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.
Nanosilver of 10-nm size was prepared by the NaBH4–sodium citrate procedure, and it was modified by a single-strand DNA (ssDNA) aptamer to fabricate an AgssDNA probe for melamine. The probe was stabile at pH 7.0 Na2HPO4–NaH2PO4 buffer solutions and in the presence of 25.0 mmol/L NaCl. Upon the addition of melamine, it interacted with the probe to aggregate big clusters, which led to the resonance scattering (RS) intensity at 470 nm increasing greatly. Under the selected conditions, the increased RS intensity (ΔI 470 nm) is linear to melamine concentration in the range of 6.31–378.4 μg/L, with a regression equation of DI470 nm = 1.124c + 10.8 \Delta {I_{{47}0{\rm{ nm}}}} = {1}.{124}c + { 10}.{8} and a detection limit of 3.1 μg/L. The aptamer-modified nanosilver RS assay has been applied for the determination of melamine in milk, with satisfactory results.  相似文献   

2.
Nanogold of 10 nm was used to label carcinoembryonic antigen antibody (CEAAb) to prepare a probe (Au-CEAAb) for carcinoembryonic antigen (CEA). In a Na2HPO4–NaH2PO4 buffer solution of pH 6.8, CEA reacted with Au-CEAAb to form a big Au-CEAAb–CEA immunocomplex that can be removed by centrifugation. The unreacted Au-CEAAb in the centrifugal supernatant exhibited catalytic effect on the Cu2O particle reaction, and the Cu2O particles displayed a resonance scattering (RS) peak at 602 nm. When CEA increased, the RS intensity at 602 nm decreased, and the decreased RS intensity (ΔI 602 nm) was linear to CEA concentration (C CEA) in the range of 0.02–12 ng mL−1, with the regression equation of ΔI 602 nm = 27.1 C CEA + 3.3, correlation coefficient of 0.9978 and detection limit of 3 pg mL−1 CEA. The proposed method was applied to detect CEA in real samples, with satisfactory results.  相似文献   

3.
Gold nanoparticle particles in size of 10 nm were used to label the thiol-modified single-stranded DNA aptamer (SH-ssDNA) to obtain an aptamer-modified gold nanoparticle probe (AussDNA) for target DNA (tDNA). In pH 7.4 NaH2PO4–Na2HPO4 buffer solution, the hybridization reaction between AussDNA and tDNA took place to form larger aptamer-modified gold nanoparticle cluster complex. The excess aptamer-modified gold nanoparticle probe in the supernatant solutions was obtained by centrifuging and can be used as nanocatalyst for the 0.276 mmol/L CuSO4-65.4 mmol/L potassium-sodium tartrate-0.37 mmol/L glucose system at 70 °C. The cubic Cu2O particles generated by the nanocatalytic reducing exhibit a strong resonance scattering (RS) peak at 620 nm. In the selected conditions, the RS intensity at 620 nm decreased with addition of tDNA, and the decreased intensity ΔI 620 nm is proportional to tDNA concentration (C tDNA) from 0.12 to 72 pM, with regress equation of ΔI 620 nm = 1.29C tDNA + 4.05, correlation coefficient of 0.9917, and detection limit of 0.084 pM tDNA.  相似文献   

4.
In the pH 6.6 Na2HPO4–NaH2PO4 buffer solutions and in the presence of urease catalyst, urea can be decomposed to form NH4 +. The NH4 + reacted with sodium tetraphenyl boron (NaTPB) to form the association particles that exhibited a resonance scattering (RS) peak at 474 nm. When the urea concentration increased, NH4 + increased, and RS intensity at 474 nm enhanced linearly. Under the chosen conditions, the increased RS intensity (ΔI 474 nm) had a linear response to the urea concentration in the range of 0.125–15 μM, with a detection limit of 0.058 μM urea, and a regression equation of ΔI 474 nm = 31.6C + 2.1, a correlation coefficient of 0.9986. This catalytic RS method was applied for the detection of urea in human serum sample, with good selectivity and sensitivity, and the results were consistent with the reference method.  相似文献   

