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 46 毫秒
1.
A novel sulfur‐utilizing perchlorate reducing bacterial consortium successfully treated perchlorate (ClO) in prior batch and bench‐scale packed bed reactor (PBR) studies. This study examined the scale up of this process for treatment of water from a ClO and RDX contaminated aquifer in Cape Cod Massachusetts. A pilot‐scale upflow PBR (~250‐L) was constructed with elemental sulfur and crushed oyster shell packing media. The reactor was inoculated with sulfur oxidizing ClO reducing cultures enriched from a wastewater seed. Sodium sulfite provided a good method of dissolved oxygen removal in batch cultures, but was found to promote the growth of bacteria that carry out sulfur disproportionation and sulfate reduction, which inhibited ClO reduction in the pilot system. After terminating sulfite addition, the PBR successfully removed 96% of the influent ClO in the groundwater at an empty bed contact time (EBCT) of 12 h (effluent ClO of 4.2 µg L?1). Simultaneous ClO and NO reduction was observed in the lower half of the reactor before reactions shifted to sulfur disproportionation and sulfate reduction. Analyses of water quality profiles were supported by molecular analysis, which showed distinct groupings of ClO and NO degrading organisms at the inlet of the PBR, while sulfur disproportionation was the primary biological process occurring in the top potion of the reactor. Biotechnol. Bioeng. 2012; 109:637–646. © 2011 Wiley Periodicals, Inc.  相似文献   

2.
We describe conditions which lead to complete helix formation of poly(I) in the presence of NH. Binding of NH is shown to be specific in the presence of Li+, which does not by itself support helix formation under these conditions. The NH–poly(I) complex is characterized by uv, CD, and ir spectroscopy. The CD spectrum is strikingly different from those of the Na+ or K+ complexes, the first extremum being changed from negative for the metal ions to positive for NH. A stereospecific model is proposed for the NH–poly(I) helix in which the N of NH is located on the axis of the four-stranded helix, midway between planar tetramers formed by the bases. The model is consistent with the tetrahedral symmetry of NH, the requirement for four acceptable hydrogen bonds, the observed stability of the helix, and the accepted geometry of the backbone.  相似文献   

3.
The relationship between cytosolic concentrations of Ca2+ (Ca) and Na+ (Na) were studied in preparations of rat submandibular and pancreatic acini loaded with the Ca2+-sensitive dye Fura-2 or the Na+-sensitive dye SBFI. Pancreatic acini showed no changes in Na during either transient or persistent changes in Ca. Increases in Ca produced by exposure of submandibular gland acini to carbachol, a muscarinic cholinergic agonist, were followed by an increase in Na after a delay of 5–10 s. When Ca2+ stores were mobilized without Ca2+ influx Na also increased, but in acini loaded with BAPTA, a nonfluorescent Ca2+ chelator, the transient increase in Ca2+ caused by mobilization of stored Ca2+ was virtually abolished, as was the increase in Na. In the presence of ionomycin, increases in Ca were followed by increases in Na. Ca2+-dependent increases in Na were abolished in Na+-free buffer and by the presence of furosemide, a blocker of Na+-K+-2Cl cotransport. In other studies, extracellular ATP (ATPo) produced an increase in Ca and Na. The steady-state increase in Ca was reduced by increasing extracellular Na+ concentrations (Na) in dose-dependent fashion (IC50 = 16.4 ± 4.7 mM Na+). Likewise, increasing Na reduced ATPo-stimulated 45Ca2+ uptake at steady state (IC50 = 15.8 ± 9.2 mM Na+). Changing Na had no effect on carbachol-stimulated increases in Ca. We conclude that, in rat submandibular gland acini, ATPo promotes an increase in Ca and Na via a common influx pathway and that, under physiologic conditions, Na+ significantly limits the ATPo-stimulated increase in Ca. In the presence of carbachol, however, Na rises in Ca-dependent fashion in submandibular gland acini via stimulation of Na+-K+-2Cl cotransport. © 1996 Wiley-Liss, Inc.  相似文献   

