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 15 毫秒
1.
 Dipeptides and tripeptides AcMet-aaH containing N-acetyl methionine, in which the group aaH is GlyH, AlaH, ValH, or Gly-GlyH, undergo hydrolytic cleavage of the Met-aaH peptide bond in the presence of the following complexes of palladium(II): cis-[Pd(en)(H2O)2]2+, cis-[Pd(tn)(H2O)2]2+, cis-[Pd(en)(CH3OH)2]2+, cis-[Pd(S,N-MetH)(H2O)2]2+, cis-[Pd(S,N-Met-GlyH)(H2O)2]2+, and cis-[Pd(S,N-Met-AlaH)(H2O)2]2+. These mononuclear complexes are precursors of binuclear palladium(II) complexes containing the substrates AcMet-aaH as bridging thioether ligands. The rate constant for cleavage is higher when the bidentate ligand in the precursor complex is ethylenediamine (which is completely displaced) than S,N-methionine (of which only the amino group is displaced), because the number of aqua ligands available for cleavage is greater in the former than in the latter case. The demonstrated dependence of the rate constant on the steric bulk (volume) of the leaving group, aaH, points the way toward achieving a degree of sequence selectivity in cleavage of peptide bonds by palladium(II) aqua complexes. One equivalent of cis-[Pd(en)(H2O)2]2+ cleaves as many as ten equivalents of AcMet-GlyH, but the rate constant decreases as the molar excess of the dipeptide over the catalyst increases. This demonstration of catalytic turnover points the way to our ultimate goal – artificial metallopeptidases. Received: 13 June 1997 / Accepted: 24 September 1997  相似文献   

2.
Six new dinuclear complexes, derived from cis-[Co(H2O)2(NH3)4]3+, cis-[Co(H2O)2(en)2]3+ and [M(CN)42? (M = Ni, Pd, Pt) were prepared and characterized by means of chemical analysis, electronic and IR measurements. The influence of the pH on the rate of the reaction was studied for the two derivatives of [Pd(CN)4]2?, showing that the best conditions to obtain the dinuclear compounds are at pH near 6, where the predominant species are cis-[Co(OH)(H2O)(amine)2]2+. The [Pt(CN)4]2? derivatives show PtPt interactions both in the solid state and in solution.  相似文献   

3.
 The palladium(II) aqua complexes [Pd(H2O)4]2+, cis-[Pd(en)(H2O)2]2+, and cis-[Pd(dtco-OH)(H2O)2]2+ effect hydrolytic cleavage of horse myoglobin in aqueous solution. The conditions were optimized with the third complex. Its structure was determined by X-ray crystallographic analysis of its precursor, the square-planar complex cis-[Pd(dtco-OH)Cl2], in which the chelating ligand adopts a boat-chair conformation. A weak interaction between the hydroxyl group and the palladium(II) atom seems to improve the stability of the reagent. The yield of cleavage after a 24-h incubation at 60  °C increases from 39% to 85% as the pH decreases from 6.2 to 3.2. The protein fragments are separated by SDS-PAGE electrophoresis and HPLC separation methods, and identified by ESIMS and MALDI-TOF mass spectrometric methods and by determination of terminal amino-acid sequences. Most of the 13 cleavage sites are clustered around the methionine, arginine, and some of the histidine residues, whose side chains can bind to palladium(II). Cleavage tends to occur at the peptide bonds one to three positions removed from the binding residues; the scissile bonds usually lie on the amino-terminal side, seldom on the carboxy-terminal side, of the binding residues. Removal of the heme and unfolding of the protein do not drastically alter the pattern of cleavage. The ability of palladium(II) aqua complexes to cleave proteins at relatively few sites, with explicable selectivity, with good to very good yield, and in weakly acidic and nearly neutral solutions, bodes well for their future use in biochemical and bioanalytical practice. Received: 23 December 1997 / Accepted: 14 April 1998  相似文献   

