Journal articles on the topic 'Ruthenium-dmso'

To see the other types of publications on this topic, follow the link: Ruthenium-dmso.

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Ruthenium-dmso.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Chandra, Manish, D. Shankar Pandey, M. Carmen Puerta, and Pedro Valerga. "Ap-cymene-ruthenium(II)–DMSO complex, [(η6-C10H14)RuCl2(DMSO)]." Acta Crystallographica Section E Structure Reports Online 58, no. 1 (December 14, 2001): m28—m29. http://dx.doi.org/10.1107/s1600536801021237.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

Zhu, Dengsen, Cong Zhao, Xuesong Wang, Wenji Wang, Baohuai Wang, and Weihong Du. "Roles of DMSO-type ruthenium complexes in disaggregation of prion neuropeptide PrP106–126." RSC Advances 6, no. 19 (2016): 16055–65. http://dx.doi.org/10.1039/c5ra21523d.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Jaswal, Jaswinder S., Steven J. Rettig, and Brian R. James. "Ruthenium(III) complexes containing dimethylsulfoxide or dimethylsulfide ligands, and a new route to trans-dichlorotetrakis(dimethylsulfoxide)ruthenium(II)." Canadian Journal of Chemistry 68, no. 10 (October 1, 1990): 1808–17. http://dx.doi.org/10.1139/v90-282.

Full text
Abstract:
The new Ru(III) complex tran-[(dmso)2H]+[RuCl4(dmso)2]− (1) (dmso = S-bonded dimethylsulfoxide) has been synthesized. New synthetic routes are reported for mer-RuX3(dms)3 complexes (dms = dimethylsulfide, X = Cl (2) and Br (3)) and for trans-RuCl2(dmso)4 (4). The complexes have been studied spectroscopically while 1, 2 and 4 have also been characterized crystallographically. Crystals of 1 are monoclinic, P2/n, a = 9.280(3), b = 16.518(4), c = 14.034(3) Å, β = 100.78(2)°, Z = 4, ρc = 1.75 g cm−3. Crystals of 2 are orthorhombic, Pca21a = 10.764(2), b = 11.375(1), c = 12.243(1) Å, Z = 4, ρc = 1.74 g cm−3.Crystals of 4 are tetragonal, I4/m, a = 9.1256(8), c = 11.184(2) Å, Z = 2, ρc = 1.73 g cm−3. The structures were determined by heavy atom methods and were refined by full-matrix least-squares procedures to R (Rw) values of 0.030 (0.036), 0.025 (0.032), and 0.051 (0.071) for 3051, 3354, and 1205 reflections with I ≥ 3σ(I), respectively. The ionic complex 1 contains the protonated dmso cation [(dmso)2H]+, with the proton asymmetrically bonded between the two O atoms (O(3)—H(1) = 1.30(6); 0(4)—H(1) = 1.12(6) Å), and the [RuCl4(dmso)2]− octahedral anion with trans-disposed S-bonded dmso ligands. The structure determined for 4 duplicates one reported recently in the literature (E. Alessio etal. Inorg. Chem. 27, 4099 (1988)), and reveals all S-bonded dmso ligands; the Ru—Cl distance, 2.432(1) Å, is significantly longer than that reported, 2.402(2) Å. The dms complexes 2 and 3 were isolated unexpectedly from reaction of Ru salts with acidic dmso solutions. Keywords: ruthenium complexes, dimethylsulfoxide, dimethylsulfide, hydrogen bonding.
APA, Harvard, Vancouver, ISO, and other styles
4

Messori, Luigi, Felix Kratz, and Enzo Alessio. "The Interaction of the Antitumor Complexes Na[trans-RuCl4(DMSO)(Im)] and Na[trans-RuCl4(DMSO)(Ind)] With Apotransferrin: a Spectroscopic Study." Metal-Based Drugs 3, no. 1 (January 1, 1996): 1–9. http://dx.doi.org/10.1155/mbd.1996.1.

Full text
Abstract:
The interaction of two antitumor ruthenium(III) complexes,-Na[trans-RuCl4(DMSO)(Im)] and Na[trans-RuCl4(DMSO)(Ind)]- with human serum apotransferrin (apoTf) was investigated through a number of spectroscopic techniques such as UV-Vis absorption, CD and H1 NMR spectroscopy. Interestingly, the hydrolysis profiles of these complexes in a physiological buffer are markedly affected by the presence, in solution, of apoTf suggesting the occurrence of a specific interaction of their respective hydrolysis products with the protein. The formation of stable adducts with apotransferrin has been demonstrated by CD spectroscopy, and additional information obtained through H1 NMR of the hyperfine shifted signals. The bound ruthenium(III) species may be detached from these adducts by addition of excess citrate at low pH. The behavior of the investigated ruthenium(III) complexes with apoTf is compared with that of the recently described and strictly related ru-im and ru-ind antitumour complexes, and discussed in the frame of general strategies of drug targeting.
APA, Harvard, Vancouver, ISO, and other styles
5

Kljun, Jakob, Saša Petriček, Dušan Žigon, Rosana Hudej, Damijan Miklavčič, and Iztok Turel. "Synthesis and Characterization of Novel Ruthenium(III) Complexes with Histamine." Bioinorganic Chemistry and Applications 2010 (2010): 1–6. http://dx.doi.org/10.1155/2010/183097.

Full text
Abstract:
Novel ruthenium(III) complexes with histamine[RuCl4(dmso-S)(histamineH)]⋅O(1a) and[RuCl4(dmso-S)(histamineH)](1b) have been prepared and characterized by X-ray structure analysis. Their crystal structures are similar and show a protonated amino group on the side chain of the ligand which is not very common for a simple heterocyclic derivative such as histamine. Biological assays to test the cytotoxicity of the compound1bcombined with electroporation were performed to determine its potential for future medical applications in cancer treatment.
APA, Harvard, Vancouver, ISO, and other styles
6

Rachford, Aaron A., Jeffrey L. Petersen, and Jeffrey J. Rack. "Efficient Energy Conversion in Photochromic Ruthenium DMSO Complexes." Inorganic Chemistry 45, no. 15 (July 2006): 5953–60. http://dx.doi.org/10.1021/ic0603398.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

Groot, Broer de, Hilary A. Jenkins, Stephen J. Loeb, and Shannon L. Murphy. "Ruthenium(II) complexes of the thiacyclophane ligands 2,5,8-trithia[9]-o-cyclophane (TT[9]OC) and 5-oxa-2,8-dithia[9]-o-cyclophane (ODT[9]OC). Structures of RuCl2(DMSO)(TT[9]OC) and RuCl2(PPh3)(ODT[9]OC)." Canadian Journal of Chemistry 73, no. 7 (July 1, 1995): 1102–10. http://dx.doi.org/10.1139/v95-136.

