Co-reporter:Debabrata Chatterjee, Namita Jaiswal, Alicja Franke and Rudi van Eldik
Chemical Communications 2014 vol. 50(Issue 93) pp:14562-14565
Publication Date(Web):07 Oct 2014
DOI:10.1039/C4CC06631F
Reported is the first example of a ruthenium(III) complex, RuIII(edta) (edta4− = ethylenediaminetetraacetate), that catalyzes the disproportion of H2O2 to O2 and water in resemblance to catalase activity, and shedding light on the possible mechanism of action of the [RuV(edta)(O)]− formed in the reacting system.
Co-reporter:Maria Oszajca, Alicja Franke, Agnieszka Drzewiecka-Matuszek, Małgorzata Brindell, Grażyna Stochel, and Rudi van Eldik
Inorganic Chemistry 2014 Volume 53(Issue 6) pp:2848-2857
Publication Date(Web):January 6, 2014
DOI:10.1021/ic402567h
The presented results cover a comparative mechanistic study on the reactivity of compound (Cpd) I and II mimics of a water-soluble iron(III) porphyrin, [meso-tetrakis(2,4,6-trimethyl-3-sulfonatophenyl)porphinato]iron(III), FeIII(TMPS). The acidity of the aqueous medium strongly controls the chemical nature and stability of the high-valent iron(IV) oxo species. Reactivity studies were performed at pH 5 and 10, where the Cpd I and II mimics are stabilized as the sole oxidizing species, respectively. The contributions of ΔH⧧ and ΔS⧧ to the free energy of activation (ΔG⧧) for the oxidation of 4-methoxybenzaldehyde (4-MB-ald), 4-methoxybenzyl alcohol (4-MB-alc), and 1-phenylethanol (1-PhEtOH) by the Cpd I and II mimics were determined. The relatively large contribution of the ΔH⧧ term in comparison to the −TΔS⧧ term to ΔG⧧ for reactions involving the Cpd II mimic indicates that the oxidation of selected substrates by this oxidizing species is clearly an enthalpy-controlled process. In contrast, different results were found for reactions with application of the Cpd I mimic. Depending on the nature of the substrate, the reaction at room temperature can be entropy-controlled, as found for the oxidation of 4-MB-alc, or enthalpy-controlled, as found for 1-PhEtOH. Importantly, for the first time, activation volumes (ΔV⧧) for the oxidation of selected substrates by both reactive intermediates could be determined. Positive values of ΔV⧧ were found for reactions with the Cpd II mimic and slightly negative ones for reactions with the Cpd II mimic. The results are discussed in the context of the oxidation mechanism conducted by the Cpd I and II mimics.
Co-reporter:Mirjana D. Đurović, Živadin D. Bugarčić, Frank W. Heinemann and Rudi van Eldik
Dalton Transactions 2014 vol. 43(Issue 10) pp:3911-3921
Publication Date(Web):21 Jan 2014
DOI:10.1039/C3DT53140F
The influence of tridentate, nitrogen donor ligands, on the stability of gold(III) complexes under physiological conditions was investigated. The interaction of [Au(terpy)Cl]2+ (terpy = 2,2′:6′2′′ terpyridine), [Au(bpma)Cl]2+ (bpma = bis(pyridyl-methyl)amine), [Au(dien)Cl]2+ (dien = diethylenetriamine) and [AuCl4]− with the biologically relevant thiols, L-cysteine (L-Cys) and glutathione (GSH), and thioether, L-methionine (L-Met), was studied using UV-Vis spectroscopy, cyclic voltammetry, 1H NMR spectroscopy and ESI-MS. In this study, the rate constants for substitution reactions between monofunctional gold(III) complexes and sulfur donor ligands in aqueous solution were determined at different initial concentrations of reactants, chloride ions, pH and constant ionic strength. The obtained second-order rate constants for the reaction with L-methionine in the absence of added chloride at pH 2.5 and 25 °C follow the sequence (7.5 ± 0.4) × 103 > (4.5 ± 0.1) × 102 > 88.3 ± 0.8 M−1 s−1 for the terpy, bpma and dien complexes, respectively, demonstrating that the substitution step could be detected prior to the reduction step. This behavior was expected due to the influence of a decreasing π-donor ability of the chelate ligands, which slows down the substitution reactions along the series of complexes studied. In order to throw more light on the mechanism of biological activity of gold(III) compounds, such a systematic study was performed for all the mentioned thiols and thioether.
Co-reporter:Debabrata Chatterjee, Alicja Franke, Maria Oszajca and Rudi van Eldik
Dalton Transactions 2014 vol. 43(Issue 8) pp:3087-3094
Publication Date(Web):16 Oct 2013
DOI:10.1039/C3DT52486H
The [RuIII(edta)(H2O)]− (edta4− = ethylenediaminetetraacetate) complex catalyzes the oxidation of azide (N3−) with H2O2, mimicking the action of metallo-enzymes such as catalase and peroxidase in biochemistry. The kinetics of the catalytic oxidation process was studied by using stopped-flow and rapid-scan spectrophotometry as a function of [RuIII(edta)], [H2O2], [N3−] and pH. The catalytic activity of the different oxidizing species produced in the reaction of [RuIII(edta)(H2O)]− with H2O2 for the oxidation of azide was compared to the oxidation of coordinated azide in [RuIII(edta)N3]2− by H2O2. Detailed reaction mechanisms in agreement with the spectroscopic and kinetic data are presented for both reaction paths.
Co-reporter:Debabrata Chatterjee, Namita Jaiswal, Matthias Schmeisser and Rudi van Eldik
Dalton Transactions 2014 vol. 43(Issue 48) pp:18042-18046
Publication Date(Web):06 Oct 2014
DOI:10.1039/C4DT02628D
Reported here is the first example of a ruthenium(III) complex [RuIII(EDTA)(H2O)]− (EDTA4− = ethylenediaminetetraacetate) that mediates S-nitrosylation of cysteine in the presence of nitrite at pH 4.5 (acetate buffer) and results in the formation of [RuIII(EDTA)(SNOCy)]−. The kinetics of the reaction was studied by stopped-flow and rapid-scan spectrophotometry as a function of [Cysteine], [NO2−] and pH (3.5–8.5). Formation of [RuIII(EDTA)(SNOCy)]−, the product of the S-nitrosylation reaction, was identified by ESI-MS experiments. A working mechanism in agreement with the spectroscopic and kinetic data is presented.
Co-reporter:Ma&x142;gorzata Brindell;Karol Dyduch;Agnieszka Adamowicz;El&x17c;bieta Urbanowicz;Maria Oszajca;Artur Michalak;Gra&x17c;yna Stochel
European Journal of Inorganic Chemistry 2014 Volume 2014( Issue 8) pp:
Publication Date(Web):
DOI:10.1002/ejic.201301154
Abstract
The ruthenium(II) complex cis-[RuCl2(dmso)4] is known for its antiproliferative properties. This has stimulated the discovery of a wide group of new ruthenium complexes considered as potential anticancer drugs. The stability of these ruthenium complexes under physiological conditions and their interaction with protein amino acid residues are particularly important because of their intravenous administration. The studies presented here are devoted to the stability of cis-[RuCl2(dmso)4] under physiological pH conditions and its reactivity towards L-methionine. Immediate dissociation of one dmso ligand from the parent complex after dilution in water leads to the formation of fac-[RuCl2(H2O)(dmso)3], which under basic conditions undergoes stepwise dissociation of chloride ions without further dmso release. The hydrolysis of fac-[RuCl2(H2O)(dmso)3] at pH 7.4 and also base-catalyzed hydrolysis were investigated in detail by application of kinetic and spectroscopic (UV/Vis, NMR) measurements. The combined experimental and theoretical study revealed that L-methionine coordinates to fac-[RuCl(OH)(H2O)(dmso)3] by substituting a water ligand with simultaneous dmso release. Detailed NMR characterization of the product indicated that methionine coordinates through the NH2 group rather than the thioether moiety. This conclusion was further supported by theoretical calculations with the application of ETS–NOCV analysis.
Co-reporter:Anna Katafias;Olga Impert;Przemys&x142;aw Kita;Joanna Fenska;Stanis&x142;aw Koter;Anna Kaczmarek-K&x119;dziera;Henryk Ró&x17c;ycki;Anna Bajek;Ma&x142;gorzata Uzarska
European Journal of Inorganic Chemistry 2014 Volume 2014( Issue 15) pp:
Publication Date(Web):
DOI:10.1002/ejic.201301556
Abstract
The kinetics of reduction of the mer-[RuIII(pic)3] complex (pic– = picolinato) by ascorbic acid (AscH2) leading to formation of a red ruthenium(II) species have been studied spectrophotometrically by using both conventional mixing and stopped-flow methods. The reaction was followed as a function of the reductant concentration over a wide pH range (1.0–7.4). Electron transfer proceeds by an outer-sphere mechanism involving three protolytic forms of ascorbic acid, AscH2, AscH– and Asc2–, for which specific rate constants have been determined. The Gibbs' energy of activation was found to correlate linearly with the HOMO energies of the protolytic forms of the reductant. The mer-[RuIII(pic)3] complex is too sparingly soluble in water to inhibit the growth of the Escherichia coli (ATCC 8739) strain. Its cytotoxicity against non-tumorigenic cells precludes its potential use as an anticancer agent.
Co-reporter:Dimitri E. Khoshtariya, Tinatin D. Dolidze, Mikhael Shushanyan, and Rudi van Eldik
The Journal of Physical Chemistry B 2014 Volume 118(Issue 3) pp:692-706
Publication Date(Web):December 26, 2013
DOI:10.1021/jp4101569
Horse muscle myoglobin (Mb) was tightly immobilized at Au-deposited ∼15-Å-thick mixed-type (1:1) alkanethiol SAMs, HS–(CH2)11–COOH/HS–(CH2)11–OH, and placed in contact with buffered H2O or D2O solutions. Fast-scan cyclic voltammetry (CV) and a Marcus-equation-based analysis were applied to determine unimolecular standard rate constants and reorganization free energies for electron transfer (ET), under variable-temperature (15–55 °C) and -pressure (0.01–150 MPa) conditions. The CV signal was surprisingly stable and reproducible even after multiple temperature and pressure cycles. The data analysis revealed the following values: standard rate constant, 33 s–1 (25 °C, 0.01 MPa, H2O); reorganization free energy, 0.5 ± 0.1 eV (throughout); activation enthalpy, 12 ± 3 kJ mol–1; activation volume, −3.1 ± 0.2 cm3 mol–1; and pH-dependent solvent kinetic isotope effect (kH0/kD0), 0.7–1.4. Furthermore, the values for the rate constant and reorganization free energy are very similar to those previously found for cytochrome c electrostatically immobilized at the monocomponent Au/HS–(CH2)11–COOH junction. In vivo, Mb apparently forms a natural electrostatic complex with cytochrome b5 (cyt-b5) through the “dynamic” (loose) docking pattern, allowing for a slow ET that is intrinsically coupled to the water’s removal from the “defective” heme iron (altogether shaping the biological repair mechanism for Mb’s “met” form). In contrary, our experiments rather mimic the case of a “simple” (tight) docking of the redesigned (mutant) Mb with cyt-b5 (Nocek et al. J. Am. Chem. Soc. 2010, 132, 6165−6175). According to our analysis, in this configuration, Mb’s distal pocket (linked to the “ligand channel”) seems to be arrested within the restricted configuration, allowing the rate-determining reversible ET process to be coupled only to the inner-sphere reorganization (minimal elongation/shortening of an Fe–OH2 bond) rather than the pronounced detachment (rebinding) of water and, hence, to be much faster.
Co-reporter:Maria Oszajca;Agnieszka Drzewiecka-Matuszek;Dr. Alicja Franke;Dr. Dorota Rutkowska-Zbik;Dr. Ma&x142;gorzata Brindell; Ma&x142;gorzata Witko; Gra&x17c;yna Stochel; Rudi vanEldik
Chemistry - A European Journal 2014 Volume 20( Issue 8) pp:2328-2343
Publication Date(Web):
DOI:10.1002/chem.201303694
Abstract
High-valent iron-oxo species have been invoked as reactive intermediates in catalytic cycles of heme and nonheme enzymes. The studies presented herein are devoted to the formation of compound II model complexes, with the application of a water soluble (TMPS)FeIII(OH) porphyrin ([meso-tetrakis(2,4,6-trimethyl-3-sulfonatophenyl)porphinato]iron(III) hydroxide) and hydrogen peroxide as oxidant, and their reactivity toward selected organic substrates. The kinetics of the reaction of H2O2 with (TMPS)FeIII(OH) was studied as a function of temperature and pressure. The negative values of the activation entropy and activation volume for the formation of (TMPS)FeIVO(OH) point to the overall associative nature of the process. A pH-dependence study on the formation of (TMPS)FeIVO(OH) revealed a very high reactivity of OOH− toward (TMPS)FeIII(OH) in comparison to H2O2. The influence of N-methylimidazole (N-MeIm) ligation on both the formation of iron(IV)-oxo species and their oxidising properties in the reactions with 4-methoxybenzyl alcohol or 4-methoxybenzaldehyde, was investigated in detail. Combined experimental and theoretical studies revealed that among the studied complexes, (TMPS)FeIII(H2O)(N-MeIm) is highly reactive toward H2O2 to form the iron(IV)-oxo species, (TMPS)FeIVO(N-MeIm). The latter species can also be formed in the reaction of (TMPS)FeIII(N-MeIm)2 with H2O2 or in the direct reaction of (TMPS)FeIVO(OH) with N-MeIm. Interestingly, the kinetic studies involving substrate oxidation by (TMPS)FeIVO(OH) and (TMPS)FeIVO(N-MeIm) do not display a pronounced effect of the N-MeIm axial ligand on the reactivity of the compound II mimic in comparison to the OH− substituted analogue. Similarly, DFT computations revealed that the presence of an axial ligand (OH− or N-MeIm) in the trans position to the oxo group in the iron(IV)-oxo species does not significantly affect the activation barriers calculated for CH dehydrogenation of the selected organic substrates.
Co-reporter:Dr. Li Ji;Dr. Alicja Franke;Dr. Ma&x142;gorzata Brindell;Dr. Maria Oszajca;Dr. Achim Zahl;Dr. Rudi vanEldik
Chemistry - A European Journal 2014 Volume 20( Issue 44) pp:14437-14450
Publication Date(Web):
DOI:10.1002/chem.201402347
Abstract
For the exploration of the intrinsic reactivity of two key active species in the catalytic cycle of horseradish peroxidase (HRP), Compound I (HRP-I) and Compound II (HRP-II), we generated in situ [FeIVO(TMP+.)(2-MeIm)]+ and [FeIVO(TMP)(2-MeIm)]0 (TMP=5,10,15,20-tetramesitylporphyrin; 2-MeIm=2-methylimidazole) as biomimetics for HRP-I and HRP-II, respectively. Their catalytic activities in epoxidation, hydrogen abstraction, and heteroatom oxidation reactions were studied in acetonitrile at −15 °C by utilizing rapid-scan UV/Vis spectroscopy. Comparison of the second-order rate constants measured for the direct reactions of the HRP-I and HRP-II mimics with the selected substrates clearly confirmed the outstanding oxidizing capability of the HRP-I mimic, which is significantly higher than that of HRP-II. The experimental study was supported by computational modeling (DFT calculations) of the oxidation mechanism of the selected substrates with the involvement of quartet and doublet HRP-I mimics (2,4Cpd I) and the closed-shell triplet spin HRP-II model (3Cpd II) as oxidizing species. The significantly lower activation barriers calculated for the oxidation systems involving 2,4Cpd I than those found for 3Cpd II are in line with the much higher oxidizing efficiency of the HRP-I mimic proven in the experimental part of the study. In addition, the DFT calculations show that all three reaction types catalyzed by HRP-I occur on the doublet spin surface in an effectively concerted manner, whereas these reactions may proceed in a stepwise mechanism with the HRP-II mimic as oxidant. However, the high desaturation or oxygen rebound barriers during CH bond activation processes by the HRP-II mimic predict a sufficient lifetime for the substrate radical formed through hydrogen abstraction. Thus, the theoretical calculations suggest that the dissociation of the substrate radical may be a more favorable pathway than desaturation or oxygen rebound processes. Importantly, depending on the electronic nature of the oxidizing species, that is, 2,4Cpd I or 3Cpd II, an interesting region-selective conversion phenomenon between sulfoxidation and H-atom abstraction was revealed in the course of the oxidation reaction of dimethylsulfide. The combined experimental and theoretical study on the elucidation of the intrinsic reactivity patterns of the HRP-I and HRP-II mimics provides a valuable tool for evaluating the particular role of the HRP active species in biological systems.
Co-reporter:Klaus Pokorny ; Matthias Schmeisser ; Frank Hampel ; Achim Zahl ; Ralph Puchta
Inorganic Chemistry 2013 Volume 52(Issue 22) pp:13167-13178
Publication Date(Web):November 4, 2013
DOI:10.1021/ic4020724
On the basis of 7Li NMR experiments, the complex-formation reaction between Li+ and the tridentate N-donor ligand terpyridine was studied in the ionic liquids [emim][NTf2] and [emim][ClO4] as solvents. For both ionic liquids, the NMR data implicate the formation of [Li(terpy)2]+. Density functional theory calculations show that partial coordination of terpyridine involving the coordination of a solvent anion can be excluded. In contrast to the studies in solution, X-ray diffraction measurements led to completely different results. In the case of [emim][NTf2], the polymeric lithium species [Li(terpy)(NTf2)]n was found to control the stacking of this complex, whereas crystals grown from [emim][ClO4] exhibit the discrete dimeric species [Li(terpy)(ClO4)]2. However, both structures indicate that each lithium ion is formally coordinated by one terpy molecule and one solvent anion in the solid state, suggesting that charge neutralization and π stacking mainly control the crystallization process.
Co-reporter:Dimitri E. Khoshtariya, Tina D. Dolidze, Tatyana Tretyakova, David H. Waldeck and Rudi van Eldik
Physical Chemistry Chemical Physics 2013 vol. 15(Issue 39) pp:16515-16526
Publication Date(Web):25 Jul 2013
DOI:10.1039/C3CP51896E
Gold electrodes were coated with alkanethiol SAM–azurin (Az, blue cupredoxin) assemblies and placed in contact with a water-doped and buffered protic ionic melt as the electrolyte, choline dihydrogen phosphate ([ch][dhp]). Fast-scan protein-film voltammetry was applied to explore interfacial biological electron transfer (ET) under conditions approaching the glass-transition border. The ET rate was studied as a function of the water amount, temperature (273–353 K), and pressure (0.1–150 MPa). Exposure of the Az films to the semi-solid electrolyte greatly affected the protein's conformational dynamics, hence the ET rate, via the mechanism occurring in the extra complicated dynamically-controlled regime, is compared to the earlier studies on the reference system with a conventional electrolyte (D. E. Khoshtariya et al., Proc. Natl. Acad. Sci. U. S. A., 2010, 107, 2757–2762), allowing for the disclosure of even more uncommon mechanistic motifs. For samples with low water content (ca. 3 or less waters per [ch][dhp]), at moderately low temperatures (below ca. 298 K) and/or high pressure (150 MPa), the voltammetric profiles systematically deviated from the standard Marcus current–overvoltage pattern, deemed as attributable to a breakdown of the linear response approximation through the essential steepening of the Gibbs energy wells near the glass-forming threshold. Electrolytes with a higher water content (6 to 15 waters per [ch][dhp]) display anomalous temperature and pressure performances, suggesting that the system crosses a broad nonergodic zone which arises from the interplay of ET-coupled large-scale conformational (highly cooperative) modes of the Az protein, inherently linked to the electrolyte's (water-doped [ch][dhp]) slowest collective relaxation(s).
Co-reporter:Rodrigo Luis Silva Ribeiro Santos, Rudi van Eldik and Denise de Oliveira Silva
Dalton Transactions 2013 vol. 42(Issue 48) pp:16796-16805
Publication Date(Web):29 Aug 2013
DOI:10.1039/C3DT51763B
Diruthenium(II,III)-tetracarboxylates have shown promising anticancer properties as metallotherapeutics. On the basis of the role that bio-reducing agents may play on the mode of action of ruthenium-based anticancer drugs, we performed detailed kinetic studies on the reaction of ascorbic acid and glutathione with the [Ru2(RCOO)4]+ paddlewheel framework by using the non-drug, diaqua complex ion [Ru2(CH3COO)4(H2O)2]+. In the presence of the reducing agents, the diaqua-Ru2 species first undergo a ligand substitution reaction by which the axially-coordinated water is displaced by the reducing agent. In both cases, this reaction is followed by an intra-molecular electron transfer process during which the metal–metal center is reduced from Ru25+ to Ru24+ and the reducing agent is oxidized. Product analyses were performed with the application of ESI-MS and 1H-NMR techniques. Rate and activation parameters are reported for the different reaction steps.
Co-reporter:Snežana Jovanović, Biljana Petrović, Živadin D. Bugarčić and Rudi van Eldik
Dalton Transactions 2013 vol. 42(Issue 24) pp:8890-8896
Publication Date(Web):16 Apr 2013
DOI:10.1039/C3DT50751C
The reduction of the Pt(IV) complexes [PtCl4(bipy)], [PtCl4(dach)] and [PtCl4(en)] by glutathione (GSH), L-cysteine (L-Cys) and L-methionine (L-Met) was investigated by stopped-flow spectrophotometry at pH 2.0 (in 0.01 M perchloric acid) and at pH 7.2 (in 25 mM Hepes buffer). Kinetic measurements were performed under pseudo-first order conditions with an excess of the reducing agent. The order of the reactivity of the studied complexes was [PtCl4(bipy)] > [PtCl4(dach)] > [PtCl4(en)], and reactivity of investigated reducing agents followed the order GSH > L-Cys > L-Met. All the reactions between the selected Pt(IV) complexes and the sulfur donor biomolecules proceeded by a reductive elimination process that included nucleophilic attack by the reducing agent on one of the mutually trans-coordinated chloride ligands, which led to a two-electron transfer process. The final products of the redox reactions were the corresponding reduced Pt(II) complexes and the oxidized form of the reducing agents.
Co-reporter:Debabrata Chatterjee, Sabine Rothbart and Rudi van Eldik
Dalton Transactions 2013 vol. 42(Issue 13) pp:4725-4729
Publication Date(Web):30 Jan 2013
DOI:10.1039/C3DT32737J
Reported here is the first example of a ruthenium complex, [RuIII(edta)(H2O)]− (edta4− = ethylenediaminetetraacetate), that catalyzes the oxidation of thiourea (TU) in the presence of H2O2. The kinetics and mechanism of this reaction were investigated in detail by using rapid-scan spectrophotometry as a function of both the hydrogen peroxide and thiourea concentrations at pH 4.9 and 25 °C. Spectral analyses and kinetic data clearly support a catalytic process in which hydrogen peroxide reacts directly with thiourea coordinated to the RuIII(edta) complex. HPLC product analyses revealed the formation of formamidine disulfide (TU2) as a major product at the end of the catalytic process, however, formation of other products like thiourea dioxide (TUO2), thiourea dioxide (TUO3) and sulfate was also observed after longer reaction times. Catalytic intermediates such as [RuIII(edta)(OOH)]2− and [RuV(edta)(O)]− were evidently found to be non-reactive in catalyzing the oxidation of thiourea by H2O2 under the specified conditions.
Co-reporter:Debabrata Chatterjee, Sabine Rothbart and Rudi van Eldik
RSC Advances 2013 vol. 3(Issue 11) pp:3606-3610
Publication Date(Web):04 Feb 2013
DOI:10.1039/C2RA22810F
The [RuIII(edta)(H2O)]− (edta4− = ethylenediaminetetraacetate) catalyzed oxidative degradation of methylene blue (MB) in the presence of H2O2 has been achieved. Complete degradation of 0.05 mM MB was noticed within 300 s in the presence of 5.0 μM of Ru(edta) catalyst under ambient conditions. Catalytic degradation of the dye was investigated by using rapid-scan spectrophotometry as a function of [Ru(edta)], [H2O2], [MB] and pH at 25 °C. Spectral analyses and kinetic data point towards a catalytic pathway that involves the rapid coordination of MB to the RuIII(edta) complex to yield [RuIII(edta)(MB)]− species that react directly with hydrogen peroxide. The catalytic-active intermediate [RuIII(edta)(OOH)]2− was also found to be operative at high H2O2 concentration over the Ru(edta) catalyst in effecting MB degradation under the specified conditions.