4.
Two dinuclear palladium(II) complexes, [{Pd(en)Cl}2(μ-pz)](NO3)2 and [{Pd(en)Cl}2(μ-pydz)](NO3)2, have been synthesized and characterized by elemental microanalysis and spectroscopic (1H and 13C NMR, IR and UV–vis) techniques (en is ethylenediamine; pz is pyrazine and pydz is pyridazine). The square planar geometry of palladium(II) metal centers in these complexes has been predicted by DFT calculations. The chlorido complexes were converted into the corresponding aqua complexes, [{Pd(en)(H2O)}2(μ-pz)]4+ and [{Pd(en)(H2O)}2(μ-pydz)]4+, and their reactions with N-acetylated l-histidylglycine (Ac–l–His–Gly) and l-methionylglycine (Ac–l–Met–Gly) were studied by 1H NMR spectroscopy. The palladium(II)-aqua complexes and dipeptides were reacted in 1:1 M ratio, and all reactions performed in the pH range 2.0 < pH < 2.5 in D2O solvent and at 37 °C. In the reactions of these complexes with Ac–l–His–Gly and Ac–l–Met–Gly dipeptides, the hydrolysis of the amide bonds involving the carboxylic group of both histidine and methionine amino acids occurs. The catalytic activities of the palladium(II)-aqua complexes were compared with those previously reported in the literature for the analogues platinum(II)-aqua complexes, [{Pt(en)(H2O)}2(μ-pz)]4+ and [{Pt(en)(H2O)}2(μ-pydz)]4+.  相似文献   

5.
The interaction of guanine, guanosine or 5-GMP (guanosine 5-monophosphate) with [Pd(en)(H2O)2](NO3)2 and [Pd(dapol)(H2O)2](NO3)2, where en is ethylenediamine and dapol is 2-hydroxy-1,3-propanediamine, were studied by UV-Vis, pH titration and 1H NMR. The pH titration data show that both N1 and N7 can coordinate to [Pd(en)(H2O)2]2+ or [Pd(dapol)(H2O)2]2+. The pKa of N1-H decreased to 3.7 upon coordination in guanosine and 5-GMP complexes, which is significantly lower than that of ∼9.3 in the free ligand. In strongly acidic solution where N1-H is still protonated, only N7 coordinates to the metal ion, but as the pH increases to pH ∼3, 1H NMR shows that both N7-only and N1-only coordinated species exist. At pH 4-5, both N1-only and N1,N7-bridged coordination to Pd(II) complexes are found for guanosine and 5-GMP. The latter form cyclic tetrameric complexes, [Pd(diamine)(μ-N1,N7-Guo]44+ and [Pd(diamine)(μ-N1,N7-5-GMP)]4Hx(4−x)−, (x=2,1, or 0) with either [Pd(en)(H2O)2](NO3)2 or [Pd(dapol)(H2O)2](NO3)2. The pH titration data and 1H NMR data agree well with the exception that the species distribution diagrams show the initial formation of the N1-only and N1,N7-bridged complexes to occur at somewhat higher pH than do the NMR data. This is due to a concentration difference in the two sets of data.  相似文献   

6.
Hydrothermal synthesis has afforded cobalt 5-substituted isophthalate complexes with 4,4′-dipyridylamine (dpa) ligands, showing different dimensionalities depending on the steric bulk and hydrogen-bonding facility of the substituent. [Co(tBuip)(dpa)(H2O)]n (1, tBuip = 5-tert-butylisophthalate) is a (4,4) grid two-dimensional coordination polymer featuring 2-fold parallel interpenetration. [Co(MeOip)2(Hdpa)2] (2, MeOip = 5-methoxyisophthalate) is organized into 3-fold parallel interpenetrated (4,4) grids through strong N-H+?O hydrogen bonding. {([Co(OHip)(dpa)(H2O)3])3·2H2O}n (3, OHip = 5-hydroxyisophthalate) possesses 1-D chain motifs. The 5-methyl derivative {[Co(mip)(dpa)]·3H2O}n (4, mip = 5-methylisophthalate) has a 3-D 658 cds topology. {[Co(H2O)4(Hdpa)2](nip)2·2H2O} (5, nip = 5-nitroisophthalate) and {[Co(sip)(Hdpa)(H2O)4]·2H2O} (6, sip = 5-sulfoisophthalate) are coordination complexes. Antiferromagnetic superexchange is observed in 1 and 4, with concomitant zero-field splitting. Thermal decomposition behavior of the higher dimensionality complexes is also discussed.  相似文献   

