首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Conformational studies of poly-L-alanine in water   总被引:5,自引:0,他引:5  
The conformational properties of poly-L -alanine have been examined in aqueous solutions in order to investigate the influence of hydrophobic interactions on the helix–random coil transition. Since water is a poor solvent for poly-L -alanine, water-soluble copolymers of the type (D , L -lysine)m–(L alanine)n-(D , L -lysine)m, having 10, 160, 450, and 1000 alanyl residues, respectively, in the central block, were synthezised. The optical rotatory dispersion of the samples was investigated in the range 190–500 mμ, and the rotation at 231 mμ was related to the α-helix content, θH, of the alanine section. In salt-free solutions, at neutral pH, the three large polymers show high θH values, which are greatly reduced when the temperature is increased from 5 to 80°C. No helicity was observed for the small (n = 10) polymer. By applying the Lifson-Roig theory, the following parameters were obtained for the transition of a residue from a coil to a helical state: ν = 0.012; ΔH = ?190 ± 40 cal./mole; ΔS = ?0.55 ± 0.12 e.u. Since ΔH and ΔS differ from the values expected for a process involving only the formation of a hydrogen bond, and in a manner predicted by theories for the influence of hydrophobic bonding on helix stability, it is concluded that a hydrophobic interaction is also involved. In the presence of salt (0.2M NaCl), or when the ε-amino groups of the lysyl residues are not protonated (pH = 12), the helical form of the two large polymers (n = 450 and n = 1000) is more stable than in water. Since the electrostatic repulsion between the lysine end blocks is greatly reduced under these conditions, the alanine helical sections fold back on themselves, and this conformation is stabilized by interchain hydrophobia bonds. This structure was predicted by the theory for the equilibrium between such interacting helices, non-interacting helices, and the random coil.  相似文献   

2.
Aims: To investigate hydrocarbon degradation by hydrophobic, hydrophilic and parental strains of Pseudomonas aeruginosa. Methods and Results: Partitioning of hydrocarbon‐degrading P. aeruginosa strain in a solvent/aqueous system yielded hydrophobic and hydrophilic fractions. Exhaustive partitioning of aqueous‐phase cells yielded the hydrophilic variants (L), while sequential fractionation of the hydrophobic phase cells yielded successive fractions exhibiting increasing cell‐surface hydrophobicity (CSH). In hydrocarbon adherence assays (bacterial attachment to hydrocarbon), L had a value of 20%, which increased from 61·7% in first hydrophobic fraction (H1) to 72·2% in the third (H3). Crude oil degradation by L was 70%, but increased from 82% in H1 to 93% in H3. L variant produced most exopolysaccharides and reduced surface tension from about 73 to 49 mN m?1. Rhamnolipid production was highest in L, but was not detected in all crude oil cultures. Conclusions: Hydrophobic subpopulations of hydrocarbon‐degrading P. aeruginosa exhibited greater hydrocarbon‐utilizing ability than hydrophilic ones, or the parental strain. Significance and Impact of the Study: Results demonstrate that a population of P. aeruginosa consists of cells with different CSH which affect hydrocarbon utilization. This potentially provides the population with the capacity to utilize different hydrophobic substrates found in petroleum. Judicious selection of such hydrophobic subpopulations can enhance hydrocarbon pollution bioremediation.  相似文献   

3.
Chain collapse and secondary structure formation are frequently observed during the early stages of protein folding. Is the chain collapse brought about by interactions between secondary structure units or is it due to polymer behavior in a poor solvent (coil‐globule transition)? To answer this question, we measured small‐angle X‐ray scattering for a series of β‐lactoglobulin mutants under conditions in which they assume a partially folded state analogous to the folding intermediates. Mutants that were designed to disrupt the secondary structure units showed the gyration radii similar to that of the wild type protein, indicating that chain collapse is due to coil‐globule transitions. © 2013 Wiley Periodicals, Inc. Biopolymers 101: 651–658, 2014.  相似文献   