Co-reporter:Dr. Matthias Schmeisser;Dr. Peter Illner;Dr. Ralph Puchta;Dr. Achim Zahl ;Dr. Rudi vanEldik
Chemistry - A European Journal 2013 Volume 19( Issue 49) pp:16835-16836
Publication Date(Web):
DOI:10.1002/chem.201302977
Co-reporter:Rodrigo L. S. R. Santos, Rudi van Eldik, and Denise de Oliveira Silva
Inorganic Chemistry 2012 Volume 51(Issue 12) pp:6615-6625
Publication Date(Web):June 1, 2012
DOI:10.1021/ic300168t
The known paddlewheel, tetrakis(acetato)chloridodiruthenium(II,III), offers a versatile synthetic route to a novel class of antitumor diruthenium(II,III) metallo drugs, where the equatorial ligands are nonsteroidal anti-inflammatory carboxylates. This complex was studied here as a soluble starting prototype model for antitumor analogues to elucidate the reactivity of the [Ru2(CH3COO)4]+ framework. Thermodynamic studies on equilibration reactions for axial substitution of water by chloride and kinetic studies on reactions of the diaqua complexes with the amino acids glycine, cysteine, histidine, and tryptophan were performed. The standard thermodynamic reaction parameters ΔH°, ΔS°, and ΔV° were determined and showed that both of the sequential axial substitution reactions are enthalpy driven. Kinetic rate laws and rate constants were determined for the axial substitution reactions of coordinated water by the amino acids that gave the corresponding aqua(amino acid)–Ru2 substituted species. The results revealed that the [Ru2(CH3COO)4]+ paddlewheel framework remained stable during the axial ligand substitution reactions and was also mostly preserved in the presence of the amino acids.
Co-reporter:Simon Kern and Rudi van Eldik
Inorganic Chemistry 2012 Volume 51(Issue 13) pp:7340-7345
Publication Date(Web):June 20, 2012
DOI:10.1021/ic300718v
Detailed kinetic studies were performed on the reaction of [RuII(terpy)(bpy)H]+ (terpy = 2,2′,6′,2″-terpyridine; bpy = 2,2′-bipyridine) with CO2 in conventional solvents (water, methanol, and ethanol) and in the ionic liquid [emim][NTf2] ([emim] = 1-ethyl-3-methyl-imidazolium; [NTf2] = bistrifluoromethylsulfonylamide). Second-order rate constants and activation parameters (ΔH⧧, ΔS⧧, and ΔV⧧) were determined for the reaction in all solvents. The second-order rate constants correlate with the acceptor number of the solvent, whereas the activation parameters support the associative nature of the reaction. The results in water, especially the activation entropy (+14 ± 2 J K–1 mol–1) and activation volume (−5.9 ± 0.6 cm3 mol–1), differ significantly from those found for the other solvents.
Co-reporter:Biljana Petrović, Živadin D. Bugarčić, Anne Dees, Ivana Ivanović-Burmazović, Frank W. Heinemann, Ralph Puchta, Stephan N. Steinmann, Clemence Corminboeuf, and Rudi van Eldik
Inorganic Chemistry 2012 Volume 51(Issue 3) pp:1516-1529
Publication Date(Web):January 20, 2012
DOI:10.1021/ic201807a
The kinetics and mechanism of substitution reactions of novel monofunctional [Pt(tpdm)Cl]+ and [Pd(tpdm)Cl]+ complexes (where tpdm = tripyridinedimethane) and their aqua analogues with thiourea (tu), l-methionine (l-met), glutathione (GSH), and guanosine-5′-monophosphate (5′-GMP) were studied in 0.1 M NaClO4 at pH = 2.5 (in the presence of 10 mM NaCl for reactions of the chlorido complexes). The reactivity of the investigated nucleophiles follows the order tu > l-met > GSH > 5′-GMP. The reported rate constants showed the higher reactivity of the Pd(II) complexes as well as the higher reactivity of the aqua complex than the corresponding chlorido complex. The negative values reported for the activation entropy as well as the activation volume confirmed an associative substitution mode. In addition, the molecular and crystal structure of [Pt(tpdm)Cl]Cl was determined by X-ray crystallography. The compound crystallizes in a monoclinic space group C2/c with two independent molecules of the complex and unit cell dimensions of a = 38.303(2) Å, b = 9.2555(5) Å, c = 27.586(2) Å, β = 133.573(1)°, and V = 7058.3(8) Å3. The cationic complex [Pt(tpdm)Cl]+ exhibits square-planar coordination around the Pt(II) center. The lability of the [Pt(tpdm)Cl]+ complex is orders of magnitude lower than that of [Pt(terpyridine)Cl]+. Quantum chemical calculations were performed on the [Pt(tpdm)Cl]+ and [Pt(terpyridine)Cl]+ complexes and their reactions with thiourea. Theoretical computations for the corresponding Ni(II) complexes clearly demonstrated that π-back-bonding properties of the terpyridine chelate can account for acceleration of the nucleophilic substitution process as compared to the tpdm chelate, where introduction of two methylene groups prevents such an effective π-back bonding.
Co-reporter:Stephanie Hochreuther
Inorganic Chemistry 2012 Volume 51(Issue 5) pp:3025-3038
Publication Date(Web):February 22, 2012
DOI:10.1021/ic202351r
A dinuclear platinum(II) complex that was recently investigated in our group was tested for its cytostatic activity and found to be active against HeLa S3 cells. The complex consists of a bidentate N,N-donor chelating ligand system in which the two platinum centers are connected by an aliphatic chain of 10 methylene groups. The complex [Pt2(N1,N10-bis(2-pyridylmethyl)-1,10-decanediamine)(OH2)4]4+ (10NNpy) is of further special interest, since only little is known about the substitution behavior of such dinuclear platinum complexes that contain a bidentate coordination sphere. The complex was investigated using different biologically relevant nucleophiles, such as thiourea (tu), l-methionine (l-Met), glutathione (GSH), and guanine-5′-monophosphate (5′-GMP), at two different pH values (2 and 7.4). The substitution of coordinated water by these nucleophiles was studied under pseudo-first-order conditions as a function of nucleophile concentration, temperature, and pressure, using stopped-flow techniques and UV–vis spectroscopy. The reactivity of 10NNpy with the selected nucleophiles was found to be tu ≫ 5′-GMP > l-Met > GSH at pH 2 and GSH > tu > l-Met at pH 7.4. The results for the dinuclear 10NNpy complex were compared to those for the corresponding mononuclear reference complex [Pt(aminomethylpyridine)(OH2)2]2+, Pt(amp), studied before in our group, by which the effect of the addition of an aliphatic chain, an increase in the overall charge, and a shift in the pKa values of the coordinated water ligands could be investigated. The reactivity order for Pt(amp) was found to be tu > GSH > l-Met at pH 7.4.
Co-reporter:Živadin D. Bugarčić, Jovana Bogojeski, Biljana Petrović, Stephanie Hochreuther and Rudi van Eldik
Dalton Transactions 2012 vol. 41(Issue 40) pp:12329-12345
Publication Date(Web):14 Aug 2012
DOI:10.1039/C2DT31045G
A brief overview of mechanistic studies on the reactions of different Pt(II) complexes with nitrogen- and sulfur-donor biomolecules is presented. The first part describes the results obtained for substitution reactions of mono-functional Pt(II) complexes with different biomolecules, under various experimental conditions (temperature, pH and ionic strength). In addition, an overview of the results obtained for the substitution reactions of bi-functional Pt(II) complexes, analogous to cisplatin, with biomolecules is given. The last part of this report deals with different polynuclear Pt(II) complexes and their substitution behaviour with different biomolecules. The purpose of this perspective is to improve the understanding of the mechanism of action of Pt(II) complexes as potential anti-tumour drugs in the human body.
Co-reporter:Ana Djeković, Biljana Petrović, Živadin D. Bugarčić, Ralph Puchta and Rudi van Eldik
Dalton Transactions 2012 vol. 41(Issue 13) pp:3633-3641
Publication Date(Web):09 Feb 2012
DOI:10.1039/C2DT11843B
The kinetics of the substitution reactions between the mono-functional Au(III) complexes, [Au(dien)Cl]2+ and [Au(terpy)Cl]2+ (dien = 3-azapentane-1,5-diamine, terpy = 2,2′;6′,2′′-terpyridine) and bi-functional Au(III) complexes, [Au(bipy)Cl2]+ and [Au(dach)Cl2]+ (bipy = 2.2′-bipyridine, dach = (1R,2R)-1,2-diaminocyclohexane) and biologically relevant ligands such as L-histidine (L-His), inosine (Ino), inosine-5′-monophosphate (5′-IMP) and guanosine-5′-monophosphate (5′-GMP), were studied in detail. All kinetic studies were performed in 25 mM Hepes buffer (pH = 7.2) in the presence of NaCl to prevent the spontaneous hydrolysis of the chloride complexes. The reactions were followed under pseudo-first order conditions as a function of ligand concentration and temperature using stopped-flow UV-vis spectrophotometry. The results showed that the mono-functional complexes react faster than the bi-functional complexes in all studied reactions. The [Au(terpy)Cl]2+ complex is more reactive than the [Au(dien)Cl]2+ complex, which was confirmed by quantum chemical (DFT) calculations. A more than 50% lower activation energy for the terpy than for the dien based complex was found. The bi-functional [Au(bipy)Cl2]+ complex is more reactive than the [Au(dach)Cl2]+ complex. The reactivity of the studied nucleophiles follows the same order for all studied systems, viz. L-His > 5′-GMP > 5′-IMP > Ino. According to the measured activation parameters, all studied reactions follow an associative substitution mechanism. Quantum chemical calculations (B3LYP/LANL2DZp) suggest that ligand substitution in [Au(terpy)Cl]2+ and [Au(dien)Cl]2+ by imidazole follows an interchange mechanism with a significant degree of associative character. The results demonstrate the strong connection between the reactivity of the complexes toward biologically relevant ligands and their structural and electronic characteristics. Therefore, the binding of gold(III) complexes to 5′-GMP, constituent of DNA, is of particular interest since this interaction is thought to be responsible for their anti-tumour activity.
Co-reporter:Tanja Soldatović, Snežana Jovanović, Živadin D. Bugarčić and Rudi van Eldik
Dalton Transactions 2012 vol. 41(Issue 3) pp:876-884
Publication Date(Web):08 Nov 2011
DOI:10.1039/C1DT11313E
The novel dinuclear Pt(II) complexes [{trans-Pt(NH3)2Cl}2(μ-pyrazine)](ClO4)2 (Pt1), [{trans-Pt(NH3)2Cl}2(μ-4,4′-bipyridyl)](ClO4)2·DMF (Pt2), and [{trans-Pt(NH3)2Cl}2(μ-1,2-bis(4-pyridyl)ethane)](ClO4)2 (Pt3), were synthesized. Acid–base titrations, and temperature and concentration dependent kinetic measurements of the reactions with biologically relevant ligands such as thiourea (Tu), glutathione (GSH) and guanosine-5′-monophosphate (5′-GMP) were studied at pH 2.5 and 7.2. The reactions were followed under pseudo-first-order conditions by stopped-flow and UV-vis spectrophotometry. 1H NMR spectroscopy was used to follow the substitution of chloride in the complex [{trans-Pt(NH3)2Cl}2(μ-4,4′-bipyridyl)](ClO4)2·DMF by guanosine-5′-monophosphate (5′-GMP) under second-order conditions. The results indicate that the bridging ligand has an influence on the reactivity of the complexes towards nucleophiles. The order of reactivity of the investigated complexes is Pt1 > Pt2 > Pt3.
Co-reporter:Stephanie Hochreuther, Sharanappa T. Nandibewoor, Ralph Puchta and Rudi van Eldik
Dalton Transactions 2012 vol. 41(Issue 2) pp:512-522
Publication Date(Web):01 Nov 2011
DOI:10.1039/C1DT11453K
The diaqua complex [Pt(2-methylthiomethylpyridine)(OH2)2]2+, Pt(mtp), was synthesized and investigated thermodynamically as well as kinetically. Spectrophotometric acid–base titrations were performed to determine the pKa values of the two coordinated water ligands. A low pKa1 value of 3.15 was observed for the water molecule trans to the pyridine donor, whereas a pKa2 value of 6.84 was found for the water molecule trans to the labilising sulphur donor. The substitution of coordinated water by a series of sterically hindered S-containing nucleophiles, viz.thiourea (tu), N,N′-dimethylthiourea (dmtu) and N,N,N′,N′-tetramethylthiourea (tmtu), was studied under pseudo first-order conditions as a function of nucleophile concentration, pH (2, 4.75, 7.4), temperature and pressure, using stopped-flow techniques and UV-vis spectroscopy. In general the first substitution reaction takes place trans to the sulphur donor. At pH 2 the nucleophiles react in the order tu (634 ± 10) > dmtu (507 ± 5) ≫ tmtu (165 ± 3 M−1 s−1 at 25 °C), which is caused by steric hindrance. The second observed reaction involves two steps, viz. the displacement of the second water ligand and dechelation of the pyridine ring with the third-order rate constants 73.3 ± 0.8 (tu), 22.1 ± 0.1 (dmtu) and 6.8 ± 0.2 M−2 s−1 (tmtu) at 25 °C. At pH 4.75 the reactions are in general slower due to the presence of aqua-hydroxo species. The same order in reactivity was found, viz. tu (106 ± 1) > dmtu (72 ± 1) ≫ tmtu (14.1 ± 0.5 M−1 s−1 at 25 °C). No evidence for ring-dechelation could be observed under these conditions. At pH 7.4 the inert dihydroxo species is predominantly present in solution and consequently no substitution reaction was observed. Quantum chemical calculations were performed to support the interpretation and discussion of the experimental results.
Co-reporter:Thomas Roth, Raquel Urpi Bertran, Manfred Pöhlein, Marion Wolf, Rudi van Eldik
Journal of Chromatography A 2012 Volume 1262() pp:188-195
Publication Date(Web):2 November 2012
DOI:10.1016/j.chroma.2012.08.070
Two methods for the determination of phosphate-based flame retardants (PFRs) and similar organophosphates (OPs) were developed. Two gas chromatographic systems were applied, one equipped with a quadrupole mass spectrometer (GC/MS), the other with a specific phosphorus–nitrogen detector (GC/PND). A procedure of ultrasonic supported extraction and precipitation (USSE) was applied and verified using in-house produced reference materials. Twelve polymer parts of electrical and electronic devices, and eight polymer references were analysed for their PFR contents. The results show that the methods are capable of identifying PFRs in concentration ranges from 3 μg L−1 (tricresyl phosphate, TCP) to 12 μg L−1 (cresyl diphenyl phosphate, CDP) on GC/PND and from 110 μg L−1 (triphenyl phosphine oxide, TPPO) to 3250 μg L−1 (tris(ethylhexyl) phosphate, TEHP) on GC/MS in reference solutions. LODs in polymer extracts range from 180 μg g−1 (triphenyl phosphate, TPP) to 670 μg g−1 (bisphenol-A-bis(diphenyl phosphate), BDP) on GC/PND and from 75 μg g−1 (TPPO) to 780 μg g−1 (BDP) on GC/MS. The overall procedure time for one sample was less than 45 min (GC/MS) and less than 65 min (GC/PND).Highlights► Methods for phosphate-based flame retardants and organophosphates were developed. ► Two gas chromatographic systems were employed, GC/MS and GC/PND. ► Ultrasonic supported extraction and precipitation was applied and verified. ► In-house produced reference materials were employed. ► 12 polymer parts of electronic devices, and 8 polymer references were analysed.
Co-reporter:Safaa S. Hassan, Mohamed M. Shoukry and Rudi van Eldik
Dalton Transactions 2012 vol. 41(Issue 43) pp:13447-13453
Publication Date(Web):25 Sep 2012
DOI:10.1039/C2DT31730C
The acid–base and complex-formation equilibria of [Ru(H2dtpa)(H2O)], where dtpa = diethylenetriaminepentaacetate, with a series of bio-relevant ligands having various functional groups, viz. thiol, amine, imidazole and carboxylate, were investigated using potentiometric and spectrophotometric techniques. The pKa values for [Ru(H2dtpa)(H2O)] were found to be 2.28 and 3.48 for the uncoordinated carboxylic acid groups and 8.83 for the coordinated water molecule. The complexes formed are of 1:1 stoichiometry and their formation-constants were determined. The effect of dioxane on the acid–base and complex-formation equilibria of the RuIII complex was studied. The displacement reaction of coordinated NO by the investigated ligands showed that thiols can compete with NO in their reaction with [RuIII(dtpa)(H2O)]2−. The results reveal that the RuIII complex is deactivated as a NO scavenger by thiolate ligands.
Co-reporter:Sabine Rothbart, Erika E. Ember and Rudi van Eldik
New Journal of Chemistry 2012 vol. 36(Issue 3) pp:732-748
Publication Date(Web):21 Dec 2011
DOI:10.1039/C2NJ20852K
A remarkably efficient, catalytic oxidative degradation of Orange II in aqueous solution was observed in the presence of simple Mn(II) salts used in combination with the green oxidant peracetic acid (PAA) under mild reaction conditions at pH 9.5 and 25 °C. In order to gain further insight into the complex reaction network responsible for the fast and efficient degradation of Orange II, detailed kinetic studies on the catalytic degradation reaction were performed. The influence of the catalyst, oxidant, and buffer concentrations on the reaction course was studied in detail by in situUV/Vis spectroscopy. The degradation rate increased with increasing catalyst and PAA concentration, decreased with increasing carbonate buffer concentration, and the reaction showed maximum reactivity at pH ≈ 9.4. Studies performed in the absence of substrate revealed the successive formation of various high valence manganese intermediates under the selected experimental conditions. The UV/Vis spectral changes observed upon addition of PAA to a Mn(II) ion containing solution at pH 9.5 showed biphasic behavior, and led to the formation of MnVIIO4− and colloidal MnIVO2 as final products. Through the combination of EPR and UV/Vis spectroscopy, the presence of a Mn(IV) species during the first phase of the reaction was observed. If the addition of a dye to the Mn(II)/PAA containing reaction mixture was carried out after different delay times during the course of the reaction, significant changes observed in the degradation performance pointed to changes in the catalyst composition with time. Thereby, the highest reactivity was reached just before the formation of the high valence intermediates and colloidal MnIVO2 occurred. Selected catalytic experiments with different in situ formed intermediates showed the essential role of a small amount of hydrogen peroxide (omnipresent in commercial peracetic acid, 6 wt% H2O2) for efficient catalytic dye decomposition. H2O2 was shown to play a crucial role as a reducing agent in the catalytic cycle, avoiding the over-oxidation of the catalyst, and thereby extending the lifetime of the catalytic system. The data were evaluated to gain mechanistic insight into the complex chemical reactions occurring during the Mn(II) catalyzed oxidative dye degradation with PAA. A comparison with earlier results concerning the use of H2O2/HCO3− as oxidant was made.
Co-reporter:Dr. Alicja Franke;Dr. Christoph Fertinger ;Dr. Rudi vanEldik
Chemistry - A European Journal 2012 Volume 18( Issue 22) pp:6935-6949
Publication Date(Web):
DOI:10.1002/chem.201103036
Abstract
The present study focuses on the formation and reactivity of hydroperoxo–iron(III) porphyrin complexes formed in the [FeIII(tpfpp)X]/H2O2/HOO− system (TPFPP=5,10,15,20-tetrakis(pentafluorophenyl)-21H,23H-porphyrin; X=Cl− or CF3SO3−) in acetonitrile under basic conditions at −15 °C. Depending on the selected reaction conditions and the active form of the catalyst, the formation of high-spin [FeIII(tpfpp)(OOH)] and low-spin [FeIII(tpfpp)(OH)(OOH)] could be observed with the application of a low-temperature rapid-scan UV/Vis spectroscopic technique. Axial ligation and the spin state of the iron(III) center control the mode of OO bond cleavage in the corresponding hydroperoxo porphyrin species. A mechanistic changeover from homo- to heterolytic OO bond cleavage is observed for high- [FeIII(tpfpp)(OOH)] and low-spin [FeIII(tpfpp)(OH)(OOH)] complexes, respectively. In contrast to other iron(III) hydroperoxo complexes with electron-rich porphyrin ligands, electron-deficient [FeIII(tpfpp)(OH)(OOH)] was stable under relatively mild conditions and could therefore be investigated directly in the oxygenation reactions of selected organic substrates. The very low reactivity of [FeIII(tpfpp)(OH)(OOH)] towards organic substrates implied that the ferric hydroperoxo intermediate must be a very sluggish oxidant compared with the iron(IV)–oxo porphyrin π-cation radical intermediate in the catalytic oxygenation reactions of cytochrome P450.
Co-reporter:Dr. Matthias Schmeisser;Dr. Peter Illner;Dr. Ralph Puchta;Dr. Achim Zahl;Dr. Rudi vanEldik
Chemistry - A European Journal 2012 Volume 18( Issue 35) pp:10969-10982
Publication Date(Web):
DOI:10.1002/chem.201200584
Abstract
We present for the first time Gutmann donor and acceptor numbers for a series of 36 different ionic liquids that include 26 distinct anions. The donor numbers were obtained by 23Na NMR spectroscopy and show a strong dependence on the anionic component of the ionic liquid. The donor numbers measured vary from −12.3 kcal mol−1 for the ionic liquid containing the weakest coordinative anion [emim][FAP] (1-ethyl-3-methylimidazolium tris(pentafluoroethyl)trifluorophosphate), which is a weaker donor than 1,2-dichloroethane, to 76.7 kcal mol−1 found for the ionic liquid [emim][Br], which exhibits a coordinative strength in the range of tertiary amines. The acceptor numbers were measured by using 31P NMR spectroscopy and also vary as a function of the anionic and cationic component of the ionic liquid. The data are presented and correlated with other solvent parameters like the Kamlet–Taft set of parameters, and compared to the donor numbers reported by other groups.
Co-reporter:Dr. Matthias Schmeisser;Dr. Peter Illner;Dr. Ralph Puchta;Dr. Achim Zahl;Dr. Rudi vanEldik
Chemistry - A European Journal 2012 Volume 18( Issue 35) pp:
Publication Date(Web):
DOI:10.1002/chem.201290149
Co-reporter:Barbara Krajewska
JBIC Journal of Biological Inorganic Chemistry 2012 Volume 17( Issue 7) pp:1123-1134
Publication Date(Web):2012 October
DOI:10.1007/s00775-012-0926-8
Urease, a Ni-containing metalloenzyme, features an activity that has profound medical and agricultural implications. The mechanism of this activity, however, has not been as yet thoroughly established. Accordingly, to improve its understanding, in this study we analyzed the steady-state kinetic parameters of the enzyme (jack bean), KM and kcat, measured at different temperatures and pressures. Such an analysis is useful as it provides information on the molecular nature of the intermediate and transition states of the catalytic reaction. We measured the parameters in a noninteracting buffer using a stopped-flow technique in the temperature range 15–35 °C and in the pressure range 5–132 MPa, the pressure-dependent measurements being the first of their kind performed for urease. While temperature enhanced the activity of urease, pressure inhibited the enzyme; the inhibition was biphasic. Analyzing KM provided the characteristics of the formation of the ES complex, and analyzing kcat, the characteristics of the activation of ES. From the temperature-dependent measurements, the energetic parameters were derived, i.e. thermodynamic ΔHo and ΔSo for ES formation, and kinetic ΔH≠ and ΔS≠ for ES activation, while from the pressure-dependent measurements, the binding ΔVb and activation \( \Updelta V_{\rm cat}^{ \ne } \) volumes were determined. The thermodynamic and activation parameters obtained are discussed in terms of the current proposals for the mechanism of the urease reaction, and they are found to support the mechanism proposed by Benini et al. (Structure 7:205–216; 1999), in which the Ni–Ni bridging hydroxide—not the terminal hydroxide—is the nucleophile in the catalytic reaction.
Co-reporter:Alicja Franke;Elisabeth Hartmann
JBIC Journal of Biological Inorganic Chemistry 2012 Volume 17( Issue 3) pp:447-463
Publication Date(Web):2012 March
DOI:10.1007/s00775-011-0867-7
The effect of pressure on the kinetics and thermodynamics of the reversible binding of camphor to cytochrome P450cam was studied as a function of the K+ concentration. The determination of the reaction and activation volumes enabled the construction of the first complete volume profile for the reversible binding of camphor to P450cam. Although the volume profiles constructed for the reactions conducted at low and high K+ concentrations are rather similar, and both show a drastic volume increase on going from the reactant to the transition state and a relatively small volume change on going from the transition to the product state, the position of the transition state is largely affected by the K+ concentration in solution. Similarly, the activation volume determined for the dissociation of camphor is influenced by the presence of K+, which reflects changes in the ease of water entering the active site of camphor-bound P450cam that depends on the K+ concentration. Careful analysis of the components that contribute to the observed volume changes allowed the estimation of the total number of water molecules expelled to the bulk solvent during the binding of camphor to P450cam and the subsequent spin transition. The results are discussed in reference to other studies reported in the literature that deal with the kinetics and thermodynamics of the binding of camphor to P450cam under various reaction conditions.
