48 results on '"Rosenberg, Edward"'
Search Results
2. Dehydrogenative coupling of alcohols to esters on a silica polyamine composite by immobilized PNN and PONOP pincer complexes of ruthenium.
- Author
-
Goni, Md Abdul, Rosenberg, Edward, Gobetto, Roberto, and Chierotti, Michele
- Subjects
- *
CATALYSTS , *RUTHENIUM , *MANNICH reaction , *PYRIDINE , *ESTERS , *AMINES - Abstract
Heterogenization of catalytically important homogeneous catalysts on solid supports has become an expanding area of research. PNN and PONOP ruthenium pincer complexes were immobilized on a silica poly(allylamine) composite, BP-1 by a two-step Mannich reaction. The complexes on BP-1 were characterized by solid state NMR, FT-IR, elemental analysis, and metal digestion studies. Model solution experiments were carried out to determine the site of electrophilic substitution on the pyridine ring of the pincer complexes and revealed substitution in both the meta - and para -position. The catalytic reactivity of immobilized (PNN)RuH(Cl)(CO) and (PONOP)RuH(Cl)(CO) on BP-1 was studied for the dehydrogenative coupling of alcohols to esters with the liberation of H 2 . Moderate to good ester yields were realized with both immobilized systems without using the base required for the homogeneous reaction and also in the presence of KOH. The homogeneous model reactions required a base for ester formation. The amine functionality on BP-1 served as the base to generate the active pincer catalyst on the BP-1 surface. Both immobilized catalysts were recycled for multiple alcohol reaction cycles. Four-step control experiments were carried out using an alcohol and both immobilized systems. The results revealed the heterogeneity of the alcohol catalysis by both BP-1-Ru-PNN and BP-1-Ru-PONOP systems. This study has opened a new catalytic methodology for reactions where base is required for catalyst activation, by using a solid support with basic functionality. [ABSTRACT FROM AUTHOR]
- Published
- 2017
- Full Text
- View/download PDF
3. Luminescence Quenching and Enhancement by Adsorbed Metal Ions with Ruthenium Diimine Complexes Immobilized on Silica Polyamine Composites.
- Author
-
Rosenberg, Edward, Abbott, Geoffrey, Ross, J. B. Alexander, McVay, Riley, and Terwilliger, Michelle
- Subjects
- *
LUMINESCENCE quenching , *METAL ions , *RUTHENIUM , *IMINES , *POLYAMINES - Abstract
Luminescent ruthenium diimine complexes have been covalently bound to the surface of a silica polyamine composite (SPC) using peptide coupling agents. The loading of the complexes using this route is quite low (∼0.01-0.04 mmol/g) leaving sufficient surface amines to coordinate added metal ions. When the composite particles containing the Ru complexes are exposed to solutions of Cu2+, Ni2+ or Zn2+, luminescence is quenched with efficiencies that follow concentration dependence and the relative binding affinities of the ions. When heavy metal ions such as mercury or lead are adsorbed onto the same surface, luminescence is enhanced by a factor of ∼3.5. When the complexes are exposed to these metals in solution, no quenching or enhancement is observed. Both phenomena were shown to be the result of adsorption of the cations onto the polyamine surface by using the Stern-Volmer relationship. The mechanism of both quenching and enhancement is discussed and the options for further development of this novel metal sensing technique are presented. [ABSTRACT FROM AUTHOR]
- Published
- 2016
- Full Text
- View/download PDF
4. A methods study of immobilization of PONOP pincer transition metal complexes on silica polyamine composites (SPC).
- Author
-
Goni, Md Abdul, Rosenberg, Edward, Meregude, Shesharao, and Abbott, Geoffrey
- Subjects
- *
TRANSITION metal complexes , *CRYSTALLOGRAPHY , *COMPOSITE materials , *SILICA , *POLYAMINES , *SURFACE chemistry , *MICROENCAPSULATION ,CATALYSTS recycling - Abstract
Immobilization of catalytically active transition metal complexes on silica polyamine composite (SPC) surfaces offers many advantages for applications in catalysis particularly for catalyst recovery and reuse. We report here the immobilization of PONOP pincer complexes of Ru, Rh, Ni and Pd on the poly(allylamine) SPC, BP-1 using the Mannich reaction. Three different methods have been investigated for synthesizing the PONOP pincer transition metal complexes on BP-1: 1) direct reaction of the preformed pincer complexes using a two step Mannich reaction; 2) immobilization of the PONOP ligand using the Mannich reaction followed by the addition of a transition metal compound of a given metal; 3) the stepwise construction of PONOP on BP-1 followed by addition of a transition metal compound. The immobilized complexes on BP-1 were characterized by FT-IR, solid-state CPMAS 13 C and 31 P NMR, as well as elemental analysis. Anchoring of the complexes on BP-1 was also evaluated by the metal loading data obtained from the digestion of the loaded composites followed by Atomic Absorption Spectroscopy (AAS) or Inductively Coupled Plasma Atomic Emission Spectroscopy (ICPAES). The results showed that method 1 works better for the loading of pincer complexes on the SPC than methods 2 and 3. In the case of the Ru and Ni pincer complexes reasonable agreement with the phosphorous analysis was realized, while for the Pd complex values were high relative to the loading predicted from the phosphorus analysis, indicating the formation of the Pd nanoparticles on the surface during immobilization. For the Rh and Ru immobilized complexes with methods 2 & 3, metal loading was lower than the phosphorous analysis and this is attributed to entrained triphenylphosphine from the starting rhodium and ruthenium complexes based on the 13 C and 31 P CPMAS NMR data. Solution experiments using the PONOP pincer ligand and the Ru(PONOP) complex with n-butyl amine were conducted to model the site of electrophilic aromatic substitution on the pyridine ring. It was found that substitution both meta - and para -to the nitrogen takes place and this helped in the interpretation of the solid-state data. [ABSTRACT FROM AUTHOR]
- Published
- 2016
- Full Text
- View/download PDF
5. Hydrogenation of nitro-compounds over rhodium catalysts supported on poly[acrylic acid]/Al2O3 composites.
- Author
-
Campos, Cristian H., Rosenberg, Edward, Fierro, José L.G., Urbano, Bruno F., Rivas, Bernabé L., Torres, Cecilia C., and Reyes, Patricio
- Subjects
- *
HYDROGENATION , *NITRO compounds , *RHODIUM catalysts , *POLYACRYLIC acid , *ALUMINUM oxide , *FREE radicals , *SOL-gel processes - Abstract
In this report, poly[acrylic acid] gels containing Al 2 O 3 were prepared by simultaneous free-radical polymerization and sol–gel chemistry using different amounts of 3-(trimethoxysilyl)propyl methacrylate (TMPM) as a compatibilizer. The hybrid materials were used as supports for a rhodium catalyst in the chemoselective hydrogenation of 3-substituted aromatic nitro-compounds. The supported rhodium catalyst was prepared by an ion-exchange process. In situ H 2 flux was used to produce active species of the catalysts. The resulting materials were characterized by infrared spectroscopy, thermogravimetric analysis, solid-state 29 Si and 13 C NMR, X-ray diffraction, transmission/scanning electron microscopy, and X-ray photoelectron spectroscopy. All materials exhibited simultaneous interpenetrating hybrid network structures (SIHNs). The morphologies and physicochemical properties depended on the amount of TMPM used. The catalysts were found to be effective for the reduction of nitrobenzene in ethanol at room temperature and a hydrogen pressure of 20 atm. The most active and selective catalyst was used in the hydrogenation of different 3-substituted aromatic nitro-compounds. The hydrogenation reactions displayed high conversion levels and promoted exclusive –NO 2 group reduction, resulting in the sole formation of the corresponding amino-compound, with the exception of 1,3-dinitrobenzene, in which over-hydrogenation was detected. The presence of electron-donating/electron-withdrawing substituents at the 3-position resulted in different rates of –NO 2 group hydrogenation. This effect was quantified in terms of the Hammett relationship, in which the catalyst displayed a linear correlation between the substituent constant ( σ i ) and the hydrogenation rate, with the exception of –OH, –NH 2 , and –OCH 3 groups. One explanation for this behavior is a proposed support-substrate hydrogen bond interaction during the catalytic reaction. [ABSTRACT FROM AUTHOR]
- Published
- 2015
- Full Text
- View/download PDF
6. Hydride Dynamics in Triosmium and Triruthenium Carbonyl Clusters: Synthesis, Reactivity and Dynamics of a Trihydrido Quinoline-4-Carboxaldehyde Triosmium Cluster.
- Author
-
Rosenberg, Edward and Kumar, Rakesh
- Subjects
- *
CARBONYL compounds , *LIGANDS (Chemistry) , *ELECTRONS , *TRIPHENYLPHOSPHINE , *TEMPERATURE , *METALLURGIC chemistry - Abstract
An overview of the dynamical processes involving the hydrido ligand in triosmium and triruthenium carbonyl clusters is presented. The relationship between the mechanisms of hydride motions and the other ligands in the cluster are discussed for mono- di- and trihydrido-clusters. In addition, the reactivity of the electron deficient 46e cluster, (μ-H)(μ-η-CHN-4-CHO)Os(CO) ( 1) with hydrogen is reported. The reaction gives two isomeric trihydrido clusters, H(μ-H)(μ-η-CHN-4-CHO)Os(CO) ( 2) and (μ-H)(μ-η-CHN-4-CHO)Os(CO) ( 2′) in low yield along with trace amounts of other hydrido clusters. Reaction of the inseparable mixture of 2 and 2′ with triphenylphosphine at ambient temperatures gives two related addition products H(μ-H)(μ-η-CHN-4-CHO)Os(CO)PPh ( 3) and (μ-H)(μ-η-CHN-4-CHO)Os(CO)PPh ( 3′) in a 5:1 ratio. These results contrast with the previously reported trihydrido-derivatives of triosmium μ-η-imidoyl clusters where only analogues of 2 and 3 are obtained. Clusters 2 and 2′ are rigid on the NMR time scale while 3 exhibits dynamical behavior in the temperature range of −50 to +25 °C. Cluster 3′ is stereochemically rigid in this temperature range. The dynamical behavior of 3 involves the exchange of the terminal and bridging hydrides coupled with tripodal motion of the phosphine substituted osmium atom, a process virtually identical to previously reported trihydrides of the μ-η-imidoyl triosmium clusters. [ABSTRACT FROM AUTHOR]
- Published
- 2014
- Full Text
- View/download PDF
7. Relative rates of reaction of Pt(en)Cl (NH2R)+ with guanosine monophosphate as a function of amino group substituent: Toward efficient labeling of DNA for TEM imaging.
- Author
-
Kumar, Rakesh, Rosenberg, Edward, Feske, Miriam Inbar, and DiPasquale, Antonio G.
- Subjects
- *
CHLOROPLATINIC acid , *CHEMICAL reactions , *GUANYLIC acid , *AMINO group , *SUBSTITUENTS (Chemistry) , *DNA , *TRANSMISSION electron microscopy , *PLATINUM isotopes - Abstract
In an attempt to understand the factors that govern the rates of reaction of the complexes [Pt(en)Cl(NH2R)]+NO3− (en = ethylene diamine) with guanosine monophosphate (dGMP) a series of amine complexes, where R=C8H9NO2 (benzo[d][1,3]dioxol-5-ylmethanamine) (1), C8H11N (phenethylamine) (2), C7H9N (benzylamine) (3), C6H7N (aniline) (4), C6H6IN (p-iodo-aniline) (5) C3H9NO (2-methoxy-ethylamine) (6) and C6H13N (cyclohexylamine) (7), were synthesized and their reactions with deoxyguanosine monophosphate (dGMP) were followed by 1H NMR. Compound 1 was initially chosen because it showed significant water solubility. Compound 1 reacted quantitatively but slowly with dGMP and a subsequent Transmission Electron Microscopy (TEM) study of the binding 1 to a GATC DNA repeat gave a TEM micrograph that showed selective labeling of DNA at guanine, using a technique that allowed the laying down of a straight single strand of DNA on a carbon platform. The TEM suggested a possible side reaction with adenine and so a study of the reaction of 1 with adenine was performed and showed slow and what appeared to be non-specific binding to deoxyadenosine monophosphate (dAMP). The reactions of compounds 2–7 with dGMP were then studied by 1H NMR and it was found that 2 reacted much faster than 1 with dGMP while the remaining complexes reacted more slowly. No reaction of 2 with dAMP was observed in the same time frame. The ultimate goal of the project was to bind a third row transition metal cluster to guanine and given the effective binding of 1 to DNA the synthesis of the complex [Os3(CO)11PPh2(CH2)2NH2(en)PtCl]NO3 (9) is also reported that contains Pt as a linker to label guanine. The synthesis was performed by reacting Os3(CO)10(CH3CN)2 with Ph2PCH2CH2NH2 which gave an η2 chelate complex Os3(CO)10PPh2(CH2)2NH2 (8). Complex 8 was reacted with [Pt(en)Cl(DMF)]NO3 in a CO atmosphere to give 9. 1H and 195Pt NMR indicate formation of an adduct with dGMP but too slowly to be of use in labeling DNA. The solid-state structure of 8 is also reported. [Copyright &y& Elsevier]