Full text
Abstract:
Ruthenium(II) complexes of the thiacyclophane ligands 2,5,8-trithia[9]-o-cyclophane (TT[9]OC) and 5-oxa-2,8-dithia[9]-o-cyclophane (ODT[9]OC) were synthesized by ligand displacement reactions employing RuCl2(DMSO)4, RuCl2(PPh3)3, and RuHCl(PPh3)3 as starting materials. X-ray crystal structures of two of these complexes, RuCl2(DMSO)(TT[9]OC) and RuCl2(PPh3)(ODT[9]OC), demonstrate how TT[9]OC and ODT[9]OC bind to Ru(II). RuCl2(DMSO)(TT[9]OC) crystallized as the DMSO solvate in the orthorhombic space group Pbca with a = 19.590(5), b = 16.849(4), c = 13.149(4) Å, V = 4340(3) Å3, and Z = 8. The structure refined to R = 5.27% and Rw = 6.27% for 2472 reflections with Fo2 > 3σ(Fo2). RuCl2(PPh3)(ODT[9]OC) crystallized as a ClCH2CH2Cl solvate in the monoclinic space group P21/c with a = 7.912(1), b = 22.419(5), c = 18.794(3) Å, β = 101.12(1),° V = 3271.2(9) Å3, and Z = 4. The structure refined to R = 4.96% and Rw = 5.14% for 1800 reflections with Fo2 > 3σ(Fo2). Both compounds are octahedral with the thiacyclophane ligand bound through three donor atoms in a facial coordination mode, cis chlorine atoms, and the unique ancillary ligand, DMSO or PPh3, bound trans to the central S or O donor of the macrocycle. The X-ray structures support 1H NMR spectral evidence which shows that the ligands are bound in an "endo" mode for L = DMSO and in an "exo" mode when L = PPh3. The reaction of RuHCl(PPh3)3 with TT[9]OC yields the ruthenium hydride complex RuHCl(PPh3)2(TT[9]OC). 1H and 31P NMR spectroscopy are consistent with an octahedral species for which the macrocycle occupies only two coordination sites acting as a bidentate η2-chelating ligand. Keywords: thioether, macrocycle, hydride, crystal structure.
APA, Harvard, Vancouver, ISO, and other styles
8

Turel, Iztok, Milena Pecanac, Amalija Golobic, Enzo Alessio, Barbara Serli, and Alberta Bergamo. "Ruthenium(III)-DMSO complexes of the antiherpes drug acyclovir." Journal of Inorganic Biochemistry 96, no. 1 (July 2003): 241. http://dx.doi.org/10.1016/s0162-0134(03)80789-4.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Chan, Peter K. L., Paul K. H. Chan, David C. Frost, Brian R. James, and Kirsten A. Skov. "Ruthenium (II) complexes of 4-nitroimidazoles: their characterization, solution chemistry, and radiosensitizing activity." Canadian Journal of Chemistry 66, no. 1 (January 1, 1988): 117–22. http://dx.doi.org/10.1139/v88-018.

Full text
Abstract:
A series of ruthenium(II) complexes of formulation RuCl2(dmso)2Ln, where dmso = S-bonded dimethylsulfoxide, L is a 4-nitroimidazole derivative, and n = 1 or 2, have been synthesized and characterized by spectroscopic methods, particularly 1H nmr and X-ray photoelectron spectroscopy. With L = 4-nitroimidazole itself (4-NO2Im), RSU-1170, -3083 or -3100, n = 2 and the six-coordinate complexes are considered to be of cis,cis,cis-geometry. The N-methyl-4-nitroimidazole (NMe-4-NO2Im) ligand (n = 1) chelates through the imidazole-N and the oxygen of NO2. The RSU-3159 ligand (n = 1) binds through the sulfur and may be chelated. The complexes are of interest because of their radiosensitizer properties toward hypoxic tumour cells; the RuCl2(dmso)2(NMe-4-NO2Im) complex has a higher sensitizing enhancement ratio than the free imidazole ligand, and shows no cytoxicity in vitro, and these data are compared to those reported previously for RuCl2(dmso)2(4-NO2Im)2. Some aspects of the aqueous solution chemistry of the complexes (aquation via loss of chloride, pKa of coordinated water) are discussed in relation to their biological activities.
APA, Harvard, Vancouver, ISO, and other styles
10

Motswainyana, William M., and Peter A. Ajibade. "Anticancer Activities of Mononuclear Ruthenium(II) Coordination Complexes." Advances in Chemistry 2015 (February 19, 2015): 1–21. http://dx.doi.org/10.1155/2015/859730.

Full text
Abstract:
Ruthenium compounds are highly regarded as potential drug candidates. The compounds offer the potential of reduced toxicity and can be tolerated in vivo. The various oxidation states, different mechanism of action, and the ligand substitution kinetics of ruthenium compounds give them advantages over platinum-based complexes, thereby making them suitable for use in biological applications. Several studies have focused attention on the interaction between active ruthenium complexes and their possible biological targets. In this paper, we review several ruthenium compounds which reportedly possess promising cytotoxic profiles: from the discovery of highly active compounds imidazolium [trans-tetrachloro(dmso)(imidazole)ruthenate(III)] (NAMI-A), indazolium [trans-tetrachlorobis(1H-indazole)ruthenate(III)](KP1019), and sodium trans-[tetrachloridobis(1H-indazole)ruthenate(III)] (NKP-1339) to the recent work based on both inorganic and organometallic ruthenium(II) compounds. Half-sandwich organometallic ruthenium complexes offer the opportunity of derivatization at the arene moiety, while the three remaining coordination sites on the metal centre can be functionalised with various coordination groups of various monoligands. It is clear from the review that these mononuclear ruthenium(II) compounds represent a strongly emerging field of research that will soon culminate into several ruthenium based antitumor agents.
APA, Harvard, Vancouver, ISO, and other styles
11

Sharutin, V., O. Sharutina, and V. Senchurin. "Synthesis and Structure of Ruthenium Complexes [Ph4P][trans-RuCl4(dmso-S)2] and [Ph4Sb(dmso-O)][trans-RuCl4(dmso-S)2]." «Bulletin of the South Ural State University series "Chemistry"» 9, no. 2 (2017): 58–64. http://dx.doi.org/10.14529/chem170208.

Full text
APA, Harvard, Vancouver, ISO, and other styles
12

Zimmermann, Teresa K., Stefan Haslinger, Alexander Pöthig, and Fritz E. Kühn. "Structure and catalytic activity of the ruthenium(I) sawhorse-type complex [Ru2{μ,η2-CF3(CF2)5COO}2(DMSO)2(CO)4]." Acta Crystallographica Section C Structural Chemistry 70, no. 4 (March 29, 2014): 384–87. http://dx.doi.org/10.1107/s2053229614006354.