5.
The resonance scattering spectral probe for Pb2+ was obtained using aptamer-modified AuPd Nanoalloy. In the pH 7.0 Na2HPO4–NaH2PO4 buffer solution, the aptamer interacted with AuPd nanoalloy particles to form stable aptamer-AuPd nanoalloy probe for Pb2+ that is stable in high concentration of salt. The probe combined with Pb2+ ions to form a G-quadruplex and to release AuPd nanoalloy particles that aggregate to form big particles which led the resonance scattering (RS) intensity enhancing. The reaction solution was filtered by 0.15 μm membrane to obtain the filtration containing aptamer-AuPd nanoalloy probe that has strong catalytic effect on the electrodeless nickel particle plating reaction between Ni(II) and PO23− that exhibited a strong RS peak at 508 nm. The RS intensity at 508 nm decreased when the Pb2+ concentration increased. The decreased intensity (ΔI 508nm) is linear to the concentration of 0.08–42 nM Pb2+, with regress equation of DI508nm = 16.3 c + 1.5 \Delta {I_{{5}0{\rm{8nm}}}} = {16}.{3}\,c + {1}.{5} , correlation coefficient of 0.9965, and detection limit of 0.04 nM Pb2+. The RS assay was applied to the analysis of Pb2+ in wastewater, with satisfactory results.  相似文献   

6.
In the medium of H2SO4 and in the presence of TiO2+, gold nanoparticles in size of 10 nm exhibited a weak surface plasmon resonance scattering (SPRS) peak at 775 nm. Upon addition of trace H2O2, the yellow complex [TiO(H2O2)]2+ formed that cause the gold nanoparticles aggregations to form bigger gold nanoparticle clusters in size of about 900 nm, and the SPRS intensity at 775 nm (I) enhanced greatly. The enhanced intensity ΔI was linear to the H2O2 concentration in the range of 0.025–48.7 μg/mL, with a detection limit of 0.014 μg/mL H2O2. This SPRS method was applied to determining H2O2 in water samples with satisfactory results.  相似文献   

7.
Aptamers interacting selectively with the anion-binding exosites 1 and 2 of thrombin were merged into dimeric oligonucleotide constructs by means of a poly-(dT)-linker of 35 nucleotides (nt) in length. Complexes of thrombin with the aptamers and their hetero- and homodimeric constructs were measured using an optical biosensor Biacore-3000. The K D values obtained for the hetero- and homodimeric constructs were correspondingly 25–30- and 2–3-fold lower than those for the primary aptamers. Analysis of temperature dependencies of the K D values within the temperature interval of 10–40°C has shown that affinity increases with the temperature decrease. The values of the enthalpy change ΔH upon formation of complexes of thrombin with the aptamers and the heterodimeric construct were basically the same. The value of the entropy change ΔS upon complex formation of thrombin with the aptamer heterodimeric construct was 1.5–2-fold higher than the ΔS values for the complexes with the aptamers. The complex formation and dissociation rates increased with the elevation of temperature from 10 to 37°C. However, at both temperatures the dissociation rate for the complex of thrombin with the heterodimeric construct was evidently lower that for the complexes with the aptamers.  相似文献   

8.
Zhang Z  Jia Y  Gao H  Zhang L  Li H  Meng Q 《Planta》2011,234(5):883-889
By simultaneously analyzing the chlorophyll a fluorescence transient and light absorbance at 820 nm as well as chlorophyll fluorescence quenching, we investigated the effects of different photon flux densities (0, 15, 200 μmol m−2 s−1) with or without 3-(3,4-dichlorophenyl)-1,1-dimethylurea (DCMU) on the repair process of cucumber (Cucumis sativus L.) leaves after treatment with low temperature (6°C) combined with moderate photon flux density (200 μmol m−2 s−1) for 6 h. Both the maximal photochemical efficiency of Photosystem II (PSII) (F v/F m) and the content of active P700 (ΔI/I o) significantly decreased after chilling treatment under 200 μmol m−2 s−1 light. After the leaves were transferred to 25°C, F v/F m recovered quickly under both 200 and 15 μmol m−2 s−1 light. ΔI/I o recovered quickly under 15 μmol m−2 s−1 light, but the recovery rate of ΔI/I o was slower than that of F v/F m. The cyclic electron transport was inhibited by chilling-light treatment obviously. The recovery of ΔI/I o was severely suppressed by 200 μmol m−2 s−1 light, whereas a pretreatment with DCMU effectively relieved this suppression. The cyclic electron transport around PSI recovered in a similar way as the active P700 content did, and the recovery of them was both accelerated by pretreatment with DCMU. The results indicate that limiting electron transport from PSII to PSI protected PSI from further photoinhibition, accelerating the recovery of PSI. Under a given photon flux density, faster recovery of PSII compared to PSI was detrimental to the recovery of PSI or even to the whole photosystem.  相似文献   