7.
The Pd(II) and Pt(II) complexes with triazolopyrimidine C-nucleosides L1 (5,7-dimethyl-3-(2′,3′,5′-tri-O-benzoyl-β-d-ribofuranosyl-s-triazolo)[4,3-a]pyrimidine), L2 (5,7-dimethyl-3-β-d-ribofuranosyl-s-triazolo[4,3-a]pyrimidine) and L3 (5,7-dimethyl[1,5-a]-s-triazolopyrimidine), [Pd(en)(L1)](NO3)2, [Pd(bpy)(L1)](NO3)2, cis-Pd(L3)2Cl2, [Pd2(L3)2Cl4] · H2O, cis-Pd(L2)2Cl2 and [Pt3(L1)2Cl6] were synthesized and characterized by elemental analysis and NMR spectroscopy. The structure of the [Pd2(L3)2Cl4] · H2O complex was established by X-ray crystallography. The two L3 ligands are found in a head to tail orientation, with a Pd?Pd distance of 3.1254(17) Å. L1 coordinates to Pd(II) through N8 and N1 forming polymeric structures. L2 coordinates to Pd(II) through N8 in acidic solutions (0.1 M HCl) forming complexes of cis-geometry. The Pd(II) coordination to L2 does not affect the sugar conformation probably due to the high stability of the C-C glycoside bond.  相似文献   

8.
A novel effect of the inhibition of the decomposition of amino acids to carbonates on addition of imidazole (HIm) to a reacting system containing equimolar amounts of copper and zinc metal powders, an amino acid [glycine (Hgly), aspartic acid (H2Asp) or glycylglycine (H2gg)] (1:1:2) and excess hydrogen peroxide (H2O2) resulting in formation of a mixed metal mixed ligand peroxo complex compound was observed, because in the absence of imidazole the corresponding reaction system yields only a mixed metal peroxo carbonate. For the resulting complex compounds, the homogeneity, i.e. [Cu(Zn)(O2 2–)(Gly)2(HIm)(H2O)], [Cu(Zn)(O2 2–)(Asp)(HIm)(H2O)2] or [Cu(Zn)2(O2 2–)2(gg)(HIm)(H2O)4], molecular formula, presence of peroxo group and coordination environment were established by combined physicochemical evidence from elemental and thermogravimetric analysis in air and argon atmospheres, electron spin resonance and electronic and IR spectral data. It is noteworthy to mention that the corresponding carboxylic acids of the above-mentioned amino acids, i.e. acetic and succinic acids, either do not decompose to carbonates in the absence of imidazole or form novel homogeneous peroxo mixed metal mixed ligand complex compounds as described above in the presence of imidazole. This suggests an important and significant mutual influence (in vitro) of biologically active chromophores like peroxo ions, imidazole and amino groups in the above-mentioned chemical reactions containing bioactive metals such as copper and zinc.  相似文献   

9.
In this work, a pair of new palladium(II) complexes, [Pd(Gly)(Phe)] and [Pd(Gly)(Tyr)], (where Gly is glycine, Phe is phenylalanine, and Tyr is tyrosine) were synthesized and characterized by UV–Vis, FT-IR, elemental analysis, 1H-NMR, and conductivity measurements. The detailed 1H NMR and infrared spectral studies of these Pd(II) complexes ascertain the mode of binding of amino acids to palladium through nitrogen of -NH2 and oxygen of -COO? groups as bidentate chelates. The Pd(II) complexes have been tested for in vitro cytotoxicity activities against cancer cell line of K562. Interactions of these Pd(II) complexes with CT-DNA and human serum albumin were identified through absorption/emission titrations and gel electrophoresis which indicated significant binding proficiency. The binding distance (r) between these synthesized complexes and HSA based on Forster?s theory of non-radiation energy transfer were calculated. Alterations of HSA secondary structure induced by complexes were confirmed by FT-IR measurements. The results of emission quenching at three temperatures have revealed that the quenching mechanism of these Pd(II) complexes with CT-DNA and HSA were the static and dynamic quenching mechanism, respectively. Binding constants (Kb), binding site number (n), and the corresponding thermodynamic parameters were calculated and revealed that the hydrogen binding and hydrophobic forces played a major role when Pd(II) complexes interacted with DNA and HSA, respectively. We bid that [Pd(Gly)(Phe)] and [Pd(Gly)(Tyr)] complexes exhibit the groove binding with CT-DNA and interact with the main binding pocket of HSA. The complexes follow the binding affinity order of [Pd(Gly)(Tyr)] > [Pd(Gly)(Phe)] with CT-DNA- and HSA-binding.  相似文献   