4.
A compound binding three Gd3+ ions, {Ph4[Gd(DTTA)(H2O)2]? 3} (where H5DTTA is diethylenetriaminetetraacetic acid), has been synthesized around a hydrophobic center made up of four phenyl rings. In aqueous solution the molecules start to self-aggregate at concentrations well below 1 mM as shown by the increase of rotational correlation times and by the decrease of the translational self-diffusion constant. NMR spectra recorded in aqueous solution of the diamagnetic analogue {Ph4[Y(DTTA)(H2O)2]? 3} show that the aggregation is dynamic and due to intermolecular π-stacking interactions between the hydrophobic aromatic centers. From estimations of effective radii, it can be concluded that the aggregates are composed of two to three monomers. The paramagnetic {Ph4[Gd(DTTA)(H2O)2]? 3} exhibits concentration-dependent 1H NMR relaxivities with high values of approximately 50 mM?1 s?1 (30 MHz, 25 °C) at gadolinium concentrations above 20 mM. A combined analysis of 1H NMR dispersion profiles measured at different concentrations of the compound and 17O NMR data measured at various temperatures was performed using different theoretical approaches. The fitted parameters showed that the increase in relaxivity with increasing concentration of the compound is due to slower global rotational motion and an increase of the Lipari–Szabo order parameter S 2.  相似文献   

5.
Backbone 15N relaxation parameters and 15N–1HN residual dipolar couplings (RDCs) have been measured for a variant of human α-lactalbumin (α-LA) in 4, 6, 8 and 10 M urea. In the α-LA variant, the eight cysteine residues in the protein have been replaced by alanines (all-Ala α-LA). This protein is a partially folded molten globule at pH 2 and has been shown previously to unfold in a stepwise non-cooperative manner on the addition of urea. 15N R2 values in some regions of all-Ala α-LA show significant exchange broadening which is reduced as the urea concentration is increased. Experimental RDC data are compared with RDCs predicted from a statistical coil model and with bulkiness, average area buried upon folding and hydrophobicity profiles in order to identify regions of non-random structure. Residues in the regions corresponding to the B, D and C-terminal 310 helices in native α-LA show R2 values and RDC data consistent with some non-random structural propensities even at high urea concentrations. Indeed, for residues 101–106 the residual structure persists in 10 M urea and the RDC data suggest that this might include the formation of a turn-like structure. The data presented here allow a detailed characterization of the non-cooperative unfolding of all-Ala α-LA at higher concentrations of denaturant and complement previous studies which focused on structural features of the molten globule which is populated at lower concentrations of denaturant.  相似文献   

6.
The optimized geometries, electron affinities, and dissociation energies of the alkylthio radicals have been determined with the higher level of the Gaussian-3(G3) theory. The geometries are fully optimized and discussed. The reliable adiabatic electron affinities with ZPVE correction have been predicted to be 1.860 eV for the methylthio radical, 1.960 eV for the ethylthio radical, 1.980 and 2.074 eV for the two isomers (n-C3H7S and i-C3H7S) of the propylthio radical, 1.991, 2.133 and 2.013 eV for the three isomers (n-C4H9S, t-C4H9S, and i-C4H9S) of the butylthio radical, and 1.999, 2.147, 2.164, and 2.059 eV for the four isomers (n-C5H11S, b-C5H11S, c-C5H11S, and d-C5H11S) of the pentylthio radical, respectively. These corrected EAad values for the alkylthio radicals are in good agreement with available experiments, and the average absolute error of the G3 method is 0.041 eV. The dissociation energies of S atom from neutral CnH2n+1S (n?=?1–5) and S- from corresponding anions CnH2n+1S- species have also been estimated respectively to examine their relative stabilities.  相似文献   

7.
The rate of reaction of [Ce(EDTA)(OH)nn] with H2O2 in 0.10 M KNO3 solution was investigated at various temperatures. The presence of a peroxy intermediate is inferred from spectrophotometric measurements. The general rate equation,
is valid for pH 7-9 with n= 1 and 2 complexes involved. The rate constants kl and k2 were determined at 25 °C to be 0.054 and 0.171 M−1 s−1 respectively. The corresponding activation enthalpies, as calculated from Arrhenius plots, were δH1#= 51.3 ± 14.8 and δH2#= 41.8 ± 5.3 kJ m−1 and the activation entropies were δS1#=-97 ± 47 and ΔS2#=−119±17 J K−1 m−1.  相似文献   