Co-reporter:Christoph Fertinger;Alicja Franke
JBIC Journal of Biological Inorganic Chemistry 2012 Volume 17( Issue 1) pp:27-36
Publication Date(Web):2012 January
DOI:10.1007/s00775-011-0822-7
Compound I, an oxo–iron(IV) porphyrin π-cation radical species, and its one-electron-reduced form compound II are regarded as key intermediates in reactions catalyzed by cytochrome P450. Although both reactive intermediates can be easily produced from model systems such as iron(III) meso-tetra(2,4,6-trimethylphenyl)porphyrin hydroxide by selecting appropriate reaction conditions, there are only a few thermal activation parameters reported for the reactions of compound I analogues, whereas such parameters for the reactions of compound II analogues have not been investigated so far. Our study demonstrates that ΔH≠ and ΔS≠ are closely related to the chemical nature of the substrate and the reactive intermediate (viz., compounds I and II) in epoxidation and C–H abstraction reactions. Although most studied reactions appear to be enthalpy-controlled (i.e., ΔH≠ > −TΔS≠), different results were found for C–H abstractions catalyzed by compound I. Whereas the reaction with 9,10-dihydroanthracene as a substrate is also dominated by the activation enthalpy (ΔH≠ = 42 kJ/mol, ΔS≠ = 41 J/Kmol), the same reaction with xanthene shows a large contribution from the activation entropy (ΔH≠ = 24 kJ/mol, ΔS≠ = −100 J/kmol). This is of special interest since the activation barrier for entropy-controlled reactions shows a significant dependence on temperature, which can have an important impact on the relative reaction rates. As a consequence, a close correlation between bond strength and reaction rate—as commonly assumed for C–H abstraction reactions—no longer exists. In this way, this study can contribute to a proper evaluation of experimental and computational data, and to a deeper understanding of mechanistic aspects that account for differences in the reactivity of compounds I and II.
Co-reporter:Colin D. Hubbard, Peter Illner and Rudi van Eldik
Chemical Society Reviews 2011 vol. 40(Issue 1) pp:272-290
Publication Date(Web):15 Nov 2010
DOI:10.1039/C0CS00043D
The focus of this article is an examination of chemical reaction mechanisms in ionic liquids. These mechanisms are compared with those pertaining to the same reactions carried out in conventional solvents. In cases where the mechanisms differ, attempts to provide an explanation in terms of the chemical and physicochemical properties of the reactants and of the components of the ionic liquids are described. A wide range of reactions from different branches of chemistry has been selected for this purpose. A sufficient background for student readers has been included. This tutorial review should also be of interest to kineticists, and to both new and experienced investigators in the ionic liquids field.
Co-reporter:Maria Oszajca ; Alicja Franke ; Małgorzata Brindell ; Grażyna Stochel
Inorganic Chemistry 2011 Volume 50(Issue 8) pp:3413-3424
Publication Date(Web):March 23, 2011
DOI:10.1021/ic1023345
The reaction of the water-soluble FeIII(TMPS) porphyrin with CN− in basic solution leads to the stepwise formation of FeIII(TMPS)(CN)(H2O) and FeIII(TMPS)(CN)2. The kinetics of the reaction of CN− with FeIII(TMPS)(CN)(H2O) was studied as a function of temperature and pressure. The positive value of the activation volume for the formation of FeIII(TMPS)(CN)2 is consistent with the operation of a dissociatively activated mechanism and confirms the six-coordinate nature of the monocyano complex. A good agreement between the rate constants at pH 8 and 9 for the formation of the dicyano complex implies the presence of water in the axial position trans to coordinated cyanide in the monocyano complex and eliminates the existence of FeIII(TMPS)(CN)(OH) under the selected reaction conditions. Both FeIII(TMPS)(CN)(H2O) and FeIII(TMPS)(CN)2 bind nitric oxide (NO) to form the same nitrosyl complex, namely, FeII(TMPS)(CN)(NO+). Kinetic studies indicate that nitrosylation of FeIII(TMPS)(CN)2 follows a limiting dissociative mechanism that is supported by the independence of the observed rate constant on [NO] at an appropriately high excess of NO, and the positive values of both the activation parameters ΔS⧧ and ΔV⧧ found for the reaction under such conditions. The relatively small first-order rate constant for NO binding, namely, (1.54 ± 0.01) × 10−2 s−1, correlates with the rate constant for CN− release from the FeIII(TMPS)(CN)2 complex, namely, (1.3 ± 0.2) × 10−2 s−1 at 20 °C, and supports the proposed nitrosylation mechanism.
Co-reporter:Svetlana Begel, Frank W. Heinemann, Grzegorz Stopa, Grazyna Stochel, and Rudi van Eldik
Inorganic Chemistry 2011 Volume 50(Issue 9) pp:3946-3958
Publication Date(Web):March 23, 2011
DOI:10.1021/ic1023357
To elucidate the applicability and properties of ionic liquids (ILs) to serve as chemical reaction media for the activation of small molecules by transition-metal complexes, detailed kinetic and mechanistic studies were performed on the reversible binding of NO to FeCl2 dissolved in the IL 1-ethyl-3-methylimidazolium dicyanamide ([emim][dca]) as a solvent. We report, for the first time, the application of laser flash photolysis at ambient and high pressure to study the kinetics of this reaction in an IL. The kinetic data and activation parameters for the “on” and “off” reactions suggest that both processes follow a limiting dissociative (D) ligand substitution mechanism, in contrast to that reported for the same reaction in aqueous solution, where this well-known “brown-ring” reaction follows an interchange dissociative (Id) ligand substitution mechanism. The observed difference apparently arises from the participation of the IL anion as a N-donor ligand, as evidenced by the formation of polymeric [Fe(dca)3Cl]x[emim]2x chains in the solid state and verified by X-ray crystallography. In addition, infrared (IR), Mössbauer, and EPR spectra were recorded for the monomeric reaction product [Fe(dca)5NO]3− formed in the IL, and the parameters closely resemble those of the {FeNO}7 unit in other well-characterized nitrosyl complexes. It is concluded that its electronic structure is best described by the presence of a high-spin Fe(III) (S = 5/2) center antiferromagnetically coupled to NO− (S = 1), yielding the observed spin quartet ground state (St = 3/2).
Co-reporter:Matthias Schmeisser, Frank W. Heinemann, Peter Illner, Ralph Puchta, Achim Zahl, and Rudi van Eldik
Inorganic Chemistry 2011 Volume 50(Issue 14) pp:6685-6695
Publication Date(Web):June 13, 2011
DOI:10.1021/ic200594e
On the basis of 7Li NMR measurements, we have made detailed studies on the influence of the ionic liquids [emim][NTf2], [emim][ClO4], and [emim][EtSO4] on the complexation of Li+ by the bidentate N-donor ligands 2,2′-bipyridine (bipy) and 1,10-phenanthroline (phen). For each of the employed ionic liquids the NMR data implicate the formation of [Li(bipy)2]+ and [Li(phen)2]+, respectively. X-ray diffraction studies were performed to determine the coordination pattern in the solid state. In the case of [emim][ClO4] and [emim][EtSO4], crystal structures confirmed the NMR data, resulting in the complexes [Li(bipy)2ClO4] and [Li(phen)2EtSO4], respectively. On the contrary, the ionic liquid [emim][NTf2] generated the Ci symmetric, dinuclear, supramolecular cluster [Li(bipy)(NTf2)]2, where the individual Li+ centers were found to be bridged by two [NTf2] anions. Density functional theory (DFT)-calculations lead to further information on the effect of stacking on the coordination geometry of the Li+ centers.
Co-reporter:Stephanie Hochreuther, Ralph Puchta, and Rudi van Eldik
Inorganic Chemistry 2011 Volume 50(Issue 18) pp:8984-8996
Publication Date(Web):August 1, 2011
DOI:10.1021/ic201151h
A series of novel dinuclear platinum(II) complexes were synthesized with bidentate nitrogen donor ligands. The two platinum centers are connected by an aliphatic chain of variable length. The selected chelating ligand system should stabilize the complex toward decomposition. The pKa values and reactivity of four synthesized complexes, viz. [Pt2(N1,N4-bis(2-pyridylmethyl)-1,4-butanediamine)(OH2)4]4+ (4NNpy), [Pt2(N1,N6-bis(2-pyridylmethyl)-1,6-hexanediamine)(OH2)4]4+ (6NNpy), [Pt2(N1,N8-bis(2-pyridylmethyl)-1,8-octanediamine)(OH2)4]4+ (8NNpy), and [Pt2(N1,N10-bis(2-pyridylmethyl)-1,10-decanediamine)(OH2)4]4+ (10NNpy), were investigated. This system is of special interest because only little is known about the substitution behavior of dinuclear platinum complexes that contain a bidentate chelate that forms part of the aliphatic bridging ligand. Spectrophotometric acid–base titrations were performed to determine the pKa values of the coordinated water ligands. The substitution of coordinated water by thiourea was studied under pseudofirst-order conditions as a function of nucleophile concentration, temperature, and pressure, using stopped-flow techniques and UV–vis spectroscopy. The results for the dinuclear complexes were compared to those for the corresponding mononuclear reference complex [Pt(aminomethylpyridine)(OH2)2]2+ (monoNNpy), by which the effect of increasing the aliphatic chain length on the bridged complexes could be investigated. The results indicated that there is a clear interaction between the two platinum centers, which becomes weaker as the chain length between the metal centers increases. In addition, quantum chemical calculations were performed to support the interpretation and discussion of the experimental data.
Co-reporter:Stephanie Hochreuther, Ralph Puchta, and Rudi van Eldik
Inorganic Chemistry 2011 Volume 50(Issue 24) pp:12747-12761
Publication Date(Web):November 15, 2011
DOI:10.1021/ic2018902
A series of novel dinuclear platinum(II) complexes were synthesized containing a mixed nitrogen–sulfur donor bidentate chelate system in which the two platinum centers are connected by an aliphatic chain of variable length. The bidentate chelating ligands were selected to stabilize the complex toward decomposition. The pKa values and reactivity of the four synthesized complexes, namely, [Pt2(S1,S4-bis(2-pyridylmethyl)-1,4-butanedithioether)(OH2)4]4+ (4NSpy), [Pt2(S1,S6-bis(2-pyridylmethyl)-1,6-hexanedithioether)(OH2)4]4+ (6NSpy), [Pt2(S1,S8-bis(2-pyridylmethyl)-1,8-octanedithioether)(OH2)4]4+ (8NSpy), and [Pt2(S1,S10-bis(2-pyridylmethyl)-1,10-decanedithioether)(OH2)4]4+ (10NSpy), were investigated. This system is of special interest because only little is known about the substitution behavior of dinuclear platinum complexes that contain a bidentate chelate that forms part of the aliphatic bridging ligand. Moreover, the ligands as well as the dinuclear complexes were examined in terms of their cytotoxic activity, and the 10NSpy complex was found to be active. Spectrophotometric acid–base titrations were performed to determine the pKa values of all the coordinated water molecules. The substitution of coordinated water by thiourea was studied under pseudo-first-order conditions as a function of nucleophile concentration, temperature, and pressure, using stopped-flow techniques and UV–vis spectroscopy. The results for the dinuclear complexes were compared to those for the corresponding mononuclear reference complex [Pt(methylthiomethylpyridine)(OH2)2]2+ (Pt(mtp)), by which the effect of the increasing aliphatic chain length of the bridged complexes could be investigated. The results indicate that there is a clear interaction between the two platinum centers, which becomes weaker as the chain length between the metal centers increases. Furthermore, differences and similarities of the N,S-system were compared to the corresponding dinuclear N,N-system studied previously in our group. In addition, quantum chemical calculations were performed to support the interpretation and discussion of the experimental data.
Co-reporter:Debabrata Chatterjee, Ujjwal Pal, Sarita Ghosh and Rudi van Eldik
Dalton Transactions 2011 vol. 40(Issue 6) pp:1302-1306
Publication Date(Web):04 Jan 2011
DOI:10.1039/C0DT01444C
The kinetics of reduction of [RuIII(edta)pz]− (edta4− = ethylenediaminetetraacetate; pz = pyrazine) by thioamino acids (RSH = cysteine, glutathione) resulting in the formation of a red [RuII(edta)pz]2− species (λmax = 462 nm) has been studied spectrophotometrically using both conventional mixing and stopped-flow techniques. The time course of the reaction was followed as a function of [RSH], pH, temperature and pressure. Alkali metal ions were found to have a positive influence (K+ > Na+ > Li+) on the reaction rate. Kinetic data and activation parameters are interpreted in terms of a mechanism involving outer-sphere electron transfer between the reaction partners. A detailed reaction mechanism in agreement with the spectral and kinetic data is presented.
Co-reporter:Matthias Schmeisser;Philipp Keil;Jörg Stierstorfer;Axel König;Thomas M. Klapötke
European Journal of Inorganic Chemistry 2011 Volume 2011( Issue 31) pp:4862-4868
Publication Date(Web):
DOI:10.1002/ejic.201100529
Abstract
The colourless room-temperature ionic liquid 1-ethyl-3-methylimidazolium perchlorate, [emim][ClO4], was revisited for reasons of its hydrophilic character and its weak coordinating ability. In this study, [emim][ClO4] was synthesized from the cheaper ionic liquid 1-ethyl-3-methylimidazolium ethyl sulfate, [emim][EtSO4], by way of direct and indirect membrane-assisted anion metathesis. Both synthetic methods, as alternatives to the metathesis of silver perchlorate and [emim] halides, are compared and discussed in detail. In addition, some physicochemical parameters were determined to characterize the ionic liquid. With respect to the hazardous nature of organic perchlorates, the thermal stability as well as the impact and friction sensitivity of [emim][ClO4] were tested according to the UN Test Series 3a–3d.
Co-reporter:Mikhael Shushanyan;Dimitri E. Khoshtariya;Tatyana Tretyakova;Maya Makharadze
Biopolymers 2011 Volume 95( Issue 12) pp:852-870
Publication Date(Web):
DOI:10.1002/bip.21688
Abstract
Catalytic mechanisms of carboxypeptidase A (CPA) are well known for their diversity and the relative inaccessibility for a decisive comprehension. Recent encouraging attempts through modern computational techniques promoted new challenges for the complementary experimental endeavors. In this work, we have applied the stopped-flow technique and the method of reaction progress curve fitting to extract kinetic parameters for the CPA-catalyzed hydrolyses of smaller (typical) peptide and ester substrates, known for their strong activating/inhibiting impact, thus to which the traditional method of “initial rates” is not applicable. Our approach that innately implies the overall constancy of the affecter (substrate plus “active” product) concentration, made it possible to rigorously determine the physically meaningful “effective” values for the catalytic and Michaelis constants under diverse experimental conditions including variable temperature and urea or trimethylamine N-oxide concentrations. Analysis of the obtained results allowed for: (i) the further substantiation of diverse mechanistic patterns for archetypal specific peptide and ester substrates, (ii) testing and disclosure of intrinsic links between the stabilizing/destabilizing and activating/inhibiting effects for the important model enzyme, CPA, and (iii) tentative explanation of a distinct activating/inhibiting impact of these substrates through the strong specific interaction of their benzyl (Bz) moiety with the substrate binding S3 subsite of CPA. We have demonstrated that stabilization of CPA either through the interaction with an extra Bz moiety (belonging to another substrate or to the product) leads to the increase of its catalytic power with respect to the specific peptide substrate and to its decrease with respect to the counterpart ester substrate. We conjecture that the catalytic mechanisms operating in these two cases include: (a) the “promoted water” mechanism for the peptide substrate that, seemingly, provides the almost “perfect induced fit” (low-barrier conformational adaptation), and (b) presumably, the “anhydride intermediate” mechanism for the ester substrate that, anyway, requires substantial conformational rearrangement (in fact, “partial or local unfolding”) of the protein environment in the course of the rate-determining step. © 2011 Wiley Periodicals, Inc. Biopolymers 95: 852–870, 2011.
Co-reporter:Rudi van Eldik, Colin D. Hubbard
Coordination Chemistry Reviews 2010 Volume 254(3–4) pp:297-308
Publication Date(Web):February 2010
DOI:10.1016/j.ccr.2009.09.009
Reactions that occur too rapidly to be monitored by rapid reaction methods at temperatures at or close to ambient can be investigated kinetically by retarding their reaction rates employing very low temperatures. A selection of reactions studied by this approach (low-temperature stopped-flow spectrophotometry) is reported. Details of the reaction mechanisms have been revealed for peroxide activation involving iron(III) porphyrins and cytochrome P450, superoxide activation involving manganese(II) complexes and iron porphyrin complexes, and dioxygen activation and binding by model mono-, and dinuclear copper(I) complexes and dioxygen activation at mono-, and dinuclear non-heme iron complexes. A final section covers progress in unravelling the mechanism of carbon–hydrogen bond activation by platinum complexes.
Co-reporter:Katharina Dürr, Norbert Jux, Achim Zahl, Rudi van Eldik, and Ivana Ivanović-Burmazović
Inorganic Chemistry 2010 Volume 49(Issue 23) pp:11254-11260
Publication Date(Web):November 8, 2010
DOI:10.1021/ic102092h
The one-electron reduced iron(II)-dioxygen adduct, {FeII−O2}−, is known to be an important intermediate in the catalytic cycle of heme (mono)oxygenases. The same type of species, considered as FeIII-peroxo, can be formed in a direct reaction between a FeII center and superoxide. In a unique high-pressure study of the reaction between superoxide and the FeII complex of a crown ether porphyrin conjugate in dimethylsulfoxide (DMSO), the overall FeII-superoxide interaction mechanism could be visualized and the nature of all species that occur along the reaction coordinate could be clarified. The equilibrium between the low-spin and high-spin forms of the starting FeII complex was quantified, which turns out to be the actual activation step toward substitution and subsequent inner-sphere electron transfer reactions. The constructed reaction volume profile demonstrates that the reaction product consists of FeIII-peroxo and FeII-superoxo species that exist in equilibrium, which can better account for the versatile reactivity of {FeII−O2}− adducts toward different substrates.
Co-reporter:Debabrata Chatterjee, Kalyan Asis Nayak, Erika Ember and Rudi van Eldik
Dalton Transactions 2010 vol. 39(Issue 7) pp:1695-1698
Publication Date(Web):15 Dec 2009
DOI:10.1039/B920839A
Reported in this paper is the first example of a ruthenium complex, [RuIII(edta)(H2O)]− (edta = ethylenediaminetetra-acetate), that catalyzes the oxidation of hydroxyurea in the presence of H2O2, mimicking the action of peroxidase or catalase and shedding light on their possible mechanism of action.
Co-reporter:Simon Kern;Peter Illner;Svetlana Begel
European Journal of Inorganic Chemistry 2010 Volume 2010( Issue 29) pp:4658-4666
Publication Date(Web):
DOI:10.1002/ejic.201000417
Abstract
Detailed kinetic studies on the ligand-substitution reactions of the very labile [Pd(terpy)Cl]+ (terpy = 2,2′:6′,2″-terpyridine) complex have been performed for neutral and anionic nucleophiles in several imidazolium-based ionic liquids. The detailed substitution mechanisms derived from the obtained rate and activation parameters differ from those expected on the basis of data aquired in aqueous medium, since the selected ionic liquids significantly affect the nucleophilic attack of the entering ligand on the complex. The rate constant for the substitution reaction with thiourea as entering nucleophile is much larger in water, viz. (7.8 ± 0.2) × 105M–1 s–1, than in the ionic liquids [emim][NTf2] and [emim][EtOSO3], viz. (2.5 ± 0.1) × 104 and (1.8 ± 0.1) × 103M–1 s–1 at 25 °C, respectively, which can be accounted for in terms of the protective interaction of the anionic components of the ionic liquids with the electrophilic metal centre.
Co-reporter:Jovana Bogojeski;&x17d;ivadin D. Bugar&x10d;i&x107;;Ralph Puchta
European Journal of Inorganic Chemistry 2010 Volume 2010( Issue 34) pp:5439-5445
Publication Date(Web):
DOI:10.1002/ejic.201000654
Abstract
Substitution reactions of the complexes cis-[Pt(NH3)2Cl2], [Pt(SMC)Cl2]–, [Pt(en)Cl2], and [Pt(dach)Cl2], where SMC = S-methyl-L-cysteine, en = ethylenediamine and dach = 1,2-diaminocyclohexane, with selected biologically important ligands, viz. guanosine-5′-monophosphate (5′-GMP), L-histidine and 1,2,4-triazole, were studied. All reactions were studied in aqueous 25 mM Hepes buffer in the presence of 5 mM NaCl at pH = 7.2 under pseudo-first-order conditions as a function of concentration at 310 K by using UV/Vis spectrophotometry. Two consecutive reaction steps, which both depend on the nucleophile concentration, were observed in all cases. The second-order rate constants for both reaction steps indicate a decrease in the order [Pt(SMC)Cl2]– > cis-[Pt(NH3)2Cl2] > [Pt(en)Cl2] > [Pt(dach)Cl2]. DFT calculations (B3LYP/LANL2DZp) showed that the Pt–N7(Guo) adduct is more stable than the Pt–S(thioether) adduct for the studied complexes cis-[Pt(NH3)2Cl2], [Pt(SMC)Cl2]–, and [Pt(en)Cl2]. The calculations collectively support the experimentally observed substitution of thioethers bound to PtII complexes by N7(5′-GMP). Finally, this result could be the first to clearly show how much the Pt–N7(Gua) adduct is more stable than the Pt–S(thioether).
Co-reporter:Colin D. Hubbard, Rudi van Eldik
Inorganica Chimica Acta 2010 Volume 363(Issue 11) pp:2357-2374
Publication Date(Web):10 August 2010
DOI:10.1016/j.ica.2009.09.042
Co-reporter:Dimitri E. Khoshtariya;Tina D. Dolidze;Mikhael Shushanyan;Kathryn L. Davis;David H. Waldeck
PNAS 2010 Volume 107 (Issue 7 ) pp:2757-2762
Publication Date(Web):2010-02-16
DOI:10.1073/pnas.0910837107
The blue copper protein from Pseudomonas aeruginosa, azurin, immobilized at gold electrodes through hydrophobic interaction with alkanethiol self-assembled monolayers (SAMs)
of the general type [-S - (CH2)n - CH3] (n = 4, 10, and 15) was employed to gain detailed insight into the physical mechanisms of short- and long-range biomolecular
electron transfer (ET). Fast scan cyclic voltammetry and a Marcus equation analysis were used to determine unimolecular standard
rate constants and reorganization free energies for variable n, temperature (2–55 °C), and pressure (5–150 MPa) conditions. A novel global fitting procedure was found to account for the
reduced ET rate constant over almost five orders of magnitude (covering different n, temperature, and pressure) and revealed that electron exchange is a direct ET process and not conformationally gated. All
the ET data, addressing SAMs with thickness variable over ca. 12 Å, could be described by using a single reorganization energy (0.3 eV), however, the values for the enthalpies and volumes
of activation were found to vary with n. These data and their comparison with theory show how to discriminate between the fundamental signatures of short- and long-range
biomolecular ET that are theoretically anticipated for the adiabatic and nonadiabatic ET mechanisms, respectively.
Co-reporter:BasamM. Alzoubi Dr.;Ralph Puchta Dr.;Rudi vanEldik Dr.
Chemistry - A European Journal 2010 Volume 16( Issue 24) pp:7300-7308
Publication Date(Web):
DOI:10.1002/chem.200902264
Abstract
The water-exchange mechanisms of [Zn(H2O)4(L)]2+⋅2 H2O (L=imidazole, pyrazole, 1,2,4-triazole, pyridine, 4-cyanopyridine, 4-aminopyridine, 2-azaphosphole, 2-azafuran, 2-azathiophene, and 2-azaselenophene) have been investigated by DFT calculations (RB3LYP/6-311+G**). The results support limiting associative reaction pathways that involve the formation of six-coordinate intermediates [Zn(H2O)5(L)]2+⋅H2O. The basicity of the coordinated heterocyclic ligands shows a good correlation with the activation barriers, structural parameters, and stability of the transition and intermediate states.
Co-reporter:Peter Illner, Svetlana Begel, Simon Kern, Ralph Puchta and Rudi van Eldik
Inorganic Chemistry 2009 Volume 48(Issue 2) pp:588-597
Publication Date(Web):December 16, 2008
DOI:10.1021/ic801561x
The effect of several imidazolium-based ionic liquids on the rate and mechanism of the substitution reaction of [Pt(terpyridine)Cl]+ with thiocyanate was investigated as a function of thiocyanate concentration and temperature under pseudo-first-order conditions using stopped-flow and other kinetic techniques. The obtained rate constants and activation parameters showed good agreement with the ion-pair stabilization energies between the anions of the ionic liquids and the cationic Pt(II) complex derived from density-functional theory calculations (RB3LYP/LANL2DZp) and with parameters derived from the linear solvation energy relationship set by the Kamlet-Taft β parameter, which is a measure of a solvent’s hydrogen bonding acceptor ability. In general, the substitution reactions followed an associative mechanism as found for conventional solvents, but the observed rate constants showed a significant dependence on the nature of the anionic component of the ionic liquid used as solvent. The second order rate constant measured in [emim][NTf2] is 2000 times higher than the one measured in [emim][EtOSO3]. This difference is much larger than observed for a neutral entering nucleophile studied before.
