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.
2.
We review an extensive body of single-crystal high-frequency electron paramagnetic resonance (HFEPR) data in order to determine the transverse spin Hamiltonian parameters that control the tunneling of the direction of magnetization in a variety of integer and half-integer-spin single-molecule magnets (SMMs). The SMMs studied are members of the following families: S = 9/2 [Mn4O3Cl]6+; S = 5 [Mn3NiO4]6+; S = 6 [Mn3ZnO4]6+; and S = 4 [Ni4(OR)4]4+. HFEPR spectra for the half-integer S = 9/2 Mn4 complexes that have C3 symmetry do not provide measurable evidence for transverse spin Hamiltonian terms. This finding is consistent with the relatively large coercive field seen in the magnetization hysteresis loops for these complexes. On the other hand, a low symmetry S = 9/2 complex exhibits a much faster rate of ground-state magnetization tunneling, in agreement with HFEPR spectra for a powder sample that gives a rhombic zero-field splitting (ZFS) parameter of E = 0.140 cm?1. The S = 5 Mn3Ni systems exhibit magnetization tunneling that is much faster than seen for the high-symmetry S = 9/2 Mn4 complexes. This can be attributed to their integer-spin ground states. Like the C3 symmetry Mn4 SMMs, the HFEPR spectra for high-symmetry Mn3Ni complexes do not provide measurable evidence for transverse ZFS terms. However, the spectra exhibit broad peaks, suggesting distributions in the local molecular environments brought about by disordered solvate molecules. This disorder likely explains the fast tunneling in the high-symmetry S = 5 Mn3Ni systems, though one cannot rule out fourth- (and higher-) order interactions that cannot be detected by HFEPR due to the broad resonances. The one S = 6 Mn3Zn complex shows an even faster rate of tunneling compared to the isostructural S = 5 Mn3Ni complex. Finally, the S = 4 [Ni(hmp)(dmb)Cl]4 complex provides unique insights into the origin of fourth- (and higher-) order interactions found for many SMMs on the basis of analysis using a giant spin Hamiltonian (GSH) approximation. We conclude that the fourth-order anisotropy found for the S = 4 ground state of [Ni(hmp)(dmb)Cl]4 originates from the second-order ZFS interactions associated with the individual NiII ions, but only as a result of higher-order processes that occur via S-mixing between the ground state and higher-lying (S < 4) spin-multiplets. The S-mixing is relatively strong in this system because of comparable exchange and anisotropy energy scales. The relatively fast tunneling is a direct consequence of this S-mixing, as opposed to any intrinsic fourth-order (spin–orbit) anisotropy associated with NiII.  相似文献   

3.
Herein we report a series of novel chloramphenicol amine derivatives as aminopeptidase N (APN)/CD13 inhibitors. All compounds were synthesized starting from commercially available (1S,2S)-2-amino-1-(4-nitrophenyl) propane-1,3-diol. The preliminary biological screening showed that some compounds exhibited potent inhibitory activity against APN. It should be noted that one compound, 13b (IC50 = 7.1 μM), possess similar APN inhibitory activity compared with Bestatin (IC50 = 3.0 μM).  相似文献   

4.
Secreted LOX from Pseudomonas aeruginosa (PA-LOX) has previously been identified as arachidonic acid 15S-lipoxygenating enzyme. Here we report that the substitution of Ala420Gly in PA-LOX leads to an enzyme variant with pronounced dual specificity favoring arachidonic acid 11R-oxygenation. When compared with other LOX-isoforms the molecular oxygen affinity of wild-type PA-LOX is 1–2 orders of magnitude lower (Km O2 of 0.4 mM) but Ala420Gly exchange improved the molecular oxygen affinity (Km O2 of 0.2 mM). Experiments with stereo-specifically deuterated linoleic acid indicated that the formation of both 13S- and 9R-HpODE involves abstraction of the proS-hydrogen from C11 of the fatty acid backbone. To explore the structural basis for the observed functional changes (altered specificity, improved molecular oxygen affinity) we solved the crystal structure of the Ala420Gly mutant of PA-LOX at 1.8 Å resolution and compared it with the wild-type enzyme. Modeling of fatty acid alignment at the catalytic center suggested that in the wild-type enzyme dioxygen is directed to C15 of arachidonic acid by a protein tunnel, which interconnects the catalytic center with the protein surface. Ala420Gly exchange redirects intra-enzyme O2 diffusion by bifurcating this tunnel so that C11 of arachidonic acid also becomes accessible for O2 insertion.  相似文献   

