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.
To compare the OH-initiated reaction mechanisms of five chlorophenoxy herbicides, density functional theory (DFT) calculations of reactions in which ·OH attacks one of three active positions on each herbicide were carried out at the MPWB1K/6-311 + G(3df,2p)//MPWB1K/6-31 + G(d,p) level. For each herbicide, the calculation results show that ·OH addition to the C1 atom, which is the nexus between the benzene ring and the side group, possesses the lowest energy barrier among the three kinds of reactions, indicating that ·OH addition–substitution of the side chain is the most energetically and kinetically favorable reaction mechanism. Comparisons among the herbicides show that the mechanisms are affected by the steric hindrance and the electronegativities of the –CH3 and –Cl groups. When comparing the addition of ·OH to the C1 site among the five herbicides, the activation energy for the reaction of ·OH with DCPP reaction is the lowest (3.61 kcal mol?1), while that for the ·OH and 4-CPA reaction was the highest (5.91 kcal mol?1). ·OH addition to the C4 site presents the highest energy barriers among the three kinds of reactions, indicating that the para Cl is difficult to break down. When comparing the H-atom abstraction reactions of the five herbicides, the H atoms in the –CH2– group of 2,4-D are the easiest for ·OH to abstract, whereas those of DCPP and MCPP are more difficult to abstract, due to the steric hindrance of the –CH3 group. Additionally, the results obtained from the PCM calculations reveal that most of the reactions occur more easily in water than in gas, though the mechanisms involved are the same as those discussed above.  相似文献   

2.
In ruminants, gastrointestinal recycling of urea is acutely enhanced by fibre-rich diets that lead to high ruminal concentration of short chain fatty acids (SCFA), while high ammonia has inhibitory effects. This study attempted to clarify if urea flux to the porcine cecum is similarly regulated. Thirty-two weaned piglets were fed diets containing protein (P) of poor prececal digestibility and fibre (F) at high (H) or low levels (L) in a 2 × 2 factorial design. After slaughter, cecal content was analyzed and the cecal mucosa incubated in Ussing chambers to measure the effect of pH, SCFA and NH4 + on the flux rates of urea, short-circuit current (I sc) and tissue conductance (G t). NH4 + significantly enhanced I sc (from 0.5 ± 0.2 to 1.2 ± 0.1 μEq cm?2 h?1). No acute effects of SCFA or ammonia on urea flux were observed. Tissue conductance was significantly lower in the high dietary fibre groups irrespective of the protein content. Only the HP-LF group emerged as different from all others in terms of urea flux (74 ± 6 versus 53 ± 3 nmol cm?2 h?1), associated with higher cecal ammonia concentration and reduced fecal consistency. The data suggest that as in the rumen, uptake of ammonia by the cecum may involve electrogenic transport of the ionic form (NH4 +). In contrast to findings in the rumen, neither a high fibre diet nor acute addition of SCFA enhanced urea transport across the pig cecum. Instead, a HP-LF diet had stimulatory effects. A potential role for urea recycling in stabilizing luminal pH is discussed.  相似文献   

3.
A density functional theory (DFT) study was performed to explore the mechanisms of the acid-catalyzed decarboxylation reaction of salicylic acids using the B3LYP method with 6-31++G(d,p) basis set in both gas phase and aqueous environment. The α-protonated cation of carboxylate acid was formed during the decarboxylation process in acidic conditions, and the presence of hydrogen ions promotes decarboxylation greatly by significantly decreasing the overall reaction energy barriers to 20.98 kcal mol?1 in gas phase and 20.93 kcal mol?1 in water, respectively. The hydrogen in the α-carbon came directly from the acid rather than from the carboxyl group in neutral state. Compared with the reaction in gas phase, water in aqueous state causes the reaction to occur more easily. Substituents of methyl group, chlorine and fluorine at the ortho-position to the carboxyl of salicylic acid could further lower the decarboxylation energy barriers and facilitate the reaction.  相似文献   