Full text
Abstract:
The title compound,cis-di-μ-perfluoroheptanoato-κ4O:O′-bis[dicarbonyl(dimethyl sulfoxide-κS)ruthenium(I)](Ru—Ru), [Ru2(C7F13O2)2(C2H6OS)2(CO)4], is a sawhorse-type dinuclear ruthenium complex with two bridging perfluoroheptanoate ligands, and with two dimethyl sulfoxide (DMSO) ligands in the axial positions coordinatingviathe S atoms. It is a new example of a compound with an aliphatic fluorinated carboxylate ligand. The Ru—Ru bond distance of 2.6908 (3) Å indicates a direct Ru—Ru interaction. The compound is an active catalyst in transvinylation of propionic acid with vinyl acetate.
APA, Harvard, Vancouver, ISO, and other styles
13

Osman, Ahmed H., Waleed A. El-Said, and M. Abd El-Shakour. "Synthesis and characterization of some new ruthenium (II) complexes as photosensitizers in dye-sensitized solar cells." JOURNAL OF ADVANCES IN CHEMISTRY 12, no. 5 (July 21, 2017): 4413–26. http://dx.doi.org/10.24297/jac.v12i5.6265.

Full text
Abstract:
New ruthenium (II) complexes, [Ru(DHZ)2(bpy)], [Ru(SCN)2(bpy)(DMSO)2], [Ru(SCN)2(dmbpy)(DMSO)2] and [RuCl2(salen)]-2, where bpy = 2,2'- bipyridine, DHZ = 1,5-diphenylthiocarbazone, dmbpy = 4,4'-dimethyl-2,2' bipyridine and salen = 2,2'- ethylenebis(nitrilomethylidene)diphenol were synthesized and characterized by elemental analysis, FTIR, UV-Vis spectroscopy and thermal analysis. From data of these investigations the structural formula and the mode of bonding were obtained. These complexes were successfully applied to sensitization of nano-crystalline TiO2 based solar cells (DSSCs). The photovoltaic efficiencies of the studied DSSCs increase in the following order [Ru(DHZ)2(bpy)]< [Ru(SCN)2(bpy)(DMSO)2]< [Ru(SCN)2(dmbpy)(DMSO)2]< [RuCl2(salen)]-2. This increase is in agreement with the light harvesting of these complexes as indicated from their absorption spectra. Ferrioxalate complex enhanced the performance of some investigated cells. Therefore, a mechanism of this improvement has been postulated. Polyaniline as well as iodine doped polyaniline modified FTO electrode has been tested as promising counter electrodes. The efficiencies of the cells using iodine doped polyaniline is higher than that of polyaniline, which is assignable to the high conductivity of iodine.
APA, Harvard, Vancouver, ISO, and other styles
14

Ducrocq, Guillaume P., Juan A. Estrada, Joyce S. Kim, and Marc P. Kaufman. "Blocking the transient receptor potential vanilloid-1 does not reduce the exercise pressor reflex in healthy rats." American Journal of Physiology-Regulatory, Integrative and Comparative Physiology 317, no. 4 (October 1, 2019): R576—R587. http://dx.doi.org/10.1152/ajpregu.00174.2019.

Full text
Abstract:
Controversy exists regarding the role played by transient receptor potential vanilloid-1 (TRPV1) in evoking the exercise pressor reflex. Here, we determine the role played by TRPV1 in evoking this reflex while assessing possible confounding factors arising from TRPV1 antagonists or from the vehicle in which they were dissolved. The exercise pressor reflex was evoked in decerebrated, anesthetized Sprague-Dawley rats by electrical stimulation of the tibial nerve to contract the triceps surae muscles statically. This procedure was repeated before and after injection of the TRPV1 blockers: capsazepine (100 μg/100 μL), ruthenium red (100 μg/100 μL), or iodoresiniferatoxin (IRTX; 1 μg/100 μL). We found that capsazepine decreased the exercise pressor reflex when the drug was dissolved in DMSO (−10 ± 9 mmHg; P = 0.015; n = 7). However, similar reduction was found when DMSO alone was injected (−8 ± 5 mmHg; P = 0.023; n = 5). Capsazepine, dissolved in ethanol (2 ± 6 mmHg; P = 0.49; n = 7), ruthenium red (−4 ± 12 mmHg; P = 0.41; n = 7), or IRTX (4 ± 18 mmHg; P = 0.56; n = 7), did not significantly decrease the exercise pressor reflex. In addition, we found that capsazepine and ruthenium red had “off-target” effects. Capsazepine decreased the pressor response evoked by intra-arterial injection of bradykinin (500 ng/kg; −12 ± 13 mmHg; P = 0.028; n = 9) and α-β-methylene ATP (10 μg/kg; −7 ± 8 mmHg; P = 0.019; n = 10), whereas ruthenium red decreased the ability of the muscle to produce and sustain force (−99 ± 83 g; P = 0.020; n = 7). Our data therefore suggest that TRPV1 does not play a role in evoking the exercise pressor reflex. Additionally, given their strong off-target effects, capsazepine and ruthenium red should not be used for studying the role played by TRPV1 in evoking the exercise pressor reflex.
APA, Harvard, Vancouver, ISO, and other styles
15

Bi, Lihua, Firasat Hussain, Ulrich Kortz, Masahiro Sadakane, and Michael H. Dickman. "A novel isopolytungstate functionalized by ruthenium: [HW9O33RuII2(dmso)6]7?" Chemical Communications, no. 12 (2004): 1420. http://dx.doi.org/10.1039/b403902e.

Full text
APA, Harvard, Vancouver, ISO, and other styles
16

Schleicher, David, Hendrik Leopold, and Thomas Strassner. "Ruthenium(II) DMSO complexes with CˆC* cyclometalated phenylimidazol NHC ligands." Journal of Organometallic Chemistry 829 (February 2017): 101–7. http://dx.doi.org/10.1016/j.jorganchem.2016.10.036.

Full text
APA, Harvard, Vancouver, ISO, and other styles
17

Chan, P. K. L., K. A. Skov, B. R. James, and N. P. Farrell. "A new ruthenium radiosensitizer: RuC12(DMSO)2(4-Nitroimidazole)2." International Journal of Radiation Oncology*Biology*Physics 12, no. 7 (July 1986): 1059–62. http://dx.doi.org/10.1016/0360-3016(86)90225-7.