9.
The biosorption equilibrium isotherms of Ni(II) onto marine brown algae Lobophora variegata, which was chemically-modified by CaCl2 were studied and modeled. To predict the biosorption isotherms and to determine the characteristic parameters for process design, twenty-three one-, two-, three-, four- and five-parameter isotherm models were applied to experimental data. The interaction among biosorbed molecules is attractive and biosorption is carried out on energetically different sites and is an endothermic process. The five-parameter Fritz–Schluender model gives the most accurate fit with high regression coefficient, R 2 (0.9911–0.9975) and F-ratio (118.03–179.96), and low standard error, SE (0.0902–0.0.1556) and the residual or sum of square error, SSE (0.0012–0.1789) values to all experimental data in comparison to other models. The biosorption isotherm models fitted the experimental data in the order: Fritz–Schluender (five-parameter) > Freundlich (two-parameter) > Langmuir (two-parameter) > Khan (three-parameter) > Fritz–Schluender (four-parameter). The thermodynamic parameters such as ΔG 0, ΔH 0 and ΔS 0 have been determined, which indicates the sorption of Ni(II) onto L. variegata was spontaneous and endothermic in nature.  相似文献   

10.
Elevation of the external potassium concentration induced a two-phase inward current in freshly isolated pyramidal hippocampal neurons. This current was voltage-dependent and demonstrated strong inward rectification. The current consisted of a leakage current and a time-dependent current (τ=40–50 msec at 21°C); the latter was designated asI ΔK. As was shown earlier, K+ is a major charge carrier in the development of slow potassium-activated current. The pharmacological properties ofI ΔK were studied using a patch-clamp technique.I ΔK was completely blocked by external 10 mM TEA or 5 mM Ba2+ (IC50=480±90mM) and exhibited low sensitivity to extracellular Cs+ (2 mM). This current was not affected by 1 mM 4-aminopyridine and was insensitive to a muscarinic agonist, carbachol (50 μM), and to 1 mM extracellular Cd2+. Elevation of external Ca2+ from 2.5 mM to 10 mM did not changeI ΔK. Our data indicate that the pharmacological properties ofI ΔK differ from those of other voltage-gated potassium currents, but more specific blockers must be used to make this evidence conclusive.  相似文献   

11.
The responses of freshly isolated hippocampal pyramidal neurons to rapid, elevations of the external potassium concentration ([K+] out ) were investigated using the whole-cell variation of a patch-clamp technique. An elevation of [K+] out induced a two-phase inward current at the membrane potentials more negative than the reversal potential for K ions. This current consisted of a leakage, current and a time-dependent current (τ=40–50 msec at 21°C), the latter designated below asI ΔK. It displayed first-order activation kinetics that showed neither voltage, nor concentration dependence. The amplitude of this current was determined by the external K+ concentration and increased with hyperpolarization. Voltage dependence ofI ΔK measured within the range from −20 to −120 mV was similar to that for inward rectifier. Activation ofI ΔK was utterly dependent on Na+; substitution of extracellular Na+ with choline chloride almost completely depressedI ΔK.I ΔK was absent in the cells freshly dissociated from the nodosal and dorsal root ganglia. This suggests that this earlier unrecognized current is instrumental in preserving densely packed hippocampal pyramidal neurons from sudden increases in [K+] out and following spontaneous over-excitation. It prevents the neurons from responding to K+-induced depolarizations by slowing down potassium influx.  相似文献   

