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 880 毫秒
1.
Carbonic anhydrases (CAs) have been given much attention as biocatalysts for CO2 sequestration process because of their ability to convert CO2 to bicarbonate. Here, we expressed codon-optimized sequence of ??-type CA cloned from Dunaliella species (Dsp-aCAopt) and characterized its catalyzing properties to apply for CO2 to calcite formation. The expressed amount of Dsp-aCAopt in Escherichia coli is about 50?mg/L via induction of 1.0?mM isopropyl-??-d-thiogalactopyranoside at 20?°C (for the case of intact Dsp-aCA, negligible). Dsp-aCAopt enzyme shows 47?°C of half-denaturation temperature and show wide pH stability (optimum pH 7.6/10.0). Apparent values of K m and V max for p-nitrophenylacetate substrate are 0.91?mM and 3.303?×?10?5???M?min?1. The effects of metal ions and anions were investigated to find out which factors enhance or inhibit Dsp-aCAopt activity. Finally, we demonstrated that Dsp-aCAopt enzyme can catalyze well the conversion of CO2 to CaCO3, as the calcite form, in the Ca2+ solution [8.9?mg/100???g (172?U/mg enzyme) with 10?mM of Ca2+]. The obtained expression and characterization results of Dsp-aCAopt would be usefully employed for the development of efficient CA-based system for CO2-converting/capturing processes.  相似文献   

2.
Highly alkaline electrolytes have been shown to improve the formation rate of C2+ products in the electrochemical reduction of carbon dioxide (CO2) and carbon monoxide (CO) on copper surfaces, with the assumption that higher OH? concentrations promote the C?C coupling chemistry. Herein, by systematically varying the concentration of Na+ and OH? at the same absolute electrode potential, we demonstrate that higher concentrations of cations (Na+), rather than OH?, exert the main promotional effect on the production of C2+ products. The impact of the nature and the concentration of cations on the electrochemical reduction of CO is supported by experiments in which a fraction or all of Na+ is chelated by a crown ether. Chelation of Na+ leads to drastic decrease in the formation rate of C2+ products. The promotional effect of OH? determined at the same potential on the reversible hydrogen electrode scale is likely caused by larger overpotentials at higher electrolyte pH.  相似文献   

3.
The role of C? C bond‐forming reactions such as aldol condensation in the degradation of organic matter in natural environments is receiving a renewed interest because naturally occurring ions, ammonium ions, NH+4, and carbonate ions, CO32?, have recently been reported to catalyze these reactions. While the catalysis of aldol condensation by OH? has been widely studied, the catalytic properties of carbonate ions, CO32?, have been little studied, especially under environmental conditions. This work presents a study of the catalysis of the aldol condensation of acetaldehyde in aqueous solutions of sodium carbonate (0.1–50 mM) at T = 295 ± 2 K. By monitoring the absorbance of the main product, crotonaldehyde, instead of that of acetaldehyde, interferences from other reaction products and from side reactions, in particular a known Cannizzaro reaction, were avoided. The rate constant was found to be first order in acetaldehyde in the presence of both CO32? and OH?, suggesting that previous studies reporting a second order for this base‐catalyzed reaction were flawed. Comparisons between the rate constants in carbonate solutions and in sodium hydroxide solutions ([NaOH] = 0.3–50 mM) showed that, among the three bases present in carbonate solutions, CO32?, HCO3?, and OH?, OH? was the main catalyst for pH ≤ 11. CO32? became the main catalyst at higher pH, whereas the catalytic contribution of HCO3? was negligible over the range of conditions studied (pH 10.3–11.3). Carbonate‐catalyzed condensation reactions could contribute significantly to the degradation of organic matter in hyperalkaline natural environments (pH ≥ 11) and be at the origin of the macromolecular matter found in these environments. © 2010 Wiley Periodicals, Inc. Int J Chem Kinet 42: 676–686, 2010  相似文献   

