Acessibilidade / Reportar erro

Nitric oxide as an activation agent for nucleophilic attack in trans-[Ru(NO)(NH3)4{P(OEt)3}](PF6)3

Abstracts

The complex trans-[Ru(NO)(NH3)4{P(OEt)3}](PF6)3 undergoes nucleophilic attack on the phosphorus ester ligand in the solid state yielding trans-[Ru(NO)(NH3)4{P(OH)(OEt)2}](PF6)3. The reaction was monitored and the products analyzed using nuclear magnetic resonance spectroscopy (31P{¹H} CP-MAS NMR and 31P{¹H} NMR), infrared spectroscopy (FTIR), electron paramagnetic resonance spectroscopy (EPR), cyclic voltammetry (CV), electronic spectroscopy (UV-Vis) and elemental analysis. According to experimental data and quantum mechanical calculations (DFT), the reaction proceeds in the solid state by the nucleophilic attack on the phosphorus ligand, promoted by the strong polarization along the P III-RuII-NO+ axis induced by the nitrosyl ligand, and takes place following the Michaelis-Arbusov type mechanism for phosphorus ester hydrolysis. In solution, the nucleophilic attack occurs simultaneously at the nitrosyl and triethylphosphite ligands, yielding trans-[Ru(H2O)(NH3)4{P(OEt)3}]2+ and trans-[Ru(NO)(H2O)(NH3)4]3+ in comparable amounts.

nitrosyl; ruthenium; phosphorus ester


Neste trabalho foi investigado o ataque nucleofílico sofrido no estado sólido pelo complexo trans-[Ru(NO)(NH3)4{P(OEt)3}](PF6)3, no éster de fósforo coordenado, originando como produto a espécie trans-[Ru(NO)(NH3)4{P(OH)(OEt)2}](PF6)3. A reação foi monitorada e os produtos caracterizados utilizando ressonância magnética nuclear (31P{¹H} CP-MAS NMR e 31P{¹H} NMR), espectroscopia no infravermelho (FTIR), ressonância paramagnética eletrônica (EPR), voltametria cíclica (CV), espectroscopia eletrônica (UV-Vis) e análise elementar. De acordo com os dados experimentais e cálculos de mecânica quântica (DFT), a reação ocorre no estado sólido pelo ataque nucleofílico no éster de fósforo coordenado devido à forte polarização no eixo P III-RuII-NO+ induzida pela presença do ligante nitrosilo. A reação segue o mecanismo tipo Michaelis-Arbusov para hidrólise de ésteres de fósforo. Em solução (pH 7,0), o ataque nucleofílico ocorre nos ligantes NO+ e P(OEt)3, gerando como produtos os complexos trans-[Ru(H2O)(NH3)4{P(OEt)3}]2+ e trans-[Ru(NO)(H2O)(NH3)4]3+ em proporções comparáveis.


ARTICLE

Nitric oxide as an activation agent for nucleophilic attack in trans-[Ru(NO)(NH3)4{P(OEt)3}](PF6)3

Gustavo MetzkerI; José C. Toledo Jr.II; Francisco C. A. LimaIII; Alvicler MagalhãesIV; Daniel R. CardosoI; Douglas W. Franco* * e-mail: douglas@iqsc.usp.br ,I

IInstituto de Química de São Carlos, Universidade de São Paulo, Av. Trabalhador Sãocarlense 400, 13560-970 São Carlos-SP, Brazil

IIDepartamento de Química, Universidade Federal do ABC, Rua Santa Adélia 166, 09210-170 Santo André-SP, Brazil

IIICentro de Ciências da Natureza, Coordenação de Química, Universidade Estadual do Piauí, Rua João Cabral, 2231, 64002-150 Teresina-PI, Brazil

IVInstituto de Química, Universidade Estadual de Campinas, Cidade Universitária Zeferino Vaz, 13083-970 Campinas-SP, Brazil

ABSTRACT

The complex trans-[Ru(NO)(NH3)4{P(OEt)3}](PF6)3 undergoes nucleophilic attack on the phosphorus ester ligand in the solid state yielding trans-[Ru(NO)(NH3)4{P(OH)(OEt)2}](PF6)3. The reaction was monitored and the products analyzed using nuclear magnetic resonance spectroscopy (31P{1H} CP-MAS NMR and 31P{1H} NMR), infrared spectroscopy (FTIR), electron paramagnetic resonance spectroscopy (EPR), cyclic voltammetry (CV), electronic spectroscopy (UV-Vis) and elemental analysis. According to experimental data and quantum mechanical calculations (DFT), the reaction proceeds in the solid state by the nucleophilic attack on the phosphorus ligand, promoted by the strong polarization along the PIII–RuII–NO+ axis induced by the nitrosyl ligand, and takes place following the Michaelis-Arbusov type mechanism for phosphorus ester hydrolysis. In solution, the nucleophilic attack occurs simultaneously at the nitrosyl and triethylphosphite ligands, yielding trans-[Ru(H2O)(NH3)4{P(OEt)3}]2+ and trans-[Ru(NO)(H2O)(NH3)4]3+ in comparable amounts.