Co-reporter:Ameli Dreher, Klaus Mersmann, Christian Näther, Ivana Ivanovic-Burmazovic, Rudi van Eldik and Felix Tuczek
Inorganic Chemistry 2009 Volume 48(Issue 5) pp:2078-2093
Publication Date(Web):February 3, 2009
DOI:10.1021/ic801952v
The reaction of [Mo(NNC5H10)(dppe)2] (BMo) with an excess of acid, HNEt3BPh4, is investigated applying temperature-dependent stopped-flow measurements. The kinetic data indicate a biphasic process with rate constants kobs(1) and kobs(2) which are both slower than the single rate constant reported by Henderson et al. (Henderson, R. A.; Leigh, G. J.; Pickett, C. J., J. Chem. Soc., Dalton Trans. 1989, 425−430). Moreover, both rate constants exhibit a linear dependence on the acid concentration with a large intercept which is attributed to acid-dependent and acid-independent components of each reaction phase, respectively. All four reaction channels exhibit temperature-dependent reaction rates. Furthermore, BMo and its Mo(IV) analogue [MoBr(NNC5H10)(dppe)2]Br (AMo) are characterized structurally and spectroscopically. Density-functional theory calculations are performed to locate possible barriers in the overall reaction scheme and determine their energies, providing additional information for the formulation of a mechanism. The temperature-dependent rate of N−N cleavage is explained by a revised mechanism which involves an α-protonated intermediate that is inert with respect to N−N cleavage and is generated from its β-protonated counterpart by a rapid 1,2-proton shift. The implications of these results with respect to N2 reduction in the Chatt cycle and the enzyme nitrogenase are discussed.
Co-reporter:Matthias Schmeisser and Rudi van Eldik
Inorganic Chemistry 2009 Volume 48(Issue 15) pp:7466-7475
Publication Date(Web):June 18, 2009
DOI:10.1021/ic9006247
To elucidate the applicability and effects of ionic liquids as reaction media for bioinorganic catalysis, detailed kinetic and mechanistic studies on the reversible binding of NO to the monohydroxo ligated iron(III) phorphyrin, (TMPS)FeIII(OH) were performed in the ionic liquid [emim][NTf2] as solvent. We report for the first time the determination of activation volumes via high pressure stopped flow methods in an ionic liquid. The studies clearly show that impurities of methylimidazole, present at the micromolar concentration level, can generate the 6-fold coordinated (TMPS)FeIII(OH)(MeIm) complex and lead to a complete changeover in mechanism from associatively activated for (TMPS)FeIII(OH) to dissociatively activated for (TMPS)FeIII(OH)(MeIm). NMR measurements on the chemical shift of the β-pyrrole protons revealed a spin state change from high spin (S = 5/2 for (TMPS)FeIII(OH)) to an intermediate spin-state (S = 5/2 and 3/2) following the coordination of methylimidazole. Because of the effect of the cationic component of the ionic liquid, FeIII(TMPS) also reacts with nitrite unlike the case in aqueous solution. Kinetic and thermodynamic studies on the reaction of (TMPS)FeIII(OH) with tetrabutylammonium nitrite allowed the determination of the equilibrium constant and thermodynamic parameters for the coordination of nitrite in [emim][NTf2].
Co-reporter:Ariane Brausam ; Joachim Maigut ; Roland Meier ; Petra Á. Szilágyi ; Hans-Jürgen Buschmann ; Werner Massa ; Zoltán Homonnay
Inorganic Chemistry 2009 Volume 48(Issue 16) pp:7864-7884
Publication Date(Web):July 20, 2009
DOI:10.1021/ic900834z
The crystal structure of the as-yet-unknown salt K[FeIII(cydta)(H2O)]·3H2O, where cydta = (±)-trans-1,2-cyclohexanediaminetetraacetate, has been resolved: orthorhombic space group Pbca with R1 = 0.0309, wR2 = 0.0700, and GOF = 0.99. There are two independent [FeIII(cydta)(H2O)]− anions in the asymmetric unit, and the ligand is (R,R)-cydta in both cases. The coordination polyhedron is a seven-coordinate capped trigonal prism where the quadrilateral face formed by the four ligand donor oxygen atoms is capped by the coordinated water molecule. The speciation of [FeIII(cydta)(H2O)]− in water was studied in detail by a combination of techniques: (i) Measurements of the pH dependence of the FeIII/IIcydta redox potentials by cyclic voltammetry enabled the estimation of the stability constants (0.1 M KNO3, 25 °C) of [FeIII(cydta)(H2O)]− (log βIII110 = 29.05 ± 0.01) and [FeII(cydta)(H2O)]2- (log βII110 = 17.96 ± 0.01) as well as pKIIIa1OH = 9.57 and pKIIa1H = 2.69. The formation enthalpy of [FeIII(cydta)(H2O)]− (△H° = −23 ± 1 kJ mol−1) was measured by direct calorimetry and is compared to the corresponding value for [FeIII(edta)(H2O)]− (△H° = −31 ± 1 kJ mol−1). (ii) pH-dependent spectrophotometric titrations of FeIIIcydta lead to pKIIIa1OH = 9.54 ± 0.01 for deprotonation of the coordinated water and a dimerization constant of log Kd = 1.07. These data are compared with those of FeIIIpdta (pdta = 1,2-propanediaminetetraacetate; pKIIIa1OH = 7.70 ± 0.01, log Kd = 2.28) and FeIIIedta (pKIIIa1OH = 7.52 ± 0.01, log Kd = 2.64). Temperature- and pressure-dependent 17O NMR measurements lead to the following kinetic parameters for the water-exchange reaction at [FeIII(cydta)(H2O)]− (at 298 K): kex = (1.7 ± 0.2) × 107 s−1, △H⧧ = 40.2 ± 1.3 kJ mol−1, △S⧧ = +28.4 ± 4.7 J mol−1 K−1, and △V⧧ = +2.3 ± 0.1 cm3 mol−1. A detailed kinetic study of the effect of the buffer, temperature, and pressure on the reaction of hydrogen peroxide with [FeIII(cydta)(H2O)]− was performed using stopped-flow techniques. The reaction was found to consist of two steps and resulted in the formation of a purple FeIII side-on-bound peroxo complex [FeIII(cydta)(η2-O2)]3−. The peroxo complex and its degradation products were characterized using Mössbauer spectroscopy. Formation of the purple peroxo complex is only observable above a pH of 9.5. Both reaction steps are affected by specific and general acid catalysis. Two different buffer systems were used to clarify the role of general acid catalysis in these reactions. Mechanistic descriptions and a comparison between the edta and cydta systems are presented. The first reaction step reveals an element of reversibility, which is evident over the whole studied pH range. The positive volume of activation for the forward reaction and the positive entropy of activation for the backward reaction suggest a dissociative interchange mechanism for the reversible end-on binding of hydrogen peroxide to [FeIII(cydta)(H2O)]−. Deprotonation of the end-on-bound hydroperoxo complex leads to the formation of a seven-coordinate side-on-bound peroxo complex [FeIII(cydta)(η2-O2)]3−, where one carboxylate arm is detached. [FeIII(cydta)(η2-O2)]3− can be reached by two different pathways, of which one is catalyzed by a base and the other by deprotonated hydrogen peroxide. For both pathways, a small negative volume and entropy of activation was observed, suggesting an associative interchange mechanism for the ring-closure step to the side-on-bound peroxo complex. For the second reaction step, no element of reversibility was found.
Co-reporter:Ariane Brausam ; Siegfried Eigler ; Norbert Jux
Inorganic Chemistry 2009 Volume 48(Issue 16) pp:7667-7678
Publication Date(Web):July 14, 2009
DOI:10.1021/ic9005955
A detailed study of the effect of pH, temperature, and pressure on the reaction of hydrogen peroxide with [FeIII(P8−)]7−, where P8− represents the octa anionic porphyrin, was performed using stopped-flow techniques. Depending on the pH, different high valent iron-oxo species were formed. At pH < 9 formation of a two-electron oxidized species [(porphyrin+•)FeIV=O] was observed. In contrast, at pH > 9 only the one electron oxidized species [(porphyrin)FeIV=O] was found to be present in solution. Under selected conditions at pH 8 it was possible to determine rate constants for both the coordination of hydrogen peroxide and subsequent heterolytic cleavage of the O−O bond. At pH 11 a composite rate constant for coordination of H2O2 and homolytic cleavage of the O−O bond could be measured. In addition, it was possible to determine the activation parameters for the overall reaction sequence leading to the formation of [(porphyrin)FeIV=O]. Careful analysis of the obtained data supports an associatively activated mechanism for the coordination of hydrogen peroxide. The catalytic properties of [FeIII(P8−)]7− in the presence of H2O2 were also investigated. Both high valent iron-oxo species turned out to be able to oxidize diammonium-2,2′-azinobis(3-ethylbenzothiazoline-6-sulfonate) (ABTS) to the radical cation ABTS+•. At higher hydrogen peroxide concentrations a reduced yield of ABTS+• was observed because of increased catalase activity of [FeIII(P8−)]7−. At high pH disproportionation of ABTS+• to ABTS and ABTS2+ occurred, which could be suppressed by an excess of unreacted ABTS. In slightly basic to acidic solutions this reaction did not play a role.
Co-reporter:Elena A. Vlasova, Natalya Hessenauer-Ilicheva, Denis S. Salnikov, Evgeny V. Kudrik, Sergei V. Makarov and Rudi van Eldik
Dalton Transactions 2009 (Issue 47) pp:10541-10549
Publication Date(Web):30 Oct 2009
DOI:10.1039/B906478H
A detailed study of the oxidation of L-ascorbic acid by dioxygen and nitrite in water at pH 5.8 and 7.0, catalyzed by the octasulfophenyltetrapyrazinoporphyrazine complex of cobalt(II), was carried out using conventional spectrophotometric, low-temperature and high-pressure stopped-flow techniques. The Co(II) complex activates L-ascorbic acid through an intramolecular one-electron oxidation step that involves the reduction of the octasulfophenyltetrapyrazinoporphyrazine. The reaction rate strongly depends on pH due to the different redox behaviour of the L-ascorbic acid/ascorbate species present in solution. Kinetic parameters for the different reaction steps of the catalytic process were determined. The final product of the reaction between L-ascorbic acid and nitrite was found to be nitrous oxide.
Co-reporter:Hanaa Asaad Gazzaz, Erika Ember, Achim Zahl and Rudi van Eldik
Dalton Transactions 2009 (Issue 43) pp:9486-9495
Publication Date(Web):14 Sep 2009
DOI:10.1039/B912911A
Rate and activation parameters for the complex-formation reaction of Ni2+ with 4-(2-pyridylazo)-N,N-dimethyl aniline (PADA) were studied as a function of pH in different buffers in both aqueous and sodium dodecyl sulfate (SDS) micelle solutions. In aqueous Tris buffer solution, the forward and backward rate constants increased with increasing pH, while the complex-formation constant decreased due to a larger increase in the backward rate constant. The activation entropy, ΔS#, and activation volume, ΔV#, changed with increasing pH from positive to negative values, suggesting an apparent changeover from a dissociative to a more associative mechanism. Complex-formation reactions with 2,2′-bipyridine in Tris buffer showed almost no increase in the forward and backward rate constants on increasing the pH, but the ΔS# and ΔV# values became more negative. N-ethylmorpholine buffer showed no pH effect on the rate constants and activation parameters. Water exchange reactions of aquated Ni2+ were also studied as a function of pH under the same conditions. The reported rate and activation parameters for water exchange in Tris and N-ethylmorpholine buffers are consistent with those found for the complex-formation reactions of Ni2+ with PADA. The observed pH and buffer effects for both the complex-formation and water exchange reactions of aquated Ni2+ can be accounted for in terms of the formation of a Ni2+-Tris complex in Tris buffer and general base catalysis by the buffer components. In SDS micelle solution, the complex-formation reaction with PADA was much faster than in aqueous solution, but the increase in rate constant with increasing pH was less significant, while ΔS# and ΔV# became more positive, pointing to a more dissociative mechanism. For SDS micelle solutions there was no effect on the water exchange rate constant or activation volume. Mechanistic interpretations are offered for all observed pH, buffer and medium effects.
Co-reporter:Tanja Soldatović, Živadin D. Bugarčić and Rudi van Eldik
Dalton Transactions 2009 (Issue 23) pp:4526-4531
Publication Date(Web):22 Apr 2009
DOI:10.1039/B822718G
The kinetics and mechanism of ligand substitution reactions of [Pt(SMC)Cl2] (SMC = S-methyl-L-cysteine) with biologically relevant ligands were studied as a function of chloride and nucleophile concentrations at pH 2.5 and 7.2. It was observed that the slope and intercept obtained from the linear dependence of the observed rate constant on the nucleophile concentration strongly depend on the [Cl−] for all the studied substitution reactions. At high [Cl−], the rate constant for the forward reaction is almost zero and that for the back reaction follows the order: L-met > GSH ∼ INO > 5′-GMP. Ion-pair formation between the positively charged Pt(II) complex and the chloride ion is suggested to account for the saturation kinetics observed for the back reaction. The results are discussed in terms of the mechanism of the anti-tumour activity of related platinum complexes.
Co-reporter:Peter Illner, Ralph Puchta, Frank W. Heinemann and Rudi van Eldik
Dalton Transactions 2009 (Issue 15) pp:2795-2801
Publication Date(Web):20 Feb 2009
DOI:10.1039/B820940E
The coordinative behaviour of the bis(trifluoromethylsulfonyl)amide (NTf2−) anion was studied in terms of its interaction with the labile [Pd(terpyridine)Cl]Cl complex. Among various attempts to coordinate NTf2− to the Pd(II) centre, the complexes [Pd(terpyridine)NO3]NTf2 and [Pd(terpyridine)H2O](NTf2)2 were isolated and characterized. The crystallographic data for these complexes are presented and the possible coordination of the NTf2− anion to the metal centre is discussed in reference to DFT predictions.
Co-reporter:Hakan Ertürk;Ralph Puchta
European Journal of Inorganic Chemistry 2009 Volume 2009( Issue 10) pp:1331-1338
Publication Date(Web):
DOI:10.1002/ejic.200801135
Abstract
The dinuclear PtII complexes bis{[(1R,2R)-(–)-1,2-diaminocyclohexane]chloroplatinum(II)}(μ-1,8-octanediamine), bis{[(1R,2R)-(–)-1,2-diaminocyclohexane]chloroplatinum(II)}(μ-1,10-decanediamine) and bis{[(1R,2R)-(–)-1,2-diaminocyclohexane]chloroplatin(II)}[μ-1,4-bis(3-pyridyl)buta-1,3-diyne] were synthesized. Acid-base titrations and concentration and temperature dependent measurements of the reactions with chloride and thiourea were performed to study the influence of the nature of the bridging ligand on the thermodynamic and kinetic properties of the complexes. The reactions with chloride and thiourea were followed under pseudo-first-order conditions by stopped-flow and UV/Vis spectrophotometry. The results indicate that the bridging ligand has an influence on the reactivity and stability of the complexes towards nucleophiles as well as on possible electronic interactions between the two PtII centres. The experimental results are discussed in reference to structures obtained by DFT (BP86/LACVP*) calculations. (© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2009)
Co-reporter:Tanja Soldatovi&x107;;Mohamed Shoukry;Ralph Puchta;&x17d;ivadin D. Bugar&x10d;i&x107;
European Journal of Inorganic Chemistry 2009 Volume 2009( Issue 15) pp:2261-2270
Publication Date(Web):
DOI:10.1002/ejic.200801229
Abstract
The kinetics and mechanism of the complex-formation reactions of [Pd(AEP)(H2O)]2+, where AEP stands for 1-(2-aminoethyl)piperazine, with biologically relevant ligands were studied as a function of selected nucleophiles and pH. The reactivity of the ligands follows the sequence L-methionine > guanosine-5′-monophosphate > glycine > inosine >> glutathione. The substitution reactions with glutathione showed two reaction steps in which the first step involves coordination through nitrogen and depends on the nucleophile concentration, whereas the second step involves intramolecular isomerization from N- to S-bonded glutathione and does not depend on the nucleophile concentration. The stoichiometry and stability constants of the formed complexes are also reported, and the concentration distribution of the various complex species was evaluated as a function of pH. The results are discussed in terms of the mechanism of antitumor activity of related platinum complexes.(© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2009)
Co-reporter:&x141;ukasz Orze&x142;;Leszek Fiedor;Gra&x17c;yna Stochel
European Journal of Inorganic Chemistry 2009 Volume 2009( Issue 16) pp:2393-2406
Publication Date(Web):
DOI:10.1002/ejic.200800662
Abstract
Chlorophyll a and pheophytin a have been treated with various metal salts in solution. As a result, metallo derivatives of photosynthetic pigments were formed, or the chlorine ring was oxidized. These studies showed that both effects can be caused by CuII ions depending on the redox potential of its complexes formed in specific media. In this report the results of studies on metal insertion and exchange reactions are presented. According to the common use of the “acetate method” in the synthesis of metallotetrapyrroles, and the fact that acetates coordinated to CuII ions protect chlorophylls from oxidative degradation, CuII acetate monohydrate was selected as main reagent. The metalation and transmetalation processes were investigated with spectroscopic (UV/Vis absorption and emission) and kinetic (conventional and high-pressure) techniques. It turned out that differences in solvent properties can significantly affect not only the rates (metalation), but also the course (transmetalation) of the reaction. The most pronounced effect was found for the transmetalation reaction in acetonitrile, which was terminated at the very initial stage. As shown elsewhere, some special agents/factors, such as excess of acetate, are required to facilitate the dissociation of Mg2+ and to push the reaction forward towards formation of the CuII derivative of chlorophyll. The mechanisms of two other reactions are proposed on the basis of the determined activation parameters, which correspond to the general mechanisms for metalation and transmetalation of tetrapyrroles with sitting-atop and bimetallic intermediates, respectively. The bimetallic complex of chlorophyll in methanol is probably the most stable complex of this type observed ever. It confirms that spontaneous exchange of Mg2+ by CuII occurring in plants must result from specific conditions and/or involve contribution from other components of the photosystem.(© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2009)
Co-reporter:Natalya Hessenauer-Ilicheva;Alicja Franke Dr.;Dominik Meyer;Wolf-D. Woggon Dr.;Rudi vanEldik Dr.
Chemistry - A European Journal 2009 Volume 15( Issue 12) pp:2941-2959
Publication Date(Web):
DOI:10.1002/chem.200801423
Co-reporter:DimitriE. Khoshtariya ;TinaD. Dolidze Dr.;Rudi vanEldik Dr.
Chemistry - A European Journal 2009 Volume 15( Issue 21) pp:5254-5262
Publication Date(Web):
DOI:10.1002/chem.200802450
Co-reporter:Alicja Franke Dr.;Maria Wolak Dr. ;Rudi vanEldik Dr.
Chemistry - A European Journal 2009 Volume 15( Issue 39) pp:10182-10198
Publication Date(Web):
DOI:10.1002/chem.200900453
Abstract
The present study focuses on the oxidation of the water-soluble and water-insoluble iron(III)–porphyrin complexes [FeIII(TMPS)] and [FeIII(TMP)] (TMPS=meso-tetrakis(2,4,6-trimethyl-3-sulfonatophenyl)porphyrinato, TMP=meso-tetrakis(2,4,6-trimethylphenyl)porphyrinato), respectively, by meta-chloroperoxybenzoic acid (m-CPBA) in aqueous methanol and aqueous acetonitrile solutions of varying acidity. With the application of a low-temperature rapid-scan UV/Vis spectroscopic technique, the complete spectral changes that accompany the formation and decomposition of the primary product of OO bond cleavage in the acylperoxoiron(III)–porphyrin intermediate [(P)FeIIIOOX] (P=porphyrin) were successfully recorded and characterized. The results clearly indicate that the OO bond in m-CPBA is heterolytically cleaved by the studied iron(III)–porphyrin complexes independent of the acidity of the reaction medium. The existence of two different oxidation products under acidic and basic conditions is suggested not to be the result of a mechanistic changeover in the mode of OO bond cleavage on going from low to high pH values, but rather the effect of environmental changes on the actual product of the OO bond cleavage in [(P)FeIIIOOX]. The oxoiron(IV)–porphyrin cation radical formed as a primary oxidation product over the entire pH range can undergo a one- or two-electron reduction depending on the selected reaction conditions. The present study provides valuable information for the interpretation and improved understanding of results obtained in product-analysis experiments.
Co-reporter:Natalya Hessenauer-Ilicheva;Alicja Franke Dr.;Maria Wolak Dr.;Tsunehiko Higuchi Dr.;Rudi vanEldik Dr.
Chemistry - A European Journal 2009 Volume 15( Issue 45) pp:12447-12459
Publication Date(Web):
DOI:10.1002/chem.200901712
Abstract
Kinetic and mechanistic studies on the formation of an oxoiron(IV) porphyrin cation radical bearing a thiolate group as proximal ligand are reported. The SR complex, a functional enzyme mimic of P450, was oxidized in peroxo-shunt reactions under different experimental conditions with variation of solvent, temperature, and identity and excess of oxidant in the presence of different organic substrates. Through the application of a low-temperature rapid-scan stopped-flow technique, the reactive intermediates in the SR catalytic cycle, such as the initially formed SR acylperoxoiron(III) complex and the SR high-valent iron(IV) porphyrin cation radical complex [(SR.+)FeIVO], were successfully identified and kinetically characterized. The oxidation of the SR complex under catalytic conditions provided direct spectroscopic information on the reactivity of [(SR.+)FeIVO] towards the oxidation of selected organic substrates. Because the catalytically active species is a synthetic oxoiron(IV) porphyrin cation radical bearing a thiolate proximal group, the effect of the strong electron donor ligand on the formation and reactivity/stability of the SR high-valent iron species is addressed and discussed in the light of the reactivity pattern observed in substrate oxygenation reactions catalyzed by native P450 enzyme systems.
Co-reporter:Christoph Fertinger;Natalya Hessenauer-Ilicheva Dr.;Alicja Franke Dr. ;Rudi vanEldik Dr.
Chemistry - A European Journal 2009 Volume 15( Issue 48) pp:13435-13440
Publication Date(Web):
DOI:10.1002/chem.200901804
Abstract
The iron(III) meso-tetramesitylporphyrin complex is a good biomimetic to study the catalytic reactions of cytochrome P450. All of the three most discussed reactive intermediates concerning P450 catalysis (namely, Cpd 0, Cpd I, and Cpd II) can be selectively produced, identified, and stabilized for many minutes in solution at low temperature by choosing appropriate reaction conditions. In this way, their reactivity against various substrates was determined by utilizing low-temperature rapid-scan UV/Vis spectroscopy. Since all reactive intermediates are derived from a single model complex, the results of these kinetic measurements provide for the first time a full comparability of the determined rate constants for the three intermediates. The rate constants reveal a significant dependence of the reactivity on the type of reaction (e.g., oxygenation, hydrogen abstraction, or hydride transfer), which closely correlates with the chemical nature of Cpds 0, I, and II. The detailed knowledge of the reactivity of these intermediates provides a valuable tool to evaluate their particular role in biological systems.
Co-reporter:Tina D. Dolidze, Dimitri E. Khoshtariya, Peter Illner and Rudi van Eldik
Chemical Communications 2008 (Issue 18) pp:2112-2114
Publication Date(Web):25 Feb 2008
DOI:10.1039/B719787J
Proven electrochemical approaches were applied to study heterogeneous electron transfer (ET) between selected redox couples and gold electrodes modified with alkanethiol self-assembled monolayers (SAMs), using the room-temperature ionic liquid (RTIL) [bmim][NTf2] as reaction medium; ferrocene as freely diffusing redox probe in the RTIL was tested for ET through both thin (butanethiol) and thick (dodecanethiol) assemblages at pressures up to 150 MPa; well behaved kinetic patterns and reproducibility of data were demonstrated for ET within the unique Au/SAM/RTIL arrays.
Co-reporter:Svetlana Begel, Peter Illner, Simon Kern, Ralph Puchta and Rudi van Eldik
Inorganic Chemistry 2008 Volume 47(Issue 16) pp:7121-7132
Publication Date(Web):July 12, 2008
DOI:10.1021/ic702420v
The effect of several imidazolium-based ionic liquids on the mechanism of a classical ligand substitution reaction of [Pt(terpyridine)Cl]+ with thiourea was investigated. A detailed kinetic study as a function of the nucleophile concentration and temperature was undertaken under pseudo-first-order conditions using stopped-flow techniques. Polarity measurements were performed for the employed ionic liquids on the basis of solvatochromic effects, and they show similarities with conventional polar solvents. Density-functional theory calculations (RB3LYP/LANL2DZp) were employed to predict the ion-pair stabilization energy between the ionic components of the ionic liquids and/or between the anions of the ionic liquids and the cationic PtII complex. These data illustrate how the anions of the ionic liquids can affect the investigated substitution reaction. In general, the substitution mechanism in ionic liquids was found to have an associative character similar to that in conventional solvents. The observed deviations reflect the influence of the ionic liquid on the interaction between the anionic component of the liquid and the positively charged complex.
