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.
A kinetic study is presented in which the effect of Zn(II) on yeast inorganic pyrophosphatase was quantitatively determined. A dual role model for metal ion effect, previously determined for the Mg(II)-pyrophosphatase system (O. A. Moe and L. G. Butler, 1972, J. Biol. Chem.247, 7308–7315), was applied successfully to the analysis of the kinetics for Zn(II)-pyrophosphatase and Zn(II), Mg(II)-pyrophosphatase systems. The model, assigning an activator role to free Zn(II) ion and a substrate role to the Zn(II)-pyrophosphate complex, gave an excellent fit to the data. Inhibition of the Mg(II)-pyrophosphatase system by Zn(II) was analyzed by a model in which competitive binding of the Mg(II)-pyrophosphate and Zn(II)-pyrophosphate complexes occurred at the enzyme active site, with both complexes undergoing reaction at different rates. Relative maximal velocities and enzymeligand dissociation constants for the Zn(II)-pyrophosphate complex were determined for the cases where the metal ion activator role was fulfilled by Zn(II) and Mg(II), respectively. The maximal velocity parameter showed a dependence on the nature of the activator metal ion, demonstrating that the role of the latter is associated both with the process of substrate binding and with the mechanism of catalysis. Values for all kinetic parameters are reported for an ionic strength of 0.2, pH 7.0, and 25.0 °C.  相似文献   

2.
A quantitative model for the slow reversible hydrophobic deactivation of alpha-chymotrypsin (alpha-CT) is proposed. Kinetic results are obtained for (1) the situation in which the inhibitor concentration, although remaining constant during the course of a run, can be varied independently of the concentration of nonself-inhibiting substrate, and for (2) the situation in which the self-inhibiting substrate concentration decreases during the course of a run, and independent variation of inhibitor and substrate concentrations is not possible. Excellent quantitative agreement between theory and experiment is obtained for a wide range of conditions using 3-(n-hexanoyl-O-benzoate (with dodecylsulfate as the inhibitor), and 3-(n-decanoyl)-O-benzoate as the self-inhibiting substrate. Activation enthalpies and entropies for the hydrophobic deactivation of alpha-CT by dodecylsulfate and tetradecyltrimethylammonium are determined. For comparison, activation enthalpies and entropies for the alpha-CT hydrolysis of 3-(n-heptanoyl)-O-benzoate are determined; evidence for a thermally induced conformational transition in alpha-CT at 30 degrees C is obtained.  相似文献   

3.
The interactions of phytic acid with Cu(II) and Zn(II) ions were examined as functions of metal ion concentrations and pH. Cu(II) ion-selective potentiometric and electron spin resonance (ESR) experiments provide strong evidence for the binding of Cu(II) ions to the phytic acid molecule at low pH (2.4–3.4) values. The relative stabilities of the copper and zinc phytates at low pH values were found to be very similar. For systems with metal ion:phytic acid molar ratios of 1:1–4:1 and 5:1–6:1 and pH values in the 3.4–5.9 and 3.4–5.0 ranges, respectively, Zn(II) ions were found to form complexes with phytic acid that were more stable than those of Cu(II) ions with phytic acid. The phytic acid molecule, however, was found to accommodate Cu(II) ions more readily than Zn(II) ions. For example, in systems containing equal amounts of Cu(II) and Zn(II) ions, 2 Zn(II) ions and 2, 3, 4, or 4.5 Cu(II) ions were found per phytic acid molecule depending upon metal ion:phytic acid molar ratios in the systems and pH. Total metal ion:phytic acid molar ratios and pH affected resultant metal ion solubilities and were factors influencing the effects of Zn(II) and Cu(II) ions on the binding of each other by phytic acid. Zn(II) and Cu(II) ions were observed to potentiate the binding of each other by phytic acid in some systems and compete with each other for phytate binding sites in others.  相似文献   

4.
Zn(II) metabolism in prokaryotes   总被引:1,自引:0,他引:1  
It is difficult to over-state the importance of Zn(II) in biology. It is a ubiquitous essential metal ion and plays a role in catalysis, protein structure and perhaps as a signal molecule, in organisms from all three kingdoms. Of necessity, organisms have evolved to optimise the intracellular availability of Zn(II) despite the extracellular milieu. To this end, prokaryotes contain a range of Zn(II) import, Zn(II) export and/or binding proteins, some of which utilise either ATP or the chemiosmotic potential to drive the movement of Zn(II) across the cytosolic membrane, together with proteins that facilitate the diffusion of this ion across either the outer or inner membranes of prokaryotes. This review seeks to give an overview of the systems currently classified as altering Zn(II) availability in prokaryotes.  相似文献   

