Network


Latest external collaboration on country level. Dive into details by clicking on the dots.

Hotspot


Dive into the research topics where Andreja Bakac is active.

Publication


Featured researches published by Andreja Bakac.


Chemical Science | 2012

pH-induced mechanistic changeover from hydroxyl radicals to iron(IV) in the Fenton reaction

Hajem Bataineh; Oleg Pestovsky; Andreja Bakac

A major pathway in the reaction between Fe(II) and H2O2 at pH 6–7 in non-coordinating buffers exhibits inverse kinetic dependence on [H+] and leads to oxidation of dimethyl sulfoxide (DMSO) to dimethyl sulfone (DMSO2). This step regenerates Fe(II) and makes the oxidation of DMSO catalytic, a finding that strongly supports Fe(IV) as a Fenton intermediate at near-neutral pH. This Fe(IV) is a less efficient oxidant for DMSO at pH 6–7 than is (H2O)5FeO2+, generated by ozone oxidation of Fe(H2O)62+, in acidic solutions. Large concentrations of DMSO are needed to achieve significant turnover numbers at pH ≥ 6 owing to the rapid competing reaction between Fe(II) and Fe(IV) that leads to irreversible loss of the catalyst. At pH 6 and ≤0.02 mM Fe(II), the ratio of apparent rate constants for the reactions of Fe(IV) with DMSO and with Fe(II) is ∼104. The results at pH 6–7 stand in stark contrast with those reported previously in acidic solutions where the Fenton reaction generates hydroxyl radicals. Under those conditions, DMSO is oxidized stoichiometrically to methylsulfinic acid and ethane. This path still plays a role (1–10%) at pH 6–7.


Inorganic Chemistry | 2010

Oxygen activation with transition-metal complexes in aqueous solution.

Andreja Bakac

Coordination to transition-metal complexes changes both the thermodynamics and kinetics of oxygen reduction. Some of the intermediates (superoxo, hydroperoxo, and oxo species) are close analogues of organic oxygen-centered radicals and peroxides (ROO(*), ROOH, and RO(*)). Metal-based intermediates are typically less reactive, but more persistent, than organic radicals, which makes the two types of intermediates similarly effective in their reactions with various substrates. The self-exchange rate constant for hydrogen-atom transfer for the couples Cr(aq)OO(2+)/Cr(aq)OOH(2+) and L(1)(H(2)O)RhOO(2+)/L(1)(H(2)O)RhOOH(2+) was estimated to be 10(1+/-1) M(-1) s(-1). The use of this value in the simplified Marcus equation for the Cr(aq)O(2+)/Cr(aq)OOH(2+) cross reaction provided an upper limit k(CrO,CrOH) <or= 10((-2+/-1)) M(-1) s(-1) for Cr(aq)O(2+)/Cr(aq)OH(2+) self-exchange. Even though superoxo complexes react very slowly in bimolecular self-reactions, extremely fast cross reactions with organic counterparts, i.e., acylperoxyl radicals, have been observed. Many of the intermediates generated by the interaction of O(2) with reduced metal complexes can also be accessed by alternative routes, both thermal and photochemical.


Inorganic Chemistry | 2010

Base-Catalyzed Insertion of Dioxygen into Rhodium−Hydrogen Bonds: Kinetics and Mechanism