4.
We have analyzed the variation of the reaction force F(ξ) and the reaction force constant κ(ξ) along the intrinsic reaction coordinates ξ of the water-assisted proton transfer reactions of HX-N = Y (X,Y = O,S). The profile of the force constant of the vibration associated with the reactive mode, k ξ (ξ), was also determined. We compare our results to the corresponding intramolecular proton transfers in the absence of a water molecule. The presence of water promotes the proton transfers, decreasing the energy barriers by about 12 – 15 kcal mol-1. This is due in part to much smaller bond angle changes being needed than when water is absent. The κ(ξ) profiles along the intrinsic reaction coordinates for the water-assisted processes show striking and intriguing differences in the transition regions. For the HS-N = S and HO-N = S systems, two κ(ξ) minima are obtained, whereas for HO-N = O only one minimum is found. The k ξ (ξ) show similar behavior in the transition regions. We propose that this fine structure reflects the degree of synchronicity of the two proton migrations in each case.
Figure
Fine Structure in the Transition Region  相似文献   

5.
To provide an in-depth insight into the molecular basis of spontaneous tautomerism in DNA and RNA base pairs, a hybrid Monte Carlo (MC)–quantum chemical (QC) methodology is implemented to map two-dimensional potential energy surfaces along the reaction coordinates of solvent-assisted proton transfer processes in guanosine and its analog acyclovir in aqueous solution. The solvent effects were simulated by explicit inclusion of water molecules that model the relevant part of the first hydration shell around the solute. The position of these water molecules was estimated by carrying out a classical Metropolis Monte Carlo simulation of dilute water solutions of the guanosine (Gs) and acyclovir (ACV) and subsequently analyzing solute–solvent intermolecular interactions in the statistically-independent MC-generated configurations. The solvent-assisted proton transfer processes were further investigated using two different ab initio MP2 quantum chemical approaches. In the first one, potential energy surfaces of the ‘bare’ finite solute–solvent clusters containing Gs/ACV and four water molecules (MP2/6-31+G(d,p) level) were explored, while within the second approach, these clusters were embedded in ‘bulk’ solvent treated as polarizable continuum (C-PCM/MP2/6-31+G(d,p) level of theory). It was found that in the gas phase and in water solution, the most stable tautomer for guanosine and acyclovir is the 1H-2-amino-6-oxo form followed by the 2-amino-6-(sZ)-hydroxy form. The energy barriers of the water-assisted proton transfer reaction in guanosine and in acyclovir are found to be very similar – 11.74 kcal mol?1 for guanosine and 11.16 kcal mol?1 for acyclovir, and the respective rate constants (k = 1.5?×?101 s?1, guanosine and k = 4.09?×?101 s?1, acyclovir), are sufficiently large to generate the 2-amino-6-(sZ)-hydroxy tautomer. The analysis of the reaction profiles in both compounds shows that the proton transfer processes occur through the asynchronous concerted mechanism.  相似文献   

6.
The conformation and the interaction of CHF2OCF2CHF2 (desflurane II) with one water molecule is investigated theoretically using the ab initio MP2/aug-cc-pvdz and DFT-based M062X/6-311++G(d,p) methods. The calculations include the optimized geometries, the harmonic frequencies of relevant vibrational modes along with a natural bond orbital (NBO) analysis including the NBO charges, the hybridization of the C atom and the intra- and intermolecular hyperconjugation energies. In the two most stable conformers, the CH bond of the F2HCO- group occupies the gauche position. The hyperconjugation energies are about the same for both conformers and the conformational preference depends on the interaction between the non-bonded F and H atoms. The deprotonation enthalpies of the CH bonds are about the same for both conformers, the proton affinity of the less stable conformer being 3 kcal mol?1 higher. Both conformers of desflurane II interact with water forming cyclic complexes characterized by CH…O and OH…F hydrogen bonds. The binding energies are moderate, ranging from ?2.4 to ?3.2 kcal mol?1 at the MP2 level. The origin of the blue shifts of the ν(CH) vibrations is analyzed. In three of the complexes, the water molecule acts as an electron donor. Interestingly, in these cases a charge transfer is also directed to the non bonded OH group of the water molecule. This effect seems to be a property of polyfluorinated ethers.  相似文献   