5.
Rat matrix-induced alkaline phosphatase is an enzyme which requires magnesium and zinc ions for its maximal activity. Two Zn(II) ions and one Mg(II) ion are bound to each subunit of native dimeric enzyme. The presence of magnesium ion (10-100 microM) or zinc ion (7-20 nM) alone is sufficient to stimulate apoenzyme activity. However maximal activity (264 U/mg) requires the presence of both ions. Binding of Zn(II) ions to the Mg(II) binding site causes a strong inhibition of the apoenzyme while the binding of Mg(II) on Zn(II) binding site is not sufficient to stimulate PNPPase activity of the apoenzyme. Binding of both ions to the enzyme molecule did not change the apparent dissociation constant for PNPP hydrolysis.  相似文献   

6.
Treatment of Escherichia coli dihydroorotase (a homodimer of subunit molecular weight 38,729) containing only the 1 active site Zn(II) ion per subunit with the sulfhydryl reagent N-(ethyl)-maleimide (NEM) blocks the two external Zn(II) sites per subunit and dramatically lessens the precipitation caused by high concentrations of Zn(II); stabilizes the enzyme partially against air oxidation and dilution inactivation; makes the active site Zn(II) easier to remove; and lowers Km and increases kcat. Treatment of NEM-blocked dihydroorotase ((NEM)dihydroorotase) with the chelator 2,6-pyridinedicarboxylic acid at pH 5.0 in the absence of oxygen and trace metal ions removes the active site Zn(II) with a half-life of 15 min, allowing the production of milligram amounts of moderately stable apo-(NEM)dihydroorotase in about 80% yield. Treatment of apo-(NEM)dihydroorotase with Co(II) at pH 7.0 produces (NEM)dihydroorotase completely substituted at the active site with Co(II) in 100% yield: analysis gives 0.95-1.1 g atoms of Co(II) per active site and 0.03-0.05 g atoms of Zn(II) per active site. This Co(II)-(NEM)dihydroorotase is hyperactive at pH 8. The electronic absorption spectrum of Co(II)-(NEM)dihydroorotase at pH 6.5 implicates an active site thiol group as a ligand to the metal ion. The spectrum is inconsistent with tetrahedral coordination of the active site metal ion and is most consistent with a pentacoordinate structure.  相似文献   

7.
Mixed aquo-N-methylimidazole complexes of Co(II) have been studied as a function of pH to gain a fuller understanding of the metal-binding site in Co(II)-carbonic anhydrase. The inherent affinity of N-methylimidazole for Co(II) has been calculated along with a species distribution for the stepwise addition of ligand to the metal ion. From these studies, it is apparent that the occurrence of Zn(II) rather than Co(II) in native carbonic anhydrase can be explained by the stronger affinity of Zn(II) for imidazole and the preference of Zn(II) for a tetrahedral geometry as offered by the enzyme. Octahedral Co(II) fails to ionize metal bound water. However, at high pH, Co(II)-N-methylimidazole complexes interact directly with the hydroxide ion, generating species with visible spectra very similar to that of Co(II)-carbonic anhydrase. Tentative structures have been proposed for these species.  相似文献   

8.
Metallo-β-lactamases (MβLs) stand as one of the main mechanisms of bacterial resistance toward carbapenems. The rational design of an inhibitor for MβLs has been limited by an incomplete knowledge of their catalytic mechanism and by the structural diversity of their active sites. Here we show that the MβL GOB from Elizabethkingia meningoseptica is active as a monometallic enzyme by using different divalent transition metal ions as surrogates of the native Zn(II) ion. Of the metal derivatives in which Zn(II) is replaced, Co(II) and Cd(II) give rise to the most active enzymes and are shown to occupy the same binding site as the native ion. However, Zn(II) is the only metal ion capable of stabilizing an anionic intermediate that accumulates during nitrocefin hydrolysis, in which the C–N bond has already been cleaved. This finding demonstrates that the catalytic role of the metal ion in GOB is to stabilize the formation of this intermediate prior to nitrogen protonation. This role may be general to all MβLs, whereas nucleophile activation by a Zn(II) ion is not a conserved mechanistic feature.  相似文献   

