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.
2.
In this work, the kinetics of short, fully complementary oligonucleotides are investigated at the single-molecule level. Constructs 6–9 bp in length exhibit single exponential kinetics over 2 orders of magnitude time for both forward (kon, association) and reverse (koff, dissociation) processes. Bimolecular rate constants for association are weakly sensitive to the number of basepairs in the duplex, with a 2.5-fold increase between 9 bp (k′on = 2.1(1) × 106 M−1 s−1) and 6 bp (k′on = 5.0(1) × 106 M−1 s−1) sequences. In sharp contrast, however, dissociation rate constants prove to be exponentially sensitive to sequence length, varying by nearly 600-fold over the same 9 bp (koff = 0.024 s−1) to 6 bp (koff = 14 s−1) range. The 8 bp sequence is explored in more detail, and the NaCl dependence of kon and koff is measured. Interestingly, konincreases by >40-fold (kon = 0.10(1) s−1 to 4.0(4) s−1 between [NaCl] = 25 mM and 1 M), whereas in contrast, koffdecreases by fourfold (0.72(3) s−1 to 0.17(7) s−1) over the same range of conditions. Thus, the equilibrium constant (Keq) increases by ≈160, largely due to changes in the association rate, kon. Finally, temperature-dependent measurements reveal that increased [NaCl] reduces the overall exothermicity (ΔΔH° > 0) of duplex formation, albeit by an amount smaller than the reduction in entropic penalty (−TΔΔS° < 0). This reduced entropic cost is attributed to a cation-facilitated preordering of the two single-stranded species, which lowers the association free-energy barrier and in turn accelerates the rate of duplex formation.  相似文献   

3.
Cold-adapted esterases and lipases have been found to be dominant activities throughout the cold marine environment, indicating their importance in bacterial degradation of the organic matter. lip2 Gene from Psychrobacter sp. TA144, a micro-organism isolated from the Antarctic sea water, was cloned and over-expressed in Escherichia coli. The recombinant protein (PsyHSL) accumulated in the insoluble fraction from which it was recovered in active form, purified to homogeneity and deeply characterised. Temperature dependence of PsyHSL activity was typical of psychrophilic enzymes, with an optimal temperature of 35 °C at pH 8.0. The enzyme resulted to be active on pNP-esters of fatty acids with acyl chain length from C2 to C12 and the preferred substrate was pNP-pentanoate showing a kcat = 26.2 ± 0.1 s−1, KM = 0.122 ± 0.006 mM and a kcat/KM = 215 ± 11 mM−1 s−1. The enzyme was strongly inhibited by Hg2+, Zn2+, Cu2+, Fe3+, Mn2+ ions and it resulted to be activated in presence of methanol and acetonitrile, with calculated C50 values of 1.98 M and 0.92 M, respectively.  相似文献   

4.
Kinetic studies of X exchange on [AuX4] square-planar complexes (where X=Cl and CN) were performed at acidic pH in the case of chloride system and as a function of pH for the cyanide one. Chloride NMR study (330-365 K) gives a second-order rate law on [AuCl4] with the kinetic parameters: (k2Au,Cl)298=0.56±0.03 s−1 mol−1 kg; ΔH2‡ Au,Cl=65.1±1 kJ mol−1; ΔS2‡ Au,Cl=−31.3±3 J mol−1 K−1 and ΔV2 Au,Cl=−14±2 cm3 mol−1. The variable pressure data clearly indicate the operation of an Ia or A mechanism for this exchange pathway. The proton exchange on HCN was determined by 13C NMR as a function of pH and the rate constant of the three reaction pathways involving H2O, OH and CN were determined: k0HCN,H=113±17 s−1, k1HCN,H=(2.9±0.7)×109 s−1 mol−1 kg and k2HCN,H=(0.6±0.2)×106 s−1 mol−1 kg at 298.1 K. The rate law of the cyanide exchange on [Au(CN)4] was found to be second order with the following kinetic parameters: (k2Au,CN)298=6240±85 s−1 mol−1 kg, ΔH2 Au,CN=40.0±0.8 kJ mol−1, ΔS2 Au,CN=−37.8±3 J mol−1 K−1 and ΔV2 Au,CN=+2±1 cm3 mol−1. The rate constant observed varies about nine orders of magnitude depending on the pH and HCN does not act as a nucleophile. The observed rate constant of X exchange on [AuX4] are two or three orders of magnitude faster than the Pt(II) analogue.  相似文献   