5.
Endocannabinoids appear to be involved in a variety of physiological processes. Lipoxygenase activity has been known to be affected by unsaturated fatty acids or phenolic compounds. In this study, we examined whether endocannabinoids containing both N-acyl group and phenolic group can affect the activity of soybean lipoxygenase (LOX)-1, similar to mammalian 15-lipoxygenase in physicochemical properties. First, N-arachidonoyl dopamine and N-oleoyl dopamine were found to inhibit soybean LOX-1-catalyzed oxygenation of linoleic acid in a non-competitive manner with a Ki value of 3.7 μM and 6.2 μM, respectively. Meanwhile, other endocannabinoids failed to show a remarkable inhibition of soybean LOX-1. Separately, N-arachidonoyl dopamine and N-arachidonoyl serotonin were observed to inactivate soybean LOX-1 with Kin value of 27 μM and 24 μM, respectively, and k3 value of 0.12 min−1 and 0.35 min−1, respectively. Furthermore, such an inactivation was enhanced by ascorbic acid, but suppressed by 13(S)-hydroperoxy-9,11-octadecadienoic acid. Taken together, it is proposed that endocannabinoids containing polyunsaturated acyl moiety and phenolic group may be efficient for the inhibition as well as inactivation of 15-lipoxygenase.  相似文献   

6.
We report for the first time kinetic and thermodynamic properties of soluble acid invertase (SAI) of sugarcane (Saccharum officinarum L.) salt sensitive local cultivar CP 77-400 (CP-77). The SAI was purified to apparent homogeneity on FPLC system. The crude enzyme was about 13 fold purified and recovery of SAI was 35%. The invertase was monomeric in nature and its native molecular mass on gel filtration and subunit mass on SDS-PAGE was 28 kDa. SAI was highly acidic having an optimum pH lower than 2. The acidic limb was missing. Proton transfer (donation and receiving) during catalysis was controlled by the basic limb having a pKa of 2.4. Carboxyl groups were involved in proton transfer during catalysis. The kinetic constants for sucrose hydrolysis by SAI were determined to be: km = 55 mg ml?1, kcat = 21 s?1, kcat/km = 0.38, while the thermodynamic parameters were: ΔH* = 52.6 kJ mol?1, ΔG* = 71.2 kJ mol?1, ΔS* = ?57 J mol?1 K?1, ΔG*E–S = 10.8 kJ mol?1 and ΔG*E–T = 2.6 kJ mol?1. The kinetics and thermodynamics of irreversible thermal denaturation at various temperatures 53–63 °C were also determined. The half -life of SAI at 53 and 63 °C was 112 and 10 min, respectively. At 55 °C, surprisingly the half -life increased to twice that at 53 °C. ΔG*, ΔH* and ΔS* of irreversible thermal stability of SAI at 55 °C were 107.7 kJ mol?1, 276.04 kJ mol?1 and 513 J mol?1K?1, respectively.  相似文献   

7.
The present study was undertaken to gain insight into the associations of mercury(II) with dicysteinyl tripeptides in buffered media at pH 7.4. We investigated the effects of increasing the distance between cysteinyl residues on mercury(II) associations and complex formations. The peptide–mercury(II) formation constants and their associated thermodynamic parameters in 3-(N-morpholino)propanesulfonic acid (MOPS) buffered solutions were evaluated by isothermal titration calorimetry. Complexes formed in different relative ratios of mercury(II) to cysteinyl peptides in ammonium formate buffered solutions were characterized by LTQ Orbitrap mass spectrometry. The results from these studies show that n-alkyl dicysteinyl peptides (CP 14), and an aryl dicysteinyl peptide (CP 5) can serve as effective “double anchors” to accommodate the coordination sites of mercury(II) to form predominantly one-to-one Hg(peptide) complexes. The aryl dicysteinyl peptide (CP 5) also forms the two-to-two Hg2(peptide)2 complex. In the presence of excess peptide, Hg(peptide)2 complexes are also detected. Notably, increasing the distance between the ligating groups or “anchor points” in CP 15 does not significantly affect their affinity for mercury(II). However, the enthalpy change (ΔH) values (ΔH1  −91 kJ mol−1 and ΔH2  −66 kJ mol−1) for complex formation between CP 4 and 5 with mercury(II) are about one and a half times larger than the related values for CP 1, 2 and 3H1  −66 kJ mol−1 and ΔH2  46 kJ mol−1). The corresponding entropy change (ΔS) values (ΔS1  −129 J K−1 mol−1 and ΔS2  −116 J K−1 mol−1) of the structurally larger dicysteinyl peptides CP 4 and 5 are less entropically favorable than for CP 1, 2 and 3S1  −48 J K−1 mol−1 and ΔS2  −44 J K−1 mol−1). Generally, these associations result in a decrease in entropy, indicating that these peptide–mercury complexes potentially form highly ordered structures. The results from this study show that dicysteinyl tripeptides are effective in binding mercury(II) and they are promising motifs for the design of multi-cysteinyl peptides for binding more than one mercury(II) ion per peptide.  相似文献   