Keywords: nitrosyl, ruthenium, phosphorus ester

RESUMO

Neste trabalho foi investigado o ataque nucleofílico sofrido no estado sólido pelo complexo trans-[Ru(NO)(NH3)4{P(OEt)3}](PF6)3, no éster de fósforo coordenado, originando como produto a espécie trans-[Ru(NO)(NH3)4{P(OH)(OEt)2}](PF6)3. A reação foi monitorada e os produtos caracterizados utilizando ressonância magnética nuclear (31P{1H} CP-MAS NMR e 31P{1H} NMR), espectroscopia no infravermelho (FTIR), ressonância paramagnética eletrônica (EPR), voltametria cíclica (CV), espectroscopia eletrônica (UV-Vis) e análise elementar. De acordo com os dados experimentais e cálculos de mecânica quântica (DFT), a reação ocorre no estado sólido pelo ataque nucleofílico no éster de fósforo coordenado devido à forte polarização no eixo PIII–RuII–NO+ induzida pela presença do ligante nitrosilo. A reação segue o mecanismo tipo Michaelis-Arbusov para hidrólise de ésteres de fósforo. Em solução (pH 7,0), o ataque nucleofílico ocorre nos ligantes NO+ e P(OEt)3, gerando como produtos os complexos trans-[Ru(H2O)(NH3)4{P(OEt)3}]2+ e trans-[Ru(NO)(H2O)(NH3)4]3+ em proporções comparáveis.

Introduction

The relevance of nitric oxide and therefore of the nitric oxide synthase enzymes (NOS) in biological processes is now widely recognized.1-3 In the absence of oxygen, i.e., hypoxia situation of living organisms, the NOS enzymes are inhibited.1 Ruthenium nitrosyl complexes of the type trans-[Ru(NO)(NH3)4(L)]3+ (L = N-heterocyclic and triethylphosphite P(OEt)3) are very robust in aqueous solutions, but when activated by one electron reduction, they release NO0 at rates that may be controlled depending on the nature of L.4,5 Therefore, these compounds would provide NO0 under circumstances where NOS are not operational.

A systematic approach aiming to use ruthenium nitrosyl complexes as NO carriers is being developed.6,7 For this, trans-[Ru(NO)(NH3)4{P(OEt)3}](PF6)3 is found to be interesting since k-NO is 0.98 s-1 at 25 ºC, the compound is water soluble, its metabolites have very low cytotoxicities, and the E0NO+/NO0 = −0.18 vs. SCE is accessible to biological reductors such as cysteine, gluthatione and ascorbic acid.8 Furthermore, this compound exhibits a quite promising behaviour in biological experiments as tripanomicyde and leishmanicyde agent.9

Phosphorus esters are known to be quite reactive.10 In acidic solutions, triethylphosphite, P(OEt)3, undergoes acid catalyzed hydrolysis yielding diethylphosphite (P(OH)(OEt)2) and diethylphosphonate (P(H)(O)(OEt)2) by the Michaelis-Arbusov reaction mechanism.11 Furthermore, in the presence of oxidants such as hydrogen peroxide or bromine, the oxidation of PIII to PV is observed to take place easily. The formation of the stable P=O bond is considered to be the driving force for this reaction.12 However, when coordinated to RuII, P(OEt)3 is stable with respect to oxidation. For example, trans-[Ru(NH3)4{(P(OEt)3}2]2+ is stable in solid state for long periods of storage in a vacuum desiccator and in the absence of light. Also, the complexes trans-[Ru(H2O)(NH3)4{P(OEt)3}]2+ and trans-[Ru(NH3)4{(P(OEt)3}2]2+ are stable in acid aqueous solution even on addition of excess Br2 or CeIV, leading only to the metal center oxidation.8

Nitrogen monoxide binds strongly to the RuII center and, in this case, the back-bonding in the [RuIINO+]3+ fragment is so strong that the metal center exhibits properties of RuIII.13,14 The RuII center in trans-[Ru(NO)(H2O)(NH3)4]3+ is not oxidized by CeIV or Br2 even in the presence of an excess of these oxidants.15 For example, the pKa value of the coordinated water molecule in trans-[Ru(NO)(H2O)(NH3)4]3+ is 3.1, thus being a stronger acid than trans-[Ru(H2O)(NH3)5]2+, pKa 4.5, by one order of magnitude.15 Thus, the presence of the nitrosyl ligand in trans-[Ru(NO)(NH3)4{P(OEt)3}](PF6)3 induces not only changes in the metal center but also promotes the activation of the trans positioned phosphorus ester for nucleophilic attack.

Herein the changes induced by nitrosyl in the solid state and solution reactivities of the P(OEt)3 ligand in trans-[Ru(NO)(NH3)4{P(OEt)3}](PF6)3 are described.

Results and Discussion

Samples of trans-[Ru(NO)(NH3)4{P(OEt)3}](PF6)3 have shown to undergo gradual degradation in the solid state, even when the complex was stored in a vacuum desiccator under controlled air moisture. Figure 1 illustrates changes in the vibrational spectrum of the complex in solid samples kept for 8 months in such conditions (p = 0.035 bar, air moisture 20 ± 3%) and protected from light at room temperature. The dashed line is the spectrum of freshly prepared trans-[Ru(NO)(NH3)4{P(OEt)3}](PF6)3, which exhibits the characteristic band for coordinated nitrosyl (ν(NO+) = 1915 cm-1).5 The band assigned to ν(NO+) was found to shift to a lower wavenumber (1885 cm-1), indicating changes in the phosphorus ester ligand as a function of time. Concomitantly, changes in the intensity of the band relative to δ(CH3) at 1390 cm-1 (Figure 1, a) were also observed in the aged sample. The area calculated for this band in the degradation product is 35% smaller than that determined for the starting material. This suggests the loss of an ethyl group (C2H5) from the original complex. Two new bands assigned as ν(P−O−H)16 are observed to appear at 1140 and 940 cm-1 (Figure 1, b and c), corroborating the nucleophilic attack on the P(OEt)3 ligand. The spectral characteristics suggest that the phosphorus ligand stays coordinated after the hydrolysis process. The NO+ ligand was not found to suffer any reaction during the period of time under analysis, as indicated by the FTIR spectra.