7.
In order to find the susceptibility of the amino-Claisen rearrangement and the next proton shift reaction of N-allyl-N-arylamine to the substituent effects in the para position, the kinetic and thermodynamic parameters were calculated at the B3LYP level using the 6-31G** basis set. The calculated activation energies for the rearrangements and proton shift reactions are close to 44.4 and 49.5 kcal mol? 1, respectively. The transition states of the rearrangement with electron-donor substituents are more stable than those with electron-withdrawing substituent groups, but for the proton shift reaction, this situation is reversed (with the exception of fluorine atom for the rearrangement and fluorine and chlorine atoms for the proton shift reaction). Negative values for the activation entropy confirm the concerted mechanism for the amino-Claisen rearrangement and proton shift reaction. The Hammett ρ values of ? 2.4172 and ? 1.7791 are obtained for σp and σ (enhanced sigma) in the amino-Claisen rearrangement, respectively. The correlation between log(k X/k H) and σp is weaker than that with σ (enhanced sigma). A negative Hammett ρ value indicates that the electron-donating groups slightly increase the rate of amino-Claisen rearrangement. A positive Hammett ρ value (2.4921) for the proton shift reaction indicates that electron-withdrawing groups increase the rate of reaction.  相似文献   

8.
A series of end-labelled isomaltose oligosaccharides was prepared by the reaction of dextran-sucrase with sucrose-14C in the presence of excess of unlabelled isomaltose saccharides as alternative acceptor. The main product of each reaction contained one more D-glucose residue than the acceptor substrate, and the label was located at the non-reducing end. The end-labelled saccharides were used to determine the specificity of a bacterial dextranase that required five or more consecutive α-(1→6)-D-glucosidic linkages in the substrate. The third linkage from the reducing end of isomaltohexaose (IM6) and of other substrates with longer chains (IM7 and IM8) was the most susceptible to attack, and the products from higher oligosaccharides were IM3, IM4, and IM5. Isomaltopentaose (IM5) was further hydrolysed to IM3 and IM2 when a 35-fold excess of enzyme was added, but there was no action on IM4, IM3, or IM2 under these conditions. It was concluded that the dextranase hydrolysed linkages penultimate to either end of the chain only with difficulty, and that end linkages were completely resistant to attack.  相似文献   

9.
S. Izawa  Donald R. Ort 《BBA》1974,357(1):127-143
NH2OH-treated, non-water oxidizing chloroplasts are shown to be capable of oxidizing ferrocyanide and I? via Photosystem II at appreciable rates (? 200 μequiv/h per mg chlorophyll). Using methylviologen as electron acceptor, ferrocyanide oxidation can be measured as O2 uptake, as ferricyanide formation, or as H+ consumption (2 Fe2+ + 2H+ + O2 → 2 Fe3+ + H2O2). I? oxidation can be measured as methylviologen-mediated O2 uptake, or spectrophotometrically, using ferricyanide as electron acceptor. The oxidation product I2 is re-reduced, as it is formed, by unknown reducing substances in the reaction system.The rate-saturating concentrations of these donors are very high: 30 mM with ferricyanide and 15 mM with I?. Relatively lipophilic Photosystem II donors such as catechol, benzidine and p-aminophenol saturate the photooxidation rate at much lower concentrations (< 0.5 mM). It thus seems that the oxidation of hydrophilic reductants such as ferricyanide and I? is limited by permeability barriers. Very likely the site of Photosystem II oxidation is embedded in the thylakoid membrane or is situated on the inner surface of the membrane.The efficiency of phosphorylation (P/e2) is 0.5 to 0.6 with ferrocyanide and about 0.5 with I?. In contrast the P/e2 ratio is 1.0 to 1.2 when water, catechol, p-aminophenol or benzidine serves as electron donor. These differences imply that only one of two phosphorylation sites operate when ferrocyanide and I? are oxidized. Ferrocyanide and I? are also chemically distinct from other Photosystem II donors in that their oxidation does not involve proton release. It is suggested that the mechanism of energy conservation associated with Photosystem II may be only operative when the removal of electrons from the donor results in release of protons (i.e. with water, hydroquinones, phenylamines, etc.).  相似文献   

