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 665 毫秒
1.
Heme‐containing catalases and catalase‐peroxidases catalyze the dismutation of hydrogen peroxide as their predominant catalytic activity, but in addition, individual enzymes support low levels of peroxidase and oxidase activities, produce superoxide, and activate isoniazid as an antitubercular drug. The recent report of a heme enzyme with catalase, peroxidase and penicillin oxidase activities in Bacillus pumilus and its categorization as an unusual catalase‐peroxidase led us to investigate the enzyme for comparison with other catalase‐peroxidases, catalases, and peroxidases. Characterization revealed a typical homotetrameric catalase with one pentacoordinated heme b per subunit (Tyr340 being the axial ligand), albeit in two orientations, and a very fast catalatic turnover rate (kcat = 339,000 s?1). In addition, the enzyme supported a much slower (kcat = 20 s?1) peroxidatic activity utilizing substrates as diverse as ABTS and polyphenols, but no oxidase activity. Two binding sites, one in the main access channel and the other on the protein surface, accommodating pyrogallol, catechol, resorcinol, guaiacol, hydroquinone, and 2‐chlorophenol were identified in crystal structures at 1.65–1.95 Å. A third site, in the heme distal side, accommodating only pyrogallol and catechol, interacting with the heme iron and the catalytic His and Arg residues, was also identified. This site was confirmed in solution by EPR spectroscopy characterization, which also showed that the phenolic oxygen was not directly coordinated to the heme iron (no low‐spin conversion of the FeIII high‐spin EPR signal upon substrate binding). This is the first demonstration of phenolic substrates directly accessing the heme distal side of a catalase. Proteins 2015; 83:853–866. © 2015 Wiley Periodicals, Inc.  相似文献   

2.
It is demonstrated that cyanobacteria (both azotrophic and non‐azotrophic) contain heme b oxidoreductases that can convert chlorite to chloride and molecular oxygen (incorrectly denominated chlorite ‘dismutase’, Cld). Beside the water‐splitting manganese complex of photosystem II, this metalloenzyme is the second known enzyme that catalyses the formation of a covalent oxygen–oxygen bond. All cyanobacterial Clds have a truncated N‐terminus and are dimeric (i.e. clade 2) proteins. As model protein, Cld from Cyanothece sp. PCC7425 (CCld) was recombinantly produced in Escherichia coli and shown to efficiently degrade chlorite with an activity optimum at pH 5.0 [kcat 1144 ± 23.8 s?1, KM 162 ± 10.0 μM, catalytic efficiency (7.1 ± 0.6) × 106 M?1 s?1]. The resting ferric high‐spin axially symmetric heme enzyme has a standard reduction potential of the Fe(III)/Fe(II) couple of ?126 ± 1.9 mV at pH 7.0. Cyanide mediates the formation of a low‐spin complex with kon = (1.6 ± 0.1) × 105 M?1 s?1 and koff = 1.4 ± 2.9 s?1 (KD ~ 8.6 μM). Both, thermal and chemical unfolding follows a non‐two‐state unfolding pathway with the first transition being related to the release of the prosthetic group. The obtained data are discussed with respect to known structure–function relationships of Clds. We ask for the physiological substrate and putative function of these O2‐producing proteins in (nitrogen‐fixing) cyanobacteria.  相似文献   

3.
The Nostoc sp (Ns) H‐NOX (heme‐nitric oxide or OXygen‐binding) domain shares 35% sequence identity with soluble guanylate cyclase (sGC) and exhibits similar ligand binding property with the sGC. Previously, our molecular dynamic (MD) simulation work identified that there exists a Y‐shaped tunnel system hosted in the Ns H‐NOX interior, which servers for ligand migration. The tunnels were then confirmed by Winter et al. [PNAS 2011;108(43):E 881–889] recently using x‐ray crystallography with xenon pressured conditions. In this work, to further investigate how the protein matrix of Ns H‐NOX modulates the ligand migration process and how the distal residue composition affects the ligand binding prosperities, the free energy profiles for nitric oxide (NO), carbon monooxide (CO), and O2 migration are explored using the steered MDs simulation and the ligand binding energies are calculated using QM/MM schemes. The potential of mean force profiles suggest that the longer branch of the tunnel would be the most favorable route for NO migration and a second NO trapping site other than the distal heme pocket along this route in the Ns H‐NOX was identified. On the contrary, CO and O2 would prefer to diffuse via the shorter branch of the tunnel. The QM/MM (quantum mechanics/molecular mechanics) calculations suggest that the hydrophobic distal pocket of Ns H‐NOX would provide an approximately vacuum environment and the ligand discrimination would be determined by the intrinsic binding properties of the diatomic gas ligand to the heme group. Proteins 2013; 81:1363–1376. © 2013 Wiley Periodicals, Inc.  相似文献   