10.
An equimolecular mixture of [Pd(RNC)2Cl2] (R = Ph, p-Me C6H4) and [Pd(MeCN)2Cl2] reacts in boiling, 1,2-dichloroethane to give the binuclear complexes [Pd(RNC)Cl2]2.These compounds undergo a variety of bridge-splitting reactions with neutral or anionic ligands yielding complexes of the type cis and trans [Pd(RNC)LX2] or [Pd(RNC)X3] (L = PPh3, pyridine, C6H11NC; X = CL, Br).By reaction of [Pd(PhNC)Cl3] with MeOH the anionic carbene complex [Pd{C(NHPh)OMe}Cl3] is obtained.[Pd(PhNC)Cl2]2 reacts with p-toluidine (excess) or o-aminopyridine to give the corresponding mononuclear carbene derivatives.In the case of the mixed derivative [Pd(p-MeC6H4NC)(C6H11NC)Cl2], only the more activated p-tolylisocyanide was found to react with p-toluidine.The complexes have been characterized by elemental analysis, conductivity measurements, i.r. and 1H n.m.r. spectra where possible.  相似文献   

11.
Hydrothermal synthesis of orotic acid (H3L) with Ni(OAc)2·4H2O gives a green 1D co-ordinative network of composition [Ni(HL)(H2O)3] (3). The kinetic product [Ni(HL)·(H2O)4]H2O (4) can be prepared by conventional crystallisation. When boiled in water it is transformed into the thermodynamically favoured trihydrate 3. An unstable blue phase 5 that could not be characterised was also observed. Hydrothermal synthesis of orotic acid and M(OAc)2·4H2O (M=Ni, Co, Mn or Zn) and either 2,2-bipyridyl (bipy), 2,2-dipyridylamine (dpa), phenanthroline (phen), methyl-3-(2-pyridyl)pyrazole (pypz) or 2,9-dimethyl-1,10-phenanthroline (dmphen) gave infinite 1D co-ordinative networks of composition [M(HL)bipy(H2O)] (M=Co or Mn) (6-7) and complexes of composition [Ni(HL)bipy (H2O)2]2H2O (8); [Ni(HL)(dpa)(H2O)2]H2O (9); [Ni(HL)(phen)(H2O)2]·2H2O (10); [Ni(HL)(C9H9N3)(H2O)2]·2H2O (11); [Ni(HL)(dmphen)(H2O)] (12); [Zn(HL)bipy(H2O)] (13) and [Ni(HL)(dpa)2]·0.5H2O (14).  相似文献   

12.
Copper(II) cations coordinated with PMDTA (pentamethyldiethylenetriamine) and TMEDA (tetramethylethylenediamine) possess a high synthetic potential. The synthesis of these cations was carried out by metathesis reactions with silver salts. The cationic copper(II) complexes, [Cu(PMDTA)(Me2CO)Cl]+, [Cu(PMDTA)(H2O)Cl]+, [Cu(PMDTA)(DMF)]+, [Cu(PMDTA)Cl]+, [Cu(PMDTA)OAc]+, [Cu(PMDTA)(MeCN)2]2+, [Cu2(TMEDA)2Cl3]+ and [Cu(TMEDA)(MeCN)3]2+ were synthesised as PF6 salts, crystallised and characterised by single-crystal X-ray diffraction.  相似文献   

