Partha Basu

Find an error

Name: Basu, Partha
Organization: Duquesne University Mellon Hall , USA
Department: Department of Chemistry and Biochemistry
Title: (PhD)
Co-reporter:Jing Yang; Benjamin Mogesa; Partha Basu;Martin L. Kirk
Inorganic Chemistry 2016 Volume 55(Issue 2) pp:785-793
Publication Date(Web):December 21, 2015
DOI:10.1021/acs.inorgchem.5b02252
Interligand charge transfer is examined in the novel metallo-dithiolene complex MoO(SPh)2(iPr2Dt0) (where iPr2Dt0 = N,N′-isopropyl-piperazine-2,3-dithione). The title complex displays a remarkable 70° “envelope”-type fold of the five-membered dithiolene ring, which is bent upward toward the terminal oxo ligand. A combination of electronic absorption and resonance Raman spectroscopies have been used to probe the basic electronic structure responsible for the large fold-angle distortion. The intense charge transfer transition observed at ∼18 000 cm–1 is assigned as a thiolate → dithione ligand-to-ligand charge transfer (LL′CT) transition that also possesses Mo(IV) → dithione charge transfer character. Strong orbital mixing between occupied and virtual orbitals with Mo(x2–y2) orbital character is derived from a strong pseudo Jahn–Teller effect, which drives the large fold-angle distortion to yield a double-well potential in the electronic ground state.
Co-reporter:Stephen C. Ratvasky, Benjamin Mogesa, Michael J. van Stipdonk, Partha Basu
Polyhedron 2016 Volume 114() pp:370-377
Publication Date(Web):16 August 2016
DOI:10.1016/j.poly.2016.01.023
Neutral complexes of zinc with N,N′-diisopropylpiperazine-2,3-dithione (iPr2Dt0) and N,N′-dimethylpiperazine-2,3-dithione (Me2Dt0) with chloride or maleonitriledithiolate (mnt2−) as coligands have been synthesized and characterized. The molecular structures of these zinc complexes have been determined using single crystal X-ray diffractometry. Complexes recrystallize in monoclinic P type systems with zinc adopting a distorted tetrahedral geometry. Two zinc complexes with mixed-valence dithiolene ligands exhibit ligand-to-ligand charge transfer bands. Optimized geometries, molecular vibrations and electronic structures of charge-transfer complexes were calculated using density functional theory (B3LYP/6-311G+(d,p) level). Redox orbitals are shown to be almost exclusively ligand in nature, with a HOMO based heavily on the electron-rich maleonitriledithiolate ligand, and a LUMO comprised mostly of the electron-deficient dithione ligand. Charge transfer is thus believed to proceed from dithiolate HOMO to dithione LUMO, showing ligand-to-ligand redox interplay across a d10 metal.Crystallographically characterized bis-dithiolene Zn complexes with ligands of differing oxidation states exhibit low energy ligand to ligand (LLCT) charge transfer band.
Co-reporter:Benjamin Mogesa; Eranda Perera; Hannah M. Rhoda; John K. Gibson; Jos Oomens; Giel Berden; Michael J. van Stipdonk; Victor N. Nemykin
Inorganic Chemistry 2015 Volume 54(Issue 16) pp:7703-7716
Publication Date(Web):August 5, 2015
DOI:10.1021/acs.inorgchem.5b00531
The syntheses of cationic nickel complexes using N,N′-dimethyl piperazine 2,3-dithione (Me2Dt0) and N,N′-diisopropyl piperazine 2,3-dithione (iPr2Dt0) ligands are reported. These ligands were used in synthesizing bis and tris(dithione)Ni(II) complexes as tetrafluoroborate or hexafluorophosphate salts, i.e., [Ni(iPr2Dt0)2][BF4]2 ([1a][BF4]2), [Ni(iPr2Dt0)2][PF6]2 ([1a][PF6]2), [Ni(Me2Dt0)2][BF4]2 ([1b][BF4]2), [Ni(iPr2Dt0)3][BF4]2 ([2a][BF4]2), and [Ni(iPr2Dt0)3][PF6]2 ([2a][PF6]2), respectively. Complex [2a][PF6]2 was isolated from a methanolic solution of [1a][PF6]2. Compound [1a][BF4]2 crystallizes in a trigonal crystal system (space group, P31/c) and exhibits unique packing features, whereas [2a][BF4]2 crystallizes in a monoclinic (P21/n) space group. Cyclic voltammograms of [1a][BF4]2 and [1b][BF4]2 are indicative of four reduction processes associated with stepwise single-electron reduction of the ligands. Spectroelectrochemical experiments on [1a][BF4]2 exhibit an intervalence charge transfer (IVCT) transition as a spectroscopic signature of the mixed-valence [Ni(iPr2Dt0)(iPr2Dt1–)]− species. Analysis of this IVCT band suggests that this ligand based mixed valence complex, [Ni(iPr2Dt0)(iPr2Dt1–)]−, behaves more like a traditional class II/III metal based mixed-valence complex. The density functional theory (DFT) and time dependent DFT calculations provide a theoretical framework for understanding the electronic structures and the nature of excited states of the target compounds that are consistent with their spectroscopic and redox properties. Vibrational spectra of [1a]2+ and [2a]2+ were investigated as discrete species in the gas phase using infrared multiple photon dissociation (IRMPD) spectroscopy.
Co-reporter:Dr. Igor V. Pimkov;Dr. Barbara Serli-Mitasev;Antoinette A. Peterson;Stephen C. Ratvasky;Dr. Bernd Hammann ; Partha Basu
Chemistry - A European Journal 2015 Volume 21( Issue 47) pp:17057-17072
Publication Date(Web):
DOI:10.1002/chem.201502845

Abstract

Molybdopterin is an essential cofactor for all forms of life. The cofactor is composed of a pterin moiety appended to a dithiolene-functionalized pyran ring, and through the dithiolene moiety it binds metal ions. Different synthetic strategies for dithiolene-functionalized pyran precursors that have been designed and synthesized are discussed. These precursors also harbor 1,2-diketone or osone functionality that has been condensed with 1,2-diaminobenzene or other heterocycles resulting in several quinoxaline or pterin derivatives. Use of additives improves the regioselectivity of the complexes. The molecules have been characterized by 1H and 13C NMR and IR spectroscopies, as well as by mass spectrometry. In addition, several compounds have been crystallographically characterized. The geometries of the synthesized molecules are more planar than the geometry of the cofactor found in proteins.