8.
Cyclooxygenases (COX) and 8R-dioxygenase (8R-DOX) activities of linoleate diol synthases (LDS) are homologous heme-dependent enzymes that oxygenate fatty acids by a tyrosyl radical-mediated hydrogen abstraction and antarafacial insertion of O2. Soybean lipoxygenase-1 (sLOX-1) contains non-heme iron and oxidizes 18:2n ? 6 with a large deuterium kinetic isotope effect (D-KIE). The aim of the present work was to obtain further mechanistic insight into the action of these enzymes by using a series of n ? 6 and n ? 9 fatty acids and by analysis of D-KIE. COX-1 oxidized C20 and C18 fatty acids in the following order of rates: 20:2n ? 6 > 20:1n ? 6 > 20:3n ? 9 > 20:1n ? 9 and 18:3n ? 3  18:2n ? 6 > 18:1n ? 6. 18:2n ? 6 and its geometrical isomer (9E,12Z)18:2 were both mainly oxygenated at C-9 by COX-1, but the 9Z,12E isomer was mostly oxygenated at C-13. A cis-configured double bond in the n ? 6 position therefore seems important for substrate positioning. 8R-DOX oxidized (9Z,12E)18:2 at C-8 in analogy with 18:2n ? 6, but the 9E,12Z isomer was mainly subject to hydrogen abstraction at C-11 and oxygen insertion at C-9 by 8R-DOX of 5,8-LDS. sLOX-1 and 13R-MnLOX oxidized [11S-2H]18:2n ? 6 with similar D-KIE (~ 53), which implies that the catalytic metals did not alter the D-KIE. Oxygenation of 18:2n ? 6 by COX-1 and COX-2 took place with a D-KIE of 3–5 as probed by incubations of [11,11-2H2]- and [11S-2H]18:2n ? 6. In contrast, the more energetically demanding hydrogen abstractions of the allylic carbons of 20:1n ? 6 by COX-1 and 18:1n ? 9 by 8R-DOX were both accompanied by large D-KIE (> 20).  相似文献   

9.
Hungate's method is a well-accepted protocol for the isolation or incubation of anaerobes with a roll tube technique. The aim of this study was to stimulate fungal enzyme production by optimizing the components of Hungate's medium for the growth of a rumen fungus Anaeromyces sp. YQ3. The organism was grown on corn stalks and incubated for 10 days in defined media with two glucose levels (G+, glucose in the Hungate's medium as a glucose control; G?, glucose removed in a modified Hungate's medium) and four N sources (N1: yeast extract + tryptone + (NH4)2SO4 in Hungate's medium (control); N2: yeast extract + (NH4)2SO4; N3: tryptone + (NH4)2SO4; and N4: tryptone + yeast extract). In the G? media, the recovered activities of feruloyl esterase (FAE) (P<0.0001), acetyl esterase (AE) (P=0.0065) and xylanase (P<0.0001) were decreased, while the G+ media with N1 nitrogen stimulated the production of FAE and xylanase (P<0.0001). The G? medium with N2 nitrogen increased the recovered activities of carboxymethyl cellulase (P=0.0001) and avicelase (P<0.0001), while the N3 and N4 media increased the recovered activity of AE (P=0.0015). The N4 medium was comparable to the N1 medium in stimulating the amount of recovered xylanase activity. The activities of FAE (P<0.0001), AE (P<0.0001), and xylanase (P<0.0001) showed a time-dependent increase and reached their peaks at day 10, while the avicelase activity peaked at day 8 (P=0.0071). The esterase activities (FAE and AE) were positively correlated with the enzyme activities of xylanase and carboxymethyl cellulase (r > 0.48, P<0.05). After a 10-day incubation, the glucose in the Hungate's media contributed to an increase in organic matter disappearance (P<0.0001) and volatile fatty acid (VFA) concentration (P<0.0001), except for molar acetate proportions. The N4 treatment increased organic matter disappearance and total VFA concentration (P=0.0002). The change in N source did not alter molar proportions of acetate, propionate and valerate, while the N2 treatment increased molar butyrate proportion (P<0.0035), and both N2 and N3 increased the molar proportion of branched chain VFAs (P<0.0041). In summary, the glucose in the Hungate's medium is beneficial for stimulating the production of esterases and xylanase, thereby promoting fungal growth. Amending the N source in Hungate's medium brings about different yields of rumen fungal esterases and polysaccharide hydrolases that have important nutritional impacts on fibre degradation in ruminant animals.  相似文献   