8.
The enthalpies of the hexokinase-catalyzed phosphorylation or glucose, mannose, and fructose by ATP to the respective hexose 6-phosphates have been measured calorimetrically in TRIS/TRIS HCl buffer at 25.0, 28.5, and 32.0°C. The effects on the measured enthalpy of the glucose/hexokinase reaction due to variation of pH (over the range 6.7 to 9.0) and ionic strength (over the range 0.02 to 0.25) have been examined. Correction for enthalpy of buffer protonation leads to δHo and δCpo values for the processes: eq-D-hexose + ATP4− = eq-D-hexose 6-phosphate2− + ADP3−+ H+. Results are δHo = −23.8 ± 0.7 kJ · mol−1 and δCpo = −156 ± 280 J·mol−1·K−1 for glucose. δHo = −21.9 ± 0.7 kJ·mol−1 and δCpo = 10 ± 140 J·mol−1·K−1 for mannose, and δHo = −15.0 ± 0.9 kJ·mol−1 and δCpo = −41 ± 160 J·mol−1·K−1 for fructose. Combination of these measured enthalpies with Gibbs energy data for hydrolysis of ATP4− and that for the hexose 6-phosphates lead to δSo values for the above hexokinase-catalyzed reactions.  相似文献   

9.
The fatty acid and retinol-binding (FAR) proteins are a family of unusual helix-rich lipid binding proteins found exclusively in nematodes, and are secreted by a range of parasites of humans, animals and plants. Na-FAR-1 is from the parasitic nematode Necator americanus, an intestinal blood-feeding parasite of humans. Sequence-specific 1H, 13C and 15N resonance assignments have been obtained for the recombinant 170 amino acid protein, using three-dimensional triple-resonance heteronuclear magnetic resonance experiments. Backbone assignments have been obtained for 99.3 % of the non-proline HN/N pairs (146 out of 147). The amide resonance of T45 was not observed, probably due to rapid exchange with solvent water. A total of 96.9 % of backbone resonances were identified, while 97.7 % assignment of amino acid sidechain protons is complete. All Hα(166), Hβ(250) and Hγ(160) and 98.4 % of the Hδ (126 out of 128) atoms were assigned. In addition, 99.4 % Cα (154 out of 155) and 99.3 % Cβ (143 out of 144) resonances have been assigned. No resonances were observed for the NHn groups of R93 NεHε, arginine, Nη1H2, Nη2H2, histidine Nδ1Hδ1, Nε1Hε1 and lysine Nζ3H3. Na-FAR-1 has a similar overall arrangement of α-helices to Ce-FAR-7 of the free-living Caeorhabditis elegans, but with an extra C-terminal helix.  相似文献   

10.
To determine the effect of denaturants [guanidine hydrochloride (GdnHCl) and urea] and polyols [with various molecular masses (62.1–600)] on calcium binding at the two hypothesized conformers (A and B forms) of the chemically equivalent sarcoplasmic reticulum Ca2+-ATPase, which bind two calcium ions in different manners, we examined the effect of these reagents on the calcium dependence of ATP-supported phosphorylation of the ATPase molecules and of their calcium-activated, acetyl phosphatate hydrolytic activity. (1) GdnHCl (~0.05 M) and urea (~0.5 M) increased the apparent calcium affinity (K 0.5) of 2–6 μM of noncooperative binding [Hill coefficient (n H) ~ 1] of the A form to 10–40 μM. (2) The employed polyols transformed the binding of the A form into cooperative binding (n H ~ 2), accompanying the approach of its K 0.5 value to that (K 0.5 = 0.04–0.2 μM) of the cooperative binding (n H ~ 2) of the B form; the transition concentration (0.025–2 M) of the polyols, above which such transformation occurs, was in inverse relation to their molecular mass. (3) The binding of the B form was resistant to these denaturants and polyols. Based on these data, a structural model of the two forms, calcium-binding domains of which are loosely and compactly folded, is presented.  相似文献   