Co-reporter:Alexander Theodoridis ; Joachim Maigut ; Ralph Puchta ; Evgeny V. Kudrik
Inorganic Chemistry 2008 Volume 47(Issue 8) pp:2994-3013
Publication Date(Web):March 20, 2008
DOI:10.1021/ic702041g
The complex [iron(III) (octaphenylsulfonato)porphyrazine]5−, FeIII(Pz), was synthesized. The pKa values of the axially coordinated water molecules were determined spectrophotometrically and found to be pKa1 = 7.50 ± 0.02 and pKa2 = 11.16 ± 0.06. The water exchange reaction studied by 17O NMR as a function of the pH was fast at pH = 1, kex = (9.8 ± 0.6) × 106 s−1 at 25 °C, and too fast to be measured at pH = 10, whereas at pH = 13, no water exchange reaction occurred. The equilibrium between mono- and diaqua FeIII(Pz) complexes was studied at acidic pH as a function of the temperature and pressure. Complex-formation equilibria with different nucleophiles (Br− and pyrazole) were studied in order to distinguish between a five- (in the case of Br−) or six-coordinate (in the case of pyrazole) iron(III) center. The kinetics of the reaction of FeIII(Pz) with NO was studied as a model ligand substitution reaction at various pH values. The mechanism observed is analogous to the one observed for iron(III) porphyrins and follows an Id mechanism. The product is (Pz)FeIINO+, and subsequent reductive nitrosylation usually takes place when other nucleophiles like OH− or buffer ions are present in solution. FeIII(Pz) also activates hydrogen peroxide. Kinetic data for the direct reaction of hydrogen peroxide with the complex clearly indicate the occurrence of more than one reaction step. Kinetic data for the catalytic decomposition of the dye Orange II by H2O2 in the presence of FeIII(Pz) imply that a catalytic oxidation cycle is initiated. The peroxide molecule first coordinates to the iron(III) center to produce the active catalytic species, which immediately oxidizes the substrate. The influence of the catalyst, oxidant, and substrate concentrations on the reaction rate was studied in detail as a function of the pH. The rate increases with increasing catalyst and peroxide concentrations but decreases with increasing substrate concentration. At low pH, the oxidation of the substrate is not complete because of catalyst decomposition. The observed kinetic traces at pH = 10 and 12 for the catalytic cycle could be simulated on the basis of the obtained kinetic data and the proposed reaction cycle. The experimental results are in good agreement with the simulated ones.
Co-reporter:Joachim Maigut ; Roland Meier ; Achim Zahl
Inorganic Chemistry 2008 Volume 47(Issue 13) pp:5702-5719
Publication Date(Web):May 30, 2008
DOI:10.1021/ic702421z
Because of our interest in evaluating a possible relationship between complex dynamics and water exchange reactivity, we performed 1H NMR studies on the paramagnetic aminopolycarboxylate complexes FeII−TMDTA and FeII−CyDTA and their diamagnetic analogues ZnII−TMDTA and ZnII−CyDTA. Whereas a fast Δ−Λ isomerization was observed for the TMDTA species, no acetate scrambling between in-plane and out-of-plane positions is accessible for any of the CyDTA complexes because the rigid ligand backbone prevents any configurational changes in the chelate system. In variable-temperature 1H NMR studies, no evidence of spectral coalescence due to nitrogen inversion was found for any of the complexes in the available temperature range. The TMDTA complexes exhibit the known solution behavior of EDTA, whereas the CyDTA complexes adopt static solution structures. Comparing the exchange kinetics of flexible EDTA-type complexes and static CyDTA complexes appears to be a suitable method for evaluating the effect of ligand dynamics on the overall reactivity. In order to assess information concerning the rates and mechanism of water exchange, we performed variable-temperature and -pressure 17O NMR studies of NiII−CyDTA, FeII−CyDTA, and MnII−CyDTA. For NiII−CyDTA, no significant effects on line widths or chemical shifts were apparent, indicating either the absence of any chemical exchange or the existence of a very small amount of the water-coordinated complex in solution. For [FeII(CyDTA)(H2O)]2− and [MnII(CyDTA)(H2O)]2−, exchange rate constant values of (1.1 ± 0.3) × 106 and (1.4 ± 0.2) × 108 s−1, respectively, at 298 K were determined from fits to resonance-shift and line-broadening data. A relationship between chelate dynamics and reactivity seems to be operative, since the CyDTA complexes exhibited significantly slower reactions than their EDTA counterparts. The variable-pressure 17O NMR measurements for [MnII(CyDTA)(H2O)]2− yielded an activation volume of +9.4 ± 0.9 cm3 mol−1. The mechanism is reliably assigned as a dissociative interchange (Id) mechanism with a pronounced dissociation of the leaving water molecule in the transition state. In the case of [FeII(CyDTA)(H2O)]2−, no suitable experimental conditions for variable-pressure measurements were accessible.
Co-reporter:Joachim Maigut ; ; Roland Meier ; ;
Inorganic Chemistry 2008 Volume 47(Issue 14) pp:6314-6321
Publication Date(Web):June 18, 2008
DOI:10.1021/ic800378c
The promising BioDeNO x process for NO removal from gaseous effluents suffers from an unsolved problem that results from the oxygen sensitivity of the Fe II−aminopolycarboxylate complexes used in the absorber unit to bind NO(g). The utilized [Fe II(EDTA)(H 2O)] 2− complex is extremely oxygen sensitive and easily oxidized to give a totally inactive [Fe III(EDTA)(H 2O)] − species toward the binding of NO(g). We found that an in situ formed, less-oxygen-sensitive mixed-ligand complex, [Fe II(EDTA)(F)] 3−, still reacts quantitatively with NO(g). The formation constant for the mixed ligand complex was determined spectrophotometrically. For [Fe III(EDTA)(F)] 2− we found log K M L F F = 1.7 ± 0.1. The [Fe II(EDTA)(F)] 3− complex has a smaller value of log K M L F F = 1.3 ± 0.2. The presence of fluoride does not affect the reversible binding of NO(g). Even over extended periods of time and fluoride concentrations of up to 1.0 M, the nitrosyl complex does not undergo any significant decomposition. The [Fe III(EDTA)(NO −)] 2− complex releases bound NO on passing nitrogen through the solution to form [Fe II(EDTA)(H 2O)] 2− almost completely. A reaction cycle is feasible in which fluoride inhibits the autoxidation of [Fe II(EDTA)(H 2O)] 2− during the reversible binding of NO(g).
Co-reporter:Hakan Ertürk, Joachim Maigut, Ralph Puchta and Rudi van Eldik
Dalton Transactions 2008 (Issue 20) pp:2759-2766
Publication Date(Web):04 Apr 2008
DOI:10.1039/B718177A
Complexes of the type [Pt2(N,N,N′,N′-tetrakis(2-pyridylmethyl)diamine(H2O)2]4+ and [Pt2(N,N,N′,N′-tetrakis(2-pyridylmethyl)diamine(Cl)2]2+ were used to study their reactions with a series of bio-relevant nucleophiles, viz.thiourea, L-methionine and guanosine-5′-monophosphate (5′-GMP2−) as a function of nucleophile concentration and temperature. The reactions with the sulfur containing nucleophiles (thiourea and L-methionine) were followed under pseudo-first-order conditions by stopped-flow and UV-Vis spectrophotometry. The reaction with 5′-GMP2− was carried out under second order conditions and studied by NMR spectroscopy. The results indicate that the bridged dinuclear complexes remain intact after coordination of the studied nucleophiles for an extended period of time, which differs significantly from that reported for other multinuclear platinum complexes in the literature.
Co-reporter:Joachim Maigut;Rol Meier;Achim Zahl
Magnetic Resonance in Chemistry 2008 Volume 46( Issue S1) pp:S94-S99
Publication Date(Web):
DOI:10.1002/mrc.2322
Abstract
The effect of temperature and pressure on the water exchange reaction of [FeII(NTA)(H2O)2]− and [FeII(BADA)(H2O)2]− (NTA = nitrilotriacetate; BADA = β-alanindiacetate) was studied by 17O NMR spectroscopy. The [FeII(NTA)(H2O)2]− complex showed a water exchange rate constant, kex, of (3.1 ± 0.4) × 106 s−1 at 298.2 K and ambient pressure. The activation parameters ΔH≠, ΔS≠ and ΔV≠ for the observed reaction are 43.4 ± 2.6 kJ mol−1, + 25 ± 9 J K−1 mol−1 and + 13.2 ± 0.6 cm3 mol−1, respectively. For [FeII(BADA)(H2O)2]−, the water exchange reaction is faster than for the [FeII(NTA)(H2O)2]− complex with kex = (7.4 ± 0.4) × 106 s−1 at 298.2 K and ambient pressure. The activation parameters ΔH≠, ΔS≠ and ΔV≠ for the water exchange reaction are 40.3 ± 2.5 kJ mol−1, + 22 ± 9 J K−1 mol−1 and + 13.3 ± 0.8 cm3 mol−1, respectively. The effect of pressure on the exchange rate constant is large and very similar for both systems, and the numerical values for ΔV≠ suggest in both cases a limiting dissociative (D) mechanism for the water exchange process. Copyright © 2008 John Wiley & Sons, Ltd.
Co-reporter:Alicja Franke Dr.;Christoph Fertinger ;Rudi vanEldik
Angewandte Chemie International Edition 2008 Volume 47( Issue 28) pp:5238-5242
Publication Date(Web):
DOI:10.1002/anie.200800907
Co-reporter:Malgorzata Brindell;Iwona Stawoska
JBIC Journal of Biological Inorganic Chemistry 2008 Volume 13( Issue 6) pp:909-918
Publication Date(Web):2008 August
DOI:10.1007/s00775-008-0378-3
A systematic study of the reduction of (ImH)[trans-RuCl4(dmso)(Im)] (NAMI-A; dmso is dimethyl sulfoxide, Im is imidazole), a promising antimetastasing agent, by l-ascorbic acid under physiological conditions is reported. Under blood plasma conditions (pH 7.4, 0.1–0.15 M NaCl , 37 °C) the rapid reduction of trans-[RuIIICl4(dmso)(Im)]− results in the formation of trans-[RuIICl4(dmso)(Im)]2− within seconds, and is followed by successive dissociation of the chloride ligands, whereas neither dmso nor imidazole ligands are released during the reaction. Under our experimental conditions, the formation of the ascorbate dianion is the rate-determining step, and once it has formed it reacts rapidly with NAMI-A. Moreover, the NAMI-A complex is very unstable at physiological pH (7.4); therefore, the hydrolysis of NAMI-A cannot be excluded as a competing reaction. During hydrolysis, aquated derivatives via stepwise dissociation of chloride and dmso ligands are formed, and most of these species have a higher redox potential and are expected to be even more easily reduced by ascorbic acid. Thus, it is very likely that the reduced form of NAMI-A or the reduction products of its hydrolytic derivatives react with albumin. The reaction of reduced NAMI-A with human serum albumin leads to the formation of stable adducts, with a binding efficiency very similar to that of the parent complex, viz., 3.2 ± 0.3 and 4.0 ± 0.4 mol of Ru(II) and Ru(III) per mole of albumin, respectively, however with a significantly higher reactivity.
Co-reporter:Małgorzata Brindell, Iwona Stawoska, Łukasz Orzeł, Przemysław Łabuz, Grażyna Stochel, Rudi van Eldik
Biochimica et Biophysica Acta (BBA) - Proteins and Proteomics 2008 Volume 1784(Issue 11) pp:1481-1492
Publication Date(Web):November 2008
DOI:10.1016/j.bbapap.2008.08.006
This article focuses on the application of high pressure laser flash photolysis for studies on selected hemoprotein reactions with the objective to establish details of the underlying reaction mechanisms. In this context, particular attention is given to the reactions of small molecules such as dioxygen, carbon monoxide, and nitric oxide with selected hemoproteins (hemoglobin, myoglobin, neuroglobin and cytochrome P450cam), as well as to photo-induced electron transfer reactions occurring in hemoproteins (particularly in various types of cytochromes). Mechanistic conclusions based on the interpretation of the obtained activation volumes are discussed in this account.
Co-reporter:Alicja Franke Dr.;Christoph Fertinger ;Rudi vanEldik
Angewandte Chemie 2008 Volume 120( Issue 28) pp:5316-5320
Publication Date(Web):
DOI:10.1002/ange.200800907
Co-reporter:Hakan Ertürk, Andreas Hofmann, Ralph Puchta and Rudi van Eldik
Dalton Transactions 2007 (Issue 22) pp:2295-2301
Publication Date(Web):04 Apr 2007
DOI:10.1039/B700770C
A series of dinuclear Pt(II) complexes of the type [Pt2(N,N,N′,N′-tetrakis(2-pyridylmethyl)diamine(H2O)2]4+ were synthesized. Acid–base titrations, and concentration and temperature dependent stopped-flow measurements of the reaction with chloride were performed to study the thermodynamic and kinetic behaviour of the dinuclear bridged complexes. The results indicate that there is a clear interaction between the two Pt(II) centres, which becomes weaker as the aliphatic chain increases in length. From a certain chain length onwards, the Pt(II) centres become independent of each other and exhibit identical thermodynamic and kinetic properties. The experimental results are discussed in reference to structures obtained by DFT (BP86/LACVP*) calculations.
Co-reporter:Azza Shoukry, Malgorzata Brindell and Rudi van Eldik
Dalton Transactions 2007 (Issue 37) pp:4169-4174
Publication Date(Web):31 Jul 2007
DOI:10.1039/B706856E
The interaction of the palladium(II) complex [Pd(Pip)(H2O)2]2+, where Pip is piperazine, with a series of biologically relevant nucleophiles including guanosine-5′-monophosphate, L-methionine and thiourea was studied under pseudo-first-order conditions as a function of nucleophile concentration and temperature, using UV-Vis spectrophotometric and stopped-flow techniques. The reactions were found to occur in two subsequent steps. For the sulfur donor containing nucleophiles thiourea and L-methionine, a third reaction step, the displacement of the labilized amine, as a result of the strong trans-effect of S-donor ligands, was observed. The activation parameters for all reactions studied suggest an associative substitution mechanism.
Co-reporter:Ewa Pasgreta;Ralph Puchta;Achim Zahl
European Journal of Inorganic Chemistry 2007 Volume 2007(Issue 19) pp:
Publication Date(Web):10 MAY 2007
DOI:10.1002/ejic.200601169
Kinetic studies on Li+ exchange between the cryptands C222 and C221 and acetone as solvent were performed as a function of ligand-to-metal ratio, temperature and pressure using 7Li NMR spectroscopy. Temperature and pressure dependence measurements were performed in the presence of an excess of Li+. The rate and activation parameters for the exchange process are: C222: k298 = (1.7 ± 0.5) × 103 s–1, ΔH# = 24 ± 1 kJ mol–1 and ΔS# = –105 ± 4 J K–1 mol–1; C221: k298 = 2.24 ± 0.09 s–1, ΔH# = 29 ± 2 kJ mol–1 and ΔS# = –141 ±5 J K–1 mol–1. The influence of pressure on the exchange rate is insignificant for both ligands, such that the activation volume is around zero within the experimental error limits. The reported activation parameters suggest that the exchange of Li+ between solvated and chelated ions follows an associative interchange mechanism.(© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2007)
Co-reporter:Ewa Pasgreta;Ralph Puchta;Achim Zahl
European Journal of Inorganic Chemistry 2007 Volume 2007(Issue 13) pp:
Publication Date(Web):23 MAR 2007
DOI:10.1002/ejic.200600930
Solutions of LiClO4 in solvent mixtures of acetonitrile and water, or acetonitrile and nitromethane, were studied by 7Li NMR spectroscopy. Measured chemical shifts indicate that the Li+ cation is coordinated by four acetonitrile molecules. In the binary water/acetonitrile mixture, water coordinates more strongly to Li+ than acetonitrile such that addition of water immediately leads to the formation of [Li(H2O)4]+. The solvent-exchange mechanism for [Li(L)4]+ (L = CH3CN and HCN) was studied by using DFT calculations (RB3LYP/6-311+G**). This process was found to follow a limiting associative mechanism involving the formation of relatively stable five-coordinate intermediates. The suggested mechanisms are discussed with reference to available experimental and theoretical data. (© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2007)
Co-reporter:Alicia Franke;Federico Roncaroli
European Journal of Inorganic Chemistry 2007 Volume 2007(Issue 6) pp:
Publication Date(Web):13 FEB 2007
DOI:10.1002/ejic.200790008
The cover picture shows how mechanistic insight into the activation of NO by biological and model metal complexes can be gained through volume profile analysis. The presented volume profile describes the binding of NO to the high-spin, five-coordinate FeIII centre in cytochrome P450 in the presence of camphor to form a low-spin (diamagnetic), six-coordinate FeII–NO+ product. Details are presented in the Microreview by R. van Eldik et al. on p. 773 ff.
Co-reporter:Alicja Franke;Federico Roncaroli
European Journal of Inorganic Chemistry 2007 Volume 2007(Issue 6) pp:
Publication Date(Web):2 FEB 2007
DOI:10.1002/ejic.200600921
In this review, the reactions of nitric oxide with selected Fe and Co complexes of biological and environmental importance are reviewed. Fundamental chemical kinetics and mechanisms of reactions that lead to the formation and decay of nitrosyl complexes are illustrated and discussed on the basis of studies on FeII chelates, FeII/III pentacyano complexes, FeIII porphyrins, cytochrome P450 and model complexes, CoII/III porphyrins and vitamin B12. Throughout the review, the focus is on mechanistic details of the binding of NO to and the release of NO from metal complexes and on the nature of the stable metal–NO complexes produced in solution. As will be seen for most of the presented reactions, the interaction of NO with the metal complex involves activation of the small NO molecule so that charge transfer occurs as a result of the formation of the metal–NO bond.(© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2007)
Co-reporter:Živadin D. Bugarčić;Jovana Rosić
JBIC Journal of Biological Inorganic Chemistry 2007 Volume 12( Issue 8) pp:1141-1150
Publication Date(Web):2007 November
DOI:10.1007/s00775-007-0283-1
The substitution reactions of [PtCl(bpma)]+, [PtCl(gly-met-S,N,N)], [Pt(bpma)(H2O)]2+ and [Pt(gly-met-S,N,N)(H2O)]+ [where bpma is bis(2-pyridylmethyl)amine and gly-met-S,N,N is glycylmethionine] with l-methionine, glutathione and guanosine 5′-monophosphate (5′-GMP) were studied in aqueous solutions in 0.10 M NaClO4 under pseudo-first-order conditions as a function of concentration and temperature using UV–vis spectrophotometry. The reactions of the chloro complexes were followed in the presence of 10 mM NaCl and at pH ~ 5, whereas the reactions of the aqua complexes were studied at pH 2.5. The [PtCl(bpma)]+ complex is more reactive towards the chosen nucleophiles than [PtCl(gly-met-S,N,N)]. Also, the aqua complexes are more reactive than the corresponding chloro complexes. The activation parameters for all the reactions studied suggest an associative substitution mechanism. The reactions of [PtCl(bpma)]+ and [PtCl(gly-met-S,N,N)] with 5′-GMP were studied by using 1H NMR spectroscopy at 298 K. The pKa value of the [Pt(gly-met-S,N,N)(H2O)]+ complex is 5.95. Density functional theory calculations (B3LYP/LANL2DZp) show that in all cases guanine coordination to the L3Pt fragment (L3 is terpyridine, bpma, diethylenetriamine, gly-met-S,N,N) is much more favorable than the thioether-coordinated form. The calculations collectively support the experimentally observed substitution of thioethers from Pt(II) complexes by N7-GMP. This study throws more light on the mechanistic behavior of platinum antitumor complexes.
Co-reporter:Nadine Summa;Tanja Soldatović
JBIC Journal of Biological Inorganic Chemistry 2007 Volume 12( Issue 4) pp:461-475
Publication Date(Web):2007 May
DOI:10.1007/s00775-006-0200-z
A set of three oxaliplatin derivatives containing 1,2-trans-R,R-diaminocyclohexane (dach) as a spectator ligand and different chelating leaving groups X–Y, viz., [Pt(dach)(O,O-cyclobutane-1,1-dicarboxylate)], or Pt(dach)(CBDCA), [Pt(dach)(N,O-glycine)]+, or Pt(dach)(gly), and [Pt(dach)(N,S-methionine)]+, or Pt(dach)(l-Met), where l-Met is l-methionine, were synthesized and the crystal structure of Pt(dach)(gly) was determined by X-ray diffraction. The effect of the leaving group on the reactivity of the resulting Pt(II) complexes was studied for the nucleophiles thiourea, glutathione (GSH) and l-Met under pseudo-first-order conditions as a function of nucleophile concentration and temperature, using UV–vis spectrophotometric techniques. 1H NMR spectroscopy was used to follow the substitution of the leaving group by guanosine 5′-monophosphate (5′-GMP2−) under second-order conditions. The rate constants indicate for all reactions a direct substitution of the X–Y chelate by the selected nucleophiles, thereby showing that the nature of the chelate, viz., O–O (CBDCA2−), N–O (glycine) or S–N (l-Met), respectively, plays an important role in the kinetic and mechanistic behavior of the Pt(II) complex. The k1 values for the reaction with thiourea, l-Met, GSH and 5′-GMP2− were found to be as follows (103k1, 37.5 °C, M−1 s−1): Pt(dach)(CBDCA) 61 ± 2, 21.6 ± 0.1, 23 ± 1, 0.352 ± 0.002; Pt(dach)(gly) 82 ± 3, 6.2 ± 0.2, 37 ± 1, 1.77 ± 0.01; Pt(dach)(l-Met) (thiourea, GSH) 62 ± 2, 24 ± 1. The activation parameters for all reactions studied suggest an associative substitution mechanism.
Co-reporter:Malgorzata Brindell;Dorota Piotrowska
JBIC Journal of Biological Inorganic Chemistry 2007 Volume 12( Issue 6) pp:809-818
Publication Date(Web):2007 August
DOI:10.1007/s00775-007-0234-x
A systematic study of the reduction of (ImH)[trans-RuCl4(dmso)(Im)] (NAMI-A;
dmso is dimethyl sulfoxide, Im is imidazole), a promising antimetastasing agent
entering phase II clinical trial, by l-ascorbic acid is reported. The rapid reduction of
trans-[RuIIICl4(dmso)(Im)]−
results in formation of trans-[RuIICl4(dmso)(Im)]2−
in acidic medium (pH = 5.0) and is followed by successive dissociation of the
chloride ligands, which cannot be suppressed even in the presence of a large
excess of chloride ions. The reduction of NAMI-A strongly depends on pH and is
accelerated on increasing the pH. Over the small pH range 4.9−5.1, the reaction
is quite pH-independent and the influence of temperature and pressure on the
reaction could be studied. On the basis of the reported activation parameters
and other experimental data, it is suggested that the redox process follows an
outer-sphere electron transfer mechanism. A small contribution from a parallel
reaction ascribed to inner-sphere reduction of aqua derivatives of NAMI-A, was
found to be favored by lower concentrations of the NAMI-A complex and higher
temperature. In the absence of an excess of chloride ions, the reduction
process is catalyzed by the Ru(II) products being formed. The reduction of
NAMI-A is also catalyzed by Cu(II) ions and the apparent catalytic rate
constant was found to be
1.5 × 106 M−2 s−1
at 25 °C.
Co-reporter:Ewa Pasgreta;Ralph Puchta;Michael Galle
Journal of Inclusion Phenomena and Macrocyclic Chemistry 2007 Volume 58( Issue 1-2) pp:81-88
Publication Date(Web):2007 June
DOI:10.1007/s10847-006-9125-y
Kinetic studies on Li+ exchange between the cryptands C222 and C221, and γ-butyrolactone as solvent were performed as a function of ligand-to-metal ratio, temperature and pressure using 7Li NMR. The thermal rate and activation parameters are: C222: k298 = (3.3 ± 0.8)×104 M−1 s−1, ΔH# = 35 ± 1 kJ mol−1 and ΔS# = −41 ± 3 J K−1 mol−1; C221: k298 = 105 ± 32 M−1 s−1, ΔH# = 48 ± 1 kJ mol−1 and ΔS# = −45 ± 2 J K−1 mol−1. Temperature and pressure dependence measurements were performed in the presence of an excess of Li+. The influence of pressure on the exchange rate is insignificant for both ligands, such that the value of activation volume is around zero within the experimental error limits. The activation parameters obtained in this study indicate that the exchange of Li+ between solvated and chelated Li+ ions follows an associative interchange mechanism.
Co-reporter:Maria Wolak Dr.;Rudi van Eldik Dr.