4.
The aim of this study was to identify gamma self-absorption correction factors for different types of Egyptian Mediterranean coastal sediments. Self-absorption corrections based on direct transmission through different thicknesses of the most dominant sediment species have been tested against point sources with gamma-ray energies of 241Am, 137Cs and 60Co with 2% uncertainties. Black sand samples from the Rashid branch of the Nile River quantitatively absorbed the low energy of 241Am through a thickness of 5 cm. In decreasing order of gamma energy self-absorption of 241Am, the samples under investigation ranked black sand, Matrouh sand, Sidi Gaber sand, shells, Salloum sand, and clay. Empirical self-absorption correction formulas were also deduced. Chemical analyses such as pH, CaCO3, total dissolved solids, Ca2+, Mg2+, CO32?, HCO3? and total Fe2+ have been carried out for the sediments. The relationships between self absorption corrections and the other chemical parameters of the sediments were also examined.  相似文献   

5.
Single-phase AB-type carbonate apatites were prepared by sintering appropriate mixtures of CaHPO4 and CaCO3 at 870°C in a CO2 atmosphere with a partial water vapor pressure of 5 mm Hg. Chemical and physical analyses indicate that at a constant CO32?/OH? ratio in the hydroxyl sublattice, carbonate substitutes for phosphate on a 1:1 mole basis. For every three PO43? ions substituted, two vacancies in the Ca2+ sublattice and one in the OH? sublattice are created. The same substitution mechanism seems to apply in pure B-type carbonate apatite.  相似文献   

6.
The hydrolysis kinetics of the dimeric complex (CuATP2? · OH2)2 {D} up to ≈40% ATP conversion at 25°C, pH 5.7–7.8, and [Cu · ATP]0 = (2.07 ± 0.03) × 10?3 mol/l is analyzed by numerical simulation. CuADP? + Pi (Pi is an inorganic phosphate) form from DOH?, and the latter forms rapidly from D. The abstraction of H+ from the coordinated H2O molecule is an irreversible reaction involving an OH? ion from the medium. The maximum possible DOH? concentration at a given pH is reached at the initial stage of hydrolysis (0.3–6.0 min after the initiation of hydrolysis). CuADP? + Pi form from D via two consecutive irreversible steps. The ADP buildup rate in the process is determined by the reversible conformational transformation of DOH? resulting in a pentacovalent intermediate (IntK). OH? ions from the medium are involved both in IntK formation and in the reverse reaction and are a hydrolysis inhibitor. AMP forms from the intermediate IntK3, which forms reversibly from DOH?, OH? ions from the medium being involved in the forward and reverse reactions. This is followed by irreversible (AMPH)? formation involving H3O+ ions from the medium. The rate and equilibrium constants are determined for the formation and decomposition of hydrolysis intermediates. The concentrations of the intermediates are plotted versus time for various pH values. The structures of the intermediates are suggested. The causes of a peak appearing in the initial ADP formation rate versus pH curve are analyzed.  相似文献   

7.
This work investigates the thermal decomposition of magnesian kutnahorite, which belongs to the dolomite group.The DTA curve measured in static air using a small amount of sample (5.0 mg) is quite different from those published previously. This difference might be due to the effect of a self-generated CO2 atmosphere.In a CO2 flow of 100 ml min?1, magnesian kutnahorite decomposes in four steps. Mg-kutnahorite → CaCO3 + Mg2MnO4 + Mn3O4 + MgO → CaCO3 + CaMnO3 + MgO → CaCO3 + CaMnO3 + Ca2MnO4 + MgO → CaMnO3 + Ca2MnO4 + MgO + CaO.However, in a mixed gas flow of CO2 at 95 ml min?1 and CO at 5 ml min?1, it decomposes, like dolomite, in two steps. Mg-kutnahorite → CaCO3 + (Mg,Mn)O- → (Ca, Mn)O + (Mg,Mn)O-.It has been found that the equilibrium redistribution of Mn between (Ca, Mn)O- and (Mg, Mn)O- is achieved at the second decomposition step. This is supported by theoretical considerations.Consequently, when the O2 partial pressure in the atmosphere is low enough to keep Mn in a bivalent state, the Mn bearing dolomite group mineral decomposes in a similar manner to dolomite itself.  相似文献   