5.
Oxidation of the title complexes with ozone takes place by hydrogen atom, hydride, and electron transfer mechanisms. The reaction with (NH3)4(H2O)RhH2+ is a two electron process, believed to involve hydride transfer with a rate constant k = (2.2 ± 0.2) × 105 M−1 s−1 and an isotope effect kH/kD = 2. The oxidation of (NH3)4(H2O)RhOOH2+ to (NH3)4(H2O)RhOO2+ by an apparent hydrogen atom transfer is quantitative and fast, k = (6.9 ± 0.3) × 103 M−1 s−1, and constitutes a useful route for the preparation of the superoxo complex. The latter is also oxidized by ozone, but more slowly, k = 480 ± 50 M−1 s−1.  相似文献   

6.
Nitric oxide (NO) has a critical role in several physiological and pathophysiological processes. In this paper, the reactions of the nitrosyl complexes of [Ru(bpy)2L(NO)]n+ type, where L = SO32− and imidazole and bpy = 2,2′-bipiridine, with cysteine and glutathione were studied. The reactions with cysteine and glutathione occurred through the formation of two sequential intermediates, previously described elsewhere, [Ru(bpy)2L(NOSR)]n+ and [Ru(bpy)2L(NOSR)2] (SR = thiol) leading to the final products [Ru(bpy)2L(H2O)]n+ and free NO. The second order rate constant for the second step of this reaction was calculated for cysteine k2(SR) = (2.20 ± 0.12) × 109 M− 1 s− 1 and k2(RSH) = (154 ± 2) M− 1 s− 1 for L = SO32− and k2(SR) = (1.30 ± 0.23) × 109 M− 1 s− 1 and k2(RSH) = (0.84 ± 0.02) M− 1 s− 1 for L = imidazole; while for glutathione they were k2(SR) = (6.70 ± 0.32) × 108 M− 1 s− 1 and k2(RSH) = 11.8 ± 0.3 M− 1 s− 1 for L = SO32− and k2(SR) = (2.50 ± 0.36) × 108 M− 1 s− 1 and k2(RSH) = 0.32 ± 0.01 M− 1 s− 1 for L = imidazole. In all reactions it was possible to detect the release of NO from the complexes, which it is remarkably distinct from other ruthenium metallocompounds described elsewhere with just N2O production. These results shine light on the possible key role of NO release mediated by physiological thiols in reaction with these metallonitrosyl ruthenium complexes.  相似文献   

7.
Human serum albumin (HSA) is a monomeric allosteric protein. Here, the effect of ibuprofen on denitrosylation kinetics (koff) and spectroscopic properties of HSA-heme-Fe(II)-NO is reported. The koff value increases from (1.4 ± 0.2) × 10−4 s−1, in the absence of the drug, to (9.5 ± 1.2) × 10−3 s−1, in the presence of 1.0 × 10−2 M ibuprofen, at pH 7.0 and 10.0 °C. From the dependence of koff on the drug concentration, values of the dissociation equilibrium constants for ibuprofen binding to HSA-heme-Fe(II)-NO (K1 = (3.1 ± 0.4) × 10−7 M, K2 = (1.7 ± 0.2) × 10−4 M, and K3 = (2.2 ± 0.2) × 10−3 M) were determined. The K3 value corresponds to the value of the dissociation equilibrium constant for ibuprofen binding to HSA-heme-Fe(II)-NO determined by monitoring drug-dependent absorbance spectroscopic changes (H = (2.6 ± 0.3) × 10−3 M). Present data indicate that ibuprofen binds to the FA3-FA4 cleft (Sudlow’s site II), to the FA6 site, and possibly to the FA2 pocket, inducing the hexa-coordination of HSA-heme-Fe(II)-NO and triggering the heme-ligand dissociation kinetics.  相似文献   