13.
《Inorganica chimica acta》1988,152(2):101-106
The interaction of potassium tetrachloropalladate(II) with theophylline in a 1:1 molar ratio resulted in the formation of the monotheophylline (K[Pd(ThH)Cl3]) or monotheophyllinato ([Pd(Th)Cl]2) complexes, depending on the solvent and the acidity conditions. In the first complex, theophylline coordinates to Pd(II) as a neutral molecule through its N9 atom, while in the second as a monoanion through both its N7 and O6 atoms. Both complexes react with nucleosides, giving the complexes [Pd(Nucl)(ThH)Cl2] and [Pd(Nucl)(Th)Cl], respectively. Those complexes with one N(1)H ionizable imino-proton undergo deprotonation and two new series of mixed ligand complexes, [Pd(Nucl − H+)(ThH)Cl] and [Pd(Nucl − H+(Th)] are formed. In the mixed ligand complexes, theophylline maintains its coordination modes. The nucleosides, on the other hand, exhibit their usual coordination sites; i.e. in the nondeprotonated complexes they coordinate only through their N7 atoms, while in the deprotonated they act as bidentate through both their N7 and O6 atoms. All complexes were characterized with elemental analyses, conductivity measurements and various spectroscopic techniques.  相似文献   

14.
Two novel, neutral and water soluble Pd(II) complexes of formula [Pd(Gly)(Ala)] (1) and [Pd(Gly)(Val)] (2) (Gly, Ala, and Val are anionic forms of glycine, alanine, and valine amino acids, respectively) have been synthesized and characterized by FT-IR, UV–Vis, 1H-NMR, elemental analysis, and molar conductivity measurement. The data revealed that each amino acid binds to Pd(II) through the nitrogen of –NH2 and the oxygen of –COO groups and acts as a bidentate chelate. These complexes have been assayed against leukemia cells (K562) using MTT method. The results indicated that both of the complexes display more cytotoxicity than the well-known anticancer drug, cisplatin. The interaction of the compounds with calf thymus DNA (CT-DNA) and human serum albumin (HSA) were assayed by a series of experimental techniques including electronic absorption, fluorescence, viscometry, gel electrophoresis, and FT-IR. The results indicated that the two complexes have interesting binding propensities toward CT-DNA as well as HSA and the binding affinity of (1) is more than (2). The fluorescence data indicated that both complexes strongly quench the fluorescence of ethidium bromide–DNA system as well as the intrinsic fluorescence of HSA via static quenching procedures. The thermodynamic parameters (ΔH°, ΔS°, and ΔG°) calculated from the fluorescence studies showed that hydrogen bonds and van der Waals interactions play a major role in the binding of the complexes to DNA and HSA. We suggest that both of the Pd(II) complexes exhibit the groove binding mode with CT-DNA and interact with the main binding pocket of HSA.

Communicated by Ramaswamy H. Sarma  相似文献   


15.
Recent ab initio studies reported in the literature have challenged the mechanistic assignments made on the basis of volume of activation data [1,2]. In addition to that ab initio molecular orbital calculations on hydrated zinc(II)-ions were used to elucidate the general role of this ion in metalloproteins [3]. Due to our interest in both inorganic reaction mechanisms and enzymatic catalysis we started a systematic investigation of solvent exchange processes on divalent zinc-ion using density functional calculations. Our investigations cover aqua complexes of the general form [Zn(H2O)n]2+·mH20 with n=3-6 and m=0-2, where n and m represent the number of water molecules in the coordination and solvation sphere, respectively.The complexes [Zn(H2O)5]2+·2H2O and [Zn(H2O)4]2+·2H2O turnend out to be the most stable zinc complexes with seven and six water molecules, respectively. This implies that a heptacoordinated zinc(II) complex, where all water molecules are located in the co-ordination sphere, should be energetically highly unfavorable and that [Zn(H2O)6]2+ can quite readily push two coordinated water molecules into the solvation sphere. For the pentaqua complex [Zn(H2O)5]2+ only one water molecule is easily lost to the solvation sphere, which makes the [Zn(H2O)4]2+·H2O complex the most favorable in order to consider the limiting dissociative and associative water exchange process of hexacoordinated zinc(II). The dehydration and hydration energies using the most stable zinc(II) complexes [Zn(H2O)4]2+·2H2O, [Zn(H2O)5]2+·2H2O and [Zn(H2O)4]2+·H2O were calculated to be 24.1 and -21.0 kcal/mol, respectively.  相似文献   