10.
Secoisolariciresinol diglucosides (SDGs) (S,S)-SDG-1 (major isomer in flaxseed) and (R,R)-SDG-2 (minor isomer in flaxseed) were synthesized from vanillin via secoisolariciresinol (6) and glucosyl donor 7 through a concise route that involved chromatographic separation of diastereomeric diglucoside derivatives (S,S)-8 and (R,R)-9. Synthetic (S,S)-SDG-1 and (R,R)-SDG-2 exhibited potent antioxidant properties (EC50 = 292.17 ± 27.71 μM and 331.94 ± 21.21 μM, respectively), which compared well with that of natural (S,S)-SDG-1 (EC50 = 275.24 ± 13.15 μM). These values are significantly lower than those of ascorbic acid (EC50 = 1129.32 ± 88.79 μM) and α-tocopherol (EC50 = 944.62 ± 148.00 μM). Compounds (S,S)-SDG-1 and (R,R)-SDG-2 also demonstrated powerful scavenging activities against hydroxyl [natural (S,S)-SDG-1: 3.68 ± 0.27; synthetic (S,S)-SDG-1: 2.09 ± 0.16; synthetic (R,R)-SDG-2: 1.96 ± 0.27], peroxyl [natural (S,S)-SDG-1: 2.55 ± 0.11; synthetic (S,S)-SDG-1: 2.20 ± 0.10; synthetic (R,R)-SDG-2: 3.03 ± 0.04] and DPPH [natural (S,S)-SDG-1: EC50 = 83.94 ± 2.80 μM; synthetic (S,S)-SDG-1: EC50 = 157.54 ± 21.30 μM; synthetic (R,R)-SDG-2: EC50 = 123.63 ± 8.67 μM] radicals. These results confirm previous studies with naturally occurring (S,S)-SDG-1 and establish both (S,S)-SDG-1 and (R,R)-SDG-2 as potent antioxidants and free radical scavengers for potential in vivo use.  相似文献   

11.
《Aquatic Botany》2007,87(2):104-110
A large-scale mesocosm (sixteen 500 L tanks) experiment was conducted to investigate the effects of hypersalinity (45–65 psu), porewater sulfide (2–6 mM) and nighttime water column hypoxia (5–3 mg L−1) on the tropical seagrass Thalassia testudinum Banks ex König. We examined stressor effects on growth, shoot survival, tissue sulfur (S0, TS, δ34S) and leaf quantum efficiencies, as well as, porewater sulfides (∑TSpw) and mesocosm water column O2 dynamics. Sulfide was injected into intact seagrass cores of T. testudinum exposing below-ground tissues to 2, 4, and 6 mM S2−, but rapid oxidation resulted in ∑TSpw < 1.5 mM. Hypersalinity at 65 psu lowered sulfide oxidation and significantly affected plant growth rates and quantum efficiencies (Fv/Fm < 0.70). The most depleted rhizome δ34S signatures were also observed at 65 psu, suggesting increased sulfide exposure. Hypoxia did not influence ∑TSpw and plant growth, but strengthened the hypersalinity response and decreased rhizome S0, indicating less efficient oxidation of ∑TSpw. Following nighttime hypoxia treatments, ecosystem level metabolism responded to salinity treatments. When O2 levels were reduced to 5 and 4 mg L−1, daytime O2 levels recovered to approximately 6 mg L−1; however, this recovery was more limited when O2 levels were lowered to 3 mg L−1. Subsequent to O2 reductions to 3 mg O2 L−1, nighttime O2 levels rose in the 35 and 45 psu tanks, stayed the same in the 55 psu tanks, and declined in the 65 psu tanks. Thus, hypersalinity at 65 psu affects T. testudinum's oxidizing capacity and places subtle demands on the positive O2 balance at an ecosystem level. This O2 demand may influence T. testudinum die-off events, particularly after periods of high temperature and salinity. We hypothesize that the interaction between hypersalinity and sulfide toxicity in T. testudinum is their synergistic effect on the critical O2 balance of the plant.  相似文献   