8.
The dissociative photoionization of molecular‐beam cooled CH2CO in a region of ?10–20 eV was investigated with photoionization mass spectrometry using a synchrotron radiation as the light source. Photoionization efficiency curves of CH2CO+ and of observed fragment ions CH2+, CHCO+, HCO+, C2O+, CO+, and C2H2+ were measured to determine their appearance energies. Relative branching ratios as a function of photon energy were determined. Energies for formation of these observed fragment ions and their neutral counterparts upon ionization of CH2CO are computed with the Gaussian‐3 method. Dissociative photoionization channels associated with six observed fragment ions are proposed based on comparison of determined appearance energies and predicted energies. The principal dissociative processes are direct breaking of C=C and C‐H bonds to form CH2+ + CO and CHCO+ + H, respectively; at greater energies, dissociation involving H migration takes place.  相似文献   

9.
The synthesis and reactivity of a CoI pincer complex [Co(?3P,CH,P‐P(CH)PNMeiPr)(CO)2]+ featuring an η2‐ Caryl?H agostic bond is described. This complex was obtained by protonation of the CoI complex [Co(PCPNMeiPr)(CO)2]. The CoIII hydride complex [Co(PCPNMeiPr)(CNtBu)2(H)]+ was obtained upon protonation of [Co(PCPNMeiPr)(CNtBu)2]. Three ways to cleave the agostic C?H bond are presented. First, owing to the acidity of the agostic proton, treatment with pyridine results in facile deprotonation (C?H bond cleavage) and reformation of [Co(PCPNMeiPr)(CO)2]. Second, C?H bond cleavage is achieved upon exposure of [Co(?3P,CH,P‐P(CH)PNMeiPr)(CO)2]+ to oxygen or TEMPO to yield the paramagnetic CoII PCP complex [Co(PCPNMeiPr)(CO)2]+. Finally, replacement of one CO ligand in [Co(?3P,CH,P‐P(CH)PNMeiPr)(CO)2]+ by CNtBu promotes the rapid oxidative addition of the agostic η2‐Caryl?H bond to give two isomeric hydride complexes of the type [Co(PCPNMeiPr)(CNtBu)(CO)(H)]+.  相似文献   

10.
By combining results from a variety of mass spectrometric techniques (metastable ion, collisional activation, collision-induced dissociative ionization, neutralization-reionization spectrometry, 2H, 13C and 18O isotopic labelling and appearance energy measurements) and high-level ab initio molecular orbital calculations, the potential energy surface of the [CH5NO]+ ˙ system has been explored. The calculations show that at least nine stable isomers exist. These include the conventional species [CH3ONH2]+ ˙ and [HO? CH2? NH2]+ ˙, the distonic ions [O? CH2? NH3]+ ˙, [O? NH2? CH3]+ ˙, [CH2? O(H)? NH2]+ ˙, [HO? NH2? CH2]+ ˙, and the ion-dipole complex CH2?NH2+ …? OH˙. Surprisingly the distonic ion [CH2? O? NH3]+ ˙ was found not to be a stable species but to dissociate spontaneously to CH2?O + NH3+ ˙. The most stable isomer is the hydrogen-bridged radical cation [H? C?O …? H …? NH3]+ ˙ which is best viewed as an immonium cation interacting with the formyl dipole. The related species [CH2?O …? H …? NH2]+ ˙, in which an ammonium radical cation interacts with the formaldehyde dipole is also a very stable ion. It is generated by loss of CO from ionized methyl carbamate, H2N? C(?O)? OCH3 and the proposed mechanism involves a 1,4-H shift followed by intramolecular ‘dictation’ and CO extrusion. The [CH2?O …? H …? NH2]+ ˙ product ions fragment exothermically, but via a barrier, to NH4+ ˙ HCO…? and to H3N? C(H)?O+ ˙ H˙. Metastable ions [CH3ONH2]+…? dissociate, via a large barrier, to CH2?O + NH3+ + and to [CH2NH2]+ + OH˙ but not to CH2?O+ ˙ + NH3. The former reaction proceeds via a 1,3-H shift after which dissociation takes place immediately. Loss of OH˙ proceeds formally via a 1,2-CH3 shift to produce excited [O? NH2? CH3]+ ˙, which rearranges to excited [HO? NH2? CH2]+ ˙ via a 1,3-H shift after which dissociation follows.  相似文献   