10.
To reduce CO2 emissions from alcoholic fermentation, Arthrospira platensis was cultivated in tubular photobioreactor using either urea or nitrate as nitrogen sources at different light intensities (60 μmol m?2 s?1?≤?I?≤?240 μmol m?2 s?1). The type of carbon source (pure CO2 or CO2 from fermentation) did not show any appreciable influence on the main cultivation parameters, whereas substitution of nitrate for urea increased the nitrogen-to-cell conversion factor (Y X/N ), and the maximum cell concentration (X m ) and productivity (P X ) increased with I. As a result, the best performance using gaseous emissions from alcoholic fermentation (X m ?=?2,960?±?35 g m?3, P X ?=?425?±?5.9 g m?3 day?1 and Y X/N ?=?15?±?0.2 g g?1) was obtained at I?=?120 μmol m?2 s?1 using urea as nitrogen source. The results obtained in this work demonstrate that the combined use of effluents rich in urea and carbon dioxide could be exploited in large-scale cyanobacteria cultivations to reduce not only the production costs of these photosynthetic microorganisms but also the environmental impact associated to the release of greenhouse emissions.  相似文献   

11.
Holger Dau  Michael Haumann 《BBA》2007,1767(6):472-483
In oxygenic photosynthesis, water is split at a Mn4Ca complex bound to the proteins of photosystem II (PSII). Powered by four quanta of visible light, four electrons and four protons are removed from two water molecules before dioxygen is released. By this process, water becomes an inexhaustible source of the protons and electrons needed for primary biomass formation. On the basis of structural and spectroscopic data, we recently have introduced a basic reaction cycle of water oxidation which extends the classical S-state cycle [B. Kok, B. Forbush, M. McGloin, Cooperation of charges in photosynthetic O2 evolution- I. A linear four-step mechanism, Photochem. Photobiol. 11 (1970) 457-475] by taking into account also the role and sequence of deprotonation events [H. Dau, M. Haumann, Reaction cycle of photosynthetic water oxidation in plants and cyanobacteria, Science 312 (2006) 1471-1472]. We propose that the outwardly convoluted and irregular events of the classical S-state cycle are governed by a simple underlying principle: protons and electrons are removed strictly alternately from the Mn complex. Starting in I0, eight successive steps of alternate proton and electron removal lead to I8 and only then the O-O bond is formed. Thus not only four oxidizing equivalents, but also four bases are accumulated prior to the onset of dioxygen formation. After reviewing the kinetic properties of the individual S-state transition, we show that the proposed basic model explains a large body of experimental results straightforwardly. Furthermore we discuss how the I-cycle model addresses the redox-potential problem of PSII water oxidation and we propose that the accumulated bases facilitate dioxygen formation by acting as proton acceptors.  相似文献   

12.
The thermally allowed electrocyclic reaction syn-cyclophanediene (CPD) to dihydropyrene (DHP) was compared with the disallowed thermal electrocyclic reaction in anti CPD through density functional theory (DFT) calculations at the B3LYP/6-31?+?G(d) level. Moreover, the results were also compared with the electrocyclization of 1,3,5 hexatriene to 1,3-cyclohexadiene . The Woodward-Hoffmann (W-H) allowed thermal reaction in syn CPD 11 has a calculated activation barrier of 6.23 kcal mol?1, compared with 29 kcal mol?1 for the electrocyclization of 1,3,5 hexatriene to 1,3-cyclohexadiene. The enhanced acceleration of electrocyclization is believed to arise from geometrically enforced spatially aligned termini of the hexatriene. Substituents at the electrocyclic terminus of cyclophanediene significantly affected (up to three fold) the activation barriers. Mono-substitution of CPD has substituent dependent acceleration or deceleration whereas di-substitution always increased the activation barrier. The activation barrier for electrocyclization in 33 is 4.44 kcal mol?1, which is the lowest activation barrier for any thermal electrocyclic reaction. Cyclophanedienes (CPDs) substituted with electron-rich substituents cyclized with high activation barriers and vice versa, a phenomenon significantly different from electrocyclic reaction of 1,3,5-hexatriene where no such trend is traceable. Comparison of W-H allowed and forbidden electrocyclization in syn and anti CPDs, respectively, revealed quite similar electronic demand, although the transition states are different in nature. The transition state for a W-H forbidden reaction is biradicaloid, with most of the spin density at the electrocyclic termini; however, the transition state for a W-H allowed reaction has no such contribution. We also believe that this is the first study of its type, where W-H allowed and forbidden reactions are compared on a similar set of molecules, and compared for electronic effect through substituents.  相似文献   