12.
《Aquatic Botany》2005,81(4):353-366
Carbon fixation and allocation were studied using 13C incubation and leaf marking techniques in mature monospecific stands of Enhalus acoroides L.f. Royle in August 1998 and January 1999 in Banten Bay, Indonesia. The highest rate of 13C uptake (>0.008 g 13C g C−1 d−1) was found in the middle to distal parts of leaves of E. acoroides. Young and senescing leaves numbers had lower 13C incorporation compared to mature leaves. The incorporation of 13C by epiphytes on the leaves was higher than that of the leaves themselves (>0.01 g 13C g C−1 d−1). The results showed that turbidity of the water influenced the leaf growth, productivity and Relative Growth Rate of E. acoroides, which were lower at Kepuh Island, the more turbid site. However, at Kepuh Island, where the water column was turbid, the plant could still harvest sufficient light for an uptake rate of 13C, higher than the uptake rates at Kubur and Panjang Islands, stations with a much more transparent water column (on average 0.0047 g 13C g C−1 d−1 at Kepuh Island, versus 0.0045 g 13C g C−1 d−1 at Panjang Island and 0.0034 g 13C g C−1 d−1 at Kubur Island). There was evidence that 13C was exported from the incubated shoots to the roots and rhizomes and to neighboring shoots of E. acoroides in clear water, but not in turbid water. We suggest that single shoots of E. acoroides are able to grow in turbid water under low light conditions. They assimilate sufficient carbon for their own maintenance but are not able to export to neighboring plant parts.  相似文献   

13.
Longan (Dimocarpus longan Lour, syn. Euphoria longan Lam.) represents an important fruit in Northern Thailand and has significant economic impact. The fruit is either consumed fresh or as commercially prepared dried and canned products. The canning industry in Thailand produces considerable quantities of waste products, in particular Longan seeds. Because these seeds may be an exploitable source of natural phenolic antioxidants, it was of interest to identify, purify and quantitate the major potential antioxidant phenolics contained therein. The polyphenolic fraction from ground Longan seeds was obtained by extraction with methanol after delipidation with hexane. The hexane extract contained predominantly long-chain fatty acids with major contributions from palmitic (35%) and oleic (28%) acids. The polyphenolic fraction (80.90 g/kg dry weight) was dominated by ellagic acid (25.84 g/kg) and the known ellagitannins corilagin (13.31 g/kg), chebulagic acid (13.06 g/kg), ellagic acid 4-O-α-l-arabinofuranoside (9.93 g/kg), isomallotinic acid (8.56 g/kg) and geraniin (5.79 g/kg). Structure elucidation was performed with mass spectrometry and complete assignment of 1H and 13C NMR signals. The methanol extracts exhibited strong antioxidant capacities with an IC50 of 154 μg/ml for reactive oxygen species attack on salicylic acid and 78 μg/ml for inhibition of xanthine oxidase in the hypoxanthine/xanthine oxidase assay. The extracts were less effective in the 2-deoxyguanosine assay (IC50 = 2.46 mg/ml), indicating that gallates along with ellagic acid and its congeners exert their potential antioxidant effects predominantly by precipitation of proteins such as xanthine oxidase. This was confirmed for the pure compounds gallic acid, methyl gallate, ellagic acid and corilagin.  相似文献   