4.
Purple non‐sulfur (PNS) bacteria produce hydrogen by photofermentation of organic acids in wastewater. However, NH in wastewater may inhibit hydrogen synthesis by repressing the expression and activity of nitrogenase, the enzyme catalyzing hydrogen production in PNS bacteria. In this study, the Rhodobacter sphaeroides 6016 glnA gene encoding glutamine synthetase (GS) was knocked out by homologous recombination, and the effects on hydrogen production and nitrogenase activity were examined. Using 3 mM glutamine as the nitrogen source, hydrogen production (1,245–1,588 mL hydrogen/L culture) and nitrogenase activity were detected in the mutant in the presence of relatively high NH concentrations (15–40 mM), whereas neither was detected in the wild‐type strain under the same conditions. Further analysis indicated that high NH concentrations greatly inhibited the expression of nifA and nitrogenase gene in the wild‐type strain but not in the glnA1? mutant. These observations suggest that GS is essential to NH repression of nitrogenase and that deletion of glnA1 results in the complete derepression of nitrogenase by preventing NH assimilation in vivo, thus relieving the inhibition of nifA and nitrogenase gene expression. Knocking out glnA1 therefore provides an efficient approach to removing the inhibitory effects of ammonium ions in R. sphaeroides and possibly in other hydrogen‐producing PNS bacteria. Biotechnol. Bioeng. 2010;106: 564–572. © 2010 Wiley Periodicals, Inc.  相似文献   

5.
The present paper deals with the analysis of the amount of oxygen utilized for oxidation of a small dose of carbon substrate in carbon limited Brevibacterium flavum culture. The ratio of the measured oxygen consumption (mo2) to the amount of added carbon substrate (ms) gives a stoichiometric coefficient of the biological oxidation equation. A linear relationship between mo2 and ms was observed. To compare the efficiency of different carbon substrate utilization there has been introduced a normalized value β = m/m. There exists a simple relationship between β and the thermodynamical growth efficiency η The theoretical considerations are proved by experimental results with β, η and Yx/s in a chemostat culture at various medium flow rates.  相似文献   

6.
A Cabani  A Paci  V Rizzo 《Biopolymers》1976,15(1):113-129
Using the formalism of nearest-neighbor Ising model and assuming that the allowed states for a monomeric unity of a polypeptide chain in solutions containing strong acids are E (helix), C (coil), and CS (solvent-bonded coil), the partition function of the system was deduced analytically. Equations were obtained which permitted the prediction of the characteristic thermodynamic behavior of the helix–coil transition under these conditions. These equations were used to examine critically the possible correlations between experimental data obtained using different techniques. Particular attention was devoted to quantities called “transition enthalpies,” obtained from the slope of the transition curves at the point where the helix fraction is one-half (ΔH), or for measurements of the heat of solution of the polymer over the total range of solvent composition (ΔH), or from heat capacity measurements taken at various temperatures (ΔH). Literature data of ΔH(j = opt, sol, cal) for the system poly-γ-benzyl-L -glutamate in mixtures of dichloroacetic acid and 1,2-dichloroethane were carefully analyzed.  相似文献   

7.
D E Goldsack 《Biopolymers》1969,7(3):299-313
A statistical analysis of the relation between the amino acid composition of proteins and the amount of helical secondary structure as characterized by the Moffitt b0 parameter has shown a high degree of correlation of the b0 parameter with those amino acids whose homopolymers can form helical structures. Using the data for 107 proteins, a linear relation was found between b0 and the sum of the residue percentages of alanine, arginine, aspartic acid, cysteine, glutamic acid, leucine, andlysine. A statistical analysis of the Moffitt a0 parameter, on the other hand, showed no statistically significant grouping of amino acids in relation to the amount of secondary β structure in a protein. A plot of b0 versus a0–a, where a represents the a0 parameter for a fully denatured protein, for 55 proteins showed distinct nonlinearity. This nonlinearity was postulated to be due to presence of β structure, and a nomagram was constructed which allowed a semiquantitative estimate of the amount of helical and β-type secondary structures from the b0 versus a0–a plot.  相似文献   