Chemistry - A European Journal 2007 Volume 13(Issue 17) pp:
Publication Date(Web):16 MAR 2007
DOI:10.1002/chem.200601148
The reactions of a water-soluble iron(III)–porphyrin, [meso-tetrakis(sulfonatomesityl)porphyrinato]iron(III), [FeIII(tmps)] (1), with m-chloroperoxybenzoic acid (mCPBA), iodosylbenzene (PhIO), and H2O2 at different pH values in aqueous methanol solutions at −35 °C have been studied by using stopped-flow UV/Vis spectroscopy. The nature of the porphyrin product resulting from the reactions with all three oxidants changed from the oxo–iron(IV)–porphyrin π-cation radical [FeIV(tmps.+)(O)] (1++) at pH<5.5 to the oxo–iron(IV)–porphyrin [FeIV(tmps)(O)] (1+) at pH>7.5, whereas a mixture of both species was formed in the intermediate pH range of 5.5–7.5. The observed reactivity pattern correlates with the E°′ versus pH profile reported for 1, which reflects pH-dependent changes in the relative positions of E°′ and E°′ for metal- and porphyrin-centered oxidation, respectively. On this basis, the pH-dependent redox equilibria involving 1++ and 1+ are suggested to determine the nature of the final products that result from the oxidation of 1 at a given pH. The conclusions reached are extended to water-insoluble iron(III)–porphyrins on the basis of literature data concerning the electrochemical and catalytic properties of [FeIII(P)(X)] species in nonaqueous solvents. Implications for mechanistic studies on [Fe(P)]-catalyzed oxidation reactions are briefly addressed.
Co-reporter:Ralph Puchta Dr.;Ewa Pasgreta Dr.;Michael Galle;Nico van Eikema Hommes Dr. Dr.;Achim Zahl Dr.
ChemPhysChem 2007 Volume 8(Issue 9) pp:1315-1320
Publication Date(Web):24 MAY 2007
DOI:10.1002/cphc.200600624
Solutions of LiClO4 in solvent mixtures consisting of dimethylsulfoxide (DMSO) and water, or DMSO and γ-butyrolactone, were studied by 7Li NMR spectroscopy (for complexation by cryptands in γ-butyrolactone as a solvent, see: E. Pasgreta, R. Puchta, M. Galle, N. J. R. van Eikema Hommes, A. Zahl, R. van Eldik, J. Incl. Phen., 2007, 58, 81–88). Chemical shifts indicate that the Li+ ion is coordinated by four DMSO molecules. In the binary solvent mixture of water and DMSO, no selective solvation is detected, thus indicating that on increasing the water content of the solvent mixture, DMSO is gradually displaced by water in the coordination sphere of Li+. The ligand-exchange mechanism of Li+ ions solvated by DMSO and water/DMSO mixtures was studied using DFT calculations. Ligand exchange on [Li(DMSO)4]+ was found to follow a limiting associative (A) mechanism. The displacement of coordinated H2O by DMSO in [Li(H2O)4]+ follows an associative interchange mechanism. The suggested mechanisms are discussed in reference to available experimental and theoretical data.
Co-reporter:Wolfgang Schiessl;Peter Pfeifer Dr. Dr. h. c. mult.
CHEMKON 2007 Volume 14(Issue 2) pp:84-89
Publication Date(Web):24 APR 2007
DOI:10.1002/ckon.200710056
Chemie ist kein beliebtes Fach. Befragungen unter Schülern gehen bis ins Jahr 1905 zurück und zeichnen ein einheitliches Bild [1]. Die mangelnde Popularität dieses Faches beschränkt sich jedoch nicht auf Schüler. Das Image der Chemie lässt in der gesamten Öffentlichkeit zu wünschen übrig. Allzu leicht bezieht man sich hier ausschließlich auf Chemie-Unfälle und negative Schlagzeilen. Es fehlt am Verständnis, inwieweit chemische Grundlagen jeden Bereich unseres täglichen Lebens durchdringen, von Atemluft über Lebensprozesse bis zu Zahnpasta. Tatsächlich gilt: „Chemie ist unser Leben!” [2]. Um hier Umdenken zu bewirken, stehen wir in der Verantwortung, vor allem Wissen um Chemie im Alltag zu vermitteln, sowohl Schülern, als auch der Öffentlichkeit. Angesichts der generell eher ablehnenden Haltung, ist dies jedoch kein leichtes Unterfangen. Gilt es doch als gesichert, dass Lernzuwachs wesentlich von der affektiven Haltung dem Inhalt gegenüber abhängt [3]. Edutainment verspricht hier einen interessanten Ansatz – unterhaltend und (quasi nebenbei) lehrreich zu sein. Der Begriff setzt sich zusammen aus „education” und „entertainment” und versucht eine Symbiose beider. Lehransätze dieser Art reichen mittlerweile von Lernsoftware bis zu aktuellen Fernsehsendungen, die Unterhaltung und Wissenstransfer zur besten Sendezeit kombinieren. Doch ist dies auch in einer abendfüllenden Veranstaltung live möglich?
Co-reporter:Ralph Puchta, Nico van Eikema Hommes, Roland Meier and Rudi van Eldik
Dalton Transactions 2006 (Issue 28) pp:3392-3395
Publication Date(Web):08 May 2006
DOI:10.1039/B602792J
Based on DFT calculations (RB3LYP/LANL2DZp), the unexpected single-line 1H NMR spectrum of ZnII(nta), nta = 2,2′,2″-nitrilotriacetate, can be ascribed to a non-dissociative enantiomerization process (δδδ
⇌
λλλ) from C3viaC3v to C3′ symmetry. The energy barrier is rather low and depends to a lesser extent on the nature of the co-ligand in [Zn(nta)X]2− (X: H−, CH3− NH2−, OCH3−, F−, Cl−, Br−, I−) and [Zn(nta)Y]− (Y: NCH, CO, N2, O(CH3)2), but more so on the overall charge of the complex. The energy barrier for enantiomerization of [Zn(nta)X]2− is between 5.7 and 6.7 kcal mol−1, and for [Zn(nta)Y]− between 2.2 and 3.1 kcal mol−1.
Co-reporter:Mario Gabričević, Erim Bešić, Mladen Biruš, Achim Zahl, Rudi van Eldik
Journal of Inorganic Biochemistry 2006 Volume 100(Issue 10) pp:1606-1613
Publication Date(Web):October 2006
DOI:10.1016/j.jinorgbio.2006.05.008
Hydroxyurea (HU) effectively reduces vanadium(V) into vanadium(IV) species (hereafter VV and VIV species, respectively) in acidic aqueous solution via the formation of a transient complex followed by an electron transfer process that includes the formation and subsequent fading out of a free radical, U (U ≡ H2N–C(=O)N(H)O). The electron paramagnetic resonance (EPR) spectra of U in H2O/D2O solutions suggest that the unpaired electron is located predominantly on the hydroxamate hydroxyl-oxygen atom. Visible and VIV–EPR spectroscopic data reveal HU as a two-electron donor, whereas formation of U, which reduces a second VV, indicates that electron transfer occurs in two successive one-electron steps. At the molarity ratio [VV]/[HU] = 2, the studied reaction can be formulated as: 2 VV + HU → 2 VIV + 0.98 CO2 + 0.44 N2O + 1.1 NH3 + 0.1 NH2OH. Lack of evidence for the formation of NO is suggested to be a consequence of the slow oxidation of HNO due to the too low reduction potential of the VV/VIV couple under the experimental conditions used.The nuclear magnetic resonance (51V-NMR) spectral data indicate protonation of (H2O)4VVO2+, and the protonation equilibrium constant was determined to be K = 0.7 M−1. Spectrophotometric titration data for the VV–HU system reveal formation of (H2O)2VVO(OH)U+ and (H2O)3VVOU2+. Their stability constants were calculated as K110 = 5 M−1 and K111 = 22 M−2, where the subscript digits refer to (H2O)4VVO2+, HU and H+, respectively.
Co-reporter:Ralph Puchta;Rol Meier;Nico J. R. van Eikema Hommes
European Journal of Inorganic Chemistry 2006 Volume 2006(Issue 20) pp:
Publication Date(Web):10 AUG 2006
DOI:10.1002/ejic.200600483
The enantiomerisation pathway for {[Pt(thiourea)4]}F+ [a model for the C4-symmetric [Pt(SU)4]SiF6 (SU = thiourea) complex] and derivatives is explored by density functional theory (B3LYP/LANL2DZp) und the activation barrier for the one-step process from C4 to C4′ via a C4 transition state is computed. The substitution of Pt2+ by Pd2+ and Ni2+ and the exchange of selenourea and tellurourea increase the barrier. ({[Pt(thiourea)4]}F+: 4.2 kcal/mol, {[Pd(thiourea)4]}F+: 4.5 kcal/mol, {[Ni(thiourea)4]}F+: 7.6 kcal/mol, {[Pt(selenourea)4]}F+: 5.3 kcal/mol, {[Pt(tellurourea)4]}F+: 8.8 kcal/mol). (© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2006)
Co-reporter:Wolfgang Schiessl;Carlos Dücker-Benfer Dr.;Nadine Summa;Ariane Brausam;Peter Pfeifer Dr. Dr. Dr. h. c.
Chemie in unserer Zeit 2006 Volume 40(Issue 3) pp:
Publication Date(Web):1 JUN 2006
DOI:10.1002/ciuz.200600383
Seit nunmehr 10 Jahren findet die Erlanger Chemie-Zaubervorlesung statt. In den letzen Jahren konnte durch die zunehmende Professionalisierung vor allem des zugrunde gelegten Informationskonzeptes die Ernsthaftigkeit dieser Veranstaltung unterstrichen werden. Wir verknüpfen Unterhaltung und Wissensvermittlung zu einem “Edutainment”-Konzept. Wir hoffen damit beispielhaft aufzuzeigen, wie sich Spaß und experimentelle Naturwissenschaften in einer Großveranstaltung live ergänzen können.
During the past 10 years the “Magic chemistry” lecture was conducted in Erlangen. Due to an increasing professionalization of the lecture and especially the underlying information concept, we could emphasize the seriousness of this event. We combine scientific education and entertainment to an “edutainment” concept. We hope to show in an exemplary way, how fun and experimental natural sciences can be combined in a big live event.
Co-reporter:Dimitri E. Khoshtariya, Tina D. Dolidze, David Sarauli,Rudi van Eldik
Angewandte Chemie International Edition 2006 45(2) pp:277-281
Publication Date(Web):
DOI:10.1002/anie.200502386
Co-reporter:Dimitri E. Khoshtariya Dr.;Tina D. Dolidze Dr.;Stefan Seifert Dr.;David Sarauli Dipl.-Phys.;Geoffrey Lee Dr. Dr.
Chemistry - A European Journal 2006 Volume 12(Issue 27) pp:
Publication Date(Web):6 SEP 2006
DOI:10.1002/chem.200690086
Co-reporter:Dimitri E. Khoshtariya Dr.;Tina D. Dolidze Dr.;Stefan Seifert Dr.;David Sarauli Dipl.-Phys.;Geoffrey Lee Dr. Dr.
Chemistry - A European Journal 2006 Volume 12(Issue 27) pp:
Publication Date(Web):3 AUG 2006
DOI:10.1002/chem.200600059
Combined kinetic (electrochemical) and thermodynamic (calorimetric) investigations were performed for an unbound (intact native-like) cytochrome c (CytC) freely diffusing to and from gold electrodes modified by hydroxyl-terminated self-assembled monolayer films (SAMs), under a unique broad range of experimental conditions. Our approach included: 1) fine-tuning of the charge-transfer (CT) distance by using the extended set of Au-deposited hydroxyl-terminated alkanethiol SAMs [-S-(CH2)n-OH] of variable thickness (n=2, 3, 4, 6, 11); 2) application of a high-pressure (up to 150 MPa) kinetic strategy toward the representative Au/SAM/CytC assemblies (n=3, 4, 6); 3) complementary electrochemical and microcalorimetric studies on the impact of some stabilizing and denaturing additives. We report for the first time a mechanistic changeover detected for “free” CytC by three independent kinetic methods, manifested through 1) the abrupt change in the dependence of the shape of the electron exchange standard rate constant (ko) versus the SAM thickness (resulting in a variation of estimated actual CT range within ca. 15 to 25 Å including ca. 11 Å of an “effective” heme-to-ω-hydroxyl distance). The corresponding values of the electronic coupling matrix element vary within the range from ca. 3 to 0.02 cm−1; 2) the change in activation volume from +6.7 (n=3), to ≈0 (n=4), and −5.5 (n=6) cm3 mol−1 (disclosing at n=3 a direct pressure effect on the protein's internal viscosity); 3) a “full” Kramers-type viscosity dependence for ko at n=2 and 3 (demonstrating control of an intraglobular friction through the external dynamic properties), and its gradual transformation to the viscosity independent (nonadiabatic) regime at n=6 and 11. Multilateral cross-testing of “free” CytC in a native-like, glucose-stabilized and urea-destabilized (molten-globule-like) states revealed novel intrinsic links between local/global structural and functional characteristics. Importantly, our results on the high-pressure and solution-viscosity effects, together with matching literature data, strongly support the concept of “dynamic slaving”, which implies that fluctuations involving “small” solution components control the proteins' intrinsic dynamics and function in a highly cooperative manner as far as CT processes under adiabatic conditions are concerned.
Co-reporter:Dimitri E. Khoshtariya Dr.;Tina D. Dolidze Dr.;David Sarauli Dr.
Angewandte Chemie 2006 Volume 118(Issue 2) pp:
Publication Date(Web):28 NOV 2005
DOI:10.1002/ange.200502386
Unter Hochdruck: Der Elektronenaustausch von Cytochrom c, das frei zu Goldelektroden diffundiert, die durch Hydroxy-terminierte selbstorganisierte Alkanthiol-Monoschichten (SAMs; siehe Bild) modifiziert sind, wurde durch kinetische Studien bei hohem Druck untersucht. Die SAM-Dicke, die Viskosität der Lösung und der hydrostatische Druck bewirken einen Wechsel des Ladungstransfermechanismus und der entsprechenden kinetischen Reaktion.
Co-reporter:Martin Schlummer, Fritz Brandl, Andreas Mäurer, Rudi van Eldik
Journal of Chromatography A 2005 Volume 1064(Issue 1) pp:39-51
Publication Date(Web):28 January 2005
DOI:10.1016/j.chroma.2004.12.016
An HPLC–UV/MS method has been developed to identify and quantify flame retardants in post-consumer plastics from waste of electric and electronic equipment (WEEE). Atmospheric pressure chemical ionisation spectra of 15 brominated and phosphate-based flame retardants were recorded and interpreted. The method was applied to detect flame retardant additives in polymer extracts obtained from pressurised liquid extraction of solid polymers. In addition, a screening method was developed for soluble styrene polymers to isolate a flame retardant fraction through the application of gel permeation chromatography (GPC). This fraction was transferred to an online-coupled HPLC column and detected by UV spectroscopy, which allowed a reliable qualitative and quantitative analysis of brominated flame retardants in the polymer solutions.
Co-reporter:Peter Illner, Achim Zahl, Ralph Puchta, Nico van Eikema Hommes, Peter Wasserscheid, Rudi van Eldik
Journal of Organometallic Chemistry 2005 Volume 690(Issue 15) pp:3567-3576
Publication Date(Web):1 August 2005
DOI:10.1016/j.jorganchem.2005.03.029
The kinetics of the formation of the active species cis-[PtII(PPh3)2Cl(SnCl3)] and cis-[PtII(PPh3)2(SnCl3)2] from the hydroformylation catalyst precursor cis-[PtII(PPh3)2Cl2] in the presence of SnCl2, was studied in two different imidazolium-based ionic liquids. A large range of different chlorostannate melts consisting of 1-butyl-3-methyl-imidazolium cations and [SnxCly](−y + 2x) anions with varying molar fraction of SnCl2, were prepared and characterized by 1H and 119Sn NMR. The observed chemical shifts point to major changes in the composition of the anionic species within the melt. The second ionic liquid employed, viz., 1-butyl-3-methyl-imidazolium-bis(trifluormethylsulfonyl)amide was prepared in a colorless quality that enabled its application in kinetic studies. The concentration and temperature dependence of the substitution of Cl− by [SnCl3]− to yield cis-[PtII(PPh3)2Cl(SnCl3)], could be studied in detail. Theoretical (DFT) calculations were employed to model the reaction progress and to resolve the role of the ionic liquid in the activation of the catalyst. The available results are presented and a plausible mechanism for the formation of the catalytically active species is suggested.The kinetics and mechanism of the formation of the active species cis-[PtII(PPh3)2Cl(SnCl3)] and cis-[PtII(PPh3)2(SnCl3)2] from the hydroformylation catalyst precursor cis-[PtII(PPh3)2Cl2] in the presence of SnCl2, was studied in two different imidazolium-based ionic liquids.
Co-reporter:Ralph Puchta;Nico van Eikema Hommes;Rudi van Eldik
Helvetica Chimica Acta 2005 Volume 88(Issue 5) pp:911-922
Publication Date(Web):18 MAY 2005
DOI:10.1002/hlca.200590085
The ligand-exchange mechanism of solvated Be2+ cations has been studied by means of DFT calculations (RB3LYP/6-311+G**). Ligand exchange around [BeL4]2+, where L=H2O, NH3, CO2, formaldehyde (H2CO), HCN, N2, and CO, was found to follow an associative interchange (Ia) process in all cases. The size of the activation barrier is almost independent of the type of donor atom, and depends mainly on the hybridization undergone by the donor atom. This, in turn, suggests that steric effects play a major role in solvent- and ligand-exchange reactions in Be2+ systems.
Co-reporter:Christian F. Weber, Ralph Puchta, Nico J. R. van Eikema Hommes, Peter Wasserscheid,Rudi van Eldik
Angewandte Chemie International Edition 2005 44(37) pp:6033-6038
Publication Date(Web):
DOI:10.1002/anie.200501329
Co-reporter:Christian F. Weber Dr.;Ralph Puchta Dr.;Nico J. R. van Eikema Hommes Dr.;Peter Wasserscheid Dr. Dr.
Angewandte Chemie 2005 Volume 117(Issue 37) pp:
Publication Date(Web):22 AUG 2005
DOI:10.1002/ange.200501329
Wenn Normalsein überrascht: Bei der Untersuchung eines Ligandenaustauschs an einem PtII-Komplex (siehe Bild; apa: 2,6-Bis(aminomethyl)pyridin) in Wasser, Methanol und der ionischen Flüssigkeit 1-Butyl-3-methylimidazolium-bis(trifluormethylsulfonyl)amid stellte sich heraus, dass sich die ionische Flüssigkeit „normal“ verhält, d. h. wie jedes andere Solvens. Änderungen in der Polarität des Übergangszustands spielen in ihr nur eine untergeordnete Rolle.
Co-reporter:Evgeny V. Kudrik, Rudi van Eldik and Sergei V. Makarov
Dalton Transactions 2004 (Issue 3) pp:429-435
Publication Date(Web):22 Dec 2003
DOI:10.1039/B311695F
The substitution reaction of the axial-coordinated water by pyridine, pyrazine and 4-CN-pyridine in the low-spin Fe(II) complex of octasulfophenyltetrapyrazinoporphyrazine was studied. Kinetic and thermodynamic parameters for the different reaction steps of the process were determined. On the basis of NMR data and spectrophotometric titrations, a pronounced non-equivalence of the two coordinated N-donor ligands was observed. The substitution of water by pyridine and 4-CN-pyridine is shown to include the formation of a precursor outer-sphere complex, whereas substitution by pyrazine follows a limiting dissociative mechanism.
Co-reporter:Marek Pawelec, Grażyna Stochel and Rudi van Eldik
Dalton Transactions 2004 (Issue 2) pp:292-298
Publication Date(Web):10 Dec 2003
DOI:10.1039/B311053B
Copper(II) ions react rapidly with sulfur from thiol groups, forming two distinct, intensely absorbing, short-lived intermediates, which decompose in a subsequent redox reaction to produce reduced copper and disulfides. In this study we report the results of a mechanistic study on the reaction between mercaptosuccinic acid, HO2CCH2CH(SH)CO2H, and Cu2+(aq) and [Cu(tren)H2O]2+, tren = tris(2-aminoethyl)amine. Spectroscopic and kinetic data indicate that in the presence of an excess of thiol, at least two distinct complexes are formed, with very different decomposition rate constants and an absorption maximum at 346 nm. Upon addition of thiol to [Cu(tren)H2O]2+
(1 ∶ 1), a transient with a maximum at 380 nm appears, whereas in an excess of thiol this complex decomposes and again the 346 nm band is observed. The use of [Cu(tren)H2O]2+ enables to study the reaction of thiol with copper also in alkaline solution, where the rate of the overall process is slowed down greatly. The reactions were studied in detail, including the effect of dioxygen, and a possible reaction mechanism for the catalysed autoxidation process is proposed and discussed in reference to available literature data.
Co-reporter:Hans Erras-Hanauer, Timothy Clark, Rudi van Eldik
Coordination Chemistry Reviews 2003 Volumes 238–239() pp:233-253
Publication Date(Web):March 2003
DOI:10.1016/S0010-8545(02)00296-5
An overview of recent molecular orbital theory and density functional theory studies on water exchange reactions of metal ions and complexes in solution is presented. The different theoretical techniques used are reviewed and representative examples are discussed. The mechanistic insight gained is discussed in reference to available experimental data and the predictive nature of the applied techniques is highlighted.
Co-reporter:Andreas Hofmann and Rudi van Eldik
Dalton Transactions 2003 (Issue 15) pp:2979-2985
Publication Date(Web):07 Jul 2003
DOI:10.1039/B305174A
A series of binuclear Pt(II) complexes were synthesized to investigate the influence of four different bridging diamines, viz. benzene-1,3-diamine, benzene-1,4-diamine, 1,3-propanediamine and 1,4-butanediamine, on the reactivity of the platinum centres in complexes of the type [Pt2(N,N,N′,N′-tetrakis(2-pyridylmethyl)diamine)(H2O)2]4+ in comparison to the corresponding mononuclear complex, [Pt(bis(2-pyridylmethyl)amine)(H2O)]2+. The pKa values for both deprotonation steps were determined, and the kinetics of the reaction with chloride was studied under pseudo-first order conditions as a function of chloride concentration and temperature. The results indicate that the electrophilicity of the binuclear complexes is higher than that of the mononuclear complex, which results in lower pKa values, higher reaction rates and a decrease in ΔH‡. In addition, the reactivity of one Pt(II) centre shows a clear dependence on the nature of the other Pt(II) centre, leading to different thermodynamic and kinetic properties for the first and second reaction steps. The interaction between the two Pt(II) centres mainly depends on the Pt–Pt distance and not on the aromatic or aliphatic nature of the bridging diamine group. Temperature dependent 1H NMR-studies were performed on the [Pt2(N,N,N′,N′-tetrakis(2-pyridylmethyl)-1,3-propanediamine)Cl2]2+ complex, and the results indicate that a non-symmetric, folded species of this complex dominates at low temperatures. This behaviour may account for the unusual behaviour of the corresponding diaqua complex.
Co-reporter:Mohamed S. A. Hamza;Xiang Zou;Kenneth L. Brown;Rudi van Eldik
European Journal of Inorganic Chemistry 2003 Volume 2003(Issue 2) pp:
Publication Date(Web):23 DEC 2002
DOI:10.1002/ejic.200390036
The base-on/base-off equilibration of a series of alkylcobalamins (XCbl) was studied as a function of the electronic properties of X. A square scheme was developed to incorporate all species that participate in this equilibration. The equilibrium between five- and six-coordinate species was studied for several alkylcobalamins (protonated base-off) and alkylcobinamides as a function of pressure by using UV/Vis spectroscopy. In addition, the kinetics of the acid-induced base-on/base-off equilibration of β-NCCH2Cbl, β-CF3Cbl and CNCbl were studied; the data obtained reveal evidence for an acid-catalysed reaction path. The results of this study and the observed correlations are discussed in reference to those reported in the past, and enable the formulation of an overall equilibration scheme. (© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2003)
Co-reporter:Živadin D. Bugarčić, Günter Liehr and Rudi van Eldik
Dalton Transactions 2002 (Issue 6) pp:951-956
Publication Date(Web):12 Feb 2002
DOI:10.1039/B106038B
The kinetics of the complex-formation reactions between monofunctional palladium(II) complexes [Pd(N–N–N)H2O]2+, where N–N–N is 2,2′:6′,2″-terpyridine (terpy), diethylenetriamine (dien) or bis(2-pyridylmethyl)amine (bpma), with L-cysteine, DL-penicillamine and glutathione, have been studied in an aqueous 0.10 M perchloric acid medium using variable-temperature and -pressure stopped-flow spectrophotometry. Second-order rate constants, k1298, varied between 2.8 × 102 and 4.4 × 104 M−1 s−1. The highest reactivity was observed for the [Pd(terpy)H2O]2+ complex, whereas glutathione is the strongest nucleophile. Activation volumes for these reactions varied between −5.6 ± 0.3 and −10.7 ± 1.0 cm3 mol−1. The
negative entropies and volumes of activation support a strong contribution from bond making in the transition state of the substitution process. The crystal structure of [Pd(bpma)Cl]Cl·H2O has been determined by X-ray diffraction at 190 K. Crystals are triclinic with space group P and consist of distorted square-planar [Pd(bpma)Cl]+ cations. The Pd–N distances are all equal to 2.005(7) Å. The Pd–Cl distance is 2.305(3) Å.