- Published
- 2013
- Full Text
- View/download PDF
8. Oxyanion removal and recovery using silica polyamine composites.
- Author
-
Kailasam, Varadharajan and Rosenberg, Edward
- Subjects
- *
OXYANIONS , *POLYAMINES , *SILICA , *COMPOSITE materials , *AMORPHOUS substances , *PHOSPHORUS acids , *ALKYLATION - Abstract
Modified silica polyamine composites (SPC) made from silanized amorphous nano-porous silica gel and polyamines have been used for the selective removal and recovery of oxyanions of heavy metals such as As and Se in the presence of excess sulfate. The SPCs were functionalized with phosphorus acid using the Mannich reaction, resulting in a phosphonic acid modified composite (BPAP) on which Zr(IV) was immobilized by complexation to the phosphonic acid group. Tungstate and molybdate anions strongly adsorbed on the ZrBPAP composite over a broad pH range of 2–10. The sorption mechanism is believed to be a complexation between the tungstate and/or molybdate and Zr(IV) immobilized-phosphonic acid complex of BPAP. However, complete regeneration of the composite was not successful. A second composite modeled after polymer based adsorbents containing tetraalkyl ammonium groups was synthesized by exhaustive alkylation of the poly(allylamine) composite, BP-1. This composite, BPQAM, composite shows a higher selectivity for tungstate over molybdate at a pH of 6 (WO42 −/MoO42 − = 3/1). Using 2MH2SO4 as the strip solution provided selectivity for molybdate (~ 90% purity). The significance of the observed selectivity is discussed in terms of the chemical properties of the oxyanion and the nature of the interactions with the immobilized metal site. Attempts to separate molybdate from nickel and vanadium using solution concentrations modeled after a contaminated creek in Western Australia are also reported. The use of ZrBPAP and the aminoacetate modified SPC, WP2, resulted in the efficient separation of molybdate but the separation of the nickel from vanadium proved more difficult. [Copyright &y& Elsevier]
- Published
- 2012
- Full Text
- View/download PDF
9. Nucleophilic Attack on an Electronically Unsaturated Triosmium Cluster Complex of 2-Methylbenzoxazole: An Unusual Oxidative Ring Opening.
- Author
-
Rosenberg, Edward, Abedin, Joynal, and Hardcastle, Kenneth
- Subjects
- *
NUCLEOPHILIC reactions , *OXIDATIVE addition , *METAL clusters , *PROTON transfer reactions , *DEUTERIUM - Abstract
In the course of our studies of nucleophilic attack on electronically unsaturated benzoheterocycle triosmium clusters we have studied the reaction of the 2-methylbenzoxazole complex (μ-H)Os(CO)(μ-η-2-CH-CHNO) ( 1) with hydride followed by protonation with acid. In sharp contrast to our previous studies with related benzoheterocycle triosmium clusters, where the nature of the heterocycle controls the regiochemistry of nucleophilic attack, we observe here an unusual ring opening of the heterocyclic ring, coupled with rearrangement of the carbocyclic ring to a 2-imino-ethyl-phenol complex (μ-H)Os(CO)(μ-η-N=CHCH-CH(OH)) ( 2). Deuterium labeling experiments verify initial attack by hydride at the 2-position followed by protonation at oxygen. Reaction of 1 with two equivalents of hydride followed by two equivalents of acid results in reduction of the C=N bond in 2 and on standing in air, oxidation of the carbocyclic ring occurs to give the 2-ethyl-amino hydroquinolyl derivative (μ-H)Os(CO)(μ-η-NHCHCH-CH(2-O)(5-OH)) ( 3). The solid-state structure of 3 is reported and a plausible mechanism, supported by deuterium labeling experiments, is presented, for the formation of 2 and 3. [ABSTRACT FROM AUTHOR]
- Published
- 2012
- Full Text
- View/download PDF
10. Surface oxidation of Co2+ and its dependence on ligand coordination number in silica polyamine composites
- Author
-
Allen, Jesse, Rosenberg, Edward, Chierotti, Michele R., and Gobetto, Roberto
- Abstract
Abstract: Coordination of CoCl2 solutions to the silica polyamine composite, WP-1, made with the branched polymer polyethylenimine (PEI) shows irreversible binding resulting from surface oxidation of the Co2+–Co3+. This is not the case for the silica polyamine composite BP-1 made with the linear polymer polyallylamine where reversible binding and no oxidation is observed. These observations suggested that oxidation of the cobalt was related to the greater coordination number available with the branched polyamine relative to the linear polyamine. A study of the kinetics of cobalt binding to WP-1 indicated initial loading of Co2+ at relatively low coordination number followed by desorption of Co2+ leading to oxidation and irreversible binding. Exclusion of oxygen from the composite-cobalt solution mixtures resulted in irreversible binding at a level that was 14% of the initial experiments. These observations prompted us to undertake a study to elucidate the coordination number around cobalt in the case of the branched polymer PEI. Towards this end, we have synthesized the model complexes [(tren)Co(H2O)2]3+3Cl− (tren=tris(2,2′,2″aminoethyl)amine) and [(dien)Co(H2O)3]3+3Cl− (dien=diethylenetriamine). The UV–Vis spectra of these model complexes were compared with Co3+ coordinated to PEI in solution and it was concluded that the UV–Vis spectrum of the tren complex was closer to that observed for the solution UV–Vis spectrum of Co3+–PEI. These data indicated that coordination of four amines was needed to drive surface oxidation under ambient conditions. In order to further elucidate the coordination number of a metal coordinated to the surfaces of WP-1 and BP-1, we reacted these composites with the probe molecule [Ru(CO)3(TFA)3]−K+ (TFA=trifluoroacetate) (1) where the carbonyl stretching frequencies could be used as a measure of coordination number and geometry of the adsorbed complex. The IR of this complex on WP-1 indicated a monocarbonyl species while the IR of this complex on BP-1 indicated the presence of dicarbonyl species on the surface. These data are consistent with a coordination number of four amines in the case of WP-1 and the coordination of two amines in the case of BP-1 based on our previous studies of the solution coordination chemistry of the 1. Subsequent 13C CPMAS solid-state NMR on 13CO enriched samples of the 1 adsorbed onto BP-1 and WP-1 were consistent with the IR data. [Copyright &y& Elsevier]
- Published
- 2010
- Full Text
- View/download PDF
11. Characterization and Applications of Poly-Acetate Modified Silica Polyamine Composites.
- Author
-
Hughes, Mark A. and Rosenberg, Edward
- Subjects
- *
SILICA gel , *SILICA , *POLYAMINES , *METAL ions , *SEPARATION (Technology) , *ETHYLAMINES , *ETHYLENEDIAMINETETRAACETIC acid , *ADSORPTION (Chemistry) , *SOLUTION (Chemistry) - Abstract
Silica polyamine composites made from silanized amorphous silica gel and a polyamine (polyallylamine (BP-1) and poly(ethyleneimine) (WP-1)) were functionalized with ethylenediamine-N,N,N′,N′-tetraacetic (EDTA) anhydride resulting in the modified composites BP-ED and WP-ED respectively. Successful immobilization of the poly-acetate ligand was confirmed by weight gain, IR and elemental analysis. The modified composites had ligand loading values of 0.78 mmol/g and 0.58 mmol/g respectively. Adsorption characteristics were investigated for BP-ED and WP-ED by pH profiles, time dependent isotherms, concentration dependent isotherms, and the Langmuir and Freundlich adsorption models. Metal ion capacities (mmol of metal adsorbed per gram of adsorbent) were shown to increase with pH for BP-ED and WP-ED, which is in contrast with silica gel modified with EDTA anhydride via amino propyl silane without the use of a polyamine where the opposite trend is observed. The adsorption of Ni(II), Zn(II) and Co(II) onto BP-ED and WP-ED was shown to fit the Langmuir adsorption model, which indicates a monolayer, non-cooperative metal ion uptake. For both composites the divalent metal ion selectivity is as follows: Cu(II)>Ni(II)>> Zn(II), Co(II)>> Mn(II). Also the trivalent metal ion selectivity is as follows: Fe(III)>Ga(III)>> Eu(III)>Al(III). This is in close agreement with reported formation constants for EDTA in solution. Breakthrough and recovery tests were carried out using a 5cc packed column. The separation and recovery of Ni(II) from Co(II) when BP-ED was challenged with a solution containing 1.5 mg/L of both metals resulted in an acidic strip solution of greater than 97% Ni(II) purity. Further, selective separation of Ni(II) from a solution containing low levels of Co(II) and Zn(II) and high level Fe(II) at pH=1 was also achieved. Ni(II) was extracted at a flow rate of 0.1 column volumes per minute (0.5 mL/min) and recovered in greater than 90% purity by an acidic strip. Impurity in the strip was a consequence of Fe(II) oxidation to Fe(III). These tests demonstrate that well resolved separations of one divalent metal ion from one or more other divalent metal ions are possible through the use of EDTA anhydride modified silica polyamine composites with higher capacities than the previously reported amino propyl silane modified silica gel. [ABSTRACT FROM AUTHOR]
- Published
- 2007
- Full Text
- View/download PDF
12. Selective covalent binding of a positively charged water-soluble benzoheterocycle triosmium cluster to single- and double-stranded DNA
- Author
-
Rosenberg, Edward, Spada, Fabrizio, Sugden, Kent, Martin, Brooke, Gobetto, Roberto, Milone, Luciano, and Viale, Alessandra
- Subjects
- *
NUCLEIC acids , *SCHIFF bases , *PIPERIDINE , *MOBILE genetic elements - Abstract
Abstract: The water-soluble triosmium cluster [Os3(CO)9(μ-η2-(4-CHO)C9H5N)(μ-H)(P(OCH2CH2N(CH3)3I)3)] (4) was tested for its reactivity with plasmid DNA. In contrast to the band retardation previously observed with a related series of positively charged clusters, an intensification and retardation of three discrete bands was observed with increasing cluster concentration. In order to further investigate the apparent modification of DNA by 4, its interaction with a 22-oligomer (sequence 5′-AGT TGT GGT GAC TTT CCC AGG C-3′) was examined. Incubation with this oligonucleotide (pH 7.4 in Tris–HCl buffer and 100 mM NaCl) followed by HPLC analysis revealed the formation of three dose dependent products assigned as covalent modifications at three sites of the oligonucleotide. Incubation of 4 with 32P-ATP labeled oligonucleotide at the 5′-end followed by treatment with piperidine and comparison with the standard Maxam–Gilbert sequencing protocol products revealed only general background cleavage, indicating that the modification products are piperidine labile and suggesting that the modification involved formation of a Schiff base. An alternative approach was then pursued which involved annealing the 4-oligonucleotide products with their complementary strand and treatment of the resulting duplex DNAwith the exonuclease, Exo III. This assay indicated three exonuclease stops, consistent with the three products observed by HPLC whose electrophoretic mobility approximately matched guanine containing fragments when compared with the Maxam–Gilbert sequencing lanes. Reduction of the 4-oligonucleotide products with borohydride reducing agents, followed by treatment with piperidine, resulted in the formation of one product (by HPLC) with the same electrophoretic mobility as the AGTT fragment based on comparison with the Maxam–Gilbert sequencing lanes. This product most likely results from reduction of an initially formed Schiff base adduct (to the corresponding amine) with the guanine of the TGT fragment of the oligonucleotide, and corresponds to the most stable of the three Schiff base adducts detected by HPLC and by incubation with the exonuclease. The other two products are less stable and competitive reduction of the free aldehyde functionality on the cluster in equilibrium with these adducts precludes their detection after treatment with the reducing agents. The formation of the Schiff base adduct is further corroborated by the model reaction of [Os3(CO)10(μ-η2-(4-CHO)C9H5N)(μ-H)] (4′) with acetylated guanine in nonaqueous solvents where disappearance of the aldehyde resonance and the appearance of several new resonances in the 6–9 ppm region of the 1H NMR of the reaction mixture is noted. [Copyright &y& Elsevier]
- Published
- 2004
- Full Text
- View/download PDF
13. Ligand dependent structural changes in the acid–base chemistry of electron deficient benzoheterocycle triosmium clusters
- Author
-
Rosenberg, Edward, Joynal Abedin, Md., Rokhsana, Dalia, Viale, Alessandra, Dastru', Walter, Gobetto, Roberto, Milone, Luciano, and Hardcastle, Kenneth
- Subjects
- *
COORDINATION compounds , *OSMIUM - Abstract
The reactions of the electron deficient benzoheterocycle clusters Os3(CO)9(μ3-η2-L&z.sbnd;H)(μ-H) (L=phenanthridine, 2; 5,6-benzoquinoline, 3; quinoxaline, 4; 2-methyl benzimidazole, 5; 2-methyl benzotriazole, 6; 2-methyl benzothiazole, 7; benzothiazole, 8; 2-methyl benzoxazole, 9; benzoxazole, 10) with n-butylamine and with the protic acids HBF4 and CF3CO2H are reported. Complexes 2 and 3 behave very similarly to the analogous quinoline complex (L=quinoline, 1) forming adducts with n-butylamine with similar spectral properties to 1 and undergoing simple protonation at the metal core with both HBF4 and CF3CO2H. Complexes 4–10 on the other hand form an additional amine adduct with different spectral characteristics than 1–3 and in the case of 5 this is the only detectable product. In addition, these compounds also form adducts with acetonitrile while 1–3 do not. Compounds 4–9 also exhibit much more complex behavior in the presence of CF3CO2H than 1–3, giving trifluoroacetate adducts subsequent to initial protonation. Although 4, 5 and 10 behave differently than 1 with n-butylamine and protic acids they form exactly the same triphenyl phosphine adduct as 1 and 3 based on multinuclear NMR data. The solid-state structure for the amine adduct of 2 is reported and suggested structures for the other amine and acid adducts, based on NMR data, are presented. The reasons for the differences in the behavior of the three distinct groups of benzoheterocycles (1–3, 4 and 5–10) are discussed. [Copyright &y& Elsevier]