12.
Shaker B potassium channels undergo rapid N-type and slow C-type inactivation. While N-type inactivation is supposed to be mediated by occlusion of the pore by the N-terminal protein structure, the molecular mechanisms leading to C-type inactivation are less well understood. Considering N-type inactivation as a model for a protein conformational transition, we investigated inactivation of heterologously expressed Shaker B potassium channels and mutants thereof, showing various degrees of C-type inactivation, under high hydrostatic (oil) pressure. In addition to the derived apparent activation and reaction volumes (ΔV), experiments at various temperatures yielded estimates for enthalpic (ΔH) and entropic (TΔS) contributions. N-type inactivation was accelerated by increasing temperature and slowed by high hydrostatic pressure yielding at equilibrium ΔH = 76 kJ/mole, TΔS = 82 kJ/mole, and ΔV = 0.18 nm3 indicating that the transition to the N-type inactivated state is accompanied by an increase in volume and a decrease in order. N-terminally deleted ShΔ6–46 constructs with additional mutations at either position 449 or 463 were used to investigate C-type inactivation. In particular at high temperatures, inactivation occurred in two phases indicating more than one process. At equilibrium the following values were estimated for the major inactivation component of mutant ShΔ6–46 T449A: ΔH = –64 kJ/mole, TΔS = –60 kJ/mole, and ΔV = –0.25 nm3, indicating that the C-type inactivated state occupies a smaller volume and is more ordered than the noninactivated state. Thus, hydrostatic pressure affects N- and C-type inactivation in opposite ways. Received: 17 May 1997 / Accepted: 18 July 1997  相似文献   

13.
 The dynamic quenching of the luminescence of racemic Eu(III)(pyridine-2,6-dicarboxylate=dpa)3 3– by the title proteins is investigated and the enantioselectivity of the proteins in the quenching of the Δ and Λ enantiomers of Eu(dpa)3 3– is determined. The two diastereomeric quenching rate constants pertaining to azurin (k q Δ=3.3×106, k q Λ=2.7×106 M–1 s–1, pH 7.2, ionic strength I=22 mM) are lower than for its Met→44Lys mutant (k q Δ=1.9×107, k q Λ=1.4×107 M–1 s–1, same pH and I), indicating that energy transfer occurs from Eu(dpa)3 3– to the Cu(II) centre when the luminophore is bound to the hydrophobic patch of the protein near residue 44. The enantioselectivity remains unaltered by the mutation: k q Δ/k q Λ=1.27±0.04, so Lys44 is probably not in direct contact with the Eu chelate. The I and pH dependence of k q indicate that the lysine residue interacts electrostatically with Eu(dpa)3 3–. For plastocyanin the quenching rates are of the order of 106 M–1 s–1; for amicyanin they are two orders of magnitude larger (k q Δ=12×107, k q Λ=11×107 M–1 s–1, pH 7.2, I=22 mM). The variation of k q is attributed to differences in the charge distribution on the proteins, which influences the binding of the luminophore to the protein surface. For amicyanin the anion binding site near Lys59 and Lys60 may be involved in the energy transfer. Received: 16 June 1998 / Accepted: 18 September 1998  相似文献   

14.
The chaperone protein CopC from Pseudomonas syringae features high-affinity binding sites (K D ~ 10−13 M) for both CuI (Met-rich) and CuII (His-rich). When presented with these sites in the apoprotein, electrospray ionisation mass spectrometry confirmed that cis-Pt(NH3)2Cl2 (cisplatin) and the fragments [PtIIL]2+ (L is 1,2-diaminoethane, 2,2′-bipyridine) occupied the CuI site specifically in the 1:1 Pt–CopC adducts (purified by cation-exchange chromatography). The cis-Pt(NH3)2 fragment was not present in these adducts (the dominant product for cisplatin was Pt–CopC in which all original ligands were displaced), while bidentate ligands L were retained in LPt–CopC adducts. In the context of the Met-rich CuI pump Ctr1 as a significant entry point for cisplatin into mammalian cells, the present work confirms the ability of Met-rich sites in proteins to remove all ligands from cisplatin. It focuses attention on the potential of proteins that are part of the natural copper transport pathways to sequester the drug. These pathways are worthy of further study at the molecular level. Electronic supplementary material  The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

15.
We have identified two types of peroxidases (POX), one ionically and one covalently bound to the particulate fraction, in stripe rust-infected and -uninfected wheat (Triticum aestivum L.) leaves. The cell walls contained a high level of POX, of which 73–76% was extractable by 1% NaCl and 24–26% by 5 mM EDTA in infected and non-infected leaves of HD 2329. The NaCl-released POX constituted the predominant fraction. Both NaCl- and EDTA-extracted POX exhibited maximum activity at pH 5.0 and had a K m (enzyme–substrate affinity measure) value of 1.61–1.70 and 1.64–1.67 mM, respectively, with o-dianisidine as the substrate. The V max (maximum catalytic rate) in the two extractions ranged between 7.06–7.45 and 6.65–7.82 μmol min−1 g−1 fresh weight. A temperature optimum of 50°C was observed for both the NaCl- and EDTA-released fractions. The two POX fractions showed a differential response to metal ions, suggesting their distinctive nature. Sodium azide inhibited POX activity markedly, which suggested the presence of heme as a prosthetic group. Inhibition of wall-bound POX by iodine and the regeneration of activity by mercaptoethanol suggested the involvement of cysteine in the active site of the enzyme. These two forms showed greater differences in terms of thermodynamic properties, such as the energy of activation (E a) and enthalpy change (ΔH), while entropy (ΔS) and free energy changes were similar. The results further show that pathogen infection of the leaves of this susceptible wheat cultivar induces an increase in the activity and kinetics of POX, which may be critical in the response of the plant cell to infection.  相似文献   