16.
A series of coordination polymers have been prepared by the combination of flexible ligand 1,1′-biphenyl-2,2′-dicarboxylic acid (H2dpa) and different types of nitrogen-containing ligands, with various metal ions such as Co(II), Zn(II) and Cd(II). The single-crystal structure analyses reveal that the above complexes possess different structure features with the introduction of different nitrogen-containing ligands. When auxiliary linear ligand 4,4′-bipyridine (4,4′-bpy) is introduced, two-dimensional layered complex, [Co2(dpa)2(4,4′-bpy)2(H2O)]n (1) is formed. Whereas if chelating ligand, 1,10-phenanthroline (1,10′-phen) and 2,2′-bipyridine (2,2′-bpy) are introduced, one-dimensional complex [Zn(dpa)(1,10′-phen)]n (2) and discrete complexes [Co2(dpa)2(2,2′-bpy)2(H2O)2] (3), [Co3(dpa)3(1,10′-phen)6(H2O)2] (4), [Cd(dpa)(1,10′-phen)2][(H2dpa)2(H2O)2] (5) are synthesized. To our interest, 1 and 2 crystallize in homochiral spacegroup. Furthermore, the magnetic property of complex 1 and the fluorescent properties of complexes 2 and 5 are studied.  相似文献   

17.
Abstract

Molecular modeling and energy minimisation calculations have been used to investigate the interaction of chromium(III) complexes in different ligand environments with various sequences of B-DNA. The complexes are [Cr(salen)(H2O)2]+; salen denotes 1, 2 bis-salicylideneaminoethane, [Cr(salprn)(H2O)2]+; salprn denotes 1, 3 bis- salicylideneamino-propane, [Cr(phen)3]3+; phen denotes 1, 10 phenanthroline and [Cr(en)3]3+; en denotes eth- ylenediamine. All the chromium(III) complexes are interacted with the minor groove and major groove of d(AT)12, d(CGCGAATTCGCG)2 and d(GC)12 sequences of DNA. The binding energy and hydrogen bond parameters of DNA-Cr complex adduct in both the groove have been determined using molecular mechanics approach. The binding energy and formation of hydrogen bonds between chromium(III) complex and DNA has shown that all complexes of chromium(III) prefer minor groove interaction as the favourable binding mode.  相似文献   

18.
An electron-rich iron(III) porphyrin complex (meso-tetramesitylporphinato)iron(III) chloride [Fe(TMP)Cl], was found to catalyze the epoxidation of olefins by aqueous 30% H2O2 when the reaction was carried out in the presence of 5-chloro-1-methylimidazole (5-Cl-1-MeIm) in aprotic solvent. Epoxides were the predominant products with trace amounts of allylic oxidation products, indicating that Fenton-type oxidation reactions were not involved in the olefin epoxidation reactions. cis-Stilbene was stereospecifically oxidized to cis-stilbene oxide without giving isomerized trans-stilbene oxide product, demonstrating that neither hydroperoxy radical (HOO·) nor oxoiron(IV) porphyrin [(TMP)FeIV=O] was responsible for the olefin epoxidations. We also found that the reactivities of other iron(III) porphyrin complexes such as (meso-tetrakis(2,6-dichlorophenyl)porphinato)iron(III) chloride [Fe(TDCPP)Cl], (meso-tetrakis(2,6-difluorophenyl)porphinato)iron(III) chloride [Fe(TDFPP)Cl], and (meso-tetrakis(pentafluorophenyl)porphinato)iron(III) chloride [Fe(TPFPP)Cl] were significantly affected by the presence of the imidazole in the epoxidation of olefins by H2O2. These iron porphyrin complexes did not yield cyclohexene oxide in the epoxidation of cyclohexene by H2O2 in the absence of 5-Cl-1-MeIm in aprotic solvent; however, addition of 5-Cl-1-MeIm to the reaction solutions gave high yields of cyclohexene oxide with the formation of trace amounts of allylic oxidation products. We proposed, on the basis of the results of mechanistic studies, that the role of the imidazole is to decelerate the O–O bond cleavage of an iron(III) hydroperoxide porphyrin (or H2O2–iron(III) porphyrin adduct) and that the intermediate transfers its oxygen to olefins prior to the O–O bond cleavage.  相似文献   