11.
An analysis of the former works devoted to the reactions of I(III) in acidic nonbuffered solutions gives new thermodynamic and kinetic information. At low iodide concentrations, the rate law of the reaction IO + I? + 2H+ ? IO2H + IOH is k+B [IO][I?][H+]2k?B [IO2H][IOH] with k+B = 4.5 × 103 M?3s?1 and k?B = 240 M?1s?1 at 25°C and zero ionic strength. The rate law of the reaction IO2H + I? + H+ ? 2IOH is k+C [IO2H][I?][H+] – k?C [IOH]2 with k+C = 1.9 × 1010 M?2s?1 and k?C = 25 M?1s?1. These values lead to a Gibbs free energy of IO2H formation of ?95 kJ mol?1. The pKa of iodous acid should be about 6, leading to a Gibbs free energy of IO formation of about ?61 kJ mol?1. Estimations of the four rate constants at 50°C give, respectively, 1.2 × 104 M?3s?1, 590 M?1s?1, 2 × 109 M?2s?1, and 20 M?1 s?1. Mechanisms of these reactions involving the protonation IO2H + H+ ? IO2H and an explanation of the decrease of the last two rate constants when the temperature increases, are proposed. © 2008 Wiley Periodicals, Inc. Int J Chem Kinet 40: 647–652, 2008  相似文献   

12.
Reversible catalysis is a hallmark of energy‐efficient chemical transformations, but can only be achieved if the changes in free energy of intermediate steps are minimized and the catalytic cycle is devoid of high transition‐state barriers. Using these criteria, we demonstrate reversible CO2/HCO2? conversion catalyzed by [Pt(depe)2]2+ (depe=1,2‐bis(diethylphosphino)ethane). Direct measurement of the free energies associated with each catalytic step correctly predicts a slight bias towards CO2 reduction. We demonstrate how the experimentally measured free energy of each step directly contributes to the <50 mV overpotential. We also find that for CO2 reduction, H2 evolution is negligible and the Faradaic efficiency for HCO2? production is nearly quantitative. A free‐energy analysis reveals H2 evolution is endergonic, providing a thermodynamic basis for highly selective CO2 reduction.  相似文献   

13.
The polarization model is applied to Li+ and F? in their interactions with water and with themselves. The model is also applied to NH3 and NH4+ in their interactions with water. Results on applying the polarization model to PO43?, HPO42?, H2PO4? and H3PO4 in their interactions with water are reported. Finally, results on CO32?, HCO3?, H2CO3 and CO2 are reported.  相似文献   

14.
Deprotonation of methyl acetoacetate yields two enolate ions MeCOC?HCO2Me (a) and C?H2COCH2CO2Me (b). On collisional activation, ions a and b fragment differently. The major fragmentation of a is specific loss of MeOH through a four-centred transition state to form ?O(Me)C?C?C?O. In contrast, ion b eliminates CH2CO to give ?CH2CO2Me. Some rearrangement of b to a is also noted. Rearrangement of a to b is very minor under single collision conditions but at high collision gas pressure rearrangement of a to b is strongly promoted. Similar effects are observed in the collisional activation spectra of MeCOC?(Me)CO2Me (c) and ?CH2COC(Me)CO2Me (d). The loss of MeOH from (c) proceeds via a six membered transition state to ?CH2? CO? C(Me)?C?O; this is a stepwise process in which the deprotonation (step two) is not rate determining. A number of other decompositions occur, these have also been studied by deuterium labelling.  相似文献   