To estimate the amount of trans-[Ru(NO)(NH3)4{P(OEt)3}](PF6)3 that suffered nucleophilic attack, deconvolution analyses of the bands relative to ν(NO+) at 1915 and 1885 cm-1 were carried out. Results suggest that, after 8 months of storage, approximately 95% of the original complex was converted into the degradation product. Furthermore, evidence of participation of air moisture in the degradation process was observed. Samples of trans-[Ru(NO)(NH3)4{P(OEt)3}](PF6)3 were also kept under argon in ampoules, where the degradation process was found to take place at the beginning and then ceased after 3 months, suggesting that the process is air-moisture dependent. FTIR experiments carried out periodically showed that 20% of the initial amount of trans-[Ru(NO)(NH3)4{P(OEt)3}](PF6)3 was converted into the degradation product. For these samples, no additional changes were observed in the FTIR spectra during a period of one year. Also, samples of trans-[Ru(NO)(NH3)4{P(OEt)3}](PF6)3 stored in Schlenk tubes under dry argon did not undergo any decomposition after three months.

Elemental analysis of the degradated samples of trans-[Ru(NO)(NH3)4{P(OEt)3}](PF6)3 showed a loss of 35% in the carbon content, but no changes in the nitrogen content. This is consistent with the loss of one ethyl group (C2H5), suggesting strongly that the P(OEt)3 ligand suffered nucleophilic attack.

Figure 2 shows the UV-Vis spectrum of the degradation product (solid line) obtained after 8 months of storage. The figure clearly depicts bands detected at 242 nm (ε = 1,429 L mol-1 cm-1), 312 nm (ε = 912 L mol-1 cm-1) and 474 nm (ε = 38 L mol-1 cm-1), indicating that the decomposition product has NO+ and a phosphorus ester in the coordination sphere. These absorption coefficient values were calculated considering 95% of conversion of the original complex into the new one. The electronic spectrum of freshly prepared trans-[Ru(NO)(NH3)4{P(OEt)3}](PF6)3 (dashed line) has three bands at 260 nm (ε = 1,291 L mol-1 cm-1), 316 nm (ε = 258 L mol-1 cm-1) and 522 nm (ε = 35 L mol-1 cm-1) attributed to ligand field (260 and 316 nm) and charge transfer (522 nm) transitions.8 This spectrum and band assignments are consistent with the described for other ruthenium nitrosyl tetraammines.17,18 No shift in the bands at 242 and 312 nm in the electronic spectra of the degradation product was observed in acetonitrile, suggesting the predominant ligand field character of the these bands. Due to the microsymmetry similarity exhibited by these two compounds, the electronic absorption bands of the product of nucleophilic attack were tentatively attributed to the same transitions assigned to trans-[Ru(NO)(NH3)4{P(OEt)3}](PF6)3.8


The 31P{1H} CP-MAS NMR analysis of the decomposition product was also carried out, confirming changes in the coordination sphere of trans-[Ru(NO)(NH3)4{P(OEt)3}](PF6)3, as suggested by FTIR and UV-Vis data. After the storage period of 8 months, more than 95% of the resonance peak recorded at 80 ppm for trans-[Ru(NO)(NH3)4{P(OEt)3}](PF6)3 (Figure 3, top) was observed to disappear, being replaced with a new one at 67 ppm (Figure 3, bottom). According to the literature,8,19 this chemical shift region for 31P can be attributed to coordinated PIII, and this indicates that the phosphorus ester remains coordinated after the degradation process. The change in the chemical shift by 13 ppm suggests the formation of a new PIII species coordinated to the RuII center, but with a weaker π acidity character than that presented by the P(OEt)3 ligand. This would explain the observed displacement of the peak to a lower chemical shift value as a consequence of the small back-donation between the RuII4dπ orbitals and PIII 3dπ orbitals. The splitting of the 31P signals in both spectra may be due to polymorphism in the solid state samples.20


The 31P{1H} NMR spectrum of the degradation product in acid aqueous solution (CH+ = 1.0×10-1 mol L-1) shows only one peak at 67 ppm, which is the same chemical shift value found in the 31P{1H} CP-NMR spectrum. No additional peaks were observed in the NMR spectra for a period of one hour, suggesting that the new complex is stable in this experimental condition.

Cyclic voltammograms of the degradated complex (after 8 months of storage) were obtained in acid aqueous solution (pH 1.0; µ = 0.1 mol L-1) at 25ºC. At first sight, the voltammograms of the degradation product (Figure 4, top) and of the original trans-[Ru(NO)(NH3)4{P(OEt)3}]3+ (Figure 4, bottom) depict similar qualitative features in the range of +0.70 to −0.70V vs. SCE. However, a close comparison of the two species in the voltammogram shows a shift of 200 mV in the half-wave potential attributed to the reversible reduction of the metal center in trans-[Ru(H2O)(NH3)4{P(OEt)3}]2+ (E01/2 a2 = 0.515V) and in the degradation product, trans-[Ru(H2O)(NH3)4{PIII}]2+ (E01/2 a1 = 0.315V). Furthermore, as described previously5,8 only one reduction wave attributable for the NO+/NO0 process is present in the voltammogram of trans-[Ru(NO)(NH3)4{P(OEt)3}]3+ at −0.18V (Figure 4, b2), whereas for the degradation product it is possible to observe two waves: one at −0.36V (b1) and another at −0.56V (c) (Figure 4). From the stand point of the observed peak current intensity, the first wave may be assigned to the reduction process NO+/NO0 (a1; E0 = −0.36V). The shift of the half-wave potential of the NO+/NO0 couple to more negative values is coherent with the reduced π acidity of the phosphorus ester ligand, corroborating the 31P{1H} CP-NMR and FTIR spectral data.