16.
In order to measure the substrate-oxidizing activity of intact cells of Acetobacter pasteurianus no. 2, a given amount of the bacterial cells was immobilized on a carbon-paste electrode, and the current at the electrode was measured in a buffer solution. When Fe(CN)3− 6 was added to the buffer solution, an anodic current was observed at 0.5 V (against Ag/AgCl). Further, when ethanol was added to the solution, the current started to increase to reach a steady-state within 3 min. The electrode had a good response to acetaldehyde and lactic acid as well as ethanol. Culture conditions affected the current response to various substances; the response of the electrode modified with the cells grown in static culture was much higher than that of the electrode with the cells grown in shaking culture, and the electrode with ethanol-grown cells had a high response to ethanol and acetaldehyde compared with that of the electrode with glucose-grown cells. The increase in the amount of the current after the addition of ethanol (ΔI EtOH) was linearly proportional to the total number of immobilized cells per electrode in the range 1.0 × 104–1.0 × 108 cells. The ΔI EtOH values were measured with the electrode prepared with a fixed volume of the cell suspensions taken from the culture at 6-h intervals; the dependence of the ΔI EtOH value on time agreed well with the cell growth measured by colony counting and turbidity in the lag and logarithmic phase. After the logarithmic phase, the value of ΔI EtOH sharply decreased, resembling to the growth measured by colony counting, rather than by turbidity. Received: 30 October 1998 / Received revision: 2 February 1999 / Accepted: 5 February 1999  相似文献   

17.
The role of the C-terminal region of Bacillus licheniformis γ-glutamyl transpeptidase (BlGGT) was investigated by deletion analysis. Seven C-terminally truncated BlGGTs lacking 581–585, 577–585, 576–585, 566–585, 558–585, 523–585, and 479–585 amino acids, respectively, were generated by site-directed mutagenesis. Deletion of the last nine amino acids had no appreciable effect on the autocatalytic processing of the enzyme, and the engineered protein was active towards the synthetic substrate L-γ-glutamyl-p-nitroanilide. However, a further deletion to Val576 impaired the autocatalytic processing. In vitro maturation experiments showed that the truncated BlGGT precursors, pro-Δ(576–585), pro-Δ(566–585), and pro-Δ(558–585), could partially precede a time-dependent autocatalytic process to generate the L- and S-subunits, and these proteins showed a dramatic decrease in catalytic activity with respect to the wild-type enzyme. The parental enzyme (BlGGT-4aa) and BlGGT were unfolded biphasically by guanidine hydrochloride (GdnCl), but Δ(577–585), Δ(576–585), Δ(566–585), Δ(558–585), Δ(523–585), and Δ(479–585) followed a monophasic unfolding process and showed a sequential reduction in the GdnCl concentration corresponding to half effect and ΔG 0 for the unfolding. BlGGT-4aa and BlGGT sedimented at ∼4.85 S and had a heterodimeric structure of approximately 65.23 kDa in solution, and this structure was conserved in all of the truncated proteins. The frictional ratio (f/f o) of BlGGT-4aa, BlGGT, Δ(581–585), and Δ(577–585) was 1.58, 1.57, 1.46, and 1.39, respectively, whereas the remaining enzymes existed exclusively as precursor form with a ratio of less than 1.18. Taken together, these results provide direct evidence for the functional role of the C-terminal region in the autocatalytic processing of BlGGT.  相似文献   