15.
Anion sensor properties of N‐alkyl‐substituted 1,4′‐diazaflavonium bromides in methanol–water were evaluated by UV–vis spectrometry. Pronounced changes were observed in the absorption spectra of all compounds for only OH?, CO32?, and CN? among F?, Cl?, Br?, I?, OH?, CO32?, NO3?, PO43?, CN?, SO42?, HSO4?, HCO3?, SCN?, NO2?, and P2O72? ions. Two new absorption bands at 385 and 685 nm accompanying the distinct color change for OH?, CO32?, and CN? ions were observed in case of all compounds. The color changes were from pink to blue for CO32? and OH? ions and from pink to purple for CN? ion. Thanks to the distinct color change, the compounds can be used as selective colorimetric anion sensors. Linear changes of absorbance of N‐heptyl‐substituted compound at 385 nm as a function of the ion concentration were used to determine CN? ion in water samples. Detection and quantification limits of the proposed method were 0.94 and 2.82 mg/L, respectively.  相似文献   

16.
A phenylenediamine‐capped conjugate of calix[4]arene ( Lamino ) was synthesized by reducing its precursor, Limino , with sodium borohydride in methanol. The Lamino sample binds to anions due to the more flexible and bent conformation of the capped aminophenolic binding core, compared to the precursor Limino . The Lamino sample showed selectivity towards H2PO4? by exhibiting a ratiometric increase in emission by about 11‐fold with a detection limit of (1.2±0.2) μm ((116±20) ppb) over 15 anions studied, including other phosphates, such as P2O74?, adenosine monophosphate (AMP2?), adenosine diphosphate (ADP2?), and adenosine triphosphate (ATP2?). The Lamino sample shows an increase in the absorbance at λ=315 nm in the presence of H2PO4?, CO32?, HCO3?, CH3CO2?, and F?. The 1H NMR spectroscopic titration of Lamino with H2PO4?, F?, and CH3CO2? showed major changes in the phenylene‐capped and salicyl moieties, and thereby, confirming the aminophenolic region as the binding core. However, the binding strength of these anions followed the trend H2PO4?>F??CH3CO2?>HSO4?. The heat changes observed by isothermal titration calorimetry support this trend. The Lamino sample showed reversible sensing towards H2PO4? and F? in the presence of Mg2+ and Ca2+, respectively. NOESY studies of Lamino , in comparison with its anionic complexes, revealed that major conformational changes occurred in the capping region to facilitate the binding of anion. ESI‐MS and the Job's method revealed 1:1 stoichiometry between Lamino and H2PO4? or F?. In the SEM micrographs of Lamino , the spherical particles are converted into spherical aggregates and further form large agglomerates and even branched sheets in the presence of anions, depending upon their binding strength.  相似文献   

17.
Thermal decomposition of supported magnesium formate has been studied by gas chro-matography.The reaction paths of decomposition of supported magnesium formate depend on thenature of the supports.For Mg(HCO_2)_2/HZSM-5,the zeolite behaves as a dehydration catalyst togive CO and H_2O at lower temperatures;when the zeolite is modified by phosphorus,the methanationreaction will be partly restrained.In the case of Mg(HCO_2)_2/AC,strong adsorption of CO_2 leadsto the formation of the shoulder peak of CO_2 at higher temperatures,however,CH_4 disappears aftermodified by phosphorus.For Mg(HCO_2)_2/Al_2O_3,the dehydrogenation of HCO_2~- takes place on thesurface of Al_2O_3.The decomposition of Mg(HCO_2)_2 on SiO_2 in hydrogen yields two peaks of COand only one appears after modified by phosphorus.When Mg(HCO_2)_2 decomposes on MgO,the firstpeak of CO_2 arises from the reaction of surface Mg~(2+) with HCO_2~- from dissociated Mg(HCO_2)_2.  相似文献   