9.
Silvaggi NR  Wilson D  Tzipori S  Allen KN 《Biochemistry》2008,47(21):5736-5745
The Clostridium botulinum neurotoxin serotype A light chain (BoNT/A-LC) is a Zn(II)-dependent metalloprotease that blocks the release of acetylcholine at the neuromuscular junction by cleaving SNAP-25, one of the SNARE proteins required for exocytosis. Because of the potential for use of the toxin in bioterrorism and the increasingly widespread application of the toxin in the medical field, there is significant interest in the development of small-molecule inhibitors of the metalloprotease. Efforts to design such inhibitors have not benefited from knowledge of how peptides bind to the active site since the enzyme-peptide structures available previously either were not occupied in the vicinity of the catalytic Zn(II) ion or did not represent the product of SNAP-25 substrate cleavage. Herein we report the 1.4 A-resolution X-ray crystal structure of a complex between the BoNT/A-LC and the inhibitory peptide N-Ac-CRATKML, the first structure of the light chain with an inhibitory peptide bound at the catalytic Zn(II) ion. The peptide is bound with the Cys S gamma atom coordinating the metal ion. Surprisingly, the cysteine sulfur is oxidized to the sulfenic acid form. Given the unstable nature of this species in solution, is it likely that oxidation occurs on the enzyme. In addition to the peptide-bound structure, we report two structures of the unliganded light chain with and without the Zn(II) cofactor bound at 1.25 and 1.20 A resolution, respectively. The two structures are nearly identical, confirming that the Zn(II) ion plays a purely catalytic role. Additionally, the structure of the Zn(II)-bound uncomplexed enzyme allows identification of the catalytic water molecule and a second water molecule that occupies the same position as the peptidic oxygen in the tetrahedral intermediate. This observation suggests that the enzyme active site is prearranged to stabilize the tetrahedral intermediate of the protease reaction.  相似文献   

10.
Characterization of an inhibitory metal binding site in carboxypeptidase A   总被引:6,自引:0,他引:6  
K S Larsen  D S Auld 《Biochemistry》1991,30(10):2613-2618
The specificity of metal ion inhibition of bovine carboxypeptidase A ([(CPD)Zn]) catalysis is examined under stopped-flow conditions with use of the fluorescent peptide substrate Dns-Gly-Ala-Phe. The enzyme is inhibited competitively by Zn(II), Pb(II), and Cd(II) with apparent KI values of 2.4 x 10(-5), 4.8 x 10(-5), and 1.1 x 10(-2) M in 0.5 M NaCl at pH 7.5 and 25 degrees C. The kcat/Km value, 7.3 x 10(6) M-1 s-1, is affected less than 10% at 1 x 10(-4) M Mn(II) or Cu(II) and at 1 x 10(-2) M Co(II), Ni(II), Hg(II), or Pt(IV). Zn(II) and Pb(II) are mutually exclusive inhibitors. Previous studies of the pH dependence of Zn(II) inhibition [Larsen, K. S., & Auld, D. S. (1989) Biochemistry 28, 9620] indicated that [(CPD)Zn] is selectively inhibited by a zinc monohydroxide complex, ZnOH+, and that ionization of a ligand, LH, in the enzyme's inhibitory site (pKLH 5.8) is obligatory for its binding. The present study allows further definition of this inhibitory zinc site. The ionizable ligand (LH) is assigned to Glu-270, since specific chemical modification of this residue decreases the binding affinity of [(CPD)Zn] for Zn(II) and Pb(II) by more than 60- and 200-fold, respectively. A bridging interaction between the Glu-270-coordinated metal hydroxide and the catalytic metal ion is implicated from the ability of Zn(II) and Pb(II) to induce a perturbation in the electronic absorption spectrum of cobalt carboxypeptidase A ([(CPD)Co]).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
Fundamental issues in zinc biology are how proteins control the concentrations of free Zn(II) ions and how tightly they interact with them. Since, basically, the Zn(II) stability constants of only two cytosolic zinc enzymes, carbonic anhydrase and superoxide dismutase, have been reported, the affinity for Zn(II) of another zinc enzyme, sorbitol dehydrogenase (SDH), was determined. Its log K is 11.2 +/- 0.1, which is similar to the log K values of carbonic anhydrase and superoxide dismutase despite considerable differences in the coordination environments of Zn(II) in these enzymes. Protein tyrosine phosphatase 1B (PTP 1B), on the other hand, is not classified as a zinc enzyme but is strongly inhibited by Zn(II), with log K = 7.8 +/- 0.1. In order to test whether or not metallothionein (MT) can serve as a source for Zn(II) ions, it was used to control free Zn(II) ion concentrations. MT makes Zn(II) available for both PTP 1B and the apoform of SDH. However, whether or not Zn(II) ions are indeed available for interaction with these enzymes depends on the thionein (T) to MT ratio and the redox poise. At ratios [T/(MT + T) = 0.08-0.31] prevailing in tissues and cells, picomolar concentrations of free Zn(II) are available from MT for reconstituting apoenzymes with Zn(II). Under conditions of decreased ratios, nanomolar concentrations of free Zn(II) become available and affect enzymes that are not zinc metalloenzymes. The match between the Zn(II) buffering capacity of MT and the Zn(II) affinity of proteins suggests a function of MT in controlling cellular Zn(II) availability.  相似文献   