Until now, we do not have a good explanation for the second peak at −0.56V. A more detailed electrochemical investigation is being carried out.

All samples analysed for 8 months were EPR silent, suggesting the presence of only RuII after the degradation process.

Taking in account all data obtained so far, it is likely that the nucleophilic attack on the coordinated phosphorus ester leads, in the solid state, to trans-[Ru(NO)(NH3)4{P(OH)(OEt)2}](PF6)3, following the equation:

Coordination of P(OH)(OEt)2 to the [Ru(NH3)4(H2O)]2+ moiety was first antecipated by Sernaglia,21 who was able to identify in solution the ion trans-[Ru(H2O)(NH3)4{P(OH)(OEt)2}]2+. Attempts to isolate trans-[Ru(NO)(NH3)4{P(OH)(OEt)2}](PF6)3 from the reaction mixture of trans-[Ru(H2O)(NH3)4{P(OH)(OEt)2}]2+ with NO2– in acidic media have been unsuccessful so far.22 On the other hand, a recent report describes the synthesis and reactions of the complex [Mo(CO)5{P(OH)(OEt)2}], in which the metal center is also in a d6 low spin configuration. In solution, this complex undergoes nucleophilic attack on the diethylphosphite ligand, yielding [Mo(CO)5{P(OH)3}].23,24

Experiments in aqueous media confirm that trans-[Ru(NO)(NH3)4{P(OEt)3}](PF6)3 suffers nucleophilic attack in solution, at pH 7.0, generating two different products: trans-[Ru(H2O)(NH3)4{P(OEt)3}]2+ and trans-[Ru(NO)(H2O)(NH3)4]3+, following the reaction scheme:

These products were identified through differential pulse voltammetry (DPV) studies (Figure 5), in which the presence of trans-[Ru(H2O)(NH3)4{P(OEt)3}]2+ and trans-[Ru(NO)(H2O)(NH3)4]3+ was observed at +0.51 (peak c) and −0.48 V (peak b) vs. SCE, respectively. Consistently, the FTIR spectrum25 from the solution shows ν(NO+) = 1880 cm-1, indicating the presence of trans-[Ru(NO)(H2O)(NH3)4]3+, while the 31P NMR spectrum (d 148 ppm) shows the presence of trans-[Ru(H2O)(NH3)4{P(OEt)3}]2+.5 According to DPV data, this reaction at pH 7.0 yields 60% of trans-[Ru(H2O)(NH3)4{P(OEt)3}]2+ and 40% of trans-[Ru(NO)(H2O)(NH3)4]3+.


The reaction of phosphorus esters in acid media is well known and recognized to take place following the Michaelis-Arbusov type mechanism, in which the nucleophilic attack occurs at the carbon atom of the P(OEt)3 ligand. The P(OEt)3 coordinated to the ruthenium in trans-[Ru(NO)(NH3)4{P(OEt)3}]3+ can be activated for nucleophilic attack, with the [Ru(NO)(NH3)4]3+ moiety playing the proton role in the Michaelis-Arbusov mechanism.25 In the solid state, water molecules are likely to be the nucleophilic agent, but in solution the behavior is quite different. Only solutions of the ion complex trans-[Ru(NO)(NH3)4{P(OEt)3}]3+ with pH > 6.0 suffered nucleophilic attack, suggesting that in solution OH– is the nucleophilic agent. In this perspective, it is likely that the reactions take place in the solid state and in solution following two different pathways (Scheme 1).


When coordinated to ruthenium in the [Ru(NH3)5]2+ or [Ru(H2O)(NH3)4]2+ complexes, P(OEt)3 is remarkably stabilized towards hydrolysis and thus stable over a wide pH range.5,8 This could be a consequence of an electron density enrichment on the phosphorus and carbon atoms by the strong back-bonding from the filled RuII 4dπ orbitals to the empty PIII 3dπ ligand orbitals.5,8,13,21 Under the influence of NO+, the back-bonding is significantly attenuated due to the competition between P(OEt)3 and NO+ for the RuII 4dπ electrons. The nearly linear RuII−NO+ angle (175º.10’) in this complex5,8 also offers a favorable symmetry for the overlapping between the empty pyπ* and pzπ* orbitals (in the NO+ ligand) and the electron rich Ru dxz and dyz orbitals.16

The changes in the NO+ ligand as a consequence of the trans influence of the phosphorus ligand can be envisaged on the plot of E0(NO+/NO0) (SCE) vs. ν(NO+) (cm1). As can be observed, the trans-[Ru(NO)(NH3)4{P(OH)(OEt)2}]3+ ion follows the linear relationship (r2 = 0.94) observed5,13 for other tetraamines (Figure 6).