Full text
APA, Harvard, Vancouver, ISO, and other styles
18

Rohmann, Kai, Jens Kothe, Matthias W. Haenel, Ulli Englert, Markus Hölscher, and Walter Leitner. "Hydrogenation of CO2to Formic Acid with a Highly Active Ruthenium Acriphos Complex in DMSO and DMSO/Water." Angewandte Chemie International Edition 55, no. 31 (June 30, 2016): 8966–69. http://dx.doi.org/10.1002/anie.201603878.

Full text
APA, Harvard, Vancouver, ISO, and other styles
19

Rohmann, Kai, Jens Kothe, Matthias W. Haenel, Ulli Englert, Markus Hölscher, and Walter Leitner. "Hydrogenation of CO2to Formic Acid with a Highly Active Ruthenium Acriphos Complex in DMSO and DMSO/Water." Angewandte Chemie 128, no. 31 (June 30, 2016): 9112–15. http://dx.doi.org/10.1002/ange.201603878.

Full text
APA, Harvard, Vancouver, ISO, and other styles
20

Sava, G., G. Salerno, A. Bergamo, M. Cocchietto, R. Gagliardi, E. Alessio, and G. Mestroni. "Reduction of Lung Metastases by Na[trans-RuCl4(DMSO)Im] is not Coupled With the Induction of Chemical Xenogenization." Metal-Based Drugs 3, no. 2 (January 1, 1996): 67–73. http://dx.doi.org/10.1155/mbd.1996.67.

Full text
Abstract:
The effects of the treatment of tumor cells of MCa mammary carcinoma and TLX5 lymphoma with the ruthenium complex Na[trans-RuCl4(DMSO)lm] for several transplant generations were studied on tumor growth and metastases formation. On TLX5 lymphoma cells, treatment was performed in vitro prior to in vivo inoculation of tumor cells in intact or immunesuppressed mice. Either considering tumor take and growth or its capacity to invade the brain of the inoculated hosts, Na[trans-RuCl4(DMSO)lm] did not induce any significant modification. Conversely, in mice with MCa mammary carcinoma, the in vivo treatment of tumor cells in immunesuppressed hosts caused a progressive increase of DNA activity and, starting from the 4th transplant generation, a significantly increased susceptibility of lung metastasis formation to a further treatment in intact mice. These data seem to suggest that Na[trans-RuCl4(DMSO)Im] does not induce chemical xenogenization of tumor cells nor its repeated treatment induces resistance in tumor cells. Conversely, it appears that Na[trans-RuCl4(DMSO)lm] may select a tumor cell population which maintains its capacity to metastasise to the lung but with enhanced sensitivity to the antimetastatic properties of this compound.
APA, Harvard, Vancouver, ISO, and other styles
21

Zhang, Si-Qi, Li-Hua Gao, Hua Zhao, and Ke-Zhi Wang. "Recent Progress in Polynuclear Ruthenium Complex-Based DNA Binders/Structural Probes and Anticancer Agents." Current Medicinal Chemistry 27, no. 22 (June 30, 2020): 3735–52. http://dx.doi.org/10.2174/0929867326666181203143422.

Full text
Abstract:
Ruthenium complexes have stood out by several mononuclear complexes which have entered into clinical trials, such as imidazolium [trans-RuCl4(1H-imidazole)(DMSO-S)] (NAMI-A) and ([Ru(II)(4,4&#039;-dimethyl-2,2&#039;-bipyridine)2-(2(2&#039;-,2&#039;&#039;:5&#039;&#039;,2&#039;&#039;&#039;-terthiophene)-imidazo[4,5-f] [1,10]phenanthroline)] 2+) (TLD-1433), opening a new avenue for developing promising ruthenium-based anticancer drugs alternative to Cisplatin. Polynuclear ruthenium complexes were reported to exhibit synergistic and/or complementary effects: the enhanced DNA structural recognition and DNA binding as well as in vitro anticancer activities. This review overviews some representative polynuclear ruthenium complexes acting as DNA structural probes, DNA binders and in vitro anticancer agents, which were developed during last decades. These complexes are reviewed according to two main categories of homo-polynuclear and hetero-polynuclear complexes, each of which is further clarified into the metal centers linked by rigid and flexible bridging ligands. The perspective, challenges and future efforts for investigations into these exciting complexes are pointed out or suggested.
APA, Harvard, Vancouver, ISO, and other styles
22

Anderson, Craig, and André L. Beauchamp. "1H NMR study of the solvolysis of the paramagnetic tetrachloro-bis(imidazole)ruthenium(III) anion in water, methanol, and dimethyl sulfoxide." Canadian Journal of Chemistry 73, no. 4 (April 1, 1995): 471–82. http://dx.doi.org/10.1139/v95-062.

Full text
Abstract:
The 1H NMR signals of the Ru(III) species present in solution are considerably broadened and shifted by paramagnetism, but they can be used to follow chloride displacement in the trans-[RuCl4Im2]− ion. This anion remains predominant for several hours at room temperature in D2O, but its signals are progressively replaced by those of a monoaqua [RuCl3(D2O)Im2] complex. Over a period of days, two new sets of peaks appear, corresponding to two isomers of [RuCl2(D2O)2Im2]+. The same behaviour is observed for the 1-methyl-and 4-methylimidazole analogues. These reactions can be driven backwards by addition of KCl, but [RuCl4Im2]− is not quantitatively regenerated in solution even for 6 M NaCl. Within several months, the [RuCl2(D2O)2Im2]+ isomers further aquate to a single species [RuCl(D2O)3Im2]2+. In CD3OD, displacement of the first chloride of [RuCl4Im2]− takes place faster, over several hours, but substitution stops at the [RuCl3(CD3OD)Im2] stage. In DMSO, substitution occurs very slowly. The [RuCl3(DMSO)Im2]:[RuCl4Im2]−mixture (1:2) obtained after 12 days starts to show very slow reduction to two Ru(II) species, one of which precipitates as yellow crystals. From X-ray diffraction work (monoclinic, P21/n, a = 9.951, b = 8.564, c = 10.527 Å, β = 92.95°, R = 0.033), the compound was identified as [RuCl2(DMSO-d6)2Im2], where the metal has a trans-trans-trans coordination and the DMSO ligands are S-bonded. Keywords: paramagnetic ruthenium anion, solvolysis, chloro complexes.
APA, Harvard, Vancouver, ISO, and other styles
23

Sarma, Uma Charan, K. P. Sarma, and Raj K. Poddar. "Complexes of Ruthenium(III) with dimethylsulphoxide—I. [Ru2Cl6(DMSO)4], fac and mer isomers of [RuCl3(DMSO)3]: Versatile starting materials for the synthesis of ruthenium(III) complexes." Polyhedron 7, no. 18 (January 1988): 1727–35. http://dx.doi.org/10.1016/s0277-5387(00)80404-0.