18.
Constitutive models describing the arterial mechanical behavior are important in the development of catheterization products, to be used in arteries with a specific radius. To prove the possible existence of a constitutive model that, provided with a generic set of material and geometric parameters, is able to predict the radius-specific mechanical behavior of a coronary artery, the passive pressure–inner radius (Pr i ) and pressure–axial force change (P–ΔF z ) relations of seven porcine left anterior descending coronary arteries were measured in an in-vitro set-up and fitted with the model of Driessen et al. in J Biomech Eng 127(3):494–503 (2005), Biomech Model Mechanobiol 7(2):93–103 (2008). Additionally, the collagen volume fraction, physiological axial pre-stretch, and wall thickness to inner radius ratio at physiological loading were determined for each artery. From this, two generic parameter sets, each comprising four material and three geometric parameters, were obtained. These generic sets were used to compute the deformation of each tested artery using a single radius measurement at physiological loading as an artery-specific input. Artery-specific Pr i and P–ΔF z relations were predicted with an accuracy of 32 μm (2.3%) and 6 mN (29% relative to ΔF z -range) on average compared to the relations measured in-vitro. It was concluded that the constitutive model provided with the generic parameters found in this study can well predict artery-specific mechanical behavior.  相似文献   

19.
We have identified two types of invertases, one bound ionically and the other covalently to the particulate fraction in grains of heat tolerant C 306 and heat susceptible WH 542 cultivars of wheat (Triticum aestivum L.). The cell walls contained a high level of invertase activity, of which 79.2–72.8% was extractable by 2 M NaCl and 14.9–21.1% by 0.5% EDTA in C 306 and WH 542, respectively. The NaCl-released invertase constituted the predominant fraction. Using 5–100 mM sucrose and pH range of 4.0–7.0, the apparent Michaelis constant (K m, enzyme substrate affinity measure) of enzyme ranged from 5.73 to 16.06 mM for C 306 and from 6.08 to 19.86 mM for WH 542. The V max (maximum catalytic rate) values at these pH were higher in C 306 (0.63–11.04 μg sucrose hydrolysed min−1) than WH 542 (0.51–8.73 μg sucrose hydrolysed min−1). By employing photo-oxidation and by studying the effect of pH on K m and V max, the involvement of histidine and α-carboxyl groups at the active site of the enzyme was indicated. The two cultivars also showed differential response in terms of thermodynamic properties of the enzyme i.e. energy of activation (E a), enthalpy change (ΔH) and entropy change (ΔS). NaCl-released invertase showed differential response to metal ions in two cultivars suggesting their distinctive nature. Mn2+, Cu2+, Hg2+, Mg2+, Zn2+ and Cd2+ were strong inhibitors in WH 542 as compared to C 306 while K+, Ca2+ were stimulators in both the cultivars. Overall the results suggest that genetic differences exist in wall bound invertase properties of wheat grains as evident in its altered kinetic behaviour.  相似文献   

20.
The present study characterizes changes in the electronic structure of reactants during chemical reactions based on the combined charge and energy decomposition scheme, ETS-NOCV (extended transition state–natural orbitals for chemical valence). Decomposition of the activation barrier, ΔE #, into stabilizing (orbital interaction, ΔE orb, and electrostatic, ΔE elstat) and destabilizing (Pauli repulsion, ΔE Pauli, and geometry distortion energy, ΔE dist) factors is discussed in detail for the following reactions: (I) hydrogen cyanide to hydrogen isocyanide, HCN → CNH isomerization; (II) Diels-Alder cycloaddition of ethene to 1,3-butadiene; and two catalytic processes, i.e., (III) insertion of ethylene into the metal-alkyl bond using half-titanocene with phenyl-phenoxy ligand catalyst; and (IV) B–H bond activation catalyzed by an Ir-containing catalyst. Various reference states for fragments were applied in ETS-NOCV analysis. We found that NOCV-based deformation densities (Δρ i) and the corresponding energies ΔE orb(i) obtained from the ETS-NOCV scheme provide a very useful picture, both qualitatively and quantitatively, of electronic density reorganization along the considered reaction pathways. Decomposition of the barrier ΔE# into stabilizing and destabilizing contributions allowed us to conclude that the main factor responsible for the existence of positive values of ΔE # for all processes (I, II, III and IV) is Pauli interaction, which is the origin of steric repulsion. In addition, in the case of reactions II, III and IV, a significant degree of structural deformation of the reactants, as measured by the geometry distortion energy, plays an important role. Depending on the reaction type, stabilization of the transition state (relatively to the reactants) originating either from the orbital interaction term or from electrostatic attraction can be of vital importance. Finally, use of the ETS-NOCV method to describe catalytic reactions allows extraction of information on the role of catalysts in determination of ΔE #.  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