12.
A multiphase model of metal ion species in human interstitial fluid was constructed under physiological conditions. The effect of Pr(III) on Zn(II) species was studied. At the normal conditions, Zn(II) species mainly distribute in [Zn(HSA)], [Zn(IgG)], and [Zn(Cys)2H]+. With the Pr(III) level increased, the apparent competition of Pr(III) for ligands lead to the redistribution of Zn(II) species.  相似文献   

13.
The water hyacinth (Eichchornia crassipes) has been successfully utilized for the removal of Zn(II) and Cd(II) as well as their admixture from samples of aqueous solutions. The growth of the plant after 16 days of exposure to the metal ions showed an increasing trend up to 2.5 ppm of Cd(II) and 6.0 ppm of Zn(II) concentrations, however, the growth became nondetectable or inhibited above these concentrations. The overall metal uptake by the plant was dependent upon the concentration of the metal and the duration of the exposure time. The metal uptake from a mixture of Cd(II) and Zn(II) was reflected by a rate constant quite different from those solutions containing only one metal ion. An analysis of metal in roots and tops of the plants showed that more Zn(II) was accumulated in the root when compared to Cd(II). However, the accumulation factor for the tops and the roots for Cd(II) and Zn(II) was higher than those obtained admixture of Zn(II) and Cd(II). The rate of metal mobility in the root was slower than that in the top of the plant for Zn(II) and Cd(II). A water hyacinth based system can be used to remove Cd(II) and Zn(II) from water/wastewater.  相似文献   

14.
The association of imidazole and auramine O to native horse-liver alcohol dehydrogenase [Zn(II)LADH] and active-site specifically cobalt(II)-substituted horse-liver alcohol dehydrogenase [Co(II)LADH], respectively, has been investigated. In all cases [except imidazole binding to Zn(II)LADH in the presence of auramine O] the association rates approached an upper limit (kmax). The kmax values were compared for the metal ligands imidazole (monodentate), 1,10-phenanthroline and 2,2'-bipyridine (bidentate; see also the preceding paper), and for auramine O which does not coordinate to the catalytic metal ion. Independent of the large differences in their structure and metal-bonding capability, all these compounds exhibit common, maximum, limiting rate constants of about 60 s-1 and 200 s-1 for Co(II)LADH and Zn(II)LADH, respectively. These results demonstrate that kmax is strongly dependent on the catalytic metal ion but not on the ligand. The absence of spectral changes in the d-d transitions of the catalytic Co(II) ion upon auramine O binding to Co(II)LADH indicates that the rate-limiting step is not accompanied by a major conformational change. Finally, it is concluded that reactions in the inner coordination sphere of the catalytic metal ion (i.e. the metal-bound water molecule) are not responsible for the step characterized by kmax. We propose the rate-limiting step to consist of the dissociation of one or several water molecules from the second coordination sphere of the catalytic metal ion in the active site of LADH in its open conformation.  相似文献   