Co-reporter:Russ Hille, James Hall, and Partha Basu
Chemical Reviews 2014 Volume 114(Issue 7) pp:3963
Publication Date(Web):January 28, 2014
DOI:10.1021/cr400443z
Co-reporter:Courtney Sparacino-Watkins, John F. Stolz and Partha Basu  
Chemical Society Reviews 2014 vol. 43(Issue 2) pp:676-706
Publication Date(Web):18 Oct 2013
DOI:10.1039/C3CS60249D
The nitrate anion is a simple, abundant and relatively stable species, yet plays a significant role in global cycling of nitrogen, global climate change, and human health. Although it has been known for quite some time that nitrate is an important species environmentally, recent studies have identified potential medical applications. In this respect the nitrate anion remains an enigmatic species that promises to offer exciting science in years to come. Many bacteria readily reduce nitrate to nitrite via nitrate reductases. Classified into three distinct types – periplasmic nitrate reductase (Nap), respiratory nitrate reductase (Nar) and assimilatory nitrate reductase (Nas), they are defined by their cellular location, operon organization and active site structure. Of these, Nap proteins are the focus of this review. Despite similarities in the catalytic and spectroscopic properties Nap from different Proteobacteria are phylogenetically distinct. This review has two major sections: in the first section, nitrate in the nitrogen cycle and human health, taxonomy of nitrate reductases, assimilatory and dissimilatory nitrate reduction, cellular locations of nitrate reductases, structural and redox chemistry are discussed. The second section focuses on the features of periplasmic nitrate reductase where the catalytic subunit of the Nap and its kinetic properties, auxiliary Nap proteins, operon structure and phylogenetic relationships are discussed.
Co-reporter:John A. Thomas, Peter Chovanec, John F. Stolz and Partha Basu  
Metallomics 2014 vol. 6(Issue 10) pp:1958-1969
Publication Date(Web):14 Aug 2014
DOI:10.1039/C4MT00185K
Alkaliphilus oremlandii strain OhILAs, a gram-positive bacterium, has been shown to ferment lactate as well as use arsenate and roxarsone as a terminal electron acceptor. This study examines the proteome expressed under four growth conditions to further elucidate the bacterial metabolism of inorganic and organic arsenic. The four growth conditions include, sodium lactate (as fermentative control), sodium lactate with 3-nitro-4-hydroxybenzenearsonic acid (roxarsone), sodium lactate with 3-amino-4-hydroxybenzenearsonic acid (3A4HBAA), and sodium lactate with sodium arsenate. Shotgun proteomics using LC-MS/MS was performed on the soluble cytoplasm as well as solubilized membrane proteins using perfluorooctanoic acid, a surfactant with properties similar to sodium dodecyl sulfate. The MS/MS data were analyzed using the Spectrum Mills Proteomic Workbench. Positive protein matches were confirmed with protein scores of 20 or greater and the presence of two or more peptides among the three technical replicates. A total of 1357 proteins (out of 2836 predicted) were identified with 791 in sodium lactate, 816 in sodium lactate and roxarsone, 715 in sodium lactate and 3A4HBAA, and 733 in sodium lactate and arsenate. The relative abundance of each protein was determined using a method called normalized spectral abundance factor (NSAF). Proteins that were identified in both the control and the experimental conditions were compared using the Power Law Global Error Model (PLGEM) to determine proteins that were significantly up or down regulated. All putative proteins were assigned functions and pathways using the COG databases. However, a large number of proteins were classified as hypothetical or had unknown function. Using the statistical information and known functionalities of the identified proteins, a pathway for the degradation of roxarsone and 3A4HBAA by A. oremlandii strain OhILAs is proposed.
Co-reporter:Igor V. Pimkov, Antoinette A. Peterson, David N. Vaccarello and Partha Basu  
RSC Advances 2014 vol. 4(Issue 37) pp:19072-19076
Publication Date(Web):11 Apr 2014
DOI:10.1039/C4RA02786H
The molybdopterin (MPT) is a complex molecule, made out of three distinctly different components. A retrosynthetic analysis provides a possible route for its synthesis that utilizes the coupling of a diamine with an osone analog. A regioselective condensation of the diamine with an osone affords the dephospho MPT, which has been characterized by NMR and IR spectroscopies, as well as high-resolution mass spectrometry.
Co-reporter:Michael J. van Stipdonk, Partha Basu, Sara A. Dille, John K. Gibson, Giel Berden, and Jos Oomens
The Journal of Physical Chemistry A 2014 Volume 118(Issue 29) pp:5407-5418
Publication Date(Web):July 2, 2014
DOI:10.1021/jp503222v
Electrospray ionization (ESI) in the negative ion mode was used to create anionic, gas-phase oxo-molybdenum complexes with dithiolene ligands. By varying ESI and ion transfer conditions, both doubly and singly charged forms of the complex, with identical formulas, could be observed. Collision-induced dissociation (CID) of the dianion generated exclusively the monoanion, while fragmentation of the monoanion involved decomposition of the dithiolene ligands. The intrinsic structure of the monoanion and the dianion were determined by using wavelength-selective infrared multiple-photon dissociation (IRMPD) spectroscopy and density functional theory calculations. The IRMPD spectrum for the dianion exhibits absorptions that can be assigned to (ligand) C═C, C–S, C—C≡N, and Mo═O stretches. Comparison of the IRMPD spectrum to spectra predicted for various possible conformations allows assignment of a pseudo square pyramidal structure with C2v symmetry, equatorial coordination of MoO2+ by the S atoms of the dithiolene ligands, and a singlet spin state. A single absorption was observed for the oxidized complex. When the same scaling factor employed for the dianion is used for the oxidized version, theoretical spectra suggest that the absorption is the Mo═O stretch for a distorted square pyramidal structure and doublet spin state. A predicted change in conformation upon oxidation of the dianion is consistent with a proposed bonding scheme for the bent-metallocene dithiolene compounds [Lauher, J. W.; Hoffmann, R. J. Am. Chem. Soc. 1976, 98, 1729−1742], where a large folding of the dithiolene moiety along the S···S vector is dependent on the occupancy of the in-plane metal d-orbital.
Co-reporter:Partha Basu, Brian W. Kail, Andrew K. Adams and Victor N. Nemykin  
Dalton Transactions 2013 vol. 42(Issue 9) pp:3071-3081
Publication Date(Web):08 Nov 2012
DOI:10.1039/C2DT32349D
The oxygen atom transfer reactivity (OAT) of dioxo-Mo(VI) complexes of hydrotrispyrazolyl borate (hydrotris(3,5-dimethylpyrazolyl)borate, TpMe2; hydrotris(3-isopropylpyrazol-1-yl)borate, TpiPr) with tertiary phosphines (PMe3, PMe2Ph, PEt3, PEt2Ph, PBun3, PMePh2, or PEtPh2) has been investigated. In acetonitrile, these reactions proceed via the formation of a phosphoryl intermediate complex that undergoes a solvolysis reaction. We report the synthesis and characterization of several phosphoryl complexes. The rates of formation of phosphoryl complexes and their solvation were determined by spectrophotometry. The rates of the reactions and the properties of the phosphoryl species were investigated using the Quantitative Analysis of Ligand Effect (QALE) methodology. The results show that, at least in this system, the first step of the reaction is controlled primarily by the steric factor, and in the second step, both electronic and steric factors are important. We also analyzed the effect of ligands on the reaction rate i.e., TpMe2vs. TpiPr.
Co-reporter:Kristine Deibler
European Journal of Inorganic Chemistry 2013 Volume 2013( Issue 7) pp:
Publication Date(Web):
DOI:10.1002/ejic.201300140

Abstract

Invited for the cover of this issue is the group of Partha Basu at Duquesne University. The cover image shows the various sources of contamination of water by lead and the detection of Pb2+ by a fluorescent sensor called Leadglow.

Co-reporter:Kristine Deibler
European Journal of Inorganic Chemistry 2013 Volume 2013( Issue 7) pp:1086-1096
Publication Date(Web):
DOI:10.1002/ejic.201200997

Abstract

In the past, Pb2+ was used in many industries, including gasoline, piping, toys, paints, and more. The use of lead has led to a natural increase in lead concentrations in the environment, especially in air and water. According to the U.S. Center for Disease Control “no level of lead in blood is considered safe.” Exposure to very low levels of lead can cause several health complications including developmental and neurological disorders. Over the past several years an emphasis has been placed on developing systems that can detect lead at very low concentrations. A great deal of work has been accomplished in the development of Pb2+ sensors that can not only detect but also quantify, and in some cases in the presence of other metal ions, the amount present. Herein, we describe current regulations, modes of exposure, and recent developments in sensing techniques.

Co-reporter:Kristine Deibler
European Journal of Inorganic Chemistry 2013 Volume 2013( Issue 7) pp:
Publication Date(Web):
DOI:10.1002/ejic.201390027
Co-reporter:Igor V. Pimkov;Archana Nigam;Kiran Venna;Fraser F. Fleming;Pavlo V. Solntsev;Victor N. Nemykin
Journal of Heterocyclic Chemistry 2013 Volume 50( Issue 4) pp:879-886
Publication Date(Web):
DOI:10.1002/jhet.1715

The bicyclic pyran thiolone tetrahydro-3αH-[1,3]dithiolo[4,5-β]pyran-2-thione (3a) engages in a highly unusual fragmentation in the presence of DDQ. The pyran thiolone, 3a, was synthesized by chlorination of 3,4-dihydro-2H-pyran (1) followed by condensing with CS2 and NaSH. Reaction of 3a with DDQ generates the isomerized pyran thiolone tetrahydro-3αH-[1,3]dithiolo[4,5-β]pyran-2-thione (3b) and 4-benzyl-5-(3-hydroxypropyl)-1,3-dithiole-2-thione (4) via a deep-seated rearrangement. The identity of 3b was confirmed by single crystal X-ray analysis: P21/c, a = 5.807(9) Å, b = 12.99(2) Å, c = 11.445(15), β = 113.23(6)°. Mechanistic experiments and computational insight is used to explain the likely sequence of events in the highly unusual formation of 4. Collectively, these results establish fundamental reactivity patterns for further research in this area.

Co-reporter:Partha Basu, Sharon J.N. Burgmayer
Coordination Chemistry Reviews 2011 Volume 255(9–10) pp:1016-1038
Publication Date(Web):May 2011
DOI:10.1016/j.ccr.2011.02.010
The molybdenum cofactor is composed of a molybdenum coordinated by one or two rather complicated ligands known as either molybdopterin or pyranopterin. Pterin is one of a large family of bicyclic N-heterocycles called pteridines. Such molecules are widely found in Nature, having various forms to perform a variety of biological functions. This article describes the basic nomenclature of pterin, their biological roles, structure, chemical synthesis and redox reactivity. In addition, the biosynthesis of pterins and current models of the molybdenum cofactor are discussed.
Co-reporter:Partha Basu, Brian W. Kail and Charles G. Young
Inorganic Chemistry 2010 Volume 49(Issue 11) pp:4895-4900
Publication Date(Web):April 30, 2010
DOI:10.1021/ic902500h
The oxygen atom transfer reactivity of the dioxo-Mo(VI) complex, TpiPrMoO2(OPh) (TpiPr = hydrotris(3-isopropylpyrazol-1-yl)borate), with a range of tertiary phosphines (PMe3, PMe2Ph, PEt3, PBun3, PEt2Ph, PEtPh2, and PMePh2) has been investigated. The first step in all the reactions follows a second-order rate law indicative of an associative transition state, consistent with nucleophilic attack by the phosphine on an oxo ligand, namely, TpiPrMoO2(OPh) + PR3 → TpiPrMoO(OPh)(OPR3). The calculated free energy of activation for the formation of the OPMe3 intermediate (Chem. Eur. J. 2006, 12, 7501) is in excellent agreement with the experimental ΔG⧧ value reported here. The second step of the reaction, that is, the exchange of the coordinated phosphine oxide by acetonitrile, TpiPrMoO(OPh)(OPR3) + MeCN → TpiPrMoO(OPh)(MeCN) + OPR3, is first-order in starting complex in acetonitrile. The reaction occurs via a dissociative interchange (Id) or associative interchange (Ia) mechanism, depending on the nature of the phosphine oxide. The activation parameters for the solvolysis of TpiPrMoO(OPh)(OPMe3) (ΔH⧧ = 56.3 kJ mol−1; ΔS⧧ = −125.9 J mol−1 K−1; ΔG⧧ = 93.8 kJ mol−1) and TpiPrMoO(OPh)(OPEtPh2) (ΔH⧧ = 66.5 kJ mol−1; ΔS⧧ = −67.6 J mol−1 K−1; ΔG⧧ = 86.7 kJ mol−1) by acetonitrile are indicative of Ia mechanisms. In contrast, the corresponding parameters for the solvolysis reaction of TpiPrMoO(OPh)(OPEt3) (ΔH⧧ = 95.8 kJ mol−1; ΔS⧧ = 26.0 J mol−1 K−1; ΔG⧧ = 88.1 kJ mol−1) and the remaining complexes by the same solvent are indicative of an Id mechanism. The equilibrium constant for the solvolysis of the oxo-Mo(V) phosphoryl complex, [TpiPrMoO(OPh)(OPMe3)]+, by acetonitrile was calculated to be 1.9 × 10−6. The oxo-Mo(V) phosphoryl complex is more stable than the acetonitrile analogue, whereas the oxo-Mo(IV) acetonitrile complex is more stable than the phosphoryl analogue. The higher stability of the Mo(V) phosphoryl complex may explain the phosphate inhibition of sulfite oxidase.
Co-reporter:Peter Chovanec, John F. Stolz and Partha Basu  
Metallomics 2010 vol. 2(Issue 2) pp:133-139
Publication Date(Web):24 Dec 2009
DOI:10.1039/B915479E
Clostridial species predominate in both chicken gastrointestinal tract as well as litter where the organoarsenical roxarsone (3-nitro 4-hydroxybenzenearsonic acid) is anaerobically transformed releasing the more recognized toxic inorganic arsenic. 2D-gel electrophoresis and mass spectrometry were used to evaluate the changes in protein expression of Alkaliphilus oremlandii in response to different growth conditions (e.g., terminal electron acceptors) in order to explore the mechanism of microbial biotransformation of roxarsone. Aldehyde ferredoxin oxidoreductase, the enzyme that belongs to the xanthine oxidase family of molybdoenzymes was significantly overexpressed in the presence of roxarsone suggesting a role in the anaerobicmetabolism of this substituted nitrophenol.
Co-reporter:Partha Basu, Archana Nigam, Benjamin Mogesa, Suzanne Denti, Victor N. Nemykin
Inorganica Chimica Acta 2010 Volume 363(Issue 12) pp:2857-2864
Publication Date(Web):15 October 2010
DOI:10.1016/j.ica.2010.04.004
A new dithiolene ligand with 3,5-dibromo substituted phenyl groups was designed and synthesized. The protected form of the ligand was reacted with a nickel salt providing neutral Ni(S2C2(3,5-C6H3Br2)2)2 and anionic [Ni(S2C2(3,5-C6H3Br2)2)2]− isolated as a Bu4N+ salt. Both were characterized by UV–Vis and IR spectroscopy and compared with the similar known molecular systems. They exhibit intense low energy transitions that are characteristic of such systems. The electrochemical behavior of these molecules was investigated by cyclic voltammetry.A new Ni-dithiolene system has been synthesized and characterized. Its redox and optical properties have been compared with known complexes.
Co-reporter:Partha Basu, Victor N. Nemykin and Raghvendra S. Sengar
Inorganic Chemistry 2009 Volume 48(Issue 13) pp:6303-6313
Publication Date(Web):June 1, 2009
DOI:10.1021/ic900579s
Dioxo molybdenum complexes of general formula Tp*MoO2(S-p-RC6H4) (1), where Tp* = hydrotris(3,5-dimethyl-1-pyrazolyl)borate and R = OMe, Me, SMe, NHCOMe, H, Cl, CF3, NO2, were reacted with trimethyl phosphine (PMe3) to convert into complexes of general formula Tp*MoO(S- p-RC6H4)(OPMe3) (2) (where R = OMe, Me, SMe, H, Cl, and CF3). These complexes were isolated and characterized by NMR, IR, UV/vis, and single crystal X-ray crystallography. Electronic and NMR spectra, as well as redox potentials vary as a function of substituent on the thiophenolato ligand. When viewed entirety of the oxygen atom transfer (OAT) reactivity, the reaction of Tp*MoO2(S-p-RC6H4) with PMe3 shows a biphasic behavior, indicating the formation of at least one intermediate. The kinetics of the both steps, that is, the formation of the phosphoryl intermediate and the formation of the solvent coordinated species have been investigated by UV−vis spectroscopy. The first step follows a second order process, first order with respect to both the complex and PMe3, and the overall second order rate constants at 25 °C range from 98.2 (±0.01) × 10−2 M−1 s−1 (for R = OMe) to 223.0 (±0.20) × 10−2 M−1 s−1 (for R = CF3); activation parameters were in the ranges ΔH‡ = 49.3(±4.1) kJ·mol−1 (for R = OMe) to 34.0 (±7.5) kJ·mol−1 (for R = CF3), ΔS‡ = −154.0 (±14.2) J·mol−1·K−1 (for R = OMe) to −184.3 (±26.1) J·mol−1 K−1 (for R = CF3), and ΔG‡ = 95.0 kJ·mol−1 (for R = OMe) to 88.7 kJ·mol−1 (for R = CF3). Formation of the acetonitrile complex from the phosphoryl complex follows a first order process with respect to the complex. The first order rate constants at 25 °C range from 3.60 (±0.01) × 10−4 sec−1 (for R = OMe) to 6.32 (±0.11) × 10−4 sec−1 (for R = CF3), and the enthalpy of activation and entropy of activation show variation; ΔH‡ = 62.5 (±2.2) to 67.8 (±1.0) kJ·mol−1, ΔS‡ = −82.5 (±3.3) to −101.3 (±7.5) J·mol−1·K−1, but the free energy of activation remains constant ΔG‡ ∼ 92 (±1) kJ·mol−1. Large entropies of activation associated with both steps are consistent with associative transition states. The comparable magnitude of the activation energy of the two steps underscores the difficulty in identifying the rate-limiting step in the overall OAT reaction. The first step, however, is more sensitive toward the substituent effects than the second step. Therefore, a change in the substituent can play an important role in deciding the rate-limiting step involved in a two-step OAT reaction.
Co-reporter:Eranda Perera and Partha Basu  
Dalton Transactions 2009 (Issue 25) pp:5023-5028
Publication Date(Web):27 May 2009
DOI:10.1039/B904113C
Using an oxidized state of a dithiolene ligand, diisopropylpiperazine-2,3-dithione (i-Pr2Pipdt), two monooxo-molybdenum complexes have been synthesized. From one of them, a desoxomolybdenum cluster, [(i-Pr2Pipdt)Mo]4[BF4]4 has been prepared. The molecular structure of this cluster reveals metal–metal interactions and weak coordination by the BF4 anion.
Co-reporter:Ryan G. Hadt, Victor N. Nemykin, Joseph G. Olsen and Partha Basu  
Physical Chemistry Chemical Physics 2009 vol. 11(Issue 44) pp:10377-10384
Publication Date(Web):30 Sep 2009
DOI:10.1039/B905554A
The EPR spectral parameters, i.e.g-tensors and molybdenum hyperfine couplings, for several d1 systems of the general formula [MoVEX4]n−, [MoVOX5]2−, and [MoVOX4(H2O)]− (E = O, N; X = F, Cl, Br; n = 1 or 2) were calculated using Density Functional Theory. The influence of basis sets, their contraction scheme, the type of exchange-correlation functional, the amount of Hartree-Fock exchange, molecular geometry, and relativistic effects on the calculated EPR spectra parameters have been discussed. The g-tensors and molybdenum hyperfine coupling parameters were calculated using a relativistic Hamiltonian coupled with several LDA, GGA, and ‘hybrid’ exchange-correlation functionals and uncontracted full-electron DGauss DZVP basis sets. The calculated EPR parameters are found to be sensitive to the MoE distance and EMo–Cl angle, and thus the choice of starting molecular geometry should be considered as an important factor in predicting the g-tensors and hyperfine coupling constants in oxo-molybdenum compounds. In the present case, the GGA exchange-correlation functionals provide a better agreement between the theory and the experiment.
Co-reporter:Lauren Marbella;Barbara Serli-Mitasev Dr. Dr.
Angewandte Chemie International Edition 2009 Volume 48( Issue 22) pp:3996-3998
Publication Date(Web):
DOI:10.1002/anie.200806297
Co-reporter:Lauren Marbella;Barbara Serli-Mitasev Dr. Dr.
Angewandte Chemie 2009 Volume 121( Issue 22) pp:4056-4058
Publication Date(Web):
DOI:10.1002/ange.200806297
Co-reporter:Raghvendra S. Sengar, Jonathan J. Miller and Partha Basu  
Dalton Transactions 2008 (Issue 19) pp:2569-2577
Publication Date(Web):27 Mar 2008
DOI:10.1039/B714386A
Presence of the hydrogen bonding near a metal center can influence the properties of the complex. Here, we describe changes in redox and spectral properties in discrete dioxo-molybdenum centers coordinated by a single thiolato ligand that can support an intra-ligand hydrogen bond. We have utilized thiophenolato ligands that can harbor hydrogen bonding between the thiophenolato sulfur with an amide functionality creating either a five- or a six-membered ring. Methylation of the amide functionality removes the NH⋯S hydrogen bonding thus providing a basis for understanding the effect of hydrogen bonding. These thiophenolato ligands have been used in synthesizing dioxo-MoVI complexes of type Tp*MoO2(S-o-RC6H4), where R = CONHMe (11), CONMe2 (12), NHCOMe (13), and N(Me)COMe (14). The complexes have been characterized by NMR, infrared, and UV-visible spectroscopy. Spectroscopic data clearly indicate the presence of hydrogen bonding in both 11 and 13, and stronger in 13, where hydrogen bonding stabilizes a five-membered ring. All complexes exhibit a MoVI/MoV redox couple and redox potentials are modulated by the nature of H-bonding. Compound 14 with the electron-releasing N(Me)COMe group has the highest reduction potential and is more difficult to reduce.
Co-reporter:Raghvendra S. Sengar, Partha Basu
Inorganica Chimica Acta 2007 Volume 360(Issue 6) pp:2092-2099
Publication Date(Web):20 April 2007
DOI:10.1016/j.ica.2006.10.041
Dioxo-MoVI complexes of general formula Tp∗MoO2(p-SC6H4Dn) (6a–6c) (where Tp∗ = hydrotris(3,5-dimethyl-pyrazol-1-yl)borate and Dn = dendritic unit) have been synthesized and characterized by spectroscopy and mass spectrometry. 1H NMR spectra of the metal complexes indicate that the Cs local symmetry about the metal core does not change by the incorporation of dendritic functionality at the thiophenolato ring. Electrochemical data show a ∼20 mV change in the redox potential in the complexes with dendritic ligands suggesting a very small perturbation in the redox orbital, which is also supported by small changes in the electronic spectra. The peak-to peak separation (ΔEp) increases from 125 mV in 6(a) to 240 mV in 6(c), suggesting sluggish electron transfer in molecules with larger dendritic ligands.Sterically encumbered diooxomolybdenum complexes have been synthesized and characterized. These complexes exhibit sluggish electrochemistry and size selective reactivity.
Co-reporter:Brian W. Kail and Partha Basu  
Dalton Transactions 2006 (Issue 11) pp:1419-1423
Publication Date(Web):21 Nov 2005
DOI:10.1039/B508543H
We have previously postulated a serine gated electron transfer hypothesis (Inorg. Chem, 2002, 41, 1281–1291) to possibly be involved in gating electron transfer between the Mo(V) and Mo(IV) states. In this study we explored the effect of solvent dielectric upon the rate and mechanism of isomerization of an oxo-Mo(V) core in attempt to understand the effect of solvent polarity to the isomerization reaction. To this end, the data suggests that there may be significant entropic contributions to the reorganization of metal center as a function of the local dielectric constant. Furthermore, we note that there is a change in the observed rate as well as the mechanism of the geometric rearrangement when it is examined in polar and non-polar environments. More specifically, in low dielectric media, the reaction proceeds either via a fast dissociation which is then followed by a twist mechanism or by a dissociation that is synchronized with the twist mechanism.
Co-reporter:Brian W. Kail;Lisa M. Pérez Dr.;Snežana D. Zarić ;Andrew J. Millar Dr.;Charles G. Young ;Michael B. Hall
Chemistry - A European Journal 2006 Volume 12(Issue 28) pp:
Publication Date(Web):25 JUL 2006
DOI:10.1002/chem.200600269

The oxygen-atom-transfer (OAT) reactivity of [LiPrMoO2(OPh)] (1, LiPr=hydrotris(3-isopropylpyrazol-1-yl)borate) with the tertiary phosphines PEt3 and PPh2Me in acetonitrile was investigated. The first step, [LiPrMoO2(OPh)]+PR3[LiPrMoO(OPh)(OPR3)], follows a second-order rate law with an associative transition state (PEt3, ΔH=48.4 (±1.9) kJ mol−1, ΔS=−149.2 (±6.4) J mol−1 K−1, ΔG=92.9 kJ mol−1; PPh2Me, ΔH=73.4 (±3.7) kJ mol−1, ΔS=−71.9 (±2.3) J mol−1 K−1, ΔG=94.8 kJ mol−1). With PMe3 as a model substrate, the geometry and the free energy of the transition state (TS) for the formation of the phosphine oxide-coordinated intermediate were calculated. The latter, 95 kJ mol−1, is in good agreement with the experimental values. An unexpectedly large O-P-C angle calculated for the TS suggests that there is significant O-nucleophilic attack on the PC σ* in addition to the expected nucleophilic attack of the P on the MoO π*. The second step of the reaction, that is, the exchange of the coordinated phosphine oxide with acetonitrile, [LiPrMoO(OPh)(OPR3)] + MeCN [LiPrMoO(OPh)(MeCN)] + OPR3, follows a first-order rate law in MeCN. A dissociative interchange (Id) mechanism, with activation parameters of ΔH=93.5 (±0.9) kJ mol−1, ΔS=18.2 (±3.3) J mol−1 K−1, ΔG=88.1 kJ mol−1 and ΔH=97.9 (±3.4) kJ mol−1, ΔS=47.3 (±11.8) J mol−1 K−1, ΔG=83.8 kJ mol−1, for [LiPrMoO(OPh)(OPEt3)] (2 a) and [LiPrMoO(OPh)(OPPh2Me)] (2 b), respectively, is consistent with the experimental data. Although gas-phase calculations indicate that the MoOPMe3 bond is stronger than the MoNCMe bond, solvation provides the driving force for the release of the phosphine oxide and formation of [LiPrMoO(OPh)(MeCN)] (3).

Co-reporter:Victor N. Nemykin, Partha Basu
Inorganica Chimica Acta 2005 Volume 358(Issue 10) pp:2876-2882
Publication Date(Web):15 June 2005
DOI:10.1016/j.ica.2004.07.025
Co-reporter:Andrew J. Millar Dr.;Christian J. Doonan Dr.;Paul D. Smith Dr.;Victor N. Nemykin Dr. ;Charles G. Young
Chemistry - A European Journal 2005 Volume 11(Issue 11) pp:
Publication Date(Web):22 MAR 2005
DOI:10.1002/chem.200401101

Intermediates in the oxygen atom transfer from MoVI to PIII, [TpiPrMoOX(OPR3)] (TpiPr=hydrotris(3-isopropylpyrazol-1-yl)borate; X=Cl, phenolates, thiolates), have been isolated from the reactions of [TpiPrMoO2X] with phosphines (PEt3, PMePh2, PPh3). The green, diamagnetic oxomolybdenum(IV) complexes possess local C1 symmetry (by NMR spectroscopy) and exhibit IR bands assigned to ν(MoO) (approximately 950 cm−1) and ν(PO) (1140–1083 cm−1) vibrations. The X-ray crystal structures of [TpiPrMoOX(OPEt3)] (X=OC6H4-2-sBu, SnBu), [TpiPrMoO(OPh)(OPMePh2)], and [TpiPrMoOCl(OPPh3)] have been determined. The monomeric complexes exhibit distorted octahedral geometries, with coordination spheres composed of tridentate fac-TpiPr and mutually cis monodentate terminal oxo, phosphoryl (phosphine oxide), and monoanionic X ligands. The electronic structures and stabilities of the complexes have been probed by computational methods, with the three-dimensional energy surfaces confirming the existence of a low-energy steric pocket that restricts the conformational freedom of the phosphoryl ligand and inhibits complete oxygen atom transfer. The reactivity of the complexes is also briefly described.

Co-reporter:Victor N. Nemykin and Partha Basu  
Dalton Transactions 2004 (Issue 13) pp:1928-1933
Publication Date(Web):01 Jun 2004
DOI:10.1039/B403964E
A stable molybdenum(V) complex, LMoOCl2 (where L is hydrotris(3,5-dimethylpyrazolyl)borate), has been oxidized under mass spectrometric conditions. The oxidized species reacts with tertiary phosphines and the products have been detected by mass spectrometry. The product distribution has been followed by isotope labeling experiments, and energy dependent electrospray mass spectrometry. These experiments reveal not only oxygen atom transfer but also loss of a chlorine atom from the resulting species.
Co-reporter:Raghvendra S. Sengar, Victor N. Nemykin and Partha Basu  
New Journal of Chemistry 2003 vol. 27(Issue 7) pp:1115-1123
Publication Date(Web):22 May 2003
DOI:10.1039/B300048F
A large number of para-substituted benzene thiols and the corresponding disulfides were synthesized and characterized by 1H NMR, 13C NMR, and IR spectroscopies. Geometries of all sixteen thiols and fourteen disulfide compounds were optimized at the B3LYP/6-31G(d) level, while the electronic structure and the 13C isotropic shifts were calculated by ab initio Hartree-Fock method coupled with the Gauge-Independent Atomic Orbital (GIAO) algorithm and a 6-31+G(d,p) basis set. The calculated 13C NMR isotropic shifts exhibit admirable agreement (δ rmsd ∼4.6 ppm) with the experimental data. The chemical shift of para-substituted carbon showed a linear correlation with Hammett constants (σp and σp+). Using this methodology the σp+ constants for the dendritic ligands have been estimated to be 0.25 and 0.24 for 2(n) and 2(o), respectively. In addition, the NBO charges on the sulfur atoms shows a latent response with the σp+ parameter. The atomic charge on the thiophenolato sulfur is invariant with the electron withdrawing ability of the substituents, however, the charge increases with increasing electron-withdrawing power.
Co-reporter:John F. Stolz Ass.
ChemBioChem 2002 Volume 3(Issue 2-3) pp:
Publication Date(Web):7 MAR 2002
DOI:10.1002/1439-7633(20020301)3:2/3<198::AID-CBIC198>3.0.CO;2-C

The biological transformation of nitrogen oxyanions is widespread in nature and gives rise to a robust biogeochemical cycle. The first step in nitrate reduction is carried out by the enzyme nitrate reductase (NR). Although NR always catalyzes the same chemical reaction (conversion of nitrate into nitrite), its location in the cell, structure, and function are organism-dependent. We use protein sequence data to determine phylogenetic relationships and to examine similarities in structure and function. Three distinct clades of NR are apparent: the eukaryotic assimilatory NR (Euk-NR) clade, the membrane-associated prokaryotic NR (Nar) clade, and a clade that includes both the periplasmic NR (Nap) and prokaryotic assimilatory NR (Nas). The high degree of sequence similarity and a phylogenetic distribution that follows taxonomic classification suggest a monophyletic origin for the Euk-NR early on in the evolution of eukaryotic cells. In contrast, sequence conservation, phylogenetic analysis, and physiology suggest that both Nar and Nap were acquired by horizontal gene transfer. Nap and Nas share a lesser degree of similarity, with Nap a subclade of Nas. Nap from strict anaerobic bacteria such as Desulfovibrio desulfuricans is ancestral to facultative species and may provide an evolutionary link between Nap and Nas. We observed conserved binding sites for molybdenum and pterin cofactors in all four proteins. In pathways involving Euk-NR, Nas, and Nar, for which ammonia is the end product, nitrite is reduced to ammonia by a siroheme nitrite reductase. Nap, however, is coupled to a pentaheme nitrite reductase. In denitrification, whether Nar or Nap is involved, nitrite is reduced to nitric oxide by either a cytochrome cd1or a copper-containing nitrite reductase. This complexity underscores the importance of nitrate reduction as a key biological process.

Co-reporter:Ryan G. Hadt, Victor N. Nemykin, Joseph G. Olsen and Partha Basu
Physical Chemistry Chemical Physics 2009 - vol. 11(Issue 44) pp:NaN10384-10384
Publication Date(Web):2009/09/30
DOI:10.1039/B905554A
The EPR spectral parameters, i.e.g-tensors and molybdenum hyperfine couplings, for several d1 systems of the general formula [MoVEX4]n−, [MoVOX5]2−, and [MoVOX4(H2O)]− (E = O, N; X = F, Cl, Br; n = 1 or 2) were calculated using Density Functional Theory. The influence of basis sets, their contraction scheme, the type of exchange-correlation functional, the amount of Hartree-Fock exchange, molecular geometry, and relativistic effects on the calculated EPR spectra parameters have been discussed. The g-tensors and molybdenum hyperfine coupling parameters were calculated using a relativistic Hamiltonian coupled with several LDA, GGA, and ‘hybrid’ exchange-correlation functionals and uncontracted full-electron DGauss DZVP basis sets. The calculated EPR parameters are found to be sensitive to the MoE distance and EMo–Cl angle, and thus the choice of starting molecular geometry should be considered as an important factor in predicting the g-tensors and hyperfine coupling constants in oxo-molybdenum compounds. In the present case, the GGA exchange-correlation functionals provide a better agreement between the theory and the experiment.
Co-reporter:Eranda Perera and Partha Basu
Dalton Transactions 2009(Issue 25) pp:NaN5028-5028
Publication Date(Web):2009/05/27
DOI:10.1039/B904113C
Using an oxidized state of a dithiolene ligand, diisopropylpiperazine-2,3-dithione (i-Pr2Pipdt), two monooxo-molybdenum complexes have been synthesized. From one of them, a desoxomolybdenum cluster, [(i-Pr2Pipdt)Mo]4[BF4]4 has been prepared. The molecular structure of this cluster reveals metal–metal interactions and weak coordination by the BF4 anion.
Co-reporter:Partha Basu, Brian W. Kail, Andrew K. Adams and Victor N. Nemykin
Dalton Transactions 2013 - vol. 42(Issue 9) pp:NaN3081-3081
Publication Date(Web):2012/11/08
DOI:10.1039/C2DT32349D
The oxygen atom transfer reactivity (OAT) of dioxo-Mo(VI) complexes of hydrotrispyrazolyl borate (hydrotris(3,5-dimethylpyrazolyl)borate, TpMe2; hydrotris(3-isopropylpyrazol-1-yl)borate, TpiPr) with tertiary phosphines (PMe3, PMe2Ph, PEt3, PEt2Ph, PBun3, PMePh2, or PEtPh2) has been investigated. In acetonitrile, these reactions proceed via the formation of a phosphoryl intermediate complex that undergoes a solvolysis reaction. We report the synthesis and characterization of several phosphoryl complexes. The rates of formation of phosphoryl complexes and their solvation were determined by spectrophotometry. The rates of the reactions and the properties of the phosphoryl species were investigated using the Quantitative Analysis of Ligand Effect (QALE) methodology. The results show that, at least in this system, the first step of the reaction is controlled primarily by the steric factor, and in the second step, both electronic and steric factors are important. We also analyzed the effect of ligands on the reaction rate i.e., TpMe2vs. TpiPr.
Co-reporter:Courtney Sparacino-Watkins, John F. Stolz and Partha Basu
Chemical Society Reviews 2014 - vol. 43(Issue 2) pp:NaN706-706
Publication Date(Web):2013/10/18
DOI:10.1039/C3CS60249D
The nitrate anion is a simple, abundant and relatively stable species, yet plays a significant role in global cycling of nitrogen, global climate change, and human health. Although it has been known for quite some time that nitrate is an important species environmentally, recent studies have identified potential medical applications. In this respect the nitrate anion remains an enigmatic species that promises to offer exciting science in years to come. Many bacteria readily reduce nitrate to nitrite via nitrate reductases. Classified into three distinct types – periplasmic nitrate reductase (Nap), respiratory nitrate reductase (Nar) and assimilatory nitrate reductase (Nas), they are defined by their cellular location, operon organization and active site structure. Of these, Nap proteins are the focus of this review. Despite similarities in the catalytic and spectroscopic properties Nap from different Proteobacteria are phylogenetically distinct. This review has two major sections: in the first section, nitrate in the nitrogen cycle and human health, taxonomy of nitrate reductases, assimilatory and dissimilatory nitrate reduction, cellular locations of nitrate reductases, structural and redox chemistry are discussed. The second section focuses on the features of periplasmic nitrate reductase where the catalytic subunit of the Nap and its kinetic properties, auxiliary Nap proteins, operon structure and phylogenetic relationships are discussed.
(3Z)-4-(2-amino-4-oxo-1,4,5,6,7,8-hexahydropteridin-6-yl)-2-hydroxy-3,4-disulfanylbut-3-en-1-yl dihydrogen phosphate
Molybdoenzymemolybdenum cofactor
4(1H)-PYRIMIDINONE, 5,6-DIAMINO-2-(4-MORPHOLINYL)-
2H-Pyran, 2-[(1-ethyl-1-methyl-2-propynyl)oxy]tetrahydro-
Ferrate(2-), [7,12-diethenyl-3,8,13,17-tetramethyl-21H,23H-porphine-2,18-dipropanoato(4-)-κN21,κN22,κN23,κN24]-, hydrogen (1:2), (SP-4-2)-
(1S,4S)-2-(5-PYRIMIDINYL)-2,5-DIAZABICYCLO[2.2.1]HEPTANE