The redox potential for the NO+/NO0 couple and the ν(NO+) in these complexes are both dependent on the Ru 4dπ → NO+ pπ* back-bonding.5,13 Based on this, it is possible to infer that the π acidity of the phosphite series follows the order:

P(OH)3 < P(OH)(OEt)2 < P(OPri)3 < P(OMe)3 < P(OEt)3 < P(OEtCl)3

Theoretical calculations were performed to confirm the hypothesis of formation of trans-[Ru(NO)(NH3)4{P(OH)(OEt)2}]3+. The electronic spectrum calculated for trans-[Ru(NO)(NH3)4{P(OH)(OEt)2}](PF6)3 using TD-DFT (time dependent DFT, see Supplementary Information) presented the same shape and band attributions of the spectrum of trans-[Ru(NO)(NH3)4{P(OEt)3}](PF6)3. The blue shift of the bands in trans-[Ru(NO)(NH3)4{P(OH)(OEt)2}](PF6)3, as compared to trans-[Ru(NO)(NH3)4{P(OEt)3}](PF6)3, may be explained by the changes in π acidity of the phosphorus ester trans to the nitrosyl ligand. A weaker π acid ligand allows a stronger Ru 4dπ → NO+ pπ* back-bonding, leading to a larger crystal field splitting between the 4d orbitals of the RuII center. From the energy difference of the ligand field bands (Figure 2), it is possible to estimate a stabilization of ca. 7 kcal mol-1 for the ruthenium 4d orbitals when P(OEt)3 is replaced with P(OH)(OEt)2.

The electronic density of the phosphorus and carbon atoms of the phosphite ligand is also expected to decrease in the nitrosyl complex. According to DFT calculations performed for the trans-[Ru(NO)(NH3)4{P(OEt)3}]3+ and trans-[Ru(NO)(NH3)4{P(OH)(OEt)2}]3+ ions, the composition of the LUMO orbitals is predominantly characteristic of the NO+ and P(OEt)3 ligands. From the calculated NBO charges for trans-[Ru(NO)(NH3)4{P(OEt)3}]3+ and for the free P(OEt)3 (Table 1), it is possible to infer that, upon coordination to the metal center, remarkable changes occur in the P(OEt)3 ligand, which becomes more electrophilic. According to the NBO charge values, the more positive is the NBO charge value, more susceptible is the atom to nucleophilic attack.26

From the NBO data (Table 1) it is possible to infer that four atoms, i.e., the metal center, the nitrogen of the nitrosyl ligand and the phosphorus and carbon atoms of the CH2 groups, are more susceptible to the nucleophilic attack. According to X-ray diffraction data27 and quantum mechanical calculations, the metal center and phosphorus atoms, but not the carbon atoms, are expected to exhibit steric hindrance, therefore making the reaction difficult.

The solid state 13C NMR spectra of the nitrosyl complexes containing P(OEt)3 and P(OH)(OEt)2 exhibited chemical shifts of 63 and 70 ppm for the CH2 groups respectively, thus suggesting that the CH2 group in the P(OEt)3 complex is more eletrophilic than in the P(OH)(OEt)2 species, being consequently more susceptible to nucleophilic attack. Corroborating the NBO data on Table 1, the 13C chemical shifts determined for the CH3 groups of the diethyl/triethylphosphite ligands did not show any expressive difference (δ(CH3) 16.0 and 16.7 ppm respectively). On this perspective, the DFT calculations, together with the 13C NMR data, strongly suggest a nucleophilic attack on the CH2 group of P(OEt)3, following the Michaelis-Arbusov type mechanism of phosphorous ester hydrolysis and leading to P(OH)(OEt)2.

Experimental

All reagents were used as purchased from Aldrich and Merck. P(OEt)3 was purified as described before.8 Solvents, purchased from Mallinckrodt Baker and Merck, were purified following literature procedures.29 Five samples of the trans-[Ru(NO)(NH3)4{P(OEt)3}](PF6)3 complex were periodically analysed by spectroscopic techniques during a period of 8 months. Samples were stored in a vacuum desiccator (p = 0.035 bar), in which air moisture was controlled with a drying agent (CaCl2·H2O, air moisture 20 ± 3%).29 Air moisture was continuosly measured using a hygrometer (Barigo, Vilingen-Schwenningen, Germany) and the samples homogenized every 15 days to avoid only superficial degradation. All manipulations of air sensitive compounds were carried out as described in literature,30 using serum caps, teflon tubes and glass syringes. The complex trans-[Ru(NO)(NH3)4{P(OEt)3}](PF6)3 was synthesized as described in the literature.5 UV-Vis measurements were performed on a Hitachi U-3501 spectrophotometer using a 1.0 cm path length quartz cell. Degassed solutions of the degradation product were transferred into quartz cells using teflon tubes. The spectra were recorded from 225 to 600 nm. FTIR spectra were recorded using a Bomem MB-102 series spectrometer, using KBr pellets, in the region of 2000-400 cm-1. Elemental analysis was performed on an EA 1110 CHNS-O Carlo Erba Instrument. 31P{1H} CP-MAS NMR measurements were carried out with a Varian Inova 400 MHz spectrometer using fine powdered samples packed into a zirconia CP rotor. The counter ion PF6 (from NH4(PF6), δ31P = −144 ppm) was used as reference. EPR spectra were recorded on a Bruker ESP 300E X-band spectrometer at 77K. DPPH• (g = 2.0037) was used as the field calibrant. Electrochemical measurements (CV and DPV) were performed on an EG&G PAR model 264A equipment, using a glass carbon electrode as a working electrode, a saturated calomel electrode (SCE) as reference and a platinum plate as auxiliary electrode. Solutions of CF3COOH/CF3COONa (µ = 0.1 mol L-1) were used as electrolyte.