Full text
APA, Harvard, Vancouver, ISO, and other styles
24

Batista, Alzir A., Salete L. Queiroz, Peter C. Healy, Robbie W. Buckley, Sue E. Boyd, Susan J. Berners-Price, Eduardo E. Castellano, and Javier Ellena. "A novel coordination mode for a pyridylphosphine ligand. X-ray structures of [RuCl2(NO)L] (I) and [RuCl2(NO)L]·DMSO (II) (L = [(2-py)2PC2H4POO(2-py)2]-)." Canadian Journal of Chemistry 79, no. 5-6 (May 1, 2001): 1030–35. http://dx.doi.org/10.1139/v01-038.

Full text
Abstract:
The ruthenium(II) complex, [RuCl2(NO)L] (I), (L = [(2-py)2PC2H4PO2(2-py)]-) was obtained from recrystallization of RuCl3NO(d2pype) (d2pype = (2-py)2PC2H4P(2-py)2) in the presence of HNO3, crystallizing in the monoclinic space group P21 (no. 4), with a = 8.012(4) Å, b = 14.454(4) Å, c = 9.353(3) Å, β = 105.77(3)°, and Z = 2. Crystals of the DMSO solvate of the complex (II) were obtained from (CD3)2SO solution, crystallizing in the monoclinic space group P21/c (no.14) with a = 9.7080(2) Å, b = 22.2920(5) Å, c = 11.5230(3) Å, β = 92.0450(10)°, and Z = 4. In both complexes, the geometry about the ruthenium atom is a distorted octahedron mainly as a result of the tridentate [P,N,O]-bonding mode of L. The ν (NO) bands at 1875 cm–1 in both complexes are consistent with the linear disposition of the NO group and the Ru atom as is observed in the X-ray crystal structure (Ru-N1-O1 angle = 178.5(4)°).Key words: pyridylphosphine, nitrosyl, ruthenium complex, X-ray structure.
APA, Harvard, Vancouver, ISO, and other styles
25

Coluccia, M., G. Sava, G. Salerno, A. Bergamo, S. Pacor, G. Mestroni, and E. Alessio. "Efficacy of 5-FU Combined to Na[trans-RuCl4(DMSO)Im] , A Novel Selective Antimetastatic Agent, on the Survival Time of Mice With P388 Leukemia, P388/DDP subline and MCa Mammary Carcinoma." Metal-Based Drugs 2, no. 4 (January 1, 1995): 195–99. http://dx.doi.org/10.1155/mbd.1995.195.

Full text
Abstract:
The combinational treatment between the selective antimetastatic agent, sodium-trans-rutheniumtetrachloridedimethylsulfoxideimidazole, Na[trans-RuCl4(DMSO)Im], and the cytotoxic drug 5-fluorouracil (5-FU) on primary tumor growth and on the survival time of experimental tumors results in an effect significantly greater than that of each single agent used alone either with the solid metastasizing MCa mammary carcinoma of the CBA mouse or with the lymphocytic leukemia P388 and its platinum resistant P388/DDP subline. Thus the inorganic compound Na[trans-RuCl4(DMSO)Im], known for its potent and selective antimetastatic effects, positively interacts with the antitumor action of an organic anticancer agent such as 5-FU on both a solid metastasizing tumor and a tumor of lymphoproliferative type. In particular, the effects of the combinational treatment on the survival time of tumor bearing mice seem to be related to the selective antimetastatic activity of the ruthenium complex that joins the potent cytotoxicity of 5-FU for the tumor. Moreover, these data show that Na[trans-RuCl4(DMSO)Im] is almost as effective on the subline of P388 made resistant to cisplatin as it was on the parental line.
APA, Harvard, Vancouver, ISO, and other styles
26

Alessio, Enzo, and Luigi Messori. "NAMI-A and KP1019/1339, Two Iconic Ruthenium Anticancer Drug Candidates Face-to-Face: A Case Story in Medicinal Inorganic Chemistry." Molecules 24, no. 10 (May 24, 2019): 1995. http://dx.doi.org/10.3390/molecules24101995.

Full text
Abstract:
NAMI-A ((ImH)[trans-RuCl4(dmso-S)(Im)], Im = imidazole) and KP1019/1339 (KP1019 = (IndH)[trans-RuCl4(Ind)2], Ind = indazole; KP1339 = Na[trans-RuCl4(Ind)2]) are two structurally related ruthenium(III) coordination compounds that have attracted a lot of attention in the medicinal inorganic chemistry scientific community as promising anticancer drug candidates. This has led to a considerable amount of studies on their respective chemico-biological features and to the eventual admission of both to clinical trials. The encouraging pharmacological performances qualified KP1019 mainly as a cytotoxic agent for the treatment of platinum-resistant colorectal cancers, whereas the non-cytotoxic NAMI-A has gained the reputation of being a very effective antimetastatic drug. A critical and strictly comparative analysis of the studies conducted so far on NAMI-A and KP1019 allows us to define the state of the art of these experimental ruthenium drugs in terms of the respective pharmacological profiles and potential clinical applications, and to gain some insight into the inherent molecular mechanisms. Despite their evident structural relatedness, deeply distinct biological and pharmacological profiles do emerge. Overall, these two iconic ruthenium complexes form an exemplary and unique case in the field of medicinal inorganic chemistry.
APA, Harvard, Vancouver, ISO, and other styles
27

Bi, Li-Hua, Bin Wang, Guang-Feng Hou, Bao Li, and Li-Xin Wu. "A novel heptatungstovanadate fragment stabilized by organo-ruthenium group: [HVW7O28Ru(dmso)3]6−." CrystEngComm 12, no. 11 (2010): 3511. http://dx.doi.org/10.1039/c0ce00087f.

Full text
APA, Harvard, Vancouver, ISO, and other styles
28

Yamamura, Tomoya, Hiroshi Nakatsuka, Shinji Tanaka, and Masato Kitamura. "Asymmetric Hydrogenation oftert-Alkyl Ketones: DMSO Effect in Unification of Stereoisomeric Ruthenium Complexes." Angewandte Chemie International Edition 52, no. 35 (July 10, 2013): 9313–15. http://dx.doi.org/10.1002/anie.201304408.

Full text
APA, Harvard, Vancouver, ISO, and other styles
29

Ponnusamy, Selvakumar, and Narayanasamy Ramasamy. "In vitro cytotoxicity and antibacterial studies of ruthenium(II) DMSO complexes with S-allyldithiocarbazate." International Journal of Materials and Product Technology 55, no. 1/2/3 (2017): 142. http://dx.doi.org/10.1504/ijmpt.2017.084958.