15.
Cd-substituted forms of the Bacillus cereus metallo-beta-lactamases (BCII) were studied by perturbed angular correlation of gamma-rays (PAC) spectroscopy. At very low [Cd]:[apo-beta-lactamase] ratios, two nuclear quadrupole interactions (NQI) were detected. For [Cd]:[apo-beta-lactamase] ratios between 0.8 and 3.0, two new NQIs appear, and the spectra show that up to 2 cadmium ions can be bound per molecule of apoenzyme. These results show the existence of two interacting Cd-binding sites in BCII. The relative populations of the two NQIs found at low [Cd]:[apo-beta-lactamase] ratios yielded a 1:3 ratio for the microscopic dissociation constants of the two different metal sites (when only one cadmium ion is bound). X-ray diffraction data at pH 7.5 demonstrate that also for Zn(II) two binding sites exist, which may be bridged by a solvent molecule. The measured NQIs could be assigned to the site with three histidines as metal ligands (three-His site) and to the site with histidine, cysteine, and aspartic acid as metal ligands (Cys site), respectively, by PAC measurements on the Cys168Ala mutant enzyme. This assignment shows that cadmium ions preferentially bind to the Cys site. This is in contrast to the preference of Zn(II) in the hybrid Zn(II)Cd(II) enzyme, where an analysis of the corresponding PAC spectrum showed that Cd(II) occupied the Cys site, whereby Zn(II) occupied the site with three histidines. The difference between Zn(II) and Cd(II) in affinity for the two sites is combined with the kinetics of hydrolysis of nitrocefin for different metal ion substitutions (Zn(2)E, ZnE, Cd(2)E, CdE, and ZnCdE) to study the function of the two metal ion binding sites.  相似文献   

16.
We previously reported the IZ-3adH peptide, which formed a triple-stranded coiled-coil after binding Ni(II), Cu(II), or Zn(II). In this paper, we report the peptide, IZ-3aH, having a new metal binding specificity. The IZ-3aH peptide was found to bind Cu(II) and Zn(II) and form a triple-stranded coiled-coil. However, it did not bind Ni(II). Metal ion titrations monitored by circular dichroism revealed that the dissociation constants, K(d) were 9 microm for Zn(II) and 10 microm for Cu(II). The bound Cu(II) ion has a planar tetragonal geometry, where the coordination positions are three nitrogens of the His residues and one H(2)O.  相似文献   

17.
The core protein of glypican-1, a glycosylphosphatidylinositol-linked heparan sulfate proteoglycan, can bind Cu(II) or Zn(II) ions and undergo S-nitrosylation in the presence of nitric oxide. Cu(II)-to-Cu(I)-reduction supports extensive and permanent nitrosothiol formation, whereas Zn(II) ions appear to support a more limited, possibly transient one. Ascorbate induces release of nitric oxide, which catalyzes deaminative degradation of the heparan sulfate chains on the same core protein. Although free Zn(II) ions support a more limited degradation, Cu(II) ions support a more extensive self-pruning process. Here, we have investigated processing of glypican-1 in rat C6 glioma cells and the possible participation of the copper-containing glycosylphosphatidylinositol-linked splice variant of ceruloplasmin in nitrosothiol formation. Confocal microscopy demonstrated colocalization of glypican-1 and ceruloplasmin in endosomal compartments. Ascorbate induced extensive, Zn(II)-supported heparan sulfate degradation, which could be demonstrated using a specific zinc probe. RNA interference silencing of ceruloplasmin expression reduced the extent of Zn(II)-supported degradation. In cell-free experiments, the presence of free Zn(II) ions prevented free Cu(II) ion from binding to glypican-1 and precluded extensive heparan sulfate autodegradation. However, in the presence of Cu(II)-loaded ceruloplasmin, heparan sulfate in Zn(II)-loaded glypican-1 underwent extensive, ascorbate-induced degradation. We propose that the Cu(II)-to-Cu(I)-reduction that is required for S-nitrosylation of glypican-1 can take place on ceruloplasmin and thereby ensure extensive glypican-1 processing in the presence of free Zn(II) ions.  相似文献   