DFT calculations for trans-[Ru(NO)(NH3)4{P(OEt)3}]3+ and trans-[Ru(NO)(NH3)4{P(OH)(OEt)2}]3+ were performed with the Gaussian 03 package.31 Optimized molecular geometries were calculated using the hybrid B3LYP exchange-correlation functional with the GDIIS algorithm,32,33 and tight self-consistent field convergence criteria were used for calculations. Solvent effects were evaluated using the Polarizable Continuum Model (PCM).34 Single-point calculations with the B3LYP and the standard 6-311+G* basis sets for H, C, N, P and O, and the LANL2DZ basis set for Ru were performed. A dielectric constant of ε = 78.39 was used for the complexes in water. Natural atomic charges35 from natural population analysis (NPA) were obtained with the NBO 3.0 program as implemented in the Gaussian 03 software. All the fully optimized geometries were characterized by vibrational frequency calculations, which showed only real frequencies. Zero-point vibrational energies were estimated based on the B3LYP frequency calculation (unscaled) using the same basis set as for geometry optimization.

Conclusions

The data presented here strongly suggest that, in solid state, a nucleophilic attack at the CH2 group of P(OEt)3 coordinated to the RuII center in trans-[Ru(NO)(NH3)4{P(OEt)3}](PF6)3 occurs, leading to the formation of trans-[Ru(NO)(NH3)4{P(OH)(OEt)2}](PF6)3. After 8 months of storage, the conversion was found to be higher than 95% and the diethylphosphite complex was stable for a period of at least one year. The degradation process was followed with five independent samples and shown to be reproducible. The nucleophilic attack follows the Michaelis-Arbusov type-reaction mechanism. In solution, the nucleophilic attack occurs on the NO+ and P(OEt)3 ligands competitively, yielding the trans-[Ru(H2O)(NH3)4{P(OEt)3}]2+ and trans-[Ru(NO)(H2O)(NH3)4]3+ complex ions. The presence of the nitrosyl on the Ru coordination sphere appears to be the driving force for the nucleophilic attack, since no reaction was observed in the same experimetal conditions for the synthetic precursors of the nitrosyl complex. The strong π electron polarization generated by NO+ may be useful as catalysis strategy to activate molecules for nucleophilic attack in biological pathways where NO is involved.

Supplementary Information

Supplementary data are available free of charge at http://jbcs.sbq.org.br, as a pdf file.

Acknowledgments

The authors acknowledge the Brazilian funding agencies FAPESP (Proc. 05/53266-4 and 06/05877-4), CNPq and CAPES for financial support. Prof. Leif H. Skibsted (University of Copenhagen) is acknowledged for reading the manuscript and Prof. Eduardo R. Azevedo (IFSC/USP) for carrying out the 31P{1H} CP-MAS NMR experiments.

References

1. Lundberg, J. O.; Weitzberg, E.; Gladwin, M. T.; Nat. Rev. Drug Discovery 2008, 7, 156.

2. Li, H.; Igarashi, J.; Jamal, J.; Yang, W.; Poulos, T. L.; J. Biol. Inorg. Chem. 2006, 11, 753.

3. Cullota, E.; Koshland, J. D. E.; Science 1992, 258, 1862.

4. Toledo, J. C.; Silva, H. A. S.; Scarpellini, M.; Mori, V.; Camargo, A. J.; Bertotti, M.; Franco, D. W.; Eur. J. Inorg. Chem. 2004, 9, 1879.

5. Tfouni, E.; Krieger, M.; McGarvey, B. R.; Franco, D. W.; Coord. Chem. Rev. 2003, 236, 57.

6. Clarke, M. J.; Coord. Chem. Rev. 2003, 236, 209.

7. McCleverty, J. A.; Chem. Rev. 2004, 104, 403.

8. Franco, D. W.; Taube, H.; Inorg. Chem. 1978, 17, 571; Lopes, L. G. F.; Castellano, E. E.; Ferreira, A. G.; Davanzo, C. U.; Clarke, M. J.; Franco, D. W.; Inorg. Chim. Acta 2005, 358, 2883.

9. Silva, J. J. N.; Osakabe, A. L.; Pavanelli, W. R.; Silva, J. S.; Franco, D. W.; Br. J. Pharmacol. 2007, 152, 112.

10. Ishmaeva, E. A.; Popova, E. V.; Mironov, V. F.; Aminova, R. M.; Vereshchagina, Y. A.; Galkin, V. I.; Moeller, K.; Schmutzler, R.; Russ. J. Org. Chem. 2004, 40, 1076; Doak, G. O.; Freedman, L. D.; Chem. Rev. 1961, 61, 31.

11. Corbridge D. C. E.; Phosphorus: An Outline of its Chemistry, Biochemistry and Technology, 4th ed., Elsevier: Amsterdam, 1990.

12. Kirby, A. J.; Warren, S. G.; The Organic Chemistry of Phosphorus, 1st ed., Elsevier: New York, 1967.

13. Toledo, J. C.; Lima-Neto, B. S.; Franco, D. W.; Coord. Chem. Rev. 2005, 249, 419.

14. Caramori, G. F.; Frenking, G.; Organometalics 2007, 26, 5815.

15. Bezerra, C. W. B.; Silva, S. C.; Gambardella, M. T. P.; Santos, R. H. A.; Plicas L. M. A.; Tfouni, E.; Franco, D. W.; Inorg. Chem. 1999, 8, 5660.

16. Nakamoto, K.; Infrared and Raman Spectra of Inorganic and Coordination Compounds, 4th ed., John Wiley & Sons: New York, 1986; Socrates, G.; Infrared and Raman Characteristic Group Frequencies, 3rd ed., John Wiley & Sons: New York, 2001.