Co-reporter:Sergei V. Makarov, Evgeny V. Kudrik, Rudi van Eldik and Ekaterina V. Naidenko
Dalton Transactions 2002 (Issue 22) pp:4074-4076
Publication Date(Web):25 Oct 2002
DOI:10.1039/B209195J
Thiourea dioxide was used as a precursor for sulfoxylate, SO22−, which is shown to reduce the methyl viologen dication to the fully reduced form, this is the first example of a direct study of the reduction with sulfoxylate; an important advantage of sulfoxylate and its parent compound thiourea dioxide, is their ability to reduce nitrite (the final product being nitrogen) and nitrous oxide in alkaline solutions in the absence of a catalyst.
Co-reporter:Debabrata Chatterjee, Anannya Mitra, Mohamed S. A. Hamza and Rudi van Eldik
Dalton Transactions 2002 (Issue 6) pp:962-965
Publication Date(Web):06 Feb 2002
DOI:10.1039/B108232A
The substitution reactions of [RuIII(edta)(H2O)]− (edta = ethylenediaminetetraacetate) with adenine, adenosine and the corresponding 5′-nucleotides (Nu), viz. adenosine-5′-monophosphate (AMP), adenosine-5′-diphosphate (ADP) and adenosine-5′-triphosphate (ATP), have been studied kinetically as a function of nucleotide concentration at various temperatures (5 to 30 °C) at a fixed pH of 4.6 to contribute to the mechanistic understanding of the binding of adenine base nucleotides. Based on the kinetic results, it is suggested that the binding of the 5′-nucleotides (AMP, ADP and ATP) takes place in a rapid nucleophile concentration-dependent step, followed by a concentration-independent ring-closure reaction. Kinetic data and activation parameters have been interpreted in terms of an associative mechanism and discussed in reference to the data reported before.
Co-reporter:Mladen Biruš;Mario Gabričević;Achim Zahl
European Journal of Inorganic Chemistry 2002 Volume 2002(Issue 4) pp:
Publication Date(Web):28 FEB 2002
DOI:10.1002/1099-0682(200203)2002:4<819::AID-EJIC819>3.0.CO;2-E
The application of 139La NMR spectroscopy to kinetic and equilibrium studies of the substitution of coordinated water by acetohydroxamic acid (HA) is described. The reaction of La3+ with HA can be represented by the following equation (with equilibrium constant K1): La3+ + HA La(A)2+ + H+ (1). The following equilibrium parameters were calculated from the 139La NMR chemical shift of the equimolar reactants measured as a function of acidity and temperature: K1(298) = 2.36(8)·10−4, ΔrH1 = −28(1) kJ·mol−1, ΔrS1 = −162(4) J·K−1·mol−1. The kinetics of the reaction between La3+ and HA can be represented by the following equations: La3+ + HA La(A)2+ + H+ (11) and La3+ + A− La(A)2+ (12). The complexation rate constants were calculated from the line widths of the corresponding signals as a function of acidity and temperature: k11(298 K) = 1.54(4)·106M−1·s−1, ΔH11‡ = 70(14) kJ·mol−1 and ΔS11‡ = +108(47) J·K−1·mol−1, k12 (298 K) = 1.6 1010M−1·s−1, ΔH12‡ = 35(6) kJ·mol−1 and ΔS12‡ = +68(20) J·K−1·mol−1. The spin-spin relaxation times of LaIII in the aqua ion and the hydroxamato complex are 4.36(4) ms and 54(1) μs at 298 K, respectively.
Co-reporter:Bernd Müller;Michael Galle;Thorsten Schneppensieper Dr.;Carlos Dücker-Benfer Dr.
Chemie in unserer Zeit 2002 Volume 36(Issue 4) pp:
Publication Date(Web):2 AUG 2002
DOI:10.1002/1521-3781(200208)36:4<246::AID-CIUZ246>3.0.CO;2-8
Jedes Jahr zum Anfang des Wintersemesters findet an der Friedrich Alexander Universität Erlangen-Nürnberg die Vorlesung „Chemische Zaubertricks” statt. Im Laufe der Jahre und mit dem Generationswechsel hat sich das Gesicht dieser Veranstaltung verändert. Aus der klassischen Experimentalvorlesung ist eine multimediale Show entstanden, die einen bemerkenswerten Erfolg beim Publikum verzeichnen darf und es bis ins Fernsehen geschafft hat. Wir geben hier unsere Erfahrungen und Informationen weiter, die es dem Interessierten möglich machen, selber eine ähnliche Werbeveranstaltung für die Chemie durchzuführen.
Every year at the start of the winter semester at the Friedrich Alexander University Erlangen-Nürnberg, the lecture „Chemical Magic Tricks” is held. In time and due to generation changes, the nature of this event has changed. From a classical experimental lecture it was transformed into a multi media show, which is very popular among the public and even made its way into television. Our aim with this article is to pass our experience and further information required on to those interested to present such an advertisement event for chemistry.
Co-reporter:Zong-Wan Mao, Frank W. Heinemann, Günther Liehr and Rudi van Eldik
Dalton Transactions 2001 (Issue 24) pp:3652-3662
Publication Date(Web):27 Nov 2001
DOI:10.1039/B103942N
A series of carbonato complexes was isolated in the reaction of [ML2(H2O)]2+ (M = Cu(II), Zn(II), L = bpy (2,2′-bipyridyl) and phen (1,10-phenanthroline) with HCO3−. Two monomeric carbonato Cu(II) complexes with different bidentate distortion, [Cu(phen)2(CO3)]·7H2O and [Cu(phen)2(CO3)]·11H2O, and three binuclear carbonato complexes with four different bridging modes, {[Cu(phen)2]2(μ2-CO3)}(ClO4)2·4.25H2O containing two different carbonate coordination modes, {[Cu(phen)2]2(μ2-CO3)}(ClO4)2·DMF·H2O and [(bpy)2(H2O)Zn(μ2-CO3)Zn(bpy)2](NO3)2·7H2O,
were characterized by X-ray crystallography. UV-Vis spectra of [CuL2(H2O)]2+ ions were recorded as a function of pH in the absence and presence of NaHCO3, and revealed evidence for a carbonation process in the pH range 6 to 10 and a hydrolysis process in the pH range 10 to 13. 13C NMR spectra of [(bpy)2(H2O)Zn(μ2-CO3)Zn(bpy)2]2+ revealed formation of polynuclear carbonato complexes in solution. Complex-formation kinetics of [CuL2(H2O)]2+ (L = bpy and phen) with HCO3− were studied by stopped-flow using a pH-jump technique. The results indicated rapid formation of [Cu(bpy)2OCO2H]+ and [Cu(phen)2OCO2H]+, followed by a slow step (induced by the pH jump) for which the observed
first order rate constants were found to be identical for both complexes and independent of the complex and carbonate concentrations employed. The thermal and pressure activation parameters for this step were also determined. The reaction is assigned to a pH-jump induced ring-closure reaction of the unstable five-coordinate monodentate bicarbonato intermediates, [Cu(bpy)2OCO2H]+ and [Cu(phen)2OCO2H]+, to form the pseudo-octahedral bidentate carbonato complexes [Cu(bpy)2O2CO] and [Cu(phen)2O2CO], respectively. A possible mechanism for the carbonation of CuN4 complexes with an in-plane coordinated water molecule in a trigonal bipyramidal structure is proposed.
Co-reporter:Joanna Macyk and Rudi van Eldik
Dalton Transactions 2001 (Issue 15) pp:2288-2292
Publication Date(Web):16 Jul 2001
DOI:10.1039/B101366L
The irreversible outer-sphere electron transfer reaction between the oxidant [Co(ox)3]3− and the redox protein horse heart cytochrome cII according to the reaction,
[CoIII(ox)3]3− + cytochrome cII6+ → [CoII(ox)3]4− + cytochrome cIII7+
was studied as a function of pH, concentration, temperature and pressure using UV–Vis and stopped-flow techniques. The concentration dependence of the observed rate constant showed a saturation of the rate constant at high concentrations of [Co(ox)3]3−, indicative of effective precursor formation between the reactants. The temperature (6–35 °C) and pressure (up to 1300 atm) dependencies at low and high complex concentrations were carried out and activation parameters (ΔH‡, ΔS‡, ΔV‡) were determined. The results are discussed in reference to earlier studies performed in our laboratories and elsewhere, and the importance of efficient precursor formation in such reactions is highlighted.
Co-reporter:Zong-Wan Mao, Günter Liehr and Rudi van Eldik
Dalton Transactions 2001 (Issue 10) pp:1593-1600
Publication Date(Web):30 Apr 2001
DOI:10.1039/B100160O
A series of binuclear bicarbonato, trinuclear carbonato and binuclear hydroxo complexes were isolated in the reaction of [M(tren)(H2O)](ClO4)2 (M = Cu(II), Zn(II) and tren = tris(2-aminoethyl)amine) with NaHCO3 at pH ca. 6.5, 8.5 and 10.0, respectively, and physically characterized. The structures of two trinuclear carbonato complexes, {[Cu(tren)]3(μ3-CO3)}(ClO4)4·H2O and {[Zn(tren)]3(μ3-CO3)}(ClO4)4·H2O were determined by X-ray analysis. UV-Vis spectra of [Cu(tren)(H2O)](ClO4)2 were recorded as a function of pH in the absence and presence of NaHCO3, and reveal evidence for a carbonation process in the pH range 6.0 to 9.5 and a hydrolysis process in the range pH 9.5 to 12. 13C NMR measurements on [Zn(tren)(H2O)](ClO4)2 as a function of pH in the presence of NaH13CO3, indicated that the Zn(II) complex behaves similarly to the Cu(II) complex in solution, and that the polynuclear carbonato complexes are only formed at low concentrations in solution. The kinetics of the reaction of [Cu(tren)(H2O)]2+ with HCO3− was studied by stopped-flow using a pH-jump technique. The results indicated that complex-formation of [Cu(tren)(H2O)]2+ with HCO3− is too fast to be resolved kinetically. The observed kinetics and second-order rate constant of 494 ± 10 M−1 s−1 were assigned to the rate-determining formation of a binuclear carbonato complex. A plausible mechanism for the carbonation of CuN4 complexes that include an axial water molecule in a trigonal bipyramidal structure, is proposed.
Co-reporter:Zhong-Lin Lu;Mohamed S. A. Hamza
European Journal of Inorganic Chemistry 2001 Volume 2001(Issue 2) pp:
Publication Date(Web):5 JAN 2001
DOI:10.1002/1099-0682(200102)2001:2<503::AID-EJIC503>3.0.CO;2-Q
Complex-formation constants were determined for the reaction of [Co(Me6tren)(H2O)]2+ with different pyridine- and imidazole-based nucleophiles, and the kinetics of these reactions was studied as a function of entering ligand concentration, temperature, and pressure. The steric hindrance associated with the methyl substituents on the trigonal-bipyramidal complex induces a sensitive control over the substitution mechanism from dissociative for the weaker and more bulky entering nucleophiles, to associative for the stronger and less bulky nucleophiles. A changeover in mechanism can be observed as a function of nucleophile concentration, in the case of 2-MeImH as the entering ligand. The mechanistic assignments are based on the observed rate laws, and the activation parameters determined under limiting concentration conditions.
Co-reporter:Thorsten Schneppensieper;Alicja Wanat;Grazyna Stochel;Sara Goldstein;Dan Meyerstein
European Journal of Inorganic Chemistry 2001 Volume 2001(Issue 9) pp:
Publication Date(Web):26 JUL 2001
DOI:10.1002/1099-0682(200109)2001:9<2317::AID-EJIC2317>3.0.CO;2-F
Rate constants for the formation and dissociation of FeII(L)NO (L = aminocarboxylato) have been determined using stopped-flow, temperature jump, flash photolysis, and pulse radiolysis techniques. For a series of ligands the formation rate constants vary between 1.6 × 106 and 2.4 × 108M−1 s−1, whereas the dissociation rate constants vary between 0.11 and 3.2 × 103 s−1 at 25 °C. These rate constants result in stability constants (KNO = kf/kd) ranging from 5.0 × 102 to 1.1 × 107M−1, which are in good agreement with values of KNO determined by a combined spectrophotometric and potentiometric technique. The results are discussed with reference to available literature data, and interpreted in terms of a generalized reaction mechanism.
Co-reporter:Thorsten Schneppensieper Dipl.-Chem.;Achim Zahl Dr. Dr. Dr. h.c.
Angewandte Chemie 2001 Volume 113(Issue 9) pp:
Publication Date(Web):3 MAY 2001
DOI:10.1002/1521-3757(20010504)113:9<1727::AID-ANGE17270>3.0.CO;2-F
Co-reporter:Deogratius Jaganyi Dr.;Andreas Hofmann Dr. Dr.
Angewandte Chemie 2001 Volume 113(Issue 9) pp:
Publication Date(Web):3 MAY 2001
DOI:10.1002/1521-3757(20010504)113:9<1730::AID-ANGE17300>3.0.CO;2-N
Co-reporter:M. Riess, H. Thoma, O. Vierle, R. van Eldik
Journal of Analytical and Applied Pyrolysis 2000 Volume 53(Issue 2) pp:135-148
Publication Date(Web):February 2000
DOI:10.1016/S0165-2370(99)00062-5
A rapid method for the qualitative identification of technical flame retardants in polymers originating from electronic waste, is described using curie-point pyrolysis gas chromatography with mass spectrometric detection. For a series of samples, the pyrolysis products of the flame retardants were studied as a function of the pyrolysis temperature. Identification of the flame retardants was achieved on the basis of a comparison with synthetic and pure flame retardant standards.
Co-reporter:Carlos Dücker-Benfer;Mohamed S. A. Hamza;Conrad Eckhardt;Rudi van Eldik
European Journal of Inorganic Chemistry 2000 Volume 2000(Issue 7) pp:
Publication Date(Web):4 JUL 2000
DOI:10.1002/1099-0682(200007)2000:7<1563::AID-EJIC1563>3.0.CO;2-V
The substitution behaviour of pentaamminemethylcobalt(III) with ethylenediamine (en) was studied in detail in aqueous ammonia solution. The displacement of four ammonia ligands by two ethylenediamine (en) chelates was followed by a subsequent slow cis to trans rearrangement of the bis(ethylenediamine)(amine)methyl complex. The dependence of the substitution reaction on [en], [NH3], temperature and pressure was studied, and the observed kinetic traces could be fitted to the sum of two exponentials. The rate law for both reaction steps showed saturation kinetics with respect to [en] and [NH3], and the activation parameters confirmed the operation of a limiting dissociative reaction mechanism under all conditions. For the first substitution reaction, ksat.(en)10 °C = 1.8 ± 0.4 s−1, ksat.(NH3)10 °C = 1.7 ± 0.3 s−1 and ΔV#15 °C = +14 ± 1 cm3·mol−1, for the second substitution reaction, ksat.(en)10 °C = 0.55 ± 0.12 s−1, ksat.(NH3)10 °C = 0.58 ± 0.10 s−1 and ΔV#15 °C = +24 ± 1 cm3·mol−1. The subsequent isomerization reaction was also studied in detail, and a complete set of activation parameters confirm the operation of a dissociative mechanism for the cis to trans rearrangement for which kiso25 °C = (1.48 ± 0.07) × 10−3 s−1; ΔH# = 115 ± 5 kJ·mol−1; ΔS# = + 86 ± 16 J·K−1·mol−1and ΔV# = +14.2 ± 0.6 cm3·mol−1.
Co-reporter:Rudi van Eldik Dr.;Colin D. Hubbard
Chemie in unserer Zeit 2000 Volume 34(Issue 5) pp:
Publication Date(Web):26 OCT 2000
DOI:10.1002/1521-3781(200010)34:5<306::AID-CIUZ306>3.0.CO;2-D
Die Untersuchung der Volumenänderung während einer chemischen Reaktion ist eine elegante Möglichkeit, Rekationsmechanismen in Lösung aufzuklären.
Co-reporter:Rudi van Eldik Dr.;Colin D. Hubbard
Chemie in unserer Zeit 2000 Volume 34(Issue 4) pp:
Publication Date(Web):20 SEP 2000
DOI:10.1002/1521-3781(200008)34:4<240::AID-CIUZ240>3.0.CO;2-3
Ist der Reaktionsmechanismus bekannt, kann eine chemische Reaktion gezielter gelenkt werden. Viele anorganische Reaktionen konnten lange Zeit mit konventionellen kinetischen Untersuchungsmethoden nicht eindeutig mechanistisch zugeordnet werden; erst unter hohem Druck gelang es, ihre Mechanismen aufzuklären.
Co-reporter:Tobias Ernst, Ralf Popp, Rudi van Eldik
Talanta 2000 Volume 53(Issue 2) pp:347-357
Publication Date(Web):1 November 2000
DOI:10.1016/S0039-9140(00)00491-4
Analytical data on element concentrations in plastics is an important prerequisite for the recycling of technical waste plastics. The chemical resistance and high additive contents of such materials place a high demand on analytical methods for quantifying elements in thermoplastics from electrotechnical applications. The applicability of three common independent analytical methods (EDXRF, AAS, ICP-AES) for the quantification of heavy metals in such technical waste plastics of varying composition was studied. Following specific sample pre-treatments, such as closed vessel microwave assisted digestion and wet ashing with H2SO4, three hazardous metals (Pb, Cd, Sb) were determined. Conditions were investigated to minimize matrix effects for all analytical techniques employed. The trueness for the quantification of Cd was checked by using the certified reference material VDA 001–004 (40–400 μg g−1 Cd in polyethylene), and no significant differences to certified values were found. The best detection limits were found to be 2, 1.3 and 7.9 μg g−1 for Cd, Pb and Sb, respectively. In technical waste polymers, Sb was detected to be in the range 1–7%, Cd in the range 80–12 000 μg g−1 and Pb in the range 90–700 μg g−1. The precision reached for the analysis of this complex material, is comparable for all methods, and can be expressed by a relative standard deviation smaller than 8%. Application of multivariate analysis of variances (MANOVA) showed no differences between the mean results, except for the ICP-AES analysis following wet ashing with H2SO4.
Co-reporter:Manuela Körner
European Journal of Inorganic Chemistry 1999 Volume 1999(Issue 10) pp:
Publication Date(Web):16 SEP 1999
DOI:10.1002/(SICI)1099-0682(199910)1999:10<1805::AID-EJIC1805>3.0.CO;2-3
The irreversible outer-sphere electron-transfer reaction between trans-bis(2-ethyl-2-hydroxybutanoato(2–))oxochromate(v) and cytochrome cII was investigated as a function of pH, concentration, temperature and pressure. The plot of the observed pseudo-first order rate constant as a function of the CrV concentration shows a clear trend towards saturation at higher CrV concentrations, from which the precursor formation constant and the electron-transfer rate constant could be separated (K = 37 ± 5 M−1 and kET = 1510 ± 180 s−1 at pH 4.8 and 279 K). In the low CrV concentration range the second-order electron-transfer rate contants were measured as a function of temperature (ΔH# = 20.9 ± 0.6 kJ mol−1; ΔS# = –79.9 ± 2.1 J K−1 mol−1; ΔG# (279 K) = 43.2 kJ mol−1). High-pressure experiments were performed at two different pH values. The kinetic (stopped-flow) and thermodynamic (electrochemical) measurements as a function of pressure enabled the construction of a volume profile for the system at 279 K. The activation volumes for the redox process are –9.2 ± 0.2 (pH 5.0) and –11.1 ± 0.8 cm3 mol−1 (pH 4.7), and the overall reaction volumes were estimated to be –7 ± 2 (pH 5.0) and –10 ± 2 cm3 mol−1 (pH 4.7) . The transition state of the redox reaction lies to a large extent on the product side and can be described as “late”. The results are discussed in comparison to earlier measurements using cobalt and ruthenium complexes as reaction partners for cytochrome c.
Co-reporter:Erika Ember, Hanaa Asaad Gazzaz, Sabine Rothbart, Ralph Puchta, Rudi van Eldik
Applied Catalysis B: Environmental (6 April 2010) Volume 95(Issues 3–4) pp:179-191
Publication Date(Web):6 April 2010
DOI:10.1016/j.apcatb.2009.12.013
Co-reporter:Debabrata Chatterjee, Namita Jaiswal, Alicja Franke and Rudi van Eldik
Chemical Communications 2014 - vol. 50(Issue 93) pp:NaN14565-14565
Publication Date(Web):2014/10/07
DOI:10.1039/C4CC06631F
Reported is the first example of a ruthenium(III) complex, RuIII(edta) (edta4− = ethylenediaminetetraacetate), that catalyzes the disproportion of H2O2 to O2 and water in resemblance to catalase activity, and shedding light on the possible mechanism of action of the [RuV(edta)(O)]− formed in the reacting system.
Co-reporter:Colin D. Hubbard, Peter Illner and Rudi van Eldik
Chemical Society Reviews 2011 - vol. 40(Issue 1) pp:NaN290-290
Publication Date(Web):2010/11/15
DOI:10.1039/C0CS00043D
The focus of this article is an examination of chemical reaction mechanisms in ionic liquids. These mechanisms are compared with those pertaining to the same reactions carried out in conventional solvents. In cases where the mechanisms differ, attempts to provide an explanation in terms of the chemical and physicochemical properties of the reactants and of the components of the ionic liquids are described. A wide range of reactions from different branches of chemistry has been selected for this purpose. A sufficient background for student readers has been included. This tutorial review should also be of interest to kineticists, and to both new and experienced investigators in the ionic liquids field.
Co-reporter:Debabrata Chatterjee, Sabine Rothbart and Rudi van Eldik
Dalton Transactions 2013 - vol. 42(Issue 13) pp:NaN4729-4729
Publication Date(Web):2013/01/30
DOI:10.1039/C3DT32737J
Reported here is the first example of a ruthenium complex, [RuIII(edta)(H2O)]− (edta4− = ethylenediaminetetraacetate), that catalyzes the oxidation of thiourea (TU) in the presence of H2O2. The kinetics and mechanism of this reaction were investigated in detail by using rapid-scan spectrophotometry as a function of both the hydrogen peroxide and thiourea concentrations at pH 4.9 and 25 °C. Spectral analyses and kinetic data clearly support a catalytic process in which hydrogen peroxide reacts directly with thiourea coordinated to the RuIII(edta) complex. HPLC product analyses revealed the formation of formamidine disulfide (TU2) as a major product at the end of the catalytic process, however, formation of other products like thiourea dioxide (TUO2), thiourea dioxide (TUO3) and sulfate was also observed after longer reaction times. Catalytic intermediates such as [RuIII(edta)(OOH)]2− and [RuV(edta)(O)]− were evidently found to be non-reactive in catalyzing the oxidation of thiourea by H2O2 under the specified conditions.
Co-reporter:Ana Djeković, Biljana Petrović, Živadin D. Bugarčić, Ralph Puchta and Rudi van Eldik
Dalton Transactions 2012 - vol. 41(Issue 13) pp:NaN3641-3641
Publication Date(Web):2012/02/09
DOI:10.1039/C2DT11843B
The kinetics of the substitution reactions between the mono-functional Au(III) complexes, [Au(dien)Cl]2+ and [Au(terpy)Cl]2+ (dien = 3-azapentane-1,5-diamine, terpy = 2,2′;6′,2′′-terpyridine) and bi-functional Au(III) complexes, [Au(bipy)Cl2]+ and [Au(dach)Cl2]+ (bipy = 2.2′-bipyridine, dach = (1R,2R)-1,2-diaminocyclohexane) and biologically relevant ligands such as L-histidine (L-His), inosine (Ino), inosine-5′-monophosphate (5′-IMP) and guanosine-5′-monophosphate (5′-GMP), were studied in detail. All kinetic studies were performed in 25 mM Hepes buffer (pH = 7.2) in the presence of NaCl to prevent the spontaneous hydrolysis of the chloride complexes. The reactions were followed under pseudo-first order conditions as a function of ligand concentration and temperature using stopped-flow UV-vis spectrophotometry. The results showed that the mono-functional complexes react faster than the bi-functional complexes in all studied reactions. The [Au(terpy)Cl]2+ complex is more reactive than the [Au(dien)Cl]2+ complex, which was confirmed by quantum chemical (DFT) calculations. A more than 50% lower activation energy for the terpy than for the dien based complex was found. The bi-functional [Au(bipy)Cl2]+ complex is more reactive than the [Au(dach)Cl2]+ complex. The reactivity of the studied nucleophiles follows the same order for all studied systems, viz. L-His > 5′-GMP > 5′-IMP > Ino. According to the measured activation parameters, all studied reactions follow an associative substitution mechanism. Quantum chemical calculations (B3LYP/LANL2DZp) suggest that ligand substitution in [Au(terpy)Cl]2+ and [Au(dien)Cl]2+ by imidazole follows an interchange mechanism with a significant degree of associative character. The results demonstrate the strong connection between the reactivity of the complexes toward biologically relevant ligands and their structural and electronic characteristics. Therefore, the binding of gold(III) complexes to 5′-GMP, constituent of DNA, is of particular interest since this interaction is thought to be responsible for their anti-tumour activity.