8.
We highly purified the Chlamydomonas inner-arm dyneins e and c, considered to be single-headed subspecies. These two dyneins reside side-by-side along the peripheral doublet microtubules of the flagellum. Electron microscopic observations and single particle analysis showed that the head domains of these two dyneins were similar, whereas the tail domain of dynein e was short and bent in contrast to the straight tail of dynein c. The ATPase activities, both basal and microtubule-stimulated, of dynein e (kcat = 0.27 s–1 and kcat,MT = 1.09 s–1, respectively) were lower than those of dynein c (kcat = 1.75 s–1 and kcat,MT = 2.03 s–1, respectively). From in vitro motility assays, the apparent velocity of microtubule translocation by dynein e was found to be slow (Vap = 1.2 ± 0.1 μm/s) and appeared independent of the surface density of the motors, whereas dynein c was very fast (Vmax = 15.8 ± 1.5 μm/s) and highly sensitive to decreases in the surface density (Vmin = 2.2 ± 0.7 μm/s). Dynein e was expected to be a processive motor, since the relationship between the microtubule landing rate and the surface density of dynein e fitted well with first-power dependence. To obtain insight into the in vivo roles of dynein e, we measured the sliding velocity of microtubules driven by a mixture of dynein e and c at various ratios. The microtubule translocation by the fast dynein c became even faster in the presence of the slow dynein e, which could be explained by assuming that dynein e does not retard motility of faster dyneins. In flagella, dynein e likely acts as a facilitator by holding adjacent microtubules to aid dynein c’s power stroke.  相似文献   

9.
Escherichiacoli RecBCD is a bipolar DNA helicase possessing two motor subunits (RecB, a 3′-to-5′ translocase, and RecD, a 5′-to-3′ translocase) that is involved in the major pathway of recombinational repair. Previous studies indicated that the minimal kinetic mechanism needed to describe the ATP-dependent unwinding of blunt-ended DNA by RecBCD in vitro is a sequential n-step mechanism with two to three additional kinetic steps prior to initiating DNA unwinding. Since RecBCD can “melt out” ∼ 6 bp upon binding to the end of a blunt-ended DNA duplex in a Mg2+-dependent but ATP-independent reaction, we investigated the effects of noncomplementary single-stranded (ss) DNA tails [3′-(dT)6 and 5′-(dT)6 or 5′-(dT)10] on the mechanism of RecBCD and RecBC unwinding of duplex DNA using rapid kinetic methods. As with blunt-ended DNA, RecBCD unwinding of DNA possessing 3′-(dT)6 and 5′-(dT)6 noncomplementary ssDNA tails is well described by a sequential n-step mechanism with the same unwinding rate (mkU = 774 ± 16 bp s− 1) and kinetic step size (m = 3.3 ± 1.3 bp), yet two to three additional kinetic steps are still required prior to initiation of DNA unwinding (kC = 45 ± 2 s− 1). However, when the noncomplementary 5′ ssDNA tail is extended to 10 nt [5′-(dT)10 and 3′-(dT)6], the DNA end structure for which RecBCD displays optimal binding affinity, the additional kinetic steps are no longer needed, although a slightly slower unwinding rate (mkU = 538 ± 24 bp s− 1) is observed with a similar kinetic step size (m = 3.9 ± 0.5 bp). The RecBC DNA helicase (without the RecD subunit) does not initiate unwinding efficiently from a blunt DNA end. However, RecBC does initiate well from a DNA end possessing noncomplementary twin 5′-(dT)6 and 3′-(dT)6 tails, and unwinding can be described by a simple uniform n-step sequential scheme, without the need for the additional kC initiation steps, with a similar kinetic step size (m = 4.4 ± 1.7 bp) and unwinding rate (mkobs = 396 ± 15 bp s− 1). These results suggest that the additional kinetic steps with rate constant kC required for RecBCD to initiate unwinding of blunt-ended and twin (dT)6-tailed DNA reflect processes needed to engage the RecD motor with the 5′ ssDNA.  相似文献   

10.
The binding affinity of the two substrate–water molecules to the water-oxidizing Mn4CaO5 catalyst in photosystem II core complexes of the extremophilic red alga Cyanidioschyzon merolae was studied in the S2 and S3 states by the exchange of bound 16O-substrate against 18O-labeled water. The rate of this exchange was detected via the membrane-inlet mass spectrometric analysis of flash-induced oxygen evolution. For both redox states a fast and slow phase of water-exchange was resolved at the mixed labeled m/z 34 mass peak: kf = 52 ± 8 s− 1 and ks = 1.9 ± 0.3 s− 1 in the S2 state, and kf = 42 ± 2 s− 1 and kslow = 1.2 ± 0.3 s− 1 in S3, respectively. Overall these exchange rates are similar to those observed previously with preparations of other organisms. The most remarkable finding is a significantly slower exchange at the fast substrate–water site in the S2 state, which confirms beyond doubt that both substrate–water molecules are already bound in the S2 state. This leads to a very small change of the affinity for both the fast and the slowly exchanging substrates during the S2 → S3 transition. Implications for recent models for water-oxidation are briefly discussed.  相似文献   