17. Schreiner, A. F.; Lin, S. W.; Hausser, P. J.; Hopcus, E. A.; Hamm, D. J.; Gunter, J. D.; Inorg. Chem. 1972, 11, 880.

18. Gorelsky, S. I.; Silva, S. C.; Lever, A. B. P.; Franco, D. W.; Inorg. Chim. Acta 2000, 300, 698.

19. Verdake, J. G.; Quin, L. D.; Phophorus-31 NMR Spectroscopy on Stereochemical Analysis, VCH Paublishers: Deerfield Beach, 1987.

20. Aguiar, M. R. M. P.; Gemal, A. L.; Gil, R. A. S. S.; Quim. Nova 1999, 22, 553; Penner, G. H.; Wasylishen, R. E.; Can. J. Chem. 1989, 67, 1909; Moynihan, H. A.; O’Hare, I. P.; Int. J. Pharmam. 2002, 247, 179.

21. Sernaglia, R. L.; Franco, D. W.; Inorg. Chem. 1989, 28, 3485.

22. Truzzi, D. R.; Franco, D. W.; Abstracts of the XIV Brazilian Meeting in Inorganic Chemistry, Foz de Iguaçu, Brazil, 2008.

23. Xi, C.; Liu, Y.; Lai, C.; Zhou, L.; Inorg. Chem. Commun. 2004, 7, 1202.

24. Han, L. B.; Tanaka, M.; J. Am. Chem. Soc. 1996, 118, 1571.

25. Brill, T. E.; Landon, S. J.; Chem. Rev. 1984, 84, 577; Bezerra, C. W. B.; Silva, S. C.; Gambardella, M. T. P.; Santos, R. H. A.; Plicas, M. L.; Tfouni, E.; Franco, D. W.; Inorg. Chem. 1999, 38, 5660.

26. Koleva, G.; Galabov, B.; Wu, J. I.; Schaefer, H. F.; Schleyer, P. R.; J. Am. Chem. Soc. 2009, 131, 14772.

27. Mazzetto, S. E.; Gambardella, M. T. P.; Santos, R. H. A.; Lopes, L. G. F.; Franco, D. W.; Polyhedron 1999, 18, 979.

28. Perrin, D. D.; Armarego, W. L. F.; Purification of Laboratory Chemicals, 3rd ed., Pergamon Press: New York, 1983.

29. Kolthoff, I. M.; Quantitative Chemical Analysis, 4th ed., Macmillan: New York, 1969.

30. Shriver, D. F.; Drezdzon M. A.; The Manipulation of Air-Sensitive Compounds, 2nd ed., Wiley: New York, 1986.

31. Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Montgomery Jr., J. A.; Vreven, T.; Kudin, K. N.; Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Gonzalez, C.; Pople, J. A.; Gaussian 03 Revision C.02; Wallingford, 2004.

32. Becke, A. D.; J. Chem. Phys. 1993, 98, 5648.

33. Farkas, O.; Schlegel, H. B.; J. Chem. Phys. 1999, 111, 10806.

34. Cances, E.; Mennucci, B.; Tomasi, J.; J. Chem. Phys. 1997, 107, 3032; Reed, A. E.; Weinhold, F.; J. Chem. Phys. 1985, 83, 1736; Reed, A. E.; Curtiss, L. A.; Weinhold, F.; Chem. Rev. 1988, 88, 899.

35. Glendening, E. D.; Badenhoop, J. K.; Reed, A. E.; Carpenter, J. E.; Weinhold, F.; NBO4.M; NBO Analysis Programs; University of Wisconsin, USA, 1999.

Received: September 18, 2009

Web Release Date: April 15, 2010

FAPESP helped in meeting the publication costs of this article.