Ewa Szajna-Fuller; Andreja Bakac

The reaction between molecular oxygen and rhodium hydrides L(OH)RhH(+) (L = (NH(3))(4), trans-L(1), and cis-L(1), where L(1) = cyclam) in basic aqueous solutions rapidly produces the corresponding hydroperoxo complexes. Over the pH range 8 < pH < 12, the kinetics exhibit a first order dependence on [OH(-)]. The dependence on [O(2)] is less than first order and approaches saturation at the highest concentrations used. These data suggest an attack by OH(-) at the hydride with k = (1.45 +/- 0.25) x 10(3) M(-1) s(-1) for trans-L(1)(OH)RhH(+) at 25 degrees C, resulting in heterolytic cleavage of the Rh-H bond and formation of a reactive Rh(I) intermediate. A competition between O(2) and H(2)O for Rh(I) is the source of the observed dependence on O(2). In support of this mechanism, there is a significant kinetic isotope effect for the initial step, L(1)(OH(D))RhH(D)(+) + OH(D)(-) k(1)/k(-1) L(1)(OH(D))Rh(I) + H(D)(2)O, k(1H)/k(1D) = 1.7, and k(-1H)/k(-1D) = 3.0. The activation parameters for k(1) for trans-L(1)(OH)RhH(+) are DeltaH(++) = 64.6 +/- 1.3 kJ mol(-1) and DeltaS(++) = 40 +/-4 J mol(-1) K(-1).


Journal of the American Chemical Society | 2008

Photochemical oxidation of halide ions by a nitratochromium(III) complex. Kinetics, mechanism, and intermediates.

Mingming Cheng; Andreja Bakac

The 266 nm laser flash photolysis of the title complex in the presence of halide ions X(-) (X = I, Br, and Cl) generates halogen atoms on nanosecond time scales, followed by the known X(*)/X(-) reactions to yield dihalide radical anions, X2(*-). Plots of k(obs) against the concentration of X(-) were linear with zero intercepts, but the yields of X2(*-) increased with increasing concentrations of [X(-)]. This result suggests that a short-lived, strongly oxidizing intermediate reacts with X(-) to generate X(*) in parallel with the decomposition to Cr(aq)O(2+) and (*)NO2, both of which were identified in steady-state photolysis experiments in the presence of selective trapping agents. Bromide was oxidized quantitatively to bromine, and a combination of molecular oxygen and methanol channeled the reaction toward superoxochromium(III) ion, Cr(aq)OO(2+). In the absence of scavengers, nitrite and chromate were produced. The proposed reaction scheme draws additional support from the good agreement between the experimental product yields and those predicted by kinetic simulations.


Inorganic Chemistry | 2010

Preparation, Crystal Structure, and Unusually Facile Redox Chemistry of a Macrocyclic Nitrosylrhodium Complex

Kathleen E. Kristian; Wenjing Song; Arkady Ellern; Ilia A. Guzei; Andreja Bakac

The reaction between NO and L(2)(H(2)O)Rh(2+) (L(2) = meso-Me(6)-1,4,8,11-tetraazacyclotetradecane) generates a sky-blue L(2)(H(2)O)RhNO(2+), a {RhNO}(8) complex. The crystal structure of the perchlorate salt features a bent Rh-N-O moiety (122.1(11)(o)), short axial Rh-NO bond (1.998(12) A) and a strongly elongated Rh-OH(2) (2.366(6) A) trans to NO. Acidic aqueous solutions of L(2)(H(2)O)RhNO(2+) are stable for weeks, and are inert toward oxygen. The complex is oxidized rapidly and reversibly with Ru(bpy)(3)(3+), k(f) = (1.9 +/- 0.1) x 10(5) M(-1) s(-1), to an intermediate believed to be L(2)(H(2)O)RhNO(3+). This unprecedented {RhNO}(7) species has a lifetime of about 90 s at room temperature at pH 0. The reverse reaction between L(2)(H(2)O)RhNO(3+) and Ru(bpy)(3)(2+) has k(r) = (1.5 +/- 0.4) x 10(6) M(-1) s(-1). The kinetic data define the equilibrium constant for the L(2)(H(2)O)RhNO(2+)/Ru(bpy)(3)(3+) reaction, K = k(f)/k(r) = 0.13, and yield a reduction potential for the L(2)(H(2)O)RhNO(3+/2+) couple of 1.31 V. Both the redox thermodynamics of L(2)(H(2)O)RhNO(3+/2+) and the kinetics of the reactions with Ru(bpy)(3)(3+/2+) are quite similar to those of uncoordinated NO(+)/NO.


Inorganic Chemistry | 2009

Oxidation of a water-soluble phosphine and some spectroscopic probes with nitric oxide and nitrous acid in aqueous solutions.