11.
The copper (II) complex of a simple pyridine- and amide-containing copolymer serves as an effective catalyst for heterogeneous hydrolysis of the prototypical phosphodiester substrate bis(p-nitrophenyl)phosphate at pH 8.0 and 25 °C. The catalysis has a first-order rate constant of kcat = 8.3 × 10−6 s−1, corresponding to a catalytic proficiency of 75-thousand folds relative to the uncatalyzed hydrolysis with a rate constant of k0 = 1.1 × 10−10 s−1 in aqueous buffer solution at pH 8.0. This observation suggests that polymers can be designed to include various functional groups feasible for effective metal-centered catalysis of phosphodiester hydrolysis.  相似文献   

12.
Human serum albumin (HSA) participates to heme scavenging, in turn HSA-heme binds gaseous diatomic ligands at the heme-Fe-atom. Here, the effect of abacavir and warfarin on denitrosylation kinetics of HSA-heme-Fe(II)-NO (i.e., koff) is reported. In the absence of drugs, the value of koff is (1.3 ± 0.2) × 10−4 s−1. Abacavir and warfarin facilitate NO dissociation from HSA-heme-Fe(II)-NO, the koff value increases to (8.6 ± 0.9) × 10−4 s−1. From the dependence of koff on the drug concentration, values of the dissociation equilibrium constant for the abacavir and warfarin binding to HSA-heme-Fe(II)-NO (i.e., K = (1.2 ± 0.2) × 10−3 M and (6.2 ± 0.7) × 10−5 M, respectively) were determined. The increase of koff values reflects the stabilization of the basic form of HSA-heme-Fe by ligands (e.g., abacavir and warfarin) that bind to Sudlow’s site I. This event parallels the stabilization of the six-coordinate derivative of the HSA-heme-Fe(II)-NO atom. Present data highlight the allosteric modulation of HSA-heme-Fe(II) reactivity by heterotropic effectors.  相似文献   

13.
Cytochromes c6 and f react by three et mechanisms under similar conditions. We report temperature and viscosity effects on the protein docking and kinetics of 3Zncyt c6 + cyt f(III) → Zncyt c6+ + cyt f(II). At 0.5-40.0 °C, this reaction occurs within the persistent (associated) diprotein complex with the rate constant kpr and within the transient (collision) complex with the rate constant ktr. The viscosity independence of kpr, the donor-acceptor coupling Hab = (0.5 ± 0.1) cm−1, and reorganizational energy λ = (2.14 ± 0.02) eV indicate true et within the persistent complex. The viscosity dependence of ktr and a break at 30 °C in the Eyring plot for ktr reveal mechanisms within the transient complex that are reversibly switched by temperature change. Kramers protein friction parameters differ much for the reactions below (σ = 0.3 ± 0.1, δ = 0.85 ± 0.07) and above (σ = 4.0 ± 0.9, δ = 0.40 ± 0.06) 30 °C. The transient complex(es) undergo(es) coupled et below ca. 30 °C and gated et above ca. 30 °C. Brownian dynamics simulations reveal two broad, dynamic ensembles of configurations “bridged” by few intermediate configurations through which the interconversion presumably occurs.  相似文献   