4.
Unlike most small globular proteins, lipocalins lack a compact hydrophobic core. Instead, they present a large central cavity that functions as the primary binding site for hydrophobic molecules. Not surprisingly, these proteins typically exhibit complex structural dynamics in solution, which is intricately modified by intermolecular recognition events. Although many lipocalins are monomeric, an increasing number of them have been proven to form oligomers. The coupling effects between self‐association and ligand binding in these proteins are largely unknown. To address this issue, we have calorimetrically characterized the recognition of dodecyl sulfate by bovine β‐lactoglobulin, which forms weak homodimers at neutral pH. A thermodynamic analysis based on coupled‐equilibria revealed that dimerization exerts disparate effects on the ligand‐binding capacity of β‐lactoglobulin. Protein dimerization decreases ligand affinity (or, reciprocally, ligand binding promotes dimer dissociation). The two subunits in the dimer exhibit a positive, entropically driven cooperativity. To investigate the structural determinants of the interaction, the crystal structure of β‐lactoglobulin bound to dodecyl sulfate was solved at 1.64 Å resolution. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

5.
Three mononuclear CuII complexes, [CuCl(naph‐pa)] ( 1 ), [Cu(bipy)(naph‐pa)]Cl ( 2 ), and [Cu(naph‐pa)(phen)]Cl ( 3 ) ((naph‐pa)=Schiff base derived from the condensation of 2‐hydroxynaphthalene‐1‐carbaldehyde and 2‐picolylamine (=2‐(aminomethyl)pyridine), bipy=2,2′‐bypiridine, and phen=1,10‐phenanthroline) were synthesized and characterized. Complex 1 exhibits square‐planar geometry, and 2 and 3 exhibit square pyramidal geometry, where Schiff base and bipy/phen act as NNO and as NN donor ligands, respectively. CT (Calf thymus)‐DNA‐binding studies revealed that the complexes bind through intercalative mode and show good binding propensity (intrinsic binding constant Kb: 0.98×105, 2.22×105, and 2.67×105 M ?1 for 1 – 3 , resp.). The oxidative and hydrolytic DNA‐cleavage activity of these complexes has been studied by gel electrophoresis: all the complexes displayed chemical nuclease activity in the presence and absence of H2O2. From the kinetic experiments, hydrolytic DNA cleavage rate constants were determined as 2.48, 3.32, and 4.10 h?1 for 1 – 3 , respectively. It amounts to (0.68–1.14)×108‐fold rate enhancement compared to non‐catalyzed DNA cleavage, which is impressive. The complexes display binding and cleavage propensity to DNA in the order of 3 > 2 > 1 .  相似文献   