13.
The biocatalyzing effect of a novel cellulose acetate immobilized redox mediators technology (CE-RM) on nitrite denitrification process was studied with anthraquinone, 1,8-dichloroanthraquinone, 1,5-dichloroanthraquinone and 1,4,5,8-tetrachloroanthraquinone. The results showed that the immobilized 1,4,5,8-tetrachloroanthraquinone presented the best biocatalyzed effect which increased nitrite denitrification rate to 2.3-fold with 12 mmol/L 1,4,5,8-tetrachloroanthraquinone. The unequal biocatalyzing effect was due to the quantity and position of –Cl substituent in anthraquinone-structure. Moreover, the nitrite denitrification rate was increased with the oxidation reduction potential (ORP) values becoming more negative during the biocatalyzing process. The stabilized ORP value with 12 mmol/L immobilized 1,4,5,8-tetrachloroanthraquinone were 81 mV lower than the control. At the same time, the more OH? was produced with the higher nitrite removal rate achieved in the nitrite denitrification process. In addition, a positive linear correlation was found between the nitrite removal reaction constants k [gNO2 ?–N/(gVSS d)] and immobilized 1,4,5,8-tetrachloroanthraquinone concentration (C 1,4,5,8-tetrachloroanthraquinone), which was k = 1.8443 C 1,4,5,8-tetrachloroanthraquinone + 33.75(R 2 = 0.9411). The initial nitrite concentration of 179 mgNO2 ?–N/L resulted in the maximum nitrite removal rate, which was 6.526[gNO2 ?–N/(gVSS d)]. These results show that the application of cellulose acetate immobilized redox mediators (CE-RM) can be valuable for increasing nitrite denitrification rate.  相似文献   

14.
Combining quantum-mechanical (QM) calculations with quantum theory of atoms in molecules (QTAIM) and using the methodology of sweeps of the energetic, electron-topological, geometric and polar parameters, which describe the course of the tautomerization along the intrinsic reaction coordinate (IRC), we showed for the first time that the biologically important A?A* base pair (Cs symmetry) formed by the amino and imino tautomers of adenine (A) tautomerizes via asynchronous concerted double proton transfer (DPT) through a transition state (TS), which is the A+?A? zwitterion with the separated charge, with Cs symmetry. The nine key points, which can be considered as electron-topological “fingerprints” of the asynchronous concerted A?A*?A*?A tautomerization process via the DPT, were detected and completely investigated along the IRC of the A?A*?A*?A tautomerization. Based on the sweeps of the H-bond energies, it was found that intermolecular antiparallel N6Н?N6 (7.01 kcal mol?1) and N1H?N1 (6.88 kcal mol?1) H-bonds are significantly cooperative and mutually reinforce each other. It was shown for the first time that the A?A*?A*?A tautomerization is assisted by the third C2H?HC2 dihydrogen bond (DHB), which, in contrast to the two others N6H?N6 and N1H?N1 H-bonds, exists within the IRC range from ?2.92 to 2.92 Å. The DHB cooperatively strengthens, reaching its maximum energy 0.42 kcal mol?1 at IRC?=??0.52 Å and minimum energy 0.25 kcal mol?1 at IRC?=??2.92 Å, and is accompanied by strengthening of the two other aforementioned classical H-bonds. We established that the C2H?HC2 DHB completely satisfies the electron-topological criteria for H-bonding, in particular Bader’s and all eight “two-molecule” Koch and Popelier’s criteria. The positive value of the Grunenberg’s compliance constant (5.203 Å/mdyn) at the TSA?A*?A*?A proves that the C2H?HC2 DHB is a stabilizing interaction. NBO analysis predicts transfer of charge from σ(C2–H) bonding orbital to σ*(H–C2) anti-bonding orbital; at this point, the stabilization energy E(2) is equal to 0.19 kcal mol?1 at the TSA?A*?A*?A.  相似文献   