18.
Bovine carbonic anhydrase (BCA) was covalently immobilized onto OAPS (octa(aminophenyl)silsesquioxane)‐functionalized Fe3O4/SiO2 nanoparticles by using glutaraldehyde as a spacer. The Fe3O4 nanoparticles were coated with SiO2, onto which was grafted OAPS, and the product was characterized using SEM, TEM, XRD, IR, X‐ray photoelectron spectroscopy (XPS), and magnetometer analysis. The enzymatic activities of the free and Fe3O4/SiO2/OAPS‐conjugated BCA (Fe? CA) were investigated by hydrolyzing p‐nitrophenylacetate (p‐NPA), and hydration and sequestration of CO2 to CaCO3. The CO2 conversion efficiency and reusability of the Fe? CA were studied before and after washing the recovered Fe? CA by applying a magnetic field and quantifying the unreacted Ca2+ ions by using ion chromatography. After 30 cycles, the Fe? CA displayed strong activity, and the CO2 capture efficiency was 26‐fold higher than that of the free enzyme. Storage stability studies suggested that Fe? CA retained nearly 82 % of its activity after 30 days. Nucleation of the precipitated CaCO3 was monitored by using polarized light microscopy, which revealed the formation of two phases, calcite and valerite, at pH 10 upon addition of serine. The magnetic nanobiocatalyst was shown to be an excellent reusable catalyst for the sequestration of CO2.  相似文献   

19.
Na-montmorillonites were exchanged with Li+, K+, Rb+, Cs+, Mg2+, Ca2+, Sr2+, and Ba2+, while Ca-montmorillonites were treated with alkaline and alkaline earth ions except for Ra2+ and Ca2+. Montmorillonites with interlayer cations Li+ or Na+ have remarkable swelling capacity and keep excellent stability. It is shown that metal ions represent different exchange ability as follows: Cs+?>?Rb+?>?K+?>?Na+?>?Li+ and Ba2+?>?Sr2+?>?Ca2+?>?Mg2+. The cation exchange capacity with single ion exchange capacity illustrates that Mg2+ and Ca2+ do not only take part in cation exchange but also produce physical adsorption on the montmorillonite. Although interlayer spacing d 001 depends on both radius and hydration radius of interlayer cations, the latter one plays a decisive role in changing d 001 value. Three stages of temperature intervals of dehydration are observed from the TG/DSC curves: the release of surface water adsorbed (36?C84?°C), the dehydration of interlayer water and the chemical-adsorption water (47?C189?°C) and dehydration of bound water of interlayer metal cation (108?C268?°C). Data show that the quantity and hydration energy of ions adsorbed on montmorillonite influence the water content in montmorillonite. Mg2+-modified Na-montmorillonite which absorbs the most quantity of ions with the highest hydration energy has the maximum water content up to 8.84%.  相似文献   

20.
Although FeO42? (ferrate(IV)) is a very strong oxidant that readily oxidizes water in acidic medium, at pH 9–10 it is relatively stable (<2 % decomposition after 1 h at 298 K). However, FeO42? is readily activated by Ca2+ at pH 9–10 to generate O2. The reaction has the following rate law: d[O2]/dt=kCa[Ca2+][FeO42?]2. 18O‐labeling experiments show that both O atoms in O2 come from FeO42?. These results together with DFT calculations suggest that the function of Ca2+ is to facilitate O–O coupling between two FeO42‐ions by bridging them together. Similar activating effects are also observed with Mg2+ and Sr2+.  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