Supplementary Information

  • 1. Lundberg, J. O.; Weitzberg, E.; Gladwin, M. T.; Nat. Rev. Drug Discovery 2008, 7, 156.
  • 2. Li, H.; Igarashi, J.; Jamal, J.; Yang, W.; Poulos, T. L.; J. Biol. Inorg. Chem. 2006, 11, 753.
  • 3. Cullota, E.; Koshland, J. D. E.; Science 1992, 258, 1862.
  • 4. Toledo, J. C.; Silva, H. A. S.; Scarpellini, M.; Mori, V.; Camargo, A. J.; Bertotti, M.; Franco, D. W.; Eur. J. Inorg. Chem. 2004, 9, 1879.
  • 5. Tfouni, E.; Krieger, M.; McGarvey, B. R.; Franco, D. W.; Coord. Chem. Rev 2003, 236, 57.
  • 6. Clarke, M. J.; Coord. Chem. Rev 2003, 236, 209.
  • 7. McCleverty, J. A.; Chem. Rev. 2004, 104, 403.
  • 8. Franco, D. W.; Taube, H.; Inorg. Chem 1978, 17, 571;
  • Lopes, L. G. F.; Castellano, E. E.; Ferreira, A. G.; Davanzo, C. U.; Clarke, M. J.; Franco, D. W.; Inorg. Chim. Acta 2005, 358, 2883.
  • 9. Silva, J. J. N.; Osakabe, A. L.; Pavanelli, W. R.; Silva, J. S.; Franco, D. W.; Br. J. Pharmacol. 2007, 152, 112.
  • 10. Ishmaeva, E. A.; Popova, E. V.; Mironov, V. F.; Aminova, R. M.; Vereshchagina, Y. A.; Galkin, V. I.; Moeller, K.; Schmutzler, R.; Russ. J. Org. Chem. 2004, 40, 1076;
  • Doak, G. O.; Freedman, L. D.; Chem. Rev. 1961, 61, 31.
  • 11. Corbridge D. C. E.; Phosphorus: An Outline of its Chemistry, Biochemistry and Technology, 4th ed., Elsevier: Amsterdam, 1990.
  • 12. Kirby, A. J.; Warren, S. G.; The Organic Chemistry of Phosphorus, 1st ed., Elsevier: New York, 1967.
  • 13. Toledo, J. C.; Lima-Neto, B. S.; Franco, D. W.; Coord. Chem. Rev. 2005, 249, 419.
  • 14. Caramori, G. F.; Frenking, G.; Organometalics 2007, 26, 5815.
  • 15. Bezerra, C. W. B.; Silva, S. C.; Gambardella, M. T. P.; Santos, R. H. A.; Plicas L. M. A.; Tfouni, E.; Franco, D. W.; Inorg. Chem. 1999, 8, 5660.
  • 16. Nakamoto, K.; Infrared and Raman Spectra of Inorganic and Coordination Compounds, 4th ed., John Wiley & Sons: New York, 1986;
  • Socrates, G.; Infrared and Raman Characteristic Group Frequencies, 3rd ed., John Wiley & Sons: New York, 2001.
  • 17. Schreiner, A. F.; Lin, S. W.; Hausser, P. J.; Hopcus, E. A.; Hamm, D. J.; Gunter, J. D.; Inorg. Chem. 1972, 11, 880.
  • 18. Gorelsky, S. I.; Silva, S. C.; Lever, A. B. P.; Franco, D. W.; Inorg. Chim. Acta 2000, 300, 698.
  • 19. Verdake, J. G.; Quin, L. D.; Phophorus-31 NMR Spectroscopy on Stereochemical Analysis, VCH Paublishers: Deerfield Beach, 1987.
  • 20. Aguiar, M. R. M. P.; Gemal, A. L.; Gil, R. A. S. S.; Quim. Nova 1999, 22, 553;
  • Penner, G. H.; Wasylishen, R. E.; Can. J. Chem. 1989, 67, 1909;
  • Moynihan, H. A.; OHare, I. P.; Int. J. Pharmam. 2002, 247, 179.
  • 21. Sernaglia, R. L.; Franco, D. W.; Inorg. Chem. 1989, 28, 3485.
  • 22. Truzzi, D. R.; Franco, D. W.; Abstracts of the XIV Brazilian Meeting in Inorganic Chemistry, Foz de Iguaçu, Brazil, 2008.
  • 23. Xi, C.; Liu, Y.; Lai, C.; Zhou, L.; Inorg. Chem. Commun. 2004, 7, 1202.
  • 24. Han, L. B.; Tanaka, M.; J. Am. Chem. Soc. 1996, 118, 1571.
  • 25. Brill, T. E.; Landon, S. J.; Chem. Rev. 1984, 84, 577;
  • Bezerra, C. W. B.; Silva, S. C.; Gambardella, M. T. P.; Santos, R. H. A.; Plicas, M. L.; Tfouni, E.; Franco, D. W.; Inorg. Chem. 1999, 38, 5660.
  • 26. Koleva, G.; Galabov, B.; Wu, J. I.; Schaefer, H. F.; Schleyer, P. R.; J. Am. Chem. Soc. 2009, 131, 14772.
  • 27. Mazzetto, S. E.; Gambardella, M. T. P.; Santos, R. H. A.; Lopes, L. G. F.; Franco, D. W.; Polyhedron 1999, 18, 979.
  • 28. Perrin, D. D.; Armarego, W. L. F.; Purification of Laboratory Chemicals, 3rd ed., Pergamon Press: New York, 1983.
  • 29. Kolthoff, I. M.; Quantitative Chemical Analysis, 4th ed., Macmillan: New York, 1969.
  • 30. Shriver, D. F.; Drezdzon M. A.; The Manipulation of Air-Sensitive Compounds, 2nd ed., Wiley: New York, 1986.
  • 31. Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Montgomery Jr., J. A.; Vreven, T.; Kudin, K. N.; Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Gonzalez, C.; Pople, J. A.; Gaussian 03 Revision C.02; Wallingford, 2004.
  • 32. Becke, A. D.; J. Chem. Phys. 1993, 98, 5648.
  • 33. Farkas, O.; Schlegel, H. B.; J. Chem. Phys. 1999, 111, 10806.
  • 34. Cances, E.; Mennucci, B.; Tomasi, J.; J. Chem. Phys 1997, 107, 3032;
  • Reed, A. E.; Weinhold, F.; J. Chem. Phys. 1985, 83, 1736;
  • Reed, A. E.; Curtiss, L. A.; Weinhold, F.; Chem. Rev. 1988, 88, 899.
  • 35. Glendening, E. D.; Badenhoop, J. K.; Reed, A. E.; Carpenter, J. E.; Weinhold, F.; NBO4.M; NBO Analysis Programs; University of Wisconsin, USA, 1999.
  • *
    e-mail:
  • Publication Dates

    • Publication in this collection
      30 July 2010
    • Date of issue
      2010

    History

    • Received
      18 Sept 2009
    • Accepted
      15 Apr 2010
    Sociedade Brasileira de Química Instituto de Química - UNICAMP, Caixa Postal 6154, 13083-970 Campinas SP - Brazil, Tel./FAX.: +55 19 3521-3151 - São Paulo - SP - Brazil
    E-mail: office@jbcs.sbq.org.br