6.
Nitric oxide reductase (NOR) catalyzes the generation of nitrous oxide (N2O) via the reductive coupling of two nitric oxide (NO) molecules at a heme/non‐heme Fe center. We report herein on the structures of the reduced and ligand‐bound forms of cytochrome c‐dependent NOR (cNOR) from Pseudomonas aeruginosa at a resolution of 2.3–2.7 Å, to elucidate structure‐function relationships in NOR, and compare them to those of cytochrome c oxidase (CCO) that is evolutionarily related to NOR. Comprehensive crystallographic refinement of the CO‐bound form of cNOR suggested that a total of four atoms can be accommodated at the binuclear center. Consistent with this, binding of bulky acetaldoxime (CH3‐CH=N‐OH) to the binuclear center of cNOR was confirmed by the structural analysis. Active site reduction and ligand binding in cNOR induced only ~0.5 Å increase in the heme/non‐heme Fe distance, but no significant structural change in the protein. The highly localized structural change is consistent with the lack of proton‐pumping activity in cNOR, because redox‐coupled conformational changes are thought to be crucial for proton pumping in CCO. It also permits the rapid decomposition of cytotoxic NO in denitrification. In addition, the shorter heme/non‐heme Fe distance even in the bulky ligand‐bound form of cNOR (~4.5 Å) than the heme/Cu distance in CCO (~5 Å) suggests the ability of NOR to maintain two NO molecules within a short distance in the confined space of the active site, thereby facilitating N‐N coupling to produce a hyponitrite intermediate for the generation of N2O. Proteins 2014; 82:1258–1271. © 2013 Wiley Periodicals, Inc.  相似文献   

7.
In our previous work, we proposed that desolvation and resolvation of the binding sites of proteins can serve as the slowest steps during ligand association and dissociation, respectively, and tested this hypothesis on two protein‐ligand systems with known binding kinetics behavior. In the present work, we test this hypothesis on another kinetically‐determined protein‐ligand system—that of p38α and eight Type II BIRB 796 inhibitor analogs. The kon values among the inhibitor analogs are narrowly distributed (104kon ≤ 105 M?1 s?1), suggesting a common rate‐determining step, whereas the koff values are widely distributed (10?1koff ≤ 10?6 s?1), suggesting a spectrum of rate‐determining steps. We calculated the solvation properties of the DFG‐out protein conformation using an explicit solvent molecular dynamics simulation and thermodynamic analysis method implemented in WaterMap to predict the enthalpic and entropic costs of water transfer to and from bulk solvent incurred upon association and dissociation of each inhibitor. The results suggest that the rate‐determining step for association consists of the transfer of a common set of enthalpically favorable solvating water molecules from the binding site to bulk solvent. The rate‐determining step for inhibitor dissociation consists of the transfer of water from bulk solvent to specific binding site positions that are unfavorably solvated in the apo protein, and evacuated during ligand association. Different sets of unfavorable solvation are evacuated by each ligand, and the observed dissociation barriers are qualitatively consistent with the calculated solvation free energies of those sets.  相似文献   

8.
Given the increase in the incidence of insulin resistance, obesity, and type 2 diabetes in children and adolescents, it would be of paramount importance to assess quantitative indices of insulin secretion and action during a physiological perturbation, such as a meal or an oral glucose‐tolerance test (OGTT). A minimal model method is proposed to measure quantitative indices of insulin secretion and action in adolescents from an oral test. A 7 h, 21‐sample OGTT was performed in 11 adolescents. The C‐peptide minimal model was identified on C‐peptide and glucose data to quantify indices of β‐cell function: static φs and dynamic φd responsivity to glucose from which total responsivity φ was also measured. The glucose minimal model was identified on glucose and insulin data to estimate insulin sensitivity, SI, which was compared to a reference measure, SIref, provided by a tracer method. Disposition indices, which adjust insulin secretion for insulin action, were then calculated. Indices of β‐cell function were φs = 51.35 ± 8.89 × 10?9min?1, φd = 1,392 ± 258 × 10?9, and φ = 82.09 ± 17.70 × 10?9min?1. Insulin sensitivity was SI = 14.19 ± 2.73 × 10?4, not significantly different from SIref = 14.96 ± 3.04 × 10?4 dl/kg·min per µU/ml, and well correlated: r = 0.98, P < 0.0001, thus indicating that SI can be accurately measured from an oral test. Disposition indices were DIs = 1,040 ± 201 × 10?14 dl/kg/min2 per pmol/l, DId = 33,178 ± 10,720 × 10?14 dl/kg/min per pmol/l, DI = 1,844 ± 522 × 10?14 dl/kg/min2 per pmol/l. Virtually the same minimal model assessment was obtained with a reduced 3 h, 9‐sample protocol. OGTT interpreted with C‐peptide and glucose minimal model has the potential to provide novel insight regarding the regulation of glucose metabolism in adolescents, and to evaluate the effect of obesity and interventions such as diet and exercise.  相似文献   

9.
Vascular endothelial growth factor receptor‐2 (VEGFR‐2) plays an important role in stimulating the proliferation of endothelial cells and improving the permeability of blood vessels, which is involved in tumor angiogenesis, a process that is essential for tumor growth and metastasis. In this study, we describe a method for high yield of recombinant extracellular domain 3 (KDR3) of human VEGFR‐2 in an Escherichia coli system with further purification by cation exchange chromatography and immobilized metal affinity chromatography (IMAC). The biological activity of recombinant KDR3 was performed by sequestering VEGF in HUVEC proliferation assay. The real‐time binding of human VEGF to immobilized KDR3 was monitored by a label‐free biosensor, Optical waveguide lightmode spectroscopy (OWLS). Under the given experimental conditions, the association rate constant ka was 4.2 × 103 M?1 s?1 and the dissociation rate kd was 5.1 × 10?3 s?1. The dissociation constant KD was then calculated to be 1.2 × 10?6 M. The obtained values will serve as baseline parameters for the design of improved versions of recombinant soluble VEGF receptors and the evaluation of developed anti‐KDR antibodies. In addition, such a scenario established by the use of OWLS will potentiate the kinetic study of ligand/receptor and antigen/antibody. The receptor discussed here, which block VEGF binding to cell membrane KDR, have potential clinical application in the treatment of cancer and other diseases where pathological angiogenesis is involved. © 2009 American Institute of Chemical Engineers Biotechnol. Prog., 2009  相似文献   

10.
Compstatin is a 13‐residue peptide that inhibits activation of the complement system by binding to the central component C3 and its fragments C3b and C3c. A combination of theoretical and experimental approaches has previously allowed us to develop analogs of the original compstatin peptide with up to 264‐fold higher activity; one of these analogs is now in clinical trials for the treatment of age‐related macular degeneration (AMD). Here we used functional assays, surface plasmon resonance (SPR), and isothermal titration calorimetry (ITC) to assess the effect of modifications at three key residues (Trp‐4, Asp‐6, Ala‐9) on the affinity and activity of compstatin and its analogs, and we correlated our findings to the recently reported co‐crystal structure of compstatin and C3c. The KD values for the panel of tested analogs ranged from 10?6 to 10?8 M. These differences in binding affinity could be attributed mainly to differences in dissociation rather than association rates, with a >4‐fold range in kon values (2–10 × 105 M?1 s?1) and a koff variation of >35‐fold (1–37 × 10?2 s?1) being observed. The stability of the C3b‐compstatin complex seemed to be highly dependent on hydrophobic effects at position 4, and even small changes at position 6 resulted in a loss of complex formation. Induction of a β‐turn shift by an A9P modification resulted in a more favorable entropy but a loss of binding specificity and stability. The results obtained by the three methods utilized here were highly correlated with regard to the activity/affinity of the analogs. Thus, our analyses have identified essential structural features of compstatin and provided important information to support the development of analogs with improved efficacy. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

11.
Capillary‐channeled polymer (C‐CP) fibers are used as a stationary phase for ion‐exchange chromatography of proteins. Collinear packing of the fibers permits operation at high linear velocities (Uo > 100 mm s?1) and low backpressure (<2,000 psi) on analytical‐scale columns. Rapid solvent transport is matched with very efficient solute mass transfer as fibers are virtually non‐porous with respect to the size of the target protein molecules. Lack of porosity of course limits the equilibrium binding capacity of stationary phases. Breakthrough curves and frontal analysis are used to better understand trade‐offs between the kinetic and thermodynamic properties as C‐CP fibers are applied in preparative situations. Fiber columns packed to different interstitial fraction values affect both the total fiber surface area (e.g., equilibrium binding capacity [EBC]) and the permittivity to flow and mass transport characteristics (e.g., dynamic binding capacity [DBC]). The EBC of the nylon 6 C‐CP fibers was found to be 1.30 mg g?1, with isotherms that were best matched by a Moreau model, showing linearity up to solute concentrations of ~0.4 mg mL?1. Isotherms generated under flow conditions were equally well approximated using Langmuir, Freundlich, and Moreau isotherm models. Fairly linear responses were seen up to the maximum load concentration of 1.2 mg mL?1. Counterintuitively, dynamic studies revealed that conditions of high column porosity yielded a DBC that is ~70% higher than the EBC. These findings point to potential advantages in terms downstream processing applications, where protein throughput and yield are critical metrics. © 2014 American Institute of Chemical Engineers Biotechnol. Prog., 31:97–109, 2015  相似文献   

12.
Studies of the kinetics of association and dissociation of the formycin nucleotides FTP and FDP with CF1 were carried out using the enhancement of formycin fluorescence. The protein used, derived from lettuce chloroplasts by chloroform induced release, contains only 4 types of subunit and has a molecular weight of 280 000.In the presence of 1.25 mM MgCl2, 1 mol of ATP or FTP is bound to the latent enzyme, with Kd = 10?7 or 2 · 10?7, respectively. The fluorescence emission (λmax 340 nm) of FTP is enhanced 3-fold upon binding, and polarization of fluorescence is markedly increased. The fluorescence changes have been used to follow FTP binding, which behaves as a bimolecular process with K1 = 2.4 · 104 M?1 · s?1. FTP is displaced by ATP in a process apparently involving unimolecular dissociation of FTP with k?1 = 3 · 10?3 s?1. The ratio of rates is comparable to the equilibrium constant and no additional steps have been observed.The protein has 3 sites for ADP binding. Rates of ADP binding are similar in magnitude to those for FTP. ADP and ATP sites are at least partly competitive with one another.The kinetics of nucleotide binding are strikingly altered upon activation of the protein as an ATPase. The rate of FTP binding increases to at least 106 M?1 · s?1. This suggests that activation involves lowering of the kinetic barriers to substrate and product binding-dissociation and has implications for the mechanism of energy transduction in photophosphorylation.  相似文献   

13.
Chemical shift perturbations (CSPs) in NMR spectra provide useful information about the interaction of a protein with its ligands. However, in a multiple‐ligand‐binding system, determining quantitative parameters such as a dissociation constant (Kd) is difficult. Here, we used a method we named CS‐PCA, a principal component analysis (PCA) of chemical shift (CS) data, to analyze the interaction between bovine β‐lactoglobulin (βLG) and 1‐anilinonaphthalene‐8‐sulfonate (ANS), which is a multiple‐ligand‐binding system. The CSP on the binding of ANS involved contributions from two distinct binding sites. PCA of the titration data successfully separated the CSP pattern into contributions from each site. Docking simulations based on the separated CSP patterns provided the structures of βLG–ANS complexes for each binding site. In addition, we determined the Kd values as 3.42 × 10−4M2 and 2.51 × 10−3M for Sites 1 and 2, respectively. In contrast, it was difficult to obtain reliable Kd values for respective sites from the isothermal titration calorimetry experiments. Two ANS molecules were found to bind at Site 1 simultaneously, suggesting that the binding occurs cooperatively with a partial unfolding of the βLG structure. On the other hand, the binding of ANS to Site 2 was a simple attachment without a significant conformational change. From the present results, CS‐PCA was confirmed to provide not only the positions and the Kd values of binding sites but also information about the binding mechanism. Thus, it is anticipated to be a general method to investigate protein–ligand interactions. Proteins 2013. © 2012 Wiley Periodicals, Inc.  相似文献   

14.
In slow mainstream flows (<4–6 cm · s?1), the transport of dissolved nutrients to seaweed blade surfaces is reduced due to the formation of thicker diffusion boundary layers (DBLs). The blade morphology of Macrocystis pyrifera (L.) C. Agardh varies with the hydrodynamic environment in which it grows; wave‐exposed blades are narrow and thick with small surface corrugations (1 mm tall), whereas wave‐sheltered blades are wider and thinner with large (2–5 cm) edge undulations. Within the surface corrugations of wave‐exposed blades, the DBL thickness, measured using an O2 micro‐optode, ranged from 0.67 to 0.80 mm and did not vary with mainstream velocities between 0.8 and 4.5 cm · s?1. At the corrugation apex, DBL thickness decreased with increasing seawater velocity, from 0.4 mm at 0.8 cm · s?1 to being undetectable at 4.5 cm · s?1. Results show how the wave‐exposed blades trap fluid within the corrugations at their surface. For wave‐sheltered blades at 0.8 cm · s?1, a DBL thickness of 0.73 ± 0.31 mm within the edge undulation was 10‐fold greater than at the undulation apex, while at 2.1 cm · s?1, DBL thicknesses were similar at <0.07 mm. Relative turbulence intensity was measured using an acoustic Doppler velocimeter (ADV), and overall, there was little evidence to support our hypothesis that the edge undulations of wave‐sheltered blades increased turbulence intensity compared to wave‐exposed blades. We discuss the positive and negative effects of thick DBLs at seaweed surfaces.  相似文献   

15.
To investigate the pharmacological mechanism of the traditional Chinese medicine, Pulsatilla decoction (PD), the levels of nitric oxide (NO), endothelin‐1 (ET‐1), tumor necrosis factor‐α (TNF‐α), and interleukin‐1α (IL‐1α) secreted by cultured rat intestinal microvascular endothelial cells (RIMECs) were determined after treatment with PD and its seven active ingredients, namely anemoside B4, anemonin, berberine, jatrorrhizine, palmatine, aesculin, and esculetin. RIMECs were challenged with lipopolysaccharide (LPS) at 1 µg ml?1 for 3 h and then treated with PD at 1, 5, and 10 mg ml?1 and its seven ingredients at 1, 5, and 10 µg ml?1 for 21 h, respectively. The results revealed that PD, anemonin, berberine, and esculetin inhibited the production of NO; PD, anemonin, and esculetin inhibited the secretion of ET‐1; PD, anemoside B4, berberine, jatrorrhizine, and aesculin downregulated TNF‐α expression; PD, anemoside B4, berberine, and palmatine decreased the content of IL‐1α. It showed that PD and its active ingredients could significantly inhibit the secretion of NO, ET‐1, TNF‐α, and IL‐1α in LPS‐induced RIMECs and suggested they would reduce inflammatory response via these cytokines. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

16.
Human serum heme–albumin (HSA–heme–Fe) displays reactivity and spectroscopic properties similar to those of heme proteins. Here, the nitrite reductase activity of ferrous HSA–heme–Fe [HSA–heme–Fe(II)] is reported. The value of the second-order rate constant for the reduction of $ {\text{NO}}_{2}^{ - } $ to NO and the concomitant formation of nitrosylated HSA–heme–Fe(II) (i.e., k on) is 1.3 M?1 s?1 at pH 7.4 and 20 °C. Values of k on increase by about one order of magnitude for each pH unit decrease between pH 6.5 to 8.2, indicating that the reaction requires one proton. Warfarin inhibits the HSA–heme–Fe(II) reductase activity, highlighting the allosteric linkage between the heme binding site [also named the fatty acid (FA) binding site 1; FA1] and the drug-binding cleft FA2. The dissociation equilibrium constant for warfarin binding to HSA–heme–Fe(II) is (3.1 ± 0.4) × 10?4 M at pH 7.4 and 20 °C. These results: (1) represent the first evidence for the $ {\text{NO}}_{2}^{ - } $ reductase activity of HSA–heme–Fe(II), (2) highlight the role of drugs (e.g., warfarin) in modulating HSA(–heme–Fe) functions, and (3) strongly support the view that HSA acts not only as a heme carrier but also displays transient heme-based reactivity.  相似文献   