11.
Toxicological and pharmacological studies demonstrated that the introduction of functional groups into the aromatic ring of diphenyl diselenide alter its effect. The aim of this study was to evaluate the in vitro effect of m-trifluoromethyl-diphenyl diselenide (m-CF3–C6H4Se)2, p-chloro-diphenyl diselenide (p-Cl–C6H4Se)2 and p-methoxyl-diphenyl diselenide (p-CH3O–C6H4Se)2 on δ-aminolevulinate dehydratase (δ-ALA-D) and Na+, K+-ATPase activities in rat brain homogenates. Diselenides inhibited δ-ALA-D activity (IC50 4–6 μM [concentration inhibiting 50%]), and dithiothreitol (DTT) restored the enzyme activity. ZnCl2 (100 μM) did not restore δ-ALA-D inhibition caused by (p-Cl–C6H4Se)2 and (m-CF3–C6H4Se)2. Na+, K+-ATPase activity was more sensitive to (p-Cl–C6H4Se)2 and (m-CF3–C6H4Se)2 (IC50 6 μM) than (p-CH3O–C6H4Se)2 and (PhSe)2 (IC50 45 and 31 μM, respectively). DTT restored the activity of Na+, K+-ATPase inhibited by diselenides. The effect of diselenides on Na+/K+-ATPase is dependent on their substitutions in the aromatic ring. The mechanism through which diselenides inhibit δ-ALA-D and Na+, K+-ATPase activities involves the oxidation of thiol groups.  相似文献   

12.
This study reports bioavailability and metabolism of fucoxanthin (FUCO) from brown algae Padina tetrastromatica in rats. Rats were divided into two groups (n = 25/group). Group one was fed basal diet (control) while the group two received retinol deficient diet (RD group) for 8 weeks. After confirmed RD in blood (0.53 μmol/l), rats were further sub-grouped (n = 5/sub group), intubated a dose of FUCO (0.83 μmol) and killed after 0, 2, 4, 6 and 8 h. The plasma levels (area under curve/8 h) of FUCO (fucoxanthinol (FUOH) + amarouciaxanthin (AAx)) was 2.93 (RD group) and 2.74 pmol/dl (control), respectively. No newly formed retinol was detected in RD rats intubated with FUCO. Besides FUOH (m/z 617 (M+H)+) and AAx (m/z 617 (M+H?)+), other deacetylated, hydrolyzed and demethylated metabolites of bearing molecular mass at m/z 600.6 (FUOH–H2O), m/z 597 (AAx–H2O), m/z 579 (AAx–2H2O+1), m/z 551 (AAx–2H2O–2CH3+2) and m/z 523 (AAx–2H2O–4CH3+4) were also detected in plasma and liver by LC-MS (APCI). Although biological functions of FUCO metabolites need thorough investigation, this is the first detailed report on FUCO metabolites in rats.  相似文献   

13.
Glycerol and glucose fermentation redox routes by Escherichia coli and their regulation by oxidizing and reducing reagents were investigated at different pHs. Cell growth was followed by decrease of pH and redox potential (E h ). During glycerol utilization at pH 7.5 ?pH, the difference between initial and end pH, was lower compared with glucose fermentation. After 8 h growth, during glycerol utilization E h dropped down to negative values (?150 mV) but during glucose fermentation it was positive (+50 mV). In case of glycerol H2 was evolved at the middle log phase while during glucose fermentation H2 was produced during early log phase. Furthermore, upon glycerol utilization, oxidizer potassium ferricyanide (1 mM) inhibited both cell growth and H2 formation. Reducing reagents dl-dithiothreitol (3 mM) and dithionite (1 mM) inhibited growth but stimulated H2 production. The findings point out the importance of reductive conditions for glycerol fermentation and H2 production by E. coli.  相似文献   

14.
We designed nine endohedral dodecahedrane heterodimers H@C20Hn-C20Hn@M (M = Cu, Ag, and Au, n = 15, 18, and 19) that may act as single-molecule spin switches, and we predicted theoretically that the ground states of the dimmers shift from low-spin states (S = 0) to the high-spin states (S = 1) under an external electric field applied parallel or perpendicular to the molecular symmetry axes, consisting well with the analyses of Stark effect. Molecular orbitals analyses provide an intuitive insight into the spin crossover behavior. This study expands the application of endohedral chemistry and provides new molecules for designing single-molecule spin switch.  相似文献   