14.
Quenching of Trp phosphorescence in proteins by diffusion of solutes of various molecular sizes unveils the frequency-amplitude of structural fluctuations. To cover the sizes gap between O2 and acrylamide, we examined the potential of acrylonitrile to probe conformational flexibility of proteins. The distance dependence of the through-space acrylonitrile quenching rate was determined in a glass at 77 K, with the indole analog 2-(3-indoyl) ethyl phenyl ketone. Intensity and decay kinetics data were fitted to a rate, k(r) = k0 exp[−(rr0)/re], with an attenuation length re = 0.03 nm and a contact rate k0 = 3.6 × 1010 s−1. At ambient temperature, the bimolecular quenching rate constant (kq) was determined for a series of proteins, appositely selected to test the importance of factors such as the degree of Trp burial and structural rigidity. Relative to kq = 1.9 × 109 M−1s−1 for free Trp in water, in proteins kq ranged from 6.5 × 106 M−1s−1 for superficial sites to 1.3 × 102 M−1s−1 for deep cores. The short-range nature of the interaction and the direct correlation between kq and structural flexibility attest that in the microsecond-second timescale of phosphorescence acrylonitrile readily penetrates even compact protein cores and exhibits significant sensitivity to variations in dynamical structure of the globular fold.  相似文献   

15.
Two 15N-labelled cis-Pt(II) diamine complexes with dimethylamine (15N-dma) and isopropylamine (15N-ipa) ligands have been prepared and characterised. [1H,15N] HSQC NMR spectroscopy is used to obtain the rate and equilibrium constants for the aquation of cis-[PtCl2(15N-dma)2] at 298 K in 0.1 M NaClO4 and to determine the pKa values of cis-[PtCl(H2O)(15N-dma)2]+ (6.37) and cis-[Pt(H2O)2(15N-dma)2]2+ (pKa1 = 5.17, pKa2 = 6.47). The rate constants for the first and second aquation steps (k1 = (2.12 ± 0.01) × 10−5 s−1, k2 = (8.7 ± 0.7) × 10−6 s−1) and anation steps (k−1 = (6.7 ± 0.8) × 10−3 M−1 s−1, k−2 = 0.043 ± 0.004 M−1 s−1) are very similar to those reported for cisplatin under similar conditions, and a minor difference is that slow formation of the hydroxo-bridged dimer is observed. Aquation studies of cis-[PtCl2(15N-ipa)2] were precluded by the close proximity of the NH proton signal to the 1H2O resonance.  相似文献   

16.
Kinetics of the reaction of octacarbonyl dicobalt with ethyl diazoacetate leading to [μ2-{ethoxycarbonyl(methylene)}-μ2-(carbonyl)-bis(tricarbonyl-cobalt)] (Co-Co) (1), dinitrogen, and carbon monoxide were investigated at 10 °C in heptane solution. The initial rate of the reaction was measured by following both the gas evolution and the decrease of the octacarbonyl dicobalt concentration. The rate is first order with respect to octacarbonyl dicobalt and a complex order with respect to ethyl diazoacetate and carbon monoxide depending on the ratio of their concentrations. This is in accord with the formation of a heptacarbonyl dicobalt reactive intermediate (k1 (10 °C) = (1.22 ± 0.06) × 10−3 s−1) for which carbon monoxide and ethyl diazoacetate compete (k−1/k2 (10 °C) = 1.34 ± 0.07).  相似文献   

17.
To gain a better understanding of the light-induced reduction of protochlorophyllide (PChlide) to chlorophyllide as a key regulatory step in chlorophyll synthesis, we performed transient infrared absorption measurements on PChlide in d4-methanol. Excitation in the Q-band at 630 nm initiates dynamics characterized by three time constants: τ1 = 3.6 ± 0.2, τ2 = 38 ± 2, and τ3 = 215 ± 8 ps. As indicated by the C13′=O carbonyl stretching mode in the electronic ground state at 1686 cm−1, showing partial ground-state recovery, and in the excited electronic state at 1625 cm−1, showing excited-state decay, τ2 describes the formation of a state with a strong change in electronic structure, and τ3 represents the partial recovery of the PChlide electronic ground state. Furthermore, τ1 corresponds with vibrational energy relaxation. The observed kinetics strongly suggest a branched reaction scheme with a branching ratio of 0.5 for the path leading to the PChlide ground state on the 200 ps timescale and the path leading to a long-lived state (>>700 ps). The results clearly support a branched reaction scheme, as proposed previously, featuring the formation of an intramolecular charge transfer state with ∼25 ps, its decay into the PChlide ground state with 200 ps, and a parallel reaction path to the long-lived PChlide triplet state.  相似文献   