- Published
- 2002
- Full Text
- View/download PDF
14. Core-shell Fe-SiO2-polyamine magnetic nanoparticles for metal recovery using a continuous flow pipeline reactor.
- Author
-
Deluca, Emile, Latterman, Ryan, and Rosenberg, Edward
- Subjects
- *
MAGNETIC nanoparticles , *METAL ions , *NANOSTRUCTURED materials , *POLYAMINES , *METAL recycling - Abstract
A series of core-shell magnetic nanomaterials have been synthesized with the intent of applying them for metal ion capture in a newly designed pipeline reactor. The synthetic chemistry is an extension of a previously developed family of materials based on amorphous silica gel, which has been used in the mining and remediation industries. The nanoparticles were characterized by infrared spectroscopy and TEM and SEM techniques. The size of the starting magnetite core was critical to the behavior of the particles under metal sequestering conditions. The capture kinetics of the resulting nanoagregates is 10 times faster than related micro composites. All of the tests performed point to the future successful development of a technology that circumvents the disadvantages associated with the use of column based microparticles. [ABSTRACT FROM AUTHOR]
- Published
- 2018
- Full Text
- View/download PDF
15. Synthesis, structure, photophysical and electrochemical properties of Ru(TFA)(CO)(PPh3)2(L) (L=2-phenylpyridine, 2-p-tolylpyridine) and Ru(CO)(PPhMe2)2(L)(L′) (L= TFA, H) (L′= bipyridine, L′= 4,4′-dimethylbipyridine) relationships between ancillary ligand structure and luminescent properties
- Author
-
Pohkrel, Shyam, Decato, Dan, Rosenberg, Edward, Ross, J.B. Alexander, and Terwilliger, Michelle
- Subjects
- *
COMPLEX compounds synthesis , *METAL complexes , *ELECTROCHEMISTRY , *RUTHENIUM compounds , *X-ray diffraction , *PYRIDINE - Abstract
The synthesis, structure and photophysical properties of the complexes [Ru[(CO)(TFA) (PPh 3 ) 2 (L)][(L = ppy = 2-phenylpyridine, ( 1a ); L = 2–(p–tolyl)pyridine] ( 1b ), are reported. The complexes were characterized by UV-VIS, IR and NMR and by single-crystal X-ray diffraction techniques. We also report the synthesis, structure and photophysical properties of [Ru(CO)(L)(PPhMe 2 ) 2 (L′)] + [PF 6 ] − [L′ = bipyridine, L = TFA, ( 3a ); L = H, ( 3b ) and L = H, L′ = 4,4′-dimethlyl bipyridine ( 3c )]. These compounds were characterized by UV-VIS, IR and NMR techniques and by a single crystal X-ray diffraction in the case of 3a. The solid state structure of [Ru(Me 2 PhP) 2 (CO) 2 (TFA) 2 (2) which is the starting material for the synthesis 3a - 3c is also reported to verify the trans relationship of the less bulky PPhMe 2 and for comparison with the previously reported PPh 3 analogs. The purpose of this study was to determine the impact, if any, of replacing bpy with ppy in the case of 1a and alkylation of the benzene ring in the case of 1b on the photophysical and electrochemical properties compared to related Ru(bpy) complexes. In contrast to the bpy analogs 1a and 1b showed reversible 1e − oxidations and blue-shifted MLCT absorptions. In the case of 3a - 3c we were interested in the effect on the photophysical properties of substituting PPh 3 with the less bulky but more electron donating PPhMe 2 . There were only minor changes in the photophysical and electrochemical properties relative to the previously reported PPh 3 analogs. [ABSTRACT FROM AUTHOR]
- Published
- 2017
- Full Text
- View/download PDF
16. Electrochemical studies of Ru(II) diimine bioconjugates.
- Author
-
Ravera, Mauro, Sharmin, Ayesha, and Rosenberg, Edward
- Subjects
- *
RUTHENIUM , *IMINES , *PHENANTHROLINE , *BIPYRIDINE , *SERUM albumin , *PLATINUM group - Abstract
The electrochemical behavior of the ruthenium(II) diimine complexes [Ru(CO)(CF 3 CO 2 )(dppene)(5-R-phen)][PF 6 ] (dppene = 1,2-diphenylphosphinoethene; phen = 1,10-phenanthroline; R = H, 1 ; R = NH 2 , 2 ; R = cholestoryl carbamate, 3 ; R = 1,2-dihexadecanoyl- sn -glycero-3-phosphoethanolamine, 4 ), [Ru(CO)(H)(4,4′-R-bpy)(R′Ph 2 P) 2 ][PF 6 ] (bpy = 2,2′-bipyridine; R = H, R′ = Ph, 5 ; R = H, R′ = 1,2-dihexadecanoyl- sn -glycero-3-phosphoethanolamine, 6 ; R = 1,2-dihexadecanoyl- sn -glycero-3-phosphoethanolamine, R′ = Ph, 7 ), [Ru(bpy) 2 (5-R-phen)][PF 6 ] 2 (R = NH 2 , 8 ; R = cholestoryl carbamate, 9 ; R = 1,2-dihexadecanoyl- sn -glycero-3-phosphoethanolamine, 10 ) is reported. Complexes 1 – 6 give cyclovoltammetric (CV) responses with multiple ill-defined reduction waves and one oxidation wave, all of which were chemically irreversible. Complexes 5 and 7 , containing axially coordinating phosphines, showed reversible oxidation and reduction CV responses, while 6 showed redox waves similar to 3 . Complexes 8 – 10 show a metal-centered irreversible oxidation around +1.4 V that, in the case of 8 and 9 , is heavily modified by adsorption phenomena. In the negative part of the CV, 9 and 10 show a single chemically and electrochemically reversible 1e − reduction both at E °′ = −1.29 V, about 500 mV cathodically shifted with respect to 8 . The interactions of complexes 1 and 2 with bovine serum albumin (BSA) and double stranded DNA (ds-DNA) were also studied by electrochemical methods. Both complexes showed strong binding to BSA. Evidence for intercalation of both complexes with DNA is presented, with 1 showing a stronger interaction than 2 . [ABSTRACT FROM AUTHOR]
- Published
- 2015
- Full Text
- View/download PDF
17. Photophysical Studies of Bioconjugated Ruthenium Metal-Ligand Complexes Incorporated in Phospholipid Membrane Bilayers.
- Author
-
Sharmin, Ayesha, Salassa, Luca, Rosenberg, Edward, Alexander Ross, J. B., Abbott, Geoffrey, Black, Labe, Terwilliger, Michelle, and Brooks, Robert
- Subjects
- *
BIOCONJUGATES , *RUTHENIUM , *LIGANDS (Chemistry) , *METAL complexes , *PHOSPHOLIPIDS , *ARTIFICIAL membranes , *BILAYER lipid membranes - Abstract
The luminescent, mono-diimine ruthenium complexes [(H)Ru(CO)(PPh3)2(dcbpy)][PF6] (1) (dcbpy = 4,4'-dicarboxy-2,2'-bipyridyl) and [(H)Ru(CO)(dppene)(5-amino-l,10-phen)] [PF6] (2) (dppene = bis(diphenyiphosphino)- ethylene; phen = phenanthroline) were conjugated with 1,2-dihexadecanoyl-sn-glycero-3-phosphoethanolamine (DPPE) and with cholesterol in the case of complex 2. Using standard conjugation techniques, compound 1 gives the bis-lipid deriv-ative [(H)Ru(CO)(PPh3)2(dcbpy-N-DPPE2)][PF6] (3), while 2 provides the monolipid conjugate [(H)Ru(CO)(dppene)-(1,10-phen-5-NHC(S)-N-DPPE)][PF6] (4) and the cholesterol derivative [(H)Ru(CO)(dppene)(1,10-phen-5-NHC(0)-Ocholesteryl)][PF6] (5). These compounds were characterized by spectroscopic methods, and their photophysical properties were measured in organic solvents. The luminescence of lipid conjugates 3 and 4 is quenched in organic solvents while compound 4 shows a weak, short-lived, blue-shifted emission in aqueous solution. The cholesterol conjugate S shows the long-lived, microsecond-time scale emission associated with triplet metal-to-ligand charge-transfer excited states. Incorporation of conjugate 3 in lipid bilayer vesicles restores the luminescence, but with blue shifts (~80 nm) accompanied by nanosecond-time scale lifetimes. In the vesicles conjugate 4 shows a short-lived and blue-shifted emission similar to that observed in solution but with increased intensity. Conjugation of the complex [(H)Ru(CO)(PhP2C2H4C(O)O-N-succinimidyl)2(bpy)][PF6] (6") (bpy = 2,2'-bipyridyl) with DPPE gives the phosphine-conjugated complex [(H)Ru(C0)(PhP2C2H4C(O)-N-DPPE)2(bpy)][PF6] (7). Complex 7 also exhibits a short-lived and blue-shifted emission in solution and in vesicles as observed for complexes 3 and 4. We have also conjugated the complex [Ru(bpy)2(5-amino-l,10-phen)][PF6]2 (8) with both cholesterol (9) and DPPE (10). Neither complex 9 nor the previously reported complex 10 exhibited the blue shifts observed for complexes 3 and 4 when incorporated into large unilamellar vesicles (LUVs). The anisotropics of the emissions of complexes 3, 4, and 7 were also measured in LUVs, and those of complex S were measured in both glycerol and LUVs. High fundamental anisotropics were observed for complexes 3, 4, and 7. [ABSTRACT FROM AUTHOR]
- Published
- 2013
- Full Text
- View/download PDF
18. Comparison of Novel and Patented Silica-Polyamine Composite Materials as Aqueous Heavy Metal Ion Recovery Materials.
- Author
-
BEATTY, SUSAN, FISCHER, ROBERT, ROSENBERG, EDWARD, and PANG, DAVID
- Subjects
- *
SILICA , *POLYAMINES , *COMPOSITE materials , *HEAVY metals , *METAL ions - Abstract
A novel silica–polyamine composite material, WP-1, was synthesized according to a new patented procedure for aqueous heavy metal ion removal. The effects of pH and initial concentration of metal ions on adsorption were investigated. Results showed that WP-1 has metal ion capacities in the order Cu(II) > Ni(II) > Fe(II) and generally works best in the 2–4 pH range. Kinetic studies showed the Cu(II) adsorption reaction occurs in a second-order fashion with a rate constant of 4.4 × 10-4 s-1. Two similar silica–polyamine materials, prepared by different procedures cited in the literature, were compared with WP-1 for Cu(II) ion capacity and mechanical durability. These tests showed that capacities were more than twice as high when the new patented procedure was used, indicating much better coverage of the silica gel. During repeated metal loading and unloading tests, WP-1 maintained up to 94% of its initial Cu(II) capacity and its mechanical integrity for 3000 cycles, whereas testing of the other materials had to be stopped prematurely due to degradation of the materials. [ABSTRACT FROM AUTHOR]
- Published
- 1999
- Full Text
- View/download PDF
19. Trans influence and substituent effects on the HOMO-LUMO energy gap and Stokes shift in Ru mono-diimine derivatives.
- Author
-
AlAbbad, Sanaa, Sardot, Tova, Lekashvili, Oliko, Decato, Daniel, Lelj, Francesco, Ross, J.B. Alexander, and Rosenberg, Edward
- Subjects
- *
STOKES shift , *BAND gaps , *INTRAMOLECULAR proton transfer reactions , *OPTICAL spectra - Abstract
The ground (S 0) and excited triplet (T 1) electronic states and corresponding optical spectra of a series of cationic complexes [RuH(CO)L(PPh 3) 2 ]+ (L = 2,2′-bipyridyl) (Rubpy), 4,4′-dicarboxylic-2,2′-bipyridyl(Rudcbpy),bis-4,4′-(N-methylamide)-2,2′-bipyridyl(Rudamidebpy),bis-4,4′-(methyl)-2,2′-bipyridyl(RudMebpy),[Ru(CO) 2 dcbpy(PPh 3) 2 ]2+(Ru(2CO)dcbpy),and [Ru(H) 2 dcbpy(PPh 3) 2 ] (Ru(2H)dcbpy) have been studied by combined Density Functional/Time-Dependent Density Functional (DFT/TDDFT) techniques using different combinations of DFT exchange-correlation functionals and basis sets. PBE0/LANL2DZ provided more accurate geometries to describe S 0 whereas B3LYP/LANL2DZ predicted spectral energies that correlated better with the available experiment data. The Ru (II) complexes with different substituents emit photons ranging from 560 to 610 nm in the series RudMebpy, Rubpy, Rudamidebpy, Rudcbpy. The calculations predicted a maximum emission at about 540 nm for the complex constructed from two carbonyl π-acceptors ligands trans to the dcbpy, while an emission in the far infrared region is calculated when two H σ-donor ligands trans to the dcbpy. Our calculation results show correlations between HOMO-LUMO energy gap, Stokes shift, and T 1 distortion, which reflect the different effects of electron-withdrawing and donating groups. We proposed that these correlations can be used to predict the photophysical properties for new complexes. The correlation between HOMO-LUMO energy gap, absorption and emission energies, Stokes shift, and the geometric distortion in the triplet excited state (T 1) as a result of introducing electron-withdrawing substituent (-COOH) (Rudcbpy) in [RuH(CO)bpy(PPh 3) 2 ]+ (Rubpy). Image 1 • Combined DFT/TDDFT calculations reproduced the spectroscopic data of Ru(bpy)(PPh 3) 2 derivatives. • Evidence of the trans influence and the substituents effects to control the HOMO-LUMO energy gap and Stokes shift. • Small HOMO-LUMO energy gap is associated with large geometric distortion in the triplet excited state and large Stokes shift. [ABSTRACT FROM AUTHOR]
- Published
- 2019
- Full Text
- View/download PDF
20. Reactivity of [Os3(CO)9(μ3-η1,κ1-C9H6N)(μ-H)] towards Ph3EH (E = Si, Ge, Sn): Synthesis and structure of triosmium clusters bearing a Ph3E ligand.
- Author
-
Parvin, Sozony, Uddin, Mohammad J., Rajbangshi, Subas, Rosenberg, Edward, Ghosh, Shishir, and Kabir, Shariff E.
- Subjects
- *
TIN , *X-ray crystallography , *MOLECULAR structure , *OXIDATIVE addition , *DECARBONYLATION , *ISOMERS - Abstract
[Display omitted] Reactions of the electron defficient triosmium [Os 3 (CO) 9 (μ 3 -η1,κ1-C 9 H 6 N)(μ-H)] with Ph 3 EH (E = Si, Ge, Sn) have been investigated. The reactions of quinoline-capped triosmium cluster [Os 3 (CO) 9 (μ 3 -η1,κ1-C 9 H 6 N)(μ-H)] (1) with Ph 3 EH (E = Si, Ge, Sn) have been investigated. Cluster 1 reacts with Ph 3 EH in refluxing benzene to afford [Os 3 (CO) 9 (EPh 3)(μ-η1,κ1-C 9 H 6 N)(µ-H) 2 ] (2a , E = Si; 2b , E = Ge; 2c , E = Sn) and [Os 3 (CO) 8 (EPh 3)(μ 3 -η1,κ1-C 9 H 6 N)(µ-H) 2 ] (3a , E = Si; 3b , E = Ge; 3c , E = Sn). Control experiments confirm that 2 is a precursor of 3. Cluster 2 is formed via oxidative-addition of E–H bond to 1 whose decarbonylation leads to the formation of 3. The molecular structures of all the new clusters (except 3c) have been determined by X-ray crystallography. Both 2 and 3 exist as a pair of isomers in solution, and their fluxionality in solution has been probed by VT NMR experiments. [ABSTRACT FROM AUTHOR]
- Published
- 2023
- Full Text
- View/download PDF
21. Corrigendum to ‘Synthesis, structure, photophysical and electrochemical properties of Ru(TFA)(CO)(PPh3)2(L) (L=2-phenylpyridine, 2-p-tolylpyridine) and Ru(CO)(PPhMe2)2(L)(L′) (L= TFA, H) (L′= bipyridine, L′= 4,4′-dimethylbipyridine) relationships between ancillary ligand structure and luminescent properties’ [J. Organomet. Chem. 849–850 (2017) 306–314]
- Author
-
Pohkrel, Shyam, Decato, Dan, Rosenberg, Edward, Ross, J.B. Alexander, and Terwilliger, Michelle
- Subjects
- *
RUTHENIUM compound synthesis , *PYRIDINE - Published
- 2018
- Full Text
- View/download PDF
22. Reaction of electron-deficient 6-methoxyquinolinate-substituted cluster [Os3(CO)9{μ3-η1,κ1-C9H5N(6-OMe)}(μ-H)] with PPh3: Thermally induced ligand isomerization, decarbonylation and orthometallation.
- Author
-
Begum, Shahin A., Chowdhury, Arshad H., Ghosh, Shishir, Tocher, Derek A., Rosenberg, Edward, Hardcastle, Kenneth I., and Kabir, Shariff E.
- Subjects
- *
ELECTRON-deficient compounds , *QUINOLINIC acid , *ISOMERIZATION , *DECARBONYLATION , *TEMPERATURE effect - Abstract
The reaction of the electron-deficient 6-methoxyquinolinate-substituted triosmium cluster [Os 3 (CO) 9 {μ 3 -η 1 ,κ 1 -C 9 H 5 N(6-OMe)}(μ-H)] ( 1 ) with PPh 3 proceeds readily at room temperature to afford the electron-precise adduct [Os 3 (CO) 9 (PPh 3 ){μ-η 1 ,κ 1 -C 9 H 5 N(6-OMe)}(μ-H)] ( 2 ), in which the PPh 3 ligand is bound to the osmium that serves as the coordination site for the hydride and the metallated-carbon atom of the benzoheterocycle. This reaction also leads to a change in the hapticity of the 6-methoxyquinolinate ligand from μ 3 -η 1 ,κ 1 to μ-η 1 ,κ 1 as confirmed by X-ray crystallography. Thermolysis of 2 in boiling toluene furnishes five new triosmium clusters ( 3 – 7 ) as a result of ligand isomerization, decarbonylation and orthometallation of the ancillary PPh 3 ligand. Clusters 3 and 4 are isomers of 2 and the location of the metal-bound hydride and PPh 3 with respect to 6-methoxyquinolinate moiety is the only difference between these isomers. Control experiments show that 5 is a decarbonylation product of 4 which converts into 6 as a result of further decarbonylation with concomitant orthometallation of one of the phenyl rings of the coordinated PPh 3 ligand, whilst 7 is formed from 6 through replacement of an equatorial carbonyl of the nitrogen-bound osmium by PPh 3 ligand. All the new clusters have been characterized by a combination of analytical and spectroscopic methods as well as by X-ray crystallography in the case of 2 , 3 , 6 and 7 . [ABSTRACT FROM AUTHOR]
- Published
- 2018
- Full Text
- View/download PDF
23. Reactions of the face-capped benzothiazolate-substituted clusters Os3(CO)9(μ3,η2-C7H3NSR)(μ-H) (R = H, Me) with PPh3: Kinetic formation of Os3(CO)9(PPh3)(μ,η2-C7H3NSR)(μ-H) and thermally induced ligand isomerization.
- Author
-
Begum, Shahin A., Chowdhury, Md. Arshad H., Ghosh, Shishir, Tocher, Derek A., Richmond, Michael G., Rosenberg, Edward, and Kabir, Shariff E.
- Subjects
- *
ISOMERIZATION , *HETEROCYCLIC compounds , *LIGANDS (Chemistry) , *TEMPERATURE effect , *X-ray crystallography , *THERMOLYSIS - Abstract
The reaction of the benzothiazolate-capped triosmium clusters Os 3 (CO) 9 (μ 3 ,η 2 -C 7 H 3 NSR)(μ-H) ( 1a , R = H; 1b , R = Me) with PPh 3 proceeds readily at room temperature with a μ 3 ,η 2 → μ,η 2 hapticity change in the benzothiazolate heterocycle to furnish Os 3 (CO) 9 (PPh 3 )(μ,η 2 -C 7 H 3 NSR)(μ-H) ( 2a , R = H; 2b , R = Me) in high yields. X-ray crystallography has confirmed the regiospecific nature of this reaction where the PPh 3 ligand is bound to the osmium atom that serves as the coordination site for the hydride and the metalated-carbon atom associated with the edge-bridged benzothiazolate ligand. The thermolysis of 2a and 2b in boiling toluene affords several new Os 3 clusters as a result of ligand isomerization, decarbonylation, and ortho metalation of the ancillary PPh 3 ligand. The new products have been isolated and characterized by a combination of spectroscopic methods and X-ray crystallography in the case of 3b , 4b , 5b and 6b . Clusters 3b and 4b are isomers of 2b and differ in the location of the hydride and PPh 3 ligands relative to the benzothiazolate moiety. Electronic structure calculations on the isomeric clusters 2b , 3b , and 4b confirm that 2b is the kinetic product of ligand substitution, accounting for its rearrangement to the latter two isomers upon heating. Cluster 5b contains a face-capping benzothiazolate moiety and is shown by DFT calculations to derive from a site-selective loss of an axial CO group in 4b . Cluster 6a,b formed from 5a,b as a result of further decarbonylation with concomitant ortho metalation of one of the phenyl groups of the coordinated PPh 3 ligand. [ABSTRACT FROM AUTHOR]
- Published
- 2017
- Full Text
- View/download PDF
24. Alkyne.
- Author
-
Karakhanov, Edward A., Maximov, Anton L., Zolotukhina, Anna V., Yatmanova, Nadezhda, and Rosenberg, Edward
- Subjects
- *
ALKYNES , *PROPENE , *DENDRIMERS , *SILICA , *CATALYSTS - Abstract
A series of Pd-Ag mixed-metal nanocatalysts were prepared by reduction of Pd-Ag salts in the presence of poly(propylene imine) dendrimers, which were covalently bound to the surface of a silica polyamine composite, BP-1 (polyallylamine covalently bound to a silanized amorphous silica gel). Three different Pd-to-Ag ratios were evaluated (50:50, catalyst 1; 40:60, catalyst 2; 60:40, catalyst 3) with the goal of determining how the amount of Ag effects selectivity, rate and conversion in the selective , such as phenylacetylene and 1- or 4-octyne, to the corresponding alkenes. Conditions for the are reported where there is improved selectivity without a serious reduction in rate when compared with the analogous Pd-only catalysts. Catalyst 2 worked best for phenylacetylene and catalyst 3 worked best for the octynes. The catalysts could be reused seven times without loss of activity. Copyright © 2015 John Wiley & Sons, Ltd. [ABSTRACT FROM AUTHOR]
- Published
- 2015
- Full Text
- View/download PDF
25. Ruthenium Nanoparticles Stabilized in Cross-Linked Dendrimer Matrices: Hydrogenation of Phenols in Aqueous Media.
- Author
-
Maximov, Anton, Zolotukhina, Anna, Murzin, Vadim, Karakhanov, Edward, and Rosenberg, Edward
- Subjects
- *
RUTHENIUM catalysts , *POLYPROPYLENE , *PHENOLS , *CROSSLINKING (Polymerization) , *HYDROGENATION , *NANOPARTICLES analysis , *DENDRIMERS , *AQUEOUS solutions - Abstract
Novel catalysts consisting of ruthenium nanoparticles encapsulated in cross-linked matrices based on the poly(propylene imine) dendrimers of the 1st and 3rd generations have been synthesized with a narrow particle size distribution (3.8 and 1.0 nm, respectively). The resulting materials showed high activity for the hydrogenation of phenols in aqueous media (specific catalytic activity reached turnover frequencies of 2975 h−1 with respect to hydrogen uptake). It has been shown that the use of water as a solvent leads to a 1.5 to 50-fold increase in the reaction rate depending upon the nature of the substrate. It has been established that unlike the traditional heterogeneous catalysts based on ruthenium, during the hydrogenation of dihydroxybenzenes, the hydrogenation rate decreases in the order: resorcinol>hydroquinone≫catechol. The maximum specific activity for resorcinol was a turnover frequency of 243150 h−1 with respect to hydrogen uptake. The catalyst based on the dendrimer of the 3rd generation containing finer particles has significantly inferior activity to the catalyst based on the dendrimer of the 1st generation by virtue of steric factors, as well as the need for prereduction of the ruthenium oxide contained on the surface. These catalysts showed resistance to metal leaching and may be reused several times without loss of activity. [ABSTRACT FROM AUTHOR]
- Published
- 2015
- Full Text
- View/download PDF
26. Experimental and computational studies on the reaction of silanes with the diphosphine-bridged triruthenium clusters Ru3(CO)10(µ-dppf), Ru3(CO)10(µ-dppm) and Ru3(CO)9{µ3- PPhCH2PPh(C6H4)}.
- Author
-
Hossain, Md. Jakir, Rajbangshi, Subas, Khan, Md. Mehedi M., Ghosh, Shishir, Hogarth, Graeme, Rosenberg, Edward, Hardcastle, Kenneth I., Richmond, Michael G., and Kabir, Shariff E.
- Subjects
- *
SILANE compounds , *METAL clusters , *COMPUTATIONAL chemistry , *CHEMICAL reactions , *DIPHOSPHINE , *LIGANDS (Chemistry) , *RUTHENIUM compounds - Abstract
Reactions of Ru3(CO)10(µ-dppf) (1) (dppf = 1,1'-bis(diphenylphosphino)ferrocene), Ru3(CO)10(µ-dppm) (2) (dppm = bis(diphenylphosphino)methane), and the orthometalated derivative Ru3(CO)9{µ3- PPhCH2PPh(C6H4)} (3) with silanes (Ph3SiH, Et3SiH, Ph2SiH2) are reported. Treatment of 1 with Ph3SiH and Ph2SiH2 at room temperature leads to facile Si--H bond activation to afford Ru3(CO)9(µ-dppf)(SiPh3)(µ-H) (4) (60% yield) and Ru3(CO)9(µ-dppf)(SiPh2H)(µ-H) (6) (53% yield), respectively. The reaction of 1 with Ph3SiH has been investigated by electronic structure calculations, and these data have facilitated the analysis of the potential energy surface leading to 4. Compound 1 does not react with Et3SiH at room temperature but reacts at 68 °C to give Ru3(CO)9(µ-dppf)(SiEt3)(µ-H) (5) in 45% yield. Reaction of 2 with Ph3SiH at room temperature yields two new products: Ru3(CO)9(µ-dppm)(SiPh3)(µ-H) (7) in 40% yield and Ru3(CO)6(µ3-O)(µ-dppm)(SiPh3)(µ- H)3 (8) in 15% yield. Interestingly, at room temperature compound 7 slowly reverts back to 2 in solution with decomposition and liberation of Ph3SiH. Complex 8 can also be prepared from the direct reaction between 7 and H2O. Similar reactions of 2 with Et3SiH and Ph2SiH2 give only intractable materials. The orthometalated compound 3 does not react with Ph3SiH, Et3SiH and Ph2SiH2 at room temperature but does react at 66 °C to give Ru3(µ-CO)(CO)7{µ3- PPhCH2PPh(C6H4)}(SiR2R¹)(µ- H)](9, R = R' = Ph, 71% yield; 10, R = R' = Et, 60% yield; 11, R = Ph, R' = H, 66% yield) by activation of the Si--H bond. Compounds 4 and 8--11 have been structurally characterized. In 4, both the dppf and the hydride bridge a common Ru--Ru vector, whereas NMR studies on 7 indicate that two ligands span different Ru--Ru edges. Compound 8 contains a face-capping oxo moiety, a terminally coordinated SiPh3 ligand, and three bridging hydride ligands, whereas 9--11 represent simple oxidative addition products. In all of the compounds examined, the triruthenium framework retains its integrity and the silyl groups occupy equatorial sites. [ABSTRACT FROM AUTHOR]
- Published
- 2014
- Full Text
- View/download PDF
27. Structural design at the polymer surface interface in nanoporous silica polyamine composites
- Author
-
Allen, Jesse, Berlin, Matthew, Hughes, Mark, Johnston, Erik, Kailasam, Varadharajan, Rosenberg, Edward, Sardot, Tova, Wood, Jessica, and Hart, Carolyn
- Subjects
- *
MOLECULAR structure , *POLYMERS , *POROUS silicon , *POLYAMINES , *NANOCOMPOSITE materials , *LEACHING , *SURFACES (Technology) , *NUCLEAR magnetic resonance , *SILICA gel , *CHEMICAL processes - Abstract
Abstract: The factors affecting the rate of silica leaching in alkaline aqueous media from surface silanized, nanoporous, amorphous, silica gels and from silanized silica gels that have been modified with polyamines to form the previously reported silica polyamine composites (SPCs), BP-1 and BP-2 have been investigated. Silanization with alkyl trichlorosilanes slows the rate of silica leaching relative to the unmodified silica gels. The use of bulkier aryl silanes somewhat decreases the silica leaching under the same conditions. Interestingly, after modification of the silanized silica with poly(allylamine) (PAA) to make BP-1, the leaching increases, but subsequent modification of the SPC with chloroacetic acid to make BP-2, quenches this increase. A mechanism explaining these results is discussed. Analogous composites have been prepared using sol–gel chemistry. These materials were characterized and their silica leaching properties were compared with the original BP-1. CPMAS 13C and 29Si NMR of the various surfaces have been applied to better understand the nature of the modified surfaces. Significant changes in the nature of the surface siloxanes are observed for the different matrices and on their conversion to the polyamine composite. Scanning electron microscopy and pore size distributions for the composites made from commercial silica gel and from sol–gel chemistry are also reported and compared. [Copyright &y& Elsevier]
- Published
- 2011
- Full Text
- View/download PDF
28. Cluster-mediated alkenyl isomerism and carbon–carbon bond formation: The reaction of the unsaturated benzothiazole cluster [Os3(CO)9(μ3-C7H4NS)(μ-H)] with dimethyl acetylenedicarboxylate
- Author
-
Mahid Uddin, Kh., Ghosh, Shishir, Raha, Arun K., Hogarth, Graeme, Rosenberg, Edward, Sharmin, Ayesha, Hardcastle, Kenneth I., and Kabir, Shariff E.
- Subjects
- *
METAL clusters , *ORGANOOSMIUM compounds , *THIAZOLES , *ISOMERISM , *CHEMICAL bonds , *HEATING , *METAL complexes , *UNSATURATED compounds - Abstract
Abstract: Upon mild heating (60–65°C), electron-deficient [Os3(CO)9(μ3-C7H4NS)(μ-H)] (1) readily adds to the activated alkyne, RChsp sp="0.25" />=CO2Me), to give two isomers of the alkenyl complexes, [Os3(CO)9(μ-C7H4NS)(μ3-RCCHR)] (2 and 3), differing in the orientation of the benzothiazole ligand. In both the alkenyl ligand acts as a five-electron donor binding through both carbons and one of the oxygen atoms of the substituents, which adopt a relative cis disposition. Heating 2 cleanly affords 3 suggesting that the former is the kinetic product and the latter is thermodynamically favoured. At higher temperatures (110°C), 3 rearranges to a third isomer [Os3(CO)9(μ-C7H4NS)(μ-RCCHR)] (4) which differs in the binding mode of the alkenyl ligand and where the alkyne substituents are in a relative trans disposition. A second product of this reaction is [Os3(CO)8(μ-OCH3){μ3-C7H4NSC(R)C}(μ-H)] (5) which results from the loss of two moles of CO, carbon–carbon coupling between hydrocarbyl and benzothiazole ligands and carbon–hydrogen addition to the triosmium centre. All new clusters are characterized by a total of 50 valence electrons and contain two metal–metal bonds as shown by single crystal X-ray diffraction studies. [Copyright &y& Elsevier]
- Published
- 2010
- Full Text
- View/download PDF
29. Tuning photophysical properties with ancillary ligands in Ru(II) mono-diimine complexes.
- Author
-
Sharmin, Ayesha, Darlington, Reuben C., Hardcastle, Kenneth I., Ravera, Mauro, Rosenberg, Edward, and Alexander Ross, J. B.
- Subjects
- *
COMPLEX compounds synthesis , *RUTHENIUM compounds , *LIGANDS (Chemistry) , *IMINES , *SOLID state chemistry , *SOLUTION (Chemistry) , *NUCLEAR magnetic resonance spectroscopy , *CHARGE transfer - Abstract
The series of complexes [XRu(CO)(L–L)(L′)2][PF6] (X = H, TFA, Cl; L–L = 2,2′-bipyridyl, 1,10-phenanthroline, 5-amino-1,10-phenanthroline and 4,4′-dicarboxylic-2,2′-bipyridyl; L′2 = 2PPh3, Ph2PC2H4PPh2, Ph2PCH=CHPPh2) have been synthesized from the starting complex K[Ru(CO)3(TFA)3] (TFA = CF3CO2) by first reacting with the phosphine ligand, followed by reaction with the L–L and anion exchange with NaPF6. In the case of L–L = phenanthroline and L′2 = 2PPh3, the neutral complex Ru(Ph3P)(CO)(1,10-phenanthroline)(TFA)2 is also obtained and its solid state structure is reported. Solid state structures are also reported for the cationic complexes where L–L = phenanthroline, L2 = 2PPh3 and X = Cl and for L–L = 2,2′-bipyridyl, L2 = 2PPh3 and X = H. All the complexes were characterized in solution by a combination of 1H and 31P NMR, IR, mass spectrometry and elemental analyses. The purpose of the project was to synthesize a series of complexes that exhibit a range of excited-state lifetimes and that have large Stokes shifts, high quantum yields and high intrinsic polarizations associated with their metal-to-ligand charge-transfer (MLCT) emissions. To a large degree these goals have been realized in that excited-state lifetimes in the range of 100 ns to over 1 μs are observed. The lifetimes are sensitive to both solvent and the presence of oxygen. The measured quantum yields and intrinsic anisotropies are higher than for previously reported Ru(II) complexes. Interestingly, the neutral complex with one phosphine ligand shows no MLCT emission. Under the conditions of synthesis some of the initially formed complexes with X = TFA are converted to the corresponding hydrides or in the presence of chlorinated solvents to the corresponding chlorides, testifying to the lability of the TFA Ligand. The compounds show multiple reduction potentials which are chemically and electrochemically reversible in a few cases as examined by cyclic voltammetry. The relationships between the observed photophysical properties of the complexes and the nature of the ligands on the Ru(II) is discussed. [Copyright &y& Elsevier]
- Published
- 2009
- Full Text
- View/download PDF
30. A comparative study of the reactivity of unsaturated triosmium clusters [Os3(CO)8{μ3-Ph2PCH2P(Ph)C6H4}(μ-H)] and [Os3(CO)9{μ3-η2-C7H3(2-Me)NS}(μ-H)] with t BuNC
- Author
-
Raha, Arun K., Ghosh, Shishir, Manzurul Karim, Md., Tocher, Derek A., Begum, Noorjahan, Sharmin, Ayesha, Rosenberg, Edward, and Kabir, Shariff E.
- Subjects
- *
UNSATURATED compounds , *MICROCLUSTERS , *CHEMICAL reactions , *PHOSPHINE , *ISOCYANIDES , *X-ray diffraction - Abstract
Abstract: Treatment of unsaturated [Os3(CO)8{μ3-Ph2PCH2P(Ph)C6H4}(μ-H)] (2) with t BuNC at room temperature gives [Os3(CO)8(CNBu t )){μ3-Ph2PCH2P(Ph)C6H4}(μ-H)] (3) which on thermolysis in refluxing toluene furnishes [Os3(CO)7(CNBu t ){μ3-Ph2PCHP(Ph)C6H4}(μ-H)2] (4). Reaction of the labile complex [Os3(CO)9(μ-dppm)(NCMe)] (5) with t BuNC at room temperature affords the substitution product [Os3(CO)9(μ-dppm)(CNBu t )] (6). Thermolysis of 6 in refluxing toluene gives 4. On the other hand, the reaction of unsaturated [Os3(CO)9{μ3-η2-C7H3(2-Me)NS}(μ-H)] (7) with t BuNC yields the addition product [Os3(CO)9(CNBu t ){μ-η2-C7H3(2-Me)NS}(μ-H)] (8) which on decarbonylation in refluxing toluene gives unsaturated [Os3(CO)8(CNBu t ){μ3-η2-C7H3(2-Me)NS}(μ-H)] (9). Compound 9 reacts with PPh3 at room temperature to give the adduct [Os3(CO)8(PPh3)(CNBu t ){μ-η2-C7H3(2-Me)NS(μ-H)] (10). Compound 8 exists as two isomers in solution whereas 10 occurs in four isomeric forms. The molecular structures of 3, 6, 8, and 10 have been determined by X-ray diffraction studies. [Copyright &y& Elsevier]
- Published
- 2008
- Full Text
- View/download PDF
31. Synthesis, structure, photophysical and electrochemical behavior of 2-amino-anthracene triosmium clusters
- Author
-
Sharmin, Ayesha, Minazzo, Agnese, Salassa, Luca, Rosenberg, Edward, Alexander Ross, J.B., Kabir, Shariff E., and Hardcastle, Kenneth I.
- Subjects
- *
ANTHRACENE , *POLYCYCLIC aromatic hydrocarbons , *X-ray diffraction , *SPECTROSCOPIC imaging - Abstract
Abstract: The reactions of 2-amino-anthracene with [Os3(CO)10(CH3CN)2] have been studied and the products structurally characterized by spectroscopic, X-ray diffraction, photophysical and electrochemical techniques. At room temperature in CH2Cl2 two major, isomeric products are obtained [Os3(CO)10(μ-η2-(N-C(1))-NH2C14H8)(μ-H)] (1, 14%) and [Os3(CO)10(μ-η2-(N-C(3))-NHC14H9)(μ-H)] (2, 35%) along with a trace amount of the dihydrido complex [Os3(CO)9(μ-η2-(N-C(3))-NHC14H8)(μ-H)2] (3). In refluxing tetrahydrofuran only complexes 2 and 3 are obtained in 24% and 28%, respectively. A separate experiment shows that complex 1 slowly converts to 2 and that the rearrangement is catalyzed by adventitious water and involves proton transfer to the anthracene ring. Complex 1 is stereochemically non-rigid; exhibiting edge to edge hydride migration while 2 is stereochemically rigid. Complex 3 is also stereochemically non-rigid showing a site exchange process of the magnetically nonequivalent hydrides typical for trinuclear dihydrides. Interestingly, 2 decarbonylates cleanly to the electronically unsaturated 46e− cluster [Os3(CO)9(μ3-η2-(N-C(3))-NHC10H9)(μ-H)] (4, 68%) in refluxing cyclohexane, while photolysis of 2 in CH2Cl2 yields only a small amount of 3 along with considerable decomposition. The mechanism of the conversion of 1 to 2 and the dependence of the product distribution on solvent are discussed. All four compounds are luminescent with compounds 1–3 showing emissions that can be assigned to radiative decay associated with the anthracene ligand. Complexes 1–3 all show irreversible 1e− reductions in the range of −1.85–2.14V while 4 shows a nicely reversible 1e− wave at −1.16V and a quasi-reversible second 1e− wave at −1.62V. Irreversible oxidations are observed in the range from +0.35 to +0.49V. The relationship between the cluster ligand configurations and the observed electrochemical and photochemical behavior is discussed and compared with that of the free ligand. [Copyright &y& Elsevier]
- Published
- 2008
- Full Text
- View/download PDF
32. Spectroscopic and Computational Studies of a Ru(II) Terpyridine Complex: The Importance of Weak Intermolecular Forces to Photophysical Properties.
- Author
-
Garino, Claudio, Gobetto, Roberto, Nervi, Carlo, Salassa, Luca, Rosenberg, Edward, Ross, J. B. Alexander, Xi Chu, Hardcastle, Kenneth I., and Sabatini, Cristiana
- Subjects
- *
RUTHENIUM , *INTERMOLECULAR forces , *CHEMICAL bonds , *ABSORPTION , *DENSITY functionals , *EMISSION spectroscopy - Abstract
The complex [Ru(tpy)(CO)2TFA]+[PF6]- (where tpy = 2,2′:6′,2″-terpyridine and TFA = CF3CO2-) (1) has been synthesized and fully characterized spectroscopically. The X-ray structure of the complex has been determined. The photopysical properties of the ruthenium complex and the free ligand tpy have been investigated at room temperature and at 77 K in acetonitrile solution and in the solid state. Their electronic spectra are highly influenced by intermolecular stacking interactions, both in solution and in the solid state. Density functional theory (DFT) and time-dependent DFT (TDDFT) calculations have been performed to characterize the electronic structure and the excited states of [Ru(tpy)(CO)2TFA]+[PF6]- and tpy. TDDFT calculations on three different conformations of free ligand have been performed as well. Absorption and emission spectra of tpy have been studied at different temperatures and concentrations in order to have a better understanding of this ruthenium derivative's properties. The absorption spectrum of 1 is characterized by metal-perturbed ligand-centered (LC) bands in the UV region. No metal-to-ligand charge transfer (MLCT) bands are observed in the visible for the complex. Only at high concentrations (10-4 M) does a very weak band appear at 470 nm. At 77 K and low concentrations, solutions of 1 exhibit a major 3LC emission band centered at 468 nm (21.4 × 10-3 cm-1). When the concentration of the complex is increased, an unstructured narrow emission at 603 nm (16.6 × 10-3 cm-1), with a lifetime of 10 μs, dominates the emission spectrum in glassy acetonitrile. This emission originates from a π-π stacked dimeric (or oligomeric) species. TDDFT calculations performed on a tail-to-tail dimer structure, similar to that seen in the solid state, ascribe the transition to a triplet excited state, where intermolecular metal (d) → ligand (π*, polypyridine) charge transfer occurs. A good estimate of the transition energy is also obtained (623 nm, 1.94 eV). [ABSTRACT FROM AUTHOR]
- Published
- 2007
- Full Text
- View/download PDF
33. Photophysical properties and computational investigations of tricarbonylrhenium(I)[2-(4-methylpyridin-2-yl)benzo[d]-X-azole]L and tricarbonylrhenium(I)[2-(benzo[d]-X-azol-2-yl)-4-methylquinoline]L derivatives (X=N–CH3, O, or S; L=Cl−, pyridine)
- Author
-
Albertino, Andrea, Garino, Claudio, Ghiani, Simona, Gobetto, Roberto, Nervi, Carlo, Salassa, Luca, Rosenberg, Edward, Sharmin, Ayesha, Viscardi, Guido, Buscaino, Roberto, Croce, Gianluca, and Milanesio, Marco
- Subjects
- *
ELECTROCHEMISTRY , *LUMINESCENCE , *X-ray diffraction , *LIGANDS (Chemistry) - Abstract
Abstract: We report a combined experimental and computational study of new rhenium tricarbonyl complexes based on the bidentate heterocyclic N–N ligands 2-(4-methylpyridin-2-yl)benzo[d]-X-azole (X=N–CH3, O, or S) and 2-(benzo[d]-X-azol-2-yl)-4-methylquinoline (X=N–CH3, O, or S). Two sets of complexes are reported. Chloro complexes, described by the general formula Re(CO)3[2-(4-methylpyridin-2-yl)benzo[d]-X-azole]Cl (X=N–CH3, 1; X=O, 2; X=S, 3) and Re(CO)3[2-(benzo[d]-X-azol-2-yl)-4-methylquinoline]Cl (X=N–CH3, 4; X=O, 5; X=S, 6) were synthesized heating at reflux Re(CO)5Cl with the appropriate N–N ligand in toluene. The corresponding pyridine set {Re(CO)3[2-(4-methylpyridin-2-yl)benzo-X-azole]py}PF6 (X=N–CH3, 7; X=O, 8; X=S, 9) and {Re(CO)3[2-(benzo[d]-X-azol-2-yl)-4-methylquinoline]py}PF6 (X=N–CH3, 10; X=O, 11; X=S, 12) was synthesized by halide abstraction with silver nitrate of 1–6 followed by heating in pyridine and isolated as their hexafluorophosphate salts. All complexes have been fully characterized by IR, NMR, electrochemical techniques and luminescence. The crystal structures of 1 and 7 were obtained by X-ray diffraction. DFT and time-dependent (TD) DFT calculations were carried out for investigating the effect of the organic ligand on the optical properties and electronic structure of the reported complexes. [Copyright &y& Elsevier]
- Published
- 2007
- Full Text
- View/download PDF
34. Computational Studies of Nucleophilic Attack and Protonation of Electron-Deficient Benzoheterocycle Triosmium Clusters.
- Author
-
Nowroozi-Isfahani, Taraneh, Musaev, Djamaladdin G., Morokuma, Keiji, and Rosenberg, Edward
- Subjects
- *
DENSITY functionals , *PROTON transfer reactions , *NUCLEOPHILIC reactions , *TRANSITION metal complexes , *ADDITION polymerization , *TAUTOMERISM - Abstract
The density functional theory method has been applied to gain insights into the regioselectivity of nucleophilic attack and protonation of electron-deficient benzoheterocycle triosmium clusters. We report our computational results on the reaction of the green 46-electron triosmium clusters Os3(CO)9(μ3-η²-(LH))(μ-H) (L = benzoxazole, 1a; benzothiazole, 1b; dihydroquinoline, 1c; 1,3-dehydroindoline, 1d; 4H-3,1-benzoxazine, 1e) with hydride (H-) and proton (H+) in order to elucidate factors affecting the observed differences in the structure of the kinetic products of these reactions. Transition-state calculations for the interconversion of the anionic tautomers resulting from H- attack on the clusters 1a-e show that the activation energies of these anionic clusters are considerably lower than the previously reported barriers for related neutral clusters. Calculations also reveal that the structures of the kinetic products resulting from sequential H-/H+ attack are determined by the protonation process. [ABSTRACT FROM AUTHOR]
- Published
- 2006
- Full Text
- View/download PDF
35. Reactions of [(μ-H)Os3(CO)10(μ-OMe)] and [(μ-H)Os3(CO)9(μ-OMe)(MeCN)] with dppm, dppe, dppp, and PPh2H: X-ray crystal structures of [(μ-H)Os3(CO)8(μ-OMe)(μ2-η2-dppm)] and [(μ-H)Os3(CO)9(μ-OMe)(PPh2H)]
- Author
-
Kabir, Shariff E., Mottalib, Md. Abdul, Hossain, G.M. Golzar, Nordlander, Ebbe, and Rosenberg, Edward
- Subjects
- *
MANURE gases , *MOLECULAR structure , *FARM manure , *MANURES - Abstract
Abstract: Treatment of [(μ-H)Os3(CO)10(μ-OMe)] (1) with dppm {dppm=bis(diphenylphosphino)methane} at 110°C gave the known compound [Os3(CO)10(μ2-η2-dppm)] (2) and the new compound [(μ-H)Os3(CO)8(μ-OMe)(μ2-η2-dppm)] (3). A similar reaction of 1 with dppe {dppe=bis(diphenylphosphino)ethane} afforded two products, [(μ-H)Os3(CO)8(μ-OMe)(μ1-η2-dppe)] (5) and [(μ-H)Os3(CO)8(μ-OMe)(μ2-η2-dppe)] (6) which differ in the mode of bonding of the dppe ligand. The reaction of 1 with dppp {dppp=bis(diphenylphosphino)propane} at 110°C gave exclusively [(μ-H)Os3(CO)8(μ-OMe)(μ2-η2-dppp)] (7) which exists as two isomers in solution. Compounds 2, 3, and 5–7 are also obtained in almost similar yields from reactions of [(μ-H)Os3(CO)9(μ-OMe)(MeCN)] (4) with the appropriate diphosphine at ambient temperature. The reaction of 1 with PPh2H at 110°C gave the new compounds [(μ-H)Os3(CO)9(μ-OMe)(PPh2H)] (8) and [(μ-H)2Os3(CO)7(μ-PPh2)2(PPh2H)] (10) and the previously reported compound [(μ-H)2Os3(CO)8(μ-PPh2)2] (9). The molecular structures of 3 and 8 were determined by single crystal X-ray diffraction. In compound 3, the dppm ligand bridges the same Os–Os edge as the hydride and the methoxide ligand. In 8, the PPh2H ligand occupies a ‘pseudo’ axial site on one of the osmium atoms bearing the bridging hydride and methoxide ligand. [Copyright &y& Elsevier]
- Published
- 2006
- Full Text
- View/download PDF
36. Triosmium clusters containing bridging dppm and EPh (E=S, Te) ligands: X-ray structures of [(μ-H)Os3(CO)7(μ-SPh){μ3-η4-Ph2PCHP(Ph)C6H4}], [Os3(CO)8(μ-SPh)2(μ-dppm)] and [Os3(CO)8(μ-TePh)2(μ-dppm)]
- Author
-
Kabir, Shariff E., Saha, Madhu S., Tocher, Derek A., Hossain, G.M. Golzar, and Rosenberg, Edward
- Subjects
- *
LIGANDS (Chemistry) , *BIOCHEMISTRY , *MOLECULAR structure , *CHEMICAL structure - Abstract
Abstract: Heating [Os3(CO)10(μ-dppm)] (1) with two equivalents of PhSSPh in toluene under reflux provided three new triosmium compounds [(μ-H)Os3(CO)7(μ-SPh){μ3-η4-Ph2PCHP(Ph)C6H4}] (2), [Os3(CO)8(μ-SPh)2(μ-dppm)] (3) and [(μ-H)Os3(CO)7(μ-η2-SC6H4)(μ-SPh)(μ-dppm)] (4) in 20%, 21% and 26% yields, respectively. In contrast, a similar reaction of 1 with two equivalents of PhTeTePh in refluxing toluene gave the binuclear compound [Os2(CO)4(μ-TePh)2(μ-dppm)] (6) in 15% yield, and two 50 electron isomeric compounds 5 and 7 with the formula [Os3(CO)8(μ-TePh)2(μ-dppm)] in 20% and 23% yields, respectively. Thermolysis of 3 at 110°C afforded 4 in 53% yield which on further thermolysis in refluxing octane at 128°C gave 2 in 45% yield. Thermolysis of 3 in refluxing octane also gave 2 in 50% yield. The new compounds, 2–7, were all spectroscopically characterized, and the X-ray structures of 2, 3 and 7 have been determined. Compound 2 contains a bridging SPh ligand and a μ3,η4-Ph2PCHP(Ph)C6H4 ligand, formed by two kinds of C–H activation, including orthometallation of a phenyl group as well as an unusual activation of the methylene group of the dppm ligand. The molecular structure of 3 reveals that two SPh groups span the open Os–Os edge of the Os3 triangle, while the dppm ligand bridges one of the closed Os–Os edges. In compound 7, one TePh group spans the open Os–Os edge, while the other spans one of the two closed Os–Os edges and the dppm ligand bridges the third Os–Os vector. [Copyright &y& Elsevier]
- Published
- 2006
- Full Text
- View/download PDF
37. Dithiolate Complexes of Manganese and Rhenium: X-ray Structure and Properties of an Unusual Mixed Valence Cluster Mn3(CO)6(μ-n2-SCH2CH2CH2S)3.
- Author
-
Begum, Noorjahan, Hyder, Iqbal, Kabir, Shariff E., Golzar Hossain, G. M., Nordlander, Ebbe, Rokhsana, Dalia, and Rosenberg, Edward
- Subjects
- *
METAL complexes , *TRANSITION metal complexes , *COMPLEX compounds , *CARBONYL compounds , *BENZENE , *ETHANES - Abstract
Treatment of Mn2(CO)10 with 3,4-toluenedithiol and 1,2-ethanedithiol in the presence of Me3NO·2H2O in CH2Cl2 at room temperature afforded the dinuclear complexes Mn2(CO)6(μ-ρ4-SC6H3(CH3)S-SC6H3(CH3)S) (1), and Mn2-(CO)6(μ-ρ4-SCH2CH2S-SCH2CH2S) (2), respectively. Similar reactions of Re2(CO)10 with 3,4-toluenedithiol, 1,2-benzenedithiol, and 1,2-ethanedithiol yielded the dirhenium complexes Re2(CO)6(μ-ρ4.SC6H3(CH3)S-SC6H3(CH3)S) (3), Re2(CO)6(μ-ρ4-SC6H4S-SC6H4S) (4), and Re2(CO)6(SCH2CH2S-SCH2CH2S) (5), respectively. In contrast, treatment of Mn2(CO)10 with 1,3-propanedithiol afforded the trimanganese compound Mn3(CO)6(μ-ρ2-SCH2CH2-CH2S)3 (6), whereas Re2(CO)10 gave only intractable materials. The molecular structures of 1, 3, and 6 have been determined by single-crystal X-ray diffraction studies. The dimanganese and dirhenium carbonyl compounds 1–5 contain a binucleating disulfide ligand, formed by interligand disulfide bond formation between two dithiolate ligands identical in structure to that of the previously reported dimanganese complex Mn2(CO)6(μ-ρ4-SC6H4S-SC6H4S). Complex 6, on the other hand, forms a unique example of a mixed-valenca trimangenese carbonyl compound containing three bridging 1,3-propanedithiolate ligands. The solution properties of 6 have been investigated by UV-vis and EPR spectroscopies as well as electrochemical techniques. [ABSTRACT FROM AUTHOR]
- Published
- 2005
- Full Text
- View/download PDF
38. Reactions of the unsaturated triosmium cluster [(μ-H)Os3(CO)8(Ph2PCH2P(Ph)C6H4)] with HX (X=Cl, Br, F, CF3CO2,CH3CO2): X-ray structures of [(μ-H)Os3(CO)7(η1-Cl)(μ-Cl)2(μ-dppm)],[(μ-H)2Os3(CO)8(Ph2PCH2P(Ph)C6H4)]+[CF3O]− and the two isomers of [(μ-H)Os3(CO)8(μ-Cl)(μ-dppm)]
- Author
-
Kabir, Shariff E., Miah, Md. Arzu, Sarker, Nitai C., Hossain, G.M. Golzar, Hardcastle, Kenneth I., Rokhsana, Dalia, and Rosenberg, Edward
- Subjects
- *
HYDROGEN-ion concentration , *CATIONS , *PHYSICAL & theoretical chemistry , *DICHLOROMETHANE - Abstract
Abstract: Treatment of the electronically unsaturated cluster [(μ-H)Os3(CO)8(Ph2PCH2P(Ph)C6H4)] (1) with HCl gas at ambient temperature in dichloromethane afforded (μ-H)Os3(CO)8(μ-Cl)(μ-dppm) (2) and [(μ-H)Os3(CO)7(η1-Cl)(μ-Cl)2(μ-dppm)] (3). Thermolysis of 2 at 110 °C led to an isomer of 2, 4. A similar reaction of 1 with HBr gas gave [(μ-H)Os3(CO)8(μ-Br)(μ-dppm)] (5) as the only product which does not isomerize at 110 °C. In sharp contrast, treatment of 1 with HF gas gave the protonated species [(μ-H)2Os3(CO)8(Ph2PCH2P(Ph)C6H4)]+ (6). Treatment of 1 with CF3CO2H also gave cation 6 whereas CH3CO2H yielded [(μ-H)Os3(CO)8(μ-η2-CH3CO2)(μ-dppm)] (7). Structures of 2, 3, 4 and 6 were established crystallographically. In 2, both the chloride and the hydride ligands simultaneously bridge the same Os–Os edge and the dppm spans another Os–Os bond whereas in 4, all the three ligands bridge the same Os–Os edge. Compound 3 is comprised of an open Os3 arrangement in which one chloride bridges the open Os⋯Os edge, another chloride and a hydride mutually bridge an Os–Os bond and the third chloride is terminally coordinated to one of the Os atoms of the dppm bridged edge. The cation 6 consists of a triangle of osmium atoms in which the shortest Os–Os edge is bridged by a hydride and the metallated phenyl ring and the longest edge is bridged by another hydride and the diphosphine ligand. [Copyright &y& Elsevier]
- Published
- 2005
- Full Text
- View/download PDF
39. Water-soluble benzoheterocycle triosmium clusters as potential inhibitors of telomerase enzyme
- Author
-
Colangelo, Donato, Ghiglia, AnnaLisa, Ghezzi, AnnaRita, Ravera, Mauro, Rosenberg, Edward, Spada, Fabrizio, and Osella, Domenico
- Subjects
- *
QUINOLINE , *CHLOROQUINE , *HYDROXYQUINOLINE , *QUINOXALINES , *TELOMERASE - Abstract
Abstract: We have studied the ability of several bioorganometallic clusters [(μ-H)Os3(CO)9(L)(μ3-η2-(Q-H))], where L=[P(C6H4SO3Na)3] or [P(OCH2CH2NMe3I)3], and Q=quinoline, 3-aminoquinoline, quinoxaline or phenanthridine, of inhibiting telomerase, a crucial enzyme for cancer progression. In general, quinolines have shown interesting biological properties, especially in inhibiting enzymes. For example, the 2,3,7-trichloro-5-nitroquinoxaline (TNQX) exhibited strong anti-telomerase activity in vitro. Among the quinoline-clusters under study, only the negatively charged ones (by virtue of the sulfonated phosphines) exhibited good anti-telomerasic activity on semi-purified enzyme in a cell-free assay, while they were ineffective in vitro on Taq, a different DNA-polymerase. On the contrary, the treatment of breast cancer MCF-7 cell line did not evidence any activity of these clusters, suggesting a low aptitude for crossing cell membrane. Furthermore, all clusters exhibited non-specific, acute cytotoxicy, probably due to accumulation on cell membranes by virtue of their amphiphilic character. A detailed study of Os uptake and accumulation in MCF-7 cells supported this hypothesis. [Copyright &y& Elsevier]
- Published
- 2005
- Full Text
- View/download PDF
40. Triruthenium clusters containing bridging dppm and capping sulfido and selenido ligands: X-ray structures of [Ru3(CO)5(μ3-CO)(μ3-Se)(μ-dppm)2], [Ru3(CO)6(μ3-CO)(μ3-Se)(μ-dppm)(η1-Ph2PCH2P(==O)Ph2)] and [Ru2(CO)4(μ-SePh)2(μ-dppm)].
- Author
-
Kabir, Shariff E., Ahmed, Syed J., Hyder, Md. Iqbal, Miah, Md. Arzu, Bennett, Dennis W., Haworth, Daniel T., Siddiquee, Tasneem A., and Rosenberg, Edward
- Subjects
- *
ORGANORUTHENIUM compounds , *METAL carbonyls , *PHOSPHORUS compounds , *BIPHENYL compounds , *PHOSPHINE , *METHANE , *LIGANDS (Chemistry) , *MOLECULAR structure , *X-ray diffraction - Abstract
The reaction of [Ru3(CO)10(mu-dppm)] (1) with dppmSe at 66 °C affords [Ru3(CO)8(mu-dppm)2] (2), [Ru3(CO)7(mu3-CO)(mu3-Se)(mu-dppm)] (3), [Ru3(CO)5(mu3-CO)(mu3-Se)(mu-dppm)2] (4) and [Ru3(CO)6(mu3-CO)(mu3-Se)(mu-dppm)(mu1-Ph2PCH2P(double bond; length as m-dashO)Ph2)] (5) in 7%, 5%, 9% and 33% yields, respectively. A similar reaction between 1 with dppmS gives [Ru3(CO)7(mu3-S)2(mu-dppm)] (6), [Ru3(CO)7(mu3-CO)(mu3-S)(mu-dppm)] (7) [Ru3(CO)5(mu3-CO)(mu3-S)(mu-dppm)2] (8) and [Ru3(CO)6(mu3-CO)(mu3-S)(mu-dppm)(mu1-Ph2PCH2P(double bond; length as m-dashO)Ph2)] (9) in 8%, 7%, 14% and 35% yields, respectively. Treatment of 1 with PhSeSePh at 66 °C affords the dinuclear compound [Ru2(CO)4(mu-SePh)2(mu-dppm)] (10) in 14% yield. Thermolysis of 5 and 9in refluxing toluene at 110 °C gives 4 and 8, respectively. The molecular structures of 4, 9 and 10 have been determined by single-crystal X-ray diffraction studies. The cores of the new clusters 4, 5, 8 and 9 consist of metal triangles capped by mu3-sulfur or selenium atoms with the bidentate ligand bridging in equatorial positions. In compounds 4 and 8, two bidentate dppm ligands bridge the Ru3 triangle in such a way that each ligand bridges two ruthenium atoms and one Ru-Ru edge remains unbridged. Compounds 5 and 9 contain one bridging dppm ligand and one dangling dppm mono-oxide ligand Ph2PCH2P(double bond; length as m-dashO)Ph2 coordinated to the rear metal atom at an equatorial position. The molecular structure of 10 shows classical "sawhorse" structure with two bridging SePh ligands as well as the dppm ligand. [ABSTRACT FROM AUTHOR]
- Published
- 2004
- Full Text
- View/download PDF
41. Reactions of benzothiazolide triosmium clusters with tetramethylthiourea
- Author
-
Begum, Noorjahan, Deeming, Antony, Islam, Mohammad, Kabir, Shariff, Rokhsana, Dalia, and Rosenberg, Edward
- Published
- 2004
- Full Text
- View/download PDF
42. Hexa- and triosmium carbonyl clusters bearing bridging dppm and capping sulfido ligands
- Author
-
Akter, Tahmina, Begum, Noorjahan, Haworth, Daniel T., Bennett, Dennis W., Kabir, Shariff E., Miah, Md. Arzu, Sarker, Nitai C., Siddiquee, Tasneem A., and Rosenberg, Edward
- Published
- 2004
- Full Text
- View/download PDF
43. Solution properties, electrochemical behavior and protein interactions of water soluble triosmium carbonyl clusters
- Author
-
Nervi, Carlo, Gobetto, Roberto, Milone, Luciano, Viale, Alessandra, Rosenberg, Edward, Spada, Fabrizio, Rokhsana, Dalia, and Fiedler, Jan
- Subjects
- *
SOLUTION (Chemistry) , *OSMIUM compounds , *TRANSITION metal compounds , *CARBON compounds , *ORGANIC chemistry - Abstract
The synthesis and solution chemistry of the water soluble clusters [Os3(CO)9(
μ-η2 -Bz)(μ-H)L+] (HBz=quinoxaline, L+=[P(OCH2CH2NMe3)3I3], 1) along with its negatively charged analog [Os3(CO)9(μ-η2 -Bz)(μ-H)L-] (L-=[P(C6H4SO3)3Na3], 2) are reported. In addition, we have examined the reduction potentials of the complexes [Os3(CO)9(μ-η2 -Bz)(μ-H)L] (HBz=phenanthridine, L=L+ (3); HBz=5,6 benzoquinoline, L=L+ (4); HBz=3-amino quinoline, L=L+ (5); HBz=3-amino quinoline, L=L- (6). The neutral analog of 1 and 2 [Os3(CO)9(μ-η2 -Bz)(μ-H) PPh3] (Bz=quinoxaline, 7) was also examined for comparison. Both compounds 1 and 2 show pH dependent NMR spectra that are interpreted in terms of the extent of protonation of the uncoordinated quinoxaline nitrogen which impacts the degree of aggregation of the clusters in aqueous solution. Compound 1 undergoes a reversible 1e- reduction in water while 2 undergoes a quasi-reversible 1e- reduction at more negative potentials as expected from the difference in charge on the phosphine ligand. Compound 7 undergoes a marginally reversible CV in methylene chloride at a potential intermediate between the positively and negatively charged clusters. The overall stability of the radical anions of 1, 2 and 7 is somewhat less than the corresponding decacarbonyl [Os3(CO)10(μ-η2 -Bz)(μ-H)] (HBz=quinoxaline). While complexes 1 and 2 show reversible 1e- reductions, all the other complexes examined show 1e- and/or two 1e- irreversible reductions in aqueous and non-aqueous solvents. The potentials for these complexes follow expected trends relating to the charge on the phosphine and the pH of the aqueous solutions. The ligand dependent trends are compared with those of the previously reported corresponding decacarbonyls. The interactions of the positively and negatively charged clusters with albumin have been investigated using the transverse and longitudinal relaxation times of the hydride resonances as probes of binding to the protein. Evidence of binding is observed for both the positive and negative clusters but the positive and negative clusters exhibit distinctly different rotational correlation times. Two additional complexes [Os3(CO)9(μ-η2 -Bz)(μ-H)L] (HBz=2-methylbenzimidazole, L=L+ (8); L=L- (10) and HBz=quinoline-4-carboxaldehyde, L=L+ (9); L=L- (11)) are reported in connection with these studies. [Copyright &y& Elsevier]- Published
- 2004
- Full Text
- View/download PDF
44. Addition of diphenyl diselenide (PhSeSePh) to the clusters [Os3(CO)10(μ-dppm)] and [(μ-H)Os3(CO)8{Ph2PCH2P(Ph)C6H4}]: X-ray structures of [Os2(CO)4(μ-SePh)2(μ-dppm)], [Os3(CO)6(μ-CO)(μ-Se)2(μ-C6H4)(μ-dppm)] and two isomers of [Os3(CO)8(μ-SePh)2(μ-dppm)]
- Author
-
Kabir, Shariff E., Begum, Noorjahan, Manjur Hassan, Md., Iqbal Hyder, Md., Nur, Hani, Bennett, Dennis W., Siddiquee, Tasneem A., Haworth, Daniel T., and Rosenberg, Edward
- Abstract
The reaction of [Os3(CO)10(μ-dppm)] (4) with diphenyldiselenide in refluxing toluene at 110 °C affords the dinuclear compound [Os2(CO)4(μ-SePh)2(μ-dppm)] (5), three 50 electron isomeric triosmium compounds 6, 7 and 8 with the formula [Os3(CO)8(μ-SePh)2(μ-dppm)], two triosmium benzyne compounds [Os3(CO)6(μ-CO)(μ-
η2 -Se)2(μ-C6H4)(μ-dppm)] (9) and [Os3(CO)9(μ-SePh)2(μ-η2 -C6H4)(μ-dppm)] (10) in 11%, 19%, 15%, 5%, 18% and 6% yields, respectively. Thermolysis of both 6 and 8 in refluxing toluene gives the dinuclear compound 5 in moderate yield, whereas a similar thermolysis of 7 yields 5 and 9 in 20% and 27% yields, respectively. Compound 10 converts to 9 refluxing in octane. Treatment of the unsaturated compound [(μ-H)Os3(CO)8{Ph2PCH2P(Ph)C6H4}] (11) with diphenyldiselenide in refluxing benzene gives 5, 6, 7 and 8 in 8%, 20%, 25% and 10% yields, respectively. The molecular structures of 5, 6, 7 and 9 have been determined by single crystal X-ray diffraction studies. The molecular structure of 5 shows classical “sawhorse” structure with two bridging phenylselenido ligands as well as a dppm ligand. The solid-state structure of 6 reveals that two SePh groups span the open Os–Os edges of the Os3 triangle, while the dppm ligand bridges one of the closed Os–Os edges. In compound 7, one SePh spans the open Os–Os site, while the other spans one of the two closed Os–Os edges and the dppm ligand bridges the third Os–Os vector. Compound 9, which exists as a mixture of two isomers in solution, contains two triply bridging selenido ligands and a benzyne ligand, which are believed to have been formed by the cleavage of Se–Se bond of the PhSeSePh ligand followed by C–Se and C–H activation of the SePh group. The 54 electron compound 10 has been characterized by elemental analysis, IR, 1H NMR, 31P{1H} NMR and mass spectroscopic data. [Copyright &y& Elsevier]- Published
- 2004
- Full Text
- View/download PDF
45. Triosmium and triruthenium clusters containing deprotonated 2-mercapto-1-methylimidazole ligand: X-ray structures of [(μ-H)Os3(CO)10{μ-SC&z.dbnd6;NCH&z.dbnd6;CHN(CH3)}], [(μ-H)Ru3(CO)9{(μ3-η2-SC&z.dbnd6;NCH&z.dbnd6;CHN(CH3)}] and [(μ-H)Os3(CO)8{μ-SC&z.dbnd6;NCH&z.dbnd6;CHN(CH3)}(μ-dppm)]
- Author
-
Azam, Kazi A., Hanif, Khandakar M., Ghosh, Amaresh C., Kabir, Shariff E., Karmakar, Shafali R., Abdul Malik, K.M., Parvin, Salina, and Rosenberg, Edward
- Subjects
- *
HETEROCYCLIC compounds , *X-ray crystallography - Abstract
The reactions of [Os3(CO)10(MeCN)2], [Ru3(CO)12], [(μ-H)Os3(CO)8{Ph2PCH2P(Ph)C6H4}], [Os3(CO)10(μ-dppm)] and [Ru3(CO)10(μ-dppm)] with 2-mercapto-1-methylimidazole are reported. The labile cluster [Os3(CO)10(MeCN)2] reacts with 2-mercapto-1-methylimidazole at 25 °C to give [(μ-H)Os3(CO)10{μ-SC&z.dbnd6;NCH&z.dbnd6;CHN(CH3)}] (1) whereas the corresponding reaction of [Ru3(CO)12] with the ligand at 68 °C gives [(μ-H)Ru3(CO)9{μ3-η2-SC&z.dbnd6;NCH&z.dbnd6;CHN(CH3)}] (2). Both [(μ-H)Os3(CO)8{Ph2PCH2P(Ph)C6H4}] and [Os3 (CO)10(μ-dppm)] react with 2-mercapto-1-methylimidazole at 25 and 110 °C, respectively, to give [(μ-H)Os3(CO)8{μ-SC&z.dbnd6;NCH&z.dbnd6;CHN(CH3)(μ-dppm)] (3). In contrast [Ru3(CO)10(μ-dppm)] reacts with 2-mercapto-1-methylimidazole at 68 °C to yield the known compound [Ru3(CO)7(μ3-CO)(μ3-S)(μ-dppm)] (4). The structures of 1–3 have been determined by X-ray crystallography. [Copyright &y& Elsevier]
- Published
- 2002
- Full Text
- View/download PDF
46. Tandem hydroformylation/hydrogenation over novel immobilized Rh-containing catalysts based on tertiary amine-functionalized hybrid inorganic-organic materials.
- Author
-
Gorbunov, Dmitry, Nenasheva, Maria, Naranov, Evgeny, Maximov, Anton, Rosenberg, Edward, and Karakhanov, Eduard
- Subjects
- *
CATALYSTS , *RHODIUM catalysts , *HYDROFORMYLATION , *HYDROGENATION , *AMINO group , *ALCOHOL , *SURFACES (Technology) , *RHODIUM - Abstract
[Display omitted] • Novel Rh-containing non-phosphorous supported catalysts for tandem hydroformylation-hydrogenation were obtained. • Tertiary nitrogen atoms of the polymer anchored on material's surface promote aldehyde hydrogenation over Rh active centers. • Catalysts showed high activity (hydroformylation TOF 206 h−1) and chemoselectivity. • Catalyst K N performed stable hydroformylation yield at least in the first 5 cycles with a decrease in alcohol selectivity. Non-phosphorous rhodium-containing catalysts for direct conversion of olefins to alcohols via tandem hydroformylation/hydrogenation have been designed and synthesized. Interaction between Rh(acac)(CO) 2 and tertiary amino groups on the surface of mesoporous hybrid organic-inorganic supports yielded materials which were successfully used in the tandem process. Data obtained for a selected catalyst K N demonstrate that rhodium is in the Rh+1 state highly dispersed on the surface and is bonded with nitrogen atoms both before and after use. Evaluation of the catalytic performance shows high activity (hydroformylation TOF 312 h−1), chemoselectivity and stable hydroformylation yield at least in the first 5 cycles with a decrease in alcohol selectivity. The influence of temperature, reaction time, total pressure, and molar CO/H 2 ratio of syngas on oxygenate yields is described. Type of the hydroformylation active sites and possible pathways for the observed decrease in hydrogenation are discussed. [ABSTRACT FROM AUTHOR]
- Published
- 2021
- Full Text
- View/download PDF
47. Silica-Polyamine Composite Materials for Heavy Metal Ion Removal, Recovery, and Recycling. II. Metal Ion Separations from Mine Wastewater and Soft Metal Ion Extraction Efficiency*.
- Author
-
FISCHER, ROBERT, PANG, DAVID, BEATTY, SUSAN, and ROSENBERG, EDWARD
- Subjects
- *
METAL ions , *COMPOSITE materials , *SILICA , *SEPARATION (Technology) - Abstract
Details a study which described two modified silica-polyamine composite materials. Separation of metal ions; Efficiency composite materials in metal ion removal; Discussion and conclusion.
- Published
- 1999
- Full Text
- View/download PDF
48. Dr W. H. R. Rivers: Siegfried Sassoon and Robert Graves' 'fathering friend'.
- Author
-
Wilson, Jean Moorcroft
- Subjects
- *
WORLD War I , *HISTORY of mental illness , *LITERARY criticism , *POETRY (Literary form) , *PHYSICIANS , *HISTORY of military medicine , *FRIENDSHIP , *HISTORY , *WAR ,HISTORY of military personnel - Published
- 2017
- Full Text
- View/download PDF
Catalog
Discovery Service for Jio Institute Digital Library
For full access to our library's resources, please sign in.