15.
Cysteine has been implicated in myocardial protection, although this is controversial and constrained by limited knowledge about the effects of cysteine at the cellular level. This study tested the hypothesis that a physiologically relevant dose of l-cysteine could be safely loaded into isolated cardiomyocytes leading to improved protection against oxidative stress. Freshly isolated adult rat ventricular cardiomyocytes were incubated for 2 h at 37°C with (cysteine incubated) or without (control) 0.5 mM cysteine prior to washing and suspension in fresh cysteine-free media. Cysteine incubated cells had higher intracellular cysteine levels compared to controls (9.6 ± 0.78 vs. 6.5 ± 0.65 nmol/mg protein, P < 0.02, n = 6 ± SE). Cell homeostasis indicators were similar in the two groups. Cysteine incubated cells had significantly higher glutathione peroxidase (GPx) activity (1.11 ± 0.23 vs. 0.54 ± 0.1 U/mg protein, P < 0.05, n = 5 ± SE) and significantly greater expression of GPx-1 (5.01 ± 0.48 vs. 3.01 ± 0.25 OD units/mm2, P < 0.05, n = 4 ± SE) compared to controls. Upon exposure to H2O2, cysteine incubated cells generated fewer reactive oxygen species and took longer to show contractile changes and undergo hypercontracture. However, when cells were exposed to H2O2 in the presence of 0.05 mM of the GPx inhibitor mercaptosuccinic acid, this increased the control cells’ susceptibility to H2O2 and completely abolished the cysteine mediated protection. These results suggest a new role for cysteine in myocardial protection involving stimulation of glutathione peroxidase.  相似文献   

16.
A simple and novel procedure for the acceleration of fungal spore production was developed. A net of hydrophobic polymer such as polypropylene (PP) and polytetrafluoroethylene (PTFE) was embedded in a nutrient agar plate, and effect of the polymer net on spore production by 6 fungal strains, such as Aspergillus terreus, Penicillium multicolor, and Trichoderma virens were estimated. The effect of hydrophobic polymer net was insufficient in a liquid-surface immobilization (LSI) system with fungal cells immobilized on a ballooned microsphere layer formed on a liquid medium surface. On the other hand, the embedding of a PTFE net in an agar plate remarkably enhanced the spore production in all 6 strains tested to produce 2.0–8.5 × 107 spores/cm2-agar plate surface. Especially, the spore production by A. terreus ATCC 20542 in the presence of a PTFE net was 7.7 times as much than that in no net. Positive correlations between the hydrophobicity of net and the spore production were observed in all 6 strains (R2, 0.653–0.999).  相似文献   

17.
Oxidative stress caused by mercury (Hg) was investigated in Pfaffia glomerata plantlets grown in nutrient solution using sand as substrate. Thirty-day-old acclimated plants were treated for 9 days with four Hg levels (0, 1, 25 and 50 μM) in the substrate. Parameters such as growth, tissue Hg concentration, toxicity indicators (δ-aminolevulinic acid dehidratase, δ-ALA-D, activity), oxidative damage markers (TBARS, lipid peroxidation, and H2O2 concentration) and enzymatic (superoxide dismutase, SOD, catalase, CAT, and ascorbate peroxidase, APX) and non-enzymatic (non-protein thiols, NPSH, ascorbic acid, AsA, and proline concentration) antioxidants were investigated. Tissue Hg concentration increased with Hg levels. Root and shoot fresh weight and δ-ALA-D activity were significantly decreased at 50 μM Hg, and chlorophyll and carotenoid concentration were not affected. Shoot H2O2 concentration increased curvilinearly with Hg levels, whereas lipid peroxidation increased at 25 and 50 μM Hg, respectively, in roots and shoots. SOD activity showed a straight correlation with H2O2 concentration, whereas CAT activity increased only in shoots at 1 and 50 μM Hg. Shoot APX activity was either decreased at 1 μM Hg or increased at 50 μM Hg. Conversely, root APX activity was only increased at 1 μM Hg. In general, AsA, NPSH and proline concentrations increased upon addition of Hg, with the exception of proline in roots, which decreased. These changes in enzymatic and non-enzymatic antioxidants had a significant protective effect on P. glomerata plantlets under mild Hg-stressed conditions.  相似文献   