Full text
APA, Harvard, Vancouver, ISO, and other styles
30

Ponnusamy, Selvakumar, and Narayanasamy Ramasamy. "In vitro cytotoxicity and antibacterial studies of ruthenium(II) DMSO complexes with S-allyldithiocarbazate." International Journal of Materials and Product Technology 55, no. 1/2/3 (2017): 142. http://dx.doi.org/10.1504/ijmpt.2017.10005667.

Full text
APA, Harvard, Vancouver, ISO, and other styles
31

Bi, Li-Hua, Ulrich Kortz, Bineta Keita, and Louis Nadjo. "The ruthenium(ii)-supported heteropolytungstates [Ru(dmso)3(H2O)XW11O39]6−(X = Ge, Si)." Dalton Trans., no. 20 (2004): 3184–90. http://dx.doi.org/10.1039/b409232e.

Full text
APA, Harvard, Vancouver, ISO, and other styles
32

Yamamura, Tomoya, Hiroshi Nakatsuka, Shinji Tanaka, and Masato Kitamura. "Asymmetric Hydrogenation of tert ‐Alkyl Ketones: DMSO Effect in Unification of Stereoisomeric Ruthenium Complexes." Angewandte Chemie 125, no. 35 (July 10, 2013): 9483–85. http://dx.doi.org/10.1002/ange.201304408.

Full text
APA, Harvard, Vancouver, ISO, and other styles
33

Alessio, E., M. Bolle, B. Milani, G. Mestroni, P. Faleschini, S. Geremia, and M. Calligaris. "Carbonyl Derivatives of Chloride-Dimethyl Sulfoxide-Ruthenium(III) Complexes: Synthesis, Crystal Structure, and Reactivity of [(DMSO)2H][trans-RuCl4(DMSO-O)(CO)] and mer,cis-RuCl3(DMSO-O)2(CO)." Inorganic Chemistry 34, no. 19 (September 1995): 4716–21. http://dx.doi.org/10.1021/ic00123a003.

Full text
APA, Harvard, Vancouver, ISO, and other styles
34

Toyama, Mari, Ryuji Suganoya, Daisuke Tsuduura, and Noriharu Nagao. "Syntheses and Crystal Structures of Mono(di-2-pyridylamine)chloro(dimethyl sulfoxide-S)ruthenium(II) Complexes [RuCl2(Hdpa)(dmso-S)2] and [RuCl(Hdpa)(dmso-O)(dmso-S)2](OTf)." Bulletin of the Chemical Society of Japan 80, no. 5 (May 15, 2007): 922–36. http://dx.doi.org/10.1246/bcsj.80.922.

Full text
APA, Harvard, Vancouver, ISO, and other styles
35

Trotter, Kasey, Navamoney Arulsamy, and Elliott Hulley. "Crystal structure ofcis,fac-{N,N-bis[(pyridin-2-yl)methyl]methylamine-κ3N,N′,N′′}dichlorido(dimethyl sulfoxide-κS)ruthenium(II)." Acta Crystallographica Section E Crystallographic Communications 71, no. 9 (August 22, 2015): m169—m170. http://dx.doi.org/10.1107/s2056989015014875.

Full text
Abstract:
The reaction of dichloridotetrakis(dimethyl sulfoxide)ruthenium(II) withN,N-bis[(pyridin-2-yl)methyl]methylamine affords the title complex, [RuCl2(C13H15N3)(C2H6OS)]. The asymmetric unit contains a well-ordered complex molecule. TheN,N-bis[(pyridin-2-yl)methyl]methylamine (bpma) ligand binds the cation through its two pyridyl N atoms and one aliphatic N atom in a facial manner. The coordination sphere of the low-spind6RuIIis distorted octahedral. The dimethyl sulfoxide (dmso) ligand coordinates to the cation through its S atom and iscisto the aliphatic N atom. The two chloride ligands occupy the remaining sites. The bpma ligand is folded with the dihedral angle between the mean planes passing through its two pyridine rings being 64.55 (8)°. The two N—Ru—N bite angles of the ligand at 81.70 (7) and 82.34 (8)° illustrate the distorted octahedral coordination geometry of the RuIIcation. Two neighboring molecules are weakly associated through mutual intermolecular hydrogen bonding involving the O atom and one of the methyl groups of the dmso ligand. One of the chloride ligands is also weakly hydrogen bonded to a pyridyl H atom of another molecule.
APA, Harvard, Vancouver, ISO, and other styles
36

Yanagisawa, Masaru, Ferenc Korodi, Jianjun He, Licheng Sun, Villy Sundström, and Björn Åkermark. "Ruthenium phthalocyanines with axial carboxylate ligands: Synthesis and function in solar cells based on nanocrystalline TiO2." Journal of Porphyrins and Phthalocyanines 06, no. 03 (March 2002): 217–24. http://dx.doi.org/10.1142/s1088424602000257.

Full text
Abstract:
The synthesis and characterization of phthalocyaninato-ruthenium ( PcRu ) complexes with potential functional axial ligands are described. The solubility of these PcRu complexes was much improved compared to their parent phthalocyanines without Ru , enabling purification by normal flash column chromatography and also NMR measurements in common solvents (e. g. DMSO - d 6 and CDCl 3). Adsorption of these phthalocyanine dyes onto the surface of a semiconductor through the carboxyl group(s) in the axial ligands prevents to some extent formation of H-aggregates, which is typical for phthalocyanines. It also prevents stacking of the dye molecules on the surface. The photovoltaic behavior of sandwich solar cells based on nanostructured TiO 2 films sensitized by these PcRu complexes was studied under irradiation with visible light. For a solar cell based on bis(4-carboxypyridine)-phthalocyaninato ruthenium(II) (1) sensitized nanoporous-nanocrystalline TiO 2, a monochromatic incident photon-to-current conversion efficiency (IPCE) of 21% was obtained at 640 nm. The overall conversion efficiency (η) was 0.61%, which is one of the best results for a solar cell based on a phthalocyanine dye. For a cell based on (4-carboxypyridine)-(4-(2-ethoxy)ethyloxycarbo-nylpyridine)-2,3,9,10,16,17,23,24-octa(n-pentyloxy)-phthalocyaninato ruthenium(II) (5) sensitized TiO 2, a IPCE of 6.6% at 640 nm and η of 0.58% were obtained.
APA, Harvard, Vancouver, ISO, and other styles
37

Cruz, Thais R., Rodolpho A. N. Silva, Antonio E. H. Machado, Benedito S. Lima-Neto, Beatriz E. Goi, and Valdemiro P. Carvalho. "New dmso–ruthenium catalysts bearing N-heterocyclic carbene ligands for the ring-opening metathesis of norbornene." New Journal of Chemistry 43, no. 16 (2019): 6220–27. http://dx.doi.org/10.1039/c9nj00810a.