18.
Protein stability remains one of the main factors limiting the realization of the full potential of protein therapeutics. Poly(ethylene glycol) (PEG) conjugation to proteins has evolved into an important tool to overcome instability issues associated with proteins. The observed increase in thermodynamic stability of several proteins upon PEGylation has been hypothesized to arise from reduced protein structural dynamics, although experimental evidence for this hypothesis is currently missing. To test this hypothesis, the model protein alpha-chymotrypsin (alpha-CT) was covalently modified with PEGs with molecular weights (M(W)) of 700, 2,000 and 5,000 and the degree of modification was systematically varied. The procedure did not cause significant tertiary structure changes. Thermodynamic unfolding experiments revealed that PEGylation increased the thermal transition temperature (T(m)) of alpha-CT by up to 6 degrees C and the free energy of unfolding [DeltaG(U) (25 degrees C)] by up to 5 kcal/mol. The increase in stability was found to be independent of the PEG M(W) and it leveled off after an average of four PEG molecules were bound to alpha-CT. Fourier-transformed infrared (FTIR) H/D exchange experiments were conducted to characterize the conformational dynamics of the PEG-conjugates. It was found that the magnitude of thermodynamic stabilization correlates with a reduction in protein structural dynamics and was independent of the PEG M(W). Thus, the initial hypothesis proved positive. Similar to the thermodynamic stabilization of proteins by covalent modification with glycans, PEG thermodynamically stabilizes alpha-CT by reducing protein structural dynamics. These results provide guidance for the future development of stable protein formulations.  相似文献   

19.
The reaction of L-ascorbic acid with the zinc group and manganese ions has been investigated in aqueous solution at pH 6-7. The solid salts of the type M (L-ascorbate)2.2H2O, where M = Zn(II), Cd(II) and Mn(II) were isolated and characterized by 13C NMR and Fourier Transform infrared (FT-IR) spectroscopy. Spectroscopic evidence showed that in aqueous solution, the bindings of the Zn(II) and Mn(II) ions are through the ascorbate anion O-3 and O(2)-H groups (chelation), while the Cd(II) ion binding is via the O-3 atom only. In the solid state, the binding of these metal ions would be through two acid anions via O-3, O-2 of the first and O-1, O-3 of the second anion as well as to two H2O molecules, resulting in a six-coordinated metal ion. The Hg(II) ion interaction leads to the oxidation of the ascorbic acid in aqueous solution.  相似文献   

20.
The synthesis, characterization and biological activity of the first zinc(II) complexes with potent inhibitors of cyclin-dependent kinases (CDKs) derived from 6-benzylaminopurine are described. Based on the results following from elemental analyses, infrared, NMR and ES+MS (electrospray mass spectra in the positive ion mode) spectroscopies, conductivity data, thermal analysis and X-ray structures, the tetrahedral Zn(II) complexes of the compositions [Zn(Olo)Cl(2)](n) (1), [Zn(iprOlo)Cl(2)](n) (2), [Zn(BohH(+))Cl(3)] x H(2)O (3) and [Zn(iprOloH(+))Cl(3)] x H(2)O (4) have been prepared, where Olo=2-(2-hydroxyethylamino)-6-benzylamino-9-methylpurine (Olomoucine), iprOlo=2-(2-hydroxyethylamino)-6-benzylamino-9-isopropylpurine (i-propyl-Olomoucine), Boh=2-(3-hydroxypropylamino)-6-benzylamino-9-isopropylpurine (Bohemine). The 1D-polymeric chain structure for [Zn(Olo)Cl(2)](n) (1) as well as the monomeric one for [Zn(BohH(+))Cl(3)] x H(2)O (3) and [Zn(iprOloH(+))Cl(3)] x H(2)O (4) have been revealed unambiguously by single crystal X-ray analyses. The 1D-polymeric chain of 1 consists of Zn(Olo)Cl(2) monomeric units in which the Zn(II) ion is coordinated by two chlorine atoms and one oxygen atom of the 2-hydroxyethylamino group of Olomoucine. The next monomeric unit is bonded to Zn(II) through the N7 atom of a purine ring. Thus, each of Zn(II) ions is tetrahedrally coordinated and a ZnCl(2)NO chromophore occurs in the complex 1. The complexes 3 and 4 are mononuclear species with a distorted tetrahedral arrangement of donor atoms around the Zn(II) ion with a ZnCl(3)N chromophore. The corresponding CDK inhibitor, i.e., both Boh and iprOlo, is coordinated to Zn(II) via the N7 atom of the purine ring in 3 and 4. The cytotoxicity of the zinc(II) complexes against human melanoma, sarcoma, leukaemia and carcinoma cell lines has been determined as well as the inhibition of the CDK2/cyclin E kinase. A relationship between the structure and biological activity of the complexes is also discussed.  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