15.
This mini-review summarizes studies my associates and I carried out that are relevant to the topic of the present volume [i.e. glutamate dehydrogenase (GDH)] using radioactive 13N (t1/2 9.96 min) as a biological tracer. These studies revealed the previously unrecognized rapidity with which nitrogen is exchanged among certain metabolites in vivo. For example, our work demonstrated that (a) the t1/2 for conversion of portal vein ammonia to urea in the rat liver is ∼10-11 s, despite the need for five enzyme-catalyzed steps and two mitochondrial transport steps, (b) the residence time for ammonia in the blood of anesthetized rats is ≤7-8 s, (c) the t1/2 for incorporation of blood-borne ammonia into glutamine in the normal rat brain is <3 s, and (d) equilibration between glutamate and aspartate nitrogen in rat liver is extremely rapid (seconds), a reflection of the fact that the components of the hepatic aspartate aminotransferase reaction are in thermodynamic equilibrium. Our work emphasizes the importance of the GDH reaction in rat liver as a conduit for dissimilating or assimilating ammonia as needed. In contrast, our work shows that the GDH reaction in rat brain appears to operate mostly in the direction of ammonia production (dissimilation). The importance of the GDH reaction as an endogenous source of ammonia in the brain and the relation of GDH to the brain glutamine cycle is discussed. Finally, our work integrates with the increasing use of positron emission tomography (PET) and nuclear magnetic resonance (NMR) to study brain ammonia uptake and brain glutamine, respectively, in normal individuals and in patients with liver disease or other diseases associated with hyperammonemia.  相似文献   

16.
An extracellular endo-dextranase has been isolated from Streptococcus mutans K1-R. Incubation of cell-free culture fluid with sucrose permitted the removal of a large proportion of the extracellular d-glucosyltransferases by irreversible adsorption onto the insoluble glucans that these enzymes synthesize from sucrose. The remaining d-glucosyltransferases were separated from dextranase by precipitation with ammonium sulphate, chromatography on hydroxylapatite and DEAE-cellulose, followed by filtration on Ultrogel. The major products of action of the purified dextranase on (1→6)-α-d-glucans were isomaltotriose (IM3), isomaltotetraose (IM4), and isomaltopentaose (IM5). Further hydrolysis of IM4 and IM5 occurred after prolonged incubation with excess of enzyme, to give d-glucose, IM2, and IM3. The relative rate of hydrolysis of isomaltose saccharides fell sharply with decreasing chainlength from IM12 to IM5. The hydrolysis of dextrans containing 96% or more of (1→6)-α-d-glucosidic linkages, expressed as apparent conversion into IM3, was virtually complete, and substrates such as Streptococcus sanguis glucan, containing sequences of (1→6)-α-d-glucosidic linkages, were also effectively hydrolyzed. Dextranase activity towards the soluble glucan of Streptococcus mutans was limited, and there was no action on the insoluble glucan synthesized by S. mutans sucrose 3-d-glucosyltransferase.  相似文献   

17.
The effects of morin and nordihydroguaiaretic acid (NDGA), two plant secondary metabolites, on porcine pancreatic phospholipase A2 (PLA2) were investigated by isothermal titration calorimetry (ITC) and in silico docking analyses. The binding energies obtained for NDGA and morin from the ITC studies are ? 6.36 and ? 5.91 kcal mol? 1, respectively. Similarly, the glide scores obtained for NDGA and morin towards PLA2 were ? 7.32 and ? 7.23 kcal mol? 1, respectively. Further the docked complexes were subjected to MD simulation in the presence of explicit water molecules to check the binding stability of the ligands in the active site of PLA2. The bound ligands make hydrogen bonds with the active site residues of the enzyme and coordinate bonds with catalytically important Ca2+ ion. The binding of ligands at the active site of PLA2 may also contribute to the reported anti-inflammatory properties of NDGA and morin.  相似文献   