14.
《Microbiological research》2014,169(4):279-286
Fatty acids as components of cuticular lipids of insects play a significant role in antifungal in protection against fungal infection. The chemical composition of cuticular and internal extracts obtained from all developmental stages of flesh flies Sarcophaga carnaria was identified. The fatty acids were detected using gas chromatography coupled with mass spectrometry and the most abundant for all examined stages were: 18:1 > 16:0 > 16:1 > 18:0 > 18:2. Polyunsaturated fatty acids (PUFA) C20 were found in both, cuticular and internal extracts. GC–MS analysis showed higher relative content of PUFA in adults than in preimaginal stages.Fatty acids alone as well as their cuticular and internal extracts obtained from larvae, pupae male and female of S. carnaria were tested according to their potential antimicrobial activity against entomopathogenic fungi: Paecilomyces lilacinus, Paecilomyces fumosoroseus, Lecanicillium lecanii, Metarhizium anisopliae, Beauveria bassiana (Tve-N39) and B. bassiana (Dv-1/07). FA presented diverse antimicrobial activity depending on the length of the chain and the presence of unsaturated bonds. Short chain and unsaturated FA (6:0, 11:0, 13:0) have shown significantly stronger activity against fungi but they were detected in lower concentrations. PUFA inhibit fungal growth more effectively than unsaturated long chain fatty acids. Cuticular and internal extracts of all living forms of S. carnaria exhibited approximately equal activity against tested entomopathogenic fungi. We presumed that the most abundant saturated long chain FA and additionally PUFA founded in our analysis are involved in protecting the flies against fungal infection.  相似文献   

15.
We have cloned, purified and investigated the catalytic activity and anion inhibition profiles of a full catalytic domain (358 amino acid residues) carbonic anhydrase (CA, EC 4.2.1.1) from Plasmodium falciparum, PfCAdom, an enzyme belonging to the η-CA class and identified in the genome of the malaria-producing protozoa. A truncated such enzyme, PfCA1, containing 235 residues was investigated earlier for its catalytic and inhibition profiles. The two enzymes were efficient catalysts for CO2 hydration: PfCAdom showed a kcat of 3.8 × 105 s−1 and kcat/Km of 7.2 × 107 M−1 × s−1, whereas PfCA showed a lower activity compared to PfCAdom, with a kcat of 1.4 × 105 s−1 and kcat/Km of 5.4 × 106 M−1 × s−1. PfCAdom was generally less inhibited by most anions and small molecules compared to PfCA1. The best PfCAdom inhibitors were sulfamide, sulfamic acid, phenylboronic acid and phenylarsonic acid, which showed KIs in the range of 9–68 μM, followed by bicarbonate, hydrogensulfide, stannate and N,N-diethyldithiocarbamate, which were submillimolar inhibitors, with KIs in the range of 0.53–0.97 mM. Malaria parasites CA inhibition was proposed as a new strategy to develop antimalarial drugs, with a novel mechanism of action.  相似文献   

16.
Alkyl hydroperoxide reductase (AhpC) is known to detoxify peroxides and reactive sulfur species (RSS). However, the relationship between its expression and combating of abiotic stresses is still not clear. To investigate this relationship, the genes encoding the alkyl hydroperoxide reductase (ahpC) from Anabaena sp. PCC 7120 were introduced into E. coli using pGEX-5X-2 vector and their possible functions against heat, salt, carbofuron, cadmium, copper and UV-B were analyzed. The transformed E. coli cells registered significantly increase in growth than the control cells under temperature (47 °C), NaCl (6% w/v), carbofuron (0.025 mg ml?1), CdCl2 (4 mM), CuCl2 (1 mM), and UV-B (10 min) exposure. Enhanced expression of ahpC gene as measured by semi-quantitative RT-PCR under aforementioned stresses at different time points demonstrated its role in offering tolerance against multiple abiotic stresses.  相似文献   