Full text
APA, Harvard, Vancouver, ISO, and other styles
38

Alagesan, M., P. Sathyadevi, P. Krishnamoorthy, N. S. P. Bhuvanesh, and N. Dharmaraj. "DMSO containing ruthenium(ii) hydrazone complexes: in vitro evaluation of biomolecular interaction and anticancer activity." Dalton Trans. 43, no. 42 (September 16, 2014): 15829–40. http://dx.doi.org/10.1039/c4dt01032a.

Full text
APA, Harvard, Vancouver, ISO, and other styles
39

Wang, Lei, Lele Duan, Beverly Stewart, Maoping Pu, Jianhui Liu, Timofei Privalov, and Licheng Sun. "Toward Controlling Water Oxidation Catalysis: Tunable Activity of Ruthenium Complexes with Axial Imidazole/DMSO Ligands." Journal of the American Chemical Society 134, no. 45 (November 5, 2012): 18868–80. http://dx.doi.org/10.1021/ja309805m.

Full text
APA, Harvard, Vancouver, ISO, and other styles
40

Turel, Iztok, Milena Pečanac, Amalija Golobič, Enzo Alessio, Barbara Serli, Alberta Bergamo, and Gianni Sava. "Solution, solid state and biological characterization of ruthenium(III)-DMSO complexes with purine base derivatives." Journal of Inorganic Biochemistry 98, no. 2 (February 2004): 393–401. http://dx.doi.org/10.1016/j.jinorgbio.2003.12.001.

Full text
APA, Harvard, Vancouver, ISO, and other styles
41

Tan, Caiping, Sheng Hu, Jie Liu, and Liangnian Ji. "Synthesis, characterization, antiproliferative and anti-metastatic properties of two ruthenium–DMSO complexes containing 2,2′-biimidazole." European Journal of Medicinal Chemistry 46, no. 5 (May 2011): 1555–63. http://dx.doi.org/10.1016/j.ejmech.2011.01.074.

Full text
APA, Harvard, Vancouver, ISO, and other styles
42

Demoro, Bruno, Andreia Bento-Oliveira, Fernanda Marques, João Costa Pessoa, Lucía Otero, Dinorah Gambino, Rodrigo F. M. de Almeida, and Ana Isabel Tomaz. "Interaction with Blood Proteins of a Ruthenium(II) Nitrofuryl Semicarbazone Complex: Effect on the Antitumoral Activity." Molecules 24, no. 16 (August 7, 2019): 2861. http://dx.doi.org/10.3390/molecules24162861.

Full text
Abstract:
The steady rise in the cancer burden and grim statistics set a vital need for new therapeutic solutions. Given their high efficiency, metallodrugs are quite appealing in cancer chemotherapy. This work examined the anticancer activity of an anti-trypanosomal ruthenium-based compound bearing the 5-nitrofuryl pharmacophore, [RuII(dmso)2(5-nitro-2-furaldehyde semicarbazone)] (abbreviated as RuNTF; dmso is the dimethyl sulfoxide ligand). The cytotoxicity of RuNTF was evaluated in vitro against ovarian adenocarcinoma, hormone-dependent breast adenocarcinoma, prostate carcinoma (grade IV) and V79 lung fibroblasts human cells. The activity of RuNTF was similar to the benchmark metallodrug cisplatin for the breast line and inactive against the prostate line and lung fibroblasts. Given the known role of serum protein binding in drug bioavailability and the distribution via blood plasma, this study assessed the interaction of RuNTF with human serum albumin (HSA) by circular dichroism (CD) and fluorescence spectroscopy. The fluorescence emission quenching from the HSA-Trp214 residue and the lifetime data upon RuNTF binding evidenced the formation of a 1:1 {RuNTF-albumin} adduct with log Ksv = (4.58 ± 0.01) and log KB = (4.55 ± 0.01). This is supported by CD data with an induced CD broad band observed at ~450 nm even after short incubation times. Importantly, the binding to either HSA or human apo-transferrin is beneficial to the cytotoxicity of the complex towards human cancer cells by enhancing the cytotoxic activity of RuNTF.
APA, Harvard, Vancouver, ISO, and other styles
43

Mondal, Ashaparna, and Priyankar Paira. "Synthesis and Biological Evaluations of Organoruthenium Scaffolds: A Comprehensive Update." Current Organic Synthesis 15, no. 2 (April 24, 2018): 179–207. http://dx.doi.org/10.2174/1570179414666170703143049.

Full text
Abstract:
Background: Currently ruthenium complexes are immerging as effective anticancer agents due to their less toxicity, better antiproliferative and antimetastatic activity, better stability in cellular environment and most importantly variable oxidation and co-ordination states of ruthenium allows binding this molecule with a variety of ligands. So in past few years researchers have shifted their interest towards organoruthenium complexes having good fluorescent profile that may be applicable for cancer theranostics. Nowadays, photodynamic therapy has become more acceptable because of its easy and effective approach towards killing cancer cells. Objective: Objective of this review article is to shed light on synthesis, characterization, stability and fluorescence studies of various ruthenium [Ru(II) and Ru(III)] complexes and different bioactivity studies conducted with the synthesized compounds to test their candidacy as potent chemotherapeutic agents. Methods: Various heterocyclic ligands containing N,O and S as heteroatom mainly were prepared and subjected to complexation with ruthenium-p-cymene moiety. In most cases [Ru(η6-p-cymene)(µ-Cl)Cl]2 was used as ruthenium precursor and the reactions were conducted in various alcohol medium such as methanol, ethanol or propanol. The synthesized complexes were characterized by 1H NMR and 13C NMR spectroscopy, GC-MS, ESI-MS, elemental analysis and single crystal X-ray crystallography methods. Fluorescence study and stability study were conducted accordingly using water, PBS buffer or DMSO. Stable compounds were considered for cell viability studies. To study the efficacy of the compounds in ROS generation as photosensitizers, in few cases, singlet oxygen quantum yields in presence of light were calculated. Suitable compounds were selected for in vitro & in vivo antiproliferative, anti-invasive activity studies. Result: Many newly synthesized compounds were found to have less IC50 compared to a standard drug cysplatin. Those compounds were also stable preferably in physiological conditions. Good fluorescence profile and ROS generation ability were observed for few compounds. Conclusion: Numerous ruthenium complexes were developed which can be used as cancer theranostic agents. Few molecules were synthesized as photosensitizers which were supposed to generate reactive singlet oxygen species in targeted cellular environment in presence of a particular type of light and thereby ceasing cancer cell growth.
APA, Harvard, Vancouver, ISO, and other styles
44

Elsayed, Shadia A., Shane Harrypersad, Heba A. Sahyon, Mohammed Abu El-Magd, and Charles J. Walsby. "Ruthenium(II)/(III) DMSO-Based Complexes of 2-Aminophenyl Benzimidazole with In Vitro and In Vivo Anticancer Activity." Molecules 25, no. 18 (September 18, 2020): 4284. http://dx.doi.org/10.3390/molecules25184284.