17.
18.
The bioactive ingredients in Semen sinapis were rapidly screened by immobilized β2‐adrenoceptor (β2‐AR) and target‐directed molecular docking. The methods involved the attachment of β2‐AR using any amino group in the receptor, the simultaneous separation and identification of the retention compounds by high‐performance affinity chromatography; the binding mechanism of the interesting compound to the receptor was investigated by zonal elution and molecular docking. Sinapine in Semen sinapis was proved to be the bioactive compound that specifically binds to the immobilized receptor. The association constant of sinapine to β2‐AR was determined to be 1.36 × 105 M?1 with a value of 1.27 × 10?6 M for the number of binding sites. Ionic bond was believed to be the driving force during the interaction between sinapine and β2‐AR. It is possible to become a powerful alternative for rapid screening of bioactive compounds from a complex matrix such as traditional Chinese medicine and further investigation on the drug–receptor interaction. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

19.
The emergence of Gram‐negative “superbugs” exhibiting resistance to known antibacterials poses a major public health concern. Low molecular weight Gram‐negative antibacterials are believed to penetrate the outer bacterial membrane (OM) through porin channels. Therefore, intracellular exposure needed to drive antibacterial target occupancy should depend critically on the translocation rates through these proteins and avoidance of efflux pumps. We used electrophysiology to study the structure‐translocation kinetics relationships of a set of carbapenem antibacterials through purified porin OmpC reconstituted in phospholipid bilayers. We also studied the relative susceptibility of OmpC+ and OmpC‐ E. coli to these compounds as an orthogonal test of translocation. Carbapenems exhibit good efficacy in OmpC‐expressing E. coli cells compared with other known antibacterials. Ertapenem, which contains an additional acidic group compared to other analogs, exhibits the fastest entry into OmpC (kon ≈ 2 × 104 M?1 s?1). Zwitterionic compounds with highly polar groups attached to the penem‐2 ring, including panipenem, imipenem and doripenem exhibit faster kon (>104 M?1 s?1), while meropenem and biapenem with fewer exposed polar groups exhibit slower kon (~5 × 103 M?1 s?1). Tebipenem pivoxil and razupenem exhibit ~13‐fold slower kon (~1.5 × 103 M?1 s?1) than ertapenem. Overall, our results suggest that (a) OmpC serves as an important route of entry of these antibacterials into E. coli cells; and (b) that the structure‐kinetic relationships of carbapenem translocation are governed by H‐bond acceptor/donor composition (in accordance with our previous findings that the enthalpic cost of transferring water from the constriction zone to bulk solvent increases in the presence of exposed nonpolar groups). Proteins 2014; 82:2998–3012. © 2014 Wiley Periodicals, Inc.  相似文献   

20.
Although sea‐ice represents a harsh physicochemical environment with steep gradients in temperature, light, and salinity, diverse microbial communities are present within the ice matrix. We describe here the photosynthetic responses of sea‐ice microalgae to varying irradiances. Rapid light curves (RLCs) were generated using pulse amplitude fluorometry and used to derive photosynthetic yield (ΦPSII), photosynthetic efficiency (α), and the irradiance (Ek) at which relative electron transport rate (rETR) saturates. Surface brine algae from near the surface and bottom‐ice algae were exposed to a range of irradiances from 7 to 262 μmol photons · m?2 · s?1. In surface brine algae, ΦPSII and α remained constant at all irradiances, and rETRmax peaked at 151 μmol photons · m?2 · s?1, indicating these algae are well acclimated to the irradiances to which they are normally exposed. In contrast, ΦPSII, α, and rETRmax in bottom‐ice algae reduced when exposed to irradiances >26 μmol photons · m?2 · s?1, indicating a high degree of shade acclimation. In addition, the previous light history had no significant effect on the photosynthetic capacity of bottom‐ice algae whether cells were gradually exposed to target irradiances over a 12 h period or were exposed immediately (light shocked). These findings indicate that bottom‐ice algae are photoinhibited in a dose‐dependent manner, while surface brine algae tolerate higher irradiances. Our study shows that sea‐ice algae are able to adjust to changes in irradiance rapidly, and this ability to acclimate may facilitate survival and subsequent long‐term acclimation to the postmelt light regime of the Southern Ocean.  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