Co-reporter:Hanaa Asaad Gazzaz, Erika Ember, Achim Zahl and Rudi van Eldik
Dalton Transactions 2009(Issue 43) pp:NaN9495-9495
Publication Date(Web):2009/09/14
DOI:10.1039/B912911A
Rate and activation parameters for the complex-formation reaction of Ni2+ with 4-(2-pyridylazo)-N,N-dimethyl aniline (PADA) were studied as a function of pH in different buffers in both aqueous and sodium dodecyl sulfate (SDS) micelle solutions. In aqueous Tris buffer solution, the forward and backward rate constants increased with increasing pH, while the complex-formation constant decreased due to a larger increase in the backward rate constant. The activation entropy, ΔS#, and activation volume, ΔV#, changed with increasing pH from positive to negative values, suggesting an apparent changeover from a dissociative to a more associative mechanism. Complex-formation reactions with 2,2′-bipyridine in Tris buffer showed almost no increase in the forward and backward rate constants on increasing the pH, but the ΔS# and ΔV# values became more negative. N-ethylmorpholine buffer showed no pH effect on the rate constants and activation parameters. Water exchange reactions of aquated Ni2+ were also studied as a function of pH under the same conditions. The reported rate and activation parameters for water exchange in Tris and N-ethylmorpholine buffers are consistent with those found for the complex-formation reactions of Ni2+ with PADA. The observed pH and buffer effects for both the complex-formation and water exchange reactions of aquated Ni2+ can be accounted for in terms of the formation of a Ni2+-Tris complex in Tris buffer and general base catalysis by the buffer components. In SDS micelle solution, the complex-formation reaction with PADA was much faster than in aqueous solution, but the increase in rate constant with increasing pH was less significant, while ΔS# and ΔV# became more positive, pointing to a more dissociative mechanism. For SDS micelle solutions there was no effect on the water exchange rate constant or activation volume. Mechanistic interpretations are offered for all observed pH, buffer and medium effects.
Co-reporter:Dimitri E. Khoshtariya, Tina D. Dolidze, Tatyana Tretyakova, David H. Waldeck and Rudi van Eldik
Physical Chemistry Chemical Physics 2013 - vol. 15(Issue 39) pp:NaN16526-16526
Publication Date(Web):2013/07/25
DOI:10.1039/C3CP51896E
Gold electrodes were coated with alkanethiol SAM–azurin (Az, blue cupredoxin) assemblies and placed in contact with a water-doped and buffered protic ionic melt as the electrolyte, choline dihydrogen phosphate ([ch][dhp]). Fast-scan protein-film voltammetry was applied to explore interfacial biological electron transfer (ET) under conditions approaching the glass-transition border. The ET rate was studied as a function of the water amount, temperature (273–353 K), and pressure (0.1–150 MPa). Exposure of the Az films to the semi-solid electrolyte greatly affected the protein's conformational dynamics, hence the ET rate, via the mechanism occurring in the extra complicated dynamically-controlled regime, is compared to the earlier studies on the reference system with a conventional electrolyte (D. E. Khoshtariya et al., Proc. Natl. Acad. Sci. U. S. A., 2010, 107, 2757–2762), allowing for the disclosure of even more uncommon mechanistic motifs. For samples with low water content (ca. 3 or less waters per [ch][dhp]), at moderately low temperatures (below ca. 298 K) and/or high pressure (150 MPa), the voltammetric profiles systematically deviated from the standard Marcus current–overvoltage pattern, deemed as attributable to a breakdown of the linear response approximation through the essential steepening of the Gibbs energy wells near the glass-forming threshold. Electrolytes with a higher water content (6 to 15 waters per [ch][dhp]) display anomalous temperature and pressure performances, suggesting that the system crosses a broad nonergodic zone which arises from the interplay of ET-coupled large-scale conformational (highly cooperative) modes of the Az protein, inherently linked to the electrolyte's (water-doped [ch][dhp]) slowest collective relaxation(s).
Co-reporter:Azza Shoukry, Malgorzata Brindell and Rudi van Eldik
Dalton Transactions 2007(Issue 37) pp:
Publication Date(Web):
DOI:10.1039/B706856E
Co-reporter:Hakan Ertürk, Andreas Hofmann, Ralph Puchta and Rudi van Eldik
Dalton Transactions 2007(Issue 22) pp:NaN2301-2301
Publication Date(Web):2007/04/04
DOI:10.1039/B700770C
A series of dinuclear Pt(II) complexes of the type [Pt2(N,N,N′,N′-tetrakis(2-pyridylmethyl)diamine(H2O)2]4+ were synthesized. Acid–base titrations, and concentration and temperature dependent stopped-flow measurements of the reaction with chloride were performed to study the thermodynamic and kinetic behaviour of the dinuclear bridged complexes. The results indicate that there is a clear interaction between the two Pt(II) centres, which becomes weaker as the aliphatic chain increases in length. From a certain chain length onwards, the Pt(II) centres become independent of each other and exhibit identical thermodynamic and kinetic properties. The experimental results are discussed in reference to structures obtained by DFT (BP86/LACVP*) calculations.
Co-reporter:Hakan Ertürk, Joachim Maigut, Ralph Puchta and Rudi van Eldik
Dalton Transactions 2008(Issue 20) pp:NaN2766-2766
Publication Date(Web):2008/04/04
DOI:10.1039/B718177A
Complexes of the type [Pt2(N,N,N′,N′-tetrakis(2-pyridylmethyl)diamine(H2O)2]4+ and [Pt2(N,N,N′,N′-tetrakis(2-pyridylmethyl)diamine(Cl)2]2+ were used to study their reactions with a series of bio-relevant nucleophiles, viz.thiourea, L-methionine and guanosine-5′-monophosphate (5′-GMP2−) as a function of nucleophile concentration and temperature. The reactions with the sulfur containing nucleophiles (thiourea and L-methionine) were followed under pseudo-first-order conditions by stopped-flow and UV-Vis spectrophotometry. The reaction with 5′-GMP2− was carried out under second order conditions and studied by NMR spectroscopy. The results indicate that the bridged dinuclear complexes remain intact after coordination of the studied nucleophiles for an extended period of time, which differs significantly from that reported for other multinuclear platinum complexes in the literature.
Co-reporter:Elena A. Vlasova, Natalya Hessenauer-Ilicheva, Denis S. Salnikov, Evgeny V. Kudrik, Sergei V. Makarov and Rudi van Eldik
Dalton Transactions 2009(Issue 47) pp:NaN10549-10549
Publication Date(Web):2009/10/30
DOI:10.1039/B906478H
A detailed study of the oxidation of L-ascorbic acid by dioxygen and nitrite in water at pH 5.8 and 7.0, catalyzed by the octasulfophenyltetrapyrazinoporphyrazine complex of cobalt(II), was carried out using conventional spectrophotometric, low-temperature and high-pressure stopped-flow techniques. The Co(II) complex activates L-ascorbic acid through an intramolecular one-electron oxidation step that involves the reduction of the octasulfophenyltetrapyrazinoporphyrazine. The reaction rate strongly depends on pH due to the different redox behaviour of the L-ascorbic acid/ascorbate species present in solution. Kinetic parameters for the different reaction steps of the catalytic process were determined. The final product of the reaction between L-ascorbic acid and nitrite was found to be nitrous oxide.
Co-reporter:Tanja Soldatović, Živadin D. Bugarčić and Rudi van Eldik
Dalton Transactions 2009(Issue 23) pp:NaN4531-4531
Publication Date(Web):2009/04/22
DOI:10.1039/B822718G
The kinetics and mechanism of ligand substitution reactions of [Pt(SMC)Cl2] (SMC = S-methyl-L-cysteine) with biologically relevant ligands were studied as a function of chloride and nucleophile concentrations at pH 2.5 and 7.2. It was observed that the slope and intercept obtained from the linear dependence of the observed rate constant on the nucleophile concentration strongly depend on the [Cl−] for all the studied substitution reactions. At high [Cl−], the rate constant for the forward reaction is almost zero and that for the back reaction follows the order: L-met > GSH ∼ INO > 5′-GMP. Ion-pair formation between the positively charged Pt(II) complex and the chloride ion is suggested to account for the saturation kinetics observed for the back reaction. The results are discussed in terms of the mechanism of the anti-tumour activity of related platinum complexes.
Co-reporter:Debabrata Chatterjee, Kalyan Asis Nayak, Erika Ember and Rudi van Eldik
Dalton Transactions 2010 - vol. 39(Issue 7) pp:NaN1698-1698
Publication Date(Web):2009/12/15
DOI:10.1039/B920839A
Reported in this paper is the first example of a ruthenium complex, [RuIII(edta)(H2O)]− (edta = ethylenediaminetetra-acetate), that catalyzes the oxidation of hydroxyurea in the presence of H2O2, mimicking the action of peroxidase or catalase and shedding light on their possible mechanism of action.
Co-reporter:Peter Illner, Ralph Puchta, Frank W. Heinemann and Rudi van Eldik
Dalton Transactions 2009(Issue 15) pp:NaN2801-2801
Publication Date(Web):2009/02/20
DOI:10.1039/B820940E
The coordinative behaviour of the bis(trifluoromethylsulfonyl)amide (NTf2−) anion was studied in terms of its interaction with the labile [Pd(terpyridine)Cl]Cl complex. Among various attempts to coordinate NTf2− to the Pd(II) centre, the complexes [Pd(terpyridine)NO3]NTf2 and [Pd(terpyridine)H2O](NTf2)2 were isolated and characterized. The crystallographic data for these complexes are presented and the possible coordination of the NTf2− anion to the metal centre is discussed in reference to DFT predictions.
Co-reporter:Debabrata Chatterjee, Ujjwal Pal, Sarita Ghosh and Rudi van Eldik
Dalton Transactions 2011 - vol. 40(Issue 6) pp:NaN1306-1306
Publication Date(Web):2011/01/04
DOI:10.1039/C0DT01444C
The kinetics of reduction of [RuIII(edta)pz]− (edta4− = ethylenediaminetetraacetate; pz = pyrazine) by thioamino acids (RSH = cysteine, glutathione) resulting in the formation of a red [RuII(edta)pz]2− species (λmax = 462 nm) has been studied spectrophotometrically using both conventional mixing and stopped-flow techniques. The time course of the reaction was followed as a function of [RSH], pH, temperature and pressure. Alkali metal ions were found to have a positive influence (K+ > Na+ > Li+) on the reaction rate. Kinetic data and activation parameters are interpreted in terms of a mechanism involving outer-sphere electron transfer between the reaction partners. A detailed reaction mechanism in agreement with the spectral and kinetic data is presented.
Co-reporter:Živadin D. Bugarčić, Jovana Bogojeski, Biljana Petrović, Stephanie Hochreuther and Rudi van Eldik
Dalton Transactions 2012 - vol. 41(Issue 40) pp:NaN12345-12345
Publication Date(Web):2012/08/14
DOI:10.1039/C2DT31045G
A brief overview of mechanistic studies on the reactions of different Pt(II) complexes with nitrogen- and sulfur-donor biomolecules is presented. The first part describes the results obtained for substitution reactions of mono-functional Pt(II) complexes with different biomolecules, under various experimental conditions (temperature, pH and ionic strength). In addition, an overview of the results obtained for the substitution reactions of bi-functional Pt(II) complexes, analogous to cisplatin, with biomolecules is given. The last part of this report deals with different polynuclear Pt(II) complexes and their substitution behaviour with different biomolecules. The purpose of this perspective is to improve the understanding of the mechanism of action of Pt(II) complexes as potential anti-tumour drugs in the human body.
Co-reporter:Stephanie Hochreuther, Sharanappa T. Nandibewoor, Ralph Puchta and Rudi van Eldik
Dalton Transactions 2012 - vol. 41(Issue 2) pp:NaN522-522
Publication Date(Web):2011/11/01
DOI:10.1039/C1DT11453K
The diaqua complex [Pt(2-methylthiomethylpyridine)(OH2)2]2+, Pt(mtp), was synthesized and investigated thermodynamically as well as kinetically. Spectrophotometric acid–base titrations were performed to determine the pKa values of the two coordinated water ligands. A low pKa1 value of 3.15 was observed for the water molecule trans to the pyridine donor, whereas a pKa2 value of 6.84 was found for the water molecule trans to the labilising sulphur donor. The substitution of coordinated water by a series of sterically hindered S-containing nucleophiles, viz.thiourea (tu), N,N′-dimethylthiourea (dmtu) and N,N,N′,N′-tetramethylthiourea (tmtu), was studied under pseudo first-order conditions as a function of nucleophile concentration, pH (2, 4.75, 7.4), temperature and pressure, using stopped-flow techniques and UV-vis spectroscopy. In general the first substitution reaction takes place trans to the sulphur donor. At pH 2 the nucleophiles react in the order tu (634 ± 10) > dmtu (507 ± 5) ≫ tmtu (165 ± 3 M−1 s−1 at 25 °C), which is caused by steric hindrance. The second observed reaction involves two steps, viz. the displacement of the second water ligand and dechelation of the pyridine ring with the third-order rate constants 73.3 ± 0.8 (tu), 22.1 ± 0.1 (dmtu) and 6.8 ± 0.2 M−2 s−1 (tmtu) at 25 °C. At pH 4.75 the reactions are in general slower due to the presence of aqua-hydroxo species. The same order in reactivity was found, viz. tu (106 ± 1) > dmtu (72 ± 1) ≫ tmtu (14.1 ± 0.5 M−1 s−1 at 25 °C). No evidence for ring-dechelation could be observed under these conditions. At pH 7.4 the inert dihydroxo species is predominantly present in solution and consequently no substitution reaction was observed. Quantum chemical calculations were performed to support the interpretation and discussion of the experimental results.
Co-reporter:Rodrigo Luis Silva Ribeiro Santos, Rudi van Eldik and Denise de Oliveira Silva
Dalton Transactions 2013 - vol. 42(Issue 48) pp:NaN16805-16805
Publication Date(Web):2013/08/29
DOI:10.1039/C3DT51763B
Diruthenium(II,III)-tetracarboxylates have shown promising anticancer properties as metallotherapeutics. On the basis of the role that bio-reducing agents may play on the mode of action of ruthenium-based anticancer drugs, we performed detailed kinetic studies on the reaction of ascorbic acid and glutathione with the [Ru2(RCOO)4]+ paddlewheel framework by using the non-drug, diaqua complex ion [Ru2(CH3COO)4(H2O)2]+. In the presence of the reducing agents, the diaqua-Ru2 species first undergo a ligand substitution reaction by which the axially-coordinated water is displaced by the reducing agent. In both cases, this reaction is followed by an intra-molecular electron transfer process during which the metal–metal center is reduced from Ru25+ to Ru24+ and the reducing agent is oxidized. Product analyses were performed with the application of ESI-MS and 1H-NMR techniques. Rate and activation parameters are reported for the different reaction steps.
Co-reporter:Safaa S. Hassan, Mohamed M. Shoukry and Rudi van Eldik
Dalton Transactions 2012 - vol. 41(Issue 43) pp:NaN13453-13453
Publication Date(Web):2012/09/25
DOI:10.1039/C2DT31730C
The acid–base and complex-formation equilibria of [Ru(H2dtpa)(H2O)], where dtpa = diethylenetriaminepentaacetate, with a series of bio-relevant ligands having various functional groups, viz. thiol, amine, imidazole and carboxylate, were investigated using potentiometric and spectrophotometric techniques. The pKa values for [Ru(H2dtpa)(H2O)] were found to be 2.28 and 3.48 for the uncoordinated carboxylic acid groups and 8.83 for the coordinated water molecule. The complexes formed are of 1:1 stoichiometry and their formation-constants were determined. The effect of dioxane on the acid–base and complex-formation equilibria of the RuIII complex was studied. The displacement reaction of coordinated NO by the investigated ligands showed that thiols can compete with NO in their reaction with [RuIII(dtpa)(H2O)]2−. The results reveal that the RuIII complex is deactivated as a NO scavenger by thiolate ligands.
Co-reporter:Snežana Jovanović, Biljana Petrović, Živadin D. Bugarčić and Rudi van Eldik
Dalton Transactions 2013 - vol. 42(Issue 24) pp:NaN8896-8896
Publication Date(Web):2013/04/16
DOI:10.1039/C3DT50751C
The reduction of the Pt(IV) complexes [PtCl4(bipy)], [PtCl4(dach)] and [PtCl4(en)] by glutathione (GSH), L-cysteine (L-Cys) and L-methionine (L-Met) was investigated by stopped-flow spectrophotometry at pH 2.0 (in 0.01 M perchloric acid) and at pH 7.2 (in 25 mM Hepes buffer). Kinetic measurements were performed under pseudo-first order conditions with an excess of the reducing agent. The order of the reactivity of the studied complexes was [PtCl4(bipy)] > [PtCl4(dach)] > [PtCl4(en)], and reactivity of investigated reducing agents followed the order GSH > L-Cys > L-Met. All the reactions between the selected Pt(IV) complexes and the sulfur donor biomolecules proceeded by a reductive elimination process that included nucleophilic attack by the reducing agent on one of the mutually trans-coordinated chloride ligands, which led to a two-electron transfer process. The final products of the redox reactions were the corresponding reduced Pt(II) complexes and the oxidized form of the reducing agents.
Co-reporter:Debabrata Chatterjee, Namita Jaiswal, Matthias Schmeisser and Rudi van Eldik
Dalton Transactions 2014 - vol. 43(Issue 48) pp:NaN18046-18046
Publication Date(Web):2014/10/06
DOI:10.1039/C4DT02628D
Reported here is the first example of a ruthenium(III) complex [RuIII(EDTA)(H2O)]− (EDTA4− = ethylenediaminetetraacetate) that mediates S-nitrosylation of cysteine in the presence of nitrite at pH 4.5 (acetate buffer) and results in the formation of [RuIII(EDTA)(SNOCy)]−. The kinetics of the reaction was studied by stopped-flow and rapid-scan spectrophotometry as a function of [Cysteine], [NO2−] and pH (3.5–8.5). Formation of [RuIII(EDTA)(SNOCy)]−, the product of the S-nitrosylation reaction, was identified by ESI-MS experiments. A working mechanism in agreement with the spectroscopic and kinetic data is presented.
Co-reporter:Tanja Soldatović, Snežana Jovanović, Živadin D. Bugarčić and Rudi van Eldik
Dalton Transactions 2012 - vol. 41(Issue 3) pp:NaN884-884
Publication Date(Web):2011/11/08
DOI:10.1039/C1DT11313E
The novel dinuclear Pt(II) complexes [{trans-Pt(NH3)2Cl}2(μ-pyrazine)](ClO4)2 (Pt1), [{trans-Pt(NH3)2Cl}2(μ-4,4′-bipyridyl)](ClO4)2·DMF (Pt2), and [{trans-Pt(NH3)2Cl}2(μ-1,2-bis(4-pyridyl)ethane)](ClO4)2 (Pt3), were synthesized. Acid–base titrations, and temperature and concentration dependent kinetic measurements of the reactions with biologically relevant ligands such as thiourea (Tu), glutathione (GSH) and guanosine-5′-monophosphate (5′-GMP) were studied at pH 2.5 and 7.2. The reactions were followed under pseudo-first-order conditions by stopped-flow and UV-vis spectrophotometry. 1H NMR spectroscopy was used to follow the substitution of chloride in the complex [{trans-Pt(NH3)2Cl}2(μ-4,4′-bipyridyl)](ClO4)2·DMF by guanosine-5′-monophosphate (5′-GMP) under second-order conditions. The results indicate that the bridging ligand has an influence on the reactivity of the complexes towards nucleophiles. The order of reactivity of the investigated complexes is Pt1 > Pt2 > Pt3.
Co-reporter:Debabrata Chatterjee, Alicja Franke, Maria Oszajca and Rudi van Eldik
Dalton Transactions 2014 - vol. 43(Issue 8) pp:NaN3094-3094
Publication Date(Web):2013/10/16
DOI:10.1039/C3DT52486H
The [RuIII(edta)(H2O)]− (edta4− = ethylenediaminetetraacetate) complex catalyzes the oxidation of azide (N3−) with H2O2, mimicking the action of metallo-enzymes such as catalase and peroxidase in biochemistry. The kinetics of the catalytic oxidation process was studied by using stopped-flow and rapid-scan spectrophotometry as a function of [RuIII(edta)], [H2O2], [N3−] and pH. The catalytic activity of the different oxidizing species produced in the reaction of [RuIII(edta)(H2O)]− with H2O2 for the oxidation of azide was compared to the oxidation of coordinated azide in [RuIII(edta)N3]2− by H2O2. Detailed reaction mechanisms in agreement with the spectroscopic and kinetic data are presented for both reaction paths.
Co-reporter:Mirjana D. Đurović, Živadin D. Bugarčić, Frank W. Heinemann and Rudi van Eldik
Dalton Transactions 2014 - vol. 43(Issue 10) pp:NaN3921-3921
Publication Date(Web):2014/01/21
DOI:10.1039/C3DT53140F
The influence of tridentate, nitrogen donor ligands, on the stability of gold(III) complexes under physiological conditions was investigated. The interaction of [Au(terpy)Cl]2+ (terpy = 2,2′:6′2′′ terpyridine), [Au(bpma)Cl]2+ (bpma = bis(pyridyl-methyl)amine), [Au(dien)Cl]2+ (dien = diethylenetriamine) and [AuCl4]− with the biologically relevant thiols, L-cysteine (L-Cys) and glutathione (GSH), and thioether, L-methionine (L-Met), was studied using UV-Vis spectroscopy, cyclic voltammetry, 1H NMR spectroscopy and ESI-MS. In this study, the rate constants for substitution reactions between monofunctional gold(III) complexes and sulfur donor ligands in aqueous solution were determined at different initial concentrations of reactants, chloride ions, pH and constant ionic strength. The obtained second-order rate constants for the reaction with L-methionine in the absence of added chloride at pH 2.5 and 25 °C follow the sequence (7.5 ± 0.4) × 103 > (4.5 ± 0.1) × 102 > 88.3 ± 0.8 M−1 s−1 for the terpy, bpma and dien complexes, respectively, demonstrating that the substitution step could be detected prior to the reduction step. This behavior was expected due to the influence of a decreasing π-donor ability of the chelate ligands, which slows down the substitution reactions along the series of complexes studied. In order to throw more light on the mechanism of biological activity of gold(III) compounds, such a systematic study was performed for all the mentioned thiols and thioether.
Co-reporter:Tina D. Dolidze, Dimitri E. Khoshtariya, Peter Illner and Rudi van Eldik
Chemical Communications 2008(Issue 18) pp:NaN2114-2114
Publication Date(Web):2008/02/25
DOI:10.1039/B719787J
Proven electrochemical approaches were applied to study heterogeneous electron transfer (ET) between selected redox couples and gold electrodes modified with alkanethiol self-assembled monolayers (SAMs), using the room-temperature ionic liquid (RTIL) [bmim][NTf2] as reaction medium; ferrocene as freely diffusing redox probe in the RTIL was tested for ET through both thin (butanethiol) and thick (dodecanethiol) assemblages at pressures up to 150 MPa; well behaved kinetic patterns and reproducibility of data were demonstrated for ET within the unique Au/SAM/RTIL arrays.
Co-reporter:Sabine Rothbart, Erika Ember and Rudi van Eldik
Dalton Transactions 2010 - vol. 39(Issue 13) pp:NaN3272-3272
Publication Date(Web):2010/02/11
DOI:10.1039/B925160J
The kinetics of the hydrogen peroxide induced oxidative degradation of the azo dye Orange II in aqueous carbonate buffered solution were studied for the oxo-bridged [Mn2III/IV(μ-O)2(bpy)4](ClO4)3 complex and its mononuclear analogue [MnII(bpy)2Cl2] as catalysts to reveal the underlying reaction mechanism and reactive intermediates participating in the catalytic cycle. Both catalysts show identical oxidative reactivity when used at equimolar manganese concentration. If a simple Mn(II) salt and a 1:2 concentration of bipyridine are added to the substrate and oxidant containing reaction mixture, the same oxidative reactivity as found for both readily prepared catalysts was observed for several investigated substrates. This demonstrates the in situ accessibility of a reactive intermediate and its precursor complex. The crucial role of bicarbonate as co-catalyst was studied. The distinct dependence of the observed rate constant for the oxidation reaction on the total carbonate concentration can be accounted for in terms of in situ generation of peroxycarbonate. EPR and rapid scan UV/Vis measurements of the reaction of hydrogen peroxide in carbonate buffered solution with [Mn2III/IV(μ-O)2(bpy)4](ClO4)3 and [MnII(bpy)2Cl2], revealed for both catalysts the presence of monomeric Mn(II) and MnIV-oxo species as the main intermediates. The proposed reaction mechanism involves two-electron oxidation of a mononuclear Mn(II) precursor complex to a high-valent Mn(IV)O intermediate as catalytically active species. Differences in the activity of in situ prepared catalyst precursors of different metal to ligand ratios are reported. The 1:2 complex was found to be the catalytically more active precursor for the oxidation of the selected substrates, whereas the 1:3 complex rather catalyzed the disproportion of hydrogen peroxide.