Full text
Abstract:
New anticancer ruthenium(II/III) complexes [RuCl2(DMSO)2(Hapbim)] (1) and [RuCl3(DMSO) (Hapbim)] (2) (Hapbim = 2-aminophenyl benzimidazole) have been synthesized and characterized, and their chemotherapeutic potential evaluated. The interaction of the compounds with DNA was studied by both UV-Visible and fluorescence spectroscopies, revealing intercalation of both the Hapbim ligand and the Ru complexes. The in vitro cytotoxicity of the compounds was tested on human breast cancer (MCF7), human colorectal cancer (Caco2), and normal human liver cell lines (THLE-2), with compound (2) the most potent against cancer cells. The cytotoxic effect of (2) is shown to correlate with the ability of the Ru(III) complex to induce apoptosis and to cause cell-cycle arrest in the G2/M phase. Notably, both compounds were inactive in the noncancerous cell line. The anticancer effect of (2) has also been studied in an EAC (Ehrlich Ascites Carcinoma) mouse model. Significantly, the activity of the complex was more pronounced in vivo, with removal of the cancer burden at doses that resulted in only low levels of hepatotoxicity and nephrotoxicity. An apoptosis mechanism was determined by the observation of increased Bax and caspase 3 and decreased Bcl2 expression. Furthermore, (2) decreased oxidative stress and increased the levels of antioxidant enzymes, especially SOD, suggesting the enhancement of normal cell repair. Overall, compound (2) shows great potential as a chemotherapeutic candidate, with promising activity and low levels of side effects.
APA, Harvard, Vancouver, ISO, and other styles
45

Prajapati, Rishikesh, Santosh Kumar Dubey, Ruchi Gaur, Raj Kumar Koiri, Brajesh Kumar Maurya, Surendra Kumar Trigun, and Lallan Mishra. "Structural characterization and cytotoxicity studies of ruthenium(II)–dmso–chloro complexes of chalcone and flavone derivatives." Polyhedron 29, no. 3 (February 2010): 1055–61. http://dx.doi.org/10.1016/j.poly.2009.11.012.

Full text
APA, Harvard, Vancouver, ISO, and other styles
46

Bergamo, Alberta, Moreno Cocchietto, Ilaria Capozzi, Giovanni Mestroni, Enzo Alessio, and Gianni Sava. "Treatment of residual metastases with Na[trans-RuCl4(DMSO)lm] and ruthenium uptake by tumor cells." Anti-Cancer Drugs 7, no. 6 (August 1996): 697–702. http://dx.doi.org/10.1097/00001813-199608000-00011.

Full text
APA, Harvard, Vancouver, ISO, and other styles
47

Sheeba, Daniel, and Allen Gnana Raj George. "Luminescence quenching of tris(4,4′-dinonyl-2,2′-bipyridyl) ruthenium(II) cation with phenolate ions in DMSO." Arabian Journal of Chemistry 10 (May 2017): S2429—S2435. http://dx.doi.org/10.1016/j.arabjc.2013.09.006.

Full text
APA, Harvard, Vancouver, ISO, and other styles
48

Yamamura, Tomoya, Hiroshi Nakatsuka, Shinji Tanaka, and Masato Kitamura. "ChemInform Abstract: Asymmetric Hydrogenation of tert-Alkyl Ketones: DMSO Effect in Unification of Stereoisomeric Ruthenium Complexes." ChemInform 45, no. 3 (January 2, 2014): no. http://dx.doi.org/10.1002/chin.201403052.

Full text
APA, Harvard, Vancouver, ISO, and other styles
49

Sharutin, V. V., O. K. Sharutina, V. S. Senchurin, and S. A. Sobalev. "Ruthenium Complexes [Ph3PCH2CH=CHCH2PPh3]2+[trans-RuCl4(Dmso)2] 2 − and [Ph3PR]+[trans-RuCl4(Dmso)2]–, R = CH2C6H4CN-4, CH2Ph, CPh3, Ph, CH2OCH3: Synthesis and Structure." Russian Journal of Inorganic Chemistry 63, no. 1 (January 2018): 48–54. http://dx.doi.org/10.1134/s0036023618010151.

Full text
APA, Harvard, Vancouver, ISO, and other styles
50

Zhang, Junda, Vadde Ramu, Xue-Quan Zhou, Carolina Frias, Daniel Ruiz-Molina, Sylvestre Bonnet, Claudio Roscini, and Fernando Novio. "Photoactivable Ruthenium-Based Coordination Polymer Nanoparticles for Light-Induced Chemotherapy." Nanomaterials 11, no. 11 (November 16, 2021): 3089. http://dx.doi.org/10.3390/nano11113089.

Full text
Abstract:
Green light photoactive Ru-based coordination polymer nanoparticles (CPNs), with chemical formula [[Ru(biqbpy)]1.5(bis)](PF6)3 (biqbpy = 6,6′-bis[N-(isoquinolyl)-1-amino]-2,2′-bipyridine; bis = bis(imidazol-1-yl)-hexane), were obtained through polymerization of the trans-[Ru(biqbpy)(dmso)Cl]Cl complex (Complex 1) and bis bridging ligands. The as-synthesized CPNs (50 ± 12 nm diameter) showed high colloidal and chemical stability in physiological solutions. The axial bis(imidazole) ligands coordinated to the ruthenium center were photosubstituted by water upon light irradiation in aqueous medium to generate the aqueous substituted and active ruthenium complexes. The UV-Vis spectral variations observed for the suspension upon irradiation corroborated the photoactivation of the CPNs, while High Performance Liquid Chromatography (HPLC) of irradiated particles in physiological media allowed for the first time precisely quantifying the amount of photoreleased complex from the polymeric material. In vitro studies with A431 and A549 cancer cell lines revealed an 11-fold increased uptake for the nanoparticles compared to the monomeric complex [Ru(biqbpy)(N-methylimidazole)2](PF6)2 (Complex 2). After irradiation (520 nm, 39.3 J/cm2), the CPNs yielded up to a two-fold increase in cytotoxicity compared to the same CPNs kept in the dark, indicating a selective effect by light irradiation. Meanwhile, the absence of 1O2 production from both nanostructured and monomeric prodrugs concluded that light-induced cell death is not caused by a photodynamic effect but rather by photoactivated chemotherapy.
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography