium family, the hydrophobic nature of the cation resulted in higher α-rhamnosidase efficiencies: [BMIM]BF4 ? [C2OHMIM]BF4 ? [BIM]BF4. Small differences in the IL cation structure resulted in important differences in the enzyme activity and stability, namely [E3-MPy] and [E2-MPy] allowed an impressive difference in the α-rhamnosidase activity and stability of almost 150%. The hydrophobic nature of the anion influenced positively α-rhamnosidase activity and stability. In the BMIM series the more hydrophobic anions (PF6, BF4 and Tf2N) led to higher activities than TFA. SEM analysis showed that the matrices are shaped lens with a film structure which varies within the lens, depending on the presence and the nature of the IL.The kinetics parameters, using naringin and prunin as substrates, were evaluated with free and naringinase encapsulated, respectively on TMOS/Glycerol@[OMIM][Tf2N] and TMOS/Glycerol@[C2OHMIM][PF6] and on TMOS/Glycerol. An improved stability and efficiency of α-l-rhamnosidase and β-glucosidase expressed by encapsulated naringinase on TMOS/Glycerol@[OMIM][Tf2N] and TMOS/Glycerol@[C2OHMIM][PF6] were achieved. In addition to these advantageous, with ILs as sol-gel templates, environmental friendly processes can be implemented.  相似文献   

7.
A new simple, sensitive and precise liquid chromatography-tandem mass spectrometry method has been developed and validated for the determination of valacyclovir-HCl and acyclovir in tsetse flies (Glossina pallipides). Tsetse flies were extracted by ultrasonication with acidified methanol/acetonitrile, centrifuged and cleaned up by solid phase dispersion using MgSO4 and MSPD C18 material. Samples were analysed using a Waters Alliance 2695 series HPLC with a C18 Gemini analytical column (150 mm × 4.6 mm × 5 μm) and a guard cartridge column connected to a Waters Quattro-Micro triple-quadrupole mass spectrometer. The isocratic mobile phase consisted of methanol:acetonitrile:water (60:30:10, v/v/v) plus formic acid (0.1%) at a flow rate of 0.25 ml/min. The precursor > product ion transition for valacyclovir (m/z 325.1 > 152) and acyclovir (m/z 226.1 > 151.9) were monitored in positive electrospray multiple reaction monitoring mode. The method was validated at fortification levels of 0.5, 1 and 2 μg/g. The range of calibration for both drugs was 0.45-4.5 μg/g. The overall accuracy of the method was 92% for valacyclovir and 95% for acyclovir with corresponding within-laboratory reproducibilities of 4.4 and 3.4%, respectively. Mean recoveries were above 80% for both drugs and repeatability ranged from 0.7 to 6.1%. For both drugs the limits of detection and quantification were 0.0625 and 0.2 μg/g, respectively. The method was applied in experiments on the mass rearing of tsetse flies for sterile insect technique (SIT) applications, in which the flies were fed with blood meals containing acyclovir or valcyclovir-HCl prior to analysis to assess effects on Glossina pallidipes Salivary Gland Hypertrophy syndrome.  相似文献   

8.
In [PtX(PPh3)3]+ complexes (X = F, Cl, Br, I, AcO, NO3, NO2, H, Me) the mutual cis and trans influences of the PPh3 groups can be considered constants in the first place, therefore the one bond Pt-P coupling constants of P(cis) and P(trans) reflect the cis and trans influences of X. The compounds [PtBr(PPh3)3](BF4) (2), [PtI(PPh3)3](BF4) (3), [Pt(AcO)(PPh3)3](BF4) (4), [Pt(NO3)(PPh3)3](BF4) (5), and the two isomers [Pt(NO2-O)(PPh3)3](BF4) (6a) and [Pt(NO2-N)(PPh3)3](BF4) (6b) have been newly synthesised and the crystal structures of 2 and 4·CH2Cl2·0.25C3H6O have been determined. From the 1JPtP values of all compounds we have deduced the series: I > Br > Cl > NO3 > ONO > F > AcO > NO2 > H > Me (cis influence) and Me > H > NO2 > AcO > I > ONO > Br > Cl > F > NO3 (trans influence). These sequences are like those obtained for the (neutral) cis- and trans-[PtClX(PPh3)2] derivatives, showing that there is no dependence on the charge of the complexes. On the contrary, the weights of both influences, relative to those of X = Cl, were found to depend on the charge and nature of the complex.  相似文献   

9.
The electrochemical behavior of the S,S-bridged adducts of square planar metalladithiolene complexes was investigated by using cyclic voltammetry and electrochemical spectroscopies (visible, near-IR, and ESR). The norbornene-bridged S,S-adduct [Ni(S2C2Ph2)2(C7H8)] (2a; C7H8=norbornene) formed by [Ni(S2C2Ph2)2] (1a) and quadricyclane (Q) was dissociated by an electrochemical reduction, and anion 1a and norbornadiene (NBD) were formed. Q was isomerized to NBD in the overall reaction. The o-xylyl-bridged S,S-adduct [Ni(S2C2Ph2)2(CH2)2(C6H4)] (3a; (CH2)2(C6H4)=o-xylyl) was also dissociated by an electrochemical reduction, and this reaction gave the o-xylyl radical (o-quinodimethane). The reduction of complex 3a in the presence of excess o-xylylene dibromide underwent the catalytic formation of o-quinodimethane. The butylene-bridged S,S-adduct [Ni(S2C2Ph2)2(CH2)4] (4a; (CH2)4=butylene) was stable on an electrochemical reduction. The lifetimes of reduced species of these adducts 2a-4a were influenced by the stability of the eliminated group (stability: NBD > o-xylyl radical (o-quinodimethane) > butylene radical). Therefore, the reduced species are stable in the sequence 4a > 3a > 2a. Although the palladium complex [Pd(S2C2Ph2)2] (1b) was easier to reduce than the nickel complex 1a or the platinum complex [Pt(S2C2Ph2)2] (1c), their S,S-adducts were easier to reduce in the order of Ni adduct > Pd adduct > Pt adduct.  相似文献   

10.
A highly sensitive liquid chromatography-tandem mass spectrometric (LC-MS/MS) method was developed for the determination of forsythiaside in rat plasma using epicatechin as internal standard. The analytes were extracted by solid-phase extraction and chromatographied on a C18 column eluted with a gradient mobile phase of acetonitrile and water both containing 0.2% formic acid. The detection was performed by negative ion electrospray ionization in multiple reaction monitoring mode, monitoring the transitions m/z 623 → 161 and m/z 289 → 109 for forsythiaside and epicatechin, respectively. The assay was linear over the concentration ranges of 2.0–50.0 and 50.0–5000.0 ng/mL with limits of detection and quantification of 0.2 and 1.0 ng/mL, respectively. The precision was <10.8% and the accuracy was >91.9%, and extraction recovery ranged from 81.3% to 85.0%. This method was successfully applied to a pharmacokinetic study of forsythiaside in rats after intravenous (20 mg/kg) and oral (100 mg/kg) administration, and the result showed that the compound was poorly absorbed with an absolute bioavailability being approximately 0.5%.  相似文献   

11.
Using a murine hypodermic air pouch infection model designed to mimic the release of bacterial products at physiological levels, 3-hydroxy fatty acid (3-OH FA) and endotoxin unit levels from Burkholderia cenocepacia isolates were assessed. The B. cenocepacia environmental isolates (n = 35) survived in the hypodermic air pouch but did not invade across the peritoneal epithelial layer during a 72-h infection. For all 35 strains, when the molar ratio of C14:0 3-OH FA to C16:0 3-OH FA in the air pouch fluid wash samples was between 1.4 and 2.5, the concentrations of C14:0 3-OH FA were correlated with the endotoxin unit levels. However, both surrogate markers exhibited different correlations to the inflammatory response. The linear regression coefficient was 0.4234 for C14:0 3-OH FA concentrations vs. NO productions, 0.223 for endotoxin unit levels vs. NO productions, 0.5008 for C14:0 3-OH FA concentrations vs. TNF-alpha productions and 0.2869 for endotoxin unit levels vs. TNF-alpha productions. Therefore, C14:0 3-OH FA concentrations, rather than endotoxin unit levels, acted as an immunostimulatory indicator for LPS in the B. cenocepacia isolates.  相似文献   

12.
Lai HT  Lin JS  Chien YH 《Bioresource technology》2011,102(9):5425-5430
This study investigated the effects of light (visible light - 5800 lux, 24 h) or dark regime and aerobic or anaerobic condition on the decay of added oxolinic acid (OA) at 5, 10 and 20 mg L−1 in eel pond sediment. An asymptotic decaying exponential model Ct = Cmin + Co × exp (−k × t) was used to facilitate quantitative approach to OA transformation, where Ct is the concentration of OA after t days, Cmin the estimated level-off concentration of OA residue, Co the concentration of added OA and k the decaying coefficient. OA decayed faster under light (Cmin = 4.6 mg L−1) than under dark (Cmin = 7.8 mg L−1) and also decayed faster under aerobic (Cmin = 4.0 mg L−1) than under anaerobic condition (Cmin = 8.5 mg L−1). Cmin increased with Co. Sundrying and tilling eel pond bottom should be able to reduce OA residue significantly.  相似文献   