19.
Abstract

The 1H NMR relaxation effects produced by paramagnetic Cr(III) complexes on nucleoside 5′-mono- and -triphosphates in D2O solution at Ph′=3 were measured. The paramagnetic probes were [Cr(III)(H2O) 6]3+, [Cr(III)(H2O)3 (HATP)], [Cr(III)(H2O)3(HCTP)] and [Cr(III) (H2O)3(UTP)?, while the matrix nucleotides (0.1 M) were H2AMP, HIMP?, and H2ATP2-. For the aromatic base protons, the ratios of the transverse to longitudinal paramagnetic relaxation rates (R2p/R1p) for the [Cr(III)(H2O)6]3+/H2ATP2-, [Cr(III)(H2O)3(HATP)]/H2ATP2-, [Cr(III)(H2O)3(HCTP)]/H2ATP2 and [Cr(III)(H2O)3(UTP)]?/H2ATP2 systems were below 2.33 so the dipolar term predominates. For a given nucleotide, R1p for the purine H(8) signal was larger than for the H(2) signal with the [Cr(III)(H2O)6]3+ probe, while R1p for the H(2) signal was larger with all the other Cr(III) probes. Molecular mechanics computations on the [Cr(III)(H2O)4(HPP)(α,β)], [Cr(III)(NH3)4(HPP)(α,β)], [Co(III)(NH3)3(H2PPP)(α,βγ)] and [Co(III)(NH3)4(HPP)(α,β)] complexes gave calculated energy-minimized geometries in good agreement with those reported in crystal structures. The molecular mechanics force constants found were then used to calculate the geometry of the inner sphere [Cr(III)(H2O)6]3+ and [Cr(III)(H2O)3(HATP)(α,βγ)] complexes as well as the structures of the outer sphere [Cr(III) (H2O)6]3+-(H2AMP) and [Cr(III)(H2O)6]-(HIMP)? species. The gas-phase structure of the [Cr(III)(H2O)3(HATP)(α,βγ)] complex shows the existence of a hydrogen bond interaction between a water ligand and the adenine N(7) (O…N = 2.82 Å). The structure is also stabilized by intramolecular hydrogen bonds involving the -O(2′)H group and the adenine N(3) (O…N = 2.80 Å) as well as phosphate oxygen atoms and a water molecule (O…O = 2.47 Å). The metal center has an almost regular octahedral coordination geometry.

The structures of the two outer-sphere species reveal that the phosphate group interacts strongly with the hexa-aquochromium probe. In both complexes, the nucleotides have a similar “anti” conformation around the N(9)-C(l′) glycosidic bond. However, a very important difference characterizes the two structures. For the (HIMP)? complex, strong hydrogen bond interactions exist between one and two water ligands and the inosine N(7) and O(6) atoms, respectively (O…O = 2.63 Å O…N = 2.72, 2.70 Å). For the H2AMP complex, the [Cr(III) (H2O)c]3+ cation does not interact with N(7) since it is far from the purine system. Hydrogen bonds occur between water ligands and phosphate oxygens. The Cr-H(8) and Cr-H(2) distances revealed by the energy-minimized geometries for the two outer sphere species were used to calculate the R1p values for the H(8) and H(2) signals for comparison with the observed R1p values: 0.92(c), 1.04(ob) (H(8)) and 0.06(c), 0.35(ob) (H(2)) for H2AMP; and 3.76(c), 4.53(ob) (H(8)) and 0.16(c), 0.77(ob) s?1 (H(2)) for HIMP?. These results suggest that the dynamic relaxation effects can be only partially understood with molecular mechanics computations, although the success of the geometry calculations suggests that future efforts in the development of computational methods are justified.  相似文献   

20.
The ternary complex [Cu(5′-IMP)(dpa)(H2O)]2 has been prepared and its structure analyzed by x-ray diffraction. It has a dimeric structure in which the 5′-IMP ligands coordinate solely through their phosphate groups. This geometry is in marked contrast to that of another Cu5′-IMP ternary complex, [Cu(5′-IMPH)(bipy)(H2O)2]+, which shows metal binding through the purine base rather than the phosphate group.  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