18.
Sickle hemoglobin forms long, multistranded polymers that account for the pathophysiology of the disease. The molecules in these polymers make significant contacts along the polymer axis (i.e., axial contacts) as well as making diagonally directed contacts (i.e., lateral contacts). The axial contacts do not engage the mutant β6 Val and its nonmutant receptor region on an adjacent molecule, in contrast to the lateral contacts which do involve the mutation site. We have studied the association process by elastic light scattering measurements as a function of temperature, concentration, and primary and quaternary structure, employing an instrument of our own construction. Even well below the solubility for polymer formation, we find a difference between the association behavior of deoxy sickle hemoglobin molecules (HbS), which can polymerize at higher concentration, in comparison to COHbS, COHbA, or deoxygenated Hemoglobin A (HbA), none of which have the capacity to form polymers. The nonpolymerizable species are all quite similar to one another, and show much less association than deoxy HbS. We conclude that axial contacts are significantly weaker than the lateral ones. All the associations are entropically favored, and enthalpically disfavored, typical of hydrophobic interactions. For nonpolymerizable Hemoglobin, ΔHo was 35 ± 4 kcal/mol, and ΔS was 102.7 ± 0.5 cal/(mol−K). For deoxyHbS, ΔHo was 19 ± 2 kcal/mol, and ΔS was 56.9 ± 0.5 cal/(mol−K). The results are quantitatively consistent with the thermodynamics of polymer assembly, suggesting that the dimer contacts and polymer contacts are very similar, and they explain a previously documented significant anisotropy between bending and torsional moduli. Unexpectedly, the results also imply that a substantial fraction of the hemoglobin has associated into dimeric species at physiological conditions.  相似文献   

19.
Summary Intracellular pH (pHi) regulation was studied in crayfish neurons with pH-, and Na+-sensitive microelectrodes. It was confirmed to involve both a HCO 3 -dependent and a HCO 3 -independent mechanism. The latter was identified as the amiloride-sensitive Na+/H+ exchange described in vertebrate cells. Its dependence on extracellular pH (pHe) and Na+ concentration ([Na+]e) was studied in CO2-free external solutions at 20°C. The steady state pHi and the rate constant (k) of the exponential pHi recovery following an acid load were determined. At pHe=7.5 and [Na+]e=200 mM, the average steady state pHi was 7.09±0.12 (as compared to 7.30±0.10 in the presence of 5 mM bicarbonate). The dependence of the rate constant of recovery on [Na+]e could be described by Michaelis-Menten kinetics; at pHe=7.5 the apparentK m andK max were 39 mM and 1.4 mmol·l–1·min–1, respectively. Decreasing pHe reduced the rate of recovery, the variations ofk with pHe conforming to a simple titration curve with an apparent pK of 7.05±0.21. These kinetic properties of the Na+/H+ exchange in crayfish neurons are similar to those described in vertebrate cells.Preliminary results were presented at the First International Congress of Comparative Physiology and Biochemistry (Liège, Belgium, 1984)  相似文献   

20.
Magnetic interactions in binuclear copper(II) complexes, [Cu2(apyhist)2Cl2](ClO4)2 (1) and [Cu2(2-pyhist)2Cl2](ClO4)2 (2) with tridentate diimine ligands and chloro-bridged groups (where apyhist=(4-imidazolyl)ethylene-2-amino-1-ethylpyridine and 2-pyhist=(4-imidazolyl)ethylene-2-aminomethylpyridine) were studied with the aim of better elucidating magneto-structural correlations in such species, both in solution and in solid state. X-ray analyses revealed that chloro-bridged ligands keep the copper(II) ion coordinated to adjacent unit, at Cu-Cl distances of 2.271 and 2.737 Å, and a Cu-Cl-Cu angle of 87.46° in compound 1. Each CuII atom is also coordinated to three N atoms from the imine ligand, in a distorted tetragonal pyramidal environment. Magnetic measurements carried out in temperatures from 0.8 to 290 K and in magnetic field up to 170 kOe indicated that besides the intramolecular magnetic coupling between the copper centers [J/k=−(1.93±0.05) K] further interactions between adjacent dimers [Jz/k=−(1.3±0.1) K] should be taken into account. Similar results were observed for compound 2, for which [J/k=−(4.27±0.05) K] and [Jz/k=−(3.7±0.1) K]. In solution, the interconversion of the dimer 1 and the related monomer species [Cu(apyhist)(H2O)2] (ClO4)2 (3) monitored by EPR spectra, was verified to be very dependent on the solvent.  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