8.
The conformational transition of poly(L -agrignine) by binding with various mono-, di-, and polyvalent anions, especially with SO, was studied by CD measurements. The intramolecular random coil-to-α-helix conformational transition and the subsequent transition to the β-turn-like structure was caused by binding with SO. The binding data obtained from equilibrium dialysis experiments showed that the α-helical conformation of poly(L -arginine) is stabilized at a 1:3 stoichiometric ratio of bound SO to arginine residue; at higher free SO concentrations, the α-helix converts to the β-turn-like structure accompanied by a decrease in amount of bound SO. The same conformaitonal transition of poly(L -arginine) also occurred in the solutions of other divalent anions (SO, CO, and HPO) and polyvalent anions (P2O, P3O). Among the monovalent anions examined, CIO and dodecyl sulfate were effective in including α-helical conformation, while the other monovalent anions (OH?, Cl?, F?, H2PO, HCO and CIO) failed to induce poly(L -arginine) to assume the α-helical conformation. Thus, we noticed that, except for dodecyl sufate, the terahedral structure is common to the α-helix-forming anions. A well-defined model to the α-helical poly(L -arginine)/anion complex was proposed, in which both the binding stoichiometry of anions to the arginine residue and the tetrahedral structure of anions were taken into consideration. Based on these results, it was concluded that the tetrahedral-type anions stabilize the α-helical conformation of poly(L -arginine) by crosslinking between two guanidinium groups of nearby side chains on the same α-helix through the ringed structures stabilized by hydrogen bonds as well as by electrostatic interaction. Throughout the study it was noticed that the structural behavior of poly(L -arginine) toward anions is distinct from that of poly(L -lysine).  相似文献   

9.
10.
G Spach  Y Trudelle  F Heitz 《Biopolymers》1983,22(1):403-407
The β- and β-helical structures do not appear consistent with the structural data and ion-transport properties of (Ala-Ala-Gly) or (Leu-Ser-Leu-Gly) oligomers. Oligoalanine derivatives also give rise to current fluctuations in bilayer lipid membranes. Bundles of molecules may explain the behavior of these various peptides in membranes.  相似文献   

11.
The hot water of Rajgir springs is used for drinking and bathing purposes by tourists. Certain physico-chemical characteristics (temperature, pH, NO, PO, etc.) of the water along with phycological parameters viz. community composition, species diversity, standing crop etc. were measured from June 1986 to April 1987. The water was deficient in Na, NO and PO ions. The hot springs were mainly dominated by algae belonging to Cyanophyceae and Bacillariophyceae. The algal community comprised 18 species, with dominance of Cyanophyceae over Bacillariophyceae. While Mastigocladus laminosus and species of Phormidium were dominant in Suraj Kund, species of Oscillaroria and Synechococcus dominated in Chandrama Kund. Diatoms comprised about 10 % of the algal community. Though there was a considerable seasonal change in species diversity of the algal community the total biomass (chlorophyll a extracted per unit area from the algal mat) remained constant.  相似文献   

12.
Biofilms are known to be robust biocatalysts. Conventionally, they have been mainly applied for wastewater treatment, however recent reports about their employment for chemical synthesis are increasingly attracting attention. Engineered Pseudomonas sp. strain VLB120ΔC biofilm growing in a tubular membrane reactor was utilized for the continuous production of (S)‐styrene oxide. A biofilm specific morphotype appeared in the effluent during cultivation, accounting for 60–80% of the total biofilm irrespective of inoculation conditions but with similar specific activities as the original morphotype. Mass transfer of the substrate styrene and the product styrene oxide was found to be dependent on the flow rate but was not limiting the epoxidation rate. Oxygen was identified as one of the main parameters influencing the biotransformation rate. Productivity was linearly dependent on the specific membrane area and on the tube wall thickness. On average volumetric productivities of 24 g L day?1 with a maximum of 70 g L day?1 and biomass concentrations of 45 gBDW L have been achieved over long continuous process periods (≥50 days) without reactor downtimes. Biotechnol. Bioeng. 2010. 105: 705–717. © 2009 Wiley Periodicals, Inc.  相似文献   

13.
The methods of optimal designing of experiments proposed by WALD (1943) are used for determination of an Aσ2-optimal concrete design for estimation of σ2′ = (σ, σ) in case of one-way analysis of variance. Starting point of definition of the optimality criterion is a quadratic loss matrix.  相似文献   

14.
A comparison of the condensation of T4 phage DNA by spermidine and Co(NH3) at pH values between 5.1 and 10.2 has been made using quasielastic light scattering to determine translational diffusion coefficients and Stokes radii. Co(NH3) is more effective than spermidine in causing condensation at all pH, indicating that the differences observed in previous work were not due to pH effects, as might have been inferred from recent theories of intermolecular forces. The DNA particles collapsed with Co(NH3) are smaller than those obtained with spermidine. The hydrodynamic radius of spermide-collapsed structures decreases slightly with increasing pH, while the size of the Co(NH3)collapsed structures is almost independent of pH. These results confirm that there are specific ion effects in DNA condensation by oligocations, in addition to the dominant general polyelectrolyte effects.  相似文献   