13.
DFT calculations with a variety of exchange-correlation functionals, including PW91, OLYP, TPSSh, B3LYP and B3LYP*, have been carried out on the low-energy spin states of chloroiron(III) porphyrin and four aryliron(III) porphyrins, viz. FeIII(P)Ph (S = 1/2), FeIII(P)C6F5 (S = 5/2), FeIII(P)(3,4,5-C6F3H2) (S = 1/2), FeIII(P)(2,4,6-C6F3H2) (S = 5/2), where the expected spin states have been indicated within parentheses. Qualitatively, OLYP reproduces all the expected ground spin states. B3LYP appears to have some difficulty yielding the observed sextet ground states. B3LYP*, TPSSh and PW91 all fail to reproduce the sextet ground states, the latter two by rather large margins of energy. As far as this study is concerned, the overall performance of the functionals appears to be OLYP/OPBE > B3LYP > B3LYP* >> TPSSh > PW91/BLYP/BP86/TPSS.  相似文献   

14.
15.
Methyl esters of fatty acids, free fatty acids, and hydrocarbons were found in the culture liquid and in the cellular lipids of the obligate methylotrophic bacterium Methylophilus quaylei under optimal growth conditions and osmotic stress. The main extracellular hydrophobic metabolite was methyl stearate. Exogenous free fatty acids C16–C18 and their methyl esters stimulated the M. quaylei growth and survivability, as well as production of exopolysaccharide under osmotic and oxidative stress, playing the role of growth factors and adaptogens. The order of hydrophobic supplements according to the ability to stimulate bacterial growth is C18: 1 > C18: 0 > C16: 0 > methyl oleate > methyl stearate > no supplements > C14: 0 > C12: 0. The mechanism underlying the protective action of fatty acids and their methyl esters is discussed.  相似文献   

16.
In the current study, capillary electrophoresis (CE)-based enzyme assay for characterization and inhibition study of bovine carbonic anhydrase II (bCA II) was developed. The developed method is the first CE assay for carbonic anhydrase (CA). The method was optimized in order to get short analysis time, minimal sample volume consumption, and high resolution of substrate and product. The CE conditions were optimized as follows: fused-silica capillary (30 cm effective length × 75 μm i.d.), pressure injection for 5 s, 20 mM sodium borate buffer (pH 9.0), constant voltage of 15 kV, constant capillary temperature of 25 °C, and detection at 260 nm. For precise measurements, uridine was used as an internal standard during optimization of the CE methods. The limits of detection and quantification for p-nitrophenyl acetate (p-NPA) were 3.01 and 9.12 μM, respectively, whereas for p-nitrophenolate they were 2.05 and 6.22 μM, respectively. The performance of the developed method was confirmed by determination of kinetic parameters (i.e., Km and Vmax of bCA for p-NPA); the inhibition constant (Ki) was determined for furosemide, a standard inhibitor of CA. The new method proved to be fast and efficient, and it can be used for the investigation of inhibitors of all isoforms of CAs.  相似文献   

17.
A range of model biochemical components, microalgae and cyanobacteria with different biochemical contents have been liquefied under hydrothermal conditions at 350 °C, ∼200 bar in water, 1 M Na2CO3 and 1 M formic acid. The model compounds include albumin and a soya protein, starch and glucose, the triglyceride from sunflower oil and two amino acids. Microalgae include Chlorella vulgaris,Nannochloropsis occulata and Porphyridium cruentum and the cyanobacteria Spirulina. The yields and product distribution obtained for each model compound have been used to predict the behaviour of microalgae with different biochemical composition and have been validated using microalgae and cyanobacteria. Broad agreement is reached between predictive yields and actual yields for the microalgae based on their biochemical composition. The yields of bio-crude are 5-25 wt.% higher than the lipid content of the algae depending upon biochemical composition. The yields of bio-crude follow the trend lipids > proteins > carbohydrates.  相似文献   

18.

Background

Mammalian GPx7 is a monomeric glutathione peroxidase of the endoplasmic reticulum (ER), containing a Cys redox center (CysGPx). Although containing a peroxidatic Cys (CP) it lacks the resolving Cys (CR), that confers fast reactivity with thioredoxin (Trx) or related proteins to most other CysGPxs.

Methods

Reducing substrate specificity and mechanism were addressed by steady-state kinetic analysis of wild type or mutated mouse GPx7. The enzymes were heterologously expressed as a synuclein fusion to overcome limited expression. Phospholipid hydroperoxide was the oxidizing substrate. Enzyme–substrate and protein–protein interaction were analyzed by molecular docking and surface plasmon resonance analysis.

Results

Oxidation of the CP is fast (k+ 1 > 103 M− 1 s− 1), however the rate of reduction by GSH is slow (k′+ 2 = 12.6 M− 1 s− 1) even though molecular docking indicates a strong GSH–GPx7 interaction. Instead, the oxidized CP can be reduced at a fast rate by human protein disulfide isomerase (HsPDI) (k+ 1 > 103 M− 1 s− 1), but not by Trx. By surface plasmon resonance analysis, a KD = 5.2 μM was calculated for PDI–GPx7 complex. Participation of an alternative non-canonical CR in the peroxidatic reaction was ruled out. Specific activity measurements in the presence of physiological reducing substrate concentration, suggest substrate competition in vivo.

Conclusions

GPx7 is an unusual CysGPx catalyzing the peroxidatic cycle by a one Cys mechanism in which GSH and PDI are alternative substrates.

General significance

In the ER, the emerging physiological role of GPx7 is oxidation of PDI, modulated by the amount of GSH.  相似文献   

19.
A hydrothermal reaction of a mixture of Gd(NO3)3, 1,2-benzenedicarboxylic acid (1,2-BDC), piperazine, NaOH and water at 180 °C for three days under autogeneous pressure gave rise to a new compound of the formula [C4N2H12][Gd2(H2O)2(C6H4(COO)2)2] (I). The connectivity between GdO8 distorted dodecahedra and 1,2-BDC units gives rise to a two-dimensional structure with large apertures. The fully protonated piperazine molecule occupies the middle of these apertures. The compound has favorable CH?π interactions between the benzene rings of adjacent layers and shows photoluminescence at room temperature. Crystal data: monoclinic, space group = P21/c (No. 14), a = 13.1671(3) Å, b = 13.7336(3) Å, c = 11.3100(1) Å, β = 115.411(1)°, v = 1847.34(6) Å3, Z = 4, R1 = 0.0238 for 2658 reflections [I > 2σ(I)].  相似文献   

20.
The silver(I) salts [AgOR] (3a, R = C9H6N; 3b, R = C6H4-2-CHO, 3c, R = C6H4-2-Cl; 3d, R = C6H4-2-CN; 3e, R = C6H4-2-NO2) are accessible by the stoichiometric reaction of [AgNO3] (1) with HOR (2a, R = C9H6N; 2b, R = C6H4-2-CHO; 2c, R = C6H4-2-Cl; 2d, R = C6H4-2-CN; 2e, R = C6H4-2-NO2) in presence of NEt3. Treatment of 3a-3e with PnBu3 (4), P(OMe)3 (5a) or P(OCH2CF3)3 (5b) in the ratios of 1:1 and 1:2, respectively, produced complexes [LmAgOR] (L = PnBu3, = 1: 6a, R = C9H6N; 6b, R = C6H4-2-CHO; 6c, R = C6H4-2-Cl; 6d, R = C6H4-2-CN; 6e, R = C6H4-2-NO2. = 2: 7a, R = C9H4; 7b, R = C6H4-2-CHO; 7c, R = C6H4-2-Cl; 7d, R = C6H4-2-CN; 7e, R = C6H4-2-NO2. L = P(OMe)3, = 1: 8a, R = C6H4-2-CHO; 8b, R = C6H4-2-NO2. = 2: 9, R = C6H4-2-NO2. L = P(OCH2CF3)3, = 1: 10, R = C6H4-2-NO2). Based on TGA, temperature-programmed and in situ molecular beam mass spectrometry metal-organic 7e was applied as CVD precursor in the deposition of silver onto glass substrates. The resulting silver films were characterized by XRD. The SEM image of a film grown from 7e at 350 °C showed a homogeneous surface with grain sizes of 40 nm. The molecular structures of 8b and 10 in the solid state were determined. They are isostructural and are cubane-like structured. Low-temperature 31P{1H} NMR studies showed that the title complexes are dynamic in solution and exchange at room temperature their ligands.  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