Andreja Bakac; Margaret Schouten; Alicia Johnson; Wenjing Song; Oleg Pestovsky; Ewa Szajna-Fuller

In acidic aqueous solutions, nitrogen monoxide oxidizes monosulfonated triphenylphosphine, TPPMS(-), to the corresponding phosphine oxide. The NO-derived product is N(2)O. This chemistry parallels that reported for the reaction of NO with the unsubstituted triphenylphosphine in nonpolar organic solvents, but the rate constant measured in this work, 5.14 x 10(6) M(-2) s(-1), is greater by several orders of magnitude. This makes TPPMS(-) a useful analytical reagent for NO in aqueous solution. The increased rate constant in the present work appears to be a medium effect, and unrelated to the introduction of a single sulfonate group in the phosphine. The reaction between nitrous acid and TPPMS(-) has a 2:1 [TPPMS(-)]/[HNO(2)] stoichiometry and generates NH(2)OH quantitatively. The rate law, rate = 4k(d)[HNO(2)](2)[TPPMS(-)], identifies the second-order self-reaction of HNO(2) as the rate-limiting step that generates the active oxidant(s) for the fast subsequent reaction with TPPMS(-). It appears that the active oxidant is N(2)O(3), although the oxides NO and NO(2) derived from it may be also involved. Bimolecular self-reaction of HNO(2) also precedes the oxidations of ABTS(2-) and TMPD. Competing with this path are the acid-catalyzed oxidations of both reagents via NO(+).


Inorganic Chemistry | 2011

Synthesis of Monomeric Fe(II) and Ru(II) Complexes of Tetradentate Phosphines

Barun Jana; Arkady Ellern; Oleg Pestovsky; Aaron D. Sadow; Andreja Bakac

rac-Bis[{(diphenylphosphino)ethyl}-phenylphosphino]methane (DPPEPM) reacts with iron(II) and ruthenium(II) halides to generate complexes with folded DPPEPM coordination. The paramagnetic, five-coordinate Fe(DPPEPM)Cl(2) (1) in CD(2)Cl(2) features a tridentate binding mode as established by (31)P{(1)H} NMR spectroscopy. Crystal structure analysis of the analogous bromo complex, Fe(DPPEPM)Br(2) (2) revealed a pseudo-octahedral, cis-α geometry at iron with DPPEPM coordinated in a tetradentate fashion. However, in CD(2)Cl(2) solution, the coordination of DPPEPM in 2 is similar to that of 1 in that one of the external phosphorus atoms is dissociated resulting in a mixture of three tridentate complexes. The chloro ruthenium complex cis-Ru(κ(4)-DPPEPM)Cl(2) (3) is obtained from rac-DPPEPM and either [RuCl(2)(COD)](2) [COD = 1,5-cyclooctadiene] or RuCl(2)(PPh(3))(4). The structure of 3 in both the solid state and in CD(2)Cl(2) solution features a folded κ(4)-DPPEPM. This binding mode was also observed in cis-[Fe(κ(4)-DPPEPM)(CH(3)CN)(2)](CF(3)SO(3))(2) (4). Addition of an excess of CO to a methanolic solution of 1 results in the replacement of one of the chloride ions by CO to yield cis-[Fe(κ(4)-DPPEPM)Cl(CO)](Cl) (5). The same reaction in CH(2)Cl(2) produces a mixture of 5 and [Fe(κ(3)-DPPEPM)Cl(2)(CO)] (6) in which one of the internal phosphines has been substituted by CO. Complexes 2, 3, 4, and 5 appear to be the first structurally characterized monometallic complexes of κ(4)-DPPEPM.