18.
Based on Hamid model of 11Å tobermorite, amorphous calcium silicates hydrates (or C-S-H) structures (Ca4Si6O14(OH)4?2H2O as the C-S-H(I) and (CaO)1.67(SiO2)(H2O)1.75 as the C-S-H(II)) with the Ca/Si ratio of 0.67 and 1.7 are concerned. Then, as the representative ‘globule’ C-S-H, two amorphous C-S-H structures with the size of 5.352 × 4.434 × 4.556 nm3 during the stretch process are simulated at a certain strain rate of 10?3 ps?1 by LAMMPS program for molecular dynamics simulation, using ClayFF force field. The tensile stress–strain curves are obtained and analysed. Besides, elastic modulus of the ‘globule’ C-S-H is calculated to assess the elastic modulus of C-S-H phases (the low-density C-S-H – LD C-S-H – and the high-density C-S-H – HD C-S-H), where the porosity is a critical factor for explaining the relationship between ‘globule’ C-S-H at nanoscale and C-S-H phases at microscale. Results show that: (1) The C-S-H(I) structure has transformed from crystalline to amorphous during the annealing process, Young’s moduli in x, y and z directions are almost the same. Besides, the extent of aggregation and aggregation path for water molecules in the structure is different in three directions. (2) Young’s modulus of both amorphous C-S-H(I) and C-S-H(II) structures with a size of about 5 nm under strain rate of 10?3 ps?1 at 300 K in three directions is averaged to be equal, of which C-S-H(II) structure is about 60.95 GPa thus can be seen as the elastic modulus of the ‘globule’ C-S-H. (3) Based on the ‘globule’ C-S-H, the LD C-S-H and HD C-S-H can be assessed by using the Self-Consistent Scheme (separately 18.11 and 31.45 GPa) and using the Mori–Tanaka scheme (29.78 and 37.71 GPa), which are close to the nanoindentation experiments by Constantinides et al. (21.7 and 29.4 GPa).  相似文献   

19.
Light-scattering studies were done to investigate the DNA collapse transition, a large and discontinuous reduction in the radius of gyration. Of particular concern was differentiating the compaction of a single DNA molecule from aggregation. Solutions of RK2 plasmid DNA (Mr = 37 × 106) or bacteriophage T7 DNA (Mr = 25 × 106) were titrated with the condensing reagents spermidine in aqueous solvent or magnesium ion in ethanol–water solvent. The transition was followed by the change in scattering at a single angle or by the change in the angular dependence of scattering. At concentrations below 1 μg/mL, only aggregation could be detected by observation at a single angle; therefore, to study the collapse transition, it was necessary to measure the angular dependence of scattering. The intensities measured between the angles 30° and 60° were fit to known scattering functions. At low concentrations of the condensing reagent, the data were consistent with the scattering function of a random coil. On the other hand, during the transition at higher reagent concentrations, the curve that fit the data required two components—the scattering function for a random coil with a large radius of gyration, plus that for a sphere with a radius about one-fifth of that of the coil. The fractional concentration of the sphere increased with increasing condensing-reagent concentration. This two-component behavior is in apparent contrast to the situation with a more flexible polymer such as polystyrene, in accord with theoretical predictions. At still higher reagent concentrations, aggregation was apparent. Condensation to a collapsed state was reversible without hysteresis, while dissolution of the aggregated state nearly always occurred with hysteresis. Qualitative agreement between the observed DNA collapse transition and the theoretical phase diagram presented in the preceding paper was found, although the light-scattering results did not show quantitative agreement with the simple theoretical model.  相似文献   

20.
Temporin-SHc (FLSHIAGFLSNLFamide) first isolated from skin extraction of the Tunisian frog Pelophylax saharica, which shows potent antimicrobial activity against Gram-positive bacteria and is highly active against yeasts and fungi without hemolytic activity at antimicrobial concentrations. The peptide adopts well-defined α-helical conformation when bound to SDS micelles. In this study, we explored the effects of residue at position 5 and the N-terminus hydrophobic character on the hydrophilic/polar face of temp-SHc, on its biological activities (antimicrobial and hemolytic) and biophysical properties (hydrophobicity, amphipathicity and helicity). Antibacterial and hemolytic properties of temporin-SHc derivatives depend strongly on physicochemical properties. Therefore, slight decreasing amphipathicity together with hydrophobicity and helicity by the substitution Ile5 → Leu decreased antimicrobial potency approximately twofold without changing of hemolytic activity. It is noteworthy that a conservative amino acid substitution decreases the antimicrobial activity, underlining the differences between Leu/Ile side chains insertion into the lipid bilayer. While the modification of N-terminal hydrophobic character by four residue inversion decreased amphipathicity (twofold) of (4-1)L5temp-SHc and resulted in an increase in antibacterial activity against E. coli, E. faecalis and C. parapsilosis of at least fourfold, its therapeutic potential is limited by its drastic increase of hemolysis (LC50 = 2 μM). We found that the percentage of helicity of temp-SHc analog is directly correlated to its hemolytic activity. Last, the hydrophobic N-terminal character is an important determinant of antimicrobial activity.  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