17.
The direct interaction of the antibiotic primycin with the plasma membrane was investigated by employing the well-characterized ergosterol-producing, amphotericin B-sensitive parental Candida albicans strain 33erg+ and its ergosterol-less amphotericin B-resistant plasma membrane mutant erg-2. The growth inhibition concentration in shaken liquid medium was 64 μg ml?1 for 33erg+ and 128 μg ml?1 for erg-2, suggesting that the plasma membrane composition influences the mode of action of primycin. To determine the primycin-induced changes in the plasma membrane dynamic, electron paramagnetic resonance (EPR) spectroscopy methods were used, the spin-labeled fatty acid 5-(4,4-dimethyloxazolidine-N-oxyl)stearic acid) being applied for the in vivo measurements. The phase transition temperatures of untreated strain 33erg+ and its mutant erg-2 were 12.5 °C and 11 °C, respectively. After 128 μg ml?1 primycin treatment, these values increased to 17.5 °C and 16 °C, revealing a significant reduction in the phospholipid flexibility. Saturation transfer EPR measurements demonstrated that, the rotational correlation times of the spin label molecule for the control samples of 33erg+ and erg-2 were 60 ns and 100 ns. These correlation times gradually decreased on the addition of increasing primycin concentrations, reaching 8 μs and 1 μs. The results indicate the plasma membrane “rigidizing” effect of primycin, a feature that may stem from its ability to undergo complex formation with membrane constituent fatty acid molecules, causing alterations in the structures of phospholipids in the hydrophobic surface near the fatty acid chain region.  相似文献   

18.
Docosahexaenoic acid (DHA) is an important polyunsatured fatty acid (PUFA) which can be purified from tuna fish oil fatty acids by selective enzymatic esterification. The present paper investigates the kinetic study for selective esterification of tuna fish oil fatty acids with butanol catalyzed by Rhizopus oryzae lipase (ROL) in biphasic solvent system. Under the most suitable reaction conditions, 76.2% esterification was achieved in 24 h. Different kinetic models for esterification given by Segel [1], Oliveira et al. [2], Gogoi et al. [3], and Kraai et al. [4] were tested for fitting the esterification data and the model given by Oliveira et al. [2] was found to be most suitable. The model given by Prazeres et al. [5] for hydrolysis was also tested for esterification and the model with second order product inhibition was found to provide better match between the predicted and experimental values than that of model by Oliveira et al. [2]. The kinetic model was fitted using MATLAB® to determine the best kinetic parameters. The average value of kinetic constants using the model given by Prazeres et al. were estimated as Km = 23.6 μmoles FFA/ml, Ki1 = 4.6 × 10−5 μmoles FFA/mg enzyme h, Ki2 = 0.0062 μmoles FFA/mg enzyme h and K2 = 149.5 μmoles FFA/mg enzyme h.  相似文献   

19.
The effects of five alternative nitrogen sources, namely, malt sprout (MS), corn steep liquor (CSL), NH4Cl, NH4NO3 and diamine citrate (DC) were investigated on the l-(+)-lactic acid (LA) production by thermophile Lactobacillus plantarum As.1.3. Through the statistical analysis of the results by three steps of response surface methodology (RSM) design, MS and CSL were found to have significant effects on the LA production and their optimal concentrations in the medium should be 16.0 g/L and 12.0 g/L, respectively. The verification of the optimized medium showed that the maximum specific growth rate (μm) was 1.09 h−1, the cell yield coefficient (YX/S) and the l-(+)-lactic acid yield coefficient (YP/S) were 0.233 (OD620/g) and 0.98 (g/g), and the maximum volumetric productivity and the average volumetric productivity were 13.0 g/L h and 3.20 g/L h, respectively. The results indicate that the LA production can also be enhanced with the inexpensive nitrogen source alternatives.  相似文献   

20.
Okadaic acid (OA) and dinophysistoxin-1 (DTX1) cause diarrheic shellfish poisoning. This article examines the biochemical interactions of the two toxins with novel okadaic acid binding proteins (OABPs) 2.1 and 2.3, originally isolated from the marine sponge Halichondria okadai. First, recombinant OABPs 2.1 and 2.3 were expressed in Escherichia coli BL21 (DE3) cells. Binding assays using [24-3H]OA and the recombinant OABP 2.1 or 2.3 demonstrated the dissociation constant Kd of 1.30 ± 0.56 nM and 1.54 ± 0.35 nM, respectively. Binding of [24-3H]okadaic acid to recombinant OABP2.1 was almost equally replaced with OA and DTX1. OA-induced cytotoxicity in mouse leukemia P388 cells was inhibited in the presence of the recombinant OABPs 2.1 and 2.3 with an EC50 of 92 ± 8.4 nM and 87 ± 13 nM, respectively. These results suggest that the blockage of OA-induced cytotoxicity by OABPs 2.1 and 2.3 may be involved in regulating symbiotic relationships present in the sponge H. okadai.  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