18.
A new bioflocculant was produced by culturing Rhodococcus erythropolis in a cheap medium. When culture pH was 7.0, inoculum size was 2 % (v/v), Na2HPO4 concentration was 0.5 g L?1, and the ratio of sludge/livestock wastewater was 7:1 (v/v), a maximum flocculating rate of 87.6 % could be achieved. Among 13 different kinds of pretreatments for sludge, the optimal one was the thermal-alkaline pretreatment. Different from a bioflocculant produced in a standard medium, this bioflocculant was effective over a wide pH range from 2 to 12 with flocculating rates higher than 98 %. Approximately, 1.6 g L?1 of crude bioflocculant could be harvested using cold ethanol for extraction. This bioflocculant showed color removal rates up to 80 % when applied to direct and disperse dye solutions, but only 23.0 % for reactive dye solutions. Infrared spectrum showed that the bioflocculant contained functional groups such as –OH, –NH2, and –CONH2. Components in the bioflocculant consisted of 91.2 % of polysaccharides, 7.6 % of proteins, and 1.2 % of DNA. When the bioflocculant and copper sulfate (CuSO4) were used together for decolorization in actual dye wastewater, the optimum decolorization conditions were specified by the response surface methodology as pH 11, bioflocculant dosage of 40 mg/L, and CuSO4 80 mg/L, under which a decolorization rate of 93.9 % could be reached.  相似文献   

19.
We previously demonstrated a transmural gradient in Na/K pump current (I P) and [Na+] i , with the highest maximum I P and lowest [Na+] i in epicardium. The present study examines the relationship between the transmural gradient in I P and Na/Ca exchange (NCX). Myocytes were isolated from canine left ventricle. Whole-cell patch clamp was used to measure current generated by NCX (I NCX) and inward background calcium current (I ibCa), defined as inward current through Ca2+ channels less outward current through Ca2+-ATPase. When resting myocytes from endocardium (Endo), midmyocardium (Mid) or epicardium (Epi) were studied in the same conditions, I NCX was the same and I ibCa was zero. Moreover, Western blots were consistent with NCX protein being uniform across the wall. However, the gradient in [Na+] i , with I ibCa = 0, should create a gradient in [Ca2+] i . To test this hypothesis, we measured resting [Ca2+] i using two methods, based on either transport or the Ca2+-sensitive dye Fura2. Both methods demonstrated a significant transmural gradient in resting [Ca2+] i , with Endo > Mid > Epi. This gradient was eliminated by exposing Epi to sufficient ouabain to partially inhibit Na/K pumps, thus increasing [Na+] i to values similar to those in Endo. These data support the existence of a transmural gradient for Ca2+ removal by NCX. This gradient is not due to differences in expression of NCX; rather, it is generated by a transmural gradient in [Na+] i , which is due to a transmural gradient in plasma membrane expression of the Na/K pump.  相似文献   

20.
Polyvinyl alcohol–sodium alginate (PVA–SA) matrix was fabricated and red algae Jania rubens was embedded for removal of lead from aqueous solutions. The Pb(II) uptake rate was rapid primarily at 1 h and equilibrium was achieved within 2 h. The optimum pH was 5, the data were well fitted by Langmuir and Freundlich models, and RL values are in the range of 0.1–0.38. The sorption capacity (qe) of PVA–calcium alginate (CA)–J. rubens matrix increased from 10.77 to 37.195 mg g?1 with increasing Pb(II) concentration from 24.86 to 98.75 mg L?1 at the temperature of 30°C and pH 5. The sorption capacity (qe) and maximum biosorption (qm) were noted as 37.179 ± 0.32 and 71.43 mg/g, respectively. The adsorption process was well described by pseudo-second-order model. The reaction is endothermic, is spontaneous, and increases in randomness. The functional groups present on matrix, i.e., –OH, –C–N, –C–O,–CO–NH, –NH2, –SH, and –C–OH, were intensely involved in the process. Scanning electron microscopy results revealed the morphological changes due to adsorption of Pb(II) on and inside of PVA–CA–J. rubens matrix. Desorption study indicates the efficient regeneration of PVA–CA–J. rubens biomass matrix for three cycles and is a promising matrix for removal of Pb(II) and can be used in continuous systems.  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