Journal of Photochemistry and Photobiology A-chemistry | 2000

Photoinduced insertion of molecular oxygen into metal–hydrogen bonds

Andreja Bakac

Abstract Rhodium(III) hydrides L(H 2 O)RhH 2+ (L=(NH 3 ) 4 and [14] aneN 4 ) are efficient quenchers for the excited uranyl ions, ∗ UO 2 2+ . In the presence of molecular oxygen, the reaction ultimately yields the hydroperoxorhodium complexes, L(H 2 O)RhOOH 2+ . The replacement of the coordinated hydride by deuteride and the solvent H 2 O by D 2 O results in a combined kinetic isotope effect of 1.5 for the macrocyclic complex. The proposed mechanism features L(H 2 O)Rh 2+ , L(H 2 O)RhOO 2+ , and UO 2 + as intermediates.


Inorganic Chemistry | 2008

Electron-transfer reactions of nitrosyl and superoxo metal complexes.

Wenjing Song; Arkady Ellern; Andreja Bakac

Novel chromium nitrosyl complexes L(H2O)CrNO(2+) (L = L(1) = 1,4,8,11-tetraazacyclotetradecane, L(2) = meso-Me 6-1,4,8,11-tetraazacyclotetradecane) are oxidized by Ru(bpy)3(3+) to LCr(H2O)2(3+) and NO with rate constants k = 2.22 M(-1) s(-1) (L(1)) and 6.83 (L(2)). Analogous reactions of the superoxo complexes L(H2O)CrOO(2+) are only slightly faster, k = 45 M(-1) s(-1) (L(1)) and 15 M(-1) s(-1) (L(2)). A related rhodium complex L(2)(H2O)RhOO(2+) has k = 15.8 M(-1) s(-1). These results, combined with our earlier data for the oxidation of Cr(aq)NO(2+) and Cr(aq)OO(2+), suggest only a modest role for thermodynamics in determining the kinetics of oxidation. This behavior is even more pronounced in the oxidation of rhodium hydrido and hydroperoxo complexes, with the latter reacting more than 10(5)-fold faster despite being thermodynamically less favored by more than 0.3 V. The X-ray crystal structure of [L(1)(H2O)CrNO](ClO4)2 supports the limiting Cr(III)-NO(-) description for the complex cation.


Chemistry: A European Journal | 2011

Visible Light-Induced Release of Nitrogen Monoxide from a Nitrosylrhodium Complex

Wenjing Song; Kathleen E. Kristian; Andreja Bakac

The important roles that nitric oxide (NO) plays in biological environments, and the need for precise and targeted delivery of NO for medicinal and other purposes have led to intense research in the area of metal nitrosyl complexes as thermal and photochemical sources of NO. Complexes with a good combination of chemical stability and high quantum yield for photochemical release of NO upon irradiation with visible light in aqueous solutions are rare. Here we report that a simple macrocyclic nitrosylrhodium complex [L(2)(H(2)O)Rh(NO)](2+) (L(2)=Me(6)[14]aneN(4)) exhibits unique chemical and photochemical properties that make it an excellent photochemical precursor of NO. The complex is highly soluble in water, thermally stable, and resistant toward O(2). Irradiation in the 648 nm band generates NO and [L(2)(H(2)O)Rh](2+) in aqueous solutions with a quantum yield of 1.00±0.07, the highest ever reported for a nitrosyl complex under any conditions. In the absence of O(2), the two fragments combine to regenerate [L(2)(H(2)O)Rh- (NO)](2+), but in O(2)-containing solutions, [L(2)(H(2)O)RhOO](2+) is formed as determined in spectral and kinetic measurements. The kinetics of the reaction of this superoxo complex with NO were measured by laser flash photolysis, k=(3.9±0.4)×10(7) M(-1) s(-1). Steady-state photolysis of [L(2)(H(2)O)Rh(NO)](2+) under O(2) yielded [L(2)(H(2)O)Rh(ONO(2))](2+), a long-lived nitrato intermediate that can also be generated in a direct reaction between NO and genuine [L(2)(H(2)O)RhOO](2+). Thus, visible-light photolysis of the [L(2)(H(2)O)Rh(NO)](2+)/O(2) system converts it to the [L(2)(H(2)O)RhOO](2+)/NO combination.

Collaboration


Dive into the Andreja Bakac's collaboration.

Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Researchain Logo
Decentralizing Knowledge