15.
We studied the interactions of the substitution-inert inversion-labile complexes Fe(bipy) and Fe(phen) [and the inversion-stable complex Ru(bipy)] with DNA. The association of these complexes to DNA is mainly electrostatic, and Fe(phen) shows a more effective binding to DNA than the two bipyridyl complexes, possibly owing to a different binding mode. The interactions are enantioselective, leading to a Pfeiffer shift in the diastereomeric inversion equilibria and an excess of the Δ-enantiomer of Fe(phen) and Fe(bipy), which is directly monitorable through CD. The partition constants for the inversion equilibrium range from 1.3 to 2.0 for Fe(bipy) and Fe(phen), depending on ionic conditions. From flow LD information about the orientation of the complexes on DNA was obtained: it is consistent with a fit of the Δ-enantiomer in the major groove of the right-handed DNA helix. The mechanisms of interaction are discussed against equilibrium, spectroscopic, and kinetic data.  相似文献   

16.
Four fundamental Raman lines were observed at 159, 111, 55 and 27 cm-1 corresponding to the I bound (I) in amyloses with DP from 20 to 100, regardless of the degree of polymerization of I and the excitation wavelength. The spectral resolution was based on the molar extinction coefficient and molar ellipticity spectra of I. Eight bands, named, S1, S2, ?, S8 from long to short wavelength, were isolated. These were found regardless of the DP. By a resonance excitation Raman study, the characteristics of S3 and S4, comprising the shoulder around 480 nm, were found to be different from those of S1 and S2, comprising the blue band. The assignment of the spectra was based on the electronic states of the monomeric I in the exciton-coupled dimeric unit. It was concluded that the blue band (S1,S2) belonged to the long-axis transitions and the shoulder band (S3,S4) to the short-axis ones on the monmeric coordinate system.  相似文献   

17.
The high-resolution nmr spectrum of baker's yeast tRNA, a recently sequenced non-denaturable tRNA, has been compared with the spectra of the native and denatured conformers of the closely related species tRNA. Because of the presence of many common base pairs in the different tRNA's, it is possible to assign most of the low-field resonances to specific secondary-structure base pairs. A comparison of the observed positions of the various resonances with those predicted by a semiempirical ring-current shift theory shows a root-mean-square deviation of 0.14, 0.11, and 0.12 ppm for tRNA (native), and tRNA (denatured), respectively. These results support the ring-current shift theory currently used to interpret the low-field nmr spectra of the tRNA molecules. Differences between the predicted and observed positions of some resonances provide new evidence for higher order effects such as shifts from second nearest neighbors, anomalous shifts exerted by G·U base pairs, and tertiary-structure effects. A model that was previously proposed for the denatured conformer of tRNA is also supported.  相似文献   

18.
For a balanced one-way classification, where the normally distributed observations obey a random model yij=μ+bi+cij with two variance components var (bi) = δ and var (cij) = δ, the probability is given that the analysis of variance estimate of δ will be negative. This probability depends on δ/δ and the degrees of freedom in the ANOVA table. Tables for this probability are given. If the normally distributed observations obey an intra-class correlation model, the probability that the Mean Square between groups is smaller than the Mean Square within groups can also be evaluated from the given tables.  相似文献   

19.
In this note it is shown that the block design with incidence matrix Ñ = [NNN], where N = c1hNh + coh (11′–Nh). coh and c1h are any non-negative integers and Nh,h = 1, 2,…,p, are incidence matrices of balanced incomplete block designs with the same number of treatments t, is a balanced block design with the block sizes exceeding the number of treatments. In derivation the matrix M0, introduced by CALIński (1971) is utilized.  相似文献   

20.
The kinetics of ethidium's intercalative binding to DNA packaged in bacteriophage T7 and two T7 deletion mutants have been determined, using enhancement of fluorescence to quantitate binding. At a constant ethidium concentration, the results can be described as first-order binding with two different rate constants, k (= k1 + k?1) and k (= k2 + k?2). The larger rate constant (k) was at least four orders of magnitude smaller than the comparable first-order forward rate constant for binding to DNA released from its capsid. At 25°C values of k decreased as the amount of DNA packaged per internal volume increased. This latter observation indicates that the rate of ethidium's binding to packaged T7 DNA is limited by an event that occurs inside of the DNA-containing region of T7, not by the crossing of T7 capsid's outer shell. Arrhenius plots of kM are biphasic, indicating a transition for packaged DNA at a temperature of 20°C. The data indicate that k s are limited by either sieving of ethidium during its passage through the packaged DNA or subsequent hindered intercalation.  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